NeuroPsychopharmacotherapy (Nov 5, 2022) - (3030620581) - (Springer)
NeuroPsychopharmacotherapy (Nov 5, 2022) - (3030620581) - (Springer)
Gerd Laux
Toshiharu Nagatsu
Weidong Le
Christian Riederer
Editors
NeuroPsychopharmacotherapy
NeuroPsychopharmacotherapy
Peter Riederer • Gerd Laux •
Toshiharu Nagatsu • Weidong Le •
Christian Riederer
Editors
NeuroPsychopharmacotherapy
Christian Riederer
Würzburg, Germany
This Springer imprint is published by the registered company Springer Nature Switzerland AG.
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Preface
Psychiatric disorders are among the most common diseases worldwide, and in many
countries, neuropsychopharmaka are among the most prescribed drugs (EMA,
FDA).
Starting in the early 1990s, Peter Riederer, Gerd Laux, and Walter Pöldinger
edited, in the German language, a series of handbooks covering (1) basic aspects of
psychopharmacological drugs, (2) tranquilizers and hypnotics, (3) antidepressants,
(4) antipsychotics, (5) drugs for Parkinson’s disease and “Nootropics,” and (6) anti-
epileptics, β-blocker, and other psychopharmacological drugs. Our goal was to
provide state-of-the-art basic knowledge, current diagnostic and differential diag-
nostic methodologies, and up-to-date therapeutic strategies.
For some years, however, fundamental scientific, political, economic, and social
dynamics have substantially influenced both the development of new drugs for
neuropsychiatric disorders and their acceptance by authorities, doctors, and patients.
The reasons for it are manifold:
v
vi Preface
focus on areas like cancer and cardiovascular disorders which promise much
higher financial revenues.
3. We also see today a much livelier field of neuropsychopharmacotherapy than
30 years ago. In a changing social environment, more reservations toward the role
of pharmaceutical companies, their financing of clinical studies, their cooperation
with physicians, and their marketed drugs themselves, exist. The balance has
shifted from a drug-oriented therapy to psychotherapy and social psychiatry.
However, in our opinion, both poles are necessary for the optimal treatment
strategy in most patients.
Volume 2 has a strong clinical focus, and classes, drugs, and special aspects of the
role of neuropsychopharmacological substances in psychiatry and neurology in
10 major neurological and psychiatric disorders.
Volume 3 contains 34 selected overview chapters on topic of applied neuropsy-
chopharmacotherapy as well as issues related to the implementation of
neuropsychopharmacotherapy.
We would like to express our sincere gratitude to our section editors as well as to
all authors, without their support it would not have been possible to publish this
handbook. We would also like to thank Springer Nature Publishers for their ongoing
support to publish another handbook on neuropsychopharmacology. We do hope it
will find high acceptance in the scientific community and establish itself as a
standard reference work in clinical and basic science.
The Editors would like to thank the Verein zur Durchführung Neurowis-
senschaftlicher Tagungen e.V. for their financial support in organizing this
publication.
ix
Contents
Volume 1
Pharmacovigilance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
Maike Scherf-Clavel
Pharmacoeconomics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
Richard Dodel, Christopher Kruse, Annette Conrads-Frank, and
Uwe Siebert
xi
xii Contents
Volume 2a
Volume 2b
Volume 2c
Volume 2d
Volume 3
Part XIII Applied Neuropsychopharmacotherapy . . . . . . . . . . . . . . 3887
Peter Riederer and Toshiharu Nagatsu
Gerd Laux is Director of the Institute of Psychological Medicine and the Center of
Neuropsychiatry MVZ Waldkraidburg in Bavaria and a professor at the Department
of Psychiatry and Psychotherapy at the University of Munich, Germany. He has
worked as a clinician at different hospitals at the Universities of Dallas (Texas USA)
and Wuerzburg and Bonn (Germany). He has authored more than 650 articles and
book chapters and written or edited more than 40 books in the field of psychiatry and
psychopharmacology. He has been principal investigator of various clinical trials
and serves as a consultant for scientific institutions, a Geriatric Hospital and is editor-
in-chief of the journal Psychopharmakotherapie.
Otto Dietmaier Ltd. Pharm. Dir. i.R. Zentren für Psychiatrie Weinsberg, Wiesloch,
Winnenden, Aulendorf, Germany
Toshiharu Nagatsu Centre for Research Promotion and Support, Fujita Health
University, Toyoake, Aichi, Japan
xxxi
xxxii Section Editors
Peter Riederer Clinic and Policlinic for Psychiatry, Psychosomatics and Psycho-
therapy, University Hospital Wuerzburg, Wuerzburg, Germany
University of Southern Denmark Odense, Odense, Denmark
Toshikazu Saito The International College of Neuropsychopharmacology, East
Kilbride, Scotland, UK
xxxiii
xxxiv Contributors
Orit Bar-Am Eve Topf and USA National Parkinson Foundation Centers of
Excellence for Neurodegenerative Diseases and Department of Pharmacology,
Technion-Rappaport Family Faculty of Medicine, Haifa, Israel
R. Baron Division of Neurological Pain Research and Therapy, University
Hopsital Schleswig-Holstein, Campus Kiel, Kiel, Germany
Thomas Bast Epilepsy Center Kork, Kehl, Germany
Pierre Baumann University Hospital and University of Lausanne, Lausanne,
Switzerland
Christoph Baumgartner Department of Neurology, General Hospital Hietzing
with Neurological Center Rosenhügel, Vienna, Austria
Karl Landsteiner Institute for Clinical Epilepsy Research and Cognitive Neurology,
Vienna, Austria
Medical Faculty, Sigmund Freud University, Vienna, Austria
Tobias R. Baumgartner Department of Epileptology, University of Bonn Medical
Centre, Bonn, Germany
Bernhard T. Baune Department of Psychiatry, University Hospital Münster, Uni-
versity of Münster, Münster, Germany
Department of Psychiatry, Melbourne Medical School, The University of Mel-
bourne, Melbourne, VIC, Australia
The Florey Institute of Neuroscience and Mental Health, The University of Mel-
bourne, Parkville, VIC, Australia
Jens Benninghoff Psychiatrische LVR-Uni-Klinik Essen, Universität Duisburg-
Essen, Essen, Germany
Chefarzt Zentrum für Altersmedizin und Entwicklungsstörungen (ZfAE), kbo-Isar-
Amper-Klinikum München-Ost, Haar, Germany
Jessica Bentley Department of Psychological Medicine, King’s College London,
London, UK
Niels Bergemann Department of Biological and Clinical Psychology, University of
Trier, Trier, Germany
Gregor Berger Department of Child and Adolescent Psychiatry, Psychiatric Uni-
versity Hospital Zurich (PUK), Zurich, Switzerland
I. Beuchat Department of Clinical Neuroscience, Lausanne University Hospital
(CHUV) and University of Lausanne, Lausanne, Switzerland
M. Beudel Department of Neurology, Amsterdam Neuroscience Institute, Amster-
dam University Medical Center, Amsterdam, The Netherlands
Contributors xxxv
Simon J. C. Davies Centre for Addiction and Mental Health, Toronto, ON, Canada
Department of Psychiatry, University of Toronto, Toronto, ON, Canada
Otto Dietmaier Ltd. Pharm. Dir. i.R. Zentren für Psychiatrie Weinsberg, Wiesloch,
Winnenden, Aulendorf, Germany
Vera Dinkelacker Fondation Rothschild Paris, Paris, France
Hôpitaux Universitaires Strasbourg, Strasbourg, France
xxxviii Contributors
Amy Hamilton Bethlem Royal Hospital, South London and Maudsley NHS Foun-
dation Trust, London, UK
Lena Hampel Central Institute of Mental Health, Department of Translational
Brain Research, Medical Faculty Mannheim; Heidelberg University, Hector Institute
for Translational Brain Research, Mannheim, Germany
German Cancer Research Center (DKFZ), Heidelberg, Germany
Ritsuko Hanajima Division of Neurology, Department of Brain and Neurosci-
ences, Faculty of Medicine, Tottori University, Yonago/Tottori, Japan
Till Hartlieb Fachzentrum für pädiatrische Neurologie, Neuro-Rehabilitation und
Epileptologie, Schön Klinik Vogtareuth, Vogtareuth, Germany
Sarah Hassan Department of Psychology, University of Illinois at Urbana-Cham-
paign, Urbana, IL, USA
Nobutaka Hattori Department of Neurology, Juntendo University School of Med-
icine, Bunkyo, Tokyo, Japan
Martin Hatzinger Psychiatric Services Solothurn and Medical Faculty of the
University of Basel, Solothurn, Switzerland
Gudrun Hefner Psychiatric Hospital, Vitos Klinik Hochtaunus, Friedrichsdorf,
Germany
Andreas Heinz Department of Psychiatry and Psychotherapy, Charité Campus
Mitte, Charité Universitätsmedizin Berlin, Freie Universität Berlin, Humboldt-
Universität zu Berlin, and Berlin Institute of Health, Berlin, Germany
Hanfried Helmchen Department of Psychiatry & Psychotherapy, CBF, Charité –
University Medicine Berlin, Berlin, Germany
Christoph Helmstaedter Department of Epileptology, University Hospital Bonn
(UKB), Bonn, Germany
Johannes M. Hennings kbo Isar-Amper-Klinikum Munich, Haar/Munich,
Bavaria, Germany
Jonathan Henssler Department of Psychiatry and Psychotherapy, Charité Campus
Mitte, Charité Universitätsmedizin Berlin, Freie Universität Berlin, Humboldt-
Universität zu Berlin, and Berlin Institute of Health, Berlin, Germany
Sebastian Herberger ProSomno Clinic for Sleep Medicine, Munich, Germany
Charité Interdisciplinary Centre of Sleep Medicine, Berlin, Germany
Marcus Herdener Department of Psychiatry, Psychotherapy and Psychosomatics,
Psychiatric Hospital, University of Zurich, Zurich, Switzerland
xlii Contributors
Peter Riederer Clinic and Policlinic for Psychiatry, Psychosomatics and Psycho-
therapy, University Hospital Wuerzburg, Wuerzburg, Germany
University of Southern Denmark Odense, Odense, Denmark
Daniela Rodrigues-Amorim Translational Neuroscience Research Group, Galicia
Sur Health Research Institute (IISGS), CIBERSAM, Vigo, Spain
Johannes Rösche Klinik und Poliklinik für Neurologie, Universitätsmedizin Ros-
tock, Rostock, Germany
Hephata Klinik, Schwalmstadt, Germany
Felix Rosenow Epilepsy Center Frankfurt Rhine-Main, Goethe-University Frank-
furt, Frankfurt am Main, Germany
Michael Rösler University of the Saarland, Neurocentre, ADHD Research Group,
Homburg/Saar, Germany
A. O. Rossetti Department of Clinical Neuroscience, Lausanne University Hospital
(CHUV) and University of Lausanne, Lausanne, Switzerland
Stephan Röttig Department of Psychiatry, Psychotherapy and Psychosomatics,
Martin Luther University of Halle-Wittenberg, Halle/Saale, Germany
Stephan Rüegg Department of Neurology, Basel University Hospital, Basel,
Switzerland
Eckart Ruether ProSomno Clinic for Sleep Medicine, Munich, Germany
Dan Rujescu Department of Psychiatry, Psychotherapy and Psychosomatics, Mar-
tin Luther University of Halle-Wittenberg, Halle/Saale, Germany
C. Rummel-Kluge Klinik und Poliklinik für Psychiatrie und Psychotherapie,
Universitätsklinikum Leipzig, Leipzig, Germany
Hans-Jürgen Rumpf Department of Psychiatry and Psychotherapy, Translational
Psychiatry Unit, University of Lübeck, Lübeck, Germany
Janusz K. Rybakowski Department of Adult Psychiatry, Poznan University of
Medical Sciences, Poznan, Poland
J. Sachau Division of Neurological Pain Research and Therapy, University
Hopsital Schleswig-Holstein, Campus Kiel, Kiel, Germany
Josemir W. Sander NIHR University College London Hospitals Biomedical
Research Centre, UCL Queen Square Institute of Neurology, London, UK
Chalfont Centre for Epilepsy, Chalfont St. Peter, Bucks, UK
Stichting Epilepsie Instellingen Nederland (SEIN), Heemstede, The Netherlands
Abel Santamaría Laboratorio de Aminoácidos Excitadores, Instituto Nacional de
Neurología y Neurocirugía Manuel Velasco Suárez, Ciudad de México, Mexico
Contributors liii
Apitharani Santhirakumar Centre for Addiction and Mental Health (CAMH) and
Department of Psychiatry, University of Toronto, Toronto, ON, Canada
Steven C. Schachter Departments of Neurology, Beth Israel Deaconess Medical
Center, Massachusetts General Hospital and Harvard Medical School, Boston, MA,
USA
Ingo Schäfer Department of Psychiatry and Psychotherapy, Center for Interdisci-
plinary Addiction Research, University Medical Center Hamburg-Eppendorf, Ham-
burg, Germany
Maximilian Schäfer Department of Psychiatry and Psychotherapy, Campus Mitte,
Charité Universitätsmedizin Berlin, Berlin, Germany
Cora Schefft Charité – Universitätsmedizin Berlin, corporate member of Freie
Universität Berlin, Humboldt-Universität zu Berlin, and Berlin Institute of Health,
Department of Psychiatry and Neurosciences, CCM, Berlin, Germany
Maike Scherf-Clavel Department of Psychiatry, Psychosomatics and Psychother-
apy, University Hospital of Würzburg, Würzburg, Germany
Samantha Schlossarek Department of Psychiatry and Psychotherapy, Transla-
tional Psychiatry Unit, University of Lübeck, Lübeck, Germany
Frank M. Schmidt Department of Psychiatry, University Hospital Leipzig, Leip-
zig, Germany
Peter Schönknecht University Hospital Leipzig and Academic Hospital Arnsdorf,
Leipzig/Arnsdorf, Germany
Georgios Schoretsanitis Psychiatry Research, The Zucker Hillside Hospital, New
York, USA
Susanne Schubert-Bast Epilepsy Center Frankfurt Rhine-Main, Goethe-Univer-
sity Frankfurt, Frankfurt am Main, Germany
Josef Schwitzer Department of Psychiatry, General Hospital Brixen, South Tyrol,
Italy
M. Seibert Department of Neurology, University Clinic Ulm, Ulm, Germany
Geriatric Center Ulm/Alb-Donau, Ulm, Germany
Zümrüt Duygu Sen Department of Psychiatry and Psychotherapy, Jena University
Hospital, Jena, Germany
Clinical Affective Neuroimaging Laboratory (CANLAB), Magdeburg, Germany
Department of Psychiatry and Psychotherapy, University Tuebingen, Tuebingen,
Germany
Alessandro Serretti Department of Biomedical and Neuromotor Sciences, Univer-
sity of Bologna, Bologna, Italy
liv Contributors
Libor Velíšek Departments of Cell Biology & Anatomy, Pediatrics, and Neurology,
New York Medical College, Valhalla, NY, USA
Jana Velíšková Departments of Cell Biology & Anatomy, Obstetrics & Gynecol-
ogy, and Neurology, New York Medical College, Valhalla, NY, USA
Olivia Verisezan Rosu Department of Neurosciences, “Iuliu Hatieganu” Univer-
sity of Medicine and Pharmacy, Cluj-Napoca, Romania
“RoNeuro” Institute for Neurological Research and Diagnostic, Cluj-Napoca,
Romania
Sasivimol Virameteekul Chulalongkorn Centre of Excellence for Parkinson’s
Disease and Related Disorders, Department of Medicine, Faculty of Medicine,
Chulalongkorn University and King Chulalongkorn Memorial Hospital, Thai Red
Cross Society, Bangkok, Thailand
Trang N. N. Vo Neurology Department, International Neurosurgery Hospital, Ho
Chi Minh City, Vietnam
Constantin Volkmann Charité – Universitätsmedizin Berlin, corporate member of
Freie Universität Berlin, Humboldt-Universität zu Berlin, and Berlin Institute of
Health, Department of Psychiatry and Neurosciences, CCM, Berlin, Germany
Hans-Peter Volz Krankenhaus für Psychiatrie, Psychotherapie und Psychoso-
matische Medizin Schloss Werneck, Werneck, Germany
C. A. F. von Arnim Division of Geriatrics, University Medical Centre, Georg
August University, Göttingen, Germany
Elena Vos University of Amsterdam, Amsterdam, The Netherlands
Jelena Vrublevska Department of Psychiatry and Narcology, Riga Stradins Uni-
versity, Riga, Latvia
Susanne Walitza Department of Child and Adolescent Psychiatry, Psychiatric
University Hospital Zurich (PUK), Zurich, Switzerland
Henriette Walter Clinical Department of Social Psychiatry, Medical University
Vienna, Vienna, Austria
Martin Walter Department of Psychiatry and Psychotherapy, Jena University
Hospital, Jena, Germany
Clinical Affective Neuroimaging Laboratory (CANLAB), Magdeburg, Germany
Department of Psychiatry and Psychotherapy, University Tuebingen, Tuebingen,
Germany
Leibniz Institute for Neurobiology, Magdeburg, Germany
Gang Wang Department of Neurology and Neuroscience Institute, Ruijin Hospital
affiliated to Shanghai Jiao Tong University School of Medicine, Shanghai, P.R.
China
lviii Contributors
Feng Zhang The First Affiliated Hospital, Dalian Medical University, Dalian,
China
Xiaojie Zhang Department of Neurology, Shanghai 6th People’s Hospital, Shang-
hai Jiaotong University School of Medicine, Shanghai, China
Hailin Zheng Department of Medicinal Chemistry, Intra-cellular Therapies Inc.,
New York, NY, USA
Peter Zwanzger Clinical Center for Psychiatry, Psychotherapy, Psychosomatic
Medicine and Neurology, Wasserburg am Inn, Germany
Part I
Basic Principles
Peter Riederer
Neurobiological Principles:
Neurotransmitters
Contents
Neurotransmitters: More than Crossing the Synaptic Cleft . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
One Neuron: Two Messengers – Neurotransmission Beyond Dale’s Principle . . . . . . . . . . . . . . . . 7
Dopamine Neurons: A Versatile Model for Joint Neurotransmitter Release . . . . . . . . . . . . . . . . . . . . 10
Maintenance of Neuronal Network Activity: A Small Insight into Giant Domains . . . . . . . . . . . . 13
A New Frontier: Serotonylation of Histones – Neurotransmitter Activity Inside the Nucleus . . . 15
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
Abstract
To integrate sensory and memory information, to control behavior, and to regulate
various metabolic processes in favor of homeostasis, the central nervous system
has to perform a huge amount of computations at once. This task is fulfilled by
intra- and intercellular signaling cascades driven by a sophisticated network of
highly connected neural cells. Most prominently, the release of neurotransmitters
into the synaptic cleft is considered to be the key element core of neuronal
intercellular communication. This core process of neuronal communication
makes the synapse the substantial unit of neurotransmission. Consequently,
neurons are thought to establish the foundation of efficient communication within
their networks by strengthening and promoting the molecular structures of this
functional unit. In the classic view of chemical neurotransmission, the synapse
requires (1) the availability for synthesis of the respective neurotransmitter; (2) an
activity-responsive, tightly regulated vesicular release machinery on the
presynaptic side; (3) postsynaptic receptors, which are selectively activated upon
binding by neurotransmitters and subsequently trigger intracellular signaling
cascades; and (4) mechanisms to successfully terminate neurotransmission by
removal of neurotransmitters from their site of action. With advanced experimen-
tal methods and interdisciplinary research approaches, our knowledge on neuro-
transmission is constantly growing. As a result, the once classical view on
neurotransmission is changing in the light of new insights into regulation and
organization of synaptic activity. The following sections first give a brief over-
view on elements determining neurotransmission, the axon initial segment, and
presynaptic cytomatrix active zone, before focusing on simultaneous release of
two chemically distinct neurotransmitters and neurotransmitter activity beyond
synaptic structures (Fig. 1).
synaptic
transmission axo-axonic synaptic transmission
at the axon initial segment
Fig. 1 Overview on neurotransmitters’ sites of action in mammalian brain. The overview intro-
duces four sites of action for neurotransmitters: (a) Neurotransmitters act at synaptic connections
between two neurons, for example, synaptic transmission between pre- and postsynapse at nerve
terminals that connect to dendrites. After crossing the synaptic cleft, neurotransmitters bind their
respective ionotropic or metabotropic receptors to trigger cellular responses. (b) An alternative
synaptic transmission occurs at axo-axonic synapses located at the axon initial segment, the site of
action potential generation. Here, the neurotransmitters dopamine, GABA, and serotonin are able to
positively or negatively interfere with action potential generation. (c) Extrasynaptic volume trans-
mission, especially for the monoamines dopamine and serotonin, was shown to occur at
somatodendritic structures as well as along the axon. These extrasynaptic release sites do not
depend on establishing connection to an opposing postsynapse. (d) Beyond neurotransmission,
monoamine neurotransmitters are substrates for transamidation of proteins, including extracellular
matrix proteins, which affect the scaffolding of these proteins, and intracellular GTPases, directly
changing the activity of small GTPases. Recently serotonin was shown to be transamidated to
histones, which influences permissive gene expression
regulation of protein trafficking but is also the critical intracellular gate of signal
transduction that determines neuronal output. The threshold for the initiation of
action potentials depends on the variable AIS position relative to the neuron’s soma,
length, and highly adaptive ion channel composition (Leterrier 2018; Petersen et al.
2017). These characteristics make the AIS a key player in fine-tuning neuronal
excitability and thereby a crucial element of neuronal plasticity. Its structural and
biophysical properties were shown to be altered in an activity-dependent manner as
well as by neurotransmitter actions, both from the surrounding extracellular space
and from direct axo-axonic synaptic innervation by GABAergic and dopaminergic
interneurons (Jamann et al. 2018; Khirug et al. 2008; Wefelmeyer et al. 2015;
Yamada and Kuba 2016). According to its central role in generating and shaping
action potentials, it is not surprising that disruptions of AIS components and
alterations in their formation are emerging as critical factors in the pathogenesis of
several important mental and neurological disorders. For example, dysregulations of
the AIS scaffold protein ankyrin-G have been associated with schizophrenia as well
as bipolar disorder, while genetic mutations of AIS voltage-dependent sodium
6 L. Hampel and T. Lau
channels are correlated with the occurrence of autism spectrum disorders (reviewed
in Hsu et al. 2014). Given that the AIS is a key structure in maintenance of neuronal
structural and functional organization, therefore AIS components and their protein-
protein interactions are promising future targets for pharmacological treatments.
Along the axon, differences in electrical charges between the extra- and intracel-
lular side of the plasma membrane are exploited to generate currents that drive
depolarization and propagation of the action potential. Axons are highly specialized
for fast and efficient conduction. In myelinated neurons the increased electrical
isolation and clustered distribution of voltage-dependent sodium channels at
the Nodes of Ranvier allow conduction of super-threshold electrical signals to the
synapses at velocities up to 150 m/s (Purves et al. 2001). Upon arrival of the
electrical signals at the presynaptic side, a tightly regulated release machinery is
activated, which is organized by a highly specialized cytoskeleton region: the
cytomatrix active zone. The cytomatrix active zone builds the framework for storage
and regulated release of synaptic vesicles at the presynapse. Its evolutionary con-
served protein complex primes and sorts neurotransmitter vesicles ready for synaptic
exocytosis, recruits voltage-gated calcium channels, and tethers them to trans-
synaptic cell adhesion molecules. In addition, its components interact in a differen-
tiated way in response to distinct bursts or trains of action potentials. Soluble matrix
proteins of the active zone initiate exocytosis in a very fast and finely tuned manner
by interacting with proteins of the presynaptic fusion machinery, consisting of
vesicular and membrane SNARE as well as calcium-sensitive proteins, which are
obligatory for all eukaryotic synaptic exocytosis events (Chia et al. 2013; Dresbach
et al. 2001; Lazarevic et al. 2013; Südhof 2012).
Incoming action potentials result in a calcium ion influx by opening voltage-
dependent calcium channels, which provides the calcium ions required for activity-
dependent exocytotic neurotransmitter release into the synaptic cleft (Nanou and
Catterall 2018). Regarding the temporal sequence and the coupling of incoming
action potentials to synaptic neurotransmitter release, three functionally different
exocytosis modes can be described. There are two different types of
stimulation-dependent neurotransmitter release and one based on spontaneous syn-
aptic exocytosis. The two stimulation-dependent release modes are based on highly
synchronized vesicular fusion within hundreds of microseconds after arrival of an
action potential at the synapse. This event is followed by a phase of asynchronous
neurotransmitter release, which complements the synchronous vesicle fusion in
order to maintain synaptic signaling strength. Synchronous neurotransmitter release
is thought to be the pacemaker in neuronal networks, while asynchronous neuro-
transmitter release may drive prolonged synaptic transmission (Bartos et al. 2002;
Chamberland and Tóth 2016; Plenz and Kitai 1998; Volman and Gerkin 2011). In
contrast to stimulation-driven synaptic release, the third mode is constituted by
neurotransmitter-containing synaptic vesicles, which undergo spontaneous exocyto-
sis. Spontaneous neurotransmitter release was shown to play a central role in
synaptic plasticity, especially since spontaneously released neurotransmitters seem
to activate a postsynaptic receptor population not targeted by stimulation-dependent
release. In this way neurons employ differential neurotransmitter signaling of
Neurobiological Principles: Neurotransmitters 7
Our general understanding of neuronal identity is closely associated with the bio-
chemical identity of its neurotransmitter. A neuron is thought to only express the
enzymes needed for the synthesis of its own set of transmitters. Consequently, the
detection of the particular enzymes allows conclusions regarding the neuron’s
signaling characteristics. The expression of tryptophan hydroxylase 2 as synthesiz-
ing enzyme of serotonin defines a neuron as “serotonergic,” while the presence of
glutamic acid decarboxylase categorizes the respective neuron as GABAergic.
Based on these biochemical features, our general view on neurotransmission is
that a genetically-biochemically defined neuron releases the same neurotransmitters
at every synapse, as stated in Dale’s principle (Strata and Harvey 1999). Therefore,
the mode of action of a presynaptic neuron on its postsynaptic counterpart is
considered to depend on the type of neurotransmitter released. For example, gluta-
mate-releasing neurons are considered as excitatory neurons, GABA-releasing neu-
rons as inhibitory neurons, and monoamine-releasing neurons, for example,
dopamine or serotonin, as neuromodulatory neurons. This view denies the potential
of a more sophisticated mode of action by the combined release of two or more
neurotransmitter molecules. In 1979 Jan and colleagues monitored the release of two
8 L. Hampel and T. Lau
A B 1
Target neurons (striatum)
dopamine Nucleus accumbens dorsal striatum
medial shell core
glutamate
+ 2
GABA + +
+ + + + + +
+
Striatum Cortex
C 1
t2
3
Substantia t1
nigra
Fig. 2 Dopamine neurons as a model for joint neurotransmitter release. (a) As part of the motor
system, the basal ganglia (striatum, globus pallidus, substantia nigra, and subthalamic nucleus)
control initiation, extent, direction, force, and speed of movements, among other things. The basal
ganglia are connected to each other by complex control loops. Initially they are activated by the
cortex and then process the signal via main and branch loops (including feedback projections) and
finally project via the thalamus to the motor cortex, which finally executes the movement. In the
overview scheme green and red arrows indicate stimulating, respectively inhibitory input. Here, a
special focus lies on the highly differentiated connection between dopamine neurons of the
substantia nigra and their target neurons in the striatum. Inside this brain region, dopamine neurons
are capable to co-transmit dopamine, GABA, and glutamate in area-specific quantities. (b and c)
Dopamine neurons have the potential to create a unique neuronal input on a postsynaptic neuron by
using combined neurotransmitter release. (b) Co-release of neurotransmitters from dopamine
neurons. (1) Dopamine and a second neurotransmitter are packed into identical synaptic vesicles
yet in different concentrations. (2) Both neurotransmitters are released simultaneously into the
synaptic cleft and bind to corresponding postsynaptic receptors. (3) The cellular response of a
dopamine target neuron is influenced by the concentrations of the second neurotransmitter at
different time points (t1, t2) during dopaminergic neurotransmission. (c) Co-transmission by
dopamine neurons: (1) the dopamine synapse contains synaptic vesicles with different neurochem-
ical identities, for example, either dopaminergic or glutamatergic. (2–3) Co-transmission features
the benefit of spatial and temporal resolution, which enables the synapse to fine-tune neurotrans-
mitter release by releasing the second neurotransmitter in a short pulse, similar duration, or
recurring during dopamine release (2). For example, the second neurotransmitter modulates the
cellular response by a short availability (t1) or a repeated exposure to itself (t2). Both ways of joint
neurotransmission enable dopamine neurons to maintain brain region-specific neurotransmitter
release patterns
10 L. Hampel and T. Lau
Dopamine neurons are well known for their responses to reward, their role in
positive motivation, and their reinforcing effect on self-harming behavior in the
development of addiction. Here, the neuronal signaling pathway constituted by
midbrain dopamine neurons that project from the ventral tegmental area to the
striatum is pivotal for mechanisms underlying associative and reward learning and,
as a component of basal ganglia, also contributes to movement generation. Based on
these exemplary essential physiological processes, it is hardly surprising that
dysregulations in dopamine signaling are assumed to contribute critically to the
pathogenesis of psychiatric and neurologic disorders. Among other psychiatric
diseases, the midbrain to striatum signaling pathway is a key factor in the emergence
of schizophrenia as well as in the onset of dysfunctional behavior in addiction. With
regard to neurodegenerative diseases, the continuous loss of dopaminergic midbrain
neurons is observed in Parkinson’s disease (Chinta and Andersen 2005).
Dopamine neurons located to the ventral tegmental area and targeting striatal
spiny neurons were shown to be capable of performing neurotransmitter co-release
(Fig. 2a). In studies, which employed electrophysiological recordings, Tritsch and
colleagues demonstrated that these neurons co-released dopamine and the fast-acting
neurotransmitter GABA (Tritsch et al. 2012, 2014). Interestingly, their studies also
provided evidence that dopamine neurons recruited GABA by active uptake via
plasma membrane-bound GABA transporter (GAT) molecules rather than de novo
synthesis of the neurotransmitter (Tritsch et al. 2012). Surprisingly, experimental
evidence shows that GABA release from dopamine neurons was diminished by
inhibition of GAT-dependent GABA uptake rather than expression of a functional
vesicular GABA transporter. Here, experimental data provided evidence for vesic-
ular monoamine transporter 2 (VMAT2)-dependent uptake of GABA into the syn-
aptic vesicles of dopamine neurons (Tritsch et al. 2014). The potential of the latter
observation implies that all types of monoamine neurons may have the potential for
GABA co-release by loading their synaptic vesicles without de novo GABA syn-
thesis via plasma membrane GAT and non-canonical VMAT2 uptake (discussed in
Granger et al. 2017; Vaaga et al. 2014). Regarding the effect of GABA co-release,
Tritsch and colleagues were able to show that GABAA receptor-evoked postsynaptic
currents in neurons, targeted by the dopamine presynapse, were diminished once
GABA uptake into the dopamine neurons was prevented (Tritsch et al. 2014).
Interestingly, the functional analysis of midbrain dopamine neurons revealed that
they are capable to simultaneously release dopamine and either glutamate or GABA
in the nucleus accumbens yet not in the dorsal striatum (Stuber et al. 2010). Based on
these observations, dopamine neurons seem to be capable to differentially control
striatal circuit function. Here, single dopamine release is viewed to affect the
neuronal activity of all striatal neurons targeted by dopamine neurons. In contrast,
simultaneous release modes comprised of dopamine-glutamate or dopamine-GABA
release, occurring exclusively at the shell and core of the nucleus accumbens,
respectively, are considered a manifold way for dopamine neurons to transmit fast,
Neurobiological Principles: Neurotransmitters 11
distinct signals to differentially affect their target neurons in defined striatal areas
(Fig. 2b and c). Furthermore, dopamine neurons may complement their own neuro-
transmission by directly modulating the activity of striatal cholinergic interneurons.
Thereby dopamine neurons have the potential to create a unique neuronal input on
general striatal function via a specific link to their target neurons and via cholinergic
interneurons (Chuhma et al. 2014). This distinctive dopaminergic neurotransmission
feature, which combines differential temporal joint neurotransmitter release and
modulation of interneuron activity, is thought to be specifically involved in
reward-related learning and the development of addictive disorders (Stuber et al.
2010; Chuhma et al. 2014).
Prior to these studies, immunofluorescence analysis as well as electrophysiolog-
ical recordings of the dopamine synapse in cultured dopamine neurons already
provided first evidence for combined synaptic dopamine-glutamate release. On the
molecular level, the expression of the vesicular glutamate transporter 2 (vGLUT2)
gene was verified in cultured dopamine neurons. Immunofluorescence analysis
revealed that vGLUT2 was localized in synaptic vesicle pools of dopamine synap-
ses. Further functional evidence was provided by electrical stimulation of dopamine
neurons, which evoked fast glutamate-dependent excitatory postsynaptic currents
(Dal Bo et al. 2004; Joyce and Rayport 2000; Onoa et al. 2010; Sulzer et al. 1998).
Evidence for dopamine-glutamate co-transmission was also found by corresponding
in vivo data: for example, vGLUT2 expression in dopamine neurons of rodent brain,
a prerequisite for specific glutamate uptake by synaptic vesicles, was verified in
midbrain dopamine neurons (Kawano et al. 2006). Interestingly, a dopamine neuron-
specific knockout of the vGLUT2 gene impaired axonal growth of dopamine
neurons and resulted in a diminished innervation of the nucleus accumbens and
even reduced dopamine neuron survival. Consequently, a diminished dopamine
signaling led to impaired motor behavior (Fortin et al. 2012). A thorough combined
immunofluorescence and electron microscopy analysis of dopamine synapses
showed that dopamine synapses in the nucleus accumbens display differentially
organized cytomatrix active zones hosting microdomains with either vGLUT2- or
VMAT2-containing synaptic vesicles (Zhang et al. 2015). Furthermore, these studies
provided evidence for an age-dependent vGLUT2 expression, with a declining
ability of combined dopamine-glutamate neurotransmission in adult rodent brain
(Mendez et al. 2008; Bérubé-Carrière et al. 2009). Optogenetic approaches
performed in nucleus accumbens brain sections provided further insight into the
characteristics of dopamine-glutamate co-transmission. One study successfully dem-
onstrated glutamate release by mesolimbic dopamine neurons in the nucleus
accumbens, which resulted in postsynaptic excitation (Tecuapetla et al. 2010).
Furthermore, the glutamatergic input could be linked to mesolimbic reward-related
dopamine signaling circuits, a first indication that synaptic glutamate release by
dopamine neurons may contribute to mesolimbic reward signaling. In addition,
optogenetic approaches verified the findings of microscopic localization analysis
of the dopamine synaptic vesicle pools. Dopamine- and glutamate-mediated post-
synaptic responses are based on similar neurotransmitter release probabilities, yet
dopamine and glutamate signaling was shown to be differentially affected by drugs
12 L. Hampel and T. Lau
targeting dopamine signaling (Adrover et al. 2014). This observation and previous
findings provide evidence for independent vesicle pools localized to the cytomatrix
active zone of dopamine synapses, which contain either dopamine or glutamate
(Adrover et al. 2014; Barker et al. 2016; Hnasko et al. 2010).
Although the relevance for joint neurotransmitter release is not fully understood
(discussed in Hnasko and Edwards 2012), the experimental data obtained by studies
covering mechanisms and structural organization of dopamine synapses provide a
constantly growing insight into co-release and co-transmission by dopamine neu-
rons. Due to their different postsynaptic or autoregulatory effects, co-transmission of
dopamine and glutamate or co-release with GABA can have versatile outcomes.
Regarding glutamate neurotransmission, dopamine may modulate glutamatergic
signaling either by strengthening or by weakening glutamate transmission,
depending on the postsynaptic dopamine receptor profile. Glutamate on the other
hand may enhance dopamine signaling by providing additional excitatory input.
Similarly, by dopamine-GABA co-release, GABA signaling may be directly mod-
ulated due to the presence of dopamine, while dopamine signaling may be dimin-
ished by simultaneous GABA release. In contrast to dopamine-glutamate co-
transmission, dopamine-GABA co-release lacks the spatial and temporal resolution
displayed by co-transmission (Barker et al. 2016; Granger et al. 2017; Svensson et al.
2019).
Due to their different postsynaptic or autoregulatory effects, co-transmission of
dopamine and glutamate or co-release with GABA can have versatile outcomes.
Regarding glutamate neurotransmission, dopamine may modulate glutamatergic
signaling either by strengthen or by weaken glutamate signaling, depending on the
postsynaptic dopamine receptor profile. Glutamate on the other hand may enhance
dopamine signaling by providing additional excitatory input. Similarly, by dopa-
mine-GABA co-release, GABA signaling may be directly modulated due to the
presence of dopamine, while dopamine signaling may be diminished by simulta-
neous GABA release. In contrast to dopamine-glutamate co-transmission, dopa-
mine-GABA co-release lacks the spatial and temporal resolution displayed by co-
transmission (Barker et al. 2016; Granger et al. 2017; Svensson et al. 2019).
Regarding neurotransmitter activity at the AIS, where dopamine and GABAergic
neurons maintain axo-axonic synapses and were shown to modulate action potential
generation, the occurance of dopamine-GABA co-release provides a highly inter-
esting synaptic input beyond singular, independent dopamine or GABA signaling.
This is especially interesting since dopamine was shown to be able to inhibit
GABAA-receptors (Hoerbelt et al. 2015).
In summary, our example for joint neurotransmitter release by dopamine neurons
shows that neurons are capable to maintain brain region-specific neurotransmitter
release patterns. This enables neurons to fine-tune synaptic communication within
well-defined neuronal circuits. Regarding psychiatric as well as neurodegenerative
diseases, established pharmacotherapies may cause additional disruptive signaling in
synapse maintaining joint neurotransmitter release if only one of multiple neuro-
transmitter signals is either enhanced or inhibited. Consequently, already impaired
signaling pathways may be further disrupted and disease conditions promoted.
Neurobiological Principles: Neurotransmitters 13
SERT
extracellular
histones matrix proteins
GTPases
regions, and some studies provided evidence for activity-dependent gene expression
to maintain dendritic spine formation (Bosch et al. 2014; Govindarajan et al. 2011).
In this scenario, neurotransmitters are able to affect gene expression by interaction
with their respective receptors and subsequent stimulation of intracellular signaling
cascades. In contrast to this, a very recent study provided first evidence for a
neurotransmitter to directly participate in gene transcription. The covalent binding
of serotonin to glutamine residues of histone H3 was shown to mediate permissive
gene expression and represents a new nonsynaptic signaling function for the neuro-
transmitter serotonin (Farrelly et al. 2019; Fig. 3).
The enzymatic covalent binding of primary amines to glutamine residues of
proteins was discovered several decades ago by Heinrich Waelsch and colleagues
(Sarkar et al. 1957). The proteins mediating this reaction were termed trans-
glutaminases, and substrates include the neurotransmitters serotonin and norepi-
nephrine among others. The transamidation of proteins with monoamines was
termed monoaminylation and specifically serotonylation for the covalent binding
of the neurotransmitter serotonin to glutamine residues (Berger et al. 2009; Bader
2019). Initially, serotonylation was shown in platelets, where serotonin was cova-
lently bound to extracellular matrix proteins as well as small cytoplasmic GTPases.
In both cases serotonylation resulted in enhanced metabolism activity of platelets
(Walther et al. 2003; Dale et al. 2002). In the context of neuronal cells, serotonylation
was first confirmed in a glia cell line, where radioactive serotonin was bound to
Neurobiological Principles: Neurotransmitters 17
transporters or organic cation transporters (Chen et al. 2004; Daws 2009). SERT
expression was recently shown to depend on temporally and spatially defined
expression in non-serotonin neurons during cortical development (Chen et al.
2016), and one may assume that intracellular serotonylation may be linked to
developmental SERT expression patterns. Independent of SERT expression, intra-
cellular serotonylation may be maintained via serotonin taken up into non-serotonin
cells by non-specific monoamine transporters (discussed in Anastas and Shi 2019;
Bader 2019; Fu and Zhang 2019). For both uptake scenarios, either the spatiotem-
poral availability of SERT or a putative constitutional uptake by PMAT and/or OCT,
serotonin uptake may clearly benefit from serotonin volume transmission without
the need to establish a synaptic connection with serotonin neurons.
In conclusion, the experimental proof of histone serotonylation reveals a new way
how a neurotransmitter may directly and independently of its neurotransmitter
receptor pathways affect gene expression. Serotonin’s covalent binding to histone
3 changes the posttranslational modification of histone and subsequently influences
transcription activity in a variety of cell types. This circumstance raises the question
to which extent other neurotransmitters may directly affect gene expression. For
example, dopamine, norepinephrine, and histamine are substrates for transamidation
of glutamine residues, too, and thereby display the potential for histone mono-
aminylation (Hummerich et al. 2012, 2015; Muma and Mi 2015; Walther et al.
2011). Monoaminylation of histones by neurotransmitters crosses a new frontier of
how we perceived neurotransmitters’ mode of action in recent years. Upcoming
research will tell us whether our view on brain diseases as axono- or synaptopathies
has to be extended to including neurotransmitter-dependent chromatin modifica-
tions. Future research will unravel whether monoaminylation crucially contributes to
the onset of psychiatric diseases or efficacy of established therapeutic treatment or
whether this fascinating mechanism is a new starting point for the discovery of novel
therapeutic approaches.
Cross-References
▶ Experimental Psychopharmacology
▶ Mood Stabilizers: Pharmacology and Biochemistry
▶ Neurobiological Principles: Psycho-Neuro-Immuno-Endocrinology
▶ Pharmacokinetic and Pharmacodynamic Principles
References
Abraham WC. How long will long-term potentiation last? Phil Trans R Soc Lond B. 2003;358:735–44.
Adrover MF, Shin JH, Alvarez VA. Glutamate and dopamine transmission from midbrain dopamine
neurons share similar release properties but are differentially affected by cocaine. J Neurosci.
2014;34:3183–92.
Anastas JN, Shi Y. Histone Serotonylation: can the brain have “happy” chromatin? Mol Cell.
2019;74(3):418–20.
Neurobiological Principles: Neurotransmitters 19
Araque A, Carmignoto G, Haydon PG, Oliet SH, Robitaille R, Volterra A. Gliotransmitters travel in
time and space. Neuron. 2014;81:728–39.
Bader M. Serotonylation: serotonin signaling and epigenetics. Front Mol Neurosci. 2019;12:288.
Ballestar E, Abad C, Franco L. Core histones are glutaminyl substrates for tissue transglutaminase. J
Biol Chem. 1996;271:18817–24.
Barker DJ, Root DH, Zhang S, Morales M. Multiplexed neurochemical signaling by neurons of the
ventral tegmental area. J Chem Neuroanat. 2016;73:33–42.
Bartos M, Vida I, Frotscher M, Meyer A, Monyer H, Geiger JR, Jonas P. Fast synaptic inhibition
promotes synchronized gamma oscillations in hippocampal interneuron networks. Proc Natl
Acad Sci U S A. 2002;99(20):13222–7.
Bazzari AH, Parri HR. Neuromodulators and long-term synaptic plasticity in learning and memory:
a steered-glutamatergic perspective. Brain Sci. 2019;9(11):300.
Bekar LK, He W, Nedergaard M. Locus coeruleus α-adrenergic–mediated activation of cortical
astrocytes in vivo. Cereb Cortex. 2008;18(12):2789–95.
Berger M, Gray JA, Roth BL. The expanded biology of serotonin. Annu Rev Med. 2009;60:355–66.
Bérubé-Carrière N, Riad M, Dal Bo G, Lévesque D, Trudeau LE, Descarries L. The dual dopamine-
glutamate phenotype of growing mesencephalic neurons regresses in mature rat brain. J Comp
Neurol. 2009;517:873–91.
Bliss TVP, Lømo T. Long-lasting potentiation of synaptic transmission in the dentate area of the
anaesthetized rabbit following stimulation of the perforant path. J Physiol. 1973;232:331–56.
Bosch M, Castro J, Saneyoshi T, Matsuno H, Sur M, Hayashi Y. Structural and molecular
remodeling of dendritic spine substructures during long-term potentiation. Neuron.
2014;82:444–59.
Bramham CR. Local protein synthesis, actin dynamics, and LTP consolidation. Curr Opin
Neurobiol. 2008;18:524–31.
Brockett AT, Kane GA, Monari PK, Briones BA, Vigneron P-A, Barber GA, Bermudez A,
Dieffenbach U, Kloth AD, Buschman TJ, Gould E. Evidence supporting a role for astrocytes
in the regulation of cognitive flexibility and neuronal oscillations through the Ca2+ binding
protein S100β. PLoS One. 2018;13(4):e0195726.
Broussard J. Co-transmission of dopamine and glutamate. J Gen Physiol. 2011;139:93–6.
Caroni P, Donato F, Muller D. Structural plasticity upon learning: regulation and functions. Nat Rev
Neurosci. 2012;13:478–90.
Chamberland S, Tóth K. Functionally heterogeneous synaptic vesicle pools support diverse syn-
aptic signalling. The Journal of Physiology 2016;594 (4):825–835.
Chen NH, Reith ME, Quick MW. Synaptic uptake and beyond: the sodium- and chloride-dependent
neurotransmitter transporter family SLC6. Pflugers Arch. 2004;447(5):519–31.
Chen X, Petit EI, Dobrenis K, Sze JY. Spatiotemporal SERT expression in cortical map develop-
ment. Neurochem Int. 2016;98:129–37.
Chia PH, Li P, Shen K. Cell biology in neuroscience: cellular and molecular mechanisms underlying
presynapse formation. J Cell Biol. 2013;203(1):11–22.
Chinta SJ, Andersen JK. Dopaminergic neurons. Int J Biochem Cell Biol. 2005;37(5):942–6.
Choquet D, Triller A. The dynamic synapse. Neuron. 2013;80(3):691–703.
Chuhma N, Mingote S, Moore H, Rayport S. Dopamine neurons control striatal cholinergic neurons
via regionally heterogeneous dopamine and glutamate signaling. Neuron. 2014;81:901–12.
Citri A, Malenka RC. Synaptic plasticity: multiple forms, functions, and mechanisms. Neuropsy-
chopharmacology. 2008;33:18–41.
Dal Bo G, St-Gelais F, Danik M, Williams S, Cotton M, Trudeau LE. Dopamine neurons in culture
express VGLUT2 explaining their capacity to release glutamate at synapses in addition to
dopamine. J Neurochem. 2004;88:1398–405.
Dale GL, Friese P, Batar P, Hamilton SF, Reed GL, Jackson KW, Clemetson KJ, Alberio L.
Stimulated platelets use serotonin to enhance their retention of procoagulant proteins on the
cell surface. Nature. 2002;415:175–9.
Daws LC. Unfaithful neurotransmitter transporters: focus on serotonin uptake and implications for
antidepressant efficacy. Pharmacol Ther. 2009;121(1):89–99.
20 L. Hampel and T. Lau
Norbert Müller
Contents
Immunological Principles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
Immunological Memory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
Definition of Psycho-neuro-immunology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
Methodological Aspects of Psycho-neuro-immunology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
Interaction Between the Immune, Endocrine, and Central Nervous Systems . . . . . . . . . . . . . . . . . . . 29
The Cytokine System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
Stimulation of Cells in the CNS by Cytokines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
Cytokine Production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
Mode of Action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
Kindling and Sensitization of the Immune Response: The Basis for a Stress-Induced
Inflammatory Immune Response in Mental Illness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
Interaction of Cytokines and Neurotransmitters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
Interleukin-1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
Interleukin-2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
Interleukin-6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
TNF-α . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
Blood-Brain Barrier . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
Immune Genetics, the HLA System, and Schizophrenia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
Cellular Immune System and Mental Disorders . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
Lymphocyte Population and Cytokine Production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
Tryptophan-Kynurenine Metabolism in Mental Disorders . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
The Role of the Immune System in Cognition and Cognitive Disorders . . . . . . . . . . . . . . . . . . . . . . . 38
Mental Disorders and Autoimmune Diseases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
Treatment with Interferon-α and Depression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
Schizophrenia and the Immune System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
Disturbances of Neurotransmitters as a Result of Pre- or Postnatal Infections . . . . . . . . . . . . . . 40
N. Müller (*)
Department of Psychiatry and Psychotherapy, Ludwig-Maximilians-Universität Munich,
Munich, Germany
e-mail: norbert.mueller@med.uni-muenchen.de
Abstract
Mental disorders have a complex pathophysiology that remains only partially
understood. In addition to changes in neurotransmitter systems and receptors –
the effector systems that act on downstream signal transduction processes – there
is increasing evidence that immunological mechanisms play an important role in
the pathophysiology of mental disorders. In recent decades, psycho-neuro-
immuno-endocrinology, which is concerned with the relationships between men-
tal functions, psychiatric disorders, the immune system, and endocrine mecha-
nisms, has become an important topic in psychiatry. Meanwhile, initial attempts
have been made to make use of such psycho-neuro-immunological mechanisms
as novel treatment strategies.
Immunological Principles
The first barrier of the immune system is monocytes which, when activated, differ-
entiate into macrophages; macrophages phagocytize foreign antigens and present
antigens on the cell surface. Monocytes are part of the non-specific, or innate,
immune system, which activates the specific immune response via the adaptive
immune system. Natural killer (NK) cells, a type of lymphocyte, are also part of
the innate immune system and play a major role in killing tumor cells and cells
infected with viruses.
Neurobiological Principles: Psycho-Neuro-Immuno-Endocrinology 27
Other lymphocytes, the B and T cells, are also crucial for the immune response.
B cells develop in the bone marrow, whereas T lymphocyte maturation and imprint-
ing take place in the thymic cortex. Both B and T cells then migrate to the other
lymphatic organs (tonsils, lymph follicles, lymph nodes, and spleen).
In an immune response, some T cells are directly involved in cytolytic cell-cell
interactions, such as in transplant rejection or graft-versus-host reactions after bone
marrow transplants. T lymphocytes are therefore referred to as facilitators of cell-
mediated immunity. They are activated when the respective specific antigen is
presented on the surface of an “accessory” cell, a so-called antigen-presenting cell
(e.g., macrophages or certain types of lymphocytes), together with a histocompati-
bility antigen (human leukocyte antigen, HLA); T-cell activation and proliferation
also require a non-antigen-specific signal from the accessory cell.
The HLA system helps differentiate between self and nonself, an essential
function of the immune system. Immune system function is critically dependent
on the HLA system in that, after antigens have been processed inside the cell, the
HLA system determines which specific parts (peptides) of them are presented to the
T lymphocytes. This system comprises about one thousandth of the human genome
and contains a number of tightly linked loci on the short arm of chromosome 6. A
number of genetic diseases of the nervous system, e.g., multiple sclerosis (MS) and
narcolepsy, are associated with HLA genes. Natural killer cells (NK cells) kill cells
non-specifically and are not restricted by the HLA system.
Antigen-presenting cells release activating cytokines, a group of proteins that
activate B lymphocytes (CD19+) and T lymphocytes. B lymphocytes differentiate
into plasma cells that produce antibodies. The cells of the immune system are defined
by their surface marker molecules and the pattern of cytokines they secrete. Thus, CD3
is a marker for the total number of T lymphocytes. T lymphocytes can be divided into
several, functionally different subpopulations that can be defined with the help of
monoclonal antibodies. The most important subpopulations are the T helper/inducer
cells (CD4+), which induce an immune response, and the cytotoxic T cells/T suppres-
sor cells (CD8+), which regulate an organism’s immune response and also have
cytotoxic effects and lyse cells. CD14+ is a surface marker for monocytes and
macrophages, CD16+/56+ for NK cells, and CD5+/ CD19+ for B lymphocytes. NK
cells are primarily activated by interferon-(INF-)γ. The function of CD4+ and CD8+
cells, which in turn are activated by CD3+ T lymphocytes, is usually balanced.
The T helper 1 (TH-1) system, which initiates the rapid immune response and
consists primarily of elements of the cellular immune system, becomes active in acute
inflammation. Characteristic cytokines of this system are INF-γ, interleukin-(IL-)2,
and IL-12. Because not only T helper cells (CD4+ cells) but also monocytes/macro-
phages and other cell types produce these cytokines, this type of immune response is
called the type 1 immune response. The T helper 2 (TH-2) system or type 2 immune
response is the humoral arm of the immune system and is activated in chronic
inflammatory processes and also in allergic reactions. T helper 2 cells (TH-2) or
type 2 monocytes/macrophages (M2) produce mainly IL-4, IL-10, and IL-13. The
differentiation to T helper 1 and T helper 2 lymphocytes (which is also defined by their
pattern of cytokine release) appears to be relevant in chronic inflammatory diseases.
28 N. Müller
Immunological Memory
The immune system is the only human system besides the central nervous system
(CNS) that has a memory. The T memory cells (CD45+) are responsible for this
function in that they “remember” a specific antigen and initiate a strong specific
immune response if re-exposed to it.
One of the historical starting points for psycho-neuro-immunology was an
immune conditioning in animals: In mice and rats, immune suppression mediated
by cyclophosphamide (unconditioned stimulus) can be conditioned by simultaneous
administration of saccharin (conditioned stimulus) (Ader and Felten 1991). The
conditioned immune response is now well elaborated in different animal models
and is an interesting approach for treatment in humans. T memory cells and other
molecules of the immunological memory may be involved in this therapeutic
approach, but so far results in humans are not convincing.
Definition of Psycho-neuro-immunology
Studies of the human immune system are known to have a range of methodolog-
ical problems because different components of the immune system are influenced
by variables that are difficult to control in human studies, including sleep, alcohol
and drug consumption, medicines, eating habits, daily rhythms, stress, smoking,
Neurobiological Principles: Psycho-Neuro-Immuno-Endocrinology 29
physical activity, infections, tumors, etc. Clinical illness factors, such as acute-
ness, course, severity, and psychopathology, also appear to influence the immune
system. These examples show that the immune system is very sensitive to various
influences; on the other hand, however, thanks to its high complexity and
variability, it is able to functionally overcome many influences and maintain a
functional homeostasis.
In recent years, numerous interactions have been described between the immune
system and the CNS. There are two main reasons for the great interest in research in
this field:
Androgens have also been shown to suppress the immune response, whereas
thyroxin, growth hormone (GH) and insulin stimulate it. At higher doses, estrogens
suppress the cellular immune response, but lower doses have a stimulatory effect.
The higher rate of autoimmune diseases such as scleroderma, rheumatoid arthritis,
and systemic lupus erythematosus in women indicates a possible involvement of sex
hormones in immune dysregulation.
Not only is the immune system influenced by the endocrine system, but it can also
regulate the endocrine system and peripheral immune processes impact the CNS in
the form of afferent effects. Thus, in experimental animals, the maximum antibody
production after an antigen injection is accompanied by a two- to threefold increase
in the blood concentration of glucocorticoids, and an immunosuppressive effect is
achieved; at the same time, the firing rate of hypothalamic nuclei reaches its
maximum (Besedovsky et al. 1986). On the basis of these findings, the immune
modulatory effect of cortisol is hypothesized to be one of the most important
physiological effects of this hormone.
On the other hand, findings indicate that an overstimulation of the hypothalamic-
pituitary-adrenal (HPA) axis through release of cytokines that stimulate corticotro-
phin-releasing factor (CRF) results in immune suppression, seen, for example, in
decreased effectiveness of hepatitis or influenza vaccinations (Pennisi 1997). Fur-
thermore, lymphocytes of the peripheral immune system were found to produce
hormones, i.e., ACTH, β-endorphins, thyroid-stimulating hormone (TSH), GH, and
prolactin. The immune system thus seems to also take over some functions of the
endocrine system.
Peptide signals of the immune and endocrine systems appear to have some joint
functions, and functions and signals of both systems appear to have many
similarities.
Cytokines convey information between cells of the peripheral immune system and
the CNS. In part, they are actively transported across the blood-brain barrier, but they
are also formed in the CNS by activated astrocytes and microglia. IL-1, IL-2, IL-6,
and TNF-α are the most important activating cytokines and have a variety of known
functions in the CNS. Findings from recent years show that the effects of cytokines
are also relevant for mental illnesses.
Cytokines in the CNS are involved in various regulatory mechanisms, including
the following:
CNS cells can be activated by cytokines in various ways: First, at least some
cytokines, e.g., IL-1, IL-2, IL-6, and TNF-α, are transported from the blood into
the CNS by active transport mechanisms; second, glia cells secrete cytokines after
being activated by antigen stimuli; and, last, cytokine secretion in the CNS can be
triggered by neurotransmitters. For example, noradrenalin was shown to stimulate
IL-6 production in astrocytes in a dose-dependent way (Norris and Benveniste
1993). Because IL-6 is closely functionally related to other cytokines, such as
IL-1, IL-2, and TNF-α, this finding indicates that the cytokine cascade can possibly
also be stimulated by neurotransmitters. This may represent an important link
between (auto)immune disorders, susceptibility to infections, state of health, and
mental disorders.
Furthermore, cytokines of course also pass through a disturbed blood-brain
barrier into the CNS.
Cytokine Production
Once activated, both astrocytes and microglia can produce and secrete cytokines.
Interestingly, the ways in which these cell types are stimulated to produce cyto-
kines and the pattern of cytokine release differ, suggesting that they have different
functions in the CNS immune response. Astrocytes, which are closely associated
with neuronal synapses, store neurotransmitters and release them when needed.
Astrocyte dysfunction, e.g., as a result of activation by cytokines, can thus easily
affect neurotransmitter balance. Microglia are stimulated by viruses to produce
cytokines, which – together with the expression of cellular surface structures –
initiates an immune reaction in the CNS in case of viral infections (Lieberman
et al. 1989).
Mode of Action
The fact that neurons have cytokine receptors suggests that cytokines have a direct
effect on neuronal functions. Neurons containing IL-1 are found in various regions
of the CNS, including the hypothalamus and hippocampus. IL-2-receptor mRNA
(the genetic information of the IL-2 receptor) has been demonstrated in neurons,
suggesting that IL-2 also has direct effects on neurons. Furthermore, in animal
studies stimuli were shown to be conducted from IL-1, i.e., from the immune
system, via the vagus nerve to critical regions in the CNS without IL-1 itself
entering the CNS. Thus, the vagus nerve is clearly a link between the immune
system and the CNS.
Physiological development of the CNS can also be significantly impaired by an
over- or underproduction of cytokines (Merrill 1992) because cytokines also func-
tion as growth factors in the CNS. This may represent a link between pre- or
32 N. Müller
perinatal damage, e.g., through birth trauma or a prenatal virus infection, and a
disturbance of brain maturation, as is postulated in schizophrenic disorders.
TNF-α Endogenous pyrogens; release of IL-1, Cytotoxic; Ubiquitous? Astrocytes; Acute: catecholamines Cognition?
activation of macrophages, cytotoxicity demyelination; microglia
fever
33
34 N. Müller
Interleukin-1
IL-1 stimulates catecholamine release both outside and within the CNS, particularly
in the brain stem and hypothalamus. Increased levels of noradrenaline and serotonin
and their metabolites were found in these regions after both intraventricular and
peripheral administration of IL-1 (Zalcman et al. 1994).
Interleukin-2
Interleukin-6
In vitro, IL-6 can stimulate neurons to secrete dopamine and possibly also other
catecholamines. In animal experiments, the peripheral administration of IL-6
increases dopamine and serotonin turnover in the hippocampus and frontal cortex,
without affecting noradrenaline metabolism (Zalcman et al. 1994). Conversely,
noradrenaline can stimulate IL-6 production in activated astrocytes (Norris and
Benveniste 1993).
Neurobiological Principles: Psycho-Neuro-Immuno-Endocrinology 35
TNF-a
TNF-α also affects the neurotransmitter balance, whereby these effects appear to
depend on the duration of administration of TNF-α: Whereas acute administration of
TNF-α was found to stimulate the catecholamine system via mechanisms in the
CNS, chronic administration diminished catecholamine secretion (Soliven and
Albert 1992). TNF-α is proposed to play a key role in dementia and also in the
cognitive impairments associated with HIV.
Blood-Brain Barrier
Genetic data from multiple, large patient cohorts found the most evidence for suscep-
tibility genes for schizophrenia on chromosome 6p22.1 (Muller et al. 2012a, b; Muller
and Schwarz 2010). This region includes various interesting genes relevant for
immune functions. The strongest evidence for an association was found in or near a
cluster of histone protein genes, which are important because of their role in regulating
DNA transcription or repair, i.e., for epigenetics (Purcell et al. 2009), or their direct
role in microbe defense (Shi et al. 2009). Furthermore, various genes of the HLA
complex are located in this region. HLA genes regulate immune function and have
been discussed as being involved in the pathophysiology of schizophrenia (Stefansson
et al. 2009). A meta-analysis showed the by far most significant association with
schizophrenia risk in this region (Schizophrenia Working Group of the Psychiatric
Genomics 2014).
Since the mid-1970s, numerous studies have evaluated the association between
HLA class I antigens (HLA-A, HLA-B, HLA-C) and schizophrenia. The results are
inconsistent, however, and a range of described associations could not be replicated.
On the other hand, several methodological factors may explain the discrepancies in
the results, e.g., ethnic and local differences, influences on “linkage disequilibrium,”
diagnostic criteria, and small sample sizes.
In various autoimmune diseases, clearer associations have been found with the
HLA class II system (HLA-DR, HLA-DQ, HLA-DP) than with the class I system.
Only a few studies have considered the class II system in schizophrenia, however.
Both a German study and a small American study found a slight increase in
HLA-DQB1 0602. This finding is particularly interesting because HLA-DQB1
0602 is also associated with narcolepsy and MS and thus may be a joint vulnera-
bility gene for several CNS diseases (Grosskopf et al. 1998).
play a key role (Hampel et al. 1996). Studies have found indications of an opening of
the blood-brain barrier in schizophrenia, depression, and dementia (Hampel et al.
1999; Muller 2019; Muller and Ackenheil 1995b; Schwarz et al. 1998).
COX-2 inhibition
Tryptophan
+ IDO - Type 2
Type 1
TDO Present in astrocytes
Kynurenic Kynurenine (KYN)
acid (KYNA) Not present
(NMDA receptor antagonist) Type 1 + - Type 2 in astrocytes
KMO
3-hydroxykynurenine (3-
HK)
Astrocytes play a key role in the production of KYNA in the CNS and are its main
source (Maes et al. 1997). Kynurenine (KYN) is predominantly metabolized in
macrophages and microglia, but also in astrocytes. Kynurenine 3-monooxygenase
(KMO; also referred to as kynurenine hydroxylase, among other names) is an
important enzyme in kynurenine metabolism that is not found in human astrocytes
(Muller et al. 1999). Consequently, astrocytes are unable to produce
3-hydroxykynurenine (3-HK) or quinolinic acid (QUIN), although they can produce
larger amounts of the earlier metabolites, such as KYN and KYNA (Muller et al.
1999). This finding is in line with the observation in animal models that inhibition of
KMO results in an increase in KYNA production in the CNS (Pollmacher et al.
2001). The complete metabolism of KNY to QUIN is mainly observed in microglia,
whereas only small amounts of QUIN are synthesized in astrocytes via a side arm of
kynurenine metabolism. Because of the lack of KMO, increased tryptophan metab-
olism to KYN can result in an accumulation of KYNA in astrocytes. Monocytic cells
that infiltrate the CNS are a second key in the metabolism of 3-HK in that they help
astrocytes in the metabolization of 3-HK to QUIN (Muller et al. 1999).
Thus, depending on the immune status, there is a preponderance either of the N-
methyl-D-aspartate (NMDA) antagonist KYNA or the NMDA agonist QUIN.
Because KYNA interferes with glutamatergic metabolism, a pro-inflammatory
immune status could explain the glutamatergic hyperfunction that has been fre-
quently described in schizophrenia. In contrast, a lack of glutamatergic neurotrans-
mission is discussed in depressive disorders (Plangar et al. 2012), which may be the
result of a preponderance of QUIN. KYNA is also an antagonist at the α7 nicotine
receptor, which is highly relevant for cognitive function; antagonism at this receptor
is associated with cognitive impairment.
QUIN and 3-HK also have neurotoxic effects, whereas KYNA has
neuroprotective properties. Because tryptophan is the starting point for the produc-
tion of serotonin and is also involved in melatonin metabolism, changes in trypto-
phan metabolism affect not only glutamatergic but also serotonergic and melatonin
metabolism. Thus, tryptophan/KYN metabolism is an important link with regard to
changes in the immune system and their effects on neurotransmitters that play a
major role in mental disorders.
A study that evaluated the association between the T-cell immune system and
cognitive ability in animals provided a clear indication that the immune system is
closely involved in cognitive processes. Mice with a severe combined immune
deficiency were compared with healthy wild-type mice by using the Morris water
maze test, which represents memory and learning. The study found that learning
behavior was worse in the T-cell-deficient study mice than in the wild-type mice, i.e.,
they learned more slowly and also more quickly forgot what they had learned. If the
study mice were substituted with T cells from other mice, they showed
Neurobiological Principles: Psycho-Neuro-Immuno-Endocrinology 39
improvements in both learning and memory, but if they were treated with the NMDA
antagonist MK801 or other substances with similar effects, their cognitive perfor-
mance deteriorated further. However, if Copaxone, a T-cell stimulator, was admin-
istered at the same time as the other substances, the study mice showed the same
learning and memory performance as the wild-type mice. This study shows that an
intact T-cell immune response is required for intact cognitive performance (Kipnis
et al. 2004). Considering that the T-cell immune response is subject to an aging
process from age 55 onward (“immunosenescence”), the immune deficiency in old
age may also explain cognitive deficits in old age. The Maastricht Aging Study
examined this hypothesis and prospectively evaluated 100 healthy participants with
a mean age of 57 years with respect to inflammatory markers and cognitive tests
(Teunissen et al. 2003). The study found a significant negative correlation between
levels of haptoglobin (an APP) and the course of cognitive abilities, measured with
the Stroop Test and the Auditory Verbal Learning Test. It also found a significant
negative correlation between high levels of C-reactive protein (CRP) and cognitive
abilities (Auditory Verbal Learning Test) after 3 and 6 years, i.e., worse cognitive
abilities were associated with higher concentrations of CRP and haptoglobin.
A similarly designed prospective study assessed the inflammatory markers CRP
and IL-6 in 4200 people and performed cognitive tests after about 7 and 12 years.
CRP and IL-6 were significantly associated with cognitive performance, particularly
in men. Higher levels of pro-inflammatory markers in midlife were moderately
correlated with poor cognitive performance and weakly correlated with the decline
in cognitive abilities (Gimeno et al. 2008). Animal studies showed that increased
release of IL-6 leads to learning and memory deficits (Heyser et al. 1997), and IL-6
knockout mice had a lower risk of forgetting what they had learned and showed
better cognitive performance overall than wild-type mice (Braida et al. 2004).
Interestingly, the intravenous administration of anti-IL-6 antibodies also improved
memory function (Balschun et al. 2004).
The impact of inflammation on cognition led to ongoing research on the role of
the immune system in cognitive decline and cognition-related disorders, such as
Alzheimer’s disease (Blum-Degen et al. 1995; Eikelenboom et al. 1991; McGeer
et al. 2016) and Parkinson’s disease (Mogi et al. 1994a, b, 1995a, b).
The fact that psychoses can result from immune processes is shown by the occur-
rence of psychotic phenomena in various autoimmune diseases in which CNS
immune processes can be demonstrated, such as lupus erythematosus, scleroderma,
Sjögren syndrome, and antiphospholipid syndrome (Kurtz and Muller 1994).
From a clinical perspective, similarities exist between autoimmune diseases and
in particular schizophrenia and affective disorders. These include the often early
onset of the disease, the genetic vulnerability, and the relapsing or phasic course.
Authors have drawn parallels between MS and the increased occurrence of both
schizophreniform syndromes (Stevens 1988) and affective disorders (Berrios and
40 N. Müller
Activation of the immune system by clinical use of interferon-α has been shown to
be associated with the occurrence of depressive symptoms and even clinical depres-
sion, including suicidality. Interferon-α is part of the standard treatment for chronic
hepatitis C and various malignant diseases, such as malignant melanoma and renal
cell carcinoma. The prevalence rates for the occurrence of a depressive syndrome
vary greatly, from 16% to 58% of treated patients. The risk of developing a
depression requiring treatment increases with higher doses of interferon-α and a
longer treatment duration. If fatigue is considered to be a symptom of depression, the
prevalence of a depressive syndrome during interferon-α treatment increases up to
80% (Capuron et al. 2002). Besides depressive mood, fatigue, insomnia, weight loss,
loss of appetite, and cognitive impairment have been described during interferon-α
treatment. Antidepressants, e.g., selective serotonin reuptake inhibitors (SSRIs),
have therapeutic and prophylactic effects in interferon-α-induced depressive
syndromes.
An infection during pregnancy, especially in the second trimester, has been repeat-
edly described in mothers of offspring with schizophrenia. The maternal immune
response, e.g., to a pathogen, appears to be responsible for the increased risk for
schizophrenia in the offspring, and a link has been shown between an increased
maternal level of the pro-inflammatory cytokine IL-8 during the second trimester of
pregnancy and the later risk for schizophrenia in the offspring (Brown et al. 2004). A
link was also shown between the increased level of IL-8 in the mother and morpho-
logical changes typical of schizophrenia in the CNS of offspring who later developed
schizophrenia (Ellman et al. 2010).
Neurobiological Principles: Psycho-Neuro-Immuno-Endocrinology 41
The findings of studies on the cellular immune system in schizophrenia are incon-
sistent (Muller and Ackenheil 1995a). A great number of researchers, however,
found increases in CD4+ T lymphocytes (Henneberg et al. 1990; Muller et al.
1991). Increases in the total number of T lymphocytes (CD3+) were also described
(DeLisi et al. 1982), which is probably due mainly to the increased number of CD4+
cells. Increased levels of CD5+ B cells have also been observed (McAllister et al.
1989b). These findings were considered to indicate that the immune system was
activated. It should be noted, however, that many of the patients were being treated
with antipsychotics, which have marked effects on the immune system. Treatment
with antipsychotics is presumed to cause an increase in certain subgroups of CD4+
and B cells.
(IFN-γ, IL-12, TNF-α, sIL-2R). One must consider, however, that the cytokines
IL-12, IL-1-β, TGF-β, and IFN-γ have only paracrine effects through cell-cell
contact and no endocrine effects, so that levels of these cytokines in circulating
blood may not actually reflect the function of the immune system. Furthermore, the
meta-analysis did not consider the time of blood sampling or the medication status of
the patients (medicated vs. not medicated), even though antipsychotic medication is
known to affect cytokine levels in schizophrenia (Potvin et al. 2008). For example,
levels of sIL-2R have been shown to increase with antipsychotic medication (Muller
et al. 1997; Potvin et al. 2008).
IFN-γ is the key cytokine of the type 1 immune response. As mentioned above,
IFN-γ is a paracrine cytokine that has effects through cell-cell contact. Thus, the
in vitro stimulation of cytokine production measured by the enzyme-linked
immunospot (ELISpot) or ELISA methods are more valid approaches to demonstrate
the response of this cytokine than measuring plasma or serum levels. The ELISpot
method allows the (unstimulated) IFN-γ content of individual immune cells to be
measured. Studies on the in vitro production of IFN-γ after stimulation in whole
blood repeatedly using ELISA showed lower IFN-γ production in patients with
schizophrenia compared with health controls (Arolt et al. 2000; Wilke et al. 1996).
Avgustin et al. (2005) stimulated peripheral blood mononuclear cells (PBMC) and
found a greater production of IFN-γ in medicated schizophrenia patients than in
healthy controls. However, stimulated PBMC did not appear to reflect the in vivo
production of IFN-γ as well as the whole blood method.
A study in unmedicated patients with schizophrenia found lower serum levels of IFN-γ
and neopterin, a product of activated monocytes/macrophages, than in healthy controls;
this finding was interpreted as indicating a reduced activation of the type 1 immune
response (Sperner-Unterweger et al. 1999). Lower neopterin levels in body fluids such as
the CSF, blood, and urine directly reflect the activation of guanosine triphosphate
cyclohydrolase, which is induced by IFN-γ but not by other pro-inflammatory factors,
such as TNF-α and IL-1-β (Schennach et al. 2002). Neopterin thus reflects the
IFN-γ-induced immune status and in clinical studies is a marker for the activation of
the cellular immune system (Murr et al. 2002). The findings on neopterin are inconsistent,
however. Higher levels were also described in schizophrenia (Chittiprol et al. 2010), and
an earlier study found no difference in the urine concentration of neopterin between
unmedicated patients with schizophrenia and controls (Duch et al. 1984). Bechter et al.
(2010) found higher neopterin levels in the CSF in about one third of a mixed sample of
patients (schizophrenia or affective disorders) being treated with multiple medications. In
another study, the neopterin concentration in the CSF of ten unmedicated schizophrenia
patients did not differ significantly from controls (Nikkila et al. 2001). An increase in
neopterin has been described during antipsychotic treatment (Korte et al. 1998; Sperner-
Unterweger et al. 1999), which underlines that medication status is an important factor
when interpreting study results.
A study that found a reduced lymphocyte response after stimulation with specific
antigens provides further support for a decreased type 1 response in schizophrenia
(Muller et al. 1991).
Neurobiological Principles: Psycho-Neuro-Immuno-Endocrinology 43
Two studies that determined IL-2 in the CSF provided interesting results. One of
them described higher IL-2 levels in the CSF of untreated schizophrenia patients
than in controls (Licinio et al. 1993). This finding attracted a good deal of attention
because IL-2 can cause schizophrenia-like symptoms in a dose-dependent manner
(Denicoff et al. 1987). The other, carefully designed study found that IL-2 in the
CSF was the only predictor for a relapse in patients with schizophrenia after
discontinuation of haloperidol; 5-HIAA and HVA in CSF and psychopathological
variables, such as anxiety, were not significant predictors. Only when the variable
IL-2 was removed from the logistic regression model did products of catechol-
amine metabolism and early symptoms of anxiety become significant predictors
(McAllister et al. 1995).
Interestingly, higher IL-6 levels were found in the CSF of patients with chronic
schizophrenia than in healthy controls (Schwieler et al. 2015). This finding has been
frequently replicated in the serum of patients with schizophrenia and underlines the
role of inflammatory components.
44 N. Müller
A number of findings suggest that in the CNS of patients with schizophrenia, there is
an increased release of activating cytokines, which is associated with a stimulation of
the catecholaminergic neurotransmitter system. A possible explanation may be that
the peripheral immune system is initially not adequately activated, so that there is
insufficient counter-regulation and associated communication between the CNS and
peripheral immune system. This could be related to a defect in antigen recognition or
presentation. Treatment with neuroleptics appears to activate the peripheral immune
system and may thus counter-regulate cytokine release in the CNS.
Immune-based treatment for schizophrenia was first proposed many decades ago: In
the 1920s, the Nobel laureate Julius Ritter Wagner von Jauregg developed a vacci-
nation treatment for psychoses (the term schizophrenia had not yet been introduced)
(Wagner-Jauregg 1926). He successfully treated patients with vaccines for tubercu-
losis, malaria, and typhoid salmonella by stimulating the type 1 immune response
(Muller et al. 2005b). Although promising, this immune-based vaccination treatment
did not become established outside German-speaking countries, especially after
introduction of electroconvulsive therapy and then antipsychotics. Except for anti-
inflammatory drugs – specifically the cyclooxygenase-2 (COX-2) inhibitors, which
are discussed below – only preliminary data exist for other immune-based therapies.
with data from another 6-week study of risperidone and celecoxib add-on, and the
resulting group of n = 90 cases was analyzed. The analysis found that patients who
had been ill for 2 years or less benefited from the celecoxib add-on, whereas patients
with a longer disease duration showed no benefit compared with the placebo add-on
group.
These findings are in line with animal studies showing that the effects of COX-2
inhibition on cytokines, hormones, and in particular abnormal behavior depend on
both the duration of the changes and the duration of administration of the COX-2
inhibitor (Casolini et al. 2002). Thus, treatment of schizophrenia patients with a
COX-2 inhibitor appears to be most beneficial in the initial period of the disease
process. Support for this hypothesis is provided by a study that found no benefit of
celecoxib in chronic schizophrenia (Rapaport et al. 2005). To further test this
hypothesis, my group studied celecoxib as an add-on to amisulpride in first-
manifestation schizophrenia (Muller et al. 2010). We found positive effects of
celecoxib add-on treatment on the Positive and Negative Syndrome Scale
(PANSS) positive, negative, and total scores and on the general psychopathology
score (Muller 2010; Muller et al. 2010). An effect on cognition could be expected on
the basis of animal experiments with COX-2 inhibitors, which found that COX-2
inhibition has a direct effect on inflammation-inducted inhibition of long-term
potentiation (LTP), an animal model for cognition (Cumiskey et al. 2007; Muller
et al. 2005a). Another study found that the more pronounced effects on cognition in
an animal model with a genetic overexpression of COX-2 could be improved by
administration of a selective COX-2 inhibitor (Melnikova et al. 2006).
Meta-analyses corroborate findings that COX-2 inhibition has clear therapeutic
effects, especially in the early stages of a schizophrenic disorder (Nitta et al. 2013;
Sommer et al. 2012). The data on chronic schizophrenia are controversial, but brief
anti-inflammatory treatment cannot usually be expected to have an effect in chronic
inflammatory clinical processes.
Acetylsalicylic acid (ASA), another anti-inflammatory agent, has also shown
positive effects in schizophrenia spectrum disorders (Laan et al. 2010). A meta-
analysis of five double-blind studies of nonsteroidal anti-inflammatory drugs
(NSAIDs) in schizophrenia (four studies of celecoxib and one of ASS) found
significant effects of the drugs on overall symptoms and both positive and negative
symptoms (Sommer et al. 2012). However, a meta-analysis of eight studies (six of
celecoxib and two of ASS) in schizophrenia found significant effects in first-episode
but not chronic patients and in inpatients but not outpatients (Nitta et al. 2013). Some
studies have evaluated omega-3 fatty acids in schizophrenia, but results are incon-
sistent and the effect size is small (Ross et al. 2007). The results of a study by
Amminger and colleagues, however, are highly interesting (Amminger et al. 2010):
In a 12-month study on individuals with prodromal symptoms, i.e., a high-risk
population for schizophrenia, the group found a significantly lower rate of transition
to psychosis in the omega-3 fatty acid group than in the control group.
The studies described above indicate that the efficacy of anti-inflammatory
treatment is associated with the disease stage, i.e., such treatment shows less efficacy
in chronic schizophrenia (which may be related to the degree of neuroprogression).
46 N. Müller
Recently, changes in the cytokine system have become a focus of interest also in
depressive disorders.
Neurobiological Principles: Psycho-Neuro-Immuno-Endocrinology 47
Maes (1995) proposed that IL-6 hypersecretion plays a role in depressive disor-
ders. In depressed patients, he found both increased serum levels of IL-6 and soluble
IL-6 receptors (sIL-6R) and other signs of immune activation, in particular APPs,
which are stimulated by IL-6. The simultaneous increase in IL-6 and sIL-6R, which
congregate as a complex and may increase the biological activity of IL-6 through an
association with a signaling protein, underlines the important role of IL-6 in major
depressive disorder.
A significant correlation between high IL-6 levels and cortisol plasma levels was
also described in depressed patients (Maes et al. 1995b). This finding could be
expected because of the known stimulatory effect of IL-6 on the HPA axis, although
suppression of IL-6 could be expected in the peripheral immune system as a counter-
regulation. In the authors’ opinion, the correlation of the higher in vitro IL-6
production from the lymphocytes of depressed patients with the lower tryptophan
plasma levels in these patients was related to the influence of IL-6 on serotonin
metabolism (Maes et al. 1995b). Serotonin synthesis in the CNS is at least partially
regulated by the availability of tryptophan in blood, so that lower tryptophan levels
in blood can result in reduced serotonin synthesis in the CNS.
The following are characteristics for immune activation in depression: a high
number of circulating lymphocytes and phagocytic cells; highly regulated serum
levels of molecules that indicate an immune activation (neopterin, sIL-2R); higher
concentrations of positive APPs and lower levels of negative APPs; and increased
production of pro-inflammatory cytokines, such as IL-1, IL-2, TNF-α, and IL-6 by
activated macrophages and INF-γ by activated T lymphocytes. Several researchers
have also described a higher number of peripheral mononuclear cells in depression
(Herbert and Cohen 1993; Rothermundt et al. 2001; Seidel et al. 1996).
CRP is the most common marker for an inflammatory process. Studies on the
inflammatory hypothesis of depression found significantly higher CRP levels in
severely depressed patients; CRP levels decreased with antidepressant treatment
(Lanquillon et al. 2000). Another study showed a significant association between
CRP levels and the severity of depression (Hafner et al. 2008). An increase in CRP is
known to be caused by many other factors that also play a role in depressive
disorders, e.g., body mass index, nicotine abuse, and aging processes. Studies
show, however, that even after controlling for these factors, the association between
depression and increased CRP levels remains statistically significant.
Higher levels of sIL-2R and IL-1 in patients with major depression than in healthy
controls were described more than 20 years ago (Maes 1995). In the meantime, many
other studies of cytokines in major depression have been published. Meta-analyses
of cytokine studies in major depression have shown higher levels of
pro-inflammatory cytokines in peripheral blood (Goldsmith et al. 2016). This find-
ing, however, is not restricted to major depression but is also found in other
diagnostic groups. Moreover, similar patterns of higher levels of pro-inflammatory
48 N. Müller
cytokines and lower levels of anti-inflammatory cytokines in the CSF have been
described in major depression, and bipolar disorder, and schizophrenia (Wang and
Miller 2018).
Autoantibodies appear to be present in a subgroup of depressed patients. For
example, “antibrain antibodies” were found in 2 of 11 patients with an affective
disorder (DeLisi et al. 1985). Findings from the Department of Psychiatry at the
University Hospital Munich of anti-DNA autoantibodies in the CSF of a depressed
patient with scleroderma (Muller et al. 1992) also indicate that autoantibodies may
be involved in the development of depressive symptoms.
The findings on the cellular immune system are also inconsistent in depressive
disorders. However, the vast majority of researchers also found signs of an activation
of the peripheral immune system, such as increased levels of CD4+ cells. An increase
in the ratio of CD4+ to CD8+ has been described more often (Syvalahti et al. 1985),
although a few researchers found no changes or even a reduction in CD4+ cells
(Denney et al. 1988). One study found interesting evidence for a positive correlation
of the Hamilton Depression Scale with the number of CD4+ cells: The more severe
the depression, the higher the number of CD4+ cells (Levy et al. 1991). This was in
line with earlier findings of a positive correlation between scores on the Hamilton
Depression Scale and the number of CD4+ cells (Irwin et al. 1987).
In addition to inhibiting IL-1 and IL-6 release, COX-2 inhibitors influence the
serotonergic system in the CNS, either directly or via CNS immune mechanisms.
In a rat model, treatment with rofecoxib was followed by an increase of serotonin in
the frontal and the temporoparietal cortex (Sandrini et al. 2002), so that COX-2
inhibitors would be expected to show a clinical antidepressant effect. In the depres-
sion animal model of the bulbectomized rat, a decrease in hypothalamic cytokine
Neurobiological Principles: Psycho-Neuro-Immuno-Endocrinology 49
levels and a change in behavior have been observed after chronic celecoxib treat-
ment (Myint et al. 2007). In another animal model of depression, however, the mixed
COX-1/COX-2 inhibitor ASA showed an additional antidepressant effect by accel-
erating the antidepressant effect of fluoxetine (Brunello et al. 2006). A significant
therapeutic effect of the COX-2 inhibitor celecoxib in major depression was also
found in a randomized, double-blind add-on pilot study of reboxetine and celecoxib
versus reboxetine versus placebo (Muller et al. 2006). Interestingly, the ratio of KYN
to tryptophan, which represents the activity of the pro-inflammatory cytokine-driven
enzyme IDO, predicted the antidepressant response to the celecoxib therapy. Patients
with high activity of IDO, i.e., high pro-inflammatory activity, responded better to
celecoxib. Another RCT in 50 patients with major depression also showed a
significantly better outcome with the COX-2 inhibitor celecoxib plus fluoxetine
than with fluoxetine alone (Akhondzadeh et al. 2008). This finding was recently
replicated in a study that used a combination of sertraline and celecoxib in 40 patients
with depression (Abbasi et al. 2012). Interestingly, the blood levels of IL-6 predicted
the antidepressant response in both the sertraline (plus placebo) and the celecoxib
(plus sertraline) groups.
A meta-analysis on the efficacy of adjunctive celecoxib treatment for patients
with major depression included 150 patients and concluded that adjunctive treatment
with NSAIDs, particularly celecoxib, can be a promising strategy for patients with
depressive disorders. However, future studies with larger sample sizes and longer
study durations are needed to confirm the efficacy and tolerability of NSAIDs in
depression (Na et al. 2014).
The results of an international multicenter add-on RCT of the selective COX-2
inhibitor cimicoxib to sertraline in patients with major depression were also inter-
esting: Although there was no benefit of cimicoxib over sertraline and placebo in the
whole group of depressed patients, there was a significantly better outcome of
cimicoxib and sertraline in the subgroup of severely depressed patients (Hamilton
Depression Score < 25 at baseline) (Müller et al., manuscript in preparation). As
discussed below, because the placebo response in depression studies is generally
very high, and even higher in mild to moderately depressed patients, a therapeutic
benefit can be shown more easily in severely depressed patients, particularly if
sample sizes are small.
In the European-wide study of the role of inflammation in mood disorders
(MOODINFLAME), a therapeutic multicenter RCT of celecoxib add-on to sertraline
compared with sertraline and placebo was performed in a subgroup of patients with
major depression (n = 53). This study found a statistically significant benefit of
COX-2 treatment only in the group of patients who completed the study (completer
analysis) and not in the last observation carried forward (LOCF) analysis (Leitner B.,
Müller N. et al., manuscript in preparation).
Although the abovementioned studies have some limitations, they represent
further small pieces in the puzzle of the effects of anti-inflammatory treatment and
in particular of COX-2 inhibition in major depression. Future studies should place
more emphasis on sample size; the high rate of placebo response, especially in
studies of an add-on to an effective antidepressant; and the severity of depression.
50 N. Müller
In a first study, the anti-TNF-α antibody infliximab, which blocks the interaction of
TNF-α with cell-surface receptors and was developed for the treatment of inflam-
matory joint disorders and psoriasis, showed a highly significant effect on
Neurobiological Principles: Psycho-Neuro-Immuno-Endocrinology 51
As mentioned above, elevated levels of IL-6 have been reported in the peripheral
blood and CSF of patients with depression. Therefore, the IL-6 complex is also an
interesting target for anti-cytokine treatment. Two studies of the anti-IL-6 antibody
tocilizumab (Gossec et al. 2015; Traki et al. 2014) were included in the meta-analysis
mentioned above (Kappelmann et al. 2018). In both open-labeled studies, the
concomitant symptoms of anxiety and depression improved. However, valid and
52 N. Müller
reliable data of anti-IL-6 treatment in patients with major depression are still
missing, and the correct target for anti-IL-6 treatment is a matter of discussion.
IL-6 levels are well-known to be higher in the CSF than in peripheral blood, and
patients with depression have higher IL-6 CSF levels than controls; therefore, CNS
IL-6 was postulated to be the most promising therapeutic target. Moreover, the
preclinical model of chronically stressed rats shows an overexpression of IL-6 in
the cortex. On the other hand, CSF IL-6 levels cannot be seen as completely distinct
from the IL-6 concentration in the CSF or IL-6 expression in the brain because there
are many interconnections between central and peripheral IL-6 concentrations.
Therefore, peripheral IL-6 might be a promising target to treat depressive symptoms
(Yang and Hashimoto 2015). This view is supported by the findings that peripheral
IL-6 promotes resilience versus susceptibility to chronic inescapable electric stress in
an animal model of depression (Yang et al. 2015a) and serum IL-6 predicts the
antidepressant response to ketamine (Yang et al. 2015b) in depressed patients. Maes
et al. (2014) propose, however, that a more promising strategy to inhibit IL-6 might
be to increase inhibition of IL-6 trans-signaling by increasing the soluble glycopro-
tein 130 (sgp 130), part of the IL-6 complex, while allowing maintenance of IL-6R
signaling. Sirukumab, a monoclonal antibody against IL-6 distinct from tocilizumab,
has also been proposed for use in depression (Zhou et al. 2017). Sirukumab targets
the IL-6 signaling pathway, inhibiting both the pro- and anti-inflammatory effects of
the pleiotropic cytokine IL-6 (Zhou et al. 2017). Beneficial effects of sirukumab have
been shown in other inflammatory diseases, such as lupus erythematosus and
rheumatoid arthritis.
Interestingly, there are also preliminary findings that angiotensin II AT1 receptor
blockade has anti-inflammatory effects in the CNS and ameliorates stress, anxiety,
and CNS inflammation (Saavedra et al. 2011; Benicky et al. 2011).
A recently published broader meta-analysis of inflammation-related therapeutic
approaches included ten publications reporting on 14 trials (6,262 participants). It
showed very interesting results of anti-inflammatory treatment in major depression:
ten trials evaluated the use of NSAIDs (n = 4,258) and four investigated cytokine
inhibitors (n = 2,004). The pooled effect estimate suggested that anti-inflammatory
treatment reduced depressive symptoms compared with placebo. This effect was
observed in studies including patients with depression and depressive symptoms.
The heterogeneity of the studies was not explained by differences in inclusion of
clinical depression vs. depressive symptoms or use of NSAIDs vs. cytokine inhib-
itors. Sub-analyses emphasized the antidepressant properties of celecoxib on remis-
sion and response. Among the six studies reporting on adverse effects, no evidence
was found of an increased number of gastrointestinal or cardiovascular events after
6 weeks or infections after 12 weeks of anti-inflammatory treatment compared with
placebo. All trials were associated with a high risk of bias because of potentially
Neurobiological Principles: Psycho-Neuro-Immuno-Endocrinology 53
Antipsychotics
(Bertini et al. 1993). In mice, chlorpromazine also protects from the toxic effects of
IL-1 and from endotoxin-induced toxic effects of TNF.
Particular attention has been paid to the immunological effects of clozapine
because such effects were implicated in its increased risk for agranulocytosis.
Clozapine was shown to have an inhibitory effect on granulocyte-macrophage
colony-stimulating factor (GM-CSF) (Sperner-Unterweger et al. 1993).
Antidepressants
Outlook
diseases, tumor immunology, and transplantation medicine. For example, the rele-
vance for processes in the CNS and in particular for neuronal processes of many of
the newly discovered cytokines, some of which are expressed in the CNS, remains
completely unknown.
A model of the immunopathogenesis of mental disorders must consider the
effects of cytokines in the CNS and the functions of both the peripheral cellular
immune system and the blood-brain barrier as well as immunogenetics, which may
contribute to an increased susceptibility. Such a model is not only a fascinating
hypothesis – an increasing number of findings support the assumption that cyto-
kines, perhaps through their regulatory effects on neurotransmitters, play an impor-
tant role in the pathogenesis of mental disorders. Immunological effects of
psychopharmaceuticals may not only be a side effect but may also contribute to
the therapeutic efficacy. This may have far-reaching effects on immuno-
psychopharmacology (Müller 1995).
Cross-References
Acknowledgments The author thanks the foundation ‘Immunität & Seele’ for its support
References
Abbasi SH, Hosseini F, Modabbernia A, Ashrafi M, Akhondzadeh S. Effect of celecoxib add-on
treatment on symptoms and serum IL-6 concentrations in patients with major depressive
disorder: randomized double-blind placebo-controlled study. J Affect Disord. 2012;141:308–14.
Ader R, Felten DL. Psychoneuroimmunology. San Diego/New York/Toronto: Academic
Press; 1991.
Ahuja N, Carroll BT. Possible anti-catatonic effects of minocycline in patients with schizophrenia.
Prog Neuro-Psychopharmacol Biol Psychiatry. 2007;31:968–9.
Akhondzadeh S, Tabatabaee M, Amini H, Ahmadi Abhari SA, Abbasi SH, Behnam B. Celecoxib as
adjunctive therapy in schizophrenia: a double-blind, randomized and placebo-controlled trial.
Schizophr Res. 2007;90:179–85.
Akhondzadeh S, Jafari S, Raisi F, Ghoreishi A. A clinical trial of adjunctive celecoxib treatment in
patients with major depression: a double blind and placebo controlled trial. Paper presented at
the WPA Section on Immunology and psychiatry – psychoneuroimmunology training work-
shop, Seeon, Munich, 11–13 July 2008.
Alonso R, Chaudieu I, Diorio J, Krishnamurthy A, Quirion R, Boksa P. Interleukin-2 modulates
evoked release of [3H]dopamine in rat cultured mesencephalic cells. J Neurochem.
1993;61:1284–90.
Amminger GP, Schafer MR, Papageorgiou K, Klier CM, Cotton SM, Harrigan SM, et al. Long-
chain omega-3 fatty acids for indicated prevention of psychotic disorders: a randomized,
placebo-controlled trial. Arch Gen Psychiatry. 2010;67:146–54.
56 N. Müller
Araujo DM, Lapchak PA, Chabot JG, Nair NP, Quirion R. Characterization and possible role of
growth factor and lymphokine receptors in the regulation of cholinergic function in the
mammalian brain. Prog Clin Biol Res. 1989;317:423–36.
Arolt V, Rothermundt M, Wandinger KP, Kirchner H. Decreased in vitro production of interferon-
gamma and interleukin-2 in whole blood of patients with schizophrenia during treatment. Mol
Psychiatry. 2000;5:150–8.
Avgustin B, Wraber B, Tavcar R. Increased Th1 and Th2 immune reactivity with relative Th2
dominance in patients with acute exacerbation of schizophrenia. Croat Med J. 2005;46:268–74.
Baker GA, Santalo R, Blumenstein J. Effect of psychotropic agents upon the blastogenic response
of human t-lymphocytes. Biol Psychiatry. 1977;12:159–69.
Balschun D, Wetzel W, Del Rey A, Pitossi F, Schneider H, Zuschratter W, et al. Interleukin-6: a
cytokine to forget. FASEB J. 2004;18:1788–90.
Bechter K, Reiber H, Herzog S, Fuchs D, Tumani H, Maxeiner HG. Cerebrospinal fluid analysis in
affective and schizophrenic spectrum disorders: identification of subgroups with immune
responses and blood-CSF barrier dysfunction. J Psychiatr Res. 2010;44:321–30.
Benicky J, Sanchez-Lemus E, Honda M, Pang T, Orecna M, Wang J, et al. Angiotensin II AT1
receptor blockade ameliorates brain inflammation. Neuropsychopharmacology.
2011;36:857–70.
Benros ME, Mortensen PB, Eaton WW. Autoimmune diseases and infections as risk factors for
schizophrenia. Ann N Y Acad Sci. 2012;1262:56–66.
Benveniste EN. Inflammatory cytokines within the central nervous system: sources, function, and
mechanism of action. Am J Phys. 1992;263:C1–16.
Berrios GE, Quemada JI. Depressive illness in multiple sclerosis. Clinical and theoretical aspects of
the association. Br J Psychiatry. 1990;156:10–6.
Bertini R, Garattini S, Delgado R, Ghezzi P. Pharmacological activities of chlorpromazine involved
in the inhibition of tumour necrosis factor production in vivo in mice. Immunology.
1993;79:217–9.
Besedovsky H, del Rey A, Sorkin E, Dinarello CA. Immunoregulatory feedback between
interleukin-1 and glucocorticoid hormones. Science. 1986;233:652–4.
Blum-Degen D, Muller T, Kuhn W, Gerlach M, Przuntek H, Riederer P. Interleukin-1 beta and
interleukin-6 are elevated in the cerebrospinal fluid of Alzheimer’s and de novo Parkinson’s
disease patients. Neurosci Lett. 1995;202:17–20.
Bohar Z, Toldi J, Fulop F, Vecsei L. Changing the face of kynurenines and neurotoxicity:
therapeutic considerations. Int J Mol Sci. 2015;16:9772–93.
Boros FA, Bohar Z, Vecsei L. Genetic alterations affecting the genes encoding the enzymes of
the kynurenine pathway and their association with human diseases. Mutat Res.
2018;776:32–45.
Braida D, Sacerdote P, Panerai AE, Bianchi M, Aloisi AM, Iosue S, et al. Cognitive function in
young and adult IL (interleukin)-6 deficient mice. Behav Brain Res. 2004;153:423–9.
Bresee C, Rapaport MH. Persistently increased serum soluble interleukin-2 receptors in continu-
ously ill patients with schizophrenia. Int J Neuropsychopharmacol. 2009;12:861–5.
Brown AS, Hooton J, Schaefer CA, Zhang H, Petkova E, Babulas V, et al. Elevated maternal
interleukin-8 levels and risk of schizophrenia in adult offspring. Am J Psychiatry.
2004;161:889–95.
Bruce LC, Peebles AMS. Clinical and experimental observations in catatonia. J Ment Sci.
1903;49:614–28.
Brunello N, Alboni S, Capone G, Benatti C, Blom JM, Tascedda F, et al. Acetylsalicylic acid
accelerates the antidepressant effect of fluoxetine in the chronic escape deficit model of
depression. Int Clin Psychopharmacol. 2006;21:219–25.
Capuron L, Gumnick JF, Musselman DL, Lawson DH, Reemsnyder A, Nemeroff CB, et al.
Neurobehavioral effects of interferon-alpha in cancer patients: phenomenology and paroxetine
responsiveness of symptom dimensions. Neuropsychopharmacology. 2002;26:643–52.
Neurobiological Principles: Psycho-Neuro-Immuno-Endocrinology 57
Casolini P, Catalani A, Zuena AR, Angelucci L. Inhibition of COX-2 reduces the age-dependent
increase of hippocampal inflammatory markers, corticosterone secretion, and behavioral impair-
ments in the rat. J Neurosci Res. 2002;68:337–43.
Chaudhry IB, Hallak J, Husain N, Minhas F, Stirling J, Richardson P, et al. Minocycline benefits
negative symptoms in early schizophrenia: a randomised double-blind placebo-controlled
clinical trial in patients on standard treatment. J Psychopharmacol. 2012;26:1185–93.
Chittiprol S, Venkatasubramanian G, Neelakantachar N, Babu SV, Reddy NA, Shetty KT, et al.
Oxidative stress and neopterin abnormalities in schizophrenia: a longitudinal study. J Psychiatr
Res. 2010;44:310–3.
Cumiskey D, Curran BP, Herron CE, O’Connor JJ. A role for inflammatory mediators in the IL-18
mediated attenuation of LTP in the rat dentate gyrus. Neuropharmacology. 2007;52:1616–23.
Dameshek W. White blood cells in dementia praecox and dementia paralytica. Arch Neurol
Psychiatr. 1930;24:855.
DeLisi LE, Goodman S, Neckers LM, Wyatt RJ. An analysis of lymphocyte subpopulations in
schizophrenic patients. Biol Psychiatry. 1982;17:1003–9.
DeLisi LE, Weber RJ, Pert CB. Are there antibodies against brain in sera from schizophrenic
patients? Review and prospectus. Biol Psychiatry. 1985;20:110–5.
Denicoff KD, Rubinow DR, Papa MZ, Simpson C, Seipp CA, Lotze MT, et al. The neuropsychiatric
effects of treatment with interleukin-2 and lymphokine-activated killer cells. Ann Intern Med.
1987;107:293–300.
Denney DR, Stephenson LA, Penick EC, Weller RA. Lymphocyte subclasses and depression.
J Abnorm Psychol. 1988;97:499–502.
Duch DS, Woolf JH, Nichol CA, Davidson JR, Garbutt JC. Urinary excretion of biopterin and
neopterin in psychiatric disorders. Psychiatry Res. 1984;11:83–9.
Ellman LM, Deicken RF, Vinogradov S, Kremen WS, Poole JH, Kern DM, Tsai WY, Schaefer CA,
Brown AS.Structural brain alterations in schizophrenia following fetal exposure to the inflam-
matory cytokine interleukin-8.Schizophr Res. 2010;121(1–3):46–54. https://doi.org/10.1016/j.
schres.2010.05.014. Epub 2010 Jun 9
Eikelenboom P, Rozemuller JM, Kraal G, Stam FC, McBride PA, Bruce ME, et al. Cerebral amyloid
plaques in Alzheimer’s disease but not in scrapie-affected mice are closely associated with a
local inflammatory process. Virchows Arch B Cell Pathol Incl Mol Pathol. 1991;60:329–36.
Frank MG, Baratta MV, Sprunger DB, Watkins LR, Maier SF. Microglia serve as a neuroimmune
substrate for stress-induced potentiation of CNS pro-inflammatory cytokine responses. Brain
Behav Immun. 2007;21:47–59.
Furukawa H, del Rey A, Monge-Arditi G, Besedovsky HO. Interleukin-1, but not stress, stimulates
glucocorticoid output during early postnatal life in mice. Ann N Y Acad Sci. 1998;840:117–22.
Gimeno D, Marmot MG, Singh-Manoux A. Inflammatory markers and cognitive function in
middle-aged adults: the Whitehall II study. Psychoneuroendocrinology. 2008;33:1322–34.
Goldsmith DR, Rapaport MH, Miller BJ. A meta-analysis of blood cytokine network alterations in
psychiatric patients: comparisons between schizophrenia, bipolar disorder and depression. Mol
Psychiatry. 2016;21:1696–709.
Gossec L, Steinberg G, Rouanet S, Combe B. Fatigue in rheumatoid arthritis: quantitative findings
on the efficacy of tocilizumab and on factors associated with fatigue. The French multicentre
prospective PEPS study. Clin Exp Rheumatol. 2015;33:664–70.
Grosskopf A, Muller N, Malo A, Wank R. Potential role for the narcolepsy- and multiple sclerosis-
associated HLA allele DQB10602 in schizophrenia subtypes. Schizophr Res. 1998;30:187–9.
Gruber L, Bunse T, Weidinger E, Reichard H, Muller N. Adjunctive recombinant human interferon
gamma-1b for treatment-resistant schizophrenia in 2 patients. J Clin Psychiatry.
2014;75:1266–7.
Haack M, Hinze-Selch D, Fenzel T, Kraus T, Kuhn M, Schuld A, et al. Plasma levels of cytokines
and soluble cytokine receptors in psychiatric patients upon hospital admission: effects of
confounding factors and diagnosis. J Psychiatr Res. 1999;33:407–18.
58 N. Müller
Hafner S, Baghai TC, Eser D, Schule C, Rupprecht R, Bondy B, et al. C-reactive protein is
associated with polymorphisms of the angiotensin-converting enzyme gene in major depressed
patients. J Psychiatr Res. 2008;42:163–5.
Hampel H, Schwarz MJ, Kötter HU, Schneider C, Müller N. Cell adhesion molecules in the central
nervous system: significance and therapeutic perspectives in neuropsychiatric disorders. Drug
News Perspect. 1996;9:69–81.
Hampel H, Kotter HU, Padberg F, Korschenhausen DA, Moller HJ. Oligoclonal bands and blood–
cerebrospinal-fluid barrier dysfunction in a subset of patients with Alzheimer disease: compar-
ison with vascular dementia, major depression, and multiple sclerosis. Alzheimer Dis Assoc
Disord. 1999;13:9–19.
Hayley S, Wall P, Anisman H. Sensitization to the neuroendocrine, central monoamine and
behavioural effects of murine tumor necrosis factor-alpha: peripheral and central mechanisms.
Eur J Neurosci. 2002;15:1061–76.
Henneberg A, Riedl B, Dumke HO, Kornhuber HH. T-lymphocyte subpopulations in schizophrenic
patients. Eur Arch Psychiatry Neurol Sci. 1990;239:283–4.
Herbert TB, Cohen S. Depression and immunity: a meta-analytic review. Psychol Bull.
1993;113:472–86.
Heyser CJ, Masliah E, Samimi A, Campbell IL, Gold LH. Progressive decline in avoidance learning
paralleled by inflammatory neurodegeneration in transgenic mice expressing interleukin 6 in the
brain. Proc Natl Acad Sci U S A. 1997;94:1500–5.
Irwin M, Smith TL, Gillin JC. Low natural killer cytotoxicity in major depression. Life Sci.
1987;41:2127–33.
Kappelmann N, Lewis G, Dantzer R, Jones PB, Khandaker GM. Antidepressant activity of anti-
cytokine treatment: a systematic review and meta-analysis of clinical trials of chronic inflam-
matory conditions. Mol Psychiatry. 2018;23:335–43.
Kipnis J, Cohen H, Cardon M, Ziv Y, Schwartz M. T cell deficiency leads to cognitive dysfunction:
implications for therapeutic vaccination for schizophrenia and other psychiatric conditions. Proc
Natl Acad Sci U S A. 2004;101:8180–5.
Knight JG. Dopamine-receptor-stimulating autoantibodies: a possible cause of schizophrenia.
Lancet. 1982;2:1073–6.
Kohler O, Benros ME, Nordentoft M, Farkouh ME, Iyengar RL, Mors O, et al. Effect of anti-
inflammatory treatment on depression, depressive symptoms, and adverse effects: a systematic
review and meta-analysis of randomized clinical trials. JAMA Psychiat. 2014;71:1381–91.
Kohler O, Gasse C, Petersen L, Ingstrup KG, Nierenberg AA, Mors O, et al. The effect of
concomitant treatment with SSRIs and statins: a population-based study. Am J Psychiatry.
2016;173:807–15.
Korte S, Arolt V, Peters M, Weitzsch C, Rothermundt M, Kirchner H. Increased serum neopterin
levels in acutely ill and recovered schizophrenic patients. Schizophr Res. 1998;32:63–7.
Kurtz G, Muller N. The antiphospholipid syndrome and psychosis. Am J Psychiatry.
1994;151:1841–2.
Laan W, Grobbee DE, Selten JP, Heijnen CJ, Kahn RS, Burger H. Adjuvant aspirin therapy reduces
symptoms of schizophrenia spectrum disorders: results from a randomized, double-blind,
placebo-controlled trial. J Clin Psychiatry. 2010;71:520–7.
Lanquillon S, Krieg JC, Bening-Abu-Shach U, Vedder H. Cytokine production and treatment
response in major depressive disorder. Neuropsychopharmacology. 2000;22:370–9.
Lapchak PA. A role for interleukin-2 in the regulation of striatal dopaminergic function.
Neuroreport. 1992;3:165–8.
Lehmann-Facius H. Serologisch-analytische Versuche mit Liquores und Seren von Schizophrenen.
Allg Z Psychiatrie. 1939;110:232–43.
Levkovitz Y, Mendlovich S, Riwkes S, Braw Y, Levkovitch-Verbin H, Gal G, et al. A double-blind,
randomized study of minocycline for the treatment of negative and cognitive symptoms in early-
phase schizophrenia. J Clin Psychiatry. 2010;71:138–49.
Neurobiological Principles: Psycho-Neuro-Immuno-Endocrinology 59
Levy EM, Borrelli DJ, Mirin SM, Salt P, Knapp PH, Peirce C, et al. Biological measures and
cellular immunological function in depressed psychiatric inpatients. Psychiatry Res.
1991;36:157–67.
Lewandowski G, Hobbs MV, Bloom FE. Alteration of intracerebral cytokine production in mice
infected with herpes simplex virus types 1 and 2. J Neuroimmunol. 1994;55:23–34.
Licinio J, Seibyl JP, Altemus M, Charney DS, Krystal JH. Elevated CSF levels of interleukin-2 in
neuroleptic-free schizophrenic patients. Am J Psychiatry. 1993;150:1408–10.
Lieberman AP, Pitha PM, Shin HS, Shin ML. Production of tumor necrosis factor and other
cytokines by astrocytes stimulated with lipopolysaccharide or a neurotropic virus. Proc Natl
Acad Sci U S A. 1989;86:6348–52.
Maes M. Evidence for an immune response in major depression: a review and hypothesis. Prog
Neuro-Psychopharmacol Biol Psychiatry. 1995;19:11–38.
Maes M, Bosmans E, Calabrese J, Smith R, Meltzer HY. Interleukin-2 and interleukin-6 in
schizophrenia and mania: effects of neuroleptics and mood stabilizers. J Psychiatr Res.
1995a;29:141–52.
Maes M, Bosmans E, Meltzer HY. Immunoendocrine aspects of major depression. Relationships
between plasma interleukin-6 and soluble interleukin-2 receptor, prolactin and cortisol. Eur
Arch Psychiatry Clin Neurosci. 1995b;245:172–8.
Maes M, Bosmans E, Kenis G, De Jong R, Smith RS, Meltzer HY. In vivo immunomodulatory
effects of clozapine in schizophrenia. Schizophr Res. 1997;26:221–5.
Maes M, Anderson G, Kubera M, Berk M. Targeting classical IL-6 signalling or IL-6 trans-
signalling in depression? Expert Opin Ther Targets. 2014;18:495–512.
McAllister CG, Rapaport MH, Pickar D, Paul SM. Effects of short-term administration of antipsy-
chotic drugs on lymphocyte subsets in schizophrenic patients. Arch Gen Psychiatry.
1989a;46:956–7.
McAllister CG, Rapaport MH, Pickar D, Podruchny TA, Christison G, Alphs LD, et al. Increased
numbers of CD5+ B lymphocytes in schizophrenic patients. Arch Gen Psychiatry.
1989b;46:890–4.
McAllister CG, van Kammen DP, Rehn TJ, Miller AL, Gurklis J, Kelley ME, et al. Increases in CSF
levels of interleukin-2 in schizophrenia: effects of recurrence of psychosis and medication
status. Am J Psychiatry. 1995;152:1291–7.
McGeer PL, Rogers J, McGeer EG. Inflammation, antiinflammatory agents, and Alzheimer’s
disease: the last 22 years. J Alzheimers Dis. 2016;54:853–7.
Melnikova T, Savonenko A, Wang Q, Liang X, Hand T, Wu L, et al. Cycloxygenase-2 activity
promotes cognitive deficits but not increased amyloid burden in a model of Alzheimer’s disease
in a sex-dimorphic pattern. Neuroscience. 2006;141:1149–62.
Merrill JE. Tumor necrosis factor alpha, interleukin 1 and related cytokines in brain development:
normal and pathological. Dev Neurosci. 1992;14:1–10.
Miller BJ, Buckley PF. The case for adjunctive monoclonal antibody immunotherapy in schizo-
phrenia. Psychiatr Clin North Am. 2016;39:187–98.
Miller AH, Lackner C. Tricyclic antidepressants and immunity. In: Miller AH, editor. Depressive
disorders and immunity. Washington: American Psychiatric Press; 1989.
Miller BJ, Buckley P, Seabolt W, Mellor A, Kirkpatrick B. Meta-analysis of cytokine
alterations in schizophrenia: clinical status and antipsychotic effects. Biol Psychiatry.
2011;70:663–71.
Mills CD, Kincaid K, Alt JM, Heilman MJ, Hill AM. M-1/M-2 macrophages and the Th1/Th2
paradigm. J Immunol. 2000;164:6166–73.
Mittleman BB, Castellanos FX, Jacobsen LK, Rapoport JL, Swedo SE, Shearer GM. Cerebrospinal
fluid cytokines in pediatric neuropsychiatric disease. J Immunol. 1997;159:2994–9.
Mizoguchi H, Takuma K, Fukakusa A, Ito Y, Nakatani A, Ibi D, et al. Improvement by minocycline
of methamphetamine-induced impairment of recognition memory in mice. Psychopharmacol-
ogy. 2008;196:233–41.
60 N. Müller
double-blind, randomized, placebo controlled, add-on pilot study to reboxetine. Mol Psychiatry.
2006;11:680–4.
Muller N, Krause D, Dehning S, Musil R, Schennach-Wolff R, Obermeier M, et al. Celecoxib
treatment in an early stage of schizophrenia: results of a randomized, double-blind, placebo-
controlled trial of celecoxib augmentation of amisulpride treatment. Schizophr Res.
2010;121:118–24.
Muller N, Myint AM, Schwarz MJ. Immunological treatment options for schizophrenia. Curr
Pharm Biotechnol. 2012a;13:1606–13.
Muller N, Myint AM, Schwarz MJ. Inflammation in schizophrenia. Adv Protein Chem Struct Biol.
2012b;88:49–68.
Murr C, Widner B, Wirleitner B, Fuchs D. Neopterin as a marker for immune system activation.
Curr Drug Metab. 2002;3:175–87.
Myint AM, Steinbusch HW, Goeghegan L, Luchtman D, Kim YK, Leonard BE. Effect of the
COX-2 inhibitor celecoxib on behavioural and immune changes in an olfactory bulbectomised
rat model of depression. Neuroimmunomodulation. 2007;14:65–71.
Na KS, Lee KJ, Lee JS, Cho YS, Jung HY. Efficacy of adjunctive celecoxib treatment for patients
with major depressive disorder: a meta-analysis. Prog Neuro-Psychopharmacol Biol Psychiatry.
2014;48:79–85.
Nair A, Bonneau RH. Stress-induced elevation of glucocorticoids increases microglia proliferation
through NMDA receptor activation. J Neuroimmunol. 2006;171:72–85.
Nemni R, Iannaccone S, Quattrini A, Smirne S, Sessa M, Lodi M, et al. Effect of chronic treatment
with recombinant interleukin-2 on the central nervous system of adult and old mice. Brain Res.
1992;591:248–52.
Nikkila HV, Muller K, Ahokas A, Rimon R, Andersson LC. Increased frequency of activated
lymphocytes in the cerebrospinal fluid of patients with acute schizophrenia. Schizophr Res.
2001;49:99–105.
Nistico G, De Sarro G. Is interleukin 2 a neuromodulator in the brain? Trends Neurosci.
1991;14:146–50.
Nitta M, Kishimoto T, Muller N, Weiser M, Davidson M, Kane JM, et al. Adjunctive use of
nonsteroidal anti-inflammatory drugs for schizophrenia: a meta-analytic investigation of ran-
domized controlled trials. Schizophr Bull. 2013;39:1230–41.
Norris JG, Benveniste EN. Interleukin-6 production by astrocytes: induction by the neurotransmit-
ter norepinephrine. J Neuroimmunol. 1993;45:137–45.
Ogawa A, Yoshizaki A, Yanaba K, Ogawa F, Hara T, Muroi E, et al. The differential role of
L-selectin and ICAM-1 in Th1-type and Th2-type contact hypersensitivity. J Invest Dermatol.
2010;130:1558–70.
Pennisi E. Neuroimmunology. Tracing molecules that make the brain-body connection. Science.
1997;275:930–1.
Plangar I, Majlath Z, Vecsei L. Kynurenines in cognitive functions: their possible role in depression.
Neuropsychopharmacol Hung. 2012;14:239–44.
Plata-Salaman CR. Immunoregulators in the nervous system. Neurosci Biobehav Rev.
1991;15:185–215.
Pollmacher T, Hinze-Selch D, Mullington J, Holsboer F. Clozapine-induced increase in plasma
levels of soluble interleukin-2 receptors. Arch Gen Psychiatry. 1995;52:877–8.
Pollmacher T, Schuld A, Kraus T, Haack M, Hinze-Selch D. On the clinical relevance of clozapine-
triggered release of cytokines and soluble cytokine-receptors. Fortschr Neurol Psychiatr.
2001;69(Suppl 2):S65–74.
Potvin S, Stip E, Sepehry AA, Gendron A, Bah R, Kouassi E. Inflammatory cytokine alterations in
schizophrenia: a systematic quantitative review. Biol Psychiatry. 2008;63:801–8.
Purcell SM, Wray NR, Stone JL, Visscher PM, O’Donovan MC, Sullivan PF, et al. Common polygenic
variation contributes to risk of schizophrenia and bipolar disorder. Nature. 2009;460:748–52.
Raison CL, Rutherford RE, Woolwine BJ, Shuo C, Schettler P, Drake DF, et al. A randomized
controlled trial of the tumor necrosis factor antagonist infliximab for treatment-resistant depres-
sion: the role of baseline inflammatory biomarkers. JAMA Psychiat. 2013;70:31–41.
62 N. Müller
Rapaport MH, Delrahim KK, Bresee CJ, Maddux RE, Ahmadpour O, Dolnak D. Celecoxib
augmentation of continuously ill patients with schizophrenia. Biol Psychiatry. 2005;57:1594–6.
Riedel M, Spellmann I, Schwarz MJ, Strassnig M, Sikorski C, Moller HJ, et al. Decreased T cellular
immune response in schizophrenic patients. J Psychiatr Res. 2007;41:3–7.
Ross BM, Seguin J, Sieswerda LE. Omega-3 fatty acids as treatments for mental illness: which
disorder and which fatty acid? Lipids Health Dis. 2007;6:21.
Rothermundt M, Arolt V, Fenker J, Gutbrodt H, Peters M, Kirchner H. Different immune patterns in
melancholic and non-melancholic major depression. Eur Arch Psychiatry Clin Neurosci.
2001;251:90–7.
Saavedra JM, Sanchez-Lemus E, Benicky J. Blockade of brain angiotensin II AT1 receptors
ameliorates stress, anxiety, brain inflammation and ischemia: therapeutic implications.
Psychoneuroendocrinology. 2011;36:1–18.
Sandrini M, Vitale G, Pini LA. Effect of rofecoxib on nociception and the serotonin system in the rat
brain. Inflamm Res. 2002;51:154–9.
Schennach H, Murr C, Gachter E, Mayersbach P, Schonitzer D, Fuchs D. Factors influencing serum
neopterin concentrations in a population of blood donors. Clin Chem. 2002;48:643–5.
Schizophrenia Working Group of the Psychiatric Genomics Consortium. Biological insights from
108 schizophrenia-associated genetic loci. Nature. 2014;511:421–7.
Schwarz MJ, Ackenheil M, Riedel M, Muller N. Blood-cerebrospinal fluid barrier impairment as
indicator for an immune process in schizophrenia. Neurosci Lett. 1998;253:201–3.
Schwarz MJ, Riedel M, Ackenheil M, Muller N. Decreased levels of soluble intercellular adhesion
molecule-1 (sICAM-1) in unmedicated and medicated schizophrenic patients. Biol Psychiatry.
2000;47:29–33.
Schwieler L, Erhardt S, Erhardt C, Engberg G. Prostaglandin-mediated control of rat brain
kynurenic acid synthesis–opposite actions by COX-1 and COX-2 isoforms. J Neural Transm
(Vienna). 2005;112:863–72.
Schwieler L, Larsson MK, Skogh E, Kegel ME, Orhan F, Abdelmoaty S, et al. Increased levels of
IL-6 in the cerebrospinal fluid of patients with chronic schizophrenia – significance for activa-
tion of the kynurenine pathway. J Psychiatry Neurosci. 2015;40:126–33.
Seidel A, Arolt V, Hunstiger M, Rink L, Behnisch A, Kirchner H. Major depressive disorder is
associated with elevated monocyte counts. Acta Psychiatr Scand. 1996;94:198–204.
Shi J, Levinson DF, Duan J, Sanders AR, Zheng Y, Pe’er I, et al. Common variants on chromosome
6p22.1 are associated with schizophrenia. Nature. 2009;460:753–7.
Sluzewska A, Rybakowski JK, Laciak M, Mackiewicz A, Sobieska M, Wiktorowicz K. Interleukin-
6 serum levels in depressed patients before and after treatment with fluoxetine. Ann N Y Acad
Sci. 1995;762:474–6.
Soliven B, Albert J. Tumor necrosis factor modulates the inactivation of catecholamine secretion in
cultured sympathetic neurons. J Neurochem. 1992;58:1073–8.
Sommer IE, de Witte L, Begemann M, Kahn RS. Nonsteroidal anti-inflammatory drugs in schizo-
phrenia: ready for practice or a good start? A meta-analysis. J Clin Psychiatry. 2012;73:414–9.
Sommer IE, van Westrhenen R, Begemann MJ, de Witte LD, Leucht S, Kahn RS. Efficacy of anti-
inflammatory agents to improve symptoms in patients with schizophrenia: an update. Schizophr
Bull. 2014;40:181–91.
Song C, Leonard BE. An acute phase protein response in the olfactory bulbectomised rat: effect of
sertraline treatment. Med Sci Res. 1994;22:313–4.
Sparkman NL, Johnson RW. Neuroinflammation associated with aging sensitizes the brain to the
effects of infection or stress. Neuroimmunomodulation. 2008;15:323–30.
Spellberg B, Edwards JE Jr. Type 1/Type 2 immunity in infectious diseases. Clin Infect Dis.
2001;32:76–102.
Sperner-Unterweger B, Gaggl S, Fleischhacker WW, Barnas C, Herold M, Geissler D. Effects of
clozapine on hematopoiesis and the cytokine system. Biol Psychiatry. 1993;34:536–43.
Sperner-Unterweger B, Miller C, Holzner B, Widner B, Fleischhacker WW, Fuchs D. Measurement
of neopterin, kynurenine and tryptophan in sera of schizophrenic patients. In: Müller N, editor.
Psychiatry, psychoimmunology, and viruses. Vienna/New York: Springer; 1999. p. 115–9.
Neurobiological Principles: Psycho-Neuro-Immuno-Endocrinology 63
Stefansson H, Ophoff RA, Steinberg S, Andreassen OA, Cichon S, Rujescu D, et al. Common
variants conferring risk of schizophrenia. Nature. 2009;460:744–7.
Stevens JR. Schizophrenia and multiple sclerosis. Schizophr Bull. 1988;14:231–41.
Swerdlow NR, van Bergeijk DP, Bergsma F, Weber E, Talledo J. The effects of memantine on
prepulse inhibition. Neuropsychopharmacology. 2009;34:1854–64.
Syvalahti E, Eskola J, Ruuskanen O, Laine T. Nonsuppression of cortisol in depression and immune
function. Prog Neuro-Psychopharmacol Biol Psychiatry. 1985;9:413–22.
Teunissen CE, van Boxtel MP, Bosma H, Bosmans E, Delanghe J, De Bruijn C, et al. Inflammation
markers in relation to cognition in a healthy aging population. J Neuroimmunol.
2003;134:142–50.
Traki L, Rostom S, Tahiri L, Bahiri R, Harzy T, Abouqal R, et al. Responsiveness of the EuroQol
EQ-5D and Hospital Anxiety and Depression Scale (HADS) in rheumatoid arthritis patients
receiving tocilizumab. Clin Rheumatol. 2014;33:1055–60.
Tyring S, Gottlieb A, Papp K, Gordon K, Leonardi C, Wang A, et al. Etanercept and clinical
outcomes, fatigue, and depression in psoriasis: double-blind placebo-controlled randomised
phase III trial. Lancet. 2006;367:29–35.
Varga G, Nippe N, Balkow S, Peters T, Wild MK, Seeliger S, et al. LFA-1 contributes to signal I of
T-cell activation and to the production of T(h)1 cytokines. J Invest Dermatol.
2010;130:1005–12.
Vecsei L, Szalardy L, Fulop F, Toldi J. Kynurenines in the CNS: recent advances and new questions.
Nat Rev Drug Discov. 2013;12:64–82.
Wagner-Jauregg J. Fieberbehandlung bei Psychosen. Wien Med Wochenschr. 1926;76:79–82.
Wang AK, Miller BJ. Meta-analysis of cerebrospinal fluid cytokine and tryptophan catabolite
alterations in psychiatric patients: comparisons between schizophrenia, bipolar disorder, and
depression. Schizophr Bull. 2018;44:75–83.
Weitz-Schmidt G. Statins as anti-inflammatory agents. Trends Pharmacol Sci. 2002;23:482–6.
Wilke I, Arolt V, Rothermundt M, Weitzsch C, Hornberg M, Kirchner H. Investigations of cytokine
production in whole blood cultures of paranoid and residual schizophrenic patients. Eur Arch
Psychiatry Clin Neurosci. 1996;246:279–84.
Yang C, Hashimoto K. Peripheral IL-6 signaling: a promising therapeutic target for depression?
Expert Opin Investig Drugs. 2015;24:989–90.
Yang C, Shirayama Y, Zhang JC, Ren Q, Hashimoto K. Peripheral interleukin-6 promotes resilience
versus susceptibility to inescapable electric stress. Acta Neuropsychiatr. 2015a;27:312–6.
Yang JJ, Wang N, Yang C, Shi JY, Yu HY, Hashimoto K. Serum interleukin-6 is a predictive
biomarker for ketamine’s antidepressant effect in treatment-resistant patients with major depres-
sion. Biol Psychiatry. 2015b;77:e19–20.
Zalcman S, Green-Johnson JM, Murray L, Nance DM, Dyck D, Anisman H, et al. Cytokine-specific
central monoamine alterations induced by interleukin-1, -2 and -6. Brain Res. 1994;643:40–9.
Zarrabi MH, Zucker S, Miller F, Derman RM, Romano GS, Hartnett JA, et al. Immunologic and
coagulation disorders in chlorpromazine-treated patients. Ann Intern Med. 1979;91:194–9.
Zhou D, Kusnecov AW, Shurin MR, DePaoli M, Rabin BS. Exposure to physical and psychological
stressors elevates plasma interleukin 6: relationship to the activation of hypothalamic-pituitary-
adrenal axis. Endocrinology. 1993;133:2523–30.
Zhou AJ, Lee Y, Salvadore G, Hsu B, Fonseka TM, Kennedy SH, et al. Sirukumab: a potential
treatment for mood disorders? Adv Ther. 2017;34:78–90.
Zhu J, Bengtsson BO, Mix E, Thorell LH, Olsson T, Link H. Effect of monoamine reuptake
inhibiting antidepressants on major histocompatibility complex expression on macrophages in
normal rats and rats with experimental allergic neuritis (EAN). Immunopharmacology.
1994;27:225–44.
Pharmacokinetic and Pharmacodynamic
Principles
Contents
Pharmacokinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
Pharmacodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
“Dose-Response Curve”: Concentration-Effect Relationships Explained . . . . . . . . . . . . . . . . . . . . . . . 70
Difficulties Encountered in Clinical Trials Aimed at Demonstrating Concentration-Effect
Curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
Why Single Center Trials Still Can Go Wrong . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
Variability of the Medication’s Concentration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
Other Examples for Clear-Cut Concentration-Effect Relationships Obtained
with Antipsychotics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
Clear-Cut Concentration-Effect Relationships Obtained with Antidepressants . . . . . . . . . . . . . . . . . 77
The Problem of the Placebo Responders . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
A Checklist to Improve Trials Investigating Concentration Effect Relationships . . . . . . . . . . . . . . 80
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
Abstract
This chapter summarizes the pharmacokinetic principles that are essential in everyday
clinical practice: Whenever initiating a dosage, changing a dosage, or upon cessation
of a treatment, the new steady state of the medication (including its de facto
elimination) is not reached before 4 elimination half-lives. Many neuropsychiatric
medications have elimination half-lives around 24 h, so the clinician has to wait at
least 4 days until the new effect intensity can be expected. On pharmacologic
principle, medications with longer elimination half-lives should be preferred because
their concentrations fluctuate less, producing less variation in effect strength and
G. Zernig (*)
Department of Psychiatry 1, Medical University of Innsbruck, Innsbruck, Tirol, Austria
e-mail: gerald.zernig@i-med.ac.at
C. Hiemke
Department of Psychiatry and Psychotherapy, University Medical Center Mainz, Mainz, Germany
e-mail: hiemke@uni-mainz.de
carrying a lower risk of adverse drug reactions (ADRs). With respect to pharmaco-
dynamics, clinical studies in psychiatric institutions have been notoriously poor in
demonstrating a relationship between the concentration of the prescribed medication
and its effect on psychiatric symptoms. This chapter identifies the very likely reasons
for this shortcoming of clinical trials: Concentration-effect relationships becomes less
and less clear-cut as one proceeds from simple outcomes that are very much upstream
in the biologic chain of events, for example, receptor occupancy or prolactin blood
concentration, to a multifactored outcome such as symptom improvement. Investi-
gators have to contend with “signal noise” contributed by placebo responders,
placebo deteriorators, and nonresponders, especially as one moves from single center
to multicenter trials. This chapter ends with a study design checklist to improve
clinical trials investigating concentration-effect relationships of neuropsychiatric
medications. With respect to the pharmacologic targets of neuropsychotropic medi-
cations (receptors, monoamine transporters, etc.), the reader is referred to the chapter
▶ “Adverse Drug Reactions, Intoxications and Interactions of Neuropsychotropic
Medications”
Abbreviations
▮ For reasons of easier accessibility through the search function of
softwares, hyphens or subscripts are given in the text. Therefore,
the 5HT1A receptor is given as “5HT1A receptor, or the com-
pound “MK-801” simply as “MK801.”
5HT 5-hydroxytryptamine ¼ serotonin
5HT2A Serotonin receptor, subtype 2A ¼ 5HT1A receptor ¼ 5HT2A
receptor
5HTT Serotonin transporter
ADR Adverse drug reaction
BPRS Brief psychiatric rating scale
CGI-I Clinical global impressions, improvement scale
D1, D2, etc. Dopamine receptor, subtype 1 ¼ D1 receptor ¼ D1 receptor
DAT Dopamine transporter
HAMD Hamilton depression rating scale
HRSD Hamilton rating scale for depression
NARI Noradrenaline reuptake inhibitor (reboxetine)
NaSSA Noradrenergic and specific serotonergic antidepressant
(mirtazapine, mianserin)
NAT Noradrenaline transporter, same as NET
PANSS Positive and negative syndrome scale
SERT Serotonin transporter
SERTI Serotonin transporter inhibitor, see also SSRI
SGA Second generation antipsychotic
SPC Summary of product characteristics
SNRI Serotonin norepinephrine reuptake inhibitor (e.g., duloxetine,
venlafaxine)
SSRI Selective serotonin reuptake inhibitor (e.g., escitalopram)
Pharmacokinetic and Pharmacodynamic Principles 67
Pharmacokinetics
Some pharmacologists like the following quip to sum up the difference between
pharmacokinetics and pharmacodynamics: “Pharmacokinetics is what the body does
to the drug and pharmacodynamics what the drug does to the body.” In the follow-
ing, we try to give a very concise overview of pharmacokinetics as relevant in
everyday clinical practice. Further readings are given at the end of this section.
When a compound is introduced into the body, be that through the mouth (“per
os,” po), by inhalation, injection, infusion, rubbing onto the skin, transdermal
patches, rectal or vaginal suppositories or enemas, it is absorbed at a speed and
completeness according to the route of administration, with the body starting to
metabolize and/or excrete the compound right away. Absorption is good for most of
the neuropsychotropic medications and independent of concomitant food intake,
with the notable exception of ziprasidone (Hiemke et al. 2018). For the oral route of
administration, a useful measure is the (oral) bioavailability, that is, the fraction
absorbed (F) of the active unmetabolized medication that can be found (usually over
the first 24 h) after oral intake as compared to an intravenous injection. Most
neuropsychotropic medications are well absorbed, so their bioavailability, F,
approaches 1 (i.e., 100%), for example, escitalopram or olanzapine with an F of
0.8 (i.e., 80% Hiemke et al. 2018). Some medications, however, are extensively
metabolized even during their first passage through the liver (“first pass effect”) so
that only a small fraction can exert its effect, for example, venlafaxine (F ¼ 0.4 or
40%) or quetiapine (F ¼ 0.09 or 9%) (Hiemke et al. 2018). Metabolism, broadly
speaking, occurs in two steps: First, the medication molecule is enzymatically
changed in its structure (phase 1 reactions: oxidation, reduction, hydrolysis), pre-
dominantly in reactions involving the cytochrome P450 (CYP) system. Often, the
medication loses its efficacy with the enzymatic change. Quite a number of psycho-
pharmacologic medications, however, retain their efficacy after enzymatic modifi-
cation, that is, become active metabolites, for example, 9-hydroxyrisperidone (also
called paliperidone (Hiemke et al. 2018)). More information on CYP isoforms
(pharmacogenetic variation) and drug-drug interactions involving the CYP system
can be found in our chapter on adverse drug reactions (ADRs) and in Hiemke et al.
(2018). Phase 1 reactions introduce a polar functional group that enables a phase
2 conjugation reaction with highly polar molecules such as glucuronic acid (forming
a glucuronide) or sulfuric acid (forming a sulfate). Both phase 1 and phase 2 reactions
take place in the liver, which has a much greater capacity for phase 2 than phase
1 reactions. These highly polar molecules are readily excreted in the liver or kidney
(Hiemke et al. 2018). Medications can also be excreted unchanged in the kidney, or
excreted through the bile and/or feces, or exhaled.
Medications can also be redistributed between different compartments of the
body (blood, muscle, fat, brain). Generally speaking, the more lipophilic a medica-
tion is (and most neuropsychotropic compounds are highly lipophilic), the more
readily it enters and “hides” in fatty compartments (including the brain), and the
larger its volume of distribution, Vd, becomes. If a medication distributes equally in
the water and tissues of the body, its Vd is 0.7 L/kg (as our bodies consist of 70%
water).
68 G. Zernig and C. Hiemke
In everyday clinical practice, it is important to know at what time after oral intake
(ingestion) a medication reaches it maximal concentration. The respective measure is
called tmax. Extended release formulations (in marketing lingo: “ER,” “XR,” “LA”
(“long acting”) or, an especially unlucky term, “retard”) have a longer tmax than the
respective “immediate release” (“IR”) formulation (sometimes not labeled as such).
For example, tmax is 6 h for Seroquel ® XR “Retardtabletten” but only 1–1.8 h in the
“normal,” that is, IR formulation of quetiapine (unfortunately, this is not explicitly
stated in the respective SPC for Seroquel ®, only in the very useful pharmacokinetics
tables in (Brunton et al. 2017) or (Hiemke et al. 2018)). Interestingly, the elimination
half-life for both Seroquel ® formulations is 7 h (respective SPC). Thus, the only
difference between the two formulations (ER ¼ XR vs IR) is that the IR formulation
peaks faster (i.e., at 1.8 h at the latest vs at 6 h; useful if quetiapine is used as a
hypnotic) than the ER formulation.
Almost all neuropsychotropic compounds are eliminated exponentially, mean-
ing that if a medication is stopped, has reached its peak concentration, and is then
eliminated from the body, at the end of the first elimination half-life (i.e., a time
span), t1/2, 50% of the medication remains. At the end of a time span of 2 elimi-
nation half-lives, 50% of 50%, that is, 25% of the original concentration remains.
After 3 t1/2, 12.5% remain, after 4 t1/2, only 6.25% remain, that is, the medication is
de facto eliminated. As Fig. 1 (corresponding to Fig. 2 of Hiemke et al. 2018)
Pharmacodynamics
SERTs), then the effect is saturable (solid curve in Fig. 2): Regardless of how high
the concentration of the medication (ligand) is pushed, the medication (ligand)
cannot produce an effect that exceeds a maximum that is defined by the system
under investigation. In the left panel (a) of Fig. 2, this is shown as a continuously
decreasing slope and, finally, a flattening of the curve describing this relationship
(called a “curvilinear” one). A pharmacologist would describe that the function
reaches its asymptote (ceiling). To emphasize, even in a plot in which the values
both of the ordinate (y axis) an abscissa (x axis) are expressed in linear (and not
logarithmic) form (panel a on the left side of Fig. 2) – a so-called “linear” plot – this
relationship is represented by a curve (solid) and not a line (small dashes). Such a
linear plot is most commonly used to describe concentration-effect relationships in
psychiatry. The saturability of the target system becomes more clearly visible if the
dose/concentration data on the abscissa (x axis) are transformed to their decadic
logarithm (panel b on the right side of Fig. 2). This type of plot is called a
semilogarithmic or “semilog” plot.
The relationship linking drug concentration to effect can be fitted to a logistic
function of the following form (see, e.g., Eq. 11.27 on p. 269 of (Kenakin 2009) or,
e.g., (Zernig et al. 1996) for a discussion):
Emax concentrationslope
E¼ ð1Þ
EC50slope þ concentrationslope
where E is an effect and Emax the maximal effect of the drug, EC50 is the agonist
concentration producing a half-maximal effect, and “slope” is the slope factor
(sometimes called “Hill coefficient” (Walker et al. 1995; Kenakin 2009)). To account
for a baseline response (e.g., placebo responders), this equation should be extended
by a factor c (for baseline “constant”; see, e.g., Zernig et al. 1995, 1996, 1997;
Walker et al. 1998):
Emax concentrationslope
E¼ þc ð2Þ
EC50slope þ concentrationslope
simple outcomes that are very much upstream in the biologic chain of events, for
example, receptor occupancy or prolactin blood concentration, to a multifactored
outcome such as symptom improvement and as investigators have to contend with
“signal noise” contributed by placebo responders, placebo deteriorators, and non-
responders. We will argue that clinical improvement as rated by different observers
may be jeopardized by such a high degree of variability that it has very often been
impossible, especially in multicenter trials, to demonstrate the meaningful
concentration-effect relationships that the natural laws as described by pharma-
cology predict.
In other words, concentration-effect relationships become less and less clear-cut
as one proceeds from outcomes that, with respect to cause-effect relationships, are
most upstream, such as binding and occupancy of dopamine D2 receptors, to
biologic events that immediately follow D2 receptor occupancy, that is, changes in
the prolactin blood concentration, to psychiatric symptom improvement, which is
very much downstream in the chain of events and which is influenced by other
factors than the pharmacologic effect of the investigated medication. The decreasing
relationship between concentration and effect is exemplified here for olanzapine.
In a positron emission tomography (PET) study (Kapur et al. 1998), the binding
of the antipsychotic olanzapine to serotonin 5HT2 and dopamine D2 receptors at
different olanzapine (blood) plasma levels (concentrations) was analyzed. Assuming
binding to a single population of D2 receptors, the data were fitted to a saturating
function, that is, a rectangular hyperbola or, in other words, to a logistic equation
with a slope of 1 (see above) with very limited “noise” (i.e., the small variability) of
the D2 receptor binding data, that is, with actual PET data points that were very close
to the fitted curve (see Fig. 1 of Kapur et al. 1998).
A similar clear-cut and statistically significant correlation can be found for
olanzapine concentrations in plasma and a laboratory parameter, that is, the prolactin
level (see Fig. 2 of Citrome et al. 2009). Prolactin is a hormone synthesized in
lactotrophs of the anterior pituitary gland. Synthesis and secretion of this hormone
are mainly under the inhibitory control of hypothalamic dopamine. Inhibition of
hypophyseal dopamine D2 receptors enhances the release of prolactin.
Going more downstream and correlating olanzapine concentrations with the
improvement of psychiatric symptoms a poor relationship is obtained at best. In a
single center clinical trial on 54 inpatients the improvement after 2 weeks of the
same dosage (Mauri et al. 2005) weak, nevertheless still significant relationship was
observed by Mauri and coworkers (see Fig. 3 of Mauri et al. 2005). An ascending
concentration-effect curve and a few data points at very high olanzapine concentra-
tions suggest a descending part of a biphasic relationship as well (please see above
for an explanation of biphasic concentration-response relationships). As detailed
above, these authors have chosen to fit the data to a mathematical function that that
does not have a pharmacologic justification, that is, does not obey the underlying
ligand-receptor interactions and signal transduction amplification processes. How-
ever, when compared to the PET binding data above, the data points are scattered to a
visibly higher degree, also depending on the psychometric test used. The olanzapine
concentration-effect relationship obtained in a multicenter trial conducted in
Pharmacokinetic and Pharmacodynamic Principles 75
The discussion of the above findings should not mislead the reader in believing that
every single center trial will yield a clear-cut concentration-effect relationship. The
following example may serve as a cautionary case, especially as one of us
(gz) contributed to that study as a coauthor (and could neither convince the other
authors to retract the paper nor retract his coauthorship once the study was published):
In a “naturalistic” treatment setting, that is, among in- and outpatients treated at a
university hospital psychiatric department, plasma concentrations of several
new-generation antipsychotics were correlated to clinical improvement, expressed
as percent change in PANSS scores (Kaufmann et al. 2016). As several different
antipsychotics were administered, their plasma concentrations were normalized to
the lower limit of their respective orienting therapeutic range (OTR, “therapeutic
reference range”) (Hiemke et al. 2018). This study demonstrated a significant effect
of treatment but did not reveal any significant relationship between concentration
and effect (see Fig. 2b of Kaufmann et al. 2016). If anything, the most pronounced
improvement occurred at concentrations that were below the lower limit of the OTR.
After publication, it turned out that steady state may not have been reached in some
cases. This meant that – also depending on the time span between the plasma level
determination and the PANSS symptom scoring relative to the antipsychotic’s
elimination half-life – overreporting or underreporting of the antipsychotic’s efficacy
may have occurred. More importantly, 14 of the 67 patients were treated with two
antipsychotics and 3 patients were treated with three antipsychotics consecutively, a
fact that had not been declared explicitly in the original article (Kaufmann et al.
2016). This mistake was corrected later (Kaufmann et al. 2020). Of importance for
the analysis of the concentration-effect relationship, 14 of the 67 patients did not
respond to the first antipsychotic or suffered from intolerable adverse drug reactions
(ADRs) and had thus to be switched to a second antipsychotic. Three of the
67 patients even had to be switched twice. Thus, underresponders or nonresponders
with respect to the desired antipsychotic effects and/or overresponders with respect
to ADRs are overrepresented – while not being declared as such – in the published
data set. With respect to interrater- and test-retest reliability, that is, the variability of
76 G. Zernig and C. Hiemke
the score of clinical response, the original article only stated that “... At each visit, a
psychiatrist belonging to a trained schizophrenia research team rated psychopathol-
ogy ...” (p. 718 of Kaufmann et al. 2016)).
Another example that a laboratory parameter, that is, prolactin concentration, is less
variable that the rating of the clinical improvement of the patient comes from a single
center trial by (Kapur et al. 2000): The prolactin concentrations of the very same
patients followed a much more orderly relationship to dopamine D2 receptor
occupancy by haloperidol than their CGI-I ratings (see Fig. 1 of (Kapur et al. 2000).
The clear-cut receptor occupancy data obtained with the PET ligand [11C]
raclopride for the D2 receptor antagonist (or “blocker”) olanzapine (see Fig. 1 of
Kapur et al. 1998) can be generalized to the partial D2 receptor partial agonist
aripiprazole as obtained in a [18F]fallypride PET study as well (see Fig. 1 of Grunder
et al. 2008).
To provide the reader with a pharmacodynamic orientation with respect to the all
these different dopamine receptor ligands: According to the International Union of
Basic and Clinical Pharmacology (IUPHAR) Subcommittee on Dopamine Recep-
tors (www.guidetopharmacology.org), dopamine receptors can be divided in two
families: a D1-like family (D1 and D5) and a D2-like family (D2, D3 and D4). The
affinities of the various dopamine receptor ligands (agonists or antagonists/blockers)
can be expressed as the negative decadic logarithm (“p”) of the concentration
(in mol/L) at which 50% of the receptors are occupied (usually quantified in
Pharmacokinetic and Pharmacodynamic Principles 77
function (i.e., y ¼ a + b*x - c*x2) that has no pharmacologic basis, that is, does not
obey the underlying ligand-receptor interactions and signal transduction amplifica-
tion processes, that is, are logistic functions (see detailed discussion above).
A similarly clear-cut concentration-effect relationship could be demonstrated by
Conca and coworkers for escitalopram (Fig. 4):
muddy the concentration effect relationship and illustrates by which type of analysis
this can be overcome, that is, by constructing bins of drug concentrations and
expressing the response as percent responders for each given bin.
We could demonstrate that this type of analysis is able to unmask hidden
concentration-effect relationships (Eggart et al. 2011) by reanalyzing data that had
been interpreted as demonstrating a “lack of correlation [of paroxetine plasma levels]
with efficacy or adverse events” (i.e., the title of (Tasker et al. 1989)). In the original
multicenter analysis, steady state paroxetine plasma concentrations were obtained
with “the timing of the blood sampled varied according to the study design,” p. 152
of (Tasker et al. 1989)) and correlated with the antidepressant response. Unfortu-
nately, no quantitative definition of the Clinical Global Impression (CGI)-based
response was given in that analysis. When paroxetine concentrations were
arranged in increasing bins of 10 ng/mL and the respective number of responders
was given for each bin, no clear-cut concentration-effect relationship was obtained
80 G. Zernig and C. Hiemke
Fig. 6 Relationship
between paroxetine plasma
concentration, serotonin
transporter occupancy, and
antidepressant effect. SERT,
serotonin transporter (5HTT).
See text for details.
Reproduced with the
publisher’s permission from
Fig. 6 of Eggart et al. (2011)
(Tasker et al. 1989)). In our analysis, we kept the original 10 ng/mL bin order but
expressed the effect as percent of responders by 10 ng/mL (Eggart et al. 2011)).
After this modest data transformation, a clear-cut and meaningful relationship
emerged (Fig. 17). We also compared the clinical response data to serotonin trans-
porter occupancy data obtained in a PET study (Meyer et al. 2004) and found that the
clinical response data corresponded well to the binding data, that is, the respective
concentration-effect relationships overlapped to a considerable degree (Fig. 6
corresponding to Fig. 6 of Eggart et al. 2011).
At the level of the study design, a placebo lead-in phase of 4–7 days had been
employed in the olanzapine trials discussed above (Beasley et al. 1996a, b; Perry
et al. 1997; Callaghan et al. 1999). We opine that this is a very commendable
approach. However, inspection of Fig. 8 above (corresponding to Fig. 10 of Calla-
ghan et al. 1999) reveals that at olanzapine concentrations below approximately
3 ng/mL (i.e., levels that are close to placebo), the BPRS change still ranges from
approximately 70% (quasi-placebo responders) to +80% (quasi-placebo
deteriorators). Still, the placebo lead-in may have improved the data (i.e., lowered
the noise of the response); we are not aware of any study that compared data
obtained without versus with the placebo lead-in.
A final note on data analysis using the receiver operating characteristic (ROC)
relationship: This type of analysis has successfully been used to derive a threshold
concentration for a significant clinical response (see, e.g., Perry et al. 1997; Spina
et al. 2000; Perry et al. 2001; Muller et al. 2007; Fellows et al. 2003) and may be
successful in cases in which the fitting of the concentration effect data to the logistic
function (see above for a detailed description) has failed. A good description of the
ROC method can be found in Fellows et al. (2003).
Based on the published studies discussed above, the following is a (most likely
incomplete) checklist for the design of clinical trials investigating concentration-
effect curves and for the analysis of such relationships:
Pharmacokinetic and Pharmacodynamic Principles 81
Cross-References
References
Bauer LA. Applied clinical pharmacokinetics. New York: McGraw Hill; 2008.
Beasley CM Jr, Sanger T, Satterlee W, Tollefson G, Tran P, Hamilton S. Olanzapine versus placebo:
results of a double-blind, fixed-dose olanzapine trial. Psychopharmacology. 1996a;124:159–67.
Beasley CM Jr, Tollefson G, Tran P, Satterlee W, Sanger T, Hamilton S. Olanzapine versus placebo
and haloperidol: acute phase results of the north American double-blind olanzapine trial.
Neuropsychopharmacology. 1996b;14:111–23.
Brunton LL, Hilal-Dandan R, Knollmann BC. Goodman and Gilman’s the pharmacological basis of
therapeutics. 13th ed. New York: McGraw-Hill; 2017.
Callaghan JT, Bergstrom RF, Ptak LR, Beasley CM. Olanzapine. Pharmacokinetic and pharmaco-
dynamic profile. Clin Pharmacokinet. 1999;37:177–93.
Citrome L, Stauffer VL, Chen L, Kinon BJ, Kurtz DL, Jacobson JG, Bergstrom RF. Olanzapine
plasma concentrations after treatment with 10, 20, and 40 mg/d in patients with schizophrenia:
an analysis of correlations with efficacy, weight gain, and prolactin concentration. J Clin
Psychopharmacol. 2009;29:278–83.
De Donatis D, Florio V, Porcelli S, Saria A, Mercolini L, Serretti A, Conca A. Duloxetine plasma level
and antidepressant response. Prog Neuro-Psychopharmacol Biol Psychiatry. 2019;92:127–32.
Eggart V, Hiemke C, Zernig G. “There is no dose-response relationship in psychopharmacotherapy”
vs “pharmacotherapy in psychiatry is based on ligand-receptor interaction”: a unifying hypoth-
esis and the need for plasma concentration based clinical trials. Psychopharmacology.
2011;217:297–300.
Fellows L, Ahmad F, Castle DJ, Dusci LJ, Bulsara MK, Ilett KF. Investigation of target plasma
concentration-effect relationships for olanzapine in schizophrenia. Ther Drug Monit.
2003;25:682–9.
Florio V, Porcelli S, Saria A, Serretti A, Conca A. Escitalopram plasma levels and antidepressant
response. Eur Neuropsychopharmacol. 2017;27:940–4.
82 G. Zernig and C. Hiemke
Spina E, Avenoso A, Facciola G, Scordo MG, Ancione M, Madia AG, Ventimiglia A, Perucca
E. Relationship between plasma concentrations of clozapine and norclozapine and therapeutic
response in patients with schizophrenia resistant to conventional neuroleptics. Psychopharma-
cology. 2000;148:83–9.
Stahl SM. Stahl’s essential psychopharmacology: neuroscientific basis and practical application.
Cambridge: Cambridge University Press; 2013.
Tanum L, Strand LP, Refsum H. Serum concentrations of citalopram – dose-dependent variation in
R- and S-enantiomer ratios. Pharmacopsychiatry. 2010;43:190–3.
Tasker TC, Kaye CM, Zussman BD, Link CG. Paroxetine plasma levels: lack of correlation with
efficacy or adverse events. Acta Psychiatr Scand Suppl. 1989;350:152–5.
Wagner JG. Pharmacokinetics for the pharmaceutical scientist. Missionsstrasse 44, CH-4055.
Basel: Technomic Publishing AG; 1993.
Walker EA, Zernig G, Woods JH. Buprenorphine antagonism of mu opioids in the rhesus monkey
tail-withdrawal procedure. J Pharmacol Exp Ther. 1995;273:1345–52.
Walker EA, Zernig G, Young AM. In vivo apparent affinity and efficacy estimates for mu opiates in
a rat tail-withdrawal assay. Psychopharmacology. 1998;136:15–23.
Welling PG. Pharmacokinetics. Processes and mathematics. Washington, DC: American Chemical
Society; 1986.
Zernig G, Issaevitch T, Broadbear J, Burke T, Lewis JW, Brine GA, Woods JH. Receptor reserve
and affinity of mu opioid agonists in mouse antinociception: correlation with receptor binding.
Life Sci. 1995;57:2113–25.
Zernig G, Issaevitch T, Woods JH. Calculation of agonist efficacy, apparent affinity and receptor
population changes after administration of insurmountable antagonists: comparison of different
analytical approaches. J Pharmacol Toxicol Methods. 1996;35:223–37.
Zernig G, Lewis JW, Woods JH. Clocinnamox inhibits the intravenous self-administration of opioid
agonists in rhesus monkeys: comparison with effects on opioid agonist-mediated anti-
nociception. Psychopharmacology. 1997;129(3):233–42.
Zernig G, Ahmed SH, Cardinal RN, Morgan D, Acquas E, Foltin RW, Vezina P, Negus SS, Crespo
JA, Stoeckl P, Grubinger P, Madlung E, Haring C, Kurz M, Saria A. Explaining the escalation of
drug use in substance dependence: models and appropriate animal laboratory tests. Pharmacol-
ogy. 2007;80:65–119.
Pharmacovigilance
Maike Scherf-Clavel
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
Core Functions of Pharmacovigilance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
How Pharmacovigilance Works . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
Pharmacovigilance in Different Countries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
Adverse Drug Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
Pharmacovigilance in Psychiatry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
Prevention and Management of Adverse Drug Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
Prevention of ADE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
Pharmacogenetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
Therapeutic Drug Monitoring . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
Managing Adverse Drug Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
Abstract
Fatal outcomes related to the use of drugs in the past, for example, most
prominently, the thalidomide tragedy, has caused the development of systematic,
structured and regulated pharmacovigilance system in the 1960s. Pharmacov-
igilance includes processes for monitoring and evaluating adverse drug effects
with the aim to improve the safety of a drug therapy. Core functions of
pharmacovigilance are case management, signal management, and benefit-risk
management. They cover the early stages in a drug’s life cycle, but also, as it is
impossible to identify all safety concerns during clinical trials, post-marketing
observational analyses. In later stages of a drug’s life cycle safety monitoring is
based on spontaneous reporting of ADEs. In the EU and the USA, good
pharmacovigilance systems are established; however, in some middle- and low-
income countries, pharmacovigilance is still at its beginning or not available.
M. Scherf-Clavel (*)
Department of Psychiatry, Psychosomatics and Psychotherapy, University Hospital of Würzburg,
Würzburg, Germany
e-mail: Scherf_M@ukw.de
Introduction
Fatal outcomes related to the use of drugs in the past have caused the development of
pharmacovigilance systems (Beninger 2018). The World Health Organization
(WHO) defines pharmacovigilance as the “science and activities relating to the
detection, assessment, understanding and prevention of adverse effects or any
other possible drug-related problem” (WHO 2004). In short, pharmacovigilance
describes processes for monitoring and evaluating adverse drug effects (WHO
2004). An adverse drug effect can be caused by every substance that produces a
therapeutic effect; however, different drugs showed different risks of producing
adverse effects (Edwards and Aronson 2000).
The thalidomide tragedy, described as the “largest men-made medical disaster in
history,” causing congenital malformation in newborns, triggered the establishment
of a systematic, structured, and regulated spontaneous reporting system (Fornasier
et al. 2018; Vargesson 2015). Probably the first pharmacovigilance program was a
careful systematic review of case reports considering adverse effects of chloram-
phenicol in 1962 (Beninger 2018). In 2001 EudraVigilance, the official European
database to manage and analyze information regarding adverse drug effects, was
founded (Fornasier et al. 2018). In 2012, a new legislation changed the pharmacov-
igilance program in Europe; in consequence, proactive strategies of risk manage-
ment replaced reactive strategies (Coleman and Pontefract 2016; Fornasier et al.
2018). Nowadays, an ADE defines any adverse event following the use of a drug,
and patients were involved in pharmacovigilance activities (Fornasier et al. 2018).
Fig. 1 Pharmacovigilance and safety concerns during clinical development of a drug. (Modified
according to WHO 2004)
88 M. Scherf-Clavel
Grootheest 2008). One example is the sequence symmetry analysis, which compares
initiation of a drug (A) after another drug (B) with initiation of B after A (Lai et al.
2017). If there is a difference, it may indicate on an ADE of one of the two drugs (Lai
et al. 2017).
Intensive monitoring uses prescription data to identify users of a drug and asking
the prescriber about adverse events (Harmark and van Grootheest 2008). Only
selected drugs during a specified time period were monitored; however, the method
can identify events that were not suspected as ADE of the studied drug and quantify
the risk of an ADE (Harmark and van Grootheest 2008).
from the WHO, and additional WHO collaboration centers were established in
Ghana and Morocco (Olsson et al. 2015).
In Europe, countries have moved toward greater harmonization of pharmacov-
igilance (Pitts et al. 2016). Since the renewal of the pharmacovigilance program in
2012, in the EU over 26,000 potential signals were reviewed resulting in 453 signals
assessed by the Pharmacovigilance Risk Assessment Committee (PRAC), and about
200 signals resulted in updates of the product information (Potts et al. 2019). Signal
detection, following the reporting of ADEs, is a collaboration between different
officials (National Competent Authorities (NCAs), European Medicines Agency
(EMA), and marketing authorization holders (MAHs)) (Potts et al. 2019). A con-
firmed signal is then reported to the PRAC for prioritization and assessment (Potts
et al. 2019). In the USA, the pharmacovigilance system is stricter than in the EU,
showing that pharmacovigilance in the twenty-first century is “the systematic mon-
itoring of the process of pre-market review and post-market surveillance, which are
linked through study design, product labelling, therapeutic outcomes, adverse
events, hospital and clinician reporting systems, the pharmacy interface, compliance,
and a complete understanding of real-world evidence” (Pitts et al. 2016).
However, as every country is individual, there is no general way to establish
pharmacovigilance in one country (Olsson et al. 2015). Moreover, no country has a
perfect pharmacovigilance system, but every country needs clear and published
national policies for pharmacovigilance (Martin et al. 2018; Olsson et al. 2015).
Pharmacovigilance in Psychiatry
Prevention of ADE
This check should include information about comorbid conditions, drug misuse,
self-harm, self-neglect, poor nutrition, alcohol/smoking, accidents, chaotic lifestyle,
social circumstances, isolation, lack of caregiver, inappropriate accommodation, and
delusional conduct (Lader 1999). Additionally, the patient should accept the risks
and benefits of the treatment option (Dodd et al. 2018). Considering specific drug
classes, pretreatment evaluation, individualization of the therapy, and regular patient
monitoring can prevent predictable adverse effects (Lader 1999). Also cotreatment
with other drugs, but also monitoring electrolytes or renal function, can also reduce
the risk for ADEs (Coleman and Pontefract 2016).
Common adverse effects and handling them during neuropsychopharma-
cotherapy are shown in Table 1 (Ables and Nagubilli 2010; Dodd et al. 2018;
Lader 1999; Toledano and Gil-Nagel 2008).
Different antipsychotic drugs show different side effect profiles (Huhn et al.
2019). In general, older antipsychotics are associated with EPS and prolactin
elevation, whereas newer antipsychotics are associated with weight gain and seda-
tion (Huhn et al. 2019).
Most of the ADEs of antiepileptic drugs are predictable and occur early after
onset of therapy, but others occur after months or years of treatment (Gaitatzis and
Sander 2013). Due to the comparable efficacy of the antiepileptic drugs, the differ-
ences in ADEs are an important factor while initiating individual therapies (Toledano
and Gil-Nagel 2008). ADEs were reported to affect the quality of life at least as much
as repetitive seizures (Toledano and Gil-Nagel 2008). However, in everyday clinical
practice, screening for ADEs is not systematically included; therefore, the preva-
lence of ADEs in antiepileptic drug treatment potentially remains underestimated
(Toledano and Gil-Nagel 2008). In general, before treatment initiation, potential
individual risks, described by age, sex, childbearing potential, comorbidities,
comedications, and individual wishes, should be taken into account (Gaitatzis and
Sander 2013). Avoiding overtreatment reduces the risk of ADEs, as most of the
reported adverse effects of antiepileptic drugs are dose-dependent (Gaitatzis and
Sander 2013).
Anticholinergic drugs include many different classes of drugs, for example,
antipsychotics, antidepressants, but also drugs for the treatment of the Parkinson’s
disease and anxiolytics (Collamati et al. 2016). Peripheral side effects of anticholin-
ergic drugs are dry mouth, nausea, vomiting, constipation, bloated feeling, abdom-
inal pain, loss of taste, and anorexia, and also ophthalmic side effects occur (blurred
vison, diplopia, mydriasis, increased ocular tension); central side effects are dizzi-
ness, lightheadedness, tingling, headache, drowsiness, weakness, nervousness,
numbness, mental confusion, excitement, dyskinesia, lethargy, syncope, speech
disturbance, and insomnia (Collamati et al. 2016). To predict the risk of known
ADEs in the treatment with anticholinergic drugs, in vitro methods and anticholin-
ergic scales have been developed (Cantudo Cuenca et al.; Villalba-Moreno et al.
2016), but none of these methods have been standardized, and there is no consensus
on how to define drug exposure (Collamati et al. 2016).
Psychoactive drugs in general affect the cardiovascular system (Brouillette and
Nattel 2017). To minimize the risk of negative cardiovascular consequences, patient-
Pharmacovigilance 93
Table 1 Common adverse drug effects, pretreatment before initiating therapy and treatment during
therapy of the most important neuropsychopharmacological drugs
Adverse effect Comments Pretreatment procedure Monitoring/treatment
Antipsychotics (Lader 1999)
Weight gain Increased risk Establish baseline Monitor weight
for weight Exercise regularly
cardiovascular Dietary advice Diet
disease,
diabetes,
osteoarthritis
Psychological
impact
Hyperprolactinemia Evaluate symptoms of Monitor symptoms
prolactinemia
Sedation Tolerance Advise patients on Take the drug as
development additive effects of nighttime dose
alcohol, risk of driving, Check if persisting for
and accidents >6 months
Anticholinergic effects Dry mouth, Exclude severe Dental hygiene
blurred vision, constipation, urologic Regular dental checks
constipation, difficulties, visual
urinary problems
retention
Hypotension Increased risk Supine and standing Monitor blood
for falls, bone blood pressure pressure
fractures, and measurement Start with a low dose
considerable and increase slowly
morbidity
Extrapyramidal Acute EPS, Assessment of Treat appropriately
symptoms (EPS) parkinsonism, preexisting EPS
akathisia,
tardive
dyskinesia
Seizures Dose-related Identify high-risk Avoid high doses and
seizure patients (family history rapid dose titration
potential of epilepsy, substantial Prophylactic use of
brain damage) anti-convulsants
Cardiac toxicity QT Check for cardiac Regular ECGs
prolongation history, irregular pulse,
undertake an ECG
Antidepressants (Ables and Nagubilli 2010; Dodd et al. 2018)
Headache, nausea, Diagnostic work up, e.g., organic causes of Monitor weight and
agitation, sedation, depression, personal and family history, waist circumference,
sexual dysfunction, physical health (weight, metabolic assess metabolic
diminished mental disorders, sexual health, alcohol, tobacco, parameters and sexual
acuity and memory, substance use/dependence, pregnancy, dysfunction, check for
weight gain, metabolic liver function), EGC, electrolytes suicidal ideation,
abnormalities, cardiac monitor renal and
toxicity, neurological hepatic function,
toxicity, hepatic electrolytes and blood
pressure, monitor
(continued)
94 M. Scherf-Clavel
Table 1 (continued)
Adverse effect Comments Pretreatment procedure Monitoring/treatment
toxicity, increased cardiac function,
suicidality therapeutic drug
monitoring
Serotonin syndrome Elucidate on symptoms of serotonin Appropriate treatment
syndrome, avoid combined use of and withdrawal of
serotonergic drugs serotonergic drugs
Antiepileptic drugs (Gaitatzis and Sander 2013; Toledano and Gil-Nagel 2008)
Somnolence Fluctuations in Tolerance development Avoid high doses,
Dizziness serum level ➔ rapid dose titration,
Ataxia exacerbation of and coadministration
Diplopia ADEs with other drugs
Blurring of vision “Start low, go slow”
Fatigue Therapeutic drug
Vertigo monitoring
Cognitive dysfunction Lower doses, if drugs
are used in
combination
Skin rush Resolved after “Start low, go slow”
discontinuation Immediate
discontinuation
Steven-Johnson- Infrequently “Start low, go slow”
Syndrome, toxic within the first Monitor: painful rash,
epidermal necrolysis 2 month of systemic
treatment inflammation, mucosal
involvement
Immediate
discontinuation
Acute hepatitis Within the first Screen for previous
3 months (drug hepatic diseases,
induced liver inborn errors of
injuries) metabolism, mental
No retardation
concentration-
dependence
Aplastic anemia
Agranulocytosis
Psychiatric adverse Common in all Screen for patient’s “Start low, go slow”
effects antiepileptic individual and family
drugs history on existing
psychiatric or
neurological disorders,
refractory epilepsy
Paradoxical Pharmacological
Aggravation of treatment should not be
Seizures based solely on the
type of seizure but
rather on the epilepsy
syndrome.
(continued)
Pharmacovigilance 95
Table 1 (continued)
Adverse effect Comments Pretreatment procedure Monitoring/treatment
Change in body weight Weight gain Establish baseline
and weight loss weight
possible
(depending on
the drugs)
Fractures, osteoporosis, No uniform consensus
rickets how to monitor and
prevent osteoporosis
periodic bone health
screening
Supplements:
bisphosphonates,
calcium, vitamin D
Sexual dysfunction,
altered sperm
morphology, motility,
concentration,
reproductive endocrine
dysfunction in women
Visual field defects
specific risk factors and the risk profile of drugs should be considered (Brouillette
and Nattel 2017). Following the low-dose treatment initiation, the applied dose
should carefully be increased according to the clinical signs (Brouillette and Nattel
2017). It is good to know that QT-prolonging drugs can be used safely, if appropriate
precautions are taken (Brouillette and Nattel 2017). Therefore, the risk of QT
prolongation should not dissuade patients from the necessary psychiatric therapy
(Brouillette and Nattel 2017).
Pharmacogenetics
Searching for genetic predictors for adverse drug reactions is another approach to
prevent ADEs (Brandl et al. 2014; Guerrini and Perucca 2018; Jaquenoud Sirot et al.
2006). Currently, for antiepileptic drugs genetic predictors were established
(Guerrini and Perucca 2018). For example, the human leukocyte antigen (HLA)-
B15:02 allele is strongly associated to carbamazepine-induced Steven-Johnson
syndrome in the Chinese Han population and other South Asian ethnic groups
(Guerrini and Perucca 2018). Also regarding antipsychotic drugs, different genetic
factors were associated with adverse effects (Brandl et al. 2014). It is recommended
to reduce doses in CYP2D6 poor metabolizers (Dean 2012), two polymorphisms in
dopamine receptor 2 were associated with a higher risk for tardive dyskinesia and the
Met allele in COMT (Val(108/158)Met) was protective against tardive dyskinesia
(Brandl et al. 2014).
96 M. Scherf-Clavel
Therapeutic drug monitoring is also a tool to prevent adverse drug effects, as the
upper limit of the therapeutic reference range is defined by the increased risk of
adverse drug reactions (Baumann 2008; Gentry and Rodvold 1995; Gerlach et al.
2016; Haen 2011; Hiemke et al. 2018; Jaquenoud Sirot et al. 2006; Schütze and
Schwarz 2016). In contrast, others argue that TDM is only of limited use for
therapy safety, as the same serum concentration in different patients can lead to
an ADE in one patient but not in the second (Dodd et al. 2018; Jaquenoud Sirot
et al. 2006).
In psychiatry the association between serum drug levels and clinical effects has
been investigated mainly for lithium and tricyclic drugs (Jaquenoud Sirot et al.
2006). In other areas, for example, immunosuppressant drugs, TDM is an
established tool to prevent ADEs (Jaquenoud Sirot et al. 2006). In epilepsy, TDM
has shown its clinical utility and in oncology, TDM is increasingly used for therapy
safety reasons (Jaquenoud Sirot et al. 2006; Verheijen et al. 2017; Westerdijk et al.
2019).
Due to the nature of the different adverse effects, various strategies for their
management were suggested. In case of serious ADEs, rapid action, for exam-
ple, emergency treatment and withdrawal of the medicine, is essential (Edwards
and Aronson 2000). Clinical benefit-risk assessment can be used to decide
which medicine should be withdrawn and which not, and a benefit-risk deci-
sion should be made whether or not the medicine is crucial to the patient
(Edwards and Aronson 2000). If it is not clear which drug was causative for
the reaction, the nonessential should be withdrawn first, and if it is a dose-
related ADR, dose reduction should be considered, rather than withdrawal of
the drug (Edwards and Aronson 2000). If a drug, causing ADR is absolutely
necessary for a patient, symptomatic treatment is obligatory (Edwards and
Aronson 2000).
Pharmacovigilance 97
References
Ables AZ, Nagubilli R. Prevention, recognition, and management of serotonin syndrome. Am Fam
Physician. 2010;81:1139–42.
Alshammari TM, Mendi N, Alenzi KA, Alsowaida Y. Pharmacovigilance systems in Arab coun-
tries: overview of 22 Arab countries. Drug Saf. 2019;42:849–68.
Awada Z, Zgheib NK. Pharmacogenovigilance: a pharmacogenomics pharmacovigilance program.
Pharmacogenomics. 2014;15:845–56.
Baumann P. Pharmacovigilance in psychiatry: pharmacogenetic tests and therapeutic drug moni-
toring are promising tools. Expert Rev Clin Pharmacol. 2008;1:183–5.
Beninger P. Pharmacovigilance: an overview. Clin Ther. 2018;40:1991–2004.
Bigi C, Bocci G. The key role of clinical and community health nurses in pharmacovigilance. Eur J
Clin Pharmacol. 2017;73:1379–87.
Biswas P. Pharmacovigilance in Asia. J Pharmacol Pharmacother. 2013;4:S7–s19.
Brandl EJ, Kennedy JL, Muller DJ. Pharmacogenetics of antipsychotics. Can J Psychiatr.
2014;59:76–88.
Brouillette J, Nattel S. A practical approach to avoiding cardiovascular adverse effects of psycho-
active medications. Can J Cardiol. 2017;33:1577–86.
Cantudo Cuenca M, Munoz Cejudo B, Mora Mora M, Fernandez Martinez G, Cantal Sanchez M.
Applying different scales for calculating the anticholinergic burden in older patients. European
Association of Hospital Pharmacists. https://www.eahp.eu/sites/default/files/4cps-180.pdf
Accessed 16 Dec 2019.
Coleman JJ, Pontefract SK. Adverse drug reactions. Clin Med (Lond). 2016;16:481–5.
Collamati A, Martone AM, Poscia A, Brandi V, Celi M, Marzetti E, Cherubini A, Landi F.
Anticholinergic drugs and negative outcomes in the older population: from biological plausi-
bility to clinical evidence. Aging Clin Exp Res. 2016;28:25–35.
Dean L. Clozapine therapy and CYP2D6, CYP1A2, and CYP3A4 genotypes. In: Pratt V, McLeod
H, Rubinstein W, Dean L, Kattman B, Malheiro A, editors. Medical genetics summaries.
Bethesda: National Center for Biotechnology Information (US); 2012.
Dodd S, Mitchell PB, Bauer M, Yatham L, Young AH, Kennedy SH, Williams L, Suppes T, Lopez
Jaramillo C, Trivedi MH, Fava M, Rush AJ, McIntyre RS, Thase ME, Lam RW, Severus E,
Kasper S, Berk M. Monitoring for antidepressant-associated adverse events in the treatment of
patients with major depressive disorder: an international consensus statement. World J Biol
Psychiatry. 2018;19:330–48.
Edwards IR, Aronson JK. Adverse drug reactions: definitions, diagnosis, and management. Lancet.
2000;356:1255–9.
FDA. Guidance for industry – good pharmacovigilance practices and pharmacoepidemiologic
assessment. Food and Drug Administration. 2005. https://www.fda.gov/regulatory-informa
tion/search-fda-guidance-documents Accessed 21 Nov 2019.
Fornasier G, Francescon S, Leone R, Baldo P. An historical overview over pharmacovigilance. Int J
Clin Pharm. 2018;40:744–7.
Gaitatzis A, Sander JW. The long-term safety of antiepileptic drugs. CNS Drugs. 2013;27:435–55.
Garon SL, Pavlos RK, White KD, Brown NJ, Stone CA Jr, Phillips EJ. Pharmacogenomics of off-
target adverse drug reactions. Br J Clin Pharmacol. 2017;83:1896–911.
Gentry CA, Rodvold KA. How important is therapeutic drug monitoring in the prediction and
avoidance of adverse reactions? Drug Saf. 1995;12:359–63.
Gerlach M, Egberts K, Dang SY, Plener P, Taurines R, Mehler-Wex C, Romanos M. Therapeutic
drug monitoring as a measure of proactive pharmacovigilance in child and adolescent psychi-
atry. Expert Opin Drug Saf. 2016;15:1477–82.
Guerrini R, Perucca E. Genetic testing to prevent adverse reactions to antiepileptic drugs: Primum
non nocere. Neurology. 2018;90:155–6.
Hadi MA, Neoh CF, Zin RM, Elrggal ME, Cheema E. Pharmacovigilance: pharmacists' perspective
on spontaneous adverse drug reaction reporting. Int Pharm Res Pract. 2017;6:91–8.
98 M. Scherf-Clavel
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
Categories of Health Economic Evaluation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
In Which Cases Is a Health Economic Evaluation Recommended? . . . . . . . . . . . . . . . . . . . . . . . 106
Parameters of Health Economic Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
Costs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
Perspective . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
Time Horizon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
Decision-Analytical Models and Procedures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
Limitations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
Institutions for Evaluating Cost-Effectiveness in Health Care . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
Health Technology Assessment and Cost-Effectiveness Thresholds . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
Pharmacoeconomic Evaluation of Drugs Prescribed for Brain Disorders . . . . . . . . . . . . . . . . . . . . . 114
Supplement: Pharmacoeconomic Evaluations of Brain Disorders . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225
Abstract
Newly developed drugs and medical devices have made a major contribution to
improving the health of populations. However, the rise in health care expenditure
has been a major concern due to the financial pressure it imposes on public
budgets. In response health authorities seek instruments to contain drug
R. Dodel (*)
Department of Geriatric Medicine, University Duisburg-Essen, Essen, Germany
e-mail: Richard.dodel@uk-essen.de
C. Kruse
Uniklinik Essen, Essen, Germany
e-mail: Christopher.kruse@uk-essen.de
A. Conrads-Frank · U. Siebert
Department of Public Health, Health Services Research and Health Technology Assessment,
UMIT – University for Health Sciences, Medical Informatics and Technology, Hall, Tyrol, Austria
e-mail: Annette.Conrads-Frank@umit.at; Uwe.siebert@umit.at
Introduction
Over the past decades, newly developed drugs and medical devices have made a
major contribution to improving the health of populations (Grootendorst et al. 2009).
However, in most countries, the rise in health care expenditure has been a major
concern due to the financial pressure it imposes on public budgets: during 1995–
2005 the OECD countries have experienced an annual average health expenditure
growth per capita of 4%, which outpaced the economic growth of 2.2%. According
to a new OECD forecast, health expenditure will outpace GDP growth over the next
15 years in almost every OECD country; health spending per capita will grow at an
average annual rate of 2.7% across the OECD and will reach 10.2% of GDP by 2030,
up from 8.8% in 2018 (OECD 2019).
This health expenditure growth has been attributed to several factors including
the advance of medical science, the ageing of societies, the increasing prevalence of
chronic conditions, health care price inflation, but also due to an increased expen-
diture on medications. The latter caused an annual average growth in per capita of
4,6% (1995–2006) exceeding the annual rise in health expenditure (OECD 2019).
From 2013–2017 spending on pharmaceuticals increased more modestly at 1.6% per
year (OECD 2019), however in 2017, retail pharmaceuticals still accounted for
almost one-fifth of all health care expenditure and represented the third largest
pending component in OECD countries after inpatient and outpatient care (OECD
2019).
In response to this development, health authorities seek instruments to contain
drug expenditures while still supporting further health improvements.
Pharmacoeconomics 103
Health economic studies are basically divided into two categories: comparative and
noncomparative study types (Fig. 1). The choice of the study type depends on the
research question and the purpose (e.g., is the use of a new drug therapy in question
in comparison with the previous standard therapy or is the aim to decide about the
allocation of financial resources to one or another patient group).
Cost studies are the simplest form of an economic evaluation and may serve
different purposes. It deals either with the costs of a specific measure (e.g., costs of a
cCT, performance of a genetic test) or, in a cost-of-illness study, all costs associated
with the disease are determined. Cost-of-illness studies are primarily used to deter-
mine the economic burden of a disease to society by measuring costs of diagnosing
and treating a disease, as well as costs arising as a result of the disease (e.g.,
productivity losses) (Onukwugha et al. 2016). These studies serve as reference
standards for economic analyses and provide important information on resource
allocation for decision-makers in the health care environment.
So-called complete health economic evaluations are studies that carry out a
comparative assessment between different procedures and put the resulting resource
consumption (e.g., costs) in relation to an outcome (Table 1). The outcome may be
expressed in improvements in the health status (clinical effectiveness):
104 R. Dodel et al.
Comparing Non-comparing
Cost-of-illness
Cost-benefit analysis analysis
Cost-effectiveness
analysis
Cost-utility analysis
Difference in cost
II I
Intervention is Intervention is
medically inferior medically
and more superior and
expensive more expensive
+
Difference in result
– IV
III
Intervention is
Intervention is
medically
medically inferior
superior and less
and less
expensive
expensive
The results of economic studies are determined by the way the underlying data/
information is collected and evaluated. In the following section, the different types of
costs, the possible perspectives, and the problem of the cost and benefit flows
occurring at different times are briefly addressed.
Pharmacoeconomics 107
Costs
Direct medical costs and nonmedical costs consist of costs directly attributable to
the treatment and care of patients (Table 2). Examples of such costs are drug costs or
hospital costs. Indirect costs mainly describe costs due to productivity loss, that is,
the loss of work potential that a society suffers due to the absence from work due to
illness or limited performance at work. Indirect costs can be calculated using the
human capital approach (Ernst 2006). Alternative approaches are, for example, the
friction cost approach (Koopmanschap et al. 1995), which, however, have not yet
been included in all recommendations for the implementation of health economic
analyses. The friction cost approach is based on the assumption that a loss of
productivity is only assumed for the duration of the average vacancy of unfilled
positions (so-called friction period). This is intended to avoid an overestimation of
costs due to the loss of work, as is conceivable in the human capital approach. For a
detailed discussion, please refer to the special literature (Krol et al. 2013; Sach and
Whynes 2003).
Intangible costs are costs due to pain, joy, suffering, disability, or a change in
quality of life that are difficult to measure and barely quantifiable in monetary terms
(Gold et al. 1996). Particularly in the case of chronic diseases, it is important for the
assessment of a service to make the changes in well-being transparent. Intangibles
costs pose a challenge in the survey, as these costs are difficult to measure and
usually cannot be evaluated. In cost-utility analyses, intangible costs are assessed in
terms of quality of life reduction and are included in the denominator of the cost-
utility ratio.
Perspective
Perspective is the viewpoint from which costs and benefits are recorded and evalu-
ated. Health economic analyses can be carried out from different perspectives, for
108 R. Dodel et al.
• Patients
• Service provider (e.g., hospital, general practitioner)
• Service providers (e.g., health insurance companies)
• Society
For example, from the patient’s perspective, costs should include direct non-
medical costs and intangible costs, expenses that are borne by the patient or are
relevant to the patient while the “costs” for hospitalization and medication (excep-
tion: co-payment) are “irrelevant” from the patient’s perspective, since he or she does
not have to pay them. This does not apply if the costs are to be determined from the
perspective of the health insurance companies; here the direct costs due to hospital
stay, rehabilitation clinic, and diagnostics are relevant.
Time Horizon
The time horizon describes the duration of the evaluation period of a study. This can
be measured in days, months, years, but also in cycles. This should be taken into
account in particular if the costs and benefits of different medical measures do not
arise within a short period of time, then direct comparability is limited. In order to
enable a comparison of costs and consequences/benefits at the current point in time,
future costs and benefits are discounted.
Data basis of economic evaluation studies. The data basis on which health
economic evaluations can be carried out can vary and depends on factors such as
time and the available budget for carrying out the analysis. In principle, it is possible
to collect information prospectively or retrospectively within the framework of field
studies (e.g., within the framework of clinical studies) or to draw on previously
published studies.
If the available cost and effectiveness data do not have sufficient external validity
because they do not fully reflect the practice of health care provision, health
economic modeling can be carried out using decision-analytical procedures
Pharmacoeconomics 109
When modeling or analyzing the decision problem, several methods are distin-
guished, including the decision tree method and Markov models. The decision tree
method is used for simple decision situations that usually have a short time horizon
in which all events related to the action strategy occur. In contrast, Markov models or
state-transition models (Siebert et al. 2012) are mainly used for more complex
problems with a longer time horizon. For studies with “queuing problems” (e.g.,
organ transplant list), discrete event simulations can be helpful (Karnon et al. 2012).
For further and more complex model methods, we refer to previous review articles
(Brennan et al. 2006; Kuntz et al. 2017). In the decision tree method, the possible
decisions, events, and outcomes are structured in the form of a tree (Fig. 3). The
decision tree contains the temporal and logical structure of the decision problem and
all relevant alternative strategies, uncertain events, their probability of occurrence
and the expected consequences. The “tree” starts at its trunk on the left side of the
diagram, where the target population of the decision problem is named. It is followed
by a so-called decision node, which represents the choice between different action
alternatives. For each action alternative, there are places in the tree that are marked
by so-called chance nodes. At a chance node, different events can occur or features
can be revealed that are not predictable and thus embody the uncertainty of the
decision problem. Various event paths lead from the trunk to the right side of the
decision tree and end there at the leaves of the tree, result nodes. At a result node, the
consequences of the decision of interest are listed, examples are epidemiological
measures, health conditions, laboratory values, costs.
An event with all subsequent event paths is called a branch of the event tree. For
each alternative strategy, the expected value of the clinical outcome can be calculated
as a weighted average of all possible outcomes by using the probabilities of
occurrence as weights. Decision tree analyses are used, especially in situations
where all relevant events occur within a short time horizon. This pattern represents
the basic scheme of all decision analyses (including nonmedical). Many decision
problems under uncertainty can be structured or broken down into subproblems in
such a way that the question arises as to whether a certain action should be taken,
whether it should be omitted, or whether further information should be obtained first
to reduce the uncertainty.
Markov models, on the other hand, are used if
Fig. 3 Hypothetical case study: Decision tree analysis for a screening test (Siebert 2003a). Disease K occurs more frequently in a risk population and, without
treatment, leads to a reduction in life expectancy and quality of life. The disease can be successfully treated and the therapy leads to a significant increase in life
expectancy and quality of life. In healthy people it leads to a reduction. There is a screening test which has a certain error rate and a risk of complications.
Therapy (€50,000) and the screening test (€4,000) lead to increased costs. The following alternatives are to be evaluated: 1. Therapy is started immediately for all
persons in a risk group; 2. Therapy is not started for all persons in the risk group; 3. A screening test is carried out once and the test positive persons are treated. In
doing so, one accepts the risk of complications and lethality and false test results
111
112 R. Dodel et al.
life expectancy (QALYs), and total costs over time can then be determined for
each alternative and compared accordingly.
Limitations
The decision analysis contains dangers if used improperly. Care must be taken to
ensure that complex interrelationships are not oversimplified based on unrealistic
assumptions and lead to a pseudo-scientific presentation of the facts. In particular,
the complex methodology has a negative effect here, since these models are no
longer transparent if the documentation is inadequate (Husereau et al. 2013). For
users, a model can present itself as a black box whose contents cannot be understood.
In summary, it can be said that desk research is the only possibility for structured and
systematic decision-making in situations where conventional study approaches reach
their limits. However, responsible communication of results is important: results
should be based on the current state of knowledge and their validity depends on the
accuracy of the model structure, the underlying assumptions, and the parameters
chosen. Recently, the Consolidated Health Economic Evaluation Reporting Stan-
dards (CHEERS) statement was introduced as an attempt to consolidate and update
previous health economic evaluation guidelines efforts into one current, useful
reporting guidance (Husereau et al. 2013). It provides a framework on which
transparent reporting is possible and should be followed.
Pharmacoeconomics 113
Finally, it should be noted that the decision analysis only has a supporting
function with regard to the decision-making process.
Health technology assessment (HTA) has become a standard policy tool for
informing decision makers who must manage the entry and use of pharmaceuticals,
medical devices, and other technologies (including complex interventions) within
health systems, for example, through reimbursement and pricing.
HTA relies on evaluations of the clinical, epidemiological, and economic data to
make decisions about the allocation of the scarce resources of public healthcare. The
centerpiece of a cost-effectiveness analysis is the incremental cost-effectiveness ratio
(ICER), which measures the differential cost for a unit of extra benefit gained from a
new therapeutic strategy. In almost all situations, new drug options that apply for
public funding present higher costs and effectiveness than the drugs currently in use.
For a new drug to be recommended based on economic assessments, several
authorities have argued that the ICER must be compared to a cost-effectiveness
threshold (CET) value that should represent the highest acceptable cost for an extra
unit of benefit.
In the UK, NICE introduced a cost per QALY threshold of £20.000–30.000 per
QALY. For cost-per QALY amounts below £20.000, NICE will generally recom-
mend coverage; for cost-per QALY amounts over £30.000, NICE will generally not
recommend the intervention to be covered. In between those amounts, NICE will
more carefully consider a coverage decision. The WHO provided similar recom-
mendations if the intervention’s cost-effectiveness ratio is less than the gross domes-
tic product per capita, it is considered very cost-effective; if between one and three
times GDP per capita, it is considered cost-effective. Above three times GDP per
capita, it is considered something to be evaluated with caution. Recently, the WHO
moved away from this characterization of thresholds and has been examining
alternative threshold estimation methods (Brock et al. 2017).
Cost-effectiveness threshold values around different healthcare system have been
summarized in a recent article by Santos et al. and the reader is referred to this
chapter for further information (Santos et al. 2018).
No group of chronic diseases burdens the world more than mental illnesses (Report
of the secretary (WHO) 2011). WHO data suggest that brain disorders cause one-
third of the burden of all diseases; according to current epidemiological estimates,
brain dis-orders account for approximately 13% of global disease prevalence,
surpassing both cardiovascular diseases and cancer (Collins et al. 2011). It represents
the largest contributor to the ‘all cause morbidity’ burden as measured by disability-
adjusted life years (DALYs). A burden of illness study from the European brain council
estimated total cost of brain disorders (19 groups of disorders) including 30 European
countries to be €798 billion, with direct health care cost 37%, direct nonmedical cost
Pharmacoeconomics 115
23%, and indirect cost 40%. The average cost per inhabitant was €5.550. The European
average cost per person with a disorder of the brain ranged between €285 for headache
and €30,000 for neuromuscular disorders (Gustavsson et al. 2011).
The spectrum of brain disorders is large, covering hundreds of disorders with a
considerable number of drugs registered. Thus, a detailed description of the various
pharmacoeconomic evaluations would go beyond the aim of this chapter. However,
to provide an overview and to identify probable gaps of currently available
pharmacoeconomic evaluations, we devised a list of each major brain disorder based
on a systematic search in PubMed for the years 2010–2020 using MesH headings. The
selection of the disorders was based on the DSM-V and the ICD-10 for neurological
disorders. Where an adequate review on pharmacoeconomic evaluations was avail-
able, this review was included in representation of earlier published research articles.
In the supplemental table, we compiled the currently available evidence.
In conclusion from the results of this systematic literature search, there are
considerable gaps in the pharmacoeconomic evaluation of the drugs used for brain
disorders. This is in contrast to the immense burden of those disorders.
As Gustavson et al. pointed out in their major work on the burden of disease in
Europe: “Disorders of the brain are one of the major economic challenges for
European healthcare now and for the future” (Gustavsson et al. 2011).
A literature search was conducted in line with a previous publication of our group in
the databases “PubMed”(Dams et al. 2011). The following keywords were
employed: “decision analysis,” “decision-analytic,” “decision model,” “health care
model,” “health care evaluation model,” “decision tree,” “Markov model,” “discrete
event simulation,” “cost-effectiveness,” “cost-utility,” “cost-benefit,” “cost-minimiz
(s)ation,” “QALY,” “and the respective disease. The following diseases were
searched: Psychiatric diseases (neurodevelopmental disorders, Schizophrenia spec-
trum and other psychotic disorders, bipolar and related disorders, depressive disor-
ders, anxiety disorders, obsessive-compulsive and related disorders, trauma- and
stressor-related disorders, dissociative disorders, somatic symptom and related dis-
orders, feeding and eating disorders, elimination disorders, sleep-wake disorders,
sexual dysfunctions, gender dysphoria, disruptive, impulse-control, and conduct
disorders substance-related and addictive disorders, neurocognitive disorders/
dementia, personality disorders, paraphilic disorders); Neurologic diseases (head-
ache, tinnitus, vertigo, epilepsy, stroke, meningitis, multiple sclerosis, myasthenia
gravis, restless legs syndrome, Parkinson’s disease, atypical parkinsonian syn-
dromes, ataxias, Wilson’s disease, amyotrophic lateral sclerosis, tremor, poly-
neuropathy, myopathies).
Furthermore, we examined reference lists of studies identified and reviewed
articles from the archives of the authors. The search was restricted to the following
three languages: English, German, Spanish. We only included published full papers
that employed a decision-analytic model or another type of mathematical healthcare
116
Table 4
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Neurodevelopmental disorders
Hassiotis et al. GBR 2018 Neurodevelopmental – Manual-assisted face- RCT QALY (among “Findings from the
(2018) disorders (intellectual to-face positive others) main study and the
disorder, ID) behavior support (PBS) naturalistic follow-up
training to therapists suggest that staff
and treatment as usual training in PBS as
(TAU) compared with delivered in this study
TAU only in the control is insufficient to
arm achieve significant
clinical gains beyond
TAU in community
ID services. Although
there is an indication
that training in PBS is
potentially cost-
effective, this is not
maintained in the
longer term. There is
increased scope to
develop new
approaches to
challenging behavior
as well as optimizing
the delivery of PBS in
routine clinical
practice”
Zimovetz et al. GBR 2018 Neurodevelopmental Lisdexamfetamine Lisdexamfetamine Decision- Cost- “From the perspective
(2018) disorders (ADHD) dimesylate (LDX), dimesylate (LDX) analytic model effectiveness of the UK NHS, LDX
methylphenidate extended versus methylphenidate is likely to provide a
release (MPH-ER), extended release cost-effective
atomoxetine (ATX) (MPH-ER) and treatment for adults
atomoxetine (ATX) with ADHD. This
R. Dodel et al.
conclusion may be
drawn with more
certainty in
comparison with ATX
than with MPH-ER”
Zimovetz et al. GBR 2016 Neurodevelopmental Lisdexamfetamine Lisdexamfetamine Decision- Cost- “From the perspective
(2016) disorders (ADHD) dimesylate (LDX), dimesylate (LDX) analytic model effectiveness of the UK NHS, LDX
Pharmacoeconomics
(continued)
118
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
lower expected health
outcomes than other
ADHD medications
in children and
adolescents who
failed prior stimulant
therapy. Furthermore,
AAPs were not a cost-
effective option”
Lachaine et al. CAN 2016 Neurodevelopmental Guanfacine extended release, GXR + stimulant versus Markov model Cost- “This economic
(2016) disorders (ADHD) amphetamine mixed salts, long-acting stimulant effectiveness evaluation
methylphenidate HCl monotherapy demonstrates that
formulations, and GXR + stimulant is
lisdexamfetamine dimesylate cost-effective
compared to stimulant
alone in the treatment
of children and
adolescents with
ADHD in Canada”
Maia et al. (2016) BRA 2016 Neurodevelopmental Methylphenidate immediate- MPH-IR versus no Markov model Cost- “MPH-IR treatment
Disorders (ADHD) release (MPH-IR) treatment effectiveness of children and
adolescents is cost-
effective for ADHD
patients from the
Brazilian public
health system
perspective. Both
patients and the
healthcare system
might benefit from
such a strategy”
R. Dodel et al.
Tilford et al. USA 2015 Neurodevelopmental Melatonin, cognitive Melatonin versus CBT Prospective Cost- “Predicted treatment
(2015) disorders (Autism) behavioral therapy (CBT) versus Melatonin and cohort study effectiveness effects for melatonin
CBT versus Parent and behavioral
based sleep education interventions were
similar in magnitude
for the child and for
the caregiver.
Accounting for
Pharmacoeconomics
caregiver spillover
effects associated
with treatments for
the child with ASD
increases treatment
benefits and improves
cost-effectiveness
profiles”
Schawo et al. NLD 2015 Neurodevelopmental Methylphenidate immediate- OROS versus IR Markov model Cost- “The results indicate
(2015) disorders (ADHD) release (IR) versus effectiveness that, for children
methylphenidate osmotic- responding
release oral system (OROS) suboptimally to
treatment with IR, the
beneficial effect of
OROS on compliance
may be worth the
additional costs of
medication. The
presented model adds
to the health
economic information
available for
policymakers and to
considerations on a
broader perspective in
cost-effectiveness
analyses”
(continued)
119
Table 4 (continued)
120
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Catala-Lopez ESP 2013 Neurodevelopmental Medications launched in Systematic Cost- “The pharmacological
et al. (2013) disorders (ADHD) Spain for treatment of ADHD review of effectiveness treatment of ADHD
in children and adolescents literature in children and
adolescents, with the
reservations arising
from the
generalization of
results to different
settings, is probably
cost-effective in the
short term. The
existing studies do not
allow the relative
efficiency of different
treatments to be
established, either in
the long-term
treatment or in patient
subgroups with
specific
characteristics or
comorbidities”
Erder et al. USA 2012 Neurodevelopmental Guanfacine extended-release GXR versus ATX Matching- Cost- “To our knowledge,
(2012) disorders (ADHD) (GXR), atomoxetine (ATX) adjusted indirect effectiveness this is the first
comparison application of the
(MAIC) novel comparative
efficacy method of
MAIC to a CEA
model. The MAIC
results indicate that
GXR (0.075–
0.12 mg/kg/day) was
more effective than
R. Dodel et al.
ATX (1.2 mg/kg/day)
in the trial population.
The CEA results
indicate that GXR is
cost effective
compared with ATX
for the treatment of
ADHD in children
Pharmacoeconomics
and adolescents”
Sikirica et al. USA 2012 Neurodevelopmental Guanfacine extended-release GXR + stimulant versus Markov model Cost- “The impairment
(2012) disorders (ADHD) (GXR) stimulant monotherapy effectiveness associated with
residual ADHD
symptoms after
stimulant therapy is
becoming
increasingly
recognized. This is
the first analysis of the
cost effectiveness of
stimulants combined
with an adjunctive
medication. This
study suggests that
the adjunctive therapy
of GXR with
stimulants is a cost-
effective treatment
based on a
willingness-to-pay
threshold of
$US50,000/QALY.
This may address an
unmet need among
patients with
suboptimal response
(continued)
121
122
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
to stimulant
monotherapy”
Tilford et al. USA 2012 Neurodevelopmental Sensitivity of two Cost- “The HUI3 was more
(2012) disorders (Autism) generic preference- effectiveness sensitive to clinical
based instruments measures used to
relative to ASD-related characterize children
conditions and with autism compared
symptoms with the Quality of
Well-Being Self-
Administered
(QWB-SA) scale. The
findings provide a
benchmark to
compare scores
obtained by
alternative methods
and instruments.
Researchers should
consider
incorporating the
HUI3 in clinical trials
and other longitudinal
research studies to
build the evidence
base for describing
the cost effectiveness
of services provided
to this important
population”
R. Dodel et al.
Wu et al. (2012) USA 2012 Neurodevelopmental Lisdexamfetamine, Pharmacotherapies for Systematic Cost- “Among children and
disorders (ADHD) guanfacine extended-release ADHD, including review of effectiveness adolescents with
and clonidine extended- stimulants and literature ADHD, there was
release nonstimulants consistent evidence
that
pharmacotherapies
are cost effective
compared with no
Pharmacoeconomics
treatment or
behavioral therapy.
Adequate data are
lacking to draw
conclusions regarding
the relative cost
effectiveness of
different
pharmacological
agents. More
economic evaluations
with standardized
methods, such as
effectiveness
measures and cost
components, are
warranted. To better
inform payers about
the economic value of
existing medications,
future studies should
also consider
identifying subgroups
that may have
heterogeneous
responses to different
treatments, including
(continued)
123
124
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
analyses of recently
approved treatments
(e.g.,
lisdexamfetamine,
guanfacine extended-
release and clonidine
extended-release) and
expanding the time
horizon to incorporate
long-term outcomes”
Schlander (2010) DEU 2010 Neurodevelopmental Child psychiatric drug Systematic Cost- “In many cases,
disorders (ADHD) treatment review of effectiveness effectiveness data
and in general: child literature came from short-term
psychiatric drug studies, and
treatment extrapolation to a
one-year time horizon
was usually based on
assumptions. Even
those evaluations
attempting to address
longer time horizons
by way of modeling
did not include the
impact of treatment
on long-term sequelae
of the conditions
studied, mainly due to
a paucity of robust
clinical data.
Nevertheless,
currently available
health economic
R. Dodel et al.
evaluations broadly
suggest an acceptable
to attractive cost
effectiveness of
medication
management of
ADHD, whereas there
is no such evidence
Pharmacoeconomics
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Prasad et al. GBR 2009 Neurodevelopmental Atomoxetine Atomoxetine versus Markov model Cost- “Atomoxetine is cost-
(2009) disorders (ADHD) treatment alternatives effectiveness effective and may
have advantages over
stimulants, including
benefits to HRQL and
no abuse liability and
is the only treatment
in the UK licensed for
continued treatment
into adulthood in
adolescents who have
shown a response
from treatment”
Romeo et al. GBR 2009 Neurodevelopmental Risperidone, haloperidol Risperidone versus RCT Cost- “The treatment of
(2009) disorders (intellectual haloperidol versus effectiveness challenging behavior
disorder) placebo in ID with
antipsychotic drugs is
not a cost-effective
option”
Hong et al. ESP 2009 Neurodevelopmental Atomoxetine, immediate- Methylphenidate versus Markov model Cost- “The economic
(2009) disorders (ADHD) release methylphenidate Atomoxetine effectiveness evaluation showed
hydrochloride (MPH), that atomoxetine is an
extended-release effective alternative
methylphenidate across a range of
hydrochloride (MPH) ADHD populations
and offers value-for
money in the
treatment of ADHD”
R. Dodel et al.
Cottrell GBR 2008 Neurodevelopmental Atomoxetine, immediate- Methylphenidate versus Markov model Cost- “The economic
et al. (2008). disorders (ADHD) release methylphenidate Atomoxetine effectiveness evaluation showed
hydrochloride (MPH), atomoxetine is an
extended-release effective alternative
methylphenidate across a range of
hydrochloride (MPH) ADHD populations
and offers value-for-
money in the
Pharmacoeconomics
treatment of ADHD”
Faber et al. NLD 2008 Neurodevelopmental Long-acting methylphenidate Long-acting Cost- “Methylphenidate-
(2008) disorders (ADHD) osmotic release oral system methylphenidate effectiveness OROS is a cost-
(OROS), immediate-release osmotic release oral effective treatment for
(IR) methylphenidate system (OROS) with youths with ADHD
youths for whom IR for whom treatment
methylphenidate is with IR
suboptimal methylphenidate is
suboptimal. Higher
medication costs of
methylphenidate-
OROS were
compensated for by
savings on resource
use, yielding similar
10-year costs
compared with
treatment with IR
methylphenidate. Our
analysis is sensitive to
both clinical
parameters and
(differences in)
resource utilization
and costs between the
groups modelled,
warranting further
(continued)
127
128
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
research within
clinical trials and
observational
databases, and into
the full scope of
costs”
King et al. (2006) GBR 2006 Neurodevelopmental Oral methylphenidate Oral methylphenidate Systematic Cost- “Drug therapy seems
disorders (ADHD) hydrochloride (MPH), hydrochloride (MPH) review of effectiveness to be superior to no
dexamfetaminesulphate versus literature drug therapy, no
(DEX) and atomoxetine dexamfetaminesulphate significant differences
(ATX (DEX) and atomoxetine between the various
(ATX) in children and drugs in terms of
adolescents (<18 years efficacy or side effects
of age) were found, mainly
owing to lack of
evidence, and the
additional benefits
from behavioral
therapy
(in combination with
drug therapy) are
uncertain. Given the
lack of evidence for
any differences in
effectiveness between
the drugs, the
economic model
tended to be driven by
drug costs, which
differed considerably.
Future trials
examining MPH,
R. Dodel et al.
DEX and ATX should
include the
assessment of
tolerability and safety
as a priority”
Donnelly et al. AUS 2004 Neurodevelopmental Dexamphetamine (DEX), Dexamphetamine RCT Cost- “MPH and DEX are
(2004) disorders (ADHD) methylphenidate (MPH) (DEX) versus effectiveness cost-effective
Pharmacoeconomics
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Schizophrenia spectrum and other psychotic disorders
Debaveye et al. BEL 2019 Schizophrenia Paliperidone Paliperidone for Markov model Benefit- “The overall
(2019) Spectrum and other 1 month (PP1M) versus analysis environmental burden
psychotic disorders 3 months (PP3M) was lower for PP1M
compared to Treatment and PP3M treatment
Interruption (TI) as a than Treatment
control group Interruption because
patients are kept more
stable, which reduces
the environmental
burden due to
hospitals. Moreover,
the Human Health
burden was
outweighed by the
Human Health
benefit”
Zhao CHN 2019 Schizophrenia Olanzapine Olanzapine orally Microsimulation Cost- “As the first-line
et al. (2019). spectrum and other disintegrating tablet model effectiveness treatment for
psychotic disorders (ODT) versus schizophrenia in
olanzapine standard China, olanzapine-
oral tablet (SOT) ODT is cost-effective
compared to
olanzapine-SOT and
olanzapine-SOT is
cost-effective
compared to
aripiprazole-SOT”
R. Dodel et al.
Potaufeu et al. FRA 2019 Schizophrenia Olanzapine pamoate Olanzapine pamoate Mirror model Cost- “By its mirror design,
(2019) spectrum and other versus the hospital cost effectiveness the study was placed
psychotic disorders differential in real conditions of
care of the patient
with schizophrenia. A
total of 61.5% of
patients maintained
treatment with
Pharmacoeconomics
olanzapine pamoate
for a minimum of 1
year. This long-acting
antipsychotics) is
more effective
without significantly
increasing the cost
compared to the
previous therapeutic
strategy (including
oral olanzapine). The
additional cost is
partly due to the
administration
restriction in a
hospital setting in
relation to risk of
Postinjection
Delirium/Sedation
Syndrome (PDSS).
There is currently no
acceptable efficiency
limit. The results of
this cost-effectiveness
analysis cannot be
extrapolated to the
other long-acting
(continued)
131
Table 4 (continued)
132
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
antipsychotics since it
is the only one with
hospital reserve
status. The current
limitations of medico-
economics in
psychiatry derive
from the
heterogeneity of
clinical forms and the
management of
mental pathologies”
Howard et al. GBR 2018 Schizophrenia Low-dose amisulpride Low-dose amisulpride RCT Cost- “Low-dose
(2018) spectrum and other versus placebo effectiveness amisulpride is
psychotic disorders effective and well
tolerated as a
treatment for Very
late-onset (aged >/¼
60 years)
schizophrenia-like
psychosis, with
benefits maintained
by prolonging
treatment. Potential
adverse events
include clinically
significant
extrapyramidal
symptoms and falls”
Wijnen et al. HLD 2018 Schizophrenia Antipsychotic medication Treatment as usual RCT Cost- “In the Dutch context
(2018) spectrum and other versus TAU augmented effectiveness where TAU for
psychotic disorders with cognitive behavior psychosis is guideline
R. Dodel et al.
therapy for social congruent and well
activation (CBTsa) implemented there
appears no added
value for adjunct
CBTsa. In other
settings where the
treatment for the
schizophrenia
Pharmacoeconomics
spectrum disorders
solely relies on
antipsychotics, add-on
CBTsa may lead to
clinically superior
outcomes, but it
should still be
evaluated if adjunct
CBTsa therapy is a
cost-effective
alternative”
Men et al. (2018) CHN 2018 Schizophrenia Amisulpride, Olanzapine Amisulpride versus Review Cost- “As the first meta-
spectrum and other Olanzapine minimization analysis and cost-
psychotic disorders analysis minimization analysis
comparing the
efficacy, safety, and
cost of amisulpride
and olanzapine within
a Chinese setting, the
study suggests that
amisulpride may be
an effective, well-
tolerated, and cost-
saving antipsychotic
drug alternative in
China”
(continued)
133
134
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Basu et al. (2018) USA 2018 Schizophrenia Long-acting injectable (LAI) One monthly and once Decision model Treatment “Our results suggest
spectrum and other formulations of paliperidone every 3 months long- effectiveness that using PP1M and
psychotic disorders palmitate, oral antipsychotics acting injectable (LAI) PP3M treatment
(Oas) formulations of strategies for patients
paliperidone palmitate with schizophrenia
versus oral receiving Medicaid
antipsychotics (Oas) could result in
reduced
hospitalizations. This
finding, along with
improvement to
patients’ health,
should be considered
when assessing the
value of these LAIs”
Girardin et al. USA 2019 Schizophrenia Clozapine Clozapine as a third-line Semi-Markovian Cost- “Up to a decision
(2019) spectrum and other antipsychotic model effectiveness threshold of
psychotic disorders medication $3.9 million per
quality-adjusted life-
year (90-fold the US
gross domestic
product per capita),
the base-case results
indicate that
compared with
current ANCM
(absolute neutrophil
R. Dodel et al.
count monitoring),
genotype-guided
blood sampling prior
to clozapine initiation
appeared cost-
effective for targeted
blood monitoring
only in patients with
Pharmacoeconomics
HLA susceptibility
alleles. Sensitivity
analysis demonstrated
that at a cost of
genotype testing of up
to USD700, HLA
genotype-guided
blood monitoring
remained a cost-
effective strategy
compared with either
current ANCM or
clozapine
substitution”
Nuhoho et al. UAE 2018 Schizophrenia Oral antipsychotics Paliperidone palmitate Decision tree Cost- “PP1M is estimated to
(2018) spectrum and other once monthly (PP1M) model effectiveness save the UAE
psychotic disorders with oral healthcare system
supplementation versus money, while at the
Paliperidone palmitate same time improving
once monthly (PP1M) patient outcomes”
without oral
supplementation
(continued)
135
136
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Barnes et al. GBR 2017 Schizophrenia Clozapine, Amisulpride Clozapine with RCT Cost- “The risk-benefit of
(2017) spectrum and other amisulpride versus effectiveness amisulpride
psychotic disorders clozapine with placebo augmentation of
clozapine for
schizophrenia that has
shown an insufficient
response to a trial of
clozapine
monotherapy is
worthy of further
investigation in larger
studies. The size and
extent of the side
effect burden
identified for the
amisulpride-
clozapine
combination may
partly reflect the
comprehensive
assessment of side
effects in this study.
The design of future
trials of such a
treatment strategy
should take into
account that a clinical
response may be not
be evident within the
R. Dodel et al.
4- to 6-week follow-
up period usually
considered adequate
in studies of
antipsychotic
treatment of acute
psychotic episodes.
Economic evaluation
Pharmacoeconomics
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Einarson et al. NLD 2017 Schizophrenia Paliperidone, haloperidol PP3M vs once-monthly Decision tree Cost- “PP3M dominated all
(2017b) spectrum and other long-acting therapy paliperidone (PP1M), model effectiveness commonly used
psychotic disorders (HAL-LAT), risperidone haloperidol long-acting drugs. It is cost-
microspheres (RIS-LAT), oral therapy (HAL-LAT), effective for treating
olanzapine (oral-OLZ) risperidone chronic schizophrenia
microspheres in the Netherlands.
(RIS-LAT), and oral Results were robust
olanzapine (oral-OLZ) over a wide range of
sensitivity analyses.
For patients requiring
a depot medication,
such as those with
adherence problems,
PP3M appears to be a
good alternative
antipsychotic
treatment”
Einarson et al. ESP 2017 Schizophrenia Paliperidone PP3M versus PP1M Decision tree Cost- “PP3M dominated
(2017a) spectrum and other model effectiveness PP1M in all analyses
psychotic disorders and was, therefore,
cost-effective for
treating chronic
relapsing
schizophrenia in
Spain. For patients
who require long-
acting therapy, PP3M
appears to be a good
alternative anti-
psychotic treatment”
R. Dodel et al.
Nemeth et al. HUN 2017 Schizophrenia Cariprazine, risperidone Cariprazine versus Markov model QALY gains “Cariprazine, which
(2017) spectrum and other risperidone showed clinically
psychotic disorders meaningful
improvement in the
symptoms, and
personal and social
performance, can also
provide significant
Pharmacoeconomics
(continued)
Table 4 (continued)
140
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Druais et al. FRA 2017 Schizophrenia Aripiprazole LAI (long- Aripiprazole LAI Markov model Cost- “This analysis, to the
(2017) spectrum and other acting injectable; ALAI), (ALAI) versus effectiveness best of our
psychotic disorders olanzapine LAI (OLAI), olanzapine LAI (OLAI) knowledge, is the first
paliperidone LAI (PLAI), versus paliperidone LAI of its kind to assess
risperidone LAI (RLAI), (PLAI) versus the cost-effectiveness
haloperidol decanoate (HD), risperidone LAI (RLAI) of antipsychotics
oral olanzapine versus haloperidol based on French
decanoate (HD) versus observational data.
oral olanzapine PLAI was associated
with the highest
probability of being
the optimal treatment
from the French
health insurance
perspective”
Correll et al. USA 2016 Schizophrenia Long-acting injectable Long-acting injectable Review Cost- “The evidence review
(2016) spectrum and other antipsychotics (LAIs) antipsychotics (LAIs) effectiveness demonstrated that
psychotic disorders versus oral LAIs are superior to
antipsychotic placebo for acute and
maintenance
treatment of
schizophrenia and, in
general, appear to be
similar to one another
in terms of
schizophrenia relapse
prevention. Study
design impacts the
demonstrated efficacy
of LAIs versus oral
antipsychotics, but
recent database and
randomized
R. Dodel et al.
controlled studies
favor the use of LAIs
in early-phase
schizophrenia
patients. LAIs vary
considerably in their
propensity to cause
certain adverse
Pharmacoeconomics
effects, including
weight gain,
metabolic effects,
extrapyramidal
symptoms, and
prolactin elevation,
and these differences
can be used to help
guide LAI selection.
Some studies, but not
all, have
demonstrated
significant reductions
in health care
utilization or overall
costs with LAIs. The
expert panel identified
several barriers to
LAI use in current
practice, including
clinician lack of
knowledge, negative
attitudes about LAIs,
and resource and cost
issues. The
participants also
identified a number of
(continued)
141
142
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
additional factors that
should be considered
when weighing the
use of LAI therapy,
including medication
adherence, relapse
risk and severity,
cognitive impairment,
ease of use, substance
misuse, access and
cost, stigma, social
support, patient
autonomy, control
over medication
dosing, fear of
needles, and the
potential for patient
harm due to relapses
and associated loss of
functioning. This
evidence review,
discussion, and
summary
recommendations
may help clinicians,
patients, families,
payers, and other
stakeholders to better
characterize the role
of LAIs in the
treatment of
schizophrenia”
R. Dodel et al.
Bipolar and related disorders
Augusto et al. USA 2018 Bipolar and related Aripiprazole, Risperidone Aripiprazole versus Markov model Cost- “AOM 400 may be
(2018) disorders LAI, Paliperidone palmitate, Ris-peri-done LAI effectiveness considered cost
Cariprazine, Asenapine, BSC versus Paliperidone effective in the
palmitate versus maintenance
Cariprazine versus Ase- monotherapy
na-pine versus BSC treatment of BP-I
Pharmacoeconomics
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Mavranezouli GBR 2017 Bipolar and related Lithium, antipsychotics (e.g., Depending on reviewed Review Cost-utility, “Pharmacological
and Lokkerbol disorders aripiprazole, asenapine, study cost- interventions are cost
(2017) haloperidol, lurasidone, effectiveness, effective, compared
olanzapine, quetiapine, cost-benefit with no treatment, in
risperidone), antidepressants and cost- the management of
(e.g., fluoxetine, imipramine, consequence BD, both in the acute
paroxetine, venlafaxine) and analyses and maintenance
antiepileptics (such as phases. However, it is
carbamazepine, lamotrigine difficult to draw safe
and valproate, either as conclusions on the
valproic acid or sodium relative cost
valproate) effectiveness between
drugs due to
differences across
studies and
limitations
characterizing many
of them. Future
economic evaluations
need to consider the
whole range of
treatment options
available for the
management of BD
and adopt appropriate
methods for evidence
synthesis and
economic modelling,
R. Dodel et al.
to explore more
robustly the relative
cost effectiveness of
pharmacological
interventions for
people with BD”
Depressive disorders
Pharmacoeconomics
Sado et al. (2019) JPN 2019 Depressive disorders Mirtazapine, SSRIs, other Mirtazapine versus Markov model Cost- “When considering
antidepressants other antidepressants effectiveness the early stage
efficacy of
mirtazapine, it
appeared to be cost-
effective compared to
selective serotonin
reuptake inhibitors,
especially for severe
depression and in the
early stage treatment
in the Japanese
setting. However, our
study has some
limitations. First,
mirtazapine is
compared with
batched selective
serotonin reuptake
inhibitors rather than
individual ones.
Second, we did not
consider
antidepressant
combination therapy
as treatment options”
(continued)
145
Table 4 (continued)
146
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Rubio-Valera ESP 2019 Depressive disorders Antidepressants Antidepressants versus RCT Cost-utility, “Incremental cost-
et al. (2019) active monitoring Cost- utility ratios favor
effectiveness pharmacological
treatment as a first-
line approach for
patients with mild-
moderate major
depressive disorder.
While our results
should be interpreted
with caution and
further real world
research is needed,
clinical practice
guidelines should
consider
antidepressant
therapy for mild-
moderate major
depressive patients as
an alternative to
active monitoring in
PC”
Yoon et al. USA 2018 Depressive disorders Aripirazole, bupropion Ariprazole versus RCT Cost- “In treatment of
(2018) bupropion versus effectiveness depression with less
ariprazole with switch than optimal
to bupropion response,
augmentation with
either aripiprazole or
bupropion was cost-
effective relative to
switching to
bupropion”
R. Dodel et al.
Kessler et al. GBR 2018 Depressive disorders SSRI, SNRI, oral mirtazapine Oral mirtazapine and RCT Cost- “This study did not
(2018) usual medication versus effectiveness find convincing
placebo and usual evidence of a
medication clinically important
benefit for
mirtazapine in
addition to a SSRI or
a SNRI antidepressant
Pharmacoeconomics
over placebo in
primary care patients
with TRD. There was
no evidence that the
addition of
mirtazapine was a
cost-effective use of
NHS resources. GPs
and patients were
concerned about
adding an additional
antidepressant”
Groessl et al. USA 2018 Depressive disorders Standard of care (SOC) Pharmacogenetic test Markov model Cost- “Pharmacogenetic
(2018) medication management (Idx) versus SOC effectiveness testing among
moderate to severe
MDD patients
improved QALYs and
resulted in cost
savings. Sensitivity
analyses supported
the robust nature of
the current findings of
the dominant IDGx
test to guide
treatment”
(continued)
147
148
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Christensen and NOR 2018 Depressive disorders Vortioxetine, Duloxetine Vortioxetine versus Cost- “Vortioxetine may be
Munro (2018) duloxetine effectiveness a cost-effective
alternative to
duloxetine, owing to
its superior ability to
improve functional
capacity. The dual-
response STP concept
introduced here
represents a more
comprehensive
analysis of the cost-
effectiveness of
antidepressants”
Lee et al. (2018) CAN 2018 Depressive disorders Agomelatine, bupropion, Depending on reviewed Review Work place “Extant data suggest
desvenlafaxine, duloxetine, study functioning that antidepressant
fluoxetine, levomilnacipran, treatment improves
paroxetine, sertraline, workplace outcomes
venlafaxine, vortioxetine in MDD. The
capability of
antidepressants in
improving measures
of workplace
functioning should be
considered in cost-
benefit analyses to
better inform cost-
modelling studies
pertaining to
antidepressant
therapy”
R. Dodel et al.
Voigt et al. USA 2017 Depressive disorders Antidepressant medication, Repetive Transcranial Markov model Cost- “rTMS was identified
(2017) Repetive Transcranial Magnetic Stimulation effectiveness as the dominant
Magnetic Stimulation (rTMS) (rTMS) Antidepressant therapy compared to
medication antidepressant
medication trials over
the life of the patient
across the lifespan of
adults with MDD,
Pharmacoeconomics
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Peterson et al. USA 2017 Depressive disorders Antidepressants Antidepressant Review Cost- “Certain
(2017) treatment versus usual effectiveness pharmacogenomics
care among others tools show promise of
improving short-term
remission rates in
women in their
mid-40s with few
comorbidities. But,
important evidence
limitations preclude
recommending their
widespread use and
indicate a need for
further research”
Rosenblat et al. CAN 2017 Depressive disorders Antidepressant medication Pharmacogemomic Review Cost- “A limited number of
(2017) testing versus unguided effectiveness studies have shown
groups promise for the
clinical utility of
pharmacogenomic
testing; however,
cost-effectiveness of
pharmacogenomics,
as well as
demonstration of
improved health
outcomes, is not yet
supported with
replicated evidence”
R. Dodel et al.
Ophuis et al. NLD 2017 Anxiety Disorders Pregabalin, SSRI, Pharmacological versus Cost- Review “Forty-two studies
(2017) venlafaxine, diazepan, pharmacological and effectiveness reporting cost-
citalopram, escitalopram, pharmacological versus effectiveness of
paroxetin psychological interventions for
anxiety disorders
were identified. iCBT
was cost-effective in
comparison with the
Pharmacoeconomics
control conditions.
Psychological
interventions for
anxiety disorders
might be more cost-
effective than
pharmacological
interventions”
Obsessive-compulsive and related disorders
Fineberg et al. GBR 2018 Obsessive- SSRI CBT versus SSRI RCT Cost- “The mean Quality
(2018) compulsive and versus SSRI+CBT effectiveness Adjusted Life Year
related disorders scores for sertraline
were 0.1823 (95%
confidence interval:
0.0447–0.3199)
greater than for CBT
and 0.1135 (95%
confidence interval:
0.0290-0.2560),
greater than for
(continued)
151
152
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
combined treatment.
Combined treatment
appeared the most
clinically effective
option, especially
over CBT, but the
advantages over SSRI
monotherapy were
not sustained beyond
16 weeks. SSRI
monotherapy was the
most cost-effective. A
definitive study can
and should be
conducted”
Skapinakis et al. GBR 2016 Obsessive- SSRI Pharmacological versus Review Cost- “In adults,
(2016) compulsive and placebo and effectiveness psychological
related disorders psychological versus interventions,
placebo and clomipramine, SSRIs
pharmacological versus or combinations of
psychological these are all effective,
whereas in children
and adolescents,
psychological
interventions, either
as monotherapy or
combined with
specific SSRIs, were
more likely to be
effective. Future
RCTs should improve
R. Dodel et al.
their design, in
particular for
psychotherapy or
combined
interventions”
Trauma- and stressor-related disorders
Painter et al. USA 2017 Trauma- and stressor- RET Cost- “Because of the
Pharmacoeconomics
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Mihalopoulos AUS 2015 Trauma- and stressor- SSRI SSRI versus trauma- Economic Cost- “The three Guideline
et al. (2015) related disorders focused cognitive modelling effectiveness recommended
behavioral therapy interventions
(TF-CBT) and current evaluated in this study
practice in Australia are likely to have a
versus TF-CBT (for positive impact on the
children) economic efficiency
of the treatment of
PTSD (posttraumatic
stress disorder) if
adopted in full. While
there are gaps in the
evidence base, policy-
makers can have
considerable
confidence that the
recommendations
assessed in the current
study are likely to
improve the
efficiency of the
mental health care
sector”
Le et al. (2014) USA 2014 Trauma- and stressor- Sertraline Sertraline versus Randomized Cost- “Giving PTSD
related disorders prolonged exposure preference trial effectiveness (posttraumatic stress
therapy disorder) patients a
choice of treatment
appears to be cost-
R. Dodel et al.
effective. When
choice is not possible,
prolonged exposure
therapy may provide a
cost-effective option
over
pharmacotherapy
Pharmacoeconomics
with sertraline”
Polak et al. NLD 2012 Trauma- and stressor- Paroxetine Parxetine versus RCT Cost- “This study is unique
(2012) related disorders psychological first line effectiveness for its direct
treatment (TF-CBT) comparison of the
most commonly used
psychological
intervention
(TF-CBT) and
pharmacological
intervention
(paroxetine) on (cost-)
effectiveness on the
short and the long
term. The anticipated
results will provide
relevant evidence
concerning long-term
effects and relapse
rates and will be
beneficial in reducing
societal costs. It may
also provide
information on who
may benefit most from
which type of
intervention. Some
methodological issues
will be discussed”
(continued)
155
156
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Somatic symptom and related disorders
Wortman et al. NLD 2018 Somatic symptom Pharmacotherapy Review Cost- “This review provides
(2018) and related disorders versus behavioral effectiveness an overview of
therapy among others 39 included studies of
interventions for
patients with MUS
and FSS (medically
unexplained
symptoms (MUS) or
functional somatic
syndromes (FUS))
and the
methodological
quality of these
studies. Considering
the limited
comparability due to
the heterogeneity of
the studies, group
interventions might
be more cost-effective
than individual
interventions”
Le et al. (2018) AUS 2018 Feeding and eating Antidepressant medication Depending on reviewed Review Cost- “Cost-effectiveness
disorders study effectiveness studies in eating
disorder appear to be
increasing in number
over the last 6 years.
Findings were
inconsistent and no
R. Dodel et al.
firm conclusion can
be drawn with regard
to comparative value-
for-money
conclusions.
However, some
promising
Pharmacoeconomics
interventions were
identified. Further
research with
improved
methodology is
required”
Elimination disorders
Ankjaer-Jensen DNK 1994 Elimination disorders Desmopressin Desmopressin versus Cost- “Treatment with a
and Sejr (1994) Desmopressin and effectiveness buzzer alarm or a
buzzer alarm combined treatment is
therefore from a
health economic point
of view preferable.
The health economic
consequences of the
introduction of new
treatments are
discussed, and it is
recommended that
health economic
analyses are
performed before the
introduction of new
treatments”
(continued)
157
158
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Sleep wake disorders
Nishimura and JPN 2018 Sleep wake disorders Suvorexant, zolpidem Suvorexant versus Decision tree Cost- “Suvorexant seemed
Nakao (2018) (insomnia) zolpidem model effectiveness to be more cost-
effective than the
alternative zolpidem.
The findings
suggested that
suvorexant might be a
viable alternative to
zolpidem for elderly
patients with
insomnia. A
sensitivity analysis
showed that outcome
varied depending on
the relative risk for
hip fractures
associated with
suvorexant. Further
investigations may be
needed for more
precise results”
Bolin et al. SWE 2017 Sleep wake disorders Sodium oxybate Sodium oxybate versus Markov model Cost-utility “The estimated cost
(2017) (narcolepsy) standard treatment and cost- per additional QALY
effectiveness for the sodium
oxybate treatment
alternative compared
with standard
treatment was
estimated above the
informal Swedish
willingness-to-pay
R. Dodel et al.
threshold (SEK
500,000). The
estimated cost per
additional QALY
obtained here is likely
to overestimate the
true cost-
Pharmacoeconomics
effectiveness ratio as
potentially beneficial
effects on
productivity of
treatment with
sodium oxybate were
not included (due to
lack of data)”
Tannenbaum CAN 2015 Sleep wake disorders Sedative hypnotics sedative-hypnotics Decision tree Cost- “Failure to consider
et al. (2015) versus CBT versus no model effectiveness drug harms such as
treatment drug-induced falls
and hospitalization
represents a growing
public health concern,
significantly
underestimating the
cost of sedative-
hypnotic therapy and
loss in quality of life
for the elderly. Public
payers should
reconsider
reimbursement of
sedative-hypnotic
drugs as first-line
treatment for
insomnia in older
adults”
(continued)
159
160
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Perraudin et al. FRA 2013 Sleep wake disorders Screening strategy with Markov model Cost- “CP involvement in
(2013) (sleep apnea) community pharmacist effectiveness OSAS (obstructive
and general practitioner sleep apnea
versus screening with syndrome) screening
general practitioner is a cost-effective
alone versus no strategy. This
screening proposal is consistent
with the trend in
Europe and the
United States to
extend the practices
and responsibilities of
the pharmacist in
primary care”
Snedecor et al. USA 2009 Sleep wake disorders Eszopiclone Eszopiclone versus Double-blind, Cost- “Our model, based on
(2009) (insomnia) placebo Placebo effectiveness efficacy data from a
controlled clinical trial,
clinical trial demonstrated
eszopiclone was cost-
effective for the
treatment of primary
insomnia in adults,
especially when lost
productivity costs
were included”
Morin and USA 2009 Sleep wake disorders Eszopiclone Eszopiclone versus Review Cost- “Eszopiclone has
Willett (2009) (insomnia) other treatments effectiveness been shown to be an
efficacious and cost-
effective option for
the treatment of
transient and chronic
insomnia in adults”
R. Dodel et al.
Lees et al. (2008) GBR 2008 Sleep wake disorders Pramipexole, Ropinirole Pramipexole versus no Markov model Cost- “Pramipexole is cost-
(restless legs) treatment versus effectiveness effective compared to
ropinirole no treatment and
ropinirole for patients
with moderate to very
severe RLS”
Hair et al. (2008) NZL 2008 Sleep wake disorders Eszopiclone Depending on reviewed Review Cost-utility “Well designed,
Pharmacoeconomics
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Matsuo and JPN 2007 Sleep wake disorders Midazolam, Flunitrazepam Midazolam versus Retrospective Cost- “Intravenous
Morita (2007) (insomnia) Flunitrazepam multicenter study effectiveness midazolam and
flunitrazepam
appeared to be almost
equal about efficacy
and safety for primary
insomnia, but
flunitrazepam is less
expensive and shows
lower risk of tolerance
development. A
future prospective
comparison study is
necessary”
Balkrishnan et al. USA 2007 sleep wake disorders Estazolam, Flurazepam, Authorization Economic model “This model showed
(2007) (insomnia) Quazepam, Temazepam, that requiring prior
Triazolam, Zolpidem, authorization for
Zaleplon, Ramelteaon, newer sleep
Eszopiclone treatments might not
be a cost-saving
strategy for managed-
care organizations”
Dundar et al. GBR 2004 Sleep wake disorders Zaleplon, zolpidem, zopiclon, Z-drugs versus Review Cost- “The short-acting
(2004) (insomnia) benzodiazepines benzodiazepines effectiveness drugs seem equally
effective and safe
with minor
differences that may
lead a prescriber to
favor one over
another in different
R. Dodel et al.
patients. There is no
evidence that one is
more cost-effective
than any other.
Analysis of the
additional costs to the
NHS, depending on
Pharmacoeconomics
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
conclusions. We
would also
recommend that any
such trial should
include a placebo
arm. It should also
collect good-quality
data around sleep
outcomes and in
particular quality of
life and daytime
drowsiness. We do
not believe that any
formal study of risk of
dependency is
feasible at present.
Finally, the
management of long-
term insomnia is
suggested for further
investigation:
considering the
frequency of this
symptom and its
recurring course, the
short-term trial of
medication and lack
of long-term follow-
up undermine
attempts to develop
R. Dodel et al.
evidence-based
guidelines for the use
of hypnotics in this
condition, or indeed
for its whole
management”
Sexual dysfunctions
Pharmacoeconomics
Martin et al. USA 2013 Sexual dysfunctions Sildenafil citrate Sildenafil versus other Review cost, cost- “The relative value of
(2013) active-treatment options effectiveness, sildenafil
cost of illness, vs. surgically
cost implanted prosthetic
consequence, devices and other
resource use, PDE5 inhibitors, is
productivity, underscored by
work loss, and patients’ WTP
willingness to (willingness to pay),
pay and cost-effectiveness
in ED patients with
comorbidities”
Aspinall et al. USA 2011 Sexual dysfunctions Vardenafil Vardenafil versus no Markov model Cost- “Although four doses
(2011) treatment and vardenafil effectiveness per month of
versus vardenafil in vardenafil was the
varying doses most cost-effective
strategy, the use of six
or eight doses per
month also compares
favorably with other
accepted medical
treatments. The
results were stable
across a range of
inputs and help to
support the current
(continued)
165
Table 4 (continued)
166
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Veterans Affairs
policy on the number
of vardenafil doses
provided per month
for erectile
dysfunction”
Stolk et al. (2000) NLD 2000 Sexual dysfunctions Sildenafil, papaverine- Sildenafil versus Trial Cost-utility, “Treatment with
phentolamine injections papaverine- cost-benefit, sildenafil is cost
phentolamine injections cost- effective. When
effectiveness considering funding
sildenafil, healthcare
systems should take
into account that the
frequency of use
affects cost
effectiveness”
Whittington and 1994 Sexual dysfunctions Oral conjugated estrogens, Overview Cost-utility, “The cost benefit and
Faulds (1994) oral ethinyl estradiol, cost-benefit, cost effectiveness of
transdermal estradiol cost- HRT in the treatment
effectiveness of menopausal
symptoms have not
been fully researched,
although preliminary
results suggest that
conjugated estrogens
and transdermal
estradiol compare
well with alternative
therapies such as
veralipride and
Chinese medicines.
A Swedish study
R. Dodel et al.
using a prevalence-
based approach
estimated that estriol
treatment in all
women with urinary
incontinence aged
greater than or equal
Pharmacoeconomics
to 65 years resulted in
monetary savings
compared with
treating 20% of
women. Cost-utility
data indicated that the
change in quality-
adjusted life years
(QALYs) with HRT
was always positive,
but the degree of
change was
determined by the
baseline assumptions.
Estimated changes in
QALYs with HRT
ranged from 0.006 for
5 years of treatment
with unopposed
estrogen in women
with intact uteri, to
0.5 for 10 years of the
same treatment in
women with severe
menopausal
symptoms following
hysterectomy.
Compliance with
(continued)
167
Table 4 (continued)
168
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
HRT is suboptimal as
5 to 50% of women
withdraw from
therapy, thereby
increasing costs per
year of life saved”
Gender dysphoria
Marfori et al. USA 2018 Gender dysphoria Review “This review
(2018) summarizes the
available literature on
surgical techniques in
addition to reporting
our institutional
outcomes using a
novel 2-port
laparoscopic
approach. Additional
preoperative and
perioperative
considerations are
needed when caring
for this patient
population and are
reviewed”
Disruptive, impulse-control, and conduct disorders
McConaghy and GBR 1988 Disruptive, impulse- Case report “Expectancy of
Blaszczynski control, and conduct improvement did not
(1988) disorders appear to play a major
role in their response,
but it appears
impossible to
disprove that
R. Dodel et al.
expectancy
determines the
response to this or any
form of
psychotherapy.
Whether or not
imaginal
Pharmacoeconomics
desensitization acted
specifically in the
present study, in view
of its cost-efficacy it
is suggested it is
worthy of trial in
impulse disorders
which have persisted
despite treatment”
Substance-related and addictive disorders
Sluiter et al. NLD 2018 Substance-related Naltrexone, acamprosate Naltrexone (G-allele Markov model Cost- “In conclusion,
(2018) and addictive carriers)/acamprosate or effectiveness pharmacogenetic
disorders (alcohol) naltrexone treatment allocation
(AA Homozygotes) of AUD (Alcohol use
versus standard care disorders) patients to
(random treatment of naltrexone, based on
acamprosate or OPRM1 (μ1 opioid
natrexone) receptor) genotype,
can be a cost-effective
strategy, and could
have potential
individual and
societal benefits.
However, more
evidence on the
impact of genotype-
guided treatment
allocation on relapse
169
(continued)
170
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
is needed to
substantiate these
conclusions, as there
is contradictory
evidence about the
effectiveness of
OPRM1 genotyping”
Bansback et al. CAN 2018 Substance-related Diacetylmorphine, injectable Diacetylmorphine RCT Cost- “In patients with
(2018) and addictive hydromorphone, methadone versus injectable effectiveness severe opioid use
disorders (opioids) hydromorphone (and disorder enrolled into
indirect versus the SALOME trial,
methadone) injectable
hydromorphone
provided similar
outcomes to
injectable
diacetylmorphine.
Modelling outcomes
during a patient’s life-
time suggested that
injectable
hydromorphone
might provide greater
benefit than
methadone alone and
may be cost-saving,
with drug costs being
offset by costs saved
from reduced
involvement in
criminal activity”
R. Dodel et al.
Krebs et al. CAN/ 2018 Substance-related Opioid agonist treatment Opioid agonist Semi-Markov Cost- “The value of
(2018) USA and addictive (OAT) treatment (OAT) versus model effectiveness publicly funded
disorders (opioids) standard of care treatment of opioid
use disorder in
California is
maximized when
OAT is delivered to
all patients presenting
Pharmacoeconomics
for treatment,
providing greater
health benefits and
cost savings than the
observed standard of
care”
Carter et al. USA 2017 Substance-related Subdermal implantable Subdermal implantable Markov model Cost- “BSI was preferred
(2017) and addictive buprenorphine (BSI), buprenorphine (BSI) effectiveness over SL-BPN from a
disorders (opioids) sublingual buprenorphine versus sublingual health-economic
(SL-BPN) buprenorphine perspective for
(SL-BPN) treatment of OUD in
clinically-stable
adults. These findings
should be interpreted
carefully, due to some
relationships having
been modeled from
inputs derived from
multiple sources, and
would benefit from
comparison with
outcomes from
studies that employ
administrative claims
data or a naturalistic
comparative design”
(continued)
171
172
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Kenworthy et al. GBR 2017 Substance-related Buprenorphine maintenance Buprenorphine Decision tree Cost- “OST can be
(2017) and addictive treatment (BMT), methadone maintenance treatment model effectiveness considered cost-
disorders (opioids) maintenance treatment (BMT) and methadone effective vs no OST
(MMT) maintenance treatment from the UK
(MMT) versus no NHS/PSS
opioid substitution perspective, with a
therapy (OST) cost per QALY well
below the UK’s
willingness-to-pay
threshold. There were
only small differences
between BMT and
MMT. The
availability of two or
more cost-effective
options is beneficial
to retaining patients in
OST programs. From
a societal perspective,
OST is estimated to
save over £14,032
and £17,174 per year
for BMT and MMT vs
no OST, respectively,
due to savings in
victim costs. Further
work is required to
fully quantify the
clinical and health
R. Dodel et al.
economic impacts of
different OST
formulations and their
societal impact over
the long-term”
Dunlop et al. AUS 2017 Substance-related Buprenorphine-naloxone Buprenorphine- RCT Cost- “When compared to
(2017) and addictive naloxone versus no effectiveness remaining on a
Pharmacoeconomics
(continued)
174
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
assured right to health
provide a formidable
implementation
framework for
updating the Russian
addiction treatment
and rehabilitation
standards to include
international,
evidence-based
addiction treatment
with opioid agonist
medication. This will
reduce the burden of
disease related to
substance use and
associated conditions
such as HIV and
TB. By reconsidering
its position on opioid
agonist therapy,
Russia could create
substantial progress in
the country’s fight
against the burden of,
OUD, HIV and other
drug-related diseases,
as well as the human
tragedies behind
them”
R. Dodel et al.
King et al. (2016) USA 2016 Substance-related Buprenorphine, Methadone Buprenorphine versus Markov model Cost- “The authors
and addictive Methadone effectiveness conclude that MMT is
disorders (opioids) cost-effective
compared with BMT
for the treatment of
patients with opioid
dependence.
However, the
Pharmacoeconomics
treatment of
substance abuse is
complex, and
decision makers
should also consider
individual patient
characteristics when
making coverage
decisions”
Laramee et al. GBR 2016 Substance-related Nalmefene Nalmefene plus Markov model Cost- “Nalmefene
(2016) and addictive psychosocial support effectiveness represents a highly
disorders (alcohol) versus psychosocial cost-effective
support alone treatment option in
this population. The
analysis shows that
integrating nalmefene
within the current UK
clinical treatment
pathway for alcohol
dependence could
reduce the economic
burden on the NHS by
limiting harmful
events and disease
progression”
(continued)
175
176
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Thursz et al. GBR 2015 Substance-related Prednisolone and Placebo/placebo versus RCT Cost- “We conclude that
(2015) and addictive pentoxifylline (PTX) placebo/prednisolone effectiveness prednisolone reduces
disorders (alcohol) versus PTX/placebo the risk of mortality at
versus 28 days, but this
PTX/prednisolone benefit is not
sustained beyond
28 days. PTX had no
impact on survival.
Future research
should focus on
interventions to
promote abstinence
and on treatments that
suppress the hepatic
inflammation without
increasing
susceptibility to
infection”
Stevenson et al. GBR 2015 Substance-related Nalmefene Depending on trial Review Cost- “The clinical
(2015) and addictive effectiveness evidence provided in
disorders (alcohol) the company’s
submission, which
was based on the
results of three pivotal
trials, confirmed the
efficacy and safety of
treatment with
nalmefene together
with psychosocial
support. However,
there were a number
R. Dodel et al.
of limitations and
uncertainties in the
evidence base which
warrant caution in the
interpretation of the
data. In particular, the
inference of treatment
Pharmacoeconomics
effects may be
confounded by the
high drop-out rates
and the use of post
hoc subgroup
analyses to define the
licensed population”
Jackson et al. USA 2015 Substance-related Injectable extended-release Injectable extended- Markov model Cost- “XR-NTX is a cost-
(2015) and addictive naltrexone (XR-NTX), release naltrexone effectiveness effective medication
disorders (opioids) methadone (XR-NTX) versus for treating opioid
methadone dependence if state
addiction treatment
payers are willing to
pay at least $72 per
opioid-free day”
Laramee et al. GBR 2014 Substance-related Nalmefene Nalmefene plus Markov model Cost- “Nalmefene can be
(2014) and addictive psychosocial support effectiveness seen as a cost-
disorders (alcohol) versus psychosocial effective treatment for
support alone alcohol dependence,
with substantial
public health
benefits”
(continued)
177
178
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Reddy et al. IND 2014 Substance-related Baclofen, chlordiazepxide, Baclofen versus RCT Cost- “Both study drugs
(2014) and addictive lorazepam chlordiazepoxide effectiveness provided relief of
disorders (alcohol) withdrawal
symptoms.
Chlordiazepoxide
was more cost-
effective than
baclofen. Baclofen
was relatively less
effective and more
expensive than
chlordiazepoxide”
Schwartz et al. USA 2014 Substance-related Methadone Interim versus standard RCT Cost-benefit “The net benefits of
(2014) and addictive methadone treatment were greater
disorders (opioids) for the IM (interim
methadone) condition
but controlling for the
baseline variables
noted above, the
difference between
conditions in net
monetary benefits was
not significant. For
the combined sample,
there was a pre- to
posttreatment net
benefit of $1470 (95%
CI: -$625; $3584) and
a benefit-cost ratio of
1.5 (95% CI: 0.8, 2.3),
but using our
R. Dodel et al.
conservative
approach to
calculating benefits,
these values were not
significant”
Ruger et al. USA 2012 Substance-related Buprenorphine, naltrexone Buprenorphine versus RCT Cost- “Buprenorphine
(2012) and addictive naltrexone versus effectiveness appears to be a cost-
Pharmacoeconomics
(continued)
Table 4 (continued)
180
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
AD. Most of the
available treatment
options for AD appear
to produce marked
economic benefits”
Saul et al. (2011) USA 2011 Substance-related Nicotine replacement Observational Cost- “Results of this
and addictive therapy with patches study effectiveness evaluation indicate
disorders (tobacco) versus nictotine that while satisfaction
replacement therapy rates increase among
with gum those receiving more
counseling and NRT,
quit rates do not, even
when controlling for
demographic and
tobacco use
characteristics”
Keating and NZL 2010 Substance-related Varenicline Depending on reviewed Review Cost- “Indeed, despite their
Lyseng- and addictive study effectiveness limitations, available
Williamson disorders (tobacco) pharmacoeconomic
(2010) analyses from
numerous countries
support the use of
varenicline for 12 or
24 weeks as a cost-
effective treatment
relative to other
smoking cessation
therapies in smokers
who wish to quit
smoking. For
example, in modelled
cost-effectiveness
R. Dodel et al.
analyses conducted
from a healthcare
payer perspective,
12 weeks’ treatment
with varenicline
consistently
dominated bupropion
Pharmacoeconomics
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Adi et al. (2007) GBR 2007 Substance-related Naltrexone Naltrexone versus Review Cost- “Following
and addictive placebo versus other effectiveness successful withdrawal
disorders (opioids) pharmacological from opioids,
treatments versus naltrexone may be
psychosocial administered on a
interventions versus no chronic basis to block
treatment any future effects of
opioids. Naltrexone
appears to have some
limited benefit in
helping formerly
opioid-dependent
individuals to remain
abstinent, although
the quality of the
evidence is relatively
poor and
heterogeneous. The
limited quality and
extent of the studies
precluded an analysis
of subgroups likely to
benefit from
naltrexone
prescribing. Oral
naltrexone is used
infrequently in
current UK practice,
and this review
suggests that this is
R. Dodel et al.
appropriate as there is
little evidence to
support its wider
implementation.
There is an important
deficit in information
about the quality of
Pharmacoeconomics
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Neurocognitive disorders
da Silva et al. BRA 2019 Neurocognitive Donepezil, Rivastigme Donepezil versus Markov model Cost- “The findings of this
(2019) disorders (dementia) Rivastigme effectiveness, study contradict the
cost-benefit, standard of care for
cost-utility mild and moderate
AD in Brazil, which is
based on
rivastigmine. A
pharmacological
treatment option
based on current
Brazilian clinical
practice guidelines for
AD suggests that
rivastigmine is less
cost-effective (0.39
QALY/BRL
32,685.77) than
donepezil.
Probabilistic analysis
indicates that
donepezil is the most
cost-effective
treatment for mild and
moderate AD”
Ebrahem and CAN 2018 Neurocognitive Cholinesterase inhibitors, Depending on reviewed Review Cost- “The literature
Oremus (2018) disorders (dementia) memantine study effectiveness suggested AD
medications generally
dominated
comparator
treatments (e.g.,
R. Dodel et al.
placebo). Expert
opinion: The authors
noted several
limitations of the
included economic
evaluations. These
limitations suggest
Pharmacoeconomics
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
were sponsored by
industry and the
results tended to
overwhelmingly
support the
manufacturer’s
product”
Carrasco et al. ESP 2016 Neurocognitive Dexmedetomidine, Dexmedetomidine RCT Cost- “In the study
(2016) disorders (delirium) haloperidol versus haloperidol effectiveness, conditions,
cost-benefit dexmedetomidine
shows to be useful as
a rescue drug for
treating agitation due
to delirium in
nonintubated patients
in whom haloperidol
has failed, and it
seems to have a better
effectiveness, safety,
and cost-benefit
profile than does
haloperidol”
Djalalov et al. CAN 2012 Neurocognitive Donepezil, Rivastigme Donepezil versus and Markov model Cost- “Using presently
(2012) disorders (amnesia) genetic screening of effectiveness available clinical
apolipoprotein versus evidence, this
standard of care exploratory study
illustrates that genetic
testing combined with
preventive donepezil
treatment for AMCI
(amnestic mild
cognitive impairment)
R. Dodel et al.
patients may be
economically
attractive. Since our
results were based on
a secondary post hoc
analysis, our study
alone is insufficient to
Pharmacoeconomics
warrant
recommending APOE
genotyping in AMCI
patients. Future
research on the
effectiveness of
preventive donepezil
as a targeted therapy
is recommended”
Personality disorders
Crawford et al. GBR 2018 Personality disorders Lamotrigine Lamotrigine versus RCT Cost- “The addition of
(2018) (borderline) placebo effectiveness lamotrigine to the
usual care of people
with BPD was not
found to be clinically
effective or provide a
cost-effective use of
resources”
Headache
Bellingham and CAN 2010 Headache Duloxetine Duloxetine versus other Review Cost- “Randomized trials
Peng (2010) antidepressants effectiveness have documented
among others significant analgesic
effects for managing
chronic pain
associated with
fibromyalgia and
diabetic peripheral
neuropathic pain.
187
(continued)
188
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Studies have also
suggested that pain
associated with major
depressive disorder
can be reduced with
this medication.
Modest effects for
headache,
osteoarthritic pain,
and pain secondary to
Parkinson disease
have also been
documented, but data
are obtained from
single-blinded or
open-label trials that
require further
corroboration with
larger randomized
studies. Duloxetine
has not yet been
directly compared
with other
antidepressants or
anti-convulsants for
the treatment of pain
syndromes”
Gracia Naya ESP 2001 Headache Tritapanes (sumatriptan, Sumatriptan versus Cost- “In chronic, recurrent
(2001) naratriptan, zolmitriptan, naratriptan versus effectiveness disorders such as
rizatriptan, almotriptan) zolmitriptan versus migraine, when the
rizatriptan versus cost of treatment
almotriptan using triptans is
R. Dodel et al.
assessed, one has not
only to evaluate the
cost per unit but also
to introduce other
parameters such as the
efficacy of the drug,
which is one of the
Pharmacoeconomics
most important”
Epilepsy
Choi and Mohit USA 2019 Epilepsy Carbamazepine Screening for Markov model Cost- “Our analysis
(2019) HLA-B*1502 versus no effectiveness confirms the 2007 US
screening Food and Drug
Administration
recommendation to
screen for
HLA-B*1502 allele
before starting
treatment with
carbamazepine in
patients of Asian
ancestry in the United
States”
Geitona et al. GRC 2019 Epilepsy Lacosamide, zonisamide Lacosamide versus Discrete event Cost- “Lacosamide is a
(2019) zonisamide simulation model effectiveness cost-effective option
at a willingness-to-
pay threshold of
€30,000 per QALY,
representing a
valuable
monotherapy
treatment option for
patients with focal
epileptic seizures in
the Greek setting”
(continued)
189
190
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Mkrtchyan and RUS 2019 Epilepsy Carbamazepine (CBZ), Carbamazepine (CBZ) “CBZ, OXZ and
Kaimovsky oxcarbazepine (OXZ), versus oxcarbazepine LCM can be used as
(2019) lamotrigine (LTG), (OXZ) versus monotherapy in
lacosamide (LCM) lamotrigine (LTG) patients with newly
versus lacosamide diagnosed
(LCM) FE. Monotherapy
with LCM is
associated with
lowest costs for
stopping focal
seizures and the
highest percentage of
seizure-free patients.
CBZ remains the
most economical
AED: the cost-
effectiveness is
minimal compared to
all other AEDs used
in the study. However,
monotherapy with
LCM is advisable if it
is necessary to
minimize side-effects
in patients with newly
diagnosed FE”
Tremblay et al. ESP 2018 Epilepsy Perampanel, other Perampanel versus Markov model Cost- “Our study
(2018) antiepileptic drugs other antiepileptic drugs effectiveness demonstrates that
perampanel is likely
to be a cost-effective
option”
R. Dodel et al.
Elliott et al. CAN 2018 Epilepsy (Dravet Stiripentol, clobazam, Stiripentol versus Markov model Cost- “From the perspective
(2018) Syndrome) valproate clobazam versus effectiveness of the Canadian
valproate public healthcare
payer, stiripentol is
not cost effective at its
current price at a
willingness-to-pay
threshold of
Pharmacoeconomics
$Can50,000. Funding
stiripentol will be
associated with
important opportunity
costs that bear
consideration”
Picot et al. (2016) FRA 2016 Epilepsy Medical group versus Markov model Cost- “Our study suggests
surgery group effectiveness that in addition to
being safe and
effective, respective
surgery of epilepsy is
cost-effective in the
medium term. It
should therefore be
considered earlier in
the development of
epilepsy”
Jacoby et al. GBR 2015 Epilepsy Carbamazepine, Lamotrigine, Carbamazepine versus RCT QoL “The choice of initial
(2015) gabapentin, oxcarbazepine, lamotrigine, treatment had no
topiramate, valproate gabapentin, significant effect on
oxcarbazepine and QoL by 2-year
topiramate, and follow-up. However,
valproate versus overall QoL was
lamotrigine and reduced with
topiramate continued seizures,
adverse events, and
failure of the initial
treatment”
191
(continued)
192
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Darba et al. ESP 2014 Epilepsy Pregabalin Depending on reviewed Review Cost- “The majority of
(2014) study effectiveness published evidence
supports the
possibility that
pregabalin could be a
cost-effective and/or
cost-saving
alternative for the
treatment of
refractory epilepsy,
GAD, and
neuropathic pain, in
both treatment-naïve
patients and in those
who have
demonstrated
inadequate response
or intolerance to
previous therapy”
Clements et al. USA 2013 Epilepsy Lamotrigine, rufinamide, Lamotrigine versus Trial-based Cost- “Over a 3-month
(2013) topiramate rufinamide versus economic model effectiveness horizon, clobazam
topiramate was more effective
and less expensive
than comparators,
with the assumption
that >0.77% of drop
seizures required
medical care. Below
this threshold,
topiramate was less
costly than clobazam.
With the base-case
R. Dodel et al.
assumption that 2.3%
of drop seizures were
medically attended,
costs for patients
receiving clobazam
totaled $30,147
versus $34,223–$
Pharmacoeconomics
35,378 for
comparators.
Clobazam was more
efficacious and less
costly than rufinamide
over a 2-year horizon.
The percentage of
medically attended
drop seizures was a
driver of results.
Clobazam treatment
may be cost-saving”
Rattanavipapong THA 2013 Epilepsy Carbamazepine Carbamazepine with Decision tree and Cost- “Universal
et al. (2013) screening versus no Markov model effectiveness HLA-B*15:02
screening screening represents
good value for the
money in terms of
preventing SJS/TEN
(Stevens-Johnson
syndrome; toxic
epidermal necrolysis)
in CBZ-treated
patients with
neuropathic pain at
the Thai ceiling ratio
of 120,000
THB/QALY gained.
However, the
prevalence of
193
(continued)
194
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
CBZ-induced
SJS/TEN in the Thai
population and the
positive predictive
value (PPV) are major
factors that influence
the cost-effectiveness
of HLA-B*15:02
screening. Therefore,
an active surveillance
system to make a
more accurate
assessment of the
prevalence
CBZ-induced
SJS/TEN in the Thai
population would
enhance the
generalizability of the
results”
Lee et al. (2013) GBR 2013 Epilepsy Buccolam Buccolam versus Decision tree Cost- “This model
diazepam and buccal effectiveness demonstrates the
midazolam possibility of
constructing a
thorough economic
case when trial or
real-world data are
not available. The
results of the model
show Buccolam to be
cost saving compared
with rectal diazepam
R. Dodel et al.
due to a reduction in
the need for
ambulance callouts
and hospital stays,
and compared with
unlicensed buccal
midazolam, through
Pharmacoeconomics
(continued)
196
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
test positive is cost-
effective for
Singaporean Chinese
and Malays, but not
for Singaporean
Indians. Population
frequency of
HLA-B*1502, PPV,
duration of treatment
relative to life
expectancy, and costs
of alternative drugs
are the key drivers
influencing cost-
effectiveness”
Bolin and SWE 2012 Epilepsy Antiepileptic drugs Depending on reviewed Review Cost- “Although failure to
Forsgren (2012) study effectiveness meet good practice
guidelines influences
the reliability of the
presented evidence
adversely, a sufficient
number of the
included studies were
found to comply
enough with the
guidelines in order for
the qualitative content
of the cost-
effectiveness results -
that some of the
newer AEDs are cost
R. Dodel et al.
effective - to be
reliable. In fact, this
conclusion is likely to
be relatively robust,
since the effect of
improved seizure
control on labor
Pharmacoeconomics
market performance
was not included in
the base-case results
in any of the included
studies and improved
seizure control need
only to have a
moderate effect on
sickness absenteeism
in order for the
corresponding
treatment to be cost
effective even when
willingness to pay for
an additional QALY is
low. However, the
cost effectiveness of
newer AEDs has only
been studied for a
small number of
settings, and hence
future studies
incorporating
additional settings are
needed”
(continued)
197
Table 4 (continued)
198
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Stroke
Bowrin et al. FRA 2020 Stroke Rivaroxaban, vitamin K Rivaroxaban versus Markov model Cost- “Although there is no
(2020) antagonists vitamin K antagonists effectiveness official willingness-
to-pay threshold in
France, these results
suggest that
rivaroxaban is likely
to be cost-effective
compared to VKA in
French patients with
AF from a national
insurance
perspective”
de Jong et al. NLD 2019 Stroke Vitamin K antagonists NOAC apixaban versus RCT Cost- “Based on RCTs
(2019) (VKAs)-non-VKA oral other NOACs effectiveness (randomized
anticoagulants (NOACs) (dabigatran, edoxaban, controlled trials) as
and rivaroxaban) and well as RWD (real-
VKA world data), we
conclude that
apixaban is generally
cost-effective or even
cost-saving (less
costly and more
effective) compared
to VKA and other
NOACs in the overall
population of patients
with atrial fibrillation”
Yagudina et al. RUS 2019 Stroke Ethylmethylhydroxypyridine Ethylmethylhydroxy- Cost- “Mexidol has the
(2019) succinate (mexidol), inosine, pyridine succinate effectiveness same efficacy as
nicotinamide, riboflavin, (mexidol), inosine + alternatives. However
succinic acid (cytoflavin) nicotinamide + mexidol is superior to
R. Dodel et al.
riboflavin + succinic cytoflavin and
acid (cytoflavin), and a actovegin in terms of
deproteinized cost minimization
hemoderivate of the analysis. The savings
blood of calves from one course of
(actovegin) alternatives will cover
costs of treatment of
Pharmacoeconomics
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Better identification
of patients not
benefiting from IVT
would optimize the
selective use of EVT
thereby improving its
effectiveness”
Kabore et al. FRA 2019 Stroke Mechanical thrombectomy IV-tPA versus MT-IV- Markov model Cost- “Although there is no
(2019) (MT) plus intravenous tissue- tPA effectiveness universally accepted
type plasminogen activator willingness-to-pay
(IV-tPA) (MT-IV-tPA) threshold in France,
our analysis suggests
that MT combined to
IV-tPA can be
considered a cost-
effective treatment
compared with
IV-tPA alone”
Reeves et al. AUS 2018 Stroke Reperfusion therapy with Multimodal computed Cost-utility “In a healthcare
(2018) alteplase tomography imaging setting where
(MMCT) versus multimodal imaging
reperfusion therapy technologies are
with alteplase available and
reimbursed, their use
in screening patients
presenting with acute
stroke to determine
eligibility for
alteplase treatment is
cost-effective given a
range of willingness-
to-pay thresholds and
R. Dodel et al.
warrants
consideration as an
alternative to routine
practice”
Pickett et al. USA 2018 Stroke Standard medical therapy Foramen ovale closure RCT Cost- “In comparison to
(2018) versus medical therapy effectiveness medical therapy
alone, PFO closure
Pharmacoeconomics
appears to be cost-
effective and
clinically efficacious”
Ruggeri et al. ITA 2018 Stroke Intravenous tissue MT and intravenous Markov model Cost- “MT plus IV t-PA for
(2018) plasminogen activation tissue plasminogen effectiveness AIS patients with
activation (MT plus IV LVO is cost-effective
t-PA) versus IV t-PA from year 1 through
year 3, and cost-
saving from year
4 onward in the Italian
context, achieving
better results, both in
terms of efficacy and
in terms of resource
consumption”
Reddy et al. USA 2018 Stroke Warfarin, non-vitamin K Left atrial appendage Markov model Cost- “Upfront procedure
(2018) antagonist oral anticoagulants closure (LAAC) effectiveness costs initially make
dabigatran, apixaban, compared with warfarin LAAC higher cost
rivaroxaban and the non-vitamin K than warfarin and the
antagonist oral non-vitamin K
anticoagulants antagonist oral
dabigatran 150 mg, anticoagulants, but
apixaban and within 10 years,
rivaroxaban LAAC delivers more
quality-adjusted life
years and has lower
total costs, making
LAAC the most cost-
201
(continued)
Table 4 (continued)
202
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
effective treatment
strategy for secondary
prevention of stroke
in atrial fibrillation”
Leppert et al. USA 2018 Stroke Percutaneous transcatheter Percutaneous Markov model Cost- “PFO closure for
(2018) closure of patent foramen transcatheter closure of effectiveness cryptogenic strokes in
ovale (PFO closure), patent foramen ovale the right setting is
antiplatelet therapy (PFO closure) plus cost-effective,
antiplatelet therapy producing benefit in
versus antiplatelet QALYs gained and
therapy potential cost savings.
However, patient
selection remains
vitally important as
marginal declines in
treatment
effectiveness can
dramatically affect
cost-effectiveness”
Roffe et al., GBR 2018 Stroke Nocturnal oxygen Low-dose oxygen vs no RCT Cost- “Routine low-dose
2018) oxygen effectiveness oxygen
supplementation in
stroke patients who
are not severely
hypoxic is safe, but
does not improve
outcome after stroke”
Joo et al. (2017) USA 2017 Stroke Intravenous recombinant Intravenous Markov model Cost- “IV rtPA saved costs
tissue plasminogen activator recombinant tissue effectiveness and improved health
plasminogen activator outcomes for patients
vs no treatment aged 18–64 years and
was cost effective for
R. Dodel et al.
those aged 65 years.
These findings
support the use of IV
rtPA”
Amiri et al. IRN 2018 Stroke Tissue plasminogen activator tPA versus no tPA Markov model Cost- “The balance of
(2018) (tPA) effectiveness hospitalization and
rehabilitation costs
Pharmacoeconomics
(continued)
204
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
prevention depends
more on the disutility
associated with pill
burden than their
degree of
cardiovascular risk”
Davies et al. USA 2017 Stroke Ezetimibe, statin Ezetimibe and statin Markov-like Cost- “Compared with
(2017) versus statin model effectiveness statin monotherapy,
ezetimibe with statin
therapy was cost-
effective for
secondary prevention
of CHD and stroke
and for primary
prevention of these
conditions in patients
whose LDL-C levels
are 100 mg/dL and
in patients with
diabetes, taking into
account a 90% cost
reduction for
ezetimibe”
Kunz et al. GER 2016 Stroke Endovascular therapy, Endovascular therapy Markov model Cost- “EVT + SC is cost-
(2016) standard care and standard care effectiveness effective in most
depending on patients subgroups. In patients
NIHSS score with ASPECTS 5 or
with M2 occlusions,
cost-effectiveness
remains uncertain
based on current data”
R. Dodel et al.
Costa et al. PRT 2015 Stroke Non-vitamin K antagonist Apixaban, dabigatran, Markov model Cost- “Apixaban is a cost-
(2015) oral anticoagulants rivaroxaban versus effectiveness effective alternative to
warfarin warfarin and
dabigatran and is
dominant over
rivaroxaban in AF
patients from the
perspective of the
Pharmacoeconomics
Portuguese national
healthcare system.
These conclusions are
based on indirect
comparisons, but
despite this limitation,
the information is
useful for healthcare
decision-makers”
Kamae et al. JPN 2015 Stroke Apixaban, warfarin Apixaban versus Markov model Cost- “Although most
(2015) warfarin effectiveness participants in the
Apixaban for
Reduction in Stroke
and Other
Thromboembolic
Events in Atrial
Fibrillation
(ARISTOTLE) trial
used for the efficacy
data of apixaban in
the model were
non-Japanese
patients, the impact of
the limitations on our
results was
considered small, and
our results were
deemed robust
205
(continued)
206
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
because of the
additional effect in
Japanese patients
compared with that in
the global population
according to the
subanalysis of
Japanese patients in
the trial. Therefore,
based on an
adaptation of a
published Markov
model, apixaban is a
cost-effective
alternative to warfarin
in Japan for stroke
prevention among
patients with NVAF”
Moretti et al. ITA 2015 Stroke Intravenous thrombolysis Depending on reviewed Review Cost- “Twenty years after
(2015) with recombinant tissue studies effectiveness the NINDS trial,
plasminogen activator (rtPA), among others i.v. rtPA therapy
endovascular treatments within 4.5 h from
(EVT), other thrombolytic stroke onset remains
agents, antiplatelet therapy the mainstay of AIS
therapy, although far
from ideal. Various
alternative strategies
and agents have been
extensively
investigated to
increase the
percentage of patients
R. Dodel et al.
treated, improve
outcome and lower
SICH. Promising
findings are emerging
from the extension of
time window beyond
4.5 h, the
Pharmacoeconomics
sonothrombolysis, the
novel thrombolytic
agents (hopefully
with a greater
commitment by the
industry), the
combination of
intravenous and
endovascular
treatments and the use
of mechanical
recanalization
devices, mainly
stents, for selected
patients in highly
specialized stroke
units. In this context,
the following
recommendations by
STAIR (Stroke
Therapy Academic
Industry Roundtable)
remain appropriate:
(i) reducing
unsubstantiated
contraindications to
treatment (based more
on regulatory
(continued)
207
208
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
requirements than
clinical data) such as
mild stroke, patient’s
age, the association of
diabetes and prior
stroke, and
(ii) improving
prehospital and
in-hospital
organization in order
to minimize delay of
treatment. There is
great expectation for
increasing the
numbers of treated
patients, and
achieving a better
quality of therapies
(in terms of efficacy,
lower adverse effects
and reduced long-
term disabilities). Not
only are these goals
crucial in order to
meet medical need,
but could also have a
substantial impact on
global costs”
R. Dodel et al.
Leppert et al. USA 2015 Stroke Intravenous tissue-type Intra-arterial treatment Decision- Cost- “Intra-arterial
(2015) plasminogen activator, intra- and intravenous tissue- analytic model effectiveness treatment after
arterial treatment type plasminogen intravenous tissue-
activator versus type plasminogen
intravenous tissue-type activator for patients
plasminogen with anterior
circulation strokes
within the 6-hour
Pharmacoeconomics
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Holmes et al. GBR 2015 Stroke Alteplase Alteplase versus Review Cost- “The incremental
(2015) standard medical and effectiveness cost-effectiveness
supportive management ratios (ICER) for all
treatment windows
were well below
accepted willingness
to pay thresholds. The
ERG had no major
concerns regarding
the completeness of
the submission or the
robustness of the
evidence presented.
For all treatment
windows considered,
alteplase was found to
be cost-effective
compared with
standard treatment”
Pan et al. (2014a) CHN 2014 Stroke Intravenous tissue-type Intravenous tissue-type Markov model Cost- “Intravenous tPA
plasminogen activator (tPA plasminogen activator effectiveness treatment within 4.5 h
(tPA) versus non-tPA- is highly cost-
treatment effective for acute
ischemic strokes in
China”
Pan et al. (2014b) CHN 2014 Stroke Clopidogrel, asprin Clopidogrel and aspirin Markov model Cost- “Early 90-day
versus aspirin effectiveness clopidogrel-aspirin
regimen for acute TIA
or minor stroke is
highly cost-effective
in China. Although
clopidogrel is generic,
R. Dodel et al.
Plavix is brand in
China. If Plavix were
generic, treatment
with clopidogrel-
aspirin would have
been cost saving”
Yang et al. (2014) CHN 2014 Stroke Aspirin, clopidogrel Aspirin and clopidogrel Markov model Cost- “To prevent recurrent
Pharmacoeconomics
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Rognoni et al. ITA 2014 Stroke Apixaban, dabigatran and Apixaban, dabigatran Markov model Cost- “Our analysis
(2014) rivaroxaban, warfarin and rivaroxaban versus effectiveness suggests that NOAs
warfarin (new oral
anticoagulants) are a
cost-effective
treatment for the
prevention of stroke
in patients with
NVAF (Nonvalvular
atrial fibrillation) in
the Italian healthcare
setting”
Doan et al. GBR 2013 Stroke OnabotulinumtoxinA Usual care plus Simulation Cost- “Based on a model,
(2013) onabotulinumtoxinA model effectiveness UC plus
versus usual care onabotulinumtoxinA
improved disability,
which translated into
greater QALYs but
also increased direct
medical costs
compared with UC
alone; however, the
resulting ICER can be
considered cost-
effective. Moreover,
UC plus
onabotulinumtoxinA
can be cost-saving if
reduction in caregiver
burden was included.
OnabotulinumtoxinA
R. Dodel et al.
offers value for
money in the
management of
ULPSS in Scotland”
Pan et al. (2012) USA 2012 Stroke Intravenpus thrombolysis, Depending on reviewed Review Cost- “Intravenous
clopidogrel, statin, warfarin, study effectiveness thrombolysis is
dabigatran consistently shown to
Pharmacoeconomics
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
different types of
treatments are
warranted. Also, few
of the published
economic evaluations
considered the
economic impact of
these treatments on
subgroups and
individuals with
different risks”
Meningitis
Tuon et al. (2019) BRA 2019 Meningitis Deoxycholate amphotericin B Deoxycholate Cost- Cost- “Treatment with
(d-AMB), lipid formulations amphotericin B minimization minimization ABLC (amphotericin
(d-AMB) versus lipid model B lipid complex)
amphotericin B would be cost saving
in comparison to
d-AMB treatment, if
early switch of
treatment occurred in
patients presenting
AKI. The change
should be as soon as
possible to avoid
further complication,
like dialysis, which is
associated with a
lower life
expectancy.2”
R. Dodel et al.
Merry and USA 2016 Meningitis Amphotericin B, flucytosine, Amphotericin B Cost- “Flucytosine is
Boulware (2016) fluconazole deoxycholate for effectiveness currently cost-
4 weeks versus effective in the United
amphotericin and States despite a
flucytosine (100 mg/kg/ dramatic increase in
day) for 2 weeks versus price in recent years.
amphotericin and Combination therapy
with amphotericin
Pharmacoeconomics
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
effectiveness and
identify optimal
treatment dosing and
implementation best
practices”
Rajasingham UGA 2012 Meningitis Fluconazole, flucytosine, Fluconazole (800– Decision- Cost- “Short-course (7-d)
et al. (2012) amphotericin 1200 mg/d) analysis effectiveness amphotericin
monotherapy versus induction therapy
fluconazole + coupled with high-
flucytosine (5FC) dose (1200 mg/d)
versus short-course fluconazole is very
amphotericin ‘cost effective’ per
(7-d) + fluconazole World Health
versus 14-d of Organization criteria
amphotericin alone vs and may be a worthy
amphotericin + investment for policy-
fluconazole versus and makers seeking cost-
amphotericin +5FC effective clinical
outcomes. More
head-to-head clinical
trials are needed on
treatments for this
neglected tropical
disease”
Trotter and GBR 2006 Meningitis Meningococcal serogroup C Alternative vaccination Transmission Cost- “Models that do not
Edmunds (2006) conjugate (MCC) vaccination strategies dynamic model effectiveness include the indirect
effects of vaccination
will underestimate the
impact of MCC
vaccination and may
lead to distorted
decision making”
R. Dodel et al.
Multiple sclerosis
Furneri et al. ITA 2019 Multiple sclerosis Natalizumab, Natalizumab versus Markov model Cost- “Adopting the Italian
(2019) immunomodulators switching among effectiveness social perspective,
(interferons/glatiramer immunomodulators, early escalation to
acetate) followed by subsequent natalizumab is
escalation to dominant versus
natalizumab switching among
Pharmacoeconomics
immunomodulators,
in RRMS patients
who do not respond
adequately to
conventional
immunomodulators”
D’Amico et al. ITA 2019 Multiple sclerosis Synthetic therapies, other Depending on reviewed Review Cost- “The emerging and
(2019) disease-modifying treatments study effectiveness more expensive
DMTs for MS
represent a
considerable
challenge for health-
care systems and
resource
consumption. Future
research should focus
on the long-term
efficacy of DMTs and
the cost of treating
MS in a real-life
setting. Future
biological and
radiological
biomarkers could help
stratify patients at
early stages of MS,
helping physicians
design a personalized
therapeutic approach
217
(continued)
Table 4 (continued)
218
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
that could have a
positive impact in
economic terms”
Melendez-Torres GBR 2017 Multiple sclerosis Beta-interferon, glatiramer Depending on reviewed Review Cost- “DMTs were
et al. (2017) acetate study effectiveness clinically effective for
RRMS and CIS but
cost-effective only for
CIS. Both RCT
evidence and RSS
data are at high risk of
bias. Research
priorities include
comparative studies
with longer follow-up
and systematic review
and meta-synthesis of
qualitative studies”
Restless legs syndrome
Reinhold et al. GER 2009 Restless legs Ropinirole, Pramipexole Ropinirole versus Review Cost- “The cost-of-illness
(2009) syndrome Pramipexole effectiveness studies were
heterogeneous but
indicated that RLS
was associated with a
substantial economic
burden, resulting in
high direct and
indirect costs to
society. Although
effective and cost-
effective treatments
appear to be available,
further research is
R. Dodel et al.
warranted, especially
regarding the
economic burden of
RLS and the cost
effectiveness of
available treatment
options”
Pharmacoeconomics
Lees et al. (2008) GBR 2008 Sleep wake disorders Pramipexole, Ropinirole Pramipexole versus no Markov model Cost- “Pramipexole is cost-
(restless legs) treatment versus effectiveness effective compared to
ropinirole no treatment and
ropinirole for patients
with moderate to very
severe RLS”
Parkinson’s disease
Wang and USA 2019 Parkinson’s disease Dopaminergic medications, Depending on reviewed Review Cost-benefit “Overall, the authors
Gunzler (2019) laxatives, blood pressure study analysis; cost- found a scarcity of
raising medications, effectiveness primary PD
antidepressants, antipsychotic analysis; cost- pharmacoeconomic
minimization literature in the
analysis; cost- twenty-first Century.
utility analysis Given the myriad of
PD motor and
nonmotor treatments,
only 24 papers
evaluating motor
treatments and two
papers evaluating
nonmotor treatments
met our search
criteria. More studies
are clearly needed to
better define the
pharmacoeconomics
of PD therapeutics”
(continued)
219
220
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
Amyotrophic lateral sclerosis
Ringel (2002) USA 2002 Amyotrophic lateral Review Cost- “Many clinical trials
sclerosis effectiveness have one or more
shortcomings that
limit the
generalizability of the
study results.4–5 For
example, the duration
of a trial is often short,
so that the value of an
agent over the long
term is seldom
assessed. High drop-
out rates are also
common in ALS
studies, because of
co-morbid conditions,
unwanted side-effects
or ineffectiveness of
the agent. Patients
selected for trials are
often unrepresentative
of the population of
patients in the real
world, since they have
to meet rigid entry
criteria and travel to
research centers. The
design of trials may
have other biases as
well, such as failure to
have a prospective
R. Dodel et al.
comparison (placebo)
group, or inadequate
blinding and
randomization. Since
we are facing
demands to limit
healthcare
Pharmacoeconomics
expenditures, we
must understand the
design and limitations
of studies to assess the
cost-effectiveness of
treatment
recommendations”
Tavakoli and GBR 2001 Amyotrophic lateral Riluzole Riluzole versus best Markov model Cost-utility “Using the Markov
Malek (2001) sclerosis supportive care model and the
transitional
probabilities the base
case cost per life year
gained was estimated
at pound sterlings
14,370 and applying
Standard Gamble
utility scores, the base
case cost per QALY
was assessed as
pound sterlings
20,904. The effect of
discounting costs and
benefits altered the
cost effectiveness
analysis to pound
sterlings 17,760 per
life year gained while
a sensitivity analysis
(continued)
221
222
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
around median or
mean scores for the
utility weight resulted
in a range of pound
sterlings 19,020 to
pound sterlings
25,794 per QALY
gained”
Polyneuropathy
Rajabally and GBR 2019 Polyneuropathy Intravenous immunoglobulin Intravenous Cost-effective “Our results indicate
Afzal (2019) (IVIg), standard dosing immunoglobulin (IVIg) that an individualized
regimes versus standard dosing IVIg treatment
regimes protocol is clinically
noninferior and 10–
25% more cost-
effective than
standard dosing
regimens in CIDP”
Bamrungsawad THA 2016 Polyneuropathy Intravenous immunoglobulin IVIG plus Markov model Cost-effective “At a threshold of
et al. (2016) (IVIG), corticosteroids corticosteroids versus US$4672 per QALY
immunosuppressants gained, IVIG plus
plus corticosteroids in corticosteroids is
steroid-resistant CIDP considered a cost-
patients effective treatment for
steroid-resistant CIDP
patients in Thailand”
R. Dodel et al.
Lazzaro et al. ITA 2014 Polyneuropathy Intravenous immunoglobulin Subcutaneous versus Cost- Cost- “Overall costs per
(2014) (IVIG), subcutaneous intravenous minimization minimization patient amount to
immunoglobulin administration of model €49,534.75 (SCIG;
immunoglobulin subcutaneous
immunoglobulins)
and €50,895.73
(IVIG); saving in
favor of SCIG reaches
Pharmacoeconomics
Table 4 (continued)
Economic
Author Country Year Indication Medication Comparator Study design Evaluation Authors conclusion
McCrone et al. GBR 2003 Polyneuropathy Intravenous immunoglobulin, Intravenous Cost-utility “Using a net-benefit
(2003) prednisolone immunoglobulin versus approach it was
prednisolone shown that the
probability of IVIg
being cost-effective in
comparison with
prednisolone was 0.5
or above (i.e., was
more likely to be cost-
effective than cost-
ineffective) only if
one QALY was
valued at over euro
250,000. The cost-
effectiveness of IVIg
is greatly affected by
the price of IVIg and
the amount
administered. The
impact of later side-
effects of
prednisolone on long-
term costs and quality
of life are likely to
reduce the cost per
QALY of IVIg
treatment”
R. Dodel et al.
Pharmacoeconomics 225
model (including decision trees, Markov models, DES, and sets of mathematical
equations), and that evaluated therapeutic interventions or diagnostic procedures,
following the definition of a “model” as suggested by Weinstein et al. (Weinstein
2006). Additionally, at least two treatment options had to be compared and the
primary aim of the studies had to be a health economic evaluation of these treat-
ments. We excluded studies using models only as an illustration of methodological
aspects or those published as abstracts.
Abbreviations of listed countries follow the code of ISO 3166 alpha-3. Authors
conclusion were taken and adapted from the abstract or the main text of the
respective article and is set in parenthesis.
References
Ademi Z, Pasupathi K, Liew D. Cost-effectiveness of apixaban compared to warfarin in the
management of atrial fibrillation in Australia. Eur J Prev Cardiol. 2015;22:344–53.
Adi Y, Juarez-Garcia A, Wang D, et al. Oral naltrexone as a treatment for relapse prevention in
formerly opioid-dependent drug users: a systematic review and economic evaluation. Health
Technol Assess. 2007;11: iii-iv, 1–85.
Amiri A, Goudarzi R, Amiresmaili M, et al. Cost-effectiveness analysis of tissue plasminogen
activator in acute ischemic stroke in Iran. J Med Econ. 2018;21:282–7.
Ankjaer-Jensen A, Sejr TE. Costs of the treatment of enuresis nocturna. Health economic conse-
quences of alternative methods in the treatment of enuresis nocturna. Ugeskr Laeger.
1994;156:4355–60.
Aspinall SL, Smith KJ, Cunningham FE, et al. Incremental cost-effectiveness of various monthly
doses of vardenafil. Value Health. 2011;14:97–101.
Augusto M, Greene M, Touya M, et al. Cost-effectiveness of long-acting injectable aripiprazole
once-monthly 400 mg in bipolar I disorder in the USA. J Comp Eff Res. 2018;7:637–50.
Balkrishnan R, Joish VN, Bhosle MJ, et al. Prior authorization of newer insomnia medications in
managed care: is it cost saving? J Clin Sleep Med. 2007;3:393–8.
Bamrungsawad N, Upakdee N, Pratoomsoot C, et al. Economic evaluation of intravenous immu-
noglobulin plus corticosteroids for the treatment of steroid-resistant chronic inflammatory
demyelinating polyradiculoneuropathy in Thailand. Clin Drug Investig. 2016;36:557–66.
Bansback N, Guh D, Oviedo-Joekes E, et al. Cost-effectiveness of hydromorphone for severe
opioid use disorder: findings from the SALOME randomized clinical trial. Addiction.
2018;113:1264–73.
Barnes TR, Leeson VC, Paton C, et al. Amisulpride augmentation in clozapine-unresponsive
schizophrenia (AMICUS): a double-blind, placebo-controlled, randomised trial of clinical
effectiveness and cost-effectiveness. Health Technol Assess. 2017;21:1–56.
Basu A, Benson C, Alphs L. Projecting the potential effect of using paliperidone palmitate once-
monthly and once-every-3-months long-acting injections among medicaid beneficiaries with
schizophrenia. J Manag Care Spec Pharm. 2018;24:759–68.
Bellingham GA, Peng PW. Duloxetine: a review of its pharmacology and use in chronic pain
management. Reg Anesth Pain Med. 2010;35:294–303.
Bolin K, Forsgren L. The cost effectiveness of newer epilepsy treatments: a review of the literature
on partial-onset seizures. PharmacoEconomics. 2012;30:903–23.
Bolin K, Berling P, Wasling P, et al. The cost-utility of sodium oxybate as narcolepsy treatment.
Acta Neurol Scand. 2017;136:715–20.
Bowrin K, Briere JB, Fauchier L, et al. Real-world cost-effectiveness of rivaroxaban compared with
vitamin K antagonists in the context of stroke prevention in atrial fibrillation in France. PLoS
One. 2020;15:e0225301.
226 R. Dodel et al.
Brennan A, Chick SE, Davies R. A taxonomy of model structures for economic evaluation of health
technologies. Health Econ. 2006;15:1295–310.
Brock D, Daniel N, Neumann PJ, et al. Ethical and distributive considerations. In: Neumann PJ,
Sanders G, Russell L, et al., editors. Cost-effectivenes in health and medicine. New York:
Oxford University Press; 2017.
Carrasco G, Baeza N, Cabre L, et al. Dexmedetomidine for the treatment of hyperactive delirium
refractory to haloperidol in nonintubated ICU patients: a nonrandomized controlled trial. Crit
Care Med. 2016;44:1295–306.
Carter JA, Dammerman R, Frost M. Cost-effectiveness of subdermal implantable buprenorphine
versus sublingual buprenorphine to treat opioid use disorder. J Med Econ. 2017;20:893–901.
Catala-Lopez F, Ridao M, Sanfelix-Gimeno G, et al. Cost-effectiveness of pharmacological treat-
ment of attention deficit hyperactivity disorder in children and adolescents: qualitative synthesis
of scientific evidence. Rev Psiquiatr Salud Ment. 2013;6:168–77.
Choi H, Mohit B. Cost-effectiveness of screening for HLA-B*1502 prior to initiation of carbamaz-
epine in epilepsy patients of Asian ancestry in the United States. Epilepsia. 2019;60:1472–81.
Christensen MC, Munro V. Cost per successfully treated patient for vortioxetine versus duloxetine
in adults with major depressive disorder: an analysis of the complete symptoms of depression
and functional outcome. Curr Med Res Opin. 2018;34:593–600.
Clements KM, Skornicki M, O’Sullivan AK. Cost-effectiveness analysis of antiepileptic drugs in
the treatment of Lennox-Gastaut syndrome. Epilepsy Behav. 2013;29:184–9.
Collins PY, Patel V, Joestl SS, et al. Grand challenges in global mental health. Nature. 2011;475:27–30.
Correll CU, Citrome L, Haddad PM, et al. The use of long-acting injectable antipsychotics in
schizophrenia: evaluating the evidence. J Clin Psychiatry. 2016;77:1–24.
Costa J, Fiorentino F, Caldeira D, et al. Cost-effectiveness of non-vitamin K antagonist oral
anticoagulants for atrial fibrillation in Portugal. Rev Port Cardiol. 2015;34:723–37.
Cottrell S, Tilden D, Robinson P, et al. A modeled economic evaluation comparing atomoxetine
with stimulant therapy in the treatment of children with attention-deficit/hyperactivity disorder
in the United Kingdom. Value Health. 2008;11:376–88.
Crawford MJ, Sanatinia R, Barrett B, et al. Lamotrigine for people with borderline personality
disorder: a RCT. Health Technol Assess. 2018;22:1–68.
D’amico E, Chisari CG, Gitto L, et al. Pharmacoeconomics of synthetic therapies for multiple
sclerosis. Expert Opin Pharmacother. 2019;20:1331–40.
Da Silva LR, Vianna CMM, Mosegui GBG, et al. Cost-effectiveness analysis of the treatment of
mild and moderate Alzheimer’s disease in Brazil. Braz J Psychiatry. 2019;41:218–24.
Dams J, Bornschein B, Reese JP, et al. Modelling the cost effectiveness of treatments for
Parkinson's disease: a methodological review. PharmacoEconomics. 2011;29:1025–49.
Darba J, Kaskens L, Perez C, et al. Pharmacoeconomic outcomes for pregabalin: a systematic
review in neuropathic pain, generalized anxiety disorder, and epilepsy from a Spanish perspec-
tive. Adv Ther. 2014;31:1–29.
Davies GM, Vyas A, Baxter CA. Economic evaluation of ezetimibe treatment in combination with
statin therapy in the United States. J Med Econ. 2017;20:723–31.
De Jong LA, Groeneveld J, Stevanovic J, et al. Cost-effectiveness of apixaban compared to other
anticoagulants in patients with atrial fibrillation in the real-world and trial settings. PLoS One.
2019;14:e0222658.
Debaveye S, De Smedt D, Heirman B, et al. Human health benefit and burden of the schizophrenia
health care pathway in Belgium: paliperidone palmitate long-acting injections. BMC Health
Serv Res. 2019;19:393.
Denchev P, Kaltman JR, Schoenbaum M, et al. Modeled economic evaluation of alternative
strategies to reduce sudden cardiac death among children treated for attention deficit/hyperac-
tivity disorder. Circulation. 2010;121:1329–37.
Djalalov S, Yong J, Beca J, et al. Genetic testing in combination with preventive donepezil
treatment for patients with amnestic mild cognitive impairment: an exploratory economic
evaluation of personalized medicine. Mol Diagn Ther. 2012;16:389–99.
Pharmacoeconomics 227
Doan QV, Gillard P, Brashear A, et al. Cost-effectiveness of onabotulinumtoxinA for the treatment
of wrist and hand disability due to upper-limb post-stroke spasticity in Scotland. Eur J Neurol.
2013;20:773–80.
Dong D, Sung C, Finkelstein EA. Cost-effectiveness of HLA-B*1502 genotyping in adult patients
with newly diagnosed epilepsy in Singapore. Neurology. 2012;79:1259–67.
Donnelly M, Haby MM, Carter R, et al. Cost-effectiveness of dexamphetamine and methylpheni-
date for the treatment of childhood attention deficit hyperactivity disorder. Aust N Z J Psychi-
atry. 2004;38:592–601.
Druais S, Doutriaux A, Cognet M, et al. Comparison of medical and economic benefits of
antipsychotics in the treatment of schizophrenia in France. L'Encéphale. 2017;43:311–20.
Dundar Y, Boland A, Strobl J, et al. Newer hypnotic drugs for the short-term management of
insomnia: a systematic review and economic evaluation. Health Technol Assess. 2004;8: iii-x,
1–125.
Dunlop AJ, Brown AL, Oldmeadow C, et al. Effectiveness and cost-effectiveness of unsupervised
buprenorphine-naloxone for the treatment of heroin dependence in a randomized waitlist
controlled trial. Drug Alcohol Depend. 2017;174:181–91.
Ebrahem AS, Oremus M. A pharmacoeconomic evaluation of cholinesterase inhibitors and
memantine for the treatment of Alzheimer’s disease. Expert Opin Pharmacother.
2018;19:1245–59.
Einarson TR, Bereza BG, Garcia Llinares I, et al. Cost-effectiveness of 3-month paliperidone
treatment for chronic schizophrenia in Spain. J Med Econ. 2017a;20:1039–47.
Einarson TR, Bereza BG, Tedouri F, et al. Cost-effectiveness of 3-month paliperidone therapy for
chronic schizophrenia in the Netherlands. J Med Econ. 2017b;20:1187–99.
Elliott J, Mccoy B, Clifford T, et al. Economic evaluation of stiripentol for Dravet syndrome: a cost-
utility analysis. PharmacoEconomics. 2018;36:1253–61.
Erder MH, Xie J, Signorovitch JE, et al. Cost effectiveness of guanfacine extended-release versus
atomoxetine for the treatment of attention-deficit/hyperactivity disorder: application of a
matching-adjusted indirect comparison. Appl Health Econ Health Policy. 2012;10:381–95.
Ernst R. Indirect costs and cost-effectiveness analysis. Value Health. 2006;9:253–61.
Faber A, Van Agthoven M, Kalverdijk LJ, et al. Long-acting methylphenidate-OROS in youths with
attention-deficit hyperactivity disorder suboptimally controlled with immediate-release methyl-
phenidate: a study of cost effectiveness in The Netherlands. CNS Drugs. 2008;22:157–70.
Feeny D, Krah M, Prosser LA, et al. Valuing health outcomes. In: Cost-effectiveness in health and
medicine. New York: Oxford University Press; 2017.
Fineberg NA, Baldwin DS, Drummond LM, et al. Optimal treatment for obsessive compulsive
disorder: a randomized controlled feasibility study of the clinical-effectiveness and cost-effec-
tiveness of cognitive-behavioural therapy, selective serotonin reuptake inhibitors and their
combination in the management of obsessive compulsive disorder. Int Clin Psychopharmacol.
2018;33:334–48.
Furneri G, Santoni L, Ricella C, et al. Cost-effectiveness analysis of escalating to natalizumab or
switching among immunomodulators in relapsing-remitting multiple sclerosis in Italy. BMC
Health Serv Res. 2019;19:436.
Galizzi MM, Ghislandi S, Miraldo M. Effects of reference pricing in pharmaceutical markets: a
review. PharmacoEconomics. 2011;29:17–33.
Geitona M, Stamuli E, Giannakodimos S, et al. Lacosamide as a first-line treatment option in focal
epilepsy: a cost-utility analysis for the Greek healthcare system. J Med Econ. 2019;22:359–64.
Gilmore A, Milne R. Methylphenidate in children with hyperactivity: review and cost-utility
analysis. Pharmacoepidemiol Drug Saf. 2001;10:85–94.
Girardin FR, Poncet A, Perrier A, et al. Cost-effectiveness of HLA-DQB1/HLA-B
pharmacogenetic-guided treatment and blood monitoring in US patients taking clozapine.
Pharm J. 2019;19:211–8.
Gold MR, Siegel JE, Lb R, et al. Cost-effectiveness in health and medicine. New York: Oxford
University Press; 1996.
228 R. Dodel et al.
Gracia Naya M. Cost effectiveness of treatment with triptanes in Spain. Rev Neurol. 2001;33:921–4.
Gregson N, Sparrowhawk K, Mauskopf J, et al. Pricing medicines: theory and practice, challenges
and opportunities. Nat Rev Drug Discov. 2005;4:121–30.
Groessl EJ, Tally SR, Hillery N, et al. Cost-effectiveness of a pharmacogenetic test to guide
treatment for major depressive disorder. J Manag Care Spec Pharm. 2018;24:726–34.
Grootendorst P, Pierard E, Shim M. Life-expectancy gains from pharmaceutical drugs: a critical
appraisal of the literature. Expert Rev Pharmacoecon Outcomes Res. 2009;9:353–64.
Gustavsson A, Svensson M, Jacobi F, et al. Cost of disorders of the brain in Europe 2010. Eur
Neuropsychopharmacol. 2011;21:718–79.
Gyrd-Hansen D, Olsen KR, Bollweg K, et al. Cost-effectiveness estimate of prehospital thrombol-
ysis: results of the PHANTOM-S study. Neurology. 2015;84:1090–7.
Hair PI, Mccormack PL, Curran MP. Eszopiclone: a review of its use in the treatment of insomnia.
Drugs. 2008;68:1415–34.
Hassiotis A, Poppe M, Strydom A, et al. Positive behaviour support training for staff for treating
challenging behaviour in people with intellectual disabilities: a cluster RCT. Health Technol
Assess. 2018;22:1–110.
Heller DJ, Coxson PG, Penko J, et al. Evaluating the impact and cost-effectiveness of statin use
guidelines for primary prevention of coronary heart disease and stroke. Circulation.
2017;136:1087–98.
Holmes M, Davis S, Simpson E. Alteplase for the treatment of acute ischaemic stroke: a NICE
single technology appraisal; an evidence review group perspective. PharmacoEconomics.
2015;33:225–33.
Hong J, Dilla T, Arellano J. A modelled economic evaluation comparing atomoxetine with
methylphenidate in the treatment of children with attention-deficit/hyperactivity disorder in
Spain. BMC Psychiatry. 2009;9:15.
Howard R, Cort E, Bradley R, et al. Amisulpride for very late-onset schizophrenia-like psychosis:
the ATLAS three-arm RCT. Health Technol Assess. 2018;22:1–62.
Husereau D, Drummond M, Petrou S, et al. Consolidated health economic evaluation reporting
standards (CHEERS) statement. Int J Technol Assess Health Care. 2013;29:117–22.
Idrisov B, Murphy SM, Morrill T, et al. Implementation of methadone therapy for opioid use
disorder in Russia – a modeled cost-effectiveness analysis. Subst Abuse Treat Prev Policy.
2017;12:4.
Jackson H, Mandell K, Johnson K, et al. Cost-effectiveness of injectable extended-release naltrex-
one compared with methadone maintenance and buprenorphine maintenance treatment for
opioid dependence. Subst Abus. 2015;36:226–31.
Jacoby A, Sudell M, Tudur Smith C, et al. Quality-of-life outcomes of initiating treatment with
standard and newer antiepileptic drugs in adults with new-onset epilepsy: findings from the
SANAD trial. Epilepsia. 2015;56:460–72.
Jofre-Bonet M, Sindelar JL, Petrakis IL, et al. Cost effectiveness of disulfiram: treating cocaine use
in methadone-maintained patients. J Subst Abus Treat. 2004;26:225–32.
Joo H, Wang G, George MG. Age-specific cost effectiveness of using intravenous recombinant tissue
plasminogen activator for treating acute ischemic stroke. Am J Prev Med. 2017;53:S205–12.
Kabore N, Marnat G, Rouanet F, et al. Cost-effectiveness analysis of mechanical thrombectomy
plus tissue-type plasminogen activator compared with tissue-type plasminogen activator alone
for acute ischemic stroke in France. Rev Neurol (Paris). 2019;175:252–60.
Kamae I, Hashimoto Y, Koretsune Y, et al. Cost-effectiveness analysis of apixaban against warfarin
for stroke prevention in patients with nonvalvular atrial fibrillation in Japan. Clin Ther.
2015;37:2837–51.
Kaplan JE, Vallabhaneni S, Smith RM, et al. Cryptococcal antigen screening and early antifungal
treatment to prevent cryptococcal meningitis: a review of the literature. J Acquir Immune Defic
Syndr. 2015;68(Suppl 3):S331–9.
Karnon J, Stahl J, Brennan A, et al. Modeling using discrete event simulation: a report of the
ISPOR-SMDM Modeling Good Research Practices Task Force-4. Med Decis Mak.
2012;32:701–11.
Pharmacoeconomics 229
Leppert MH, Campbell JD, Simpson JR, et al. Cost-effectiveness of intra-arterial treatment as an
adjunct to intravenous tissue-type plasminogen activator for acute ischemic stroke. Stroke.
2015;46:1870–6.
Leppert MH, Poisson SN, Carroll JD, et al. Cost-effectiveness of patent foramen ovale closure
versus medical therapy for secondary stroke prevention. Stroke. 2018;49:1443–50.
Linden K, Jormanainen V, Linna M, et al. Cost effectiveness of varenicline versus bupropion and
unaided cessation for smoking cessation in a cohort of Finnish adult smokers. Curr Med Res
Opin. 2010;26:549–60.
Maia CR, Stella SF, Wagner F, et al. Cost-utility analysis of methylphenidate treatment for children
and adolescents with ADHD in Brazil. Braz J Psychiatry. 2016;38:30–8.
Malhotra A, Wu X, Payabvash S, et al. Comparative effectiveness of endovascular Thrombectomy
in elderly stroke patients. Stroke. 2019;50:963–9.
Marfori CQ, Wu CZ, Katler Q, et al. Hysterectomy for the transgendered male: review of
perioperative considerations and surgical techniques with description of a novel 2-port laparo-
scopic approach. J Minim Invasive Gynecol. 2018;25:1149–56.
Martin AL, Huelin R, Wilson D, et al. A systematic review assessing the economic impact of
sildenafil citrate (Viagra) in the treatment of erectile dysfunction. J Sex Med. 2013;10:1389–
400.
Matsuo N, Morita T. Efficacy, safety, and cost effectiveness of intravenous midazolam and
flunitrazepam for primary insomnia in terminally ill patients with cancer: a retrospective
multicenter audit study. J Palliat Med. 2007;10:1054–62.
Mavranezouli I, Lokkerbol J. A systematic review and critical appraisal of economic evaluations of
pharmacological interventions for people with bipolar disorder. PharmacoEconomics.
2017;35:271–96.
Mcconaghy N, Blaszczynski A. Imaginal desensitization: a cost-effective treatment in two
shop-lifters and a binge-eater resistant to previous therapy. Aust N Z J Psychiatry.
1988;22:78–82.
Mccrone P, Chisholm D, Knapp M, et al. Cost-utility analysis of intravenous immunoglobulin and
prednisolone for chronic inflammatory demyelinating polyradiculoneuropathy. Eur J Neurol.
2003;10:687–94.
Melendez-Torres GJ, Auguste P, Armoiry X, et al. Clinical effectiveness and cost-effectiveness of
beta-interferon and glatiramer acetate for treating multiple sclerosis: systematic review and
economic evaluation. Health Technol Assess. 2017;21:1–352.
Men P, Yi Z, Li C, et al. Comparative efficacy and safety between amisulpride and olanzapine in
schizophrenia treatment and a cost analysis in China: a systematic review, meta-analysis, and
cost-minimization analysis. BMC Psychiatry. 2018;18:286.
Merry M, Boulware DR. Cryptococcal meningitis treatment strategies affected by the explosive cost
of flucytosine in the United States: a cost-effectiveness analysis. Clin Infect Dis. 2016;62:1564–8.
Mihalopoulos C, Magnus A, Lal A, et al. Is implementation of the 2013 Australian treatment
guidelines for posttraumatic stress disorder cost-effective compared to current practice? A cost-
utility analysis using QALYs and DALYs. Aust N Z J Psychiatry. 2015;49:360–76.
Mkrtchyan VR, Kaimovsky IL. Pharmacoeconomic aspects of monotherapy of focal epilepsy. Zh
Nevrol Psikhiatr Im S S Korsakova. 2019;119:92–8.
Moretti A, Ferrari F, Villa RF. Pharmacological therapy of acute ischaemic stroke: achievements
and problems. Pharmacol Ther. 2015;153:79–89.
Morin AK, Willett K. The role of eszopiclone in the treatment of insomnia. Adv Ther. 2009;26:500–18.
Nemeth B, Molnar A, Akehurst R, et al. Quality-adjusted life year difference in patients with
predominant negative symptoms of schizophrenia treated with cariprazine and risperidone. J
Comp Eff Res. 2017;6:639–48.
Neumann PJ, Thorat T, Zhong Y, et al. A systematic review of cost-effectiveness studies reporting
cost-per-DALY averted. PLoS One. 2016;11:e0168512.
Neumann PJ, Russell LB, Siegel J, et al. Using cost-effectiveness analysis in health and medicine.
Experiences since the original panel. New York: Oxford University Press; 2017a.
Neumann PJ, Sanders G, Russell L, et al. Cost-effectivenes in health and medicine. New York:
Oxford University Press; 2017b.
Pharmacoeconomics 231
Reeves P, Edmunds K, Levi C, et al. Cost-effectiveness of targeted thrombolytic therapy for stroke
patients using multi-modal CT compared to usual practice. PLoS One. 2018;13:e0206203.
Reinhold T, Muller-Riemenschneider F, Willich SN, et al. Economic and human costs of restless
legs syndrome. PharmacoEconomics. 2009;27:267–79.
Report of the Secretary (WHO). Global burden of mental disorders and the need for a comprehen-
sive, coordinated response from health and social sectors at the country level WHO. 2011.
https://apps.who.int/iris/handle/10665/78898
Ringel SP. Cost effectiveness: summary. Amyotroph Lateral Scler Other Motor Neuron Disord.
2002;3(Suppl 1):S67–9.
Roffe C, Nevatte T, Bishop J, et al. Routine low-dose continuous or nocturnal oxygen for people
with acute stroke: three-arm Stroke Oxygen Supplementation RCT. Health Technol Assess.
2018;22:1–88.
Rognoni C, Marchetti M, Quaglini S, et al. Apixaban, dabigatran, and rivaroxaban versus warfarin
for stroke prevention in non-valvular atrial fibrillation: a cost-effectiveness analysis. Clin Drug
Investig. 2014;34:9–17.
Romeo R, Knapp M, Tyrer P, et al. The treatment of challenging behaviour in intellectual
disabilities: cost-effectiveness analysis. J Intellect Disabil Res. 2009;53:633–43.
Rosenblat JD, Lee Y, Mcintyre RS. Does pharmacogenomic testing improve clinical outcomes for
major depressive disorder? A systematic review of clinical trials and cost-effectiveness studies. J
Clin Psychiatry. 2017;78:720–9.
Rubio-Valera M, Penarrubia-Maria MT, Iglesias-Gonzalez M, et al. Cost-effectiveness of antide-
pressants versus active monitoring for mild-to-moderate major depressive disorder: a multisite
non-randomized-controlled trial in primary care (INFAP study). Eur J Health Econ.
2019;20:703–13.
Ruger JP, Chawarski M, Mazlan M, et al. Cost-effectiveness of buprenorphine and naltrexone
treatments for heroin dependence in Malaysia. PLoS One. 2012;7:e50673.
Ruggeri M, Basile M, Zini A, et al. Cost-effectiveness analysis of mechanical thrombectomy with
stent retriever in the treatment of acute ischemic stroke in Italy. J Med Econ. 2018;21:902–11.
Sach TH, Whynes DK. Measuring indirect costs: is there a problem? Appl Health Econ Health
Policy. 2003;2:135–9.
Sado M, Wada M, Ninomiya A, et al. Does the rapid response of an antidepressant contribute to
better cost-effectiveness? Comparison between mirtazapine and SSRIs for first-line treatment of
depression in Japan. Psychiatry Clin Neurosci. 2019;73:400–8.
Santos AS, Guerra-Junior AA, Godman B, et al. Cost-effectiveness thresholds: methods for
setting and examples from around the world. Expert Rev Pharmacoecon Outcomes Res.
2018;18:277–88.
Saul JE, Lien R, Schillo B, et al. Outcomes and cost-effectiveness of two nicotine replacement treatment
delivery models for a tobacco quitline. Int J Environ Res Public Health. 2011;8:1547–59.
Schawo S, Van Der Kolk A, Bouwmans C, et al. Probabilistic Markov model estimating cost
effectiveness of methylphenidate osmotic-release oral system versus immediate-release meth-
ylphenidate in children and adolescents: which information is needed? PharmacoEconomics.
2015;33:489–509.
Schlander M. The pharmaceutical economics of child psychiatric drug treatment. Curr Pharm Des.
2010;16:2443–61.
Schöffski O, Von Der Schulenburg JM. Gesundheitsökonomische Evaluationen. Berlin: Springer;
2011.
Schwartz RP, Alexandre PK, Kelly SM, et al. Interim versus standard methadone treatment: a
benefit-cost analysis. J Subst Abus Treat. 2014;46:306–14.
Shearer J, Shanahan M, Darke S, et al. A cost-effectiveness analysis of modafinil therapy for
psychostimulant dependence. Drug Alcohol Rev. 2010;29:235–42.
Siebert U. Transparente Entscheidungen in Public Health mittels systematischer Entscheidung-
sanalyse. In: Schwartz Fw BB, Busse R, Leidl R, Raspe H, Siegrist J, editors. Das public health
Buch. Munich: Urban and Fischer; 2003a.
Pharmacoeconomics 233
Siebert U. When should decision-analytic modeling be used in the economic evaluation of health
care? Eur J Health Econ. 2003b;4:143–50.
Siebert U, Alagoz O, Bayoumi AM, et al. State-transition modeling: a report of the ISPOR-SMDM
Modeling Good Research Practices Task Force-3. Value Health. 2012;15:812–20.
Siegel JE, Weinstein MC, Russell LB, et al. Recommendations for reporting cost-effectiveness
analyses. Panel on cost-effectiveness in health and medicine. JAMA. 1996;276:1339–41.
Sikirica V, Haim Erder M, Xie J, et al. Cost effectiveness of guanfacine extended release as an
adjunctive therapy to a stimulant compared with stimulant monotherapy for the treatment of
attention-deficit hyperactivity disorder in children and adolescents. PharmacoEconomics.
2012;30:e1–15.
Simon J, Geddes JR, Gardiner A, et al. Comparative economic evaluation of quetiapine plus
lamotrigine combination vs quetiapine monotherapy (and folic acid vs placebo) in patients
with bipolar depression (CEQUEL). Bipolar Disord. 2018;20:733–45.
Skapinakis P, Caldwell D, Hollingworth W, et al. A systematic review of the clinical effectiveness
and cost-effectiveness of pharmacological and psychological interventions for the management
of obsessive-compulsive disorder in children/adolescents and adults. Health Technol Assess.
2016;20:1–392.
Sluiter RL, Kievit W, Van Der Wilt GJ, et al. Cost-effectiveness analysis of genotype-guided
treatment allocation in patients with alcohol use disorders using naltrexone or acamprosate,
using a modeling approach. Eur Addict Res. 2018;24:245–54.
Smith MD, Drummond M, Brixner D. Moving the QALY forward: rationale for change. Value
Health. 2009;12(Suppl 1):S1–4.
Snedecor SJ, Botteman MF, Bojke C, et al. Cost-effectiveness of eszopiclone for the treatment of
adults with primary chronic insomnia. Sleep. 2009;32:817–24.
Sohn M, Talbert J, Moga DC, et al. A cost-effectiveness analysis of off-label atypical antipsychotic
treatment in children and adolescents with ADHD who have failed stimulant therapy. Atten
Defic Hyperact Disord. 2016;8:149–58.
Stevenson M, Pandor A, Stevens JW, et al. Nalmefene for reducing alcohol consumption in people
with alcohol dependence: an evidence review group perspective of a NICE single technology
appraisal. PharmacoEconomics. 2015;33:833–47.
Stolk EA, Busschbach JJ, Caffa M, et al. Cost utility analysis of sildenafil compared with
papaverine-phentolamine injections. BMJ. 2000;320:1165–8.
Tannenbaum C, Diaby V, Singh D, et al. Sedative-hypnotic medicines and falls in community-
dwelling older adults: a cost-effectiveness (decision-tree) analysis from a US Medicare perspec-
tive. Drugs Aging. 2015;32:305–14.
Tavakoli M, Malek M. The cost utility analysis of riluzole for the treatment of amyotrophic lateral
sclerosis in the UK. J Neurol Sci. 2001;191:95–102.
Thursz M, Forrest E, Roderick P, et al. The clinical effectiveness and cost-effectiveness of STeroids
Or Pentoxifylline for Alcoholic Hepatitis (STOPAH): a 2 2 factorial randomised controlled
trial. Health Technol Assess. 2015;19:1–104.
Tilford JM, Payakachat N, Kovacs E, et al. Preference-based health-related quality-of-life outcomes
in children with autism spectrum disorders: a comparison of generic instruments. PharmacoE-
conomics. 2012;30:661–79.
Tilford JM, Payakachat N, Kuhlthau KA, et al. Treatment for sleep problems in children with autism
and caregiver spillover effects. J Autism Dev Disord. 2015;45:3613–23.
Tremblay G, Howard D, Tsong W, et al. Cost-effectiveness of perampanel for the treatment of
primary generalized tonic-clonic seizures (PGTCS) in epilepsy: a Spanish perspective. Epilepsy
Behav. 2018;86:108–15.
Trotter CL, Edmunds WJ. Reassessing the cost-effectiveness of meningococcal serogroup C conju-
gate (MCC) vaccines using a transmission dynamic model. Med Decis Mak. 2006;26:38–47.
Tuon FF, Florencio KL, Rocha JL. Burden of acute kidney injury in HIV patients under
deoxycholate amphotericin B therapy for cryptococcal meningitis and cost-minimization anal-
ysis of amphotericin B lipid complex. Med Mycol. 2019;57:265–9.
234 R. Dodel et al.
Vincent PD, Demers MF, Doyon-Kemp V, et al. One year mirror-image study using paliperidone
palmitate for relapse prevention of schizophrenia in four university hospitals in Canada.
Schizophr Res. 2017;185:96–100.
Voigt J, Carpenter L, Leuchter A. Cost effectiveness analysis comparing repetitive transcranial
magnetic stimulation to antidepressant medications after a first treatment failure for major
depressive disorder in newly diagnosed patients – a lifetime analysis. PLoS One. 2017;12:
e0186950.
Walley T, Haycox A. Pharmacoeconomics: basic concepts and terminology. Br J Clin Pharmacol.
1997;43:343–8.
Wang AS, Gunzler SA. Systematic review of the pharmacoeconomics of Parkinson disease
medications. Expert Opin Pharmacother. 2019;20:1659–70.
Weinstein MC. Recent developments in decision-analytic modelling for economic evaluation.
PharmacoEconomics. 2006;24:1043–53.
Whittington R, Faulds D. Hormone replacement therapy: I. A pharmacoeconomic appraisal of its
therapeutic use in menopausal symptoms and urogenital estrogen deficiency. PharmacoE-
conomics. 1994;5:419–45.
Wijnen BFM, Pos K, Velthorst E, et al. Economic evaluation of brief cognitive behavioural therapy
for social activation in recent-onset psychosis. PLoS One. 2018;13:e0206236.
Wortman MSH, Lokkerbol J, Van Der Wouden JC, et al. Cost-effectiveness of interventions for
medically unexplained symptoms: a systematic review. PLoS One. 2018;13:e0205278.
Wu EQ, Hodgkins P, Ben-Hamadi R, et al. Cost effectiveness of pharmacotherapies for attention-
deficit hyperactivity disorder: a systematic literature review. CNS Drugs. 2012;26:581–600.
Yagudina RI, Kulikov AY, Krylov VA, et al. Pharmacoeconomic analysis of the neuroprotective
medicines in the treatment of ischemic stroke. Zh Nevrol Psikhiatr Im S S Korsakova.
2019;119:60–8.
Yang J, Chen L, Chitkara N, et al. A Markov model to compare the long-term effect of aspirin,
clopidogrel and clopidogrel plus aspirin on prevention of recurrent ischemic stroke due to
intracranial artery stenosis. Neurol India. 2014;62:48–52.
Yoon J, Zisook S, Park A, et al. Comparing cost-effectiveness of aripiprazole augmentation with
other “next-step” depression treatment strategies: a randomized clinical trial. J Clin Psychiatry.
2018;80:18m12294.
Young AH, Evitt L, Brignone M, et al. Cost-utility evaluation of vortioxetine in patients with Major
Depressive Disorder experiencing inadequate response to alternative antidepressants in the
United Kingdom. J Affect Disord. 2017;218:291–8.
Zhao J, Jiang K, Li Q, et al. Cost-effectiveness of olanzapine in the first-line treatment of
schizophrenia in China. J Med Econ. 2019;22:439–46.
Zimovetz EA, Beard SM, Hodgkins P, et al. A cost-utility analysis of lisdexamfetamine versus
atomoxetine in the treatment of children and adolescents with attention-deficit/hyperactivity
disorder and inadequate response to methylphenidate. CNS Drugs. 2016;30:985–96.
Zimovetz EA, Joseph A, Ayyagari R, et al. A cost-effectiveness analysis of lisdexamfetamine
dimesylate in the treatment of adults with attention-deficit/hyperactivity disorder in the UK. Eur
J Health Econ. 2018;19:21–35.
Compliance and Psychoeducation
Stefan Unterecker
Contents
Compliance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235
Psychoeducation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240
Abstract
Noncompliance or nonadherence is a well-known but still underestimated prob-
lem in pharmacological treatment in general and in neuropsychopharmacological
therapy in detail. Approximately 50% of all patients with chronic diseases exhibit
poor compliance regarding drug therapy. Poor adherence is a worldwide problem
of striking magnitude. Increasing the effectiveness of adherence interventions
might have a far greater impact than any improvement in specific medical
treatments. Psychoeducation is effective and has the potential to reduce the rate
of nonadherence significantly.
Compliance
The term “compliance” is used in different disciplines; still its use in medicine
is most widespread with the meaning of obeying an order, rule, or request. More
and more, in medicine instead of “compliance” the term “adherence” is used, which
in contrast to compliance implies aspects of self-management and self-responsibility
of the patient. Even if nonadherence is the most important factor for suboptimal
S. Unterecker (*)
Center of Mental Health, Department of Psychiatry, Psychosomatics and Psychotherapy, University
Hospital of Würzburg, Würzburg, Germany
e-mail: Unterecker_S@ukw.de
outcomes and yet underestimated and underassessed (Hatch et al. 2017), there is
rather little scientific activity regarding this research topic since several years (De
Las Cuevas and De Leon 2017). Besides reports on a high medical service utilization
and medical costs in schizophrenia and related disorders by nonadherence (Joe and
Lee 2016), there are also unequivocal study results regarding economic aspects of
nonadherence (Pennington and McCrone 2018).
The World Health Organization (WHO 2003) defined adherence as “the extent
to which the patient follows medical instructions,” even if the term “medical” was
discussed to be insufficient in describing the wide range of interventions in the
treatment of chronic diseases. Moreover, it is noteworthy that the term “instructions”
underestimates the aspect of active collaboration of the patient and partnership
between healthcare provider and patient.
Comprehensively, the WHO in 2003 discussed the issue of adherence to long-
term therapies according to different chronic diseases. Poor adherence is a world-
wide problem of striking magnitude with growing in the same way as chronic
diseases increase worldwide. Adherence is an important modifier of effectiveness
in the health system and increasing the effectiveness of adherence interventions
might have a far greater impact than any improvement in specific medical treatments
(WHO 2003). Different influence factors on adherence need to be considered like
social and economic factors, characteristics of the disease, disease therapies, and
patient-related factors. Adherence is a dynamic process that needs to be followed up,
and health professionals need to be trained in adherence. A key factor for success in
improving adherence lies in the family, in the community, and in patients’ organi-
zations, so that a multidisciplinary approach is needed (WHO 2003).
Obviously, nonadherence is highly relevant in clinical medicine and is still an
underestimated problem in public health. Approximately 50% of all patients with
chronic diseases exhibit poor compliance regarding drug therapy (Gold and
McClung 2006). In patients with depressive disorders, several studies have
shown prevalence rates of nonadherence being dependent on the drug class: in
primary care setting 40% discontinued a treatment with tricyclics within 12 weeks
(Peveler et al. 1999). While only 20% of patients to whom tricyclics had been
prescribed filled four or more prescriptions within 6 months, 34% of patients under
newer antidepressants did (Katon et al. 1992). Regarding drug type several meta-
analyses described a small, but mostly significant difference between tricyclic
antidepressants and SSRIs with lower discontinuation rates for the latter (Mont-
gomery and Kasper 1995; Anderson and Tomenson 1995; Hotopf et al. 1997), so
that drug type is interpreted as a highly influential factor on drug adherence. The
co-medication of benzodiazepines for up to 8 weeks reduced the discontinuation
rates of antidepressant drug therapy only marginally (Furukawa et al. 2001).
Moreover, comorbid somatoform symptoms increase the rate of nonadherence
(Keeley et al. 2000), and sensation-seeking personality traits seem to do so
(Ekselius et al. 2000).
In a recent Spanish study on long-term antidepressant treatment, only 21% of
anxiety patients and 28% of depressed patients were compliant with higher rates
in patients under polypharmacy (Serna et al. 2015).
Compliance and Psychoeducation 237
The adherence to antidepressant drug therapy is more likely in patients who are
well-educated regarding the following aspects: daily medication intake, no benefit
by the treatment for the first 2–4 weeks, necessity of continuation even if feeling
better, and no stop of drug intake without consultation of the physician (Lin et al.
1995). The reasons for noncompliance in the treatment with antidepressant drugs
must be discussed in categories of illness, physician, patient, and antidepressant drug
(Demyttenaere 1998).
In a previous literature review on patients with bipolar disorder, a nonadherence
rate between 45% and 50% was found and good adherence was discussed as an
exception (Greene et al. 2018). In the context of adherence, the parameter of
medication possession ratio (MPR) often is used. The definition of MPR as ratio
between days of medication supply and the total period of observation within the
study in bipolar patients revealed that an MPR of at least 80% ranged from 8.3%
to 54.1% with a median of 28% and the mean MPR ranged from 19% to 77% with
a median of 47.1%. In bipolar patients no difference in adherence was found
between second-generation antipsychotics and mood stabilizers (Greene et al.
2018). Defining MPR as ratio between days of medication supply and the period
between first and last day of medication supply comes to higher adherence rates with
MPR of at least 80% in 58–62% and a mean MPR of 68–71% in patients under
different second-generation antipsychotics. Most often younger age and comorbid
substance use disorder in bipolar disorder are associated with poor adherence
(Greene et al. 2018).
In the treatment with antipsychotic drugs in contrast to other psychotropic drug
classes, different options of long-acting injectables are available, regarding both
first- and second-generation antipsychotic drugs. These formulations help to increase
drug adherence and avoid drug discontinuation, not alone by improved symptom
remission (Anderson et al. 2017). This is necessary as in schizophrenia patients after
1 year with oral antipsychotic therapy, the discontinuation rate is 51% – in contrast
to injectable paliperidone palmitate (27%) (Anderson et al. 2017). This difference
of higher adherence rates in patients under long-acting injectables in comparison to
oral atypical antipsychotics has been proved for different antipsychotic drugs like
paliperidone, risperidone, aripiprazole, and olanzapine (Pilon et al. 2017).
Regarding the assessment of adherence with antipsychotic medication in schizo-
phrenia, an overview on objective and subjective methods is available (Haddad et al.
2014; Table 1).
In contrast to the described adherence rates in affective disorder and schizophre-
nia, the 1-year compliance rate of patients under anti-dementia drugs in Germany
with about 60% is higher, and a treatment by specialist physician led to an increase of
the compliance rate (Bohlken et al. 2015).
Furthermore, in patients with epilepsy drug adherence must be considered. The
prevalence of adherence to antiepileptic drugs in epilepsy patients varies from 20%
to 80% (Hargrave and Remler 1996; Buck et al. 1997; Lannon 1997; Getnet et al.
2016) and seems to be lower in children in comparison to adult patients (Michaelis et
al. 2018). The factors of nonadherence in epilepsy treatment are socioeconomic-
related, health system-related, condition-related, treatment-related, and patient-
238 S. Unterecker
Psychoeducation
“pharmacophobia” which doubles the low adherence rate (De Leon and De Las
Cuevas 2017). Regarding an improvement of adherence, therapeutic drug monitor-
ing has the potential to optimize pharmacotherapy (Hiemke et al. 2018), not only by
controlling adherence via drug concentration measurement but also by including
information about the mode of action of pharmacotherapy in psychoeducation.
Obviously, patient information and education is effective and reduces the rate
of nonadherence, even if 6 months after patient education one third of patients is
not compliant (Gold and McClung 2006). Therefore, more attention is needed to
identify patients who are least likely to be adherent to drug therapy and intensify
psychoeducation individually (Gold and McClung 2006; Berk et al. 2010).
A recent randomized clinical trial showed that psychoeducation including a
treatment initiation and participation program is effective and increases the
adherence rate for three to five times in antidepressant treated patients (Sirey
et al. 2017).
After the identification of barriers of adherence, it is important to consider
facilitators and to include them in psychoeducation because targeted interventions
are necessary (Ho et al. 2017).
Therefore, psychoeducation has the potential of a meaningful impact on public
health.
References
Aagaard J, Foldager L, Makki A, Hansen V, Müller-Nielsen K. The efficacy of psychoeducation
on recurrent depression: a randomized trial with a 2-year follow-up. Nord J Psychiatry. 2017;
71:223–9.
Anderson IM, Tomenson BM. Treatment discontinuation with selective serotonin reuptake inhib-
itors compared with tricyclic antidepressants: a meta-analysis. Br Med J. 1995;310:1433–8.
Anderson JP, Icten Z, Alas V, Benson C, Joshi K. Comparison and predictors of treatment adherence
and remission among patients with schizophrenia treated with paliperidone palmitate or atypical
oral antipsychotics in community behavioral health organizations. BMC Psychiatry. 2017;
17:346.
Arlt AD, Nestoriuc Y, Rief W. Why current drug adherence programs fail: addressing psychological
risk factors of nonadherence. Curr Opin Psychiatry. 2017;30:326–33.
Bäuml J, Behrendt B, Henningsen P, Pitschel-Walz G. Handbuch der Psychoedukation. Stuttgart:
Schattauer; 2016.
Berk L, Hallam KT, Colom F, Vieta E, Hasty M, Macneil C, Berk M. Enhancing medication
adherence in patients with bipolar disorder. Human Psychopharmacology: Clinical and Exper-
imental 2010;25(1):1–16.
Bohlken J, Weber S, Rapp MA, Kostev K. Continuous treatment with antidementia drugs in
Germany 2003–2013: a retrospective database analysis. Int Psychogeriatr. 2015;27:1335–42.
Buck D, Jacoby A, Baker GA, Chadwick DW. Factors influencing compliance with antiepileptic
drug regimes. Seizure. 1997;6:87–93.
Chen R, Zhu X, Capitao LP, Zhang H, Luo J, Wang X, Xi Y, Song X, Feng Y, Cao L, Malhi GS.
Psychoeducation for psychiatric inpatients following remission of a manic episode in bipolar I
disorder: a randomized controlled trial. Bipolar Disord. 2019;21:76–85.
Claxton A, Klerk E. de, Parry M, Robinson JM, Schmidt ME. Patient compliance to a new enteric-
coated weekly formulation of fluoxetine during continuation treatment of major depressive
disorder. The Journal of Clinical Psychiatry 2000;61(12):928–932.
Compliance and Psychoeducation 241
Cloyd JC, Kriel RL, Jones-Saete CM, Ong BY, Jancik JT, Remmel RP. Comparison of sprinkle
versus syrup formulations of valproate for bioavailability, tolerance, and preference. J Pediatr.
1992;120:634–8.
Conradi HJ, Bos EH, Kamphuis JH, de Jonge P. The ten-years course of depression in primary care
and long-term effects of psychoeducation, psychiatric consultation and cognitive behavioral
therapy. J Affect Disord. 2017;217:174–82.
De Las Cuevas C, de Leon J. Reviving research on medication attitudes for improving pharmaco-
therapy: focusing on adherence. Psychother Psychosom. 2017;86:73–9.
De Las Cuevas C, Penate W, Cabrera C. Perceived health control: a promising step forward in our
understanding of treatment adherence in psychiatric care. J Clin Psychiatry. 2016;77:e1233–9.
De Leon J, de Las Cuevas C. The art of pharmacotherapy – reflections on pharmacophobia. J Clin
Psychopharmacol. 2017;37:131–7.
Demyttenaere K. Noncompliance with antidepressants: who’s to blame? Int Clin Psychopharmacol.
1998;13:S19–25.
Ekselius L, Bengtsson F, von Knorring L. Non-compliance with pharmacotherapy of depression is
associated with a sensation seeking personality. Int Clin Psychopharmacol. 2000;15:273–8.
Furukawa TA, Streiner DL, Young LT. Is antidepressant-benzodiazepine combination therapy
clinically more useful? A meta-analytic study. J Affect Disord. 2001;5:173–7.
Garnett WR. Antiepileptic drug treatment: outcomes and adherence. Pharmacotherapy. 2000;20:
S191–9.
Getnet A, Woldeyohannes SM, Bekana L, Mekonen T, Fekadu W, Menberu M, Yimer S, Assaye A,
Belete A, Belete H. Antiepileptic drug nonadherence and its predictors among people with
epilepsy. Behav Neurol. 2016;2016:3189108.
Gold DT, McClung B. Approaches to patient education: emphasizing the long-term value of
compliance and persistence. Am J Med. 2006;119:S32–7.
Greene M, Paladini L, Lemmer T, Piedade A, Touya M, Clark O. Systematic literature review
on patterns of pharmacological treatment and adherence among patients with bipolar disorder
type I in the USA. Neuropsychiatr Dis Treat. 2018;14:1545–59.
Gutierrez PM, Wortzel HS, Forster JE, Leitner RA, Hostetter TA, Brenner LA. Blister packaging
medication increases treatment adherence in psychiatric patients. J Psychiatr Pract. 2017;23:
320–7.
Haddad PM, Brain C, Scott J. Nonadherence with antipsychotic medication in schizophrenia:
challenges and management strategies. Patient Relat Outcome Meas. 2014;5:43–62.
Hargrave R, Remler MP. Noncompliance. J Natl Med Assoc. 1996;88:7, 11
Hatch A, Docherty JP, Carpenter D, Ross R, Weiden PJ. Expert consensus survey on medication
adherence in psychiatric patients and use of a digital medicine system. J Clin Psychiatry. 2017;
78:e803–12.
Hiemke C, Bergemann N, Clement HW, Conca A, Deckert J, Domschke K, Eckermann G, Egberts
K, Gerlach M, Greiner C, Gründer G, Haen E, Havemann-Reinecke U, Hefner G, Helmer R,
Janssen G, Jaquenoud E, Laux G, Messer T, Mössner R, Müller MJ, Paulzen M, Pfuhlmann B,
Riederer P, Saria A, Schoppek B, Schoretsanitis G, Schwarz M, Gracia MS, Stegmann B,
Steimer W, Stingl JC, Uhr M, Ulrich S, Unterecker S, Waschgler R, Zernig G, Zurek G,
Baumann P. Consensus guidelines for therapeutic drug monitoring in neuropsychophar-
macology: update 2017. Pharmacopsychiatry. 2018;51:9–62.
Ho SC, Jacob SA, Tangiisuran B. Barriers and facilitators of adherence to antidepressants among
outpatients with major depressive disorder: a qualitative study. PLoS One. 2017;12:e0179290.
Hotopf M, Hardy R, Lewis G. Discontinuation rates of SSRIs and tricyclic antidepressants: a meta-
analysis and investigation of heterogeneity. Br J Psychiatry. 1997;170:120–7.
Jawad I, Watson S, Haddad PM, Talbot PS, McAllister-Williams RH. Medication nonadherence
in bipolar disorder: a narrative review. Ther Adv Psychopharmacol. 2018;8:349–63.
Joe S, Lee JS. Association between non-compliance with psychiatric treatment and non-
psychiatric service utilization and costs in patients with schizophrenia and related disorders.
BMC Psychiatry. 2016;16:444.
Katon W. Collaborative management to achieve treatment guidelines. JAMA 1995;273(13):1026.
242 S. Unterecker
Katon W, von Korff M, Lin E, Bush T, Ormel J. Adequacy and duration of antidepressant treatment
in primary care. Med Care. 1992;30:67–76.
Katon W, Von Korff M, Lin E, Simon G, Walker E, Unützer J, Bush T, Russo J, Ludman E. Stepped
collaborative care for primary care patients with persistent symptoms of depression: a random-
ized trial. Arch Gen Psychiatry. 1999;56:1109–15.
Keeley R, Smith M, Miller J. Somatiform symptoms and treatment nonadherence in depressed
family medicine outpatients. Arch Fam Med. 2000;9:46–54.
Krijnen-de Bruin E, Muntingh AD, Hoogendoorn AW, van Straten A, Batelaan NM,
Maarsingh OR, van Balkom AJ, van Meijel B. The GET READY relapse prevention pro-
gramme for anxiety and depression: a mixed-methods study protocol. BMC Psychiatry.
2019;19:64.
Lannon SL. Using a health promotion model to enhance medication compliance. J Neurosci Nurs.
1997;29:170–8.
Lin EH, Von Korff M, Katon W, Bush T, Simon GE, Walker E, Robinson P. The role of the primary
care physician in patients’ adherence to antidepressant therapy. Med Care. 1995;33:67–74.
Lin EH, Simon GE, Katon WJ, Russo JE, Von Korff M, Bush TM, Ludman EJ, Walker EA. Can
enhanced acute-phase treatment of depression improve long-term outcomes? A report of
randomized trial in primary care. Am J Psychiatry. 1999;156:643–5.
Lorig KR, Sobel DS, Stewart AL, Brown BW, Bandura A, Ritter P, Gonzalez VM, Laurent DD,
Holman HR. Evidence suggesting that a chronic disease self-management program can improve
health status while reducing hospitalization: a randomized trial. Med Care. 1999;37:5–14.
Mathes T, Pieper D, Antoine SL, Eikermann M. 50% adherence of patients suffering chronic
conditions – where ist he evidence? Ger Med Sci. 2012;10:Doc16.
Melin EO, Svensson R, Thulesius HO. Psychoeducation against depression, anxiety, alexithymia
and fibromyalgia: a pilot study in primary care for patients on sick leave. Scand J Prim Health
Care. 2018;36:123–33.
Michaelis R, Tang V, Goldstein LH, Reuber M, LaFrance WC, Lundgren T, Modi AC, Wagner JL.
Psychological treatments for adults and children with epilepsy: evidence-based recom-
mendations by the international league against epilepsy psychology task force. Epilepsia.
2018;59:1282–302.
Montgomery SA, Kasper S. Comparison of compliance between serotonin reuptake inhibitors and
tricyclic antidepressants: a meta-analysis. Int Clin Psychopharmacol. 1995;9:33–40.
Myers ED, Branthwaite A. Out-patient compliance with antidepressant medication. Br J Psychiatry.
1992;160:83–6.
National Institute for Health and Care Excellence. Medicines adherence: involving patients in
decisions about prescribed medicines and supporting adherence. NICE guidelines CG76.
London: Royal College of General Practitioners; 2009.
Pennington M, McCrone P. Does non-adherence increase treatment costs in schizophrenia?
Pharmacoeconomics. 2018;36:941–55.
Peveler R, George C, Kinmonth AL, Campbell M, Thompson C. Effect of antidepressant drug
counselling and information leaflets on adherence to drug treatment in primary care: randomised
controlled trial. BMJ. 1999;319:612–5.
Phan SV. Medication adherence in patients with schizophrenia. Int J Psychiatry Med.
2016;51:211–9.
Pilon D, Tandon N, Lafeuille MH, Kamstra R, Emond B, Lefebvre P, Joshi K. Treatment pattern,
health care resource utilization, and spending in Medicaid beneficiaries initiating second-
generation long-acting injectable agents versus oral atypical antipsychotics. Clin Ther. 2017;
39:1972–85.
Pompili M, Giordano G, Luciano M, Lamis DA, Del Vecchio V, Serafini G, Sampogna G, Erbuto D,
Falkai P, Fiorillo A. Unmet needs in schizophrenia. CNS Neurol Disord Drug Targets.
2017;16:870–84.
Compliance and Psychoeducation 243
Profit D, Rohatagi S, Zhao C, Hatch A, Docherty JP, Peters-Strickland TS. Developing a digital
medicine system in psychiatry: ingestion detection rate and latency period. J Clin Psychiatry.
2016;77:e1095–100.
Rohatagi S, Profit D, Hatch A, Zhao C, Docherty JP, Peters-Strickland TS. Optimization of a digital
medicine system in psychiatry. J Clin Psychiatry. 2016;77:e1101–7.
Sendt KV, Tracy DK, Bhattacharyya S. A systematic review of factors influencing adherence to
antipsychotic medication in schizophrenia-spectrum disorders. Psychiatry Res. 2015;225:
14–30.
Serna MC, Real J, Cruz I, Galvan L, Martin E. Monitoring patients on chronic treatment with
antidepressants between 2003 and 2011: analysis of factors associated with compliance. BMC
Public Health. 2015;15:1184.
Sirey JA, Banerjee S, Marino P, Bruce ML, Halkett A, Turnwald M, Chiang C, Liles B, Artis A,
Blow F, Klaes HC. Adherence to depression treatment in primary care – a randomized clinical
trial. JAMA Psychiatry. 2017;74:1129–35.
Wenze SJ, Armey MF, Weinstock LM, Gaudiano BA, Miller IW. An open trial of a smartphone-
assisted, adjunctive intervention to improve treatment adherence in bipolar disorder. J Psychiatr
Pract. 2016;22:492–504.
World Health Organization. Adherence to long-term therapies: evidence for action. Geneva:
World Health Organization; 2003. http://www.who.int/chp/knowledge/publications/adherence_
report/en/
Psychopathology: Rating Scales for the
Assessment of Mental Status
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 246
Assessment Aims . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 246
Characteristics of Instruments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247
Evaluation Criteria and Selection of Rating Scales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
Selected Areas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251
Side Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 254
Scales for Measuring Sexual Functioning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 256
Scales for Measuring Extrapyramidal Motor Side Effects (EPS) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
Other Domains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 259
Ascertainment of Progression and Changes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 259
Requirements for the Use of Interviews and Rating Scales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 261
Examples of Practical Application . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 261
Conclusion and Perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 263
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 264
Abstract
In the context of psychopharmacotherapy, diagnostic instruments have always
played a prominent role, in order either to assist the classification of a patient or to
measure the degree of severity of psychopathological syndromes at a given stage
as well as during the course of the disorder. In both these cases, but also in
connection with other clinical areas, rating scales and diagnostic interviews are
Introduction
Assessment Aims
Characteristics of Instruments
Unidimensional rating scales, for example, the HDRS, only produce one scale
value representing the sum score of individual items, thus limiting themselves to a
single indicator of degree of severity for a specific syndrome (e.g., depressive
syndrome). Multidimensional rating scales, by contrast, such as the AMDP-System
(Broome et al. 2018), allow for a wide spectrum of different syndromes to be
assessed and are therefore applicable to various groups of disorders. Both types of
procedure involve mostly additive rating scales where the scale value or values are
calculated by simply adding up the individual scores of items assessed. The obtained
value or values are then used as indicator(s) of the constructs examined for. Global
rating scales, as their name suggests, aim to facilitate assessment of particular
phenomena based on a global impression rather than the sum total of individual
symptoms. The most widely used such scale is the Clinical Global Impressions Scale
(CGI; cf. AMDP and CIPS 1990). Its first component, the CGI-Severity (CGI-S)
Scale, contains a scale giving ratings from 1 (normal, not at all ill) to 7 (among the
most severely ill patients), which allows for a global evaluation of the patient’s
overall condition (unrelated to a specific disorder). The second component, the CGI
Improvement (CGI-I) Scale, contains treatment response ratings ranging from 1
(marked improvement since initiation of treatment) to 7 (marked deterioration
since initiation of treatment), which allows for a global assessment of change. The
CGI is one of the most widely applied scales in (psychopharmacological) trial
studies.
A further global assessment scale frequently referred to in the literature is the
Sheehan Disability Scale (SDS, Sheehan 1983), a self-rating scale. In a 10-point
Likert Scale (scored from 0 not at all to 10 extremely), the level of impairment to
functionality during the preceding month is assessed in the following three areas:
work and occupation; social contacts and leisure activities; family life and household
duties.
The distinction between short and long rating scales stems from attempts to
develop shorter and more concise versions of existing, valid, and reliable scales.
For the Symptom Check List – Revised (SCL-90-R), a short version, the Brief
Psychopathology: Rating Scales for the Assessment of Mental Status 249
Symptom Inventory (BSI), has been available for many years. Instead of the original
90-item scale, this abbreviated version is limited to 53 items, which are divided into
the same 9 subscales as in the long version.
In most self- and observer-rating scales, assessment options take the form of
verbal descriptors (generally ranging from not present to mild, moderate, and severe)
or operationalized levels of assessment (as in, e.g., the Montgomery Åsberg Depres-
sion Rating Scale, MADRS). Differentiated from such scales are visual analogue
rating scales. This group of scales follows the general principle of inviting the rater
to mark, on a roughly 100 mm long line ending on either side in polar opposites, the
position corresponding with the extent of agreement or disagreement with either of
the two opposing extremes given. The most commonly used scale in this category is
Aitken’s (1969) VAS to examine for depressive moods. This scale was formerly
predominantly applied in research settings, where repeated assessment was a fre-
quent requirement (e.g., in the treatment of sleep deprivation).
These publications, however, chiefly discuss general and basic considerations and
recommendations (e.g., evaluation of therapy success; types of scale to be applied).
Selected Areas
Given the large number of rating scales available, only a limited selection can be
addressed here. Initially, the first consideration is whether an instrument is applicable
with different groups of disorders (so-called general rating scales) or whether it is
specifically designed for one particular group only (disorder-specific rating scales).
As expected, most rating scales fall into the latter category.
Table 3 contains an overview of the best known self- and observer-rating scales
which find application in more than one group of disorders. A distinction is made
between rating scales screening for specific syndromes and those screening for more
general aspects of psychiatric impairment (i.e., mood).
On account of their prevalence, the following groups of disorders are of particular
interest and discussed below: schizophrenic disorders, affective disorders, and
anxiety disorders.
Schizophrenic disorders. As a result of the close link between the discovery of
psychotropic drugs and the development of rating scales, instruments screening for
schizophrenic symptoms were among the first to be developed (cf. Stieglitz et al.
252 R.-D. Stieglitz and H.-P. Volz
2017). In order to measure and evaluate the efficacy of newly discovered substances,
instruments have been continuously developed since the 1960s, a number of which
remain in use today. Table 4 contains a summary of the most important of these
instruments. They serve to assess the primary symptom dimensions in schizophrenic
disorders such as delusions, hallucinations, or thought disorders, which are, in some
instances, grouped according to negative and positive symptom domains. The
instruments listed remain the most important instruments in neuroleptic (antipsy-
chotic) studies.
Psychopathology: Rating Scales for the Assessment of Mental Status 253
Affective disorders. Rating scales in the area of affective disorders are grouped
into scales for measuring depressive symptoms and scales for measuring manic
symptoms. While there are over a hundred instruments available for depressive
symptoms, only a small number exists for manic symptoms. Table 5 again provides
a summary of the most important instruments still in use today. In respect to
depression rating scales, the following points are of particular interest:
Side Effects
In the context of clinical studies, side effects are now largely known as adverse
events (AEs) or adverse reactions (ARs). The term AE refers to any unintended
medical occurrence in patients or participants in clinical studies after administration
of a particular drug, independent of any causal relation to the medication in question.
An AR, on the other hand, denotes any unintended and adverse reaction to a trial
substance independent of the size of dosage administered (Harnisch et al. 2012), thus
indicating a clear causal relation. ARs therefore closely approximate what would
traditionally have been referred to as side effects.
Accurate assessment of AEs and ARs is decisive in the effective evaluation of
benefit and risk potentials of particular therapeutic measures. Assessment of AEs
tends to be restricted to medical therapies, while the practice of recording AEs or
ARs in the context of psychotherapeutic intervention remains in its infancy (see,
e.g., Linden 2013). A benefit-risk assessment is of particular importance in the
clinical development of pharmaceutical substances, especially during Phase II
and Phase III trials. Usually, this takes the form of group statistical comparisons
with placebos or a reference substance. The assessment of side effects continues
to play a pivotal role during later administration of a substance to individual
patients.
256 R.-D. Stieglitz and H.-P. Volz
It is thus all the more surprising to find that while there exist – at least in the
context of clinical studies – various conventions of assessing the efficacy of medical
substances (e.g., the Hamilton Depression Rating Scale or the Montgomery-Åsberg
Depression Rating Scale), there appears to be almost no consensus on how the
presence of AEs (or ARs) should be systematically ascertained.
Admittedly, it is far simpler to assess and classify AEs as ARs, since the causal
attributions which qualify AEs as ARs are relatively unreliable.
1) On the one hand, patients are asked in an open question whether any AEs were
noticed. Where this is confirmed, further questions follow as to the nature of
these AEs, so they can be attributed to various organ classes based on a
universally applicable classification system. In some cases, assessment includes
the level of severity of AEs experienced. A system frequently applied to match
AEs to specific organ classes is the MedDRA-System (Medical Dictionary for
Regulatory Activities). MedDRA follows a multiaxial structure, in which a
preferred term (PT) may be matched to one or more System Organ Classes
(SOCs). However, in order to prevent double counting, every PT has a primary
allocation. MedDRA is hierarchically organized. The SOCs represent the
highest of five levels (i.e., the first level with 26 organ classes), with PTs
representing the fourth level. Based on this, tables are established (e.g., in
final reports on clinical trials) which comprise the SOCs (the first level) and,
for example, show individual AEs as PTs (fourth level) (thus, e.g., SOC = ner-
vous system, PT = headache).
This way of recording AEs may lead to instances of “underreporting” and AEs
being recorded to a lesser extent than actually present. Conversely, explicit
questioning as to AEs, or the use of rating scales, may lead to cases of “over-
reporting.” This dilemma remains unresolved.
2) Alternatively, systematic AE screening instruments are used. These, however, are
not available for every spectrum of side effects (exception the scale “Somatic
Symptoms” in the AMDP system; Stieglitz et al. 2017; Broome et al. 2018). The
best and most extensive rating scales are those designed to assess sexual dys-
functions and extrapyramidal motor symptoms (EPS) (after administration of
antipsychotic drugs). Both types of scales are investigated in more detail in the
following.
The most commonly used scales in this area are summarized in Table 7.
The Arizona Sexual Experience Scale (ASEX). The ASEX (McGahuey et al.
2000) consists of five questions aiming at sexual interest, arousal, lubrification/
erection, orgasm, and satisfaction (Clayton 2001). The scale can be used in hetero-
or homosexual populations irrespective of the availability of a sexual partner. Global
Psychopathology: Rating Scales for the Assessment of Mental Status 257
Table 7 Comparison of assessment instruments for sexual functioning. (Adapted from Clayton
2001)
ASEX CFSQ-C DISF-SR RSI
No. of questions 5 14 25 29 M/18 F
<10 min Yes Yes No Yes
Validated Moderate Good Good Good
Measures change Yes Yes No Yes
Explicit/intrusive No No Yes Yes
Likert scale Yes Yes Yes Only 5 items
Abbreviations: ASEX arizona sexual experiences scale, CSFQ changes in sexual functioning
questionnaire, DISF-SR derogatis interview for sexual functioning-short report, RSI rush sexual
inventory
scores above 18 or single item scores above 4 indicate sexual dysfunction and may
be used to monitor change in function over time.
The Changes in Sexual Functioning Questionnaire-Clinical version (CSFQ-C).
The CSFQ-C (Clayton et al. 1997) uses 14 questions to measure sexual activity in five
domains: sexual pleasure, sexual desire/frequency, sexual desire/interest, arousal, and
orgasm (Clayton 2001). Since a definition of sexual activity (intercourse, masturbation,
sexual fantasy) is initially given, the patient is not required to report specific sexual
behaviors. Global scores below 48 for men and below 42 for women are indicative of
sexual dysfunction.
The Derogatis Interview for Sexual Functioning – Short Report (DISF-SR).
The DISF-SR (Derogatis 1997) measures quality of sexual functioning with 25
questions in 5 domains: sexual cognition, arousal, sexual behavior and experiences,
orgasm, and sexual drive (Clayton 2001). Questions aiming at frequency of specific
sexual behaviors might be perceived as intrusive by the patient.
The Rush Sexual Inventory (RSI). The RSI (Zajecka et al. 1997) measures
sexual functioning over time with 29 questions addressed to men and 18 questions to
women (Clayton 2001). Questions about the frequency of specific sexual behavior
may be perceived as intrusive. Five items use a visual analogue scale; the other items
require a yes/no answer.
A systematic literature overview (Knol et al. 2010) identified a total of 17 scales for
evaluating EPS. Table 8 contains a summary of the scales most used.
The Drug-Induced Extra Pyramidal Symptoms Scale (DIEPSS). The DIEPSS
(Knol et al. 2010; Inada et al. 1996, 2003) consists of nine items, five for parkin-
sonism and one each for akathisia, dystonia, dyskinesia, and severity. Items are
scored on a five-point scale. The scale showed good inter- and intrarater reliability.
The parkinsonism items show high agreement with the Simpson-Angus Scale
(SAS).
258 R.-D. Stieglitz and H.-P. Volz
Table 8 Comparison of assessment instruments for extrapyramidal symptoms. (EPS; adapted from
Knol et al. 2010)
DIP Manual Duration
Scale Total items Items available Scoring Interpretation [in mins]
DIEPSS 9 5 Yes Clear Clear ?
ESRS
Questionnaire 7 16 Yes Clear Unclear 15
complex
Examination 34
Global 4
impression
SAS 10 10 Yes Clear Clear 10
UKU
Single items 48 3 Yes Clear Unclear 30
complex
Global 2X2 (patient
assessment and physician)
BAR 4 0 Yes Clear Clear 5
complex
Abbreviations: BAR bar akathisia rating scale, DIEPSS drug-induced extra pyramidal symptoms
scale, DIP drug-induced parkinsonism, EPS extrapyramidal symptoms, ESRS extrapyramidal
symptom rating scale, SAS simpson angus scale, UKU udvalg for kliniske undersogelse
Other Domains
Besides the assessment of various aspects of psychopathology and side effects in the
context of psychopharmacology, assessing other domains, such as quality of life as
an additional outcome, is important. Again, there is now a wide variety of reliable
instruments available in this respect (cf. also Bech 1993). Depending on the nature of
a specific disorder, further domains require evaluation. With schizophrenia, for
example, important areas are cognitive functioning (e.g., Barnett et al. 2010) or
adherence or insight (e.g., Keefe 2012).
Special rating scales to measure degrees of severity are particularly useful also for
the assessment of change. There are several options available, as shown in Table 9.
While therapy response is traditionally defined as a 50% decrease in the score
from baseline of a rating scale, increasingly complex and demanding approaches are
coming into use. Some of these are briefly discussed below, in particular the concept
of clinical significance and ways to operationalize remission and recovery.
Where both these requirements are met, a clinically significant change is deemed
to have occurred (cf. Ogles 2013). Unfortunately, this rather conservative approach
to evaluating therapy success has hitherto found little application in research.
There are numerous suggestions for ways to approach the operationalization of
response or remission, some of which are briefly outlined here:
Table 10 Definitions of response, remission, and recovery based on the Hamilton Depression
Rating Scale. (HDRS; adapted from O’Donavan 2004)
Definition
Nonresponse 20% reduction
Partial response 21–49% reduction
Response 50% reduction
Residual symptoms Response and total score 8
Remission Response and total score 7
Recovery Remission and return of function
Table 11 Remission criteria for social phobia. (Adapted from Ballenger 1999)
Symptomatology Operationalization
Abolish core symptoms e.g., 70% improvement LSAS
Reduction anxiety e.g., HAMA 7–10
Reduction depression e.g., HAMD 7
Resolve functional impairments SDS 1 (mildly disabled)
LSAS liebowitz social anxiety scale, HAMA hamilton anxiety scale, HAMD hamilton depression
scale, SDS sheehan disability scale
Psychopathology: Rating Scales for the Assessment of Mental Status 261
Table 12 Remission criteria for schizophrenia based on the Positive and Negative Syndrome Scale
PANSS. (Adapted from Andreasen et al. 2005)
PANSS items all 3
P1 Delusions
G9 Unusual thought content
P3 Hallucinatory behavior
P2 Conceptual disorganization
G5 Mannerisms/posturing
N1 Blunted affect
N4 Social withdrawal
N6 Lack of spontaneity
• Selection of instruments. Where the aim is not simply to rely on past recom-
mendations, a knowledge and understanding of the psychometric quality of
assessment instruments is mandatory in order to be able to evaluate these effec-
tively. Only an evaluation of available evidence as to the reliability and validity of
a chosen instrument allows for a valid decision on the appropriacy of the
instrument in a given case.
• Conducting the examination. Particularly in the context of interviews and
observer-rating scales, formal training and practical experience in the administra-
tion of specific instruments is an essential prerequisite. For this purpose, a number
of carefully designed concepts are available (e.g., Kobak et al. 2004, 2005). In
clinical practice, and to some extent also in research, the need for this indisputably
essential requirement fails to be adequately recognized. Especially in routine
clinical practice, screening procedures are being deployed by medical practi-
tioners who have undergone no formal training in their use and administration.
• Interpretation. The use of rating scales also requires at least a basic knowledge
of standardization methods (e.g., the significance of percentile ranks and T-
scores). As already outlined above, rating scale values alone are meaningless.
Only their evaluation in relation to benchmark values and reliable standards of
comparison allows for meaningful interpretation of scores attained.
For the purpose of illustrating their specific benefits, there follow some examples of
interviews and rating scales in practical application.
Example 1 Figure 1 demonstrates the use of two rating scales applied at various
points during therapy (here the BDI and the BRMS). At the start of therapy, the
262 R.-D. Stieglitz and H.-P. Volz
Cut-Off-Value
BDI
severe
SEVERITY
moderate
BRMS
mild
no depressive syndrome
1 4 6 8 12 14 16 18 end
Therapy in number of weeks
Fig. 1 Course of therapy: data obtained with the BDI and the BRMS
example shows relatively little agreement between respective scores of self- and
observer-rating scales. This is a widely observed phenomenon, and agreement
between both data sources usually improves only during the course of treatment. It
is further recognizable that at therapy-end, both rating scales provide equally clear
data in relation to therapy success (Möller 2009).
80
70
60
50
T-Score
40
Discharge
30
Admission
20
10
0
ParHal Depres Psyorg Mani Host Veget Apa Obs
• At the level of test theory, no approaches to measuring change exist which are
included in the construction of scales. Another point of concern is the fact that
classical test theory, which forms the basis for all scale development, offers
no scope for weighting symptoms differently. As a result, symptoms such as
disturbed concentration and suicidal behavior are afforded equal weight in the
sum score of a scale.
• For certain domains, there remains a continued shortage of suitable instruments
(e.g., for the assessment of compliance).
• Another general deficiency in the field of diagnosis is the absence of approaches
to evidence based assessment (EBA; Hunsley and Mash 2007; analogous to
evidence based medicine) as well as the lack of explicit guidelines for diagnosing
psychiatric disorders at syndrome and categorical levels.
References
Addington D, Addington J, Schissel B. A depression rating scale for schizophrenics. Schizophr Res.
1990;3:247–51.
Adli M, Berghöfer A, Linden M, Helmchen H, Müller-Oerlinghausen B, Mackert A, Stamm T.
Bauer, M. Effectiveness and feasibility of a standardized stepwise drug treatment regimen
algorithm for inpatients with depressive disorders: Results of a 2-year observational algorithm
study. J Cin Psychiatry; 2002;63:782–790.
Aitken RCB. Measuring of feelings using analogue scales. Proc R Soc Med. 1969;62:989–993.
AMDP, CIPS. Rating scales for psychiatry. Weinheim: Beltz; 1990.
Andreasen NC. Scale for the assessment of negative symptoms (SANS). Iowa City: University of
Iowa; 1981.
Andreasen NC. Scale for the assessment of positive symptoms (SAPS). Iowa City: University of
Iowa; 1984.
Andreasen NC, Carpenter WT, Kane JM, et al. Remission in schizophrenia: proposed criteria and
rational for consens. Am J Psychiatry. 2005;162:441–9.
Angst J, Bech P, Boyer P, Bruinvels J, Engel R, Helmchen H, Hippius H, Lingjaerde O, Racagni G,
Saletu B, Sedvall G, Silverstone JT, Stefanis CN, Stoll K, Woggon B. Consensus conference
on the methodology of clinical trials of antidepressants, Zurich, march 1988: report of the
consensus committee. Pharmacopsychiat. 1989;22:3–7.
Angst J, Bech P, Bobon D, Engel R, Hippius H, Janzen GJ, Lecrubier Y, Lingjaerde O, Möller HJ,
Montgomery SA, Paes de Sousa M, Rossi A, Saletu B, Sedvall G, Stefanis C, Stoll KD, Woggon
B. Report on the third consensus conference on the methodology of clinical trials with
antipsychotic drugs. Pharmacopsychiat. 1991;24:149–52.
Angst J, Bech P, Bruinvels J, Engel RR, Ferner U, Guelfi JD, Lingjaerde O, Müller-Oerlinghausen
B, Paes de Sousa M, Paykel E, Rimon R, Rzewuska M, Saletu B, Spiegel R, Stassen HH, Stoll
KD, Wiesel FA, Woggon B, Zvolsky P. Report on the fifth consensus conference: methodology
of long-term clinical trials in psychiatry. Pharmacopsychiat. 1994;27:101–7.
Antony MM, Barlow DH, editors. Handbook of assessment and treatment planning for
psychological disorders. 2nd ed. New York: Guilford; 2010.
Ayearst LE, Bagby RM. Evaluating the psychometric properties of psychological measures.
In: Antony MM, Barlow DH, editors. Handbook of assessment and treatment planning for
psychological disorders. 2nd ed. New York: Guilford; 2010. p. 23–61.
Ballenger JC. Clinical guidelines for establishing remission in patients with depression and anxiety.
J Clin Psychiatry. 1999;50(suppl 22):29–34.
Barnes TR. The Barnes akathisia rating scale – revisited. J Psychopharmacol. 2003;17:365–70.
Barnett JH, Eobbins TW, Leeson VC, Sahakian BJ, Joyce EM, Blackwell AD. Assessing cognitive
function in clinical trials of schizophrenia. Neurosci Biobehav R. 2010;34:1161–77.
Psychopathology: Rating Scales for the Assessment of Mental Status 265
Bech P. Rating scales for psychopathology, health status and quality of life. Berlin: Springer; 1993.
Bech P, Rafaelsen OJ. The melancholia scale: development, consistency, validity and utility. In:
Sartorius N, Ban TA, editors. Assessment of depression. Berlin: Springer; 1986. p. 259–69.
Bech P, Rafaelsen OJ, Kramp P, Bolwig TG. The mania rating scale: scale construction and inter-
observer agreement. Neuropharmacology. 1978;17:430–1.
Bech P, Rasmussen NA, Olsen LR, Noerholm V, Abildgaard W. The sensitivity and specificity of
the major depression inventory, using the present state examination as the index of diagnostic
validity. J Affect Dis. 2001;66:159–64.
Beck AT, Steer RA. Beck anxiety inventory manual. San Antonio: The Psychological Corporation;
1993.
Beck AT, Steer RA, Brown GK. Beck depression inventory-II (BDI-II). San Antonio: Psychological
Corporation; 1996.
Broome MR, Bottlender R, Rösler M, Stieglitz RD. The AMDP system. Manual for the assessment
and documentation of psychopathology in psychiatry 9th ed. Göttingen: Hogrefe; 2018.
Brown TA, Barlow DH. Anxiety and related disorders interview schedule for DSM-5 (ADIS-5) –
adult version. New York: Oxford University Press; 2014.
Chambless DL, Caputo GC, Bright P, Gallagher R. Assessment of fear in agoraphobics: the body
sensations questionnaire and the agoraphobic cognition questionnaire. J Consult Clin Psychol.
1984;52:1090–7.
Chambless DL, Caputo GC, Jasin FE, Gracley EJ, Williams C. The mobility inventory for
agoraphobia. Behav Res Ther. 1985;23:35–44.
Chouinard G, Margolese HC. Manual for the extrapyramidal symptom rating scale (ESRS).
Schizophr Res. 2005;76:247–65.
Clayton AH. Recognition and assessment of sexual dysfunction associated with depression. J Clin
Psychiatry. 2001;62(Suppl 3):5–9.
Clayton AH, McGarvey EL, Clavet GJ. The changes in sexual functioning questionnaire (CSFQ):
development, reliability, and validity. Psychopharmacol Bull. 1997;33:731–45.
Conner KM, Davidson JRT, Churchill LE, Sherwood A, Foa E, Weisler RH. Psychomatric
properties of the social phobia inventory (SPIN): a new self-rating scale. Br J Psychiatry.
2000;176:379–86.
Cox JL, Holden JM, Sagovsky R. Detection of postnatal depression. Development of the 10-item
Edingurgh postnatal depression scale. Br J Psychiatry. 1987;150:782–6.
Day JC, Wood G, Dewey M, Bentall RP. A self-rating scale for measuring neuroleptic side-effects.
Validation in a group of schizophrenic patients. Br J Psychiatry. 1995;166:650–3.
Derogatis LR. SCL-90-R, administration, scoring & procedures manual-II for the R(evised)
version and other instruments of the psychopathology rating scales series. Towson: Clinical
Psychometric Research; 1994.
Derogatis LR. The Derogatis interview for sexual functioning (DISF/DISF-SR): an introductory
report. J Sex Marital Ther. 1997;23:291–304.
First MB, Williams JBW, Karg RS, Spitzer RL. Structured clinical interview for DSM-5©
disorders – clinical version (SCID-5-CV). Washington, DC: American Psychiatric Associa-
tion; 2016a.
First MB, Williams JBW, Smith Benjamin L, Spitzer RL. Structured clinical interview for DSM-5©
disorders – personality disorders (SCID-5-PD). Washington, DC: American Psychiatric
Association; 2016b.
Goodman WK, Price LH, Rasmussen SA, Mazure C, Fleischmann RL, Hill CL, Heninger GR,
Charney DS. The Yale-Brown obsessive compulsive scale. I. Development, use, and reliability.
Arch Gen Psychiatr. 1989;46:1006–11.
Hamilton M. The assessment of anxiety states by rating. Br J Med Psychol. 1959;32:50–5.
Hamilton M. A rating scale for depression. J Neurol Neurosurg Psychiatry. 1960;23(1):56–62.
Harnisch S, Schade-Brittinger S, Rief W. Assessing adverse reactions in clinical trials [Deutsch].
Dtsch Med Wochenschr. 2012;137:1421–5.
Hawley C, Fineberg N, Roberts A, Baldwin D, Sahadevan A, Sharman V. The use of the Simpson
Angus scale for the assessment of movement disorder: a training guide. Int J Psychiatry Clin
Pract. 2003;7:349–2257.
266 R.-D. Stieglitz and H.-P. Volz
Honigfeld G, Gillis RD, Klett JC. NOSIE. Nurses‘ observation scale for inpatient evaluation. In:
Guy W, editor. ECDEU assessment manual for psychopharmacology (rev.ed.). Rockville:
National Institute of Mental Health; 1976. p. 265–73.
Hunsley J, Mash EJ. Evidence-based assessment. Ann Rev Clin Psychol. 2007;3:29–51.
Inada T, Yagi G, Gardos G. Inter-rater reliability and the drug-induced-extrapyramidal symptoms
scale (DIEPSS). Eur Neuropsychopharmacol. 1996;6:62.
Inada T, Beasley CM Jr, Tanaka Y, Walker DJ. Extrapyramidal symptom profiles assessed with the
drug-induced extrapyramidal symptom scale: comparison with western scales in the clinical
double-blind studies of schizophrenic patients treated with either olanzapine or haloperidol. Int
Clin Psychopharmacol. 2003;18:39–48.
Kay SR, Fiszbein A, Opler LA. The positive and negative syndrome scale (PANNS) for schizo-
phrenia. Schiz Bull. 1987;13:261–76.
Keefe R. Assessment scales in schizophrenia. London: Springer Health Care; 2012.
Kessler RC, Ustün TB. The World Mental Health (WMH) Survey Initiative Version of the World
Health Organization (WHO) Composite International Diagnostic Interview (CIDI). Int J
Methods Psychiatr Res. 2004;13:93–121.
Knol W, Keijsers CJ, Jansen PA, van Marum RJ. Systematic evaluation of rating scales for drug-
induced parkinsonism and recommendations for future research. J Clin Psychopharmacol.
2010;30:57–63.
Kobak KA, Engelhardt N, Williams JBW, Lipsitz JD. Rater training in multicenter clinicla trials:
issues and recommendations. J Clin Psychopharmacol. 2004;24:113–7.
Kobak KA, Lipsitz JD, Williams JBW, Engelhardt N, Bellew KM. A new approach to rater training
and certification in a multicenter clinical trial. J Clin Psychopharmacol. 2005;25:407–12.
Liebowitz MR. Social phobia. Mod Probl Pharmacopsychiatry. 1987;22:141–73.
Linden M. How to define, find and classify side effects in psychotherapy: from unwanted events to
adverse treatment reactions. Clin Psychol Psychother. 2013;20:286–96.
Lingjaerde O, Ahlfors UG, Bech P, Dencker SJ, Elgen K. The UKU side effect rating scale. A new
comprehensive rating scale for psychotropic drugs and a cross-sectional study of side effects in
neuroleptic-treated patients. Acta Psychiatr Scand Suppl. 1987;334:1–100.
Lorr M, Klett JC. Inpatient multidimensional psychiatric scale (IMPS). Manual. Palo ALTO:
Consulting Psychologists Press; 1967.
Mason BJ, Kocsis JH, Leon AC, et al. Measurement of severity and treatment response in
dysthymia. Psychiatic Ann. 1993;23:625–31.
McGahuey CA, Gelenberg AJ, Laukes CA, Moreno FA, Delgado PL, McKnight KM, Manber R.
The Arizona sexual experience scale (ASEX): reliability and validity. J Sex Marital Ther.
2000;26:25–40.
McNair DM, Lorr M, Droppleman LF. Manual for the profile of mood sates (POMS). San Diego:
Educational Testing Services; 1971.
Möller HJ. Standard rating scales in psychiatry: methodological basis, their possibilities and
limitations and description of important rating scales. World J Biol Psychiatry. 2009;10:6–26.
Montgomery SA, Åsberg M. A new depression rating scale designed to be sensitive to change.
Br J Psychiatry. 1979;134:382–9.
O’Donavan C. Achieving and sustaining remission in depression and anxiety disorder: introduc-
tion. Can J Psychiatr. 2004;49(suppl. 1):5–9.
Ogles BM. Measuring change in psychotherapy research. In: Lambert MJ, editor. Handbook of
psychotherapy and behavior change. 6th ed. New York: Wiley; 2013. p. 134–66.
Olsen LR, Jensen DV, Noerholm V, Martiny K, Bech P. The internal and external validity of the
major depression inventory in measuring severity of depressive states. Psychol Med.
2003;33:351–6.
Østergaard SD, Bech P, Trivedi MH, Wisniewski SR, Rush AJ, Fava M. Brief, unidimensional
melancholia rating scales are highly sensitive to the effect of citalopram and may be have
biological validity: implications for the research domain criteria (RDoC). J Affect Dis.
2014;163:18–24.
Psychopathology: Rating Scales for the Assessment of Mental Status 267
Overall JE, Gorham DR. 047 BPRS. Brief psychiatric rating scale. In: Guy W, editor. ECDEU
assessment manual for psychopharmacology (rev.ed.). Rockville: National Institute of Mental
Health; 1976. p. 157–69.
Rush AJ, First MB, Blacker D. Handbook of psychiatric measures. 2nd ed. Washington, DC:
American Psychiatric Association; 2010.
Schennach R, Obermeier M, Seemüller F, Jäger M, Schmauss M, Laux G, Pfeiffer H, Naber D,
Schmidt LG, Gaebel W, Klosterkötter J, Heuser I, Maier W, Lemke MR, Rüther E, Klingberg S,
Gastpar M, Riedel M, Möller HJ. Evaluating depressive symptoms in schizophrenia: a
psychometric comparison of the Calgary depression scale for schizophrenia and the Hamilton
depression rating scale. Psychopathology. 2012;45:276–85.
Sheehan DV. The anxiety disease. New York: Charles Scribner & Sons; 1983.
Sheehan DV, Lecrubier Y, Sheehan KH, Amorim P, Janavs J, Weiller E, Hergueta T, Baker R,
Dunbar GC. The Mini-International Neuropsychiatric Interview (M.I.N.I.): the development
and validation of a structured diagnostic psychiatric interview for DSM-IV and ICD-10. J Clin
Psychiatry. 1998;59(Suppl 20):22–33.
Simpson GM, Angus JW. A rating scale for extrapyramidal side effects. Acta Psychiatr Scand
Suppl. 1970;212:11–9.
Snaith RP, Hamilton M, Morley S, Humayan A, Hargreaves D, Trigwell P. A scale for the
assessment of hedonic tone the Snaith-Hamilton pleasure scale. Br J Psychiatry.
1995;167:99–103.
Stieglitz RD. Multimodal assessment (including triangulation). In: Fernandez-Ballesteros R, editor.
Encyclopedia of psychological assessment. London: Sage; 2003. p. 606–10.
Stieglitz RD, Haug A, Fähndrich E, Rösler M, Trabert W. Comprehensive psychopathological
assessment based on the Association for Methodology and Documentation in Psychiatry,
(AMDP) system: development. Methodological foundation, application in clinical routine,
and research. Front Psych. 2017;8:45. https://doi.org/10.33389/fpsyt.2017.00045.
van Riezen H, Segal M. Comparative evaluation of rating scales for clinical psychopharmacology.
Amsterdam: Elsevier; 1988; 1988
Ventura J, Lukoff D, Nuechterlein KH, Liberman RP, Green MF, Shaner A. Brief Psychiatric Rating
Scale (BPRS) expanded version (4.0): scales, anchor points, and administration manual. Int J
Methods Psych Res. 1993;3:227–43.
Wing JK, Babor T, Brugha T, Burke J, Cooper JE, Giel R, Jablenski A, Regier D, Sartorius N.
SCAN. Schedules for clinical assessment in neuropsychiatry. Arch Gen Psychiatry.
1990;47:589–93.
Young RC, Biggs JT, Ziegler VE, Meyer DA. A rating scale for mania: reliability, validity and
sensitivity. Br J Psychiatry. 1978;133:429–35.
Zajecka J, Mitchell S, Fawcett J. Treatment-emergent changes in sexual function with selective
serotonin reuptake inhibitors as measured with the Rush sexual inventory. Psychopharmacol
Bull. 1997;33:755–60.
Xenobiotic Interactions in
Psychopharmacotherapy: Classification and
Handling
Ekkehard Haen
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 270
Xenobiotic Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 272
Potentially Inadequate Medication (PIM), Potentially Inadequate Prescription (PIP) . . . . . . . . . 272
Classification of Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 273
Pharmacodynamic Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 275
Pharmacokinetic Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 275
PD1 Interactions (So-Called “Double Prescriptions”) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 275
PD2 Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 278
PD3 Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 280
PD4 Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 282
PK1 Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283
PK2 Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 284
PK3 Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 287
PK3c Interactions: Prodrugs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 292
Number of Potential Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 294
How to Handle Xenobiotic Interactions? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 295
Data Bases to Check a Medication for Risk of Interaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 295
Drug Information Services (AID) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 297
Therapeutic Drug Monitoring . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 297
The Clinical Pharmacological TDM Report . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 299
KONBEST . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 299
AMBEW . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 300
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 301
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 301
E. Haen (*)
Clinical Pharmacology, Institut AGATE gGmbH Pentling, Pentling, Germany
Department of Psychiatry and Psychotherapy, Department of Pharmacology and Toxicology,
University of Regensburg, Regensburg, Germany
e-mail: ekkehard.haen@klinik.uni-regensburg.de
Abstract
Drug-drug-interactions (DDI) are a major topic in programs for continuous
medical education (CME). Lectures on this topic usually attract a great audience.
Many physicians are afraid of being trapped in charges with malpractice. The idea
that a drug-drug interaction is evidence of medical malpractice supports this
notion. Some computer databases accessible via internet feed this anxiety by
presenting a red exclamation mark if a “drug-drug interaction” is detected in a
patient’s medication. Even worse is to categorize this warning like a traffic light.
The medical problem is not the drug-drug interaction (DDI) but the adverse drug
effect (ADE) that arises from a drug-drug interaction in the individual patient!
DDI belong to routine medical practice, and it is often impossible to avoid them.
Moreover, they do not just occur between drugs but between any kind of foreign
substances (xenobiotica), such as food as well as legal (tobacco smoke, caffeine,
alcohol) and illegal drugs. Therefore, the medical challenge is not to just avoid
any interaction. Instead, the physician faces the question how to proceed with
drug treatment in the presence of such interactions. Based on their medical
education, the physician has to judge first of all by themself whether there is a
risk for interactions in the prescription they are planning for an individual patient.
The classification of interactions into just seven categories proposed in this
chapter might hereby help as a sort of checklist. For information that is more
detailed, the physician can then consult one of the databases that address the risk
for ADE, such as PSIAC (www.psiac.de) or MediQ (www.mediq.ch). Pharma-
cokinetic interactions may be easily assessed, monitored, and controlled by
therapeutic drug monitoring (TDM). Besides these tools, it is important to keep
in mind that nobody knows everything; even physicians do not know everything.
So take pride in asking someone who might help. For this purpose, AGATE offers
its drug information service AID (www.amuep-agate.de). Just good for nothing
are computer programs without any kind of medical basis that judge prescriptions
without taking into account a patient’s individual peculiarities. In case these types
of programs produce red exclamation marks or traffic lights to underline their
judgment, they might even work contrapuntal by just eliciting insecurity and
terror.
Introduction
“Drug-Drug Interactions” (DDI) is a very exciting and in many instances and for
many persons alarming topic. Lectures on this topic usually attract a great audience.
Physicians are afraid one of their patients might be harmfully caught by a drug-drug
interaction. The idea that a drug-drug interaction is evidence of medical malpractice
supports this notion. Computer databases accessible via Internet feed this anxiety by
presenting a red exclamation mark if a drug-drug interaction is detected in a patient’s
medication. Even worse is to categorize this warning like a traffic light without
taking into account specific aspects of the individual case: Red ¼ risk of a very
Xenobiotic Interactions in Psychopharmacotherapy: Classification and Handling 271
2018
14 (91)
13
12 i = (n2-n)/2 (66)
number of prescribed drugs
11 (55)
10 (45)
9 (36)
8 (28)
7 (21)
6 (15)
5 (10)
4 ( 6)
3 ( 3)
2 ( 1)
1 (0)
0
0 5 10 15 20 25
proportion of patients [%]
might not be at risk at all by the same DDI. This chapter is intended to explain the
facts, to organize all occurring drug-drug interactions into just seven different
categories, and to compile recommendations on how to handle interactions.
Xenobiotic Interactions
To begin with, it has to be stated that the term “drug-drug interaction” is incorrect.
The body does not care, whether a foreign substance (i.e., a “xenobiotic”) adminis-
tered into the body, is intended
The body treats all of these substances in the same way; all of them therefore give
rise to interactions. Consequently, the correct term is “interaction among xenobi-
otics” or “xenobiotic interactions.”
One fascinating aspect of psychopharmacology is that patients suffering from a
psychiatric disease also suffer from all diseases the apparently sane population can
suffer as well. So, a psychopharmacologist must not be familiar with just
psychopharmaca but with all other pharmaca, too.
for example, contains 83 such active substances available on the German drug
market with a particular risk for people >65 years of age (Holt et al. 2010). These
lists can be easily integrated in computer software to warn a prescribing physician;
however, neither the lists nor the respective software addresses the individual case.
The following classification of PIM has been suggested in 2018 (Haen et al.
2018a), prescriptions bearing a risk for xenobiotic interactions are potentially inad-
equate medications of type 7 (PIM7):
Other medication errors are classified as PIM even if the drugs itself have a
positive risk-benefit evaluation when properly used:
• PIM4: Inappropriate duration of use: Drugs are prescribed not long enough to
become effective or longer than necessary (e.g., antidepressants for one week or
less, benzodiazepines for weeks).
• PIM5: inappropriate application (e.g., certain penicillins per os) or inappropriate
route of administration (e.g., sustained-release capsules for patients with prob-
lems in swallowing).
• PIM6: Underdosing/overdosing: The drug should be applied in a dosage or
frequency to bring the concentration in the individual patient over the lower limit
but not over the upper limit of the therapeutic reference range (TRR) (Hiemke
et al. 2017).
• PIM7: Xenobiotic interaction. Prescriptions bearing a risk for xenobiotic inter-
actions. To this type of PIM also belong “double prescriptions,” that is, the
prescription of two pharmacologically identical drugs (pharmacodynamics inter-
action type PD1, see below).
• PIM8: Contraindication: A drug with an otherwise positive risk-benefit evalu-
ation is inappropriate in certain conditions (e.g., pregnancy, young or old age,
certain comorbidities).
• PIM9: Cascades of polypharmacy: A drug is not recognized as cause of an ADE
and a new drug is prescribed against the new complain again with other health
risks.
Classification of Interactions
Xenobiotic interactions are one out of these nine classes of potentially inappropriate
medications (PIM7). The huge number of theoretically (i.e., mathematically) possi-
ble interactions and the often frightening stories of what might happen can be cut
274 E. Haen
a
pharmacodynamic pharmacokinetic
Interaction Interaction at
at the receptor distribution
Similar effects
Interaction at
elimination
Interaction
of different effects and metabolism
(prodrugs)
b
pharmacodynamic pharmacokinetic
a
PD1 PK1
b
a
PD2 PK2
b
a
PD3
b a
PK3
b
PD4 +
Prodrugs c
-
Pharmacodynamic Interactions
Pharmacokinetic Interactions
theophylline bronchodilatator
15
12
caffeine concentration (µg / ml)
0
8 12 16 20 24 4 8
time of day (clock hours)
Fig. 4 Caffeine concentration in plasma of six healthy subjects. Blood was drawn every 4 h for
24 h. (Reproduced with permission from Haen E, Frankewitsch T, Decker W, Fichtl B, Emslander
HP (1991): Plasmakonzentrationen von Methylxanthinen unter Therapie mit Theophyllin bei
chronisch-obstruktiven Atemwegserkrankungen. Atemw.-Lungenkrkh. 17, 421–423, permission
granted by Jörg Feistle, Dustri-Verlag Dr. Karl Feistle GmbH & Co. KG, 19.10.2020)
hours after people usually had their caffeinated beverages for breakfast and at lunch
time. It is not surprising that the interindividual variation is huge, there are subjects
with very high caffeine concentrations and others with almost no caffeine in their
blood. However, because of its pharmacological effects, it is important for clinical
routine to pay attention to the caffeine concentration in each individual caffeine
consumer.
A third example of PD1 interactions is the prescription of sulphonamides against
microbial infections in combination with antidiabetics. The potential of
sulphonamides to decrease blood sugar was discovered by alerted physicians in
the early fifties who noticed hypoglycemia in patients they treated for bacterial
infections. This led to the development of sulphonyl urea derivatives as oral antidi-
abetics. Today the prescription of sulphonamides is rather rare because better
effective and tolerable drugs are available. Sulphonamides, however, are still com-
ponents of drugs like cotrimoxazol (an antifolic combination of trimethoprim and
sulfamethoxazole prescribed against urinary infections) and Fansidar ® (an antifolic
combination of pyrimethamin and sulfadoxine prescribed against malaria
infections).
278 E. Haen
PD2 Interactions
PD2 interactions are the classic type of drug-drug interactions: At the level of
pharmacological receptors agonists stimulate the receptors, antagonists block them
which results in loss of effectiveness of the physiological ligand or an agonist. These
receptors are not just part of a theoretical pharmacological concept; they are real
existing biochemical structures, usually proteins. Physiologically, the receptors are
part of the information system of the body (Fig. 5) that constitutes of the nerve
K+
Gs Gq
Gi
K+ Na+
Fig. 5 The information system of the human body. Information is transmitted throughout the body
via the nervous system (depicted as bulb like nerve endings in different colors ¼ different neuro-
transmitters), the endocrine system (depicted as blood vessel, upper left corner), and the immune
system (depicted as different immune cells, lymphocyte, granulocyte, macrophage). The differently
colored dots in between these structures symbolize the various first information messengers, that is,
neurotransmitters, hormones, and cytokines, transmitting their information strictly on the scene
(neurotransmitters in the synaptic cleft) or over a more or less close or longer distance (hormones
and cytokines). The curved line at the bottom symbolizes a postsynaptic membrane or another cell
membrane bearing various receptor types. Upon interaction of a first information messenger with
one of these receptor types a second information messenger molecule is formed inside the cell or an
ion channel is opened/closed that conveys the information upon the biochemical reactions inside the
cell altering the cell function. The figure is meant as a “confusion diagram” to create an idea of the
huge and complex information of the human body. The dot in the blood vessel symbolizes a
psychopharmacon arriving in this physiological information system of the human body. It interacts
with the existing receptor systems thereby modifying cell functions. At present knowledge, it seems
to be a characteristic of psychopharmaca to interact not selectively with just one receptor system but
at the same time with several of these receptor systems more or less effectively. Because of this
unselectivity pharmacologists call psychopharmaca “dirty drugs”
Xenobiotic Interactions in Psychopharmacotherapy: Classification and Handling 279
system, the hormonal system, and the immune system (Haen 2018b). Each of these
systems uses messenger molecules (“first messenger”) to convey their message to a
cell, a tissue, or an organ. Upon interaction of the first messenger with one of these
receptors, the message of the molecule is transmitted into the cell.
The risk is particularly high in psychopharmacology that PD2 interactions occur:
In contrast to most of the drugs used in internal medicine psychopharmaca are not
very selective in their affinity to pharmacological “receptors”: Psychopharmaca are
pharmacologically “dirty drugs.” Both large groups of psychopharmaca, antipsy-
chotics (receptor profile see Fig. 1.4.5 in Haen 2018b), and antidepressants (Fig. 6)
interact with more than just one receptor system, and the receptor is usually blocked.
Antidepressants can block the same receptors as antipsychotics thereby additively
increasing the effect usually transmitted by this receptor (interaction of type PD2a).
An example is the anticholinergic effect that is additively increased in a combination
of the antipsychotic quetiapine and the tricyclic antidepressant amitriptyline. The
resulting clinical effect might be urinary retention, obstipation up to an ileus, or an
anticholinergic delirium. Such a situation might be even more aggravated if
biperiden is used to treat extrapyramidal symptoms (EPMS) in the patient or a
urologist has issued a prescription of a spasmolytic drug like oxybutynin or
tolterodine against urinary incontinence.
Particularly difficult to handle is a psychosis in a patient suffering from
Parkinson’s disease. Dopaminergic drugs like levodopa, an agonist at dopamine
inhibition antagonism
SR NR DR 5-HT2A H1 M1 α1 α 2
TCA + + + + + + + +
α2-antagonists - - - + + - - +
SSRI + - - - - - - -
SNRI - + - - - - - -
SSNRI + + - - - - - -
SNDRI - + + - - - - -
Fig. 6 Receptor profile of antidepressants. The table gives the information on which class of
antidepressants (TCA: Tricyclic antidepressants; SSRI: Selective serotonin reuptake inhibitor;
SNRI: Selective norepinephrine reuptake inhibitor; SSNRI: Selectivly combined serotonin and
norepinephrine reuptake inhibitor; SNDRI: Selectivly combined norepinephrine and dopamine
reuptake inhibitor) inhibits a certain neurotransmitter reuptake mechanism from the synaptic cleft
(SR: Serotonin reuptake; NR: Norepinephrine reuptake; DR: Dopamine reuptake). Some, mostly older
antidepressants antagonize receptor systems, too, which might contribute to the desired drug effects; it
definitely contributes to the adverse drug effects (5-HT2A: Serotonin receptor type 2A; H1: Histamine
receptor type 1; M1: muscarine receptor type 1; α1: alpha-adrenergic receptor type 1; α2: alpha-
adrenergic receptor type 2). Many of the older antidepressants (TCA) unselectively inhibit neurotrans-
mitter reuptake mechanisms that were not studied at the time of their development. Newer antidepres-
sants were designed on the drawing board guided by a certain idea of interaction with
neurotransmitter reuptake mechanisms or receptor systems
280 E. Haen
PD3 Interactions
Some drugs exert similar effects via different pharmacological mechanisms. Such an
example is the antihypertensive effect of drugs. Five different pharmacological
mechanisms are at present accepted for first-line treatment of hypertension (The
Task Force for the management of arterial hypertension of the European Society of
Cardiology (ESC) and the European Society of Hypertension (ESH) 2018):
b
am ount of f ood vitamin K content of 100 g food [µg]
containing 1 µ g vitam in K [g]
0
50
100
150
200
250
300
350
400
450
500
550
600
650
700
750
800
850
live
r (c
hi
0
1000
2000
3000
4000
5000
6000
7000
8000
9000
10000
11000
12000
13000
live cken )
liver r(
(ch live beef )
liver icken ) r (p
live ork )
liver (beef ) r
( m ea (calf )
t
liver pork ) m ea (beef)
mea (calf ) t (p
t ork)
mea (beef) whi
t (po te c
rk) ab
whit
e ca sau bage
sau bageb Bru
red erkrau
c t
erk sse abbag
Bru red cab raut ls s
sse prou e
cau
ls spbage liflo ts
cau routs wer
liflow b
er Chi roccoli kale
b nes
e ca (raw)
Chi roccol kale bba
nes i (ra
e ca w ge
food
bba )
ge
food
spin
s w he a
at g ch
whe pinach
at g soy erm
soy erm a flo
a flo wer lettuce
(full
Fig. 7 Vitamin K content of food. (a) Vitamin K content of 100 g various foodstuffs, cross lined
taken from (Souci et al. 2016). (b) Amount of various foodstuffs that contain 1 μg vitamin K, data
bar: Daily need (60–80 μg vitamin K1) according to the German Society of Nutrition (DGE), data
281
282 E. Haen
PD4 Interactions
Some drugs give rise to interactions because their effects interact though being
completely different. An example is the long-known interaction between cardiac
glycosides (such as digoxin and digitoxin) and saluretic drugs (such as hydrochlo-
rothiazide or chlortalidone among others) or laxatives (such as bisacodyl or sorbitol
among others). Cardiac glycosides block the ATP-dependent sodium/potassium
pump that plays an important role for repolarization of the membrane potential in
the electrical conduction system of the heart (Fig. 8). Important ADEs of cardiac
glycosides include all kinds of cardiac arrhythmia (mechanism 1). Saluretic drugs
and laxatives lead to a loss of potassium thereby lowering the extracellular potassium
concentration (mechanism 2). A low extracellular potassium concentration, how-
ever, increases the risk for cardiac arrhythmia, thereby increasing toxicity of cardiac
glycosides (interaction of both mechanisms, Fig. 9). Because cardiac glucosides
have been replaced nowadays in most of their indications by ACE-1 inhibitors or AT-
II1A receptor antagonists, the clinical relevance of this particular type of PD4
interaction example is almost negligible today. However, low potassium concentra-
tions increase the risk of all kinds of cardiac arrhythmias. Prolongation of the QT-
interval in the ECG has become a clinically very important cardiac arrhythmia,
because a long QT-Interval bears the risk of a sudden onset of tachycardia, a so-
called torsade-de-pointes tachycardia (TdP) (Haen 2020). TdP tachycardia may be
self-limiting within a few seconds or it leads to nausea, palpitation, syncope,
ischemia triggered grand-mal seizure, or even sudden cardiac death. Among others,
there are almost no psychopharmaca not bearing the risk for QT prolongation
(Wenzel-Seifert et al. 2011). The list of QT prolonging drugs becomes longer and
Xenobiotic Interactions in Psychopharmacotherapy: Classification and Handling 283
[K+] 2 K+
[Ca2+] P
[Na+] U
DM
P
3 Na+
exchange site
2Na+ Ca2+
Fig. 8 Mode of action of cardiac glycosides. The action potential on a cell membrane is driven by
influx and efflux of ions. A fast sodium influx drives the steep increase, a slow potassium efflux the
repolarization. The original intra-/extracellular ion gradient (intracellular: low sodium/high potas-
sium concentration, extracellular: high sodium/low potassium concentration) is restored by an
energy (ADP) consuming ion pump. It transports three sodium ions/ATP molecule outwards of
the cell, by bringing two potassium ions/ATP molecules back inside. Cardiac glycosides block this
ion pump thereby lowering the intracellular potassium concentration
longer, a continuously updated list of these drugs is available via Internet (Woosley
et al. 2020). Therefore, type PD4 interactions are nowadays clinically more relevant
than ever.
PK1 Interactions
saluretics
laxatives cardiac glycosides
PD4 QTc
Fig. 9 Scheme of type PD4 interactions (interaction of drug mode of actions). Type 4 interactions
were originally described for saluretics or laxatives and cardiac glycosides. Important ADEs of
cardiac glycosides include all kinds of cardiac arrhythmia (mechanism 1). Saluretic drugs and
laxatives lead to a loss of potassium thereby lowering the extracellular potassium concentration
(mechanism 2). A low extracellular potassium concentration, however, increases the risk for cardiac
arrhythmia thereby increasing toxicity of cardiac glycosides (interaction of both mechanisms).
Because cardiac glucosides have been replaced nowadays in most of their indications by ACE-1
inhibitors or AT-II1A receptor antagonists the clinical relevance of this particular type PD4 interac-
tion example is almost negligible today. However, low potassium concentrations increase the risk of
all kinds of cardiac arrhythmias. Prolongation of the QT-interval in the ECG has become a clinically
very important cardiac arrhythmia, because a long QT-Interval bears the risk of a sudden onset of
tachycardia, a so-called torsade-de-pointes tachycardia (TdP). Almost all psychopharmaca bear the
risk for QT prolongation. For this reason PD4 type interaction are still clinically highly important
time influences the absorbed amount of drugs which might be accelerated or slowed
down by other drugs (e.g., laxatives and opioids, respectively). A clinically relevant
example for the effect of food in psychopharmacology is the antipsychotic
ziprasidone: If ziprasidone is orally administered into an empty stomach (e.g.,
after an overnight fast of 8 h), the resulting drug peak plasma concentration is
50% lower compared to be taken together with food (Fig. 10) (Hamelin et al. 1998).
Besides these few general considerations, it is very difficult to foresee the effect
food may have on the uptake of a particular active agent into the body. It has to be
evaluated for each active agent separately. Wunderer (1998) gives an excellent
overview, and Welling published an English review on the topic (Welling 1996).
Oral administration of the agents concerned 2 h apart may alleviate or even prevent
PK1 type interactions (Fig. 10).
PK2 Interactions
The typical and up to now almost exclusively displayed example of a PK2 type
interaction is the displacement of a drug from its plasma protein binding (PPB).
Among others, proteins in body fluids (most of all albumin and alpha1-acid-
Xenobiotic Interactions in Psychopharmacotherapy: Classification and Handling 285
120
100
80
60
40
20
0
after an overnight fast of 8h directly after standard breakfast 2h after standard breakfast
Germany 2018) with 5% of the total concentration as active fraction. If another drug
(e.g., valproic acid) displaces amitriptyline by reducing the bound fraction, for
example, to 90% of the total concentration, the free or active fraction is increased
to 10% of the total concentration, that is, the active concentration is doubled!
This example sounds very dramatic – and, it is still very dramatically presented
nowadays! Nevertheless, the problem for drug therapy is much more complicated
and lies in a different area: Human bodies are no in vitro reaction jars! Figure 11
displays an example with real data that demonstrates the order of magnitude of what
800
600
400
200
0
diazepam monotherapy diazepam + valproic acid diazepam + valproic acid
in vitro human body
Fig. 11 Schematic illustration of plasma protein binding (PPB) as PK2 type interaction. 98% of the
total diazepam concentration is normally bound to plasma albumin (Roche Pharma AG Grenzach-
Wyhlen, Germany 2013), that is, in a total diazepam concentration of 1000 ng/ml 980 ng/ml are
bound to plasma proteins, 20 ng/ml (¼2% of the total diazepam concentration) are free (active
fraction, left bar). If a drug like valproic acid displace diazepam from these proteins thereby
increasing the free diazepam concentration up to 25 ng/ml, the bound fraction is just lowered by
0.5% to 97.5% of total concentration, but the active fraction of the drug is increased by 25% of the
original free concentration (Desitin Arzneimittel GmbH Hamburg, Germany 2016) (situation in an
in-vitro reaction jar, middle panel). In a human body with undisturbed function of the excretion
organs, most drugs are eliminated with first order elimination kinetics, that is, the molecules of the
increased free concentration are ready to enter hepatocytes and tubulocytes, in which they are
metabolized or from which they are filtrated into renal tubules, that is, they will be eliminated until
the steady state concentration with 20 ng/ml free concentration is re-established. The difference is
now that 20 ng/ml free concentration represent 2.5% of total concentration, which means that the
total diazepam concentration is lowered to 800 ng/ml (situation in a human body, right panel). If
drug therapy is adapted to the individual need of a patient by therapeutic drug monitoring (TDM),
which nowadays usually determines the total concentration of a drug, this lowered total drug
concentration must not be raised by increasing the prescribed dosage.
Xenobiotic Interactions in Psychopharmacotherapy: Classification and Handling 287
is going on. As mentioned before, only free molecules enter hepatocytes and
tubulocytes, in which they are metabolized or from which they are filtrated into
renal tubules. In an otherwise healthy body with undisturbed elimination organs,
such a free concentration will increase just for a short period, if at all, because most
drugs follow first order elimination kinetics, that is, the amount of drug eliminated is
proportional to its free concentration (Fichtl 2001). Therefore, the increased free
concentration is immediately eliminated down to the normal free concentration.
Therefore, the patient will not experience an increased, potentially adverse drug
effect. However, if the patient’s drug therapy is individualized by therapeutic drug
monitoring (TDM), it has to be kept in mind that TDM does not usually distinguish
between free and bound fraction, and the therapeutic reference ranges refer to the
total drug concentration. In case of displacement from plasma protein binding total
drug concentration decreases with the normal free drug concentration being now the
higher free fraction, that is, the drug now has another lower therapeutic reference
range in this patient, his dosage must not be increased.
This effect is the more important the higher the bound fraction of a drug to plasma
proteins. Up to now no clinical evidence is available as to its relevance in daily
routine. Oie and Levy reported in 1979 an animal experiment with rats (Oie and
Levy 1979): Endogenous bilirubin is bound 99.9% to albumin. Sulfonamides
displace bilirubin from its protein binding, and Oie and Levy used sulfisoxazole in
their experiments. Upon rapid injection of sulfisoxazole free bilirubin concentration
rapidly increased accompanied by a drop of total bilirubin concentration within
15 min. Free bilirubin concentration sharply decreased again within the next
15 min, the increase lasted for 60 min in total. Free bilirubin concentration did not
increase, if sulfisoxazole was slowly infused; nevertheless, total bilirubin concen-
tration still decreased slowly. If sulfisoxazol concentration slowly increases during
infusion the bilirubin is so slowly displaced from its protein binding sites that there is
enough time to redispose in the body into other compartments.
PK3 Interactions
PK3 type interactions include all kinds of interactions involving metabolic and or
elimination mechanisms. Xenobiotic interactions in the kidneys were within the
focus of interest in former decades, namely the 1970s and 1980s. Nevertheless, they
are still clinically very relevant! Examples (e.g. case report 1) are interactions of
lithium ions (prescribed for bipolar affective disorders) and all kinds of active agents
stimulating the excretion of sodium such as diuretics, laxatives, and inhibitors of the
renin-angiotensin-aldosterone-system (RAAS, such as angiotensin-converting-
enzyme inhibitors (ACE inhibitors) and angiotensin-II1A receptor antagonists).
Everything that leads to a decrease in sodium concentration is eventually coupled
to an increased lithium retention by the kidneys leading to an increase in clinical
effectiveness of lithium (PK3a interaction). The onset of this effect might take up to
six weeks (Ortlieb et al. 2012).
288 E. Haen
Case Report 1
A 58-years-old man was diagnosed 20 years ago with bipolar disease. For the
last 19 years the same community specialist for psychiatry and psychotherapy
is attending him. The patient has further health problems including lumbar
osteochondrosis, spinal disc herniation, arterial hypertension, and
nephrolithiasis. With regard to drug-drug interaction the diagnosis “arterial
hypertension” need special attention, because those patients usually have
antihypertensive prescriptions. However, the psychiatrist just filed amitripty-
line, desipramine, nitrazepame, and lithium carbonate, that is, the drugs she
had prescribed.
In the beginning of December, the psychiatrist noted that the psychopatholog-
ical situation of her patient had deteriorated. So she decided to increase the lithium
dose from 2 tablets/day up to 3 tablets/day, an increase by 50%. Before the patient
started his new dosage the physician draw blood for lithium quantification:
0.88 mval/l. The next entry into the patient’s file dates from next February:
Apparently, the psychopathological situation had not improved. The psychiatrist
decided to stop the double medication of amitriptyline and desipramine and
switched the patient to just one antidepressant (mirtazapine). Two weeks later
she saw the patient again: The switch to mirtazapine had not changed the
psychopathological situation. The patient reported diarrhea and a strange, metallic
taste since beginning of February. His general practitioner had prescribed “tablets”
(no further details) against diarrhea. The next file entry dates to April 28, when the
patient was admitted to a hospital emergency room because of increasing drows-
iness and movement problems. The patient was diagnosed with increasing som-
nolence and pneumonia and transferred to the intensive care unit (ICU).
Laboratory test revealed a white blood cell count (WBC) of 17.100 Zellen/μl,
C-reactive protein (CRP) of 39.1 mg/l, and lithium concentration of 3.24 mval/l.
In the intensive care unit, the complete medication was recorded: In addition to
the abovementioned psychopharmaca, the patient was on enalapril (an ACE
inhibitor, ACE ¼ angiotensin-converting enzyme), candesartan (an angiotensin-
II-receptor1A inhibitor) + hydrochlorothiazide (a thiazid diuretic), piretanide (a
loop diuretic), and nitrendipine (a calcium channel blocker). This means the
patient was on at least four drugs that are known to increase lithium concentration
by type PK3a interactions.
We do not know when this polymedication started. Judged from the
patient’s complaints it must have been by end of January or beginning of
February. For the case, it was only important that the psychiatrist basically
treated the patient according to the rules. She drew blood, before she increased
the lithium dose. Because of the linear correlation between dose and concen-
tration, she was allowed to expect a lithium concentration around 1.32 mval/l
after dose increase. She just had to confirm this calculation. However, it was
Christmas time, so the psychiatrist was on holiday. When she had returned, the
(continued)
Xenobiotic Interactions in Psychopharmacotherapy: Classification and Handling 289
patient was not available. In February, both psychiatrist and patient had
forgotten. Later the psychiatrist and the general practitioner argued who was
responsible for controlling the lithium concentration.
Conclusion: Communicate with colleagues that are involved in the care of
your patients. Note all medications of your patient in your files not just the
medication of your speciality. Lithium concentration might increase in com-
bination with a multitude of drugs. Keep in mind that lithium concentration
might increase for up to 2 months after a change in medication.
A diminished kidney function may have the same effect, whether resulting from
aging or induced by other drugs, such as cephalosporines (case report 2, Geisslinger
et al. 2020).
Case Report 2
A woman, 58 years of age, was diagnosed some 30 years ago with bipolar
disease. The same community specialist for psychiatry and psychotherapy
took care of her for 24 years, before he turned his ambulatory over to a
younger colleague. He had set the woman on lithium 10 years after the
diagnosis, which she had well tolerated now for 20 years. Shortly before
Christmas, when the problems started, the woman was on trimipramine,
hydroxyzine, and lithium carbonate. For the last 3 months, the younger
physician had tried to wean the patient from a long-standing medication of
zopiclone (that she had started because of severe sleeping problems) over
chloral hydrate to the combination of trimipramine and hydroxyzine. On
December 22, she was admitted to the emergency room of a local hospital
because of respiratory insufficiency, bronchial infection, and an unclear speech
difficult to understand. Lithium concentration in blood was 2.44 mval/l. The
patient was transferred to the intensive care unit (ICU) for 3 days, lithium was
stopped until January 4, and lithium concentration fell over 0.8 mval/l on
December 25 down to 0.0 mval/l on January 4. On December 27, the patient
had to be referred back to the ICU, where she was intubated and artificially
ventilated because of CO2 narcosis until January 4.
Explaining to the patient what had happened the hospital physicians asked
the woman, why she took all these sedating medications: “Did you want to kill
yourself?” The hospital report lists as admittance diagnosis “lithium intoxica-
tion due to drug-drug interaction.” Of course, the patient had not intended to
kill herself! Her husband wrote to the young community specialist for psychi-
atry and psychotherapy asking him with respect to the hospital report why he
had issued this prescription of sedating agents in combination with lithium. In
his reply, the community psychiatrist argues: “The drugs I had prescribed,
(continued)
290 E. Haen
PK3 type interactions in the liver are mainly due to metabolic interactions. The
body always tries to avoid the effect of a xenobiotic. One successful strategy among
others is to fasten inactivation and/or elimination: Enzymes metabolizing a xenobi-
otic increase in amount and activity. Such a process involves genetic alterations
(gene amplification, induced transcription), and it takes a couple of days to become
effective. Once metabolism is more effective, the enzymes may catalyze other
substrates than the originator as well, a mechanism that is called “enzyme induc-
tion.” Enzyme induction causes a diminished clinical effect (PK3b interaction).
“Enzyme inhibition” originates from the same phylogenetic mechanisms (Fig. 12)
besides that the originator’s affinity to the active center of the enlarged enzyme is
higher than the victim’s affinity. The victim agent is not able to displace the
originator (also called “perpetrator”) from the active center of the enzyme, so that
the latter metabolism is slowed down and clinical effectiveness increases (PK3a
interaction). In contrast to enzyme induction, enzyme inhibition becomes evident
parallel to the increase of the xenobiotic concentration in the body. Enzyme inhibi-
tion is concentration dependent: It depends on both the concentration of the origi-
nator and the victim agent. Often enzyme inhibitors are compiled in lists, for
Xenobiotic Interactions in Psychopharmacotherapy: Classification and Handling 291
enzyme induction
xenobiotic induces transcription of the genetic code
enzyme meabolises faster
other xenobiotics as well
enzyme inhibition
xenobiotic with high affinity to the active center
blocks enzyme activity even in induced enzymes
for other xenobiotics with lower affinity
Fig. 12 Schematic presentation of type PK3 interactions at the level of metabolism. Enzyme
induction: Xenobiotics induce their inactivation by metabolism, that is, transcription of the genetic
code for metabolizing enzymes is stimulated. This process takes a couple of days. The induced
enzyme is then ready to metabolize other xenobiotics with affinity to its active center, too. Enzyme
inhibition: Xenobiotics block as high affinity substrates the active center of their metabolizing
enzymes. The metabolism of other substrates with lower affinity to the active center of the enzyme
is blocked as well, depending on their affinities and concentrations. This process accompanies the
increase in substrate concentration, that is, it just takes minutes up to hours. Xenobiotics might also
block enzymes that have been induced before. Both processes are reversible over a similar time as
their activation
Case Report 3
We received a blood specimen for determination of citalopram concentration
drawn from a 30-years-old man at 07h30. The patient was on 30 mg
(continued)
292 E. Haen
If prodrugs are applied in a medication, the active agents are only formed in the
metabolism, whereas the mother substances might be more or less inactive. The
adaptation processes presented above now apply to the formation of the active agents
but with opposite consequences: Enzyme induction leads to an increase in clinical
effectiveness (PK3c+ interaction), and enzyme inhibition to a decrease in clinical
effectiveness (PK3c interaction). Among others, clinically relevant examples are
clopidogrel, an anticoagulant that inhibits the aggregation of thrombocytes, and
tamoxifen, a selective estrogen-receptor modulator used in breast cancer patients
to stop estrogen-sensitive breast cell growth.
Clopidogrel is metabolized by CYP1A2, CYP2B6, CYP2C9, CYP2C19, and
CYP3A4 (Kazui et al. 2010). The active metabolite is a thiolderivative of the
Xenobiotic Interactions in Psychopharmacotherapy: Classification and Handling 293
CYP
20 out of 26 subfamilies
known in men
isoforms
33 bekannt AA-sequence homology > 55 %
12 relevant for drug metabolism
(es)citalopram 1
1 A 2
A 6
B 6
CYP 8
2 C 9
18
19
D 6
E 1
4
3 A
5
Fig. 14 Metabolic pathway of citalopram and escitalopram. Please read the figure like a book from
left to right: (Es)citalopram is metabolized by CYP2C19 and CYP2D6
It is important to emphasize that this theoretical risk estimation does not mean,
that each of these mathematically possible interactions will also give rise to a
clinically relevant adverse drug effect (ADE). What will eventually happen in the
patient does not have anything to do with this type of theoretical risk calcuation. On
the contrary, the clinical risk evaluation reveals in some instances that one of these
mathematically possible interactions causes several different complaints of the
patient instead of just one ADE. For example, one often applied combination of
two drugs in the pharmacotherapy of psychoses, the combination of benzodiazepine
antipsychotics (e.g., clozapine) with a benzodiazepine (e.g., diazepam), may lead to
four different clinical situations: (1) Fatigue and sedation until unconsciousness and
Xenobiotic Interactions in Psychopharmacotherapy: Classification and Handling 295
coma, (2) depression of breathing until arrest, (3) hypotension, and 4. malignant
neuroleptic syndrome (MNS) (Haen 2019).
Nobody can avoid drug-drug interactions in prescriptions of more than one drug.
The interaction problem for the treating physician does not consist of avoiding an
interaction; he or she has to find a way to deal with the interaction. For this purpose,
physicians may
• Check databases
• Consult drug information services
• Quantify xenobiotic concentrations (therapeutic drug monitoring, TDM)
Many Internet sources list (iso)enzymes together with their substrates and informa-
tion of inducers and inhibitors. However, the bedside physician needs the informa-
tion the other way round: The bedside physician, the general practitioner, the
community specialist, and the pharmacist have a list of prescriptions to a patient.
Göpfert and Haen therefore started in 2006 to compile the available information in
alphabetical order of the INN (elimination pathway table). Unfortunately, the infor-
mation available via Internet is very scarcely validated by references of the interna-
tional literature. The 2017 version is available in print (Haen and Göpfert 2018). The
table is continuously updated in KONBEST and AMBEW (see below) PSIAC and,
by now independently, in KONBEST and AMBEW (see below) PSIAC. The PSIAC
table is more and more backed with international references that remarkably
decrease the number of entries. PSIAC is meant to check for medication risks. The
entries are therefore carefully double-checked to avoid overalerting. The KONBEST
and AMBEW platform is compiled for interpretation of a measured drug concentra-
tion; the elimination pathway table therefore contains all the available internet
information.
Databases accessible via internet offer an interaction check of a medication
(Quick et al. 2018). Several electronic data banks list all kind of pharmacological
and pharmaceutical data of drugs. Some of them are available by registration or as
part of electronic communication systems in outpatient practices, hospitals, and
pharmacies:
Few data banks are especially designed and compiled to yield information on
drug-drug interactions or, in a broader sense, xenobiotic interactions:
People, healthcare professionals, and patients may contact drug information services
to get information on questions regarding medications. AGATE (the working group
to support both a rational and economic drug therapy in [psychiatric] diseases,
German: Arbeitsgemeinschaft zur Unterstützung einer sowohl rationalen wie
rationellen Arzneimitteltherapie bei [psychiatrischen] Erkrankungen) is a nonprofit
assembly of healthcare professionals devoted to support a rational and economical
drug therapy (Haen and Laux 2011). It offers unbiased information on drugs
independent from any lobby interest in the healthcare system. Everybody may ask
any question and is encouraged to do so. Its budget comes from membership fees,
tutorials, scholarships, and donations free of healthcare lobby interests.
The dose-related reference range (DRR) is obtained by putting the total clearance
as mean + 1 standard deviation (SD) and mean – 1 SD into Eq. (1).
The thus defined DRR contains statistically 68.27 % of drug concentrations built
up in compliant “normal patients” under the constant dose De after the
298 E. Haen
DRR calculated with Eq. (1) is usually not too much different from the results
of Eq. (2). Differences are larger the shorter the elimination half-live of the
xenobiotic. The TDM consensus guideline suggested an elimination half-live
above 24 h as a limit (Hiemke et al. 2017) to use Eqs. (1) and (2) for shorter
elimination half-lives. However, the difference between the two equations
depends linearly from the length of the elimination half-life; therefore, it is better
to use Eq. (2) whenever applicable. Note: Eq. (2) is only mathematically valid if
the daily dose is either applied once daily or at equal time intervals within
24 hours in equal partial doses.
In clinical routine, however, physicians prescribe drugs at every possible admin-
istration schedule. In our experience, it is reasonable to use Eq. (2) for once daily
application or twice daily every 12 h. In all other cases, particularly with dosing
emphasis in the second half of the day, we use Eq. (1). The discussion goes on. If the
“correct” (whatever that is?) concentration is to be predicted, an iterative summation
is to be used with an extra term for each partial dose at its particular application time
(Endres and Haen 2020).
Unfortunately, DRC factors in Table 5 of the TDM consensus guideline (Hiemke
et al. 2017) are struck by two calculation errors: (1) The absorptive phase, i.e. Tmax,
is not taken into account. (2) The factors have not been calculated in two steps (total
clearance mean + and – one standard deviation) but after calculating the arithmetic
mean, the result has been then adapted with the coefficient of variation to the low and
Xenobiotic Interactions in Psychopharmacotherapy: Classification and Handling 299
high factors. The respective Table 1.7.3 in Klein and Haen (2018) has ameliorated
these errors (Haen et al. 2018b).
KONBEST
It is impossible for bedside physicians to keep in mind all the pharmacological data
and patient relevant information necessary to evaluate properly and comprehen-
sively a drug concentration. There are softwares to support the composition of a
clinical pharmacological TDM report: Köstlbacher (software programmer) and Haen
(algorithms, databases) created between 2008 and 2010 KONBEST, an Internet-
based platform that supports compiling clinical pharmacological TDM reports
(Köstlbacher and Haen 2008; Köstlbacher 2012). KONBEST contains pharmaco-
logical databases. Upon entering therapeutically relevant, anonymous patient data,
KONBEST sorts along pharmacological algorithms the pharmacological to the
individual patient data. The result is presented to a clinical pharmacological expert
300 E. Haen
Fig. 15 Clinical pharmacological report set up by KONBEST. The example shows KONBEST
case no. 48412. The report gives all clinically relevant information contained in a drug blood
concentration (Haen 2018d, 2012). Furthermore, the clinical pharmacological expert answers
questions formulated on the TDM request form.
who checks the information, evaluates it, and adds the therapeutic recommendation.
The finalized and issued TDM reports are stored in the KONBEST reports database;
by May 2020 this database contains some 65,000 clinical pharmacological TDM
reports. KONBEST is property of AGATE (www.amuep-agate.de).
AMBEW
AMBEW was created by Herdt & Klon (software programmers) and Haen (algo-
rithms, databases). It is available since March 13, 2017. AMBEW is property of
AGATE (www.amuep-agate.de).
Cross-References
References
Ahmad A, Shahabuddin S, Sheikh S, Kale P, Krishnappa M, Rane RC, Ahmad I. Endoxifen, a new
cornerstone of breast cancer therapy: demonstration of safety, tolerability, and systemic bio-
availability in healthy human subjects. Clin Pharmacol Ther. 2010;88(6):814–7.
ABDATA for pharmacies Avoxa – Mediengruppe Deutscher Apotheker GmbH, Eschborn. n.d.-a.
Accessible via https://abdata.de/datenangebot/abda-datenbank. Accessed 26 Jan 2020.
ABDAMED for medical practitioners Avoxa – Mediengruppe Deutscher Apotheker GmbH,
Eschborn. n.d.-b. Accessible via https://abdata.de/datenangebot/abdamed. Accessed 26 Jan
2020.
Barry PJ, O’Keefe N, O’Connor KA, O’Mahony D. Inappropriate prescribing in the elderly: a
comparison of the Beers criteria and the improved prescribing in the elderly tool (IPET) in
acutely ill elderly hospitalized patients. J Clin Pharm Ther. 2006;31(6):617–26.
Bayer Vital GmbH Leberkusen/Germany. Fachinformation (Summary of Product Characteristics)
Saroten ® retard Tabs 75 mg. (2018). issued Nov 2018.
Boehr S, Haen E. Development of an UHPLC-UV-method for quantification of direct oral antico-
agulants: apixaban, rivaroxaban, dabigatran, and its prodrug dabigatran etexilate in human
serum. Ther Drug Monit. 2017;39(1):66–76.
Corsonello A, Pranno L, Garasto S, Fabietti P, Bustacchini S, Lattanzio F. Potentially inappropriate
medication in elderly hospitalized patients. Drugs Aging. 2009;26(Suppl 1):31–9.
Desitin Arzneimittel GmbH Hamburg, Germany. Fachinformation (Summary of Product Informa-
tion) Orfiril ® magensaftresistentes (gastro-fluid resistant) Dragee 155, 300 und 600 mg. 2016,
issued Nov 2016.
Deutsches Institut für Medizinische Dokumentation und Information (DIMDI), Köln. 2020. Acces-
sible via https://www.dimdi.de/dynamic/de/arzneimittel. Accessed 26 Jan 2020.
Eckermann G, Haen E, Hiemke C, editors. PSIAC – polymedication under control. Heidelberg:
Springer; 2020. Accessible via https://www.psiac.de. Accessed 27 Jan 2020.
Endres K, Haen E. The dose-related reference range – a new approach with improved predictive
quality. Pharmacopsychiatry. 2020;53(3):140–1.
Fichtl B. Arzneistoffkonzentration im Organismus in Abhängigkeit von der Zeit: Pharmakokinetik
im engeren Sinn. In: Forth W, Henschler D, Rummel W, Förstermann U, Starke K, editors.
Allgemeine und Spezielle Pharmakologie und Toxikologie. 8th ed. München/Jena: Urban &
Fischer Verlag; 2001. p. 72–4.
Geisslinger G, Menzel S, Gudermann T, Hinz B, Ruth P, editors. Mutschler Arzneimittelwirkungen.
11th ed. Stuttgart: Wissenschaftliche Verlagsgesellschaft mbH; 2020. p. 1002.
Gex-Fabry M, Balant-Gorgia AE, Balant LP. Therapeutic drug monitoring of olanzapine: the
combined effect of age, gender, smoking, and comedication. Ther Drug Monit. 2003;25:46–53.
Haen E. Bedeutung der klinisch-pharmakologischen Befundung von Wirkstoffkonzentrations-
messungen zur Therapieleitung. Psychopharmakotherapie. 2005;12:138–43.
302 E. Haen
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 306
A Brief History of the Randomized Controlled Trial . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 307
The Role of Governmental Oversight of Medical Products . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 307
Key Characteristics of Modern Clinical Trial Designs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 309
Protection of Human Subjects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 309
Trial Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 310
Duration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 312
Participant Selection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 312
Confirmation of Diagnosis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 312
Control Group . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 313
Randomization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 313
Stratification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 314
Blinding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 314
Outcomes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 315
Study Monitoring . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 316
Statistical Considerations of RCTs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 316
Effect Size . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 316
P-value . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 317
Confidence Interval . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 317
Intention-to-Treat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319
Missing Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319
Internal Validity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 320
External Validity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 321
Challenges to Detecting Efficacy of Drugs in RCTs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 321
Placebo Response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 321
Nonadherence to Treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 324
Treatment Resistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 324
Abstract
Randomized controlled trials (RCTs) are the most important component of the
evidence base justifying the use of psychiatric medications in clinical practice.
Establishing efficacy of an investigational new drug through RCTs requires
careful consideration of factors related to trial design, patient enrollment, study
conduct, and data analysis. Each disorder has unique characteristics that require
tailoring of RCT designs in terms of duration, assessment, and treatment setting.
Placebo effects and patient nonadherence to study medication present significant
challenges to identifying truly efficacious drugs. Although RCTs have been
criticized for being vulnerable to various forms of bias, they remain the best
means for identifying treatments to relieve mental suffering. Mental health
practitioners can benefit from a deeper understanding of clinical research pro-
cedures and the standards used for evaluating a medication’s efficacy, safety, and
tolerability. This understanding can inform clinicians’ judgment regarding the
generalizability of RCT results to their clinical practice.
Introduction
Complementing the enhancements in clinical trial methodology has been the expan-
sion of the role of government in managing the public release of pharmaceuticals,
though this has been primarily a reactive process. Before governments began to take
308 B. W. Dunlop and C. Medeiros Da Frota Ribeiro
an active role, the American Medical Association formed the Council on Pharmacy
and Chemistry in 1905 as a reaction to the growth of “patent” medicines. These
formulations promised a variety of health benefits, typically without any supporting
studies. The Council would, for a fee, award the AMA’s Seal of Acceptance to the
drug if it passed their ingredient quality testing, and the Council members adjudged
the product to produce symptomatic relief. This initial step to bring some semblance
of order to the free-for-all in marketing drugs to the public was followed in 1906 by
the passage of the Pure Food and Drugs Act by the United States. The statute made it
illegal for manufacturers to print “false and misleading” statements and required
listing potentially dangerous ingredients on the product label, but overall the statute
provided minimal public health protections because government interventions could
only occur after drugs had been marketed.
The subsequent United States’ Food, Drug, and Cosmetic Act of 1938 was the
most important single step toward effective government oversight of pharmaceu-
tical development. Spurred by a tragedy in which over 100 people died from an
antibiotic that was suspended in toxic diethylene glycol, this was the first
legislation to require drug manufacturers to submit both efficacy and safety
data, as part of a New Drug Application (NDA), to the newly formed Food and
Drug Administration (FDA) prior to marketing their products. The act thus
endowed the FDA with the power to prevent dangerous drugs from reaching
the market rather than removing them once harms had emerged. Notably, the
1938 law did not require testing on animals prior to human use and did not
empower the FDA to require clinical trials demonstrating efficacy of the product.
National regulatory bodies’ ability to regulate pharmaceutical development was
substantially increased again in 1962 following the infamous tragedy of thalid-
omide, an anti-nausea medication that was given during pregnancy and resulted
in thousands of infants to be born with severe limb malformations. The FDA
formalized regulations following the 1962 Kefauver-Harris Amendments in the
United States to require drug manufacturers to submit an Investigational New
Drug (IND) application in order to be permitted to conduct studies in humans for
a drug under development and established the clinical trial phases of drug
development. For the first time, drug manufacturers now were required to dem-
onstrate that their products were efficacious for the condition being studied. In
the subsequent decades, the FDA continued to refine its role in overseeing drug
development and today actively works with drug manufacturers in designing
their clinical trials to ensure the efficacy and safety aspects are addressed to its
satisfaction (Junod 2008).
In Europe, government regulation of drugs proceeded more slowly. Each Euro-
pean nation originally had its own set of regulations for approving drugs for public
use, making the marketing and distribution of medications across the continent
highly problematic and expensive. Following the founding of the European Union
in 1993, the European Medicines Agency (EMA) was formed in 1995 with the goal
of harmonizing drug approval processes and reducing costs for drug companies. The
Clinical Trials Directive (EU Directive 2001/20/EC) established the framework for
Randomized Controlled Trials and the Efficacy of Psychotropic Medications 309
clinical trials in Europe, but implementation varied across the member states.
Consequently, Clinical Trials Regulation (Regulation (EU) 536/2014) was adopted,
which aimed to further harmonize clinical trials rules across the EU and is expected
to go into force in 2019. Although the terminology of the oversight process differs
between the EMA and FDA (e.g., “Clinical Trial Application” versus the IND of the
FDA), the overall components and procedures are very similar. Perhaps the most
important difference is that in the United States, the only route to marketing approval
is through the FDA, but in Europe drugs can be approved through one of four routes,
including a centralized process that applies to all member states (though no psychi-
atric drugs can be approved via this route), as well as decentralized or nation-level
processes (van Norman 2016). Other important national drug regulatory agencies
include the Australian Therapeutic Goods Administration, founded in 1989, and the
Pharmaceuticals and Medical Devices Agency of Japan, founded in 2004. The
national regulatory agencies differ regarding the degree to which they consider
evidence from clinical trials conducted outside their regions of oversight when
making decisions about marketing approval in their countries. Consequently, many
psychotropic drugs are available in ex-US nations that have either not undergone
FDA review or failed approval, including widely prescribed antidepressants
(agomelatine, moclobemide, milnacipran, reboxetine) and antipsychotics
(amisulpride, zotepine, zuclopenthixol).
At the forefront of all clinical research is the need to ensure the rights of each study
participant are protected. The 1964 World Medical Association Declaration of
Helsinki and its subsequent amendments lay out the ethical principles governing
research with human subjects (World Medical Association 2013). The process of
obtaining the informed consent of a participant prior to performing any research
procedures is a crucial component of the respect for persons and is a foundational
component in the conduct of clinical trials. Adherence to the principles of ethical
research is the responsibility of the investigators, with oversight for safeguarding the
rights and welfare of research participants provided by Ethics Committees or
Institutional Review Boards. Participants in clinical trials are always free to with-
draw from studies at any time and for any reason.
310 B. W. Dunlop and C. Medeiros Da Frota Ribeiro
Trial Design
The primary aim of pivotal RCTs is to evaluate a drug’s efficacy for a disease.
Determinations of efficacy require a comparison between the investigational (some-
times called “active”) treatment and another intervention, typically placebo in
psychiatric RCTs. Perhaps the ideal RCT design is a crossover study. These trials
possess the great advantage of having each participant serve as their own control,
thereby eliminating most potential confounding factors. In a crossover design, a
subject is randomly assigned to first receive either the active treatment or the
comparator for a specified period of time and then subsequently given the other
treatment for an equivalent period. Often a “washout” period of 1–2 weeks is used
between the treatment periods to allow for the effects of the first treatment to resolve.
Despite their appeal, crossover designs have found limited application in psychiatry
because treatments change people; that is, the first treatment can result in carry-over
effects that may last far longer than the washout period (e.g., once a person’s
depression remits, it may not return for years), so participants are not in the same
state when they start the second treatment as they were when they started the first.
Chronic insomnia is the psychiatric disorder perhaps best suited for study in cross-
over designs.
Randomized Controlled Trials and the Efficacy of Psychotropic Medications 311
The randomized, parallel-group trial has become the most widely used and
accepted trial design to evaluate the efficacy of psychiatric drugs. In this design,
“parallel group” means each subject is assigned to receive only one treatment so that
the two arms run “in parallel”; the patients never cross between treatments (Atkins
2009). The comparison made to determine efficacy is based on the mean (average)
outcomes of the subjects in the treatment arms. Dosing may be either fixed (patient
remains on the same dose throughout the trial) or flexible (prescriber may adjust the
patient’s dose based on tolerability and response). Fixed-dose studies are used in
phase II trials to determine the dose-response relationships and plan doses for phase
III trials, which may use fixed or flexible dosing regimens. Flexible dosing is
associated with reduced frequency of dropout in trials of antipsychotics (Rabinowitz
et al. 2008) but not antidepressants (Khan et al. 2003).
A more complicated design is the factorial trial, which evaluates the effect of two
or more treatments simultaneously using combinations of the treatments. This design
has more statistical power than a standard parallel-group design, but if the two
treatments have interaction effects, the interpretation of the results can be difficult.
In psychiatric trials, these designs are typically used for drugs that already have
marketing approval, such as combining an antidepressant and an agent to reduce the
use of substances in patients with comorbid mood and substance use disorders.
The trial design most commonly used to evaluate a medication’s efficacy in
preventing recurrence of an episode of a disorder is a double-bind placebo-controlled
discontinuation trial. These trials typically use an open-label stabilization treatment
phase for 1–6 months, depending on the disorder, in which all patients receive the
active medication. Those who respond to the medication are then randomized in a
double-blind manner to continue on the medication or switch to placebo using a
short taper, with outcomes assessed over follow-up periods of 6 months to 2 years
(Glue et al. 2010; Lindström et al. 2017; Leucht et al. 2012; Batelaan et al. 2017).
Outcomes in these trials in mood disorders are usually recurrence of a full episode or
an increase in symptom rating scale score above a certain threshold; outcomes for
other disorders may additionally include time to hospitalization or need for change in
treatment.
Recently, clinical trials designed to identify long-term protective effects of med-
ications have been applied to psychiatric illnesses. Such “disease modification” trials
evaluate whether a treatment slows the progression of clinical symptoms in associ-
ation with a significant effect on a validated biomarker for the illness (European
Medicines Agency 2018). When a treatment is believed to directly or indirectly
restore or prevent the loss of functioning of neurons, it may be considered
“neuroprotective” and contribute to disease modification. Among neuropsychiatric
disorders, Alzheimer’s disease and Parkinson’s disease, where the pathobiology is
best established, have received the most study, but no drugs have proven to be
disease modifying to date. Many approaches to demonstrating disease modification
have been applied, including washin, washout, delayed start, and time-to-event
clinical trial designs, though it has been difficult to clearly distinguish a treatment’s
312 B. W. Dunlop and C. Medeiros Da Frota Ribeiro
effects on reducing symptom expression versus its effects on the progression of the
illness (McGhee et al. 2016). Currently, schizophrenia and bipolar disorder are the
psychiatric disorders for which disease modification and neuroprotection studies are
of greatest interest, utilizing neuroinflammation and gray and white matter brain
volumes as potential biomarkers (Robertson et al. 2019).
Duration
The length of an RCT depends on the question to be answered, the naturalistic course
of the illness studied, and the pharmacokinetics and mechanism of action of the
investigational drug. RCTs assessing acute efficacy may be as short as 4–6 weeks,
whereas trials evaluating relapse prevention may last a year or more. For illnesses
such as obsessive-compulsive disorder (OCD) or post-traumatic stress disorder
(PTSD), where substantial gains may not emerge until 2–3 months of treatment,
shorter trials risk falsely concluding that a drug is ineffective (Soomro et al. 2008;
Stein et al. 2006). Alternatively, trials may be excessively long for conditions that
can improve naturalistically on shorter time frames, such as the mood episodes of
bipolar disorder (Thase et al. 2008).
Participant Selection
Confirmation of Diagnosis
For most psychiatric RCTs, the presence of the targeted disorder in an individual is
assessed through a combination of a psychiatrist’s evaluation and administration of a
structured or semi-structured clinical interview. These assessments also serve to
ensure that excluded disorders are absent. The most commonly used diagnostic
instruments in RCTs are the Structured Clinical Interview for DSM (SCID), avail-
able in versions reflecting the DSM-IV criteria and the DSM-V criteria (First et al.
2015); the Mini Neuropsychiatric Interview (MINI), with multiple versions
reflecting corrections and revised DSM criteria (Sheehan et al. 1998); and the
World Health Organization’s Composite International Diagnostic Interview v2.1
(Robins et al. 1988), which assesses disorders defined by ICD and DSM criteria.
The World Health Organization’s World Mental Health Composite International
Diagnostic Interview (WMH-CIDI) was developed to replace the WHO CIDI,
Randomized Controlled Trials and the Efficacy of Psychotropic Medications 313
incorporating risk factors, treatment histories, and other queries in addition to the
diagnostic criteria assessments (Kessler and Ustün 2004).
Control Group
The control group in pivotal RCTs in psychiatry is nearly always placebo. Pill
placebos should have similar appearance, odor, and taste to maximize blinding. If
the investigational drug proves superior to placebo, the RCT is considered a “pos-
itive trial”; if not, it is a “negative trial.” Some RCTs employ three arms, with the
third arm being an “active control,” i.e., a medication that is already marketed for the
condition. Having an active control arm allows for assessment of the “assay sensi-
tivity” of the enrolled sample. Specifically, if the investigational treatment and the
active control treatment do not prove superior to placebo, then the RCT is considered
a “failed trial.” With this result, the investigational medication cannot be concluded
to lack efficacy, but rather that the RCT itself, either through its design or the subjects
enrolled, was inadequate for detecting the hypothesized effects. Despite the advan-
tages of having an active control, the additional arm greatly increases the number of
required subjects, thereby reducing feasibility of achieving the necessary enrollment,
and may also increase placebo response rates through increases in patient expecta-
tion for improvement (Sinyor et al. 2010; Woods et al. 2005). Notably, these RCTs
are usually not designed to test whether the investigational drug is superior to the
active comparator, and head-to-head comparisons between the two drugs are rarely
reported.
Randomization
Stratification
Blinding
Blinding (also known as “masking”) refers to efforts to prevent patients, their family
members, physician-investigators, or outcome assessors from knowing which treat-
ment the patient has been assigned to in an RCT. Blinding of participants is
important to minimize expectancy of improvement, which is a major driver of
placebo response. Similarly, blinding of investigators and assessors is important to
prevent biases arising from ratings assessments or early withdrawal of patients due to
potential allegiance to the investigational treatment (Mora et al. 2011). “Single-
blind” usually means that the patient is unaware, but that the physician knows the
treatment; in trials where patient blinding is not possible (e.g., psychotherapy versus
medication trials), the term is used to indicate that the outcome assessor was blind.
“Double-blind” refers to both the patient and the physicians/assessors being unaware
of treatment assignment; sometimes this is referred to as “triple-blind” to emphasize
that the outcome assessor (if different from the investigator) is unaware.
“Allocation concealment” refers to blinding of investigators to the randomization
sequence. Lack of concealment could lead an investigator to defer a certain patient’s
randomization visit for various reasons (e.g., wanting to ensure they received the
active treatment), thereby undermining the goal of randomization. Maintaining the
blind after the study treatment is initiated is also very important and challenging.
During treatment, it may be possible for patients and investigators to “see through
the blind,” often due to the emergence of side effects within an individual. This
concern is particularly great for medications with side effects that are well-known
and common, such as sedation with quetiapine (Calabrese et al. 2005). Although
rarely used, “active placebos” (i.e., medications that can mimic the investigational
drug’s side effect) have been proposed, such as using low-dose diphenhydramine as
a placebo for investigational drugs with the side effect of sedation. However, there is
a potential ethical concern about giving clinical trial participants a product with
known adverse effects without expected therapeutic benefit. Despite their potential
utility, active placebos have received scant use in modern trials (Perlis et al. 2010).
An exception is RCTs of depression or PTSD that evaluate sub-anesthetic doses of
Randomized Controlled Trials and the Efficacy of Psychotropic Medications 315
ketamine (which produces rapid and obvious mental effects), for which midazolam
has been used as an active placebo (Fava et al. 2018).
Outcomes
include their dependence on the rater’s range of experience with patients and the
unanchored, coarse nature of the scoring levels. Inter-rater reliability training or
assessment for the CGI scales is rarely conducted prior to beginning the RCTs. For
these reasons, CGI outcomes are usually used as secondary measures of treatment
efficacy. Efforts to bring greater standardization to CGI scoring across disorders
have recently been published (Dunlop et al. 2017).
Study Monitoring
Before, during and after the completion of a clinical trial, sites conducting the study
undergo external monitoring to ensure the study protocol is being followed correctly
and that the documentation of clinical data, study medication accountability, and
regulatory procedures are all appropriate. Government authorities have the power to
audit research sites and are particularly vigilant for concerns around a site’s data
quality and protection of research subjects.
Effect Size
The magnitude of the difference between treatments is often referred to as effect size.
Many different measures can be considered effect sizes, the simplest being the
difference in rating scale scores between treatment arms. For many disorders,
different scales may be used across trials, making direct comparisons difficult. For
this reason, a difference between treatment arm scores may be converted to a
standardized mean difference (SMD), also known as Cohen’s d. This measure of
effect size is calculated by dividing the mean difference between the treatments by
the pooled standard deviation. A d = 1 indicates that two compared groups differ by
one standard deviation. By convention, an SMD (d) of 0.2 is considered to be a small
effect, 0.5 a medium effect, and 0.8 a large effect, though Cohen cautioned that these
thresholds may not apply in all contexts (Cohen 1988). Another SMD measure,
Hedges’ g, is calculated similarly to Cohen’s d but uses pooled weighted standard
deviations (as opposed to unweighted) and is considered a better measure when
sample sizes are small (Hedges 1981). When the outcome of an RCT is not a
continuous measure (e.g., score change) but rather a categorical outcome (e.g.,
response, remission, relapse), then relative risk may be used as the effect size. In
this usage, “risk” does not imply a bad outcome, but simply a ratio: the proportion of
Randomized Controlled Trials and the Efficacy of Psychotropic Medications 317
active drug-treated patients with the categorical outcome divided by the proportion
of comparator-treated patients with the categorical outcome. Across all psychiatric
disorders, the relative risk for active medications compared to placebo in preventing
relapse/recurrence is larger than for achieving response or remission with acute
treatment.
Perhaps the most clinically applicable measure of effect size is the number needed
to treat (NNT), which gives an absolute measure: the number of patients that would
need to receive the active drug to achieve one more positive outcome than if the same
number had received the comparator. Number needed to harm (NNH) is the oppo-
site: the number of patients treated who would need to receive the active drug to
experience one additional harm (typically a specific side effect) above the number
occurring in the control condition. The specific harms evaluated by this metric may
range from discomforting side effects to severe adverse reactions or death. Effect
sizes derived from placebo-controlled RCTs for selected medications for psychiatric
treatments are presented in Table 2.
P-value
Much emphasis is placed on the p-value (i.e., the probability that the study results
would have occurred if the null hypothesis was true, i.e. that there is no difference in
efficacy between the treatments). The results are declared to be “statistically signif-
icant” if the p-value is below a certain threshold, traditionally p < 0.05 (meaning that
there was <1/20 chance that the results could have arisen if the drug is truly no better
than placebo). Small-magnitude differences between treatments, which may have
little clinical relevance, can reach “statistical significance” if the sample size is large
enough, or the standard deviations of the means are very narrow. Conversely, a large
difference between groups may appear to have high clinical relevance, but if derived
from a small sample, the possibility that the results arose simply by chance is greater,
and statistical significance not achieved. When multiple outcomes are assessed, a
downward adjustment needs to be applied to the threshold p-value because with each
additional statistical comparison the probability of a chance finding increases.
Confidence Interval
Table 2 (continued)
SMD
Disorder State Drug class (95% CI) NNT (95% CI) Reference
Social anxiety Acute SSRI 0.44 Curtiss et al.
disorder (0.37–0.51) (2017)
Venlafaxine 0.45
(0.35–0.55)
MAOI 0.36
(0.21–0.51)
Post- Acute SSRI 0.23 Hoskins
traumatic (0.12–0.33) et al. (2015)
stress disorder Venlafaxine 0.20
(0.05–0.35)
Obsessive- Acute SSRI 3.2 Soomro et al.
compulsive (3.8–2.6) (2008)
disorder YBOCS
points
Adult ADHD Acute Amphetamine 0.79 Cortese et al.
(0.58–0.99) (2018)
Methylphenidate 0.49
(0.34–0.65)
Atomoxetine 0.45
(0.58–0.32)
2nd Gen AP second-generation antipsychotic, ADHD attention deficit hyperactivity disorder, ADM
antidepressant medications, MAOI monoamine oxidase inhibitor, Olanx + Fluox olanzapine-fluox-
etine combination, NNT number needed to treat to achieve one more response or prevent one more
relapse, SMD standardized mean difference, SNRI serotonin-norepinephrine reuptake inhibitor,
SSRI selective serotonin reuptake inhibitor, TCA tricyclic antidepressant, YBOCS Yale-Brown
Obsessive Compulsive Scale
Intention-to-Treat
Missing Data
Missing data are a significant challenge in analyzing RCTs. Study participants who miss
scheduled trial visits, or who dropout prior to completion are sources of missing data.
320 B. W. Dunlop and C. Medeiros Da Frota Ribeiro
There is no single agreed upon approach to handling missing data in statistical analyses,
other than to emphasize trial designs and procedures should minimize the factors that
contribute to missing data. For purposes of analysis, missing data can be broadly
subdivided into data that are “missing at random” versus “missing not at random”
(Little and Rubin 2002). The former applies to data that are missing due to reasons
unrelated to their actual values. The later applies when the cause for the data to be
missing is related to the actual values of the missing data. The distinction is important
because the results may be biased by the causes for the missing data, thereby impacting
analytic approaches.
The last observation carried forward (LOCF) principle is a means of imputing
final outcomes for missing data, based on the last measure for the outcome obtained
during the trial. For patients who complete the trial, no imputation is necessary; their
last measure is the protocol-specified outcome. However, for patients who drop out,
the last measure obtained is used as their outcome. Although regulatory agencies
historically preferred LOCF analyses of RCTs, considering them to be more con-
servative than analyses of complete cases (Harmer and Simpson 2009), there are
many statistical concerns about using LOCF. In particular, it uses the unrealistic
assumption that participants who drop out would continue responding at exactly the
same level at the end of the trial as they did at the time of dropout, which may
unpredictably bias results either toward or away from the null hypothesis.
Mixed model repeated measure (MMRM) approaches to data analysis are a
superior means of handling missing data in RCTs, now accepted by regulatory
agencies. With MMRM, likelihood-based methods using data collected prior to
dropout are used to predict the outcomes that would have been likely to occur
after dropout. The term “mixed models” is used because these analyses generally
incorporate both fixed effects (variables not thought to vary between participants,
such as treatment arm) and random effects (variables that differ between participants,
such as baseline severity). MMRM approaches make more efficient use of the data
and are robust when data are missing at random. They remain susceptible to bias
when data are missing not at random, in which case sensitivity analyses are required
to evaluate the robustness of the results under various assumptions about the data
structure (Mallinckrodt et al. 2008). Simulation and evaluation of NDA datasets
comparing MMRM versus LOCF approaches have found that they generally pro-
duce similar results in terms of estimating effect sizes, but that LOCF analysis of
variance approaches tend to underestimate standard errors and increase the risk of
Type I error (Siddiqui et al. 2009).
Internal Validity
Internal validity of RCTs refers to the whether the observed (measured) difference
between treatment arms reflects a true difference in the efficacy of the medications,
and not some other factor distorting the measured difference. Internal validity
emerges from the quality of the design and conduct of the clinical trial. Low internal
validity can arise from either a systemic bias or from uncontrolled confounding
Randomized Controlled Trials and the Efficacy of Psychotropic Medications 321
External Validity
Placebo Response
The greatest challenge in assessing the efficacy of drugs for psychiatric disorders is
the degree of placebo response. Placebo response rates in RCTs for depression,
anxiety, and schizophrenia have been increasing over the past three decades, in
322 B. W. Dunlop and C. Medeiros Da Frota Ribeiro
synchrony with decreasing effect sizes of active medications (Walsh et al. 2002;
Dunlop et al. 2012; Rutherford et al. 2014, 2015). Response to placebo in an RCT
derives from three factors that can drive improvement: (1) natural waning of
symptoms over time as part of an episodic illness, (2) regression to the mean
(which is particularly relevant when high severity scores are required for study
entry) (Khan et al. 2007); and (3) the placebo effect, i.e., positive psychological
effects and an expectation for improvement patients experience simply by receiving
care. Important elements of the placebo effect include instillation of hope,
psychoeducation resulting from repeated assessments of symptoms and an explan-
atory model for understanding their illness, and high frequency (often weekly) of
contact with concerned research staff. (Frank 1973). In antidepressant RCTs,
response rates decrease and dropouts increase with increasing probability of receiv-
ing placebo, which may stem from expectancy effects (Salanti et al. 2018). Nocebo
responses can also occur, when expectations of harm result in adverse events or
symptomatic worsening with inert treatment.
An insidious contributor to apparent placebo response is selective score
inflation (Landin et al. 2000). Because RCT investigators receive payment for
enrolling eligible subjects, there can be an incentive for over-rating (i.e., inflat-
ing) symptoms at the screening or baseline visits to achieve the necessary
symptom severity threshold required for the patient to qualify. After randomiza-
tion, this pressure to over-rate resolves and more accurate ratings may be
conducted at subsequent visits, leading to a rapid drop in the severity score. A
visual demonstration of these effects in is displayed in Fig. 1a. This factor, along
with the other drivers of placebo response, can result in the frequently observed
graph of symptom change over time in RCTs for major depression, where the
largest reductions in symptoms are observed in the first 2 weeks after starting
treatment (Fig. 1b), which is not consistent with trajectories of improvement
observed in clinical practice.
Several approaches have been developed to minimize the effect of placebo
response in RCTs. A single-blind placebo-lead in phase, in which a 1–2-week
treatment with placebo is given prior to the randomization visit, may be used to
exclude patients who show 25% improvement with the placebo. To address
selective score inflation, centralized raters have been increasingly incorporated into
psychiatric RCTs. With centralized rating, after a patient has been identified by an
investigator at a trial site as appropriate for the study, an independent off-site clinical
assessor conducts a second evaluation via telephone or internet to confirm the
diagnosis and ensure the patient meets the symptom severity criteria. Centralized
ratings may also be used for evaluating symptom scale assessments of efficacy
through the trial, which may suffer from poor inter-rater reliability at trial sites and
be vulnerable to bias arising from unblinding occurring at the site (Kobak et al.
2010).
The ethics of continuing to test investigational drugs against placebo in mental
health disorders despite the existence of effective treatments has been criticized, on
the basis that use of placebo amounts to an unethical withholding of treatment
(Michels and Rothman 2003). However, there are strong arguments for continuing
Randomized Controlled Trials and the Efficacy of Psychotropic Medications 323
Fig. 1 Expectation, change, and the potential effect of selective score inflation on placebo-
controlled RCT results. (a) Theoretical figure representing patient expectations and symptom
ratings during the early phases of an RCT. Speech bubbles contain thoughts patients may have as
they enter an RCT. The solid curve represents patient’s “true” symptom level scores over time, with
a sharp decline between the screening and baseline visits reflecting increases in hope and expec-
tation. It is possible that this improvement is sufficiently great that the patient’s true severity score
may drop below the minimum score required to be eligible to be randomized at the baseline visit.
The dotted lines between screening and week 1 reflect the site rater’s actual scores. At the baseline
visit, selective score inflation occurs if site raters inflate their ratings to ensure that the patient
remains eligible to continue in the study. The difference between the “true” score and the inflated
score is the selective score inflation amount. At week 1, no minimum score is required for the
patient to remain in the study, so the incentive to over-rate the score is removed and the “true” score
is reported, resulting in an apparent steep drop in score from baseline to week 1. Other factors
related to placebo effects may also contribute to this rapid decline. After starting treatment, an
efficacious investigational drug (solid line) will show increasing separation from placebo (dashed
line) over time, but the statistical significance of this difference may be reduced due to the steep
324 B. W. Dunlop and C. Medeiros Da Frota Ribeiro
Nonadherence to Treatment
Treatment Resistance
Fig. 1 (continued) declines in recorded scores that occur in both treatment arms at the week 1 visit.
(b) Outcome data from a placebo-controlled RCT of desvenlafaxine and placebo (Dunlop et al.
2011). It is notable that 65% of the 10-point change in the placebo arm occurred in the first 2 weeks,
with the other 35% gradually accruing over the next 10 weeks. A similar pattern is observed in the
desvenlafaxine arm, which is not consistent with the slower rates of improvement usually observed
in clinical practice. Although the degree of selective score inflation in this trial could not be
assessed, it is a possible contributor to this improvement curve, a pattern also found in many
other RCTs of mood and anxiety disorders
Randomized Controlled Trials and the Efficacy of Psychotropic Medications 325
Major depressive disorder (MDD) is a leading cause of disability around the globe,
with lifetime prevalence estimates ranging from 10% to 18% (Kessler and Bromet
2013). Roughly 80% of patients with MDD experience recurrent episodes (Burcusa
and Iacono 2007), so both acute treatment and relapse prevention are important
treatment targets. MDD is the most common psychiatric disorder studied in RCTs;
placebo effects present a major challenge, with even proven antidepressants failing
to beat placebo in about half of trials (Khin et al. 2011). Psychotherapy and
pharmacotherapy have been shown to be roughly equally efficacious for mild-to-
moderate depression (Weitz et al. 2015), though the presence of psychotic symptoms
(APA 2010) or suicidal ideation (Dunlop et al. 2018; Boschloo et al. 2019) warrants
treatment with medication.
The earliest RCTs with tricyclic antidepressants were usually 4 weeks in
duration and showed large effect sizes over placebo (Undurraga and Baldessarini
2012). With the development of the first selective serotonin reuptake inhibitor
(SSRI), fluoxetine, a belief developed that these agents took 4 weeks or more to
work and the length of the double-blind phase of RCTs trials for modern antide-
pressants is now typically 6–8 weeks. However, the slow time to onset of effect of
fluoxetine likely derives from its long elimination half-life (and that of its primary
metabolite, norfluoxetine), which prolongs the time required to reach effective
plasma and cerebral spinal fluid concentrations of the drug (Altamura et al. 1994;
Henry et al. 2005). Other SSRIs or serotonin-norepinephrine reuptake inhibitors
(SNRIs) have half-lives of a day or less, leading to steady-state levels within the first
week of treatment (Anderson 2001). Meta-analyses have demonstrated that these
shorter half-life agents achieve separation from placebo by the first or second week
of treatment (Hieronymus et al. 2016), though for some patients the specific antide-
pressant effect may not emerge until week 8 (Henssler et al. 2018).
The primary outcome in MDD trials is change from baseline on a clinician-rated
continuous measure of depression severity. The most commonly used scales are the
clinician-rated Hamilton Rating Scale for Depression (HAM-D) (Hamilton 1967) or
the Montgomery-Asberg Depression Rating Scale (MADRS) (Montgomery and
Asberg 1979). Although both scales are quite old, newer scales developed to align
with the DSM diagnostic criteria for MDD have not proven to have better psycho-
metrics than these older scales. Self-report measures, such as the Beck Depression
Inventory (BDI) (Beck et al. 1961) or the Quick Inventory of Depressive Symptom-
atology Self-Report (Rush et al. 2003), are often used as secondary assessments of
antidepressant efficacy.
Several versions of the HAM-D exist, ranging from 17 to 31 items. The 17-item
version is the original and most commonly used version, with item scores ranging
from 0 to 4 or 0 to 2 depending on the item, and a total score range of 0–54. Factor
analyses of the HAM-D have identified 2–6 factors, indicating it is not a unidimen-
sional measure of depression (Bagby et al. 2004). Briefer six-item versions of the
326 B. W. Dunlop and C. Medeiros Da Frota Ribeiro
HAM-D (Bech et al. 1981; Maier and Philipp 1985) may be more sensitive to
detecting differences between active and placebo treatments, particularly for drugs
with minimal anticholinergic or antihistaminergic effects (Boessen et al. 2013). The
MADRS is composed of 10 items, each scored 0–6 with scores ranging from 0 to 60.
Severity eligibility thresholds for MDD RCTs typically use a score of 18 22 on
the HAM-D or 25 30 on the MADRS, with the higher scores representing a
moderate-to-severe level of depression severity.
“Response” and “remission” are secondary categorical outcomes that reveal the
proportion of patients who benefit within each treatment arm. Response in MDD is
defined as a 50% reduction in score from baseline. The outcome of remission is
important because it implies a near-complete absence of ongoing symptoms, which
is associated with better psychosocial functioning and lower relapse risk (Judd et al.
1998). The MacArthur criteria define the most widely employed remission thresh-
old: 7 on the 17-item HAM-D (Frank et al. 1991). Concerns have been raised that
this threshold is too high and should be lowered to achieve greater specificity for
positive long-term outcomes (Dunlop and Rapaport 2016; Zimmerman et al. 2012a).
For the MADRS, the remission threshold is less well-agreed upon; total scores 10
appear to correlate best with the HAM-D score of 7, though the scales’ differing
items will result in a lack of agreement on remission for some patients (Zimmerman
et al. 2004). Response rates in short-term trials are 50–60% with antidepressants and
30–40% with placebo (Walsh et al. 2002); remission rates are typically 25–35% vs
15–20% (Depression Guideline Panel 1993), respectively.
The sustained benefits of maintenance antidepressant treatment, particularly in
patients with recurrent depression, have been consistently demonstrated in double-
blind discontinuation studies (Borges et al. 2014). The absolute reduction in recur-
rence risk among individuals maintained on the active antidepressant compared to
those switched to placebo is roughly 20% over 1–2 years of follow-up (Borges et al.
2014; Glue et al. 2010). Although some have raised concern that the higher rates of
depressive relapse with placebo in these trials are artificially elevated due to antide-
pressant withdrawal/discontinuation effects (Fava et al. 2018), this explanation
would require that a substantial proportion of observed recurrences occur during
the first month following the switch to placebo. However, recurrence rates remain
relatively constant over the duration of the follow-up phase, indicating that antide-
pressant withdrawal symptoms do not explain these medications’ protective effects
during long-term treatment (Borges et al. 2014).
Bipolar Disorder
Grobler 2015). Bipolar disorder is diagnosed categorically, but there is strong evidence
for conceptualizing its symptomatology on a spectrum, with lifetime prevalence of
bipolar disorder being roughly 1% in adults, and another 1.5% have subsyndromal
symptoms (Merikangas et al. 2011). Due to its episodic nature, most bipolar disorder
trials target specific phases of the illness, specifically treatment of an acute mixed or
manic episode, an acute depressive episode, or maintenance treatment to prevent
recurrence of mood episodes. Regulatory approval of medications is given specifically
for each of these components.
Depressive episodes cause the majority of disability, and time spent ill in patients
with bipolar disorder (Judd et al. 2005). The importance of maintenance treatment is
evident from the 20% to 30% average annual risk for recurrence of a mood episode
(Vázquez et al. 2015). The mainstay of treatment is pharmacotherapy, though
individual and carer-focused therapies can enhance medication adherence and
reduce mood episode recurrences (Chatterton et al. 2017).
Nearly all trials for the treatment of acute mania have used a duration of 3 weeks,
typically enrolling patients in hospital, who may transition to outpatient care before
the end of the trial. The majority have used the Young Mania Rating Scale (YMRS)
as the primary outcome measure. The YMRS consists of 11 items, with 7 items
scored from 0 to 4 and 4 scored from 0 to 8, for a total range from 0 to 60 (Young
et al. 1978). Less commonly used are the Bech-Rafaelsen Mania Rating Scale (11
items rated 0–4, range 0–4) (Bech et al. 1979) and the Mania Rating Scale (11 items,
range 0–52) derived from the Schedule for Affective Disorders and Schizophrenia
(Spitzer and Endicott 1987).
A score of 20 on the YMRS is a common threshold used to select patients for
modern mania trials. Response is defined as a 50% reduction in total score from
baseline. Remission from mania is less well-defined, with many studies using a
YMRS <12 to indicate remission, though the International Society for Bipolar
Disorders has recommended a threshold of <8 or <5 be used, based on the
association of these thresholds with high-functioning status (Tohen et al. 2009).
The mood stabilizers lithium, valproic acid, carbamazepine and several atypical
antipsychotics have marketing approval for the treatment of acute manic episodes.
Clinical trials for acute major depressive episodes of bipolar patients have a
duration of 6–8 weeks and employ the HAM-D or MADRS as the primary outcome
measure, using the same thresholds for eligibility, response, and remission as in
MDD trials. In addition, excessive mood elevation or switch to mania is assessed
with a Mania Rating Scale. Only quetiapine, lurasidone, cariprazine, and the com-
bination of olanzapine-fluoxetine have marketing approval from the US FDA for the
treatment of acute depressive episodes in bipolar disorder. The efficacy of antide-
pressant medications for bipolar depressive episodes has been subject to substantial
debate (Sidor and MacQueen 2011; Gitlin 2018). Perhaps the most consistent finding
is that antidepressants given to patients who are not on a stable dose of a mood
stabilizer increase the risk of manic or mixed episode induction. This risk is
particularly prominent among patients with rapid cycling or depression with mixed
features (Gitlin 2018; Yatham et al. 2018); these patients are typically excluded from
RCTs of medications for bipolar depressive episodes.
328 B. W. Dunlop and C. Medeiros Da Frota Ribeiro
The two most commonly employed rating scales in schizophrenia RCTs are the
Brief Psychiatric Rating Scale (BPRS) and the Positive and Negative Syndrome
Scale (PANSS). The original BPRS, published in 1962, has 16 items, but a revised
version with 18 items (which added the symptoms of “excitement” and “disorienta-
tion”) has become the standard for RCTs (Overall and Gorham 1962). An expanded
version of the BPRS with 24 items has also been used. The BPRS was not
specifically designed for psychosis studies but received widespread use in schizo-
phrenia due to the absence of alternative scales and its brief length. Each item is
scored from 0 to 7, with total scores ranging from 0 to 126 on the 18-item version.
An 18-item BPRS score 45 is often used as a minimum severity entry criterion for
modern schizophrenia RCTs (Canadian Agency for Drugs and Technologies in
Health 2011).
The PANSS, published in 1987, was specifically developed to be sensitive to drug
treatments in schizophrenia and consists of 30 items scored 1–7. Three subscales,
including general psychopathology (score range 16–112), positive symptoms
(PANSS-P, range 7–49), and negative symptoms (PANSS-N, range 7–49) can be
derived from the scale (Kay et al. 1987). Threshold scores to include patients in most
RCTs are usually set between 60 and 80. A limitation of the PANSS is the time
taken to administer it; a six-item version of the PANSS has been developed but
awaits validation (∅stergaard et al. 2016). Less commonly used are the Scale for the
Assessment of Positive Symptoms (Andreasen and Olsen 1982) and the Scale for the
Assessment of Negative Symptoms (Andreasen 1982).
Unlike mood disorders, response in schizophrenia RCTs lacks a consistent
definition. Investigators have used reductions on the BPRS or PANSS as low as
20% and as high as 50% to define responders, while others have used CGI-I score of
2 (“much improved”) or a combination of CGI-I and rating scale scores (Samara
et al. 2019). Schizophrenia remission criteria are not widely used but have been
proposed to require low levels of eight key symptoms for at least 6 months
(Andreasen et al. 2005).
The psychosis rating scales do not capture the cognitive impairments of schizo-
phrenia. In order to develop a pathway for the US FDA to approve treatments for the
cognitive impairments of schizophrenia, the Measurement and Treatment Research
to Improve Cognition in Schizophrenia (MATRICS) initiative was developed by the
NIMH (Nuechterlein et al. 2004). The resulting MATRICS Consensus Cognitive
Battery (MCCB) consists of ten tests assessing seven cognitive domains that may be
responsive to treatment: verbal learning and memory, visual learning and memory,
working memory, attention and vigilance, processing speed, reasoning and problem-
solving, and social cognition (Nuechterlein et al. 2008). Currently approved anti-
psychotics have not proven to have robust effects on MCCB measures.
Long-acting injectable (LAI) versions of antipsychotics offer important advan-
tages over oral forms, including more consistent blood concentrations and improved
treatment adherence (Fagiolini et al. 2017; Kane et al. 2013; Alphs et al. 2014).
Demonstrating advantages of LAI over oral therapy in traditional RCTs has been
very challenging, with meta-analyses finding no significant benefit over oral therapy,
in large part because the requirements of participating in an RCT (frequent visits and
330 B. W. Dunlop and C. Medeiros Da Frota Ribeiro
monitoring, which increase treatment adherence) run directly counter the major
advantage LAIs offer in routine clinical settings (Kishimoto et al. 2014). Marketing
approval for LAIs may be achieved through RCTs demonstrating noninferiority or
superiority of the LAI formulation compared to oral therapy for symptoms (Euro-
pean Medicines Agency 2012) or real-world outcomes (Alphs et al. 2016).
Evidence derived from RCTs of LAIs has been supplemented with naturalistic
studies, such as mirror-image studies and cohort studies. In mirror-image studies,
patients undergo a retrospective evaluation of their illness course during oral therapy
and then are prospectively assessed after initiating LAI treatment, so that each
patient serves as their own control. These designs are susceptible to recall, selection,
and expectation biases, as well as regression to the mean effects. In cohort studies of
LAIs, patients nonrandomly assigned to LAI or oral therapy are evaluated, prospec-
tively or retrospectively, for differences in outcomes. Confounding arising from the
lack of randomization, as well as attrition or recall bias, can impact cohort studies.
Meta-analyses have found large effect sizes for LAIs in reducing hospitalization for
both mirror-image and cohort designs (Kishimoto et al. 2013, 2018). To address the
discrepancy between RCT and naturalistic trials, effectiveness trials, using few
exclusion criteria and which minimally alter the usual setting for treatment delivery,
have shown efficacy for LAIs (Naber et al. 2015; Schreiner et al. 2015), and likely
represent the best approach for assessing LAI moving forward (Kane et al. 2013).
be excluded from GAD trials, with presence of a current major depressive episode
the most important exclusion criterion (Hoertel et al. 2012).
The classes of medications with marketing indications for GAD include benzo-
diazepines, SSRIs, SNRIs, pregabalin, and hydroxyzine. Benzodiazepines were
developed before GAD was defined by the DSM; their labeled indication for the
management of anxiety disorders may be construed to include GAD. Efficacy of
SSRIs and SNRIs in GAD RCTs may be underestimated because several somatic
symptoms of anxiety assessed on the HAM-A (gastrointestinal, genitourinary, and
autonomic symptoms) may also arise as side effects of these medications, leading to
elevations in HAM-A scores by blinded raters. Although benzodiazepines have the
largest effect sizes for GAD, clinical practice guidelines recommend SSRIs or SNRIs
as first-line agents due to benzodiazepines’ higher risk of abuse, lack of efficacy for
comorbid depression, and potential for cognitive impairments (Gomez et al. 2018).
Panic attacks consist of the rapid-onset periods of fear or anxiety associated with
multiple physiological symptoms. When these attacks occur unexpectedly (i.e.,
without an obvious triggering event or situation), are recurrent, and are followed
by persistent worry of having additional attacks or avoidance of situations associated
with the attacks, panic disorder can be diagnosed. Panic disorder affects roughly 5%
of Western adults in their lifetime, with lower rates observed in other global regions
(Lewis-Fernández et al. 2010) and is often comorbid with mood disorders or other
anxiety disorders (Kessler et al. 2005). Psychological therapies, specifically CBT-
based treatments, and pharmacotherapy demonstrate roughly equal efficacy in the
treatment of panic disorder with or without agoraphobia (Imai et al. 2016).
RCTs for panic disorder typically range from 8 to 12 weeks. In comparison to other
psychiatric disorders, there has been little consistency in the measures used to assess
efficacy in panic disorder (Weise et al. 1996). Many RCTs have used primary outcome
measures that were not specific for panic attacks, such as the HAM-A or the Clinical
Anxiety Scale, a six-item scale derived from the HAM-A (Snaith et al. 1982). There
are several panic-specific questionnaires available, all of which contain one or more
questions about the frequency of panic attacks and associated somatic or avoidance
symptoms. A count of full panic attacks in the prior week, or number of patients being
free of panic attacks at trial endpoint, are the simplest outcome measures and have
been used as a primary or secondary outcome measure in most trials. Due to the
heterogeneity in outcome measures, meta-analyses have used CGI-C scores (Imai
et al. 2016; Bighelli et al. 2018). The Panic Disorder Severity Scale (PDSS), (7 items
rated 0–4, range 0–28) is available in clinician-rated and self-report versions and has
undergone the most complete psychometric evaluation of panic-specific measures,
making it the best option for use in future studies (Shear et al. 2007). An endpoint
score 5 on the PDSS has been proposed as a remission threshold, with a 40%
reduction from baseline representing response (22). Interest in the pharmacologic
332 B. W. Dunlop and C. Medeiros Da Frota Ribeiro
treatment of panic disorder has diminished greatly since the number of RCTs peaked in
the late 1990s, with no placebo-controlled trials conducted in the past decade.
Agoraphobia, which by DSM-IV criteria could only be diagnosed in patients with
panic disorder, was revised to be a free-standing illness in DSM-V. Agoraphobia
symptoms improve with treatment among patients with panic disorder, but there
have been no pharmacotherapy RCTs of patients with agoraphobia who do not have
panic disorder (Perna et al. 2011).
Social anxiety disorder is among the most common psychiatric disorders, affecting
up to 12% of adults. Fear of public speaking is the most common manifestation of
this illness, and many patients forgo pharmacotherapy and choose employment
where presentations are not required or simply use an anxiolytic on an as-needed
basis. In more generalized forms of social anxiety disorder, patients may have
difficulty speaking up in small groups, to authority figures, or engaging in behaviors
such as eating or writing while being observed. The adverse life consequences for
these individuals are much greater, with lower likelihood of marriage, promotion,
and increased risk for major depression (Wong et al. 2011). CBT can effectively treat
social anxiety disorder and is roughly equally as effective as pharmacotherapy. For
patients with more generalized forms of social anxiety, several medications have
proven effective, including SSRIs and venlafaxine, which have marketing indica-
tions, phenelzine, moclobemide, benzodiazepines, and anti-convulsants (Williams
et al. 2017).
Response to pharmacotherapy in social anxiety disorder is slow, with most
placebo-controlled RCTs lasting 10–12 weeks. The Liebowitz Social Anxiety
Scale (LSAS), which is available in well-correlated self-rated and clinician-rated
forms (Fresco et al. 2001), has emerged as the preferred primary symptom measure.
The LSAS lists 24 social scenarios, each of which is scored twice: once for fear of
the scenario and once for avoidance of it. Fear and avoidance are both scored from
0 to 3, resulting in total scores ranging from 0 to 144. The LSAS score can also be
divided into sub-scores assessing performance situations and social interaction
situations. A threshold of LSAS 50 or 60 is usually used as the severity inclusion
score for RCTs.
A score of <30 has been used to define remission on the LSAS, as this threshold
has been found to distinguish between people who meet criteria for social anxiety
disorder versus those who do not (von Glischinski et al. 2018). However, social
anxiety disorder, perhaps more than any other psychiatric disorder, is impacted by
cultural norms. Indeed, LSAS mean scores among those clinically diagnosed with
social anxiety vary substantially across cultures (von Glischinski et al. 2018),
suggesting that remission scores may warrant individualizing by culture. Response
definitions in RCTs have usually used a 50% reduction from baseline, but this
amount of change does not align with CGI-I scores (Bandelow et al. 2006). Although
an optimal percent improvement for defining response in social anxiety disorder
Randomized Controlled Trials and the Efficacy of Psychotropic Medications 333
trials has not been determined, some work suggests a 30% reduction from baseline
may be most appropriate (Bandelow et al. 2006; von Glischinski et al. 2018).
the CAPS-5 was released, incorporating the new symptoms and revising the scoring
of the symptom items (Weathers et al. 2018). On the CAPS-5, each of the 20 items is
only scored once from 0 to 4, incorporating severity and frequency together,
resulting in a range of 0–80. Current studies using the CAPS-5 use scores of
28–32 as the inclusion threshold.
Other scales commonly used in PTSD RCTs are the PTSD Checklist for DSM-5
(PCL-5), a self-report of 20 items rated 0–4 that reflect the DSM-5 symptom criteria,
and the Impact of Events Scale-Revised (IES-R), a self-report scale assessing 22
symptoms from 0 to 4 over the past week. The CAPS-IV and PCL-5 have shown
good correlation with the CAPS-5, indicating compatibility across trials using these
instruments (Weathers et al. 2018). On the CAPS-IV, response in RCTs has been
most commonly defined as a reduction in total CAPS scores of 30%, with
remission defined as an endpoint score <20 (Stein et al. 2006). Improvement
thresholds for the CAPS-5 have not yet been established. Many other measures for
assessing PTSD symptoms have been developed but have not received widespread
use (American Psychological Association 2017).
The variety of symptoms subsumed within the PTSD diagnostic criteria results in
heterogeneous samples in RCTs, particularly in the levels of re-experiencing or
hyperarousal symptoms across patients. The broad variability in symptomatology
can produce highly variable outcomes of pharmacotherapy trials. The SSRIs and
venlafaxine broadly improve the syndrome but appear less effective for
re-experiencing symptoms than for the other clusters (Stein et al. 2006). Prazosin
is helpful for insomnia and nightmares, and atypical antipsychotics such as risper-
idone and quetiapine have also proven to be particularly efficacious for re-experienc-
ing symptoms either as monotherapy or when added to an SSRI (Dunlop and
Davidson 2019).
The importance of patient selection for PTSD RCT outcomes is evident from the
remarkable differences in efficacy of agents tested in civilians with PTSD versus US
military veterans with PTSD. SSRIs, prazosin, and risperidone all have demon-
strated efficacy in multiple PTSD RCTs enrolling civilians, but all three have failed
in large trials administered through the US Veterans Affairs (VA) health system.
Reasons for these discrepancies remain uncertain but may result from the greater
chronicity or severity of PTSD in military samples, secondary gain concerns, or
aspects of the VA trial designs, including broad inclusion criteria, including signif-
icant medical comorbidity, and allowing multiple concomitant psychotropic medi-
cations during the trial. These trial design features all act to reduce the probability of
identifying differences in efficacy between an active intervention and placebo. Trials
conducted in the VA may be more appropriately considered effectiveness trials in a
chronic and severely ill population, rather than classic RCTs assessing efficacy.
Obsessive-Compulsive Disorder
OCD involves the experience of obsessive thoughts or compulsive behaviors that are
distressing or impairing or occupy the patient for more than 1 h per day. OCD affects
Randomized Controlled Trials and the Efficacy of Psychotropic Medications 335
Thresholds for response and remission are not well-established for ADHD. Reductions
of 25–30% from baseline on rating scales have been proposed to define response,
which aligns with a one-level change on the CGI-C scale (Steele et al. 2006; Goodman
et al. 2010). RCTs for adult ADHD are similar in design to those for children and
adolescents, most often using the ADHD-RS or Conners Adult ADHD Rating Scale as
symptom outcome measures (Cunill et al. 2016).
Eating Disorders
Anorexia nervosa, bulimia nervosa, and binge eating are not common, with preva-
lence in Western nations of 0.5%, 2%, and 4%, respectively (with lower rates in other
regions), but they carry high rates of morbidity and mortality (Hoek 2016;
Westmoreland et al. 2016). Psychotherapeutic treatments are the primary modality
of treatment for eating disorders, though medications may be used adjunctively. No
medications carry a marketing indication for anorexia nervosa; fluoxetine is the only
drug with an FDA indication for bulimia nervosa, and lisdexamfetamine is the only
drug with an indication for binge eating disorder.
One 11-week phase II and two 12-week phase III trials have demonstrated the
efficacy of lisdexamfetamine for binge eating disorder. The trials used as the primary
outcome the change in days per week of eating binges, with a secondary outcome
being the proportion of patients with complete cessation of binge eating for 4 weeks
(McElroy et al. 2015, 2016). A single 6-month, double-blind discontinuation trial
with this drug, which defined relapse as 2 binge eating days per week for 2
consecutive weeks and a 2-point CGI-S score increase, found significantly fewer
patients continuing on the medication relapsed (4%), compared to those assigned to
placebo (32%) (Hudson et al. 2017). Although lisdexamfetamine significantly
reduced body mass index in the acute treatment trial of binge eating disorder
(McElroy et al. 2015), the effects on BMI in the long-term study were not reported.
As a continuous outcome measure, a 10-item version of the Y-BOCS modified for
binge eating (Y-BOCS-BE) has demonstrated sensitivity to treatment effects (Deal
et al. 2015).
Most pharmacotherapy RCTs for bulimia are 6–52 weeks in duration. SSRIs
(particularly fluoxetine) and TCAs are the most studied medications, though MAOIs
have also been tested (Svaldi et al. 2019). Remission in bulimia has been defined as
either the abstinence of bulimia-related symptoms for 2 weeks, or as no longer
meeting diagnostic criteria for bulimia nervosa (including the cognitive elements)
(Slade et al. 2018). Reduction of binge eating episodes and compensatory behaviors
are additional outcomes assessed. Meta-analyses have found pharmacotherapy to be
substantially less effective than psychotherapy, though the combination together
improve outcomes (Slade et al. 2018; Svaldi et al. 2019).
Pharmacotherapy RCTs for anorexia nervosa are conducted as add-on treatments
to a psychosocial treatment program. Trials have ranged from 7 to 52 weeks in
duration, with most having very small sample sizes and only two analyzing >100
patients. Weight gain is the primary outcome for most studies, though hormonal
Randomized Controlled Trials and the Efficacy of Psychotropic Medications 337
Throughout the world, substance use disorders (SUD) are widespread and are
leading causes of disability, though the specific patterns of drugs abuse vary by
nation (Peacock et al. 2018). Medication-assisted treatment (MAT), in which med-
ications are prescribed within a psychosocial treatment approach, is considered the
optimal form of treatment for SUD, and pharmacotherapy RCTs for SUD are
conducted within this paradigm. Unlike other psychiatric disorders, SUD outcomes
can be objectively assessed through laboratory testing using biochemical verification
to supplement self-reported measures. Unfortunately, overall success rates in treating
SUD remain unsatisfactory. A novel approach under development is to use vaccines
for SUD, most of which act by inducing production of drug-specific polyclonal
antibodies, thereby retaining the abused drug in the systemic circulation, preventing
or slowing its penetration into the central nervous system (Heekin et al. 2017).
Because substance use disorders are often characterized by binge use or frequent
relapses, the primary outcome in substance abuse RCTs is typically sustained
abstinence, which may be defined as 6 months or longer without use. Other out-
comes commonly assessed include (1) number or percent of days/weeks abstinent
from the substance; (2) proportion of the sample abstinent for a minimum period
(e.g., 3 weeks); (3) proportion of sample abstinent at end-of-treatment; (4) changes
on a scale assessing drug addiction (e.g., the drug scale of the Addiction Severity
Index-5) (McLellan et al. 1992); and (5) biochemical assessments, such as the
number or percent of negative drug screens during treatment. The most rigorous
assessments of abstinence use a combination of self-reported use and biochemical
verification measures. Biochemical verification has particular value when studying
populations who may feel pressure to under-report their use, such as pregnant
women or post-surgical patients when queried about recent smoking (SRNT Sub-
committee on Biochemical Verification 2002).
Nicotine, alcohol, and opiate use disorders each have medications with marketing
approval for their treatment. After alcohol and tobacco, cannabis is the most widely
used substance of abuse, and the efforts to legalize cannabis products proceeding in
many nations are likely to increase its usage and associated harms (Ammerman et al.
2015). RCTs for medications to address withdrawal symptoms or maintenance
treatment for cannabis (Sherman and McRae-Clark 2016), cocaine (Shorter and
Kosten 2011), or amphetamines (Lee et al. 2018) have thus far failed to identify
medications that safely and effectively enhance treatment outcomes.
Nicotine
Although the prevalence of nicotine use disorder has been declining for several
decades, it remains the leading cause of premature death and preventable disease
338 B. W. Dunlop and C. Medeiros Da Frota Ribeiro
around the world. Smoking rates vary widely across nations, ranging from 10% to
>50%, with 12% of all deaths among people over 30 years old attributable to
tobacco (Drope et al. 2018). Roughly one-third of adults with mental illness
smoke cigarettes, compared to 15% without mental illness, and they smoke more
intensively (Drope et al. 2018). A variety of psychotherapeutic interventions have
proven helpful, though their efficacy is enhanced when combined with pharmaco-
therapy (Stead and Lancaster 2012). Pharmacological interventions for nicotine use
disorders can be considered aid-to-cessation trials, which test a treatment in smokers
currently willing to cease use, using long-term abstinence rate as the primary
outcome measure. Medications with marketing approval as aids to smoking cessa-
tion include varenicline (a nicotinic acetylcholine receptor partial agonist),
bupropion (a noradrenergic and dopaminergic antidepressant), and nicotine replace-
ment therapy (NRT, delivered via patch, tablets, inhalers or sprays, or gum).
The self-assessment Fagerstrom Test for Nicotine Dependence (Heatherton et al.
1991) is the most widely used assessment for diagnosis of nicotine use disorder,
consisting of six items that total from 0 to 10. Scores of 1–2 reflect low dependence;
scores of 8–10 reflect very high dependence. Duration of RCTs for nicotine use
disorder is usually 12 weeks, and often have participants set a target quit date.
Because patients commonly fail to fully quit on the specified date, the Society for
Research on Nicotine and Tobacco suggests a 2-week grace period beyond the quit
date (Hughes et al. 2003), during which any smoking does not count toward
abstinence failure. Thus, the targeted quit date needs to be clearly defined.
A wide variety of outcome measures have been used in pharmacotherapy RCTs
for nicotine use disorder. The optimal outcome is abstinence, which has been
characterized in three constructs:
(1) Continuous abstinence refers to sustained abstinence from the quit date through
the end of the follow-up period.
(2) Prolonged abstinence is another measure of sustained abstinence but begins
from the end of the grace period following the quit date through the end of
follow-up
(3) Point-prevalence abstinence is abstinence in a specific time window (usually
7 days) prior to the follow-up assessment (Hughes et al. 2003).
for these trials, using the number of cigarettes per day (CPD) or the proportion of
participants reducing CPD by 50% as the primary outcome measure. Whether such
reductions actually reduce the health harms from smoking has not been adequately
resolved (Lindson-Hawley et al. 2016).
Biochemical verification, using cotinine or exhaled carbon monoxide, may be
used to confirm self-reported abstinence. Cotinine is a major metabolite of nicotine
and can be detected in blood, urine, and saliva. Plasma and salivary cotinine levels
are closely correlated, and the half-life of cotinine in plasma is 15–20 h (compared to
1–2 h for nicotine), allowing for detection of nicotine exposure 3–7 days after last
use. Urinary cotinine concentrations are fourfold to sixfold greater than in plasma,
making urine cotinine the most sensitive biomarker of exposure. However, most
RCTs that employ cotinine measures use salivary assessments due to the ease of
collection. A limitation of cotinine is that abstinent patients on NRT will continue to
test positive for it (Jatlow et al. 2008). Exhaled carbon monoxide (CO) may also be
used for verification, with the advantage of being valid for NRT trials. Although
exhaled CO is only able to detect very recent nicotine use (i.e., up to 9 h prior to
testing), the ease of administration with a breathalyzer, which provides real-time
results, offers high utility. The Russell standard for smoking cessation recommends
the use of self-reported prolonged abstinence with biochemical validation using
exhaled CO (West et al. 2005).
Alcohol
The diagnosis of alcohol use disorder (AUD) requires that a patient meets at least 2
of 11 DSM-V criteria and can be classified as mild, moderate, or severe depending
on the number of symptoms endorsed. The global 1-year prevalence of AUD is
nearly 5%, affecting more than 100 million people (GBD 2016 Alcohol and Drug
Use Collaborators 2018). AUD has devastating health impacts (Kranzler and Soyka
2018) and drives 5.1% of global disability-adjusted life years (World Health Orga-
nization 2018). Combination of psychotherapy (often provided through 12-step
groups, such as Alcoholics Anonymous or outpatient therapists) combined with
comprehensive medical care and pharmacological approaches are considered opti-
mal care (43).
The duration of most pharmacotherapy trials for AUD is between 12 and
17 weeks, though several have treatment periods up to 1 year. Studies typically
evaluate medications added on to a psychosocial intervention, enrolling patients
immediately after detoxification who have a minimum of 3 days of sobriety. The
primary outcome is usually the proportion of patients remaining abstinent from
alcohol (Jonas et al. 2014). Although abstinence is the most desirable result and
associated with the best long-term health, many patients wish to reduce drinking
without becoming completely abstinent. Based on data from the COMBINE trial
(Anton et al. 2006), the US FDA also accepts the surrogate outcome of “no heavy
drinking days,” with a heavy drinking day defined for men as >4 standard drinks and
for women as >3 standard drinks (FDA 2015; Falk et al. 2010).
Three medications have US FDA marking approval for alcohol use disorder,
disulfiram, acamprosate, and naltrexone, which is available in oral and long-acting
340 B. W. Dunlop and C. Medeiros Da Frota Ribeiro
injectable formulations. Despite their efficacy, <10% of patients who could benefit
from these medications receive them in the United States (Mark et al. 2009).
Disulfiram was approved in 1949 by the US FDA, though meta-analyses of RCTs
have found no benefit over control treatments placebo (Skinner et al. 2014). Because
the mechanism of disulfiram depends on the psychological fear of experiencing a
disulfiram-ethanol reaction, and the risk is similarly present in placebo arm, identi-
fying a specific effect of disulfiram is challenging (Suh et al. 2006).
Nalmefene, an antagonist at μ- and δ-opioid receptors and partial agonist at
κ-opioid receptors, is approved for treatment of alcohol use disorder in Europe for
a highly specific subset of drinkers. When used as needed when feeling tempted to
drink, nalmefene reduced total alcohol consumption and number of heavy drinking
days per month compared to placebo (Jonas et al. 2014). Several concerns exist
about the data used to justify approval of nalmefene, including post hoc outcome
definitions and a retrospectively defined subset of patients analyzed (Fitzgerald et al.
2016).
Opiates
Opioid use disorder (OUD) is considered to be at an epidemic level in the United
States (Scholl et al. 2018) and is the fastest growing substance use disorder globally
(GBD 2016 Alcohol and Drug Use Collaborators 2018). Unlike other substance use
disorders, behavioral treatments without medication assistance show low success
and are rarely used. Pharmacotherapy for opioid dependence is divided between the
acute treatment phase, designed to minimize opioid withdrawal symptoms, and the
maintenance treatment phase, where the primary goals are to prevent relapse and
minimize harms resulting from use. A key consideration in opiate use disorder RCTs
is determining the population to enroll based on the form of opiate abused. Com-
pared to heroin users, prescription opioid users have better treatment outcomes on
average, likely resulting in large part from better a variety of positive predictive
factors (Potter et al. 2013; Weiss and Rao 2017).
Four drugs carry FDA indications for use in the treatment of OUD: methadone, a
long-acting mu-receptor agonist approved for both detoxification and maintenance
treatment; buprenorphine, a mu-opioid receptor partial agonist, which is also
marketed as a sublingual film combination product of buprenorphine and naloxone
(a mu-receptor antagonist) approved for maintenance treatment; naltrexone, an
opioid receptor antagonist approved for maintenance when delivered in a long-
acting injectable form; and lofexidine, a structural analog of clonidine that agonizes
central α2 receptors, approved for the mitigation of opiate withdrawal symptoms.
Although lofexidine was approved for the treatment of opiate withdrawal in the
United Kingdom in 1992, it did not receive approval in the United States until 2018.
Additionally, an extended-release version of buprenorphine delivered as a weekly or
monthly injection has received tentative approval from the FDA (FDA 2018b).
Acute treatment is designed to minimize opioid withdrawal symptoms. Patients
entering acute treatment may complete detoxification or undergo transition to long-
acting opioid treatment with methadone or buprenorphine as part of agonist main-
tenance. RCTs vary in their design depending on whether the intent is for patients to
Randomized Controlled Trials and the Efficacy of Psychotropic Medications 341
Safety and tolerability are important considerations in investigational drug trials but
historically have received less attention than efficacy assessments. Risk is unavoid-
able in evaluating new medications, and the potential benefit to the individual should
be commensurate with the known risks. The risk-benefit ratio may change through
the trial as data accumulates. Particularly in phase I, when a drug is being adminis-
tered to humans for the first time, risks not seen from animal studies may emerge. In
later phases, when the drug is given to hundreds of people with the psychiatric illness
being studied, rarer medical risks may emerge. Monitoring for emergent risks in an
RCT may be conducted by the study sponsor or by an independent Data Safety
Monitoring Committee, which reviews the safety data at intervals while the trial is
ongoing. New safety concerns, such as liver or cardiac toxicity, that emerge during
the trial may sufficiently increase the risk to result in trial termination. In addition,
the potential “futility” of the intervention may be evaluated by a statistical analysis of
the data at an interim assessment, which may determine that even with continued
342 B. W. Dunlop and C. Medeiros Da Frota Ribeiro
When a drug receives marketing approval, the number of patients who have been
exposed to the compound in phase I–III trials is usually only on the order of several
hundred. Thus, rare side effects or other complications of treatment may not have
been detected, leading to the need for post-marketing pharmacovigilance for risks
emerging in the broader population. For example, the US FDA may require the
sponsor to conduct additional studies or surveillance after giving approval to market
a drug. In addition, systems such as the FDA Adverse Event Reporting System
(comprised of safety reports from manufacturers and MedWatch reports clinicians
and patients), the FDA’s Sentinel system (monitoring electronic health records and
administrative claims data), and the World Health Organization’s Program for
International Drug Monitoring all provide means for detecting emerging risks once
a drug is widely prescribed. Pharmacovigilance for liver toxicity has led to warnings
or market removal for several psychiatric medications, including amineptine,
nefazodone, and pemoline (Spina and Trifirò 2016). In addition, the FDA Center
for Drug Evaluation and Research conducts a scheduled safety analysis after 10,000
patients have been prescribed the drug or 18 months after approval, whichever is
later.
Despite the success of RCTs in identifying efficacious medications that are widely
adopted by clinicians, several criticisms have been levied against the widespread use
of psychopharmacological drugs and the RCT evidence base supporting their use,
particularly for antidepressants (Moncrieff 2008). These concerns can be broadly
divided into aspects of internal validity, external validity, and safety.
Concerns about internal validity stem from potential for bias in the design,
conduct, or analysis of RCTs. Potential sources of bias in modern phase II or III
clinical trials are listed in Table 3. To mitigate many of the potential biases of RCTs,
the Consolidated Standards of Reporting Trials (CONSORT) have been developed
(http://www.consort-statement.org/), to support the full reporting of trial elements
that could lead to bias or confounding.
Additional concerns have been raised regarding the internal validity of relapse
prevention studies. Because patients who do not benefit from the study medication
during the open-label phase are dropped from the study, these “enrichment” designs
have been critiqued for selection bias, in that only patients who benefit from the
medication are studied (Ghaemi and Selker 2017). Attrition bias (dropout) during the
maintenance phase can be a more significant factor that has unpredictable conse-
quences for effect sizes, in that patients who dropout prior to a relapse are usually
censored in the statistical analysis (i.e., contribute no further data beyond the time of
dropout); the effects of medication may be inflated if these subjects are dropping out
for reasons related to the medication (Nunan et al. 2018).
Generalizability of psychiatric RCT results is a widespread concern (Wisniewski
et al. 2009; Licht 2002; Gilbody et al. 2002). External validity is threatened by the
aforementioned inclusion and exclusion criteria, which greatly narrow the subset of
344 B. W. Dunlop and C. Medeiros Da Frota Ribeiro
Table 3 Sources of bias threatening the internal validity of randomized controlled trials
Type of bias Description Example
Design bias Writing the study protocol in a Using an active comparator at high
way that will artificially enhance doses so that tolerability appears
the efficacy or tolerability of the better for the investigational drug
investigational drug versus the
comparator
Selection bias Enrollment of patients specifically Offering trial participation only to
more likely to show benefit of the patients with good responses in the
investigational drug versus the past to a treatment similar to the
comparator one under study
Measurement The scoring of clinical measures is The clinician assessing side effects
(ascertainment) bias influenced by the investigator/ is the same one rating efficacy,
assessor which increases the likelihood of
unblinding
Analytic bias Statistical analysis of the study Analyzing as the primary outcome
data in a way that differs from the efficacy detected on a self-report
originally specified plan, including measure of symptoms instead of
changes in the primary outcome the pre-specified primary
reported, exclusion of certain clinician-rated scale
subjects from analysis, and sub-
group analyses without indicating
the testing to be exploratory
Competing interest Investigators or sponsors with a Decisions made during the trial,
bias financial interest in the results of such as dose increases in a flexible
the trial, usually acting via one of dose trial, may be influenced by
the other sources of bias investigators with a financial
interest in the trial’s outcome
Lewis and Warlow (2004), Gul (2016)
patients with a disorder who are studied. For example, two widely reported meta-
analyses concluded that antidepressants are not meaningfully better than placebo for
non-severe forms of major depression (Kirsch et al. 2008; Fournier et al. 2010).
However, individual patient-level meta-analyses have found that SSRIs and SNRIs
are roughly equally effective across the range of severity of MDD (Gibbons et al.
2012; Furukawa et al. 2018). Furthermore, the effect size of antidepressants is even
larger in placebo-controlled RCTs for dysthymia (a form of milder, chronic depres-
sion) than for MDD (Levkovitz et al. 2011). Effectiveness studies (also known as
large simple trials) are used to assess the broader clinical utility of drugs established
by RCTs to have therapeutic efficacy.
Another criticism is the emphasis on symptom change outcome measures in
RCTs, which may inappropriately focus the goals of treatment away from those
that patients most care about: improvements in quality of life and functional status
(Zimmerman et al. 2012b). Symptom improvement and functional gains are corre-
lated, though symptom improvement typically precedes gains in functioning and
quality of life (McKnight and Kashdan 2009; Sheehan et al. 2017). The categorical
outcome of “response” in depression trials has also been criticized as a statistically
Randomized Controlled Trials and the Efficacy of Psychotropic Medications 345
misused measure to create an illusion of drug efficacy (Kirsch and Moncrieff 2007).
Notably, the categorical outcome of remission has not been similarly criticized,
perhaps given its clear association with better functioning and duration of wellness
(Möller 2008).
Another form of bias is the selective publication of trial results, which can convey
false impression of a drug’s efficacy and distort meta-analyses (Kirkham et al. 2010;
Dal-Ré et al. 2017). These forms of bias include failing or delaying publication of
negative trials or reporting a trial as positive using a secondary outcome measure
when the primary outcome did not find efficacy; these problems have been identified
in mood, anxiety, and psychotic disorder clinical trials (Turner et al. 2008; Roest et
al. 2015; Lancee et al. 2017). The effect of publication bias on overall estimates of
drug efficacy appears to be small (roughly 15% relative increase in effect size), at
least for antidepressants used for anxiety disorders (Roest et al. 2015). Pharmaceu-
tical industry sponsorship and investigator conflicts of interest in RCTs in particular
have been identified as a source of publication bias (Lundh et al. 2012). Concerns
about publication and analytic bias produced pressures for more transparency about
clinical trials. In the United States, the 1997 FDA Modernization Act required the
NIH to create a registry of clinical trials; the subsequent 2007 FDA Amendments Act
greatly expanded the types of trials to be included and required submission of results
after trial completion (Avorn et al. 2018). From this legislation was born
clinicaltrials.gov, the first and largest registry of RCTs. The European Union
launched the EU Clinical Trials Register (clinicaltrialsregister.eu) in 2011. Many
other registries now exist around the globe, allowing patients to easily search for
studies looking for participants and to access summarized results of prior research.
Among the safety concerns is the underreporting of side effects as discussed
previously, particularly if the side effect is similar to a symptom that occurs as part of
the disorder (e.g., fatigue is a symptom of depression, but can also be a side effect of
antidepressant medications) (Hughes et al. 2014). There is also a relative paucity of
safety evaluations of chronic medication use, which is how many psychotropic
agents are prescribed in clinical practice. Specific concerns have been raised about
chronic use of antidepressants worsening the long-term course of major depression
(Fava 1994; El-Mallakh et al. 2011) and of antipsychotics deleteriously affecting
brain volume (Ho et al. 2011). Unfortunately, observed long-term adverse outcomes
with psychotropic agents may be confounded by the natural progression of the
disease process, which may be associated with chronic use of the prescribed
medications; adequate data to resolve this issue does not exist (Targum 2014; Goff
et al. 2017). An additional concern is that many drugs studied in adults are used also
in children and adolescents, though little safety data is available for these age groups.
“Me-too” Drugs
costs without adding value to patient care (Angell 2005). Some critics assert that new
drugs should not be approved unless they demonstrate superiority over an active
comparator already on the market. Common examples of me-too drugs cited in
psychiatry include SSRIs and atypical antipsychotics. While there is no question that
the pharmaceutical industry has engaged in shameful and egregious promotional
practices (Qureshi et al. 2011), whether drugs labeled as “me-too” products are
problematic is less clear (Huskamp 2006).
The challenge of designating a new psychiatric drug as a “me-too” agent is that
the pathophysiology of psychiatric disorders is complex and poorly understood.
Drugs that do not show overall better efficacy may nevertheless prove highly
valuable for certain patients, due to biological characteristics that are relevant to a
specific drug response but not yet known to science. For example, if bupropion
(a marketed antidepressant) was in fact being tested as an investigational medication
today, it would fail to pass the superior efficacy test because it does not produce
greater mean change or remission rates than SSRIs in patients with depression
(Cipriani et al. 2018). However, the medication is highly valuable in clinical settings,
as some patients can respond better to it than SSRIs or find it more tolerable or better
treats a comorbid disorder, though they may be a smaller proportion of the overall
depressed population.
It is also unclear what level of pharmacological difference justifies a drug to be
considered novel versus a copy of an existing drug. For example, vilazodone is
classified as an SSRI but also acts as a 5HT1a receptor partial agonist. This
secondary action does not produce a superior antidepressant effect but appears to
cause less sexual dysfunction than the older SSRIs (Cipriani et al. 2018; Clayton et
al. 2017). Whether such a difference should exclude it from a “me-too” designation
is an open question. Furthermore, drugs with similar pharmacodynamic mechanisms
of action may differ greatly in their pharmacokinetic properties, including metabo-
lism pathways, half-lives, and susceptibility to extrusion from the CNS via p-gly-
coprotein, which may all may be affected by an individual’s specific genotypes. For
example, fluoxetine and paroxetine were two of the first SSRIs. Both are very strong
inhibitors of cytochrome P450 2D6, and sertraline is a moderate inhibitor of this
enzyme. Without the development of citalopram, which has minimal inhibitory
effects on CYP2D6, patients who were on other drugs with a narrow therapeutic
index that are metabolized through CYP2D6 (e.g., tricyclic antidepressants) or a
prodrug that needs CYP2D6 activity to create an active metabolite (e.g. tamoxifen)
would face difficult clinical options. Finally, patients who do not respond to one
SSRI often respond to a second one, suggesting that classifying “me-too” drugs
simply on their presumed mechanism of action is simplistic and carries a significant
risk of excluding potentially useful medications from reaching clinical practice.
Summary
must be used to minimize bias and confounding that could lead to false conclusions.
Phase II and III RCTs designed to assess efficacy necessarily use inclusion and
exclusion criteria that may reduce generalizability of the results; large simple trials of
effectiveness in broader populations are needed to determine how efficacious med-
ications are best used in clinical settings. Despite the limitations of psychiatric RCTs,
they remain the best means for identifying treatments that truly work to improve
mental health.
References
Alonso J, Lepine JP, Committee for the European Study of Mental Disorders. Overview of key data
from the European Study of the Epidemiology of Mental Disorders (ESEMeD). J Clin Psychi-
atry. 2007;68(suppl 2):3–9.
Alphs L, Schooler N, Lauriello J. How study designs influence comparative effectiveness out-
comes: the case of oral versus long-acting injectable antipsychotic treatments for schizophrenia.
Schizophr Res Treat. 2014;156(2–3):228–32.
Alphs L, Mao L, Lynn Starr H, Benson C. A pragmatic analysis comparing once-monthly
paliperidone palmitate versus daily oral antipsychotic treatment in patients with schizophrenia.
Schizophr Res. 2016;170(2–3):259–64.
Altamura AC, Moro AR, Percudani M. Clinical pharmacokinetics of fluoxetine. Clin
Pharmacokinet. 1994;26(3):201–14.
Altman DG, Gore SM, Gardner MJ, Pocock SJ. Statistical guidelines for contributors to medical
journals. Br Med J. 1983;286:1489–93.
American Psychiatric Association. Practice guideline for the treatment of patients with major
depressive disorder. 3rd ed. Washington, DC: American Psychiatric Association; 2010.
American Psychological Association. PTSD Assessment Instruments. 2017. https://www.apa.org/
ptsd-guideline/assessment/index.aspx. Accessed 19 Dec 2018.
Ammerman S, Ryan S, Adelman WP, Committee on Substance Abuse, the Committee on Adoles-
cence. The impact of marijuana policies on youth: clinical, research, and legal update. Pediat-
rics. 2015;135(3):e769–85.
Anderson IM. Meta-analytical studies on new antidepressants. Br Med Bull. 2001;57:161–78.
Andreasen NC. Negative symptoms in schizophrenia: definition and reliability. Arch Gen Psychi-
atry. 1982;39:784–8.
Andreasen NC, Olsen S. Negative v positive schizophrenia: definition and validation. Arch Gen
Psychiatry. 1982;39:789–94.
Andreasen NC, Carpenter WT Jr, Kane JM, Lasser RA, Marder SR, Weinberger DR. Remission in
schizophrenia: proposed criteria and rationale for consensus. Am J Psychiatry. 2005;162
(3):441–9.
Angell M. The truth about the drug companies: how they deceive us and what to do about it. New
York: Random House; 2005.
Angst J, Grobler C. Unipolar mania: a necessary diagnostic concept. Eur Arch Psychiatry Clin
Neurosci. 2015;265(4):273–80.
Anton RF, O’Malley SS, Ciraulo DA, Cisler RA, Couper D, Donovan DM, et al. Combined
pharmacotherapies and behavioral interventions for alcohol dependence: the COMBINE
study: a randomized controlled trial. JAMA. 2006;295(17):2003–17.
Atkins DC. Clinical trials methodology: randomization, intent-to-treat, and random-effects regres-
sion. Depress Anxiety. 2009;26:697–700.
Avorn J, Kesselheim A, Sarpatwari A. The FDA Amendments Act of 2007 – assessing its effects a
decade later. N Engl J Med. 2018;379(12):1097–9.
Bagby RM, Ryder AG, Schuller DR, Marshall MB. The Hamilton Depression Rating Scale: has the
gold standard become a lead weight? Am J Psychiatry. 2004;161(12):2163–77.
348 B. W. Dunlop and C. Medeiros Da Frota Ribeiro
Bandelow B, Baldwin DS, Dolberg OT, Andersen HF, Stein DJ. What is the threshold for
symptomatic response and remission for major depressive disorder, panic disorder, social
anxiety disorder, and generalized anxiety disorder? J Clin Psychiatry. 2006;67(9):1428–34.
Barbano AC, van der Mei WF, Bryant RA, Delahanty DL, deRoon-Cassini TA, Matsuoka YJ, et al.
Clinical implications of the proposed ICD-11 PTSD diagnostic criteria. Psychol Med.
2018;14:1–8.
Barnes TR. A rating scale for drug-induced akathisia. Br J Psychiatry. 1989;154:672–6.
Batelaan NM, Bosman RC, Muntingh A, Scholten WD, Huijbregts KM, van Balkom AJLM. Risk
of relapse after antidepressant discontinuation in anxiety disorders, obsessive-compulsive
disorder, and post-traumatic stress disorder: systematic review and meta-analysis of relapse
prevention trials. BMJ. 2017;358:j3927.
Bech P, Bolwig TG, Kramp P, Rafaelsen OJ. The Bech-Rafaelsen Mania Scale and the Hamilton
Depression Scale. Acta Psychiatr Scand. 1979;59:420–30.
Bech P, Allerup P, Gram LF, Reisby N, Rosenberg R, Jacobsen O, et al. The Hamilton depression
scale. Evaluation of objectivity using logistic models. Acta Psychiatr Scand. 1981;63:290–9.
Beck AT, Ward CH, Mendelson M, Mock J, Erbaugh J. An inventory for measuring depression.
Arch Gen Psychiatry. 1961;4:561–71.
Bighelli I, Salanti G, Huhn M, Schneider-Thoma J, Krause M, Reitneir C, et al. Psychological
interventions to reduce positive symptoms in schizophrenia: systematic review and network
meta-analysis. World Psychiatry. 2018;17(3):316–29.
Blake DD, Weathers FW, Nagy LM, Kaloupek DG, Gusman FD, Charney DS, et al. The develop-
ment of a clinician-administered PTSD scale. J Trauma Stress. 1995;8:75–90.
Bloch MH, Landeros-Weisenberger A, Kelmendi B, Coric V, Bracken MB, Leckman JF.
A systematic review: antipsychotic augmentation with treatment refractory obsessive-compul-
sive disorder. Mol Psychiatry. 2006;11(7):622–32.
Boessen R, Groenwold RH, Knol MJ, Grobbee DE, Roes KC. Comparing HAMD(17) and HAMD
subscales on their ability to differentiate active treatment from placebo in randomized controlled
trials. J Affect Disord. 2013;145(3):363–9.
Borges S, Chen YF, Laughren TP, Temple R, Patel HD, David PA, et al. Review of maintenance
trials for major depressive disorder: a 25-year perspective from the US Food and Drug
Administration. J Clin Psychiatry. 2014;75(3):205–14.
Boschloo L, Bekhuis E, Borsboom D, Weitz ES, Reijnders M, DeRubeis RJ, et al. The symptom-
specific efficacy of cognitive behavioral therapy versus antidepressant medication in the treat-
ment of depression: results from an individual patient data meta-analysis. World Psychiatry.
2019;18(2):183–91.
Brendel DH, Miller FG. A plea for pragmatism in clinical research ethics. Am J Bioeth. 2008;8
(4):24–31.
Buchanan RW, Kreyenbuhl J, Kelly DL, Noel JM, Boggs DL, Fischer BA, et al. Schizophrenia
patient outcomes research team (PORT). The 2009 schizophrenia PORT psychopharmacolog-
ical treatment recommendations and summary statements. Schizophr Bull. 2010;36(1):71–93.
Burcusa SL, Iacono WG. Risk for recurrence in depression. Clin Psychol Rev. 2007;27:959–85.
Calabrese JR, Keck PE, Macfadden W, Minkwitz M, Ketter TA, Weisler RH, et al. A randomized,
double-blind, placebo-controlled trial of quetiapine in the treatment of bipolar I or II depression.
Am J Psychiatry. 2005;162:1351–60.
Canadian Agency for Drugs and Technologies in Health. Appendix 9, Inclusion and exclusion
criteria reported in included RCTs. In: A Systematic Review of Combination and High-Dose
Atypical Antipsychotic Therapy in Patients with Schizophrenia. Canadian Agency for Drugs
and Technologies in Health. 2011. https://www.ncbi.nlm.nih.gov/books/NBK169687/.
Accessed 21 Jan 2019.
Catalá-López F, Hutton B, Núñez-Beltrán A, Page MJ, Ridao M, Macías Saint-Gerons D, et al. The
pharmacological and non-pharmacological treatment of attention deficit hyperactivity disorder
in children and adolescents: a systematic review with network meta-analyses of randomised
trials. PLoS One. 2017;12(7):e0180355.
Randomized Controlled Trials and the Efficacy of Psychotropic Medications 349
Dunlop BW, Kaye JL, Youngner C, Rothbaum B. Assessing treatment-resistant posttraumatic stress
disorder: the Emory Treatment Resistance Interview for PTSD (E-TRIP). Behav Sci (Basel).
2014;4(4):511–27.
Dunlop BW, Gray J, Rapaport MH. Transdiagnostic clinical global impression scoring for routine
clinical settings. Behav Sci (Basel). 2017;7(4):40.
Dunlop BW, Polychroniou P, Rakofsky JJ, Nemeroff CB, Craighead WE, Mayberg HS. Suicidal
ideation and other persisting symptoms after CBT or antidepressant medication treatment for
major depressive disorder. Psychol Med. 2018; epub ahead of print.
DuPaul GJ, Power TJ, Anastopoulos AD, Reid R. ADHD rating scale-IV: checklists, norms, and
clinical interpretation. New York: Guilford Press; 1998.
El-Mallakh RS, Gao Y, Jeannie RR. Tardive dysphoria: the role of long term antidepressant use in-
inducing chronic depression. Med Hypotheses. 2011;76(6):769–73.
Erford BT, Jackson J, Bardhoshi G, Duncan K, Atalay Z. Selecting suicide ideation assessment
instruments: a meta-analytic review. Meas Eval Couns Dev. 2018;51:42–59.
European Medicines Agency. Guideline on clinical investigation of medicinal products, including
depot preparations in the treatment of schizophrenia. In: Committee for Medicinal Products for
Human Use; 2012.
European Medicines Agency. Guideline on the clinical investigation of medicines for the treatment
of Alzheimer’s disease. In: Committee for Medicinal Products for Human Use; 2018.
Fagiolini A, Rocca P, De Giorgi S, Spina E, Amodeo G, Amore M. Clinical trial methodology to
assess the efficacy/effectiveness of long-acting antipsychotics: randomized controlled trials vs
naturalistic studies. Psychiatry Res. 2017;247:257–64.
Falk D, Wang XQ, Liu L, Fertig J, Mattson M, Ryan M, et al. Percentage of subjects with no heavy
drinking days: evaluation as an efficacy endpoint for alcohol clinical trials. Alcohol Clin Exp
Res. 2010;34:2022–34.
Faraone SV, Biederman J, Mick E. The age-dependent decline of attention deficit hyperactivity
disorder: a meta-analysis of follow-up studies. Psychol Med. 2006;36(2):159–65.
Fava GA. Do antidepressant and antianxiety drugs increase chronicity in affective disorders?
Psychother Psychosom. 1994;61(3–4):125–31.
Fava M, Freeman MP, Flynn M, Judge H, Hoeppner BB, Cusin C, et al. Double-blind, placebo-
controlled, dose-ranging trial of intravenous ketamine as adjunctive therapy in treatment-
resistant depression (TRD). Mol Psychiatry. 2018; epub ahead of print.
Fekadu A, Donocik JG, Cleare AJ. Standardisation framework for the Maudsley staging method for
treatment resistance in depression. BMC Psychiatry. 2018;18(1):100.
Ferrer P, Ballarín E, Sabaté M, Vidal X, Rottenkolber M, Amelio J, et al. Antiepileptic drugs and
suicide: a systematic review of adverse effects. Neuroepidemiology. 2014;42(2):107–20.
Fineberg NA, Pampaloni I, Pallanti S, Ipser J, Stein DJ. Sustained response versus relapse: the
pharmacotherapeutic goal for obsessive-compulsive disorder. Int Clin Psychopharmacol.
2007;22(6):313–22.
Fineberg NA, Brown A, Reghunandanan S, Pampaloni I. Evidence-based pharmacotherapy of
obsessive-compulsive disorder. Int J Neuropsychopharmacol. 2012;15(8):1173–91.
First MB, Williams JBW, Karg RS, Spitzer RL. Structured clinical interview for DSM-5 – research
version. Washington, DC: American Psychiatric Association; 2015.
Fitzgerald N, Angus K, Elders A, de Andrade M, Raistrick D, Heather N, et al. Weak evidence on
nalmefene creates dilemmas for clinicians and poses questions for regulators and researchers.
Addiction. 2016;111(8):1477–87.
Food and Drug Administration. Guidance for industry: alcoholism: developing drugs for treatment.
Silver Spring: Department of Health and Human Services; 2015.
Food and Drug Administration. Step 3: clinical research. 2018a. https://www.fda.gov/patients/drug-
development-process/step-3-clinical-research. Accessed 11 Sept 2019.
Food and Drug Administration. Guidance for industry: opioid use disorder: endpoints for demon-
strating effectiveness of drugs for medication-assisted treatment. Silver Spring, Maryland:
Department of Health and Human Services; 2018b.
Randomized Controlled Trials and the Efficacy of Psychotropic Medications 351
Fournier JC, DeRubeis RJ, Hollon SD, Dimidjian S, Amsterdam JD, Shelton RC, et al. Antide-
pressant drug effects and depression severity: a patient-level meta-analysis. JAMA.
2010;303:47–53.
Frank JD. Persuasion & healing. Baltimore: Johns Hopkins University Press; 1973.
Frank E, Prien RF, Jarrett RB, Keller MB, Kupfer DJ, Lavori PW, et al. Conceptualization and
rationale for consensus definitions of terms in major depressive disorder. Remission, recovery,
relapse, and recurrence. Arch Gen Psychiatry. 1991;48(9):851–5.
Fresco DM, Coles ME, Heimberg RG, Liebowitz MR, Hami S, Stein MB, et al. The Liebowitz
Social Anxiety Scale: a comparison of the psychometric properties of self-report and clinician-
administered formats. Psychol Med. 2001;31(6):1025–35.
Furukawa TA, Maruo K, Noma H, Tanaka S, Imai H, Shinohara K, et al. Initial severity of major
depression and efficacy of new generation antidepressants: individual participant data meta-
analysis. Acta Psychiatr Scand. 2018;137(6):450–8.
GBD 2016 Alcohol and Drug Use Collaborators. The global burden of disease attributable to
alcohol and drug use in 195 countries and territories, 1990–2016: a systematic analysis for the
global burden of disease study 2016. Lancet Psychiatry. 2018;5(12):987–1012.
Ghaemi SN, Selker HP. Maintenance efficacy designs in psychiatry: randomized discontinuation
trials – enriched but not better. J Clin Transl Sci. 2017;1(3):198–204.
Gianarris WJ, Golden CJ, Greene L. The Conners’ Parent Rating Scales: a critical review of the
literature. Clin Psychol Rev. 2001;21(7):1061–93.
Gibbons RD, Hur K, Brown CH, Davis JM, Mann JJ. Benefits from antidepressants: synthesis of 6-
week patient-level outcomes from double-blind placebo-controlled randomized trials of fluox-
etine and venlafaxine. Arch Gen Psychiatry. 2012;69(6):572–9.
Gibertini M, Nations KR, Whitaker JA. Obtained effect size as a function of sample size in
approved antidepressants: a real-world illustration in support of better trial design. Int Clin
Psychopharmacol. 2012;27(2):100–6.
Gilbody S, Wahlbeck K, Adams C. Randomized controlled trials in schizophrenia: a critical
perspective on the literature. Acta Psychiatr Scand. 2002;105(4):243–51.
Gitlin MJ. Antidepressants in bipolar depression: an enduring controversy. Int J Bipolar Disord.
2018;6(1):25.
Glue P, Donovan MR, Kolluri S, Emir B. Meta-analysis of relapse prevention antidepressant trials
in depressive disorders. Aust N Z J Psychiatry. 2010;44:697–705.
Goff DC, Falkai P, Fleischhacker WW, Girgis RR, Kahn RM, Uchida H, et al. The long-term effects
of antipsychotic medication on clinical course in schizophrenia. Am J Psychiatry. 2017;174
(9):840–9.
Gomez AF, Barthel AL, Hofmann SG. Comparing the efficacy of benzodiazepines and serotonergic
anti-depressants for adults with generalized anxiety disorder: a meta-analytic review. Expert
Opin Pharmacother. 2018;19(8):883–94.
Goodman WK, Price LH, Rasmussen SA, Mazure C, Fleischmann RL, Hill CL, et al. The Yale-
Brown Obsessive Compulsive Scale. I. Development, use, and reliability. Arch Gen Psychiatry.
1989;46(11):1006–11.
Goodman D, Faraone SV, Adler LA, Dirks B, Hamdani M, Weisler R. Interpreting ADHD rating
scale scores: linking ADHD rating scale scores and CGI levels in two randomized controlled
trials of lisdexamfetamine dimesylate in ADHD. Primary Psychiatry. 2010;17(3):44–52.
Gul M. Bias in a randomized controlled trial and how these can be minimized. J Psychiatry.
2016;19:2.
Guy W. Clinical global impressions. ECDEU assessment manual for psychopharmacology. Rock-
ville: US Department of Health, Education, and Welfare; 1976. p. 217–22.
Hamer RM, Simpson PM. Last observation carried forward versus mixed models in the analysis of
psychiatric clinical trials. Am J Psychiatry. 2009;166(6):639–41.
Hamilton M. The assessment of anxiety states by rating. Br J Med Psychol. 1959;32:50–5.
Hamilton M. Development of a rating scale for primary depressive illness. Br J Soc Clin Psychol.
1967;6:278–96.
352 B. W. Dunlop and C. Medeiros Da Frota Ribeiro
Hasan A, Falkai P, Wobrock T, Lieberman J, Glenthoj B, Gattaz WF, et al. World Federation of
Societies of Biological Psychiatry (WFSBP) task force on treatment guidelines for schizo-
phrenia. World Federation of Societies of Biological Psychiatry (WFSBP) guidelines for
biological treatment of schizophrenia, part 1: update 2012 on the acute treatment of schizo-
phrenia and the management of treatment resistance. World J Biol Psychiatry. 2012;13
(5):318–78.
Heatherton TF, Kozlowski LT, Frecker RC, Fagerstrom KO. The Fagerstrom test for nicotine
dependence: a revision of the Fagerstrom tolerance questionnaire. Br J Addict.
1991;86:1119–27.
Heekin RD, Shorter D, Kosten TR. Current status and future prospects for the development of
substance abuse vaccines. Expert Rev. Vaccines. 2017;16(11):1067–77.
Hedges L. Distribution theory for Glass’s estimator of effect size and related estimators. J Educ Stat.
1981;6(2):107–28.
Henry ME, Schmidt ME, Hennen J, Villafuerte RA, Butman ML, Tran P, et al. A comparison of
brain and serum pharmacokinetics of R-fluoxetine and racemic fluoxetine: a 19-F MRS study.
Neuropsychopharmacology. 2005;30(8):1576–83.
Henssler J, Kurschus M, Franklin J, Bschor T, Baethge C. Trajectories of acute antidepressant
efficacy: how long to wait for response? A systematic review and meta-analysis of long-term,
placebo-controlled acute treatment trials. J Clin Psychiatry. 2018;79(3):pii:17r11470.
Hieronymus F, Nilsson S, Eriksson E. A mega-analysis of fixed-dose trials reveals dose-depen-
dency and a rapid onset of action for the antidepressant effect of three selective serotonin
reuptake inhibitors. Transl Psychiatry. 2016;6(6):e834.
Hindmarch I. Beyond the monoamine hypothesis: mechanisms, molecules and methods. Eur
Psychiatry. 2002;17(3):294–9.
Hirschtritt ME, Bloch MH, Mathews CA. Obsessive-compulsive disorder: advances in diagnosis
and treatment. JAMA. 2017;317(13):1358–67.
Ho BC, Andreasen NC, Ziebell S, Pierson R, Magnotta V. Long-term antipsychotic treatment and
brain volumes: a longitudinal study of first-episode schizophrenia. Arch Gen Psychiatry.
2011;68(2):128–37.
Hoek HW. Review of the worldwide epidemiology of eating disorders. Curr Opin Psychiatry.
2016;29(6):336–9.
Hoertel N, Le Strat Y, Blanco C, Lavaud P, Dubertret C. Generalizability of clinical trial results for
generalized anxiety disorder to community samples. Depress Anxiety. 2012;29(7):614–20.
Hoskins M, Pearce J, Bethell A, Dankova L, Barbui C, Tol WA, et al. Pharmacotherapy for post-
traumatic stress disorder: systematic review and meta-analysis. Br J Psychiatry. 2015;206
(2):93–100.
Howes OD, McCutcheon R, Agid O, de Bartolomeis A, van Beveren NJ, Birnbaum ML, et al.
Treatment-resistant schizophrenia: treatment response and resistance in psychosis (TRRIP)
working group consensus guidelines on diagnosis and terminology. Am J Psychiatry.
2017;174(3):216–29.
Hudson JI, McElroy SL, Ferreira-Cornwell MC, Radewonuk J, Gasior M. Efficacy of lisdexam-
fetamine in adults with moderate to severe binge-eating disorder: a randomized clinical trial.
JAMA Psychiat. 2017;74(9):903–10.
Hughes JR, Keely JP, Niaura RS, Ossip-Klein DJ, Richmond RL, Swan GE. Measures of abstinence
in clinical trials: issues and recommendations. Nicotine Tob Res. 2003;5(1):13–25.
Hughes S, Cohen D, Jaggi R. Differences in reporting serious adverse events in industry sponsored
clinical trial registries and journal articles on antidepressant and antipsychotic drugs: a cross
sectional study. BMJ Open. 2014;4:e005535.
Huskamp HA. Prices, profits, and innovation: examining criticisms of new psychotropic drugs'
value. Health Aff (Millwood). 2006;25(3):635–46.
Imai H, Tajika A, Chen P, Pompoli A, Furukawa TA. Psychological therapies versus pharmaco-
logical interventions for panic disorder with or without agoraphobia in adults. Cochrane
Database Syst Rev. 2016;10:CD011170.
Randomized Controlled Trials and the Efficacy of Psychotropic Medications 353
Insel TR, Murphy DL, Cohen RM, Alterman I, Kilts C, Linnoila M. Obsessive compulsive disorder.
A double-blind trial of clomipramine and clorgyline. Arch Gen Psychiatry. 1983;40:605–12.
Jatlow P, Toll BA, Leary V, Krishnan-Sarin S, O'Malley SS. Comparison of expired carbon
monoxide and plasma cotinine as markers of cigarette abstinence. Drug Alcohol Depend.
2008;98(3):203–9.
Jonas DE, Amick HR, Feltner C, Bobashev G, Thomas K, Wines R, et al. Pharmacotherapy for
adults with alcohol use disorders in outpatient settings: a systematic review and meta-analysis.
JAMA. 2014;311(18):1889–900.
Judd LL, Akiskal HS, Maser JD, Zeller PJ, Endicott J, Coryell W, et al. A prospective 12-year study
of subsyndromal and syndromal depressive symptoms in unipolar major depressive disorders.
Arch Gen Psychiatry. 1998;55(8):694–700.
Judd LL, Akiskal HS, Schettler PJ, Endicott J, Leon AC, Solomon DA, et al. Psychosocial disability
in the course of bipolar I and II disorders: a prospective, comparative, longitudinal study. Arch
Gen Psychiatry. 2005;62(12):1322–30.
Junod SW. FDA and clinical drug trials: a short history. In: Davies M, Kerimani F, editors. A quick
guide to clinical trials. Washington, DC: Bioplan, Inc; 2008. p. 25–55.
Kane JM, Kishimoto T, Correll CU. Non-adherence to medication in patients with psychotic
disorders: epidemiology, contributing factors and management strategies. World Psychiatry.
2013;12(3):216–26.
Katzman MA, Bleau P, Blier P, Chokka P, Kjernisted K, Van Ameringen M, et al. Canadian clinical
practice guidelines for the management of anxiety, posttraumatic stress and obsessive-compul-
sive disorders. BMC Psychiatry. 2014;14(Suppl 1):S1.
Kay SR, Fiszbein A, Opler LA. The positive and negative syndrome scale (PANSS) for schizo-
phrenia. Schizophr Bull. 1987;13(2):261–76.
Kessler RC, Bromet EJ. The epidemiology of depression across cultures. Annu Rev Public Health.
2013;34:119–38.
Kessler RC, Ustün TB. The World Mental Health (WMH) survey initiative version of the World
Health Organization (WHO) Composite International Diagnostic Interview (CIDI). Int J
Methods Psychiatr Res. 2004;13(2):93–121.
Kessler RC, Chiu WT, Demler O, Merikangas KR, Walters EE. Prevalence, severity, and comor-
bidity of 12-month DSM-IV disorders in the National Comorbidity Survey Replication. Arch
Gen Psychiatry. 2005;62(6):617–27.
Khan A, Khan SR, Walens G, Kolts R, Giller EL. Frequency of positive studies among fixed and
flexible dose antidepressant clinical trials: an analysis of the food and drug administration
summary basis of approval reports. Neuropsychopharmacology. 2003;28(3):552–7.
Khan A, Schwartz K, Kolts RL, Ridgway D, Lineberry C. Relationship between depression severity
entry criteria and antidepressant clinical trial outcomes. Biol Psychiatry. 2007;62(1):65–71.
Khin NA, Chen YF, Yang Y, Yang P, Laughren TP. Exploratory analyses of efficacy data from major
depressive disorder trials submitted to the US Food and Drug Administration in support of new
drug applications. J Clin Psychiatry. 2011;72(4):464–72.
Kirkham JJ, Dwan KM, Altman DG, Gamble C, Dodd S, Smyth R, et al. The impact of outcome
reporting bias in randomised controlled trials on a cohort of systematic reviews. BMJ. 2010;340:
c365.
Kirsch I, Moncrieff J. Clinical trials and the response rate illusion. Contemp Clin Trials.
2007;28:348–51.
Kirsch I, Deacon BJ, Huedo-Medina TB, et al. Initial severity and antidepressant benefits: a meta-
analysis of data submitted to the Food and Drug Administration. PLoS Med. 2008;5(2):e45.
Kishimoto T, Nitta M, Borenstein M, Kane JM, Correll CU. Long-acting injectable versus oral
antipsychotics in schizophrenia: a systematic review and meta-analysis of mirror-image studies.
J Clin Psychiatry. 2013;74(10):957–65.
Kishimoto T, Robenzadeh A, Leucht C, Leucht S, Watanabe K, Mimura M, et al. Long-acting
injectable vs oral antipsychotics for relapse prevention in schizophrenia: a meta-analysis of
randomized trials. Schizophr Bull. 2014;40(1):192–213.
354 B. W. Dunlop and C. Medeiros Da Frota Ribeiro
Kishimoto T, Hagi K, Nitta M, Leucht S, Olfson M, Kane JM, et al. Effectiveness of long-acting
injectable vs oral antipsychotics in patients with schizophrenia: a meta-analysis of prospective
and retrospective cohort studies. Schizophr Bull. 2018;44(3):603–19.
Kobak KA, Leuchter A, DeBrota D, Engelhardt N, Williams JB, Cook IA, et al. Site versus
centralized raters in a clinical depression trial: impact on patient selection and placebo response.
J Clin Psychopharmacol. 2010;30(2):193–7.
Kranzler HR, Soyka M. Diagnosis and pharmacotherapy of alcohol use disorder: a review. JAMA.
2018;320(8):815–24.
Lancee M, Lemmens CMC, Kahn RS, Vinkers CH, Luykx JJ. Outcome reporting bias in random-
ized-controlled trials investigating antipsychotic drugs. Transl Psychiatry. 2017;7(9):e1232.
Landin R, DeBrota DJ, DeVries TA, Potter WZ, Demitrack MA. The impact of restrictive entry
criterion during the placebo lead-in period. Biometrics. 2000;56(1):271–8.
Lee NK, Jenner L, Harney A, Cameron J. Pharmacotherapy for amphetamine dependence: a
systematic review. Drug Alcohol Depend. 2018;191:309–37.
Leucht S, Tardy M, Komossa K, Heres S, Kissling W, Salanti G, et al. Antipsychotic drugs versus
placebo for relapse prevention in schizophrenia: a systematic review and metaanalysis. Lancet.
2012;379:2063–71.
Leucht S, Leucht C, Huhn M, Chaimani A, Mavridis D, Helfer B, et al. Sixty years of placebo-
controlled antipsychotic drug trials in acute schizophrenia: systematic review, bayesian meta-
analysis, and meta-regression of efficacy predictors. Am J Psychiatry. 2017;174(10):927–42.
Levine J, Schooler NR. SAFTEE: a technique for the systematic assessment of side effects in
clinical trials. Psychopharmacol Bull. 1986;22(2):343–81.
Levkovitz Y, Tedeschini E, Papakostas GI. Efficacy of antidepressants for dysthymia: a meta-
analysis of placebo-controlled randomized trials. J Clin Psychiatry. 2011;72:509–14.
Lewis SC, Warlow CP. How to spot bias and other potential problems in randomised controlled
trials. J Neurol Neurosurg Psychiatry. 2004;75(2):181–7.
Lewis-Fernández R, Hinton DE, Laria AJ, Patterson EH, Hofmann SG, Craske MG, et al. Culture
and the anxiety disorders: recommendations for DSM-V. Depress Anxiety. 2010;27(2):212–29.
Licht RW. Limits of the applicability and generalizability of drug trials in mania. Bipolar Disord.
2002;4(Suppl 1):66–8.
Lindson-Hawley N, Hartmann-Boyce J, Fanshawe TR, Begh R, Farley A, Lancaster T. Interven-
tions to reduce harm from continued tobacco use. Cochrane Database Syst Rev. 2016;10:
CD005231.
Lindström L, Lindström E, Nilsson M, Höistad M. Maintenance therapy with second generation
antipsychotics for bipolar disorder – a systematic review and meta-analysis. J Affect Disord.
2017;213:138–50.
Lingjaerde O, Ahlfors UG, Bech P, Dencker SJ, Elgen K. The UKU side effect rating scale. A new
comprehensive rating scale for psychotropic drugs and a cross-sectional study of side effects in
neuroleptic-treated patients. Acta Psychiatr Scand Suppl. 1987;334:1–100.
Little RJA, Rubin DB. Statistical analysis with missing data. Hoboken: Wiley; 2002.
Lundh A, Sismondo S, Lexchin J, Busuioc OA, Bero L. Industry sponsorship and research
outcome. Cochrane Database Syst Rev. 2012;12:MR000033.
Maier W, Philipp M. Improving the assessment of severity of depressive states: a reduction of the
Hamilton Depression Scale. Pharmacopsychiatry. 1985;18:114–5.
Mallinckrodt CH, Lane PW, Schnell D, Peng Y, Mancuso JP. Recommendations for the primary
analysis of continuous endpoints in longitudinal clinical trials. Drug Inf J. 2008;42:303–19.
Mark TL, Kassed CA, Vandivort-Warren R, Levit KR, Kranzler HR. Alcohol and opioid depen-
dence medications: prescription trends, overall and by physician specialty. Drug Alcohol
Depend. 2009;99(1–3):345–9.
McCann DJ, Petry NM, Bresell A, Isacsson E, Wilson E, Alexander RC. Medication nonadherence,
“professional subjects,” and apparent placebo responders: overlapping challenges for medica-
tions development. J Clin Psychopharmacol. 2015;35(5):566–73.
Randomized Controlled Trials and the Efficacy of Psychotropic Medications 355
Perna G, Daccò S, Menotti R, Caldirola D. Antianxiety medications for the treatment of complex
agoraphobia: pharmacological interventions for a behavioral condition. Neuropsychiatr Dis
Treat. 2011;7:621–37.
Posner K, Brown GK, Stanley B, Brent DA, Yershova KV, Oquendo MA, et al. The Columbia-
Suicide Severity Rating Scale: initial validity and internal consistency findings from three
multisite studies with adolescents and adults. Am J Psychiatry. 2011;168(12):1266–77.
Potter JS, Marino EN, Hillhouse MP, Nielsen S, Wiest K, Canamar CP, et al. Buprenorphine/
naloxone and methadone maintenance treatment outcomes for opioid analgesic, heroin, and
combined users: findings from starting treatment with agonist replacement therapies (START).
J Stud Alcohol Drugs. 2013;74(4):605–13.
Qureshi ZP, Sartor O, Xirasagar S, Liu Y, Bennett CL. Pharmaceutical fraud and abuse in the United
States, 1996–2010. Arch Intern Med. 2011;171(16):1503–6.
Rabinowitz J, Levine SZ, Barkai O, Davidov O. Dropout rates in randomized clinical trials of
antipsychotics: a meta-analysis comparing first- and second-generation drugs and an examina-
tion of the role of trial design features. Schizophr Bull. 2008;35(4):775–88.
Rabkin JG, Markowitz JS, Ocepek-Welikson K, Wager SS. General versus systematic inquiry about
emergent clinical events with SAFTEE: implications for clinical research. J Clin
Psychopharmacol. 1992;12(1):3–10.
Rickels K, Schweizer E, Case WG, Greenblatt DJ. Long-term therapeutic use of benzodiazepines. I.
Effects of abrupt discontinuation. Arch Gen Psychiatry. 1990;47:899–907.
Rickels K, Downing R, Schweizer E, Hassman H. Antidepressants for the treatment of generalized
anxiety disorder. A placebo-controlled comparison of imipramine, trazodone, and diazepam.
Arch Gen Psychiatry. 1993;50(11):884–95.
Rickels K, Garcia-Espana F, Mandos LA, Case GW. Physician Withdrawal Checklist (PWC-20).
J Clin Psychopharmacol. 2008;28(4):447–51.
Robertson OD, Coronado NG, Sethi R, Berk M, Dodd S. Putative neuroprotective pharmacother-
apies to target the staged progression of mental illness. Early Interv Psychiatry. 2019;13
(5):1032–49.
Robins LN, Wing J, Wittchen HU, Helzer JE, Babor TF, Burke J, et al. The Composite International
Diagnostic Interview: an epidemiologic instrument suitable for use in conjunction with different
diagnostic systems and in different cultures. Arch Gen Psychiatry. 1988;45:1069–77.
Roest AM, de Jonge P, Williams CD, de Vries YA, Schoevers RA, Turner EH. Reporting bias in
clinical trials investigating the efficacy of second-generation antidepressants in the treatment of
anxiety disorders: a report of 2 meta-analyses. JAMA Psychiat. 2015;72:500–10.
Rosenbaum JF, Fava M, Hoog SL, Ascroft RC, Krebs WB. Selective serotonin reuptake inhibitor
discontinuation syndrome: a randomized clinical trial. Biol Psychiatry. 1998;44:77–87.
Rush AJ, Trivedi MH, Ibrahim HM, Carmody TJ, Arnow B, Klein DN, et al. The 16-item Quick
Inventory of Depressive Symptomatology (QIDS), clinician rating (QIDS-C), and self-report
(QIDS-SR): a psychometric evaluation in patients with chronic major depression. Biol Psychi-
atry. 2003;54:573–83.
Rush AJ, Fava M, Wisniewski SR, Lavori PW, Trivedi MH, Sackeim HA, et al. Sequenced
treatment alternatives to relieve depression (STARD): rationale and design. Control Clin Trials.
2004;25:119–42.
Rutherford BR, Pott E, Tandler JM, Wall MM, Roose SP, Lieberman JA. Placebo response in
antipsychotic clinical trials: a meta-analysis. JAMA Psychiat. 2014;71(12):1409–21.
Rutherford BR, Bailey VS, Schneier FR, Pott E, Brown PJ, Roose SP. Influence of study design on
treatment response in anxiety disorder clinical trials. Depress Anxiety. 2015;32(12):944–57.
Sackett DL. Bias in analytic research. J Chronic Dis. 1979;32:51–63.
Salanti G, Chaimani A, Furukawa TA, Higgins JPT, Ogawa Y, Cipriani A, et al. Impact of placebo
arms on outcomes in antidepressant trials: systematic review and meta-regression analysis. Int J
Epidemiol. 2018;47(5):1454–64.
Samara MT, Nikolakopoulou A, Salanti G, Leucht S. How many patients with schizophrenia do not
respond to antipsychotic drugs in the short term? An analysis based on individual patient data
from randomized controlled trials. Schizophr Bull. 2019;45(3):639–46.
Schildkraut JJ, Kety SS. Biogenic amines and emotion. Science. 1967;156(3771):21–30.
Randomized Controlled Trials and the Efficacy of Psychotropic Medications 357
Scholl L, Seth P, Kariisa M, Wilson N, Baldwin G. Drug and opioid-involved overdose deaths –
United States, 2013–2017. WR Morb Mortal Wkly Rep. 2018;67(5152):1419–27.
Schreiner A, Aadamsoo K, Altamura AC, Franco M, Gorwood P, Neznanov NG, et al. Paliperidone
palmitate versus oral antipsychotics in recently diagnosed schizophrenia. Schizophr Res.
2015;169(1–3):393–9.
Selle V, Schalkwijk S, Vázquez GH, Baldessarini RJ. Treatments for acute bipolar depression:
meta-analyses of placebo-controlled, monotherapy trials of anticonvulsants, lithium and anti-
psychotics. Pharmacopsychiatry. 2014;47(2):43–52.
Shear MK, Barlow D, Brown T, et al. Panic Disorder Severity Scale (PDSS). In: Rush AJ, First MB,
Blacker D, editors. Handbook of psychiatric measures. 2nd ed. Washington, DC: American
Psychiatric Association; 2007. p. 542–3.
Sheehan DV, Lecrubier Y, Sheehan KH, Amorim P, Janavs J, Weiller E, et al. The Mini-Interna-
tional Neuropsychiatric Interview (M.I.N.I.): the development and validation of a structured
diagnostic psychiatric interview for DSM-IV and ICD-10. J Clin Psychiatry. 1998;59
(20):22–33.
Sheehan DV, Nakagome K, Asami Y, Pappadopulos EA, Boucher M. Restoring function in major
depressive disorder: a systematic review. J Affect Disord. 2017;215:299–313.
Sherman BJ, McRae-Clark AL. Treatment of cannabis use disorder: current science and future
outlook. Pharmacotherapy. 2016;36(5):511–35.
Shiovitz TM, Bain EE, McCann DJ, Skolnick P, Laughren T, Hanina A, et al. Mitigating the effects
of nonadherence in clinical trials. J Clin Pharmacol. 2016;56(9):1151–64.
Shorter E. A brief history of placebos and clinical trials in psychiatry. Can J Psychiatr. 2011;56
(4):193–7.
Shorter D, Kosten TR. Novel pharmacotherapeutic treatments for cocaine addiction. BMC Med.
2011;9:119.
Siddiqui HM, Hung J, O’Neill R. MMRM vs. LOCF: a comprehensive comparison based on
simulation study and 25 NDA datasets. J Biopharm Stat. 2009;19:227–46.
Sidor MM, MacQueen GM. Antidepressants for acute treatment of bipolar depression: a systematic
review and meta-analysis. J Clin Psychiatry. 2011;72:156–67.
Sigmon SC, Bisaga A, Nunes EV, O'Connor PG, Kosten T, Woody G. Opioid detoxification and
naltrexone induction strategies: recommendations for clinical practice. Am J Drug Alcohol
Abuse. 2012;38(3):187–99.
Sigmon SC, Dunn KE, Saulsgiver K, Patrick ME, Badger GJ, Heil SH, et al. A randomized, double-
blind evaluation of buprenorphine taper duration in primary prescription opioid abusers. JAMA
Psychiat. 2013;70(12):1347–54.
Sim K, Lau WK, Sim J, Sum MY, Baldessarini RJ. Prevention of relapse and recurrence in adults
with major depressive disorder: systematic review and meta-analyses of controlled trials. Int J
Neuropsychopharmacol. 2015;19(2):pii:pyv076.
Simpson GM, Angus JW. A rating scale for extrapyramidal side effects. Acta Psychiat Scand.
1970;212(Suppl):11–9.
Sinyor M, Levitt AJ, Cheung AH, Schaffer A, Kiss A, Dowlati Y, et al. Does inclusion of a
placebo arm influence response to active antidepressant treatment in randomized con-
trolled trials? Results from pooled and meta-analyses. J Clin Psychiatry. 2010;71
(3):270–9.
Skinner MD, Lahmek P, Pham H, Aubin HJ. Disulfiram efficacy in the treatment of alcohol
dependence: a meta-analysis. PLoS One. 2014;9(2):e87366.
Slade E, Keeney E, Mavranezouli I, Dias S, Fou L, Stockton S, et al. Treatments for bulimia
nervosa: a network meta-analysis. Psychol Med. 2018;48(16):2629–36.
Snaith RP, Baugh SJ, Clayden AD, Husain A, Sipple MA. The Clinical Anxiety Scale: an
instrument derived from the Hamilton Anxiety Scale. Br J Psychiatry. 1982;141:518–23.
Soomro GM, Altman DG, Rajagopal S, Oakley BM. Selective serotonin re-uptake inhibitors
(SSRIs) versus placebo for obsessive compulsive disorder (OCD). Cochrane Database Syst
Rev. 2008;1:CD001765.
Spilker B. Guide to clinical trials. Philadelphia: Lippincott Williams & Wilkins; 1984.
Spina E, Trifirò G. Pharmacovigilance in psychiatry. Cham: Adis; 2016.
358 B. W. Dunlop and C. Medeiros Da Frota Ribeiro
Spitzer RL, Endicott J. Schedule for affective disorders and schizophrenia – change version. 3rd ed.
New York: Biometrics Research; 1987.
SRNT Subcommittee on Biochemical Verification. Biochemical verification of tobacco use and
cessation. Nicotine Tob Res. 2002;4(2):149–59.
Stead LF, Lancaster T. Combined pharmacotherapy and behavioural interventions for smoking
cessation. Cochrane Database Syst Rev. 2012;10:CD008286. Update in: Cochrane Database
Syst Rev. 2016;3:CD008286.
Steele M, Jensen PS, Quinn DMP. Remission versus response as the goal of therapy in ADHD: a
new standard for the field? Clin Ther. 2006;28(11):1892–908.
Stein DJ, Ipser JC, Seedat S, Sager C, Amos T. Pharmacotherapy for post-traumatic stress disorder
(PTSD). Cochrane Database Syst Rev. 2006;1:CD002795.
Stein DJ, McLaughlin KA, Koenen KC, Atwoli L, Friedman MJ, Hill ED, et al. DSM-5 and ICD-11
definitions of posttraumatic stress disorder: investigating “narrow” and “broad” approaches.
Depress Anxiety. 2014;31:494–505.
Stone M, Laughren T, Jones ML, Levenson M, Holland PC, Hughes A, et al. Risk of suicidality in
clinical trials of antidepressants in adults: analysis of proprietary data submitted to US Food and
Drug Administration. BMJ. 2009;339:b2880.
Suh JJ, Pettinati HM, Kampman KM, O'Brien CP. The status of disulfiram: a half of a century later.
J Clin Psychopharmacol. 2006;26:290–302.
Svaldi J, Schmitz F, Baur J, Hartmann AS, Legenbauer T, Thaler C, et al. Efficacy of psychother-
apies and pharmacotherapies for bulimia nervosa. Psychol Med. 2019;49(6):898–910.
Targum SD. Identification and treatment of antidepressant tachyphylaxis. Innov Clin Neurosci.
2014;11(3–4):24–8.
Thase ME, Jonas A, Khan A, Bowden CL, Wu X, McQuade RD, et al. Aripiprazole monotherapy in
nonpsychotic bipolar I depression: results of 2 randomized, placebo-controlled studies. J Clin
Psychopharmacol. 2008;28(1):13–20.
Tohen M, Frank E, Bowden CL, Colom F, Ghaemi SN, Yatham LN, et al. The International Society
for Bipolar Disorders (ISBD) task force report on the nomenclature of course and outcome in
bipolar disorders. Bipolar Disord. 2009;11(5):453–73.
Trevino K, McClintock SM, McDonald Fischer N, Vora A, Husain MM. Defining treatment-
resistant depression: a comprehensive review of the literature. Ann Clin Psychiatry. 2014;26
(3):222–32.
Turner EH, Matthews AM, Linardatos E, Tell RA, Rosenthal R. Selective publication of antide-
pressant trials and its influence on apparent efficacy. N Engl J Med. 2008;358:252–60.
Undurraga J, Baldessarini RJ. Randomized, placebo-controlled trials of antidepressants for acute
major depression: thirty-year meta-analytic review. Neuropsychopharmacology. 2012;37
(4):851–64.
Van Norman GA. Drugs and devices: comparison of European and U.S. approval processes. JACC
Basic Transl Sci. 2016;1(5):399–412.
Vázquez GH, Holtzman JN, Lolich M, Ketter TA, Baldessarini RJ. Recurrence rates in bipolar
disorder: systematic comparison of long-term prospective, naturalistic studies versus random-
ized controlled trials. Eur Neuropsychopharmacol. 2015;25(10):1501–12.
Ventura J, Hellemann GS, Thames AD, Koellner V, Nuechterlein KH. Symptoms as mediators of
the relationship between neurocognition and functional outcome in schizophrenia: a meta-
analysis. Schizophr Res. 2009;113:189–99.
von Glischinski M, Willutzki U, Stangier U, Hiller W, Hoyer J, Leibing E, et al. Liebowitz Social
Anxiety Scale (LSAS): optimal cut points for remission and response in a German sample. Clin
Psychol Psychother. 2018;25(3):465–73.
Walsh TB, Seidman SN, Sysko R, Gould M. Placebo response in studies of major depression.
JAMA. 2002;287(14):1840–7.
Weathers FW, Bovin MJ, Lee DJ, Sloan DM, Schnurr PP, Kaloupek DG, et al. The Clinician-
Administered PTSD Scale for DSM-5 (CAPS-5): development and initial psychometric evalu-
ation in military veterans. Psychol Assess. 2018;30(3):383–95.
Randomized Controlled Trials and the Efficacy of Psychotropic Medications 359
Weise RE, Shear MK, Maser JD. On the need for standardization in panic disorder treatment
research: survey of the literature, 1980–1992. Anxiety. 1996;2:257–64.
Weiss RD, Rao V. The prescription opioid addiction treatment study: what have we learned. Drug
Alcohol Depend. 2017;173(Suppl 1):S48–54.
Weitz E, Hollon SD, Twisk J, van Straten A, David D, DeRubeis RJ, et al. Does baseline depression
severity moderate outcomes between CBT and pharmacotherapy? An individual participant data
meta-analysis. JAMA Psychiat. 2015;72:1102–9.
Wesson DR, Ling W. The Clinical Opiate Withdrawal Scale (COWS). J Psychoactive Drugs.
2003;35(2):253–9.
West R, Hajek P, Stead L, Stapleton J. Outcome criteria in smoking cessation trials: proposal for a
common standard. Addiction. 2005;100(3):299–303.
Westmoreland P, Krantz MJ, Mehler PS. Medical complications of anorexia nervosa and bulimia.
Am J Med. 2016;129(1):30–7.
Williams T, Hattingh CJ, Kariuki CM, Tromp SA, van Balkom AJ, Ipser JC, et al. Pharmacotherapy
for social anxiety disorder (SAnD). Cochrane Database Syst Rev. 2017;10:CD001206.
Winhusen T, Winstanley EL, Somoza E, Brigham G. The potential impact of recruitment method on
sample characteristics and treatment outcomes in a psychosocial trial for women with co-
occurring substance use disorder and PTSD. Drug Alcohol Depend. 2011;120(1–3):225–8.
Wisniewski SR, Rush AJ, Nierenberg AA, Gaynes BN, Warden D, Luther JF, et al. Can phase III
trial results of antidepressant medications be generalized to clinical practice? A STARD report.
Am J Psychiatry. 2009;166(5):599–607.
Wittchen HU, Zhao S, Kessler RC, Eaton WW. DSM-III-R generalized anxiety disorder in the
National Comorbidity Survey. Arch Gen Psychiatry. 1994;51(5):355–64.
Wong N, Sarver DE, Beidel DC. Quality of life impairments among adults with social phobia: the
impact of subtype. J Anxiety Disord. 2011;26(1):50–7.
Woods SW, Gueorguieva RV, Baker CB, Makuch RW. Control group bias in randomized atypical
antipsychotic medication trials for schizophrenia. Arch Gen Psychiatry. 2005;62(9):961–70.
World Health Organization. Global status report on alcohol and health 2018. Geneva: World Health
Organization; 2018.
World Medical Association. Declaration of Helsinki: ethical principles for medical research involv-
ing human subjects. JAMA. 2013;310(20):2191–4.
Yatham LN, Kennedy SH, Parikh SV, Schaffer A, Bond DJ, Frey BN, et al. Canadian Network for
Mood and Anxiety Treatments (CANMAT) and International Society for Bipolar Disorders
(ISBD) 2018 guidelines for the management of patients with bipolar disorder. Bipolar Disord.
2018;20(2):97–170.
Yehuda R, Hoge CW, McFarlane AC, Vermetten E, Lanius RA, Nievergelt CM, et al. Post-
traumatic stress disorder. Nat Rev Dis Primers. 2015;1:15057.
Yildiz A, Vieta E, Leucht S, Baldessarini RJ. Efficacy of antimanic treatments: meta-analysis of
randomized, controlled trials. Neuropsychopharmacology. 2011;36(2):375–89.
Young RC, Biggs JT, Ziegler VE, Meyer DA. A rating scale for mania: reliability, validity and
sensitivity. Br J Psychiatry. 1978;133:429–35.
Zimmerman M, Posternak MA, Chelminski I. Derivation of a definition of remission on the
Montgomery-Asberg depression rating scale corresponding to the definition of remission on
the Hamilton rating scale for depression. J Psychiatr Res. 2004;38(6):577–82.
Zimmerman M, Martinez J, Attiullah N, Friedman M, Toba C, Boerescu DA, et al. Further evidence
that the cutoff to define remission on the 17-item Hamilton Depression Rating Scale should be
lowered. Depress Anxiety. 2012a;29(2):159–65.
Zimmerman M, Martinez JA, Attiullah N, Friedman M, Toba C, Boerescu DA, et al. Why do some
depressed outpatients who are in remission according to the Hamilton Depression Rating Scale
not consider themselves to be in remission? J Clin Psychiatry. 2012b;73(6):790–5.
Adverse Drug Reactions, Intoxications
and Interactions of Neuropsychotropic
Medications
Contents
Abbreviations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 362
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 364
Adverse Drug Reactions and Intoxications: Definition, Prevalence, Severity . . . . . . . . . . . . . . . . . 367
Information on ADRs on the Internet . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 396
ADRs Due to the Pharmacologic Target Profile of the Prescribed Medication . . . . . . . . . . . . 397
ADRs Due to Enantioselectivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 400
Enantiomer Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 402
ADRs Due to Chemical Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 402
ADRs Due to Treatment Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 403
Clinically Most Impressive Toxidromes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 403
ADRs Due to Pharmacologic Target(s) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 403
Clozapine ADRs and Clozapine Toxicity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 405
Medications Not Covered in the Present Chapter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 407
Further Reading . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 408
Pharmacovigilance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 409
Optimization of Pharmacotherapy in Clinical Practice: An Example . . . . . . . . . . . . . . . . . . . . . . . . . . 409
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 410
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 410
G. Zernig (*)
Department of Psychiatry 1, Medical University of Innsbruck, Innsbruck, Tirol, Austria
e-mail: gerald.zernig@i-med.ac.at
S. Bischinger
Hospital Pharmacy, University Hospital of Innsbruck, Innsbruck, Austria
e-mail: sabine.bischinger@tirol-kliniken.at
C. Hiemke
Department of Psychiatry and Psychotherapy, University Medical Center Mainz, Mainz, Germany
e-mail: hiemke@uni-mainz.de
Abstract
Any medication has desirable effects (e.g., alleviation of depressive symptoms)
and may produce undesirable, unintended effects (e.g., headache or vomiting)
which are called ‘adverse reactions’ or ‘adverse drug reactions’ (ADRs). ADRs
can be ‘severe’ / ‘serious’ and have to be differentiated from ‘side effects’ and
‘adverse events.’ Definitions for these terms are given in this chapter. Many
ADRs can be explained by the pharmacologic target(s) of the prescribed medi-
cation. This chapter gives a comprehensive review of these molecular target-
based ADRs. To emphasize, some ADRs cannot be explained yet in sufficient
detail by the pharmacologic target profile of the prescribed drug. Fortunately, we
live in a great era with respect to wealth of and ease of access to information
(i.e., the Internet). As no single individual can hold all the ADR information
currently available and as it is not advisable to rely on a single database only, this
chapter also contains a comprehensive description of available databases
on ADRs, pharmacogenetic profiles of drug metabolism, drug interactions –
including pharmacogenetic ones – and toxicologic information on medications.
Abbreviations
To ease access of the text through the search function of various software, no
hyphens are given in the text. To increase simplicity, no subscripts are given either.
Therefore, the compound “MK-801” is given simply as “MK801,” or the “5-HT1A
receptor” is given simply as “5HT1A receptor.”
Introduction
Different terms and definitions (see below) exist for a phenomenon that we will in
the following refer to as an ‘adverse drug reaction’ (ADR). ADRs are of keen interest
for the prescribing physician, not only the ADRs caused by the medication itself, but
also ADRs to drug interaction and ADRs due to the pharmcogenetic disposition of
the patient.
We live in a great era with respect to the wealth of information and the ease of
access to it. This also means that no single individual (whom we are aware of) can
hold all the information available on ADRs. To emphasize, some ADRs cannot be
explained yet in sufficient detail by the pharmacologic target profile (receptor
binding profile) of the prescribed drug. Consequently, a major focus of the present
review is a short description of Internet databases available to the single prescribing
physician or patient (Table 1). For an overview of the clinically most impressive
toxidromes, see Table 2 below. A comprehensive overview of ADRs that can be
explained on the basis of the pharmacologic target(s) of the medication is given in
Table 3.
As stated above, the wealth of information is overwhelming. In order to help the
individual user orient her/himself, some databases have tried to consolidate ADRs in
a limited number of categories. For example, diagnosia (enterprise.diagnosia.com)
Adverse Drug Reactions, Intoxications and Interactions of. . . 365
Table 1 Internet databases on adverse drug reactions (ADRs), drug drug interactions, and
pharmacogenetic profiles
SPC or website
(in alphabetical order)
yearly costs for single user Advantages Disadvantages
SPC (many SPCs are Legally binding document Inconsistencies between
available free of charge on different SPCs (i.e., between
the Internet) originator brand and generics
or between different generics)
Difficult to find relevant
information for a distinct
combination
Lack of explicit information on
the escitalopram + quetiapine
combination
Apotheken-Umschau German language only
(https://www.apotheken-
umschau.de/Medikamente/
Wechselwirkungscheck)
Austria Codex (shop. Collection of SPCs only
apoverlag.at/austria-codex, No drug interaction software
160 EUR)
Diagnosia Beneficial interaction between
escitalopram and quetiapine
(i.e., augmentation of
antidepressant effect) not
mentioned
drugs.com (free of charge) Information for both patients References NOT unequivocally
and health-care professionals linked to claim of ADR
given in tailored form Beneficial interaction between
References given escitalopram or SSRIs in
When searching for general and quetiapine
“amlodipine candesartan (i.e., augmentation of
ezetimibe,” the website alerts antidepressant effect) not
to a drug interaction (liver mentioned
damage, myopathy) given for a Drug interaction given for
combination of simvastatin and pharmacologic class (e.g.,
(1) amlodipine (CYP3A4 benzodiazepines and opioids)
inhibition) or (2) ezetimibe but NOT for specific
(cause unknown) compounds
Lexicomp ¼ online. No interaction between
lexi.com (free trial upon escitalopram and quetiapine
request, 669 USD) given
Interaction between
hydrocodone and oxazepam
given IF oxazepam BUT NOT
IF hydrocodone is search item
References NOT unequivocally
linked to claim of ADR or drug
interaction
(continued)
366 G. Zernig et al.
Table 1 (continued)
SPC or website
(in alphabetical order)
yearly costs for single user Advantages Disadvantages
mediq.ch (250 CHF/EUR, References (with PubMed link; German language only
free trial access, 15 days) no references for interaction References NOT unequivocally
between hydrocodon and linked to claim of ADR
oxazepam) Link to complete SPC
Partial SPC directly shown unwieldy
Pharmacogenetics table Beneficial interaction between
Pharmacodynamic table (but escitalopram and quetiapine
not for escitalopram) (i.e., augmentation of
Pharmacodynamic drug antidepressant effect) not
interactions mentioned
Orienting therapeutic range and
laboratory alert level
(conflicting values for
aripiprazole, Kirschbaum
2008)
Recommended dosage
Pharmacologic targets
New interactions alert, monthly
update
micromedexsolutions.com Recommended dosage Beneficial interaction between
(8900 EUR, access to free Drug–pregnancy interaction escitalopram and quetiapine
trial unwieldy) Drug–drug of abuse test (i.e., augmentation of
interaction antidepressant effect) not
Phytopharmaceuticals mentioned
Pediatric patients
Elderly patients
Toxicology
Physicochemical
incompatibility
Global coverage of availability/
trade name
psiac.de (120 EUR, free References (with PubMed link) German language only
trial access, 7 + 7 days) references NOT unequivocally Focused on psychiatric and
linked to claim of ADR neurologic medications
Pharmacogenetics table References NOT unequivocally
including drug interactions linked to claim of ADR
exceeding search profile Interaction between
Pharmacodynamic drug hydrocodone and oxazepam
interactions not mentioned
Main emphasis on drug
interactions
ADRs grouped in only three
domains: delirium/
anticholinergic, QTc
prolongation/TdP, 5HT effects
Links to only some SPCs
(continued)
Adverse Drug Reactions, Intoxications and Interactions of. . . 367
Table 1 (continued)
SPC or website
(in alphabetical order)
yearly costs for single user Advantages Disadvantages
Wikipedia (free of charge) ADRs, drug interactions and Depth of information varies
pharmacogenetics often given considerably between drugs
ADRs, drug interactions, and
pharmacogenetics listed but
mostly not detailed
References given sparingly
gives seven ADR categories: (1) anticholinergic, (2) sodium level changes, (3) QTc
prolongation, (4) orthostasis, (5) seizures, (6) serotoneric, and (7) bleeding/potas-
sium level changes, obstipation, renal toxicity, and sedation; whereas psiac (www.
psiac.de) gives three ADR categories: (1) delirogenic/anticholinergic, (2) torsade de
pointes (TdP)/QTC prolongation, and (3) serotonergic.
Ideally, wide accessibility to ADR information can greatly improve a patient’s
knowledge about the prescribed medication and, hence, hopefully increase her/his
treatment adherence. It may, however, also lead to less desirable phenomena that
warrant a prescribing physician’s close attention to ADRs, for example, legal
liability issues. To illustrate, one of the authors (GZ) served as an expert court
witness in a claim of damages by a patient who was mistakenly handed escitalopram
by the pharmacist instead of the prescribed esomeprazole. The patient sued the
pharmacy and claimed to have most of the ADRs known in the summary of product
characteristics of escitalopram. Prolonged and costly litigation ensued.
To access to the information of this chapter easily, we strongly encourage the
reader to make ample use of the search function of her/his reading software. We have
therefore taken pains to give as many synonyms for the most important terms and
concepts as feasible.
Table 2 Differential diagnosis of the toxidromes presenting with agitation, delirium, or coma. This
table is a modification of the excellent Table 3 of Scotton et al. (2019)
Other clinical
Toxidrome Caused by Timecourse Vital signs features
Malignant Inhalational Very sudden Severe Rigor mortis –
hyperthermia anesthetics or onset (minutes to hyperthermia like rigidity,
succinylcholine hours), resolves up to 46 C, hyporeflexia,
(a depolarizing within 24–48 h tachycardia, hypoactive bowel
muscle relaxant) with treatment hypertension, sounds, rising
in genetically tachypnea end-tidal CO2,
susceptible mottled skin with
individuals flushing,
cyanosis, normal
pupils;
rhabdomyolysis,
high blood
potassium
Serotonin Serotonergic Sudden onset Hyperthermia Neuromuscular
syndrome drugs, i.e., 5HT <24 h, most >41.1 C, hyperactivity:
receptor resolve within tachycardia, tremor, clonus,
agonists or 5HT 24 h with hypertension, myoclonus,
transporter treatment (25% tachypnea hyperreflexia;
inhibitors develop diaphoresis;
symptoms hyperactive
>24 h) bowel sounds
Anticholinergic Acetylcholine Sudden onset Mil Normal muscle
syndrome receptor <24 h, resolves hyperthermia tone and reflexes,
blockers (e.g., within hours to usually mumbling,
M1 blockers) days with <38.8 C, slurred speech,
treatment tachycardia, picking behavior,
mild dry flushed skin
hypertension, and mucous
tachypnea membranes,
hypoactive bowel
sounds, urinary
retention,
hyperkinesis
(myoclonus,
choreoathetosis),
seizures (rare)
Neuroleptic Dopamine D2 Slower onset Hyperthermia Neuromuscular
malignant receptor (days to weeks), >41.1 C, hypoactivity
syndrome blockers or resolves within tachycardia, (lead pipe
“dopamine 10 days with hypertension, rigidity),
withdrawal” treatment tachypnea bradykinesia,
(Scotton et al. hypoactive bowel
2019), i.e., sounds
abrupt cessation
of levodopa or
DA agonists
(Brunton et al.
2017)
Table 3 Desired effects (DXEs) and adverse drug reactions (ADRs) explained by pharmacologic target. Whenever a consensus for DXEs and/or ADRs
was found, all respective references are given at the same place in the text. When a source gave additional DXEs and/or ADRs, this is stated separately in the
text. If DXEs and/or ADRs were obtained from a Summary of Product Characteristics, preferably from the originator, this is declared as SPC information (eg
Lyrica ® SPC). In some cases, different sources vary consideriably in their description of a certain toxidrome. These different descriptions are given in order to
empower the reader with greater flexibiliy.
Adverse Drug Reactions, Intoxications and Interactions of. . .
369
370 G. Zernig et al.
Adverse Drug Reactions, Intoxications and Interactions of. . . 371
372 G. Zernig et al.
Adverse Drug Reactions, Intoxications and Interactions of. . . 373
374 G. Zernig et al.
Adverse Drug Reactions, Intoxications and Interactions of. . . 375
376 G. Zernig et al.
Adverse Drug Reactions, Intoxications and Interactions of. . . 377
378 G. Zernig et al.
Adverse Drug Reactions, Intoxications and Interactions of. . . 379
380 G. Zernig et al.
Adverse Drug Reactions, Intoxications and Interactions of. . . 381
382 G. Zernig et al.
Adverse Drug Reactions, Intoxications and Interactions of. . . 383
384 G. Zernig et al.
Adverse Drug Reactions, Intoxications and Interactions of. . . 385
386 G. Zernig et al.
Adverse Drug Reactions, Intoxications and Interactions of. . . 387
388 G. Zernig et al.
Adverse Drug Reactions, Intoxications and Interactions of. . . 389
390 G. Zernig et al.
Adverse Drug Reactions, Intoxications and Interactions of. . . 391
392 G. Zernig et al.
Adverse Drug Reactions, Intoxications and Interactions of. . . 393
394 G. Zernig et al.
Interestingly, the precise reference for “(WHO 1972)” is not given in the WHO
glossary cited above. An ADR is called “unerwuenschte Arzneimittelwirkung”
(UAW) in German and “effet indésirable médicamenteux” (EIM) in French. To
complicate matters, the term “side effect” (“Nebenwirkung,” “effet secondaire”) is
still often used when describing an ADR. In fact, a “side effect” may very well be a
desirable effect under certain circumstances (e.g., the delay of ejaculation by selec-
tive serotonin reuptake inhibitors (SSRIs) when treating premature ejaculation or the
sedative/hypnotic or muscle relaxant effect of “old,” i.e., centrally acting, antihista-
minic drugs like diphenhydramine). A “side effect” can thus be defined as an
unintended effect occurring at a recommended dosage and is most likely related to
the pharmacologic properties of the prescribed medication during on-label use
(“intended use,” “bestimmungsgemaesser Gebrauch,” “avec autorisation de mise
sur le marché, avec AMM, selon l’AMM”).
A medication may also be used “off-label,” that is, outside the approved indica-
tion/s (“Off-label-Gebrauch,” “hors AMM,” “sans AMM”). In the neuropsychiatric
field, off-label use is notorious in the treatment of children or adolescents (Hiemke
et al. 2018), arguably due to the scarcity of (costly) clinical trials in these patient
populations. Often, governing bodies allow off-label use but require the prescribing
physician to scientifically justify her/his off-label use and exercise extra caution and
monitoring.
The WHO (https://www.who.int › coordination › English_Glossary, pdf auto-
matically downloaded 5 Nov 2019 and 13 Dec 2019) differentiates an ADR from
an “adverse event” which occurs during treatment with a drug but is not suspected
to be causally related to the medicine. An adverse event (AE) is, especially in the
context of clinical trials, is sometimes called a “treatment emergent adverse event”
(TEAE).
The WHO further defines a “serious adverse reaction” as an ADR requiring
inpatient hospitalization or prolongation of existing hospitalization or an ADR that is
life-threating or fatal or results in persistent or significant disability or incapacity, or
is a congenital anomaly/birth defect caused by a medication. The WHO also defines
an “unexpected adverse reaction” as an adverse reaction “the nature, severity or
outcome of which is not consistent with the summary of product characteristics.”
ADRs can also be differentiated according to their causality as “certain,” “probable,”
“possible,” “unlikely,” or “conditional/unclassified” (Edwards and Biriell 1994).
All summaries of product characteristics (SPCs or SmPCs, Fachinformationen,
résumés des caractéristiques du produit) list ADRs and group and rank them
according to their prevalence (in Europe, this information can be found in section
4.8 of the SPC). The SPC is a legally binding document that must be available for
every approved medicine and has undergone scrutiny by regulatory governmental
bodies. The SPC should be based on controlled clinical trials during development
Adverse Drug Reactions, Intoxications and Interactions of. . . 395
Very common (sehr haeufig, très fréquemment) in at least 10%, i.e., at least 10 cases
per 100 patients
Common (haeufig, fréquemment): in less than 10%, i.e., 1/10 but at least 1%, i.e.,
1/100
Uncommon (gelegentlich, peu fréquemment): in less than 1%, i.e., 1/100 but at least
0.1%, i.e., 1/1000
Rare (selten, rarement): in less than 0.1%, i.e., 1/1000 but more than 0.01%, i.e.,
1/10,000
Very rare (sehr selten, très rarement): in less than 0.01%, i.e., 1/10,000 patients
[Frequency] not known ([Haeufigkeit] unbekannt, frequence indeterminée): Fre-
quency cannot be estimated from the available data
As stated above, ADRs and the frequency of their occurrence during on-label use
have to be listed in the SPC of the prescribed medication (specifically, in section
4.8 of the European SPC or section 6 of the FDA-approved medication guide in
the USA). SPCs are widely available on the Internet. Although SPCs are, to our
knowledge, meant for the use of health-care professionals only, SPCs for many
medications can be downloaded from the World Wide Web by patients as well. As
a further example of SPC availability for patients on the Internet, the website
www.drugs.com gives information for both the “consumer” and the “health-care
professional.” Patient-tailored information is added to the medication package as
a paper handout in form of a “medication package insert” or “label”
(wikipedia.org) or “medication guide” (FDA term) or “consumer information”
(drugs.com). German or French terms are: “Patienteninformation,”
“Packungsbeilage,” “Beipackzettel,” “notice patient,” and “notice d’emballage.”
In addition, databases that focus on drug interactions and/or pharmacogenetic
disposition (see below) give information on ADRs, sometimes precisely as listed
and frequency-ranked in the SPC. Information on ADRs is therefore easily
available for any approved medicine. In clinical practice, however, SPCs are
rarely used by general practitioners, especially under routine conditions and in
case of polypharmacy (Gahr et al. 2020).
Thus, for the prescribing physician whose pharmacologic knowledge has become
a little rusty (see Table 3 below for a comprehensive refresher), Internet databases
offer a convenient way to get abundant information on ADRs, especially because, as
stated above, some ADRs cannot be explained yet in sufficient detail by the
pharmacologic target profile of the prescribed drug, such as clozapine-induced
agranulocytosis, which is considered an immunologic and/or toxic ADR (see
below). A downside of the plethora of information available on ADRs and drug–
drug interactions is “alert fatigue,” leading clinicians to override the alerts. An
analysis of 213,253 alerts issued between 2009 and 2012 to physicians at a
tertiary-care teaching facility of Harvard Medical School (Nanji et al. 2018) found
that in 73%, the alert was overriden by the prescribing physician. When the over-
riding physician declared that the patient would be monitored as recommended, did
actually monitor the patient, and documented this monitoring, the override was
considered “appropriate.” Such an appropriate alert override, however, occurred in
only 61% of the overrides (Nanji et al. 2018).
Despite incurring the risk of alert fatigue, our analysis of the available Internet
databases (Table 1 below) indicates that it is not advisable to rely on a single
database only. Even within the same databank, drug interactions may be given
when searching for one medication of a pair but not when searching for the other.
Table 1 gives an overview of the databases tested by us from October to December
2019 by searching for information on ADRs, drug interactions, and pharmacogenetic
disposition for the following drug combinations: (1) escitalopram + quetiapine,
(2) hydromorphone + oxazepam, and (3) amlodipine + candesartan + ezetimibe. These
are the reasons for our choice of drug combinations: (1) The escitalopram + quetiapine
Adverse Drug Reactions, Intoxications and Interactions of. . . 397
combination was chosen as it is, on the one side, risky because of an at least additive
prolongation of the corrected QT interval (QTc), a very common ADR of psychophar-
macologic agents, but most likely also conveys the beneficial drug interaction of an
increase in the antidepressant effect of the SSRI escitalopram by quetiapine (see, e.g.,
McIntyre et al. 2007; Quante et al. 2013; Wen et al. 2014). (2) The
hydromorphone + oxazepam combination was chosen because it is very common (at least
at the university hospital in Innsbruck; SB, personal communication) and the (at least)
additive sedation may cause serious problems for the patient and, consequently, the
prescribing physician. (3) The amlodipine + candesartan + ezetimibe combination was
chosen to test if non-psychopharmacologic medications are also covered in the tested
database and because the consensus of the authors was that this combination is essentially
free of adverse drug interactions, allowing us to probe for overreporting of ADRs and
adverse drug interactions by the database. The preference for a certain type of presenta-
tion of information (i.e., the layout) is highly subjective; we have therefore refrained from
giving any information on that aspect.
Most ADRs can be logically explained on the basis of the pharmacologic target
profile of the medication just like the desired effects (DXEs) of this medication.
For the understanding of an ADR, it is therefore very worthwhile to consider the
pharmacologic profile of the medication causing the ADR. By “pharmacologic
target profile” or “receptor (binding) profile,” we mean the sum total of the target
structures (e.g., dopamine receptors vs. dopamine transporters) which have recep-
tor(s) for the medication. Very broadly speaking, the overwhelming majority of
psychopharmacologic mediations bind to neurotransmitter receptors (e.g., dopa-
mine D2 receptors and 5HT2A receptors in the case of most antipsychotics) or
monoamine transporters (e.g., the serotonin [5HT] transporter in the case of
selective serotonin reuptake inhibitors, SSRIs) (Stahl 2013). Only three enzymes
are known as targets for neuropsychiatric medications, that is, acetylcholine
esterase (AChE), monoamine oxidase (MOA), and glycogen synthase kinase
(GSK) (Stahl 2013). Conventional antipsychotic activity is conferred by dopamine
D2 receptor inhibition (blockade; more on that below). Also very broadly speak-
ing, antidepressants seem to elevate the concentrations of monoamine neurotrans-
mitters in the synaptic cleft, mostly by blocking transporters that move the
neurotransmitter back into the neuron.
With respect to the pharmacologic target profile, it also has to be considered with
which respective affinities the medication binds to the receptors of different target
structures. In that context, remember that a transporter, such as the dopamine
transporter (DAT), also has a receptor (binding site) for medications, for example,
the antidepressant bupropion. This receptor on the DAT transporter is not to be
confused with the dopamine receptor, for example, the D2 receptor that mediates the
antipsychotic effect of antipsychotics/neuroleptics.
398 G. Zernig et al.
Enantiomer Nomenclature
Whereas explaining the clinical effect(s) and the ADRs by the pharmacologic
target of the drug of interest currently seems to be the most effective strategy,
explaining ADRs according to chemical structure, for example, butyrophenones
versus benzamides, has, in our opinion, always been unpopular among prescrib-
ing physicians, not least because navigating the chemical names requires
an expertise that is far removed from the clinical routine. More importantly,
chemically very diverse drugs can produce very similar desired effects and
ADRs, for example, haloperidol, a butyrophenone, and risperidone, a
benzisoxazole (and its active metabolite 9-hydroxyrisperidone ¼ paliperidone).
Adverse Drug Reactions, Intoxications and Interactions of. . . 403
On the other hand, chemically similar drugs can produce very different effects,
like amitriptyline, a tricyclic antidepressant, versus flupentixol, a tricyclic anti-
psychotic from the thioxanthene group. For a very helpful overview of different
chemical/pharmacological categorization systems, see the German article by
Laux (2010).
Grouping the drug according to the psychiatric disease category the drug can be
used to treat, for example, antidepressant versus antipsychotic, has retained more
differentiating power, although borders have become blurred (see above). Many
groups in the field use the following categories: antidepressant, antipsychotic
(conventional/first generation vs. atypical/second generation), anti-convulsant
and mood stabilizing, anxiolytic/tranquilizer/used for the treatment of sleep disor-
ders (hypnotic), anti-dementia (nootropic), drugs used for the treatment of
substance-related disorders (including anticraving), drugs used for the treatment
of the attention deficit hyperactivity syndrome (psychostimulant) (see, e.g., Laux
2010; Hiemke et al. 2018).
For the sake of completeness, it should be mentioned that there is an “Anatomical
Therapeutic Chemical” (ATC) classification system adopted by the WHO (https://www.
whocc.no/atc_ddd_index/, accessed 7 Nov 2019), which differentiates “psycholeptics”
(ATC code N05), that is, drugs that produce a calming effect, encompassing antipsy-
chotics (neuroleptics), anxiolytics, and hyponotics/sedatives from “psychoanalep-
tics,”that is, psychostimulants (N06), that is, antidepressants, psychostimulants/agents
used for attention deficit hyperactivity disorder/nootropics; psycholeptics and
psychoanaleptics in combination, and anti-dementia drugs and from “other nervous
system drugs” (N07), with, among others, drugs used in addictive disorders.
Table 3 gives an overview of the desirable and adverse drug reactions (ADRs) of
most psychopharmacologic compounds and lists their pharmacologic targets. To
enhance understanding of physiologic brain functions, the targets are listed alpha-
betically according to the neurotransmitter that activates them (e.g., the NMDA
404 G. Zernig et al.
receptor can be found under “glutamate receptor”). One receptor for the neurotrans-
mitter glutamate, that is, the NMDA receptor, and one receptor for the neurotrans-
mitter gamma-aminobutyric the GABAA receptor, are (at least in our view) so
formidably complex pharmacologic targets that we have supplemented the text of
Table 3 with two figures, Fig. 1 for the GABAA receptor, and Fig. 2 for the NMDA
receptor, below.
With respect to the desired effects and ADRs of dopamine receptor agonists,
Table 3 lists dose of the levodopamine, that is, the precursor of the physiologic
agonist dopamine itself. A refresher: oral dopamine does not cross the blood–brain
barrier but its precursor levodopa does, and as levodopa is readily converted to
dopamine peripherally by aromatic L-amino acid decarboxylase (AADC), oral
levodopa is combined with an AAC inhibitor, typically either benserazid (e.g., in
Madopar ®) or carbidopa (e.g., in Sinemet ®). Please bear in mind that the desired
effects and ADRs of oral levodopa and, hence, dopamine in the brain have been
Fig. 1 Schematic cross sections of the GABAA receptor and the associated chloride channel.
Shown are two cross sections, that is, (1) various binding sites within the transmembrane domain
(i.e., the chloride pore) on the left and (2) the binding sites on the extracellular domain. Reproduced
with the publisher’s permission from Olsen (2018). For more details on the ethanol binding site, see
Wallner et al. (2014). Abbreviations: Barbs, barbiturates; BZ, benzodiazepines; EtOH, ethanol
(alcohol); Eto, etomidate; Pro, propofol, Pyr, pyrazoloquinolines; volatiles, volatile anesthetics.
Examples for benzodiazepines (BZ, BZD) are: alprazolam, bromazepam, brotizolam, chlordiaz-
epoxide, clobazam, clonazepam, diazepam, desmethyldiazepam, flunitrazepam, flurazepam, loraz-
epam, lormetazepam, medazepam, midazolam, nitrazepam, nordazepam, oxazepam, temazepam,
triazolam. Another pharmacologic group that binds to the BZD site is the alpha1-over-alpha3
preferring (Soderhielm et al. 2018) so-called Z drugs (“Z-drugs”) like zolpidem, zopiclone,
eszopiclone, or zaleplon. Examples for barbiturates are: methohexital (methohexital), phenobarbital
(phenobarbitone), primidone (desoxyphenobarbital), and thiopental (thiopentone)
Adverse Drug Reactions, Intoxications and Interactions of. . . 405
Fig. 2 Schematic cross sections of the NMDA receptor and psychotropic drugs modulating its
function. Abbreviations: 5,7-DCKA, 5,7-dichloro-kynurenic acid, a glycine site blocker; AP5, see
D-AP5; Ca2+, calcium ions; D-AP5, dextrorotatory, that is, R-D()-2-amino-5-phosphonovalerate
or phosphonopentanoate; Mg2+, magnesium ions; N, asparagine N-site for channel blockers; Na+,
sodium ions; NMDA, N-methyl-D-aspartate; NR1, NR2A, NR2B, subunits of the NMDA receptor;
Zn2+, zinc ions. (Reproduced with the publisher’s permission from Parsons et al. (2007))
described in patients with Parkinson’s disease, and therefore may not reflect what
dopamine’s physiologic effects. However, as a first approximation they are worth of
our interest.
Clozapine is, in the eyes of many prescribing psychiatrists, the most effective
antipsychotic medication and, for some, still the only “true” atypical (second
generation) antipsychotic, with a 20-fold higher affinity for 5HT2A receptors than
D2 receptors (Schotte et al. 1996, see above). Unfortunately, clozapine’s use is
jeopardized by a considerable risk of severe ADRs, most notably neutropenia.
Accordingly, from 28 Feb 2019 on, the FDA requires prescribing health-care pro-
fessionals as well as pharmacies dispensing clozapine, to be certified in a Clozapine
Risk Evaluation and Mitigation Strategy (REMS) Program (https://www.fda.gov/
drugs/postmarket-drug-safety-information-patients-and-providers/information-
clozapine, accessed 17 Mar 2020). The SPC for Clozaril ® in the USA (version Feb
406 G. Zernig et al.
2017) lists the following ADRs in its boxed warning (“black box warning”): severe
neutropenia, orthostatic hypotension, bradycardia and syncope, seizure, myocarditis
and cardiomyophathy, increased mortality in elderly patients with dementia-related
psychosis. The SPC for Clozaril ® in the UK (UK, version 6 June 2019) states the
following: “As a consequence of a recent European regulatory initiative, the Clozaril
Summary of Product Characteristics (SmPC) has been harmonised across Europe.
The SmPC states that blood monitoring should be carried out in accordance with
national-specific official recommendations. . . .” The SPC for Clozaril ® in the UK
then proceeds to summarize, in a boxed warning (which is identical to the German
version in Austrian SPCs of clozapine products:
Prescribing physicians must comply fully with the required safety mea-
sures. At each consultation, a patient receiving Clozaril must be reminded to
contact the treating physician immediately if any kind of infection begins to
develop. Particular attention must be paid to flu-like complaints such as fever
or sore throat and to other evidence of infection, which may be indicative of
neutropenia (see section 4.4).
Clozaril must be dispensed under strict medical supervision in accordance
with official recommendations (see section 4.4).
Myocarditis
Clozapine is associated with an increased risk of myocarditis which has, in
rare cases, been fatal. The increased risk of myocarditis is greatest in the first
2 months of treatment. Fatal cases of cardiomyopathy have also been reported
rarely (see section 4.4).
Myocarditis or cardiomyopathy should be suspected in patients who expe-
rience persistent tachycardia at rest, especially in the first 2 months of
treatment, and/or palpitations, arrhythmias, chest pain and other signs and
symptoms of heart failure (e.g., unexplained fatigue, dyspnoea, tachypnoea)
or symptoms that mimic myocardial infarction (see section 4.4).
(continued)
Adverse Drug Reactions, Intoxications and Interactions of. . . 407
From the perspective of therapeutic drug monitoring (TDM, see, e.g., Zernig et al.
2009; Hiemke et al. 2011, 2018) it must be noted that the definition of “non-
responsiveness” to antipsychotic medication or “treatment resistance” in the cloza-
pine SPCs is defined only “as a lack of satisfactory clinical improvement despite the
use of adequate doses of at least two different antipsychotic agents, including an
atypical antipsychotic agent, prescribed for adequate duration” does not include a
requirement to quantify the blood concentration of the respective antipsychotic. It is
well known in field that the treatment adherence (compliance) of psychiatric patients
suffering from schizophrenia, depression, or bipolar disorder is notoriously low, with
nonadherence ranging from 10% to 69% (Hiemke et al. 2018), with average
discontinuation rates of 44% for antipsychotics (range, 18–70%), and with patients
taking, on average, only 65% of their prescribed antipsychotic dosage (range,
40–90%) (Zernig et al. 2009). We would argue that before switching a patient to
clozapine, proper TDM of the two previously prescribed antipsychotics be
performed to clarify if these two pre-clozapine antipsychotics have indeed reached
blood concentrations in the orienting therapeutic range for a long enough time span
so that clinical improvement and acceptable tolerability can be expected.
The pathophysiology of clozapine-induced neutropenia (incidences around 3%
for milder cases and 1% fullblown agranulocytosis) has not been clarified in detail
yet and has been attributed to an immune response (Brunton et al. 2017; Regen et al.
2017) and a direct toxic effect on the myeloid cell line (Pick and Nystrom 2014).
In addition to neutropenia/agranulocytosis, the SPC for Clozaril ® in the UK
(version 6 June 2019) lists the following ADRs: eosinophilia, thrombocytopenia,
orthostatic hypotension, myocardial infarction, QT interval prolongation, cerebro-
vascular adverse events, thromboembolism, seizures, anticholinergic effects, fever,
falls, metabolic changes such as hyperglycemia, dyslipidemia, weight gain; rebound,
that is, withdrawal effects, hepatic impairment with preexisting liver disorders, and
increased mortality in elderly people with dementia. The SPC lists drowsiness/
sedation, dizziness, tachycardia, constipation, and hypersalivation as the most com-
mon ADRs of clozapine. The pharmacologic targets for these effects can be found in
Table 3 of the present chapter.
next section “Further Reading” we will suggest references that have successfully
provided such information. For the sake of transparency, we will list the medi-
cations that we did not cover, based on the comprehensive list in our TDM
update (Hiemke et al. 2018) which gives the pharmacokinetic data for the
medications listed below, their orienting therapeutic ranges, and their metabolic
pathways, including the respective isoforms of cytochrome p450 (CYP)
involved.
A number of medications can be used either as a mood stabilizer or an anti-
convulsant. We tried to cover those compounds but, with a few exceptions, not
medications that are considered to be antiepileptics only (Hiemke et al. 2018).
Therefore, we did not cover brivaracetam, ethosuximide, eslicarbazepine, felbamate,
lacosamide, levetiracetam (mechanism of action not fully understood (Brunton et al.
2017)), methsuximide, perampanel, phenytoin, retigabine, rufinamide, sulthiame,
tiagabine, topiramate, vigabatrin, and zonisamide.
We also did not cover the following antiparkinson drugs (Hiemke et al. 2018):
amantadine, bornaprine, cabergoline, carbidopa, entacapone, rotigotine, and
tolcapone. The following anxiolytics and drugs for the treatment of sleep disorders
(Hiemke et al. 2018) were not covered either: modafinil (unknown mechanism of
action and some abuse liability according to the SPCs, and a schedule IV controlled
drug in the USA (Brunton et al. 2017)), opipramol, prothipendyl, and promethazine.
Finally, we did not cover the tricyclic antidepressant tianeptine the mechanism of
action of which is not clear (Stablon ® SPC) or clomethiazole (Distraneurin ®,
Heminevrin ®), a medication not covered in Stahl (2013) and Brunton et al. (2017),
previously popular in some countries for the treatment of alcohol withdrawal but
jeopardized by considerable ADRs including tolerance, withdrawal, and abuse
liability (SPCs).
Further Reading
For further information on the pharmacologic basis of the desired effects and ADRs
of neuropsychotropic medications, the reader is referred to the informative and
popular book by Stahl which is pleasant to read (Stahl 2013). “Goodman and
Gilman’s: The Pharmacological Basis of Therapeutics)” is a pharmacologic classic
which is currently in its 13th edition and also contains a useful table on pharmaco-
kinetic parameters of selected pharmaceuticals (Brunton et al. 2017). Another
source, in German, is Degner et al. (2004). Over the last decades, SPCs have become
more comprehensive with respect to giving clinical proof of efficacy in randomized
controlled trials (RCTs) and with respect to the pharmacodynamics (i.e., the phar-
macologic target(s)) of some medications. In comparison to pharmacologic text-
books and scholarly articles (including the present chapter), SPCs also have the
advantage of being legally binding documents with respect to the information given
therein.
For an excellent, extremely informative, and authoritative overview of pharmaco-
logic targets, receptor classification, function, selective agonists and blockers etc., the
Adverse Drug Reactions, Intoxications and Interactions of. . . 409
reader can consult the joint website of the International Union of Basic and Clinical
Pharmacology (IUPHAR), and the British Pharmacological Society, www.guidetophar
macology.org. Wikipedia (*wikipedia.org) in its different language version is an
informative source, too, and unsurpassed with respect to speed of access and scope.
Finally, a word of caution: Wikipedia’s information sometimes needs to be cross-
checked – first and most quickly, by cross-checking the different language versions
of the same article. The anonymity of its contributors is problematic with respect to
effective source checking. Sometimes, information from Stahl (2013) and Goodman
and Gilman’s (Brunton et al. 2017) is discrepant. To enable the reader to form an
independent judgment, we have taken care to present both views in the present
chapter and have tried to resolve apparent discrepancies. A shortcoming of both
sources is that claims are only rarely supported by citations. For example, Stahl
claims that guanfacine is 15–60 times more selective for the alpha2A receptor than
for alpha2B- and alpha2C receptors (Stahl 2013) without giving a reference for this
claim. The IUPHAR/BPS website www.guidetopharmacology (accessed 21 Mar
2020) gives guanfacine’s pKi values as 7.1–7.3 for alpha2A, 5.8–6.5 for alpha2B,
and 5.4–6.2 for alpha2C. This translates to a 4–32-fold selectivity for alpha2A/2B
and an 8–79-fold difference for alpha2A/2C.
Pharmacovigilance
Cross-References
Conflicts of Interest Christoph Hiemke is one of the editors of the commercially available
database www.psiac.de. Sabine Bischinger and Gerald Zernig declare no conflicts of interest.
References
Anonymous. Wikipedia article 2020. Available from: en.wikipedia.org
Barnes NM, Sharp T. A review of central 5-HT receptors and their function. Neuropharmacology.
1999;38(8):1083–152.
Bear MF, Connors BW, Paradiso MA. Neuroscience. 4th ed. Philadelphia: Wolters Kluwer; 2016.
Beaulieu JM, Espinoza S, Gainetdinov RR. Dopamine receptors – IUPHAR review 13. Br
J Pharmacol. 2015;172(1):1–23.
Bortolotti F, De Paoli G, Gottardo R, Trattene M, Tagliaro F. Determination of gamma-
hydroxybutyric acid in biological fluids by using capillary electrophoresis with indirect detec-
tion. J Chromatogr B Anal Technol Biomed Life Sci. 2004;800(1–2):239–44.
Brunton LL, Hilal-Dandan R, Knollmann BC. Goodman and Gilman’s the pharmacological basis of
therapeutics. 13th ed. New York: McGraw-Hill; 2017.
Busardo FP, Gottardi M, Tini A, Minutillo A, Sirignano A, Marinelli E, et al. Replacing GHB with
GBL in recreational settings: a new trend in chemsex. Curr Drug Metab. 2018;19(13):1080–5.
Bymaster FP, Dreshfield-Ahmad LJ, Threlkeld PG, Shaw JL, Thompson L, Nelson DL, et al.
Comparative affinity of duloxetine and venlafaxine for serotonin and norepinephrine trans-
porters in vitro and in vivo, human serotonin receptor subtypes, and other neuronal receptors.
Neuropsychopharmacology. 2001;25(6):871–80.
Carruthers SG. Adverse effects of alpha 1-adrenergic blocking drugs. Drug Saf. 1994;11(1):12–20.
Chae YJ, Jeon JH, Lee HJ, Kim IB, Choi JS, Sung KW, et al. Escitalopram block of hERG
potassium channels. Naunyn Schmiedeberg’s Arch Pharmacol. 2014;387(1):23–32.
Chevillard L, Megarbane B, Baud FJ, Risede P, Decleves X, Mager D, et al. Mechanisms of
respiratory insufficiency induced by methadone overdose in rats. Addict Biol. 2010;15(1):62–80.
Chiara DC, Hamouda AK, Ziebell MR, Mejia LA, Garcia G 3rd, Cohen JB. [(3)H]chlorpromazine
photolabeling of the torpedo nicotinic acetylcholine receptor identifies two state-dependent
binding sites in the ion channel. Biochemistry. 2009;48(42):10066–77.
Degner D, Grohmann R, Kropp S, Ruther E, Bender S, Engel RR, et al. Severe adverse drug
reactions of antidepressants: results of the German multicenter drug surveillance program
AMSP. Pharmacopsychiatry. 2004;37(Suppl 1):S39–45.
Adverse Drug Reactions, Intoxications and Interactions of. . . 411
Regenthal R, Krueger M, Koeppel C, Preiss R. Drug levels: therapeutic and toxic serum/plasma
concentrations of common drugs. J Clin Monit. 1999;15:529–44.
Repetto MR, Repetto M. Habitual, toxic, and lethal concentrations of 103 drugs of abuse in humans.
J Toxicol Clin Toxicol. 1997;35(1):1–9.
Salous AK, Ren H, Lamb KA, Hu XQ, Lipsky RH, Peoples RW. Differential actions of ethanol and
trichloroethanol at sites in the M3 and M4 domains of the NMDA receptor GluN2A (NR2A)
subunit. Br J Pharmacol. 2009;158(5):1395–404.
Sanchez C. The pharmacology of citalopram enantiomers: the antagonism by R-citalopram on the
effect of S-citalopram. Basic Clin Pharmacol Toxicol. 2006;99(2):91–5.
Savelyeva MV, Baldenkov GN, Kaverina NV. Receptor binding potencies of chlorpromazine,
trifluoperazine, fluphenazine and their 10-N-substituted analogues. Biomed Biochim Acta.
1988;47(12):1085–7.
Schotte A, Janssen PFM, Gommeren W, Luyten WHML, van Gompel P, Lesage AS, et al.
Risperidone compared with new and reference antipsychotic drugs: in vitro and in vivo receptor
binding. Psychopharmacology. 1996;124:57–73.
Schulz M, Schmoldt A. Therapeutic and toxic blood concentrations of more than 800 drugs and
other xenobiotics. Pharmazie. 2003;58(7):447–74.
Schulz M, Iwersen-Bergmann S, Andresen H, Schmoldt A. Therapeutic and toxic blood concen-
trations of nearly 1,000 drugs and other xenobiotics. Crit Care. 2012;16(4):R136.
Scotton WJ, Hill LJ, Williams AC, Barnes NM. Serotonin syndrome: pathophysiology, clinical
features, management, and potential future directions. Int J Tryptophan Res.
2019;12:1178646919873925.
Soderhielm PC, Balle T, Bak-Nyhus S, Zhang M, Hansen KM, Ahring PK, et al. Probing the
molecular basis for affinity/potency- and efficacy-based subtype-selectivity exhibited by
benzodiazepine-site modulators at GABAA receptors. Biochem Pharmacol. 2018;158:339–58.
Stahl SM. Mirror, mirror on the wall, which enantiomer is fairest of them all? J Clin Psychiatry.
2002;63(8):656–7.
Stahl SM. Stahl’s essential psychopharmacology: neuroscientific basis and practical application. 4th
ed. Cambridge: Cambridge University Press; 2013.
Ulrich S, Hiemke C, Laux G, MuellerOerlinghausen B, HavemannReinecke U, Riederer P, et al.
Value and actuality of the prescription information for therapeutic drug monitoring of
psychopharmaceuticals: a comparison with the medico-scientific evidence. Pharmacop-
sychiatry. 2007;40:121–7.
Vollenweider FX, Leenders KL, Oye I, Hell D, Angst J. Differential psychopathology and patterns
of cerebral glucose utilisation produced by (S)- and (R)-ketamine in healthy volunteers using
positron emission tomography (PET). Eur Neuropsychopharmacol. 1997;7(1):25–38.
Wallner M, Hanchar HJ, Olsen RW. Alcohol selectivity of beta3-containing GABAA receptors:
evidence for a unique extracellular alcohol/imidazobenzodiazepine Ro15-4513 binding site at
the alpha+beta- subunit interface in alphabeta3delta GABAA receptors. Neurochem Res.
2014;39(6):1118–26.
Wen XJ, Wang LM, Liu ZL, Huang A, Liu YY, Hu JY. Meta-analysis on the efficacy and tolerability
of the augmentation of antidepressants with atypical antipsychotics in patients with major
depressive disorder. Braz J Med Biol Res. 2014;47(7):605–16.
Witchel HJ, Pabbathi VK, Hofmann G, Paul AA, Hancox JC. Inhibitory actions of the selective
serotonin re-uptake inhibitor citalopram on HERG and ventricular L-type calcium currents.
FEBS Lett. 2002;512(1–3):59–66.
Wong DT, Bymaster FP, Mayle DA, Reid LR, Krushinski JH, Robertson DW. LY248686, a new
inhibitor of serotonin and norepinephrine uptake. Neuropsychopharmacology. 1993;8(1):23–33.
Zernig G, Issaevitch T, Woods JH. Calculation of agonist efficacy, apparent affinity and receptor
population changes after administration of insurmountable antagonists: comparison of different
analytical approaches. J Pharmacol Toxicol Methods. 1996;35:223–37.
Zernig G, Giacomuzzi S, Riemer Y, Wakonigg G, Sturm K, Saria A. Intravenous drug injection
habits: drug users’ self-reports vs researchers’ perception. Pharmacology. 2003;976:49–56.
414 G. Zernig et al.
Zernig G, Ahmed SH, Cardinal RN, Morgan D, Acquas E, Foltin RW, et al. Explaining the
escalation of drug use in substance dependence: models and appropriate animal laboratory
tests. Pharmacology. 2007;80(2–3):65–119.
Zernig G, Hiemke C, Havemann-Reinecke U, Laux G, Riederer P, Rabl W, et al. Empfehlungen fuer
die gutachterliche Bewertung von Medikamentenspiegeln in der Psychiatrie im gerichtsan-
haengigen Schadensfall. Psychopharmakotherapie. 2009;16(2):57–65.
Zvosec DL, Smith SW, McCutcheon JR, Spillane J, Hall BJ, Peacock EA. Adverse events,
including death, associated with the use of 1,4-butanediol. N Engl J Med. 2001;344(2):87–94.
TCM Substances in
Neuropsychopharmacotherapy: Basic
Aspects with a Focus on Depression
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 417
Diet Decreases Incidence of Depression and Improves Symptoms: Meta-analyses and
Clinical Intervention with Healthy Diet . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 419
Antidepressant Function of Herbal Medication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 419
Antidepressant Phytochemicals: Structure, Activity, and Molecular Mechanism . . . . . . . . . . . . . 424
Structure and Basic Properties of Phytochemicals for Antidepressant Activity . . . . . . . . . . . . 424
Diet and Phytochemicals Enhance 5-HT Level to Exert Antidepressant Activity . . . . . . . . . 432
Phytochemicals Inhibit Type A Monoamine Oxidase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 434
Phytochemicals Bind to NTF Receptors, Activate Signal Transduction, Induce NTF
Expression, and Function as NTF-Mimics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 435
Anti-Inflammatory Functions of Phytochemicals in Depression . . . . . . . . . . . . . . . . . . . . . . . . . . . 438
Phytochemicals Modulate the HPA Axis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 439
Phytochemicals Show Anxiolytic Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 439
Novel Antidepressants Synthesized Based on Phytochemical Scaffold . . . . . . . . . . . . . . . . . . . . . . . 440
Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 441
Financial and Competitive Interests Disclosure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 442
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 442
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 443
Abstract
Depression is one of the most common complexes of heterogeneous psychiatric
disorders. Etiopathogenic factors of depression are deregulation of serotonin,
noradrenaline and dopamine systems, increased activity of type A monoamine
oxidase, deficit of brain-derived neurotrophic factor, and impaired neurogenesis
in the hippocampus and neuroinflammation. Current antidepressant medicines,
such as tricyclic antidepressants, selective serotonin, or noradrenaline reuptake
inhibitors and monoamine oxidase inhibitors, are effective, but about one-third of
the patients are resistant to the therapy. Traditional Chinese medicine, herb, and
plant-derived phytochemicals are proposed as alterative therapeutic agents in
depression. This review presents molecular mechanisms underlying antidepres-
sant effects of diet habits, traditional herbs, St John’s wort, saffron, passion
flower, lavender, valerian, kava, and Ginkgo biloba, and their phytochemical
constituents. Phytochemicals have been proved to modify multiple pathogenic
factors of depression and ameliorate the symptoms in preclinical models. The
epidemiological and clinical intervention studies have presented some beneficial
effects of herb and phytochemicals. However, the intensive rapid metabolism and
poor bioavailability in the brain prevent the clinical application of phytochemi-
cals. Synthesis of more effective and stable compounds based on phytochemical
scaffold and establishment of effective delivery systems are discussed in order to
develop novel therapeutic strategy for depressive disorders by use of
phytochemicals.
Abbreviations
7,8-DHF 7,8-Dihydroxyflavone
BD Bipolar depression
CUMS Chronic un-predictable mild stress
EO Essential oil
ER Estrogen receptor
FST Forced swimming test
GR Glucocorticoid receptor
HDRS Hamilton Depression Rating Scale
HPA Hypothalamic-pituitary-adrenal
IDO Indoleamine 2,3-dioxygenase
MADRS Montgomery–Asberg Depression Rate Scale
MAO-A and –B Type A and B monoamine oxidase
MDD Major depressive disorder
NLRP3 Nucleotide-binding domain, leucine-rich repeat, pyrin
domain containing protein 3
RCT Randomized controlled trial
RPCT Randomized, placebo control trial
TCM Traditional Chinese medicine
THM Traditional herb medicine
TCM Substances in Neuropsychopharmacotherapy: Basic Aspects with a Focus. . . 417
Introduction
Depression is the most prevalent mental illness and several forms are distinguished:
major depressive disorder (MDD), bipolar depression (BD), seasonal affective
disorder, psychotic, postpartum, premenstrual dysphoric, and situational depression.
MDD itself is a heterogeneous disease characterized by depressive mood, inability to
experience pleasure, cognitive impairment, and autonomic disturbance. The etiology
of depression remains elusive because of lack of identification of disease-responsible
genes and large inter-individual variability. Multiple genetic, social, psychological,
and neurochemical factors are involved in the etiopathogenesis of depression.
Population-based study estimated the hereditability of depression at 42% in
women and 29% in men (Kendler et al. 2006). The impact of negative environmental
factors, such as life stress, traumatic experiences, social and psychological stress,
and neuroinflammation are implicated in depression, particularly in late life, which,
however, might be effected by genetically regulated vulnerability of neuronal circuit
to stress. These results suggest the interaction between genetic and environmental
factors might be the therapeutic target of depression, and epidemiological studies
present that lifestyle factors, such as dietary pattern, sleep, and exercise affect
development of depression (Lopresti et al. 2013). The Mediterranean dietary pattern
reduced risk of developing depression by over 30% (Sanchez-Villegas et al. 2013).
These results indicate that multidisciplinary approach with dietary factors might be
an alternative therapy strategy for depression.
Finding of antidepressant activity of iproniazid (N0 -isopropyl-iso-
nicotinohydazide) and its inhibition of type A monoamine oxidase (MAO-A), a
major enzyme catabolizing serotonin (5-hydroxytryptamine, 5-HT), noradrenaline
(NA) and dopamine (DA), and deregulated serotonergic signal pathways in the
postmortem brain led to the “monoamine hypothesis” and “MAO hypothesis” for
depression pathogenesis (Birkmayer and Riederer 1975). Tricyclic antidepressants
(TCAs), selective serotonin and noradrenaline reuptake inhibitors (SSRIs, SNRIs),
and reversible MAO-A inhibitors (moclobemide, befloxatone) have been developed
based on these hypotheses. Neuropathological analyses and structural MRI studies
have shown the cortical alterations, decrease of the total volume and gray matter of
the hippocampus and functional and structural disruption of neuronal circuit related
in emotional processing in MDD patients (He et al. 2017). Brain-derived
neurotrophic factor (BDNF) decreases in the hippocampus and serum of patients
with MDD, which impairs neurogenesisin the hippocampus and causes atrophy with
neuronal loss of the paraventricular nucleus. The atrophy progresses depending on
disease duration, reduced 5-HT level, increased stress, and neuroinflammation.
418 M. Naoi et al.
Effects of dietary patterns and nutrients on the incidence of depression have been
presented in clinical studies (Lassale et al. 2019). The degree of adherence to the
healthy diet, in particular Mediterranean diet, was reversely correlated with the rate
of incidence. Inverse dose–response relationship was found for intake of fruit,
vegetables, legumes, nuts, ω-3 unsaturated fatty acid–rich foods, fish and whole
grains, and for avoiding processed foods containing high amounts of refined carbo-
hydrate or sugar. Cohort analyses of diet quality scores [Mediterranean Diet Score,
Pro-vegetarian Dietary Pattern and Alternative Healthy Eating Index-2010] pre-
sented the effects on incidence of depression in the Seguimiento Universidat de
Navarra (SUN) project (Sanchez-Villegas et al. 2015). The SUN cohort study
subjected healthy 15,093 participants for 8.5 years, and depression risk decreased
significantly in subjects with modulate versus lower adherence. Recently an “Anti-
depressant Food Score” was proposed as a scale of antidepressant potency in food,
based on the content of antidepressant nutrients (“Antidepressant Nutrient Density”)
(LaChance and Ramsey 2018). Twelve beneficial nutraceuticals were reported for
the prevention and treatment of depression: vitamins (folic acid, A, B6, B12, C, and
thiamine), minerals (magnesium, potassium, selenium, iron, and zinc), and ω-3 fatty
acids. The highest scored plant foods were leafy greens, lettuces, peppers, and
cruciferous vegetables.
However, a phase II/III, 8-week, double-blind trial of nutraceutical combination
[S-adenosylmethionine, folic acid, ω-3 fatty acids, 5-hydroxytryptophan (5-HTP),
and zinc] did not improve symptoms assessed with Montgomery–Asberg Depres-
sion Rate Scale (MADRS) score in MDD patients (Sarris et al. 2019). A randomized
controlled trial (RCT) of dietary improvement intervention was reported for patients
with MDD (Jacka et al. 2017). The SMILES (Supporting the Modification of
lifestyle In Lowered Emotional State) study was a 12-week, parallel-group, single-
blind trial of an adjunctive dietary intervention in the treatment of moderate-to-
severe depression. The dietary support group demonstrated significantly greater
improvement of MADRS score. About a half of intervention studies of adjunctive
diet reported the positive effects in treatment of moderate-to-severe depression, but
statistical difference was not significant in most studies. The results of clinical
studies are still inconsistent, because of the absence of adequate size and design of
the trials and rational quality control of diet.
Depression
(kava), and Parax ginseng (ginseng) confer therapeutic benefits in the treatment of
depressive and anxiety disorders. A meta-analysis of clinical trials of herbs has
presented the evidence-based efficacy of herbs: St John’s wort and saffron for
unipolar depression, and kava for generalized anxiety disorder (Sarris et al. 2011).
St John’s wort is used to treat emotional distress in TCM for more than 2000 years
in China and since the late fifteenth century in European countries, and the only
herbal alternative to synthetic antidepressants in therapy of mild-to-moderate depres-
sion. It is available as tablet, tea, oil, and tincture. A large meta-analysis showed that
the efficacy and safety of St John’s wort were comparable to SSRIs in mild and
moderate depressive disorders (Ng et al. 2017). Hypericum perforatum extract, such
as WS ® 5570, LI160, PM23, and ZE117, are applied in treatment of depression, and
WS ® 5570 was effective as paroxetine in mild-to-moderate MDD. The major
ingredients are napthhodianthrones (hypericin, psudohypericin), phloroglucinol
derivatives (hyperforin, adehyperforin), flavonoids, proanthrocyanidins, and
chlorogenic acid. Hyperforin, rutin, and hypericinmainly contribute to antidepres-
sant efficacy. The extract inhibits monoamine reuptake and enhances the synaptic
availability in the hippocampus and hypothalamus, upregulates 5-HT1A and 5-HT2A
receptors, and also has affinity to DA, β-adrenergic, GABA, and δ-opioid receptors.
Hyperforin activates nonselective cation transient receptor potential channel (TRPC)
6, inhibits reuptake of 5-HT and NA, modulates dendritic spine morphology in
pyramidal neurons of hippocampus, and shows BDNF-mimic activity (Zirak et al.
2019). Hyperforin also inhibits tumor cell proliferation, angiogenesis and inflam-
mation, and disaggregates amyloid deposits in Alzheimer’s disease.
TCM Substances in Neuropsychopharmacotherapy: Basic Aspects with a Focus. . . 421
Saffron (Crocus sativus L.) has been used as a medical antidepressant herb in
Arabic and Islamic traditional medicine (Shafiee et al. 2018). Meta-analysis of nine
RCTs presented the effectiveness in treatment of mild-to-moderate depression also in
mothers with postpartum depression (Toth et al. 2019). It improved clinical symp-
toms assessed with Hamilton Depression Rating Scale (HDRS) effectively as fluox-
etine in short-term therapy. A randomized, placebo control trial (RPCT) of
standardized extract from saffron affron® showed improvement of depressive symp-
toms and anxiety in youth assessed in the Revised Child Anxiety and Depression
Scale (RCADS). Major bioactive constituents of saffron are crocetin (a carotenoid
dicarboxylic acid precursor of crocin), crocin (family of six mono- or di-glycosyl
polyene ester), picrocrocin (monoterpene glycoside precursor of safranal), lipophilic
safranal (2,3-dihydro-2,2,6-trimethlbenzaldehyde), and flavonoids (quercetin and
kaempferol). Crocin and safranal are the major ingredients to contribute the antide-
pressant effects of saffron, and crocin significantly improved depression scores in
patients with MDD as an adjunct to SSRI. Saffron can modulate the HPA axis and
induce BDNF and has anti-inflammatory and antioxidant and antidepressant
effects. Saffron extract and crocin inhibit MAO-A and monoamine reuptake and
bind to N-methyl-D-aspartate (NMDA) receptor as an antagonist and to GABA-α
receptor as an agonist, and crocin to 5-HT2C receptor as an antagonist (Leone et al.
2018). Safranal and crocin prevented increase in plasma corticosterone levels and
depressive-like symptoms in rats exposed to chronic restraint stress. Saffron has
anti-inflammatory, anticancer, antioxidant, and neuroprotective properties and is
traditionally used as anti-convulsant, memory enhancer, analgesic, and sedative in
schizophrenia and anxiety disorder.
Passion flower (Passiflora incarnata L.) is a widely used mild sedative and
anxiolytic agent in most European counties. Beneficial effects have been reported
for treatment of insomnia, generalized anxiety disorder, nervous restlessness, atten-
tion-deficit hyperactivity disorder (ADHD), and menopausal symptoms (Miroddi
et al. 2013). It contains flavonoids (apigenin, chrysin, and kaempferol), C-glycosyl
flavones (vitexin, isovitexin, orientin, and isoprientin), MAO-inhibiting indole alka-
loids (harman, harmin, and harmalin), and also a certain amount of GABA, inhibits
GABA uptake and antagonizes GABAB receptors (Appel et al. 2011). Passion
flower flavonoids require metabolic activation by intestinal microflora into
hydroxyphenylacetic acids for anxiolytic effects in preclinical studies.
Clinical trials of aromatherapy show the promising therapeutic effects in depres-
sion. Essential oils (EOs) extracted from lavender (Lavandula angustifolia), lemon
balm (Melissa officinalis L.), and rosemary (Rosmarinus officinalis L.) and their
constituents are applied for treatment of anxiety, stress, and depression (de Sousa
et al. 2017). Lavender tea and the EO are approved by the European Medicines
Agency as anxiolytic and stress-relieving agents, and used for treatment of stress,
anxiety, postpartum depression, agitated behavior in dementia, premenstrual symp-
toms, neurasthenia, and post-traumatic stress disorder or somatization disorder.
Silexan (WS ® 1265) is an oil product of lavender flower approved as a drug in
Germany for the treatment of restlessness related to anxious mood, and in RPCTs it
was effective in generalized anxiety disorder assessed according to DSM-5 criteria
422 M. Naoi et al.
and with the Hamilton Anxiety Rating Scale score (Kasper et al. 2014). Lavender
EO ameliorated depression-like behavior in rats treated with corticosterone and
increased neurogenesis in the hippocampus and subventricular zone. Lavender EO
contains lynalyl acetate (30–50%), linalool (20–35%), garanyl acetate (5%), and
β-caryophyliene (5%) as the most important compounds. Lavender EO antagonized
NMDA receptors and inhibited 5-HT transporter, which mainly contributed to
antidepressant activity, in addition to inhibition of MAO-A and voltage-gated
calcium channel, calcium influx, and binding to 5-TH1A receptor (Lopez et al.
2017). Anxiolytic effect of lavender EO was mediated by serotonergic transmission,
not GABAergic. Linalool-rich EOs from Aniba rosaeodrora (pau-rosa, rosewood),
Aniba paviflara (macacaporanga), and Aellanthus suaveolens (catinga-de-mulata)
have shown antidepressant effects in rodent model of depression.
Lemon balm (Melissa officinalis L.) is a medical plant used in European tradi-
tional medicine and contains flavonoids [quercetin, rhamnocitrin (3,40 ,5-trihydroxy-
7-methoxyflavone), luteolin], phenolic acids (rosmarinic, caffeic, and protocatechuic
acid), triterpenes (ursolic and oleanolic acid), sesquiterpenes, and EOs. In clinical
trials it improved mood, cognition, and memory. In a RCT the extract improved the
insomnia severity score, the Beck Depression Inventory, and Beck Anxiety Inven-
tory, in insomnia, anxiety, and depression (Ranjbar et al. 2018). Cyracos ® Melissa
officinalis L. leaf extract reduced sleep disturbance and anxiety manifestations in
stressed volunteers with mild-to-moderate anxiety disorders. Inhibition of DA and
GABA transmission and activity of MAO-A and acetylcholinesterase, and increase
in serotonergic activity via 5-HT1A receptors contributed antidepressant and anxio-
lytic activity (Shakeri et al. 2016). The EO has anti-agitation effects via GABAergic
transmission.
Rosemary (Rosmarinus officinalis L.) tea, infusions, alcohol extract, and the EO
are used for treatment of depression, nervous agitation, physical and mental
fatigue, and inflammatory diseases. A RCT of oral rosemary (500 mg twice daily
for 1 month) in university students reduced scores of anxiety and depression
(assessed by Hospital Anxiety and Depression Scale) and memory performance
(Prospective and Retrospective Memory Quetionaire) (Nematolahi et al. 2018).
The active components include terpenoids (carnosol, rosmanol, carnosic,
oleanolic, and ursolic acid), flavonoids [diosmin (a glycosyloxylflavone), luteolin,
apigenin, quercetin], and phenolic acids (caffeic, rosmarinic acids). The rosemary
polyphenols (luteolin, carnosic acid, and rosmarinic acid) enhanced 5-HT, DA,
NA, and acetylcholine and expression of genes involving in monoamine synthesis,
modulated GABAA receptors and showed antidepressant activity in a mouse
depression model (Sasaki et al. 2013). The EO, carnosol, and betulinic acid [3β-
hydoxyl-lup-20(29)-en-28-oic acid] had also antidepressant effects in a rodent
model of depression.
Valerian (Valeriana officinalis L.) extract reduces stress and anxiety, improves
sleep, relieves premenstrual syndrome, and has anti-convulsant effect and sedation.
In generalized anxiety disorder valerian extract reduced the total the Hamilton
Anxiety Rating Scale score and the psychiatric factors in a RCT (Andreatini et al.
2002). In patients with psychophysiological insomnia, valerian demonstrated
TCM Substances in Neuropsychopharmacotherapy: Basic Aspects with a Focus. . . 423
positive effects on sleep structure and perception, and valerian extract LI 156 was as
efficacious as oxazepan in treatment of nonorganic insomnia. Extract of valerian root
showed some anti-obsessive and anti-compulsive effects in patients with obsessive–
compulsive disorder. The constituents include alkaloids, flavonoids, and a
sequiterpene valerenic acid. Valeriana wallichii extract increased NA and DA levels
in the forebrain and had antidepressant effects in a mouse model of depression.
Valerenic acid interacted with β2/3 subunits of GABAA receptor as a potent allosteric
modulator and exhibited anxiolytic effects (Khom et al. 2010).
Kava (Piper methysticum G. Forster) has the largest evidences base from in vitro,
in vivo, and clinical studies for use in anxiety disorders, stress, fear, and menstrual
disorder. The major constituents are lipophilic kavalactonees, including kavain
[(R)-6,6,-dihydro-4-methoxy-6-styryl-2H-pyran-2-one] and dihydrokavain [4-meth-
oxy-6-(2-phenylethyl)-5,6-dihydro-2H-pyran-2-one], which have the strongest anxi-
olytic activity. The anxiolytic effects are due to positive allosteric modulation of
GABAA receptors, inhibition of excitatory neurotransmitter release and neuronal
reuptake of NA and DA by blockade of voltage-gated sodium and calcium ion
channels (Sarris et al. 2013). In generalized anxiety disorder, kava extract LI 150 was
effective as opipramol and buspirone (Boerner et al. 2003), and WS ® 1490 improved
anxiety, tension, and restlessness states in RCTs. Treatment with kava tablets
containing 50 mg kavalactones for 3 weeks (5 tablets/day) significantly reduced
Beck Anxiety Inventory and MADR scores (Sarris et al. 2009). The phase III
monotherapy with extract of kava cultivar demonstrated positive effects in subjects
with generalized anxiety disorder (Savage et al. 2015).
Ginkgo biloba extract EGb761 is prepared from dried leaves of Ginkgo tree and
approved in Germany for the treatment of dementia syndromes of neurodegenerative
or vascular origin, cardiovascular impairment, and neurosensory disorders via
neuroprotective, antioxidant, and free radical-scavenging potency. The extract con-
tains ginkgo-flavone glycosides and diterpene trilactones (ginkgolide A, B, C, and
bilobalides). Ginkgolides play a major role in neuroprotection and also antidepres-
sant actions. Ginkgo biloba extract was used as adjunct in elderly depressive patients
treated with citalopram and could relieve anxiety and improve daily living and mood
in elderly patients (Dai et al. 2018). EGb 761 had antioxidant function, enhanced
BDNF expression, and exhibited antidepressant-like effect in mice exposed against
forced swimming test (FST). Bilobalide increased expression of glucocorticoid
receptor (GR) in the mouse hippocampus, and modulate the HPA axis and
ginkgolide B and bilobalide were noncompetitive inhibitors of GABA, glycine,
and 5-HT3 receptors, and exerted anxiolytic and antidepressant effects (Wu et al.
2016a). Ginkgolide B activated NT-3/TrkA (neurotrophin- 3/tropomycin-related
kinase A) and Ras/MAPK (mitogen-activated protein kinase) pathways, induced
BDNF expression and protected hippocampal neurons, and exerted antidepressant
effects.
Ginseng Panax ginseng Meyer is one of the most popular THMs and the major
ingredients ginsenodises have been confirmed to exert antidepressant, anxiolytic,
and anti-fatigue activities in depressive disorders. The details of the pharmacological
properties are present in another chapter of this book (Naoi et al. 2020).
424 M. Naoi et al.
Table 1 (continued)
Flavanones
Table 1 (continued)
Isoflavones
HO O
O OH
HO R
HO CH3
O O
(continued)
TCM Substances in Neuropsychopharmacotherapy: Basic Aspects with a Focus. . . 429
Table 2 (continued)
Diterpene lactones
Ginkgolide A, or B (R= H, or OH) Bilobalide
O
O
HO
R O HO
H tBu O
tBu
O O O
O OH
O H O
OH O O
H
O
O
O
H
O
O
O
Lavender essential oil, linalyl acetate, linanol Binding to glutamate NMDA receptor
5-HT transporter
Lavender essential oil 5-HT1A receptor voltage-gated Ca2+ channels
Hippocampal neurogenesis, neuroprotection
Linalool 3,7-dimethyl-1,6- octadien-3-ol Binding to 5-TH1A, D1 receptors
Sequiterpene
Valerenic acid
Table 2 (continued)
HO
O
(continued)
TCM Substances in Neuropsychopharmacotherapy: Basic Aspects with a Focus. . . 431
Table 2 (continued)
of an aliphatic unsaturated heptene linker with two aromatic rings attached at the
both ends. In animal depression models, curcumin inhibited MAO activity, increased
5-HT, NA, and DA and 5-HT1A expression, activated MAPK/ERK (extracellular
signal-related kinase) and NF-κB pathways, and induced BDNF and neurogenesis in
the hippocampus (Kaufmann et al. 2016). It has antioxidant, anti-inflammatory, anti-
cancer, anti-viral, and neuroprotective activities. The hydroxyl and methoxyl groups
can scavenge ROS and reactive nitrogen species (RNS), and prevent lipid peroxi-
dation in vivo and in vitro. Clinical intervention studies with curcumin (500 mg
twice a day) for 8 weeks ameliorated depressive and anxiety symptoms in patients
with MDD and especially with atypical depression (Lopresti et al. 2014). Curcumin
monotherapy (1000 mg/day) presented antidepressant activity measured with
HDRS-17 effectively as fluoxetine monotherapy in a RCT (Sanmukhani et al.
2014). Curcumin supplementation in MDD patients enhanced the efficacy of current
antidepressants by increase of plasma BDNF and decrease of inflammatory cyto-
kines, interleukin(IL)-1,β and tumor necrosis factor-α (TNF-α) (Yu et al. 2015).
Resveratrol is found in red grapes (Vitis vinifera L.), berries, and peanuts and a
compound conferring anti-aging, neuroprotective, anti-cancer, and calorie-restric-
tion mimic functions. It is composed of two phenol rings connected through a
styrene double bond. The trans-3,5,40 -trihydroxystilbene is more bioactive than
the cis-form, and it can directly scavenge hydroxyl radicals with its hydroxyl groups,
and 40 -hydroxyl group is the most reactive. Resveratrol exerts antidepressant effects
in animal models by increasing 5-HT and NA, inhibiting MAO-A expression and
activity, inducing BDNF expression in the hippocampus, suppressing phosphoryla-
tion of Akt/mTOR (mammalian target of rapamycin) pathway, and modulating the
HPA axis (de Oliveira et al. 2018). Tetrahydroxystilbene-2-O-D-glucoside isolated
from Polygonum multiforum induced BDNF and neurogenesis in the hippocampus
and presented antidepressant, antioxidant, and anti-inflammation functions in mouse
432 M. Naoi et al.
Vitamin B6 5-HT
TRPC 5-HT transporter MAO-A
Polyphenols 5-HT Hyperforin, silymarin
NE, DA NE, DA
5-HT, NE, DA, Transporter
Resveratrol,
GABA
BDNF genistein Polyphenols
BDNF BDNF
activity (Yanez et al. 2006). Piperine, a major alkaloid of Piper nigrum L. (black
pepper), increased 5-HT synthesis in the hippocampus and frontal cortex, inhibited
MAO and activated 5-HT1A and 5-HT1B receptors, and presented antidepressant
effects in mouse models of depression (Mao et al. 2011).
L-Tryptophan is metabolized also by indoleamine 2,3-dioxygenase (IDO) into
kynurenine and further kynurenic acid and quinolinic acid, a strong agonist of
NMDA receptor. These tryptophan catabolites (TRYCATs) can induce anxiety and
depression in animal models. Pro-inflammatory INF-α, β, -γ, IL-6, TNF-α and
oxidative stress upregulate IDO expression and deplete plasma L-tryptophan. The
negative relationship was reported between the levels of kynurenic acid and C-reac-
tive protein in the blood and the volumes of hippocampal cornu ammonis and
subiculum in patients with MDD, suggesting the association of IDO with the
pathogenesis (Doolin et al. 2018). Trans-Astaxanthin, a red carotenoid pigment
rich in algae and ferulic acid, inhibited IDO in the hippocampus, frontal cortex,
striatum, and hypothalamus, suppressed L-tryptophan-kynurenine pathway, and
increased 5-HT synthesis in a mouse model of depression (Koshiguchi et al. 2017).
Flavonoids, phenolic acids, and curcumin bind to 5-HT1A and 5-HT2A receptors,
activate intracellular signal pathways, and function as antidepressants in animal models.
Chronic curcumin administration increased 5-HT levels, activated cAMP-cAMP response
element binding protein (CREB) pathway, and increased expression of 5-HT1A receptor
and BDNF in the hippocampal CA1, cortex and hypothalamus (Li et al. 2009). Vitexin,
nobiletin, amentoflavone and gallic acid increased monoamines in synaptic cleft and
interacted with 5-HT1A, noradrenergic α2, and D1, D2, and D3 receptors.
During specific development periods, 5-HT affects maturation of the brain archi-
tecture associated with emotional behavior in postnatal life (Suri et al. 2015). In the
adult hippocampus, 5-HT signals from the dorsal and median raphe modulate
neuroplasticity and neurogenesis, and prevent the cortical atrophy and behavioral
and cognitive abnormalities in depression (Kraus et al. 2017). Antidepressants, elec-
troconvulsive shock, exercise, and enriched environment promote proliferation and
survival of newborn neurons in the hippocampus. Postmortem studies of brains from
patients with MDD confirmed that antidepressant treatment increased the total dentate
cell number and dentate gyrus size (Boldrini et al. 2013). 5-HT stimulates 5-HT2A
receptor, increases cAMP production, activates CREB, and induces transcription of
BDNF. Vice versa BDNF promotes the survival and differentiation of 5-HT neurons in
the dentate gyrus. CREB1 polymorphisms were associated with decreased volume and
activity of the hippocampus in BD and MDD. These results indicate that 5-HT
functions not only as a neurotransmitter, but also as a modifier of cell signal transduc-
tion to regulate hippocampal neurogenesis and incidence of depression.
behaviors, MDD, BD, and ADHD. MAO-A level in the brain is determined before
the birth, and MAO-A and 5-HT regulate development of neuronal architecture.
MAO-B appears only in the postnatal stage and increases with age, and induces
oxidative stress, suggesting its association with the pathogenesis of age-dependent
neurodegeneration. In the brain of MDD patients, MAO-A binding measured using
[11C]harmine PET significantly increased in the prefrontal and anterior cingulate
cortex according to greater disease severity (Chiuccariello et al. 2014). In the
postmortem prefrontal cortex of untreated MDD patients, expression of R1, a
transcriptional repressor of MAO-A protein, significantly decreased and MAO-A
protein increased (Johnson et al. 2011). Chronic stress enhanced glucocorticoid and
activated a transcription factor Krüpple-like factor 11 (KLF11), upregulated Mao-A
expression, and caused hippocampal atrophy (Harris et al. 2015). In the postmortem
prefrontal cortex of patients with MDD, protein expression of KLF11 and MAO-A
significantly increased by 36% and 44% from control, respectively. In the carriers
with longer variant of an upstream variable-number-tandem-repeat (uVNTR) pro-
moter region with high transcription function, enhanced expression of MAO-A was
implicated in depression, insomnia, and suicide attempt (Ziegler and Domschke
2018).
Phytochemicals, especially flavonoids (acacetin, apigenin, chrysin, luteolin, quer-
cetin, and wogonin) are MAO-A inhibitors, which increase available 5-HT, NA, and
DA, suppress ROS/RNS production, and improve depressive behaviors. The inhi-
bition is reversible and in a competitive way to substrates, and does not cause
hypertension crisis called “cheese effect,” a serious side effect of irreversible
MAO-A inhibitors (Carradori et al. 2016). Catechol (ortho-dihydroxy group) struc-
ture of the flavonoid B ring is required for MAO inhibition, whereas a hydroxyl
group at position 3 in the C ring decreases the inhibitory potential. Hydroxyl groups
at position 5 in the A ring and at position 3 of the C ring are required to access to the
substrate binding site, as shown by comparison of genistein (40 ,5,7-tri-
hydroxyisoflavone) with daidzein (40 ,7-dihydroxyisoflavone), and kaempferol
(3,5,7,40 -tetrahydroxyflavonol) with luteolin (5,7,30 ,40 -tetrahydroflavone). The
order of inhibitory potency of flavonoids against MAO is flavone, flavonol >
flavone glycoside > flavanonol (Guan and Liu 2016). Eugenol (4-allyl-2-
methoxyphenol, a major active component of Rhizoma acori graminei), purpurin
(1,2,4-trihydroxanthroquinone), hispidol (6,40 -dihydroxyaurone), chelerythrine (an
isoquinoline alkaloid), and osthenol (a hydroxycoumarin isolated from the roots of
Angelica pubescens) are potent MAO-A inhibitors. On the other hand, non-flavo-
noid polyphenols, curcumin, tetrahyrocurcumin, ellagic acid, trans-resveratrol, and
Ginkgo biloba extract (EGb 761) inhibit MAO-B in preference to MAO-A and show
neuroprotective activities in animal models of neurodegenerative diseases.
Rosmarinic acid,
EGCG, EC, 7,8-DHF, 7,8,3’-THF, Curcumin, hesperidin,
6-Hydroxyflavone,
Hesperetin Fisetin, curcumin Quercetin, resveratrol
6-Methoylflavonee
EGCG, genistein,
Curcumin, resveratrol,
EGCG, quercetin
Berberine, PI3K PKC Baicalein
Amentiflavone
MAPK cascades
Naringen, apigenin, curcumin, hesperidin,
ferulic acid, rosmarinic acid
Naringen, hesperetin,
Genistein, kaempferol,
Berberine, curcumin, AKT/PKB Rutin, liquiritin, icaritin
ERK1/2
Ferulic acid, resveratrol,
Salvianolic acid
and promote structural and functional plasticity of neurons and synapses in the
prefrontal cortex and hippocampus (Ignacio et al. 2016). BDNF, VEGF, and insu-
lin-like growth factor (IGF) activate mTOR pathway and increase protein synthesis
via eukaryotic initiation factor 4E (eIF4E)-binding protein (eIF4BP) and ribosomal
protein S6 kinases (S6Ks) protein. In the brains of MDD patients, mTOR signaling
pathway was impaired and phosphorylated eIF4BP and S6Ks and mTOR/p70S6K/
eIF4B function decreased (Jernigan et al. 2011). Yueju (a TCM consisting of five
herbs) and radix polygalate isolated from the root of Polygala tenuifolia have been
used to treat depression, anxiety, and irritability in China, exerted ketamine-like
rapid antidepressant activity through reversing the reduction of phosphorylated
mTOR, phospho-p70S6K, and immediately increased BDNF expression in the
hippocampus, as shown in rodent depression models (Tang et al. 2015). In a RCT
for patients with MDD, Yueju plus fluoxetine treatment exerted fast-onset antide-
pressant function with significantly decreased HDRS-24 score from day 3 to 7 in
correlation with increase in serum BDNF level (Wu et al. 2015). Among constituent
herbs of Yueju, Gardenia jasminoides J. Ellis exerted rapid antidepressant potency,
to which the ingredients, crocins, gardenosides, and iridoid glycosides (cornin)
contributed (Wu et al. 2016b).
(Derosa et al. 2016). Icariin, ferulic acid, and trans-astaxanthin downregulated expres-
sion of inflammatory cytokines in the hippocampus, increased BDNF, normalized GR
function, HPA hyperactivity, and glucocorticoid sensitivity, and exert antidepressant
activity in a rat model of depression (Liu et al. 2015). Resveratrol inhibited NF-κB and
activator protein-1 (AP-1) and suppressed cytokine expression. Stilbenoids
(picearannol, 5-methoxy-3-stilbenol) inhibited PI3K/AKT pathway and downregulated
production of IL-6 and monocyte chemotactic peptide-1 (MCP1) (Eräsolo et al. 2018).
Fisetin and chiisanoside from the leaves of Acanthopanax sessiliflorus inhibited iNOS
and COX-2, decreased IL-6 and TNF-α levels in serum, activated BDNF/TrkB/NF-κB
pathway in the hippocampus and prefrontal cortex, and showed antidepressant activity
in mouse models of depression.
Stress, early-life experiences, adverse life events, and physical abuse cause long-
lasting epigenetic alteration of DNA and histone, and are powerful risk factors
influencing the vulnerability to develop depression in adulthood. Under stressful
conditions, corticotropin-releasing hormoneis released from neurons in the para-
ventricular nucleus, and stimulates the synthesis and release of adenocorticotropic
hormone (ACTH) and glucocorticoids. Prolonged severe stress sustains glucocorti-
coid elevation, damages hippocampal neurons, inhibits neurogenesis, and influences
the activity of several cortical and subcortical structures involving in endocrine,
motor, effective, and cognitive functions. Abnormal excessive activation of the HPA
and alterations of the negative feedback were observed in approximately half of the
individuals with depression, and antidepressant treatment could restore the dysfunc-
tion (Surget et al. 2011). In postpartum depression, functional changes in the HPA
axis were correlated with metabolite levels in the anterior cingulate gyrus (de
Rezende et al. 2018). Alterations in the hippocampal GRs and the HPA axis activity
have close relationship with cytokines and inflammatory signal pathways, such as
MAPK-NF-κB pathway.
Puerarin,resveratrol and flavonoids extracted from Xiaobuxin-tang (a TCH
decoction) modulated the HPA axis hyperactivity, decreased corticotropin-releasing
hormone, corticosterone, and ACTH, and inhibited hippocampal GR expression in
depression rodent models (Yang et al. 2017). Apigenin, baicalin, and salvianolic acid
B extracted from Salvia miltiorrhiza inhibited the activation of NLRP3
inflammasome in the prefrontal cortex and IL-1β production, reversed HPA hyper-
activity and showed antidepressant effects in rats exposed to CUMS (Li et al. 2016).
Anxiety disorders share many symptoms with depression and anxiety is associated
with depression either as risk markers or causal risk factors, and anxiety occurs in
46% and 52% of patients with MDD and BD. Piper methysticum, Melissa officinalis,
440 M. Naoi et al.
Discussion
In long history of TCM, medical herbs and the formulas, Chaihu-Shugan-San, Ban-
xia-hou-pu-tang, Gan-mai-da-zao-tang, and Xia-yao-san have been used for the
treatment for depression (Zhang and Cheng 2019). Diverse functions of TCM herbs
and the formulas exert therapeutic effects additively or synergistically, as evaluated
based on the philosophical holism and TCM symptom criteria. However, the clinical
442 M. Naoi et al.
Cross-References
References
Andreatini R, Sartoni VA, Seabra ML, Leite JR. Effect of valepotriates (valerian extract) in
generalized anxiety disorder: a randomized placebo-controlled pilot study. Phytother Res.
2002;16(7):650–4.
Appel K, Rose T, Fiebich B, Kammier T, Hoffmann C, Weiss G. Modulation of the γ-aminobutyric
acid (GABA) system by Passiflora incarnata L. Phytother Res. 2011;25(6):838–43.
Atteritano M, Mazzaferro S, Bitti A, et al. Genistein effects on quality of life and depression
symptoms in osteopenic postmenopausal women: a 2-year randomized, double-blind, controlled
study. Osteoporos Int. 2014;25(3):1123–9.
Badavath VN, Baysal I, Ucar G, Sinha BN, Jayaprakash V. Monoamine oxidase inhibitory activity
of ferulic acid amides: curcumin based design and synthesis. Arch Pharm Life Sci. 2016;349
(1):9–19.
Birkmayer W, Riederer P. Biochemical post-mortem findings in depressed patients. J Neural
Transm. 1975;37(2):95–109.
Boerner RJ, Sommer H, Berger W, Kuhn U, Schmidt U, Mannel M. Kava-kava extract LI 150 is as
effective as opioramol and buspirone in generalised, double-blind multi-centre clinical trial in
129 out-patients. Phytomedicine. 2003;10(Suppl 4):38–49.
Boldrini M, Santiago A, Hen R, Dwork AJ, Rosoklija GB, Tamir H, Arango V, Mann JJ.
Hippocampal granule neuron number and dentate gyrus volume in antidepressant-treated and
untreated major depression. Neuropsychopharmacology. 2013;38(6):1068–77.
Capra JC, Cunha MP, Machado DG, Zomkowski AD, Mendes BG, Santos AR, Pizzolatti MG,
Rodrigues AL. Antidepressant-like effect of scopoletin, a coumarin isolated from Polygala
sablola (Polygalacase) in mice: evidence for the involvement of monoaminergic system. Eur J
Pharmacol. 2010;643(2–3):232–8.
Carradori S, Gidaro MC, Petzer A, Costa G, Gulglielmi P, Chimenti P, Alcaro S, Petzer JP.
Inhibition of human monoamine oxidase: biological and molecular modeling studies on selected
natural flavonoids. J Agric Food Chem. 2016;64(47):9004–11.
Chang SC, Cassidy A, Willett WC, Rimm EB, O’Reilley EJ, Okereke OI. Dietary flavonoids intake
and risk of incident depression in midlife and older women. Am J Clin Nutr. 2016;104
(3):704–14.
Chimenti F, Fioravanti R, Bolasco A, et al. A new series of flavones, thioflavones, and flavanones as
selective monoamine oxidase-B inhibitors. Bioorg Med Chem. 2010;18(3):1273–9.
Chiuccariello L, Houle S, Miler L, et al. Elevated monoamine oxidase A binding during major
depression episodes is associated with greater severity and reverse neurovegetative symptoms.
Neuropsychopharmacology. 2014;39(4):973–80.
Dai CX, Hu CC, Shang YS, Xie J. Role of Ginkgo biloba extract as an adjunctive treatment of
elderly patients with depression and on the expression of serum S100B. Medicine (Baltimore).
2018;97(39):e12421.
Davinelli S, Scapagnini G, Marzatico F, Nobile V, Ferrara N, Corbi G. Influence of equal and
resveratrol supplementation on health-related quality of life in menopausal women: a random-
ized, placebo-controlled study. Maturitas. 2017;96:77–83.
de Oliveira MR, Chenet AL, Duarte AR, Scaini G, Quevedo J. Molecular mechanisms
underlying the anti-depressant effects of resveratrol: a review. Mol Neurobiol.
2018;55(6):4543–39.
de Rezende MG, Rosa CE, Garcia-Leal C, et al. Correlation between changes in the hypothalamus-
pituitary-adrenal axis and neurochemistry of the anterior cingulate gyrus in postpartum depres-
sion. J Affect Disord. 2018;239:274–81.
de Sousa DP, Silva RHN, da Silva EF, Gavioli EC. Essential oils and their constituents: an
alternative source for novel antidepressants. Molecules. 2017;22(8):E1290.
Derosa G, Maffioli P, Simental-Mendia LE, Bo S, Sahebkar A. Effect of curcumin on circulating
interleulin-6 concentrations: a systematic review and meta-analysis of randomized controlled
trials. Pharmacol Res. 2016;111:394–404.
Deyama S, Bang E, Wohleb ES, et al. Role of neuronal VEGF signaling in the prefrontal cortex in
the rapid antidepressant effects of ketamine. Am J Psychiatry. 2019;176(5):388–400.
444 M. Naoi et al.
Jacka FN, O’Neil A, Opie R, et al. A randomised controlled trial of dietary improvement for adults
with major depression (the ‘SMILES’ trial). BMC Med. 2017;15:23.
Jang SW, Okada M, Sayeed I, Xiao G, Stein D, Jin P, Ye K. Gambogic amide, a selective agonist for
TrkA receptor that possesses robust neurotrophic activity, prevents neuronal cell death. Proc
Natl Acad Sci USA. 2007;104(41):16329–34.
Jernigan CS, Goswami DB, Austin MC, Iyo AH, Chandran A, Stockmeier CK, Karolewicz B. The
mTOR signaling pathway in the prefrontal cortex is compromised in major depressive disorder.
Prog Neuro-Psychopharmacol Biol Psychiatry. 2011;35(7):1774–9.
Jeschke E, Ostermann T, Vollmar HC, Tabali M, Matthes H. Depression, comorbidities, and
prescriptions of antidepressants in a German network of GPs and specialists with sub-
specialisation in anthroposophic medicine: a longitudinal observational study. Evid Based
Complement Alternat Med. 2012;2012:508623.
Jin X, Liu P, Yang F, Zhang Y, Miao D. Rosmarinic acid ameliorates depression-like behaviors in a
rat model of CUS and up-regulates BDNF levels in the hippocampus and hippocampal-derived
astrocytes. Neurochem Res. 2013;38(9):1828–37.
Johnson S, Stockmeier CA, Meyer JH, et al. The reduction of R1, a novel repressor protein for
monoamine oxidase A, in major depressive disorder. Neuropsychopharmacology. 2011;36
(10):2139–48.
Johnston GAR. Flavonoid nutraceuticals and ionotropic receptors for the inhibitory neurotransmit-
ter GABA. Neurochem Int. 2015;89:120–5.
Karim N, Curmi J, Gavande N, Johnston BAR, Hanrahen JR, Tierney ML, Chbib M. 20 -Methoxy-6-
methylflavone: a novel anxiolytic and sedative with subtype selective activating and modulating
actions at GABAA receptors. Br J Pharmacol. 2012;165(4):880–96.
Kasper S, Gastpar M, Müller WE, Volz HP, Möller HJ, Schläfke S, Dienel A. Lavender oil
preparation Silexan is effective in generalized anxiety disorder – a randomized, double-blind
comparison to placebo and paroxetine. Int J Neuropsychopharmacol. 2014;17(6):859–69.
Kaufmann FN, Gazal M, Bastos CR, Kaster MP, Ghisleni G. Curcumin in depressive disorders: an
overview of potential mechanisms, preclinical and clinical findings. Eur J Pharmacol.
2016;784:192–8.
Kaur A, Garg S, Shielh BA, Singh N, Singh P, Bhatti R. In Silico studies and in vivo MAOA
inhibitory activity of coumarins isolated from Angelica archangelica extract: approach towards
antidepressant activity. ACS Omega. 2020;5:15069–76.
Kendler KS, Gatz M, Gardner CO, Redersen M. A Swedish national twin study of lifetime major
depression. Am J Psychiatry. 2006;163(1):109–14.
Khan H, Pervitz S, Sureda A, Nabavi S, Tajada S. Current standing of plant derived flavonoids as an
antidepressant. Food Chem Toxicol. 2018;119:176–88.
Khom S, Stommer B, Ramaharter J, et al. Valerenic acid derivatives as novel subunit-selective GABAA
receptor ligands –in vitro and in vivo characterization. Br J Pharmacol. 2010;16(1):65–78.
Koshiguchi M, Komazaki H, Hirai S, Egashira Y. Ferulic acid suppresses expression of trypto-
phan metabolic ken enzyme indoleamine 2,3-dioxygenase via NF-κB and p38MAPK in
lipopolysaccharide-stimulated microglial cells. Biomed Biotechnol Biochem. 2017;81
(5):966–71.
Kraus C, Castren E, Kasper S, Lanzenberger R. Serotonin and neuroplasticity – links between
molecular, functional and structural pathophysiology in depression. Neurosci Biobehav Rev.
2017;77:317–26.
LaChance LR, Ramsey D. Antidepressant foods: an evidence-based nutrients profiling system for
depression. World J Psychiatry. 2018;8(3):97–104.
Lassale C, Batty GD, Baghdadli A, Jacka F, Sanchez-Villegas A, Kivimäki M, Akbaraly T. Healthy
dietary indices and risk of depressive outcomes: a systematic review and meta-analysis of
observational studies. Mol Psychiatry. 2019;24(7):965–86.
Leonard BE. Inflammation and depression: a causal or coincidental link to the pathophysiology?
Acta Neuropsychiatr. 2018;30(1):1–16.
446 M. Naoi et al.
Sarris J, Byrne GJ, Stough C, et al. Nutraceuticals for major depressive disorder- more is not
merrier: an 8-week double-blind, randomised, controlled trial. J Affect Disord. 2019;245:
1007–15.
Sasaki K, El Omri E, Kondo S, Han J, Isoda H. Rosmarinus officinalis polyphenols produce anti-
depressant like effect through monoaminergic and cholinergic functions modulation. Behav
Brain Res. 2013;238:86–94.
Sashidhara KV, Modukuri RK, Singh S, Rao KB, Teja GA, Gupta S, Shukla S. Design and synthesis
of new series of coumarin-aminopyran derivatives possessing potential anti-depressant-like
activity. Bioorg Med Chem Lett. 2015;25(2):337–41.
Savage KM, Stough CK, Byrne GJ, et al. Kava for the treatment of generalized anxiety disorder (K-
GAD): study protocol for a randomised controlled trial. Trials. 2015;16:493.
Savage K, Firth J, Stough C, Sarris J. GABA-modulating phytomedicines for anxiety: a systematic
review of preclinical and clinical evidences. Phytother Res. 2018;32(1):3–18.
Scheid V. Depression, constraint and the liver: (Dis)assembling the treatment of emotion-related
disorders in Chinese medicine. Cult Med Psychiatry. 2013;37(1):30–58.
Shafiee M, Arekhi A, Omaranzadeh A, Sahebkar A. Saffron in the treatment of depression, anxiety
and other mental disorders: current evidence and potential mechanisms of action. J Affect
Disord. 2018;227:330–7.
Shakeri A, Sahebkar A, Javadi B. Melissa officinalis L. – a review of its traditional uses,
phytochemistry and pharmacology. J Ethnopharmacol. 2016;188:204–28.
Sowa-Kucma M, Styczen K, Siwek M, Misztak P, Nowak RJ, Dudek D, Rybakowski J, Nowak G,
Maes M. Lipid peroxidation and immune biomarkers are associated with major depression and
its phenotypes, including treatment-resistant depression and melancholia. Neurotox Res.
2018;33(2):448–60.
Squillaro T, Cimini A, Peluso G, Giordano A, Melone MAB. Nano-delivery systems for encapsu-
lation of dietary polyphenols: an experimental approach for neurodegenerative diseases and
brain tumors. Biochem Pharmacol. 2018;154:303–17.
Sun J, Kong L, Wu F, Wei Y, Zhu Y, Yin Z, Deng X, Jiang X, Tang Y, Wang F. Decreased plasma
glial cell line-derived neurotrophic factor level in major depressive disorder is associated with
age and clinical severity. J Affect Disord. 2019;245:602–7.
Surget A, Tanti A, Leonardo ED, Laugeray A, Rainer Q, Touma C, Palme R, Griebel G, Ibarguen-
Vargas Y, Hen R, Belzung C. Antidepressants recruit new neurons to improve stress response
regulation. Mol Psychiatry. 2011;16(12):1177–88.
Suri D, Teixeira CM, Cagliostro MKC, Mahadevia D, Ansorge MS. Monoamine-sensitive devel-
opmental periods impacting adult emotional and cognitive behaviors. Neuropsychophar-
macology. 2015;40:88–112.
Tang J, Xue W, Xia B, et al. Involvement of normalized NMDA receptor and mTOR-related
signaling in rapid antidepressant effects of Yueju and ketamine on chronically stressed mice.
Sci Rep. 2015;5:13573.
Toth B, Hegyi P, Lantos T, et al. The efficacy of saffron in the treatment of mild to moderate
depression: a meta-analysis. Planta Med. 2019;85(1):24–31.
Wang F, Xu Z, Yuen CT, Chow CY, Lui YL, Tsang SY, Xue H. 6,20 -Dihydroxyflavone, a subtype-
selective partial inverse agonist of GABAA receptor benzodiazepine site. Neuropharmacology.
2007;53(4):574–82.
Wang Y, Wang B, Lu J, Shi H, Gong S, Wang Y, Hamdy RC, Chua BHL, Yang L, Xu X. Fisetin
provides antidepressant effects by activating the tropomycin receptor kinase B signal pathway in
mice. J Neurochem. 2017;143(5):561–8.
Wang Y, Shi Y, Xu Z, Fu H, Zeng H, Zheng G. Efficacy and safety of Chinese herbal medicine for
depression: a systematic review and meta-analysis of randomized controlled trials. J Psychiatr
Res. 2019;117:74–91.
Wu R, Zhu D, Xia Y, et al. A role of Yueju in fast-onset antidepressant action on major depressive
disorder and serum BDNF expression: a randomly double-blind, fluoxetine-adjunct, placebo-
controlled, plot clinical study. Neuropsychiatr Dis Treat. 2015;11:2013–21.
TCM Substances in Neuropsychopharmacotherapy: Basic Aspects with a Focus. . . 449
Wu R, Shui L, Wang S, Song Z, Tai F. Bilobalide alleviates depression-like behavior and cognitive
deficit induced by chronic unpredictable mild stress in mice. Behav Pharmacol. 2016a;27
(7):596–605.
Wu R, Tao W, Zhang H, et al. Instant and persistent antidepressant response of Gardenia yellow
pigment is associated with acute protein synthesis and delayed upregulation of BDNF expres-
sion in the hippocampus. ACS Chem Neurosci. 2016b;7(8):1068–76.
Yanez M, Fraiz N, Cano E, Orallo F. Inhibitory effects of cis- and trans-resveratrol on noradrenaline
and 5-hydroxytryptamne uptake and on monoamine oxidase activity. Biochem Biophys Res
Commun. 2006;344(2):688–95.
Yang XH, Song SQ, Xu Y. Resveratrol ameliorates chronic unpredictable mild stress-induced
depression-like behavior: involvement of the HPA axis, inflammatory markers, BDNF, and
Wnt/β-catenin pathway in rats. Neuropsychiatr Dis Treat. 2017;13:2727–36.
Yu Q, Chang Q, Liu X, Gang S, Ye K, Lin X. 7,8,30 -Trihydroxyflavone, a potent small molecule
TrkB receptor agonist, protects spiral ganglion neurons from degeneration both in vitro and in
vivo. Biochem Biophys Res Commun. 2012;422(3):387–92.
Yu JJ, Pei LB, Zhang Y, Wen ZY, Yang JL. Chronic supplementation of curcumin enhances the
efficacy of antidepressants in major depressive disorder: a randomized, double-blind, placebo-
controlled pilot study. J Clin Psychopharmacol. 2015;35(4):406–10.
Zhang YW, Cheng YC. Challenge and prospect of traditional Chinese medicine in depression
treatment. Front Neurosci. 2019;13:190.
Zhen HH, Quan YC, Peng Z, Han Y, Zheng ZJ, Guan LP. Design, synthesis and potential
antidepressant-like activity of 7-prenyloxyl-2,3-dihydroxyflavone derivatives. Chem Biol
Drug Des. 2016;87(6):858–66.
Ziegler C, Domschke K. Epigenetic signature of MAOA and MAOB genes in mental disorders.
J Neural Transm. 2018;125(11):1581–8.
Zirak N, Shafiee M, Soltani G, Mirzaei M, Sahebkar A. Hypericum perforatum in the treatment of
psychiatric and neurodegenerative disorders: current evidence and potential mechanisms of
action. J Cell Physiol. 2019;234(6):8496–508.
TCM Substances in
Neuropsychopharmacotherapy
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 452
Medicinals of TCM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 453
Ginkgo biloba L. (银杏) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 453
Radix et Rhizoma Ginseng/Ginseng (人参) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 454
Stigma Croci (藏红花) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 458
Gardeniae Fructus (栀子) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 459
Rhizoma Curcumae Longae (姜黄) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 461
Radix et Rhizoma Glycyrrhizae (甘草) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 464
Radix Rehmanniae Raw (生地黄) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 467
Radix Rehmanniae Praeparata (熟地黄) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 467
Bulbus Lilii (百合) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 470
Hypericum perforatum/St.-John’s Wort (贯叶连翘/圣约翰草) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 471
Formulae of TCM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 473
Yueju Pill (越鞠丸) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 474
Xiao Buxin Tang Decoction (小补心汤) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 476
Bulbus Lilii and Radix Rehmanniae Tang Decoction (百合地黄汤) . . . . . . . . . . . . . . . . . . . . . . 477
Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 479
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 479
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 480
Abstract
With the unique and comprehensive understanding of human’s body and vari-
ous pathological conditions, Traditional Chinese Medicine (TCM) has undoubt-
edly become a field of increasing interests from doctors, researchers, and even
patients these years. A number of specialties benefited from the combination of
Y. Wang
Tianjin Mental Health Institute, Tianjin Anding Hospital, Tianjin, China
e-mail: ashleywy@126.com
J. Li (*)
Institute of Mental Health, Tianjin Anding Hospital, Tianjin, China
e-mail: jieli@tmu.edu.cn
Introduction
For years, precision medicine has prevailed not only in clinical medicine, but also
researches from fields related to medicine. Increasing resources have been attracted
to the search for the precise explanations, mechanisms, and treatments for abnormal
conditions, diseases, pathogens, etc. Neither psychiatry nor psychology was excep-
tion. Accumulating breakthroughs and findings have been gathered along the way,
most of which were also proved beneficial for patients. Yet the more breakthroughs
we make, the more intensely we realize that “one drug” or “one formula” seems
reasonable to equalize “one disease,” but not anymore once we switched “one
disease” with “one patient.” Human’s body is a complex system. In physiology,
we learned there are eight systems in a human’s body, yet, we cannot comprehend a
patient’s condition in merely eight aspects accordingly, considering the complexity
and wholism of the human’s body. This could partially explain why Traditional
Chinese Medicine (TCM) caught so much attention these days, aside from its being
equally effective for various diseases and well-acknowledged fewer adverse effects.
According to the theory of TCM, diseases are generally related to disorders of
Zang-fu (脏腑) functions and imbalance between Yin and Yang in human body.
Different pathogenic factors affecting the human body can cause excess or defi-
ciency of Yin or Yang, and the dysfunction of Zang-fu as well as Meridians and
Collaterals (经络). The basic principle of treating diseases with traditional Chinese
materia medica is to enhance the body resistance, expel pathogenic factors, and
restore the normal functions of Zang-fu organs, so that the excess or deficiency of
Yin or Yang can be corrected (Chen et al. 2017).
The symptoms related to psychiatric disorders, such as dysthymia or elation, had
been recognized and documented in different articles and books, which could also be
considered as the imbalance between Yin and Yang and the dysfunction of Zang-fu.
TCM Substances in Neuropsychopharmacotherapy 453
Hence, treatments for these symptoms, or diseases, are focused on the restoration of
the normal balance of Yin-Yang and function of Zang-fu. There are generally two
different ways to adjust Yin and Yang:
Removing the excess: The so-called excess refers to the surplus syndrome of either
Yin or Yang, which can be treated essentially with “removal.” “Treating heat with
cold” (热者寒之), or “treating cold with hot” (寒者热之) as one of the basic
principle of the treatment has been widely accepted in the clinical prescription of
TCM (Chen et al. 2017).
Supplementing the deficiency: This strategy is used to treat different kinds of
deficiency syndromes. It is usually carried out by using medicines that tonify
the Qi, blood, Yin or Yang. In many chronic diseases, deficiency of Yin or Yang
often involves or exists in more than one organ of the body, which is usually
treated with different tonic herbs (Chen et al. 2017).
There are thousands of herbs and medicinals of TCM; the most commonly used
ones are more than 500 nowadays. For practical considerations, this chapter includes
characteristic herbs, medicinals, and formulae of TCM that has been acknowledged
to be effective for psychiatric disorders, and also with abundant evidences form
different studies in recent years.
Medicinals of TCM
Ginkgo biloba L., family of Ginkgoaceae, has a bitter, sweet, and astringent flavor.
According to TCM, it is capable of activating blood circulation and resolving stasis,
freeing the collateral vessels and relieving pain, astringing the lung and calming
panting, resolving turbidity (Chen et al. 2017).
Recommended dosage: Decoction, 9–12 g. It is not suggested for patient with
excessive syndrome. It is contraindicated to eat fish when taking Ginkgo Folium.
Cautions and interactions
Case reports have illustrated that Ginkgo biloba may potentiate bleeding when
combined with warfarin or aspirin, increase blood pressure when combined with
thiazide diuretics, and has even led to a coma when combined with trazodone, a
serotonin antagonist and reuptake inhibitor used for depression (Chen et al. 2011).
Researches in recent years:
Previous studies proved that ginkgo biloba extract (GBE) could affect nucleotide
metabolism and protein biosynthesis to enhance cell proliferation, as well as the
metabolism of glutamate and aspartate, which plays an important role in the antide-
pressant effects (Bai et al. 2017; Hu et al. 2018; Liang et al. 2016). The mechanism
of GBE in relieving depressive behaviors was also demonstrated to be related with
reducing inflammatory response which is involved in the development of depression
(Liang et al. 2016). To date, studies available have illustrated that extract from
Ginkgo biloba leaves could reverse depression-associated gut dysbiosis and
454 Y. Wang and J. Li
increased the richness of lactobacillus species which had been proven to be a path to
relieve depression (Chen et al. 2019). Meantime, it was suggested that GBE pro-
duced an antioxidant effect against oxidative stress, which may be partly responsible
for its well-observed neuroprotective effects as well (Rojas et al. 2011).
GBE has been used in treating various diseases, including angina pectoris (Sun et
al. 2015), glaucoma (Kang and Lin 2018), brain injury (Tulsulkar et al. 2016),
depression (Dai et al. 2018), and dementia (Napryeyenko et al. 2009; Scripnikov
et al. 2007). Its antiapoptotic, antioxidative, anti-inflammatory properties have been
widely acknowledged (Diamond and Bailey 2013).
These years, Ginkgo biloba has mostly been applied as augmentative treatment for
dementia, stroke, and cardiovascular diseases, most of which are usually seen among
geriatric population. Increasing concerns about the comorbid depression have caught
more attention from the public. Considering the specific and unusual condition of this
population, treatment of choice should be very cautious. As such, natural herbs are
more preferred than conventional medications. It has been suggested that a dose of
240 mg/d of Ginkgo biloba extract was able to stabilize or slow decline in cognition,
function, behavior at 22–26 weeks in cognitive impairment and dementia, especially
for those with neuropsychiatric symptoms (Tan et al. 2015). Researchers also found
that the GBE was a good augmentation of venlafaxine in treating Post stroke depres-
sion (PSD). Meanwhile, the addition of GBE could significantly reduce the dose of
venlafaxine, which was very important for elderly patients (Liang et al. 2019). Also,
GBE did benefit antidepressant-like behaviors and improved cardiac functions in mice
with heart failure (Zhang et al. 2019).
estrogen receptor (ERs) ligands and Rh2 and K as glucocorticoid receptor (GRs)
ligands may affect the function of HPA axis mediated by ERs or GRs (Lee and Ji
2014). Furthermore, Rb1 and Rg3 could attenuate neuronal toxicity, which is related
to glucocorticoid activity (Kim et al. 2010).
Ginseng prevents various diseases by ameliorating tissue injury and immune cell
death, while modulating immune cells to limit inflammatory responses. The HPA
axis is the major pathway regulating the immune response to stress (Lee and Rhee
2017). In the event of severe stress-induced dysfunction of the HPA axis, endocrine
homeostasis is disturbed, which can predispose the patient to a number of diseases.
Patients with depression and anxiety often exhibit increased production of HPA
hormones, including CRH and AVP, as compared to healthy individuals (Scott and
Dinan 1998). Researchers found that chronic pretreatment with ginsenoside-Rg1
prior to stress exposure could suppress inflammatory pathway activity through
alleviating the over-expression of proinflammatory cytokines and the activation of
microglia and astrocytes (Fan et al. 2018).
Amyloid-β peptides (Aβ), regarded as a main pathogenic factor of Alzheimer’s
disease, is playing an emerging role in stress response as well as in depression. An
increasing number of reports have shown that many other kinds of ginsenosides,
such as Re, Rg1, and Rf, had the capability to down-regulate the Aβdeposition and
the tau protein phosphorylation. These results suggest that these ginsenosides might
have antidepressant effects through regulating Aβ (Cao et al. 2016; Du et al. 2018;
Shim et al. 2017).
Different methods of processing the ginseng were demonstrated to affect the
absorption. The rate and extent of absorption of Rg1, Re, Rb1, and Rd (all
ginsenosides) appeared to be affected by the different methods used in processing
the ginseng samples. The areas under the plasma drug concentration-time curves
(AUCs) of Rg1, Re, Rb1, and Rd were significantly higher than those of the pure
ginsenosides (unprocessed). In addition, the AUCs of Rg1, Re, Rb1, and Rd were
different for WG (white ginseng), FG (frozen ginseng), and RG (red ginseng). The
amounts of Rg1, Re, Rd, and Rb1 were significantly (p < 0.05) higher in the
tissues than those of the pure ginsenosides (Chen et al. 2018). Also it showed that
the blood-brain barrier was poorly permeable to the ginsenosides as the concen-
trations of Rg1, Re, Rb1, and Rd in the brain tissues was much lower than those in
the plasma. These results are in agreement with those obtained in a previous study
(Han et al. 1986).
Another study also demonstrated that white ginseng has an antidepressant effi-
cacy in an animal model of menopausal depression (Jang et al. 2019). Ginsenoside
Rb1 and its metabolite compound K ameliorated depression-like behaviors during a
menopausal depressive-like state in female mice through the 5-hydroxytryptamine
2A-receptor (Yamada et al. 2011).
Even with the abundant evidence collected from various models, it is over-
optimistic to conclude that ginseng can be beneficial for depressive patients, espe-
cially without enough support from carefully and strictly conducted randomized
clinical trials or high-quality meta-analyses. Clinical studies with a large sample size,
longer duration, and follow-up should be considered in the future.
458 Y. Wang and J. Li
Stigma Croci is the dried flower chapiter of Crocus sativus L., pertaining to
Iridaceae, and has a unique slightly irritating smell and is slightly bitter in taste. It
is sweet in flavor and slightly cold in nature. It has the same functions with Flos
Carthami (hong hua), such as invigorating blood, dispelling stasis, and promoting
menstruation. However, its medical efficacy is much stronger. It can also cool the
blood and resolve toxins, for which it is especially used for purpura caused by heat
entering the ying and blood levels of warm diseases, and dark macula due to heat
stagnancy and blood stasis (Teng et al. 2019). Crocus sativus L.(Saffron) is an herb
used for “blood stasis syndrome” according to the theory of TCM; it also happens to
be the world’s most expensive spice. Apart from its traditional value as a food
additive, recent studies indicate its potential as an anticancer agent and memory
enhancer (Abe and Saito 2000; Abdullaev Jafarova et al. 2002).
Recommended dosage: Decoction, 1–3 g.
Precautions: It should be used with caution for pregnant women.
Researches in recent years:
Saffron exhibits antidepressant activity in mild and moderate depression (Shafiee
et al. 2018) and the meta-analysis of nine RCTs presented the effectiveness of saffron
in treatment of mild and moderate depression, with noninferior to SSRI antidepres-
sants (fluoxetine, citalopram) (Tóth et al. 2019). It improved clinical symptoms
assessed with Hamilton Depression Rating Scale (HDRS) almost same effective as
fluoxetine in short-term therapy (Kashani et al. 2017). Even for patients with MDD,
saffron supplementation still ameliorated symptoms of depression (Hausenblas et al.
2013). A standardized extract from saffron affron® showed improvement in depres-
sive symptoms and anxiety assessed with the Revised Child Anxiety and Depression
Scale (RCADS) in youth treated for 8 weeks (Lopresti et al. 2018). Antidepressant
activity of saffron was also demonstrated in mothers with postpartum depression
(Tabeshpour et al. 2017).
Saffron extract and crocin have exhibited antidepressant function through inhi-
bition of monoamine reuptake and binding to N-methyl-D-aspartate (NMDA) recep-
tor as an antagonist and to GABA-α receptor (Georgiadou et al. 2012). Safranal and
crocin prevented increase in plasma corticosterone levels and depressive-like symp-
toms in rats exposed to chronic restraint stress (CRS) (Ghadrdoost et al. 2011;
Ghalandari-Shamami et al. 2019). Saffron has also showed its antioxidant, anti-
inflammatory, anxiolytic, neuroprotective, anticancer, and memory improving func-
tion (Lopresti and Drummond 2014).
Crocin, one of the main glycosylated carotenoids of saffron, has been found to
have numerous pharmacological activities and reported to be associated with
neuroprotective effects. A study showed that crocin-I exerts significant antidepres-
sant effects in a model of chronic corticosterone (CORT)-induced depression. The
antidepressant activity of crocin-I was probably achieved through the suppression of
neuroinflammation (IL-1β) and oxidative stress in the mouse hippocampus. Addi-
tionally, the oral administration of crocin-I at a dose of 40 mg/kg reduced the CORT-
induced accumulation of nicotinamide in the liver of the mice to improve the
TCM Substances in Neuropsychopharmacotherapy 459
1. Pain in the chest and hypochondriac region caused by qi stagnation and blood
stasis, and abdominal pain caused by menstrual block.
2. Painful bi syndrome (cold bi) due to wind and dampness; it is especially effec-
tively to circulate qi in the limbs and relieve painful bi due to wind and dampness.
In addition, it can be applied externally after ground into powder together with
Radix et Rhizoma Rhei (da huang), Radix Angelicae Dahuricae (bai zhi), and Raidx
Trichosanthis (tian hua fen). It can also treat yang syndromes manifested as furuncle
with redness, swelling, feverish sensation and pain caused by abscess, ulcer, and
sores at early stage.
It contains curcumin and volatile oil and has anticoagulative and antimyocardial
ischemia actions. Besides, it can lower blood pressure, reduce blood lipid, suppress
462 Y. Wang and J. Li
bacteria, resist inflammation, protect the liver, and promote gallbladder function.
Moreover, it can prevent early pregnancy and excite the uterus.
Recommended dosage: Decoction, 3–10 g. An appropriate amount is used
externally. It is contraindicated for pregnant women (Teng et al. 2019).
Researches in recent years:
Curcuma longa, also known as turmeric, is a perennial herb, belonging to
Zingiberaceae family with a total of 133 different species worldwide (Bhat et al.
2019). Curcuma longa is commonly used as spice, food preservative, and coloring
material (Kalaycıoğlu et al. 2017), also as a cosmetic agent, dying agent, and
medicinal herb to treat ailments like skin infections, liver and gastrointestinal
disorders, etc. (Yallapu et al. 2012). Curcuma has high concentrations of poly-
phenols and flavonoids (Hossen et al. 2017). Its biosynthesis begins with phenyl-
alanine, a common precursor in the biosynthesis of flavonoids (Kita et al. 2008;
Sandhu et al. 2011). Flavonoids are crucial for the development of effective
therapeutic agents to treat neurodegenerative diseases; along with their antioxi-
dant properties, these compounds exhibited neuroprotective properties through
interactions with different cellular signaling pathways which is followed by
transcriptions, along with translations that mediate cellular functions (Solanki
et al. 2015). Curcumin, the main polyphenol from the Curcuma longa (turmeric
curry), has exhibited a variety of beneficial effects: anti-inflammatory, antitumor,
immunomodulatory, and neuroprotective properties (Venigalla et al. 2015;
Daverey and Agrawal 2016, 2018). Up till recently, the antidepressant effects of
curcumin have been acknowledged, not only owing to its effectiveness in pre-
venting genesis of depression-like behaviors in different animal models (Andrade
2014; Lopresti et al. 2014; Zhang et al. 2014a), but also available clinical trials
supporting the antidepressant effects of curcumin administration in depressed
patients (Ng et al. 2017a).
It has been evidenced that chronic mild stress (CMS) procedure increased serum
interleukin-6 (IL-6) and tumor necrosis factor-alpha (TNF-α) levels, along with
reduction of natural killer cell (NK cell) activity in splenocytes. CMS-treated rats
exhibited higher corticotropin-releasing factor in serum and medulla oblongata, also
an elevated cortisol levels in serum. A reduction in sucrose intake was observed in
the meantime (Xia et al. 2006).
Inflammatory cytokines are considered as significant pro-apoptotic factors which
are involved in the progression of neurological disorders (Tsai 2017). Researches
have demonstrated that activation of pro-inflammatory factors, including interleu-
kin-1β(IL-1β), interleukin-6 (IL-6), and tumor necrosis factor-α(TNF-α), might play
an important role in neuronal damage seen among patients with major depressive
disorder (MDD) (Maes et al. 2012; Rawdin et al. 2013).
The IL-1-type cytokines have been acknowledged as major mediators of neuro-
inflammation and increasing evidence suggested that IL-1βis a key contributor to the
neuronal deterioration in depression (Maes et al. 2012; Kaufmann et al. 2017).
Increased IL-1β was detected in the brain in response to chronic stress (Pan et al.
2014) and inhibiting IL-1β could reverse stress-induced social avoidance in rats
(Ramirez et al. 2015).
TCM Substances in Neuropsychopharmacotherapy 463
Increased serum IL-6 has been observed in depressed patients, which could be
reversed after being treated with antidepressant agents, like fluoxetine (Maes et al.
1995). It has also been reported that the reduction of natural killer (NK) cell activity
could reversed by antidepressant, along with clinical improvement (Jozuka et al.
2003; Frank et al. 1999). Elevated corticotropin-releasing factor (CRF) in cerebro-
spinal fluid and different regulation of adrenocorticotropin hormone (ACTH), along
with abnormal cortisol secretion, have been detected in depressive patients (Plotsky
et al. 1998). Effective antidepressant treatment with fluoxetine could normalize these
hormones (Stout et al. 2002).
Sleep deprivation for 72 h caused weight loss, anxiety like behavior, impaired
locomotor activity. Oxidative damage and weakened antioxidative defense system
were found in sleep deprivation models (Gopalakrishnan et al. 2004; Ramanathan
et al. 2002; Kalonia and Kumar 2007; Kumar and Kalonia 2007). It has been
reported sleep deprivation increased lipid peroxidation, and nitrite level, depleted
reduced glutathione, as well as catalase activity, indicating an oxidative damage in
sleep-deprived animals.
Alzheimer’s disease (AD) is a neural degenerative disorder, which usually leads
to damages to memory, thinking, and behavior (Cole and Frautschy 2006). The main
characteristic pathological changes of AD include aggregations of amyloid-β(Aβ)
and tau proteins (da Costa et al. 2019).
Aβ accumulation leads to the formation of senile plaques, activation of oxidative
damage or even apoptosis to neurons, neuroinflammation, and cognitive deficits
(Zhang et al. 2015b; Huang et al. 2012; Zheng 2005).
Curcuma longa possesses neuroprotective effects against ischemic damage (Jiang
et al. 2007; Dohare et al. 2008; Rathore et al. 2008), Alzheimer’s disease (Frautschy
2001; Yang et al. 2004; Garcia-Alloza et al. 2007), Parkinson’s disease (Mansouri et al.
2012; Jiang et al. 2013), and parathion-induced damage (Canales-Aguirre et al. 2012).
Curcuma longa extracts also exhibited powerful antioxidant, anti-inflammatory,
lipid-reducing, chemo-preventive, immunomodulatory, and sedative properties
(Miquel et al. 2002; Joe et al. 2004).
The ethanolic extracts of Curcuma longa reversed the CMS-induced reduction in
sucrose intake, elevations in serum IL-6, TNF-α, and cortisol levels, along with
reducing corticotropin-releasing factor (CRF) in serum and medulla oblongata, the
decrease in splenic NK cell activity as well (Xia et al. 2006).
Curcuma longa extract contains curcumin, a polyphenolic nonflavonon compound,
which is considered as a pharmacologically active substance (Ganguli et al. 2000).
It can cross the blood-brain barrier and enter brain tissue, regardless of its poor
bioavailability, which was mainly distributed in the hippocampus (Tsai et al. 2011).
Curcumin possesses a strong antidepressant activity in depressive animal models
(Andrade 2014; Hurley et al. 2013; He et al. 2016; Lee and Lee 2018). It acts through
inhibiting the expression of MAO-A and MAO-B enzymes which leads to increase
of norepinephrine, serotonin, and dopamine (Kulkarni and Dhir 2010). It was
reported that curcumin’s antidepressant activity was induced by ERK regulated
increase in the expression of brain-derived neurotrophic factor (BDNF) in the
amygdale (Zhang et al. 2012). Meanwhile, curcumin was also witnessed to promote
464 Y. Wang and J. Li
It is the root and rhizome of Glycyrrhiza uralensis Fisch., Glycyrrhiza inflata Bat., or
Glycyrrhiza glabra L., pertaining to Leguminosae. It is mainly produced in Inner
Mongolia Autonomous Region, Gansu, and Heilongjiang provinces in China. It is
collected in spring and autumn. After the fibrils are removed, it is dried in the sun and
sliced into thick pieces. It has mild fragrance and tastes sweet and unique. It is used
raw or honey-fried. According to the theory of TCM, it is able to supplement the
TCM Substances in Neuropsychopharmacotherapy 465
spleen and boost qi, clear heat and resolve toxins, dispel phlegm and relieve cough,
relax spasms and relieve pain, and harmonize all medicinals.
Clinical applications:
Recommended dosage: Decoction, 2–10 g. It should be used raw for clearing heat
and resolving toxins, and it should be dry-fried with liquid adjuvant for
supplementing the center and relaxing spasms.
Precautions: Radix et Rhizoma Glycyrrhizae (gan cao) is not proper to be used
together with sargassum pallidum (turn.) c.ag./ Sargassum fusiforme (Harv.) Setch.
(hai zao), Radix Euphorbiae Pekinensis (jing da ji), Flos Genkwa (yuan hua), and
Radix Kansui (gan sui). It has the side effects of assisting dampness accumulation
and stagnation, so it is not proper for distension, fullness, and edema due to
dampness exuberance. Long-term application in large dose may induce water and
sodium retention and lead to edema (Teng et al. 2019).
Oral administration for a long term may cause hypertension, edema, lowered
blood potassium, headache, dizzy, and palpitation. When orally administered over
500 mg per day for1 month, glycyrrhizin may produce pseudohyperaldosteronism,
the symptoms will be improved or disappeared after withdrawal or treatment with
spironolactone (Chen et al. 2017).
Researches in recent years:
The pharmacopoeia of the People’s Republic of China contains G. uralensis,
G. glabra, and G. inflata as medicinal materials. Their roots and rhizomes are the
medicinal parts, which mainly contain triterpenoids and flavonoids. To date, more
than 400 compounds have been isolated from licorice (Bao et al. 2019), including
flavonoids, triterpenoid saponins, coumarins, lignans, and polysaccharides
466 Y. Wang and J. Li
(Hosseinzadeh and Nassiri-Asl 2015; Yang et al. 2015; Pastorino et al. 2018). It is
reported that licorice showed significant efficacy in relieving coughing and asthma
(Kuang et al. 2018), along with hepatoprotective (Abdel-Kader et al. 2018), anti-
inflammatory (Wang et al. 2017b), antivirus (Fukuchi et al. 2016), antiulcer (Liu
et al. 2018), and antidiabetes (Bai et al. 2018) properties.
Stresses can cause hyperactivity of the HPA axis, increased level of corticoste-
rone, and also have a negative effect on the neurogenesis, especially neurons from
the hippocampus, which are closely related to the onset of depression (Garcia 2002;
Sapolsky 2000).
Flavonoids are secondary metabolites of the plant. It is demonstrated that the total
flavonoids extract from the cultivated Glycyrrhiza uralensis Fisch. could produce
antidepressive effect on chronic unpredictable stress of depression model rats, along
with its neurogenesis protective effect under large dosage (Fan et al. 2012). Licorice
flavonoids extract is evidenced to have an inhibitory effect on Caspase-3 (a key
enzyme having promoting effect in the apoptosis process), to increase Bcl-xl
(a protein having protective effect on the neuronal apoptosis through inhibiting the
activity of adaptor needed for the activation of Caspase-3) (Cheng et al. 2014), and
the number of the new born BrdU positive progenitor cells at the subgranular zone
(SGZ) of dentate gyrus (DG) region in hippocampus, through which licorice flavo-
noids displayed antidepressive effect (Fan et al. 2012).
Researches found that the effective antidepressive components from Glycyrrhiza
also include liquirtin and isoliquirtin (Zhao et al. 2008). Liquirtin is an important
monomer active component among Glycyrrhiza flavonoids, also reversed the
decrease of sucrose consumption among CUMS depressive rats (Su et al. 2011;
Zhao et al. 2006a). Liquirtin and isoliquirtin exhibited antidepression effect, through
increasing the concentration of neurotransmitter 5-HT, norepinephrine in the hippo-
campus, hypothalamus and cortex, inhibiting the metabolism of 5-HT, upregulating
the level of BDNF, enhancing the activity of SOD, removing the free radicals,
preventing lipid peroxidation, reducing the production of MDA (malondialdehyde),
and alleviating the oxidative damage of the brain caused by chronic stress (Ma 2009;
Zhao et al. 2006b, c, 2008).
Processing with honey could increase the content of liquirtin and isoliquirtin
(Zhao et al. 2008). The higher the temperature was, the higher level of liquirtin,
isoliquiritigenin, isoliquirtin. The increase of liquirtin made Glycyrrhiza more suit-
able for treating depression (Zhang et al. 2009).
Glycyrrhizic acid, a rich content triterpenoid saponin, is one of the most important
pharmacologically active components in it. Glycyrrhizic acid have significant effects
similar to adrenal cortical hormone and can be used in clinical for anti-inflammatory,
antiaging, decompression, enhancing body immunity, improving physiological func-
tion, and restraining cancer cells growth, showing really curative effect (Wang et al.
2013). Combining with glycyrrhizic acid, as an anti-inflammatory method, could
provide a more efficient treatment for depression and accompanying cognitive
impairment. For depressive patients with C-reactive protein (CRP) higher than
3 mg/L, glycyrrhizic acid could enhance the therapeutic effects of traditional anti-
depressant agents greatly (Cao 2019).
TCM Substances in Neuropsychopharmacotherapy 467
including wine, Fructus Amomi (sha ren), Pericarpium Citri Reticulatae (chen pi)
until both the internal and external become black, oily, and moistening and the
texture becomes soft, sticky, and greasy. Thus Radix Rehmanniae Praeparata is
obtained. It has a mild smell and tastes sweet. It is sliced or dry-fried until charred
for application. Unlike the unprepared Radox Rehmanniae, the prepared, according
to the theory of TCM, it is believed to supplement blood and enrich yin, boost
essence and replenish marrow (Teng et al. 2019).
Clinical applications according to TCM:
Ajugol, gentistic acid (Liu 2013), and oleanolic acid (Anouar el et al. 2015) have the
ability of antioxidation. Versulin (Zhang et al. 2015b) and oleanolic acid (Zhang
et al. 2014b) have the capacity of inhibiting tumor growth. Oleanolic acid (Mengoni
et al. 2002) has been demonstrated to exhibit weak anti-HIV activity.
Aside from different pharmacological effects Rehmannia glutinosa polysaccha-
rides of various constituents from Radix rehmanniae, it is still seen as a herb with
tonic effects these days. It has been acknowledged that Radix rehmanniae bears with
effect of nourishing yin and is usually applied in formulae that nourishes yin to treat
patient with symptoms of yin deficiency, which partially explains the lack of clinical
trials using Radix rehmanniae for patients with psychiatric disorders, since these
disorders are usually seen as excessive syndrome or Yu syndrome (obstruction
syndrome), rather than deficient ones. Hopefully, researchers or clinicians will
consider the administration of Radix rehmanniae as a potential method for patients
with psychiatric disorders.
Bulbus Lilii is mainly produced in Anhui, Zhejiang, and Jiangsu provinces in China.
It is collected in autumn and washed clean. After its scale leaves are peeled off. It is
slightly scaled in boiling water and then dried. It has mild smell and tastes mildly
bitter. It is used raw or honey-fried. In TCM, it is able to enrich yin and moisten the
lung; clear heart heat and calm the mind (Teng et al. 2019).
Clinical applications in TCM
1. Syndrome of lung yin deficiency: It is used for yin deficiency and lung dryness
with heat manifested as dry cough with less sputum, hemoptysis, dry throat, and
hoarse voice.
2. Syndrome of heart yin deficiency: (1) To treat heart yin deficiency manifested as
vexation and agitation, insomnia, and profuse dreaming. (2) To treat lily disease
due to deficiency of heart yin and lung yin and malnutrition of the heart spirit,
which is manifested as blurred mind, uncontrolled emotions, bitter taste in the
mouth, dark urine, faint, and rapid pulse. Additionally, it is able to enrich stomach
yin and clear stomach heat, so it is used to treat stomachache due to stomach yin
deficiency with heat.
DA, 5-HT in the brain of depressive rat model, and improving the dysfunction of
monoamine neurotransmitters; 2. decreasing the cortisol, adrenocorticotrophic hor-
mone in blood, reducing the expression of corticotrophin-releasing factor mRNA
(CRFmRNA) in hypothalamus; and 3. enhancing the expression of glucocorticoid
receptor mRNA (GRmRNA) and mineralcorticoid receptor mRNA (MRmRNA) in
hippocampus; attenuating the hyperactivity of the hypothalamic-pituitary-adrenal axis
(Guo et al. 2009; Guo and Li 2009).
It has also been (Huang et al. 2011) detected the changes of substance P (SP),
vasoactive intestinal peptide (VIP), and gastrins (Gas) in blood, stomach, and intestine
of depressive rats. The application of total saponins of Bulbus Lilii could increase
plasma VIP, serum, and colon tissue Gas and SP, as it alleviated the depressive
symptoms in the meantime, indicating that the total saponins of Bulbus Lilii could
improve depressive symptoms through regulation of the brain-gut peptides.
Like Radix rehmanniae, Bulbus Lilii is commonly applied in formulae nourishes
yin. In some parts of China, people even use it as a food supplement. But still,
nothing conclusive, concerning the comprehensive pharmacological effects of
Bulbus Lilii and its definite pharmacological effects among patients, has been
extracted from clinical trials or meta-analyses yet. The detailed responses from
patients with psychiatric disorders to Bulbus Lilii will require more evidence from
clinical trials with large sample size, longer duration, and follow-ups.
St. John’s wort (Hypericum perforatum L.) is named mainly due to the fact that it
flowers at the time of the summer solstice on or around St. John’s day on 24 June.
Having been administered as a remedy by a Roman military doctor as early as the
first century AD (Pöldinger 2000), St. John’s wort is stated to possess sedative and
astringent properties and has been used traditionally for the treatment of excitability,
neuralgia, sciatica, menopausal neurosis, anxiety, depression, and as a nerve tonic,
and in preparations for the treatment of wounds. More recently, in a number of
clinical double-blind trials against placebo and other antidepressants, the whole
extract of St.-John’s wort, for example, as in Jarsin coated tablets, has proved to
be effective for mild and moderate depression. Meta-analyses of reported clinical
trials of herbs in depression, anxiety, and insomnia has presented the evidence-based
efficacy of herbs: St John’s wort and saffron for unipolar depression (Sarris et al.
2011; Apaydin et al. 2016). A large meta-analysis showed that the efficacy and
safety of St John’s wort were comparable to SSRIs in mild and moderate depressive
disorder (Ng et al. 2017b) as well. Hypericum perforatum extract, including WS ®
5700, LI160, PM23, and ZE117, has been applied in clinical trials for treatment of
depression (Zirak et al. 2019). WS 5570 has been demonstrated to be effective as
paroxetine in mild to moderate MDD (Seifritz et al. 2016).
Recent researches:
The major constituents are napthodianthrones (hypericin, pseudohypericin),
phloroglucinol derivatives (hypericin, adehyperforin), flavonoids, biflaonoids,
472 Y. Wang and J. Li
discussed at that time that cyclosporine concentrations could indicate dosage adjust-
ment with therapeutic drug monitoring; in that case the interaction would be con-
sidered clinically significant.
A plethora of clinical interaction studies and case reports were published in causal
association St John’s wort’s extracts with high-hyperforin content (Chrubasik-
Hausmann et al. 2019; Soleymani et al. 2017; ESCOP 2018). As concluded by the
EMA/HMPC, hyperforin is mainly responsible for pharmacokinetic interactions
with other drug substances, which are metabolised by certain CYP450 isoenzymes
and transported by ABCB1 (P-Glycoprotein, P-gp). Hence, with regard to pharma-
cokinetic interactions, St John’s wort’s products have to consider the daily hyper-
forin dose, which led to the different low-hyperforin preparations (1 mg/d) and
high-hyperforin ones (>1 mg/d) (ESCOP 2018). In a recently finalized risk assess-
ment of the EMA, it was stated that adequate studies with extracts with low-
hyperforin content are available which could justify exemptions with regard to
contraindications.
Pharmacokinetic interactions with St John’s wort’s preparations correlate directly
with the daily dose of hyperforin (Mueller et al. 2006). The induction of PXR-related
metabolic enzymes and transporters cannot be excluded at daily dosages >1 mg
hyperforin. To avoid pharmacokinetic interactions and to contribute to St John’s
Wort’s product safety, low-hyperforin SJW extracts should be recommended for the
therapeutic use. At daily dosages of maximum 1 mg hyperforin, no clinically
relevant pharmacokinetic interactions are to be expected (ESCOP 2018; EMA/
HMPC 2018; Zahner et al. 2019).
Precautions: Cautions should be taken when applied for pregnant and breast-
feeding women.
St. John’s wort (Hypericum perforatum) contains hypericin and hyperforin as
well as flavonoids, such as quercetin. It has been recommended by midwives for
postpartum depression (Allaire et al. 2000; Dennehy et al. 2010). Both hypericin and
hyperforin are poorly excreted into breast milk; no other components have been
measured in milk. One study found a slightly increased frequency of colic, drows-
iness, and lethargy among breastfed infants whose mothers were taking St. John’s
wort, but none of the effects were severe or required treatment. Most reports were
related to breastfeeding older infants, rather than during the first 2 months postpar-
tum or preterm infants when they are more susceptible to adverse reactions.
Conflicting information still exists on whether St. John’s wort can reduce prolactin
levels or the maternal milk supply. Since there is little publications concerning St.
John’s wort application during breastfeeding, an alternate drug may be preferred,
especially while nursing a newborn or preterm infant (Clauson et al. 2008).
Formulae of TCM
needs for different syndromes and diseases and is guided by certain combination
principles of TCM.
The reasons that Chinese medicinals are usually combined are listed as follows:
The efficacy of one medicinal is too limited to treat complex conditions and is not
powerful enough to deal with severe conditions; some medicinals are toxic, so it is
not safe to use singly;
Individual medicinal may have different efficacies, which may do harm to human
body if it is not necessary for the condition.
Hence, the combination of Chinese medicinals, based on the need of conditions and
requirement of medicinals and guided by certain combination rules, can enhance the
medicinal power to promote clinical efficacy. On the other hand, it can relieve or
eliminate the toxicity of medicinals to restrict the harmful effects induced by herbs, but
unnecessary for the conditions and to make medicinals safer (Chen et al. 2017). Some
experts consider the different herbs or medicinals from a formula of TCM as pieces of
a chess; with the right coordination and strategy, these pieces could overwhelm the
counterpart (the diseases). Yet in clinical practice, the strategies applied are far more
complicated than that in the metaphor, considering the fact that different origins,
processing methods, or different dosage of even one herb/medicinal could change
the entire therapeutic effect of a formula, not to mention different perspectives about
the original dosage and production area of genuine medicinals held by experts
nowadays, since many formulae have been documented for thousands of years. As
such, it remains a challenge to elaborate the underlying mechanism of a formula of
TCM, even with the help from the most advanced and sophisticated equipment.
The following formulae were carefully selected, with the acknowledgment of
bearing a therapeutic effect for psychiatric disorders. We hope the following content
may shed light on understanding the formation and the application of formulae of
TCM in psychiatry.
the depression-like responses in the chronic mild stress exposed mice and remained
effective at 5 and 6 days postdrug administration, respectively (Tang et al. 2015).
Studies available have demonstrated that the routine dose of Yueju has significant
antidepressant efficacy and effectively attenuated behavioral deficits in chronic
learned helpless animals. Restoration of neural plasticity via CREB signaling is
crucially involved in the antidepressant effect of Yueju, whereas the inhibition of
NMDA signaling is required and part of the mechanism of antidepressant efficacy of
Yueju (Zou et al. 2017).
Source: Auxiliary
Verse On Drugs And Methods For Zangfu Organs (Fu xing jue wu zang yong yao
fa yao)
Ingredients:
Haematitum (dai zhe shi)
Flos Inulae (xuan fu hua)
Folium Phyllostachydis Henonis (zhu ye)
Semen Sojae Preparatum (dan dou chi)
Preparation: Heat the Haematitum till it turns red, then spray it with vinegar.
Grind the Haematitum after heating and spaying for 3 times. 18 g of Haematitum,
Flos Inulae, Folium Phyllostachydis Henonis, respectively, along with 9 g of Semen
Sojae Preparatum are decocted together.
Indications: palpitation due to blood and qi deficiency, irritation, sweating,
periodic low mood, and irregular pulse.
Mechanisms:
Xiaonbuxin-Tang is a decoction with nearly 1000-year history. These years, up to
22 compounds have been isolated from the total flavonoids (XBXT-2) isolated from
the extract of Xiaobuxin-Tang(XBXT) including 21 flavone compounds: japonicins A
and B, onpordin, 30 -O-methylorobol, glycitein, nepetin, patuletin, genistein, luteolin,
daidzein, quercetin, apigenin, isoquercitrin, genistin, nepitrin, quercimeritrin, daidzin,
patulitrin, quercetagitrin, 3-glucosyl isorhamnetin, isoorientin, and an organic acid
protocatechuic acid. Additionally, the rough HPLC fingerprint of the extract demon-
strated that the major constituents of XBXT-2 were flavones, flavonols, isoflavones,
and their glycosides. HPLC analysis also revealed that the flavones and flavonols were
primary derived from Flos Inulae, as well as the isoflavones were derived from Semen
Sojae Preparatum, and the rest of flavones came from Folium Phyllostachydis Henonis
(An et al. 2008a).
Studies indicated that acute treatment with XBXT-2 produced serotonergic, but
not noradrenergic activation. In addition, chronic XBXT-2 (25, 50 mg/kg, p.o.,
28 days) treatments significantly reversed the depressive-like behaviors in chron-
ically mildly stressed (CMS) rats, including the reduced sucrose preference,
deficient locomotor activity, and prolonged latency to novelty-suppressed feeding
(An et al. 2008b).
TCM Substances in Neuropsychopharmacotherapy 477
In animal models, researchers found that the stress hormones including cortico-
sterone and ACTH levels in rats were significantly increased by chronic stress
exposure, as well as the decreased hippocampal glucocorticoid receptors (GRs)
expression, which together manifested the hyperactivity and impaired feedback
inhibition of HPA axis in chronically stressed rats. Chronic administration of
XBXT-2 (25, 50 mg kg1, p.o., 28 days, the effective doses for behavioral
responses) significantly decreased serum corticosterone level and its upstream stress
hormone adrenocorticotropic hormone (ACTH) level in chronically stressed rats.
Furthermore, western blotting result demonstrated XBXT-2 treatment ameliorated
stress-induced decrease of GRs expression in hippocampus, an important target
involved in the hyperactivity of HPA axis (An et al. 2011).
The pretreatment with lipopolysaccharide (LPS) significantly increased the
immobility time in TST and FST in mice, as well as the brain levels of IL-1β and
TNF-α. XBXT-2 (25, 50, and 100 mg/kg, p.o.) administration decreased the duration
of immobility in TST and FST and normalized the cytokines levels. Findings also
demonstrated that XBXT-2 at dose of 100 mg/kg was more efficacious than that of
200 mg/kg, which was also in line with our behavioral results which showed that
XBXT-2 exhibited an inverse U-shaped dose-response curve in acute and subchronic
models (An et al. 2015).
The increase of neurogenesis, as well as expression of BDNF and pCREB in
hippocampus, may be one of the molecular and cellular mechanisms underlying the
antidepressant action of XBXT-2. Also, immunohistochemistry results showed that
concomitant administration of XBXT-2 (25, 50 mg/kg, p.o., 28 days, the effective doses
for behavioral responses) significantly increased hippocampal neurogenesis in chroni-
cally stressed rats. Furthermore, XBXT-2 treatment reserved stress-induced decrease of
hippocampal BDNF and pCREB (Ser133) expression, two important factors which
were closely related to hippocampal neurogenesis. The increase in hippocampal BDNF
and pCREB expression induced by XBXT-2 treatment might be the early potent
stimulation for the increase of hippocampal neurogenesis (An et al. 2008a).
Xiao Buxin Tang Decoction was designed to treat palpitation, irritation, sweating,
irregular pulse, and accompanied periodic low mood, most of which are symptoms
related to heart diseases. It is not surprising that studies and clinical trials concerning
Xiao Buxin Tang Decoction have been focused on symptoms, like palpitation,
arrhythmia, and the underlying mechanisms. Nevertheless, it is not uncommon to
see patients with all the aforementioned symptoms and a normal EKG, which may
indicate a possible diagnosis of somatoform disorder. It is suggested that future
clinical research should put more emphasis on the application of Xiao Buxin Tang
Decoction on patients with psychiatric disorders.
been sharply increasing, mainly resulting from factors, such as excessive stress of
work and life styles, social failure stress (Dunstan et al. 2013).
Researchers has proved that the SCL-90 scores of PSHS persons could be
significantly decreased after decoction of Bulbus Lilii and Radix Rehmanniae
administrated for 4 weeks, indicating that the decoction was of a significant thera-
peutic effect on the improvement of PSHS (Meng et al. 2018).
It has been noticed that constituents of different herbs and medicinals in a
formula change because of the intermolecular chemical reactions that form novel
molecules or nanostructures, which are mainly governed by electrostatic and
hydrophobic interactions during the decoction procedure (Li et al. 2019). Bulbus
Lilii and Radix Rehmanniae Tang Decoction, therefore, are not appropriate to be
comprehended or applied as merely combination of Bulbus Lilii and Radix
Rehmanniae. Clinical trials with well-designed methodology, large sample size,
and enough follow-ups are required to determine the effectiveness of Bulbus Lilii
and Radix Rehmanniae Tang Decoction for psychiatric disorders.
Conclusion
This chapter makes a brief introduction and description of a few herbs, medicinals, and
formulae of TCM and their applications in psychiatry. This chapter not only includes the
conventional understanding of herbs and formulae, but also provides information of the
latest researches accordingly. It has been the complexity and diversity of herbs, medic-
inals, and formulae that make it remains a huge challenge to depict the underlying
mechanism of herbs and formulae of TCM in psychiatry, some of which stays myste-
rious, let alone other well-known, and less-studied methods of TCM that have been
applied to alleviates symptoms of psychiatric disorders. For instance, acupuncture,
moxibustion, and cupping have been suggested to ease the somatic symptoms of
depression, hysteria, etc. Additional large scale randomized controlled clinical trials
and sophisticated pharmacology studies are still need to be performed at present.
Opportunities permitting, we do hope that more knowledge concerning theories and
clinical methods of TCM could be shared, understood, and accepted in shortly coming
future.
Cross-References
References
Abdel-Kader MS, Abulhamd AT, Hamad AM, et al. Evaluation of the hepatoprotective effect of
combination between Hinokiflavone and Glycyrrhizin against CCl4 induced toxicity in rats.
Saudi Pharm J. 2018;26:496–503.
Abdullaev Jafarova F, Caballero-Ortega H, Riveron-Negrete L, et al. In vitro evaluation of the
chemopreventive potential of saffron. Rev Investig Clin. 2002;54:430–6.
Abe K, Saito H. Effects of saffron extract and its constituent crocin on learning behaviour and long-
term potentiation. Phytother Res. 2000;14:149–52.
Aguiar AS Jr, Castro AA, Moreira EL, et al. Short bouts of mild-intensity physical exercise improve
spatial learning and memory in aging rats: involvement of hippocampal plasticity via AKT,
CREB and BDNF signaling. Mech Ageing Dev. 2011;132(11–12):560–7. https://doi.org/10.
1016/j.mad.2011.09.005.
Allaire AD, Moos MK, Wells SR. Complementary and alternative medicine in pregnancy: a survey
of North Carolina certified nurse-midwives. Obstet Gynecol. 2000;95:19–23.
An L, Zhang YZ, Yu NJ, et al. The total flavonoids extracted from Xiaobuxin-Tang up-regulate the
decreased hippocampal neurogenesis and neurotrophic molecules expression in chronically
stressed rats. Prog Neuro-Psychopharmacol Biol Psychiatry. 2008a;32(6):1484–90. https://doi.
org/10.1016/j.pnpbp.2008.05.005.
An L, Zhang YZ, Yu NJ, et al. Role for serotonin in the antidepressant-like effect of a flavonoid
extract of Xiaobuxin-Tang. Pharmacol Biochem Behav. 2008b;89(4):572–80. https://doi.org/10.
1016/j.pbb.2008.02.014.
An L, Zhang YZ, Liu XM, et al. Total flavonoids extracted from Xiaobuxin-Tang on the hyperac-
tivity of hypothalamic-pituitary-adrenal axis in chronically stressed rats. Evid Based Comple-
ment Alternat Med. 2011:367619. https://doi.org/10.1093/ecam/nep218.
An L, Li J, Yu ST, et al. Effects of the total flavonoid extract of Xiaobuxin-Tang on depression-like
behavior induced by lipopolysaccharide and proinflammatory cytokine levels in mice. J
Ethnopharmacol. 2015;163:83–7. https://doi.org/10.1016/j.jep.2015.01.022.
Andrade C. A critical examination of studies on curcumin for depression. J Clin Psychiatry. 2014;75
(10):e1110–2. https://doi.org/10.4088/JCP.14f09489.
Anouar el H, Zakaria NS, Alsalme A, Shah SA. α-Glucosidase activity of oleanolic acid and its
oxidative metabolites: DFT and Docking studies. Mini-Rev Med Chem. 2015;15(14):1148–58.
https://doi.org/10.2174/1389557515666150724154044.
Apaydin EA, Maher AR, Shanman R, et al. A systematic review of St. John’s wort for major
depressive disorder. Syst Rev. 2016;5(1):148. Published 2016 Sep 2. https://doi.org/10.1186/
s13643-016-0325-2.
Ates-Alagoz Z, Adejare A. NMDA receptor antagonists for treatment of depression. Pharmaceuti-
cals (Basel). 2013;6(4):480–499. Published 2013 Apr 3. https://doi.org/10.3390/ph6040480
Attele AS, Wu JA, Yuan CS. Ginseng pharmacology: multiple constituents and multiple actions.
Biochem Pharmacol. 1999;58(11):1685–93. https://doi.org/10.1016/s0006-2952(99)00212-9.
Bai S, Zhang X, Chen Z, et al. Insight into the metabolic mechanism of Diterpene Ginkgolides on
antidepressant effects for attenuating behavioural deficits compared with venlafaxine. Sci Rep.
2017;7(1):9591. https://doi.org/10.1038/s41598-017-10391-1.
Bai M, Yao GD, Ren Q, et al. Triterpenoid saponins and flavonoids from licorice residues with anti-
inflammatory activity. Ind Crop Prod. 2018;125:50–8.
Bao Q, Shen X, Qian L, Gong C, Nie M, Dong Y. Anti-diabetic activities of catalpol in db/db
mice. Korean J Physiol Pharmacol. 2016;20(2):153–60. https://doi.org/10.4196/kjpp.2016.
20.2.153.
Bao F, Bai HY, Wu ZR, Yang ZG. Phenolic compounds from cultivated Glycyrrhiza uralensis and
their PD-1/PD-L1 inhibitory activities [Published online ahead of print, 2019 Mar 25]. Nat Prod
Res. 2019;1–8. https://doi.org/10.1080/14786419.2019.1586698
Bhat A, Mahalakshmi AM, Ray B, et al. Benefits of curcumin in brain disorders. Biofactors. 2019;
https://doi.org/10.1002/biof.1533.
TCM Substances in Neuropsychopharmacotherapy 481
Bishnoi M, Chopra K, Rongzhu L, et al. Protective effect of curcumin and its combination with
piperine (bioavailability enhancer) against haloperidol-associated neurotoxicity: cellular and
neurochemical evidence. Neurotox Res. 2011;20(3):215–25. https://doi.org/10.1007/s12640-
010-9229-4.
Boonlert W, Benya-Aphikul H, Umka Welbat J, Rodsiri R. Ginseng extract G115 attenuates
ethanol-induced depression in mice by increasing brain BDNF levels. Nutrients. 2017;9
(9):931. Published 2017 Aug 24. https://doi.org/10.3390/nu9090931
Caccamo A, Maldonado MA, Bokov AF, et al. CBP gene transfer increases BDNF levels and
ameliorates learning and memory deficits in a mouse model of Alzheimer’s disease. Proc Natl
Acad Sci U S A. 2010;107(52):22687–92. https://doi.org/10.1073/pnas.1012851108.
Canales-Aguirre AA, Gomez-Pinedo UA, Luquin S, et al. Curcumin protects against the oxidative
damage induced by the pesticide parathion in the hippocampus of the rat brain. Nutr Neurosci.
2012;15(2):62–9. https://doi.org/10.1179/1476830511y.0000000034.
Cao ZY. The clinical effect of glycyrrhizic acid based on the inflammatory mechanism of depres-
sion. Shanghai: Naval Medical University; 2019.
Cao G, Su P, Zhang S, et al. Ginsenoside Re reduces Aβ production by activating PPARγ to inhibit
BACE1 in N2a/APP695 cells. Eur J Pharmacol. 2016;793:101–8.
Castrén E, Rantamäki T. The role of BDNF and its receptors in depression and antidepressant drug
action: reactivation of developmental plasticity. Dev Neurobiol. 2010;70(5):289–97. https://doi.
org/10.1002/dneu.20758.
Chen XW, Serag ES, Sneed KB, et al. Clinical herbal interactions with conventional drugs: from
molecules to maladies. Curr Med Chem. 2011;18(31):4836–50. https://doi.org/10.2174/
092986711797535317.
Chen ML, Gao J, He XR, Chen Q. Involvement of the cerebral monoamine neurotransmitters
system in antidepressant-like effects of a chinese herbal decoction, baihe dihuang tang, in mice
model. Evid Based Complement Alternat Med. 2012;2012:419257. https://doi.org/10.1155/
2012/419257.
Chen CX, Wang XB, Nie H, et al. Pharmacology of Chinese Meteria Medica. Beijing: China Press
of Traditional Chinese Medicine; 2017.
Chen J, Li M, Chen L, et al. Effects of processing method on the pharmacokinetics and tissue
distribution of orally administered ginseng. J Ginseng Res. 2018;42(1):27–34. https://doi.org/
10.1016/j.jgr.2016.12.008.
Chen P, Hei M, Kong L, et al. One water-soluble polysaccharide from Ginkgo biloba leaves with
antidepressant activities via modulation of the gut microbiome. Food Funct. 2019;10(12):8161–
71. https://doi.org/10.1039/c9fo01178a.
Cheng RF, Hua B, Jing J, Xue MQ, Zhao WH, Fan ZZ, et al. Modulation of the apoptotic protein
expression in hippocampus is associated with the antidepressant effects of licorice
flavonoids from Glycyrrhiza uralensis in rats. Pharmacol Clin Chinese Materia Medica.
2014;30(2):69–72.
Chi X, Wang S, Baloch Z, et al. Research progress on classical traditional Chinese medicine formula
Lily Bulb and Rehmannia Decoction in the treatment of depression. Biomed Pharmacother.
2019;112:108616. https://doi.org/10.1016/j.biopha.2019.108616.
Choi S, Lee JH, Oh S, Rhim H, Lee SM, Nah SY. Effects of ginsenoside Rg2 on the 5-HT3A
receptor-mediated ion current in Xenopus oocytes. Mol Cell. 2003;15(1):108–13.
Choi YJ, Yoon Y, Choi HS, et al. Effects of medicinal herb extracts and their components on
steatogenic hepatotoxicity in Sk-hep1 cells. Toxicol Res. 2011;27(4):211–6. https://doi.org/10.
5487/TR.2011.27.4.211.
Chrubasik-Hausmann S, Vlachojannis J, McLachlan AJ. Understanding drug interactions with St
John’s wort (Hypericum perforatum L.): impact of hyperforin content. J Pharm Pharmacol.
2019;71(1):129–38. https://doi.org/10.1111/jphp.12858.
Clauson KA, Santamarina ML, Rutledge JC. Clinically relevant safety issues associated with St.
John’s wort product labels. BMC Complement Altern Med. 2008;8:42. Published 2008 Jul 17.
https://doi.org/10.1186/1472-6882-8-42
482 Y. Wang and J. Li
Cole GM, Frautschy SA. Docosahexaenoic acid protects from amyloid and dendritic pathology in
an Alzheimer’s disease mouse model. Nutr Health. 2006;18(3):249–59. https://doi.org/10.1177/
026010600601800307.
Conboy L, Foley AG, O’Boyle NM, et al. Curcumin-induced degradation of PKC delta is
associated with enhanced dentate NCAM PSA expression and spatial learning in adult and
aged Wistar rats. Biochem Pharmacol. 2009;77(7):1254–65. https://doi.org/10.1016/j.bcp.
2008.12.011.
Czéh B, Simon M. Neuroplaszticitás és depresszió [Neuroplasticity and depression]. Psychiatr
Hung. 2005;20(1):4–17.
da Costa IM, Freire MAM, de Paiva Cavalcanti JRL, et al. Supplementation with Curcuma longa
reverses neurotoxic and behavioral damage in models of Alzheimer’s disease: a systematic
review. Curr Neuropharmacol. 2019;17(5):406–21. https://doi.org/10.2174/
0929867325666180117112610.
Dai MM, Wu H, Li H, et al. Effects and mechanisms of Geniposide on rats with adjuvant arthritis.
Int Immunopharmacol. 2014;20(1):46–53. https://doi.org/10.1016/j.intimp.2014.02.021.
Dai CX, Hu CC, Shang YS, et al. Role of Ginkgo biloba extract as an adjunctive treatment of elderly
patients with depression and on the expression of serum S100B. Medicine. 2018;97(39):e12421.
https://doi.org/10.1097/MD.0000000000012421.
Daverey A, Agrawal SK. Curcumin alleviates oxidative stress and mitochondrial dysfunction in
astrocytes. Neuroscience. 2016;333:92–103. https://doi.org/10.1016/j.neuroscience.2016.07.012.
Daverey A, Agrawal SK. Pre and post treatment with curcumin and resveratrol protects astrocytes after
oxidative stress. Brain Res. 2018;1692:45–55. https://doi.org/10.1016/j.brainres.2018.05.001.
de Sousa CN, Meneses LN, Vasconcelos GS, et al. Reversal of corticosterone-induced BDNF
alterations by the natural antioxidant alpha-lipoic acid alone and combined with desvenlafaxine:
emphasis on the neurotrophic hypothesis of depression. Psychiatry Res. 2015;230(2):211–9.
https://doi.org/10.1016/j.psychres.2015.08.042.
Dennehy C, Tsourounis C, Bui L, King TL. The use of herbs by California midwives. J Obstet Gynecol
Neonatal Nurs. 2010;39(6):684–93. https://doi.org/10.1111/j.1552-6909.2010.01193.x.
Diamond BJ, Bailey MR. Ginkgo biloba: indications, mechanisms, and safety. Psychiatr Clin North
Am. 2013;36(1):73–83. https://doi.org/10.1016/j.psc.2012.12.006.
Dohare P, Garg P, Sharma U, et al. Neuroprotective efficacy and therapeutic window of curcuma oil:
in rat embolic stroke model. BMC Complement Altern Med. 2008;8:55. https://doi.org/10.1186/
1472-6882-8-55. PMID: 18826584; PMCID: PMC2573880
Du Y, Fu M, Wang YT, Dong Z. Neuroprotective effects of ginsenoside Rf on amyloid-β-induced
neurotoxicity in vitro and in vivo. J Alzheimers Dis. 2018;64(1):309–22. https://doi.org/10.
3233/JAD-180251.
Dunstan RH, Sparkes DL, Roberts TK, Crompton MJ, Gottfries J, Dascombe BJ. Development of a
complex amino acid supplement, Fatigue Reviva™, for oral ingestion: initial evaluations of
product concept and impact on symptoms of sub-health in a group of males. Nutr J.
2013;12:115. Published 2013 Aug 8. https://doi.org/10.1186/1475-2891-12-115
EMA/HMPC. Assessment report on Hypericum perforatum L., herba Draft EMA/HMPC/244315/
2016. 2018. Available from: https://www.ema.europa.eu/en/documents/herbal-report/draft-assess
ment-report-hypericum-perforatum-l-herba-revision-1_en.pdf
Erickson KI, Prakash RS, Voss MW, et al. Brain-derived neurotrophic factor is associated with age-
related decline in hippocampal volume. J Neurosci. 2010;30(15):5368–75. https://doi.org/10.
1523/JNEUROSCI.6251-09.2010.
ESCOP. Hyperici herba – St. John’s Wort. European Scientific Cooperative on Phytotherapy
(ESCOP). 2018;2018:1–87. Available from: https://escop.com/downloads/hypericum-2018/
Fan ZZ, Zhao WH, Guo J, et al. Antidepressant activities of flavonoids from Glycyrrhiza uralensis
and its neurogenesis protective effect in rats. Acta Pharm Sin. 2012;47(12):1612–7.
Fan C, Song Q, Wang P, Li Y, Yang M, Yu SY. Neuroprotective effects of ginsenoside-Rg1 against
depression-like behaviors via suppressing glial activation, synaptic deficits, and neuronal
apoptosis in rats. Front Immunol. 2018;9:2889. https://doi.org/10.3389/fimmu.2018.02889.
TCM Substances in Neuropsychopharmacotherapy 483
Han BH, Park MH, Kim DH, Hong SK. Studies on metabolic fates of ginsenosides. Korean
Biochem J. 1986;19:213–8.
Hao K, Gong P, Sun SQ, et al. Beneficial estrogen-like effects of ginsenoside Rb1, an active
component of Panax ginseng, on neural 5-HT disposition and behavioral tasks in ovariectomized
mice. Eur J Pharmacol. 2011;659(1):15–25. https://doi.org/10.1016/j.ejphar.2011.03.005.
Hashimoto K, Shimizu E, Iyo M. Critical role of brain-derived neurotrophic factor in mood
disorders. Brain Res Brain Res Rev. 2004;45(2):104–14. https://doi.org/10.1016/j.brainresrev.
2004.02.003.
Hausenblas HA, Saha D, Dubyak PJ, Anton SD. Saffron (Crocus sativus L.) and major depressive
disorder: a meta-analysis of randomized clinical trials. J Integr Med. 2013;11(6):377–83. https://
doi.org/10.3736/jintegrmed2013056.
He ML, Cheng XW, Chen JK, et al. Simultaneous determination of five major biological active
ingredients in different parts of Gardenia jasminoides fruits by HPLC with diode-array detec-
tion. Chromatographia. 2006;64:713–7.
He X, Zhu Y, Wang M, et al. Antidepressant effects of curcumin and HU-211 coencapsulated solid
lipid nanoparticles against corticosterone-induced cellular and animal models of major depression.
Int J Nanomed. 2016;11:4975–4990. Published 2016 Oct 3. https://doi.org/10.2147/IJN.S109088
Hosseinzadeh H, Nassiri-Asl M. Pharmacological effects of Glycyrrhiza spp. and its bioactive constit-
uents: update and review. Phytother Res. 2015;29(12):1868–86. https://doi.org/10.1002/ptr.5487.
Hossen MS, Tanvir EM, Prince MB, et al. Protective mechanism of turmeric (Curcuma longa) on
carbofuran-induced hematological and hepatic toxicities in a rat model. Pharm Biol. 2017;55
(1):1937–45. https://doi.org/10.1080/13880209.2017.1345951.
Hu Q, Shen P, Bai S, et al. Metabolite-related antidepressant action of diterpene ginkgolides in the
prefrontal cortex. Neuropsychiatr Dis Treat. 2018;14:999–1011. https://doi.org/10.2147/NDT.
S161351.
Hu Y, Liu X, Xia Q, et al. Comparative anti-arthritic investigation of iridoid glycosides and crocetin
derivatives from Gardenia jasminoides Ellis in Freund’s complete adjuvant-induced arthritis in
rats. Phytomedicine. 2019;53:223–33. https://doi.org/10.1016/j.phymed.2018.07.005.
Huang JJ, Li WM, Gao Y. Studies on quality standard and antidepression effect of the effective parts
of Bulbus Lilii. Guangzhou: Guangzhou University of Chinese Med; 2011.
Huang HC, Xu K, Jiang ZF. Curcumin-mediated neuroprotection against amyloid-β-induced
mitochondrial dysfunction involves the inhibition of GSK-3β. J Alzheimers Dis. 2012;32
(4):981–96. https://doi.org/10.3233/JAD-2012-120688.
Huang Y, Jiang C, Hu Y, et al. Immunoenhancement effect of rehmannia glutinosa polysaccharide
on lymphocyte proliferation and dendritic cell. Carbohydr Polym. 2013;96(2):516–21. https://
doi.org/10.1016/j.carbpol.2013.04.018.
Huang Y, Wu C, Liu Z, et al. Optimization on preparation conditions of Rehmannia glutinosa
polysaccharide liposome and its immunological activity. Carbohydr Polym. 2014;104:118–26.
https://doi.org/10.1016/j.carbpol.2014.01.022.
Huang JZ, Wu J, Xiang S, et al. Catalpol preserves neural function and attenuates the pathology of
Alzheimer’s disease in mice. Mol Med Rep. 2016;13(1):491–6. https://doi.org/10.3892/mmr.
2015.4496.
Hurley LL, Akinfiresoye L, Nwulia E, et al. Antidepressant-like effects of curcumin in WKY rat
model of depression is associated with an increase in hippocampal BDNF. Behav Brain Res.
2013;239:27–30. https://doi.org/10.1016/j.bbr.2012.10.049.
Jang D, Lee HJ, Lee K, et al. White ginseng ameliorates depressive behavior and increases hippo-
campal 5-HT level in the stressed ovariectomized rats. Biomed Res Int. 2019;2019:5705232.
Published 2019 Feb 11. https://doi.org/10.1155/2019/5705232
Jiang LD. Discussion on concepts of the health, subhealth, before sickness and prevention. China J
Tradit Chin Med Pharm. 2010;25(2):167–70.
Jiang J, Wang W, Sun YJ, et al. Neuroprotective effect of curcumin on focal cerebral ischemic rats
by preventing blood–brain barrier damage. Eur J Pharmacol. 2007;561(1–3):54–62. https://doi.
org/10.1016/j.ejphar.2006.12.028.
TCM Substances in Neuropsychopharmacotherapy 485
Jiang TF, Zhang YJ, Zhou HY, et al. Curcumin ameliorates the neurodegenerative pathology in
A53T α-synuclein cell model of Parkinson’s disease through the downregulation of mTOR/
p70S6K signaling and the recovery of macroautophagy. J NeuroImmune Pharmacol. 2013;8
(1):356–69. https://doi.org/10.1007/s11481-012-9431-7.
Jiang N, Zhang BY, Dong LM, et al. Antidepressant effects of dammarane sapogenins in chronic
unpredictable mild stress-induced depressive mice. Phytother Res. 2018;32(6):1023–9. https://
doi.org/10.1002/ptr.6040.
Jin Y, Cui R, Zhao L, Fan J, Li B. Mechanisms of Panax ginseng action as an antidepressant. Cell
Prolif. 2019;52(6):e12696. https://doi.org/10.1111/cpr.12696.
Joe B, Vijaykumar M, Lokesh BR. Biological properties of curcumin-cellular and molecular
mechanisms of action. Crit Rev Food Sci Nutr. 2004;44(2):97–111. https://doi.org/10.1080/
10408690490424702.
Jozuka H, Jozuka E, Takeuchi S, Nishikaze O. Comparison of immunological and endocrinological
markers associated with major depression. J Int Med Res. 2003;31(1):36–41. https://doi.org/10.
1177/147323000303100106.
Kalaycıoğlu Z, Gazioğlu I, Erim FB. Comparison of antioxidant, anticholinesterase, and anti-
diabetic activities of three curcuminoids isolated from Curcuma longa L. Nat Prod Res.
2017;31(24):2914–7. https://doi.org/10.1080/14786419.2017.1299727.
Kalonia H, Kumar A. Protective effect of melatonin on certain behavioral and biochemical
alterations induced by sleep-deprivation in mice. Indian J Pharm. 2007;39(1):48–51.
Kang JM, Lin S. Ginkgo biloba and its potential role in glaucoma. Curr Opin Ophthalmol. 2018;29
(2):116–20. https://doi.org/10.1097/ICU.0000000000000459.
Kaniakova M, Krausova B, Vyklicky V, et al. Key amino acid residues within the third membrane
domains of NR1 and NR2 subunits contribute to the regulation of the surface delivery of N-
methyl-D-aspartate receptors. J Biol Chem. 2012;287(31):26423–34. https://doi.org/10.1074/
jbc.M112.339085.
Kashani L, Eslatmanesh S, Saedi N, et al. Comparison of saffron versus fluoxetine in treatment of
mild to moderate postpartum depression: a double-blind, randomized clinical trial. Pharmacop-
sychiatry. 2017;50(2):64–8. https://doi.org/10.1055/s-0042-115306.
Kaufmann FN, Costa AP, Ghisleni G, et al. NLRP3 inflammasome-driven pathways in depression:
clinical and preclinical findings. Brain Behav Immun. 2017;64:367–83. https://doi.org/10.1016/
j.bbi.2017.03.002.
Kim SO, You JM, Yun SJ, Son MS, Nam KN, Hong JW, Kim SY, Choi SY, Lee EH. Ginsenoside
rb1 and rg3 attenuate glucocorticoid-induced neurotoxicity. Cell Mol Neurobiol. 2010;30
(6):857–62. https://doi.org/10.1007/s10571-010-9513-0. Epub 2010 Mar 25. PMID:
20336484.
Kim JS, Kim Y, Han SH, et al. Development and validation of an LC-MS/MS method for
determination of compound K in human plasma and clinical application. J Ginseng Res.
2013;37(1):135–41. https://doi.org/10.5142/jgr.2013.37.135.
Kim HJ, Park SD, Lee RM, et al. Gintonin attenuates depressive-like behaviors associated with
alcohol withdrawal in mice. J Affect Disord. 2017;215:23–9. https://doi.org/10.1016/j.jad.2017.
03.026.
Kiss JP, Szasz BK, Fodor L, et al. GluN2B-containing NMDA receptors as possible targets for the
neuroprotective and antidepressant effects of fluoxetine. Neurochem Int. 2012;60(2):170–6.
https://doi.org/10.1016/j.neuint.2011.12.005.
Kita T, Imai S, Sawada H, et al. The biosynthetic pathway of curcuminoid in turmeric (Curcuma
longa) as revealed by13C-labeled precursors. Biosci Biotechnol Biochem. 2008;72(7):1789–98.
https://doi.org/10.1271/bbb.80075.
Koo HJ, Lim KH, Jung HJ, Park EH. Anti-inflammatory evaluation of gardenia extract,
geniposide and genipin. J Ethnopharmacol. 2006;103(3):496–500. https://doi.org/10.1016/j.
jep.2005.08.011.
Kuang Y, Li B, Fan J, Qiao X, Ye M. Antitussive and expectorant activities of licorice and its major
compounds. Bioorg Med Chem. 2018;26(1):278–84. https://doi.org/10.1016/j.bmc.2017.11.046.
486 Y. Wang and J. Li
Kulkarni SK, Dhir A. An overview of curcumin in neurological disorders. Indian J Pharm Sci.
2010;72(2):149–54. https://doi.org/10.4103/0250-474X.65012.
Kumar A, Kalonia H. Protective effect of Withania somnifera Dunal on the behavioral and
biochemical alterations in sleep-disturbed mice (Grid over water suspended method). Indian J
Exp Biol. 2007;45(6):524–8.
Kumar A, Singh A. Possible nitric oxide modulation in protective effect of (Curcuma longa,
Zingiberaceae) against sleep deprivation-induced behavioral alterations and oxidative damage
in mice. Phytomedicine. 2008;15(8):577–86. https://doi.org/10.1016/j.phymed.2008.02.003.
Kumar A, Prakash A, Dogra S. Protective effect of curcumin (Curcuma longa) against D-galactose-
induced senescence in mice. J Asian Nat Prod Res. 2011;13(1):42–55. https://doi.org/10.1080/
10286020.2010.544253.
Laakmann G, Dienel A, Kieser M. Clinical significance of hyperforin for the efficacy of Hypericum
extracts on depressive disorders of different severities. Phytomedicine. 1998;5(6):435–42.
https://doi.org/10.1016/S0944-7113(98)80039-1.
Lai N, Zhang J, Ma X, et al. Regulatory effect of catalpol on Th1/Th2 cells in mice with bone loss
induced by estrogen deficiency. Am J Reprod Immunol. 2015;74(6):487–98. https://doi.org/10.
1111/aji.12423.
Lee KJ, Ji GE. The effect of fermented red ginseng on depression is mediated by lipids. Nutr
Neurosci. 2014;17(1):7–15. https://doi.org/10.1179/1476830513Y.0000000059.
Lee B, Lee H. Systemic administration of curcumin affect anxiety-related behaviors in a rat model of
posttraumatic stress disorder via activation of serotonergic systems. Evid Based Complement
Alternat Med. 2018;2018:9041309. Published 2018 Jun 19. https://doi.org/10.1155/2018/9041309
Lee S, Rhee DK. Effects of ginseng on stress-related depression, anxiety, and the hypothalamic-
pituitary-adrenal axis. J Ginseng Res. 2017;41(4):589–94. https://doi.org/10.1016/j.jgr.2017.01.010.
Lee B, Sur B, Lee H, Oh S. Korean Red Ginseng prevents posttraumatic stress disorder-triggered
depression-like behaviors in rats via activation of the serotonergic system. J Ginseng Res.
2020;44(4):644–54. https://doi.org/10.1016/j.jgr.2019.09.005.
Leuner K, Kazanski V, Müller M, et al. Hyperforin – a key constituent of St. John’s wort specifically
activates TRPC6 channels. FASEB J. 2007;21(14):4101–11. https://doi.org/10.1096/fj.07-8110com.
Leung AY. Traditional toxicity documentation of Chinese Materia Medica – an overview. Toxicol
Pathol. 2006;34(4):319–26. https://doi.org/10.1080/01926230600773958.
Li HW, Meng XL. Research progress on chemical constituents and pharmacological activities of
Rehmannia glutinosa. Drug Eval Res. 2015;38(2):218–28.
Li T, Wang P, Guo W, et al. Natural berberine-based Chinese herb medicine assembled nano-
structures with modified antibacterial application. ACS Nano. 2019;13(6):6770–81. https://doi.
org/10.1021/acsnano.9b01346.
Liang Z, Bai S, Shen P, et al. GC-MS-based metabolomic study on the antidepressant-like effects of
diterpene ginkgolides in mouse Hippocampus. Behav Brain Res. 2016;314:116–24. https://doi.
org/10.1016/j.bbr.2016.08.001.
Liang ZH, Jia YB, Wang ML, et al. Efficacy of ginkgo biloba extract as augmentation of
venlafaxine in treating post-stroke depression. Neuropsychiatr Dis Treat. 2019;15:2551–2557.
Published 2019 Sep 3. https://doi.org/10.2147/NDT.S215191
Lim CY, Moon JM, Kim BY, et al. Comparative study of Korean White Ginseng and Korean Red
Ginseng on efficacies of OVA-induced asthma model in mice. J Ginseng Res. 2015;39(1):38–
45. https://doi.org/10.1016/j.jgr.2014.07.004.
Linde K, Ramirez G, Mulrow CD, Pauls A, Weidenhammer W, Melchart D. St John’s wort for
depression – an overview and meta-analysis of randomised clinical trials. BMJ. 1996;313
(7052):253–8. https://doi.org/10.1136/bmj.313.7052.253.
Liu YF. Studies on the active substances and function of Rehmannia Radix. Beijing: Chinese
Academy of Medical Science Peking Union Medical Colloge; 2013.
Liu T. Clinical application of LBRD. Nei Mongol. J Tradit Chin Med. 2014;33(22):94–6.
Liu JY, Zhang DJ. Amelioration by catalpol of atherosclerotic lesions in hypercholesterolemic
rabbits. Planta Med. 2015;81(3):175–84. https://doi.org/10.1055/s-0034-1396240.
TCM Substances in Neuropsychopharmacotherapy 487
Liu C, Wu F, Liu Y, Meng C. Catalpol suppresses proliferation and facilitates apoptosis of MCF-7
breast cancer cells through upregulating microRNA-146a and downregulating matrix meta-
lloproteinase-16 expression. Mol Med Rep. 2015;12(5):7609–14. https://doi.org/10.3892/mmr.
2015.4361.
Liu Z, Qi Y, Cheng Z, Zhu X, Fan C, Yu SY. The effects of ginsenoside Rg1 on chronic stress induced
depression-like behaviors, BDNF expression and the phosphorylation of PKA and CREB in rats.
Neuroscience. 2016a;322:358–69. https://doi.org/10.1016/j.neuroscience.2016.02.050.
Liu X, Guo H, Sayed MD, et al. cAMP/PKA/CREB/GLT1 signaling involved in the antidepressant-
like effects of phosphodiesterase 4D inhibitor (GEBR-7b) in rats. Neuropsychiatr Dis Treat.
2016b;12:219–227. Published 2016 Jan 21. https://doi.org/10.2147/NDT.S90960
Liu P, Lin ZJ, Zhang B. Research progress on chemical constituents and pharmacological effect of
Lilii Bulbus. Chin J Exp Tradit Med Formulae. 2017;23(23):201–11.
Liu D, Huo X, Gao L, Zhang J, Ni H, Cao L. NF-κB and Nrf2 pathways contribute to the protective
effect of Licochalcone A on dextran sulphate sodium-induced ulcerative colitis in mice. Biomed
Pharmacother. 2018;102:922–9. https://doi.org/10.1016/j.biopha.2018.03.130.
Lopresti AL. Curcumin for neuropsychiatric disorders: a review of in vitro, animal and human
studies. J Psychopharmacol. 2017;31(3):287–302. https://doi.org/10.1177/0269881116686883.
Lopresti AL, Drummond PD. Saffron (Crocus sativus) for depression: a systematic review of
clinical studies and examination of underlying antidepressant mechanisms of action. Hum
Psychopharmacol. 2014;29(6):517–27. https://doi.org/10.1002/hup.2434.
Lopresti AL, Maes M, Maker GL, Hood SD, Drummond PD. Curcumin for the treatment of major
depression: a randomised, double-blind, placebo controlled study. J Affect Disord.
2014;167:368–75. https://doi.org/10.1016/j.jad.2014.06.001.
Lopresti AL, Drummond PD, Inarejos-García AM, Prodanov M. affron®, a standardised extract
from saffron (Crocus sativus L.) for the treatment of youth anxiety and depressive symptoms: a
randomised, double-blind, placebo-controlled study. J Affect Disord. 2018;232:349–57. https://
doi.org/10.1016/j.jad.2018.02.070.
Ma YT. Synthesis and Antitumor Activity of novel Isoliquiritigenin Analogues. Harbin, Heilong-
jiang Province, China. Northeast Forestry University. 2009.
Ma R, Zhu R, Wang L, et al. Diabetic osteoporosis: a review of its traditional Chinese medicinal
use and clinical and preclinical research. Evid Based Complement Alternat Med.
2016;2016:3218313. https://doi.org/10.1155/2016/3218313.
Maes M, Bosmans E, Meltzer HY. Immunoendocrine aspects of major depression. Relationships
between plasma interleukin-6 and soluble interleukin-2 receptor, prolactin and cortisol. Eur
Arch Psychiatry Clin Neurosci. 1995;245(3):172–8. https://doi.org/10.1007/BF02193091.
Maes M, Song C, Yirmiya R. Targeting IL-1 in depression. Expert Opin Ther Targets. 2012;16
(11):1097–112. https://doi.org/10.1517/14728222.2012.718331.
Malberg JE, Blendy JA. Antidepressant action: to the nucleus and beyond. Trends Pharmacol Sci.
2005;26(12):631–8. https://doi.org/10.1016/j.tips.2005.10.005.
Mancuso C, Santangelo R. Panax ginseng and Panax quinquefolius: from pharmacology to toxicol-
ogy. Food Chem Toxicol. 2017;107(Pt A):362–72. https://doi.org/10.1016/j.fct.2017.07.019.
Mansouri Z, Sabetkasaei M, Moradi F, et al. Curcumin has neuroprotection effect on homocysteine
rat model of Parkinson. J Mol Neurosci. 2012;47(2):234–42. https://doi.org/10.1007/s12031-
012-9727-3.
Meng Y, Jia Y, Wu YW, Xiang H, Qin XM, Tian JS. Research progress on Baihe Dihuang decoction
in nervous-mental system. Chin Tradit Herb Drug. 2018;49(1):251–5.
Mengoni F, Lichtner M, Battinelli L, et al. In vitro anti-HIV activity of oleanolic acid on infected
human mononuclear cells. Planta Med. 2002;68(2):111–4. https://doi.org/10.1055/s-2002-20256.
Miao M, Peng M, Chen H, Liu B. Effects of Baihe Dihuang powder on chronic stress depression rat
models. Saudi J Biol Sci. 2019;26(3):582–8. https://doi.org/10.1016/j.sjbs.2018.12.002.
Miquel J, Bernd A, Sempere JM, Díaz-Alperi J, Ramírez A. The curcuma antioxidants: pharmaco-
logical effects and prospects for future clinical use. A review. Arch Gerontol Geriatr. 2002;34
(1):37–46. https://doi.org/10.1016/s0167-4943(01)00194-7.
488 Y. Wang and J. Li
Molteni R, Calabrese F, Bedogni F, et al. Chronic treatment with fluoxetine up-regulates cellular
BDNF mRNA expression in rat dopaminergic regions. Int J Neuropsychopharmacol. 2006;9
(3):307–17. https://doi.org/10.1017/S1461145705005766.
Motin VG, Yasnetsov VV. Effect of NMDA, a specific agonist to NMDA receptor complex, on rat
hippocampus. Bull Exp Biol Med. 2015;159(6):704–7. https://doi.org/10.1007/s10517-015-3053-z.
Mueller SC, Majcher-Peszynska J, Uehleke B, et al. The extent of induction of CYP3A by St.
John’s wort varies among products and is linked to hyperforin dose. Eur J Clin Pharmacol.
2006;62(1):29–36. https://doi.org/10.1007/s00228-005-0061-3.
Murakami S, Imbe H, Morikawa Y, Kubo C, Senba E. Chronic stress, as well as acute stress,
reduces BDNF mRNA expression in the rat hippocampus but less robustly. Neurosci Res.
2005;53(2):129–39. https://doi.org/10.1016/j.neures.2005.06.008.
Nakagawasai O, Oba A, Sato A, et al. Subchronic stress-induced depressive behavior in ovariec-
tomized mice. Life Sci. 2009;84(15–16):512–6. https://doi.org/10.1016/j.lfs.2009.01.009.
Napryeyenko O, Sonnik G, Tartakovsky I, et al. Efficacy and tolerability of Ginkgo biloba extract
EGb 761 by type of dementia: analyses of a randomised controlled trial. J Neurol Sci. 2009;283
(1–2):224–9. https://doi.org/10.1016/j.jns.2009.02.353.
Ng QX, Koh SSH, Chan HW, Ho CYX. Clinical use of curcumin in depression: a meta-analysis. J
Am Med Dir Assoc. 2017a;18(6):503–8. https://doi.org/10.1016/j.jamda.2016.12.071.
Ng QX, Venkatanarayanan N, Ho CY. Clinical use of Hypericum perforatum (St John’s wort) in
depression: a meta-analysis. J Affect Disord. 2017b;210:211–21. https://doi.org/10.1016/j.jad.
2016.12.048.
Nibuya M, Nestler EJ, Duman RS. Chronic antidepressant administration increases the expression
of cAMP response element binding protein (CREB) in rat hippocampus. J Neurosci. 1996;16
(7):2365–72. https://doi.org/10.1523/JNEUROSCI.16-07-02365.1996.
Oh J, Jeon SB, Lee Y, et al. Fermented red ginseng extract inhibits cancer cell proliferation and
viability. J Med Food. 2015;18(4):421–8. https://doi.org/10.1089/jmf.2014.3248.
Pan R, Qiu S, Lu DX, Dong J. Curcumin improves learning and memory ability and its
neuroprotective mechanism in mice. Chin Med J. 2008;121(9):832–9.
Pan Y, Chen XY, Zhang QY, et al. Microglial NLRP3 inflammasome activation mediates IL-1β-
related inflammation in prefrontal cortex of depressive rats. Brain Behav Immun. 2014;41:90–
100. https://doi.org/10.1016/j.bbi.2014.04.007.
Pastorino G, Cornara L, Soares S, Rodrigues F, Oliveira MBPP. Liquorice (Glycyrrhiza glabra): a
phytochemical and pharmacological review. Phytother Res. 2018;32(12):2323–39. https://doi.
org/10.1002/ptr.6178.
Pathak AK, Bhutani M, Nair AS, et al. Ursolic acid inhibits STAT3 activation pathway leading to
suppression of proliferation and chemosensitization of human multiple myeloma cells
[published correction appears in Mol Cancer Res. 2018 Sep;16(9):1442]. Mol Cancer Res.
2007; 5(9):943–955. https://doi.org/10.1158/1541-7786.MCR-06-0348.
Patil SS, Schlick F, Höger H, et al. Involvement of individual hippocampal signaling protein levels
in spatial memory formation is strain-dependent [published correction appears in Amino Acids.
2010 Jul;39(2):617–618]. Amino Acids. 2010;39(1):75–87. https://doi.org/10.1007/s00726-
009-0379-8.
Plotsky PM, Owens MJ, Nemeroff CB. Psychoneuroendocrinology of depression. Hypothalamic-
pituitary-adrenal axis. Psychiatr Clin North Am. 1998;21(2):293–307. https://doi.org/10.1016/
s0193-953x(05)70006-x.
Pöldinger W. Zur Geschichte des Johanniskrauts [History of St. Johns wort]. Praxis (Bern 1994).
2000;89(50):2102–9.
Pollier J, Goossens A. Oleanolic acid. Phytochemistry. 2012;77:10–5. https://doi.org/10.1016/j.
phytochem.2011.12.022.
Pyrzanowska J, Piechal A, Blecharz-Klin K, et al. The influence of the long-term administration of
Curcuma longa extract on learning and spatial memory as well as the concentration of brain
neurotransmitters and level of plasma corticosterone in aged rats. Pharmacol Biochem Behav.
2010;95(3):351–8. https://doi.org/10.1016/j.pbb.2010.02.013.
TCM Substances in Neuropsychopharmacotherapy 489
Shim JS, Song MY, Yim SV, Lee SE, Park KS. Global analysis of ginsenoside Rg1 protective
effects in β-amyloid-treated neuronal cells. J Ginseng Res. 2017;41(4):566–71. https://doi.org/
10.1016/j.jgr.2016.12.003.
Shin BK, Kwon SW, Park JH. Chemical diversity of ginseng saponins from Panax ginseng. J
Ginseng Res. 2015;39(4):287–98. https://doi.org/10.1016/j.jgr.2014.12.005.
Shirayama Y, Chen AC, Nakagawa S, Russell DS, Duman RS. Brain-derived neurotrophic factor
produces antidepressant effects in behavioral models of depression. J Neurosci. 2002;22
(8):3251–61. https://doi.org/10.1523/JNEUROSCI.22-08-03251.2002.
Shishodia S, Majumdar S, Banerjee S, Aggarwal BB. Ursolic acid inhibits nuclear factor-kappaB
activation induced by carcinogenic agents through suppression of IkappaBalpha kinase and p65
phosphorylation: correlation with down-regulation of cyclooxygenase 2, matrix meta-
lloproteinase 9, and cyclin D1. Cancer Res. 2003;63(15):4375–83.
Shukla Y, Arora A, Taneja P. Antimutagenic potential of curcumin on chromosomal aberrations in
Wistar rats. Mutat Res. 2002;515(1–2):197–202. https://doi.org/10.1016/s1383-5718(02)00016-5.
Solanki I, Parihar P, Mansuri ML, et al. Flavonoid-based therapies in the early management of
neurodegenerative diseases. Adv Nutr. 2015;6(1):64–72. Published 2015 Jan 15. https://doi.org/
10.3945/an.114.007500
Soleymani S, Bahramsoltani R, Rahimi R, Abdollahi M. Clinical risks of St John’s Wort (Hyper-
icum perforatum) co-administration. Expert Opin Drug Metab Toxicol. 2017;13(10):1047–62.
https://doi.org/10.1080/17425255.2017.1378342.
Song X, Zhang W, Wang T, et al. Geniposide plays an anti-inflammatory role via regulating TLR4
and downstream signaling pathways in lipopolysaccharide-induced mastitis in mice. Inflamma-
tion. 2014;37(5):1588–98. https://doi.org/10.1007/s10753-014-9885-2.
Stout SC, Owens MJ, Nemeroff CB. Regulation of corticotropin-releasing factor neuronal systems
and hypothalamic-pituitary-adrenal axis activity by stress and chronic antidepressant treatment.
J Pharmacol Exp Ther. 2002;300(3):1085–92. https://doi.org/10.1124/jpet.300.3.1085.
Su GL, Liu G, Liu YC, Dong HR. Progress on extraction and purification of liquiritin and its
pharmacological effects. Mod Chin Med. 2011;13(10):48–51.
Sun T, Wang X, Xu H. Ginkgo biloba extract for angina pectoris: a systematic review. Chin J Integr
Med. 2015;21(7):542–50. https://doi.org/10.1007/s11655-015-2070-0.
Sung JY, Goo JS, Lee DE, et al. Learning strategy selection in the water maze and hippocampal
CREB phosphorylation differ in two inbred strains of mice. Learn Mem. 2008;15(4):183–188.
Published 2008 Mar 19. https://doi.org/10.1101/lm.783108
Tabeshpour J, Sobhani F, Sadjadi SA, et al. A double-blind, randomized, placebo-controlled
trial of saffron stigma (Crocus sativus L.) in mothers suffering from mild-to-moderate
postpartum depression. Phytomedicine. 2017;36:145–52. https://doi.org/10.1016/j.
phymed.2017.10.005.
Takei S, Morinobu S, Yamamoto S, et al. Enhanced hippocampal BDNF/TrkB signaling in response
to fear conditioning in an animal model of posttraumatic stress disorder. J Psychiatr Res.
2011;45(4):460–8. https://doi.org/10.1016/j.jpsychires.2010.08.009.
Tan MS, Yu JT, Tan CC, et al. Efficacy and adverse effects of Ginkgo biloba for cognitive
impairment and dementia: a systematic review and meta-analysis. J Alzheimers Dis. 2015;43
(2):589–603. https://doi.org/10.3233/JAD-140837.
Tang J, Xue W, Xia B, et al. Involvement of normalized NMDA receptor and mTOR-related
signaling in rapid antidepressant effects of Yueju and ketamine on chronically stressed mice.
Sci Rep. 2015;5:13573. Published 2015 Aug 28. https://doi.org/10.1038/srep13573
Tao W, Zhang H, Xue W, et al. Optimization of supercritical fluid extraction of oil from the fruit of
Gardenia jasminoides and its antidepressant activity. Molecules. 2014;19(12):19350–19360.
Published 2014 Nov 25. https://doi.org/10.3390/molecules191219350
Taylor MJ, Freemantle N, Geddes JR, Bhagwagar Z. Early onset of selective serotonin reuptake
inhibitor antidepressant action: systematic review and meta-analysis. Arch Gen Psychiatry.
2006;63(11):1217–23. https://doi.org/10.1001/archpsyc.63.11.1217.
Teng GL, Zhang YX, Wu QG, Chen Z, Huang XW, et al. Chinese Materia Medica. Beijing:
People’s Medical Publishing House; 2019.
TCM Substances in Neuropsychopharmacotherapy 491
Tóth B, Hegyi P, Lantos T, et al. The efficacy of saffron in the treatment of mild to moderate
depression: a meta-analysis. Planta Med. 2019;85(1):24–31. https://doi.org/10.1055/a-0660-9565.
Tsai SJ. Effects of interleukin-1beta polymorphisms on brain function and behavior in healthy and
psychiatric disease conditions. Cytokine Growth Factor Rev. 2017;37:89–97. https://doi.org/10.
1016/j.cytogfr.2017.06.001.
Tsai YM, Chien CF, Lin LC, et al. Curcumin and its nano-formulation: the kinetics of tissue
distribution and blood-brain barrier penetration. Int J Pharm. 2011;416(1):331–8. https://doi.
org/10.1016/j.ijpharm.2011.06.030.
Tulsulkar J, Glueck B, Hinds TD Jr, et al. Ginkgo biloba extract prevents female mice from
ischemic brain damage and the mechanism is independent of the HO1/Wnt pathway. Transl
Stroke Res. 2016;7(2):120–31. https://doi.org/10.1007/s12975-015-0433-7.
Venigalla M, Gyengesi E, Münch G. Curcumin and Apigenin – novel and promising therapeutics
against chronic neuroinflammation in Alzheimer’s disease [published correction appears in
Neural Regen Res. 2015 Dec;10(12):2017]. Neural Regen Res. 2015;10(8):1181–1185.
https://doi.org/10.4103/1673-5374.162686.
Voehringer DW. BCL-2 and glutathione: alterations in cellular redox state that regulate apoptosis
sensitivity. Free Radic Biol Med. 1999;27(9–10):945–50. https://doi.org/10.1016/s0891-5849
(99)00174-4.
Wahlström B, Blennow G. A study on the fate of curcumin in the rat. Acta Pharmacol Toxicol
(Copenh). 1978;43(2):86–92. https://doi.org/10.1111/j.1600-0773.1978.tb02240.x.
Wang P, Wang WP, Sun-Zhang, et al. Impaired spatial learning related with decreased expression of
calcium/calmodulin-dependent protein kinase IIα and cAMP-response element binding protein
in the pentylenetetrazol-kindled rats. Brain Res. 2008;1238:108–17. https://doi.org/10.1016/j.
brainres.2008.07.103.
Wang J, Gao W, Zhang L, Huang L. Establishment and quality assessment of tissue cultures in
Glycyrrhiza uralensis Fisch. Appl Biochem Biotechnol. 2013;169(2):588–94. https://doi.org/
10.1007/s12010-012-0012-2.
Wang J, Jing L, Toledo-Salas JC, Xu L. Rapid-onset antidepressant efficacy of glutamatergic system
modulators: the neural plasticity hypothesis of depression. Neurosci Bull. 2015a;31(1):75–86.
https://doi.org/10.1007/s12264-014-1484-6. Epub 2014 Dec 6.
Wang Z, Wei G, Ma S. Chemical and pharmacological effects of Rehmanniae Radix polysaccha-
rides. Chin J Exp Tradit Med Formulae. 2015b;21(16):231–5.
Wang GL, He ZM, Zhu HY, et al. Involvement of serotonergic, noradrenergic and dopaminergic
systems in the antidepressant-like effect of ginsenoside Rb1, a major active ingredient of Panax
ginseng C.A. Meyer. J Ethnopharmacol. 2017a;204:118–24.
Wang XR, Hao HG, Chu L. Glycyrrhizin inhibits LPS-induced inflammatory mediator production
in endometrial epithelial cells. Microb Pathog. 2017b;109:110–3. https://doi.org/10.1016/j.
micpath.2017.05.032.
Wise K, Selby-Pham S, Bennett L, Selby-Pham J. Pharmacokinetic properties of phytochemicals in
Hypericum perforatum influence efficacy of regulating oxidative stress. Phytomedicine.
2019;59:152763. https://doi.org/10.1016/j.phymed.2018.11.023.
Woelk H, Burkard G, Grünwald J. Benefits and risks of the hypericum extract LI 160: drug
monitoring study with 3250 patients. J Geriatr Psychiatry Neurol. 1994;7(Suppl 1):S34–8.
https://doi.org/10.1177/089198879400700110.
Wu A, Ying Z, Gomez-Pinilla F. The interplay between oxidative stress and brain-derived neurotrophic
factor modulates the outcome of a saturated fat diet on synaptic plasticity and cognition. Eur J
Neurosci. 2004;19(7):1699–707. https://doi.org/10.1111/j.1460-9568.2004.03246.x.
Wu PS, Wu SJ, Tsai YH, Lin YH, Chao JC. Hot water extracted Lycium barbarum and Rehmannia
glutinosa inhibit liver inflammation and fibrosis in rats. Am J Chin Med. 2011;39(6):1173–91.
https://doi.org/10.1142/S0192415X11009482.
Wu R, Zhu D, Xia Y, et al. A role of Yueju in fast-onset antidepressant action on major depressive
disorder and serum BDNF expression: a randomly double-blind, fluoxetine-adjunct, placebo-
controlled, pilot clinical study. Neuropsychiatr Dis Treat. 2015;11:2013–2021. Published 2015
Aug 6. https://doi.org/10.2147/NDT.S86585
492 Y. Wang and J. Li
Wu R, Tao W, Zhang H, et al. Instant and persistent antidepressant response of gardenia yellow
pigment is associated with acute protein synthesis and delayed upregulation of BDNF expres-
sion in the hippocampus. ACS Chem Neurosci. 2016;7(8):1068–76. https://doi.org/10.1021/
acschemneuro.6b00011.
Xia X, Pan Y, Zhang WY, et al. Ethanolic extracts from Curcuma longa attenuates behavioral,
immune, and neuroendocrine alterations in a rat chronic mild stress model. Biol Pharm Bull.
2006;29(5):938–44. https://doi.org/10.1248/bpb.29.938.
Xiao Q, Xiong Z, Yu C, et al. Antidepressant activity of crocin-I is associated with amelioration of
neuroinflammation and attenuates oxidative damage induced by corticosterone in mice. Physiol
Behav. 2019;212:112699. https://doi.org/10.1016/j.physbeh.2019.112699.
Xu Y, Ku B, Tie L, et al. Curcumin reverses the effects of chronic stress on behavior, the HPA axis,
BDNF expression and phosphorylation of CREB. Brain Res. 2006;1122(1):56–64. https://doi.
org/10.1016/j.brainres.2006.09.009.
Xue W, Zhou X, Yi N, et al. Yueju pill rapidly induces antidepressant-like effects and acutely
enhances BDNF expression in mouse brain. Evid Based Complement Alternat Med.
2013;2013:184367. https://doi.org/10.1155/2013/184367.
Xue W, Wang W, Gong T, et al. PKA-CREB-BDNF signaling regulated long lasting antidepressant
activities of Yueju but not ketamine. Sci Rep. 2016;6:26331. Published 2016 May 20. https://
doi.org/10.1038/srep26331
Yallapu MM, Jaggi M, Chauhan SC. Curcumin nanoformulations: a future nanomedicine for cancer.
Drug Discov Today. 2012;17(1–2):71–80. https://doi.org/10.1016/j.drudis.2011.09.009.
Yamada N, Araki H, Yoshimura H. Identification of antidepressant-like ingredients in ginseng root
(Panax ginseng C.A. Meyer) using a menopausal depressive-like state in female mice: partic-
ipation of 5-HT2A receptors. Psychopharmacology. 2011;216(4):589–99. https://doi.org/10.
1007/s00213-011-2252-1.
Yan YX, Liu YQ, Li M, et al. Development and evaluation of a questionnaire for measuring
suboptimal health status in urban Chinese. J Epidemiol. 2009;19(6):333–41. https://doi.org/10.
2188/jea.je20080086.
Yang L, Dong X. Inhibition of inflammatory response by crocin attenuates hemorrhagic shock-
induced organ damages in rats. J Interf Cytokine Res. 2017;37(7):295–302. https://doi.org/10.
1089/jir.2016.0137.
Yang F, Lim GP, Begum AN, et al. Curcumin inhibits formation of amyloid β oligomers and fibrils,
binds plaques, and reduces amyloidin vivo. J Biol Chem. 2004;280(7):5892–901. https://doi.
org/10.1074/jbc.m404751200.
Yang R, Wang LQ, Yuan BC, Liu Y. The pharmacological activities of licorice. Planta Med.
2015;81(18):1654–69. https://doi.org/10.1055/s-0035-1557893.
Yosri H, Elkashef WF, Said E, Gameil NM. Crocin modulates IL-4/IL-13 signaling and ameliorates
experimentally induced allergic airway asthma in a murine model. Int Immunopharmacol.
2017;50:305–12. https://doi.org/10.1016/j.intimp.2017.07.012.
Zahner C, Kruttschnitt E, Uricher J, et al. No clinically relevant interactions of St. John’s Wort
extract Ze 117 low in Hyperforin with cytochrome P450 enzymes and P-glycoprotein. Clin
Pharmacol Ther. 2019;106(2):432–40. https://doi.org/10.1002/cpt.1392.
Zhang YL, Meng FJ, Tian Y, et al. Study on chemical composition and pharmacological action of
licorice. Chem Eng. 2009;8:60–63, 66
Zhang L, Xu T, Wang S, et al. Curcumin produces antidepressant effects via activating MAPK/
ERK-dependent brain-derived neurotrophic factor expression in the amygdala of mice. Behav
Brain Res. 2012;235(1):67–72. https://doi.org/10.1016/j.bbr.2012.07.019.
Zhang L, Luo J, Zhang M, et al. Effects of curcumin on chronic, unpredictable, mild, stress-induced
depressive-like behaviour and structural plasticity in the lateral amygdala of rats. Int J
Neuropsychopharmacol. 2014a;17(5):793–806. https://doi.org/10.1017/S1461145713001661.
Zhang W, Men X, Lei P. Review on anti-tumor effect of triterpene acid compounds. J Cancer Res
Ther. 2014b;10(Suppl 1):14–9. https://doi.org/10.4103/0973-1482.139746.
TCM Substances in Neuropsychopharmacotherapy 493
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 496
Evidence for NMDA Receptor Hypofunction in Schizophrenia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 498
NMDA Receptor Antagonists . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 498
Anti-NMDA Receptor Antibodies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 500
NMDA Receptor Hypofunction and Onset of Schizophrenia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 500
NMDA Receptor System in Schizophrenia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 501
Dopaminergic Transmission Under NMDA Receptor Hypofunction . . . . . . . . . . . . . . . . . . . . . . 502
NMDA Receptor Function Enhancement Therapy for Schizophrenia: Rationale
for Application of D-Serine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 503
Preclinical Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 503
Clinical Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 505
Issues and Improvement of NMDA Receptor Function Enhancement Therapy . . . . . . . . . . . 506
Modulation of Brain D-Serine Signaling for Development of Next Generation
Antipsychotics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 506
Detection and Distribution of Endogenous D-Serine in Mammalian Brains . . . . . . . . . . . . . . . 507
Biosynthesis of D-Serine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 508
Storage of D-Serine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 509
Extracellular D-Serine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 510
Uptake of D-Serine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 513
Degradation of D-Serine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 513
D-Serine Action on GluD2 and GluN1/GluN3 Glutamate Receptors . . . . . . . . . . . . . . . . . . . . . . 514
D-Serine Systems in Schizophrenia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 515
Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 516
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 516
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 516
Abstract
Based on the glutamate hypothesis of schizophrenia postulating hypofunction of
the N-methyl-D-aspartate type glutamate receptor (NMDA receptor) consisting of
GluN1 and GluN2 subunits, D-serine has been selected as a candidate agent that
is expected to satisfy an unmet need to ameliorate antipsychotic-resistant negative
and cognitive symptoms in schizophrenia by recovering the NMDA receptor
disturbance as one of its coagonists. Meta-analyses of randomized controlled
trials of D-serine given with currently available antipsychotic drugs in patients
with schizophrenia have consistently shown a significant reduction in the rating
scores of their negative symptoms although D-serine has not yet been approved
for clinical use. This chapter begins with verification of the evidence for the
involvement of diminished NMDA receptor-mediated transmission in the patho-
physiology of schizophrenia, accumulated data illustrating the plausible mecha-
nisms from the hypofunction to the positive, negative, and cognitive symptoms
and the relationships between the NMDA dysfunction and the well-established
dopamine hypothesis assuming cerebral hyperdopaminergic activities. The con-
secutive parts verify the rationales for application of D-serine and other agonists
acting at the glycine site of the NMDA receptor for development of NMDA
receptor activity-enhancing therapy of schizophrenia and outline the preclinical
and clinical studies using these agonists. The NMDA receptor targeting strategy
is challenged by the results that the add-on treatments with augmentation of the
extracellular glycine concentrations have failed to improve the antipsychotic-
refractory symptoms of schizophrenia patients. To address this issue, the latter
parts raise the possibility to restore functioning of the forebrain NMDA receptor
by upregulating endogenous D-serine signaling at the synapses by modifying the
molecules that compose the metabolic pathways and regulatory systems of the
extracellular levels of D-serine.
Introduction
Schizophrenia is a serious brain disorder with a high prevalence of about 1%, which
causes a variety of mental and behavioral disturbances leading to a chronic disability
in individuals and social life such as a very low rate of full employment (Bouwmans
et al. 2015; McCutcheon et al. 2020). The characteristic disturbances fall into three
major categories positive, negative, and cognitive symptoms (Fig. 1).
Of these, the currently used antipsychotic-resistant negative and cognitive symp-
toms are attributed to the difficult prognosis (Fig. 1). An urgent unmet need in
schizophrenia, thus, is to develop a highly effective treatment for these two symptom
groups including blunted affect, alogia, social withdrawal, avolition, anhedonia,
executive dysfunction, working memory, attention deficits, etc. However, in contrast
to the ameliorating actions of antipsychotics on the positive symptoms, which have
been shown to occur through dopamine receptor blockade, pharmacotherapeutic
clues to improve these refractory symptoms were not available for a long time until
D-Serine: Basic Aspects with a Focus on Psychosis 497
Fig. 1 Treatment strategies for hypothetical molecular pathophysiology underlying the positive,
negative, and cognitive symptoms of schizophrenia and drug-induced schizophrenia-like psychosis
The neurotransmitter and neuromodulator mechanisms underlying the positive, negative, and
cognitive symptoms of schizophrenia and substance-induced psychosis, as inferred from clinical
and molecular pharmacological data, are described. Conventional pharmacotherapy by antipsy-
chotics with dopamine antagonist properties ameliorates the positive symptoms of schizophrenia
and novel pharmacotherapy by NMDA receptor glycine site agonists and D-serine signaling
enhancers is expected to possess efficacies on both of its antipsychotic-responsive and -resistant
symptoms.
Abbreviations: AMP amphetamine, D-Ser D-serine, GABA γ-aminobutyric acid, KET ketamine, MAP
methamphetamine, NMDAR N-methyl-D-aspartate type glutamate receptor, PCP phencyclidine
the 1980s. The discovery in 1983 that phencyclidine (PCP), which had been known
to induce schizophrenia-like negative and cognitive symptoms as well as positive
symptoms, is a noncompetitive antagonist for the N-methyl-D-aspartate
(NMDA)-type glutamate receptor (NMDA receptor) (Anis et al. 1983) (Fig. 1) has
made this receptor a target for the development of agents possessing therapeutic
efficacy on the intractable schizophrenic symptoms.
This section outlines the evidence for the involvement of the NMDA receptor
hypofunction in the pathophysiology of schizophrenia stemming from the basic and
clinical research on PCP, the road by which NMDA receptor function-enhancing
drugs, including D-serine, were introduced as novel therapeutic agents for this
disorder, and their clinical effects (Nishikawa 2011). In addition, the process from
these studies to the uncovering of the presence of endogenous D-serine in mamma-
lian brains, and the molecular mechanisms underlying the brain D-serine metabolism
and dynamics will be reviewed. Finally, the possible application of the molecules
498 T. Nishikawa et al.
that regulate the D-serine signaling for the creation of new methods for the recovery
of the NMDA receptor hypofunction in schizophrenia will be discussed.
Fig. 2 Schematic representation of the relationships among D-serine and kynurenic acid metabolic
pathways and GluN1/GluN2-type NMDA receptor in mammalian brains
The regulatory sites of the GluN1/GluN2-type NMDA receptor and the candidate molecules and
cells for the metabolic or functional processes of D-serine and kynureate in the mammalian brains
are illustrated. The NMDA receptor complex has the multiple binding sites for L-glutamate
(L-GLU), glycine/D-serine (GLY), magnesium ion (Mg++), phencyclidine (PCP), zing ion (Zn++),
and polyamines (POLY). The heteromeric NMDA receptor forms a tetrameric ligand-gated channel
comprising two copies each of the GluN1 and GluN2 subunits. D-Serine has been demonstrated to
be mainly synthesized by SRR and to act as a coagonist for the NMDA receptor. Kynurenate occurs
via the kynurenine pathway and attenuates the NMDA receptor activity by blocking the glycine-
binding site. The interaction of D-serine with the NMDA receptor containing the GluN1 and GluN3
subunits and δ2 glutamate receptor are not shown in this scheme
Abbreviations: AMPAR α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor, Asc-1
Na+-independent alanine-serine-cysteine transporter 1, ASCT1 Na+-dependent broad-spectrum
neutral amino acid transporter 1, ASCT2 Na+-dependent broad-spectrum neutral amino acid trans-
porter 2, Bergmann Bergmann glia, DAO D-amino acid oxidase, D-Ser D-serine, GABAAR
γ-aminobutyric acid A receptor, Gly glycine, L-Glu L-glutamate, Müller Müller cells, NMDAR N-
methyl-D-aspartate type glutamate receptor, KYNA kynurenate, PAT1 proton/amino acid transporter
1, PCP phencyclidine, POLY polyamines, SNARE soluble N-methylmaleimide susceptibility factor
attachment protein receptor, SNAT2 sodium-coupled neutral amino acid transporter 2, SRR serine
racemase
Symbols: dotted square and line, antagonist; solid squares and lines, agonists
observations that in the stereoisomers of ketamine, the S form of ketamine, which has
a higher affinity for the NMDA receptors, is more potent to cause psychotic symptoms
than the R form, which has a lower affinity (Vollenweider et al. 1997). Furthermore, a
single challenge of small doses of PCP (Luby et al. 1959) or ketamine (Malhotra et al.
1997) has been found to elicit higher degrees and longer durations of induced
psychotic symptoms in schizophrenia patients than in healthy volunteers, suggesting
a lower level of the NMDA receptor function in schizophrenia.
500 T. Nishikawa et al.
medication-free patients with schizophrenia (Iwata et al. 2018). The glycine modu-
latory site has been recognized to play a critical role in the physiological functioning
of the GluN1/GluN2 type NMDA receptor (Fig. 2), as the glycine site agonists, such
as glycine, D-serine, and D-alanine, are essential for activation of the glutamate
receptor without provoking EPSPs alone (Danysz and Parsons 1998). Accordingly,
insufficient stimulation of the glycine site following reduced or increased concen-
trations of its endogenous agonist or antagonist, respectively, may be a factor that
attenuates the NMDA receptor activity. Kynurenic acid, which is biosynthesized by
the kynurenine pathway from tryptophan and acts as an antagonist at the glycine
regulatory site (Fig. 2), has been reported in meta-analyses to be increased in the
cerebrospinal fluid of schizophrenia patients (Plitman et al. 2017). The findings
concerning the intrinsic glycine site agonists, glycine and D-serine, in biological samples
obtained from patients with schizophrenia are discussed later in section “D-Serine
Systems in Schizophrenia.”
The important question then arises as to the relationships between the NMDA
receptor hypofunction and cerebral hyperdopaminergic activity that has been well
established to be involved in the pathophysiology of schizophrenia. A body of
evidence has been accumulated arguing that disruption of the GluN1/GluN2 type
NMDA receptor-mediated glutamate transmission is accompanied by accelerated
dopaminergic transmission in the cerebrum.
Early animal studies have shown that noncompetitive antagonists for the NMDA
receptor, such as PCP and MK-801 (dizocilpine), which bind to the PCP site of the
receptor (Fig. 2), and competitive antagonists that act on the glutamate site (Fig. 2)
augment dopamine metabolism in the cerebral cortex and limbic regions including
the nucleus accumbens and olfactory tubercle (Deutch et al. 1987; Hata et al. 1990;
Rao et al. 1990; Umino et al. 1998). Furthermore, in vivo dialysis experiments have
revealed that both of these types of antagonists elicit an increase in the extracellular
dopamine concentrations in the frontal cortex and striatum in a nerve impulse-
dependent manner (Kashiwa et al. 1995; Nishijima et al. 1994, 1996). Interestingly,
the GABAA receptor agonists blocked the ability of PCP to produce an upregulation
of the extracellular dopamine levels accompanied by reduction in the extracellular
GABA concentrations in the frontal cortex (Yonezawa et al. 1998) (Fig. 1). In
conjunction with the expression of NMDA receptors on GABAergic interneurons
(Belforte et al. 2010), these results suggest the pathological changes in the cortical
neuron circuits in that blockade of NMDA receptors interrupts the tonic facilitatory
influence of glutamate transmission on the GABAergic interneurons, resulting in an
enhancement of dopaminergic activity by attenuation of the inhibitory GABAergic
transmission on the dopamine neurons (disinhibition) in the frontal cortex (Fig. 1).
In agreement with these observations, an in vivo human study using PET
(positron emission tomography) (Aalto et al. 2005), healthy subjects treated with
ketamine showed decreased radioligand binding to dopamine receptors in the
D-Serine: Basic Aspects with a Focus on Psychosis 503
Given the above data from the animal and human studies, an impaired NMDA
receptor function may elicit positive symptoms by indirectly increasing dopamine
transmission, and the negative and cognitive symptoms by affecting the extra-
dopamine systems in the brain of schizophrenia patients (Fig. 1). This assumption
is supported by observations that the NMDA receptor antagonists, PCP (Luby et al.
1959) and ketamine (Vollenweider et al. 1997), exacerbate all symptom groups
(positive, negative, and cognitive) in schizophrenia patients, whereas the dopamine
agonist, amphetamine (Luby et al. 1959), aggravates only the positive symptoms
(Fig. 1).
From this point of view, it is rationale to postulate that recovery of the NMDA
receptor function by its agonistic drugs may improve both the antipsychotic-respon-
sive positive symptoms and pharmacotherapy-resistant negative and cognitive
symptoms in schizophrenia. To this end, for a method to facilitate the NMDA
receptor function, glycine-binding site agonists (Fig. 2) were selected from agents
acting on various regulatory sites of the glutamate receptor because overstimulation
of the glutamate-binding site, but not the glycine site, has been found to induce
cytotoxicity and convulsions (Johnson and Jones 1990).
Preclinical Studies
Clinical Studies
Since the previously mentioned test drugs have problems, such as poor
peripheral-to-brain transferability (need for excessive peripheral doses), side effects
from long-term continuous use, nonselectivity or low potency of action on the
glycine modulatory site, they are not suitable for long-term clinical use and the
creation of new drugs is awaited. To overcome these concerns, bitopertin, a blood–
brain barrier (BBB)-permeable and selective type I glycine transporter inhibitor, was
developed and large add-on double-blind randomized placebo-controlled trials were
conducted (Bugarski-Kirola et al. 2017). However, they failed to prove any statis-
tically significant improvement in the negative and/or cognitive symptoms of
schizophrenia patients with antipsychotic medication, challenging the NMDA recep-
tor disfunction hypothesis of schizophrenia (Bugarski-Kirola et al. 2017)
Taking into account the lack of the ameliorating effects of the selective glycine
transporter inhibitor, the below data extrapolate that it is more rationale to upregulate
the extracellular concentrations of D-serine as a mean of facilitating the NMDA
receptor function than to increase those of glycine: (i) glycine requires about
30 times the dose of D-serine to improve the schizophrenic symptoms (Goh et al.
2021; Singh and Singh 2011) despite a similar BBB permeability and ED50s for the
GluN1/GluN2 heteromeric NMDA receptor function-enhancing effects between the
two NMDA receptor coagonists (Matsui et al. 1995), (ii) it has been suggested that
D-serine and glycine binds to the synaptic and extrasynaptic NMDA receptors,
respectively (Papouin et al. 2012), (iii) glycine, unlike D-serine, is a selective agonist
for the inhibitory glycine receptor (Danysz and Parsons 1998), (iv) D-serine, but not
bitopertin, amended in schizophrenia patients the impairment of mismatch negativity
that has been indicated to be related to the NMDA receptor dysfunction and
disturbed information processing (Kantrowitz et al. 2016, 2017, 2018).
Since long-term administration of high doses of D-serine itself in the periphery poses
a risk of nephrotoxicity and since D-serine, but not glycine, has been shown to be an
intrinsic coagonist for the NMDA receptor as described in the next section, one of the
future strategies for developing drugs for augmentation of the NMDA receptor function
is to search for their target molecules that regulate the amount of D-serine at the
synapses.
It had generally been believed that D-amino acids and L-sugars do not occur in nature.
The homochirality theories were challenged by the fact that D-amino acid oxidase
(DAO)activity was uncovered in mammalian tissues in 1935 by Krebs et al. This
intriguing discovery suggested the existence of certain D-optimal isomers of amino
D-Serine: Basic Aspects with a Focus on Psychosis 507
acids as substrates for the enzyme. D-Serine is one of the substrates with a high affinity
for this DAO, and was observed in free form in high concentrations in the earthworm
(Rosenberg and Ennor 1960) and silkworm (Srinivasan et al. 1962) tissues. However,
the free D-isomer of serine was not found in mammals for a long time afterward. The
coagonist action of D-serine on the NMDA receptor, thus, had been considered to be
an artificial phenomenon under in vitro experimental conditions until the detection of
endogenous D-serine in mammalian brains in 1992 (Hashimoto et al. 1992a, b). The
presence of D-serine raises the possibility that the metabolic processes of D-serine and
control pathways of synaptic D-serine signaling to the NMDA receptor may be
suitable targets for the creation of agents that ameliorate the schizophrenic symptoms
by positive modulation of the NMDA receptor function (Fig. 1).
The author’s research group first proved the natural occurrence of D-serine in
mammals by using gas chromatography-mass spectrometry in 1992 (Hashimoto
et al. 1992a, b) during the course of investigation in the rat to develop a novel
treatment for schizophrenia with enhancers of the NMDA receptor function includ-
ing D-serine and D-alanine that bind to the NMDA receptor glycine site as men-
tioned in the previous section “NMDA Receptor Function Enhancement Therapy for
Schizophrenia: Rationale for Application of D-Serine” (3.1~3.3). These D-amino
acids are highly polar and do not readily cross the BBB, making them unsuitable for
peripheral administration as therapeutic agents. Therefore, their fatty acid com-
pounds, N-myristoyl-D-serine and N-myristoyl-D-alanine, were prepared and their
efficacy as potential therapeutic agents for schizophrenia was indicated by their anti-
PCP properties in the rat (Tanii et al. 1991). The glycine site antagonist-reversible
nature of their anti-PCP properties and the lack of their direct binding to the glycine
site (Tanii et al. 1991) pointed out the plausibility that the fatty acid compounds
could liberate free D-serine and D-alanine in the brain, which stimulate the glycine
site. To examine this assumption, the author’s research group established the gas-
chromatgraphic method and HPLC with fluorometric detection for separation and
quantitation of chiral amino acids and discovered the presence of a large amount of
D-serine in the forebrain regions of the control rats without peripheral injections of
(0.2~0.3 μ mol/g wet weight) (Hashimoto et al. 1992a, b, 1993a, b).
The concern that D-serine could be an artifact product by nonspecific procedures
(e.g., nonenzymatic conversion of L- to D-serine during sample manipulation) may
be negated because: (i) the D-serine concentrations vary markedly among tissues and
brain regions or in the same brain portion at different developmental stages, and
(ii) the measured values are alike even though sample processing differs among the
detection methods (Hashimoto et al. 1992a, b, 1993a, b; Nishikawa et al. 1994).
Endogenous D-serine is enriched in the brain and at low or trace levels in the
peripheral organs and blood (Hashimoto et al. 1993b). In the brain of the adult
period, the concentrations of D-serine are high in the forebrain portions including the
508 T. Nishikawa et al.
cerebral cortex, hippocampus, striatum, and limbic system (containing the olfactory
tubercle, nucleus accumbens, and septum), moderate in the diencephalon and mid-
brain, and marginal in the pons-medulla and cerebellum (Hashimoto et al. 1993b).
The uneven distribution of D-serine is closely correlated with those of the radio-
ligand binding densities to L-glutamate, glycine, and PCP site of the NMDA
receptor (Hashimoto et al. 1993b). In contrast, the D-serine levels in each brain
region are almost homogeneous at birth (Hashimoto et al. 1995). The developmental
changes in the distribution patterns of D-serine are especially similar to those of the
GRIN2B mRNA expression in the brain (Hashimoto et al. 1993b, 1995, Nishikawa
2011). Together with the fact that D-serine is a selective agonist of the glycine
regulatory site (Fig. 2) that is indispensable for NMDA receptor activation, the very
similar distribution patterns throughout postnatal development between D-serine and
NMDA receptors agrees with the prediction that D-serine is an intrinsic coagonist of
the NMDA receptors.
Biosynthesis of D-Serine
In mammalian brains, endogenous D-serine has been considered not to originate from
enteric bacteria because the tissue D-serine concentrations are similar between normal
and germ-free rats (Hashimoto et al. 1993b). Early studies of D-serine synthesis
clarified in the rat that (i) the intraperitoneal administration of L-serine elevated the
D-serine level in the cerebral neocortex in 8-day-old rats, and vice versa (Takahashi
et al. 1997), and (ii) in the experiments using radioactive tracers, brain D-serine was
converted from L-serine applied intraventricularly whereas D-glucose and glycine
administered centrally or L-serine loaded peripherally failed to become D-serine
(Dunlop and Neidle 1997). These phenomena suggest that serine racemase (SRR),
which is involved in the interconversion of D- and L-serine, is present and important
for the D-serine synthesis (Fig. 2). Subsequently, the SRR protein and its cDNA were
identified and isolated by Wolosker et al. (1999a, b). Four types of prepared SRR gene-
knockout mice consistently display that the D-serine levels in various forebrain brain
regions decreased to 9~22 % of the level in the corresponding wild-type mice (Basu
et al. 2009; Horio et al. 2011; Labrie et al. 2009; Miyoshi et al. 2012), indicating that
SRR is the main D-serine-synthesizing enzyme in the D-serine-rich brain areas. SRR
is predominantly localized in the neurons under physiological conditions (Balu et al.
2014; Miya et al. 2008). However, brain injury by controlled cortical impact causes
enhanced expression of SRR in the astroglia (Perez et al. 2017).
The D-serine concentrations in both tissues and extracellular fluid are drastically
reduced in parallel in the SRR knockout mice (Horio et al. 2011) and the distribution
pattern of SRR is similar to that of the GluN2B subunit of the NMDA receptor
throughout life (Miya et al. 2008). These observations support the idea that drugs
that increase the SRR activity could augment the NMDA receptor function by
elevating the extracellular D-serine levels in the brain.
On the other hand, in the brain of mice in which D-3-phosphoglycerate
dehydrogenase, involved in the L-serine synthesis, was selectively knocked out
D-Serine: Basic Aspects with a Focus on Psychosis 509
in the glial fibrillary acidic protein (GFAP)-expressing cells (mainly astroglia), the
brain D-serine concentrations markedly decreased with a depletion in the L-serine
levels (Yang et al. 2010). This finding accords the view that D-serine in the brain is
dependent on L-serine supplied by the GFAP-expressing cells. The glycine cleav-
age system may also participate in the regulation and/or synthesis of brain
D-serine because the neocortical D-serine concentrations were downregulated in
the patients with non-ketotic hyperglycinemia deficient in the cleavage system
activity and rats treated with the inhibitor of the glycine degrading system (Iwama
et al. 1997). Consequently, facilitatory agents for the D-3-phosphoglycerate dehy-
drogenase and glycine cleavage system could be candidates as enhancers of the
NMDA receptor function by their possible increasing effects on the extracellular
D-serine levels.
Storage of D-Serine
Extracellular D-Serine
Uptake of D-Serine
Degradation of D-Serine
Degradation of brain D-serine by D-amino acid oxidase (DAO) has been demon-
strated by the experiments revealing that the cerebellar tissue concentrations of D-
serine and D-alanine, which are proved to be stereoselective substrates for DAO
(Konno and Yasumura 1984), markedly increase in mature mutant mice with
deficiency of the enzyme activity (Hashimoto et al. 1993a) (Fig. 2). DAO was
514 T. Nishikawa et al.
Not only the GluN1/G luN2A~2D heteromeric NMDA receptor but also the GluD2
glutamate receptor (Naur et al. 2007; Kakegawa et al. 2011) and GluN1/
GluN3A~3B type NMDA receptor (Danysz and Parsons 1998; Pérez-Otaño et al.
2016) have been shown to harbor physiological binding sites of intrinsic D-serine
with KD values in the 107 and 104 M ranges, respectively. The GluD2 receptor-D-
serine interaction has been demonstrated to regulate the cerebellar LTD and motor
coordination in mice (Kakegawa et al. 2011). The GluN1/GluN3 receptors are
manifested to take part in control of the LTP formation and synaptic maturation
although the exact electrophysiological properties of action of D-serine on the
glutamate-insensitive receptors are still elusive due to the reports that they are
excitatory, inhibitory, or have no significant influence (Pérez-Otaño et al. 2016).
On the basis of these data, D-serine could be expected to display therapeutic
efficacies on certain cognitive deficits in schizophrenia by the regulation of glutamate
D-Serine: Basic Aspects with a Focus on Psychosis 515
synaptic plasticity through the actions on the GluD2 and GluN1/GluN3 receptors.
To obtain evidence for these relationships requires further experimental and clinical
investigations.
The coagonist property of D-serine intimates that the assumed NMDA receptor dys-
function in schizophrenia could be caused by reduced D-serine signaling in the brain. In
support of this idea, a recent GWAS study on big data has disclosed a significant
association with schizophrenia of the chromosomal locus encoding the SRR gene
(Schizophrenia Working Group of the Psychiatric Genomics Consortium 2014).
Indeed, SRR gene knockout mice, in which the tissue and extracellular D-serine
levels markedly decreased in the brain, exhibit various changes serving as schizo-
phrenia models. These changes include the NMDA receptor hypofunction as
revealed by a diminished inward current and LTP formation via the receptor,
augmentation of NMDA receptor blockade-induced abnormal behavior, and distur-
bance of prepulse inhibition (Basu et al. 2009; Labrie et al. 2009). Furthermore, the
electrophysiological and behavioral impairments, at least in part, were improved by
D-serine or antipsychotic drugs.
In the postmortem brains of schizophrenia patients, however, no studies have so far
reported significant changes in the tissue D-serine concentrations in the prefrontal,
temporal, and parietal cortical areas (Kumashiro et al. 1995; Bendikov et al. 2007).
The expression amounts of proteins and mRNAs for the D-serine metabolic enzymes,
SRR and DAO, in the postmortem brain tissues are controversial among research
groups (Hu et al. 2015; Bendikov et al. 2007; Keller et al. 2018; Steffek et al. 2006).
The blood and cerebrospinal fluid D-serine levels in schizophrenia are also debatable
matters (Brouwer et al. 2013; Cho et al. 2016). Some investigations have cast doubt on
the significance of the blood D-serine levels as an index of the brain D-serine
pathology. The D-serine concentrations in the blood have been demonstrated not to
reflect those in the brain tissues by the experiments that brain-specific depletion of the
L-serine synthesizing enzyme resulted in a marked decrease in brain, but not blood,
D-serine levels in the adult genetically modified mice (Yang et al. 2010). Deletion of
the SRR gene yields a drastic reduction in the D-serine levels in various brain portions
without affecting those in the serum of the mouse (Miyoshi et al. 2012)
Despite the lack of direct evidence for D-serine signal deficit in biological samples
of schizophrenia patients, it is of interest to note that an increase in the density of the
NMDA receptor glycine-binding site, on which D-serine selectively acts, was observed
in various regions, such as the somatosensory and prefrontal cortices, supramarginal
and angular gyri, and visual cortex of postmortem schizophrenia brains (Ishimaru et al.
1994) because this phenomenon could be a consequence of the compensatory
upregulation of the NMDA receptor to inadequate D-serine signaling. This assumption
is in line with the recent PET observations indicating the reduced frequency of D-
serine-induced channel opening of the NMDA receptor that the binding ability of [123I]
CNS-1261, a ligand selectively binding to the PCP site of the NMDA receptor ion
516 T. Nishikawa et al.
Conclusions
Attenuated transmission via the NMDA receptor in the brain of patients with
schizophrenia has been corroborated by the previous observations that
schizophrenia-like positive, negative, and cognitive symptoms are caused by com-
petitive and noncompetitive NMDA receptor antagonists, and anti-NMDA receptor
autoantibodies with internalizing NMDA receptors from the cell surface. This
concept is further supported by the hypersensitivity of schizophrenia patients to
the psychotomimetic action of the NMDA receptor antagonist as compared to
healthy volunteers, suggesting reduced basal activity levels of the NMDA receptor
in the patients. Moreover, the NMDA receptor blockade results in overactivation of
the cortico-limbic dopaminergic neurons that have been considered to produce
antipsychotic-sensitive positive symptoms in an NMDA receptor glycine site
agonist-reversible manner. Taken together, functional restoration of the NMDA
receptor by stimulation of its glycine site has been expected to improve not only
antipsychotic-responsive positive symptoms but also antipsychotic-resistant nega-
tive and cognitive symptoms in schizophrenia patients. Among the glycine site
agonists, the intrinsic NMDA receptor coagonists glycine and D-serine have been
shown to ameliorate the negative symptoms in the meta-analyses of relatively large-
scale data of add-on clinical trials of each amino acid in schizophrenia patients
treated with conventional antipsychotic drugs except clozapine. The greater potent
therapeutic efficacy of D-serine than glycine and lack of significant palliation by a
glycine transporter inhibitor in the RCT raise the future direction for the develop-
ment of NMDA receptor potentiating therapy targeting molecules that participate in
the regulation of the metabolic pathways and synaptic concentrations of D-serine.
Cross-References
References
Aalto S, Ihalainen J, Hirvonen J, Kajander J, Scheinin H, Tanila H, et al. Cortical glutamate-
dopamine interaction and ketamine-induced psychotic symptoms in man. Psychopharmacology
(Berl). 2005;182:375–83.
American Psychiatric Association. Schizophrenia spectrum and other psychotic disorders. In:
American Psychiatric Association, editor. Diagnostic and statistical manual of mental disorders
(DSM-5 ®). 5th ed. Washington, DC: Amer Psychiatric Pub Inc; 2013. p. 87–122.
D-Serine: Basic Aspects with a Focus on Psychosis 517
Anis NA, Berry SC, Burton NR, Lodge D. The dissociative anaesthetics, ketamine and phencycli-
dine, selectively reduce excitation of central mammalian neurones by N-methyl-aspartate. Br J
Pharmacol. 1983;79:565–575.
Balu DT, Takagi S, Puhl MD, Benneyworth MA, Coyle JT. D-serine and serine racemase are localized
to neurons in the adult mouse and human forebrain. Cell Mol Neurobiol. 2014;34:419–35.
Basu AC, Tsai GE, Ma CL, Ehmsen JT, Mustafa AK, Han L, et al. Targeted disruption of serine
racemase affects glutamatergic neurotransmission and behavior. Mol Psychiatry. 2009;14:
719–27.
Belforte JE, Zsiros V, Sklar ER, Jiang Z, Yu G, Li Y, et al. Postnatal NMDA receptor ablation in
corticolimbic interneurons confers schizophrenia-like phenotypes. Nat Neurosci. 2010;13:76–83.
Bendikov I, Nadri C, Amar S, Panizzutti R, De Miranda J, Wolosker H, et al. A CSF and
postmortem brain study of D-serine metabolic parameters in schizophrenia. Schizophr Res.
2007;90:41–51.
Benneyworth MA, Li Y, Basu AC, Bolshakov VY, Coyle JT. Cell selective conditional null
mutations of serine racemase demonstrate a predominate localization in cortical glutamatergic
neurons. Cell Mol Neurobiol. 2012;32:613–24.
Bodner O, Radzishevsky I, Foltyn VN, Touitou A, Valenta AC, Rangel IF, et al. D-serine signaling
and NMDAR-mediated synaptic plasticity are regulated by system A-type of glutamine/D-
serine dual transporters. J Neurosci. 2020;40:6489–502.
Bouwmans C, de Sonneville C, Mulder CL, Hakkaart-Van RL. Employment and the associated
impact on quality of life in people diagnosed with schizophrenia. Neuropsychiatr Dis Treat.
2015;11:2125–42.
Brouwer A, Luykx JJ, van Boxmeer L, Bakker SC, Kahn RS. NMDA-receptor coagonists in serum,
plasma, and cerebrospinal fluid of schizophrenia patients: a meta-analysis of case-control studies.
Neurosci Biobehav Rev. 2013 Sep;37(8):1587–96. https://doi.org/10.1016/j.neubiorev.2013.
Bugarski-Kirola D, Blaettler T, Arango C, Fleischhacker WW, Garibaldi G, Wang A, et al.
Bitopertin in negative symptoms of schizophrenia-results from the phase III FlashLyte and
DayLyte studies. Biol Psychiatry. 2017;82:8–16.
Castillo-Gómez E, Oliveira B, Tapken D, Bertrand S, Klein-Schmidt C, Pan H et al. All naturally
occurring autoantibodies against the NMDA receptor subunit NR1 have pathogenic potential
irrespective of epitope and immunoglobulin class. Mol Psychiatry. 2017;22:1776–1784.
Chang CH, Lin CH, Liu CY, Chen SJ, Lane HY. Efficacy and cognitive effect of sarcosine
(N-methylglycine) in patients with schizophrenia: a systematic review and meta-analysis of
double-blind randomised controlled trials. J Psychopharmacol. 2020;34:495–505.
Chen CH, Cheng MC, Liu CM, Liu CC, Lin KH, Hwu HG. Seroprevalence survey of selective anti-
neuronal autoantibodies in patients with first-episode schizophrenia and chronic schizophrenia.
Schizophr Res. 2017;190:28–31.
Cho SE, Na KS, Cho SJ, Kang SG. Low d-serine levels in schizophrenia: a systematic review and
meta-analysis. Neurosci Lett. 2016;634:42–51.
Chumakov I, Blumenfeld M, Guerassimenko O, Cavarec L, Palicio M, Abderrahim H, et al. Genetic
and physiological data implicating the new human gene G72 and the gene for D-amino acid
oxidase in schizophrenia. Proc Natl Acad Sci U S A. 2002;99:13675–80.
Connel PH. Amphetamine Psychosis. Maudsley monographs number five. London: Chapman &
Hall Ltd; 1958. p. 123.
Contreras PC. D-serine antagonized phencyclidine- and MK-801-induced stereotyped behavior and
ataxia. Neuropharmacology. 1990;29:291–3.
Curcio L, Podda MV, Leone L, Piacentini R, Mastrodonato A, Cappelletti P, et al. Reduced D-serine
levels in the nucleus accumbens of cocaine-treated rats hinder the induction of NMDA receptor-
dependent synaptic plasticity. Brain. 2013;136:1216–30.
Dalmau J, Tüzün E, Wu HY, Masjuan J, Rossi JE, Voloschin A, et al. Paraneoplastic anti-N-methyl-D-
aspartate receptor encephalitis associated with ovarian teratoma. Ann Neurol. 2007;61:25–36.
Dalmau J, Armangué T, Planagumà J, Radosevic M, Mannara F, Leypoldt F, et al. An update on
anti-NMDA receptor encephalitis for neurologists and psychiatrists: mechanisms and models.
Lancet Neurol. 2019;18:1045–57.
518 T. Nishikawa et al.
Kidd LR, Lyons SC, Lloyd G. Paediatric procedural sedation using ketamine in a UK emergency
department: a 7 year review of practice. Br J Anaesth. 2016;116:518–23.
Konno R, Yasumura Y. Brain and kidney D-amino acid oxidases are coded by a single gene in the
mouse. J Neurochem. 1984;42:584–6.
Krebs HA. Metabolism of amino-acids: The synthesis of glutamine from glutamic acid and ammonia,
and the enzymic hydrolysis of glutamine in animal tissues. Biochem J. 1935;29:1951–69.
Kumashiro S, Hashimoto A, Nishikawa T. Free D-serine in post-mortem brains and spinal cords of
individuals with and without neuropsychiatric diseases. Brain Res. 1995;681:117–25.
Kuppili PP, Menon V, Sathyanarayanan G, Sarkar S, Andrade C. Efficacy of adjunctive
D-Cycloserine for the treatment of schizophrenia: a systematic review and meta-analysis of
randomized controlled trials. J Neural Transm (Vienna). 2021;128:253–62.
Labrie V, Fukumura R, Rastogi A, Fick LJ, Wang W, Boutros PC, et al. Serine racemase is
associated with schizophrenia susceptibility in humans and in a mouse model. Hum Mol
Genet. 2009;18:3227–43.
Lahti AC, Holcomb HH, Gao X-M, Tamminga CA. NMDA-sensitive glutamate antagonism: a
human model for psychosis. Neuropsychopharmacol. 1999;21:S158–69.
Lane HY, Huang CL, Wu PL, Liu YC, Chang YC, Lin PY, et al. Glycine transporter I inhibitor,
N-methylglycine (sarcosine), added to clozapine for the treatment of schizophrenia. Biol
Psychiatry. 2006;60:645–9.
Lane HY, Lin CH, Green MF, Hellemann G, Huang CC, Chen PW, et al. Add-on treatment of
benzoate for schizophrenia: a randomized, double-blind, placebo-controlled trial of D-amino
acid oxidase inhibitor. JAMA Psychiatry. 2013;70:1267–75.
Laruelle M. Imaging dopamine transmission in schizophrenia. A review and meta-analysis. Q J
Nucl Med. 1998;42:211–21.
Le Bail M, Martineau M, Sacchi S, Yatsenko N, Radzishevsky I, Conrod S, et al. Identity of the
NMDA receptor coagonist is synapse specific and developmentally regulated in the hippocam-
pus. Proc Natl Acad Sci U S A. 2015;112:E204–13.
Lin CH, Lin CH, Chang YC, Huang YJ, Chen PW, Yang HT, et al. Sodium benzoate, a D-amino
acid oxidase inhibitor, added to clozapine for the treatment of schizophrenia: a randomized,
double-blind, placebo-controlled trial. Biol Psychiatry. 2018;84:422–32.
Luby ED, Cohen BD, Rosenbaum G, Gottlieb JS, Kelley R. Study of a new schizophrenomimetic
drug; sernyl. AMA Arch Neurol Psychiatry. 1959;81:363–9.
Malhotra AK, Pinals DA, Adler CM, Elman I, Clifton A, Pickar D, et al. Ketamine-induced
exacerbation of psychotic symptoms and cognitive impairment in neuroleptic-free schizo-
phrenics. Neuropsychopharmacology. 1997;17:141–50.
Martineau M, Shi T, Puyal J, Knolhoff AM, Dulong J, Gasnier B, et al. Storage and uptake of
D-serine into astrocytic synaptic-like vesicles specify gliotransmission. J Neurosci. 2013;33:
3413–23.
Matsui T, Sekiguchi M, Hashimoto A, Tomita U, Nishikawa T, Wada K. Functional comparison of
D-serine and glycine in rodents: the effect on cloned NMDA receptors and the extracellular
concentration. J Neurochem. 1995;65:454–8.
Matsuura A, Fujita Y, Iyo M, Hashimoto K. Effects of sodium benzoate on pre-pulse inhibition
deficits and hyperlocomotion in mice after administration of phencyclidine. Acta
Neuropsychiatr. 2015;27:159–67.
McCutcheon RA, Reis Marques T, Howes OD. Schizophrenia – an overview. JAMA Psychiatry.
2020;77:201–10.
Miya K, Inoue R, Takata Y, Abe M, Natsume R, Sakimura K, et al. Serine racemase is predomi-
nantly localized in neurons in mouse brain. J Comp Neurol. 2008;510:641–54.
Miyoshi Y, Konno R, Sasabe J, Ueno K, Tojo Y, Mita M, et al. Alteration of intrinsic amounts of
D-serine in the mice lacking serine racemase and D-amino acid oxidase. Amino Acids. 2012;43:
1919–31.
Moghaddam B, Adams BW. Reversal of phencyclidine effects by a Group II metabotropic
glutamate receptor agonist in rats. Science. 1998;281:1349–52.
D-Serine: Basic Aspects with a Focus on Psychosis 521
Mothet JP, Parent AT, Wolosker H, Brady RO Jr, Linden DJ, Ferris CD et al. D-serine is an
endogenous ligand for the glycine site of the N-methyl-D-aspartate receptor. Proc Natl Acad Sci
U S A. 2000;97:4926–4931.
Naur P, Hansen KB, Kristensen AS, Dravid SM, Pickering DS, Olsen L, et al. Ionotropic
glutamate-like receptor delta2 binds D-serine and glycine. Proc Natl Acad Sci U S A.
2007;104:14116–21.
Neame S, Safory H, Radzishevsky I, Touitou A, Marchesani F, Marchetti M, et al. The NMDA
receptor activation by d-serine and glycine is controlled by an astrocytic Phgdh-dependent
serine shuttle. Proc Natl Acad Sci U S A. 2019;116:20736–42.
Nishijima K, Kashiwa A, Nishikawa T. Preferential stimulation of extracellular release of dopamine
in rat frontal cortex to striatum following competitive inhibition of the N-methyl-D-aspartate
receptor. J Neurochem. 1994;63:375–8.
Nishijima K, Kashiwa A, Hashimoto A, Iwama H, Umino A, Nishikawa T. Differential effects of
phencyclidine and methamphetamine on dopamine metabolism in rat frontal cortex and striatum
as revealed by in vivo dialysis. Synapse. 1996;22:304–12.
Nishikawa T. Analysis of free D-serine in mammals and its biological relevance. J Chromatogr
B Analyt Technol Biomed Life Sci. 2011;879:3169–83.
Nishikawa T, Hashimoto A, Tanii Y, Umino A, Kashiwa A, Kumashiro S, et al. Disturbed
neurotransmission via the N-methyl-D-aspartate receptor and schizophrenia. In: Moroji T,
Yamamoto K, editors. The biology of schizophrenia. Development of psychiatry series. Amster-
dam: Elsevier; 1994. p. 197–207.
Papouin T, Ladépêche L, Ruel J, Sacchi S, Labasque M, Hanini M, et al. Synaptic and extrasynaptic
NMDA receptors are gated by different endogenous coagonists. Cell. 2012;150:633–46.
Perez EJ, Tapanes SA, Loris ZB, Balu DT, Sick TJ, Coyle JT, et al. Enhanced astrocytic d-serine
underlies synaptic damage after traumatic brain injury. J Clin Invest. 2017;127:3114–25.
Pérez-Otaño I, Larsen RS, Wesseling JF. Emerging roles of GluN3-containing NMDA receptors in
the CNS. Nat Rev Neurosci. 2016;17:623–35.
Petersen RC, Stillman RC, editors. Phencyclidine (PCP) abuse: an appraisal. National Institute on
Drug Abuse Research Monographs, superintendent of documents. Washington, DC: U. S.
Government Printing Office; 1978. p. 313.
Pilowsky LS, Bressan RA, Stone JM, Erlandsson K, Mulligan RS, Krystal JH, et al. First in vivo
evidence of an NMDA receptor deficit in medication-free schizophrenic patients. Mol Psychi-
atry. 2006;11:118–9.
Plitman E, Iwata Y, Caravaggio F, Nakajima S, Chung JK, Gerretsen P, et al. Kynurenic acid in
schizophrenia: a systematic review and meta-analysis. Schizophr Bull. 2017;43:764–77.
Potkin SG, Jin Y, Bunney BG, Costa J, Gulasekaram B. Effect of clozapine and adjunctive high-
dose glycine in treatment-resistant schizophrenia. Am J Psychiatry. 1999;156:145–7.
Rao TS, Kim HS, Lehmann J, Martin LL, Wood PL. Interactions of phencyclidine receptor
agonist MK-801 with dopaminergic system: regional studies in the rat. J Neurochem. 1990;54:
1157–62.
Reich DL, Silvay G. Ketamine: an update on the first twenty-five years of clinical experience. Can
J Anaesth. 1989;36:186–97.
Rosenberg H, Ennor AH. Occurrence of free D-serine in the earthworm. Nature. 1960;187:617–8.
Sato D, Umino A, Kaneda K, Takigawa M, Nishikawa T. Developmental changes in distribution
patterns of phencyclidine-induced c-Fos in rat forebrain. Neurosci Lett. 1997;239:21–4.
Scheer S, John RM. Anti-N-methyl-D-aspartate receptor encephalitis in children and adolescents.
J Pediatr Health Care. 2016;30:347–58.
Schell MJ, Molliver ME, Snyder SH. D-Serine, an endogenous synaptic modulator: localization to
astrocytes and glutamate-stimulated release. Proc Natl Acad Sci U S A. 1995;92:3948–52.
Schizophrenia Working Group of the Psychiatric Genomics Consortium. Biological insights from
108 schizophrenia-associated genetic loci. Nature. 2014;511:421–7.
Schwartz RH, Einhorn A. PCP intoxication in seven young children. Pediatr Emerg Care. 1986;2:
238–41.
522 T. Nishikawa et al.
Sershen H, Hashim A, Dunlop DS, Suckow RF, Cooper TB, Javitt DC. Modulating NMDA
receptor function with D-amino acid oxidase inhibitors: understanding functional activity in
PCP-treated mouse model. Neurochem Res. 2016;41:398–408.
Shimazu D, Yamamoto N, Umino A, Ishii S, Sakurai S, Nishikawa T. Inhibition of D-serine
accumulation in the Xenopus oocyte by expression of the rat ortholog of human 30 -phosphoa-
denosine 50 -phosphosulfate transporter gene isolated from the neocortex as D-serine modulator-
1. J Neurochem. 2006;96:30–42.
Singh SP, Singh V. Meta-analysis of the efficacy of adjunctive NMDA receptor modulators in
chronic schizophrenia. CNS Drugs. 2011;25:859–85.
Srinivasan NG, Corrigan JJ, Meister A. D-Serine in the blood of the silkworm Bombyx mori and
other lepidoptera. J Biol Chem. 1962;237:3844–5.
Steffek AE, Haroutunian V, Meador-Woodruff JH. Serine racemase protein expression in cortex and
hippocampus in schizophrenia. Neuroreport. 2006;17:1181–5.
Steiner J, Walter M, Glanz W, Sarnyai Z, Bernstein HG, Vielhaber S, et al. Increased prevalence of
diverse N-methyl-D-aspartate glutamate receptor antibodies in patients with an initial diagnosis
of schizophrenia: specific relevance of IgG NR1a antibodies for distinction from N-methyl-D-
aspartate glutamate receptor encephalitis. JAMA Psychiatry. 2013;70:271–8.
Takahashi K, Hayashi F, Nishikawa T. In vivo evidence for the link between L- and D-serine
metabolism in rat cerebral cortex. J Neurochem. 1997;69:1286–90.
Tanii Y, Nishikawa T, Hashimoto A, Takahashi K. Stereoselective inhibition by D- and L-alanine of
phencyclidine-induced locomotor stimulation in the rat. Brain Res. 1991;563:281–4.
Tanii Y, Nishikawa T, Hashimoto A, Takahashi K. Stereoselective antagonism by enantiomers of
alanine and serine of phencyclidine-induced hyperactivity, stereotypy and ataxia in the rat.
J Pharmacol Exp Ther. 1994;269:1040–8.
Tanii Y, Nishikawa T, Hibino H, Takahashi K. Effects of allosteric agonists for N-methyl-D-
aspartate receptor and their derivatives on phencyclidine-induced abnormal behavior in the
rat. Brain Sci Mental Disord (Present: Jpn J Biol Psychiatry), 1991b; 2:497–502 (in Japanese
with English abstract). Erratum in: Jpn J Biol Psychiatry. 2010;21:126.
Tsai GE, Lin PY. Strategies to enhance N-methyl-D-aspartate receptor-mediated neurotransmission
in schizophrenia, a critical review and meta-analysis. Curr Pharm Des. 2010;16:522–37.
Tsai GE, Yang P, Chung LC, Tsai IC, Tsai CW, Coyle JT. D-serine added to clozapine for the
treatment of schizophrenia. Am J Psychiatry. 1999;156:1822–5.
Umino A, Takahashi K, Nishikawa T. Characterization of the phencyclidine-induced increase in
prefrontal cortical dopamine metabolism in the rat. Br J Pharmacol. 1998;124:377–85.
Umino A, Ishiwata S, Iwama H, Nishikawa T. Evidence for tonic control by the GABAA receptor of
extracellular D-Serine concentrations in the medial prefrontal cortex of rodents. Front Mol
Neurosci. 2017;10:240.
Umino M, Umino A, Nishikawa T. Effects of selective calcium-permeable AMPA receptor block-
ade by IEM 1460 on psychotomimetic-induced hyperactivity in the mouse. J Neural Transm
(Vienna). 2018;125:705–11.
Vollenweider FX, Leenders KL, Oye I, Hell D, Angst J. Differential psychopathology and patterns
of cerebral glucose utilisation produced by (S)- and (R)-ketamine in healthy volunteers using
positron emission tomography (PET). Eur Neuropsychopharmacol. 1997;7:25–38.
Weimar WR, Neims AH. The development of D-amino acid oxidase in rat cerebellum.
J Neurochem. 1977;29:649–56.
Welch MJ, Correa GA. PCP intoxication in young children and infants. Clin Pediatr (Phila).
1980;19:510–4.
White PF, Way WL, Trevor AJ. Ketamine – its pharmacology and therapeutic uses. Anesthesiology.
1982;56:119–36.
Wolosker H, Blackshaw S, Snyder SH. Serine racemase: a glial enzyme synthesizing D-serine to
regulate glutamate-N-methyl-D-aspartate neurotransmission. Proc Natl Acad Sci U S A. 1999a;96:
13409–14.
D-Serine: Basic Aspects with a Focus on Psychosis 523
Wolosker H, Sheth KN, Takahashi M, Mothet JP, Brady RO Jr, Ferris CD, et al. Purification of serine
racemase: biosynthesis of the neuromodulator D-serine. Proc Natl Acad Sci U S A. 1999b;96:721–5.
Wong JM, Folorunso OO, Barragan EV, Berciu C, Harvey TL, Coyle JT, et al. Postsynaptic serine
racemase regulates NMDA receptor function. J Neurosci. 2020;40:9564–75.
Yagi K, Nagatsu T, Ozawa T. Inhibitory action of chlorpromazine on the oxidation of d-amino-acid
in the diencephalon part of the brain. Nature. 1956;177:891–2.
Yamamoto N, Tomita U, Umino A, Nishikawa T. Uptake of D-serine by synaptosomal P2 fraction
isolated from rat brain. Synapse. 2001;42:84–6.
Yang Y, Ge W, Chen Y, Zhang Z, Shen W, Wu C, et al. Contribution of astrocytes to hippocampal long-
term potentiation through release of D-serine. Proc Natl Acad Sci U S A. 2003;100:15194–9.
Yang JH, Wada A, Yoshida K, Miyoshi Y, Sayano T, Esaki K, et al. Brain-specific Phgdh deletion
reveals a pivotal role for L-serine biosynthesis in controlling the level of D-serine, an N-methyl-
D-aspartate receptor co-agonist, in adult brain. J Biol Chem. 2010;285:41380–90.
Yonezawa Y, Kuroki T, Kawahara T, Tashiro N, Uchimura H. Involvement of gamma-aminobutyric
acid neurotransmission in phencyclidine-induced dopamine release in the medial prefrontal
cortex. Eur J Pharmacol. 1998;341:45–56.
Amine Precursors in Depressive Disorders
and the Blood-Brain Barrier
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 526
Biogenic Amines and Depression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 528
Serotonin Precursor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 529
Dopamine Precursor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 530
Histamine Precursor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 530
Pathophysiology of Depression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 531
Traumatic Brain Injury and Depression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 531
Concussive Head Injury and Depression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 532
Depression in Military . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 533
Depression in Alzheimer’s Disease . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 534
Depression in Parkinson’s Disease . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 534
Psychostimulant Abuse and Depression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 535
Stress and Depression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 536
Neurotrophins and Depression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 537
BDNF and Serotonin Interaction in the CNS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 538
Cytokines and Depression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 538
Unifying Concepts of Depression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 539
Monoamines and Neurotrophin Interaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 539
Monoamine and Cytokine Interaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 540
Blood-Brain Barrier Is the Gateway of Depression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 541
Conclusion and Future Perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 548
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 548
Abstract
Depression caused by either genetic factors or environmental stimulates such as
chronic psychological or physiological stress, traumatic or sports-related brain
injuries, neurodegenerative diseases, and/or substance abuse and drug
H. S. Sharma (*) · A. Sharma
International Experimental Central Nervous System Injury & Repair (IECNSIR), Department of
Surgical Sciences, Anesthesiology & Intensive Care Medicine, Uppsala University Hospital,
Uppsala University, Uppsala, Sweden
e-mail: sharma@surgsci.uu.se; aruna.sharma@surgsci.uu.se
Introduction
Depression is one of the most severe diseases resulting in mortality and morbidity
among population (Blazer 2003). A rough estimate suggests that about 40 k people
die every year in the USA caused by either self-inflicted injury or suicide related to
depressive illness (Read et al. 2017). At the time of suicidal activity in depressive
illness, about 30% of victims are on antidepressive treatments (Brent 2016). This
suggests that treatment strategies for depression prescribed so far are still not quite
effective in containing the disease. In spite of recent progresses in depressive
therapy, the basic mechanisms or causes of depressive illnesses are not well
known. Thus, new avenues for treating depression beyond current strategies of
therapeutic advice are needed.
Available evidences now suggest that depression is not caused by any single
neurotransmitter such as serotonin or a group of closely related neurochemicals
like biogenic amines comprising dopamine, norepinephrine, epinephrine, and hista-
mine metabolism in the brain (Birkmayer et al. 1972; Ménard et al. 2016). There are
reasons to believe that several other agents and factors also contribute to the
mechanism of depressive illness. Thus, involvement of cytokines such as tumor
Amine Precursors in Depressive Disorders and the Blood-Brain Barrier 527
Several lines of evidences suggest that in most of these conditions, the BBB is
compromised (Sharma 2004a). Thus, the need of the hour is to restore the BBB
function in depression to maintain a healthy brain and mind (Dudek et al. 2020;
Shalev et al. 2009).
In clinics, depression is treated with several biogenic amine precursor agents that
seem to control the episode of depression for some time but not completely (Hamon
and Blier 2013). Thus, the current treatment strategy using amine precursors in
depression requires further investigation. Alternatively, the root cause of depressive
illnesses requires to be investigated in detail so that development of suitable thera-
peutic strategies to treat depression may be worked out more effectively.
This review discusses the role of amine pressures in depression in relation to the
BBB disturbances. In addition, other possible therapeutic angles of depression are
explored to develop suitable strategies to treat the disease more effectively in clinics.
the biogenic amines in the brain of depressive patients is the administration of amine
precursors that after entering into the brain endogenously converted into amines.
That amine precursors are working to elevate the amine in question physiologically
inside the brain is supported by the success of catecholamine precursor
l-dihydroxyphenylalanine (levodopa) in successfully treating Parkinson’s disease
(PD) (Birkmayer and Hornykiewicz 1961). This finding has revolutionized interest
in amine precursors as potential treatment in depressive illnesses.
Serotonin Precursor
L-Tryptophan
Results from several investigators suggest that L-tryptophan treatment is very
effective as an antidepressant in comparison to imipramine hydrochloride or elec-
troconvulsive therapy (Coppen et al. 1967, 1972; Bowers Jr. 1970; Kline and Shah
1973). However, large doses of L-tryptophan were found ineffective in depression
(Carroll et al. 1970; Bunney Jr et al. 1971; Gayford et al. 1973; Murphy et al. 1973;
Mendlewicz and Youdim 1980).
Studies further show that L-tryptophan potentiated the antidepressant effects of
MAOI as compared to MAOI alone (Coppen 1967; Pare and Sandler 1959; Pare
1963; Coppen et al. 1963; Gutierrez and Alino 1971). However, the combination of
L-5hydroxytryptamine with tricyclic antidepressant was ineffective as compared to
tricyclic alone (Pare 1973; Glassman and Platman 1969; Shaw et al. 1972). Inter-
estingly, L-tryptophan exerted anti-manic effects in two double-blind investigations,
indicating the interrelationship between mania and depression (Whybrow et al.
1969; Mendels and Stinnett 1973).
5-Hydroxytryptophan
The use of the immediate serotonin precursor 5-hydroxytryptophan (5-HTP) is
recommended for the treatment of depression either alone or in combination with
MAOI (Pare and Sandler 1959; Kline and Sacks 1963; Kline et al. 1964; Persson
and Roos 1967; Sano 1972; Brodie et al. 1973). However, it appears that 5-HTP is
definitely superior than placebo regarding its antidepressant effects; the effects are
weak on rating scales of mood elevation and psychiatric rated depression scale
(Mendels et al. 1975). Interestingly, when probenecid was given before 5-HTP
administration, the CSF accumulation of 5-hydroxyindoleacetic acid (5-HIAA)
was significantly reduced. This suggests that 5-HTP has reduced the serotonin
deficiency in these patients (Fuxe et al. 1971). However, 5-HTP administration not
only resulted in 5-HT production in serotoninergic neurons but also seen in the
530 H. S. Sharma and A. Sharma
Dopamine Precursor
Catecholamine precursor levodopa has been extensively studied for the treatment of
depressive illnesses (Goodwin et al. 1971; Matussek 1971; Persson and Walinder
1971; Náhunek et al. 1972). Studies in 22 depressive patients receiving either
levodopa alone or in combination with alpha-methyldopahydrazine, an inhibitor of
the enzyme l-aromatic amino acid decarboxylase, showed improved conditions in
6 patients (Goodwin et al. 1971; Bunney Jr et al. 1970). The alpha-methylhydrazine
does not cross the BBB and prevents the peripheral conversion of levodopa to
dopamine. This will result in more levodopa to cross the BBB, indicating a potential
role of levodopa in treating depression. The results suggest that only the retarded
depressive patients and not agitated depressive cases were benefitted by combined
treatment with levodopa and alpha-methylhydrazine (Goodwin et al. 1971). How-
ever, the hydrazine compounds were shown to be liver toxic (Riederer et al. 1973;
Fernstrom 1979).
Further study showed that levodopa treatment resulted in hypomania in seven of
the depressed patients with levodopa treatment (Murphy et al. 1971; Goodwin et al.
1971). Studies by Matussek (1971) showed self-improvement score on depression in
36 retarded depressive patients with levodopa and benserazide hydrochloride, a
decarboxylase inhibitor. However, the differences between levodopa treated and
placebo were not significantly different (Matussek 1971).
Histamine Precursor
Histamine is another biogenic amine that is synthesized by the amino acid histidine
using histamine decarboxylase present in histaminergic neurons in the CNS (Panula
and Nuutinen 2013; Simons 2004). Although several indications suggest a role of
histamine receptors in the pathology of depression, histamine precursors have not
been systematically examined as possible therapy for depressive illness.
Previous report suggests that histidine, the precursor of histamine in the CNS,
reduces immobility time during forced swimming (FS) in mice, indicating a role of
histamine precursor in depression (Lamberti et al. 1998). In addition, animals fed on
histidine-free diet exhibit low histamine content in the brain and show several
symptoms of depression such as fatigue, sleeplessness, and other anxiety-related
behaviors, indicating a clear role of histamine precursor in depression (Yoshikawa
et al. 2014).
Amine Precursors in Depressive Disorders and the Blood-Brain Barrier 531
Pathophysiology of Depression
Traumatic brain injury (TBI) is one of the leading causes of depression that
progresses with severity and duration of the insult (Fann et al. 2009). The
depressive illness is further aggravated in TBI by associated anxiety and alcohol
abuse in victims (Hibbard et al. 1998; Seel and Kreutzer 2003). More than 30% of
TBI cases show major depressive disorder (MDD) or organic defective disorder
ICD-10, F06.3 (mood disorder due to known physiological condition)
(Schoenhuber and Gentilini 1988; van Zomeren and van den Burg 1985).
According to rough estimate MDD is seen following TBI in 33% to 42% follow-
ing first year of insult and developed in more than 61% after 7 years of injury
(Jorge et al. 2004; Hibbard et al. 1998). Recent data shows that out of 559 subjects
hospitalized for TBI, 52% develop MDD within the first year of primary insult
(Fann et al. 2003). Interestingly the increased risk of MDD following TBI is not
limited to mild TBI as there were no subtle differences between severe cases or
mild TBI in precipitating MDD (Hoge et al. 2008). Apart from MDD in cases of
TBI, increased risk of suicide is also prevalent in 10% of cases 1 year post-injury
and about 15% 5 years post-injury (Brooks et al. 1986). Depressed TBI patients
are also susceptible to enhanced post-injury symptoms such as memory impair-
ment, blurred vision, headache, and dizziness as compared to non-depressed
patients (Fann et al. 1995).
There are reasons to believe that depression in TBI cases occurs due to a direct or
indirect secondary injury to brain tissues (Fann et al. 2009; Merritt et al. 2018).
532 H. S. Sharma and A. Sharma
TBI depression is linked to either injured or depressed frontal lobe and basal ganglia
circuit together with ascending monoaminergic pathways (Levin and Kraus 1994;
Rosenthal et al. 1998). Furthermore, the diffused or focal injury after TBI affected
the frontal and temporal lobe circuitry and reduced metabolism (Mayberg 1994). The
MDD can further be affected adversely depending on the post-TBI socioeconomic
factors (Seel et al. 2003).
Recent reports suggest that treatment with selective serotonin reuptake inhibitor
(SSRIs) or serotonin-norepinephrine reuptake inhibitor (SNRIs) could be the first
line of treatment of MDD associated with TBI (Warden et al. 2006). However,
further research is needed to enhance a combination of different drug combinations
including MAOI (DeRubeis et al. 2005).
Patients with concussion caused by sports injury such as either football and rugby or
amateur horse riding develop depression at a higher rate than the average population
(Yrondi et al. 2017). Concussion induces sudden trauma to the head caused by a
blow that transmits string inertial forces to the brain distorting memory and cognitive
and mental state (Brooks et al. 1986; Carman et al. 2015; Holm et al. 2005; Sharma
et al. 2016a, b).
Rough estimate suggests that incidences of concussion could occur in American
football players ranging from 0.2 per 1000 h and 4.1 to 8 per 1000 h in rugby players
as compared to 95 per 1000 h in amateur horse riding or training (Hollis et al. 2012;
Kemp et al. 2008; Koh et al. 2003; McCrory et al., 2017; Tommasone and Valovich
McLeod 2006). Several studies show that these repetitive concussions lead to the
development of depressive episodes associated with neurodegeneration and other
psychiatric diseases (Young et al. 2016; Omalu et al. 2005; Lehman et al. 2012;
McKee et al. 2013). In such concussive episodes MDD could occur in about 14% to
77% of cases that account for immediate after blow and also associated with longer-
term survival (Solomon et al. 2016). This suggests a clear link between MDD and
sports-related repetitive concussions (Yrondi et al. 2017).
Short-term concussion and development of MDD studies show that about 20% of
17 concussed athletes exhibited depressive symptoms within 5 days as compared to
non-concussed players (Vargas et al. 2015). Another study demonstrated 19.8%
incidences of depressive illness in concussed athletes, among which 33.8% have
both depression and anxiety (Yang et al. 2015a, b). In young high school athletes,
concussion-induced depressive symptoms appear within 1 week in more than 9% of
subjects as compared to non-concussed group (Roiger et al. 2015). In a study of
collegiate versus high school athletes 2 weeks after concussion, MDD was apparent
that is greater in collegiate students as compared to high school athletes (Kontos
et al. 2012).
Several lines of incidences suggest that concussion-induced depressive illnesses
are about 77% higher in rugby players as compared to other players. Likewise
American football players exhibited MDD that is 2 to 3 times higher than the general
Amine Precursors in Depressive Disorders and the Blood-Brain Barrier 533
population (Makdissi 2010; Makdissi et al. 2014; McCrory et al. 2013). It has also
emerged that lifetime repetitive concussion, e.g., 77 in rugby players versus
13 non-rugby players, exhibited higher degree of MDD associated with other
psychiatric and anxiety problems (Decq et al. 2016). It was found that athletes
who suffer at least three concussions are susceptible to develop three times higher
MDD as compared to those without any concussion. The magnitude and severity of
MDD in concussion is further exacerbated by alcohol intake, other co-morbidity
factors such as ligament injury, and anxiety-related disorders prior to concussion
(Guskiewicz et al. 2007, Guskiewicz and Broglio 2015).
There are evidences that sports concussion and neuroinflammation play key roles
associated with stress factors that are responsible for the increase in circulating
cytokines (Dowlati et al. 2010; Leonard and Maes 2012). Among the cytokines,
tumor necrosis factor-alpha (TNF-α) and interleukin-6 (IL-6) are particularly
involved in MDD caused by concussion (Plata-Salaman 2002). These cytokines
modulate the regulation of monoaminergic neurotransmitters including serotonin
(Hasse and Brown 2015). This suggests that neuropsychotherapy of depression
requires treatment needs regulating cytokines and monoamines together for effective
therapy in clinics.
Depression in Military
Traumatic brain injury (TBI) caused by either penetrating forces or concussive head
injury (CHI) by blunt forces inflicted in military is one of the most leading causes of
post-trauma cognitive, behavioral, and affective disorders (Kennedy et al. 2019).
According to the US Defense Medical Surveillance System data, about 384 k
service members received mild to severe brain injury in the last 18 years (2000–
2018), out of which more than 25% of cases showed chronic disorders ranging from
neurological to neuropsychiatric dysfunction (DoD Worldwide Numbers for TBI,
http://dvbic.dcoe.mil/).
Military personnel after TBI are prone to significantly higher episodes (40% to
77%) of developing MDD (Perry et al. 2016) as compared to personnel without
brain injury (Swanson et al. 2017). Symptoms of depression in TBI cases appear
within 2 weeks after trauma and progressing up to 6 months after the insult (Hou
et al. 2012).
The depression rate after brain injury was significantly higher in female service
members (57.9%) as compared to their male counterparts (Kennedy et al. 2019). The
rate of depression is further exacerbated in military personnel who are depleted to
combat operations more than three times (Shen et al. 2012).
Treatment with SSRI like sertraline is able to partially help depressive episodes in
these service members (Jorge et al. 2016). Since MDD could be a great risk of
suicide, it is important to treat MDD cases with TBI in military (Kambe et al. 2018).
Thus, TBI associated with military in relation to MDD should be identified earlier to
decrease cases of suicide and given suitable treatment to cope with these unwanted
psychiatric diseases.
534 H. S. Sharma and A. Sharma
other tricyclic antidepressants (Giannini and Billett 1987; Barbosa Méndez and
Salazar-Juárez 2019; Ru et al. 2019;Shabani et al. 2019; Quin et al. 2019). The
dopamine D2 receptor agonist apomorphine also rapidly reduces withdrawal symp-
toms of psychostimulants (Hollander et al. 1990; Kampman et al. 2000).
Morphine also induces dependence in individuals using it to suppress either
pain or pleasure. Morphine dependence and withdrawal both cause severe MDD
symptoms (Schulteis et al. 1995). However, morphine withdrawal affecting both
opiates and monoaminergic pathways play significant role. Accordingly, naloxone,
an opiate antagonist, and serotonin 5-HT3 receptor antagonist ondansetron are
capable to alleviate morphine withdrawal symptoms (Schulteis et al. 1995; Sharma
et al. 2019).
These observations suggest that psychostimulant-induced MDD is mediated
through monoaminergic and opioid mechanisms. Further studies on the use of
combination drugs and nanodelivery may improve the clinical efficacy in these
areas.
Rasmusson et al. 2002; Pizarro et al. 2004; Greenwood et al. 2007). This down-
regulation of BDNF at mRNA and/or protein level was restored following antide-
pressant treatment (Tsankova et al. 2006). This suggests that BDNF mechanisms
are somehow involved in the development of depressive behaviors (Jiang and
Salton 2013).
Several lines of evidences suggest that BDNF promotes the development and
functioning of serotoninergic neuromas (see Martinowich and Lu 2008). Serotonin
is phylogenetically the oldest neurotransmitter in the CNS and has the highest
number of receptors compared to any neurotransmitters (Bockaert et al. 2006).
BDNF and its TrK receptor are co-expressed in serotoninergic neurons in the dorsal
and median raphe (Mamounas et al. 2000). The BDNF is retrogradely transported
from the serotonin terminals in the striatum and hippocampus to the cell bodies
located in the raphe nuclei (Anderson et al. 1995). The BDNF promotes expression
of serotoninergic expression in raphe neurons (Siuciak et al. 1996).
Likewise, serotoninergic neurotransmission exerts powerful control on BDNF
expression. This effect is one of the key mechanisms underlying the therapeutic
effects of BDNF enhancement following antidepressant therapy (Szapacs et al.
2004). This is supported by the fact that administration of SSRIs increases the
extracellular concentration of serotonin that could enhance the BDNF level probably
via cAMP response-binding protein (CREB) phosphorylation (Szapacs et al. 2004;
Martinowich and Lu 2008).
Although nearly all antidepressants increase BDNF mRNA levels and enhance
BDNF expression controlling depression, the underlying mechanisms are still not
well known and require further investigation. It appears that the genetic mechanisms
are involved in the BDNF and serotonin interaction (Rybakowski et al. 2007). This is
an interesting feature of research that needed additional investigation.
et al. 1999). Further studies show that antidepressants are able to reduce the activated
immune system in depressive illness (Castanon et al. 2001; Yirmiya et al. 2001).
Cytokines induce downregulation of serotonin synthesis and influence dopami-
nergic system in basal ganglia (Lestage et al. 2002). The hippocampus is highly
innervated by serotoninergic nerve terminals (Hensler 2006). Also the density of
IL-1β, IL-6, IL-2, and TNF-α, the pro-inflammatory cytokines, is highest in the
hippocampus (Wilson et al. 2002). BDNF and TrK levels are downregulated in the
prefrontal cortex that is heavily innervated by serotoninergic neurons (Autry and
Monteggia 2012). This suggests that serotonin regulation of immune modulators
actively plays key roles in depression. Serotonin receptors are actively involved in
immune modulation (Baganz and Blakely 2013; Ahern 2011). Serotonin receptor
5-HT7 induces activation of cytokines in astrocytes (Lieb et al. 2005). These
observations indicate an intricate relationship between serotonin, cytokines, and
the neurotrophins in depression.
The monoamine theory is based on the fact that decreased availability of neurotrans-
mitters serotonin and noradrenaline in the CNS is responsible for the pathogenesis of
depression (Krishnan and Nestler 2008). This theory gets support from the results
based on the antidepressant drugs SSRIs and SNRIs that alleviate depression in
patients (Trivedi et al. 2006). However, not all patients respond to these treatments.
This indicates that monoamines could be a part of the depression process. Additional
research showed that synaptic plasticity and neurogenesis are other important
aspects of depressive illness (Lee and Kim 2010; Eisch and Petrik 2012). The
antidepressive action of the N-methyl-D-aspartate (NMDA) receptor antagonist
ketamine indicates the involvement of glutamatergic receptors in depression
(Berman et al. 2000). However, serotonin is actively involved in the pathology of
depression because the monoamine neurotransmitters interact with several other
agents in the CNS to regulate mood, psychology, and behavioral functions.
Studies in clinics and in animal models of depression show that SSRIs enhance
neurogenesis and reverse hippocampal atrophy (Malberg et al. 2000). BDNF mRNA
and protein levels in the hippocampus increased after SSRI treatment (Autry and
540 H. S. Sharma and A. Sharma
Monteggia 2012). These observations clearly indicate the strong interaction between
BDNF, depression, and serotoninergic neurotransmission (Martinowich and Lu
2008; Hasse and Brown 2015). In addition, extracellular serotonin levels influence
BDNF expression and BDNF regulates serotonin levels through SERT activity
(Daws et al. 2007). These results suggest that reciprocal regulatory mechanisms
are operating between serotonin and BDNF as a feedback loop to maintain both
agents in balance (Hasse and Brown 2015).
There are reasons to believe that cerebrovascular reactions play key roles in the
development of depressive illnesses following stress-induced neuroinflammation
following neurotransmitter release (Selye 1976). Stressors induce the release of
several neurotransmitters in the circulation as well as in the brain microenvironment
leading to cerebrovascular alteration in blood flow and disrupt BBB function (Mora
et al. 2012; Sharma et al. 1990; Shalev et al. 2009; Sharma and Westman 2004). In
this process monoaminergic neurotransmission, harmful cytokines, and other factors
play key roles (Elenkov and Chrousos 2006). Neurotrophins and other molecules are
also released to counteract the harmful vascular response (Smith et al. 1995). Thus, a
balance between neuroprotective and neurodestructive elements finally regulates the
state of the BBB leading to disease formation (Sharma 2009).
Recently, clinical and preclinical studies showed that neuroinflammation and
cerebrovascular dysfunction play key roles in the pathology of MDD (Dudek et al.
2020). It is well known that chronic stress alters the BBB permeability (Fig. 1,
Table 1) probably by enhancing the vesicular transport as well as by disrupting tight
junctional proteins (Sharma 2004b, 2009; Sharma and Sharma 2010). Disruption of
the BBB following stress allows circulating pro-inflammatory cytokines in the brain
resulting in depression-like behaviors, e.g., social avoidance, anhedonia, and help-
lessness (Dudek et al. 2020). One of the main mechanisms of stress-induced BBB
breakdown appears to be mediated by loss of claudin-5 present in the tight junction
of the cerebral endothelial cell (Ménard et al. 2017). Interestingly, the claudin-5
expression is severely reduced in depressive patients in nucleus accumbens region
(Ménard et al. 2017; Dudek et al. 2020).
Experiments from our laboratory showed that various stressors induce the
breakdown of the BBB to protein tracers leading to vasogenic edema formation
associated with depressive behaviors (Table 1, Sharma 1999; Sharma and Westman
2004; Sharma 2005; Sharma and Sharma 2010). At the ultrastructural level
increased vesicular permeability to lanthanum is clearly seen without widening
of the tight junctions (Figs. 2 and 3) (Sharma 2004a, 2009a; Sharma and Sharma
2010). This suggests that endothelial cell membrane forming the tight junctions is
permeable in stress allowing the passage of lanthanum across the tight junctions
without widening them (Sharma 2009a, b; Sharma and Sharma 2010). This effect is
seen in almost all cases of BBB leakage at the ultrastructural level (Figs. 2 and 3).
Stress is also capable to increase blood-cerebrospinal fluid barrier (BCSFB)
breakdown (Fig. 4) allowing cytokines and other immunologic elements to enter
542 H. S. Sharma and A. Sharma
Fig. 1 Showing stages of stress (a) and relation of stress severity with brain pathology (b). There
are three stages of stress, namely, alarm reaction, followed by state of resistance and after that stress-
induced exhaustion (Selye 1976). There are reasons to believe that blood-brain barrier (BBB)
permeability occurs during alarm reaction to the state of resistance (a). Brain pathology following
stress depends on the magnitude and duration of stress (b). It is likely that when moderate to severe
stress prolonged for some time, the BBB disruption occurs. Further continuation of stress results in
alteration of synaptic plasticity and brain pathology that could be irreversible in nature. Depression
may occur after BBB breakdown and alteration in synaptic plasticity associated with cerebrovas-
cular alteration and brain pathology (Sharma 1999, 2004a, 2009a). (Data modified after Sharma
1999, 2004a, 2009a, b; Sharma and Sharma 2010)
into the CSF and affect depressive behaviors (Dudek et al. 2020; Sharma 2009a, b;
Sharma and Sharma 2010).
When the serotonin precursor 5-HTP was injected acutely, the BBB disruption
occurs in several brain regions (Sharma et al. 2019a, b). This suggests that 5-HTP
by enhancing the serotonin levels in the brain and plasma is responsible for the
breakdown of the BBB (Sharma et al. 1990). These observations in part may
explain that 5-HTP treatment alone alters the BBB breakdown; therefore, this
treatment has mixed effects in treating depression-like behaviors in some
patients.
Treating stressful situations with cerebrolysin, a balanced composition of several
neurotrophic factors and/or neurotrophins, e.g., BDNF or IGF-1, significantly
reduced the BBB leakage resulting in attenuation of depressive symptoms
Amine Precursors in Depressive Disorders and the Blood-Brain Barrier 543
Fig. 2 High-power transmission electron micrograph of cerebral capillaries showing tight junc-
tional permeability in stress-induced depression. In control tight junction (a) is closely apposed
endothelial cells that are tight to small molecule lanthanum (molecular diameter 12 Å) is very tight
(arrowhead). The dark black particles are lanthanum present exclusively in the lumen (L). Follow-
ing heat stress-induced depression, the endothelial cell is permeable to lanthanum and could be seen
within the endothelial cell cytoplasm (b, d) without widening of the tight junction (arrows). In (b)
lanthanum is seen penetrating into the endothelial cells around the tight junction (arrows) without
widening them. In (d) the tight junction stops the passage of lanthanum into the basal lamina
(arrowhead) while the endothelial cell cytoplasm is filled with lanthanum. The tight junction is not
widened. That lanthanum could penetrate without widening of the tight junction is clearly seen in
(e) where the cell membrane apposing the tight junction is permeable to lanthanum but the tight
junctions are not widened in heat stress-induced depressive episode (arrows). The tight junctions
between the cerebral capillaries in immobilization (c) and forced swimming (f) are intact and not
widened. The passage of lanthanum was stopped at the tight junctions (arrowheads) while the
cerebral endothelium shows infiltration of lanthanum. Bar 500 nm. (Data adapted from Sharma
et al. 1998; Sharma 2004a, 2009a; Sharma and Sharma 2010)
(Table 2, Sharma HS, unpublished observation; Sharma 1982, 1999, 2004a, 2009a, b).
We have used TiO2-nanowired cerebrolysin and related drugs to induce the
superior neuroprotective effects on the BBB breakdown and brain pathology. Thus,
the need of the hour is to investigate the effects of SSRIs and SNRIs using nano-
delivery in depression. However, these observations are in line with the idea that
depression results from the breakdown of the BBB and restoration of the BBB
structure and the function is likely to alleviate depressive illnesses (Dudek et al.
2020; Nation et al. 2019).
Amine Precursors in Depressive Disorders and the Blood-Brain Barrier
Fig. 3 Low-power transmission electron micrograph (TEM) showing cerebral capillary alterations at the ultrastructural level in depressive episodes following
545
heat stress (a, b, f), immobilization stress (c), and forced swimming (e) as compared to control (d). The microvessel collapse following heat stress in the
546 H. S. Sharma and A. Sharma
Fig. 4 Light microscopy images showing neuronal loss and damages in stress-induced depressive
episode following heat stress (HS, a, b), immobilization (IMZ, c, e, f, g), and forced swimming (FS,
d, h). Hippocampus degeneration (b) in dentate gyrus and CA-4 regions (arrowheads) with
edematous swelling and sponginess (*) is clearly seen as compared to control (a). Following
immobilization (IMZ) induced depressive episodes lead to neuronal damages showing dark neurons
(arrows) with perineuronal edema, and expansion of the neuropil in the cerebral cortex (c) and in the
dorsal thalamus (arrows, g). Immobilization stress also affected blood-cerebrospinal fluid-barrier
(BCSFB) showing degeneration of choroidal epithelial cells and ependymal cells (arrows, f) as
compared to control (e). Likewise forced swimming (FS)-induced depression also resulted in
neuronal damages in the cerebral cortex (d) and ventral thalamus (h) showing dark and distorted
neurons (arrows) with perineuronal edema and expansion of the neuropil (d, h). The BCSFB is also
compromised in FS (k) as evident in the degeneration of choroidal epithelial cells and ependymal
cells (arrows) in FS. (Data modified from Sharma 2004a; Sharma and Sharma 2010. Paraffin
sections H&E stain (a, b); Nissl stain (c-k). Bar ¼ 30 μm)
Fig. 3 (continued) hippocampus (a) and cerebral cortex (f) with perivascular edema and sponginess
is quite apparent. In the thalamic region (b), two microvessels are partially collapsed, showing
lanthanum extravasation across the endothelial cells (arrows) following depression in heat stress.
Membrane vacuolation and edema (*) are clearly seen around the perivascular areas after heat stress
(a, b, f). Immobilization stress (c) and forced swimming (e)-induced microvascular alteration
showed partially collapsed cerebral capillary with perivascular edema (*) with dark and altered
pericytes (arrow) and membrane vacuolation. Bar ¼ 1 μm. (Data modified after Sharma 2004a,
2009; Sharma and Sharma 2010)
Amine Precursors in Depressive Disorders and the Blood-Brain Barrier 547
Table 2 (continued)
Spinal cord injury Cerebrolysin, Reduced No Reduced Sharma
BDNF (2004a,
2009, 2018,
2019)
IGF-1, Growth
hormone
Diabetes Cerebrolysin, Reduced No Reduced Sharma
MSCs et al. (2007,
2015b)
MSCc, mesenchymal stem cells
a
unpublished observation
There are reasons to believe that BBB is a gateway of depression and drugs that are
capable to restore BBB function are suitable therapy in treating depression in
clinic. Amine precursors are used for treating depression and work for some time
and then its effects are dwindling. To this end a combination of drugs antagonizing
pro-inflammatory cytokines and enhancing neurotrophins in the brain could be
used in addition to antidepressants for better clinical efficacy. Alternatively, with
the recent advancement in nanobiotechnology, these drugs may be delivered using
nanoformulation for enhanced therapeutic value in treating depressive illnesses.
The currently using available drugs or their combinations may be used as nano-
formulation in depression or superior effects. This is a new feature in depressive
therapy that is being examined in our laboratory currently.
Acknowledgments The authors’ research reported here are supported in part by grants from the
Air Force Office of Scientific Research (EOARD, London, UK) and Air Force Materiel Command,
USAF, under grant number FA8655-05-1-3065; Swedish Medical Research Council
(Nr 2710-HSS); Göran Gustafsson Foundation, Stockholm, Sweden (HSS); and Astra Zeneca,
Mölndal, Sweden (HSS/AS). We thank Suraj Sharma, Blekinge Institute of Technology, Karls-
krona, Sweden, for computer and graphic support. The US Government is authorized to reproduce
and distribute reprints for Government purpose notwithstanding any copyright notation thereon.
The views and conclusions contained herein are those of the authors and should not be interpreted as
necessarily representing the official policies or endorsements, either expressed or implied, of the Air
Force Office of Scientific Research or the US Government.
References
Aarsland D, Pahlhagen S, Ballard CG, et al. Depression in Parkinson disease -epidemiology,
mechanisms, and management. Nat Rev Neurol. 2012;8:35–47.
Ahern GP. 5-HT and the immune system. Curr Opin Pharmacol. 2011;11(1):29–33.
American Psychiatric Association. Diagnostic and statistical manual of mental disorders.
Washington, DC: American Psychiatric Association; 1952.
Amine Precursors in Depressive Disorders and the Blood-Brain Barrier 549
American Psychiatric Association. Diagnostic and statistical manual of mental disorders (DSM-IV).
Washington, DC: American Psychiatric Association; 1994.
Andersen J, Aabro E, Gulmann N, et al. Anti-depressive treatment in Parkinson’s disease.
A controlled trial of the effect of nortriptyline in patients with Parkinson's disease treated with
L-DOPA. Acta Neurol Scand. 1980;62:210–9.
Anderson KD, Alderson RF, Altar CA, DiStefano PS, Corcoran TL, Lindsay RM, Wiegand
SJ. Differential distribution of exogenous BDNF, NGF, and NT-3 in the brain corresponds to
the relative abundance and distribution of high-affinity and low-affinity neurotrophin receptors.
J Comp Neurol. 1995;357(2):296–317.
Anisman H, et al. Serotonin receptor subtype and p11 mRNA expression in stress-relevant brain
regions of suicide and control subjects. J Psychiatry Neurosci. 2008;33:131–41.
Arnsten AF. Stress signalling pathways that impair prefrontal cortex structure and function. Nat Rev
Neurosci. 2009;10(6):410–22.
Audet MC, Anisman H. Interplay between pro-inflammatory cytokines and growth factors in
depressive illnesses. Front Cell Neurosci. 2013;7:68.
Autry AE, Monteggia LM. Brain-derived neurotrophic factor and neuropsychiatric disorders.
Pharmacol Rev. 2012;64:238–58.
Baganz NL, Blakely RD. A dialogue between the immune system and brain, spoken in the language
of serotonin. ACS Chem Neurosci. 2013;4:48–63.
Banerjee S, Hellier J, Dewey M, Romeo R, Ballard C, Baldwin R, Bentham P, Fox C, Holmes C,
Katona C, Knapp M, Lawton C, Lindesay J, Livingston G, McCrae N, Moniz-Cook E,
Murray J, Nurock S, Orrell M, O’Brien J, Poppe M, Thomas A, Walwyn R, Wilson K, Burns
A. Sertraline or mirtazapine for depression in dementia (HTA-SADD): a randomised, multi-
centre, double-blind, placebo-controlled trial. Lancet. 2011;378:403–11.
Barbacid M. Neurotrophic factors and their receptors. Curr Opin Cell Biol. 1995;7(2):148–55.
Barbosa Méndez S, Salazar-Juárez A. Mirtazapine attenuates anxiety- and depression-like behav-
iors in rats during cocaine withdrawal. J Psychopharmacol. 2019;33(5):589–605. https://doi.
org/10.1177/0269881119840521. Epub 2019 Apr 23
Barone P, Scarzella L, Marconi R, Antonini A, Morgante L, Bracco F, Zappia M, Musch B,
Depression/Parkinson Italian Study Group. Pramipexole versus sertraline in the treatment of
depression in Parkinson’s disease: a national multicenter parallel-group randomized study.
J Neurol. 2006;253(5):601–7.
Barone P, Poewe W, Albrecht S, et al. Pramipexole for the treatment of depressive symptoms in
patients with Parkinson’s disease: a randomised, double-blind, placebo-controlled trial. Lancet
Neurol. 2010;9:573–80.
Barr AM, Markou A. Psychostimulant withdrawal as an inducing condition in animal models of
depression. Neurosci Biobehav Rev. 2005;29(4–5):675–706.
Barr AM, Markou A, Phillips AG. A ‘crash’ course on psychostimulant withdrawal as a model of
depression. Trends Pharmacol Sci. 2002;23(10):475–82.
Barrientos RM, Sprunger DB, Campeau S, Higgins EA, Watkins LR, Rudy JW, Maier
SF. Brain-derived neurotrophic factor mRNA downregulation produced by social isolation
is blocked by intrahippocampal interleukin-1 receptor antagonist. Neuroscience.
2003;121(4):847–53.
Belmaker RH, Agam G. N Major depressive disorder. Engl J Med. 2008;358(1):55–68.
Berman RM, Cappiello A, Anand A, Oren DA, Heninger GR, Charney DS, et al. Antidepressant
effects of ketamine in depressed patients. Biol Psychiatry. 2000;47:351–4.
Birkmayer W, Hornykiewicz O. The L-3,4-dioxyphenylalanine (DOPA)-effect in Parkinson-
akinesia. Wien Klin Wochenschr. 1961;73:787–8.
Birkmayer W, Riederer P. Biochemical post-mortem findings in depressed patients. J Neural
Transm. 1975;37:95–109. https://doi.org/10.1007/BF01663627.
Birkmayer W, Danielczyk W, Neumayer E, Riederer P. The balance of biogenic amines as condition
for normal behavior. J Neural Transm. 1972;33:163–78.
Birkmayer W, Riederer P, Linauer W, et al. L-deprenyl plus l-phenylalanine in the treatment of
depression. J Neural Transm. 1984;59:81–7. https://doi.org/10.1007/BF01249880.
550 H. S. Sharma and A. Sharma
Blazer DG. Depression in late life: review and commentary. J Gerontol A Biol Sci Med Sci.
2003;58(3):249–65.
Bockaert J, Claeysen S, Bécamel C, Dumuis A, Marin P. Neuronal 5-HT metabotropic receptors:
fine-tuning of their structure, signaling, and roles in synaptic modulation. Cell Tissue Res.
2006;326(2):553–72. https://doi.org/10.1007/s00441-006-0286-1. Epub 2006 Aug 1.
Bowers MB Jr. Cerebrospinal fluid 5-hydroxyindoles and behavior after L-tryptophan and pyri-
doxine administration to psychiatric patients. Neuropharmacology. 1970;9(6):599–604.
Brent DA. Antidepressants and suicidality. Psychiatr Clin North Am. 2016;39(3):503–12.
Brodie HKH, Sack R, Siever L. Clinical studies of L-5-hydroxytryptophan in depression. In:
Barchas J, Usdin E, editors. Serotonin and behavior. New York: Academic; 1973. p. 549–59.
Brooks N, Campsie L, Symington C, Beattie A, McKinlay W. The five year outcome of severe blunt
head injury: a relative’s view. J Neurol Neurosurg Psychiatry. 1986;49(7):764–70. https://doi.
org/10.1136/jnnp.49.7.764.
Bunney WE Jr, Murphy DL, Brodie HKH, et al. L-DOPA in depressed patients. Lancet.
1970;1:352.
Bunney WE Jr, Brodie HK, Murphy DL, Goodwin FK. Studies of alpha-methyl-para-tyrosine,
L-dopa, and L-tryptophan in depression and mania. Am J Psychiatry. 1971;127(7):872–81.
Burton, R. (1621/2001). The anatomy of melancholy. New York: New York Review Books.
Cacabelos R. Parkinson’s disease: from pathogenesis to pharmacogenomics. Int J Mol Sci.
2017;18(3):551.
Carman AJ, Ferguson R, Cantu R, Comstock RD, Dacks PA, DeKosky ST, et al. Expert consensus
document: mind the gaps – advancing research into short-term and long-term neuropsycholog-
ical outcomes of youth sports-related concussions. Nat Rev Neurol. 2015;11:230–44.
Carroll BJ, Mowbray RM, Davies B. L-tryptophan in depression. Lancet. 1970;2(7676):776.
Castanon N, Bluthé RM, Dantzer R. Chronic treatment with the atypical antidepressant tianeptine
attenuates sickness behavior induced by peripheral but not central lipopolysaccharide and
interleukin-1beta in the rat. Psychopharmacology. 2001;154(1):50–60.
Chen B, Dowlatshahi D, MacQueen GM, Wang JF, Young LT. Increased hippocampal BDNF
immunoreactivity in subjects treated with antidepressant medication. Biol Psychiatry.
2001;50(4):260–5.
Coppen A. The biochemistry of affective disorders. Br J Psychiatry. 1967;113:1237–64.
Coppen A, Shaw DM, Farrell JP. Potentiation of the antidepressive effect of a monoamine-oxidase
inhibitor by tryptophan. Lancet. 1963;1:79–81.
Coppen A, Shaw DM, Herzberg B, et al. Tryptophan in the treatment of depression. Lancet. 1967;2:
1178–80.
Coppen A, Whybrow PC, Noguera R, et al. The comparative antidepressant value of L-tryptophan
and imipramine with and without attempted potentiation by liothyronine. Arch Gen Psychiatry.
1972;26:234–41.
Couch Y, Anthony DC, Dolgov O, Revischin A, Festoff B, Santos AI, Steinbusch HW, Strekalova
T. Microglial activation, increased TNF and SERT expression in the prefrontal cortex define stress-
altered behaviour in mice susceptible to anhedonia. Brain Behav Immun. 2013;29:136–46.
Coury A, Blaha CD, Atkinson LJ, Phillips AG. Cocaine induced changes in extracellular levels of
striatal dopamine measured concurrently by microdialysis with HPLC-EC and chronoam-
perometry. Ann N Y Acad Sci. 1992;654:424–7.
Curzon G. Influence of plasma tryptophan on brain 5HT synthesis and serotonergic activity. In:
Haber B, Gabay S, Issidorides MR, Alivisatos SGA, editors. Serotonin. Advances in experi-
mental medicine and Biology, vol. 133. Boston: Springer; 1981. https://doi.org/10.1007/978-1-
4684-3860-4_11.
Davis JM. Theories of biological etiology of affective disorders. In: Pfeiffer CC, Smythies JR,
editors. International review of neurobiology, vol. 12. New York: Academic; 1970. p. 145–75.
Daws LC, Munn JL, Valdez MF, Frosto-Burke T, Hensler JG. Serotonin transporter function, but
not expression, is dependent on brain-derived neurotrophic factor (BDNF): in vivo studies in
BDNF-deficient mice. J Neurochem. 2007;101(3):641–51.
Amine Precursors in Depressive Disorders and the Blood-Brain Barrier 551
de Kloet ER, Joëls M, Holsboer F. Stress and the brain: from adaptation to disease. Nat Rev
Neurosci. 2005;6(6):463–75.
Decq P, Gault N, Blandeau M, Kerdraon T, Berkal M, ElHelou A, et al. Long-term con- sequences
of recurrent sports concussion. Acta Neurochir. 2016;158:289–300.
DeRubeis RJ, Hollon SD, Amsterdam JD, Shelton RC, Young PR, Salomon RM, O’Reardon JP,
Lovett ML, Gladis MM, Brown LL, Gallop R. Cognitive therapy vs. medications in the
treatment of moderate to severe depression. Arch Gen Psychiatry. 2005;62:409–16.
Deuschle M, Gilles M, Scharnholz B, Lederbogen F, Lang UE, Hellweg R. Changes of serum
concentrations of brain-derived neurotrophic factor (BDNF) during treatment with venlafaxine
and mirtazapine: role of medication and response to treatment. Pharmacopsychiatry. 2013;46(2):
54–8.
Devos D, Dujardin K, Poirot I, Moreau C, Cottencin O, Thomas P, Destée A, Bordet R, Defebvre
L. Comparison of desipramine and citalopram treatments for depression in Parkinson’s
disease: a double-blind, randomized, placebo-controlled study. Mov Disord. 2008;23(6):
850–7.
Dougherty KD, Dreyfus CF, Black IB. Brain-derived neurotrophic factor in astrocytes, oligo-
dendrocytes, and microglia/macrophages after spinal cord injury. Neurobiol Dis. 2000;7
(6 Pt B):574–85.
Dowlati Y, Herrmann N, Swardfager W, Liu H, Sham L, Reim EK, et al. A meta-analysis of
cytokines in major depression. Biol Psychiatry. 2010;67:446–57.
Dudek KA, Dion-Albert L, Lebel M, LeClair K, Labrecque S, Tuck E, Ferrer Perez C, Golden SA,
Tamminga C, Turecki G, Mechawar N, Russo SJ, Menard C. Molecular adaptations of the
blood-brain barrier promote stress resilience vs. depression. Proc Natl Acad Sci U S
A. 2020;117(6):3326–36.
Duman RS, Aghajanian GK, Sanacora G, Krystal JH. Synaptic plasticity and depression: new
insights from stress and rapid-acting antidepressants. Nat Med. 2016;22(3):238–49.
Dwivedi Y, Rizavi HS, Conley RR, Roberts RC, Tamminga CA, Pandey GN. Altered gene
expression of brain-derived neurotrophic factor and receptor tyrosine kinase B in postmortem
brain of suicide subjects. Arch Gen Psychiatry. 2003;60(8):804–15.
Eisch AJ, Petrik D. Depression and hippocampal neurogenesis: a road to remission? Science.
2012;338(6103):72–5.
Elenkov IJ, Chrousos GP. Stress system–organization, physiology and immunoregulation.
Neuroimmunomodulation. 2006;13(5–6):257–67.
Estcourt MJ, Ramshaw LA, Ramsay AJ. Cytokine responses in virus infections: effects on
pathogenesis, recovery and persistence. Curr Opin Microbiol. 1998;1(4):411–8.
Eyre H, Baune BT. Neuroplastic changes in depression: a role for the immune system. Psychoneur-
oendocrinology. 2012;37(9):1397–416.
Fann JR, Katon WJ, Uomoto JM, Esselman PC. Psychiatric disorders and functional disability in
outpatients with traumatic brain injuries. Am J Psychiatry. 1995;152:1493–9.
Fann JR, Bombardier CH, Temkin NR, Esselman P, Pelzer E, Keough M, Romero H, Dikmen
S. Incidence, severity, and phenomenology of depression and anxiety in patients with moderate
to severe traumatic brain injury. Psychosomatics. 2003;44:161.
Fann JR, Hart T, Schomer KG. Treatment for depression after traumatic brain injury: a systematic
review. J Neurotrauma. 2009;26:2383–402.
Fernstrom JD. Hydrazine compounds were shown to be liver toxic. J Neural Transm. 1979;15:55.
Fernstrom JD, Wurtman RJ. Brain serotonin content: physiological dependence on plasma trypto-
phan levels. Science. 1971;173:149.
Fernstrom JD, Wurtman RJ. Brain serotonin content: physiological regulation by plasma neutral
amino acids. Science. 1972;178:414.
Fibiger HC, Phillips AG. Mesocorticolimbic dopamine systems and reward. Ann N Y Acad Sci.
1988;537:206–15.
Freud, S. (1917/1957). Mourning and melancholia. In J. Strachey (Ed. & Trans.), Standard edition
of the complete works of Sigmund Freud (Vol. 14, pp. 237–258). London: Hogarth Press.
552 H. S. Sharma and A. Sharma
Horwitz AV, Wakefield JC, Lorenzo-Luaces L. History of depression. In: RJ DR, Strunk RD,
editors. Oxford handbooks; 2016. https://doi.org/10.1093/oxfordhb/9780199973965.013.2.
https://www.oxfordhandbooks.com/. Visited on Feb 02, 2021.
Hou R, Moss-Morris R, Peveler R, Mogg K, Bradley BP, Belli A. When a minor head injury results
in enduring symptoms: a prospective investigation of risk factors for postconcussional syn-
drome after mild traumatic brain injury. J Neurol Neurosurg Psychiatry. 2012;83(2):217–23.
Huang J, Pickel VM. Serotonin transporters (SERTs) within the rat nucleus of the solitary tract:
subcellular distribution and relation to 5HT2A receptors. J Neurocytol. 2002;31(8–9):667–79.
Huang EJ, Reichardt LF. Neurotrophins: roles in neuronal development and function. Annu Rev
Neurosci. 2001;24:677–736.
Hurley LL, Tizabi Y. Neuroinflammation, neurodegeneration, and depression. Neurotox Res.
2013;23(2):131–44.
Imamura K, Okayasu N, Nagatsu T. The relationship between depression and regional cerebral
blood flow in Parkinson’s disease and the effect of selegiline treatment. Acta Neurol Scand.
2011;124(1):28–39.
Ishihara L, Brayne C. A systematic review of depression and mental illness preceding Parkinson’s
disease. Acta Neurol Scand. 2006;113:211–20.
Jiang C, Salton SR. The role of neurotrophins in major depressive disorder. Transl Neurosci.
2013;4(1):46–58.
Joëls M, Karst H, Krugers HJ, Lucassen PJ. Chronic stress: implications for neuronal morphology,
function and neurogenesis. Front Neuroendocrinol. 2007;28(2–3):72–96.
Jorge RE, Robinson RG, Moser D, Tateno A, Crespo-Facorro B, Arndt S. Major depression
following traumatic brain injury. Arch Gen Psychiatry. 2004;61:42–50.
Jorge RE, Acion L, Burin DI, Robinson RG. Sertraline for preventing mood disorders following
traumatic brain injury: a randomized clinical trial. JAMA Psychiat. 2016;73(10):1041–7.
Kambe T, Yasuda A, Kinoshita S, Shigeta M, Kinoshita T. Severity of depressive symptoms and
Volume of Superior Temporal Gyrus in People who visit a memory clinic unaccompanied.
Dement Geriatr Cogn Dis Extra. 2018;8(2):207–13. https://doi.org/10.1159/000489008.
eCollection 2018 May-Aug
Kampman KM, Rukstalis M, Pettinati H, et al. The combination of phentermine and fenfluramine
reduced cocaine withdrawal symptoms in an open trial. J Subst Abus Treat. 2000;19(1):77–9.
Karege F, Perret G, Bondolfi G, Schwald M, Bertschy G, Aubry JM. Decreased serum brain-derived
neurotrophic factor levels in major depressed patients. Psychiatry Res. 2002;109(2):143–8.
Karege F, Bondolfi G, Gervasoni N, Schwald M, Aubry JM, Bertschy G. Low brain-derived
neurotrophic factor (BDNF) levels in serum of depressed patients probably results from lowered
platelet BDNF release unrelated to platelet reactivity. Biol Psychiatry. 2005;57(9):1068–72.
Kekuda R, Leibach FH, Furesz TC, Smith CH, Ganapathy V. Polarized distribution of interleukin-1
receptors and their role in regulation of serotonin transporter in placenta. J Pharmacol Exp Ther.
2000;292(3):1032–41.
Kemp SPT, Hudson Z, Brooks JHM, Fuller CW. The epidemiology of head injuries in English
professional rugby union. Clin J Sport Med. 2008;18:227–34.
Kennedy JE, Lu LH, Reid MW, Leal FO, Cooper DB. Correlates of depression in U.S. military
service members with a history of mild traumatic brain injury. Mil Med. 2019;184(3/4):148–54.
Kettenmann H, Hanisch UK, Noda M, Verkhratsky A. Physiology of microglia. Physiol Rev.
2011;91(2):461–553.
Kitahama K, Jouvet A, Fujimiya M, Nagatsu I. Arai R.5-Hydroxytryptophan (5-HTP) uptake and
decarboxylation in the kitten brain. J Neural Transm (Vienna). 2002;109(5–6):683–9.
Kiyatkin EA, Brown PL, Sharma HS. Brain edema and breakdown of the blood-brain barrier during
methamphetamine intoxication: critical role of brain hyperthermia. Eur J Neurosci. 2007;26(5):
1242–53.
Klein AB, Williamson R, Santini MA, Clemmensen C, Ettrup A, Rios M, Knudsen GM, Aznar
S. Blood BDNF concentrations reflect brain-tissue BDNF levels across species. Int
J Neuropsychopharmacol. 2011;14(3):347–53.
554 H. S. Sharma and A. Sharma
Kline NS, Sacks W. Relief of depression within one day using an M.A.O. inhibitor and intravenous
5-HTP. Am J Psychiatry. 1963;120:274–5.
Kline NS, Shah BK. Comparable therapeutic efficacy of tryptophan and imipramine: average
therapeutic ratings versus “true” equivalence: an important difference. Curr Ther Res.
1973;15:484–7.
Kline NS, Sacks W, Simpson GM. Further studies on one day treatment of depression with 5-HTP.
Am J Psychiatry. 1964;121:379–81.
Knott P, Curzon G. Free tryptophan in plasma and brain tryptophan metabolism. Nature. 1972;239:
452–3. https://doi.org/10.1038/239452a0.
Koh JO, Cassidy JD, Watkinson EJ. Incidence of concussion in contact sports: a systematic review
of the evidence. Brain Inj. 2003;17:901–17.
Kohman RA, Rhodes JS. Neurogenesis, inflammation and behavior. Brain Behav Immun.
2013;27(1):22–32.
Kontos AP, Covassin T, Elbin RJ, Parker T. Depression and neurocognitive performance after
concussion among male and female high school and collegiate athletes. Arch Phys Med Rehabil.
2012;93:1751–6.
Koo JW, Russo SJ, Ferguson D, Nestler EJ, Duman RS. Nuclear factor kappaB is a critical mediator
of stress-impaired neurogenesis and depressive behavior. Proc Natl Acad Sci U S A. 2010;107:
2669–74.
Koob GF, Caine SB, Parsons L, Markou A, Weiss F. Opponent process model and psychostimulant
addiction. Pharmacol Biochem Behav. 1997;57(3):513–21.
Kosten TR, Markou A, Koob GF. Depression and stimulant dependence: neurobiology and
pharmacotherapy. J Nerv Ment Dis. 1998;186(12):737–45.
Krishnan V, Nestler EJ. The molecular neurobiology of depression. Nature. 2008;455(7215):
894–902.
Lamberti C, Ipponi A, Bartolini A, Schunack W, Malmberg-Aiello P. Antidepressant-like effects of
endogenous histamine and of two histamine H1 receptor agonists in the mouse forced swim test.
Br J Pharmacol. 1998;123(7):1331–6.
Lee BH, Kim YK. The roles of BDNF in the pathophysiology of major depression and in
antidepressant treatment. Psychiatry Investig. 2010;7(4):231–5.
Lee BH, Kim H, Park SH, Kim YK. Decreased plasma BDNF level in depressive patients. J Affect
Disord. 2007;101(1–3):239–44.
Leentjens AF, Dujardin K, Marsh L, et al. Symptomatology and markers of anxiety disorders in
Parkinson's disease: a cross-sectional study. Mov Disord. 2011;26:484–92.
Lehman EJ, Hein MJ, Baron SL, Gersic CM. Neurodegenerative causes of death among retired
National Football League players. Neurology. 2012;79:1970–4.
Lehmann J. TRYPTOPHAN MALABSORPTION IN LEVODOPA-TREATED PARKINSONIAN
PATIENTS. Effect of tryptophan on mental disturbances. Acta Med Scand. 1973;194(3):181–9.
https://doi.org/10.1111/j.0954-6820.1973.tb19428.x.
Leonard B, Maes M. Mechanistic explanations how cell-mediated immune activation, inflammation
and oxidative and nitrosative stress pathways and their sequels and concomitants play a role in
the pathophysiology of unipolar depression. Neurosci Biobehav Rev. 2012;36:764–85.
Lestage J, Verrier D, Palin K, Dantzer R. The enzyme indoleamine 2,3-dioxygenase is induced in
the mouse brain in response to peripheral administration of lipopolysaccharide and super-
antigen. Brain Behav Immun. 2002;16(5):596–601.
Levin H, Kraus MF. The frontal lobes and traumatic brain injury. J Neuropsychiatr Clin Neurosci.
1994;6:443–54.
Levine J, Barak Y, Chengappa KN, Rapoport A, Rebey M, Barak V. Cerebrospinal cytokine levels
in patients with acute depression. Neuropsychobiology. 1999;40:171–6.
Lewin GR, Barde YA. Physiology of the neurotrophins. Annu Rev Neurosci. 1996;19:289–317.
Li Y, Luikart BW, Birnbaum S, Chen J, Kwon CH, Kernie SG, et al. TrkB regulates hippocam-
pal neurogenesis and governs sensitivity to antidepressive treatment. Neuron. 2008;59:
399–412.
Amine Precursors in Depressive Disorders and the Blood-Brain Barrier 555
Mendels J, Stinnett J. Biogenic amine metabolism, depression and mania. In: Mendels J, editor.
Biological psychiatry. New York: Interscience-John Wiley & Sons Inc; 1973. p. 99–131.
Mendels J, Stinnett JL, Burns D, Frazer A. Amine precursors and depression. Arch Gen Psychiatry.
1975;32(1):22–30.
Mendlewicz J, Youdim MBH. Antidepressant potentiation of 5-hydroxytryptophan by L-deprenil in
affective illness. J Affect Disord. 1980;2(2):137–46. https://doi.org/10.1016/0165-0327(80)
90013-0.
Menza M, Dobkin RD, Marin H, et al. The impact of treatment of depression on quality of life,
disability and relapse in patients with Parkinson’s disease. Mov Disord. 2009;24:1325–32.
Merritt VC, Clark AL, Sorg SF, Evangelista ND, Werhane M, Bondi MW, Schiehser DM, Delano-
Wood L. Apolipoprotein E ε4 Genotype Is Associated with Elevated Psychiatric Distress in
Veterans with a History of Mild to Moderate Traumatic Brain Injury. J Neurotrauma. 2018;35
(19):2272–2282. https://doi.org/10.1089/neu.2017.5372. Epub 2018 Jun 7.
Meyerson LR, Wennogle LP, Abel MS, Coupet J, Lippa AS, Rauh CE, Beer B. Human brain
receptor alterations in suicide victims Pharmacol Biochem Behav 1982;17(1):159–63. https://
doi.org/10.1016/0091-3057(82)90279-9.
Mills KA, Greene MC, Dezube R, Goodson C, Karmarkar T, Pontone GM. Efficacy and tolerability
of antidepressants in Parkinson’s disease: A systematic review and network meta-analysis. Int J
Geriatr Psychiatry. 2018;33(4):642–651. https://doi.org/10.1002/gps.4834. Epub 2017 Dec 13.
Mora F, Segovia G, Del Arco A, de Blas M, Garrido P. Stress, neurotransmitters, corticosterone and
body-brain integration. Brain Res. 2012;1476:71–85.
Murphy DL, Brodie HK, Goodwin FK, Bunney WE Jr. Regular induction of hypomania by L-dopa
in “bipolar” manic-depressive patients. Nature. 1971;229(5280):135–6. https://doi.org/10.1038/
229135a0.
Murphy DL, Baker M, Kotin J, et al. Behavioral and metabolic effects of L-tryptophan in unipolar
depressed patients. In: Barchas J, Usdin E, editors. Serotonin and behavior. New York: Aca-
demic Press Inc; 1973. p. 529–37.
Murray JB. Depression in Parkinson’s disease. The Journal of Psychology. 1996;130(6):659–667.
https://doi.org/10.1080/00223980.1996.9915039.
Náhunek K, Svestka J, Kamenická V, Rodová A. Preliminary clinical experience with L-dopa in
endogenous depressions. Act Nerv Super (Praha). 1972;14(2):101–2.
Nation DA, Sweeney MD, Montagne A, Sagare AP, D'Orazio LM, Pachicano M, Sepehrband F,
Nelson AR, Buennagel DP, Harrington MG, Benzinger TLS, Fagan AM, Ringman JM,
Schneider LS, Morris JC, Chui HC, Law M, Toga AW, Zlokovic BV. Blood-brain barrier
breakdown is an early biomarker of human cognitive dysfunction. Nat Med. 2019;25(2):270–6.
Nibuya M, Nestler EJ, Duman RS. Chronic antidepressant administration increases the expression
of cAMP response element binding protein (CREB) in rat hippocampus. J Neurosci. 1996;16:
2365–72.
O’Brien FE, Dinan TG, Griffin BT, Cryan JF. Interactions between antidepressants and P-glyco-
protein at the blood-brain barrier: clinical significance of in vitro and in vivo findings. Br J
Pharmacol. 2012;165(2):289–312. https://doi.org/10.1111/j.1476-5381.2011.01557.x.
Oh YS, Gao P, Lee KW, Ceglia I, Seo JS, Zhang X, Ahn JH, Chait BT, Patel DJ, Kim Y, Greengard
P. SMARCA3, a chromatin remodeling factor, is required for p11-dependent antidepressant
action. Cell. 2013;152(4):831–43. https://doi.org/10.1016/j.cell.2013.01.014.
Omalu BI, DeKosky ST, Minster RL, Kamboh MI, Hamilton RL, Wecht CH. Chronic traumatic
encephalopathy in a National Football League player. Neurosurgery. 2005;57:128–34. [Discus-
sion 128–34]
Orgeta V, Tabet N, Nilorooshan R, Howard R. Efficacy of antidepressants for depression in
Alzheimer’s disease: systematic review and meta-analysis. J Alzheimer Disease. 2017;58:
725–33.
Ozkizilcik A, Sharma A, Muresanu DF, Lafuente JV, Tian ZR, Patnaik R, Mössler H, Sharma
HS. Timed release of cerebrolysin using drug-loaded titanate nanospheres reduces brain pathology
and improves behavioral functions in Parkinson’s disease. Mol Neurobiol. 2018;55(1):359–69.
Amine Precursors in Depressive Disorders and the Blood-Brain Barrier 557
Ozkizilcik A, Sharma A, Lafuente JV, Muresanu DF, Castellani RJ, Nozari A, Tian ZR, Mössler H,
Sharma HS. Nanodelivery of cerebrolysin reduces pathophysiology of Parkinson’s disease. Prog
Brain Res. 2019;245:201–46.
Panula P, Nuutinen S. The histaminergic network in the brain: basic organization and role in disease.
Nat Rev Neurosci. 2013;14(7):472–87.
Panula P, Rinne J, Kuokkanen K, Eriksson KS, Sallmen T, Kalimo H, Relja M. Neuronal histamine
deficit in Alzheimer’s disease. Neuroscience. 1998;82(4):993–7.
Pardridge WM. Tryptophan transport through the blood-brain barrier: in vivo measurement of free
and albumin-bound amino acid. Life Sci. 1979;25:1519.
Pardridge WM, Oldendorf WH. Kinetic analyses of blood-brain barrier transport of amino acids.
Biochem Biophys Acta. 1975;401:128.
Pare CM. Potentiation of monoamine-oxidase inhibitors by tryptophan. Lancet. 1963;2(7306):527–8.
Pare CM. Psychiatric complications of everyday drugs. Practitioner 1973;210(255):120–6.
Pare CMB, Sandler M. A clinical and biochemical study of a trial of iproniazid in the treatment of
depression. J Neurol Neurosurg Psychiatry. 1959;22:247–51.
Pary R, Scarff JR, Jijakli A, Tobias C, Lippmann S. A review of psychostimulants for adults with
depression. Fed Pract. 2015;32(Suppl 3):30S–7S.
Peña E, Mata M, López-Manzanares L, Kurtis M, Eimil M, Martínez-Castrillo JC, Navas I, Posada
IJ, Prieto C, Ruíz-Huete C, Vela L, Venegas B. en nombre del grupo de trastornos del
movimiento de la Asociación Madrileña de NeurologíaAntidepressants in Parkinson’s disease.
Recommendations by the movement disorder study group of the Neurological Association of
Madrid. Neurologia. 2016;19:S0213–4853(16)00055-4. https://doi.org/10.1016/j.nrl.2016.02.
002.
Perry DC, Sturm VE, Peterson MJ, et al. Association of traumatic brain injury with subsequent
neurological and psychiatric disease: a meta-analysis. J Neurosurg. 2016;124(2):511–26.
Persson T, Roos BE. 5-Hydroxytryptophan for depression. Lancet. 1967;2:987–8.
Persson T, Walinder J. L-DOPA in the treatment of depressive symptoms. Br J Psychiatry.
1971;119:277–8.
Pizarro JM, Lumley LA, Medina W, Robison CL, Chang WE, Alagappan A, Bah MJ, Dawood MY,
Shah JD, Mark B, Kendall N, Smith MA, Saviolakis GA, Meyerhoff JL. Acute social defeat
reduces neurotrophin expression in brain cortical and subcortical areas in mice. Brain Res.
2004;1025(1–2):10–20.
Plata-Salaman CR. Brain cytokines and dis-ease. Acta Neuropsychiatr. 2002;14:262–78.
Politis M, Niccolini F. Serotonin in Parkinson’s disease. Behav Brain Res. 2015;277:136–45.
Pollin W, Cardon PV Jr, Kety SS. Effects of amino acid feedings in schizophrenic patients treated
with iproniazid. Science. 1961;133(3446):104–5.
Popoli M, Yan Z, McEwen BS, Sanacora G. The stressed synapse: the impact of stress and
glucocorticoids on glutamate transmission. Nat Rev Neurosci. 2012;13:22–37.
Prange AJ Jr. The pharmacology and biochemistry of depression. Dis Nerv Syst. 1964;25:217–21.
Qian Z, Wu X, Qiao Y, Shi M, Liu Z, Ren W, Han J, Zheng Q. Downregulation of mGluR2/3
receptors during morphine withdrawal in rats impairs mGluR2/3- and NMDA receptor-
dependent long-term depression in the nucleus accumbens. Neurosci Lett. 2019;690:76–82.
https://doi.org/10.1016/j.neulet.2018.10.018. Epub 2018 Oct 11.
Rakel RE. Depression. Prim Care. 1999;26(2):211–24.
Ramamoorthy S, Ramamoorthy JD, Prasad PD, Bhat GK, Mahesh VB, Leibach FH, Ganapathy
V. Regulation of the human serotonin transporter by interleukin-1 beta. Biochem Biophys Res
Commun. 1995;216(2):560–7.
Rasmussen K, Kendrick WT, Kogan JH, Aghajanian GK. A selective AMPA antagonist, LY293558,
suppresses morphine withdrawal-induced activation of locus coeruleus neurons and behavioral signs
of morphine withdrawal. Neuropsychopharmacology. 1996;15(5):497–505.
Rasmusson AM, Shi L, Duman R. Downregulation of BDNF mRNA in the hippocampal dentate
gyrus after re-exposure to cues previously associated with footshock. Neuropsychophar-
macology. 2002;27(2):133–42.
558 H. S. Sharma and A. Sharma
Rawson KA, Gunstad J, Hughes J, Spitznagel MB, Potter V, Waechter D, Rosneck J. The METER:
a brief, self-administered measure of health literacy. Gen Intern Med. 2010;25(1):67–71.
Rawson KS, Dixon D, Nowotny P, Ricci WM, Binder EF, Rodebaugh TL, Wendleton L, Doré P,
Lenze EJ. Association of functional polymorphisms from brain-derived neurotrophic factor and
serotonin-related genes with depressive symptoms after a medical stressor in older adults. PLoS
One. 2015;10(3):e0120685. https://doi.org/10.1371/journal.pone.0120685. eCollection 2015.
PMID: 25781924
Read JR, Sharpe L, Modini M, Dear BF. Multimorbidity and depression: a systematic review and
meta-analysis. J Affect Disord. 2017;221:36–46.
Reijnders JS, Ehrt U, Weber WE, et al. A systematic review of prevalence studies of depression in
Parkinson’s disease. Mov Disord. 2008;23:183–9. quiz 313
Richard IH, McDermott MP, Kurlan R, et al. A randomized, double-blind, placebo-controlled trial
of antidepressants in Parkinson disease. Neurology. 2012;78:1229–36.
Riederer P, Birkmayer W, Neumayer E. The tyrosine-tryptophan-diagram in a longtime study with
depressed patients. J Neural Transm. 1973;34(1):31–48. https://doi.org/10.1007/BF01244825.
PMID: 4714592
Roiger T, Weidauer L, Kern B. A longitudinal pilot study of depressive symptoms in con-cussed
and injured/nonconcussed National Collegiate Athletic Association Division I student-athletes.
J Athl Train. 2015;50:256–61.
Rosenthal M, Christensen BK, Ross TP. Depression following traumatic brain injury. Arch Phys
Med Rehabil. 1998;79:90–103.
Ru Q, Xiong Q, Zhou M, Chen L, Tian X, Xiao H, Li C, Li Y. Withdrawal from chronic treatment
with methamphetamine induces anxiety and depression-like behavior in mice. Psychiatry Res.
2019;271:476–83. https://doi.org/10.1016/j.psychres.2018.11.072. Epub 2018 Dec 3
Ryan M, Eatmon CV, Slevin JT. Drug treatment strategies for depression in Parkinson disease.
vExpert Opin Pharmacother. 2019;20(11):1351–1363. https://doi.org/10.1080/14656566.2019.
1612877. Epub 2019 May 23.
Rybakowski JK, Suwalska A, Skibinska M, Dmitrzak-Weglarz M, Leszczynska-Rodziewicz A,
Hauser J. Response to lithium prophylaxis: interaction between serotonin transporter and BDNF
genes. Am J Med Genet B Neuropsychiatr Genet. 2007;144B(6):820–3.
Ryding E, Lindström M, Träskman-Bendz L. The role of dopamine and serotonin in suicidal
behaviour and aggression. Prog Brain Res. 2008;172:307–15.
Saha RN, Liu X, Pahan K. Up-regulation of BDNF in astrocytes by TNF-alpha: a case for the
neuroprotective role of cytokine. J NeuroImmune Pharmacol. 2006;1(3):212–22.
Sano I. L-5-hydroxytryptophan(L-5-HTP) therapy in endogenous depression. 1. Munch Med
Wochenschr. 1972;114(40):1713–6.
Sasahara I, Fujimura N, Nozawa Y, Furuhata Y, Sato H. The effect of histidine on mental fatigue and
cognitive performance in subjects with high fatigue and sleep disruption scores. Physiol Behav.
2015;147:238–44.
Schildkraut JJ. The catecholamine hypothesis of affective disorders: a review of supporting
evidence. Am J Psychiatry. 1965a;112:509–22.
Schildkraut JJ. The catecholamine hypothesis of affective disorders: a review of supporting
evidence. Am J Psychiatry. 1965b;122(5):509–22.
Schildkraut JJ, Watson R, Draskoczy PR. Amphetamine withdrawal: depression and
M.H.P.G. excretion. Lancet. 1971;2(7722):485–6.
Schmidt EF, et al. Identification of the cortical neurons that mediate antidepressant responses. Cell.
2012;149:1152–63.
Schoenhuber R, Gentilini M. Anxiety and depression after mild head injury: a case control study.
J Neurol Neurosurg Psychiatry. 1988;51:722–4.
Schulteis G, Markou A, Cole M, Koob GF. Decreased brain reward produced by ethanol with-
drawal. Proc Natl Acad Sci U S A. 1995;92(13):5880–4.
Seel RT, Kreutzer JS. Depression assessment after traumatic brain injury: an empirically based
classification method. Arch Phys Med Rehabil. 2003;84:1621–8.
Amine Precursors in Depressive Disorders and the Blood-Brain Barrier 559
Seel RT, Kreutzer JS, Rosenthal M, Hammond FM, Corrigan JD, Black K. Depression after
traumatic brain injury: a National Institute on Disability and Rehabilitation Research Model
Systems multicenter investigation. Arch Phys Med Rehabil. 2003;84(2):177–84. https://doi.org/
10.1053/apmr.2003.50106.
Selye H. Stress and disease. Science. 1955;122(3171):625–31.
Selye H. Forty years of stress research: principal remaining problems and misconceptions. Can Med
Assoc J. 1976;115(1):53–6.
Seo JS, Wei J, Qin L, Kim Y, Yan Z, Greengard P. Cellular and molecular basis for stress-induced
depression. Mol Psychiatry. 2017;22(10):1440–7.
Seo JS, Zhong P, Liu A, Yan Z, Greengard P. Elevation of p11 in lateral habenula mediates
depression-like behavior. Mol Psychiatry. 2018;23(5):1113–9.
Seppi K, Weintraub D, Coelho M, et al. The Movement Disorder Society evidence-based medicine
review update: treatments for the non-motor symptoms of Parkinson’s disease. Mov Disord.
2011;26(Suppl 3):S42–80.
Shabani S, Schmidt B, Ghimire B, Houlton SK, Hellmuth L, Mojica E, Phillips TJ. Depression-like
symptoms of withdrawal in a genetic mouse model of binge methamphetamine intake. Genes
Brain Behav. 2019;18(3):e12533. https://doi.org/10.1111/gbb.12533. Epub 2018 Nov 26
Shalev H, Serlin Y, Friedman A. Breaching the blood-brain barrier as a gate to psychiatric disorder.
Cardiovasc Psychiatry Neurol. 2009;2009:278531.
Sharma HS. Blood-brain barrier in Stress, Ph D Thesis, May 1982, Banaras Hindu University Press,
Varanasi, India.
Sharma HS. Pathophysiology of blood-brain barrier, brain edema and cell injury following hyper-
thermia: new role of heat shock protein, nitric oxide and carbon monoxide. an experimental
study in the rat using light and electron microscopy. Acta Universitatis Upsaliensis. 1999;830:1–
94. Sharma 2004a,b,c
Sharma HS. Blood-brain and spinal cord barriers in stress. In: Sharma HS, Westman J, editors. The
blood-spinal cord and brain barriers in health and disease. San Diego: Elsevier Academic Press;
2004a. p. 231–98.
Sharma HS. Histamine influences the blood-spinal cord and brain barriers following injuries to the
central nervous system. In: Sharma HS, Westman J, editors. The blood-spinal cord and brain
barriers in health and disease. San Diego: Elsevier Academic Press; 2004b. p. 159–90.
Sharma HS. Pathophysiology of blood-spinal cord barrier in traumatic injury and repair. Curr
Pharm Des. 2005;11(11):1353–89. Review.
Sharma HS. Hyperthermia influences excitatory and inhibitory amino acid neurotransmitters in the
central nervous system. An experimental study in the rat using behavioural, biochemical,
pharmacological, and morphological approaches. J Neural Transm (Vienna). 2006;113(4):
497–519.
Sharma HS. Neurotrophic factors in combination: a possible new therapeutic strategy to influence
pathophysiology of spinal cord injury and repair mechanisms. Curr Pharm Des. 2007;13(18):
1841–74.
Sharma HS. Blood–central nervous system barriers: the gateway to neurodegeneration,
neuroprotection and neuroregeneration. In: Lajtha A, Banik N, Ray SK, editors. Handbook of
neurochemistry and molecular neurobiology: brain and spinal cord trauma. Berlin, Heidelberg,
New York: Springer; 2009a. p. 363–457.
Sharma HS. New concepts of psychostimulants induced neurotoxicity. Int Rev Neurobiol. vol. 89.
San Diego, USA, Oxford, UK: Academic Press. 2009b; pp. 1–435.
Sharma HS. A combination of tumor necrosis factor-alpha and neuronal nitric oxide synthase
antibodies applied topically over the traumatized spinal cord enhances neuroprotection and
functional recovery in the rat. Ann N Y Acad Sci. 2010;1199:175–85.
Sharma HS. New perspectives of central nervous system injury and neuroprotection. Int Rev
Neurobiol. 2012;102:1–424. https://doi.org/10.1016/B978-0-12-386986-9.00013-2.
Sharma HS, Ali SF. Alterations in blood-brain barrier function by morphine and methamphetamine.
Ann N Y Acad Sci. 2006;1074:198–224.
560 H. S. Sharma and A. Sharma
Sharma A, Muresanu DF, Castellani RJ, Nozari A, Lafuente JV, Sahib S, Tian ZR, Buzoianu AD,
Patnaik R, Wiklund L, Sharma HS. Mild traumatic brain injury exacerbates Parkinson’s disease
induced hemeoxygenase-2 expression and brain pathology: neuroprotective effects of
co-administration of TiO2 nanowired mesenchymal stem cells and cerebrolysin. Prog Brain
Res. 2020b;258:157–231.
Shaw DM, Johnson AL, MacSweeney DA. Tricyclic antidepressants and tryptophan in unipolar
affective disorder. Lancet. 1972;2:1245.
Shen YC, Arkes J, Williams TV. Effects of Iraq/Afghanistan deployments on major depression and
substance use disorder: analysis of active duty personnel in the US military. Am J Public Health.
2012;102(Suppl 1):S80–7.
Shimizu E, Hashimoto K, Okamura N, Koike K, Komatsu N, Kumakiri C, Nakazato M, Watanabe H,
Shinoda N, Okada S, Iyo M. Alterations of serum levels of brain-derived neurotrophic factor
(BDNF) in depressed patients with or without antidepressants. Biol Psychiatry. 2003;54(1):70–5.
Simons FE. Advances in H1-antihistamines. N Engl J Med. 2004;351(21):2203–17.
Siuciak JA, Boylan C, Fritsche M, Altar CA, Lindsay RM. BDNF increases monoaminergic activity
in rat brain following intracerebroventricular or intraparenchymal administration. Brain Res.
1996;710(1–2):11–20.
Skapinakis P, Bakola E, Salanti G, Lewis G, Kyritsis AP, Mavreas V. Efficacy and acceptability of
selective serotonin reuptake inhibitors for the treatment of depression in Parkinson’s disease: a
systematic review and meta-analysis of randomized controlled trials. BMC Neurol. 2010;10:49.
Smith RS. The macrophage theory of depression. Med Hypotheses. 1991;35:298–306.
Smith MA, Makino S, Kvetnansky R, Post RM. Stress and glucocorticoids affect the expression of
brain-derived neurotrophic factor and neurotrophin-3 mRNAs in the hippocampus. J Neurosci.
1995;15:1768–77.
Solomon RL, Corbit JD. An opponent-process theory of motivation. I. Temporal dynamics of
affect. Psychol Rev. 1974;81(2):119–45.
Solomon GS, Kuhn AW, Zuckerman SL. Depression as a modifying factor in sport- related
concussion: a critical review of the literature. Phys Sportsmed. 2016;44:14–9.
Starkstein SE, Jorge R, Mizrahi R, Robinson RG. The construct of minor and major depression in
Alzheimer’s disease. Am J Psychiatry. 2005;162:2086–93.
Svenningsson P, et al. Alterations in 5-HT1B receptor function by p11 in depression-like states.
Science. 2006;311:77–80.2006.
Svenningsson P, Kim Y, Warner-Schmidt J, Oh YS, Greengard P. p11 and its role in depression and
therapeutic responses to antidepressants. Nat Rev Neurosci. 2013;14(10):673–80.
Swanson TM, Isaacson BM, Cyborski CM, French LM, Tsao JW, Pasquina PF. Traumatic brain
injury incidence, clinical overview, and policies in the US military health system since 2000.
Public Health Rep. 2017;21(9):1–9.
Szapacs ME, Mathews TA, Tessarollo L, Ernest Lyons W, Mamounas LA, Andrews AM. Exploring
the relationship between serotonin and brain-derived neurotrophic factor: analysis of BDNF
protein and extraneuronal 5-HT in mice with reduced serotonin transporter or BDNF expression.
J Neurosci Methods. 2004;140(1–2):81–92.
Thomas AJ, Hendriksen M, Piggott M, Ferrier IN, Perry E, Ince P, O’Brien JT. A study of the
serotonin transporter in the prefrontal cortex in late-life depression and Alzheimer’s disease with
and without depression. Neuropathol Appl Neurobiol. 2006;32:296–303.
Tommasone BA, Valovich McLeod TC. Contact sport concussion incidence. J Athl Train. 2006;41:
470–2.
Trivedi MH, Fava M, Wisniewski SR, Thase ME, Quitkin F, Warden D, Ritz L, Nierenberg AA,
Lebowitz BD, Biggs MM, Luther JF, Shores-Wilson K, Rush AJ, STAR*D Study Team.
Medication augmentation after the failure of SSRIs for depression. N Engl J Med.
2006;354(12):1243–52.
Troubat R, Barone P, Leman S, Desmidt T, Cressant A, Atanasova B, Brizard B, El Hage W,
Surget A, Belzung C, Camus V. Neuroinflammation and depression: a review. Eur J Neurosci.
2021;53(1):151–71.
Amine Precursors in Depressive Disorders and the Blood-Brain Barrier 563
Tsankova NM, Berton O, Renthal W, Kumar A, Neve RL, Nestler EJ. Sustained hippocampal
chromatin regulation in a mouse model of depression and antidepressant action. Nat Neurosci.
2006;9(4):519–25.
Ueyama T, Kawai Y, Nemoto K, Sekimoto M, Toné S, Senba E. Immobilization stress reduced the
expression of neurotrophins and their receptors in the rat brain. Neurosci Res. 1997;28(2):
103–10.
van Heesch F, Prins J, Korte-Bouws GA, Westphal KG, Lemstra S, Olivier B, Kraneveld AD, Korte
SM. Systemic tumor necrosis factor-alpha decreases brain stimulation reward and increases
metabolites of serotonin and dopamine in the nucleus accumbens of mice. Behav Brain Res.
2013a;253:191–5.
van Heesch F, Prins J, Konsman JP, Westphal KGC, Olivier B, Kraneveld AD, Korte
SM. Lipopolysaccharide-induced anhedonia is abolished in male serotonin transporter knockout
rats: an intracranial self-stimulation study. Brain Behav Immun. 2013b;29:98–103.
van Heesch F, Prins J, Konsman JP, Korte-Bouws GA, Westphal KG, Rybka J, Olivier B, Kraneveld
AD, Korte SM. Lipopolysaccharide increases degradation of central monoamines: an in vivo
microdialysis study in the nucleus accumbens and medial prefrontal cortex of mice. Eur
J Pharmacol. 2014;725:55–63.
van Zomeren AH, van den Burg W. Residual complaints of patients two years after severe head
injury. J Neurol Neurosurg Psychiatry. 1985;48:21–8.
Vargas G, Rabinowitz A, Meyer J, Arnett PA. Predictors and prevalence of postconcussion
depression symptoms in collegiate athletes. J Athl Train. 2015;50:250–5.
Verhoeff NP, Christensen BK, Hussey D, et al. Effects of catecholamine depletion on D2 receptor
binding, mood, and attentiveness in humans: a replication study. Pharmacol Biochem Behav.
2003;74(2):425–32.
Vollenweider FX, Liechti ME, Gamma A, Greer G, Geyer M. Acute psychological and neurophys-
iological effects of MDMA in humans. J Psychoactive Drugs. 2002;34(2):171–84.
Walker FR, Nilsson M, Jones K. Acute and chronic stress-induced disturbances of microglial
plasticity, phenotype and function. Curr Drug Targets. 2013;14(11):1262–76.
Warden DL, Gordon B, McAllister TW, Silver JM, Barth JT, Bruns J, Drake A, Gentry T, Jagoda A,
Katz DL, Kraus J, Labbate LA, Ryan LM, Sparling MB, Walters B, Whyte J, Zapata A, Zitnay
G. Guidelines for the pharmacologic treatment of neurobehavioral sequelae of traumatic brain
injury. J Neurotrauma. 2006;23:1468–501.
Warner-Schmidt JL, et al. Role of p11 in cellular and behavioral effects of 5-HT4 receptor
stimulation. J Neurosci. 2009;29:1937–46.
Warner-Schmidt JL, et al. A role for p11 in the antidepressant action of brain-derived neurotrophic
factor. Biol Psychiatry. 2010;68:528–35.
Warner-Schmidt JL, et al. Antidepressant effects of selective serotonin reuptake inhibitors (SSRIs)
are attenuated by antiinflammatory drugs in mice and humans. Proc Natl Acad Sci U S
A. 2011;108:9262–7.
Watanabe T, Taguchi Y, Shiosaka S, Tanaka J, Kubota H, Terano Y, Tohyama M, Wada
H. Distribution of the histaminergic neuron system in the central nervous system of rats; a
fluorescent immunohistochemical analysis with histidine decarboxylase as a marker. Brain Res.
1984;295(1):13–25.
Weintraub D, Newberg AB, Cary MS, et al. Striatal dopamine transporter imaging correlates with
anxiety and depression symptoms in Parkinson’s disease. J Nucl Med. 2005;46:227–32.
Welcome MO, Mastorakis NE. Stress-induced blood brain barrier disruption: molecular mecha-
nisms and signaling pathways. Pharmacol Res. 2020;157:104769.
Whybrow PC, Prange AJ Jr, Treadway CR. Mental changes accompanying thyroid gland dysfunc-
tion. A reappraisal using objective psychological measurement. Arch Gen Psychiatry.
1969;20(1):48–63.
Wilson AA, Ginovart N, Hussey D, Meyer J, Houle S. In vitro and in vivo characterisation of [11C]-
DASB: a probe for in vivo measurements of the serotonin transporter by positron emission
tomograph. Nucl Med Biol. 2002;29(5):509–15.
564 H. S. Sharma and A. Sharma
Yang J, Peek-Asa C, Covassin T, Torner JC. Post-concussion symptoms of depression and anxiety
in division I collegiate athletes. Dev Neuropsychol. 2015a;40:18–23.
Yang L, Zhao Y, Wang Y, Liu L, Zhang X, Li B, Cui R. The effects of psychological stress on
depression. Curr Neuropharmacol. 2015b;13(4):494–504.
Yirmiya R, Pollak Y, Barak O, Avitsur R, Ovadia H, Bette M, Weihe E, Weidenfeld J. Effects of
antidepressant drugs on the behavioral and physiological responses to lipopolysaccharide (LPS)
in rodents. Neuropsychopharmacology. 2001;24(5):531–44.
Yoshida T, Ishikawa M, Niitsu T, Nakazato M, Watanabe H, Shiraishi T, Shiina A, Hashimoto T,
Kanahara N, Hasegawa T, Enohara M, Kimura A, Iyo M, Hashimoto K. Decreased serum levels
of mature brain-derived neurotrophic factor (BDNF), but not its precursor proBDNF, in patients
with major depressive disorder. PLoS One. 2012;7(8):e42676.
Yoshikawa T, Nakamura T, Shibakusa T, Sugita M, Naganuma F, Iida T, Miura Y, Mohsen A,
Harada R, Yanai K. Insufficient intake of L-histidine reduces brain histamine and causes
anxiety-like behaviors in male mice. J Nutr. 2014;144(10):1637–41.
You Z, Luo C, Zhang W, Chen Y, He J, Zhao Q, et al. Pro- and antiinflammatory cytokines
expression in rat's brain and spleen exposed to chronic mild stress: involvement in depression.
Behav Brain Res. 2011;225:135–41.
Young SN, Lal S, Feldmuller F, Sourkes TL, Ford RM, Kiely M, Martin JB. Parallel variation of
ventricular CSF tryptophan and free serum tryptophan in man. J Neurol Neurosurg Psychiatry.
1976;39:61–5.
Young JS, Hobbs JG, Bailes JE. The impact of traumatic brain injury on the aging brain. Curr
Psychiatry Rep. 2016;18:81.
Yrondi A, Brauge D, LeMen J, Arbus C, Pariente J. Depression and sports-related concussion: a
systematic review. Presse Med. 2017;46:890–902.
Zhu CB, Blakely RD, Hewlett WA. The proinflammatory cytokines interleukin-1beta and tumor
necrosis factor-alpha activate serotonin transporters. Neuropsychopharmacology. 2006;31(10):
2121–31.
The Endocannabinoid System in the Central
Nervous System: Emphasis on the Role
of the Mitochondrial Cannabinoid Receptor
1 (mtCB1R)
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 566
The Endocannabinoid System (ECS) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 567
Historical Background of the ECS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 567
Endocannabinoids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 570
Receptors for Cannabinoids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 571
The ECS and the Modulation of Functions in the CNS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 574
Regulation of Mitochondrial Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 575
The Cannabinoid 1 Mitochondrial Receptor (mtCB1R) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 577
Effects of Cannabinoids on Mitochondria . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 580
Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 582
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 583
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 583
Abstract
The study of the endocannabinoid system (ECS) emerges formally from the
chemical characterization of the psychoactive component of Cannabis sativa,
M. Maya-López
Posgrado en Ciencias Biológicas y de la Salud, DCBS, Universidad Autónoma Metropolitana-
Iztapalapa, Ciudad de México, Mexico
Laboratorio de Aminoácidos Excitadores, Instituto Nacional de Neurología y Neurocirugía Manuel
Velasco Suárez, Ciudad de México, Mexico
C. Zazueta
Departamento de Biomedicina Cardiovascular, Instituto Nacional de Cardiología Ignacio Chávez,
Ciudad de México, Mexico
S. Retana-Márquez
Departamento de Biología de Reproducción, Universidad Autónoma Metropolitana-Iztapalapa,
Ciudad de México, Mexico
S. F. Ali
Division of Neurotoxicology, National Center for Toxicological Research, United States Food and
Drug Administration, Jefferson, AR, USA
Introduction
C. Karasu
Cellular Stress Response and Signal Transduction Research Laboratory, Faculty of Medicine,
Department of Medical Pharmacology, Gazi University, Ankara, Turkey
E. S. Onaivi
Department of Biology, William Paterson University, Wayne, NJ, USA
M. Aschner
Albert Einstein College of Medicine, Bronx, NY, USA
IM Sechenov First Moscow State Medical University, Moscow, Russia
A. Santamaría (*)
Laboratorio de Aminoácidos Excitadores, Instituto Nacional de Neurología y Neurocirugía Manuel
Velasco Suárez, Ciudad de México, Mexico
e-mail: absada@yahoo.com
The Endocannabinoid System in the Central Nervous System: Emphasis on the. . . 567
and neuroprotective responses and prevent several mechanisms of cell death and
damage occurring in neurodegenerative disorders. In turn, these pathologies are
characterized by progressive neuronal loss in brain tissue, leading to metabolic
dysfunction accompanied by behavioral, cognitive, and motor alterations (Warby
et al. 2011; Domaradzki 2015; Ren et al. 2020). The neuroprotective mechanisms
evoked by the ECS may emerge, among several processes, from the inhibition of
excessive glutamatergic transmission, thereby reducing excitotoxicity through the
induction of hypofunction of the subtype of glutamatergic receptors N-methyl-D-
aspartate receptors (NMDAR) (Sánchez-Blázquez et al. 2014), favoring neuronal
survival via increased production of the brain-derived neurotrophic factor (BDNF)
(Blázquez et al. 2015) through the activation of cannabinoid receptors at the cell
membrane. It is noteworthy that only a few studies have paid attention to the
expression and regulation of mitochondrial cannabinoid 1 receptors (mtCB1R) and
the consequent reduction of the energetic metabolism (Bénard et al. 2012; Hebert-
Chatelain et al. 2016). In this chapter, we discuss the cumulative general knowledge
concerning the role of the ECS in the CNS and explore the contribution of mtCB1R
to the energetic metabolism as a physiological function in neurons, with emphasis on
whether this function can be considered as a mechanism accounting for the regula-
tion of mitochondrial activity linked to neuroprotection or is merely a triggering
signal for pathophysiological events. Thus, the aim of this chapter is to provide an
overview on the role of the ECS in optimal CNS since its first descriptions and up to
the present, with emphasis on those reports exploring the modulation of mitochon-
drial energetic metabolism and signaling by mtCB1R.
Since ancient times, the presence of cannabis in different civilizations has shown the
utility of this plant for several purposes, including the fabrication of ropes and cloths,
568 M. Maya-López et al.
the preparation of food, and its use as a substrate for preparation de medicinal
infusions and rituals. The characterization of the psychoactive components of the
plant began in the XIX century with the lipophilic extraction of the plant compo-
nents. In 1964, Raphael Mechoulam and Yechiel Gaoni isolated Δ9-tetrahydrocan-
nabinol (Δ9-THC), the main psychoactive compound of cannabis (Mechoulam and
Gaoni 1965), and other components such as cannabidiol (CBD), which combined
were termed as “cannabinoids.” The characterization of the effects of Δ9-THC
catapulted the study of a considerable number of effects of other components of
cannabis, leading to a most recent description of an endogenous regulatory system
activated by phytocannabinoids.
In 1988, this story took a step forward when the Allyn Howlett’s group charac-
terized brain receptors located on the surface of the cell membrane which were
activated by THC in rats (Devane et al. 1988). Once cloned, the receptor was named
of cannabinoid receptor 1 (CB1R), which is one of the most abundant receptors
coupled to G proteins (GPCR) in the brain. Further studies revealed that CB1R is
also present in the skin and several other organs. Then, a key question emerged: why
does the body possess receptors capable of responding to phytocannabinoids?
Reflections on this topic led to the conclusion on the existence of endogenous
compounds similar to THC referred to as endocannabinoids (eCBs), as discussed
below.
N-arachidonoylethanolamine (anandamide or AEA) was discovered in 1992 by
Raphael Mechoulam’s group (Devane et al. 1992) and formally identified as the first
endogenous cannabinoid neurotransmitter that can bind to CB1R. Later on, AEA
and THC were found to bind CB1R and generate similar effects. Both are ligands
capable of triggering cascades of intracellular events regulating several physiolog-
ical processes, including appetite, glucose metabolism, mood, nociception, and
fertility, among several others. AEA is produced by the body upon demand and is
known to facilitate neurogenesis and neuroprotection.
Later, in 1993, a second type of cannabinoid receptor, CB2R, was identified.
These receptors are mainly located in the immunological system, the peripheral
nervous system (PNS), and in metabolic tissue and inner organs, suggesting, for the
first time, an active role of cannabinoids in the modulation of inflammatory events in
autoimmune alterations (Munro et al. 1993). Nowadays, CB2R is known to be
located in immune cells in the CNS, including microglia and astrocytes. In this
regard, AEA was shown to exhibit low affinity for CB2R, suggesting the existence
yet of another endocannabinoid capable of activating this receptor.
2-Arachidonoylglycerol (2-AG) was simultaneously identified for the first time in
1995 by Raphael Mechoulam’s group (Mechoulam et al. 1995) and by a Japanese
group (Sugiura et al. 1995). This compound is more widely synthesized than AEA
throughout the body, showing higher content in the brain and exhibiting affinity for
both CB1R and CB2R. Both AEA and 2-AG are considered lipidic neurotransmit-
ters capable of maintaining homeostasis and reducing oxidative stress, though only
2-AG can reduce directly the expression of proinflammatory cytokines.
Endocannabinoids possess a well-defined cell metabolism which is orchestrated
by several synthesis and degradative enzymes, acting upon demand. While AEA is
The Endocannabinoid System in the Central Nervous System: Emphasis on the. . . 569
degraded by the enzyme characterized by Ben Cravatt’s group in 1997, fatty acid
amide hydrolase (FAAH) (Giang and Cravatt 1997), 2-AG is catabolized by the
enzyme described by Di Marzo’s group in 1997, monoacylglycerol lipase (MAGL)
(Bisogno et al. 1997). It is noteworthy that the partial inhibition of these enzymes
may be responsible for augmenting the circulating levels of these molecules, which
may afford neuroprotection in some neurodegenerative processes.
The formal characterization of the endocannabinoids AEA and 2-AG and other
lipid derivatives, as well as of CB1R and CB2R and the synthesis and degradative
enzymes for these agents, established the basis for the concept of a canonical
endocannabinoid system (ECS), a complex neurotransmitter system responsible
for the regulation of several major physiological functions.
Taking advantage of CB1R antagonists, evidence emerged demonstrating that not
all the effects of AEA are mediated by these receptors. In 2000, a British group
(Smart et al. 2000) described the role of AEA as agonist of the vanilloid receptors
TRPV1. Such receptors have been related to the regulation of body temperature
and inflammatory pain and are mainly activated by 2-AG. In addition, it is
known that TRP channels can be modulated by both endocannabinoids and
phytocannabinoids.
On the other hand, the concept of retrograde signaling, through which AEA and
2-AG – in contrast to typical neurotransmitters – cross the synaptic cleft from the
postsynaptic terminal to bind to their receptors at the presynaptic terminal, emerged
from observations of several groups in the early 2000s (Wilson et al. 2001; Yoshida
et al. 2002). Their actions as retrograde messengers allow endocannabinoids to
modulate excitatory and inhibitory neurotransmission and reduce excessive
glutamatergic excitation and the subsequent neuronal inflammation; however, at this
point it is pertinent to mention that for the case of AEA, this endocannabinoid can also
be released from presynaptic terminals to bind postsynaptic receptors for regulating
the activity of glutamatergic receptors (reviewed by Joshi and Onaivi 2019).
In 2004, Ethan Russo formally established the concept of “clinical deficiency of
endocannabinoids” based on observations suggesting that disorders such as
migraine, irritable colon, fibromyalgia, and clinical depression share the condition
abovementioned (Russo 2004). Nowadays, the number of disorders sharing this
deficiency has been broadened to include epilepsy, autism, and several neurodegen-
erative diseases.
Another mechanism of action of cannabinoids occurring independently of CB1R
and CB2R was described in 2005. It involves direct activation of the nuclear
receptors known as peroxisome proliferator-activated receptor gamma (PPARγ),
which regulate the lipidic metabolism, genetic expression, and inflammatory
responses (Bouaboula et al. 2005). Both AEA and 2-AG can activate these receptors.
In this regard, cumulative evidence demonstrating these actions led to consider the
existence of a transporting mechanism responsible for conducting endocannabinoids
to the nucleus to bind this nuclear receptor. In 2009, a group from New York
identified a fatty-acid-binding protein (FABP) endowed with transporting endo-
cannabinoids from the cytoplasmic domain to the nucleus or to other intracellular
structures (Kaczocha et al. 2009).
570 M. Maya-López et al.
A few years later, in 2012, a French group described the presence of CB1R on the
mitochondrial membranes (Bénard et al. 2012), suggesting that cannabinoids might
regulate mitochondrial functions such as the energetic metabolism, neurotransmis-
sion, and redox activity.
Considering the mechanistic elements that have been added to the conventional
concept of the canonical ECS, several groups have suggested the adoption of a new
concept, the “extended ECS,” which includes other lipid derivatives besides AEA
and 2-AG. Given this cumulative amount of evidence in favor of a more complex
ECS, in 2013 Vincenzo Di Marzo’s group established the concept of “endo-
cannabinoidome” in direct reference to a complex hypersystem including the
lipidome and microbiome since the ECS signaling facilitates a cross-talk between
the gut bacteria and the brain (Maione et al. 2013).
Therapies based on cannabinoids are gaining relevance and attention in modern
medicine. Specifically, phytocannabinoids have emerged as the first source of
neuroactive compounds producing a broad therapeutic spectrum in the CNS through
their interactions with the ECS. Although several of these effects remain unclear, this
broad efficacy spectrum may be mostly attributed to the pharmacological properties
of these compounds as modulators of the ECS (Rohleder and Müller 2020). Among
these promising phytocompounds, Δ9-THC, cannabidiol (CBD), cannabigerol
(CBG), and cannabidivarin (CBDV) have shown notorious effects on pain percep-
tion, energy homeostasis, appetite, lipid metabolism, cardiovascular functions, ther-
moregulation, immune response, sleep-wake rhythm, psychomotor activity,
memory, and stress response, though their differential effects might be linked to
diverse mechanisms mediated by interactions with specific molecular targets and a
variety of receptors and ion channels (Rohleder and Müller 2020) that will be revised
in this chapter.
Endocannabinoids
The most studied endocannabinoids, on the bases of their levels, distribution, and
physiological relevance, are AEA, 2-AG, and oleamide (ODA). AEA and ODA are
considered fatty acid amides (FAA). The chemical name of AEA is N-arachidonoy-
lethanolamine, which describes an N-acylethanolamine (NAE) composed by an acyl
group and ethanol group and an amino group. The physiological responses evoked
by AEA in the CNS are commonly associated to a positive mental status, which is
implicit in its composed name, combining the Sanskrit term “Ananda,” equivalent to
happiness or plenitude, with the generic term amide. AEA has been shown to exert
similar effects to those of Δ9-THC, the first molecule extracted from Cannabis
sativa. It is also known that AEA exhibits affinity for both CB1R and CB2R, as
well as for TRPV, modulating the Ca2+ exchange (McKinney and Cravat 2005).
ODA is one of the fatty acid primary amides (FAPAs). It is an unsaturated amide
of 18 carbon groups derived from oleic acid. Its chemical name is cis-9-10-octa-
decenamide. ODA is also considered as endocannabinoid due to its affinity for
CB1R, as well as for other receptors for neurotransmitters, such as the GABA-A
The Endocannabinoid System in the Central Nervous System: Emphasis on the. . . 571
receptor and serotonin receptor. This amide has also affinity for GAP junctions to
antagonize cellular communication (McKinney and Cravatt 2005). ODA was found
for the first time in the cerebrospinal fluid of cats subjected to sleep deprivation
(Mueller and Driscoll 2009; Tripathi 2020).
In turn, 2-AG is a lipid derivative from the cell membrane which is synthesized
from the activation of phospholipase C and diacylglycerol lipase. This endo-
cannabinoid is an ester formed from omega-6 arachidonic acid and glycerol and
has been characterized as a ligand for CB1R and CB2R in the CNS, the PNS, and the
immunological system (Mechoulam et al. 1995). 2-AG is accumulated in neurons in
response to Ca2+ mobilization stimuli – in contrast to AEA which presents increased
levels in the brain of mammals – being the most abundant endocannabinoid in the
CNS (Piomelli et al. 1998).
function in the CNS (Joshi and Onaivi 2019). Cannabinoid receptors are known to
be widely distributed throughout the body, consisting of receptor subtypes
which most of them are coupled to G proteins, with CB1R and CB2R being the
most studied. There is also a wide variety of endocannabinoids which can be
considered as retrograde messengers since they are synthesized and released from
postsynaptic terminals to complete their function at presynaptic terminals as
described above.
Other Receptors
Cannabinoids may interact with several other receptors, thus generating a variety of
effects. For instance, CBG has been shown to activate α2-adrenoceptors and block G
protein-coupled CB1R and 5-HT1A receptors (Cascio et al. 2010; reviewed by De
Petrocellis et al. 2011). Additional studies are needed to elucidate whether other
cannabinoids might also act on α2-adrenoceptors and which kind of effects can be
derived from these interactions, since these GPCR receptors are commonly related to
sedation, muscle relaxation, and analgesia. In addition, an interaction of CBD with
the 5-hydroxytriptamine (5-HT)1A receptor led to propose this cannabinoid as a
serotoninergic agonist (Russo et al. 2005), which might support a role of CBD in
relief of anxiety, pain, headache, and thermoregulation through the activation of
these receptors. It is noteworthy that the regulatory role of CBD on nausea and
vomiting via 5-HT1A receptor activation can be suppressed by an interaction
between CBD and CGB (Rock et al. 2011), suggesting complex pharmacological
effects of these compounds at the serotonergic level.
The G protein-coupled receptor (GPCR) 55 regulates the triggering of intracel-
lular Ca2+ mobilization, playing a crucial role in neurotransmitter release in hippo-
campal CA3-CA1 synapses. It has been shown that CBD, acting as an antagonist on
GPR55, inhibits Ca2+ release from the presynaptic stores, thus modulating the effects
of this receptor and contributing to the regulation of hippocampal functions
(Sylantyev et al. 2013). The impact of this evidence has been extended to other
levels, since CBD has also been demonstrated to reduce seizures and autistic-like
social deficits in a mouse model of Dravet syndrome through the recovery of the
function of inhibitory interneurons in the hippocampal dentate gyrus, with CDB
acting as a GPR55 antagonist (Kaplan et al. 2017).
Endogenous and exogenous cannabinoids have been shown to allosterically
modulate glycine receptors (GlyRs) (Xiong et al. 2012). Both AEA and Δ9-THC
increased glycine-activated currents in spinal-cultured neurons through α1 and α3
subunits, potentiating GlyRs. Such receptors are ligand-gated chloride ion channels
mediating inhibitory transmission in the spinal cord and brainstem. They are also
involved in motor activity and pain perception (Avila et al. 2013). GlyR regulation
by cannabinoids suggests a role of these molecules not only in the mentioned
functions but also in cellular and molecular mechanisms controlling the brain
development.
Collectively, the role of cannabinoids (either phyto-, endo- or synthetic com-
pounds) as ligands of several receptors located in the CNS needs more detailed
characterization in light of the broad regulatory functions that these complex inter-
actions might exert at the central level.
Recent research on the role of the ECS in the CNShas generated a considerable
interest in its physiological functions and its promising therapeutic potential through
the use of drugs interacting with several of its elements (Robson 2014). Among its
The Endocannabinoid System in the Central Nervous System: Emphasis on the. . . 575
Table 1 Reports on the use of cannabinoids (agonists of CB1R) in isolated mitochondria and other
biological preparations in studies about the function of mitochondrial cannabinoid receptor
1 (mtCB1R)
Experimental Cannabinoid Effect
Source model use Concentrations (summary)
Hebert- Mice post-natal WIN 55,212-2 2 μM Decreased cellular
Chatelain fibroblasts respiration
et al. Mice post-natal HU210 1 μM Reduced mitochondrial
(2016) primary mobility
hippocampal
neurons
In vivo model in WIN 55,212-2 5 mg/kg, i.p. Reduced performance in the
mice NOR test
Hippocampal Δ9-THC 800 nM Decreased adenylate cyclase
mitochondria and cellular respiration
isolated from Decreased expression of
mice NDUFS2 subunit of
complex I
Mice HU210 2.5 μM Reduced fEPSP
hippocampal
slices
Bénard Hippocampal WIN 55,212-2 20–100 nM Decreased mitochondrial
et al. mitochondria respiration in a
(2012) isolated from concentration-dependent
mice manner
100 nM Decreased complex I
activity, cAMP, and PKA
Δ9-THC 400–800 nM Reduced mitochondrial
respiration, cAMP, and PKA
In vivo model in Δ9-THC 5 mg/kg, i.p. Decreased activity of the
C57BL6/N mice mitochondrial respiratory
chain
In these studies, the presence of CB1R in mitochondria isolated from the hippocampus was
demonstrated, and the signaling pathway involved in their activity was explored, providing
evidence of a reduced activity in mitochondrial electron transport chain
Abbreviations: NOR, novel object recognition; fEPSP, field excitatory postsynaptic potentials;
cAMP, cyclic AMP; PKA, protein kinase A; NDUFS2, NADH dehydrogenase [ubiquinone] iron-
sulfur protein 2
(Chacinska et al. 2009); in turn, this process implies the cooperation with chaperons
and the assembly of molecular complexes to direct mitochondrial proteins to their
final destinations. Other functions linked to mtCB1R are yet to be fully characterized
in order to establish the precise role of these receptors in the physiology of the CNS.
In regard to NDUFS2, it has been demonstrated that a decrease in the levels of
mitochondrial hydrogen peroxide (H2O2) may be mediated by inhibition of this
subunit from the mitochondrial complex I due to the lack of oxygen (O2) under
hypoxic conditions, as it has been shown in pulmonary artery smooth muscle cells
(PASMC), as well as by rotenone, which binds this subunit to inhibit the activity of
NADH dehydrogenase complex. This evidence has been collected from a study
where the expression of NDUFS2 was inhibited through the use of small interfering
RNA (siRNA). It is known that H2O2 formation is important for lung vasodilation.
This radical is formed in the mitochondrial matrix by the reduction of O2•– through
the activity of superoxide dismutase 2 (SOD2) (Dunham-Snary et al. 2019; Wolin
et al. 2019). Through this mechanism, we hypothesize that mitCB1R might exert
relevant physiological functions in the mitochondria from the CNS since the inhibi-
tion of the subunit NDUFS2 reduces the respiratory chain, thus decreasing quinone
over-reduction to prevent RET and modulate harmful ROS formation in mitochon-
dria. Figure 1 summarizes the series of events underlying mtCB1R activation.
Conclusion
Since the first description of THC in 1964 by Raphael Mechoulam, and the ensuing
characterization of receptors with affinity for THC by Allyn Howlett’s group, the
ECS has been intensely investigated, revealing so far an unexpected complexity
linked to its broad distribution throughout the body, particularly in the brain. Such
level of complexity is enriched by a considerable amount and diversity of receptors
and endogenous ligands (endocannabinoids), the many pathways for their synthesis
and degradation, and the molecular and cellular mechanisms regulated by the
membrane receptors CB1R, CB2R, and TRPV1, as well as by activation of nuclear
receptors PPAR-γ, for the coordination of various physiological, neuroprotective,
and antineoplastic events. This chapter aims to update the reader on the state-of-the-
art knowledge about the many faces of the ECS by providing a brief description of
key findings on the functions of this complex system, highlighting the presence and
function of CB1R on the membranes of organelles in charge of the cellular bioen-
ergetics, the mitochondria. This relatively novel field of research deserves consider-
able attention given the wide distribution of mitochondria in neurons, which are
responsible for maintaining several functions such as cell survival, synaptic com-
munication, cell plasticity, and axonal growth, to name a few. Thus, the regulation of
the energetic metabolism depends on constant ATP availability, and mitochondria
are not only in charge of this process, but also of the regulated ROS production for an
adequate cellular signaling. Through the formal description of the signaling pathway
linked to the activation of mtCB1R by Bénard and Hebert-Chatelain, we have
learned that these receptors activate inhibitory oxidative phosphorylation signals
through the inhibition of adenylate cyclase in hippocampal mitochondria, thus
potentially reducing cognitive skills; however, based on evidence discussed in this
chapter, we suggest that, while the contributions made by these groups are essential
for the comprehension of mtCB1R functions, these findings remain limited to the
extent that they only address hippocampal function, thereby lacking broader support
to explain the full complexity of functions that may be elicited by these receptors in
other brain regions, as well as justifying their real physiological role. So far, it seems
that these receptors are abundant in glutamatergic and GABAergic neurons and their
activity contributes to regulatory processes yet poorly explored, also participating in
deleterious events such as episodic amnesia in rodents. These findings are of major
relevance, as they link the function of these receptors to pathophysiological events,
partially supporting the concept that cannabinoids, mostly those with psychoactive
properties, may alter cognitive functions through the described mechanism. None-
theless, several questions remain in regard to the function of mtCB1Rs, necessitating
an integrative approach to characterizing their function and brain distribution, as
well as the development of binding studies with endogenous ligands to establish
specific regional pharmacological criteria such as affinity and potency. In addition,
this chapter highlights the need to explore the distribution of these receptors, not
only in neurons but also in other cell types such as glia, exploring their physiological
contribution and therapeutic potential. Combined, these concerns remain as perspec-
tives to be explored in a near future.
The Endocannabinoid System in the Central Nervous System: Emphasis on the. . . 583
Cross-References
Declaration of Interest
The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this chapter. The authors are responsible
for content and writing of the manuscript and do not necessarily reflect the position of the US Food
and Drug Administration nor do mention of trade names and commercial products constituting
endorsement or recommendation for use and do not represent agency position or policy.
References
Aguilera-Portillo G, Rangel-López E, Villeda-Hernández J, Chavarría A, Castellanos P,
Elmazoglu Z, Karasu Ç, Túnez I, Pedraza G, Königsberg M, Santamaría A. The pharmacolog-
ical inhibition of fatty acid amide hydrolase prevents excitotoxic damage in the rat striatum:
possible involvement of CB1 receptors regulation. Mol Neurobiol. 2019;56:844–56.
Ahmadian M, Suh JM, Hah N, Liddle C, Atkins AR, Downes M, Evans RM. PPARγ signaling and
metabolism: the good, the bad and the future. Nat Med. 2013;19:557–66.
Avila A, Nguyen L, Rigo J-M. Glycine receptors and brain development. Front Cell Neurosci.
2013;7:184.
Bénard G, Massa F, Puente N, Lourenço J, Bellocchio L, Soria-Gómez E, Matias I, Delamarre A,
Metna-Laurent M, Cannich A, Hebert-Chatelain E, Mulle C, Ortega-Gutiérrez S, Martín-
Fontecha M, Klugmann M, Guggenhuber S, Lutz B, Gertsch J, Chaouloff F, López-Rodríguez
ML, Grandes P, Rossignol R, Marsicano G. Mitochondrial CB1 receptors regulate neuronal
energy metabolism. Nat Neurosci. 2012;15:558–64.
Bisogno T, Sepe N, Melck D, Maurelli S, De Petrocellis L, Di Marzo V. Biosynthesis, release and
degradation of the novel endogenous cannabimimetic metabolite 2-arachidonoylglycerol in
mouse neuroblastoma cells. Biochem J. 1997;322:671–7.
Blázquez C, Chiarlone A, Bellocchio L, Resel E, Pruunsild P, García-Rincón D, Sendtner M,
Timmusk T, Lutz B, Galve-Roperh I, Guzmán M. The CB1 cannabinoid receptor signals striatal
neuroprotection via a PI3K/Akt/mTORC1/BDNF pathway. Cell Death Differ. 2015;22:
1618–29.
Bouaboula M, Poinot-Chazel C, Marchand J, Canat X, Bourrié B, Rinaldi-Carmona M, Calandra B,
Le Fur G, Casellas P. Signaling pathway associated with stimulation of CB2 peripheral
cannabinoid receptor. Involvement of both mitogen-activated protein kinase and induction of
Krox-24 expression. Eur J Biochem. 1996;237:704–11.
Bouaboula M, Hilairet S, Marchand J, Fajas L, Le Fur G, Casellas P. Anandamide induced
PPARgamma transcriptional activation and 3T3-L1 preadipocyte differentiation. Eur J
Pharmacol. 2005;517:174–81.
Cabral GA, Griffin-Thomas L. Emerging role of the cannabinoid receptor CB2 in immune regula-
tion: therapeutic prospects for neuroinflammation. Exp Rev Mol Med. 2009;11:e3.
Cabral GA, Raborn ES, Griffin L, Dennis J, Marciano-Cabral F. CB2 receptors in the brain: role in
central immune function. Br J Pharmacol. 2008;153:240–51.
Cardone L, de Cristofaro T, Affaitati A, Garbi C, Ginsberg MD, Saviano M, Varrone S, Rubin CS,
Gottesman ME, Avvedimento EV, Feliciello A. A-kinase anchor protein 84/121 are targeted to
584 M. Maya-López et al.
mitochondria and mitotic spindles by overlapping amino-terminal motifs. J Mol Biol. 2002;320:
663–75.
Cascio MG, Gauson LA, Stevenson LA, Ross RA, Pertwee RG. Evidence that the plant cannabi-
noid cannabigerol is a highly potent alpha2-adrenoceptor agonist and moderately potent 5HT1A
receptor antagonist. Br J Pharmacol. 2010;159:129–41.
Catanzaro G, Rapino C, Oddi S, Maccarrone M. Anandamide increases swelling and reduces
calcium sensitivity of mitochondria. Biochem Biophys Res Commun. 2009;388:439–42.
Chacinska A, Koehler CM, Milenkovic D, Lithgow T, Pfanner N. Importing mitochondrial pro-
teins: machineries and mechanisms. Cell. 2009;138:628–44.
Chiarlone A, Bellocchio L, Blázquez C, Resel E, Soria-Gómez E, Cannich A, Ferrero JJ,
Sagredo O, Benito C, Romero J, Sánchez-Prieto J, Lutz B, Fernández-Ruiz J, Galve-Roperh I,
Guzmán M. A restricted population of CB1 cannabinoid receptors with neuroprotective activity.
Proc Natl Acad Sci U S A. 2014;111:8257–62.
Clapham DE, Julius D, Montell C, Schultz G. International Union of Pharmacology. XLIX.
Nomenclature and structure-function relationships of transient receptor potential channels.
Pharmacol Rev. 2005;57:427–50.
Console-Bram L, Marcu J, Abood ME. Cannabinoid receptors: nomenclature and pharmacological
principles. Prog Neuro-Psychopharmacol Biol Psychiatry. 2012;38:4–15.
Cristino L, Bisogno T, Di Marzo V. Cannabinoids and the expanded endocannabinoid system in
neurological disorders. Nat Rev Neurol. 2020;16:9–29.
Cui M, Honore P, Zhong C, Gauvin D, Mikusa J, Hernandez G, Chandran P, Gomtsyan A,
Brown B, Bayburt EK, Marsh K, Bianchi B, McDonald H, Niforatos W, Neelands TR,
Moreland RB, Decker MW, Lee C-H, Sullivan JP, Faltynek CR. TRPV1 receptors in the CNS
play a key role in broad-spectrum analgesia of TRPV1 antagonists. J Neurosci. 2006;26:9385–93.
De Petrocellis L, Ligresti A, Moriello AS, Allara M, Bisogno T, Petrosino S, Stott CG, Di Marzo
V. Effects of cannabinoids and cannabinoid-enriched Cannabis extracts on TRP channels and
endocannabinoid metabolic enzymes. Br J Pharmacol. 2011;163:1479–94.
Devane WA, Dysarz FA 3rd, Johnson MR, Melvin LS, Howlett AC. Determination and character-
ization of a cannabinoid receptor in rat brain. Mol Pharmacol. 1988;34:605–13.
Devane WA, Hanus L, Breuer A, Pertwee RG, Stevenson LA, Griffin G, Gibson D, Mandelbaum A,
Etinger A, Mechoulam R. Isolation and structure of a brain constituent that binds to the
cannabinoid receptor. Science. 1992;258:1946–9.
Djeungoue-Petga M-A, Hebert-Chatelain E. Linking mitochondria and synaptic transmission: the
CB1 receptor. BioEssays 2017;1700126.
Domaradzki J. The impact of Huntington disease on family carers: a literature overview. Psychiatr
Pol. 2015;49:931–44.
Dreyer C, Keller H, Mahfoudi A, Laudet V, Krey G, Wahli W. Positive regulation of the peroxi-
somal beta-oxidation pathway by fatty acids through activation of peroxisome proliferator-
activated receptors (PPAR). Biol Cell. 1993;77:67–76.
Dunham-Snary KJ, Wu D, Potu F, Sykes EA, Mewburn JD, Charles RL, Eaton P, Sultanian RA,
Archer SL. Ndufs2, a core subunit of mitochondrial complex I, is essential for acute oxygen-
sensing and hypoxic pulmonary vasoconstriction. Circ Res. 2019;124:1727–46.
Elbrecht A, Chen Y, Cullinan CA, Hayes N, Leibowitz MD, Moller DE, Berger J. Molecular
cloning, expression and characterization of human peroxisome proliferator activated receptors
gamma 1 and gamma 2. Biochem Biophys Res Comm. 1996;224:431–7.
Elmazoglu Z, Rangel-López E, Medina-Campos ON, Pedraza-Chaverri J, Túnez I, Aschner M,
Santamaría A, Karasu Ç. Cannabinoid-profiled agents improve cell survival via reduction of
oxidative stress and inflammation, and Nrf2 activation in a toxic model combining hypergly-
cemia+Aβ1-42 peptide in rat hippocampal neurons. Neurochem Int. 2020;140:104817. https://
doi.org/10.1016/j.neuint.2020.104817.
Everaerts W, Gees M, Alpizar YA, Farre R, Leten C, Apetrei A, Dewachter I, van Leuven F,
Vennekens R, De Ridder D, Nilius B, Voets T, Talavera K. The capsaicin receptor TRPV1 is a
crucial mediator of the noxious effects of mustard oil. Curr Biol. 2011;21:316–21.
The Endocannabinoid System in the Central Nervous System: Emphasis on the. . . 585
Smart D, Gunthorpe MJ, Jerman JC, Nasir S, Gray J, Muir AI, Chambers JK, Randall AD, Davis
JB. The endogenous lipid anandamide is a full agonist at the human vanilloid receptor (hVR1).
Br J Pharmacol. 2000;129:227–30.
Sugiura T, Kondo S, Sukagawa A, Nakane S, Shinoda A, Itoh K, Yamashita A, Waku
K. 2-Arachidonoylglycerol: a possible endogenous cannabinoid receptor ligand in brain.
Biochem Biophys Res Commun. 1995;215:89–97.
Sylantyev S, Jensen TP, Ross RA, Rusakov DA. Cannabinoid- and lysophosphatidylinositol-
sensitive receptor GPR55 boosts neurotransmitter release at central synapses. Proc Natl Acad
Sci. 2013;110:5193–8.
Taylor CT, Moncada S. Nitric oxide, cytochrome C oxidase, and the cellular response to hypoxia.
Arterioscler Thromb Vasc Biol. 2010;30:643–7.
Tripathi RKP. A perspective review on fatty acid amide hydrolase (FAAH) inhibitors as potential
therapeutic agents. Eur J Med Chem. 2020;188:111953.
Velez-Pardo C, Jimenez-Del-Rio M, Lores-Arnaiz S, Bustamante J. Protective effects of the
synthetic cannabinoids CP55,940 and JWH-015 on rat brain mitochondria upon paraquat
exposure. Neurochem Res. 2010;35:1323–32.
Warby SC, Visscher H, Collins JA, Doty CN, Carter C, Butland SL, Hayden AR, Kanazawa I, Ross
CJ, Hayden MR. HTT haplotypes contribute to differences in Huntington disease prevalence
between Europe and East Asia. Eur J Hum Genet. 2011;19:561–6.
Wilson RI, Kunos G, Nicoll RA. Presynaptic specificity of endocannabinoid signaling in the
hippocampus. Neuron. 2001;31:453–62.
Wolin MS, Alruwaili N, Sharath K. Studies on hypoxic pulmonary vasoconstriction detect a novel
role for the mitochondrial complex I subunit Ndufs2 in controlling peroxide generation for
oxygen-sensing. Circ Res. 2019;124:1683–5.
Xin Q, Xu F, Taylor DH, Zhao J, Wu J. The impact of cannabinoid type 2 receptors (CB2Rs) in
neuroprotection against neurological disorders. Acta Pharmacol Sin. 2020;41:1507–18.
Xiong W, Wu X, Li F, Cheng K, Rice KC, Lovinger DM, Zhang L. A common molecular basis for
exogenous and endogenous cannabinoid potentiation of glycine receptors. J Neurosci. 2012;32:
5200–8.
Xu Z, Lv XA, Dai Q, Ge YQ, Xu J. Acute upregulation of neuronal mitochondrial type-1
cannabinoid receptor and its role in metabolic defects and neuronal apoptosis after TBI. Mol
Brain. 2016;9:75.
Yoshida T, Hashimoto K, Zimmer A, Maejima T, Araishi K, Kano M. The cannabinoid CB1
receptor mediates retrograde signals for depolarization-induced suppression of inhibition in
cerebellar Purkinje cells. J Neurosci. 2002;22:1690–7.
Experimental Psychopharmacology
Nicola Simola
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 590
Use of Instrumental Techniques in Experimental Psychopharmacology . . . . . . . . . . . . . . . . . . . . . . . 590
Use of Animal Models in Experimental Psychopharmacology: General Considerations . . . . . 593
Overview of Animal Models Used in Experimental Psychopharmacology . . . . . . . . . . . . . . . . . . . . 600
Animal Models Used to Evaluate the Antidepressant Potential of Drugs . . . . . . . . . . . . . . . . . . . . . 600
Animal Models Used to Evaluate the Anxiolytic Potential of Drugs . . . . . . . . . . . . . . . . . . . . . . . . . . 603
Animal Models Used to Evaluate Rewarding and Addictive Properties of Drugs . . . . . . . . . . . . 604
Animal Models Used to Evaluate the Antimanic Potential of Drugs . . . . . . . . . . . . . . . . . . . . . . . . . . 606
Animal Models Used to Evaluate the Antipsychotic Potential of Drugs . . . . . . . . . . . . . . . . . . . . . . 607
Animal Models Used to Evaluate the Effects of Drugs on Cognitive and Executive
Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 609
Animal Models Used to Evaluate New Drugs for the Treatment of Neurodegenerative
Diseases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 611
Towards Improved Behavioral Animal Models: The Example of Ultrasonic Vocalizations . . . 612
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 614
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 614
Abstract
Studies of psychopharmacology in experimental animals are a major source of
information on the effects that drugs elicit on superior brain functions. Moreover,
studies of psychopharmacology are a key step in the refinement of existing drugs
and development of new drugs for the treatment of psychiatric diseases, such as
anxiety, depression, and psychosis. The present chapter provides an overview of
the techniques and animal models that can be used in studies of experimental
psychopharmacology and discusses their strengths and limitations.
N. Simola (*)
Department of Biomedical Sciences, University of Cagliari, Monserrato University Campus,
Monserrato, Italy
e-mail: nicola.simola@unica.it; nicola.simola@gmail.com
Introduction
Table 1 Strengths and limitations of some animal models commonly used in studies of
psychopharmacology
Disease Model Major strengths Major limitations
Anxiety Conditioned models Modifications in the May not always
(i.e., conditioned foot experimental protocol discriminate between
shocks) may allow to study anxiety and fear
different types of
anxiety (i.e.,
anticipatory anxiety,
discomfort caused by
an aversive event)
Reproducible
Anxiety Conflict models (i.e., Reproducible Employ deprivation
Vogel test) of food or water
which may interfere
with the emotional
state of the animals
and with the effects
of the drugs
evaluated
Anxiety Maze-based models - Cost-effective - Are based on
(i.e., elevated plus - Easy to implement exploratory drive
maze; elevated zero - Reproducible - Should not be
maze) replicated in the
same animal to
preserve accuracy
of results
Anxiety Models based on - Cost-effective The specificity of
defensive burying (i. - Easy to implement these models to
e., conditioned probe - Reproducible anxiety is questioned
burying; marble
burying)
Depression Chronic mild stress - Induces behavioral - May be time-
changes reminiscent consuming and
of the symptoms of difficult to
human depression implement
that are reverted by - Results may have
chronic, but not low reproducibility
acute, treatment with across laboratories
antidepressants
- Fulfills all the
validity criteria of
animal models
- May reproduce the
anhedonia observed
in human depression
(continued)
Experimental Psychopharmacology 597
Table 1 (continued)
Disease Model Major strengths Major limitations
Depression Learned helplessness May reproduce certain - May be sensitive to
aspects of the so- acute antidepressants
called stress-coping - The results obtained
depression may be significantly
affected by the
individual variability
in developing the
learned helplessness
behavior and/or by
the individual
sensitivity to pain
induced by foot
shocks
Depression Models based on - Cost-effective - Are sensitive to the
acute stress (i.e., - Easy to implement acute administration
forced swim test, tail - Reproducible of antidepressant
suspension test) drugs
- The readout of these
models (immobility)
is based on motor
activity
- The significance of
the immobility
measured in these
models is disputed
Depression Olfactory bulbectomy Induces behavioral - Is based on the
changes that are induction of
reverted by chronic, hyperactivity,
but not acute, which is not a
treatment with symptom of human
antidepressants depression
- The neurobiological
mechanisms of
hyperactivity
measured in this
model are unknown
Depression Social defeat - Induces behavioral - Results may have
changes reminiscent low reproducibility
of the symptoms of across laboratories
human depression - Specificity of the
that are reverted by model to depression
chronic, but not is questioned
acute, treatment with
antidepressants
- Induces modifications
in the hypothalamic-
pituitary-adrenal axis
reminiscent of those
featuring human
depression
(continued)
598 N. Simola
Table 1 (continued)
Disease Model Major strengths Major limitations
Drug addiction Conditioned models - Cost-effective Drugs that induce
(i.e., conditioned - Easy to implement conditioning may not
place preference; - Reproducible necessarily have
conditioned taste addiction potential
aversion)
Drug addiction Locomotor - Cost-effective Specificity of the
sensitization - Easy to implement model to drug
- Reproducible addiction is
questioned
Drug addiction Models based on - Drug self- - Interpretation of
operant behavior (i.e., administration results may be
drug self- predicts the abuse affected by
administration; drug potential of drugs differences in
discrimination) - Drug self- preclinical and
administration clinical endpoints
reproduces in used to define drug
animals several addiction,
stages of human abstinence, and
addiction relapse
- Drug discrimination - May be time-
allows to study the consuming and
subjective effects of difficult to
drugs implement
- Possess marked
translational value
Impairment in Models based on non- - Allow to measure - Are influenced by
cognitive and operant behavior learning, nonspatial exploratory drive
executive functions (i.e., object memory, and spatial and/or lateralization
recognition, memory of animals
exploration of mazes) - Cost-effective - May require the use
- Easy to implement of stressful stimuli
- Reproducible (i.e., food
deprivation,
swimming) that may
interfere with the
emotional state of
the animals
- Use readouts that
may have limited
translational value
Impairment in Models based on - Allow to measure - Implementation may
cognitive and operant behavior (i.e., attention, motivation, be complex and
executive functions go/no-go, stop-signal, and impulsivity time-consuming
five-choice serial - Are analogous to or - Require extensive
reaction time task) derived from tests training of animals
that are used to - The predictive
measure cognitive validity of these
and executive models may greatly
functions in humans vary according to
- Possess marked the class of drugs
translational value studied
(continued)
Experimental Psychopharmacology 599
Table 1 (continued)
Disease Model Major strengths Major limitations
Mania Environmental - May induce in - In the sleep
models (i.e., resident- animals behavioral deprivation model
intruder; sleep alterations observed animals are exposed
deprivation) in human mania (i.e., to stressful stimuli
aggressiveness, (i.e., isolation) that
agitation) may affect the
- The mania-like emotional state
phenotype can be - Sleep deprivation
reverted by chronic does not induce
treatment with manic-like states in
clinically effective healthy humans
antimanic drugs
Mania Pharmacological - Easy to implement - Are based solely on
models (i.e., - Sensitive to the acute hyperactivity
amphetamine, administration of - Limited and
ouabain, quinpirole) clinically effective contrasting evidence
antimanic drugs exists on the effects
of chronic treatment
with antimanic
drugs in these
models
Neurodegenerative Toxin-based models - Easy to implement Neurodegeneration
diseases - Cost-effective obtained in these
- Induce phenotypes in models is usually
animals that based on a single
effectively mimic one mechanism of
or more symptoms of toxicity
human disease
- Reproducible
Psychosis Catalepsy model - Easy to implement - Does not reproduce
- Cost-effective in animals any
- Reproducible symptoms of
human psychosis
- Possesses only
predictive validity
that allows to
identify the “typical
profile” of
antipsychotic drugs
Psychosis Neurodevelopmental - Are consistent with - Difficult to
models (i.e., the hypothesis that implement
hippocampal neonatal schizophrenia is a - Low reproducibility
lesion; perinatal disorder of of results
insults) neurodevelopment - Specificity of these
- Reproduce in models to psychosis
animals anatomical is questioned
and behavioral
deficits that mimic
the abnormalities
featuring human
schizophrenia
(continued)
600 N. Simola
Table 1 (continued)
Disease Model Major strengths Major limitations
Psychosis Pharmacological - Easy to implement Are based on
models (i.e., - Cost-effective behavioral readouts
dopaminomimetic - Reproducible that reproduce in
drugs, glutamate animals
receptor antagonists) predominantly the
“positive” symptoms
of human
schizophrenia
The most common animal models used for studying the antidepressant potential of
drugs are based on the exposure of rodents to stress, either acute or repeated, in order
to induce changes in behavior that are thought to reproduce certain symptoms
featuring human depression.
Experimental models of depression that use acute exposure to stress are the FST
and the TST. In the FST, rats or mice are placed in a water cylinder (Porsolt et al.
1977), whereas in the TST mice are hung by the tail upside down (Steru et al. 1985).
In either model, the animal will initially try to escape from the stressful situation, but
it will eventually keep an immobile posture as soon as it realizes that the situation is
inescapable. Immobile posture of rodents in the FST and TST is often referred to as
“behavioral despair” and can be attenuated by clinically effective antidepressant
Experimental Psychopharmacology 601
drugs (Porsolt et al. 1977; Steru et al. 1985). On this basis, drugs that reduce
immobility in the FST and TST are envisaged as potential antidepressants (Cryan
et al. 2005; McGonigle 2014). Both the FST and TST are easy to implement,
cost-effective, and allow reproducibility of results but do also have some limitations
(Cryan et al. 2005; McGonigle 2014). Thus, the FST and TST use a readout that is
based on decreased motor activity. Accordingly, drugs that increase or decrease
motor activity may lead to false-positive or false-negative results in those models.
Moreover, the significance of immobility observed in the FST and TST is disputed.
Immobility has originally been proposed to mimic symptoms of despair that may be
observed in human depression (Porsolt et al. 1977; Steru et al. 1985). However,
others have questioned this hypothesis and have suggested that immobility may be
a strategy that rodents use to cope with stressful stimuli (Cryan et al. 2005).
Additionally, it has been hypothesized that immobility may be analogous to reluc-
tance of depressed patients to maintain sustained expenditure of effort, rather than
to despair (Cryan et al. 2005). Furthermore, immobility observed in the FST and
TST can be reversed by the acute administration of antidepressants, which contrasts
with clinical evidence showing that the therapeutic effects of antidepressants
emerge only after repeated treatment (Cryan et al. 2005). Finally, the FST and
TST may have varying sensitivity towards different classes of antidepressants. For
example, in mice the TST may be more effective in detecting the efficacy of
selective serotonin reuptake inhibitors (SSRI), compared with the FST (Cryan
et al. 2005).
Experimental models of depression that use repeated exposure to stress are the
CMS and the chronic social stress. In the CMS model, rodents are exposed for a
period of 3–4 consecutive weeks to unpredictable stressful stimuli of mild intensity
and various nature (i.e., temporary food or water deprivation, soiled cage, small
modifications in light or temperature) (Willner 1997). Rodents subjected to the CMS
model eventually manifest a series of behavioral changes that are thought to mimic
symptoms of human depression, for example, reduction in grooming, exploratory,
and sexual behaviors (Willner 1997). Moreover, rodents subjected to the CMS
model display a decreased responsiveness to rewards, which is thought to model
anhedonia observed in major depression (Willner 1997). For example, rodents
subjected to CMS drink lower amounts of solutions containing 1% sucrose, com-
pared with non-stressed animals. Moreover, rats subjected to the CMS and treated
with amphetamine emit lower numbers of 50-kHz ultrasonic vocalizations (USVs),
a behavioral marker of positive emotional states (Brudzynski 2013), compared with
amphetamine-treated rats not exposed to stress (Kõiv et al. 2016). Finally, behavioral
changes observed in the CMS model can be effectively reversed by repeated, but not
acute, administration of antidepressant drugs with clinical efficacy. Based on all the
above considerations, CMS is considered a highly valuable animal model of depres-
sion, since it shows construct validity, face validity, and predictive validity (Willner
1997). Nevertheless, the potential of the CMS model in the study of antidepressant
drugs may be limited by time-demanding implementation and difficulties in
reproducing results across laboratories (Micale et al. 2013). Changes in behavior
similar to those featuring the CMS model can also be observed in rodents subjected
602 N. Simola
Experimental models of anxiety are all based on the induction of a state of conflict in
animals, which can be defined as the condition where an organism is both attracted
and repelled by the same stimulus (Ito and Lee 2016). Nevertheless, experimental
models of anxiety use different strategies to induce conflict in animals and can be
broadly categorized into ethological models, conflict-based models properly defined,
and models based on conditioned learning (McGonigle 2014).
Ethological models of anxiety evaluate spontaneous behaviors that animals
perform either in normal conditions or in the presence of natural aversive stimuli.
Very popular ethological models of anxiety are the elevated plus maze, the elevated
zero maze, the open field, and the light-dark compartment test. Those models
measure the reduction in the exploratory drive of rodents for novel environments
under conditions of bright light and/or elevation from the floor that may ultimately
cause an anxiety-like state of conflict (reviewed in Harro 2018). Two other popular
ethological models used in the screening of anxiolytic drugs are the social interaction
test and the defensive burying (reviewed in Harro 2018). The social interaction test
evaluates the time spent in nonaggressive contacts by two unfamiliar rats that are
placed together in a novel environment under bright light, whereas the defensive
burying measures the tendency of rodents to bury objects (i.e., marbles) that do not
have any emotional values. A conditioned version of the defensive burying model
also exists which evaluates the tendency of rodents to bury a non-electrified probe
that was previously associated with the delivery of shocks. However, the signifi-
cance of defensive burying is disputed, since the model has been suggested to better
recapitulate the features of obsessive-compulsive disorder rather than anxiety
(Albelda and Joel 2012). Differently from ethological models, conflict models of
anxiety do not evaluate the animals’ spontaneous behavior but rather performance of
motivated behaviors that are punished by mild electric shocks. The most common
model in this category is the Vogel test, which measures the suppression of drinking
in water-deprived rodents with free access to a bottle provided with an electrified
spout (Vogel et al. 1971). Finally, conditioned models of anxiety are based on the
Pavlovian association between a neutral stimulus, such as a light, and an aversive
stimulus, usually an electric shock (reviewed in Harro 2018). Animals that are
trained in conditioned paradigms of anxiety and subsequently exposed to the
formerly neutral stimulus display anticipatory behaviors which are thought to
indicate an anxiety-like state, such as increased levels of freezing behavior, startle
response, and emission of 22-kHz USVs, a behavioral marker of negative emotional
states in rats (Brudzynski 2013).
Several classes of drugs with clinical efficacy on anxiety attenuate behavioral
abnormalities in most of the abovementioned models, which makes those models
suited for the development of new anxiolytic agents. Nevertheless, animal models of
anxiety may have varying sensitivity for different classes of drugs with putative
anxiolytic potential. For example, the defensive burying test may disclose the
anxiolytic potential of SSRIs more effectively than other tests, like the elevated
plus maze (McGonigle 2014). Hence, a careful selection of animal models should be
604 N. Simola
done when evaluating novel anxiolytic agents, in order not to dismiss drugs that
may be potentially effective. Finally, caution is advisable when using conflict and
conditioned animal models which rely on procedures such as delivery of foot shocks,
presentation of visual or audio stimuli, and deprivation of water or food. In fact,
drugs that interfere with one or more sensorial processes may lead to false-positive or
false-negative findings in those models.
Studies in experimental animals are still a major source of information about the
subjective effects of drugs that possess rewarding and addictive properties. Animal
models used in this respect can be based on the evaluation of locomotor, condi-
tioned, or operant behavior.
Several drugs that possess rewarding properties stimulate locomotor activity in
experimental animals after acute administration, an effect that depends on the
activation of striatal and limbic dopaminergic transmission (Wise 1987). Moreover,
rodents repeatedly and intermittently treated with addictive drugs eventually develop
the so-called locomotor sensitization, which consists in the progressive increase of
locomotor activity in response to escalating administrations of the same dose of a
given drug. Locomotor sensitization has been envisaged as a behavioral correlate of
the neuroadaptive changes that addictive drugs induce in the dopaminergic circuits
that regulate reward and motivation (Robinson and Berridge 2000). On this basis,
locomotor sensitization has long been used as an animal model with which to study
drug addiction and its neurobiological mechanisms. However, it is still unclear
whether human addicts develop sensitization to the behavioral effects of drugs and
whether behavioral sensitization plays a role in drug dependence (Vanderschuren
and Pierce 2010). Moreover, it should be considered that locomotor sensitization
may also reflect the psychotomimetic and/or manic potential of drugs, particularly
when psychostimulants are administered (Seeman 2009a; Logan and McClung
2016). In fact, rodents that develop sensitization to the locomotor stimulating effects
of psychostimulants also display neurochemical changes similar to those observed in
experimental models of schizophrenia (Seeman 2009a).
Another popular rodent model that is used to study the rewarding and addictive
properties of drugs is the conditioned place preference (CPP), which is based on
drug-induced conditioning. The CPP model is performed in an apparatus made of
two compartments with different visual and/or tactile characteristics and involves a
conditioning phase and an expression phase. In the conditioning phase, whose
duration varies according to the drug tested, rodents are confined on alternate days
in either compartment of the apparatus and treated with the drug or its vehicle,
respectively. In the expression phase, rodents are left free to explore the whole
apparatus and should spend more time in the compartment previously paired with
drug administration. The preference of rodents for the drug-paired compartment in
the CPP model depends on the conditioning properties of drugs and is considered
Experimental Psychopharmacology 605
a behavioral marker of the positive effects that drugs may induce on the emotional
state (Tzschentke 1998). Nevertheless, it should be noted that drugs that induce CPP
definitively possess rewarding properties but do not necessarily have addiction poten-
tial (Tzschentke 1998). Conditioned behavior of rodents can also be used to study the
aversive properties of drugs in the conditioned place aversion model, which is similar
to CPP but evaluates avoidance of rodents for the drug-paired compartment
(Tzschentke 1998). Another popular rodent model used to evaluate the aversive
properties of drugs is the conditioned taste aversion, which evaluates the avoidance
of palatable fluids (i.e., solutions containing saccharin) that were previously paired
with the administration of drugs with aversive properties (Davis and Riley 2010).
Rewarding and addictive properties of drugs can also be evaluated by means of
paradigms of operant behavior, and a very reliable model in this regard is drug self-
administration. Several studies have demonstrated that rodents and primates may
perform drug self-administration, which can be based either on the principle of
operant reinforcement or on non-operant behavior (i.e., in the case of ethanol)
(reviewed in Sanchis-Segura and Spanagel 2006). The drug self-administration
model may reproduce in animals several stages of drug dependence observed in
human addicts and may predict the abuse potential of drugs. In fact, experimental
animals readily self-administer drugs that are highly addictive in humans (i.e.,
cocaine) but scarcely self-administer drugs that have a negligible abuse potential
(i.e., caffeine). The drug discrimination model is also widely used to evaluate the
rewarding and addictive properties of drugs (Stolerman et al. 2011) and involves a
training phase and a testing phase. In the training phase, animals must learn to
discriminate between the subjective effects of drugs and saline in a reinforced lever-
pressing procedure, where food is usually used as reinforcer. In the testing phase,
animals will press the lever associated with food in response to the injection of a drug
that possesses subjective effects similar to those of the drug used in the training
phase. Animals readily discriminate rewarding drugs from saline, and drugs with
similar subjective properties (i.e., amphetamine and cocaine) fully substitute for each
other in the drug discrimination model (Stolerman et al. 2011). Another model of
operant behavior that is widely used in studies of the rewarding and addictive
properties of drugs is the intracranial self-stimulation. In that model, rodents are
trained to self-administer electric pulses in brain regions that regulate reward and
motivated behavior. Pretreatment of animals with drugs that possess rewarding and/
or addictive properties reduces the intensity of pulses that sustain intracranial self-
stimulation behavior (reviewed in Sanchis-Segura and Spanagel 2006).
Animal models have greatly contributed to elucidate the neurobiological bases of
goal-directed behavior and drug addiction, and novel animal models are being
developed in order to investigate non-pharmacological human addictions, such as
pathological gambling (van den Bos et al. 2014). In addition, it is noteworthy that the
drug discrimination and drug self-administration models can be implemented in
humans, which confers them high translational value. Finally, animal models have
been extensively used to investigate new therapeutics for drug addiction although
with limited success, likely due to differences in preclinical and clinical endpoints
used to define drug addiction, abstinence, and relapse (Paterson 2011).
606 N. Simola
Animal models used to evaluate the antipsychotic potential of drugs can be broadly
subdivided in pharmacological and developmental (reviewed in Jones et al. 2011).
Pharmacological models of psychosis rely on the use of drugs that act on
dopaminergic or glutamatergic receptors, in order to reproduce in animals certain
behavioral abnormalities that are reminiscent of the symptoms of human psychosis.
This approach has both a theoretical and an empirical rationale. Thus, it has been
hypothesized that schizophrenia may stem from hyperfunction of the dopaminergic
mesolimbic system and hypofunction of the glutamatergic system (Seeman 2009b).
Besides, overt psychotic episodes have been consistently documented in healthy
humans after the intake of either dopaminomimetic psychostimulants (i.e., amphet-
amines, cocaine) or phencyclidine (PCP), an antagonist of glutamate N-methyl-D-
aspartate (NMDA) receptors (Steinpreis 1996). Models of psychosis that rely on
dopaminomimetic drugs use acute or repeated administration of either amphetamine
or apomorphine in rodents, in order to induce hyperactivity and stereotyped behavior
and/or to impair associative learning and avoidance behavior (reviewed in Jones
et al. 2011). Similar to what observed in models based on dopaminomimetic agents,
acute or repeated administration of PCP may induce hyperactivity in rodents, which
is envisaged to reproduce the hyperactivity and agitation featuring human psychosis
(Steinpreis 1996; Bubeníková-Valesová et al. 2008). Hyperactivity has also been
observed in rodents treated with other NMDA receptor antagonists, like dizocilpine
and ketamine (Bubeníková-Valesová et al. 2008). Moreover, administration of either
dopaminomimetic drugs or NMDA receptor antagonists to experimental animals
induces deficits in PPI, providing an experimental model with marked face validity
608 N. Simola
that reproduces a deficit in sensorimotor gating that may be observed in patients with
schizophrenia (reviewed in Jones et al. 2011). Experimental models of psychosis
based on dopaminomimetics or NMDA receptor antagonists are very popular but are
thought to reproduce in animals only the “positive” symptoms of schizophrenia.
However, administration of PCP to rodents has also been shown to induce behav-
ioral deficits that could reproduce the “negative” symptoms of schizophrenia. For
example, a recent study has demonstrated that PCP reduced the emission of 50-kHz
USVs in rats that engaged in heterospecific play with a familiar human, a procedure
known as “tickling” (Boulay et al. 2013). Since 50-kHz USVs are thought to mark
positive emotional states in rats (Brudzynski 2013), suppression of calling behavior
by PCP has been proposed to model the social withdrawal that may be observed in
schizophrenic patients (Boulay et al. 2013). Animal models of psychosis that rely on
dopaminomimetics or NMDA receptor antagonists can disclose the effects of anti-
psychotic drugs that have either a “typical” or “atypical” profile (reviewed in Jones
et al. 2011). Another pharmacological model exists that is widely used in the
screening of antipsychotic drugs and consists in the evaluation of drug-induced
catalepsy (Hoffman and Donovan 1995). In the catalepsy model, antipsychotic
drugs to be tested are not supposed to correct psychotic-like behavioral abnormal-
ities that are induced by previous pharmacological treatment. Conversely, an anti-
psychotic that is effective in the catalepsy model should induce in animals a
behavioral phenotype characterized by muscular stiffness and immobility, termed
as catalepsy, which stems from the antagonism of dopaminergic receptors in the
nigrostriatal tract. Accordingly, the catalepsy model is used to screen antipsychotic
drugs that are thought to possess a “typical” profile of action or, alternatively, to
clarify whether antipsychotic drugs effective in other animal models possess a
“typical” profile of action. The catalepsy model has neither face validity nor con-
struct validity for psychosis but is still widely used in drug discovery because it is
easy to implement, allows high reproducibility of results, has predictive validity, and
is cost-effective.
Neurodevelopmental animal models of psychosis have been characterized in
recent years starting from the evidence obtained in epidemiological and imaging
studies which suggests that schizophrenia may be a neurodevelopmental disorder
caused by the interaction between inherited factors and factors acquired during the
pre- and perinatal phases of development (reviewed in Mouri et al. 2013). Popular
neurodevelopmental models of psychosis are the lesion of the ventral hippocampus
in neonatal rodents, the treatment of pregnant rodents with the mitotic inhibitor
methylazoxymethanol, which reduces the proliferation of neuroblasts, and post-
weaning isolation of rodents. Other neurodevelopmental models of psychosis that
have recently emerged are the administration of polyriboinosinic-polyribocytidilic
acid or lipopolysaccharide to pregnant rodents (reviewed in Jones et al. 2011). The
latter models are thought to recapitulate in the offspring the neurochemical and
behavioral effects elicited by systemic viral or bacterial infection during pregnancy,
which has been suggested to be an environmental factor that may favor the mani-
festation of schizophrenia (reviewed in Mouri et al. 2013). Neurodevelopmental
models of psychosis all reproduce in animals enduring anatomical and behavioral
Experimental Psychopharmacology 609
Animal models used to study how drugs affect cognitive and executive functions can
be broadly subdivided in models based on non-operant or operant behavior.
Animal models based on non-operant behavior allow to study several aspects of
cognition, such as general cognitive function, acquisition and retention of spatial and
nonspatial memory, and learning. In detail, those models evaluate behaviors that are
readily performed by animals and that may be either spontaneous or motivated by
obtaining a reward or escaping from an unpleasant situation. For example, a model
commonly used to evaluate episodic memory in rodents is the novel object recog-
nition test, which is based on the innate exploratory drive that rodents have for
objects. In that test, the inability of rodents to discriminate between a novel object
and a familiar object previously encountered is thought to reflect deficits in episodic
memory (Ennaceur 2010). Models commonly used to study general cognitive
function evaluate the sequence of arm entries that rodents perform in Y-shaped or
T-shaped mazes; these models may also provide information about the integrity of
spatial memory and learning (Hughes 2004; van der Staay et al. 2011). Furthermore,
spatial learning and retention of spatial memory may be studied by means of the
radial arm maze or the Morris water maze, two models that evaluate the ability of
rodents to recognize baited arms previously visited or to locate and retrieve an
escape platform in a water tank, respectively (Morris 1984; van der Staay et al.
2011). Animal models for the study of cognition that evaluate non-operant behaviors
can be easily implemented in rats and mice, are cost-effective, and allow reproduc-
ibility of results across laboratories.
Animal models based on operant behavior may be used to study certain aspects of
executive functions, such as inhibitory control of behavior and impulsivity. Popular
models in this respect are the go/no-go, the stop-signal, the five-choice serial reaction
time (5-CSRT) tasks, and the delay discounting (reviewed in Bari and Robbins
2013). In all those models, rodents are requested to make lever-pressing or nose-
poking in order to obtain a reward and to adjust their behavior according to the
stimuli that are delivered during trials. The go/no-go task consists in a series of trials
where animals are exposed to stimuli that must either promote or inhibit response. In
some of the trials only the “go” signal is presented, which prompts animals to
respond. Conversely, in the remaining trials, the “go” signal is either presented
simultaneously with or preceded by another stimulus, the “no-go” signal that should
prevent animals from responding. Inability of animals to adjust responding after the
610 N. Simola
requires considerable expertise, which may somehow limit the use of those models
across different laboratories.
Animal models can be used to investigate the mechanisms that promote and sustain
the neuronal demise underlying neurodegenerative diseases. Moreover, animal
models can allow to reproduce at the preclinical level some of the cardinal behav-
ioral symptoms observed in patients suffering from neurodegenerative diseases, such
as the memory deficits featuring Alzheimer’s disease or the motor impairment
featuring Parkinson’s disease. Popular models used in preclinical studies of neuro-
degenerative diseases are based on the induction of damage in specific neuronal
populations of the animal brain, in order to reproduce the neurodegeneration featur-
ing human disease. For example, these models include the use of dopaminergic
neurotoxins, such as 6-hydroxydopamine (6-OHDA) or 1-methyl-4-phenyl-1,2,3,6-
tetrahydropyridine (MPTP), in the case of Parkinson’s disease, or the use of
either toxins or peptides that induce cholinergic dysfunction (i.e., β-amyloid,
streptozotocin) in the case of Alzheimer’s disease (Duty and Jenner 2011; Iqbal
et al. 2013; Salkovic-Petrisic et al. 2013; Simola et al. 2007). Several genetic models
of Alzheimer’s, Huntington’s, and Parkinson’s disease have also been developed
(Harvey et al. 2011). Although a detailed description of the numerous animal models
used in preclinical research on neurodegenerative diseases falls outside the aims of
the present chapter, it is relevant to mention that experimental models have recently
been developed which allow to study the neuropsychiatric-like behavioral abnor-
malities that may precede and/or accompany neurodegenerative diseases. For exam-
ple, new models have utilized discrete and bilateral neurotoxic lesions of the
nigrostriatal dopaminergic system induced by 6-OHDA in order to reproduce in
rodents the motivational deficits, anxiety-like, and depression-like behavioral abnor-
malities that may be observed in Parkinson’s disease (Magnard et al. 2016). Notably,
rodents subjected to discrete dopaminergic denervation do not display the overt
motor impairment featuring classic animal models of Parkinson’s disease that are
based on complete and unilateral denervation of the nigrostriatal tract (Magnard et al.
2016). Hence, rodent models based on discrete dopaminergic lesions may allow to
study non-motor symptoms of Parkinson’s disease at the preclinical level without the
interference of motor impairment. Indeed, the presence of motor deficits may
significantly affect the interpretation of results in preclinical studies of non-motor
symptoms of Parkinson’s disease that are performed in classic animal models of
dopaminergic denervation (Magnard et al. 2016). New animal models that reproduce
neuropsychiatric-like abnormalities associated with neurodegenerative diseases
appear highly relevant to experimental psychopharmacology. In fact, those models
may allow to study the cognitive and emotional symptoms of disease that are often
neglected and may also help to develop new drugs for treating those symptoms.
612 N. Simola
of 22-kHz USVs in rats conditioned to foot shocks (Molewijk et al. 1995). Measur-
ing 22-kHz USVs may also have relevance in the preclinical study of analgesic
drugs, since calling behavior could provide an experimental model with which to
study the affective dimension of pain (Oliveira and Barros 2006).
Emission of 50-kHz USVs is thought to be a marker of positive affect in young/
adult rats (Brudzynski 2013). Previous studies have demonstrated that drugs with
rewarding and addictive properties, such as amphetamines, cocaine, morphine, and
nicotine, may modify the number and acoustic parameters (bandwidth, duration,
maximum frequency) of 50-kHz USVs in rats (Barker et al. 2015; Costa et al. 2015;
Knutson et al. 1999; Simola et al. 2010, 2012, 2014, 2016; Simola and Morelli 2015;
Wright et al. 2010). Emission of 50-kHz USVs is increasingly being used to study
the motivational effects of drugs, either alone or combined with CPP or drug self-
administration paradigms (reviewed in Simola 2015). Notably, evaluating 50-kHz
USVs during self-administration procedures is a promising strategy that may dis-
close the emotional components inherent in drug abuse and the factors that may
promote relapse and reinstatement of drug taking (Barker et al. 2015). Moreover, it is
noteworthy that measuring 50-kHz USVs may disclose effects of drugs on the
emotional state that are not, or are only partially, revealed by other paradigms that
are classically used to study addiction (Barker et al. 2015; Simola and Morelli 2015).
Emission of 50-kHz USVs may also be relevant to fields of experimental psycho-
pharmacology other than drug addiction. Thus, it has been demonstrated that rats
selectively bred for low levels of 50-kHz USV emission showed blunted affectivity
and emitted fewer 50-kHz USVs in response to tickling by a familiar human,
compared with rats that had normal affectivity. This finding was thought to indicate
impairment in social communication and has led to a novel model of autism-like
social withdrawal (Harmon et al. 2008). Notably, the model has been shown to have
predictive validity, since it was able to detect the beneficial effects of the NDMA
receptor partial agonist GLYX-13, a drug that was reported to ameliorate social
withdrawal in autistic individuals (Moskal et al. 2011). Furthermore, as discussed in
the paragraph on animal models of psychosis, a reduced emission of 50-kHz USVs
has been observed in rats subjected to tickling that were treated with PCP (Boulay
et al. 2013). Deficits in calling behavior displayed by PCP-treated rats have been
proposed to model social withdrawal and could be useful to study the negative
symptoms of schizophrenia at the preclinical level. This hypothesis may be strength-
ened by the finding that buspirone, a drug that may attenuate the negative symptoms
of schizophrenia when added to antipsychotic therapy, restored calling behavior in
PCP-treated rats subjected to tickling (Boulay et al. 2013). Moreover, emission of
50-kHz USVs may be useful in studies of the modifications in the emotional state of
rats subjected to models of depression, as suggested by recent findings showing that
rats previously exposed to CMS emitted low numbers of 50-kHz USVs in response
to amphetamine administration (Kõiv et al. 2016).
Use of rat USVs as a marker of emotional states in experiments of psychophar-
macology is not devoid of limitations. For example, certain aspects related to the
behavioral significance of rat USVs and their relevance to the subjective effects of
psychoactive drugs are still to be elucidated (Simola and Costa 2018). Moreover,
614 N. Simola
certain drugs that possess clinically relevant effects may have a negligible influence
on the emission of USVs. Nevertheless, the evidence obtained in studies concerned
with rat USVs supports the idea that expanding the repertoire of behaviors that are
evaluated in animals is an urgent need in experimental psychopharmacology. This
approach shall hopefully be combined with other emerging techniques (i.e., set up of
developmental models, omics techniques, optogenetics), in order to obtain models
that adequately reproduce human disease in animals and that increase the reliability
and translational validity of results obtained in experiments of psychopharmacology.
Cross-References
References
Albelda N, Joel D. Animal models of obsessive-compulsive disorder: exploring pharmacology and
neural substrates. Neurosci Biobehav Rev. 2012;36:47–63.
Bari A, Robbins TW. Inhibition and impulsivity: behavioral and neural basis of response control.
Prog Neurobiol. 2013;108:44–79.
Barker DJ, Simmons SJ, West MO. Ultrasonic vocalizations as a measure of affect in preclinical
models of drug abuse: a review of current findings. Curr Neuropharmacol. 2015;13:193–210.
Bech P. Rating scales in depression: limitations and pitfalls. Dialogues Clin Neurosci.
2006;8:207–15.
Bogdanova OV, Kanekar S, D’Anci KE, Renshaw PF. Factors influencing behavior in the forced
swim test. Physiol Behav. 2013;118:227–39.
Boulay D, Ho-Van S, Bergis O, Avenet P, Griebel G. Phencyclidine decreases tickling-induced 50-
kHz ultrasound vocalizations in juvenile rats: a putative model of the negative symptoms of
schizophrenia? Behav Pharmacol. 2013;24:543–51.
Brooks SP, Jones L, Dunnett SB. Comparative analysis of pathology and behavioural phenotypes in
mouse models of Huntington’s disease. Brain Res Bull. 2012;88:81–93.
Brudzynski SM. Ethotransmission: communication of emotional states through ultrasonic vocali-
zation in rats. Curr Opin Neurobiol. 2013;23:310–7.
Bubeníková-Valesová V, Horácek J, Vrajová M, Höschl C. Models of schizophrenia in humans and
animals based on inhibition of NMDA receptors. Neurosci Biobehav Rev. 2008;32:1014–23.
Cappeliez P, Moore E. Effects of lithium on an amphetamine animal model of bipolar disorder. Prog
Neuro-Psychopharmacol Biol Psychiatry. 1990;14:347–58.
Carli M, Robbins TW, Evenden JL, Everitt BJ. Effects of lesions to ascending noradrenergic
neurones on performance of a 5-choice serial reaction task in rats; implications for theories of
dorsal noradrenergic bundle function based on selective attention and arousal. Behav Brain Res.
1983;9:361–80.
Experimental Psychopharmacology 615
Cechinel-Recco K, Valvassori SS, Varela RB, Resende WR, Arent CO, Vitto MF, Luz G,
de Souza CT, Quevedo J. Lithium and tamoxifen modulate cellular plasticity cascades in animal
model of mania. J Psychopharmacol. 2012;26:1594–604.
Costa G, Morelli M, Simola N. Involvement of glutamate NMDA receptors in the acute, long-term,
and conditioned effects of amphetamine on rat 50 kHz ultrasonic vocalizations. Int J Neuropsy-
chopharmacol. 2015;18:pyv057.
Coull JT. Getting the timing right: experimental protocols for investigating time with functional
neuroimaging and psychopharmacology. Adv Exp Med Biol. 2014;829:237–64.
Cryan JF, Mombereau C. In search of a depressed mouse: utility of models for studying depression-
related behavior in genetically modified mice. Mol Psychiatry. 2004;9:326–57.
Cryan JF, Mombereau C, Vassout A. The tail suspension test as a model for assessing antidepressant
activity: review of pharmacological and genetic studies in mice. Neurosci Biobehav Rev.
2005;29:571–625.
D’Hooge R, De Deyn PP. Applications of the Morris water maze in the study of learning and
memory. Brain Res Rev. 2001;36:60–90.
Davis CM, Riley AL. Conditioned taste aversion learning: implications for animal models of drug
abuse. Ann N Y Acad Sci. 2010;1187:247–75.
de Bartolomeis A, Buonaguro EF, Iasevoli F, Tomasetti C. The emerging role of dopamine-
glutamate interaction and of the postsynaptic density in bipolar disorder pathophysiology:
implications for treatment. J Psychopharmacol. 2014;28:505–26.
Dencker D, Husum H. Antimanic efficacy of retigabine in a proposed mouse model of bipolar
disorder. Behav Brain Res. 2010;207:78–83.
Drinkenburg WH, Ahnaou A, Ruigt GS. Pharmaco-EEG studies in animals: a history-based
introduction to contemporary translational applications. Neuropsychobiology. 2015;72:139–50.
Duty S, Jenner P. Animal models of Parkinson’s disease: a source of novel treatments and clues to
the cause of the disease. Br J Pharmacol. 2011;164:1357–91.
Einat H. Establishment of a battery of simple models for facets of bipolar disorder: a practical
approach to achieve increased validity, better screening and possible insights into endo-
phenotypes of disease. Behav Genet. 2007;37:244–55.
Ennaceur A. One-trial object recognition in rats and mice: methodological and theoretical issues.
Behav Brain Res. 2010;215:244–54.
Fischer DA, Ferger B, Kuschinsky K. Discrimination of morphine- and haloperidol-induced
muscular rigidity and akinesia/catalepsy in simple tests in rats. Behav Brain Res.
2002;134:317–21.
Gessa GL, Pani L, Fadda P, Fratta W. Sleep deprivation in the rat: an animal model of mania.
Eur Neuropsychopharmacol. 1995;5:89–93.
Geyer MA, Krebs-Thomson K, Braff DL, Swerdlow NR. Pharmacological studies of prepulse
inhibition models of sensorimotor gating deficits in schizophrenia: a decade in review. Psycho-
pharmacology. 2001;156:117–54.
Gutheil TG. Reflections on ethical issues in psychopharmacology: an American perspective. Int J
Law Psychiatry. 2012;35:387–91.
Harmon KM, Cromwell HC, Burgdorf J, Moskal JR, Brudzynski SM, Kroes RA, Panksepp J.
Rats selectively bred for low levels of 50 kHz ultrasonic vocalizations exhibit alterations in early
social motivation. Dev Psychobiol. 2008;50:322–31.
Harro J. Animals, anxiety, and anxiety disorders: how to measure anxiety in rodents and why.
Behav Brain Res. 2018;352:81–93.
Harvey BK, Richie CT, Hoffer BJ, Airavaara M. Transgenic animal models of neurodegeneration
based on human genetic studies. J Neural Transm. 2011;118:27–45.
Hendriksen H, Groenink L. Back to the future of psychopharmacology: a perspective on animal
models in drug discovery. Eur J Pharmacol. 2015;759:30–41.
Hendriksen H, Korte SM, Olivier B, Oosting RS. The olfactory bulbectomy model in mice and rat:
one story or two tails? Eur J Pharmacol. 2015;753:105–13.
Hoffman DC, Donovan H. Catalepsy as a rodent model for detecting antipsychotic drugs with
extrapyramidal side effect liability. Psychopharmacology. 1995;120:128–33.
616 N. Simola
Hughes RN. The value of spontaneous alternation behavior (SAB) as a test of retention in
pharmacological investigations of memory. Neurosci Biobehav Rev. 2004;28:497–505.
Iqbal K, Bolognin S, Wang X, Basurto-Islas G, Blanchard J, Tung YC. Animal models of the
sporadic form of Alzheimer’s disease: focus on the disease and not just the lesions. J Alzheimers
Dis. 2013;37:469–74.
Ito R, Lee AC. The role of the hippocampus in approach-avoidance conflict decision-making:
evidence from rodent and human studies. Behav Brain Res. 2016;313:345–57.
Jobert M, Wilson FJ, Ruigt GS, Brunovsky M, Prichep LS, Drinkenburg WH. IPEG Pharmaco-
EEG Guidelines Committee. Guidelines for the recording and evaluation of pharmaco-EEG data
in man: the International Pharmaco-EEG Society (IPEG). Neuropsychobiology.
2012;66:201–20.
Jones CA, Watson DJ, Fone KC. Animal models of schizophrenia. Br J Pharmacol.
2011;164:1162–94.
Keeney AJ, Hogg S. Behavioural consequences of repeated social defeat in the mouse: preliminary
evaluation of a potential animal model of depression. Behav Pharmacol. 1999;10:753–64.
Kennedy RT. Emerging trends in in vivo neurochemical monitoring by microdialysis. Curr Opin
Chem Biol. 2013;17:860–7.
Knutson B, Burgdorf J, Panksepp J. High-frequency ultrasonic vocalizations index conditioned
pharmacological reward in rats. Physiol Behav. 1999;66:639–43.
Kõiv K, Metelitsa M, Vares M, Tiitsaar K, Raudkivi K, Jaako K, Vulla K, Shimmo R, Harro J.
Chronic variable stress prevents amphetamine-elicited 50-kHz calls in rats with low positive
affectivity. Eur Neuropsychopharmacol. 2016;26:631–43.
Kokras N, Dalla C. Sex differences in animal models of psychiatric disorders. Br J Pharmacol.
2014;171:4595–619.
Lidow MS, Williams GV, Goldman-Rakic PS. The cerebral cortex: a case for common site of action
of antipsychotics. Trends Pharmacol Sci. 1998;19:136–40.
Logan RW, McClung CA. Animal models of bipolar mania: the past, present and future.
Neuroscience. 2016;321:163–88.
Magnard R, Vachez Y, Carcenac C, Krack P, David O, Savasta M, Boulet S, Carnicella S. What can
rodent models tell us about apathy and associated neuropsychiatric symptoms in Parkinson’s
disease? Transl Psychiatry. 2016;6:e753.
McGonigle P. Animal models of CNS disorders. Biochem Pharmacol. 2014;87:140–9.
Micale V, Kucerova J, Sulcova A. Leading compounds for the validation of animal models of
psychopathology. Cell Tissue Res. 2013;354:309–30.
Millan MJ, Goodwin GM, Meyer-Lindenberg A, Ove Ögren S. Learning from the past and looking
to the future: emerging perspectives for improving the treatment of psychiatric disorders.
Eur Neuropsychopharmacol. 2015;25:599–656.
Molewijk HE, van der Poel AM, Mos J, van der Heyden JA, Olivier B. Conditioned ultrasonic
distress vocalizations in adult male rats as a behavioural paradigm for screening anti-panic
drugs. Psychopharmacology. 1995;117:32–40.
Monterosso J, Ainslie G. Beyond discounting: possible experimental models of impulse control.
Psychopharmacology. 1999;146:339–47.
Moresco RM, Fazio F. SPET in psychopharmacology. Q J Nucl Med Mol Imaging.
2005;49:205–14.
Morris R. Developments of a water-maze procedure for studying spatial learning in the rat.
J Neurosci Methods. 1984;11:47–60.
Moskal JR, Burgdorf J, Kroes RA, Brudzynski SM, Panksepp J. A novel NMDA receptor glycine-
site partial agonist, GLYX-13, has therapeutic potential for the treatment of autism. Neurosci
Biobehav Rev. 2011;35:1982–8.
Mouri A, Nagai T, Ibi D, Yamada K. Animal models of schizophrenia for molecular and pharma-
cological intervention and potential candidate molecules. Neurobiol Dis. 2013;53:61–74.
Oliveira AR, Barros HM. Ultrasonic rat vocalizations during the formalin test: a measure of the
affective dimension of pain? Anesth Analg. 2006;102:832–9.
Experimental Psychopharmacology 617
Page GG, Opp MR, Kozachik SL. Sex differences in sleep, anhedonia, and HPA axis activity in a rat
model of chronic social defeat. Neurobiol Stress. 2016;3:105–13.
Paterson NE. Translational research in addiction: toward a framework for the development of novel
therapeutics. Biochem Pharmacol. 2011;81:1388–407.
Porsolt RD, Le Pichon M, Jalfre M. Depression: a new animal model sensitive to antidepressant
treatments. Nature. 1977;266:730–2.
Portfors CV. Types and functions of ultrasonic vocalizations in laboratory rats and mice. J Am
Assoc Lab Anim Sci. 2007;46:28–34.
Robbins TW. Cross-species studies of cognition relevant to drug discovery: a translational
approach. Br J Pharmacol. 2017;174:3191–9.
Robinson TE, Berridge KC. The psychology and neurobiology of addiction: an incentive-sensiti-
zation view. Addiction. 2000;95(Suppl 2):S91–S117.
Rodeberg NT, Sandberg SG, Johnson JA, Phillips PE, Wightman RM. Hitchhiker’s guide to
voltammetry: acute and chronic electrodes for in vivo fast-scan cyclic voltammetry. ACS
Chem Neurosci. 2017;8:221–34.
Salkovic-Petrisic M, Knezovic A, Hoyer S, Riederer P. What have we learned from the
streptozotocin-induced animal model of sporadic Alzheimer’s disease, about the therapeutic
strategies in Alzheimer’s research. J Neural Transm. 2013;120:233–52.
Sánchez C. Stress-induced vocalisation in adult animals. A valid model of anxiety? Eur J
Pharmacol. 2003;463:133–43.
Sanchis-Segura C, Spanagel R. Behavioural assessment of drug reinforcement and addictive
features in rodents: an overview. Addict Biol. 2006;11:2–38.
Sarnyai Z, Alsaif M, Bahn S, Ernst A, Guest PC, Hradetzky E, Kluge W, Stelzhammer V,
Wesseling H. Behavioral and molecular biomarkers in translational animal models for neuro-
psychiatric disorders. Int Rev Neurobiol. 2011;101:203–38.
Seeman P. Dopamine D2High receptors measured ex vivo are elevated in amphetamine-sensitized
animals. Synapse. 2009a;63:186–92.
Seeman P. Glutamate and dopamine components in schizophrenia. J Psychiatry Neurosci.
2009b;34:143–9.
Seligman ME, Beagley G. Learned helplessness in the rat. J Comp Physiol Psychol.
1975;88:534–41.
Shannon RJ, Carpenter KL, Guilfoyle MR, Helmy A, Hutchinson PJ. Cerebral microdialysis in clinical
studies of drugs: pharmacokinetic applications. J Pharmacokinet Pharmacodyn. 2013;40:343–58.
Simola N. Rat ultrasonic vocalizations and behavioral neuropharmacology: from the screening of
drugs to the study of disease. Curr Neuropharmacol. 2015;13:164–79.
Simola N, Brudzynski SM. Rat 50-kHz ultrasonic vocalizations as a tool in studying neurochemical
mechanisms that regulate positive emotional states. J Neurosci Methods. 2018;310:33–44.
Simola N, Costa G. Emission of categorized 50-kHz ultrasonic vocalizations in rats repeatedly
treated with amphetamine or apomorphine: possible relevance to drug-induced modifications in
the emotional state. Behav Brain Res. 2018;347:88–98.
Simola N, Morelli M. Repeated amphetamine administration and long-term effects on 50-kHz
ultrasonic vocalizations: possible relevance to the motivational and dopamine-stimulating
properties of the drug. Eur Neuropsychopharmacol. 2015;25:343–55.
Simola N, Morelli M, Carta AR. The 6-hydroxydopamine model of Parkinson’s disease. Neurotox
Res. 2007;11:151–67.
Simola N, Ma ST, Schallert T. Influence of acute caffeine on 50-kHz ultrasonic vocalizations in
male adult rats and relevance to caffeine-mediated psychopharmacological effects. Int J
Neuropsychopharmacol. 2010;13:123–32.
Simola N, Fenu S, Costa G, Pinna A, Plumitallo A, Morelli M. Pharmacological characterization of
50-kHz ultrasonic vocalizations in rats: comparison of the effects of different psychoactive
drugs and relevance in drug-induced reward. Neuropharmacology. 2012;63:224–34.
Simola N, Frau L, Plumitallo A, Morelli M. Direct and long-lasting effects elicited by repeated drug
administration on 50-kHz ultrasonic vocalizations are regulated differently: implications for the
618 N. Simola
Peter Riederer
Psychopharmacology: A Brief Overview of
Its History
Paul Foley
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 622
The Ancient and Medieval Periods: Opium, Hellebore, and Chandra . . . . . . . . . . . . . . . . . . . . . . . . 622
The Nineteenth Century: First Attempts at Scientific Pharmacology, and
the Need for Restraint . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 624
Tentative Explorations During the First Half of the Twentieth Century . . . . . . . . . . . . . . . . . . . . . . . 627
Chlorpromazine and Reserpine: The Beginning of the 1950s Psychopharmacology
Revolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 631
Haloperidol and Depot Antipsychotic Medications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 635
After Haloperidol: The Atypical Antipsychotics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 636
The Antidepressants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 638
Mood Stabilizers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 641
Anxiolytics and Minor Tranquilizers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 642
Hypnotica . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 643
Psychostimulants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 643
Nootropics (Cognitive Enhancers) and Anti-dementia Agents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 645
Antiparkinsonian Drugs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 646
Psychopharmacology and the Emergence of Modern Neurochemistry . . . . . . . . . . . . . . . . . . . . . . . . 647
Assessing the Effects of Modern Psychopharmacology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 651
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 653
Abstract
People have used diverse types of drugs and other preparations to ease their
physical pain and mental distress for thousands of years. The search for new,
more effective drugs gathered pace during the nineteenth century, before a
confluence of factors culminated in the birth of modern psychopharmacology
in the early 1950s, heralded by the discovery of the first neuroleptic and
P. Foley (*)
Medical Journal of Australia, Sydney, NSW, Australia
e-mail: pfoley@mja.com.au
Introduction
Pain has been eased and sleep facilitated in the Middle East and Europe by the juice
of the opium poppy since the Neolithic period, roles that opium and its chemical
derivatives have never completely surrendered. Its effects on mood were also noticed
Psychopharmacology: A Brief Overview of Its History 623
early; it has been suggested, for instance, that the drug Helen added to her guests’
wine in the Odyssey (in an early instance of polypharmacy!) to raise their spirits,
nepenthe (“without sorrow”), was opium. Wine and opium were consistently
recommended by Greek doctors for similar purposes, but also as calmatives,
reflecting their bivalent effects on mood. The most respected physician of classical
antiquity, Galen (c. AD 129–216), recognized that drugs could both cause (as
excitants) and lessen mental symptoms (as sedatives), and endorsed opium, man-
drake, and henbane as effective in some cases, but placed a greater emphasis on
dietary approaches to modifying mental states (Siegel 1973, pp. 275f.). Mandrake
(Mandragora) and henbane (Hyoscyamus), both members of the Solanaceae family,
continued to be used alongside the related belladonna as intoxicants or hallucinogens
and as medications until the early twentieth century. St John’s wort, on the other
hand, was long included in composita (theriacs) for treating wounds, but also as a
sedative and antidepressant (Tschupp 2004).
Galen also mentioned the most eminent psychoactive substance of the period,
hellebore, the popular reputation of which was so great that the adjective elleborosus
– literally, “in need of hellebore” – was used for “mad”; poets, satirists, and
philosophers frequently referred to the plant that grew especially well in Antikyra
on the Gulf of Corinth (Oppenheimer 1928, pp. 330f.). Two types were distin-
guished: black (various Helleborus species) and the unrelated white hellebore
(Veratrum album), the latter preferred for treating the symptoms of madness, partic-
ularly hallucinations, gloomy thoughts, and agitation. Hellebore, also used to reduce
fever and to slow the pulse, retained a place in the therapy of mental disorders until
the nineteenth century, when the first of a range of alkaloids were isolated from the
plant, many of which are now known to activate neuronal sodium channels (Barroso
2015).
The persistent humoral model of disease was long invoked to justify herbal
therapy for mental disorders; as it assumed the unity of mind and body, this model
did not distinguish between psychiatric and somatic disorders. Such therapies were
retained for centuries not, however, because they were rational, but because their
effects were convincing. Their perceived efficacy consequently provided no insights
into the nature of mental function and disturbances.
Of herbal preparations used outside Europe, the most relevant to later develop-
ments was Sarpagandha, or snakeroot (Rauvolfia serpentina), used in India, among
other purposes, for calming excited patients (Somers 1958). The root contains a
range of alkaloids that were isolated by Indian researchers in the twentieth century,
one of which – reserpine – played a major role in Western psychopharmacology
during the 1950s, as will be discussed below.
The psychotropic effects of naturally occurring substances were all too clear to
those who consumed grain contaminated by the ergot fungus (Claviceps purpurea),
particularly prominent in Europe from the early Middle Ages. Ergotism has been
held responsible for a range of bizarre epidemics, including outbreaks characterized
by manic or psychotic symptoms now attributed to the serotonergic actions of its
alkaloids, which include lysergic acid (Eadie 2003). It is also relevant that German
beer purity laws (including the Bavarian regulation of 1516) were partly motivated
624 P. Foley
by the desire to prevent brewers adulterating their products with psychotropic plants,
such as henbane (supplanted by the safer hops).
At the beginning of the nineteenth century, French alienist and reformer Philippe
Pinel (1745–1826) argued that prescribing drugs should not be the primary approach
to treating psychiatric patients, because “. . . waiting, assisted by a moral or physical
regime, often suffices, and in other cases the problem is beyond all help.” The doctor
should “reserve for extreme cases, hitherto regarded as incurable, the application of
certain active remedies, which other circumstances would render superfluous, harm-
ful, or reckless” (Pinel 1809, pp. 343f.). Pharmacological approaches were often
defended by practitioners by appeal to systems regarded by Pinel as outdated or
misguided, and the pharmacopeia of the early nineteenth century was indeed little
changed since Dioscorides had composed his Materia medica in the first century, a
limited catalog of sedatives, stimulants, purgatives, and emetics (Weber 1999).
The popularity of hellebore was in decline, but opium retained its place as an
antidepressant, supplied from the sixteenth century as a component of laudanum, a
composite of varying constitutions taken for a broad range of somatic and mental
indications, including pain, insomnia, and restlessness, and as a general panacea
(Dormandy 2012). Italian psychiatrist and reformer Vincenzo Chiarugi (1759–1820)
had employed opium as a sedative for asylum patients but was aware that long
term use entailed the risk of mental dulling (Schmitz 1925). In Germany, Friedrich
Engelken (1744–1815) treated agitated patients with opium in his private sanatorium
near Bremen, with an approach that his successors kept secret until Hermann
Engelken (1807–1881) finally published details of their increasing dose strategy in
1844. Mid-century, German psychiatrist Albrecht Erlenmeyer (1849–1926) classed
opium as the most valuable of the psychopharmaca, and it was widely used to treat
depression (including postnatal melancholy), mania, and hyperactivity in children
(Schmitz 1925; Weber 1987). By the end of the century, tinctures of opium were the
preferred forms of the drug, and they remained popular for early depression well into
the twentieth century, although in his 1925 review Bremen Nervenarzt Hans Schmitz
concluded that unpleasant side effects were common and cures not achieved
(Schmitz 1925). Hashish, digitalis, and alcohol also found applications as sedative
medications in psychiatry, as did ether and chloroform, solanaceous plant extracts,
and prussic acid (cyanide), while camphor, musk, and ammonia salts were employed
as “excitants” (see, for instance, Schneider 1824; Griesinger 1861, pp. 481–495).
Alcohol (as spirits, wine, or beer), on the other hand, lost favor in medicine, both
because of its unpredictable mixture of stimulating and sedative effects and because
of growing recognition of the deleterious social and mental effects of its misuse.
But a new era was beginning, one in which specific chemical substances were
preferred to herbal preparations, in the hope that more specific and controllable
effects could be achieved. In 1804, German pharmacist Friedrich Sertürner isolated
Psychopharmacology: A Brief Overview of Its History 625
morphine (from “Morpheus,” the god of dreams; Sertürner 1806) from the opium
poppy plant, the first of the alkaloids that would increasingly supplant less consistent
vegetable extracts, particularly for treating states of excitation and restlessness.
Morphine was used both for sedation and analgesia, as well as for treating opium
and alcohol addiction, until its own addictive properties were recognized. From the
late 1870s, cocaine, first isolated from the coca plant in 1855, was in turn employed
to treat morphinism until cocainism was decried as even more pernicious (Maier
1926). Morphine was soon joined by the belladonna alkaloids – hyoscyamine,
hyoscine (scopolamine), and atropine – which replaced their corresponding plant
preparations as mental and motor calmatives (Debreyne 1852; Harley 1869;
Kalischer 1910).
A series of further drugs were introduced during the nineteenth century for
reducing agitation and inducing sleep. Bromides were discovered to have sedative
(first major report: Behrend 1864) and anti-epileptic properties mid-century (Eadie
2012). Potassium bromide enjoyed massive popularity as an hospital and over-the-
counter sedative even after the deleterious effects of chronic intake (“bromism”)
were recognized in the 1870s (Sourkes 1991); it has been estimated that 40% of all
prescriptions by British general practitioners between the World Wars were for
bromides (Balme 1976). Bromides were also combined with other substances, as
in the treatment for epilepsy introduced by German psychiatrist Paul Flechsig
(1847–1929), in which bromide therapy was preceded by several weeks of slowly
elevated opium treatment (Flechsig 1893).
The first synthetic drug used in psychiatry was also a sedative, chloral hydrate.
First synthesized in 1832, Berlin pharmacologist Oscar Liebreich (1839–1908)
described its hypnotic properties in 1869, erroneously believing it could be
converted in vivo to chloroform (Butler 1970). Its ease of use and reliability led to
its quickly displacing opiates in both psychiatric and general practice, but it was in
turn largely displaced by morphine as a sedative after the 1870s and by barbiturates
for overcoming insomnia in the early twentieth century, although it was retained in
asylums into the 1940s because it was both effective and inexpensive (Sourkes
1992). Wilhelm Griesinger (1817–1868), the German Nervenarzt who famously
declared that mental disorders were essentially brain disorders, noted that patients
recovering from ether or chloroform narcosis often experienced remission of mel-
ancholia or mania, but also that the effect diminished with repeated attempts, ruling
out regular use of this early sleep therapy (Griesinger 1861, p. 489).
Inspired by insights gained during his 1836–1840 voyage thérapeutique through
the Middle East, French psychiatrist Jacques-Joseph Moreau (1804–1884) investi-
gated hashish both as a neuropsychiatric drug and as a tool for simulating the
symptoms of insanity – the first model psychosis, in which he related drug-induced
delirium to “mental pathogeny” – including self-experimentation that allowed him to
experience these symptoms from within. He reported some success in displacing
pathologic or disordered thoughts in patients with mania with more governable drug-
induced symptoms, but his limited supply of hashish limited his studies (Moreau de
Tours 1845). Moreau’s systematic investigations and his foundation of the Annales
médico-psychologiques as a forum for medical investigations of mental disorders
626 P. Foley
My goal was exclusively to formulate methods for precisely measuring the mental effects of
drugs in certain respects, and to show the possibility of their practical application, as well as
the usefulness of their results, with selected examples. (Kraepelin 1892, pp. 229f.)
. . . many derangements of the brain and mind, which are at present obscure, will become
accurately definable and amenable to precise treatment, and what is now an object of anxious
empiricism will become one for the proud exercise of exact science. (Thudichum 1884,
pp. 259f.)
Psychopharmacology: A Brief Overview of Its History 627
At about the same time, Kraepelin emphasized the probable role of metabolic
disturbances in the development of psychiatric disorders (for instance, Kraepelin
1899, p. 40); in 1914, Bleuler conceived schizophrenia as a “somatic, anatomic, or
chemical disorder” in which no significant brain lesion was evident (Bleuler 1914).
In 1904, Swedish-American chemist Otto Folin (1867–1934) wrote of the “wide-
spread hope that chemical investigations will, perhaps, sooner or later yield a clearer
understanding of [mental diseases] than can be looked for from the other sciences
that have been brought to bear on them” (Folin 1904). These and similarly minded
authors were, however, aware that knowledge in this area was as yet inadequate to
understand the chemistry of mental processes, let alone to provide insights into
rational therapies for dysfunctions. With few exceptions, the nascent field of neuro-
chemistry would remain largely restricted to separate chemo-anatomic and physio-
logical analyses until the mid-twentieth century. The first major textbook on
“neurochemistry,” by American Irvine Page in 1937, was, apart from a speculative
final chapter, one of general physiological chemistry (Page 1937).
Hypnagogic agents lacking the cardio-circulatory side effects of chloral hydrate were
identified during the final quarter of the nineteenth century, including paraldehyde,
urethane (ethyl carbamide), sulfonmethane, and their derivatives. Of greatest signif-
icance was the synthesis by Emil Fischer (1852–1919) and Josef von Mering
(1849–1908) of barbital (Veronal) in 1902, the comparative safety and tolerability
of which rapidly won it popularity in psychiatry, quickly displacing alternatives. The
potential for dependence and misuse of barbital (including for suicide, albeit at much
larger than therapeutic doses) were, however, also quickly recognized; during the
1950s, chlorpromazine was initially used primarily to facilitate withdrawal from
barbiturates or morphine. Barbital also served as the prototype for a range of
commercialized derivatives, including phenobarbital (Luminal; 1912), pentobarbital
(Neodorm, Nembutal; 1915), and amobarbital (sodium amytal, “truth serum”; 1923),
thereby presaging the post-Second World War approach to finding new agents
(Dundee and McIlroy 1982; López-Muñoz et al. 2004b).
Until the 1950s, barbiturate infusions roused patients from catatonic immobility
for hours, but with declining effectiveness as tolerance developed (Fink 2009). A
more radical use for barbiturates was in the “prolonged sleep therapies” introduced
during the early twentieth century, most famously the method of Swiss psychiatrist
Jakob Klaesi (1883–1980) in Bleuler’s Zürich clinic. In 1922, he described treating
patients with schizophrenia with a combination of Somnifen (barbital/
diallylbarbituric acid) and morphine, with the aim of providing a respite that might
allow stabilization of nervous function that promoted openness to psychotherapy on
waking, what might be termed today “re-booting the soul”; 12 of 30 patients
improved, but 7 died (Haenel 1979). In Europe, the approach was soon regarded
628 P. Foley
The effect of drugs on psychological functions has been the subject of remarkably little
investigation on the part of either psychologists or pharmacologists . . . the number of
contributions to what we may be permitted to call “psychopharmacology” is certainly very
meagre. (Macht 1920)
The torrent of sedative and hypnotic drugs that have flooded the market over the past decade
shows no sign of abating, and new ones are constantly commended under cleverly
constructed names, while others are already forgotten. Even of those that have held their
place in medical practice, few have been able to gain more general recognition; in almost
every case we see the same pattern in the literature: the initial news was extremely favorable
in every respect, but it soon became clear that they did not satisfy the old requirements of
cito, tuto et jucunde; that even in small doses they elicited all sorts of unpleasant or harmful
side effects. (Seige 1912)
From the 1930s until the 1950s, the somatic or shock therapies attracted broader
attention than pharmacological therapy for treating serious mental disorders. The
alleged benefits of sharp physical interventions for mental health had been described
since ancient times, but modern shock therapy commenced with Julius Wagner-
Jauregg (1857–1940) systematically exploiting the long recognized influence of
acute febrile disease on mood by artificially inducing fever, initially with tuberculin
and later with malaria parasites, in patients with syphilitic psychoses; his success
was recognized by the Nobel Prize committee in 1927. In 1933, Austrian-American
psychiatrist Manfred Sakel (1900–1957) described insulin coma therapy for treating
the symptoms of psychosis, depression, and morphine dependence: insulin-induced
hypoglycemia led to altered consciousness (not necessarily genuine coma) and
convulsions until terminated a few minutes later by the administration of glucagon.
A short time later, Hungarian psychiatrist Ladislas von Meduna (1896–1964)
induced convulsions in patients with schizophrenia with camphor, then with
pentetrazol (Cardiazol), motivated by the perceived antagonism between epilepsy
and schizophrenia (Fink 1984; Braslow 1999).
Chemical shock therapies were always controversial because of the risks of
permanent nervous or physical injury, including death (particularly in insulin coma
therapy), and because it was unclear whether the claimed success rates in terms of
the alleviation of mental symptoms were justified. Critics argued that the physical
exhaustion associated with these methods could explain short term symptomatic
improvements in behavior, while the nature of psychiatric disease meant that spon-
taneous remission was always a possibility. That these risky options were abandoned
in the mid-1950s (but not the successor technique, electroconvulsive therapy) was
attributable less to dismay with their undesirable effects as to the fact that the new
Psychopharmacology: A Brief Overview of Its History 631
The Second World War facilitated the rise of large commercial chemical and
pharmaceutical firms, with a string of achievements during the 1940s (antibiotics,
steroids, anti-tuberculosis drugs) fostering confidence that the industrial approach to
the systematic development of new agents would dominate future advances rather
than ad hoc investigations by individual researchers. Nevertheless, individual phy-
sicians and researchers and chance naturally played critical roles in the post-War
revolution in psychopharmacology. Even should developments during this period
not constitute a scientific revolution in the Kuhnian sense (Baumeister and Hawkins
2005), the seismic shifts in both theory and practice were undoubtedly revolutionary
by any other measure. It was during this period that the term “psychopharmacology”
entered both scientific and popular discourse as an everyday word, with a sharp rise
in the number of publications indexed under this word in MEDLINE, peaking in
1965 before slowing until a second rise from 1976, a rise that has not yet reached its
peak. The authors of an early review confessed that the “task of reviewing the
development of psychopharmacology is a formidable one, which only dedicated,
highly involved, or foolhardy individuals undertake. The problem arises from the
thousands of reports which have appeared in the past five years or so on the effects of
drugs as related to some aspect of behavior” (Ross and Cole 1960).
The roots of the revolution reach back to the first flowering of industrial organic
chemistry in the nineteenth century, particularly the interest in synthetic dyes. In
1876, methylene blue was synthesized at the German chemical giant BASF, its core
structure the phenothiazine nucleus (itself synthesized in 1886). The dye stained
several cell types, including nerve cells, and was also vasoconstrictive, prompting
investigation of its analgesic, hypnagogic, and psychotropic effects; Pietro Bodoni in
Genoa, for instance, reported the value of intramuscular injections of the dye for
calming patients with mania. Overshadowed by the new barbiturates, methylene
blue and phenothiazine were both largely forgotten by human medicine for many
years, although in 1938 the oxidative effects of methylene blue were reported to
ameliorate catatonic symptoms. More recently identified interesting features of the
dye have led to investigations of its usefulness for treating bipolar and neurodegen-
erative disorders (Howland 2016).
632 P. Foley
The results of our treatment of 38 patients were thoroughly encouraging. For most, a rapid
and spectacular effect could be achieved. And several cures or solid remissions were
observed within periods comparable with those of shock therapy. (Delay and Deniker 1952)
A patient who has been in restraint for two or three years continuously, only out for the legal
number of minutes per day under very close supervision, does not need restraint any more,
and yet he is not a well individual. He needs a great deal of extra attention from the
occupational therapy department, from the rehabilitation service, possibly from teachers
and other people who can reeducate him into useful channels. (Duval and Goldman 1956)
Sedation without sleepiness, less excited but not benumbed, inner peace without
paralysis: the optimism engendered by the initial experiences with neuroleptic
agents encouraged the search for further psychotropic phenothiazines. Pro-
chlorperazine (Stemetil; 1955) was much more potent as an antipsychotic than
chlorpromazine, but was not sedative. This first example of an “incisive agent,” as
psychiatrists in Lyon termed them, suggested that drugs with a more restricted set of
Psychopharmacology: A Brief Overview of Its History 635
specific effects were possible. But prochlorperazine also elicited excitomotor reac-
tions (dyskinesias) that could be influenced by mood and attention, reminiscent of
signs attributed to hysteria, thereby blurring the line assumed to divide psychiatry
from neurology (Lambert and Revol 1960; Brouillot et al. 2000).
This phenomenon, and growing recognition of the similarities of encephalitis
lethargica and phenothiazine therapy, led Delay and Deniker in 1955 to denote the
phenothiazine drugs (and reserpine) as “neuroleptic” agents; that is, drugs that
produce psychomotor indifference, useful for treating excitement and agitation, but
which also elicit extrapyramidal symptoms, with both the desired and “side” effects
explained by their action on the basal ganglia (Deniker 1989). “Neuroleptic”
was preferred to alternative suggestions (“tranquilizer,” “ataraxic”) at the Second
International Congress of Psychiatry in Zürich in 1957 and was adopted in Europe;
North Americans, however, favored “(major) tranquilizers” (King and Voruganti
2002).
Further phenothiazine neuroleptics soon followed, including perphenazine
(Trilafon; marketed 1957) and trifluoperazine (Stelazine, 1958), and in the 1960s
by the related thioxanthene antipsychotic agents, including clopenthixol (Sordinol;
1961), flupenthixol (Fluanxol; 1965), and thiothixene (Navane; 1967).
phenothiazines and incisive in terms of their lower propensity for sedating patients;
haloperidol consequently displaced chlorpromazine during the 1960s as the bench-
mark for antipsychotic efficacy.
Depot preparations for simplifying antipsychotic therapy, particularly for patients
with poor compliance, were explored during the early 1960s. The first commercially
available items were fluphenazine enanthate (Moditen) in 1963 and fluphenazine
decanoate (Prolixin) in 1965; a depot form of haloperidol became available in 1981.
Zuclopenthixol, the cis-isomer of the typical antipsychotic clopenthixol (introduced
1961), was itself introduced in 1978; it was supplied as a regular depot preparation
(Clopixol) from the 1980s, but also as a shorter acting formulation (2–4 days)
(Clopixol-Acuphase) for treating acute psychotic states (Healy 2002, pp. 231–234).
On the other hand, the dopamine hypothesis of schizophrenia and the apparent high
tolerance of many patients for antipsychotic drugs encouraged some psychiatrists to
administer very large doses of haloperidol, particularly as extrapyramidal reactions
could cease beyond a certain drug level. This approach was largely discredited in 1988
by separate reports that D2 dopamine receptors were effectively completely blocked at
low drug doses, and that clinical benefits were not increased beyond daily dosages of
600 mg chlorpromazine or 30 mg haloperidol (Healy 2002, pp. 234–238).
The Antidepressants
Opium and alcohol had long found favor for raising spirits; “nerves” might be treated
with laudanum, tonics such as strychnine, or anticholinergic alkaloids, especially
hyoscine; cocaine and (later) amphetamines could be used to boost mental energy.
Psychopharmacology: A Brief Overview of Its History 639
But these remedies were hardly antidepressants as the term is now understood
(Moncrieff 2008).
At about the same time that chlorpromazine was being discovered, Ernst Zeller
(1907–1987) and his Chicago colleagues reported that the anti-tuberculosis drug
isoniazid and an analog, iproniazid, each inhibited liver and kidney amine oxidases.
In the same year (1952), it was found that iproniazid markedly improved mood in
tuberculosis patients: they were more vivacious (“dancing in the halls tho’ there were
holes in their lungs”; Robitzek et al. 1952), their appetite, sleep, and sociability
improved – if in some cases to a more than desirable degree, causing psychomotor
agitation, hypersexuality, and psychotic symptoms. Clinical and public attention on
both sides of the Atlantic grew, and in 1958 chlorpromazine was marketed as
Marsilid. Cincinnati psychiatrist Max Lurie, one of the early investigators of isoni-
azid, had termed the new type of drug “antidepressant,” but Nathan S. Kline
(1916–1983), the head of research at Rockland State Hospital in New York who
had already shared the 1957 Lasker Clinical Medical Research Award for his
neuroleptic work and would win the 1964 prize for his antidepressant research,
initially achieved more traction with “psychic energizer” (López-Muñoz and Álamo
2009). Iproniazid, however, was withdrawn in many countries only a few years later
because of its hepatotoxicity. By this time, more potent but safer alternatives had
become available, including phenelzine (Nardil; described 1959) and isocarboxazid
(Marplan; 1960).
The monoamine oxidase (MAO)-inhibiting properties of tranylcypromine
(Parnate), a substituted amphetamine first synthesized in 1948, were not recognized
until 1959. It was first marketed in 1960, but was temporarily withdrawn a few years
later after a number of deaths were attributed to its use. It had also been found that
therapy with MAO inhibitors required avoiding tyramine-containing foods, includ-
ing red wine and aged cheese, as the accumulation of tyramine could lead to
potentially fatal hypertensive crises (“cheese effect”). As a result, and because of
their interactions with several other medications, MAO inhibitors lost favor as
antidepressants. But in the 1980s, it was discovered that MAO occurs in two
forms, and that the antidepressant effect is primarily associated with inhibition
of type A MAO. Reversible inhibitors of this isozyme not requiring stringent
dietary restrictions, such as moclobemide (Aurorix) – originally examined in 1972
as a lipid-lowering or antibiotic drug – have recently been investigated (Yáñez et al.
2012).
Reservations about MAO inhibitors were exacerbated by an alternative class of
antidepressants developed at about the same time. In 1956, Swiss psychiatrist
Roland Kuhn (1912–2005), searching for a medication more effective than opium
for “vital depressive mood disorders,” discovered the value of imipramine (Tofranil)
as a “thymoleptic.” Imipramine, the first tricyclic antidepressant, was originally
developed by Geigy (Basel) as an antihistamine, then as a neuroleptic (it includes
the chlorpromazine side chain, and its dibenzazepine nucleus resembles phenothia-
zine). Like opium, imipramine had some depressive effects, but Kuhn was satisfied
in 1958 that it specifically elevated mood in patients with depression of any etiology,
consistent with his view of depression as a “psychophysical phenomenon” that was
part of various conditions, although the effects of imipramine were less reliable in
640 P. Foley
Mood Stabilizers
The primary aim of a mood stabilizing agent is to reduce the severity of the mood
shifts in manic-depressive disorder, but they are also employed to manage borderline
personality and schizo-affective disorders. They are more effective for treating
mania than depression, but their mechanisms of action remain controversial.
The first (and hitherto only) specialist mood stabilizer was lithium. In 1886,
Danish psychologist Carl Lange (1834–1900) recommended lithium salts for
patients with manic-depressive illness on the basis that they suffered “cerebral
gout,” the effects of uric acid accumulation on nervous function, and lithium had
been found to dissolve urate stones (Schioldann 2011). Mineral spas including
lithium-containing waters had been increasingly praised for their health-promoting
effects through the nineteenth century. Lange’s treatment was as effective as any, but
was abandoned after its rationale lost favor, although the soft drink 7 Up was still
promoted as a “lithiated lemon-lime soda” when launched in 1929, and included
lithium citrate until 1948 (Aita et al. 1990). Lithium was still used for treating gout
until its sometimes fatal cardiac effects led to its being banned in the United States in
1949. In the same year, Australian psychiatrist John Cade (1912–1980) treated
people with schizophrenia and mania with the metal salt, having observed the
sedating effect of lithium urate in laboratory animals and of lithium carbonate in
himself. Danish psychiatrist Mogens Schou (1918–2005) undertook further studies,
including the first double-blind, placebo-controlled trial of a psychotropic agent,
which revived interest in the usefulness of lithium for treating mania. Schou and
colleagues later also provided evidence that lithium had prophylactic effects, but
these findings were more controversial. Nevertheless, the US Food and Drug
Administration (FDA) approved lithium for long-term therapy in 1974, and it is
recognized as effective for treating bipolar disorder, despite the original rationale
being discredited (Schou 2001).
A range of other agents have been employed as mood stabilizers, including
atypical antipsychotic agents also employed for treatment-resistant depression
(aripiprazole, risperidone and its analogs, olanzapine, quetiapine) and anti-convul-
sants (valproate, lamotrigine, carbamazepine) (overviews: Denicoff et al. 1998;
López-Muñoz et al. 2018).
642 P. Foley
Hypnotica
Barbiturates were the mainstay of hypnotic therapy until the 1960s; glutethimide
(introduced 1954) and meprobamate were regarded as safer alternatives until their
potential for misuse was also recognized (Hollister 1983). Methaqualone, originally
an antimalarial drug, was introduced in 1965 as Quaalude (in the United States; also
as Mandrax elsewhere), but its misuse (as “ludes,” “mandies”) led to benzodiaze-
pines being preferred for medical purposes from the late 1970s (Herzberg 2011).
Several quinazolinone analogs of methaqualone, all positive allosteric modulators of
GABAA receptors, have been marketed for the treatment of insomnia. Benzodiaze-
pines themselves have considerable misuse potential, and the risk of dependence is
great. The desire for safer alternatives led to development during the late 1980s of
non-benzodiazepines with similar effects on GABAA receptors and efficacy, but with
lower propensity for inducing tolerance and dependence: the “Z-drugs” zopiclone
(Imovane; marketed 1989), zolpidem (Ambien; 1992), and zaleplon (Sonata; 1999)
(Becker and Somiah 2015).
The orexin receptor antagonist suvorexant (Belsomra) has been available for
treating insomnia in Japan and the United States since 2014; orexin is a hypotha-
lamic neuropeptide involved in arousal and wakefulness. The pineal hormone
melatonin (discovered 1958) is critical to the regulation of circadian rhythms,
including the sleep-wake cycle; it has been reported to reduce time to sleep onset
without increasing total sleep time and to ameliorate the effects of shift work and jet
lag. Melatonin (and the melatonin receptor agonist ramelteon) may have greater
benefits for sleep in specific populations, such as children with neurodevelopmental
problems (Auld et al. 2017; Abdelgadir et al. 2018). The sedative effects of other
drug types have also been exploited as mild hypnotics, including antidepressants
(such as mirtazapine), atypical antipsychotics, antihistamines (particularly diphen-
hydramine), the α2 receptor agonist clonidine, and gabapentinoids (such as pre-
gabalin) (Atkin et al. 2018).
Psychostimulants
Stimulant substances have long been used informally in most cultures to improve
mood, endurance, and alertness, including tea, coffee, cocaine, khat, and tobacco.
Until the mid-twentieth century, stimulants also played a major role in psychophar-
macology, including “tonics” for alleviating melancholy (such as St John’s wort),
reducing nervousness, stimulating the appetite, or increasing mental energy. Since
the 1950s, they have largely fallen into disfavor because of misuse (amphetamines,
for instance) but also because they have been supplanted by agents directed at
specific symptoms, although these have also changed with time: in the Western
644 P. Foley
world, the neurasthenia of the nineteenth century became the nervousness of the
early twentieth; anxiety then dominated the post-Second World War, before depres-
sion became more common in the 1980s, followed by burnout in the early twenty-
first century. These conditions were absolute synonyms neither clinically nor
socially, but all would once have been indications for agents that stimulate the
nervous system.
Apart from their general stimulatory effects, amphetamines had been employed to
treat narcolepsy and as appetite suppressants since the 1920s. The first controlled
psychopharmacologic trial found in 1939 that amphetamines were efficacious for
treating depression (Dub and Lurie 1939; Rasmussen 2015), and they were used to
treat depression until the 1970s, but were ineffective in this regard for institutional-
ized patients and those with pychoses (Rasmussen 2006, 2015). It has been noted
that the same applies to later agents defined as being specifically antidepressant,
including the most recent SSRIs, whereas more sedative drugs can be effective in
these patients (Schatzberg 2003); indeed, it was noted as early as 1962 that it was
“well known that antidepressant drugs are of most use in the mild-to-moderate
degrees of depression” (Pare et al. 1962). As late as the 1980s, the usefulness of
amphetamines and methylphenidate for treating intransigent depression was advo-
cated by some authors (Chiarello and Cole 1987), but more recently doubt has been
cast on the evidence for their efficacy (Malhi et al. 2016).
A more surprising and enduring role for stimulants has been their use for
managing hyperactivity and improving attention, particularly in children and ado-
lescents. In 1937, American physician Charles Bradley (1902–1979) reported that
“a single dose of Benzedrine [could] produce a greater improvement in school
performance than the combined efforts of a capable staff working in a most
favorable setting,” but noted that the child’s environment and psychotherapy
were still critical. Concern about the effects of amphetamines was already suffi-
cient to prevent immediately pursuing this direction further. In 1944, methylphe-
nidate (Ritalin) was synthesized, and was used from the early 1950s for treating
depression, narcolepsy, and parkinsonism, before Bradley’s approach was resumed
in the 1960s by administering it to children then designated as having “minimal
brain dysfunction,” roughly equivalent to the attention deficit/hyperactivity disor-
ders defined in the DSM-III in 1980. Prevalence of this latter diagnosis rose
rapidly, particularly in the United States, from the 1990s, and correspondingly
the prescribing of methylphenidate. Its effectiveness for this indication is generally
accepted (and was tested in one of the first controlled drug trials in children), but
the question of whether the diagnosis is too broadly applied remains the subject of
ongoing discussion in medical, educational, and lay circles. Methylphenidate
occurs in four distinct stereoisomers, of which D-threo-methylphenidate
(dexmethylphenidate) is the pharmacologically active form of the norepineph-
rine/dopamine reuptake inhibitor (commercially available as Focalin) (Heal et al.
2013; Foley 2014). Lisdexamfetamine, an amphetamine prodrug with longer
action and lower abuse potential than standard amphetamines, has been approved
for treating attention deficit/hyperactivity disorder and binge eating in adults
(Roncero and Álvarez 2014).
Psychopharmacology: A Brief Overview of Its History 645
Drinamyl was an interesting product marketed by Smith, Kline and French from
1950 until the 1970s for relieving depression and anxiety. The combination of its two
components, amobarbital (50 mg) and D-amphetamine (5 mg), was designed to
balance their effects, but it was regarded as more stimulating than amphetamine
alone (for which reason it was popular with British mods as “purple hearts”).
Drinamyl was as effective as imipramine for treating depression in a controlled
trial, leading the authors to conclude that imipramine itself was not specifically
effective in this regard (Hare et al. 1964).
The newest category of psychotropic agent is also the most controversial. In 1972,
Romanian psychologist and pharmacologist Corneliu Giurgea (1923–1995) intro-
duced the concept of the “nootropic” as a drug that directly activated integrative
brain functions, enhancing cognitive performance by acting on the highest cortical
regions (Margineanu 2011). Giurgea applied the term to a drug he had synthesized in
the early 1960s, piracetam, the prototype for the racetam group, which includes
oxiracetam and aniracetam. Their nootropic properties are disputed, and sale of
piracetam as a nootropic is permitted in Europe but not in the United States. Its
mechanism of action is unclear, but increased cholinergic function in the prefrontal
cortex may be involved (Winblad 2005; Malykh and Sadaie 2010).
Claims of cognitive enhancement properties have been made for a diverse range
of other medicinal agents (Froestl et al. 2014). For example, modafinil (1998) and
armodafinil (2007), which promote wakefulness and alertness (eugeroics) and are
therefore primarily prescribed for treating narcolepsy, have been reported to improve
cognitive function (Battleday and Brem 2015). Recent analyses suggest that low
doses of amphetamine or methylphenidate may improve cognitive functions, includ-
ing inhibitory control, memory, and aspects of attention, as well as increase moti-
vation (Ilieva et al. 2015; Marraccini et al. 2016).
The various nootropic candidates exhibit a diverse range of pharmacologic
properties, and their effectiveness is contentious. A recent meta-analysis found
some evidence of mild benefits in certain cognitive domains, but the authors noted
that the number of investigations has been limited by ethical considerations regard-
ing “brain doping,” and that many studies have returned negative results (Fond et al.
2015).
Other putative nootropics achieve some degree of symptomatic cognitive
improvement for patients with Alzheimer dementia, but none slow the course of
the disease (reviewed: Aisen et al. 2012):
Antiparkinsonian Drugs
At the end of the 1960s, Robert Schwab discovered that amantadine, an antiviral
agent used for treating influenza, was also antiparkinsonian. Only much later was it
established that amantadine is an NMDA glutamate receptor antagonist; it also
increases dopamine release and blocks its reuptake. It has found particular use as
an adjunct to L-DOPA in later stage disease and for relieving akinetic crises.
Budipine is a further antiparkinsonian agent that antagonizes NMDA receptors,
but amantadine may also be anticholinergic and enhance dopamine release. The
potential role of NMDA receptor antagonists in managing the psychiatric symptoms
of Parkinson disease has also been recently explored (Vanle et al. 2018).
motivated by the problem of surgical shock. The corollary of this circumstance was
that the neuroleptics were accepted because they worked, not because they fitted
with concepts of how psychiatric disorders develop; nor was it known how drugs
modulated brain processes, let alone mental functions. Whatever their mechanism,
there was renewed hope that specific psychiatric processes might be amenable to
pharmacologic modulation beyond the traditional poles of sedation and stimulation,
a possibility entertained by Freud toward the end of his life:
Here we are concerned with therapy only with respect to its use of psychological means: at
the moment we have no alternatives. The future may teach us how to directly influence
energy resources and their distribution in the psychic apparatus by means of particular
chemical substances. (Freud 1940, p. 44)
On the other hand, antihistamine research had shown that new agents could be
systematically sought by modifying the structures of existing agents and screening
the results for interesting properties.
Another gradual but key change throughout the slow revolution initiated by
chlorpromazine was the shift in emphasis from the safety to the efficacy of psycho-
tropic drugs; for a new agent to be sanctioned by industry, clinicians, and regulatory
authorities, it was no longer sufficient that it do no harm: it must effectively improve
specific aspects of mood and behavior. This was reflected by the move from ad hoc
testing of new agents in single institutions to multicenter controlled trials. In the
United States, the National Institute of Mental Health created the Psychopharma-
cology Research Center in 1956 to formalize clinical testing of psychotropic agents,
leading to several investigations during the early 1960s that confirmed their efficacy,
including the 1964 nine-hospital, double-blind controlled trial of chlorpromazine
therapy (Healy 2002, pp. 99f.).
Further, rather than reflecting neurochemical models of the brain, the new drugs
informed the development of such models. For example, reports that chlorpromazine
could resolve the psychosis-like states induced by hallucinogens validated model
psychoses based on these effects, including transmethylation and serotonin hypoth-
eses, while amine depletion by reserpine provided clues about the roles played by
chemical modulators in mood (Baumeister et al. 2003; Baumeister and Hawkins
2005). The psychopharmacology revolution was stoked by an interaction between
basic and applied biomedical research that was unprecedented in the history of
psychiatry. The combination of initial serendipity exploited by perceptive observers
with systematic assessment and follow-up studies provided short term successes in
the clinic, but also longer term insights that would eventually reform psychophar-
macology by providing a rational (if incomplete) basis for its endeavors.
That neurochemistry and psychopharmacology should both blossom during the
1950s partially reflected their conceptual cross-fertilization. Two concepts were key to
their interaction: that mental processes were somehow related to chemical transmis-
sion in the brain (not accepted by all pharmacologists or physiologists at the start of the
decade), and that the effects of psychotropic agents on mental and cognitive function
provided clues about how chemical transmission operated in health and in disease.
Psychopharmacology: A Brief Overview of Its History 649
The 1960s were the golden age of the monoamine transmitters, so much enthusiasm, so
much hope. Many tools were developed and much knowledge accumulated, with rapid
progress on all frontiers — pharmacology (virtually every step of monoamine transmission
could be influenced by a drug), pathology (Parkinson’s disease, schizophrenia, depression),
metabolites in cerebrospinal fluid and, not least, the histochemistry. (Hökfelt 2010)
even as a means of social control. The aim must be to constantly improve how
psychoactive drugs are deployed rather than to damn them altogether, to apply the
same careful clinical acuity and care that initiated the rapid advance of psychophar-
macology in the first place.
Marked differences between the effects experienced by people taking a particular
medication mean that listening to the individual taking them is critical. Personalized
medicine is a higher priority when treating patients with mental or cognitive
problems than in other areas of medicine; as Roland Kuhn reflected in 2005:
Instead of large statistical analyses with numerous patients and always applying the same
methods and asking the same questions, a return to the individual patient is needed. This
happens in open conversation between the patient and doctor who asks about past symptoms
and the effects of treatments already tried . . . In this manner, a unity of pharmacologic and
psychological therapy can be achieved that allows the selection, dosing and administration
of drugs to be evaluated in as short a time as possible. The more accurately the patient is
examined, the greater are the chances of success with minimal disturbing side effects. (Kuhn
2005)
Strangely enough, [its impact] is often difficult to define, although most observers —
doctors, nurses and, last not least, the patient — agree that a change has occurred. The
essence of the transformation, it seems, is not so much reflected in what the patient
experiences in a novel manner as in what ceases to disturb him; not the spontaneous
expression of new thoughts and feelings, but the attenuation of depressive self-concern is
indicative of the pharmacotherapeutic effect. Conversely, if [imipramine] fails, the patient
remains as he was. (Freyhan 1959)
Symptoms, not diseases, are treated in psychiatry: there are no genuinely “anti-
schizophrenia” or “anti-anxiety” drugs; a medication from one class will probably be
useful for some patients with other problems. The exercise is complicated by the
importance of individual differences in set and setting, of the psychology and
physiology of the person taking a drug and the circumstances in which they are
taking it. Further, a drug that benefits “only” a small number of patients in a clinical
trial may not be a failure, but rather a tool that identifies people with a specific
treatable dysfunction. We have learned a great deal about the brain and the mind
from the effects of psychotropic agents in patients, but they may also point to
shortcomings of the Kraepelinian nosology of psychiatric disorders or the concept,
for instance, of “depressive illness.” Re-examination of older concepts of syndromes
or symptom complexes in psychiatry, rather than diseases, may be rewarding.
Psychopharmacology: A Brief Overview of Its History 653
One of the speakers yesterday mentioned that the court of last appeal is the practical effect on
patients. We must not forget, however, that basic research can also be done in man, and
clinical studies can contribute a great deal to a fundamental understanding of how these
agents act and what are the roles of the monoamines in the brain. (Kety 1962)
Too stringent a focus on pursuing what is already unknown can divert attention from
what lies just outside one’s field of view. Goal-directed development is indubitably
valuable, but breakthroughs and radically new insights usually come from unex-
pected quarters.
References
Abdelgadir IS, Gordon MA, Akobeng AK. Melatonin for the management of sleep problems in
children with neurodevelopmental disorders: a systematic review and meta-analysis. Arch Dis
Child. 2018;103:1155–62.
Aisen PS, Cummings J, Schneider LS. Symptomatic and nonamyloid/tau based pharmacologic
treatment for Alzheimer disease. Cold Spring Harb Perspect Med. 2012;2:a006395.
Aita JF, Aita JA, Aita VA. 7-Up anti-acid lithiated lemon soda or early medicinal use of lithium.
Nebr Med J. 1990;75:277–80.
Angrist BM. The neurobiologically active benzamides and related compounds: some historical
aspects. In: Rotrosen J, Stanley M, editors. The benzamides: pharmacology, neurobiology, and
654 P. Foley
clinical effects (Advances in biochemical psychopharmacology; 35). New York: Raven Press;
1982. pp. 1–6.
Ashok AH, Marques TR, Jauhar S, Nour MM, Goodwin GM, Young AH, Howes OD. The
dopamine hypothesis of bipolar affective disorder: the state of the art and implications for
treatment. Mol Psychiatry. 2017;22:666–79.
Atkin T, Comai S, Gobbi G. Drugs for insomnia beyond benzodiazepines: pharmacology, clinical
applications, and discovery. Pharmacol Rev. 2018;70:197–245.
Auld F, Maschauer EL, Morrison I, Skene DJ, Riha RL. Evidence for the efficacy of melatonin in
the treatment of primary adult sleep disorders. Sleep Med Rev. 2017;34:10–22.
Awouters FH, Lewi PJ. Forty years of antipsychotic drug research – from haloperidol to
paliperidone – with Dr. Paul Janssen. Arzneimittelforschung. 2007;57:625–32.
Ayd FJ. Fatal hyperpyrexia during chlorpromazine therapy. J Clin Exp Psychopathol.
1956;17:189–92.
Baldwin DS, Anderson IM, Nutt DJ, Christer A, Bandelow B, den Boer JA, Christmas DM, Davies
S, Fineberg N, Lidbetter N, Malizia A, McCrone P, Nabarro D, O’Neill C, Scott J, van der Wee
N, Wittchen H-U. Evidence-based pharmacological treatment of anxiety disorders, post-trau-
matic stress disorder and obsessive-compulsive disorder: a revision of the 2005 guidelines from
the British Association for Psychopharmacology. J Psychopharmacol. 2014;28:403–39.
Balme R. Early medicinal uses of bromides. J R Coll Physicians Edinb. 1976;10:205–8.
Ban TA. Fifty years chlorpromazine: a historical perspective. Neuropsychiatr Dis Treat.
2007;3:495–500.
Barbara J-G. History of psychopharmacology: from functional restitution to functional enhance-
ment. In: Clausen J, Levy N, editors. Handbook of neuroethics. Dordrecht: Springer; 2015.
pp. 489–504.
Barroso MS. The hellebore, the plant beloved by the Greeks: the reasons behind a myth. Vesalius.
2015;21:30–7.
Battleday RM, Brem AK. Modafinil for cognitive neuroenhancement in healthy non-sleep-deprived
subjects: a systematic review. Eur Neuropsychopharmacol. 2015;25:1865–81.
Baumeister AA. The chlorpromazine enigma. J Hist Neurosci. 2013;22:14–29.
Baumeister AA, Francis JL. Historical development of the dopamine hypothesis of schizophrenia.
J Hist Neurosci. 2002;11:265–77.
Baumeister AA, Hawkins MF. Continuity and discontinuity in the historical development of
modern psychopharmacology. J Hist Neurosci. 2005;14:199–209.
Baumeister AA, Hawkins MF, Uzelac SM. The myth of reserpine-induced depression: role in the
historical development of the monoamine hypothesis. J Hist Neurosci. 2003;12:207–20.
Becker PM, Somiah M. Non-benzodiazepine receptor agonists for insomnia. Sleep Med Clin.
2015;10:57–76.
Bennett MR. Monoaminergic synapses and schizophrenia: 45 years of neuroleptics.
J Psychopharmacol. 1998;12:289–304.
Berger H. Zur Pathogenese des katatonischen Stupors. Münch med Wschr. 1921;68:448–50.
Beringer K. Der Meskalinrausch: Seine Geschichte und Erscheinungsweise. Berlin: Julius Springer;
1927.
Bleuler E. Die Kritiken der Schizophrenien. Zschr f d ges Neurol u Psychiat. 1914;22:19–44.
Bossong F. Erinnerung an Roland Kuhn (1912–2005) und 50 Jahre Imipramin. Nervenarzt.
2008;79:1080–6.
Braslow JT. History and evidence-based medicine: lessons from the history of somatic treatments
from the 1900s to the 1950s. Ment Health Serv Res. 1999;1:231–40.
Brouillot P, Broussolle P, Greffe J, Guyotat J, Lambert P, Lemoine P. The birth of
psychopharmacotherapy: explorations in a new world 1952–1968. In: Healy D, editor. The
psychopharmacologists, vol. 3. London: Arnold; 2000. pp. 1–54.
Brown WA, Rosdolsky M. The clinical discovery of imipramine. Am J Psychiatr. 2015;172:426–9.
Butler TC. The introduction of chloral hydrate into medical practice. Bull Hist Med.
1970;44:168–72.
Psychopharmacology: A Brief Overview of Its History 655
Carlsson M, Carlsson A. Interactions between glutamatergic and monoaminergic systems within the
basal ganglia – implications for schizophrenia and Parkinson’s disease. Trends Neurosci.
1990;13:272–6.
Caroff SN. The neuroleptic malignant syndrome. J Clin Psychiatry. 1980;41:79–83.
Carpenter WT, Davis JM. Another view of the history of antipsychotic drug discovery and
development. Mol Psychiatry. 2012;17:1168–73.
Castro Caldas A, Teodoro T, Ferreira JJ. The launch of opicapone for Parkinson’s disease: negatives
versus positives. Expert Opin Drug Saf. 2018;17:331–7.
Chiarello RJ, Cole JO. The use of psychostimulants in general psychiatry: a reconsideration.
Arch Gen Psychiatry. 1987;44:286–95.
Cipriani A, Furukawa TA, Salanti G, Chaimani A, Atkinson LZ, Ogawa Y, Leucht S, Ruhe HG,
Turner EH, Higgins JPT, Egger M, Takeshima N, Hayasaka Y, Imai H, Shinohara K, Tajika A,
Ioannidis JPA, Geddes JR. Comparative efficacy and acceptability of 21 antidepressant drugs for
the acute treatment of adults with major depressive disorder: a systematic review and network
meta-analysis. Lancet. 2018;391:1357–66.
Clouston TS. Clinical lectures on mental diseases. London: J. & A. Churchill; 1887.
Colonna L. Antideficit properties of neuroleptics. Acta Psychiatr Scand. 1994;89(Suppl.
380):77–82.
Contestabile A. The history of the cholinergic hypothesis. Behav Brain Res. 2011;221:334–40.
Coppen A. The biochemistry of affective disorders. Br J Psychiatry. 1967;113:1237–64.
Corponi F, Fabbri C, Bitter I, Montgomery S, Vieta E, Kasper S, Pallanti S, Serretti A. Novel
antipsychotics specificity profile: a clinically oriented review of lurasidone, brexpiprazole,
cariprazine and lumateperone. Eur Neuropsychopharmacol. 2019;29:971–85.
Costentin J. Une nouvelle approche de la prise en charge de la schizophrénie: les agonistes partiels
des récepteurs D2 de la dopamine. Ann Pharm Fr. 2009;67:310–9.
Crilly J. The history of clozapine and its emergence in the US market: a review and analysis. Hist
Psychiatry. 2007;18:39–60.
D’Abreu A, Akbar U, Friedmanac JH. Tardive dyskinesia: epidemiology. J Neurol Sci.
2018;389:17–20.
Dale E, Bang-Andersen B, Sánchez C. Emerging mechanisms and treatments for depression beyond
SSRIs and SNRIs. Biochem Pharmacol. 2015;95:81–97.
Davis KL, Kahn R, Ko G, Davidson M. Dopamine in schizophrenia: a review and
reconceptualization. Am J Psychiatr. 1991;148:1474–86.
Debreyne. Des vertus thérapeutiques de la belladone. Paris: J.-B. Ballière; 1852.
Degkwitz R. Leitfaden der Psychopharmakologie für Klinik und Praxis. Stuttgart:
Wissenschaftliche Verlags-Gesellschaft; 1967.
Delay J, Deniker P. Trente-huit cas de psychoses traitées par la cure prolongée et continue de 4560
R. P. C. R. In: Congrès des médecins aliénistes et neurologistes de France et des pays de langue
française, Le session, Luxembourg, 21–27 juillet 1952. Paris: Masson & Cie; 1952. pp. 503–13.
Delay J, Pichot P, Lemperiere T, Glissalde B, Peigne F. Le neuroleptique majeur non
phenothiazinique et non reserpinique, l’haloperidol, dans le traitement es psychoses [Séance
du 21 Decembre 1959]. Annales médico-psychologiques (Paris). 1960;118:145–52.
Denicoff KD, Frye MA, Dunn RT, Leverich GS, Osuch E, Speer A. A history of the use of
anticonvulsants as mood stabilizers in the last two decades of the 20th century. Neuropsychobiology.
1998;38:152–66.
Deniker P. From chlorpromazine to tardive dyskinesia (brief history of the neuroleptics). Psychiatr
J Univ Ott. 1989;14:253–9.
Dormandy T. Opium: reality’s dark dream. New Haven: Yale University Press; 2012.
Douchamps V, Mathis C. A second wind for the cholinergic system in Alzheimer’s therapy. Behav
Pharmacol. 2017;28:112–23.
Dub LA, Lurie LA. Use of benzedrine in the depressed phase of the psychotic state. Ohio State Med
J. 1939;35:39–45.
Dundee JW, McIlroy PDA. The history of the barbiturates. Anaesthesia. 1982;37:726–34.
656 P. Foley
Duval AM, Goldman D. The new drugs (chlorpromazine & reserpine): administrative aspects.
Ment Hosp. 1956:7, 30–34. (reprinted: Psychiatric Services 2000; 2051:2327–2331)
Eadie MJ. Convulsive ergotism: epidemics of the serotonin syndrome? Lancet Neurol.
2003;2:429–34.
Eadie MJ. Sir Charles Locock and potassium bromide. J R Coll Physicians Edinb. 2012;42:274–9.
Fangmann P, Assion H-J, Juckel G, González CÁ, López-Muñoz F. Half a century of antidepressant
drugs. On the clinical introduction of monoamine oxidase inhibitors, tricyclics, and tetracyclics.
Part II: tricyclics and tetracyclics. J Clin Psychopharmacol. 2008;28:1–4.
Fink M. Meduna and the origins of convulsive therapy. Am J Psychiatr. 1984;141:1034–41.
Fink M. Catatonia: a syndrome appears, disappears, and is rediscovered. Can J Psychiatr Rev Can
Psychiatr. 2009;54:437–45.
Flechsig P. Über eine neue Behandlungsmethode der Epilepsie. Neurol Centralbl. 1893;12:229–31.
Foley PB. Beans, roots and leaves. A history of the chemical therapy of parkinsonism. Marburg:
Tectum Verlag; 2003.
Foley P. Succi nervorum: a brief history of neurochemistry. J Neural Transm Suppl. 2008;72:5–15.
Foley PB. Sons and daughters beyond your control: episodes in the prehistory of the attention
deficit/hyperactivity syndrome. Atten Defic Hyperact Disord. 2014;6:125–51.
Foley PB. Encephalitis lethargica. The mind and brain virus. New York: Springer; 2018.
Folin O. Some metabolism studies. With special reference to mental disorders. Am J Insanity.
1904;60:699–732.
Fond G, Micoulaud-Franchi J-A, Brunel L, Macgregor A, Miot S, Lopez R, Richieri R, Abbar M,
Lancon C, Repantis D. Innovative mechanisms of action for pharmaceutical cognitive
enhancement: a systematic review. Psychiatry Res. 2015;229:12–20.
Freeman W. Psychochemistry: some physicochemical factors in mental disorders. J Am Med Assoc.
1931;97:293–6.
Freud S. Abriß der Psychoanalyse (aus dem Nachlaß). Internationalen Zeitschrift für Psychoanalyse
und Imago. 1940;25:7–67.
Freyhan FA. Comments on the biological and psychopathological basis of individual variation in
chlorpromazine therapy. L’Encéphale. 1956;45:913–9.
Freyhan FA. Clinical effectiveness of Tofranil in the treatment of depressive psychoses.
Can Psychiatr Assoc J. 1959;4(Suppl):S86–99.
Froestl W, Pfeifer A, Muhs A. Cognitive enhancers (nootropics). Update 2014. J Alzheimers Dis.
2014;42:1–68,961–1019,1079–1149
Gjessing R. Beiträge zur Kenntnis der Pathophysiologie periodisch katatoner Zustände. IV.
Mitteilung. Versuch einer Ausgleichung der Funktionsstörungen. Arch Psychiatr Nervenkr.
1939;109:525–95.
Granger B, Albu S. The haloperidol story. Ann Clin Psychiatry. 2005;17:137–40.
Griesinger W. Die Pathologie und Therapie der psychischen Krankheiten für Ärzte und Studirende.
2nd ed. Stuttgart: Adolph Krabbe; 1861.
Haddad PM, Correll CU. The acute efficacy of antipsychotics in schizophrenia: a review of recent
meta-analyses. Ther Adv Psychopharmacol. 2018;8:303–18.
Haenel T. Jakob Klaesi: Schlafkur und Antieidodiathese. Gesnerus. 1979;36:246–65.
Hanzlik PJ. Purkinje’s pioneer self-experiments in psychopharmacology. Cal West Med.
1938;49:140–2.
Hare EH, McCance C, McCormick WO. Imipramine and “Drinamyl” in depressive illness: a
comparative trial. Br Med J. 1964;1:818.
Harley J. The old vegetable neurotics. Hemlock, opium, belladonna and henbane. Their
physiological action and therapeutical use alone and in combination. Being the Gulstonian
lectures of 1868, extended and including a complete examination of the active constituents of
opium. London: Macmillan and Co.; 1869.
Heal DJ, Smith SL, Gosden J, Nutt DJ. Amphetamine, past and present – a pharmacological and
clinical perspective. J Psychopharmacol. 2013;27:479–96.
Healy D. The creation of psychopharmacology. Cambridge, MA: Harvard University Press; 2002.
Psychopharmacology: A Brief Overview of Its History 657
Herzberg D. Blockbusters and controlled substances: Miltown, Quaalude, and consumer demand
for drugs in postwar America. Stud Hist Phil Biol Biomed Sci. 2011;42:415–26.
Hökfelt T. Looking at neurotransmitters in the microscope. Prog Neurobiol. 2010;90:101–18.
Hollister LE. The pre-benzodiazepine era. J Psychoactive Drugs. 1983;15:9–13.
Howland RH. Methylene blue: the long and winding road from stain to brain. J Psychosoc Nurs
Ment Health Serv. 2016;10(9):21–4, (10) 21–26
Huehnerfeld J. The hematoporphyrin treatment of melancholia and endogenous depression.
Am J Psychiatr. 1936;92:1323–30.
Ilieva IP, Hook CJ, Farah MJ. Prescription stimulants’ effects on healthy inhibitory control, working
memory, and episodic memory: a meta-analysis. J Cogn Neurosci. 2015;27:1069–89.
Jaspers K. Allgemeine Psychopathologie. 4th ed. Berlin/Heidelberg: Springer; 1946.
Jones PB, Barnes TRE, Davies L, Dunn G, Lloyd H, Hayhurst KP, Murray RM, Markwick A,
Lewis SW. Randomized controlled trial of the effect on quality of life of second- vs first-
generation antipsychotic drugs in schizophrenia: Cost Utility of the latest Antipsychotic Drugs
in Schizophrenia Study (CUtLASS 1). Arch Gen Psychiatry. 2006;63:1079–87.
Kalischer S. Medikamentöse Therapie. In: Lewandowsky M, editor. Handbuch der Neurologie.
Erster Band: Allgemeine Neurologie, zweiter Teil. Berlin: Julius Springer; 1910. pp. 1481–515.
Kendler KS, Schaffner KF. The dopamine hypothesis of schizophrenia: an historical and
philosophical analysis. Philos Psychiatry Psychol. 2011;18:41–63.
Kety SS. Résumé: I. Biochimie. In: de Ajuriaguerra J, editor. Monoamines et système nerveux
central. Symposium Bel-Air, Genève, Septembre 1961. Genève: Georg et Cie; 1962. pp. 263–8.
King C, Voruganti LNP. What’s in a name? The evolution of the nomenclature of antipsychotic
drugs. J Psychiatry Neurosci. 2002;27:168–75.
Kraepelin E. Über die Beeinflussung einfacher psychischer Vorgänge durch einige Arzneimittel.
experimentelle Untersuchungen. Jena: Gustav Fischer; 1892.
Kraepelin E. Klinische Psychiatrie. Psychiatrie: ein Lehrbuch für Studirende und Ärzte. Bd. 2.
Leipzig: Barth; 1899.
Kuhn R. Psychopharmakologie gestern – heute – morgen. Vortrag zur Verleihung der Hans-
Prinzhorn-Medaille in Berlin am 28. Oktober 2004. Schweiz Arch Neurol Psychiatr.
2005;156:267–9.
Lader M. History of benzodiazepine dependence. J Subst Abus Treat. 1991;8:53–9.
Laignel-Lavastine M. Des troubles psychiques par perturbations des glandes à sécrétion interne.
Congrès des aliénistes et neurologistes de France et des pays de langue française, XVIIIe
session, Dijon, août 1908. Paris: G. Masson; 1908.
Lambert PA, Revol L. Classification psychopharmacologique et clinique des différents
neuroleptiques. Indications thérapeutiques générales dans les psychoses. Presse Med.
1960;68:1509–11.
Ledermann F. Pharmacie, médicaments et psychiatrie vers 1850: le cas de Jacques-Joseph Moreau
de Tours. Rev Hist Pharm. 1988;35:67–76.
Lehmann HE. Before they called it psychopharmacology. Neuropsychophrmacology.
1993;8:291–303.
Leucht S, Cipriani A, Spineli L, Mavridis D, Örey D, Richter F, Samara M, Barbui C, Engel RR,
Geddes JR, Kissling W, Stapf MP, Lässig B, Salanti G, Davis JM. Comparative efficacy and
tolerability of 15 antipsychotic drugs in schizophrenia: a multiple-treatments meta-analysis.
Lancet. 2013;382:951–62.
Lewin L. Phantastica: die betäubenden und erregenden Genussmittel, für Ärzte und Nichtärzte.
Berlin: Georg Stilke; 1924.
Loera-Valencia R, Cedazo-Minguez A, Kenigsberg PA, Page G, Duarte AI, Giusti P, Zusso M,
Robert P, Frisoni GB, Cattaneo A, Zille M, Boltze J, Cartier N, Buee L, Johansson G, Winblad
B. Current and emerging avenues for Alzheimer’s disease drug targets. J Intern Med.
2019;86:398–437.
López-Muñoz F, Alamo C. The consolidation of neuroleptic therapy: Janssen, the discovery of
haloperidol and its introduction into clinical practice. Brain Res Bull. 2009;79:130–41.
658 P. Foley
Oppenheimer H. Medical and allied topics in Latin poetry. London: John Bale, Sons & Danielson;
1928.
Page IH. Chemistry of the brain. Springfield/Baltimore: Charles C. Thomas; 1937.
Paladin A. Recherches sur la biochimie du cerveau. Bull Soc Chim Biol. 1934;16:1193–210.
Pare CMB, Rees L, Sainsbury MJ. Differentiation of two genetically specific types of depression by
the response to anti-depressants. Lancet. 1962;2:1340–3.
Paulson GW. Historical comments on tardive dyskinesia: a neurologist’s perspective. J Clin
Psychiatry. 2005;66:260–4.
Perez-Caballero L, Torres-Sanchez S, Bravo L, Mico JA, Berrocoso E. Fluoxetine: a case history of
its discovery and preclinical development. Expert Opin Drug Discovery. 2014;9:567–78.
Pinel P. Traité medico-philosophique sur l’aliénation mentale. Paris: J. Ant. Brosson; 1809.
Ramchandani D, López-Muñoz F, Álamo C. Meprobamate: tranquilizer or anxiolytic? A historical
perspective. Psychiatry Q. 2006;77:43–53.
Rasmussen N. Making the first anti-depressant: amphetamine in American medicine, 1929–1950.
J Hist Med Allied Sci. 2006;61:288–323.
Rasmussen N. Amphetamine-type stimulants: the early history of their medical and non-medical
uses. Int Rev Neurobiol. 2015;120:9–25.
Reiter PJ. Behandlung von Dementia praecox mit Metallsalzen a. m. Walbum. I. Mangan.
Vorläufige Mitteilung. Zschr f d ges Neurol u Psychiat. 1929;108:464–80.
Ring J, Grosber M, Brockow K, Bergmann K-C. Anaphylaxis. In: Bergmann K-C, Ring J, editors.
History of allergy. Basel: Karger; 2014. pp. 54–61.
Robbins TW. Pharmacological treatment of cognitive deficits in nondementing mental health
disorders. Dialogues Clin Neurosci. 2019;21:301–8.
Robitzek EH, Selikoff IJ, Ornstein GG. Chemotherapy of human tuberculosis with hydrazine
derivatives of isonicotinic acid. Q Bull Sea View Hosp. 1952;13:27–51.
Roncero C, Álvarez FJ. The use of lisdexamfetamine dimesylate for the treatment of ADHD and
other psychiatric disorders. Expert Rev Neurother. 2014;14:849–65.
Ross S, Cole JO. Psychopharmacology. Annu Rev Psychol. 1960;11:415–38.
Rupniak NMJ, Kramer MS. NK1 receptor antagonists for depression: why a validated concept was
abandoned. J Affect Disord. 2017;223:121–5.
Schatzberg AF. New approaches to managing psychotic depression. J Clin Psychiatry. 2003;64
(Suppl. 1):19–23.
Schildkraut JJ. The catecholamine hypothesis of affective disorders: a review of supporting
evidence. Am J Psychiatr. 1965;509:509–22.
Schioldann J. ‘On periodical depressions and their pathogenesis’ by Carl Lange (1886). Hist
Psychiatry. 2011;22:108–30.
Schmitz H. Die Opiumbehandlung bei Geisteskrankheiten, insbesondere bei Melancholie, ihre
Geschichte, ihr heutiger Stand und eigene Erfahrungen. Allg Zschr Psychiat psych-gerichtl
Med. 1925;83:92–112.
Schneider PJ. Entwurf zu einer Heilmittellehre gegen psychische Krankheiten, oder Heilmittel in
Beziehung auf psychische Krankheitsformen. Tübingen: Heinrich Laupp; 1824.
Schou M. Lithium treatment at 52. J Affect Disord. 2001;67:21–32.
Seeman P. Dopamine receptors and the dopamine hypothesis of schizophrenia. Synapse.
1987;1:133–52.
Seige M. Klinische Erfahrungen mit Neuronal. Dtsch Med Wochenschr. 1912;38:1828–30.
Sertürner F. Darstellung der reinen Mohnsäure (Opiumsäure) nebst einer chemischen Untersuchung
des Opiums mit vorzüglicher Hinsicht auf einen darin neu entdeckten Stoff und die dahin
gehörigen Bemerkungen. Journal der Pharmacie für Ärzte, Apotheken und Chemisten.
1806;14:47–93.
Shen WW. A history of antipsychotic drug development. Compr Psychiatry. 1999;40:407–14.
Siegel RE. Galen on psychology, psychopathology, and function and diseases of the nervous
system. Basel: S. Karger; 1973.
Somers K. Notes on rauwolfia and ancient medical writings of India. Med Hist. 1958;2:87–91.
660 P. Foley
Sourkes TL. Early clinical neurochemistry of CNS-active drugs. Bromides. Mol Chem
Neuropathol. 1991;14:131–42.
Sourkes TL. Early clinical neurochemistry of CNS-active drugs. Chloral hydrate. Mol Chem
Neuropathol. 1992;17:21–30.
Speaker SL. From “happiness pills” to “national nightmare”: changing cultural assessment of minor
tranquilizers in America, 1955–1980. J Hist Med Allied Sci. 1997;52:338–76.
Steck H. Le syndrome extrapyramidal et diencéphalique au cours de traitements au largactil et ou
serpasil. Ann Med Psychol. 1954;112:737–43.
Sternbach LH. The benzodiazepine story. J Psychoactive Drugs. 1983;15:15–7.
Szökő É, Tábi T, Riederer P, Vécsei L, Magyar K. Pharmacological aspects of the neuroprotective
effects of irreversible MAO-B inhibitors, selegiline and rasagiline, in Parkinson’s disease.
J Neural Transm. 2018;125:1735–49.
Tarsy D. History and definition of tardive dyskinesia. Clin Neuropharmacol. 1983;6:91–9.
Thorner MW. Psycho-pharmacology of sodium amytal. J Nerv Ment Dis. 1935;81:161–7.
Thudichum JLW. A treatise on the chemical constitution of the brain. London: Ballière, Tindall &
Cox; 1884.
Toda M, Abi-Dargham A. Dopamine hypothesis of schizophrenia: making sense of it all.
Curr Psychiatry Rep. 2007;9:329–36.
Tschupp C. Johanniskraut, Hypericum perforatum L.: vom Hexenkraut zum modernen Arzneimittel
(Veröffentlichungen der Schweizerischen Gesellschaft für Geschichte der Pharmazie; 26).
Brugg: Schweizerischen Gesellschaft für Geschichte der Pharmazie; 2004.
van Enkhuizen J, Janowsky DS, Olivier B, Minassian A, Perry W, Young JW, Geyer MA. The
catecholaminergic-cholinergic balance hypothesis of bipolar disorder revisited. Eur J
Pharmacol. 2015;753:114–26.
Vanle B, Olcott W, Jimenez J, Bashmi L, Danovitch I, IsHak WW. NMDA antagonists for treating
the non-motor symptoms in Parkinson’s disease. Transl Psychiatry. 2018;8:117.
Vogt C, Vogt O. Zur Lehre der Erkrankungen des striären Systems. J Psychol Neurol. 1920;25
(Suppl 3):627–846.
Walther-Buël H. Drei Dezennien Narkotherapie. Monatsschr Psychiatr Neurol. 1953;125:718–31.
Walton M. Deep sleep therapy and Chelmsford private hospital: have we learnt anything? Australas
Psychiatry. 2013;21:206–12.
Weber MM. Die „Opiumkur” in der Psychiatrie: Ein Beitrag zur Geschichte der
Psychopharmakotherapie. Sudhoffs Arch. 1987;71:31–61.
Weber MM. Die Entwicklung der Psychopharmakologie im Zeitalter der naturwissenschaftlichen
Medizin Ideengeschichte eines psychiatrischen Therapiesystems. München: Urban und Vogel;
1999.
Weiden PJ. EPS profiles: the atypical antipsychotics are not all the same. J Psychiatr Pract.
2007;13:13–24.
Wilkins RW. Clinical usage of Rauwolfia alkaloids, including reserpine (Serpasil). Ann N Y Acad
Sci. 1954;59:36–44.
Williams M. Antianxiety agents: a historical perspective. In: Meldrum BS, Williams M, editors.
Current and future trends in anticonvulsant, anxiety, and stroke therapy: proceedings of a
symposium held at Princeton, New Jersey, May 21–23, 1989 (Progress in clinical and biological
research, 361). New York: Wiley-Liss; 1990. pp. 131–44.
Winblad B. Piracetam: a review of pharmacological properties and clinical uses. CNS Drug Rev.
2005;11:169–82.
Wolf MA, Yassa R, Llorca PM. Complications extrapyramidales induites par les neuroleptiques.
Perspectives historiques. Encephalé. 1993;19:657–61.
Wong DT, Perry KW, Bymaster FP. The discovery of fluoxetine hydrochloride (Prozac). Nat Rev
Drug Discov. 2005;4:764–74.
Wyatt RJ, Gillin JC. The transmethylation hypothesis: a quarter of a century later. Psychiatr Ann.
1976;6:33–49.
Yáñez M, Padín JF, Arranz-Tagarro JA, Camina M, Laguna R. History and therapeutic use of
MAO-A inhibitors: a historical perspective of MAO-A inhibitors as antidepressant drug. Curr
Top Med Chem. 2012;12:2275–82.
Historical Overview of Psychiatric
Diagnostics in Japan
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 662
Psychiatric Diagnostics Before DSM-III in Japan (Conventional Psychiatric Diagnostics) . . . 662
Problems of Conventional Psychiatric Diagnostics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 663
Turning Point of Psychiatric Diagnostics in Japan . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 664
Psychiatric Diagnostics After Introduction of DSM in Japan . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 664
Reaffirming the Basic Idea of DSM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 665
Domestic Opinion About DSM-5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 666
Major Depressive Disorder Versus Bereavement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 666
Neurocognitive Disorder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 667
Autism Spectrum Disorder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 667
Toward Future Development of Psychiatric Diagnostics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 667
Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 668
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 669
Abstract
When Japanese psychiatry was in its infancy, the disease concepts held by
individual physicians varied, and even the same diagnosis could sometimes
refer to a completely different group of symptoms. Based on such reliability
problems of psychiatric diagnostics, DSM-III (Diagnostic and Statistical Manual
of Mental Disorders – Third Edition) was introduced to Japan in 1982. Since then,
the reliability of psychiatric diagnostics has increased. Fields such as clinical
practice, education, and research in Japan have also benefited from DSM. How-
ever, it is also true that the misuse of DSM by inexperienced users has caused
adverse effects. The misuse of DSM is attributed mainly to the misunderstanding
that “it’s a disease simply because it’s specified in DSM diagnostic criteria.” DSM
diagnostic criteria have been created to stipulate the core of each disease. Thus, in
using those criteria, sufficient understanding of the ideal type of the disease, the
characteristics and symptoms including those not stated in DSM, is a prerequisite.
Attempting to extract a homogeneous population based on such an appropriate
use has been a challenge of DSM to date. Reaffirming this basic idea of DSM is
necessary for the future development of psychiatric diagnostics. In addition, we
must not forget that psychiatric diagnostics strive not only to provide diagnosis of
a disease but also to understand human beings from a multifaceted viewpoint,
bearing in mind each patient’s personality as well as the social and environmental
background.
Introduction
The DSM (Diagnostic and Statistical Manual of Mental Disorders) and ICD (Inter-
national Classification of Diseases) are both taught in Japanese medical schools in
addition to traditional psychiatry, and both DSM and ICD are widely used in
psychiatric care in Japan. The DSM in particular has widely permeated clinical
practice, education, and research in Japan, and virtually all clinical research and
clinical trials reference it for the definition of the target disease (Kuroki et al. 2016).
In this chapter, we summarize the changes in psychiatric diagnostics in Japan,
focusing on the DSM and clarifying the important points for moving forward with
future diagnostic development.
In Japan, the systematic study of psychiatry is said to have begun in 1886, with the
establishment of Japan’s first psychiatry course in the Medical College of the
Imperial University (currently the Faculty of Medicine, University of Tokyo).
With the landing of the United States Navy East India Fleet (in Japan, this is
commonly known as the Black Ships) commanded by Perry in 1853, there ended
215 years of national isolation (a self-imposed policy prohibiting transport and trade
between Japan and other countries), and henceforth the Japanese government rapidly
introduced modern Western technology. In particular, it was officially announced
that German medicine would be introduced into Japan. Hajime Sasaki, the first
lecturer appointed to teach psychiatry in the Medical College of the Imperial
University, had studied in Germany, and German army surgeons were also recruited
as professors at the same university. Doctors who studied psychiatry in this course
were subsequently appointed to psychiatry departments in university medical facul-
ties newly established in various regions throughout Japan. Thus, German psychiatry
strongly influenced the emerging field of psychiatry in Japan (Omata 2005).
Historical Overview of Psychiatric Diagnostics in Japan 663
the following criteria exist in the past 12 months in order to make a diagnosis.” By
doing this, presentation of symptom list is inevitably carried out, and information
variance is greatly improved at the same time. On the other hand, for observation/
interpretation variance, sufficient training for symptomatology is necessary.
many benefits to fields such as clinical practice, education, and research in Japan
(Kuroki 2012).
In the clinical practice, using DSM as a common language enabled experts in
the different areas of psychiatry to engage in discussion. Additionally, it allowed
psychiatry to utilize diagnostic inference methods such as the algorithmic
approach, the pattern recognition method, and the hypothetical-deductive method
(Kitamura 2016).
DSM also served an important purpose in education. DSM outlines the core
symptoms of each disease, allowing beginners to learn about psychiatric diseases in
a systematic way. Furthermore, it also allowed beginners to study a precise diagnos-
tic process.
DSM also contributed significantly to research. The clear disease categories,
coupled with the progress in imaging technology and molecular genetic techniques,
contributed to the acceleration of both epidemiological and biological research into
psychiatric disease.
Nonetheless, due to the usefulness of DSM and its multifarious dissemination, the
misunderstanding that “it’s a disease simply because it’s specified in DSM diagnostic
criteria” became common among users who had not acquired sufficient training in
psychiatric diagnostics. As a result, adverse effects arose in every field where DSM
had been used. For example, in the clinical practice, DSM was used as a bible, and
the abundant descriptions of diseases and symptoms for which psychopathology had
been cultivated up to that time came to be unfortunately neglected. In education, it
became common for students to simply memorize DSM diagnostic criteria without
studying the disease concepts. Additionally, this kind of flawed use made it impos-
sible to extract homogeneous populations. It is possible that the negative effects
arising from misuse of DSM were one reason why research aimed at clarifying the
pathology of psychiatric disease was not advancing.
The DSM-5, which is the most recent revision of the manual for psychiatric
diagnostics, was published in May of 2013 and is already receiving criticism for
different reasons. The most common reason pertains to concern about overdiagnosis
as a result of the current revisions. In particular, elimination of the bereavement
exclusion from major depression has frequently been the target of criticism (Fried-
man 2012). In addition to depression, there are also concerns about overdiagnosis of
conditions such as mild cognitive impairment and substance use disorders.
Let’s see what kind of opinions in Japan about some diagnoses of DSM-5.
DSM considers a condition to be major depressive disorder if the criteria for major
depressive disorder are met from the points of severity, duration, and functional
impairment, even in the case of a major life event, such as bereavement. As the
positive opinion, this would decrease the amount of missed diagnoses. On the
contrary, as the negative opinions, there is a concern regarding overdiagnosis,
resulting in unnecessary drug therapy.
It is important to note that the aim of outlining the characteristics of major
depressive episode in DSM is established on the basis of the prerequisite of “careful
consideration is given to the delineation of normal sadness and grief from a major
depressive episode.” As such, there is a note in DSM-5 regarding discrimination
Historical Overview of Psychiatric Diagnostics in Japan 667
between grief and major depressive episode. As described in the note, investigating
the characteristics of the symptoms and of the progression in detail was the
natural approach in traditional psychiatry, which determined the cause by using
the concept of understanding. In light of the basic idea of DSM, if diagnostic criteria
are applied appropriately in this kind of approach, the extraction of a uniform major
depressive disorder group can be expected, even from among patients with the
experience of loss.
Neurocognitive Disorder
From the viewpoint of detailed examination of the symptoms, the abundant expla-
nation of neurocognitive disorders is also worth noting. On the contrary, as a
disadvantage, there is a concern regarding insufficient introduction of biological
indicators to the diagnostic criteria.
For diagnosis of a neurocognitive disorder using DSM-5, detailed evaluation of
the cognitive domains defined as complex attention, executive function, learning and
memory, language, perceptual-motor function, and social cognition is required
(Sachdev et al. 2014). And the practical guidelines are provided, outlining what
kinds of episodes in daily life can be observed in case of the impairment of the
various domains, as well as concrete assessment methods to evaluate these episodes.
Similar to the note for major depressive disorder, review of details regarding the
nature of symptoms, and an attempt to extract a homogeneous neurocognitive
disorder population using those particular details, conform to the basic idea of DSM.
Conclusion
are sufficiently obtained. It is important for psychiatrists to not merely follow the
diagnostic criteria but to understand the essence of them and to use them
therapeutically.
References
American Psychiatric Association. Quick reference to the diagnostic criteria from DSM-III.
Washington, DC: American Psychiatric Association; 1980. (Translated into Japanese by
Takahashi S, Hanada K and Fujinawa A. Tokyo: Igaku Shoin; 1982.)
Cooper JE. Psychiatric diagnosis in New York and London. Maudsley monograph no. 20. London:
Oxford University Press; 1972.
Feighner JP, Robins E, Guze SB, et al. Diagnostic criteria for use in psychiatric research. Arch Gen
Psychiatry. 1972;26:57–63.
Friedman RA. Grief, depression, and the DSM-5. N Engl J Med. 2012;366:1855–7.
Kasahara Y, Kimura B. Utsu jōtai no rinshōteki bunrui ni kansuru kenkyū (Study on clinical
classification of depressive state). Seishin Shinkeigaku Zasshi. 1975;77:715–73.
Kendler KS. Phenomenology of schizophrenia and the representativeness of modern diagnostic
criteria. JAMA Psychiat. 2016;73:1082–92.
Kitamura H. DSM recommendation to refine psychiatric diagnosis. Arch Psychiatr Diagn Clin Eval.
2016;9:46–52.
Kitamura T, Shima S, Sakio E, et al. Psychiatric diagnosis in Japan. I. A study on diagnostic labels
used by practitioners. Psychopathology. 1989a;22:239–49.
Kitamura T, Shima S, Sakio E, et al. Psychiatric diagnosis in Japan. II. Reliability of conventional
diagnosis and discrepancies with RDC diagnosis. Psychopathology. 1989b;22:250–9.
Kuroki T. DSM to gendai no seishin igaku – Doko kara kite, doko e mukau no ka (DSM and modern
psychiatry – where do we come from and where are we going). In: Kanba S, Matsuhita M,
editors. Senmoni no tame no seishinka ryumiēru 30 (Psychiatric Lumière for specialists 30).
Tokyo: Nakayama Shoten; 2012. p. 123–36.
Kuroki T, Ishitobi M, Kamio Y, et al. Current viewpoints on DSM-5 in Japan. Psychiatry Clin
Neurosci. 2016;70:371–93.
Morrison J, Winokur G, Crowe R, Clancy J. The Iowa 500. The first follow-up. Arch Gen
Psychiatry. 1973;29:678–82.
Omata W. Seishin igaku no rekishi (History of psychiatry). Tokyo: Daisan Bunmeisha; 2005.
Sachdev PS, Blacker D, Blazer DG, et al. Classifying neurocognitive disorders: the DSM-5
approach. Nat Rev Neurol. 2014;10:634–42.
Spitzer RL, Endicott J, Robins E. Clinical criteria for psychiatric diagnosis and DSM-III.
Am J Psychiatry. 1975;132:1187.
Spitzer RL, Endicott J, Robins E. Research diagnostic criteria: rationale and reliability. Arch Gen
Psychiatry. 1978;35:773–82.
Neuro-psychopharmacotherapy and the
Differential Diagnostic Approaches in
Europe
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 672
Differential Diagnostic Approach to Establish the Diagnosis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 673
Neuro-psychopharmacological Recommendations According to Treatment Phases . . . . . . . . . . 674
Treatment Resistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 685
Differential Diagnostic Approach According to Predominant Symptomatology and
Subtypes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 690
Differential Diagnostic Approach According to Special Circumstances . . . . . . . . . . . . . . . . . . . . . . . 690
Summary and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 697
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 698
Abstract
At present there are no Pan-European guidelines regarding the differential diag-
nostic approach for most psychiatric diseases. Exceptions are guidelines for the
diagnosis of Tourette syndrome or Alzheimer’s disease published by European
societies. By contrast, national guidelines give recommendations on how to
approach the patient with initial or recurrent presentation of, e.g., psychotic
symptoms, which diagnostic measures to take or which kind of first-line treatment
to choose for common clinical scenarios.
In this chapter we focus on the differential diagnostic approach to the patient
with psychotic symptoms and have chosen the guidelines of the World Federation
of Societies for Biological Psychiatry for schizophrenia as a reference and
compared these to six national guidelines available in German or English. Tables
summarize commonalities and disparities on how to establish a diagnosis, of
neuro-psychopharmacotherapeutical recommendations according to different
Introduction
Psychiatric diagnoses in Europe are usually based upon the International Classification
of Diseases (ICD) in its current version, which is the tenth revision at the moment
(Akiskal and Benazzi 2005). In 2012 the European Psychiatric Association founded
the “Guidance Project,” which should fill the gap of clinically important situations and
other aspects of treatment not covered in national or international guidelines (Gaebel
and Moller 2012). The differential diagnostic approach is not yet covered by any of the
guidelines published (European Psychiatric Association 2018).
For most of the psychiatric disorders, there is no Pan-European guideline.
Exceptions are, e.g., the guidelines for the diagnosis of Tourette syndrome from
the European Society for the Study of Tourette Syndrome (ESSTS) (Cath et al. 2011)
or guidelines for diagnosis and treatment of Alzheimer’s disease from the European
Federation of Neurological Societies (EFNS) (Hort et al. 2010).
In the late 1990s, European societies promoted development of European guide-
lines for diagnosis and management of dementias and harmonization of European
approaches with those of the United States and Japan (Harvey et al. 1998).
Since then, identification of factors influencing early diagnosis of dementia
(Vernooij-Dassen et al. 2005), developing screening guidelines for early detection
(Visser et al. 2008), and computer-based educational programs to train diagnosis and
treatment in dementia (Degryse et al. 2009) were in the focus of European harmo-
nization processes.
Nowadays, European recommendations exist for the use of neuroimaging in the
diagnostic approach of dementias (Filippi et al. 2012), early diagnosis for dementia
(Brooker et al. 2014), and assessment of dementia of patients from ethnic minorities
(Nielsen et al. 2011), and currently researchers are making efforts to create European
guidelines for outcome parameters in Alzheimer’s prevention under the Horizon
2020/IMI European Prevention of Alzheimer’s Dementia (EPAD) project (Ritchie
et al. 2017).
Although in 1986 a survey of the different approaches toward the assessment of
schizophrenia was published (Berner et al. 1986) and expert recommendations from
European researchers followed focusing on treatment of schizophrenia (Altamura et al.
2000) or use of antipsychotic depot medication throughout Europe (Kane et al. 1998),
there is still no European guideline for diagnosis and treatment of schizophrenia.
Neuro-psychopharmacotherapy and the Differential Diagnostic Approaches in. . . 673
Most psychiatric diseases are diagnosed based on clinical assessment and not on any
laboratory, imaging, or other biological marker. However, ruling out any somatic causes
of the symptomatology presented by the patient is crucial, and this will be one of the first
steps to the differential diagnostic approach in any country. An exception might be the
differential diagnostic approach to establish a diagnosis of dementia, as neuroimaging
and laboratory findings are largely approaching a useful state to be incorporated into
clinical practice. The reader is kindly asked to refer to the abovementioned European
guidelines of diagnosis and treatment of Alzheimer’s disease and other dementias.
After any somatic causes have been established to be very unlikely, the next
important step is to gain a clearer picture of the psychiatric symptoms themselves
and to group these symptoms according to defined syndromes. Together with time
criteria, a psychiatric diagnosis can be made.
In Table 1 recommendations of the six national guidelines for the approach to the
patient with first psychotic symptoms and investigations to rule out organic brain
disorders are presented as well as initial setup after establishing a diagnosis.
Not all national guidelines cover recommendations for ruling out organic brain
disorders. And some go into a lot of detail, including individual lab values, whereas
others remain superficial. Not surprisingly, the most comprehensive one is the latest
published national guideline, the current version of the S3 DGPPN guideline, which
was published only a few months ago and thus incorporates all the latest data.
This guideline also provides a dedicated differential diagnostic algorithm to
establish a diagnosis of schizophrenia or other related disorders and to rule out
somatic diseases as secondary causes for psychotic symptoms. Furthermore, it lists
in detail organic diseases that may mimic schizophrenia-like symptoms such as
674 R. Musil and P. Falkai
Now we are reaching the point where first therapeutic decisions have to be made.
The diagnosis is established, a treatment plan worked out, and a therapeutic alliance
reached. The differential diagnostic approach will now separate between phases of a
disorder or according to specific symptomatology. This means that treatment will
most likely differ in a patient with a first episode of any disorder, with a recurrent
episode, or with a chronic condition.
In Table 2 treatment recommendations for patients with a first psychotic episode
are listed and followed by recommendations for a post-acute stabilization phase,
patients in remission, and relapsing cases. Although specific recommendations will
Table 1 Differential diagnostic approaches across European national guidelines for establishing a diagnosis of schizophrenia and initial approach
Danish
Austrian consensus national Clinical practice SIGN Scottish
Guideline WFSBP DGPPN S3 statement guideline NICE guidelines (CPG) guideline
Year 2012 2019 2016 2016 2014 2009 2013
Approach “Careful • Diagnosis should • Physical and Not • Comprehensive • Perform At baseline
to the diagnostic be based on neurological covered multidisciplinary psychiatric and physical
patient evaluation”: operationalized examination assessment general medical monitoring
with first laboratory criteria • Lab (incl. complete should be history, should include
psychotic investigation, • Differentiation blood count, signs of performed psychosocial assessment of
symptoms screening for from other psychotic inflammation (CRP), including history and family individual and
drug abuse disturbance or liver and kidney psychiatric psychiatric history, family history
organic and function, thyroid comorbidities, examination of the of physical
substance-induced hormones, fasting medical history mental state, illness, smoking
disorders should be glucose, HbA1c, and full physical physical assessment history, BMI,
made cholesterin, HDL- examination, including weight, waist
• Perform physical cholesterin, lifestyle factors neurological circumference,
examination (incl. triglycerides, drug incl. smoking, examination, blood pressure,
weight, size, temp., screen weight, physical necessary HbA1c, fasting
RR, HR) • ECG activity, nutrition, complementary glucose, fasting
• Lab (incl. complete • MRT psychosocial examinations to rule lipids (prolactin,
blood count, liver • Lues serology situation, out other disorders ECG as
and kidney • HIV screen developmental • Differentiate clinically
functioning, CRP, • Waist circumference history, social symptomatology indicated)
TSH, drug screen, • Neuropsychological activities, from other
fasting glucose tests occupational psychiatric
• Structural imaging • Pregnancy tests situation, quality disorders (e.g.,
MRT (T1, T2, of life and mood disorders
Neuro-psychopharmacotherapy and the Differential Diagnostic Approaches in. . .
(continued)
676
Table 1 (continued)
Danish
Austrian consensus national Clinical practice SIGN Scottish
Guideline WFSBP DGPPN S3 statement guideline NICE guidelines (CPG) guideline
Year 2012 2019 2016 2016 2014 2009 2013
• Initial assessment
should include
MRI,
neurocognitive
assessment,
neurological exam,
ECG, height,
weight, illegal drug
screen
How to Imaging • CSF in case of Optional: ANA, Not Not covered • Consider Not covered
exclude techniques suspected coeruloplasmin, CSF, covered consumption of
organic (MRI, if not secondary causes EEG, X-ray thorax, drugs or specific
brain available for psychotic serum-cortisol levels, medications (e.g.,
disorders CCT); CSF symptoms drug plasma levels, steroids)
only if organic • Neuropsychology screen for toxins, • Lab should
brain disorders testing (attention, genetic testing, include toxicology
are clinically learning, memory, prolactin screen, general
expected executive function., biochemistry,
social cognition) complete
• EEG in case of hemogram and
suspected urine analysis
neurological • Consider
disorders pregnancy test,
R. Musil and P. Falkai
• Screening for ECG, CT or MRI,
dementia in higher neuropsychological
ages assessment, and
• Lab: CK, general
antibodies, iron, psychometry
copper, vit. B1, B6,
B12, serological
screening (HIV,
Hep., Lues, etc.) in
case of suspected
secondary causes
After • Establish • At any new • At any new relapse: Not • Establish a care • Involve patient • Treatment
diagnosis therapeutic relapse, offer physical and covered plan in and family in the should be
has been alliance recommended neurological cooperation with treatment plan in an offered within
established • Formulate diagnostic steps for examination, lab primary service active collaboration the context of a
treatment plan initial presentation (complete blood providers • Develop an specialist early
• Reevaluate if not yet performed count, CRP, liver, and integrated and intervention
diagnosis and • Establish overall kidney function, coordinated model of care
treatment plan treatment plan with recheck pathological treatment plan involving
periodically participation of findings, ECG, fasting • Establish a assertive
• Involve patients and glucose, HbA1c, therapeutic alliance outreach
family and professionals in cholesterin, HDL- and stimulate approaches,
significant cooperation with cholesterin, treatment family
others with relatives and triglycerides, waist compliance involvement,
patients’ significant others circumference • Provide education access to
permission and involve • Optional: drug for patient and psychological
nonprofessional plasma levels, TDM, family interventions,
helping systems neuropsychological • Treat comorbid vocational
Neuro-psychopharmacotherapy and the Differential Diagnostic Approaches in. . .
Table 1 (continued)
Danish
Austrian consensus national Clinical practice SIGN Scottish
Guideline WFSBP DGPPN S3 statement guideline NICE guidelines (CPG) guideline
Year 2012 2019 2016 2016 2014 2009 2013
plan and offered on circumstances into antipsychotic
an individual and account medication
phase-specific
manner
(multiprofessional
and close to home)
• Facilitate access to
healthcare system
and providers
• Facilitate
empowerment
R. Musil and P. Falkai
Table 2 Differential diagnostic approaches across European national guidelines with respect to treatment phases
Austrian Danish
consensus national Clinical practice SIGN Scottish
Guideline WFSBP DGPPN S3 statement guideline NICE guidelines (CPG) guideline
Year 2012 2019 2016 2016 2014 2009 2013
Acute • Beware of • Early intervention • Focus put on Not • Oral • Recommendations • Individual
treatment in EPS risk with antipsychotics treatment of covered antipsychotics of different prescribing
first episode • Gradual • Depending on positive together with international should consider
introduction psychopathology symptoms, psychological guidelines are given benefits and
of and patient agitation, and interventions, (e.g., Canadian harms
antipsychotic preferences, anxiety Treatment offer family guidelines, PORT, • Adverse effects
treatment watchful waiting of with SGAs interventions and TMAP, APA, and should be
• Aim at some days to combined with CBT, provide RANZCP: discussed in
lowest weeks can be additional regular antipsychotic detail
possible considered at first measures assessments of medication should • Medication
effective dose episode and should • Psychiatric and treatment plan, be prescribed in a should be
• Choose be integrated into somatic and monitor noncoercive manner continued for at
FGA or SGA overall treatment comorbidities symptoms combined with least 2 weeks
according to plan should be regularly psychosocial • In case of poor
patient’s • SGAs should be identified and • Decide upon interventions incl. response to
mental and preferred treated medication Aadherence- medication
somatic considering • SGAs should be together with the promoting adherence and
condition, substance-specific preferred patients strategies. inter-current
risk of EPS, side effects (considering side considering • A 24–48- h substance misuse
other side • Antipsychotics effect profile, possible side observation period is should be
effects, and should be used on patients‘ effects. Start with recommended assessed
comorbid lowest effective preferences, lower doses and • Choice of • After 4 weeks of
conditions dosage earlier gradual titrate medication should non-response and
Neuro-psychopharmacotherapy and the Differential Diagnostic Approaches in. . .
Table 2 (continued)
Austrian Danish
consensus national Clinical practice SIGN Scottish
Guideline WFSBP DGPPN S3 statement guideline NICE guidelines (CPG) guideline
Year 2012 2019 2016 2016 2014 2009 2013
clinical symptoms • In patients with each patient medication
aimed at; first episode, • Weight and BMI should be
experiences with lower doses may should be measured, considered
prior treatments; be sufficient consultation with • Minimum
risk and benefits of • Once-daily dietitian is advisable, effective dose of
the specific agents; peroral physical activity either first- or
metabolic, motor, application is the should be second-
cardiovascular, and first choice encouraged generation
endocrine/sexual • Specific • Monitor fasting antipsychotics
adverse events; risk recommendations plasma glycemia, should be used
and benefits of for patients with lipid profiles
abstinence from first episode are • Antipsychotics
AP treatment; provided (citing should be initialized
preferences of the WFSBP and gradually after
patients and DGPPN careful explanation
gender-, age-, and guidelines) • SGAs are
comorbidity- recommended at first
specific aspects time
• Symptomatology
should be
monitored on a
regular basis using
disease-specific
rating scales
Treatment in • Antipsychotic • Psychoeducation • Encourage • In case of response • Following
post-acute treatment for should be offered patients to write to antipsychotic remission of the
R. Musil and P. Falkai
stabilization relapse prevention • Cognitive an account of their treatment, first episode,
phase should be offered remediation illness after each maintenance of duration of
in a continuous should be initiated acute episode 12 months and maintenance
form • Family members • Inform patients gradual decrease treatment with
should be about risk of thereafter is antipsychotics
involved relapse after recommended should be at least
cessation of • CBT, intensive 18 months
medication within case management, • All service users
first 1–2 years family, and should be offered
after acute episode community support a mental health
should be provided review at regular
• It is recommended intervals
that care be given in
least restrictive and
coercive settings
possible
Treatment • In case of stability • Sociotherapeutic • Offer • In case of non- • Maintenance
for patients and reasons against interventions opportunity to response to initial treatment with an
in remission continuing AP stand at the point return to primary treatment offer antipsychotic
treatment, stepwise of interest care facilities after switching to medication
reduction should be stabilization of (another) SGA should be offered
offered and symptoms • Provide (usually the one
intermittent psychosocial used during the
treatment with interventions (early last acute episode
early intervention intervention in case of efficacy
in case of programs, supported and tolerability)
prodromes of employment, for a minimum of
threatened relapse cognitive 2 years; for
Neuro-psychopharmacotherapy and the Differential Diagnostic Approaches in. . .
Table 2 (continued)
Austrian Danish
consensus national Clinical practice SIGN Scottish
Guideline WFSBP DGPPN S3 statement guideline NICE guidelines (CPG) guideline
Year 2012 2019 2016 2016 2014 2009 2013
all treatment phases biopsychosocial risperidone
if discontinuation care) should be
of treatment is considered (low
wished by patients to moderate
• Regular dosing of
monitoring of 300–400 mg
symptoms for any chlorpromazine
relapse being part equivalents)
of an overall
treatment plan
should be offered
for at least 2 years
Treatment FGAs and • Antipsychotic • Last effective Not • Offer crisis • Immediately • In case of acute
for relapsing SGAs can be treatment for dose should be covered resolution and initiate exacerbation,
patients chosen relapse prevention aimed for home treatment pharmacological amisulpride,
according to should be offered teams as a first- treatment (inform olanzapine, or
patient’s • Antipsychotics line service and consult patient risperidone
previous previously leading • Crisis houses and family members should be offered
experience, to remission should and acute day on different agents • Chlorpromazine
treatment be offered for facilities should taking into account and other low-
effects, side relapse prevention be considered previous potency FGAs are
effects, unless tolerability • Consider the experiences) considered
patient’s was low impact of hospital • Consider SGAs in alternatives
preference, • Choice of interventions case of non-response • Previous
comorbid medication for • Offer oral to FGAs response to
disorders, and relapse prevention antipsychotics • Promote CBT individual
R. Musil and P. Falkai
potential should take together with • After remission is antipsychotic
interactions preferences, psychological achieved, continue medication and
with other previous interventions for 12 months, with relative adverse
medication experiences, and taking into gradual decrease of effect profiles
different adverse account several weeks should be
event profile recommendations thereafter considered
(tardive for starting • Medication
dyskinesias, treatment should be
sedation, cardial, continued for at
metabolic, least 4 weeks
endocrine, and • In partial
other effects) into remission
account medication should
be reassessed after
8 weeks (unless
significant adverse
effects)
• Patients, relatives,
and significant
others should be
informed about
higher relapse risks
with discontinuing
antipsychotic
treatment
• Duration of
treatment has to
take several aspects
Neuro-psychopharmacotherapy and the Differential Diagnostic Approaches in. . .
into account
• Depot medication
should be offered
(continued)
683
684
Table 2 (continued)
Austrian Danish
consensus national Clinical practice SIGN Scottish
Guideline WFSBP DGPPN S3 statement guideline NICE guidelines (CPG) guideline
Year 2012 2019 2016 2016 2014 2009 2013
due to secure
application and
high bioavailability
• Choice of depot
medication should
be based on side
effects profile and
intended treatment
interval
R. Musil and P. Falkai
Neuro-psychopharmacotherapy and the Differential Diagnostic Approaches in. . . 685
differ, this approach is similar in patients with other severe mental illnesses like
major depressive disorder or bipolar disorder.
As can be depicted, first choice of medication is in most cases a second-
generation antipsychotic given at the lowest effective dose and slowly up-titrated
if needed. Patients should be involved in a coercive manner, and choice of the
specific drug is based more on adverse event profiles than effects on any specific
symptomatology. Some guidelines highlight the possibility of watchful waiting
for some hours to a few weeks, and most give specific recommendations for
monitoring.
Differences are greater in the post-acute stabilization phase. Some guidelines
recommend continuous antipsychotic treatment for a specific time frame, while
others focus on additional interventions such as psychoeducation or cognitive
behavioral therapy. The same applies to the approach toward patients in remission.
While the latest S3 DGPPN guidelines get very specific about recommendations for
the process of discontinuing medication, other guidelines shift the focus on non-
medication strategies.
In the case of relapse, most guidelines recommend reinitiation of antipsychotic
treatment, some recommend specific compounds, while most stress the fact that
selection should be based on previous experiences and adverse event profiles. The
British guidelines put the focus on different kinds of early intervention programs.
Treatment Resistance
Austrian
consensus Danish national Clinical practice SIGN Scottish
Guideline WFSBP DGPPN S3 statement guideline NICE guidelines (CPG) guideline
Year 2012 2019 2016 2016 2014 2009 2013
Treatment Def.: No significant Def.: Moderate • Clozapine is Def.: No In case of In case of no Failure to
resistance improvement with severity and < 20% recommended response to at inadequate response, low respond to an
two different symptom • Reevaluate least two response: adherence or adequate trial of
antipsychotics from improvement on diagnosis, different SGAs reconsider persistent suicide two different
two different standardized comorbidities and clozapine diagnosis risk, use of antipsychotics
chemical classes at symptom scale (substance • Maintaining • Revise clozapine is (including a
recommended (PANSS, BPRS, abuse), treatment with adherence to recommended SGA)
doses for at least SANS, or SAPS) compliance, and antipsychotics is medication, dose, • If there is no • Clozapine
2–8 weeks per drug during 6-week metabolism recommended for and duration of response to first should be offered
• Multidimensional treatment periods non-remitted antipsychotic choice of • Diagnosis and
evaluation of • Overall treatment patients with treatment antipsychotic, comorbidity
persisting positive duration was previous • Review consider switch should be
or negative 12 weeks incl. at least response to engagement with to (another) reviewed
symptoms, 6-week treatment antipsychotic psychological SGA; evaluate • Clozapine
cognitive periods with two treatment treatments, offer adherence; plasma levels
dysfunctions, different • In patients with interventions that switch to should be
bizarre behavior, antipsychotics no previous have not been clozapine in case monitored in
recurrent affective • Mean dosage was response to performed (CBT of non-response certain cases
symptoms, deficits 600 mg several or family to two adequate -Augmentation
in vocational and chlorpromazine antipsychotic intervention) trials of with a second
social functioning, equivalents, and at compounds • Consider any antipsychotic SGA (or
and a poor quality least 80% of dosage including comorbid treatment (at lamotrigine)
of life has been taken clozapine, a disorders or least one SGA) should be
• Ensure adherence, • Pseudotherapeutic gradual taper of substance use as considered with
consider clozapine resistance should be any ongoing causes for non- inadequate
as first choice, ruled out before antipsychotic response
R. Musil and P. Falkai
consider other treatment resistance treatment with • Offer clozapine response to
SGAs, is diagnosed: close monitoring in case of non- clozapine alone
augmentation consider adherence, for any relapse is response to a
strategies, use of illegal recommended to least two
combination of substances, severe reduce adverse adequate trials of
drugs, adverse events, events different drugs
electroconvulsive comorbidities, (at least one other
therapy effective dosage SGA)
• Consider (incl. plasma levels), • Consider
psychoeducation to and environmental measuring
ensure treatment factors (stress, high therapeutic drug
adherence expressed emotions) levels and wait
• In case of for an adequate
established treatment time in case of
resistance, clozapine any
should be offered augmentation
after risk benefit- strategy
evaluation, consent,
and accompanying
screening tests
• If clozapine is not
tolerated, olanzapine
and risperidone
should be considered
• Dose escalation
beyond
recommended use
should be avoided
Neuro-psychopharmacotherapy and the Differential Diagnostic Approaches in. . .
• Monotherapy
should be preferred,
and combination of
two substances only
considered after
687
(continued)
688
Table 3 (continued)
Austrian
consensus Danish national Clinical practice SIGN Scottish
Guideline WFSBP DGPPN S3 statement guideline NICE guidelines (CPG) guideline
Year 2012 2019 2016 2016 2014 2009 2013
monotherapy with
three substances incl.
clozapine was not
successful
• Augmentation with
carbamazepine,
lamotrigine, or
valproic acid to
improve symptoms
should not be offered
on a routine basis
• ECT can be
considered in case of
pharmacotherapeutic
non-response
• rTMS should be
offered in case of
persisting acoustic
R. Musil and P. Falkai
hallucinations or
negative symptoms
• Treatment should be
continued for at least
10 weeks
• High-dose
antipsychotics should
only be considered
after adequate trials
of antipsychotic
monotherapy, and
augmentation,
including a trial of
clozapine, has failed
• ECT should only be
considered in those
individuals for whom
other approaches
have failed
Neuro-psychopharmacotherapy and the Differential Diagnostic Approaches in. . .
689
690 R. Musil and P. Falkai
Table 4 (continued)
Austrian
consensus Danish national Clinical practice SIGN Scottish
Guideline WFSBP DGPPN S3 statement guideline NICE guidelines (CPG) guideline
Year 2012 2019 2016 2016 2014 2009 2013
• Optimize
pharmacological
treatment and
consider
inpatient
treatment if
needed
Differential Consider ECT in • Lorazepam can Regular Not covered Follow national Consider • Consider
treatment patients with be used in a evaluation for treatment comorbidity with sedative
according to catatonic time-restricted depressive and guidelines in other mental antipsychotic
subtypes features (when fashion (in manic features is case of suspected disorders (e.g., medication in
an initial combination recommended affective depression, individuals in
lorazepam trial with psychosis or OCD, remission with
fails) antipsychotics bipolar disorder dissociative comorbid
with low risk for symptoms, anxiety
malignant anxiety symptoms
neuroleptic symptoms, etc.) • Treatment in
syndrome) anxiety state
• ECT is the first- should be
choice treatment offered in
in case of accordance with
malignant CPG for anxiety
catatonia and panic
disorders
R. Musil and P. Falkai
• Patients with
criteria of
depressive
disorder should
be treated in
accordance with
CPG for
depression incl.
use of
antidepressant
medication
• SGA treatment
should be
considered in
case of comorbid
depressive
symptoms
Differential • Differentiate • In patients with Standard • In patients • In stable phases In case of
treatment primary and predominant instruments for with persisting consider persistent
according to secondary negative evaluation of negative treatment for negative
predominant negative symptoms, use psychotic symptoms, depressive or symptoms
symptomatology symptoms amisulpride or symptoms, SSRIs/SNRIs OCD symptoms despite
• Consider SGAs olanzapine and clinical should only be • SGAs are adherence to
• Consider avoid strong D2 symptomatology, used cautiously recommended antipsychotic
treatment of blockade and cognitive • CBT should be for treatment of medication,
cognitive deficits • In case of non- functions are considered in depressive augmentation
• Consider response to listed without a patients with symptoms with an
administration of antipsychotic clear persisting • Use antidepressant,
Neuro-psychopharmacotherapy and the Differential Diagnostic Approaches in. . .
(continued)
694
Table 4 (continued)
Austrian
consensus Danish national Clinical practice SIGN Scottish
Guideline WFSBP DGPPN S3 statement guideline NICE guidelines (CPG) guideline
Year 2012 2019 2016 2016 2014 2009 2013
should be • In case of
favored) persisting
negative
symptoms, high-
frequency rTMS
can be offered
• Regularly
check for
depressive
symptoms;
optimize
antipsychotic
medication in
case of
depressive
symptoms, offer
CBT, consider
use of
antidepressants
R. Musil and P. Falkai
Table 5 Differential diagnostic approaches for other factors in schizophrenic patients across European national guidelines
Austrian
consensus Danish national Clinical practice SIGN Scottish
Guideline WFSBP DGPPN S3 statement guideline NICE guidelines CPG guideline
Year 2012 2019 2016 2016 2014 2009 2013
Differential • Check drug • Offer • Antipsychotics • Consider use of • Consider • Maintenance
treatment in concentrations therapeutic drug in fluid or fast long-acting intensive case treatment with depot
patients with monitoring in dissolvable injectable management for antipsychotics should
poor case of assumed forms should be antipsychotics patients likely to be offered
adherence poor adherence preferred disengage from • Service users should
• Causes for non- treatment or be given the option of
compliance are services oral or depot
given in table • Consider depot/ medication in line
form long-acting with their preference
injectable
medication if
avoiding
nonadherence is a
clinical priority
Differential • SGAs • Cognitive • Acetylcholinesterase
treatment in should be remediation inhibitors may be
patients with favored should be offered considered as
cognitive in combination adjunctive therapies to
impairment with other antipsychotic
psychosocial and medication
rehabilitative • Cognitive
measurements remediation therapy
Neuro-psychopharmacotherapy and the Differential Diagnostic Approaches in. . .
Austrian
consensus Danish national Clinical practice SIGN Scottish
Guideline WFSBP DGPPN S3 statement guideline NICE guidelines CPG guideline
Year 2012 2019 2016 2016 2014 2009 2013
functions are
discussed
Differential • Explore use of • In patients with • Discuss the use of Patients should • Comorbid substance
treatment in addictive cannabis and/or alcohol, tobacco, receive specific misuse should not
patients with substances central stimulant and illicit drugs and treatment for exclude people with
comorbid • Use an abuse, CBT and their possible alcohol or drug schizophrenia from
substance integrative motivational effects abuse (harm services or
abuse approach in interviewing reduction, interventions
patients with should be abstinence, relapse • Management may
comorbid considered prevention, require a joint
substance abuse rehabilitation) consultive approach
• Use
antipsychotics
with low
anticholinergic
and EPMS
adverse events
risk
• Offer substance-
specific
psychotherapy
and psychosocial
interventions
• Consider
specific
guidelines for
treatment of
R. Musil and P. Falkai
comorbid
substance abuse
Neuro-psychopharmacotherapy and the Differential Diagnostic Approaches in. . . 697
Again, the national guidelines show differences in their approach toward these
situations. Though cognitive impairment is known as one of the most important
factors hampering integration of affected patients into work and should be addressed
specifically, not all guidelines give specific recommendations for treatment. And the
diagnostic approach for these topics is mostly not explained in greater detail, and
some guidelines refer to other guidelines (like guidelines for substance abuse).
In the new version of the S3 guideline, common comorbidities of schizophrenia
are listed, namely, substance abuse or dependency, depression and suicidality,
obsessive-compulsive disorder, posttraumatic stress disorder, anxiety disorders,
agitation, and sleeping disorders. As such, screening for pertinent symptoms consti-
tutes an important aspect in the differential diagnostic approach to schizophrenia.
References
Akiskal HS, Benazzi F. Atypical depression: a variant of bipolar II or a bridge between unipolar and
bipolar II? J Affect Disord. 2005;84(2–3):209–17.
Altamura AC, Bobes J, Owens DC, Gerlach J, Hellewell JS, Kasper S, et al. Principles of practice
from the European expert panel on the contemporary treatment of schizophrenia. Int J Psychi-
atry Clin Pract. 2000;4(1):1–11.
Neuro-psychopharmacotherapy and the Differential Diagnostic Approaches in. . . 699
Schultze-Lutter F, Michel C, Schmidt SJ, Schimmelmann BG, Maric NP, Salokangas RK, et al. EPA
guidance on the early detection of clinical high risk states of psychoses. Eur Psychiatry. 2015;30
(3):405–16.
Scottish Intercollegiate Guidelines Network (SIGN). Management of schizophrenia. SIGN publi-
cation no. 131. Edinburgh; 2013. http://www.sign.ac.uk
Vernooij-Dassen MJ, Moniz-Cook ED, Woods RT, De Lepeleire J, Leuschner A, Zanetti O, et al.
Factors affecting timely recognition and diagnosis of dementia across Europe: from awareness
to stigma. Int J Geriatr Psychiatry. 2005;20(4):377–86.
Visser PJ, Verhey FR, Boada M, Bullock R, De Deyn PP, Frisoni GB, et al. Development of
screening guidelines and clinical criteria for predementia Alzheimer’s disease. The DESCRIPA
study. Neuroepidemiology. 2008;30(4):254–65.
Neuropsychopharmacotherapy
and the Differential Diagnostic Approaches
in China
Contents
Mental Health in China . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 702
Brief History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 702
Prevalence of Mental Illness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 703
Lack of Qualified Staff . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 703
Diagnostic Classification CCMD System in China . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 704
Brief History and Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 704
CCMD-3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 705
Diagnosis of Mental Disorders . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 706
Psychiatric Diagnosis Principles and Ideas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 707
Standardized and Diagnostic Mental Examination Tool . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 710
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 711
Abstract
Mental health in China is a growing issue and has emerged as an important
specialty. Over the past four decades, China’s economic reforms have achieved
great success. However, rapid urbanization and economic growth are generating
new challenges for the country and its mental health system. The mental health
care system in China has undergone substantial changes during the last 40 years.
Most international experts have assessed China’s mental health system and
diagnostic approaches according to western standards. It is necessary to introduce
J. Li (*)
Institute of Mental Health, Tianjin Anding Hospital, Tianjin, China
e-mail: jieli@tjmhc.com
S. Li
Department of Psychiatry, Tianjin Medical University, Tianjin, China
e-mail: lishen@tmu.edu.cn
Mental health in China is a growing issue, with experts estimating that 173 million
people live with a mental disorder (Xiang et al. 2012). Social stigma related to
religious and cultural beliefs of upholding social harmony and maintaining personal
reputation contribute to a lack of desire to seek treatment. While the Chinese
government has committed to expanding mental health care services and legislation,
the country struggles with a lack of mental health professionals and access to
specialists in rural areas.
Brief History
China’s first mental institutions were introduced before 1849 by western missionar-
ies. Missionary and doctor John Kerr opened the first psychiatric hospital in 1898,
with the goal of providing care to people with mental illness and treating them in a
more humane way (Blum and Fee 2008).
In 1949, the country began developing its mental health resources by building
psychiatric hospitals and facilities for training mental health professionals. However,
many community programs were discontinued during the Cultural Revolution (Liu
et al. 2011).
In a 1999 meeting jointly held by Chinese ministries and the World Health
Organization, the Chinese government committed to creating a mental health action
plan and a national mental health law, among other measures, to expand and improve
care (Liu et al. 2011). The action plan, adopted in 2002, outlined China’s priorities of
enacting legislation, educating its people on mental illness and mental health
resources, and developing a stable and comprehensive system of care.
In 2000, the Minority Health Disparities Research and Education Act was
enacted. This act helped raised national awareness to health issues through research,
health education, and data collection.
Though China continues to develop its mental health services, it continues to
have a large number of untreated people with mental illness. Intense stigma
associated with mental illness, a lack of mental health professionals and special-
ists, and culturally specific expressions of mental illness may play a role in the
disparity.
Neuropsychopharmacotherapy and the Differential Diagnostic Approaches. . . 703
In the past 30 years, China – with almost a fifth of the world’s population – has
undergone unprecedented economic development and social change. This transfor-
mation has led to tremendous changes in population structure, urbanization, migra-
tion, education, transportation, culture, leisure, social concepts, and disease
epidemiology. Researchers estimate that roughly 173 million people out of 1.4
billion people in China have a mental disorder (Xiang et al. 2012). Over 90% of
people with a mental disorder have never been treated. A lack of government data on
mental illness makes it difficult to estimate the prevalence of specific mental
disorders (Xiang et al. 2012).
Conducted between 2001 and 2005, a nongovernmental survey of 63,000 Chi-
nese adults found that 16% of the population had a mood disorder, including 6% of
people with major depressive disorder (Phillips et al. 2009). Thirteen percent of the
population had an anxiety disorder, and 9% had an alcohol use disorder. Women
were more likely to have a mood or anxiety disorder compared to men, but men were
significantly more likely to have an alcohol use disorder. People living in rural areas
were more likely to have major depressive disorder or alcohol dependence.
The suicide rate in China was approximately 23 per 100,000 between 1995 and
1999 (Xiang et al. 2012).Since then, the rate is thought to have fallen to roughly 7 per
100,000, according to government data. WHO states that the rate of suicide is
thought to be three to four times higher in rural areas than in urban areas.
Till now, it took more than 10 years for Chinese psychiatrists, related profes-
sionals, and the Chinese Government to set up the China Mental Health Survey
(CMHS) . Launched in 2012, the CMHS addressed the limitations of the previous
studies by using a consistent methodology to do a nationally representative survey,
which relied on coherent diagnostic nomenclature, fully structured diagnostic inter-
views, and sophisticated household survey technology (Huang et al. 2016; Liu et al.
2016).
The latest results were published in the Lancet Psychiatry. Huang et al. found the
prevalence of most mental disorders in China in 2013 is higher than in 1982 (point
prevalence 1.1% and lifetime prevalence 1.3%), 1993 (point prevalence 1.1% and
lifetime prevalence 1.4%), and 2002 (12-month prevalence 7.0% and lifetime
prevalence 13.2%), but lower than in 2009 (1-month prevalence 17.5%). The
evidence from this survey poses serious challenges related to the high burdens of
disease identified, but also offers valuable opportunities for policy makers and
healthcare professionals to explore and address the factors that affect mental health
in China (Huang et al. 2019).
China has 17,000 certified psychiatrists, which is 10% of that of other developed
countries per capita. China averages one psychologist for every 83,000 people, and
some of these psychologists are not board-licensed or certified to diagnose illness.
704 J. Li and S. Li
Individuals without any academic background in mental health can obtain a license
to counsel following several month of training through the National Exam for
Psychological Counselors (Huang 2015), even if most of them study psychology
for personal use and do not intend to pursue a career in counseling (Hizi 2017).
Patients leave the clinics with false diagnoses and often do not return for follow-up
treatments, which is detrimental to the degenerative nature of many psychiatric
disorders (Chen 2016).
The disparity between psychiatric services available between rural and urban
areas (Chen 2016) partially contributes to this statistic, as rural areas have tradition-
ally relied on barefoot doctors since the 1970s for medical advice. These doctors, one
of the few modes of healthcare able to reach isolated parts of rural China, are unable
to obtain modern medical equipment and, therefore, reliable diagnoses for psychi-
atric illness. Furthermore, the nearest psychiatric clinic may be hundreds of kilome-
ters away, and families may be unable to afford professional psychiatric treatment for
the afflicted.
Before the foundation of the People’s Republic of China in 1949, in China, there
wasn’t its own classification system for mental illness. In June 1958, the Ministry of
Health held the first national conference on the prevention and treatment of mental
illness in Nanjing, referring to the etiological classification of the Soviet Union;
mental illness was classified into 14 categories, including (1) infectious psychosis;
(2) toxic psychosis; (3) psychosis induced by physical diseases; (4) traumatic psy-
chosis; (5) psychosis induced by brain tumors; (6) cerebrovascular psychosis;
(7) mental illness in the early stage of old age or in old age; (8) epileptic psychosis;
(9) schizophrenia; (10) manic and depressive psychosis; (11) psychogenic psycho-
sis; (12) paranoia; (13) pathological personality; and (14) mental retardation. Among
these 14 categories, mental disorders due to brain organic diseases, physical dis-
eases, and mental retardation account for the majority, while the proportion of
“functional” mental disorders is less.
In July 1978, the second academic annual conference of the Chinese Society of
Neuropsychiatry was held in Nanjing, and a special group was set up to revise the
draft of classification (in 1958). In 1979, the Chinese Journal of Neuropsychiatry
published the revised Classification of Mental Diseases (trial draft), which classified
mental diseases into ten categories: (1) brain organic psychosis, (2) psychosis
induced by physical diseases, (3) schizophrenia, (4) affective psychosis, (5) reactive
psychosis, (6) other psychosis, (7) neuroses, (8) abnormal personality, (9) mental
retardation, and (10) childhood psychosis.
At the academic conference on schizophrenia held in Suzhou in 1981, the
diagnostic criteria for clinical work of schizophrenia in our country were discussed
and formulated, with more than 3 months as the course criterion. In the criteria of
Neuropsychopharmacotherapy and the Differential Diagnostic Approaches. . . 705
CCMD-3
paying attention to international integration; and being concise and easy to operate.
The classification was mainly close to ICD-10, considering both etiological classi-
fication and symptomatology classification. The classification and ranking of mental
disorders in CCMD-3 were subject to hierarchical diagnosis and the classification
principle of ICD-10. CCMD-3 emphasizes the traditional, scientific, comprehensi-
ble, acceptable, operable, and relatively stable feature of classification diagnosis.
Large classes and small classes maintained a master-slave logical relationship. Some
categories considered being necessary by Chinese scholars have been retained or
added, including reservations of neuroses (but separating hysteria from neuroses),
recurrent mania, homosexuality, etc.; “Mental disorders closely related to culture,
i.e., mental disorders caused by Qigong, superstition and witchcraft, shrink phobia,
other or cultural related mental disorders to be classified” has been retained. In
China, the concept of simple subtype of schizophrenia was retained, and the
concepts of remission period, residual period, and decline period were still used in
the course of disease, which was different from the classification of remission type
and residual type of ICD-10. A more detailed classification of the mental disorders
caused by other encephalopathy in organic mental disorders reflects the importance
of the concept of etiological diagnosis. According to China's social and cultural
characteristics and traditional classification, some mental disorders were not
included in CCMD-3, such as in the ICD-10 (F 52.7 hypersexuality, F 60.31
borderline personality disorder, F 64.2 childhood gender identity disorder, F 66
some subtypes of psychological and behavioral disorders related to sexual develop-
ment and sexual orientation, F 68.0 rendering somatic symptoms for psychological
reasons, and F 93.3 sibling competition disorder).
Since 2002, China has officially used ICD-10 system for disease classification
statistics, and all kinds of disease codes reported to health administrative depart-
ments by psychiatric medical institutions have been required to use ICD-10 code. At
present, ICD and DSM classification systems have been basically adopted in clinical
and scientific research work.
In short, as mentioned earlier, the existing classification system, whether ICD,
DSM or CCMD, has many problems and is far from a strictly scientific classification.
Like other medical disciplines, the diagnostic classification system of mental disor-
ders also needs to be revised with the progress of medical research and clinical
practice.
Clinical medicine (including internal medicine and surgery) and clinical psychiatry
belong to the same category of natural sciences. The common goal they explore is
for human health and development. Clinical medicine belongs to the practical
science system, and practical experience is of paramount importance. A qualified
clinician must have professional theoretical knowledge and must be a practicing
doctor with certain clinical experience and certain clinical skills and methods. Most
of the exact causes of mental disorders are still unknown, so far it has not been found
Neuropsychopharmacotherapy and the Differential Diagnostic Approaches. . . 707
that can help to clarify the diagnostic well and objective biological indicators.
Clinical psychiatrists also lack tools for assisted diagnosis such as physical diagno-
sis, chemical diagnosis, and imaging diagnosis as physicians and surgeons. Clinical
psychiatrists need to practice hard in their daily work to master the scientifically
correct and reliable skills and methods that are consistent with objective reality and
scientific thinking methods and then can make a correct diagnosis.
(1) the collection of medical history is not detailed and reliable; (2) the manifestation
of the disease is not sufficient; (3) the observation of the condition is not objective
enough, and the identification of symptoms is not correct; (4) the diagnostic criteria
used are not perfect or cannot be used correctly; (5) the process of diagnostic
thinking is unscientific, for example, adopting a fixed and exclusive way of thinking
about the initial diagnosis hypothesis, which makes oneself fall into the subjective
prejudice of “preemption”; and (6) the level of scientific development is limited, and
some diseases cannot be recognized well.
At present, the causes of mental disorders are complex and symptoms are diverse
and often rely on the diagnosis of symptom groups. However, there are often
overlaps between mild mental symptoms and normal mental activities, so the
diagnosis of some diseases is characterized by varying degrees of rigidity and
yardstick. Based on this fact, some scholars have put forward the characteristic
“diagnostic symptoms” for a disease, such as the “4A” fundamental symptoms
proposed by Eugen Bleuler and the “first rank symptoms” proposed by Kurt
Schneider when schizophrenia is diagnosed, although this view, once approved by
more experts, has not been widely used in clinical practice and is increasingly being
“challenged.” In view of this, WHO and APA of the United States have successively
formulated a unified diagnostic standard for each mental disorder in accordance with
the definition of disease and have continuously supplemented and revised it
according to the development of the discipline, becoming an internationally accepted
ICD and DSM diagnostic system.
Basis of Disease
It included personal information, family disease history, pre-disease personality, past
disease history, and so on. These related factors often affect the clinical manifesta-
tion, progression, etiology, or incentive of the disease. The main points should be
noted:
Neuropsychopharmacotherapy and the Differential Diagnostic Approaches. . . 709
1. In terms of the occupation of the patient, attention should be paid to the patients’
exposure to harmful substances, the history of pesticide exposure by farmers, the
history of chemical exposure of workers, and so on.
2. Attention should be paid to the existence and progression of acute and chronic
somatic diseases in the history of past diseases, the relationship between physical
diseases and mental disorders, the characteristics of the course of disease, the
treatment situation, and the current curative effect. Sometimes physical symp-
toms before the onset of mental disorders, such as fever, oral herpes, and upper
respiratory tract infections may be prodromal symptoms of sporadic encephalitis.
3. Should pay attention to the influence of pre-illness personality, family, and school
education on the formation and development of patients’ personality. The healthy
personality or some bias often has a certain relationship with a certain disease.
4. The history of mental illness, epilepsy, mental retardation, and personality abnor-
mality in the family members can be used as a reference for the diagnosis and
analysis of mental disorders.
Clinical Manifestation
According to SSD thinking method, we should first determine the mental symptoms,
then determine the syndrome according to the combination of symptoms, and
compare each symptom or syndrome with similar phenomena to find out the
relationship among its nature, psychological background, and environment. Through
extensive and meticulous analysis, propose reasoning and then make it a diagnostic
basis. As such disturbance of consciousness or dementia (including the
corresponding syndrome) often suggests mental disorder caused by brain organic
disorder or somatic disease. Generally, a symptom or syndrome can be seen in a
variety of mental disorders. For example, the encephalasthenia syndrome may be an
early symptom of schizophrenia, a preexisting manifestation of cerebral arterioscle-
rosis, or just a neurosis. In order to understand the real connotation and essence of
the encephalasthenia syndrome through the external appearance, it is necessary to
repeatedly analyze the principal and secondary relationship from the clinical practice
and to identify according to the other characteristic manifestations of different
diseases.
710 J. Li and S. Li
The World Health Organization has studied the reliability and consistency of mental
disorders in different social and cultural contexts. The study showed that there is
difference for diagnosis among doctors. Analyze the cause of the difference: the
source of the data collected is different; terminology that doctor used and under-
standing the terminology are different; and the methods of talking examinations,
disease classification, and diagnostic criteria are different. In order to improve the
diagnosis and reliability of the disease, while developing diagnostic criteria, foreign
psychiatrists have also developed standardized mental examination tools and com-
puter diagnostic systems for clinical diagnosis and research. This tool is designed by
a psychiatrist with clinical experience based on diagnostic criteria and diagnostic
criteria. This tool consists of a series of entries, each of which represents a symptom
or clinical variable, a prescribed examination procedure, a questioning method and a
scoring standard, with an explanation of the terms of the tool. This is a fixed or semi-
timed interview tool. The doctor or the researcher conducts inquiries and examina-
tions in strict accordance with the regulations. The results are scored according to the
definition of the entry to determine whether the symptoms exist and determine the
severity. Different doctors use this standardized inspection tool to check patients,
which greatly improves the consistency of diagnosis. Currently used diagnostic
mental examination tools are Composite International Diagnostic Interview-Core
Version (CIDI) and Structured Clinical Interview for DSM-IV (SCID), which is
fixed clinical examination. The former can separately infer the diagnosis of ICD-10
Neuropsychopharmacotherapy and the Differential Diagnostic Approaches. . . 711
and DSM-IV. The former can be operated by a non-psychiatrist, while the latter must
be operated by a trained psychiatrist.
In recent years, in some epidemiological studies, a more simplified diagnostic
tool has been adopted, namely, mini-international neuropsychiatric interview
(MINI). MINI is a simple, effective, and reliable structured interview tool developed
by Sheehan and Lecrubier. It is mainly used for screening and diagnosing 16 kinds of
axis mental illness and personality disorder in DSM-IV and ICD-10, including
130 questions. Like Structured Clinical Interview for DSM-IV-Patients (SCID-P)
and CIDI, each diagnosis in the MINI is a question group, and most diagnoses have
screening problems that exclude diagnosis. There have been studies comparing the
reliability and validity of MINI with SCID-P and CIDI, and the results show that
MINI has a very acceptable reliability and validity score. The MINI has been
translated into a variety of languages and is widely used in clinical trials and clinical
practice. In recent years, China has been participating more and more in international
clinical research. The English version 5.0.0 (2004) of MINI has been translated into
Chinese version, and the reliability and validity test has been carried out. The results
show that the MINI Chinese version has a high consistency in the diagnosis of
depressive episodes, anxiety disorders, substance dependence, psychotic disorders
and the diagnosis made by SCID-P. The use of MINI ensures the accuracy and
consistency of the diagnostic process and can reveal potential psychiatric
comorbidities. Due to the short interview process, the problem is concise and easy
to be accepted by patients, and it can be widely used in clinical practice.
References
Blum N, Fee E. The first mental hospital in China. Am J Public Health. 2008;98:1593.
Chen M. Abuse of mentally ill in China. Mental Health in China. Weekly; 2016.
Hizi G. “Developmental” therapy for a “modernised” society: the sociopolitical meanings of
psychology in urban China. China. 2017;15:98–119.
Huang H-Y. From psychotherapy to psycho-boom: a historical overview of psychotherapy in China.
Psychoanal Psychother China. 2015;1(1):1–30.
Huang Y, Liu Z, Wang H, Guan X, Chen H, Ma C, Li Q, Yan J, Yu Y, Kou C, Xu X, Lu J, Wang Z,
Liu L, Xu Y, He Y, Li T, Guo W, Tian H, Xu G, Xu X, Lv S, Wang L, Wang L, Yan Y, Wang B,
Xiao S, Zhou L, Li L, Tan L. The China mental health survey (CMHS): I. Background, aims and
measures. Soc Psychiatry Psychiatr Epidemiol. 2016;51:1559–69.
Huang Y, Wang Y, Wang H, Liu Z, Yu X, Yan J, Yu Y, Kou C, Xu X, Lu J, Wang Z, He S, Xu Y,
He Y, Li T, Guo W, Tian H, Xu G, Xu X, Ma Y, Wang L, Wang L, Yan Y, Wang B, Xiao S,
Zhou L, LI L, Tan L, Zhang T, Ma C, Li Q, Ding H, Geng H, Jia F, Shi J, Wang S, Zhang N,
Du X, Du X, Wu Y. Prevalence of mental disorders in China: a cross-sectional epidemiological
study. Lancet Psychiatry. 2019;6:211–24.
Liu J, Ma H, He YL, Xie B, Xu YF, Tang HY, Li M, Hao W, Wang XD, Zhang MY, Ng CH,
Goding M, Fraser J, Herrman H, Chiu HF, Chan SS, Chiu E, YU X. Mental health system in
China: history, recent service reform and future challenges. World Psychiatry. 2011;10:210–6.
Liu Z, Huang Y, Lv P, Zhang T, Wang H, Li Q, Yan J, Yu Y, Kou C, Xu X, Lu J, Wang Z, Qiu H,
Xu Y, He Y, Li T, Guo W, Tian H, Xu G, Xu X, Ma Y, Wang L, Wang L, Yan Y, Wang B, Xiao S,
712 J. Li and S. Li
Zhou L, Li L, Tan L, Chen H, Ma C. The China mental health survey: II. design and field
procedures. Soc Psychiatry Psychiatr Epidemiol. 2016;51:1547–57.
Phillips MR, Zhang J, Shi Q, Song Z, Ding Z, Pang S, Li X, Zhang Y, Wang Z. Prevalence,
treatment, and associated disability of mental disorders in four provinces in China during 2001-
05: an epidemiological survey. Lancet. 2009;373:2041–53.
Xiang YT, Yu X, Sartorius N, Ungvari GS, Chiu HF. Mental health in China: challenges and
progress. Lancet. 2012;380:1715–6.
Ethical Issues in
Neuropsychopharmacotherapy: US
Perspective
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 714
Autonomy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 715
Beneficence and Nonmaleficence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 717
Justice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 718
Iatrogenic Properties and Trust . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 719
Nontherapeutic Use and Cosmetic Neuroenhancement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 722
Socioenvironmental Determinants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 723
Fluoxetine (Prozac ®) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 724
Clinical Indications and Side Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 724
Mechanism of Action of Fluoxetine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 726
Additional Ethical Considerations with Fluoxetine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 726
Modafinil (Provigil ®) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 728
Clinical Indications and Side Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 728
Mechanism of Action of Modafinil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 730
Additional Ethical Considerations with Modafinil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 731
Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 733
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 735
Abstract
The patient rights movement has designated medical autonomy as the predomi-
nant principle within the contemporary biomedical ethics paradigm. The
principles that govern the clinical delivery of medicine and clinical research
have, as of late, elicited numerous ethical concerns within the realm of neuropsy-
chopharmacotherapy that can only be addressed by reviewing the principles of
M. Menconi Jr.
Bioethics and Public Health, Columbia University, New York, NY, USA
e-mail: m.menconi@columbia.edu
V. Dubljević (*)
Department of Philosophy and Religious Studies, and Science Technology and Society Program,
North Carolina State University, Raleigh, NC, USA
e-mail: veljko_dubljevic@ncsu.edu
Introduction
Autonomy
Respect for autonomy refers to the right of the patient or individual to make personal
decisions regarding their own medical care, including the right to refuse unwanted
treatment (Beauchamp and Childress 2013). This right is recognized in medical and
legal contexts and is contingent on the patient’s ability to become fully informed of
proposed medical treatments and interventions. If a patient is to exercise their
medical autonomy directly, they must also hold the cognitive capacity – or compe-
tence – to understand the full array of risks and benefits of personal medical
decisions. If a patient is determined to lack this capacity by a medical provider, a
healthcare proxy or agent (typically a next of kin or family member) is appointed to
undergo the informed consent process for them. Today, the safeguard of medical
autonomy is informed consent, which constitutes a medicolegal contract (verbal or
written) between clinician and patient. A thorough and competent informed consent
process outlines the purpose of the proposed therapeutic intervention(s), relevant
risks, and expected therapeutic benefits. These are discussed between the clinician
and patient, during which time the patient may ask questions regarding the proposed
treatment. Once mutual understanding is reached, whether or not the proposed
intervention(s) are carried out is solely contingent on the patient’s explicitly stated
preferences.
The widespread integration of the informed consent process into the practice of
medicine and clinical research was the product of public and governmental
responses to the human rights violations observed in the Tuskegee Syphilis study.
In October 1932, poor black men in Tuskegee, Alabama were coerced into partic-
ipating in a study that sought to observe the effects of untreated Syphilis. Specifi-
cally, investigators sought to observe the neurological effects of untreated Syphilis.
There was no therapeutic purpose or motive for the study, which continued for over
40 years despite the discovery of penicillin as a curative treatment. The study
operated largely out of the public eye until a reporter published a whistleblowing
report on the front page of the Washington Star in 1972 (Brandt 1978). The
716 M. Menconi Jr. and V. Dubljević
consequent public outcry over such inhumane research resulted in the creation of a
federal ad hoc ethics committee to review the study. The review, led by the United
States Centers for Disease Control and Prevention (CDC), ultimately led to the
study’s termination in 1973. While aftermath of the Tuskegee Syphilis Study spurred
many of the informed consent standards used today and sparked an international
conversation on racial bias in medicine, the application of these standards to
contemporary healthcare delivery remains a complicated endeavor.
The Tuskegee Syphilis Study instigated national and international conversations
about patient rights, raising questions about what information – in the context of
clinical research or otherwise – individuals (or patients) seeking treatment are
entitled to receive. This marked the first time informed consent occupied a signifi-
cant platform in the medical ethics community. While the post-World War II
Nuremberg Trials elucidated important ethical guidelines surrounding the treatment
of human research subjects on a global scale, the domestic reality of Tuskegee
imparted crucial lessons that permeated the practice of clinical medicine and
research in the United States. In response to Tuskegee, the National Research Act
was passed by the U.S. Congress in 1973, in which autonomy – or “respect for
persons” as it is called in the law – was codified (Reverby 2011). The subsequent five
decades witnessed a monumental erosion of medical paternalism as a result. Since
the advent of primary care medicine, patients largely accepted physician recommen-
dations, including prognoses and treatment plans. Today however, between 20 and
40% of patients seek second opinions (Payne et al. 2014). As will be elucidated later
in this chapter, this newly empowered patient – a consumer of healthcare that is not
only acutely aware of their medical autonomy but chooses to maximize it – elicits
numerous ethical implications for neuropsychopharmacotherapy.
Unfortunately, the operationalization of the necessary elements of the informed
consent process is not always straightforward in the context of neuropsychopharma-
cotherapy. First, the side effects of psychotropic medications – particularly the
frequency, specificity, and risk of these side effects – are not always communicated
appropriately by the prescribing clinician. For example, many clinicians do not
disclose the well-documented risk of increased appetite that accompanies many
SSRIs, thus falsely implying that the risk of weight gain is insignificant (Gafoor
et al. 2018). However, even in instances where side effects and therapeutic expec-
tations are articulated properly by the prescribing clinician, clinicians and patients
frequently utilize data from industry-funded clinical trials to inform their decisions.
The use and interpretation of such data, sometimes published via misleading
industry-authored reports that minimize drug risks while exaggerating therapeutic
benefits, influence the clinician’s decision to offer and prescribe the medication, as
well as the patient’s decision to accept and comply with the proposed treatment. In
such instances where the information guiding medical decision-making is suspect,
medical autonomy is threatened because the patient’s consent, if offered, is not fully
and/or accurately informed.
Ethical Issues in Neuropsychopharmacotherapy: US Perspective 717
The principles of beneficence (do good) and nonmaleficence (do no harm) originate
from the earliest versions of the Hippocratic Oath authored many centuries ago.
Today, these elements guide clinicians in their administration of medical treatments
and interventions. While beneficence and nonmaleficence are presented in founda-
tional bioethics texts as separate principles, both principles jointly operate as a dyad
in clinical practice and research. For example, clinicians are obligated to present
treatment options to patients that stand to produce more benefit than harm. This is, of
course, if the treatments are administered correctly and unexpected complications
are avoided. In essence, this mode of reasoning involves ensuring the benefits of a
proposed medical intervention outweigh the known risks. However, all risk assess-
ments in the clinical environment remain highly contingent on the type of medical
intervention being considered. In the context of surgery, for example, surgeons
consider the complete array of possible complications that may arise during the
procedure, in addition to the predicted severity of the postoperative recovery course.
Patents are informed of these elements during the informed consent process. Alter-
natively, practitioners prescribing medication must be reasonably confident that a
drug’s potential side effects represent an acceptable harm. This determination is
made by weighing these potential harms against the drug’s expected therapeutic
benefits. This standard also holds for clinical trials. In all cases, the patient performs
their own risk-benefit analysis under the guidance of their clinician during the
informed consent process.
While this ethical analysis may appear elementary, it becomes vastly more
complicated in the context of psychotropic medications. For example, some studies
claim placebo effects are responsible for alleviating symptoms of depression and
anxiety in over 50% of patients that are prescribed SSRIs (Weimer et al. 2015).
However, side effects of the same medications, including sexual dysfunction and
suicidal ideation, are neither insignificant nor rare (Boseley 2000). Suppose a patient
experiences significant therapeutic relief as a result of a placebo effect (this phe-
nomenon is impossible to determine definitively outside of a randomized controlled
trial). If the clinician were able to determine that the placebo effect was producing
the symptom relief instead of the prescribed psychotropic, it would be unethical to
continue prescribing the drug if any possibility of serious side effects exists. In this
hypothetical case, since a transition to a harmless sugar pill would impart identical
therapeutic benefit with zero risk, the clinician would be obligated to transition the
patient to this zero-risk option. Of course, this would require that the patient be
blinded to the fact that the psychotropic drug they was originally prescribed is now in
fact a sugar pill. While such deception is not ethical or practical, the logic surround-
ing the risk-benefit analysis remains valid. Consequently, the outcome of a risk-
benefit analysis in this context is entirely contingent on the prescribing clinician’s
interpretation of risk, the data they employ to assess such risk, and the quality and
scope of relevant clinical training necessary to integrate these elements into practice.
718 M. Menconi Jr. and V. Dubljević
Justice
The principle of justice – or the degree to which healthcare resources are fairly
distributed among society – is largely dictated by the economic environment in
which a healthcare delivery system operates. Notions and theories of justice vary
widely across academic and political disciplines, with the most polarizing differ-
ences deriving from libertarian and utilitarian spheres. Such debates often include
discussion of oppositional ideological forces (i.e., collectivism vs. individualism),
and what rights ought to be recognized (and thus protected) by government and
society. While the perennial (and important) international debate regarding the merit
and feasibility of codifying health as a fundamental human right lies outside the
scope of this discussion, it is necessary to acknowledge that in societies without
universal healthcare coverage (such as the USA), socioeconomic status often deter-
mines the degree to which, if at all, neuropsychopharmacotherapy is accessible.
Recall our previous discussion of the modern patient’s newfound affinity for
second opinions. In certain instances, this phenomenon is associated with “doctor
shopping,” a term used to describe patient attempts at seeking out physicians with
permissive prescribing practices. This phenomenon – which does not occur exclu-
sively in the context of clinician malpractice – often manifests via clinicians’ low
diagnostic threshold for diagnosing attention deficit hyperactivity disorder (ADHD).
For example, some clinicians may require (and administer) a comprehensive psy-
choanalytic assessment before diagnosing ADHD, while others may simply inter-
view the patient and screen for relevant symptoms. While inconsistency in
diagnostic approaches may be a result of the clinician’s personal preferences or
training, such variability may alternatively arise from the diagnostic manual
referenced by the clinician. For example, US clinicians utilize the Diagnostic and
Statistical Manual for Mental Disorders (DSM), while European practitioners and
others utilize the International Classification of Diseases (ICD). Diagnostic criteria
found in the DSM are associated with higher levels of specificity, thus allowing less
room for interpretation. Diagnostic criteria found in the ICD are considered to be
more inclusive and flexible, thus allowing more room for clinician judgment (Dalal
and Sivakumar 2009).
In cases where diagnostic criteria are less stringent and/or the clinician adopts an
approach characterized by a low diagnostic threshold, the possibility of exploitation
by drug seeking patients increases. This includes university students seeking
amphetamines for cognitive enhancement purposes. Currently, no research has
investigated the ability of psychiatric clinicians to detect dishonesty during ADHD
assessments, thus the extent of the problem is largely unknown. Regardless, “doctor
shopping” in the United States requires an expensive and comprehensive private
insurance plan, or the ability to pay out of pocket for multiple visits. Patients with
Ethical Issues in Neuropsychopharmacotherapy: US Perspective 719
few resources, such as those enrolled in government insurance plans (i.e., Medicaid),
are assigned clinicians and therefore not afforded this flexibility. As such, the
socioeconomic status of the patient influences how, and to what degree, psychotropic
medications are accessed.
In addition to economic access barriers, the high level of subjectivity inherent in
diagnostic approaches to psychopathology raises concerns about establishing a
uniform standard of care. The standard of care is defined as what a “reasonable
clinician” with relevant specialty training considers a competent approach to a
particular clinical situation (Cohen et al. 2017). The DSM, currently in its fifth
edition, also exhibits subjectivity and variability through its constant revisions,
which continue to generate significant controversy among US practitioners (Amer-
ican Psychiatric Association 2013). For example, the American Psychiatric Associ-
ation is currently debating the introduction of new diagnostic revisions to the
DSM-V. Most notably, proposed additions include expanded “dimensional mea-
sures,” which allow clinicians to diagnose mental disorders on a spectrum based
on the severity and prevalence of symptoms (American Psychiatric Association
2020). It is unclear how clinicians will interpret and implement these new criteria,
thus future effects on prescribing patterns are difficult to predict. Such clinical
inconsistency, now poised to increase as a result of proposed expansions of the
less precise “dimensional” DSM criteria, raises significant justice-related ethical
concerns, namely fairness. For example, it increases the probability that a patient
presenting with identical symptoms to two different psychiatric clinicians will
receive disparate diagnoses and treatments. This reality is problematically charac-
terized by decreased clinical objectivity and leaves the treatment course (and patient)
more susceptible to clinician biases of all types (clinical, social, etc.). These concerns
are exacerbated for patients of low socioeconomic status who lack the ability to
choose their prescribing clinician.
While all four pillars of biomedical ethics are relevant to our discussion, the
principalist biomedical ethics paradigm as it operates today remains predominantly
driven by medical autonomy. The confluence of social, economic, and cultural forces
within the psychotherapeutic enterprise raises significant ethical questions for the
research and practice of neuropsychopharmacotherapy. These dilemmas – collec-
tively representing the ethical aspects of neuropsychopharmacotherapy – are most
logically taxonomized into three primary categories: iatrogenic properties and trust,
nontherapeutic use and cosmetic neuroenhancement, and socioenvironmental deter-
minants of health. Through this manner of organization, we seek to examine the
ethical implications of neuropsychopharmacotherapy as they affect patients, clini-
cians, and the public at large.
The global history of medicine and clinical research has been consistently afflicted
with instances of gross human rights violations at the hands of clinicians. It is
therefore unsurprising that many of the same issues continue to plague the practice
720 M. Menconi Jr. and V. Dubljević
are significant and require serious self-awareness on the part of the patient to ensure
that the onset of some side effects, such as suicidal thoughts, do not become
catastrophic (Appelbaum 2007). This raises serious medico-legal questions sur-
rounding the capacity of the psychiatric patient to undergo a proper informed
consent process in nonemergent contexts. For example, consider an instance in
which a patient in emotional turmoil presents to a psychiatrist following the loss
of a loved one. Whether or not this patient is able to properly internalize the risks of a
proposed course of psychotropic medication is up to the clinician’s sole discretion. If
this encounter is the clinician’s first with the patient, such an assessment is compli-
cated due to the clinician’s lack of familiarity with the patient’s baseline emotional
function and coping mechanisms.
The notion of capacity (used interchangeably in both legal and medical literature
with “competence”) is further complicated by the subjective nature of cognitive
evaluations. It is not uncommon for psychiatric clinicians to reach different conclu-
sions following capacity assessments of the same patient (Appelbaum 2007). These
assessments are less complex in instances where the patient presents with psychosis.
In these cases, clinicians are legally empowered to file for a medical hold (commonly
understood as civil commitment), effectively restraining and detaining the patient
until forced treatment restores the patient’s cognitive capacity.
Questions regarding criteria for determination of cognitive capacity (the degree to
which an individual is deemed competent to make personal healthcare decisions) in
clinical settings additionally raise ethical concerns. In extreme cases, psychiatric
practitioners are able to temporarily restrict medical autonomy entirely by forcing
treatment via state-specific civil commitment statutes if they determine patients are a
danger to themselves or others. In such cases, no informed consent process is
required, as the patient has been clinically determined to lack capacity (Appelbaum
2007). Regardless, the complexities and clinical inconsistencies surrounding the
informed consent process in the context of psychotropic medications provide
ammunition for those skeptical of the psychotherapeutic enterprise.
Furthermore, the questionable practices of corporate pharmaceutical advertising
of psychotropic medications have only furthered ethical concerns surrounding trust.
While recent public focus has revolved around the marketing of prescription opioids,
antidepressants were the most heavily marketed drug class in the mid 2000’s
(McHenry 2006). Significant levels of spending on direct-to-consumer advertising
and massive distributions of free samples to providers has been linked to increased
fluoxetine prescription for children, despite a lack of evidence of its efficacy in a
number of pediatric clinical trials (Boseley 2000). Specifically, experts and regula-
tory authorities have questioned the grounds on which pharmaceutical companies
such as Eli Lilly (the original manufacturer of Prozac) make claims of efficacy in
advertisements. Opposition to these claims of efficacy takes many forms, from direct
challenges to the neuroscience underlying various mechanisms of action (e.g., the
Serotonin hypothesis) to allegations of systemic rigging of the clinical trial process
and selective publication of subsequent trial results (McHenry 2006). The onslaught
of direct to consumer advertising in conjunction with the development of a powerful,
industry-dependent clinical research enterprise has led some critics to coin the
722 M. Menconi Jr. and V. Dubljević
umbrella term “corporate psychiatry” (McHenry 2006). Both of these elements will
be elucidated further in the specific discussion of fluoxetine.
Socioenvironmental Determinants
The previous two decades have produced an extensive body of public health
research that attempts to identify the underlying causes of disease, illness, and
mortality. In an attempt to focus on disease prevention, the field of public health
has integrated crucial elements of sociomedical science into its theory and practice.
As such, a paradigm shift has resulted that supplants the individualistic health
behavior model (i.e., focus on personal responsibility) with a holistic psychosocial
scheme that evaluates the patient’s environment for factors that instigate and rein-
force maladaptive health behaviors. Consequently, when examining trends in public
health such as increased use of psychotropic medications, we are simultaneously
obligated to seek out potential drivers of patient (or consumer) behavior, particularly
at the societal level. In the context of neuropsychopharmacotherapy, the socio-
environmental determinants that begin to explain the increased use of psychotropic
medications are found in two primary realms. First, the ways in which the market
economy affects and often dictates healthcare delivery provides insight into how the
concept of the “patient” has been reframed to meet market demands. This phenom-
enon has produced an environment that has readily increased the availability of
prescription psychotropic medications via licit and illicit means. Second, contem-
porary societal trends in academia and occupational productivity represent social
forces that encourage exploitation of their widespread availability.
The reduction from patient to consumer has promulgated resounding effects on
the credibility of the healthcare system. In the context of psychoactive pharmaceu-
ticals, an ongoing ethical debate persists – both in theory and clinical practice –
regarding whether or not the conceptualization of psychotropics merely includes
therapeutic interventions or if the scope includes cosmetic neuroenhancers (Jotterand
and Dubljević 2016). If the patient is primarily treated as a customer, the prescribing
provider’s fiduciary obligation to the patient now includes the subjective, market-
driven element of patient satisfaction as an aside to beneficent therapeutic care. It is
therefore unsurprising that over 40% of SSRI prescriptions issued to patients by
general practitioners are the result of patient requests (Mercier et al. 2014). Recent
efforts to combat financial abuse of the pay-per-service physician reimbursement
model have resulted in a nationwide increase in the use of patient satisfaction
surveys, the results of which are compiled by hospital administrators and used as a
performance benchmark. Physician compensation is either increased or decreased
based on these results.
This increased incentive for prescribers – including physicians and nurse practi-
tioners – to satisfy their patients has resulted in a decreased deterrence of requests for
medications patients believe they need. This phenomenon is evidenced by the
problematic over-prescription of antibiotics, where drugs like penicillin are pre-
scribed for antibiotic-resistant viral infections for the purposes of meeting the
patient’s conception of a successful medical appointment. This nearly universally
involves the expectation of receiving a prescription at the conclusion of the visit
(Smith 2012). This affront to medical paternalism has promulgated a public health
724 M. Menconi Jr. and V. Dubljević
Fluoxetine (Prozac ®)
Fluoxetine and some other SSRI’s have garnered attention for their unusually high
placebo effects. Knorr et al. (2019) completed a systematic literature review of SSRI
effects found in 51 clinical trials. Of the 249 outcome tests that were performed on
this data, only 4 revealed a statistically significant change in patient outcome
Ethical Issues in Neuropsychopharmacotherapy: US Perspective 727
following SSRI intervention compared to effects with a placebo (Knorr et al. 2019).
The high placebo rate holds significant ethical implications for the risk-benefit
analysis that providers and patients are obligated to undertake prior to treatment. If
more than half of fluoxetine prescriptions are alleviating symptoms in patients via a
mere placebo effect, prescribers must question whether administering a drug with
such serious potential side effects (e.g., suicidal ideation) outweighs its benefits.
Additionally, placebo data is not included as part of the informed consent process in
conventional psychiatric treatment settings (outside of a clinical trial), and thus the
patient may be left with incomplete and/or misleading information with which to
guide their treatment decisions, including the option of in-person therapy – usually
cognitive behavioral therapy – as a first-line treatment option with minimal risk
before beginning fluoxetine.
Additionally, the prescription of fluoxetine has promulgated a series of scope of
practice issues among nonpsychiatric clinicians. Currently, general practitioners
(physicians or nurse practitioners who did not undergo specialty psychiatric training
post clinical degree) are responsible for 80% of SSRI prescriptions worldwide
(Mercier et al. 2014). This phenomenon coincides with a systematic over-
prescription of antidepressant medications; a recent analysis found that up to 18%
of fluoxetine prescriptions were for the purposes of off-label, nonpsychiatric condi-
tions (Mercier et al. 2014). While professional organizations such as the American
Psychiatric Association (APA) have not openly asserted that general practitioners
lack the specialized expertise for the prescription of psychotropic medications, there
is some evidence that prescriptions issued by general practitioners have not met the
standard of care. This includes instances of terminating treatment before therapeutic
effects may be realized and prescription of subtherapeutic dosages (Mercier et al.
2014). The APA has issued treatment guidelines for general practitioners to address
some of these issues. However, it remains unclear whether the primary care delivery
system is appropriately suited for neuropsychopharmacotherapy. Areas of concern
include short appointment times, limited psychosocial understanding of patients, and
high incidence of complex comorbidities. Furthermore, primary care clinicians are
not trained in non-pharmaceutical psychiatric therapies, which are frequently co-
indicated with fluoxetine and other SSRIs for the treatment of anxiety and depres-
sion. In many instances, anxiety and depression may be more appropriately treated
with cognitive behavioral therapy (CBT) or alternative non-pharmaceutical therapies
alone. However, such treatments often require access to a psychiatric or psycholog-
ical specialist, which is problematic for patients whose geographic location or
insurance status represent a barrier to access. Finally, the exponential increase in
fluoxetine prescription and use since the early 1990s may be at least partially
attributed to the extensive and aggressive corporate marketing campaigns commis-
sioned by Eli Lilly & Company, the original manufacturer of Prozac ®. For example,
in 2003 Eli Lilly was sued in a Florida court for invasion of privacy. The company
was accused of using confidential medical information from patient charts to send
unsolicited samples of Prozac ® to their homes without a prescription (Dyer 2003).
Ely Lilly later fired the company officials who contrived the scheme. Such aggres-
sive and unlawful marketing tactics were most frequent from 1996 to 2005, during
728 M. Menconi Jr. and V. Dubljević
1
In this section, we draw on Dubljević 2016. We are also grateful to Anirudh Nair for assisting
the corresponding author in the updated literature review of modafinil effects.
Ethical Issues in Neuropsychopharmacotherapy: US Perspective 729
Contrary to the results in the domain of attention and working memory, Kredlow
et al. (2019) noted that improvements in executive functioning and processing
speed were, in fact, significant (g ¼ 0.1, g ¼ 0.2). Battleday and Brem (2015)
adopted a much broader approach to executive functioning, evaluating tests involv-
ing complex tasks. Most of the studies reviewed for “complex tasks” involved
administering tests that demanded a range of abilities such as higher-order function-
ing, decision-making, creativity, and fluid intelligence. Firstly, Battleday and Brem
(2015) note that no improvement was found in their review of cognitive flexibility
tests, and participants actually performed worse following modafinil intake. How-
ever, they also found that in studies that tested for fluid intelligence, all showed
improvements in participants, particularly those involving medium difficulty tasks.
When examining studies that tested creativity, Battleday and Brem (2015) found that
for the most part, studies that tested convergent-thinking-based creativity, which
involves problem strategies, showed no improvements in performance.
Currently, modafinil requires a prescription to be obtained in theory, but clinician
involvement can be circumvented in practice. In fact, the ease with which it can be
obtained via online marketplaces has been the subject of ethical scrutiny. In part, this
may be a result of policy decisions: the U.S. Food and Drug Administration (FDA)
extended the modafinil label to help alleviate symptoms of sleepiness associated
with shift-work syndrome (Food and Drug Administration 2007), and the U.S. Drug
Enforcement Agency (DEA) classified it as a Schedule 4 drug, i.e., a drug with low
abuse potential. Increases in on-label prescriptions and policy designations ostensi-
bly based on safety contribute to the popularity of modafinil as a go-to stimulant. An
additional explanatory factor might be the unfettered availability for online purchase
and complete lack of enforcement. Hockenhull et al. (2019) reported that online
searches using popular search engines yielded 73 websites that sold modafinil in the
United Kingdom for the purpose of enhancement without the need for a prescription.
A majority of these websites listed possible desired effects such as increased
alertness, improved memory, increased concentration and productivity and perfor-
mance enhancement. Similarly, Dursun et al. (2019) found that 13 websites sold
modafinil to consumers in Australia without prescription, with a median price of
118 Australian dollars (81 USD) for 90,150 mg pills. Such availability of
unprescribed modafinil only fuels the already existing ethical controversy.
2005). At first, Provigil (the first commercial brand of modafinil in the United States)
was approved to treat narcolepsy, but the label was subsequently expanded to
include treatment of sleep apnea and shift work sleep disorder. From 2001 through
2006, Cephalon allegedly promoted Provigil as a nonstimulant drug for the treatment
of sleepiness, tiredness, decreased activity, lack of energy and fatigue. In 2002, the
FDA sent Cephalon a letter, warning the company to cease and desist promoting
Provigil off-label. Cephalon apparently ignored this warning and continued to
undertake its promotional practices via a variety of techniques, such as training its
sales force to disregard or downplay restrictions of the FDA-approved label. The
activities of Cephalon resulted in a lawsuit which was settled in 2008 for 425 million
US$ (Department of Justice 2008). However, the government fine still did nothing to
curb Cephalon’s unethical practices. The effectiveness of these questionable promo-
tional strategies is evidenced in the steady rise of patients filling prescriptions for on-
and off-label uses of modafinil; 90% of all modafinil prescriptions are issued for
off-label purposes (Cahill 2005), and this rate continues to increase (Kasselheim
et al. 2012). In 2016, there were almost 1.5 million modafinil prescriptions in the
United States.2
The relatively permissive attitudes regarding modafinil are a result of its desig-
nation as “relatively safe,” especially compared to traditional stimulants. Indeed, the
wakefulness promoting properties of modafinil might be very beneficial for the
society at large by alleviating the effects of fatigue during work and even freeing
up more time for leisure activities (see Tannenbaum 2012). However, modafinil is
not without detrimental effects that may impact safety of patients (and other users).
In 2007, Health Canada – the health agency with historically the most cautious
approach toward the emergence of serious health issues associated with modafinil
use – warned about the possibility of severe allergic reactions to modafinil and a host
of associated skin problems, including rashes, hives, and sores (CBC 2007). The
probability of allergic and other types of dermatological reactions to modafinil seems
to be greater than for other comparable medications. Furthermore, the FDA findings
on the long-term side effects of modafinil use are inconclusive, yet the severity of
rashes and dermatological toxicity observed in younger patients fuels doubts as the
potential impact is uncertain (Kumar 2008). Life-threatening side effects were also
reported, including instances of extreme allergic reactions resulting in myocarditis
(Bauml et al. 2019). These concerns were corroborated in a study conducted by
Davies et al. (2013). They observed that incidences of skin conditions among
patients in the United Kingdom that were prescribed modafinil were responsible
for 33% of patients discontinuing treatment. Finally, sleep deprivation may be
exacerbated by the long-term use of modafinil, which may lead to immune system
failure as well as emotional and physical damage (Dubljević 2016; Eliyahu et al.
2007).
Evidence suggests that significant placebo effects are associated with modafinil.
Battleday and Brem (2015) comment on the “ceiling effects” they observed in their
2
See https://clincalc.com/DrugStats/Drugs/Modafinil
Ethical Issues in Neuropsychopharmacotherapy: US Perspective 733
review, and they claim that “when these ceiling effects were lessened by only
analyzing data from low baseline performers, many studies actually did detect
significant differences between modafinil and placebo groups.” It is important to
note this recurring trend of selective efficiency of modafinil in low-baseline per-
formers. It would appear that any putative “enhancement” effects of modafinil are in
fact placebo effects or instances of “self-medication” with nonprescribed psychotro-
pics to handle social pressures or life-style choices. Arguably, the noticeable cogni-
tive benefits of modafinil observed in the “high-pressure” populations of students,
medical professionals, and shift workers are likely attributable to the prevalence of
sleep deprivation in those populations, which means that, in reality, they are typically
performing below their baseline. As Battleday and Brem (2015) note: “robust effects
on these same tasks are seen with sleep deprivation, when all participants effectively
become low baseline performers.” In light of this information, the purported benefits
of modafinil should not be considered proven, and the inconsistencies between
modafinil’s supposed enhancement effects raise the need to devise and implement
more innovative testing methods that account for these discrepancies and can
accurately ascertain the efficacy of modafinil in enhancing cognition in populations
of normal and high functioning individuals. It also highlights the importance of
conducting these tests on populations of healthy individuals who neither suffer from
cognitive impairments nor are suffering from any health issues when being tested,
including sleep deprivation.
Concluding Remarks
(Klein et al. 2016). Such skepticism is only exacerbated by the reality that the
mechanisms of action responsible for the therapeutic effects of many psychotropic
medications remain unidentified or only partially explained. The traditional
approach to this question by practitioners and researchers within the psychothera-
peutic enterprise has involved designating the diagnosed mental illness (depression,
for example) as a departure from one’s authentic self, thus neuropsychopharma-
cotherapy is required to restore authenticity. While many patients accept this
approach to treatment, new neurotechnologies such as deep brain stimulation
(DBS) have resulted in patients expressing concern that manipulation of mood and
emotion is “artificial” and correspondingly results in an unacceptable loss of agency
(Klein et al. 2016). This is in contrast to those who licitly and illicitly obtain
amphetamines and other stimulants for the purposes of cosmetic neuroenhancement.
The large majority of these individuals embrace the notion of artificial optimization
of neurochemistry for the purpose of achieving a competitive edge in high cognitive
demand environments. By doing so, they implicitly acknowledge that their authentic
self is in need of enhancement and therefore little attention is paid to the potential
loss of agency or authenticity. Regardless of the existential position adopted in this
context, serious implications for autonomy lie in the balance.
Furthermore, the question of access to neuropsychopharmacotherapy continues to
vault to the forefront of national debate as U.S. healthcare remains a top priority for
voters. Currently, the United States is in the midst of a severe mental health services
shortage, with over three-quarters of counties meeting less than half of psychotropic
prescription demand (Thomas et al. 2009). While the shortage of prescribers is
substantial on its own, some studies estimate that the nonprescriber (those delivering
psychotherapy only) shortfall is ten times greater than the prescriber shortfall
(Thomas et al. 2009). As such, current rates of accessibility for neuropsychophar-
macotherapy are alarmingly low, signaling a healthcare delivery crisis in the arena of
mental health. Access rates are even lower for low SES populations in both urban
and rural areas. Such reduced provider numbers reflect only the status quo – if
current proposals for universal healthcare in the United States were to succeed,
addition of the anticipated 30 million newly insured beneficiaries to the healthcare
marketplace will likely have wide-ranging consequences, the effects of which may
disproportionately burden vulnerable populations. Even in the event the healthcare
system remains unchanged, solutions to the existing challenges continue to elude
experts. Such unfortunate realities promulgate serious questions of justice – how do
we distribute therapeutic mental health resources fairly?
There are no simple solutions to the ethical issues raised in this chapter – the
difficult conceptual work of balancing normative values with critical bioethical
reasoning will only grow more rigorous as neuropsychopharmacotherapy continues
to advance. However, with the lessons of history as guiding light, it is necessary that
we do not cede ground to the polarized hyperbole that characterizes much of the
contemporary public debate surrounding these issues. With the critical lens articu-
lated in this chapter, it is our sincere hope that current and future practitioners,
researchers, innovators, and consumers of neuropsychopharmacotherapy will be
more attuned to the ethics of their endeavors. In bringing these ethical implications
Ethical Issues in Neuropsychopharmacotherapy: US Perspective 735
References
Academy of Medical Sciences [AMA]. 2008. Brain science, addiction and drugs, An Academy of
Medical Sciences working group report chaired by Professor Sir Gabriel Horn FRS FRCP,
Available at: http://www.acmedsci.ac.uk/p99puid126.html. Accessed 5 Mar 2013.
Academy of Medical Sciences [AMA]. 2012. Human enhancement and the future of work. Joint
report of the Academy of Medical Sciences, the British Academy, the Royal Academy of
Engineering and the Royal Society, Available at: http://royalsociety.org/uploadedFiles/Royal_
Society_Content/policy/projects/human-enhancement/2012-11-06-Human-enhancement.pdf.
Accessed 10 Nov 2012.
American Psychiatric Association (Ed.). (2013). Diagnostic and statistical manual of mental
disorders Dsm-5. American Psychiatric Association.
American Psychiatric Association. 2017, April 27. APA Public Opinion Poll – Annual Meeting
2017. APA Newsroom. https://www.psychiatry.org/newsroom/apa-public-opinion-poll-annual-
meeting-2017
Andrews AM. Does chronic antidepressant treatment increase extracellular serotonin? Front
Neurosci. 2009;3:246–7. (n.d.).
Appel JM. When the boss turns pusher: a proposal for employee protections in the age of cosmetic
neurology. J Med Ethics. 2008;34(8):616–8.
Appelbaum PS. Assessment of patients’ competence to consent to treatment. N Engl J Med.
2007;357(18):1834–40.
Ballon JS, Feifel D. A systematic review of Modafinil: potential clinical uses and mechanisms of
action. J Clin Psychiatry. 2006;67(4):554–66.
Battleday RM, Brem A-K. Modafinil for cognitive neuroenhancement in healthy non-sleep-
deprived subjects: a systematic review. Eur Neuropsychopharmacol. 2015;25(11):1865–81.
https://doi.org/10.1016/j.euroneuro.2015.07.028.
Bauml M, Udi J, Klingel K, Bode C, Warnatz K, Zirlik A, et al. Severe eosinophilic myocarditis
associated with modafinil in a patient with normal peripheral eosinophil count. Clinical
Research in Cardiology: Official Journal of the German Cardiac Society. 2019,
August;108:963–6. https://doi.org/10.1007/s00392-019-01434-w.
Beauchamp TL, Childress JF. Principles of biomedical ethics. 7th ed. New York [u.a.]: Oxford
University Press; 2013.
Boseley S. Happy drug Prozac can bring on impulse to suicide, study says. The Guardian. 2000,
May 22. https://www.theguardian.com/science/2000/may/22/drugs.uknews
Brandt AM. Racism and research: the case of the Tuskegee syphilis study. Hast Cent Rep. 1978;8
(6):21. https://doi.org/10.2307/3561468.
Busch SH, Barry CL. Pediatric antidepressant use after the black-box warning. Health Aff. 2009;28
(3):724–33. https://doi.org/10.1377/hlthaff.28.3.724.
Cahill M. The ethical consequences of modafinil use. Penn Bioethics Journal. 2005;1(1):1–3.
Caldwell JA, Smythe NK, Caldwell JL et al. The Effects of Modafinil on Aviator Performance
During 40 Hours of Continuous Wakefulness: A UH-60 Helicopter Simulator Study, U.S. Army
Aeromedical Research Laboratory, Report No. 99–17. 1999.
Cohen IG, Lynch HF, Sepper E. Law, religion, and health in the United States: Cambridge
University Press; 2017.
Dalal PK, Sivakumar T. Moving towards ICD-11 and DSM-V: concept and evolution of psychiatric
classification. Indian J Psychiatry. 2009;51(4):310–9.
Davies M, Wilton L, Shakir S. Safety profile of modafinil across a range of prescribing indications,
including off-label use, in a primary care setting in England: results of a modified prescription-
736 M. Menconi Jr. and V. Dubljević
Kanal MG, Ozkan C, Doganavsargil O, Eryilmaz M. Late onset mania possibly related to modafinil
use: a case report. Bulletin of Clin Psychopharmacol. 2012;22(1):71–74.
Kasselheim AS, Myers JA, Solomon DH, Winkelmayer WC, Levin R, Avorn J. The prevalence and
cost of unapproved uses of top-selling orphan drugs. PLoS One. 2012;7(2):e31894. https://doi.
org/10.1371/journal.pone.0031894.
Kim D. Practical use and risk of Modafinil, a novel waking drug, Environ Health Toxicol. 2012.
https://doi.org/10.5620/eht.2012.27.e2012007.
Klein E, Goering S, Gagne J, Shea CV, Franklin R, Zorowitz S, Dougherty DD, Widge AS. Brain-
computer interface-based control of closed-loop brain stimulation: attitudes and ethical considerations.
Brain-Comput Interfaces. 2016;3(3):140–8. https://doi.org/10.1080/2326263X.2016.1207497.
Knorr U, Madsen JM, Kessing LV. The effect of selective serotonin reuptake inhibitors in healthy
subjects revisited: a systematic review of the literature. Exp Clin Psychopharmacol. 2019;
27(5):413–32. https://doi.org/10.1037/pha0000264.
Kredlow MA, Keshishian A, Oppenheimer S, Otto MW. The efficacy of Modafinil as a cognitive
enhancer: a systematic review and meta-analysis. J Clin Psychopharmacol. 2019;39(5).
Retrieved from https://journals.lww.com/psychopharmacology/Fulltext/2019/09000/The_Effi
cacy_of_Modafinil_as_a_Cognitive_Enhancer_.7.aspx
Kumar R. Approved and investigational uses of modafinil: an evidence-based review. Drugs.
2008;68(13):1803–39. https://doi.org/10.2165/00003495-200868130-00003.
Lagarde D, Batejet D, Van Beers P, Sarafian D, Pradella S. Interest of Modafinil, a new
psychostimulant, during a sixty-hour sleep deprivation experiment. Fundam Clin Pharmacol.
1995;9(3):271–9.
Lieb K. Hirndoping: Warum wir nicht alles schlucken sollten. Mannheim: Artemis & Winkler;
2010.
Machado-Vieira R, Baumann J, Wheeler-Castillo C, Latov D, Henter I, Salvadore G, Zarate C. The
timing of antidepressant effects: a comparison of diverse pharmacological and somatic treat-
ments. Pharmaceuticals. 2010;3(1):19–41. https://doi.org/10.3390/ph3010019.
Maher B. Poll results: look who’s doping. Nature. 2008;452(7188):674–5. https://doi.org/10.1038/
452674a.
Mariani JD, Hart CL. Psychosis associated with modafinil and shift work. Am J Psychiatr. 2005;
162(10):18.
Marshall KP, Georgievskava Z, Georgievsky I. Social reactions to valium and Prozac: a cultural lag
perspective of drug diffusion and adoption. Res Soc Adm Pharm. 2009;5(2):94–107. https://doi.
org/10.1016/j.sapharm.2008.06.005.
McHenry L. Ethical issues in psychopharmacology. J Med Ethics. 2006;32(7):405–10. https://doi.
org/10.1136/jme.2005.013185.
Mercier A, Auger-Aubin I, Lebeau J-P, Schuers M, Boulet P, Van Royen P, Peremans L. Why do
general practitioners prescribe antidepressants to their patients? A pilot study. Bio Psycho Soc
Med. 2014;8(1):17. https://doi.org/10.1186/1751-0759-8-17.
Minzenberg MJ, Carter CS. Modafinil: a review of neurochemical actions and effects on cognition.
Neuropsychopharmacology. 2008;33:1477–502.
Mohamed AD. Modafinil has the potential for addiction. AJOB Neurosci. 2012;3(2):36–8.
Mohamed AD, Sahakian BJ. The ethics of elective psychopharmacology. Int
J Neuropsychopharmacol. 2012;15:559–71.
Murillo-Rodriguez E, Veras A, Rocha N, Budde H, Machado S. An overview of the clinical uses,
pharmacology, and safety of Modafinil. ACS Chem Neurosci; 2017; 9. https://doi.org/10.1021/
acschemneuro.7b00374.
National Alliance on Mental Illness. 2017, December. Anxiety disorders. Mental health conditions.
https://www.nami.org/learn-more/mental-health-conditions/anxiety-disorders
National Institutes of Health. 2019, February. Major depression. Mental health information. https://
www.nimh.nih.gov/health/statistics/major-depression.shtml
Nulman I, Rovet J, Stewart DE, Wolpin J, Pace-Asciak P, Shuhaiber S, Koren G. Child development
following exposure to tricyclic antidepressants or fluoxetine throughout fetal life: a prospective,
738 M. Menconi Jr. and V. Dubljević
Hanfried Helmchen
Contents
The Normative Framework and Its Change Over Time . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 740
Informed Consent . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 741
Information . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 741
Capacity to Consent . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 741
Coercive Medication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 742
Acute Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 742
Chronic Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 742
Preventive Medication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 744
Primary Prevention of Schizophrenia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 745
Primary Prevention of Dementia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 746
Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 749
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 749
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 749
Abstract
Acknowledgement of human rights and its translation into psychiatric action is an
increasingly extensive achievement of the past decades. Orientation towards the
individual patient means to recognize both his/her right and capacity of self-
determination as well as to consider his/her preferences. This is particularly valid
in the use of psychotropic medication in problematic situations such as in
coercive and in preventive medication. In view of coercive medication in life-
threatening psychiatric emergencies, psychiatrists must convert a seemingly
antagonism between respecting the autonomy of the mentally ill and their
H. Helmchen (*)
Department of Psychiatry & Psychotherapy, CBF, Charité – University Medicine Berlin, Berlin,
Germany
e-mail: hanfried.helmchen@charite.de
Keywords
Self-determination · Informed consent to medication · Preventive medication ·
Change of normative framework · Coercive medication · Dementia ·
Schizophrenia
Two hundred years ago a dominating psychiatric conviction was that “the mentally
ill is under age as a child”; and this was the first sentence of the influential book Die
Irrenanstalt in allen ihren Beziehungen (The lunatic asylum in all its relationships);
it was published 1831 by Carl Friedrich Roller, the very influential founding director
of the then most modern asylum Illenau in south-west Germany (Roller 1831).
Today, psychiatry had implemented the principle of informed consent into clinical
practice and is on the way to recognize the right of self-determination of the mentally
ill as is required by the UN Convention on the Rights of Persons with Disabilities
(UN 2006); the CRPD is interpreted by the Central Ethics Committee at the German
Federal Board of Physicians as: “the assistant does not substitute the decisions of the
patient with his own decisions but assists him/her in the exercise of his/her right of
self-determination.” (Zentrale Ethikkommission bei der Bundesärztekammer 2016).
The British Supreme Court expressed this in a decision in 2016 as follows: “The UK
Supreme Court ruled that the standard for what physicians should inform patients
about the risks, benefits, and alternatives of treatment will no longer be determined
by what a responsible body of physicians deems important but rather by what a
reasonable patient deems important” (Spatz et al. 2016).
Such a change of the normative framework corresponds to a change of psychiatric
attitudes towards the patient (Helmchen 2019b). But it needs continuous educational
efforts across generations; for example, although the psychiatrist Wilhelm
Griesinger in 1845 had already written about the use of general human rights for
the mentally ill (Griesinger 1845), it was deemed necessary by the German Society
for Psychiatry and Psychotherapy, Psychosomatics and Neuroscience (DGPPN) to
publish a “plan of action” in 2018 in order to implement the CRPD into clinical
reality (DGPPN 2018a).
This understanding of the patient’s right of self-determination will be discussed
by the following aspects of psychopharmacotherapy: 1. Informed consent to medi-
cation, 2. Coercive medication, 3. Preventive medication.
Ethics of Informed Consent: Coercive and Preventive Medication 741
Informed Consent
In order to realize his right of self-determination, the patient must understand all
information that is relevant for his decision to accept or to reject the drug treatment
recommended by his psychiatrist. Accordingly it is the twofold responsibility of the
psychiatrist to provide to the patient all relevant evidence-based information as well
as to test whether the patient can use this information for a rational decision, that is,
the patient‘s capacity of self-determination (Capacity to consent is included in the
broader term capacity of self-determination; the latter term is clearer in implying also
the capacity to reject as well as being a human right).
Information
Respect for autonomy requires improving the patient’s understanding of the disorder
and of the benefits and risks of both the recommended and alternative therapeutic
interventions as well as the elucidation of their meaning for the patient’s personal
values, especially his expectations towards the therapy (Helmchen 2010). The infor-
mation must include the aim of psychotropic medication, that is, to mitigate the present
symptoms, or to prevent their recurrence – not the least in order to strengthen the
patient’s capacity to use psychotherapeutic and rehabilitative interventions (Helmchen
2018). Besides these intended benefits, the medication may also have unwanted effects
that must be named; the psychiatrist should explain that nevertheless he recommends
the medication because in his view its probable benefits will outweigh its possible
risks; in some situations also the timing will be of relevance: some unwanted effects of
drug treatment such as tardive hyperkinesias and a metabolic syndrome become
manifest only in long-term treatment; therefore, information about these risks must
be given at the latest when the acute treatment will proceed to maintenance treatment.
Of course, the psychiatrist must inform the patient also about alternative treatments,
and why he recommends them only as a second or additional or later choice. Also the
patient should be informed on the procedure of medication: how often, how long, as a
pill or by injection, mode of control (monitoring for efficiency, safety, and adherence),
and with regard to poor adherence not only about the risk of relapse but also about
withdrawal symptoms or “discontinuation syndromes” of antipsychotics (Buffo 2020;
Keks et al. 2019) as well as of antidepressants that could be prevented by a monitored
slow reduction over weeks (Henssler et al. 2019).
Recently a first schedule for measuring the quality of this information has been
published (Spatz et al. 2020). It seems to be a further step to the standardization of
medical procedures.
Capacity to Consent
If there are doubts about the patient’s capacity of self-determination the psychiatrist
should test it by asking the patient to repeat the given information, that is, what will
742 H. Helmchen
be done, why will it be done, and what it means for the patient him-/her self. In case
of questionable capacity of self-determination, the patient should be assisted to
improve his understanding and decision-making capacity (Zentrale
Ethikkommission bei der Bundesärztekammer 2016); (Szmukler and Rose 2014),
for example, by corrective feedback in case of misrepresenting the given information
(Prince et al. 2013), or by assistance in drawing up a “facilitated psychiatric advance
directive” (F-PAD (Swanson et al. 2006)). Information needs time and should not be
a single event but a process in which the patient should have time to consider and ask
again for clarification (Helmchen 2019a).
Coercive Medication
The psychiatrist is entitled to medicate the patient against his will only under two
conditions: first, if the patient lacks the capacity for self-determination, especially the
capacity to consent to the needed medication in question, and, second, if the patient‘s
violent or suicidal behavior leads to acute life-threatening situations. But the assess-
ment of the capacity for self-determination may be difficult and may lead only to a
questionable result; in such cases the physician’s obligation of caring for the best
interest of the patient (and the safety of others) may conflict with the physician’s
obligation to respect the autonomous will of the patient. Therefore, a consent of the
patient should be attempted at by all gentler measures, and coercion must be only the
last possibility, and according to the law controlled by a judge (DGPPN 2014)
(Deutsche Gesellschaft für Psychiatrie und Psychotherapie Psychosomatik und
Nervenheilkunde (DGPPN) (Task Force “Ethik in Psychiatrie und Psychotherapie”)
2018); (Steinert and Hirsch 2020). All measures of coercion should be fairly
balanced against their expected potential benefits (Appelbaum and Redlich 2006).
Acute Conditions
Commonly the patient voluntarily accepts the medication as help against his painful
experiences. However, sometimes he might not understand the reason for medica-
tion, especially in case his disturbed and dangerous behavior will be experienced
mainly by others and, therefore, he rejects any medication. Nevertheless, the psy-
chiatrist has the task to inform the patient about the negative or even dangerous
consequences of his behavior and that it can be controlled by the medication.
However, if the patient refuses a medication due to previously experienced unwanted
drug effects or other rational reasons, the psychiatrist has to accept the rejection in
case the patient has the capacity of self-determination.
Chronic Conditions
Preventive Medication
been under discussion since the 2000s (Mcgorry 1998) especially due to ethical
questions with regard to the accuracy and validity of prediction: the probability of
transition from a clinically premorbid or mildly morbid state to a severe mental
disorder must be evaluated against the risk of unwanted drug effects in order to avoid
them particularly in patients who despite their high-risk state never will develop a
severe mental disorder. This will be explained by the examples of preventive
utilization of psychotropic drugs in early stages of schizophrenia and dementia.
Because of the early beginning in adolescence, the often progressive course, and the
perhaps lifelong duration of the disabling symptoms and lower functioning of
schizophrenia, it seems natural to start a treatment as early as possible. However,
the possible benefits of a preventive treatment, that is, the exclusion or postponing of
a transition from a high-risk state to a psychosis, or at least mitigation of disabling
symptoms, must be assessed against their possible risks (Klosterkoetter and
Schultze-Lutter 2010).
(a) Predictive intervention: The first step is to recognize people with the risk of
developing schizophrenia, for example, healthy persons with single psychotic
symptoms. A state-of-the-art review concluded that “clinical high-risk (CHR)
state for psychosis has evolved to capture the prepsychotic phase, describing
people presenting with potentially prodromal symptoms” (Fusar-Poli et al.
2013). According to fairly heterogeneous criteria of CHR and different follow-
up periods, a correct prediction of the manifestation of schizophrenia varies
between less than 20 and almost 90% (Ruhrmann et al. 2010), or 1/3 in 3 years
(Fusar-Poli et al. 2012), that is, almost 2/3 of people with high risk do not
develop a psychosis (Bechdolf et al. 2015). Even if the accuracy of prediction
will be further improved by combination of criteria for early and late clinical
high-risk stages and inclusion of biological and environmental risk factors
(Klosterkötter 2014), it may not reach 100%, at least on the individual level.
Therefore, early detection (ED) of healthy people at risk of schizophrenia
implies the risk of provoking unnecessary fears and stigmatization particularly
in those who never will get a manifest psychosis.
These psychological risks should be countered by starting such predictive
intervention as an offer of information and help in an individualized and
nonstigmatizing mode only for those seeking help (Bechdolf et al. 2015) who
are competent to consent after full information, particularly about the probabil-
ities of benefits and risks of preventive interventions (Helmchen 2014).
(b) Preventive intervention: The second step is to consider with regard to the
pharmacotherapeutic component of such preventive intervention not only pos-
sible unwanted drug effects but also the possibility that they may occur in
persons who would never have developed the psychosis even without any
746 H. Helmchen
Specialized services have been developed to translate the new findings into
practice, e.g. associated centers for early intervention and therapy for young adults
with beginning psychosis (HÄfner et al. 2012). They provide early detection (ED) as
well as early intervention (EI) by a comprehensive offer of treatment, particularly
cognitive-behavioral psychotherapy, support of the patients dealing with their ail-
ments, promotion of stigma coping and empowerment, and medication as one
component (Schmidt et al. 2015).
Controlled studies show an improvement of symptoms and functioning of the
high-risk state as well as a reduction of conversion rates, for example, from 1/3 to
1/10 (Mcgorry et al. 2002) for both cognitive behavioral therapy and psychophar-
macotherapy, but no significant differences between them. Therefore, the benefit-
risk-assessment of preventive medication has to consider on the one hand that
psychotherapeutic interventions are more acceptable by patients and less stigmatiz-
ing, have less unwanted effects, and are first choice treatments for anxiety and
depressive disorders that are frequent in these patients (Bechdolf et al. 2015; Michel
et al. 2018); on the other hand, psychotropic drugs will be effective more quickly
than psychotherapy and are indicated in cases of rapid progression of symptoms or
acute suicidality or strong social disability in order to make psychosocial interven-
tions possible (Schmidt et al. 2015; Lambert and Correll 2015). The risk of unwanted
effects (body weight gain, tardive hyperkinesia) of antipsychotic drugs such as
Risperidon, Amisulprid, Aripiprazol, Olanzapin should be avoided by low doses
and tight monitoring (Lambert and Correll 2015).
In general people with high-risk for psychosis should be accompanied and
empowered to deal with their ailments and stress and be offered risk-adapted
(Klosterkötter 2014) preventive interventions, but due to the unsatisfying state of
phase-specific preventive interventions with convincing and unequivocal efficacy
“caution is required when making clinical recommendations for the prevention of
psychosis in individuals at risk” (Fusar-Poli et al. 2019).
(a) Predictive intervention, that is, prognosis of the onset of dementia in a healthy or
at least nondemented person, must consider the expense and problems (risks of
taking body fluids or subjective discomfort by imaging procedures) of biologi-
cally based interventions. They may be necessary for research studies, but have
almost no or only very limited indication in practice and care. In any case the
concerned person must be competent to understand the comprehensive informa-
tion about the aim and risks and missing therapeutic consequences of this
predictive intervention. If a person is willing to defer to such predictive inter-
vention, he also should be informed about possibilities to write an advance
research directive for his willingness to participate in a future research study.
However, predictive intervention on a less invasive level is helpful for
counseling the patient. Less invasive are anamnestic (family and genetic) data
and findings of neuropsychological tests. Predicted high probability of clinical
dementia clearly indicates preventive interventions, as primary prevention on a
general level and as early secondary prevention on the individual level.
(b) Preventive interventions sensu interception with drugs are not available and
RCTs with new investigative drugs have to overcome methodological difficul-
ties (Rujescu and Maul 2019) and to apply specific ethical evaluations
(Helmchen 2017; Vollmann 2017; Winkler 2019).
Conclusion
The appraisal of this information and balancing according to the patient’s prefer-
ences may be difficult and needs assistance from the informing physician. Accord-
ingly his orientation towards the patient will be expressed by taking time enough for
listening and taking seriously the patient’s comments and questions, also giving time
for repeated questions.
Cross-References
References
Appelbaum PS, Redlich A. Impact of decisional capacity on the use of leverage to encourage
treatment adherence. Community Ment Health J. 2006;42:121–30.
750 H. Helmchen
Petra Netter
Contents
Definition of Placebos and Areas of Use . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 756
Factors Involved in Drug and Placebo Responses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 758
Influences of Attributes of Medication, Accompanying Information, and “Ambiente”
(Factors c and S) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 760
Factor “c” Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 760
Factor “S” Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 761
Target Variables of Placebo Responses and Their Methods of Assessment
(Factors R and m) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 761
Functions and Symptoms Responding to Placebo (Factor R) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 761
Methods of Assessment (Factor m) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 762
Factors Associated with Patient/Participant Variables (Factors Pi and Pj and Their
Interactions) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 762
Considerations on General Placebo Response Rates in Psychiatric Drug Studies . . . . . . . . . . . . 764
Mechanisms Operating in Placebo Responses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 765
Cognitive Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 765
Biological Mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 768
Use of Placebos for Therapy and the Role of Nocebo in Clinical Trials and
Medical Practice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 771
Nocebo Effects in Therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 772
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 773
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 773
Abstract
Clinical and experimental examples of placebo effects are presented. Placebo
effects including spontaneous remission are also part of response to active
treatment. Placebo effects result from complex interactions between accompany-
ing information about positive and negative (¼ nocebo) drug effects, character-
istics of the experimenter or physician, and experimental setting. They can be
P. Netter (*)
Department of Psychology, University of Giessen, Giessen, Germany
e-mail: petra.netter@psychol.uni-giessen.de
The original translation of the term placebo means “I will please.” The first time the
word placebo appears in a medical context was in a dictionary of the eighteenth
century where placebo was defined as “an unspecific method which is meant to
please a patient for a while,” but phenomena of healing after some kind of “treat-
ment” which modern placebo researchers would term placebo effects have been
reported across many previous centuries (De Craen et al. 1999; Finniss 2018).
A later more specific definition given by Shapiro (1971) as “a therapeutic device
applied by a physician intentionally or without his information, exerting an effect on
a patient or a symptom, but without an objective effect” according to present
knowledge is not valid either anymore, since:
• Placebos are not only applied by physicians in patients or diseases but refer to any
kind of effects on emotional, performance, or social activity also in healthy
persons.
• Placebo is not only related to specific actions but may be effective with respect to
more global well-being as assessed after recreational or physiotherapy.
• Placebo does show objective biological effects in the autonomic as well as in the
central nervous system.
• in therapeutic contexts.
• as control conditions in clinical trials.
• as a device for elucidating underlying factors of placebo effects in research.
In the present chapter, the first two aspects will only be briefly touched, since they
may be partly covered by the chapters ▶ “Randomized Controlled Trials and the
Efficacy of Psychotropic Medications” and “Significance of and Attitude to Psy-
chopharmacological Agents.” Instead, the focus will be on the third aspect from a
more experimental approach. Since the factors of influence and mechanisms of
placebo and nocebo are principally the same, only examples for placebo will be
addressed. However, after briefly discussing aspects of placebo effects also in
clinical practice, a respective section on the role of nocebo effects in therapy will
be added.
Although placebo effects are also operating on non-pharmacological types of
therapy, this contribution will primarily illustrate aspects of placebos used in drug
studies.
The responses to a placebo do not only imitate the size of the active drug effect
(e.g., on reduction of pain) but also the inverted u-shape course of its effects
in observation periods across time after application (rise to a maximum and
consecutive decline), accumulation effects after consecutive drug applications, or
number of applications required according to severity of symptoms (Lasagna et al.
1958).
It has to be considered that a placebo response is also part of active drug responses
as nowadays emphasized in reviews on placebo (e.g., Colloca 2018; Evers et al.
2018). This is depicted in Fig. 1.
In both conditions, baseline levels of the outcome variable as well as elapse of
time (¼ “natural history,” spontaneous remission) largely contribute to therapeutic
effects. A meta-analysis on 37 controlled clinical trials on patients suffering from
different diseases of minor severity (phobia, insomnia, nausea, acute and chronic
pain) compared a placebo, a no treatment, and a treatment group represented in each
of the 37 trials with respect to standardized symptom reduction (Krogsbøll et al.
2009). The overall result across different psychological, physical, and pharmacolog-
ical treatments and across different diseases revealed that symptom reduction in the
no treatment group (representing the part ET of Fig. 1) was about half the size of the
placebo effect (part P of the figure) and that the relative contributions of spontaneous
improvement and of the placebo effect to that of the active interventions were 24%
and 20%, respectively.
758 P. Netter
Fig. 1 Composition of drug effects in active compounds and placebos. (Modified from
Klebelsberg (1974)
Part ET also comprises statistical bias like regression to the mean (the tendency of
a repeated measurement to be shifted to the population mean). Placebo effects (part
P), i.e., subjective factors involved in judgment (see later), are also adding to the true
pharmacodynamic effects of the active drug to a considerable degree. This aspect is
often neglected in controlled clinical trials as well as in clinical practice. A way to
subtract the placebo effect from the outcome of an active drug is, for instance,
achieved by hidden application (infusion) of the drug. The reported relief from pain
after open administration of an analgesic was considerably reduced when the drug
was infused (Amanzio et al. 2001). In order to judge the pure net effect of a placebo,
a control condition has to be run which only consists of a baseline and a repeated
measurement obtained at the same time point as assessment of the placebo effect.
Placebo and active treatment responses measured within the same individuals are
often correlated because the placebo parts as well as the baseline levels of a variable
in the treatment and the placebo group are correlated.
Fig. 2 Schematic representation of factors involved in drug and placebo responses (legend see
text)
In the beginning of placebo research, it has been reported that color, shape, and
taste as well as mode of application of a placebo are relevant for the outcome, for
example, blue pills seemed to induce more sedative, red ones rather stimulating
effects, green ones were effective against states of anxiety, yellow ones against
depression, fluids were better in case of sleeplessness than solid pills, and the
number of pills showed a dose response curve (Buckalew and Ross 1981).
More important is the large number of publications concerned with effects of the
accompanying information provided along with drug application in clinical set-
tings as well as in drug experiments (Benedetti et al. 2003; Kaptchuk and Miller
2015); this concerns:
In clinical settings, the placebo declared as the active drug (D) and the active drug
declared as the placebo (P) had similar effects on pain and attacks of migraine in the
conditions D and P, whereas the condition, in which the active drug was combined
with the information that it was a powerful analgesic, elicited significantly higher
success rates in pain reduction than the application without any concomitant infor-
mation or compared with conditions D and P (Colloca et al. 2004).
Similarly, in an experiment testing expected stimulating effects of caffeine on
performance in healthy participants, the placebo announced as caffeine elicited even
higher increases in mental performance than the reverse condition of an active
compound declared as a placebo (Lienert 1956).
Experiments declaring a placebo either as a tranquilizer or as a stimulant indicated
that a stimulating effect was experienced in 78% of 68 heathy participants and
tranquilizing effects in 38% (Janke 1967, unpublished). Comparisons of effects on
arousing versus relaxing emotional states revealed that changes after the “sedative”
were more pronounced than after the “stimulant” and that increases of certain states
like calm, relaxed for the “sedative” and alert, talkative for the “stimulant” were
more readily reported than decreases of states like feeling less anxious and tense
with the “stimulant” and less tired or drowsy for the “sedative” (Netter 1977).
Finally, information about possible side effects, explained to patients or participants
(nocebo effects), have been reported mostly in the predicted direction of the active
drug, mostly less frequently or as less intensive. Occasionally, when side effects are
assessed as open questions, e.g., in a study on effects of indomethacin on relief from
pain in rheumatoid arthritis, side effects of vomiting, diarrhea, and edemas were even
reported at a higher rate in the placebo than in the indomethacin group (Schindel
1967).
Neuropsychopharmacotherapy: Aspects of Placebo Effects 761
With regard to target areas of response (factor R), emotional states like anxiety,
relaxation, or arousal as well as perception of pain are more susceptible to placebo
effects than physiological responses of the autonomic nervous system like heart rate,
blood pressure, or sweating and than psychomotor behavior and mental perfor-
mance. There are also reports of changes of the central nervous system (EEG
activity) indicating onset of slow wave sleep upon administration of placebo
sleeping pills. But these somatic changes, invisible to the individual, and consider-
ably less pronounced, are often accompanied by improvement of subjective well-
being. In all cases of somatic changes, the accompanying information is essential
and may exert very specific effects, as demonstrated by an experiment in which an
“antihypertensive” placebo distinctly lowered systolic blood pressure, but neither
diastolic blood pressure nor heart rate, heart rate variability, skin conductance, or
EEG activity were affected (Meissner and Ziep 2011).
In the clinical setting, symptoms of almost all diseases have been shown to
respond to placebos (Schedlowski et al. 2015). When ranking diseases according
to susceptibility to placebo effects, any disease related to pain perception like
headaches, migraine, rheumatism, gastrointestinal pain, dysmenorrhea, general
pains of limbs and back during flu, and fibromyalgia are ranging highest, followed
by psychiatric diseases (in particular anxiety and depression-related disorders and
addiction) which again are more “placebo-prone” than neurological diseases, like
motor disturbances as in Parkinson’s disease. These again are influenced to a greater
extent than allergies or infections and parameters of the endocrine, immunological,
or metabolic system. Yet, several studies demonstrate that not only responses of
762 P. Netter
heart rate or bowel activity, but also immune cell responses may be modified by
previous treatment or physician-induced expectations (Hadamitzky et al. 2018;
Schedlowski et al. 2015).
(older men and younger women reporting better results (Interaction Pi Pj) (Raskin
1974)), it could be concluded in a large meta-analysis of more than 500 placebo-
controlled clinical trials derived from 31 meta-analyses on placebo effects in psy-
chiatric disorders that age and gender could not be identified to be good general
predictors of placebo responsiveness, in spite of being relevant in drug- or disease-
specific studies (Weimer et al. 2015). Also drug taking history and attitudes toward
drugs were identified as predictors only in some clinical or experimental studies,
depending on the type of drug or disease. A certain universal finding correlated to
placebo responsiveness was low severity of symptoms at baseline (component B in
Fig. 1). This did not only refer to neuropsychiatric disorders but also to gastrointes-
tinal, cardiovascular, and allergic diseases (Weimer et al. 2015).
The modifying influences of individual differences (Pi, Pj) also depend on the
type of target symptoms or functions (Rk, Rl) and mode of assessment (m) which
again are differently affected by factors c and S as shown by an experiment which
tested the influences of different personality factors on the effects of a “stimulant
placebo” with respect to subjective and objective indicators of placebo-induced
changes in activation (Janke 1967, unpublished). The personality factor of neurot-
icism was positively correlated with an increase of activation rated on a checklist of
emotional states but negatively with speed of performance achieved in a psychomo-
tor test, while in extraverts this relationship was reversed.
Similar individual differences appeared when comparing respective patterns
following the information about a “sedative placebo” (c P R m).
This means the expectation induced by the accompanying information affected
feelings and psychomotor performance differently depending on basic temperament
traits. Statistically speaking this means that with a given information provided along
with the placebo (type c variable), an interaction was observed between the person-
ality factors (Pi and Pj) and the dependent variable (Rk, Rl) and its mode of
assessment (mk ml).
Induction of expectations, which is a major factor in placebo effects, as demon-
strated above, is also a leading component in the trait of suggestibility (De Pascalis
et al. 2002; Netter 1989; Nitzan et al. 2015). It can be measured by a behavioral
device based on detection of nonverbal faked sensory stimuli (Gheorghiu et al. 1975;
Gheorghiu 1989) and has been shown to largely influence pain perception, for
instance, in patients suffering from headaches treated with metamizole or placebo,
respectively (Fig. 3).
It may be concluded that, as indicated also in several similar experiments, the
predictive power of the personality variable for placebo responsiveness (in this case
sensory suggestibility) depends on its similarity or closeness to the target variable
(sensory experience of pain reduction).
Defining “placebo responders” by special characteristics of the individual has
been tried for several decades in order to exclude placebo responders from controlled
clinical trials for increasing “true” drug effects. However, defining a placebo
responder as a general trait requires (a) consistence across time, that is, reproduc-
ibility of results upon consecutive trials (t1. . .tn) and (b) generalizability across
different drugs, drug informations, symptoms, or diseases assessed by different
764 P. Netter
15 Influence of personality
factor suggestibility
11
3
= low suggestible patients
= highly suggestible patients
1
tools of measurement. That means: (a) will the placebo effect be detectable across
repeated trials? and (b) will a person responding, for instance, positively to reduction
of dental pain by an “analgesic” placebo also respond positively to a “sleeping pill”
placebo? Will this work in case he or she is participating in a controlled clinical trial
or being treated by a clinical practitioner? Will this depend on whether or not the
accompanying information was about positive effects or side effects?
It is obvious that this degree of generalizability cannot be achieved, so the
concept of a general placebo responder was already abandoned in the 1980s
(Brown 1988) and is still critically discussed (Kaptchuk et al. 2008). Not even
stability across time with the same drug, the same information, experimenter, and
clinical setting can be completely achieved in consecutive trials as shown by an old
experiment on frequent migraine attacks (Jellinek 1946). The reason is that in
particular in drug studies habituation occurs due to downregulation of receptors,
learning and adjusting expectations according to preceding experience. This is also
confirmed by declining placebo response rates during longitudinal observations
(Posternak and Zimmerman 2007).
differences between response rates of antidepressants and placebos and that severity
of symptoms at baseline may be a source of bias artificially increasing active drug
response rates (Kirsch et al. 2008). However, authors of later meta-analyses and
consensus papers (Baghai et al. 2012; Möller et al. 2012; Melander et al. 2008) ruled
out the bias of differences in baseline severity of depression and criticized that
average differences in means of response rates between placebo and active treatment
are not the correct parameter for treatment success. Rather, a number of patients with
a critical reduction of symptom severity (usually defined by 2 or 3 points on accepted
depression rating scales) and numbers of patients needed to treat (NNT, Cook and
Sackett 1995) were considered as more relevant parameters (this is the inverted value
of the difference between percentages of patients showing responses in the active
treatment versus in the placebo group, i.e., the higher the value, the more patients are
needed to be treated in order to obtain one true response to the active drug). The
average placebo response rates vary with length of observation period, types of
antidepressants included, and homogeneity of samples, but summaries for differ-
ences between placebo and active drug response rates obtained in several meta-
analyses of controlled clinical trials have been reported to be between 13.5% and
19.5% (Baghai et al. 2012), and the average NNT values across different antide-
pressants range between 5 and 7 for acute treatment of depression and 4 for
maintenance treatment (Möller et al. 2012).
Since it was claimed that placebo response rates had been rising since the 1970s, a
critical meta-analysis by Furukawa et al. (2016) investigated the influence of pub-
lication year on placebo response rates in 252 placebo-controlled clinical trials on
more than 26,000 patients. Studies include first and second generation of antide-
pressants applied for acute treatment of major depression between 1978 and 2015.
Responses were defined by log transformed placebo responses based on 50% or
more reduction of depression severity assessed by valid depression scales. It became
evident that after 1991 the average placebo response rates have remained constant
between 35% and 40%.
However, these gross figures require detailed inspection of placebo response rates
separately analyzed according to their interactions with factors discussed above.
Cognitive Functions
Expectancy
It has long been known that expectation of success is correlated with drug-induced
increase of performance as well as with subjectively experienced improvement, as
766 P. Netter
Learning
The second approach is based on learning theory. However, learning from previous
experience is supposed to be mediated by expectations (Price et al. 1999). A
previously taken active drug serves as the unconditioned stimulus; attributes of the
drug, accompanying information, and surrounding environment or behavior of the
physician as described above (factors c and S) serve as conditioned stimuli which
transfer the unconditioned response to the active drug (relief from symptoms) into
the corresponding conditioned response, when receiving the placebo.
This process of classical conditioning is supported by an experiment on pain
perception by Voudouris et al. (1985) in which pain intensity levels were experi-
mentally manipulated: in the first session, participants received two trials, one with
and one without an analgesic placebo with identical objective stimulus intensities of
5. In the second session, the stimulus intensity without the placebo was increased to
8 in one group and decreased to 2 in the second group, while in both groups the
stimulus intensity with the placebo was kept constant at 5. So the first group
experienced a positive placebo effect and the second group an increase of pain
(inefficiency of placebo or even a nocebo effect). After this conditioning procedure,
a third session, identical to the first, was applied to both groups, showing that group
one experienced a significant pain reduction with placebo compared to no drug,
while the second group judged pain intensity after the placebo as even stronger than
in the first trial (see Fig. 4).
There was a long fight between placebo researchers whether expectation or
learning is the dominating process. A possibility for decision was achieved by
adding experimental sessions in which (a) the deception in session 2 (experimental
manipulation of the pain stimulus) was disclosed to participants or not and
(b) extinction of the conditioned placebo responses was tested (Montgomery and
Kirsch 1997).
Conditioning theory was supported if (a) disclosure information had no effect and
(b) extinction of the conditioned placebo response upon repeated trials was possible,
while expectation theory was supported if disclosing the stimulus manipulation
(cognitive information processing) reduced the placebo response and (b) extinction
of the placebo response upon repeated trials was not possible (since the uncondi-
tioned stimulus keeps reinforcing expectations).
Classical conditioning of placebo effects also takes place in experiments not
related to pain but, for instance, in experimental induction of anxiety: in an exper-
iment in which aversiveness ratings of threatening pictures (from the standardized
Neuropsychopharmacotherapy: Aspects of Placebo Effects 767
3,0 Conditioning
difference between pain ratings in
the condition with minus without
2nd session:
2,5 stimulus
intensity =5
2,0 with and
intensity = 8
the placebo
without the
1,5 placebo
= positive
1,0 placebo effect
2nd session:
0,5 stimulus
intensity =5
with and
0,0
intensity = 3
1st session: stimulus 3rd session: ( after without the
placebo
intensity with and without conditioning) stimulus = negative
the placebo analgesic intensity with and withourt placebo effect
identical ( = 5) the placebo analgesic
identical (= 5)
Biological Mechanisms
Not only investigators of clinical trials but also doctors in primary care or medical
specialists have to outline possible side effects of drugs and non-pharmacological
therapies to patients which can cause nocebo effects by inducing negative expecta-
tions in patients. The package inserts of pharmaceuticals even list frequency prob-
abilities of each type of side effect. This often leads to discontinuation of drug use in
patients and to denial of participation or dropout in drug studies. Nocebo effects not
only occur by intentionally provided information about therapeutic devices (see
“Examples of Factors “c” with Placebo Effects”) but also by the way a doctor
phrases accompanying comments before or during therapy, for instance, by giving
too much reassurance, by emphasizing the low rate of detrimental effects of the
therapy, by advising strictly to adhere to the therapeutic regimen, by asking if
patients feel nausea or headaches after therapy, or by using technical terms for
Neuropsychopharmacotherapy: Aspects of Placebo Effects 773
Cross-References
References
Amanzio M, Benedetti F. Neuropharmacological dissection of placebo analgesia: expectation-
activated opioid systems versus conditioning-activated specific subsystems. J Neurosci.
1999;19(1):484–94.
Amanzio M, Pollo A, Maggi G, Benedetti F. Response variability to analgesics: a role for
non-specific activation of endogenous opioids. Pain. 2001;90(3):205–15.
Baghai TC, Blier P, Baldwin DS, Bauer M, Goodwin GM, Fountoulakis KN, . . . Möller
HJ. Executive summary of the report by the WPA section on pharmacopsychiatry on general
and comparative efficacy and effectiveness of antidepressants in the acute treatment of depres-
sive disorders. Eur Arch Psychiatry Clin Neurosci. 2012;262(1):13–22.
Benedetti F. Placebo effects. 3rd ed. Oxford: Oxford University Press; 2020.
Benedetti F, Amanzio M. Mechanisms of the placebo response. Pulm Pharmacol Ther. 2013;26(5):
520–3.
Benedetti F, Maggi G, Lopiano L, Lanotte M, Rainero I, Vighetti S, Pollo A. Open versus hidden
medical treatments: the patient’s knowledge about a therapy affects the therapy outcome. Prev
Treat. 2003;6(1):1–18.
Benedetti F, Amanzio M, Rosato R, Blanchard C. Non-opioid placebo analgesia is mediated by
CB1 cannabinoid receptors. Nat Med. 2011a;17(10):1228–30.
Benedetti F, Amanzio M, Thoen W. Disruption of opioid-induced placebo responses by activation
of cholecystokinin type-2 receptors. Psychopharmacology. 2011b;213(4):791–7.
Berridge KC, Robinson TE. What is the role of dopamine in reward: hedonic impact, reward
learning, or incentive salience? Brain Res Rev. 1998;28(3):309–36.
774 P. Netter
Kelley JM, Lembo AJ, Ablon JS, Villanueva JJ, Conboy LA, Levy R, . . . Kaptchuk TJ. Patient and
practitioner influences on the placebo effect in irritable bowel syndrome. Psychosom Med.
2009;71(7):789. definition 18(4), 318–322.
Kirsch I, Deacon BJ, Huedo-Medina TB, Scoboria A, Moore TJ, Johnson BT. Initial severity and
antidepressant benefits: a meta-analysis of data submitted to the Food and Drug Administration.
PLoS Med. 2008;5(2):e45. https://doi.org/10.1371/journal.pmed.0050045
Klebelsberg D. On the psychology of the placebo effect. Psychol Beitrage. 1974;16(2):168–87.
Krogsbøll LT, Hróbjartsson A, Gøtzsche PC. Spontaneous improvement in randomised clinical
trials: meta-analysis of three-armed trials comparing no treatment, placebo and active interven-
tion. BMC Med Res Methodol. 2009;9(1):1–10.
Lasagna L, Laties VG, Dohan JL. Further studies on the “pharmacology” of placebo administration.
J Clin Invest. 1958;37(4):533–7.
Levine JD, Gordon NC, Jones RT, Fields HL. The narcotic antagonist naloxone enhances clinical
pain. Nature. 1978;272(5656):826–7.
Lidstone SC, Schulzer M, Dinelle K, Mak E, Sossi V, Ruth TJ, . . . Stoessl AJ. Effects of expectation
on placebo-induced dopamine release in Parkinson disease. Arch Gen Psychiatry. 2010;67(8):
857–65.
Lienert GA. Die Bedeutung der Suggestion in pharmakopsychologischen Untersuchungen. Z Exp
Angew Psychol. 1956;3:418–38.
Meissner K, Ziep D. Organ-specificity of placebo effects on blood pressure. Auton Neurosci.
2011;164(1–2):62–6.
Melander H, Salmonson T, Abadie E, van Zwieten-Boot B. A regulatory Apologia – a review of
placebo-controlled studies in regulatory submissions of new-generation antidepressants. Eur
Neuropsychopharmacol. 2008;18(9):623–7.
Möller HJ, Bitter I, Bobes J, Fountoulakis K, Höschl C, Kasper S. Position statement of the
European Psychiatric Association (EPA) on the value of antidepressants in the treatment of
unipolar depression. Eur Psychiatry. 2012;27(2):114–28.
Montgomery GH, Kirsch I. Classical conditioning and the placebo effect. Pain. 1997;72(1–2):107–13.
Nash MM, Zimring FM. Prediction of reaction to placebo. J Abnorm Psychol. 1969;74(5):568.
Netter P. The placebo effect (author’s transl). MMW Munch Med Wochenschr. 1977;119(7):203–8.
Netter P. Sensory suggestibility: measurement, individual differences, and relation to placebo
and drug effects. In: Suggestion and suggestibility. Berlin/Heidelberg: Springer; 1989.
p. 123–33.
Netter P, Heck S, Müller H. What selection of patients is achieved by requesting informed consent
in placebo controlled drug trials? Pharmacopsychiatry. 1986;19:336–7.
Nitzan U, Chalamish Y, Krieger I, Erez HB, Braw Y, Lichtenberg P. Suggestibility as a predictor of
response to antidepressants: a preliminary prospective trial. J Affect Disord. 2015;185:8–11.
Peciña M, Bohnert AS, Sikora M, Avery ET, Langenecker SA, Mickey BJ, Zubieta JK. Association
between placebo-activated neural systems and antidepressant responses: neurochemistry of
placebo effects in major depression. JAMA Psychiat. 2015;72(11):1087–94.
Petrovic P, Dietrich T, Fransson P, Andersson J, Carlsson K, Ingvar M. Placebo in emotional
processing – induced expectations of anxiety relief activate a generalized modulatory network.
Neuron. 2005;46(6):957–69.
Posternak MA, Zimmerman M. Therapeutic effect of follow-up assessments on antidepressant and
placebo response rates in antidepressant efficacy trials: meta-analysis. Br J Psychiatry. 2007;190
(4):287–92.
Price DD, Milling LS, Kirsch I, Duff A, Montgomery GH, Nicholls SS. An analysis of factors that
contribute to the magnitude of placebo analgesia in an experimental paradigm. Pain. 1999;83(2):
147–56.
Raskin A. Age-sex differences in response to antidepressant drugs. J Nerv Ment Dis. 1974;159:
120–30.
Reyna VF, Nelson WL, Han PK, Dieckmann NF. How numeracy influences risk comprehension
and medical decision making. Psychol Bull. 2009;135(6):94.
776 P. Netter
Rief W, Nestoriuc Y, Weiss S, Welzel E, Barsky AJ, Hofmann SG. Meta-analysis of the placebo
response in antidepressant trials. J Affect Disord. 2009;118(1–3):1–8.
Rollman GB. Signal detection theory measurement of pain: a review and critique. Pain. 1977;3(3):
187–211.
Schedlowski M, Enck P, Rief W, Bingel U. Neuro-bio-behavioral mechanisms of placebo and
nocebo responses: implications for clinical trials and clinical practice. Pharmacol Rev. 2015;67
(3):697–730.
Schindel L. Placebo und Placeboeffekte in Klinik und Forschung, [Placebos and placebo effects in
clinical practice and research]. Arzneim Forsch. 1967;17:892–917.
Shapiro AK. Placebo effects in medicine, psychotherapy, and psychoanalysis. In: Handbook of
psychotherapy and behavior change: empirical analysis. Wiley Hoboken, NJ; 1971. p. 439–73.
Skyt I, Moslemi K, Baastrup C, Grosen K, Benedetti F, Petersen GL, . . . Vase L. Dopaminergic tone
does not influence pain levels during placebo interventions in patients with chronic neuropathic
pain. Pain. 2018;159(2):261–72.
Uhlenhuth EH, Rickels K, Fisher S, Park LC, Lipman RS, Mock J. Drug, doctor’s verbal attitude
and clinic setting in the symptomatic response to pharmacotherapy. Psychopharmacologia.
1966;9(5):392–418.
Voudouris NJ, Peck CL, Coleman G. Conditioned placebo responses. J Pers Soc Psychol. 1985;48
(1):47.
Wager TD, Rilling JK, Smith EE, Sokolik A, Casey KL, Davidson RJ, . . . Cohen JD. Placebo-
induced changes in FMRI in the anticipation and experience of pain. Science. 2004;303(5661):
1162–7.
Weimer K, Colloca L, Enck P. Placebo effects in psychiatry: mediators and moderators. Lancet
Psychiatry. 2015;2(3):246.
Zubieta JK, Stohler CS. Neurobiological mechanisms of placebo responses. Ann N Y Acad Sci.
2009;1156:198.
Neuropsychopharmacotherapy:
International Regulations and Regulating
Laws in Japan
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 778
Single Convention on Narcotic Drugs, 1961 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 778
Convention on Psychotropic Substances, 1971 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 779
Convention Against Illicit Traffic in Narcotic Drugs and Psychotropic Substances, 1988 . . . 781
The Laws Regulating Drug Control in Japan . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 781
Some Specific Topics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 786
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 787
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 787
Abstract
Psychopharmaceuticals contain various drugs of which chemical structures and
pharmacological effects are different. Each substance varies in usefulness for
medication purposes as well as risk of abuse. Various substance abuse counter-
measures have thus been implemented internationally and within each country. In
this chapter, the three main International Conventions for controlling and regu-
lating narcotics and psychotropics are outlined: the Single Convention on Nar-
cotic Drugs of 1961, the Convention on Psychotropic Drugs of 1971, and the
Convention against Illicit Traffic in Narcotic Drugs and Psychotropic Substances
of 1988. The history on the abuse of narcotics and psychotropics in Japan is
explained, and their countermeasures including the four main domestic laws are
then presented: the Stimulants Control Act, the Narcotics and Psychotropic Drugs
Control Act, the Cannabis Control Act, and the Opium Control Act. Finally,
specific topics including the issues on benzodiazepines including etizolam and on
synthetic cannabinoids in Japan are described.
Introduction
The first international regulation on psychoactive drugs was the 1912 Hague Interna-
tional Opium Convention, which consisted of 6 chapters and 25 articles. In addition to
opium and morphine, which were already under extensive international discussion, the
Hague Convention also included two new substances: cocaine and heroin [United
Nations Office of Drugs and Crime (UNODC) 2020]. Although loopholes existed
within the Hague Convention, it contained many elements of a comprehensive drug
control treaty. The 1925 International Opium Convention in Geneva was intended to
impose global control over a wider range of drugs, including, for the first time,
cannabis, which was referred to as “Indian hemp” (marijuana) in Article 11 of the
Convention (The Senate Special Committee on Illegal Drugs, Canada 2002). Several
treaties were signed thereafter. In 1961, the previous Conventions were integrated and
signed as the Single Convention on Narcotic Drugs and came into force in 1964. As of
July 2020, the Single Convention has 186 state Parties.
The preamble of the Single Convention on Narcotic Drugs, 1961, describes the
treaty as follows: “Recognizing that the medical use of narcotic drugs continues to be
indispensable for the relief of pain and suffering and that adequate provision must be
made to ensure the availability of narcotic drugs for such purposes, and recognizing
that addiction to narcotic drugs constitutes a serious evil for the individual and is
fraught with social and economic danger to mankind,” and based on “the compe-
tence of the United Nations in the field of narcotics control and desirous that the
international organs concerned should be within the framework of that Organiza-
tion,” “a generally acceptable international convention replacing existing treaties on
narcotic drugs, limiting such drugs to medical and scientific use, and providing for
continuous international co-operation and control for the achievement of such aims
and objectives” (UNODC 2013).
Neuropsychopharmacotherapy: International Regulations and Regulating Laws. . . 779
In Article 2, substances under legal control are categorized into four groups as
follows, from most restrictive to least restrictive: Schedule IV, Schedule I, Schedule
II, and Schedule III. Schedule I lists 124 substances including cannabis, carfentanyl,
heroin, morphine, and opium. Schedule II designates ten substances including
codeine and its derivatives. In Schedule III, lighter subsets of substances already
categorized in Substance II are listed, for when they are included as a minor
constituent of another medication. Schedule IV lists 20 drugs which are already
listed under Schedule I, but are particularly liable to abuse and produce adverse
effects: acetylfentanyl, cannabis, carfentanyl, and heroin [International Narcotics
Control Board (INCB) 2019a].
The main contents of the Single Convention are summarized as follows. Because
the medical use of these controlled substances is important, Articles 1, 2, 4, 9, 12, 19,
and 49 contain specifications for medical and scientific use of the drugs. Article
33 prohibits the possession of drugs except under legal authority. Article 36 stipulates
penal provisions; it requests that Parties adopt measures against “cultivation, pro-
duction, manufacture, extraction, preparation, possession, offering, offering for sale,
distribution, purchase, sale, delivery on any terms whatsoever, brokerage, dispatch,
dispatch in transit, transport, importation and exportation of drugs contrary to the
provisions of this Convention,” and “intentional participation in, conspiracy to
commit and attempts to commit, any of such offences, and preparatory acts and
financial operations in connexion with the offences referred to in this article”
(UNODC 2013). Article 38 stipulates measures against the abuse of drugs, including
treatment, after-care, rehabilitation, and social integration of drug addicts (UNODC
2013). Article 39 requests Parties to adopt measures of control “more strict or
severe” than those provided by the Convention.
for such purposes should not be unduly restricted, Believing that effective measures
against abuse of such substances require co-ordination and universal action,
Acknowledging the competence of the United Nations in the field of control of
psychotropic substances and desirous that the international organs concerned should
be within the framework of that Organization, Recognizing that an international
convention is necessary to achieve these purposes,” agree to the articles outlined in
the Convention (UNODC 2013).
Article 1 defines the terms used in the Convention. Article 2 restricts the sub-
stances that should be controlled under the Convention and need to fulfill the
following items: (a) That the substances have the capacity to produce (i) (1) a state
of dependence and (2) CNS stimulation or depression, resulting in hallucinations or
disturbances in motor function or thinking or behavior or perception or mood or
(ii) similar abuse and similar ill effects as a substance in Schedule I, II, III, or IV, and
(b) that there is sufficient evidence that the substance is being or is likely to be
abused so as to constitute a public health and social problem warranting the placing
of the substance under international control (UNODC 2013). In 2019, 78 substances
including lysergic acid diethylamide (LSD), 3,4-methylenedioxymethamphetamine
(MDMA), and the isomers of tetrahydrocannabinol, posing serious risk to the public
health while not having therapeutic value, were designated in Schedule I; 46 sub-
stances including amphetamine, gamma-hydroxyl butylic acid (GHB), methamphet-
amine, phencyclidine, secobarbital, and some synthetic cannabinoids were
designated in Schedule II, as having serious risk but also having some therapeutic
value; 9 substances including amobarbital, buprenorphine, flunitrazepam, and pen-
tobarbital having moderate risk and average therapeutic value were designated in
Schedule III; 62 substances including alprazolam, barbital, diazepam, mazindol,
phenobaribital, triazolam, and zolpidem having high therapeutic value and some
possibility of abuse were designated in Schedule IV (INCB 2019b).
The main contents of the Convention are summarized as follows. The usage of the
designated drugs is limited to medical and scientific purposes. Licensing or other
similar control measure is required for the manufacture of, trade in, and distribution
of the substances listed in Schedules II, III, and IV. The substances in Schedules II,
III, and IV need to be supplied or dispensed for use by individuals pursuant to
medical prescription only. Article 11 lists the guidelines for all Parties with regard to
recordkeeping of the substances in respective Schedules, such as the details of the
manufactured quantity, acquisition and disposal, and supplier–recipient information.
Article 13 provides descriptions regarding prohibition of and restriction on export/
import of the substances. Article 15 stipulates an inspection system of manufacturers
and all other bodies that use the substances for medical and scientific purposes. In
Article 16, the Parties are instructed to furnish their reports, such as annual reports
and any incident of illicit traffic, to the Secretary-General of the UN. Article
20 specifies measures against the abuse of psychotropic substances, including
preventative measures, personnel training, and assistance for the general public to
gain widespread understanding on the abuse of psychotropic substances (Kato
1999).
Neuropsychopharmacotherapy: International Regulations and Regulating Laws. . . 781
Since the Convention on Psychotropic Substances was ratified, the abuse of nar-
cotics or psychotropic drugs was rising in developed countries. In 1984, the General
Assembly of the UN requested the Commission on Narcotic Drugs to initiate
preparation of a draft convention against illicit traffic in narcotic drugs; the Conven-
tion against Illicit Traffic in Narcotic Drugs and Psychotropic Substances, 1988, was
then adopted on December 20, 1988. The purpose of this Convention is to promote
co-operation among the Parties for dealing with illicit trafficking in narcotic drugs
and psychotropic substances. As of July 2020, the Convention has 191 Parties.
Some of the statements in the permeable include the following:“(r)ecognizing
also that illicit traffic is an international criminal activity, the suppression of which
demands urgent attention and the highest priority,” “(d)etermined to deprive persons
engaged in illicit traffic of the proceeds of their criminal activities and thereby
eliminate their main incentive for so doing,” “(r)ecognizing the need to reinforce
and supplement the measures provided in the Single Convention on Narcotic Drugs,
1961, that Convention as amended by the 1972 Protocol Amending the Single
Convention on Narcotic Drugs, 1961, and the 1971 Convention on Psychotropic
Substances, in order to counter the magnitude and extent of illicit traffic and its grave
consequences,”and claim the necessity of “a comprehensive, effective and operative
international convention that is directed specifically against illicit traffic and that
considers the various aspects of the problem as a whole, in particular those aspects
not envisaged in the existing treaties in the field of narcotic drugs and psychotropic
substances” (UNODC 2013).
Article 3 stipulates the offenses to which countermeasures should be taken and
the sanctions for the offenses. Article 5 requires that the proceeds from drug offenses
as well as the drugs and any other equipment intended for the offenses be confis-
cated. Article 11 stipulates that controlled delivery be made by the Parties at the
international level, taking necessary measures and arranging mutual agreements in
view of identifying the persons involved in the drug offenses. Article 12 provides the
substances frequently used in the illicit market and manufacturers of narcotic drugs
and psychotropic substances; the International Narcotics Control Board designates
18 compounds including acetic anhydride, ephedrine, lysergic acid, norephedrine,
and pseudoephedrine in Table I, and 8 compounds including acetone, ethyl ether,
piperidine, and toluene in Table II, as controlled drug precursors (INCB 2018).
1
Note: Many of the translated terms used in this manuscript are mainly taken from the “Japanese
Law Translation” Website of the Ministry of Justice, Japan (http://www.japaneselawtranslation.go.
jp/); however, these terms are not officially authorized. Researchers used different translation terms:
“Stimulant Control Act” or “Stimulant Control Law.” In this chapter, we basically adopted the
translation terms which are currently effective based on “Japanese Law Translation.”
2
In 2015, the Pharmaceutical Act was abolished, and the Act on Securing Quality, Efficacy and
Safety of Products Including Pharmaceuticals and Medical Devices newly came into effect.
Neuropsychopharmacotherapy: International Regulations and Regulating Laws. . . 783
60,000
Stimulants Control Act
50,000
40,000
30,000
20,000
10,000
-
1951
1954
1957
1960
1963
1966
1969
1976
1979
1982
1985
1988
1994
1997
2000
2003
2006
2009
2012
2015
2018
1972
1991
Fig. 1 The number of arrests for violation of the Stimulants Control Law. Based on the data of
White Paper on Crime 1997 and 2019 (Ministry of Justice 1997, 2019a, b)
cultivation and possession of opium were prohibited in 1840, under the influence of
the Opium War in 1940 (Nagahama 1968). In the Meiji Era, Japan continued to
control opium and participated in the International Opium Convention signed at The
Hague in 1912 and the International Opium Convention signed at Geneva in 1925;
the government promulgated the Narcotics Control Regulations in 1930. After
World War II, crimes related to narcotics increased; the number of offenses reached
to approximately 2000 in 1960 (Nagahama 1968, Ministry of Justice 1982). Heroin
abuse especially exploded in urban areas (the miserable situation of heroin abusers in
the Yokohama District was vividly depicted in Akira Kurosawa’s movie, High and
Low). Although the Narcotics Control Act had already been enacted in 1953 and was
amended several times, the laws did not curb the narcotics-abuse-related offenses. In
1962, the government amended the Narcotics Control Act, which consisted of three
pillars. First, penalties against offenders were strengthened: maximum penalty was
extended from 7 years to life imprisonment for heroin, and the maximum fine was
raised from 500,000 yen to 5,000,000 yen (equivalent to USD 1390–13,900 in
1962). Second, a compulsory hospitalization system was established. Third, licenses
for narcotics uses became restricted to only medical and academic purposes. The
number of offenders plummeted from 2571 in 1963 to 792 in 1964; the offenders
decreased to less than 500 in the late 1960s.
From the early 1960s, abuse of “sleeping drugs” and tranquilizers became
apparent among juvenile delinquents; the Ministry of Health and Welfare designated
those drugs as “habit-forming drugs” under the Pharmaceutical Affairs Law in 1961
(Nagahama 1968; Ministry of Justice 1964). The “habit-forming drugs” include
bromovalerylurea, propofol, flunitrazepam, triazolam, and pentazocine. Some of
widely abused “sleeping drugs,” such as methaqualone, were available over the
counter. These drugs were then designated as “deleterious drugs” under the Phar-
maceutical Act in 1963; the abuse cases of those drugs dropped after the amendment
784 A. Ishii and Y. Natori
of the Law (Ministry of Justice 1980). The offenders of the Narcotics Control Act
were kept in the range of 100 and 450 in the 1970s. In 1990, Japan adopted the
Convention on Psychotropic Substances, 1971; the new Narcotics and Psychotropics
Control Act, which encompassed not only narcotic drugs but also psychotropics, was
endorsed in the same year. It took 19 years for Japan to ratify the Convention on
Psychotropic Substances, 1971; it took 9 years in the US, 10 years in Italy, and
15 years in the UK to do the same. The long delay until ratification in many countries
has been assumed to be due to difficulties in regulating Schedule III and Schedule IV
drugs (Kato 1999). Offenses under the Narcotics Control Act (and the Narcotics and
Psychotropic Drugs Control Act) have been relatively regulated (see Fig. 2); how-
ever, in the mid-2010s, emergence of new psychoactive substances (NPSs) including
synthetic cannabinoids became a new problematic phenomenon in Japan. The
problems of these newly synthesized psychotropic drugs are discussed later.
Article 1 of the Narcotics and Psychotropic Drugs Control Act provides its
purpose: “to prevent the health and sanitation hazards caused by the abuse of
narcotics and psychotropics and to thereby promote the public welfare, by setting
in place the necessary controls on the import, export, manufacture, formulation of
pharmaceutical preparations, transfer, and other handling of narcotics and psycho-
tropics, as well as by taking action with regard to narcotics addicts such as
establishing measures to provide them with the necessary medical treatment” (Min-
istry of Justice 2016). Article 3 provides licensing for appropriate narcotics handlers;
a person not licensed cannot handle narcotics and psychotropics. Articles 12–29
stipulate prohibition and restrictions for narcotics. Article 50 provides licensing and
registration for appropriate psychotropic handlers. Articles 58–2 to 58–19 describe
3500
3000
Narcotics and Psychotropics
Cannabis
2500
2000
1500
1000
Opium
500
0
1951
1954
1957
1960
1963
1966
1969
1972
1976
1982
1985
1988
1991
1994
1997
2000
2003
2006
2009
2012
2015
1979
2018
Fig. 2 The number of arrests for violation of the Narcotics and Psychotropic Drugs Control Act
(dotted line with circles), the Cannabis Control Act (solid line), and the Opium Control Act (dotted
line with rectangles). Based on the data of White Paper on Crime 1997 and 2019 (Ministry of Justice
1997, 2019a, b)
Neuropsychopharmacotherapy: International Regulations and Regulating Laws. . . 785
Abuse of Etizolam
Etizolam, a thienotriazolobenzodiazepine derivative originally developed in Japan,
became commercially available in 1984; it has marketing authorization only in India,
Italy, and Japan [World Health Organization (WHO) 2019]. Etizolam has highly
anxiolytic potency and was not designated as any “psychotropic”; not only psychi-
atrists but also general physicians have thus prescribed etizolam for insomnia or
other psychosomatic symptoms. Although the risk of benzodiazepine abuse has been
acknowledged since the 1980s, benzodiazepine prescription cases were 6–20 times
higher in Japan compared to those prescribed in European countries from 1998 to
1999 (Murasaki 2001). The number of patients who were prescribed benzodiaze-
pines that developed abuse and dependence symptoms is not small; Matsumoto
pointed out that flunitrazepam, triazolam, etizolam, zolpidem, and brotizolam were
most abused in Japan (Matsumoto 2012). The Ministry of Health, Labour and
Welfare finally designated etizolam as a Type III psychotropic in 2016. Today,
etizolam is under national control in Denmark, Germany, Japan, Switzerland,
Poland, the United Arab Emirates, and the United Kingdom (WHO 2019).
Cross-References
References
Edström B. The forgotten success story: Japan and the methamphetamine problem. Japan Forum.
2015;27:519–43.
Hasegawa K, Wurita A, Minakata K, Gonmori K, Yamagishi I, Nozawa H, Watanabe K, Suzuki
O. Identification and quantitation of 5-fluoro-ADB, one of the most dangerous synthetic
cannabinoids, in the stomach contents and solid tissues of a human cadaver and in some herbal
products. Forensic Toxicology. 2015;33:112–21.
International Narcotics Control Board (INCB). List of precursors and chemicals frequently used in
the illicit manufacture of narcotic drugs and psychotropic substances under International
Control (Red List). 18th ed. 2018.
International Narcotics Control Board (INCB). List of narcotic drugs under international control
(Yellow List). 58th ed. 2019a.
International Narcotics Control Board (INCB). List of Psychotropic Substances under International
Control (Green List). 30th ed. 2019b.
Kato N. International conventions on psychotropics. (In Japanese). In: Matsushita M, Asai M,
Ushijima S, Kurachi M, Koyama T, Nakane H, Miyoshi Y, editors. Encyclopedia of clinical
psychiatry, Drugs and alcohol related disorders, vol. 8: Tokyo, Nakayama Shoten; 1999.
p. 109–23.
Kusano M, Zaitsu K, Taki K, Hisatsune K, Nakajima J, Moriyasu T, Asano T, Hayashi Y,
Tsuchihashi H, Ishii A. Fatal intoxication by 5F–ADB and diphenidine: detection, quantifica-
tion, and investigation of their main metabolic pathways in humans by LC/MS/MS and LC/Q-
TOFMS. Drug Test Anal. 2018;10:284–93.
788 A. Ishii and Y. Natori
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 790
A Broad View of Health Economics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 790
General Classification of the Objects of Study . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 791
The Association Between Health Economics and Neurology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 792
Economic Burden on Neurodegenerative Diseases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 793
Parkinson’s Disease . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 793
Alzheimer’s Disease . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 796
Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 801
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 803
Abstract
As an important economy branch dealing with the allocation of health-related
resource, health economics mainly covers studies concentrating on issues related
to multiple medical actions and disease researches concerning efficiency, effec-
tiveness, value, and behavior in the production and consumption of both health
and health care. The focus of this chapter is to give a brief introduction on the
notions and definitions of the modern health economies, as well as the general
containment of the works underlying this specific notion of study. Two commonly
seen neuropsychological or neurodegenerative diseases (Parkinson’s disease and
Alzheimer’s disease), both inconvertible and with relatively lengthy process, are
specifically described in order to illustrate the work of health economies in a well-
understood manner. The calculation of the economic burden and the process of
financial evaluation of the two examples would help demonstrate the importance
of economic planning of health problems.
Introduction
The focus of this chapter is on a rather doom-laden scope of subject that is not strictly
constrained within the field of pure medicine. In particular, it is the linkage between
economic analysis and the impact of means of medicine on which our life lifestyle
and life circumstances have exerted. As the discipline of economics concerned with
the efficient allocation of health-care resources, health economics is essentially
trying to maximize health benefits to society contingent upon available resources
generally. One may see the merits of the application of economic tools effortlessly,
since those proved general economic principles offer a theoretical foundation of
hygienic resource allocation dealing with an environment with a scarcity of
resources. Back to the early period of human society, medicine, as well as the act
of the medical treatment, seemed to be an intuition of moral salvation with a
mutually reciprocal effect on individuals; the advent of capitalistic era, however,
had brought with itself the private ownership and market economy, which redefined
the vocation of the medical workers previously thought to be unselfish saviors. For
the time being, medicine has gradually become an independent means of profession
and livelihood. At the meantime, along with the unprecedented promotion of world
socialization since the end of the World War II, the expenditures on personal medical
services, as well as the increasingly complex situation resulted from the ever-
increasing mobility of population and the restless urban development, have exerted
great pressure on global control and management of various diseases. From this
perspective, the derivation of economic section in medical field seems to be an
inevitable trend.
Economics rests on two basic hypotheses: that the economic entities are rationally
guided (i.e., the optimization of economic costs or interests) and that the economic
resources are relatively scarce but versatile. As a consequence, the efficient alloca-
tion of the limited medical resources turns out to be the fundamental researching
purpose. Meanwhile, economics renders two basic tools and a set of criteria in order
to analyze the issues of efficiency and distribution: the techniques of optimization
and the determination of equilibrium situations. Both tools could be used by
authorities in charge of health and other organizations to determine the most efficient
allocation of medical or nonmedical resources, as well as the total expenditures and
policies governing the redistribution of varied services, such as the national health
insurance.
Past decades have witnessed a more rapid growth of expenditures on personal
medical services than that of other goods or services in the economy on a global
respect (Fan and Savedoff 2014). Statistics of the United States showed a doubled
health’s proportion of GDP between 1980 and 2011, with the fastest increased
health-care expenditure (2.9%/year) compared to any other industries during the
2000–2010 period (Anderson and Chalkidou 2008). Part of this increase in total
medical care expenditures is the result of the enlarged aging population and pro-
longed life span (Durrani 2016). Other reasons, such as rising income, changes in the
Health Economics for Neuropsychological Diseases in China 791
Health economics from its outset has been frequently misinterpreted to health-care
economics and has been used without specification on a wide range of topics. What
792 H.-l. Cui and G. Wang
exactly does this subject concentrate on, and to what extent this discipline of study
may apply to germane territory, is still under debate. Forerunners of this field have
tried to conceptualize this specific area more than once; and so far, the comprehen-
sively accepted encapsulation of this indistinctively delimited discipline was con-
cluded by Allan Williams in his schematic representation of the main elements in
health economics, which includes mainly 8 categories as follows: (A) What influ-
ences health (other than health care); (B) What is health, What is its value; (C)
Demand for health care; (D) Supply of health care; (E) Micro-economic evaluation
at treatment level; (F) Market equilibrium; (G) Evaluation at whole system level; (H)
Planning, budgeting and monitoring mechanisms (Williams 1987).
After a close look at the contents, we may notice that only two (A and B) of eight
categories are particularly related to health itself, while other domains are predom-
inantly occupied by factors such as health determinants and the evaluation, as well as
the operative features of health service organizations. However, since the aims for
health policies and researches are to improve the state of health of particular
individuals, of groups and of nations, it is essential that we combine analytical
insights with practical application to evaluate the real straits of the sick ones and
their caregivers.
Health services, like other goods, should be produced efficiently, and resources be
spread across various diseases so as to maximize benefits at the greatest extent; and
personal care should be given to reduce the burden of illness. Equity enters through-
out the health problem, across both its healing and care functions, but what is most
needed is a formulation of the problem in manageable form. The ethical and
philosophical issues, such as entitlement to equal health, the right to access proper
treatments, and the right to live or die, must also be faced at the end, and this is the
underlying core that health economics has pointed out. But the basics of measuring
health consumption, essential for the guidance of policy making, must be addressed
as primary task.
Parkinson’s Disease
using the questionnaire designed for this study. The questionnaire included two main
parts. Part I requested PD patient characteristics such as demographic details,
information on concurrent conditions, disease course, and disease stage according
to Hoehn & Yahr (H&Y). PD resource use and cost data for the 1-year period were
derived from Part II, which consisted of two sections: Part 2a assessed resource use
including drug treatment, number of outpatient visits and inpatient admissions,
means of transportation to the clinic, and special equipment; Part 2b directly
assessed the costs of inpatient care, outpatient visits, PD drugs (excluding drugs
for secondary symptoms such as constipation, sleep disorders, etc.), physical treat-
ment, purchase of equipment needed because of PD symptoms, medically related
travel expenses, professional home care, and productivity loss due to sick leave, and
early disability retirement. To ensure the accuracy of the recall information during
the interview, particularly for some patients with cognitive forfeit, PD patients were
asked to be accompanied by a caregiver and to bring their medical records, including
outpatient and inpatient records. The study was approved by the Research Ethics
Committee, the Medical School of Shanghai Jiaotong University, China. Voluntary
completion of the questionnaires was interpreted as consent to participate.
Calculation of Costs
The cost per patient was determined for each of the five H&Y stages of disease
progression. Unit costs were derived from the Shanghai public health administra-
tions tariff for 2004 (annual listing of all licensed drugs in Shanghai and their prices),
local taxi and public transportation published fares, Shanghai city guidelines for
mileage reimbursement, and the local fees for home care and home help service
provided by the local authority. Indirect costs were estimated using the human
capital approach. Evaluations of productivity loss due to long-term sick leave and
early retirement were based on the Shanghai city average annual income according
to the 2004 annual statistics almanac of Shanghai. For health-care cost and produc-
tion loss, only PD-related resource use was recorded. Monetary values were reported
as 2004 Chinese yuan, renminbi (RMB). For reference, the exchange rate was
USD 1 = RMB 8.3 or EUR 1 = RMB 10.5 in 2004.
To provide a measure of dosage change in dopaminergic drug use, the daily
L-dopa equivalent dose (LED) was calculated on the basis of the following equiv-
alences: 100 mg of standard L-dopa = 140 mg of controlled release L-dopa = 10 mg
of bromocriptine = 1 mg of pergolide = 5 mg of ropinirole = 1 mg of
pramipexole = 10 mg of selegiline.
Results
Characteristics of Patients
We interviewed 201 patients with PD, and all participants were permanent residents
in Shanghai. Valid data were collected from 190 patients, giving a response rate of
94.5%. There were 116 men and 74 women with an overall mean age of 67.9 years
(range, 36–85 years) a mean disease duration of 6.0 years (range, 1–23 years). With
regard to H&Y classification, there were 88 patients in Stage I, 80 in Stage II, 18 in
Health Economics for Neuropsychological Diseases in China 795
Stage III, 4 in Stage IV, and none in Stage V. Of these, 87 patients had at least one
additional chronic disease such as hypertension or diabetes. The mean annual
income per patient was RMB 13,462 (range, RMB 2,400–150,000).
Neurological Care
The average number of outpatient contacts with the Department of Neurology was
13.8 per patient per year. Visiting the PD clinic was the single most frequent form of
contacts, accounting for 99% of all contacts. There was a trend toward increasing
outpatient care across H&Y stages. During the year in question, 16 patients were
admitted for neurological care on a total of 16 occasions (0.1 per patient in total).
Except for one patient who received electrode implantation for deep brain stimula-
tion (DBS), the other 15 patients were hospitalized for treatment of severe symptoms
such as motor fluctuations and freezing of gait.
Home Care
In addition to getting help from family members, mainly from spouses and/or
children, 23.2% of patients also required a private nursemaid. To improve their
medical treatment, 25.8% of patients bought preventive health-care foods by
themselves. Special equipments, including wheel chairs and walking sticks, were
needed by 7.4% and 9.5% of patients, respectively.
796 H.-l. Cui and G. Wang
Loss of Production
There were 33 patients of working age, that is, under the age of 60 in China. Many
patients did not work full time before developing PD symptoms due to other factors,
including having other illnesses. Three people were on sick leave occasionally, and
one patient had retired early.
Costs
The mean costs per patient for 1 year were summarized in Table 1. Direct medical
care costs averaged approximately RMB 4,305 per year per patient, of which drugs
accounted for the major costly component. Annual drug costs averaged RMB 2,677
per patient and increased with disease progression (P = 0.018). Nonmedical direct
costs were much less than direct health-care costs, averaging approximately RMB
3,301. Costs due to loss of productivity averaged approximately RMB 73 per patient
per year. Taken together, the overall mean annual cost for PD in our series was
approximately RMB 7,679. The number of clinic visits and the patient’s H&Y stage
were found to be the factors significantly associated with the total cost. Meanwhile,
LED and the number of drugs taken were found to be associated with the drug cost in
the multivariable regression analysis (Table 2). A majority of the patients (86.8%)
had some kind of medical insurance, which covered a part of or the entire medical
cost. No significant difference of the total cost between the patients with medical
insurance and those without was found.
Alzheimer’s Disease
Table 1 Mean annual PD-related costs (RMB)/(USD) per patient for 1 year (the exchange rate was
USD 1 = RMB 8.3 in 2004)
H&Y I H&Y II H&Y III H&Y IV H&Y V
Costs (n = 88) (n = 80) (n = 18) (n = 4) (n = 190)
Direct medical 4,862/586 3,757/453 3,822/460 5,179/624 4,305/519
costs
Outpatient 159/19 181/22 158/19 192/23 169/20
costs
Impatient 2,291/276 446/55 444/53 0/0 1,329/160
costs
Western drug 2,202/265 2,812/339 3,041/366 4,986/601 2,597/313
costs
Physical 94/11 271/33 95/11 0/0 167/20
treatment
TCM drug 114/14 45/5 83/10 0/0 80/10
costs
Direct 2,146/259 3,342/403 7,871/948 7,196/867 3,301/398
nonmedical costs
Transportation 116/14 109/13 159/19 6/0.7 115/14
Home care 940/113 1,264/152 3,836/462 6,690/806 1,473/177
Preventive 1,089/131 1,908/230 3,500/422 0/0 1,640/198
care foods
Special 1/0 61/7 376/45 500/60 73/9
equipment
Indirect costs 42/5 101/12 115/14 0/0 73/9
Sick leave 42/5 79/10 115/14 0/0 64/8
Early 0/0 21/3 0/0 0/0 9/1
retirement
Overall mean 7,050/849 7,200/867 11,809/1423 12,375/ 7,679/925
total cost 1491
PD, Parkinson’s disease; RMB, renminbi; USD, United States dollar; TCM, traditional Chinese
medicine
countries (Prince 2004; Wong and Leung 2012; Wang et al. 2014). To better
understand the status of the economic impact of patients with dementia, and provide
baseline data for future evaluations of health economic impacts of AD in China, we
investigated the consumption of resource and analyzed total, direct, and indirect
costs and its variations associated with the disease severity in patients with AD in
Shanghai, China.
Table 2 AD-related costsa (RMB)/(USD) per patient for 1 year (mean SD) (the exchange rate
was USD 1 = RMB 8.0 in 2006)
Disease severity according to MMSE
Moderate
Costs Mild (n = 13) (n = 37) Severe (n = 16) Total (n = 66)
Direct medical 5,333 4,650/ 5,510 5,311/ 6,191 4,524/ 5,640 4,944/
costs 667 581 688 664 774 566 705 618
Outpatient 229 148/ 183 79/23 10 252 150/ 208 117/
costs 29 19 32 19 26 15
Inpatient costs 192 693/ 0 0/0 0 625 2,500/ 189 1,264/
24 87 78 313 24 158
Medications 4,911 4,710/ 5,291 5,216/ 5,230 4,599/ 5,202 4,906/
costs 614 589 661 652 654 575 650 613
Physical 0 0/0 0 70 286/9 36 123 245/ 69 247/9 31
therapy 15 31
Direct 2,829 3,471/ 2,667 1,750/ 3,050 1,965/ 2,792 2,199/
nonmedical 354 434 333 219 381 246 349 275
costs
Transportation 155 163/ 205 134/ 229 158/ 201 146/
19 20 26 17 29 20 25 18
Formal care 1,866 1,958/ 2,462 1,615/ 2,758 1,898/ 2,416 1,753/
233 245 308 202 345 237 302 219
Special 807 2,765/ 0 0/0 0 62 170/8 21 174 1,232/
equipments 101 346 22 154
Indirect costs 4,653 2,672/ 9,329 7,281/ 18,238 7,891/ 10,568 8,209/
582 334 1,166 910 2,280 986 1,321 1,026
Caregiver 4,653 2,672/ 9,329 7,281/ 18,238 7,891/ 10,568 8,209/
unpaid cost 582 334 1,166 910 2,280 986 1,321 1,026
Overall 12,816 4,843/ 17,507 10,922/ 27,480 10,365/ 19,001 11,037/
average cost 1,602 605 2,188 1,365 3,435 1,296 2,375 1,380
a
Mean annual costs per patient in 2006 RMB
study was approved by the Research Ethics Committee, Ruijin Hospital, affiliated to
Shanghai Jiaotong University School of Medicine, Shanghai, China. Voluntary
completion of the questionnaires was interpreted as consent to participate.
Examination (MMSE). AD resource use and cost data for the 1-year period were
derived from Part II which consisted of two sections: Part 2a assessed resource use
including drug treatment (including the drugs only for direct treatment of cognitive
deficit and not including drugs for secondary symptoms like constipation, sleep
disorders, etc.), number of outpatient visits and inpatient admissions, and means of
transportation to the clinic and assistive devices; Part 2b directly assessed the costs
of inpatient care, outpatient visits, mediations, assistive devices, medical-related
travel expenses, formal or professional care, and unpaid caregiving for the patients.
The caregivers and patients were also asked to provide investigators with their
medical records and inpatient and/or outpatient invoices.
We used the MMSE, activity of daily living (ADL)/instrumental activity of daily
living (IADL) scales, and neuropsychiatric inventory (NPI) to examine the patients’
cognitive function, functional capacity, psychotic symptoms, and behavioral prob-
lems as well as the Zarit Burden Interview (ZBI) to assess the informal care burden
of caregivers. The assessment tools were employed for patients in the present study:
(1) MMSE was used to assess patients’ cognitive functions. Three levels of cognitive
impairment were classified according to the MMSE score: mild (21–26), moderate
(11–20), and severe (0–10); (2) ADL and IADL were employed to estimate the
functional capacity of patients; and (3) NPI was applied to evaluate psychopathology
in dementia patients. It covers 12 neuropsychiatric disturbances common in
dementia: delusions, hallucinations, agitation, dysphoria, anxiety, apathy, irritability,
euphoria, disinhibition, aberrant motor behavior, night-time behavior disturbances,
and appetite and eating abnormalities. The severity and frequency of each neuro-
psychiatric symptom were rated on the basis of scripted questions administered
to the patient’s caregiver. A total NPI score that ranges from 0 to 144 was calculated,
in addition to the scores for the individual symptom domains ranging from 0 to 12.
The assessment tools were applied for caregivers: The ZBI comprised 22 items.
Caregivers were required to indicate the level of distress caused by each item,
ranging from “never” to “nearly always” distressing, on a scale of 0–4 and summed
with a total score ranging from 0 to 88, and a high score correlated with a high level
of burden.
Calculation of Costs
The cost per patient was determined for each of the three levels of the disease
severity. (1) Unit costs were derived from the Shanghai public health administra-
tion’s tariff for 2006 (annual listing of all licensed drugs in Shanghai and their
prices), local taxi and public transportation published fares, Shanghai city guidelines
for mileage reimbursement, and the local fees for home care and home help service
provided by the local authority. (2) Indirect costs were estimated using the replace-
ment wage approach (Guerriere et al. 2008; Harrow et al. 2004). Hours of unpaid
care provided per day for total caregiving task according to the tasks of ADL and
IADL were asked in the following categories: 0, 1–2 h, 3–5 h, 6–8 h, 9–11 h, and
12 h or more. We transformed the categories into continuous values using the mean
of each category as the estimated hours of care provided. For subjects who reported
12 h or more per day for a particular type of task, we top-coded the value as 12 h.
800 H.-l. Cui and G. Wang
Evaluations of informal caregiving cost were based on the minimum hourly earnings
from Shanghai Municipal Labor and Social Security Bureau as the hourly wage rate
to estimate unpaid care-giving cost. (3) For health-care cost, only AD-related
resource use was recorded. (4) Monetary values were reported as 2006 Chinese
Yuan, Renminbi (RMB). For reference, the exchange rate was 1 USD = 7.97 RMB
or 1 EUR = 10.01 RMB in 2006.
Results
Characteristics of Patients
We interviewed 67 patients with AD and their caregivers. All participants
were community-dwelling patients in Shanghai. Valid data were collected from 66
patients, giving a response rate of 98.5%. There were 23 men and 43 women with the
overall mean age of 74.0 years (range: 53–90 years) and the mean disease duration
was 2.7 years (range: 1–12 years). With regard to MMSE classification, there were
13 patients in mild stage (score: 21–26), 37 moderate (11–20), and 16 severe (0–10).
Of these, 36 patients had at least 1 additional chronic disease such as hypertension or
diabetes.
Resources use was summarized as follows:
1. Neurological care: The average number of outpatient visits to and contacts with
an AD clinic was 14.9 per patient per year. During the year in question, two
patients were hospitalized to get treatment for pneumonia and diabetes,
respectively.
2. Medications: 51 of 66 patients received anti-dementia drugs for therapy. The most
frequently used drug was acetylcholinesterase inhibitors (62.1% of patients took
huperzine A, 28.8% donepezil, and 4.5% rivastigmine), 6.1% of patients took
memantine, 13.6% Ginkgo biloba extract (EGB), and 4.5% traditional Chinese
medicine (TCM).
3. Transportation and non-neurological care: All but 12 patients took a transporta-
tion service to reach the AD clinic. Public transportation and taxi were used by 39
(59.1%) and 7 (10.6%) patients, while use of personal car was 8 (12.1%) patients.
Fifty-eight (87.8%) patients visited the AD clinic usually with a companion and
8 (12.1%) patients had 2 or more companions who usually were spouse or
offspring of patients. Beyond the health care provided by neurologists, few
patients sought alternative traditional Chinese health care, and only one received
massage (1.5%) provided by physiotherapists.
4. Formal care: In addition to getting help from family members, mainly from
spouses and/or offspring, 9.0% of patients also required a private nursemaid.
To improve their condition, 12.1% of patients took preventive health-care foods
from their relatives.
Special equipments including wheelchairs were needed by 6.0% of patients.
5. Informal care: There were 33 men and 33 women with the overall mean age of
64.8 years among the primary caregivers. 68.2% of caregivers were patients’
spouses, 30.3% offspring, and 1.5% other relatives. The average time spent (h/
Health Economics for Neuropsychological Diseases in China 801
day) was 4.45 (1.5–12) for each patient with AD. The mean costs per patient for
1 year are summarized in Table 2.
Direct medical care costs averaged approximately RMB 5,640 per year per
patient, of which drugs accounted for the major costly component. Annual drug
costs averaged RMB 5,202 per patient. Nonmedical direct costs were much less than
direct health-care costs, averaging approximately RMB 2,792. Caregiver cost aver-
aged approximately RMB 10,568 per patient per year and was the most significant
item in the overall cost. Taken together, the overall mean annual cost for AD in our
series was approximately RMB 19,001.
Discussion
clinical features associated with the costs, the associations found in the study should
be cautiously interpreted. Second, although the indirect cost for AD and PD was
estimated to be substantial, it should be noted that we underestimated the actual
indirect cost by only including the cost of caregivers’ time spent on caregiving. For
example, indirect cost may also include those related to caregivers’ lost productivity
and reduced hours of work and income. In our study of AD, the majority (68.2%) of
the primary caregivers were spouses of patients and old retired persons; therefore
costs from reduced work hours were relatively minor. At the same time, reductions
in work hours due to caregiving and hours spent on caregiving may overlap and lead
to possible double counting of caregivers’ cost. Third, the study is limited by a
relatively small sample size. These limitations were compensated by the general
representative and refinement of diagnosis and richness of clinical variables in
our study. Still, it is believed that our results will be referred to as a baseline for
large-scale investigations in the future.
As the disease progresses, there is a loss of cognitive and physical functions
which ultimately leads to complete dependency, as is believed to be applicable for all
types of neurodegenerative diseases. For example, previous studies have revealed
a significant relationship between total costs of care and the level of dementia
severity in AD (Jonsson et al. 2006; Park et al. 2015). The association of the cost
and the severity of dementia were assessed by the level of cognitive impairment
(e.g., MMSE), the occurrence of behavioral disorders (e.g., NPI), and the physical
abilities (e.g., ADL and IADL) of our patients. As cognitive and physical functions
are highly correlated, reduced cognitive function is likely responsible for the impor-
tant share of the physical dependency. By contrast, our studies revealed that the
caregivers’ burden was not a significant factor associated with the total cost. This
result possibly indicated that the caregiver burden is not always paralleled with the
severity of disease and may be explained by our negligence in calculations of the
intangible time cost which related to pain and suffering endured by the caregivers.
Based on these, the total cost increased with the disease severity should be under-
standable. The mean total cost for patients in the severe stage was almost double or
threefold of that for patients in the mild stage, which was consistent with the results
from previous studies. In addition, because we used the city minimum hourly
earnings for all private industries as the hourly wage rate to estimate unpaid
caregiving costs, the cost of caregiver’s time which increased with the disease
severity was also easily understandable. Compared with the situation of economic
burden of PD, there are different component features of total cost. For PD and some
other types of neuropathy, previous studies indicated that the indirect cost is only a
minority of the total cost (Wang et al. 2006). However, the indirect cost rather than
the direct cost is the majority of the total cost in AD cases. These characteristics
possibly reflect different roles of caregivers’ care and lost productivity in the cost of
care determined by the features of disease itself. In addition, AD-related direct health
costs were compared between our study and that from developed countries such as
the United States, the United Kingdom, and developing areas like Latin America and
Turkey (Yang et al. 2013; Morris et al. 2015; Manes 2016). Although the absolute
values reported in our study are far lower than that from developed countries and
Health Economics for Neuropsychological Diseases in China 803
close to that from developing countries, the differences can be partly explained by
the substantial differences in hourly wage rates and actual income between western
countries and China. Differences in income level between countries might be able to
explain the difficulty to compare the actual economic burden for patients around the
world.
The findings of this study indicate that the economic burden for Chinese patients
with neurodegenerative diseases is heavy and increases dramatically with the disease
severity. The cost estimates provided by this study can be useful in future economic
evaluations of interventions aimed at reducing the progression of AD and PD in
China.
References
Allegri RF, Butman J, Arizaga RL, Machnicki G, Serrano C, Taragano FE, Sarasola D, Lon L.
Economic impact of dementia in developing countries: an evaluation of costs of Alzheimer-type
dementia in Argentina. Int Psychogeriatr. 2007;19(4):705–18.
Anderson GF, Chalkidou K. Spending on medical care: more is better? JAMA. 2008;299(20):
2444–5.
Bragaglia P, O’Brien L. Chronic disease management: the primary care perspective. Healthc Pap.
2007;7(4):26–8; discussion 68–70.
Dodel RC, Haacke C, Zamzow K, Paweilik S, Spottke A, Rethfeldt M, Siebert U, Oertel WH,
Schoffski O, Back T. Resource utilization and costs of stroke unit care in Germany. Value
Health. 2004;7(2):144–52.
Durrani H. Healthcare and healthcare systems: inspiring progress and future prospects. mHealth.
2016;2:3.
Fan VY, Savedoff WD. The health financing transition: a conceptual framework and empirical
evidence. Soc Sci Med. 2014;105:112–21.
Fan J, Dawson TM, Dawson VL. Cell death mechanisms of neurodegeneration. Adv Neurobiol.
2017;15:403–25.
Garibaldi P, Martins JO, van Ours J. Ageing, health, and productivity: the economics of increased
life expectancy. Oxford: Oxford University Press; 2010.
Gooch CL, Pracht E, Borenstein AR. The burden of neurological disease in the United States:
a summary report and call to action. Ann Neurol. 2017;81(4):479–84.
Guerriere DN, Tranmer JE, Ungar WJ, Manoharan V, Coyte PC. Valuing care recipient and family
caregiver time: A comparison of methods. International Journal of Technology Assessment in
Health Care. 2008;24(01):52–9.
Hall RE, Jones CI. The value of life and the rise in health spending. Q J Econ. 2007;122(1):39–72.
Harrow BS, Mahoney DF, Mendelsohn AB, Ory MG, Coon DW, Belle SH, Nichols LO. Variation
in cost of informal caregiving and formal-service use for people with Alzheimer’s desease.
American Journal of Alzheimer’s Disease & Other Dementias. 2004;19(5):299–308.
Jia J, Zhou A, Wei C, Jia X, Wang F, Li F, Wu X, Mok V, Gauthier S, Tang M, Chu L, Zhou Y, Zhou
C, Cui Y, Wang Q, Wang W, Yin P, Hu N, Zuo X, Song H, Qin W, Wu L, Li D, Jia L, Song J,
Han Y, Xing Y, Yang P, Li Y, Qiao Y, Tang Y, Lv J, Dong X. The prevalence of mild cognitive
impairment and its etiological subtypes in elderly Chinese. Alzheimers Dement. 2014;10
(4):439–47.
Jonsson L, Eriksdotter Jonhagen M, Kilander L, Soininen H, Hallikainen M, Waldemar G, Nygaard
H, Andreasen N, Winblad B, Wimo A. Determinants of costs of care for patients with
Alzheimer’s disease. Int J Geriatr Psychiatry. 2006;21(5):449–59.
Kowal SL, Dall TM, Chakrabarti R, Storm MV, Jain A. The current and projected economic burden
of Parkinson’s disease in the United States. Mov Disord. 2013;28(3):311–8.
804 H.-l. Cui and G. Wang
Langa KM, Larson EB, Crimmins EM, Faul JD, Levine DA, Kabeto MU, Weir DR. A comparison
of the prevalence of dementia in the United States in 2000 and 2012. JAMA Intern Med.
2017;177(1):51–8.
Liu N, Zhang J, Guo L. Alzheimer’s disease epidemiological situation. J Liaoning Univ Tradit Chin
Med. 2011;1:35–6.
Manes F. The huge burden of dementia in Latin America. Lancet Neurol. 2016;15(1):29.
Morris S, Patel N, Baio G, Kelly L, Lewis-Holmes E, Omar RZ, Katona C, Cooper C, Livingston G.
Monetary costs of agitation in older adults with Alzheimer’s disease in the UK: prospective
cohort study. BMJ Open. 2015;5(3):e007382.
Muangpaisan W, Hori H, Brayne C. Systematic review of the prevalence and incidence of
Parkinson’s disease in Asia. J Epidemiol. 2009;19(6):281–93.
Park M, Sung M, Kim SK, Kim S, Lee DY. Multidimensional determinants of family caregiver
burden in Alzheimer’s disease. Int Psychogeriatr. 2015;27(8):1355–64.
Prince M. Care arrangements for people with dementia in developing countries. Int J Geriatr
Psychiatry. 2004;19(2):170–7.
Rajput AH, Offord KP, Beard CM, Kurland LT. Epidemiology of parkinsonism: incidence,
classification, and mortality. Ann Neurol. 1984;16(3):278–82.
Sorenson C, Drummond M, Bhuiyan Khan B. Medical technology as a key driver of rising health
expenditure: disentangling the relationship. Clinicoecon Outcomes Res. 2013;5:223–34.
von Campenhausen S, Bornschein B, Wick R, Botzel K, Sampaio C, Poewe W, Oertel W, Siebert U,
Berger K, Dodel R. Prevalence and incidence of Parkinson’s disease in Europe. Eur
Neuropsychopharmacol. 2005;15(4):473–90.
Wang G, Cheng Q, Zheng R, Tan YY, Sun XK, Zhou HY, Ye XL, Wang Y, Wang Z, Sun BM,
Chen SD. Economic burden of Parkinson’s disease in a developing country: a retrospective cost
analysis in Shanghai, China. Mov Disord. 2006;21(9):1439–43.
Wang G, Cheng Q, Zhang S, Bai L, Zeng J, Cui PJ, Zhang T, Sun ZK, Ren RJ, Deng YL, Xu W,
Wang Y, Chen SD. Economic Impact of Dementia in Developing Countries: An Evaluation of
Alzheimer-Type Dementia in Shanghai, China. Journal of Alzheimer’s Disease. 2008;15
(1):109–15.
Wang J, Xiao LD, He GP, De Bellis A. Family caregiver challenges in dementia care in a country
with undeveloped dementia services. J Adv Nurs. 2014;70(6):1369–80.
Williams A. Health and Economics. Palgrave Macmillan, London. 1987.
Wong YC, Leung J. Long-term care in China: issues and prospects. J Gerontol Soc Work. 2012;
55(7):570–86.
Wu C, Gao L, Chen S, Dong H. Care services for elderly people with dementia in rural China: a case
study. Bull World Health Organ. 2016;94(3):167–73.
Yang Z, Lin PJ, Levey A. Monetary costs of dementia in the United States. N Engl J Med. 2013;
369(5):489.
Zencir M, Kuzu N, Beser NG, Ergin A, Catak B, Sahiner T. Cost of Alzheimer’s disease in
a developing country setting. Int J Geriatr Psychiatry. 2005;20(7):616–22.
Zhang Y, Xu Y, Nie H, Lei T, Wu Y, Zhang L, Zhang M. Prevalence of dementia and major
dementia subtypes in the Chinese populations: a meta-analysis of dementia prevalence surveys,
1980–2010. J Clin Neurosci. 2012;19(10):1333–7.
Zhang H, Loi SM, Zhou S, Zhao M, Lv X, Wang J, Wang X, Lautenschlager N, Yu X, Wang H.
Dementia literacy among community-dwelling older adults in urban China: a cross-sectional
study. Front Public Health. 2017;5:124.
Zhu CW, Scarmeas N, Torgan R, Albert M, Brandt J, Blacker D, Sano M, Stern Y. Clinical features
associated with costs in early AD: baseline data from the Predictors Study. Neurology. 2006;
66(7):1021–8.
Zou YM, Liu J, Tian ZY, Lu D, Zhou YY. Systematic review of the prevalence and incidence
of Parkinson’s disease in the People’s Republic of China. Neuropsychiatr Dis Treat.
2015;11:1467–72.
Causal Inference Methods in
Pharmacoepidemiology
Contents
Pharmacoepidemiology and Causal Inference . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 806
Common Study Designs in Pharmacoepidemiologic Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 807
Cohort Study Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 808
New-User Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 808
Active Comparator Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 809
Case-Control Study Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 809
Within-Subjects Designs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 810
Exposures in Pharmacoepidemiologic Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 810
Confounding in Pharmacoepidemiologic Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 812
Confounding Adjustment Methods in Pharmacoepidemiologic Studies . . . . . . . . . . . . . . . . . . . . . . . 814
Potential Outcomes Framework and Causal Assumptions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 814
Methods for Time-Fixed Confounding Adjustment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 815
Methods for Complex Time-Varying Confounding Adjustment . . . . . . . . . . . . . . . . . . . . . . . . . . . 816
Applications in Neuro-Psycho-Pharmacoepidemiology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 817
Applicability of Causal Inference Methods in Pharmacoepidemiology . . . . . . . . . . . . . . . . . . . . . . . . 818
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 820
Abstract
This chapter provides an introduction to pharmacoepidemiology, with a special
focus on some of the key concepts related to the application of causal inference
methods in pharmacoepidemiology, particularly in the time-varying setting. The
chapter first describes frequently used study designs in pharmacoepidemiologic
studies of treatment effects. The chapter then follows with drug exposures and
L. Pazzagli (*)
Centre for Pharmacoepidemiology, Department of Medicine, Karolinska Institutet, Stockholm,
Sweden
e-mail: laura.pazzagli@ki.se
X. Li
Department of Population Medicine, Harvard Medical School and Harvard Pilgrim Health Care
Institute, Boston, MA, USA
e-mail: Xiaojuan_Li@HarvardPilgrim.org
years, important efforts have been made to integrate evidence coming from both
experimental and observational settings and to enhance the use of retrospectively
collected observational data in pharmacoepidemiology (ElZarrad and Corrigan-
Curay 2019; Franklin et al. 2019; Schneeweiss et al. 2016). This has led to increased
use of observational studies to assess the effectiveness and safety of a drug after
regulatory approval, particularly when RCTs are unfeasible or cannot be done in a
timely manner.
Use of routinely collected observational data for evidence generation has many
advantages. Using readily available data, a research question can be answered in a
timely manner and at lower cost with an observational study relative to using an
RCT. The use of large heterogeneous populations, combined with appropriate
analyses, has the potential of wider generalizability of the estimated treatment
effects. However, challenges exist to provide valid evidence with observational
data. Because the treatment is not randomized in the observational setting,
researchers need to rely on the information collected in the data source to adjust
for potential confounding. However, because the data are routinely collected for
administrative or clinical reasons rather than for a specific research question, it is
possible that information is unmeasured on potentially important factors for the
exposure-outcome relationship of interest, leading to potentially biased estimates of
treatment effects. Moreover, when investigating time-varying treatment effects, the
presence of complex time-varying confounding, which may be affected by prior
treatments, necessitates the use of proper adjustment methods other than conven-
tional regression approaches. These challenges warrant the use of robust and flexible
causal inference methods in studies using observational data.
Causal inference is a set of graphical and mathematical tools developed to answer
causal questions using observational data. These methods allow to link causal
structures between the variables involved in a research question to the observed
data via causal assumptions and modeling needed to draw statistical inference.
Specifically, the potential outcomes framework (Rubin 1974) allows the estimation
of causal effects of treatments by comparing the potential outcomes that would have
been observed if all individuals in the study population were exposed to each
treatment of interest. Under this framework, several methods have been proposed
and used to assess drug effectiveness and safety in pharmacoepidemiology and will
be explored in the rest of this chapter.
important to consider all potential study design options before choosing the most
appropriate one for any given research question.
In general, epidemiologic studies are defined by how study participants are sampled.
In cohort studies, individuals are sampled based on their treatment (i.e., exposure) status
at a certain point in time and followed over time for the occurrence of outcome(s).
Cohort entry is often defined by a meaningful event such as initiation of a medication
treatment, as is the case for the new-user design discussed below (Ray 2003). A group
of individuals who have not been exposed to the medication under study or exposed to
a comparator drug are also included in the cohort to use as a comparison group. The
probabilities of developing the outcome of interest in the two groups are then compared
to estimate the treatment-outcome association.
A major advantage of the cohort study design is that it measures exposure and
outcome, and exposure and potential confounders in temporal sequence, making it
possible to elucidate causal relationships. Cohort studies also allow for investigation
into multiple outcomes from given treatments in one study. Compared to the design
and conduct of an RCT, cohort studies, especially with retrospective collection of
data from existing electronic health databases or registries, are also timelier and more
competitive in terms of economic resources.
The cohort study design becomes costly and less timely when study participants
need to be recruited first and followed over time such as in prospective cohort
studies. It can also be inconvenient for studies of rare outcomes (e.g., cancer) that
require a long follow up. The common challenges that affect all observational studies
also apply to the cohort study design, which include the inability to control for all
confounders that distort the treatment-outcome relationship under interest as well as
the potential for selection bias due to the way participants are selected into the cohort
and differential loss to follow up between the comparison groups due to migration,
death, or drop-out of the study.
New-User Design
Within cohort studies for investigating pharmacological treatment effects, the con-
ventional prevalent-user design includes both prevalent and new users of a drug.
However, prevalent users who have been taking a medication for some time are
likely to be more tolerant to the drug, perceive some therapeutic benefits, have better
health in general, or experience long-term adverse effects due to the accumulation of
the treatment over time (Glynn et al. 2001). Consequently, the prevalent-user design
is subject to selection bias and confounding bias. In addition, covariates selected for
confounding adjustment in prevalent-user studies may have been affected by prior
treatment. Necessary adjustment for these covariates to remove confounding will,
however, remove some treatment effect under study.
Causal Inference Methods in Pharmacoepidemiology 809
The new-user design (Ray 2003) provides a solution to remove the confounding
and selection bias associated with the prevalent-user design by restricting study
cohorts to all individuals in a defined population who initiate a course of treatment
with the study medication. This design ensures the appropriate temporal ordering of
baseline confounders, exposures, and outcomes and anchors the timescale for
analyses at time since treatment initiation for all individuals in the study, conse-
quently avoiding selection bias and adjustment for intermediate variables that may
be on the causal pathway between exposure and outcome. The new-user design is
widely used in pharmacoepidemiologic studies.
In cohort studies, another important design aspect to consider is the choice of the
comparator group. When comparing treated individuals with untreated individuals,
the study is prone to confounding by indication (Epidemiology 1996) and
confounding by frailty, which are difficult to control for. The active-comparator
design (Lund et al. 2015) (Schneeweiss et al. 2007) uses a group of individuals
initiating on a different therapeutic alternative for the same indication to help
mitigate biases from both types of confounding. When coupled with the new-user
design, this design can also reduce the potential for immortal-time bias (Suissa
2003). The active-comparator, new-user study design is considered to be the stan-
dard for comparative effectiveness research in pharmacoepidemiology (Johnson
et al. 2013). Researchers can use current treatment guidelines for the condition of
interest to select active comparator treatments.
In contrast to the cohort study design, the case-control study design (Martinez et al.
2010) samples individuals into a study based on their outcome status. With this
design, individuals who have developed the outcome of interest are first identified
(i.e., cases). Their treatment exposure history is then compared with that of individ-
uals without the outcome of interest (i.e., controls) to estimate the odds ratio, the key
measure of treatment-outcome association in case-control studies. Due to the lack of
denominators for the two treatment groups, case-control studies cannot directly
estimate incidence or incidence rates of the outcome in the two groups, thus cannot
directly estimate the relative risk, unless additional information from the underlying
cohort is available. However, when the outcome under study is a rare condition, the
estimated odds ratio in case-control studies approximates the relative risk.
With the case-control study design, it is possible to examine multiple treatment
exposures in one single study. By oversampling of individuals with the outcome,
case control studies are efficient compared to traditional cohort studies in that they
generate the same adequate information using a smaller sample size, reducing the
cost and burden of prospective data collection. This efficiency gain, however, is
810 L. Pazzagli and X. Li
Within-Subjects Designs
Fig. 1 A simplified directed acyclic graph for the relationship between a treatment A and an
outcome Y
Fig. 2 A simplified directed acyclic graph for the relationship between a time-varying treatment
A and an outcome Y
812 L. Pazzagli and X. Li
neuropsychiatric programs have adopted the chronic care model to provide care,
such as treatments for depression care in elderly patients (Mcevoy and Barnes 2007).
Other examples of time-varying treatments include intermittent therapies such as
selective serotonin reuptake inhibitors (SSRIs) for the treatment of premenstrual
symptoms.
In this context, dynamic treatment strategies offer a way to operationalize the
sequential decision-making process involved in adaptive clinical practice. A
dynamic treatment strategy is a sequence of decision rules, one per each stage of
treatment/intervention depending on the clinical progress of the patient (Hernan et al.
2009). Each decision rule considers a patient’s individual characteristics and treat-
ment history observed up to a given stage and offers a recommended treatment at
that stage (recommendations can include treatment type, dosage, and timing).
Conceptually, a dynamic treatment strategy can be viewed as a decision support
system, which is one of the six elements of the chronic care model. By allowing
providers to adjust treatment due to side effects or lack of effectiveness, dynamic
treatment strategies are more realistic compared to static strategies or recommenda-
tions that require the same treatment over time regardless. Accordingly, more and
more clinical guidelines provide treatment recommendations in the format of
dynamic treatment strategies.
The relationship between a drug treatment and a health outcome may be confounded
by factors which affect both the treatment and the outcome. In the presence of
confounding, the true effects of the treatment under study on the outcome of interest
are mixed with effects of factors other than the drug treatment itself (Greenland et al.
1999b). Without appropriately accounting for the presence of confounding, the
estimated treatment-outcome associations will be confounded. Special attention is
needed on the analytical methods to estimate unbiased causal effects in the presence
of confounding factors.
To identify confounding, a useful tool is provided and formalized in the theory of
DAGs (Greenland et al. 1999a; Pearl 1995). DAGs are directed in that the direction
of the arrowhead indicates the directions of the hypothesized relationships between
the variables. Because a variable cannot be caused by itself, and the future cannot
affect the past, DAGs are acyclic. The use of these diagrams allows one to visualize
causal links between variables in an intuitive way and, together with background
clinical knowledge, helps to identify potential confounders for the treatment-
outcome association under study. The presence of confounding, which exists if
the drug treatment remains associated with the outcome also after removing all the
arrows representing treatment effects, is intuitively represented in DAGs by the
presence of a common cause for the drug treatment and the health outcome (Fig. 3).
Background knowledge is often needed to draw the DAG for a research question.
In the example of a pharmacoepidemiologic study that concerns the effect of a drug
treatment on the health outcome of interest, background knowledge or assumptions
Causal Inference Methods in Pharmacoepidemiology 813
are needed with respect to the mechanism of action. These assumptions are qualita-
tive rather than quantitative because causal links represented in causal diagrams do
not contain information about the functional form that connects variables.
The simplified DAG in Fig. 3 represents the scenario where confounding is
present for the relationship between a time-fixed drug treatment A and an outcome
Y. In order to estimate the true causal effect of the treatment A on Y, appropriate
adjustment for the confounder C is necessary.
When investigating long-term treatment effects using observational data, effect
modification may be introduced by time-varying covariates, and incorrect identifi-
cation of the causal pathways may occur (Newsome et al. 2018). Moreover, the DAG
gets more complex in the setting of time-varying treatments, especially in the
presence of time-varying confounding; the levels of the variables at different time
points affect the levels of the other variables involved in the study (Fig. 4). There-
fore, to identify drug treatment effects, analyses need to account for changes in
treatments and confounders across time.
In the setting of dynamic strategies with treatments changing over time, the
scenario becomes even more complex because some confounding variables may
also be affected by previous treatment. Thus, these time-varying confounders
become intermediates on the causal pathway between the treatment and the outcome,
introducing a treatment-confounder feedback (as shown by the direct arrow between
A0 and C1 in Fig. 5).
An example of a time-varying confounder that also acts as causal intermediate is a
drug adverse effect caused by the previous treatment received. Patients experiencing
adverse effects may be advised to stop or change the therapy (influencing the
814 L. Pazzagli and X. Li
treatment at a later time t1) with consequences for the health outcome of interest.
This complex time-varying confounding deems traditional regression methods inva-
lid. Fortunately, causal inference methods are available to deal with this specific
problem and offer a great opportunity to estimate causal effects of time-varying
treatments in the presence of time-varying confounders and treatment-confounder
feedback (Robins and Hernán 2009).
In causal inference, the potential outcomes framework, also known as the Neyman-
Rubin Potential Outcomes, or the Rubin causal model, provides an approach to
assess causality in nonrandomized studies (Rubin 1974, 1977, 1978). A potential
outcome is the outcome for an individual under a potential treatment or treatment
strategy. The causal effect of a treatment is the difference between the potential
outcome under the treatment and the potential outcome under no treatment (or the
comparator treatment). However, in reality, all potential outcomes for an individual
cannot be observed. Instead, only one outcome can be observed for any individual,
the outcome under the treatment actually received, making it impossible to estimate
individual causal effects. The missingness of potential outcomes under all possible
drug treatment strategies for individuals in a study is the fundamental problem of
causal inference, which relates causal inference from observational data to a problem
of missing data. Even though the estimation of individual causal effects is not
possible with the observed data, other contrasts of interest, such as the population
average treatment effect, can be estimated using observational data under the
potential outcomes framework.
The potential outcomes framework needs several assumptions to draw causal
inference using observational data. The stable-unit treatment value assumption
(SUTVA) (Rubin 1986) requires that the outcome of an individual is independent
from the treatments assigned to the other individuals in the study. This assumption is
often referred to as two parts: no interference and no multiple versions of a treatment
Causal Inference Methods in Pharmacoepidemiology 815
to balance exposure groups with respect to the set of covariates used to estimate the
score (Rosenbaum et al. 2005). After estimation of the propensity score using a
regression model for the treatment, different confounding adjustment approaches
including matching, weighting, or stratification can then be used (Desai et al. 2017;
Glymour et al. 2019; Park et al. 2018). One can also adjust for confounding by
including the propensity score in the outcome regression model; however, this
approach requires the specification of the outcome model in relation to the propen-
sity score and that the relationship must be correctly specified. Moreover, the method
estimates a conditional treatment effect rather than a marginal treatment effect as
estimated via other propensity score-based methods. With propensity score
matching, individuals treated with the drug of interest are matched with individuals
from the comparison group on the estimated propensity scores, resulting in a group
of individuals with different treatment status balanced with respect to the covariates
used to estimate the propensity score. The treatment effect can then be estimated
using the matched groups in the outcome analysis. With propensity score stratifica-
tion, individuals in a study are grouped into strata based on their propensity score
values before effect estimation (Desai et al. 2017). With propensity score weighting,
the study population is weighed based on an inverse function of their estimated
propensity scores to create a pseudo-population in which the distributions of
covariates used in the propensity score estimation are balanced (which is similar to
standardization). The suite of propensity score-based approaches model the con-
founders in the model for the treatment rather than for the outcome, making them
particularly advantageous in the context of rare outcomes (Jackson et al. 2017).
Originally proposed for binary treatment, the propensity score methods have been
extended to the case of comparisons of multiple drug treatments where a generalized
propensity score can be used (Imai and Van Dyk 2004).
In studies involving time-fixed drug treatments, all these aforementioned
approaches can be used to estimate causal treatment effects. The choice of approach
depends on the research question and targeted treatment effect (e.g., average treat-
ment effect in the treated, or average treatment effect in the entire population). In
longitudinal studies with time-varying drug treatment and confounders, the methods
are still suitable, but regression models need to account for treatment and con-
founders as time-varying covariates. However, in situations where time-varying
confounders are affected by prior treatment, special problems arise and more
advanced approaches are needed.
affects the decision on the next drug treatment level. This type of complex
time-varying confounding is very common in dynamic drug treatment strategies.
For example, in a study of effectiveness of continuous antidepressant treatments, the
severity of other comorbidities associated with the response to the treatments may be
affected by the antidepressants received earlier on and may affect the decision on the
future use of antidepressants. In the presence of complex time-varying confounding,
traditional regression methods that adjust for the severity of comorbidities (used as a
covariate in the model) would fail to estimate the part of the effect of antidepressants
use on the outcome of interest mediated by comorbidities severity.
The complexity introduced by the treatment-confounder feedback requires spe-
cific methods for confounding adjustment. Robins’ g-methods are a special class of
methods meant to deal with this type of time-varying confounding. They include
inverse probability weighting estimation of marginal structural models, the paramet-
ric g formula, and g-estimation of structural nested models. Marginal structural
models can be estimated via weighted regression models. Each individual in the
study will receive a weight, which is a function of the probability of receiving the
treatment conditional on the history of the time-varying treatments and confounders.
The weighting process allows simulation of the potential outcomes for the individ-
uals under all treatment strategies. This in turn allows for estimation of the marginal
effect of exposure to the treatment on the outcome without directly adjusting in the
regression model for time-varying confounders that have resulted in treatment-
confounder feedback. The parametric g formula uses parametric models for the
distribution of the outcomes conditional on treatment and confounders and the
joint distributions of the confounders to simulate the potential outcomes under
different treatment strategies (Robins 1986). G-estimation of structural nested
models allows to calculate the potential outcomes for each individual under all
possible treatment strategies using the observed data. The effect of the exposure to
the treatment on the outcome is estimated from the observed data independently at
each time point adjusting for the history of the time-varying treatments and con-
founders. The estimated effects can be used to calculate all the potential outcomes
and the difference between the potential outcomes that would be observed if
individuals were always exposed and the potential outcomes if individuals were
never exposed (or exposed to another treatment strategy) correspond to a causal
effect (Robins 1994). The g-methods differ in implementation and computational
intensity, but they all aim to simulate the potential outcomes that would have been
observed under all the potential dynamic treatment strategies (Daniel et al. 2013; Li
et al. 2017; Naimi et al. 2017; Pazzagli et al. 2018).
Applications in Neuro-Psycho-Pharmacoepidemiology
Applications of these causal inference methods have increased in the field of neuro-
psycho-pharmacoepidemiology over recent years. For example, an investigation of
the risk of suicide attempts and suicide in adults in relation to antidepressant agents
used a propensity score adjusted analysis (Schneeweiss et al. 2010). Another
818 L. Pazzagli and X. Li
example studied the risk of death in elderly users of conventional versus atypical
antipsychotic medications using traditional multivariate Cox models, propensity-
score adjustments, and an instrumental-variable analysis (Wang et al. 2005). The use
of antidepressants during pregnancy in relation to the risk of cardiac defects was
examined using a propensity score stratification approach (Huybrechts et al. 2014).
Propensity score stratification and high-dimensional propensity score have been
used to assess the risk of antidepressant use during late pregnancy and persistent
pulmonary hypertension in the newborn (Huybrechts et al. 2015). Anti-convulsant
use and the risk of attempted suicide, suicide, and violent death has been investigated
using Cox proportional hazards models and propensity score–matched analyses
(Patorno et al. 2010). The adjuvant use of gabapentinoids with opioids and the
association with the risk of opioid-related adverse events in patients undergoing
surgery has been studied using propensity score stratification (Bykov et al. 2020).
These studies are examples of a more extensive literature on causal inference
applications in Neuro-Psycho-Pharmacoepidemiology studies using observational
data.
expected to increasingly become available. Together with evidence form RCTs, they
shall inform clinical decisions on appropriate treatment use in the field of neuro-
psycho-pharmacoepidemiology.
References
What is Pharmacoepidemiology? Pharmacoepidemiology. 2019a:1–26.
Basic principles of clinical pharmacology relevant to Pharmacoepidemiologic studies. Pharmacoe-
pidemiology. 2019b:27–43.
Basic principles of clinical epidemiology relevant to Pharmacoepidemiologic studies. Pharmacoe-
pidemiology. 2019c:44–59.
AM W. confounding by indication. Epidemiology. 1996;7(4):335–6.
Bodenheimer T, Wagner EH, Grumbach K. Improving primary care for patients with chronic
illness. JAMA. 2002a;288(14):1775–9.
Bodenheimer T, Wagner EH, Grumbach K. Improving primary care for patients with chronic
illness: the chronic care model, part 2. JAMA. 2002b;288(15):1909–14.
Bykov K, Bateman BT, Franklin JM, Vine SM, Patorno E. Association of Gabapentinoids with the
risk of opioid-related adverse events in surgical patients in the United States. JAMA Netw Open.
2020;3(12):e2031647-e.
Caniglia EC, Rebecca Z, Jacobson DL, Diseko M, Mayondi G, Lockman S, et al. Emulating a target
trial of antiretroviral therapy regimens started before conception and risk of adverse birth
outcomes. AIDS (London, England). 2018;32(1):113.
Caniglia EC, Rojas-Saunero LP, Hilal S, Licher S, Logan R, Stricker B, et al. Emulating a target trial
of statin use and risk of dementia using cohort data. Neurology. 2020;95(10):e1322–e32.
Cochran WG. Observational studies. Ames Iowa State University Press; 1972.
Daniel RM, Cousens SN, De Stavola BL, Kenward MG, Sterne JA. Methods for dealing with time-
dependent confounding. Stat Med. 2013;32(9):1584–618.
Dawid AP. Causal inference without counterfactuals. J Am Stat Assoc. 2000;95(450):407–24.
Desai RJ, Rothman KJ, Bateman BT, Hernandez-Diaz S, Huybrechts KF. A propensity score based
fine stratification approach for confounding adjustment when exposure is infrequent. Epidemi-
ology. 2017;28(2):249.
Dickerman BA, García-Albéniz X, Logan RW, Denaxas S, Hernán MA. Emulating a target trial in
case-control designs: an application to statins and colorectal cancer. Int J Epidemiol. 2020;49(5):
1637–46.
Dorn HF. Philosophy of inferences from retrospective studies. Am J Public Health Nations Health.
1953;43(6_Pt_1):677–83.
ElZarrad MK, Corrigan-Curay J. The US Food and Drug Administration’s real-world evidence
framework: a commitment for engagement and transparency on real-world evidence. Clin
Pharmacol Ther. 2019;106(1):33–5.
Farrington C. Control without separate controls: evaluation of vaccine safety using case-only
methods. Vaccine. 2004;22(15–16):2064–70.
Feinstein AR. XI. Sources of ‘chronology bias’ in cohort statistics. Clin Pharmacol Ther. 1971;12
(5):864–79.
Franklin JM, Glynn RJ, Martin D, Schneeweiss S. Evaluating the use of nonrandomized real-world
data analyses for regulatory decision making. Clin Pharmacol Ther. 2019;105(4):867–77.
Glymour MM, Gibbons LE, Gilsanz P, Gross AL, Mez J, Brewster PW, et al. Initiation of
antidepressant medication and risk of incident stroke: using the adult changes in thought cohort
to address time-varying confounding. Ann Epidemiol. 2019;35:42-7. e1.
Glynn RJ, Knight EL, Levin R, Avorn J. Paradoxical relations of drug treatment with mortality in
older persons. Epidemiology. 2001:682–9.
Causal Inference Methods in Pharmacoepidemiology 821
Greenland S, Pearl J, Robins JM. Causal diagrams for epidemiologic research. Epidemiology.
1999a:37–48.
Greenland S, Robins JM, Pearl J. Confounding and collapsibility in causal inference. Stat Sci.
1999b;14(1):29–46.
Hernán MA, Alonso A, Logan R, Grodstein F, Michels KB, Stampfer MJ, et al. Observational
studies analyzed like randomized experiments: an application to postmenopausal hormone
therapy and coronary heart disease. Epidemiology. 2008;19(6):766.
Hernan MA, McAdams M, McGrath N, Lanoy E, Costagliola D. Observation plans in longitudinal
studies with time-varying treatments. Stat Methods Med Res. 2009;18(1):27–52.
Hernán MARJ. Causal inference: what if. Boca Raton: Chapman & Hall/CRC; 2020.
Hernán MA, Robins JM. Using big data to emulate a target trial when a randomized trial is not
available. Am J Epidemiol. 2016;183(8):758–64.
Huybrechts KF, Bateman BT, Palmsten K, Desai RJ, Patorno E, Gopalakrishnan C, et al. Antide-
pressant use late in pregnancy and risk of persistent pulmonary hypertension of the newborn.
JAMA. 2015;313(21):2142–51.
Huybrechts KF, Palmsten K, Avorn J, Cohen LS, Holmes LB, Franklin JM, et al. Antidepressant use
in pregnancy and the risk of cardiac defects. N Engl J Med. 2014;370(25):2397–407.
Imai K, Van Dyk DA. Causal inference with general treatment regimes: generalizing the propensity
score. J Am Stat Assoc. 2004;99(467):854–66.
Jackson JW, Schmid I, Stuart EA. Propensity scores in pharmacoepidemiology: beyond the horizon.
Curr Epidemiol Rep. 2017;4(4):271–80.
Johnson ES, Bartman BA, Briesacher BA, Fleming NS, Gerhard T, Kornegay CJ, et al. The incident
user design in comparative effectiveness research. Pharmacoepidemiol Drug Saf. 2013;22(1):1–6.
Li X, Young JG, Toh S. Estimating effects of dynamic treatment strategies in Pharmacoepidemiologic
studies with time-varying confounding: a primer. Curr Epidemiol Rep. 2017;4(4):288–97.
Lund JL, Richardson DB, Stürmer T. The active comparator, new user study design in pharmacoe-
pidemiology: historical foundations and contemporary application. Curr Epidemiol Rep. 2015;2
(4):221–8.
Maclure M. The case-crossover design: a method for studying transient effects on the risk of acute
events. Am J Epidemiol. 1991;133(2):144–53.
Martinez C, Assimes TL, Mines D, Dell’aniello S, Suissa S. Use of venlafaxine compared with
other antidepressants and the risk of sudden cardiac death or near death: a nested case-control
study. BMJ. 2010;340:c249.
Mcevoy P, Barnes P. Using the chronic care model to tackle depression among older adults who
have long-term physical conditions. J Psychiatr Ment Health Nurs. 2007;14(3):233–8.
Moura LM, Westover MB, Kwasnik D, Cole AJ, Hsu J. Causal inference as an emerging statistical
approach in neurology: an example for epilepsy in the elderly. Clin Epidemiol. 2017;9:9.
Naimi AI, Cole SR, Kennedy EH. An introduction to g methods. Int J Epidemiol. 2017;46(2):756–62.
Newsome SJ, Keogh RH, Daniel RM. Estimating long-term treatment effects in observational data:
a comparison of the performance of different methods under real-world uncertainty. Stat Med.
2018;37(15):2367–90.
Palmsten K, Hernández-Díaz S, Huybrechts KF, Williams PL, Michels KB, Achtyes ED, et al. Use
of antidepressants near delivery and risk of postpartum hemorrhage: cohort study of low income
women in the United States. BMJ. 2013;347
Park Y, Bateman BT, Kim DH, Hernandez-Diaz S, Patorno E, Glynn RJ, et al. Use of haloperidol
versus atypical antipsychotics and risk of in-hospital death in patients with acute myocardial
infarction: cohort study. BMJ. 2018;360
Patorno E, Bohn RL, Wahl PM, Avorn J, Patrick AR, Liu J, et al. Anticonvulsant medications and
the risk of suicide, attempted suicide, or violent death. JAMA. 2010;303(14):1401–9.
Pazzagli L, Linder M, Zhang M, Vago E, Stang P, Myers D, et al. Methods for time-varying
exposure related problems in pharmacoepidemiology: an overview. Pharmacoepidemiol Drug
Saf. 2018;27(2):148–60.
Pearl J. Causal diagrams for empirical research. Biometrika. 1995;82(4):669–88.
822 L. Pazzagli and X. Li
Pearl J, Glymour M, Jewell NP. Causal inference in statistics: a primer. John Wiley & Sons; 2016.
Prentice RL. A case-cohort design for epidemiologic cohort studies and disease prevention trials.
Biometrika. 1986;73(1):1–11.
Ray WA. Evaluating medication effects outside of clinical trials: new-user designs. Am J
Epidemiol. 2003;158(9):915–20.
Robins J. A new approach to causal inference in mortality studies with a sustained exposure
period—application to control of the healthy worker survivor effect. Math Model. 1986;7
(9–12):1393–512.
Robins JM. Correcting for non-compliance in randomized trials using structural nested mean
models. Commun Stat Theory Methods. 1994;23(8):2379–412.
Robins JM, Hernán MA. Estimation of the causal effects of time-varying exposures. Longitudinal
Data Anal. 2009;553:599.
Rosenbaum P, Armitage P, Colton T. Encyclopedia of biostatistics. 2005.
Rosenbaum PR, Rubin DB. The central role of the propensity score in observational studies for
causal effects. Biometrika. 1983;70(1):41–55.
Rubin DB. Comment: which ifs have causal answers. J Am Stat Assoc. 1986;81(396):961–2.
Rubin DB. Bayesian inference for causal effects: the role of randomization. Ann Stat. 1978:34–58.
Rubin DB. Assignment to treatment group on the basis of a covariate. J Educ Stat. 1977;2(1):1–26.
Rubin DB. Estimating causal effects of treatments in randomized and nonrandomized studies.
J Educ Psychol. 1974;66(5):688–701.
Schneeweiss S, Eichler HG, Garcia-Altes A, Chinn C, Eggimann AV, Garner S, et al. Real world
data in adaptive biomedical innovation: a framework for generating evidence fit for decision-
making. Clin Pharmacol Ther. 2016;100(6):633–46.
Schneeweiss S, Patrick AR, Solomon DH, Mehta J, Dormuth C, Miller M, et al. Variation in the risk
of suicide attempts and completed suicides by antidepressant agent in adults: a propensity score–
adjusted analysis of 9 years’ data. Arch Gen Psychiatry. 2010;67(5):497–506.
Schneeweiss S, Patrick AR, Sturmer T, Brookhart MA, Avorn J, Maclure M, et al. Increasing levels
of restriction in pharmacoepidemiologic database studies of elderly and comparison with
randomized trial results. Med Care. 2007;45(10 Supl 2):S131–42.
Schuemie MJ, Ryan PB, Man KK, Wong IC, Suchard MA, Hripcsak G. A plea to stop using the
case-control design in retrospective database studies. Stat Med. 2019;38(22):4199–208.
Suissa S. Effectiveness of inhaled corticosteroids in chronic obstructive pulmonary disease: immor-
tal time bias in observational studies. Am J Respir Crit Care Med. 2003;168(1):49–53.
Suissa S. The case-time-control design. Epidemiology. 1995:248–53.
Sun JW, Hernández-Díaz S, Haneuse S, Bourgeois FT, Vine SM, Olfson M, et al. Association of
Selective Serotonin Reuptake Inhibitors with the risk of type 2 diabetes in children and
adolescents. JAMA Psychiat. 2021;78(1):91–100.
VanderWeele TJ. Concerning the consistency assumption in causal inference. Epidemiology.
2009;20(6):880–3.
Wang PS, Schneeweiss S, Avorn J, Fischer MA, Mogun H, Solomon DH, et al. Risk of death in
elderly users of conventional vs. atypical antipsychotic medications. N Engl J Med. 2005;353
(22):2335–41.
Whitaker HJ, Paddy Farrington C, Spiessens B, Musonda P. Tutorial in biostatistics: the self-
controlled case series method. Stat Med. 2006;25(10):1768–97.
Neuropsychopharmacotherapy: Guidelines
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 824
Guideline Developing Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 825
Existing Guidelines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 825
Potential Benefits from Clinical Guidelines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 826
Limitations of Clinical Guidelines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 826
Review of the Important Guidelines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 827
Guidelines for the Treatment of Schizophrenia (Displayed in Table 1) . . . . . . . . . . . . . . . . . . . . 827
Guidelines for the Treatment of Major Depression (Displayed in Table 2) . . . . . . . . . . . . . . . . 832
Guidelines for the Treatment of Bipolar Disorder (Displayed in Tables 3, 4, and 5) . . . . . . 840
Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 849
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 849
Abstract
Practicing evidence-based medicine (EBM) requires the ability to evaluate rele-
vant evidence for the purpose of making an evidence-based treatment decision.
The need to standardize different therapeutic approaches has encouraged the
development of guidelines to advice on the treatment, management, and assess-
ment of psychiatric conditions. Treatment guidelines serve the purpose of gath-
ering and evaluating research data and transform them into comprehensive
treatment algorithm. Based on the current status of evidence, algorithms have
been combined to form a unified guideline for management. Clinical practice
guidelines form the basis of standard clinical practice for all disciplines of
medicine, including psychiatry. However, a significant challenge for psychiatry
is the incorporation of this new knowledge into the daily work of clinicians.
Moreover, there are several potential risks related with use of algorithms and
guidelines. The evidence might be insufficient and the consensus panels might
express biased opinions. They are often not read or followed because of poor
quality or because of barriers to implementation due to either lack of agreement or
ambiguity. We briefly review practice guidelines for the treatment of schizophre-
nia, major depression, and bipolar disorder.
Introduction
During the past 150 years, two significant factors have shaped the modern health
system: first, the growth of scientific knowledge, and, second, the growth of public
acceptance of disease control. Throughout the years, much has been learned from
efficacy and effectiveness of various treatments – biologic, psychotherapeutic, and
social (Hanlon 2012).
Guidelines have existed for centuries, and recommendations about appropriate
care can be found even in ancient writings. However, until recent, they had no
provided evidence-based recommendations for the assessment and treatment of
psychiatric disorders. Moreover, the process used in their review was not
documented, and there were no process of revision (McIntyre 2002).
Only over the past few decades, the role of guidelines was described variously. As
defined by the Institute of Medicine (Amsterdam et al. 1996) report, practice
guidelines are “systematically developed statements to assist practitioner and patient
decisions about appropriate health care for specific clinical circumstances” (Field
et al. 1992). The IOM reported two objectives: first, to encourage constructive
expectations from guidelines, and, second, to promote the care and rigor for their
development, application, evaluation, and revision. To accept criteria for validity of
guidelines, the essential elements defined by the IOM were set out first. These
recommended attributes included validity, reliability, clinical applicability, clinical
flexibility, clarity, multidisciplinary process, scheduled review, and documentation
(Field and Lohr 1990).
Another definition of clinical practice guidelines (CPGs) states “CPGs are able to
enhance clinician and patient decision making by clearly describing and appraising
the scientific evidence and reasoning (the likely benefits and harms) behind clinical
recommendations, making them relevant to the individual patient encounter” (Field
and Lohr 1990). Modern guidelines are developed by professional organizations,
government, even insurance companies, and third parties, as well by providers of
care. Part of these guidelines are evidence-based, part of them based on expert
opinions or based on opinions or one more authors (Woolf 1992; Woolf et al. 1999).
In 1988, the American Medical Association organized a Practice Guideline
Partnership comprised of 14 specialty organizations including the American
Psychiatric Association (APA). This partnership also defined “good” guidelines
and identified five criteria. They: (a) are developed by physicians in active clinical
practice; (b) integrate relevant research and clinical expertise; (c) describe specific
treatment approaches, including indicators, efficacy, safety, and alternative treatment
strategies; (d) are reviewed and revised at regular intervals not longer than 5 years;
Neuropsychopharmacotherapy: Guidelines 825
and (e) after approval, are widely disseminated. They set the standard of care and
training for health professionals and they also identify priority areas for further
research, since they are based primarily on the available evidence, but also in
areas where evidence is not available, on expert opinion. Clinicians, policymakers,
and payers see guidelines as a tool for making care more consistent and efficient and
for closing the gap between what clinicians do and what scientific evidence supports
(Woolf et al. 1999).
A major challenge for psychiatry is to implement evidence-based knowledge
into the daily work and practice of clinicians (McIntyre 2002). A good guideline
should be able to identify the key decisions (e.g., diagnosis, assessment strategy, and
treatment choice), review the relevant, valid evidence on the benefits, risks, and costs
of alternative decisions, and present recommendations in a concise, updated format
(Woolf 1992; McIntyre 2002; Hasan et al. 2019).
Existing Guidelines
The principal benefit from guidelines is to improve the quality of care received by
patients. For patients, the greatest benefit that could be achieved by guidelines is to
improve health outcomes. Guidelines can also improve the consistency of care;
studies around the world show that the frequency with which procedures are
performed varies dramatically among doctors, specialties, and even geographical
regions. Clinical guidelines can help patients by influencing public policy.
Moreover, they call attention to under-recognized health problems, clinical services,
and preventive interventions and to neglected patient populations and high-risk
groups. Guidelines aim to assist to clinicians and also policymakers (Woolf et al.
1999; Kredo et al. 2016).
Clinical guidelines can improve the quality of clinical decisions. They offer
explicit recommendations for clinicians who are uncertain about how to proceed,
improve the consistency of care, and provide authoritative recommendations
that reassure practitioners about the appropriateness of their treatment policies.
Guidelines based on a critical appraisal of scientific evidence clarify which inter-
ventions are of proved benefit and document the quality of the supporting data
(McIntyre 2002; Stasevic et al. 2019).
Health care systems that provide services, government bodies and private insurers
that pay for them, have found that clinical guidelines may be effective in improving
efficiency (Carnett 1999; Woolf et al. 1999).
to the logic of the randomized controlled trial (RCT), patients’ experience of and
active participation in psychotherapy systematically influence the content of the
intervention they receive. It has therefore been argued that RCTs are not optimal
methodologies for validating psychotherapies, and instead process evaluation is
required (Moore et al. 2015). Therefore while RCTs carry significant weight, it
must be remembered that they and their meta-analyses are of variable quality and
validity. The scientific literature on psychological therapies is also influenced by
publication bias that positive outcome studies will be published (Driessen et al.
2013, 2015) and researcher allegiance effects (Munder et al. 2013).
Until the present time, it does not appear that the use of practice guidelines has
increased medical liability. Practice guidelines might have the potential to reduce the
number of malpractice cases and the costs of settling them. However, for practice
guidelines to exert any influence, they must be assumed to be developed for
conditions or procedures that frequently lead to events for which negligence claims
are filed; widely accepted in the medical profession; fully integrated into clinical
practice; and straightforward and readily interpreted in a litigation setting. Because
the validity of each of these assumptions can be questioned, the idea that inserting
practice guidelines into the existing litigation process will generate large savings is
overoptimistic. Another issue is that the recommendations of a guideline cannot be
followed because of a lack of adequate resources (Garnick et al. 1991; McIntyre
2002).
Finally, the guidelines are made with a methodology that takes little account of
the opinions of experts in daily practice; they refer to ideally selected patients in ideal
care settings; they do not take into account the complex patients with comorbid
psychiatric conditions in daily practice; they can be influenced by economic prior-
ities; they are different in the indications between them; often they change their
principles over time; they can be used for a defensive psychiatry that does not
privilege the patient’s benefit (Nivoli et al. 2017).
psychopharmacology; CINP, The International College of Neuropsychopharmacology. SGAs, second-generation antipsychotics; FGAs: first-generation antipsychotics
Neuropsychopharmacotherapy: Guidelines 829
For the treatment of the acute phase of schizophrenia, the guideline recommends
to select the first-generation or the second-generation antipsychotic medication, and
the selection of medication should be guided by the patient’s previous experience
with antipsychotics, including the degree of symptom response, past experience of
side effects, and preferred route of administration. In addition, the psychoeducation
of the patient and the family members is advised during the acute phase of the
treatment. The treatment duration of the acute phase is determined as 4–6 weeks.
During the stabilization phase, the guideline recommends to continue the same dose
of the antipsychotics and to provide psychoeducation of the patient and the family
members up to 6 months with a goal to reduce stress on the patient and provide
support to minimize the likelihood of relapse.
There is a clear suggestion that antipsychotics and psychosocial therapies
should be used in the stable phase up to 1–1.5 years after a first episode, up to
5–10 years in case of two or more episodes, and indefinite for multiple prior
episodes or more than two episodes in 5 years; this should ensure that symptom
remission or control is sustained. The psychosocial treatments that have demon-
strated effectiveness during the stable phase include family intervention,
supported employment, assertive community treatment, skills training, and
cognitive behaviorally oriented psychotherapy.
The guidelines for the treatment of patients with major depressive disorder (MDD)
are based on the criteria provided in the Diagnostic and Statistical Manual of Mental
Disorders of the American Psychiatric Association, 4th or 5th Edition (DSM–IV or
V) (APA 1994, 2013). All guidelines consider MDD, though some consider persis-
tent depressive disorder or dysthymia. They use a severity-based model of depres-
sion (e.g., subthreshold, mild, moderate, and severe). The approach to modelling of
the MDD varies in parameters of depressive subtype (e.g., melancholic, catatonic,
psychotic, atypical, seasonal, anxious), chronicity (e.g., chronic or recurrent MDD,
dysthymia, PDD), complexity (e.g., “uncomplicated,” “complex,” treatment-resis-
tant depression), and comorbidity (e.g., psychiatric or medical). Some guidelines
consider multiple domains concurrently. The guidelines will be analyzed in depth
taking into account their specific recommendations with a focus on the pharmaco-
logical treatment of MDD.
depressive
episodes
CANMAT SSRIs, TCAs, MAOIs, Psychosocial Medication – ADs and Acute phase:
2016 SNRIs, quetiapine, reboxetine and and antipsychotics or 6–12 weeks
agomelatine, trazodone, psychological psychological ECT Maintenance
bupropion, moclobemide, treatments or treatments With catatonic phase:
mirtazapine, selegiline, medication features: 6–9 month. If
vortioxetine levomilnacipran Benzodiazepines risk factors for
vilazodone recurrence,
extend
treatment to 2
or more years
NICE SSRIs – – Psychosocial ADs and ADs and ADs and Acute phase:
2018 and psychological psychological antipsychotics and till remission
psychological treatments treatments or psychological Continuation
treatments or ECT treatments or ECT phase:
ADs 6 months after
remission
833
(continued)
834
Table 2 (continued)
Pharmacological treatment Severity of major depressive disorder
Moderate to
severe without
First-line Second-line Third-line Mild to psychotic Severe with Duration of
Guideline medication medication medication moderate features Severe psychotic features treatment
Maintenance
phase: 2 years
if they are at
risk of relapse
RANZCP SSRIs, SNRIs, TCAs, MAOIs, Psychosocial ADs and ADs and ADs and Acute phase:
2015 NARIs serotonin reversible and psychological psychological antipsychotics 6–12 weeks.
(reboxetine), modulator MAOIs psychological treatments treatments ADs, ECT Maintenance
NaSSAs, (vortioxetine) (Moclobemide), treatments or phase:
NDRIs SARIs ADs 6 months and
(Bupropion), Trazadone up to 1 year
agomelatine
WFSPB No single – – SSRIs and ADs (as for – TCAs, SSRIs, Acute phase:
2013, class of ADs other “newer mild) or SNRIs, MAOIs, 8–10 weeks
2015 has proven to ADs” (except psychosocial complex Continuation
be more reboxetine) or and psychotherapeutic phase: 6–
effective psychosocial psychological treatment, 9 months
and treatments and combination Maintenance
psychological ADs treatments or ECT phase: from
3 years to
J. Vrublevska and L. Renemane
treatments and lifetime, if
ADs residual
symptoms or
psychotic
features;
5–10 years,
if greater risk
of relapse
BAP SSRIs TCAs, MAOIs – Psychosocial ADs and – ADs and Acute phase:
2015 and psychological antipsychotics, 6–8 weeks.
psychological treatments psychological Maintenance
treatments or treatments or ECT phase after full
ADs remission:
6–9 months, if
Neuropsychopharmacotherapy: Guidelines
low risk of
relapse, 1 year
if any
increased risk
of relapse,
2 years in
higher-risk
patients
APA, American Psychiatric Association; CANMAT, Canadian Network for Mood and Anxiety Treatments; NICE, National Institute for Health and Care
Excellence; RANZCP, Royal Australian and New Zealand College of Psychiatrists; WFSPB, World Federation of Societies of Biological Psychiatry; BAP,
British Association of Psychopharmacology; SSRIs: selective serotonin reuptake inhibitors; NARIs: noradrenaline reuptake inhibitors; NaSSAs: noradrenaline
and specific serotonergic antidepressants; NDRIs: noradrenaline dopamine reuptake inhibitors; SNRIs: serotonin and noradrenaline reuptake inhibitors; TCAs:
tricyclic antidepressants; MAOIs, monoamine oxidase inhibitors; SARIs, serotonin antagonist and reuptake inhibitors; ADs, antidepressants; ECT, electrocon-
vulsive therapy
835
836 J. Vrublevska and L. Renemane
assessment, support, psychoeducation, active monitoring, and referral for further risk
assessment, monitoring, and interventions. The second step focuses on persons with
recognized depression characterized by persistent subthreshold depressive symp-
toms or mild to moderate depression. Low-intensity psychosocial interventions,
psychological interventions, medication, and referral for further assessment and
interventions are recommended for the second step. The guidelines offer low-
intensity psychosocial interventions such as individual guided self-help based on
the principles of cognitive behavioral therapy (CBT), computerized cognitive behav-
ioral therapy (CCBT), and a structured group physical activity program. Drug
treatment is recommended for people either with a history of moderate or severe
depression or initial presentation of subthreshold depressive symptoms that have
been present at least for 2 years, or subthreshold depressive symptoms or mild
depression that persist(s) after other interventions. St John’s wort may be of benefit
in mild or moderate depression. The third step provides recommendations for
patients with persistent subthreshold depressive symptoms or mild to moderate
depression with inadequate response to initial interventions and moderate and severe
depression. The treatment options include antidepressants (SSRI), high-intensity
psychological interventions, combined treatments, collaborative care, and referral
for further assessment and interventions. Step four focuses on patients with severe
and complex depression, who are at significant risk of self-harm and have psychotic
symptoms. For this severity level of depressed patients, the guideline recommends
medication, high-intensity psychological interventions, electroconvulsive therapy
(ECT), crisis service, combined treatments, and multiprofessional and inpatient
care. For the people who have depression with psychotic symptoms, it is
recommended to augment the current treatment plan with antipsychotic medication.
The NICE guidelines specify a need to continue antidepressant treatment for at
least 6 months after remission taking into account episodes of depression. For those
who are at risk of relapse, it is advised to continue medical treatment for at least
2 years keeping the level of medication at which acute treatment was effective.
The APA Practical Guideline for the Treatment of Patients with Major
Depressive Disorder
The APA Practice Guideline for the Treatment of Patients with MDD, 3rd Edition,
was originally published in October 2010 (APA 2010). This guideline provides
evidence-based recommendations for the assessment and treatment of MDD and is
intended to assist in clinical decision-making by presenting systematically devel-
oped patient care strategies in a standardized format.
In the APA guideline that was produced in 2010, there are three categories of
endorsement identified for each recommendation. These categories represent vary-
ing levels of clinical confidence. The first level address recommendations with
substantial clinical confidence, the second level is attributed to recommendations
with moderate clinical confidence, and the third one – recommendations made on the
basis of individual circumstances.
The guideline suggests that acute phase treatment may include pharmacotherapy,
depression-focused psychotherapy, the combination of medications and psychotherapy,
Neuropsychopharmacotherapy: Guidelines 837
The 2016 CANMAT guidelines, to treat the acute phase of MDD, suggest
following first-line recommendations: SSRIs, SNRIs, agomelatine, bupropion,
mirtazapine, or vortioxetine. Second line includes TCAs, quetiapine, trazodone,
moclobemide, selegiline, levomilnacipran, and vilazodone; the third line of pharma-
cological treatment includes MAOIs and reboxetine.
Regarding the severity-based model of depression, they recommend to choose
between psychosocial and psychological treatments and medication for the mild to
moderate level of depression. For moderate to severe depression, a combination of
pharmacotherapy and psychological treatments is advised. The guidelines suggest
using a combination of antidepressants and antipsychotics or ECT for the treatment
of severe depression with psychotic features. For depression with catatonic features,
they recommend benzodiazepines. These guidelines define specific types of the
psychological interventions for acute and maintenance phases. The psychological
interventions are categorized in three lines and are shown in Table.
In terms of the duration of treatment, the guidelines recommend to provide the
treatment for 6–12 weeks in the acute phase and to continue it from 6 up to 9 months
in the maintenance phase. If the patient has risk factors for recurrence of disease, the
guidelines advise to extend the treatment up to 2 or more years.
For the severe cases of MDD with psychotic features, the guideline recommends to
combine antidepressants with antipsychotics or indicate ECT approach.
This guideline specifies the duration of acute phase treatment from 6 up to
12 weeks and recommends continuation treatment in the maintenance phase for
6 months and up to 1 year, to prevent the relapse.
The quality of evidence was graded from category I to category IV, and strength
of recommendation was graded from A–D levels and S level was based on standard
of good practice.
This guideline cover the nature and detection of depressive disorders, acute
treatment with antidepressant drugs, choice of drug versus alternative treatment,
practical issues in prescribing and management, next-step treatment, relapse preven-
tion, treatment of relapse, and stopping treatment. The first-line antidepressant
treatment includes SSRIs. As second line, TCAs and MAOIs were recommended.
The guideline recommended antidepressants as a first-line treatment option for
moderate and severe MDD in adults irrespective of environmental factors and depres-
sion symptom profile, and for depression of any severity that has persisted for 2 years
or more. The guideline suggested antidepressants as a treatment option in short-
duration mild MDD and should be considered if there is a prior history of moderate
to severe recurrent depression or the depression persists for more than 2–3 months.
In terms of duration of treatment, they considered to continue the treatment at
least 6–9 months after full remission in patients at lower risk of relapse (e.g., first-
episode patients without other risk factors), duration in other cases should be tailored
to the individual relapse risk. They also note that a duration of at least 1 year after full
remission should to be provided in patients with any increased risk of relapse. In
higher-risk patients (e.g., more than five lifetime episodes and/or two episodes in the
last few years), the duration of treatment at least 2 years should be advised and
for most long-term treatment should be considered.
Table 3 (continued)
Acute treatment Maintenance treatment
First-line Second-line First-line Second-line
Guidelines treatment treatment treatment treatment
risperidone Aripiprazole + Risperidone Olanzapine (or
Valproate Li/Val (including RLAI) aripiprazole) + MS
Haloperidol + Li/
Val Olanzapine +
Li/Val
Allopurinol + Li
Val + TAP
Val + celecoxib
BAP Not taking Li/Val + dopamine Li Combination
2016 long-term antagonist/ therapy ADs
treatment: Aripiprazole Clozapine
Haloperidol Maintenance ECT
Olanzapine
Risperidone
Aripiprazole
Quetiapine
Val Li
Taking long-
term
treatment:
optimization
Haloperidol
Olanzapine
Risperidone
Quetiapine
BAP, The British Association of Psychopharmacology; CANMAT, Canadian Network for Mood
and Anxiety Treatments Clinical Guidelines for the Management of Patients with Bipolar Disorder;
CINP-BD The International College of Neuropsychopharmacology Treatment Guidelines for
Bipolar Disorder; NICE, National Institute for Health and Clinical Excellence Clinical Guideline
for Bipolar Disorder; WFSPB, The World Federation of Societies of Biological Psychiatry
(WFSBP) Guidelines for the Biological Treatment of BD; RANZCP, Royal Australian and New
Zealand College of Psychiatrists clinical practice guidelines for mood disorders; Ads, antidepres-
sants, Li, lithium; Val, valproate; ECT, electroconvulsive therapy; DVP, divalproex; TAP, typical
antipsychotics; RLAI, risperidone long-acting injection; MS, mood stabilizers
also recommended to consider ECT for patients with high suicidal risk, treatment
resistance psychosis, severe depression during pregnancy, or life-threatening inanition.
For the maintenance phase, these guidelines recommend the choice of treatment based
on predominant polarity. Maintenance treatment is recommended with lithium,
aripiprazole, quetiapine, valproate, olanzapine if the predominant polarity is mania,
and lamotrigine or quetiapine if predominant polarity is depression.
Canadian Network for Mood and Anxiety Treatments Guidelines for the
Management of Patients with Bipolar Disorder (CANMAT)
The CANMAT previously published guidelines for the first time in 1997
(Kusumakar et al. 1997). These guidelines were revised in 2005, along with
Neuropsychopharmacotherapy: Guidelines 843
Table 4 (continued)
Pharmacological treatment Maintenance treatment
First-line Second-line First-line Second-line
Guidelines treatment medication treatment medication
FEWP +
carbamazepine
CINP-BP Quetiapine Val Predominant polarity Add FLX or Li
2017 Lurasidone Li Li Carbamazepine
OFC Lurasidone+MS Aripiprazole + Li
Consider add Modafinil + MS Olanzapine Quetiapine +
on CBT Pramipexole + Paliperidone Li/ Val
MS Quetiapine Olanzapine or
Pioglitazone + Li Risperidone aripiprazole +
Carbamazepine (including RLAI) MS
plus FEWP
Add escitalopram
or FLX
BAP 2016 Not taking Add antimanic Li Lamotrigine
long-term agent Quetiapine
treatment: ECT Lurasidone
Quetiapine
Olanzapine
Olanzapine +
FLX
Lurasidone
Lamotrigine as
combination
BAP, The British Association of Psychopharmacology; CANMAT, Canadian Network for Mood
and Anxiety Treatments Clinical Guidelines for the Management of Patients with Bipolar Disorder;
CINP-BD The International College of Neuropsychopharmacology Treatment Guidelines for
Bipolar Disorder; NICE, National Institute for Health and Clinical Excellence Clinical Guideline
for Bipolar Disorder; WFSPB, The World Federation of Societies of Biological Psychiatry
(WFSBP) Guidelines for the Biological Treatment of BD; RANZCP, Royal Australian and New
Zealand College of Psychiatrists clinical practice guidelines for mood disorders; AD, antidepres-
sant, Li, lithium; Val, valproate; DVP, divalproex; TAP, typical antipsychotics; RLAI, risperidone
long-acting injection; MS, mood stabilizers; FLX, fluoxetine; Adj., adjunctive; FEWP, herbal
remedy “Free and easy wanderer plus”; OFC, olanzapine-fluoxetine combination
Table 5 (continued)
Pharmacological treatment Maintenance treatment
First-line Second-line First-line
Guidelines treatment medication treatment Second-line Treatment
BAP 2016 Same as for Same as for Same as for Same as for mania
mania mania mania
BAP, The British Association of Psychopharmacology; CANMAT, Canadian Network for Mood
and Anxiety Treatments Clinical Guidelines for the Management of Patients with Bipolar Disorder;
CINP-BD The International College of Neuropsychopharmacology Treatment Guidelines for
Bipolar Disorder; NICE, National Institute for Health and Clinical Excellence Clinical Guideline
for Bipolar Disorder; WFSPB, The World Federation of Societies of Biological Psychiatry
(WFSBP) Guidelines for the Biological Treatment of BD; RANZP, Royal Australian and New
Zealand College of Psychiatrists clinical practice guidelines for mood disorders; AD, antidepres-
sant, Li, lithium; Val, valproate; DVP, divalproex; TAP, typical antipsychotics; RLAI, risperidone
long-acting injection; MS, mood stabilizers; FLX, fluoxetine; Adj., adjunctive; OFC, olanzapine-
fluoxetine combination; MM, manic mixed; DM, depressed mixed; MxIE, mixed index episode
emergent switch with pharmacological agents were categorized, ranging from Level
1 (the most powerful evidence) to Level 4 (the weakest, uncontrolled trial, anecdotal
reports, or expert opinion). Treatment recommendations were categorized into four
levels based on the strength of supporting evidence. These guidelines recommend
monotherapy of quetiapine, asenapine, paliperidone, risperidone, and cariprazine,
and adjunctive quetiapine or aripiprazole or risperidone or asenapine with lithium or
divalproex as first-line treatment strategies for treating mania. As second-line treat-
ment, they recommend olanzapine, carbamazepine, ziprasidone, haloperidol, and
combination of olanzapine and lithium/divalproex and combination of lithium and
divalproex. As the first-line treatment of bipolar depression, the CANMAT recom-
mends monotherapy of quetiapine, lithium, lamotrigine, lurasidone, and combina-
tion of lurasidone and lithium/divalproex. As second-line treatment, they
recommend divalproex, cariprazine, olanzapine-fluoxetine, adjunctive SSRI or
bupropion, and ECT.
offered five levels of recommendations for treatment. For the treatment of acute
mania, the CINP-BD guidelines suggest aripiprazole, asenapine, cariprazine,
paliperidone, quetiapine, risperidone, or valproate monotherapy. If the treatment
during the first step fails, CINP suggests olanzapine, lithium, carbamazepine, halo-
peridol, or ziprasidone monotherapy or combinations of lithium or valproate plus
asenapine, aripiprazole, haloperidol, olanzapine or lithium plus allopurinol,
valproate plus a first-generation antipsychotic, valproate plus celecoxib. For the
treatment of acute bipolar depression, the CINP-BD recommends quetiapine or
lurasidone or consider CBT as add-on to medication. For the second step of the
treatment of acute depression, they suggest monotherapy with olanzapine or OFC,
combination of a mood stabilizer with lurasidone, modafinil, or pramipexole, lithium
plus lamotrigine or pioglitazone, or add escitalopram or fluoxetine to ongoing
therapy. In the maintenance phase, the CINP suggests to take into account predom-
inant polarity and start with lithium, aripiprazole, olanzapine, paliperidone,
quetiapine, or risperidone, including risperidone long-acting injectable, mono-
therapy depending on predominant polarity (Fountoulakis et al. 2016a).
published in 2009 for mania (Grunze et al. 2009), in 2010 for acute depression
(Grunze et al. 2010), for maintenance treatment in 2013 (Grunze et al. 2013), and
for the acute and long-term treatment of mixed states in 2018 (Grunze et al. 2018).
Their scientific rigor was categorized into six levels of evidence (A–F). The
scientific evidence was finally assigned different grades of recommendation to
ensure practicability.
For the treatment of acute mania based on the category of evidence and recom-
mendation grade as first-choice agents, the WFSBP recommends aripiprazole,
risperidone, valproate, and ziprasidone. As second choice, asenapine, haloperidol,
olanzapine, quetiapine, carbamazepine, and lithium were suggested.
For the treatment of acute depression, the WFSB guidelines as first-choice
treatment recommend only quetiapine, as second-choice agents fluoxetine,
olanzapine, lamotrigine, valproate as monotherapy. As combination and augmenta-
tion treatments, they suggest lamotrigine plus lithium, olanzapine fluoxetine
combination, modafinil plus ongoing treatment, N-acetylcysteine plus lithium or
valproate, FEWP + carbamazepine.
During the maintenance phase, they recommend aripiprazole, lithium, lamotrigine,
quetiapine but not all of them prevent episodes of both depression and mania.
For the treatment of manic symptoms in bipolar mixed states appeared responsive
to treatment with several atypical antipsychotics, the best evidence had olanzapine.
For depressive symptoms, addition of ziprasidone to treatment as usual may be
beneficial; however, the evidence base is much more limited than for the treatment of
manic symptoms. Valproate and lithium should also be considered for recurrence
prevention.
Conclusions
The guidelines that were discussed in the chapter provide the evidence-based
recommendations that grow up from systematic process of gathering and critical
evaluation of the research data regarding a particular treatment. Randomized control
studies, cohort studies, observational case studies, and retrospective studies
were analyzed and categorized in the guidelines to prepare the levels of evidence.
However, a different level of agreement in recommendations for managing psychi-
atric disorders was found across the guidelines. Therefore, the clinician should also
recognize that there is no perfect treatment practice guideline that applies to all
patients. There is a need for internationally acceptable recommendations to be
developed and then set the framework for further development on a national basis.
International organizations such as the WPA or WHO may help formulate a unified
guideline, which may then be modified to meet the needs. This need, however, must
not underestimate the limits of the guidelines and must lead the specialist to a critical
acceptance of them in everyday clinical practice.
References
American Psychiatric Association (APA). Practice guideline for the treatment of patients with major
depressive disorder. 3rd ed. Washington, DC: APA; 2010.
Amsterdam JD, Fava M, Maislin G, Rosenbaum J, Hornig-Rohan M. TRH stimulation test as a
predictor of acute and long-term antidepressant response in major depression. J Affect Disord.
1996;38(2–3):165–72.
APA. Diagnostic and statistical manual of mental disorders. 4th ed. Washington: American
Psychiatric Association; 1994.
APA. Diagnostic and statistical manual of mental disorders. 5th ed. Arlington: American Psychi-
atric Association; 2013.
APA. American Psychiatric Association Practice Guidelines. https://psychiatryonline.org/
guidelines
Barnes TR. Evidence-based guidelines for the pharmacological treatment of schizophrenia: recom-
mendations from the British Association for Psychopharmacology. J Psychopharmacol (Oxford,
England). 2011;25(5):567–620.
Bauer M, Pfennig A, Severus E, Whybrow PC, Angst J, Moller HJ. World Federation of Societies of
Biological Psychiatry (WFSBP) guidelines for biological treatment of unipolar depressive
disorders, part 1: update 2013 on the acute and continuation treatment of unipolar depressive
disorders. World J Biol Psychiatry. 2013;14(5):334–85.
Bauer M, Severus E, Kohler S, Whybrow PC, Angst J, Moller HJ. World Federation of Societies of
Biological Psychiatry (WFSBP) guidelines for biological treatment of unipolar depressive
disorders. Part 2: maintenance treatment of major depressive disorder-update 2015. World J
Biol Psychiatry. 2015;16(2):76–95.
Brouwers MC, Kho ME, Browman GP, Burgers JS, Cluzeau F, Feder G, et al. AGREE II: advancing
guideline development, reporting and evaluation in health care. CMAJ. 2010;182(18):E839–42.
850 J. Vrublevska and L. Renemane
Carnett WG. Clinical practice guidelines: a tool to improve care. Qual Manag Health Care.
1999;8(1):13–21.
CINP. The International College of Neuropsychopharmacology. Guidelines. https://cinp.org/Guide
lines. Accessed April 2020.
Cleare A, Pariante CM, Young AH, Anderson IM, Christmas D, Cowen PJ, et al. Evidence-based
guidelines for treating depressive disorders with antidepressants: a revision of the 2008 British
Association for Psychopharmacology guidelines. J Psychopharmacol (Oxford, England).
2015;29(5):459–525.
Coleman K, Norris S, Weston A, Grimmer-Somers K, Hilier S, Merlin T, et al. NHMRC additional
levels of evidence and grades for recommendations for developers of guidelines. Evidence
Translation Section, and supported by National Institute of Clinical Studies Officers of the
NHMRC. Australia; 2009. https://www.mja.com.au/sites/default/files/NHMRC.levels.of.evi
dence.2008-09.pdf. Accesed April 2020.
Driessen E, Van HL, Don FJ, Peen J, Kool S, Westra D, et al. The efficacy of cognitive-behavioral
therapy and psychodynamic therapy in the outpatient treatment of major depression: a random-
ized clinical trial. Am J Psychiatry. 2013;170(9):1041–50.
Driessen E, Hollon SD, Bockting CLH, Cuijpers P, Turner EH. Does publication bias inflate the
apparent efficacy of psychological treatment for major depressive disorder? A systematic review
and meta-analysis of US National Institutes of Health-Funded Trials. Plos One. 2015;10(9):
e0137864.
Field MJ, Lohr KN. Clinical practice guidelines: directions for a new program. Washington, DC:
National Academies Press 1990.
Field MJ, Lohr KN, Institute of Medicine. Guidelines for clinical practice : from development to
use. Washington, DC: National Academies Press; 1992.
Forsner T, Hansson J, Brommels M, Wistedt AA, Forsell Y. Implementing clinical guidelines in
psychiatry: a qualitative study of perceived facilitators and barriers. BMC Psychiatry.
2010;10:8.
Fountoulakis KN. Treatment guidelines bipolar disorder: an evidence based guide to manic
depression. 2015.
Fountoulakis KN, Grunze H, Vieta E, Young A, Yatham L, Blier P, et al. The International College
of Neuro-Psychopharmacology (CINP) treatment guidelines for bipolar disorder in adults
(CINP-BD-2017), part 3: the clinical guidelines. Int J Neuropsychopharmacol.
2016a;20(2):180–95.
Fountoulakis KN, Yatham L, Grunze H, Vieta E, Young A, Blier P, et al. The International College
of Neuro-Psychopharmacology (CINP) treatment guidelines for bipolar disorder in adults
(CINP-BD-2017), part 2: review, grading of the evidence, and a precise algorithm. Int J
Neuropsychopharmacol. 2016b;20(2):121–79.
Fountoulakis KN, Young A, Yatham L, Grunze H, Vieta E, Blier P, et al. The International
College of Neuropsychopharmacology (CINP) treatment guidelines for bipolar disorder in
adults (CINP-BD-2017), part 1: background and methods of the development of guidelines.
Int J Neuropsychopharmacol. 2016c;20(2):98–120.
Fountoulakis KN, Vieta E, Young A, Yatham L, Grunze H, Blier P, et al. The International
College of Neuropsychopharmacology (CINP) treatment guidelines for bipolar disorder in
adults (CINP-BD-2017), part 4: unmet needs in the treatment of bipolar disorder and recom-
mendations for future research. Int J Neuropsychopharmacol. 2017;20(2):196–205.
Galletly C, Castle D, Dark F, Humberstone V, Jablensky A, Killackey E, et al. Royal Australian and
New Zealand College of Psychiatrists clinical practice guidelines for the management of
schizophrenia and related disorders. Aust N Z J Psychiatry. 2016;50(5):410–72.
Garnick DW, Hendricks AM, Brennan TA. Can practice guidelines reduce the number and costs of
malpractice claims? JAMA. 1991;266(20):2856–60.
Goodwin GM. Evidence-based guidelines for treating bipolar disorder: revised second edition –
recommendations from the British Association for Psychopharmacology. J Psychopharmacol
(Oxford, England). 2009;23(4):346–88.
Neuropsychopharmacotherapy: Guidelines 851
Goodwin GM, Young AH. The British Association for Psychopharmacology guidelines for treat-
ment of bipolar disorder: a summary. J Psychopharmacol (Oxford, England). 2003;17(4
Suppl):3–6.
Goodwin GM, Haddad PM, Ferrier IN, Aronson JK, Barnes T, Cipriani A, et al. Evidence-based
guidelines for treating bipolar disorder: Revised third edition recommendations from the
British Association for Psychopharmacology. J Psychopharmacol (Oxford, England).
2016;30(6):495–553.
Grunze H, Kasper S, Goodwin G, Bowden C, Baldwin D, Licht R, et al. World Federation of
Societies of Biological Psychiatry (WFSBP) guidelines for biological treatment of bipolar
disorders. Part I: treatment of bipolar depression. World J Biol Psychiatry. 2002;3(3):115–24.
Grunze H, Kasper S, Goodwin G, Bowden C, Baldwin D, Licht RW, et al. The World Federation of
Societies of Biological Psychiatry (WFSBP) guidelines for the biological treatment of bipolar
disorders, part II: treatment of mania. World J Biol Psychiatry. 2003;4(1):5–13.
Grunze H, Kasper S, Goodwin G, Bowden C, Moller HJ. The World Federation of Societies of
Biological Psychiatry (WFSBP) guidelines for the biological treatment of bipolar disorders, part
III: maintenance treatment. World J Biol Psychiatry. 2004;5(3):120–35.
Grunze H, Vieta E, Goodwin GM, Bowden C, Licht RW, Moller HJ, et al. The World Federation
of Societies of Biological Psychiatry (WFSBP) guidelines for the biological treatment of
bipolar disorders: update 2009 on the treatment of acute mania. World J Biol Psychiatry.
2009;10(2):85–116.
Grunze H, Vieta E, Goodwin GM, Bowden C, Licht RW, Moller HJ, et al. The World Federation of
Societies of Biological Psychiatry (WFSBP) guidelines for the biological treatment of bipolar
disorders: update 2010 on the treatment of acute bipolar depression. World J Biol Psychiatry.
2010;11(2):81–109.
Grunze H, Vieta E, Goodwin GM, Bowden C, Licht RW, Moller HJ, et al. The World Federation of
Societies of Biological Psychiatry (WFSBP) guidelines for the biological treatment of bipolar
disorders: update 2012 on the long-term treatment of bipolar disorder. World J Biol Psychiatry.
2013;14(3):154–219.
Grunze H, Vieta E, Goodwin GM, Bowden C, Licht RW, Azorin JM, et al. The World Federation of
Societies of Biological Psychiatry (WFSBP) guidelines for the biological treatment of bipolar
disorders: acute and long-term treatment of mixed states in bipolar disorder. World J Biol
Psychiatry. 2018;19(1):2–58.
Hanlon P. The future public health. Maidenhead: McGraw-Hill Education; 2012.
Hasan A, Falkai P, Wobrock T, Lieberman J, Glenthoj B, Gattaz WF, et al. World Federation of
Societies of Biological Psychiatry (WFSBP) guidelines for biological treatment of schizophre-
nia, part 1: update 2012 on the acute treatment of schizophrenia and the management of
treatment resistance. World J Biol Psychiatry. 2012;13(5):318–78.
Hasan A, Falkai P, Wobrock T, Lieberman J, Glenthoj B, Gattaz WF, et al. World Federation of
Societies of Biological Psychiatry (WFSBP) guidelines for biological treatment of schizophre-
nia, part 2: update 2012 on the long-term treatment of schizophrenia and management of
antipsychotic-induced side effects. World J Biol Psychiatry. 2013;14(1):2–44.
Hasan A, Falkai P, Wobrock T, Lieberman J, Glenthøj B, Gattaz WF, et al. World Federation of
Societies of biological psychiatry (WFSBP) guidelines for biological treatment of schizophre-
nia. Part 3: update 2015 management of special circumstances: depression, suicidality, sub-
stance use disorders and pregnancy and lactation. World J Biol Psychiatry. 2015;16(3):142–70.
Hasan A, Bandelow B, Yatham LN, Berk M, Falkai P, Möller H-J, et al. WFSBP guidelines on how
to grade treatment evidence for clinical guideline development. World J Biol Psychiatry.
2019;20(1):2–16.
Kamens SR, Cosgrove L, Peters SM, Jones N, Flanagan E, Longden E, et al. Standards and
guidelines for the development of diagnostic nomenclatures and alternatives in mental health
research and practice. J Humanist Psychol. 2019;59(3):401–27.
Kennedy SH, Lam RW, McIntyre RS, Tourjman SV, Bhat V, Blier P, et al. Canadian Network for
Mood and Anxiety Treatments (CANMAT) 2016 clinical guidelines for the management of
852 J. Vrublevska and L. Renemane
adults with major depressive disorder: section 3. Pharmacological treatments. Can J Psychiatry.
2016;61(9):540–60.
Kredo T, Bernhardsson S, Machingaidze S, Young T, Louw Q, Ochodo E, et al. Guide to clinical
practice guidelines: the current state of play. Int J Qual Health Care. 2016;28(1):122–8.
Kusumakar V, Yatham LN, Haslam DR, Parikh SV, Matte R, Silverstone PH, et al. Treatment
of mania, mixed state, and rapid cycling. Can J Psychiatr Rev can psychiatr. 1997;42(Suppl
2):79s–86s.
Lam RW, McIntosh D, Wang J, Enns MW, Kolivakis T, Michalak EE, et al. Canadian Network for
Mood and Anxiety Treatments (CANMAT) 2016 clinical guidelines for the management of
adults with major depressive disorder: section 1. Disease burden and principles of care. Can J
Psychiatry. 2016;61(9):510–23.
Lehman AF, Lieberman JA, Dixon LB, McGlashan TH, Miller AL, Perkins DO, et al. Practice
guideline for the treatment of patients with schizophrenia, second edition. Am J Psychiatry.
2004;161(2 Suppl):1–56.
Leucht S, Arango C, Fleischhacker WW, Kapur S, Stroup S, van Os J, et al. CINP SCHIZOPHRE-
NIA GUIDELINE. 2013. The International College of Neuropsychopharmacology (CINP)
https://www.cinp.org/resources/Documents/CINP-schizophrenia-guideline-24.5.2013-A-C-
method.pdf. Accessed April 2020.
MacQueen GM, Frey BN, Ismail Z, Jaworska N, Steiner M, Lieshout RJ, et al. Canadian Network
for Mood and Anxiety Treatments (CANMAT) 2016 clinical guidelines for the management of
adults with major depressive disorder: section 6. Special populations: youth, women, and the
elderly. Can J Psychiatry. 2016;61(9):588–603.
Malhi GS, Bassett D, Boyce P, Bryant R, Fitzgerald PB, Fritz K, et al. Royal Australian and New
Zealand College of Psychiatrists clinical practice guidelines for mood disorders. Aust N Z J
Psychiatry. 2015;49(12):1087–206.
McGorry P, Killackey E, Lambert T. Royal Australian and New Zealand College of Psychiatrists
clinical practice guidelines for the treatment of schizophrenia and related disorders. Aust N Z J
Psychiatry. 2005;39:1–30.
McIntyre JS. Usefulness and limitations of treatment guidelines in psychiatry. World Psychiatry.
2002;1(3):186–9.
Milev RV, Giacobbe P, Kennedy SH, Blumberger DM, Daskalakis ZJ, Downar J, et al. Canadian
Network for Mood and Anxiety Treatments (CANMAT) 2016 clinical guidelines for the
management of adults with major depressive disorder: section 4. Neurostimulation treatments.
Can J Psychiatry. 2016;61(9):561–75.
Mulrow CD. Rationale for systematic reviews. BMJ. 1994;309(6954):597–9.
Munder T, Brütsch O, Leonhart R, Gerger H, Barth J. Researcher allegiance in psychotherapy
outcome research: an overview of reviews. Clin Psychol Rev. 2013;33(4):501–11.
NICE. National Institute for Health and Care Excellence. Depression in adults: recognition and
management (Clinical guideline 90). www.nice.org.uk/guidance/cg90. 2009.
NICE. Bipolar disorder: assessment and management clinical guideline. www.nice.org.uk/guid
ance/cg185. 2014
Nivoli G, Lorettu L, Carpiniello B, Milia P, Pinna F, Lepretti A, et al. [Charges and convictions of
psychiatrists for the violent behavior of the patient: psychiatric-forensic remarks]. Riv Psichiatr.
2017;52(3):101–8.
Parikh SV, Quilty LC, Ravitz P, Rosenbluth M, Pavlova B, Grigoriadis S, et al. Canadian Network
for Mood and Anxiety Treatments (CANMAT) 2016 clinical guidelines for the management of
adults with major depressive disorder: section 2. Psychological treatments. Can J Psychiatry.
2016;61(9):524–39.
Ravindran AV, Balneaves LG, Faulkner G, Ortiz A, McIntosh D, Morehouse RL, et al. Canadian
Network for Mood and Anxiety Treatments (CANMAT) 2016 clinical guidelines for the
management of adults with major depressive disorder: section 5. Complementary and alternative
medicine treatments. Can J Psychiatry. 2016;61(9):576–87.
Neuropsychopharmacotherapy: Guidelines 853
Contents
Prescription Patterns of Neuropsychopharmacotherapy in China . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 856
Brief Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 856
Antipsychotic Polypharmacy in China . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 856
Clozapine in China . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 858
Brief History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 858
Prevalence of Clozapine Use in China . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 859
Product Information . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 861
Pattern of Utilization of Clozapine in China . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 862
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 863
Abstract
Prescription habits are affected by a variety of factors, including the healthcare
system, financing schemes, medical traditions, consumers’ choices, and the
overall sociocultural background of the respective countries. In the treatment of
psychosis, many Western countries have published national guidelines or con-
sensus statements that refer to the prescribing of neuropsychopharmacotherapy.
All advise against the use of high dosage other than in exceptional circumstances.
However, there are discrepancies between these guidelines or algorithms and
prescription patterns and that practice varies substantially according to patient,
prescriber, and institution.
J. Li (*)
Institute of Mental Health, Tianjin Anding Hospital, Tianjin, China
e-mail: jieli@tjmhc.com
S. Li
Department of Psychiatry, Tianjin Medical University, Tianjin, China
e-mail: lishen@tmu.edu.cn
Brief Introduction
Clozapine in China
Brief History
metabolic syndrome, lowering treatment adherence (Iqbal et al. 2003). It was found
that many patients discontinue clozapine in clinical practice because of its adverse
effects (Mustafa et al. 2015). Nowadays, the prevalence of clozapine-associated
agranulocytosis is low, and agranulocytosis-related death appears rare (Li et al.
2019). Thirty years of both ignorance and clinical experience suggest that clozapine
intoxication during co-occurring infections and inflammation may have higher
morbidity and mortality than is currently believed (Ruan and De Leon 2019).
Serious infections or inflammations have been associated with serum clozapine
concentration increases and sometimes with clozapine toxicity (Ruan et al. 2018).
Clozapine is frequently and increasingly prescribed for schizophrenia in Asia,
with large variation across countries and territories (Xu et al. 2019). Unlike most
countries around the world in which clozapine is considered a second-line medica-
tion reserved for use in treatment-resistant schizophrenia, China embraced clozapine
as a first-line treatment for schizophrenia and other psychotic disorders for many
years from the early 1980s (Tang et al. 2008; Potter et al. 1989). For a time, clozapine
was even the first choice for acute schizophrenia (Tang et al. 2008; Mao et al. 2007).
The major reason for the initial interest in clozapine was the widespread belief of
clinicians that its use was associated with a reduced risk of tardive dyskinesia.
Furthermore, the profound efficacy of clozapine in the treatment of schizophrenia
(Lieberman et al. 1989) and its minimal cost further contributed to its popularity
among clinicians (Mao et al. 2007; Tang 2000). The use of clozapine is cost efficient
because of the reduction in hospitalizations (Nielsen et al. 2012b; Aitchison and
Kerwin 1997). Increasingly, however, Chinese clinicians have become more aware
of the potentially severe side effects associated with clozapine and now tend to use it
mostly for treatment-resistant patients, particularly since a wide variety of new
second-generation antipsychotics (SGAs) has become available (Mao et al. 2007).
As China has the largest international population on clozapine treatment, its
experience and research findings with this compound are of keen interest to Western
psychiatrists. However, there is a substantial lack of scientific communication
between the Chinese and Western psychiatric communities. In particular, most
of the reports on clozapine have been published only in the Chinese language,
with very few papers published in international journals. To our knowledge, the
first report on the Chinese experience in English (Potter et al. 1989) was published
nearly 30 years ago, while another brief letter on the use of clozapine in China was
published in French in 1996 (Stip et al. 1996).
as high as 60% in some parts of China to as low as 2–3% in some parts of the USA
(Nielsen et al. 2016).
A national, cross-sectional pharmacoepidemiological survey project on pre-
scription trends for psychotropic drugs in schizophrenia was initiated by the
Chinese Society of Psychiatry in 2002 (Si et al. 2004). Approximately one-third
of Chinese schizophrenia patients received clozapine in the pooled sample with a
steadily decreasing prescription trend from 2002 to 2012 of 13.3%, representing a
relative drop by 33.5% (Li et al. 2015b). The figures found in this study are also
higher than those reported in the Western literature. For example, the frequency of
clozapine prescription was 8% in Afro-American and 15% in Caucasian patients
in the USA (Kuno and Rothbard 2002). Also clozapine was prescribed to less than
10% of schizophrenia patients in the USA between 1991 and 1996 (Rothbard et al.
2003).
Similar to the findings in the REAP surveys (Xiang et al. 2011), clozapine
prescriptions gradually, but steadily decreased in China (39.7% in 2002, 32.5% in
2006 and 26.4% in 2012), which could be due to the introduction of strict guidelines
for clozapine use, starting with the Chinese Psychiatric Association guidelines in
2003, and the introduction of other SGAs (Xiang et al. 2013). In addition, while
in the rest of the world clozapine underutilization is a large concern, clozapine was
probably overused in China (Xiang et al. 2011, 2013). The influence of recent
publications and presentations at academic conferences may have enhanced Chinese
clinicians’ awareness of rational prescription practices resulting in reduced and,
possibly, more appropriate clozapine prescriptions.
Clozapine prescription was associated with a number of demographic and clinical
variables. Clozapine is mainly used for treating TRS, which could explain the
association of frequent clozapine use with inpatient status, family history of psychi-
atric disorders, CGI-based severity of psychopathology, earlier age of onset, and
longer duration of illness, features characteristic of TRS (Yang et al. 2005). Unlike
positive symptoms, negative symptoms are far more difficult to treat, and they tend
to occur more frequently in TRS, increasing the likelihood of clozapine use
(Buchanan et al. 2010; Iqbal et al. 2003; Xiang et al. 2007). In this study, patients
receiving clozapine presented with less severe delusions and hallucinations, yet, they
had higher global illness scores on the CGI. The reason for this apparent discrepancy
is that the CGI reflects patients’ global functioning determined by the patients’
personal and psychiatric history, psychosocial circumstances, psychiatric symptoms,
and the impact of symptoms on patients’ functioning. Thus, the CGI is a more
comprehensive measure than the severity of delusions or hallucinations (Busner and
Targum 2007). Therefore, although the severity of delusions and hallucinations may
be lower, patients with treatment-resistant schizophrenia are rated as having a more
severe illness and as being more impaired overall, which is consistent with the
families’ lower satisfaction with treatment.
In conclusion, approximately one-third of Chinese schizophrenia patients
received clozapine in this study during the decade of 2002–2012. Considering
the increased likelihood of drug-induced side effects related to clozapine and
families’ dissatisfaction with psychiatric treatment in general and possibly
Neuropsychopharmacotherapy: Differential Dose Regimes in China 861
Product Information
Orally disintegrating tablets are only available in the USA and China, but the oral
suspension form is available in China, Denmark, Ireland, The Netherlands,
New Zealand, the UK, and the USA. Injectable clozapine is only available in The
Netherlands as an unlicensed medicinal product at a hospital pharmacy (clozapine
injection 125 mg = 5 ml; supplier Brocacef Ziekenhuisfarmacie). The maximum
licensed dosage is 900 mg/day in all countries, with the exception of Japan and
Romania, where it is only 600 mg (Nielsen et al. 2016).
According to the summaries of product characteristics, clozapine is licensed
in all countries for treatment-resistant schizophrenia, the criteria which vary
among the included countries. In the USA, treatment-resistant schizophrenia is
defined as “only severely ill patients not responding to standard antipsychotic
treatment,” whereas in most other countries, it is defined as inadequate response
to at least two antipsychotics, with at least one of them being a second-
generation antipsychotic (SGA). Minimum duration of treatment or dose is not
otherwise specified.
Clozapine is licensed for treatment of psychosis in Parkinson’s disease patients in
most countries, with the exception of China, Japan, the USA, and New Zealand,
whereas it is only licensed for suicidal behaviors in patients with schizophrenia in the
USA and Romania.
The Chinese Medical Association also recommends clozapine, in addition to the
abovementioned indication, for patients with schizoaffective disorder, treatment-
resistant mania or severe treatment-resistant psychotic depression, and for other
treatment-resistant psychiatric disorders, i.e., patients with pervasive developmental
disorder, autism, or obsessive-compulsive disorder.
Although it is poorly investigated, Asian patients may have a slower metabolism
and may warrant lower doses (Ng et al. 2005). In general, psychiatrists are reluctant
to use clozapine above 600 mg/day (Nielsen et al. 2010). Clozapine provokes large
interindividual variation in pharmacokinetic properties, and although response
occurs across the whole range of plasma levels, the chance of response is twice as
high above a clozapine plasma level of approximately 400 ng/ml (Schulte 2003).
Therefore, some patients may even warrant higher doses than 900 mg because of
ultrahigh activity in the cytochrome P450 (CYP) 1A2 enzyme (Maccall et al. 2009).
Therapeutic drug monitoring (TDM) is not mandatory in any country, although it
may be of important value in the optimization of clozapine treatment (Nielsen et al.
2011). It is worth noting that a large number of patients respond at much lower
concentrations than suggested above, and clozapine should always be used at the
lowest effective dose in order to minimize the burden of adverse effects (Nielsen
et al. 2011).
862 J. Li and S. Li
In China, the national guidelines for the prevention and treatment of psychiatric
disorders in 2003 stipulated that (1) clozapine should be considered only after
FGAs or other SGAs have been deemed ineffective following 6–8 weeks of
treatment and (2) blood monitoring for patients on clozapine weekly or biweekly
during the first 6 months and biweekly or monthly afterwards is mandatory
(Chinese Medical Association 2003). As the most effective antipsychotic for
TRS, clozapine is often the last resort in clinical practice, at least in countries
other than China (Esposito et al. 2005; Malalagama et al. 2011; Schulte 2003).
Approximately 15–25% of schizophrenia patients can be characterized as TRS
(Conley and Buchanan 1997), all of whom are potential candidates for clozapine
(Howes et al. 2012). The common prescription of clozapine in China could be
accounted for by the following reasons: clozapine has continuously been used
since its introduction in 1976 in China, yielding extensive clinical experience
(Xiang et al. 2007) and making it one of the most frequently used antipsychotics
in China, even as the first-line treatment in some areas (Liu and Li 2003). In
addition, clozapine is one of the cheapest antipsychotics in China, costing only
US$0.08 for 300 mg. Moreover, clozapine and risperidone were the only SGAs
covered by public health insurance in China before 2005.
In China, a study showed the mean clozapine daily dose was 206.1 mg/day
(Li et al. 2015b), which was lower than the recommended doses for Western
patients (Barnes and Schizophrenia Consensus Group of British Association for
Psychopharmacology 2011; Buchanan et al. 2010). Ethnic differences may account
for the dose difference between Chinese and Western patients (Wang et al. 2010).
In addition, in most Western countries clozapine is only prescribed for treatment-
resistant schizophrenia, while in China clozapine is more widely used and even
prescribed as a first-line treatment (Liu and Li 2003), which may also contribute to
the lower doses of clozapine in China.
With the exception of China, Denmark, Romania, and The Netherlands, all the
included countries have a database for hematological monitoring that is run by the
manufacturer.
Male patients were more likely to receive clozapine, which could be related to
males being overrepresented in this study with poorer treatment response (Carbon
and Correll 2014) and/or males having more severe negative symptoms compared
to females (Abel et al. 2010).
Prior surveys also found that adjunctive FGAs were less frequently co-prescribed
with clozapine, which is consistent with the current treatment recommendation
(Chinese Medical Association 2003) based on the lack of compelling evidence of
the superiority of this combination (Remington et al. 2005), and the greater possi-
bility of tardive dyskinesia with FGAs (Correll et al. 2004) APP, particularly the
combination of clozapine with risperidone or sulpiride, may provide some additional
clinical benefit in TRS (Josiassen et al. 2005; Shiloh et al. 1997), although the
findings are equivocal (Correll et al. 2009; Honer et al. 2006; Miyamoto et al. 2014,
2015). A further problem with APP is that it could increase the risk of side effects
Neuropsychopharmacotherapy: Differential Dose Regimes in China 863
and also make it difficult to gauge which medication really works (Gallego et al.
2012; Henderson and Goff 1996; Kapur et al. 2001).
Patients without health insurance in China are usually treated in Level-II hospitals
because of the relatively low treatment expenses. Clozapine is one of the cheapest
antipsychotics in China, thus it is more frequently used in uninsured patients.
Clozapine was associated with more frequent and severe side effects in this study,
which could be one of the reasons for the increasing likelihood of lower antipsy-
chotic doses in the clozapine group.
Clozapine use was associated with families’ poor satisfaction with treatment,
which may be attributed to the greater likelihood of clozapine use in TRS patients
whose continued illness severity albeit reduced by clozapine with regard to halluci-
nations and delusions remains a source of frustration and creates ongoing family
stress and burden, resulting in dissatisfaction with the overall effectiveness of the
treatment.
In summary, some important aspects of clinical practice with clozapine in China
are also different from those occurring in Western settings. For example, the dosing
schedule of clozapine by Chinese psychiatrists is more aggressive. Furthermore,
there is evidence supporting the benefit of clozapine in both mania- and treatment-
resistant depression. Overall, the authors believe that clozapine is still an important
antipsychotic agent for Chinese patients and psychiatrists. However, due to its
potentially severe side effects, we also strongly recommend that Chinese psychia-
trists need to pay more attention to this safety issue and should primarily consider
clozapine only for treatment-resistant patients with schizophrenia or chronic patients
with a low risk for diabetes.
References
Abel KM, Drake R, Goldstein JM. Sex differences in schizophrenia. Int Rev Psychiatry.
2010;22:417–28.
Aitchison KJ, Kerwin RW. Cost-effectiveness of clozapine. A UK clinic-based study. Br J
Psychiatry. 1997;171:125–30.
Amsler HA, Teerenhovi L, Barth E, Harjula K, Vuopio P. Agranulocytosis in patients treated with
clozapine. A study of the Finnish epidemic. Acta Psychiatr Scand. 1977;56:241–8.
Bai Z, Wang G, Cai S, Ding X, Liu W, Huang D, Shen W, Zhang J, Chen K, Yang Y, Zhang L,
Zhao X, Ouyang Q, Zhao J, Lu H, Hao W. Efficacy, acceptability and tolerability of 8 atypical
antipsychotics in Chinese patients with acute schizophrenia: a network meta-analysis. Schizophr
Res. 2017;185:73–9.
Barnes TR, Schizophrenia Consensus Group of British Association for Psychopharmacology.
Evidence-based guidelines for the pharmacological treatment of schizophrenia: recommenda-
tions from the British Association for Psychopharmacology. J Psychopharmacol. 2011;25:
567–620.
Buchanan RW, Kreyenbuhl J, Kelly DL, Noel JM, Boggs DL, Fischer BA, Himelhoch S, Fang B,
Peterson E, Aquino PR, Keller W, Schizophrenia Patient Outcomes Research Team. The 2009
schizophrenia PORT psychopharmacological treatment recommendations and summary state-
ments. Schizophr Bull. 2010;36:71–93.
Busner J, Targum SD. The clinical global impressions scale: applying a research tool in clinical
practice. Psychiatry (Edgmont). 2007;4:28–37.
864 J. Li and S. Li
Ifteni P, Correll CU, Nielsen J, Burtea V, Kane JM, Manu P. Rapid clozapine titration in treatment-
refractory bipolar disorder. J Affect Disord. 2014;166:168–72.
Iqbal MM, Rahman A, Husain Z, Mahmud SZ, Ryan WG, Feldman JM. Clozapine: a clinical
review of adverse effects and management. Ann Clin Psychiatry. 2003;15:33–48.
Ito H, Okumura Y, Higuchi T, et al. International variation in antipsychotic prescribing
for schizophrenia: pooled results from the research on East Asia psychotropic prescription
(REAP) studies. Open J Psychiatry. 2012;2:340–6.
Jeon SW, Kim YK. Unresolved issues for utilization of atypical antipsychotics in schizophrenia:
antipsychotic polypharmacy and metabolic syndrome. Int J Mol Sci. 2017;18:2174.
Josiassen RC, Joseph A, Kohegyi E, Stokes S, Dadvand M, Paing WW, Shaughnessy RA.
Clozapine augmented with risperidone in the treatment of schizophrenia: a randomized,
double-blind, placebo-controlled trial. Am J Psychiatry. 2005;162:130–6.
Kane J, Honigfeld G, Singer J, Meltzer H. Clozapine for the treatment-resistant schizophrenic.
A double-blind comparison with chlorpromazine. Arch Gen Psychiatry. 1988;45:789–96.
Kapur S, Roy P, Daskalakis J, Remington G, Zipursky R. Increased dopamine d(2) receptor
occupancy and elevated prolactin level associated with addition of haloperidol to clozapine.
Am J Psychiatry. 2001;158:311–4.
Kochi K, Sato I, Nishiyama C, Tanaka-Mizuno S, Doi Y, Arai M, Fujii Y, Matsunaga T, Ogawa Y,
Furukawa TA, Kawakami K. Trends in antipsychotic prescriptions for Japanese outpatients
during 2006–2012: a descriptive epidemiological study. Pharmacoepidemiol Drug Saf.
2017;26:642–56.
Krakowski MI, Czobor P, Citrome L, Bark N, Cooper TB. Atypical antipsychotic agents in
the treatment of violent patients with schizophrenia and schizoaffective disorder. Arch Gen
Psychiatry. 2006;63:622–9.
Kuno E, Rothbard AB. Racial disparities in antipsychotic prescription patterns for patients with
schizophrenia. Am J Psychiatry. 2002;159:567–72.
Leucht S, Cipriani A, Spineli L, Mavridis D, Orey D, Richter F, Samara M, Barbui C, Engel RR,
Geddes JR, Kissling W, Stapf MP, Lassig B, Salanti G, Davis JM. Comparative efficacy and
tolerability of 15 antipsychotic drugs in schizophrenia: a multiple-treatments meta-analysis.
Lancet. 2013;382:951–62.
Li Q, Xiang YT, Su YA, Shu L, Yu X, Chiu HF, Correll CU, Ungvari GS, Lai KY, Ma C, Wang GH,
Bai PS, Li T, Sun LZ, Shi JG, Chen XS, Mei QY, Li KQ, Si TM. Antipsychotic polypharmacy
in schizophrenia patients in China and its association with treatment satisfaction and quality
of life: findings of the third national survey on use of psychotropic medications in China. Aust
N Z J Psychiatry. 2015a;49:129–36.
Li Q, Xiang YT, Su YA, Shu L, Yu X, Correll CU, Ungvari GS, Chiu HF, Ma C, Wang GH, Bai PS,
Li T, Sun LZ, Shi JG, Chen XS, Mei QY, Li KQ, Si TM, Kane JM. Clozapine in schizophrenia and
its association with treatment satisfaction and quality of life: findings of the three national surveys
on use of psychotropic medications in China (2002–2012). Schizophr Res. 2015b;168:523–9.
Li XH, Zhong XM, Lu L, Zheng W, Wang SB, Rao WW, Wang S, Ng CH, Ungvari GS, Wang G,
Xiang YT. The prevalence of agranulocytosis and related death in clozapine-treated patients:
a comprehensive meta-analysis of observational studies. Psychol Med. 2019;12:1–12.
Lieberman JA, Kane JM, Johns CA. Clozapine: guidelines for clinical management. J Clin
Psychiatry. 1989;50:329–38.
Lin CH, Lin CY, Wang HS, Lane HY. Long-term use of clozapine is protective for bone density in
patients with schizophrenia. Sci Rep. 2019;9:3895.
Liu TL, Li CY. The comparison of use of antipsychotics in first-episode patients with schizophrenia
(in Chinese). Med J Chin People Health. 2003;15:289–90.
Lung SLM, Lee HME, Chen YHE, Chan KWS, Chang WC, Hui LMC. Prevalence and correlates of
antipsychotic polypharmacy in Hong Kong. Asian J Psychiatr. 2018;33:113–20.
Maccall C, Billcliff N, Igbrude W, Natynczuk S, Spencer EP, Flanagan RJ. Clozapine: more than
900 mg/day may be needed. J Psychopharmacol. 2009;23:206–10.
Malalagama G, Bastiampillai T, Dhillon R. Clozapine prescription patterns in Australia over the last
10 years. Aust N Z J Psychiatry. 2011;45:498–9.
866 J. Li and S. Li
Mao PX, Tang YL, Wang ZM, Jiang F, Gillespie CF, Cai ZJ. Antipsychotic drug use in 503 Chinese
inpatients with schizophrenia. Int J Psychiatry Clin Pract. 2007;11:29–35.
Meltzer HY, Alphs L, Green AI, Altamura AC, Anand R, Bertoldi A, Bourgeois M, Chouinard G,
Islam MZ, Kane J, Krishnan R, Lindenmayer JP, Potkin S, International Suicide Prevention
Trial Study Group. Clozapine treatment for suicidality in schizophrenia: International Suicide
Prevention Trial (InterSePT). Arch Gen Psychiatry. 2003;60:82–91.
Miyamoto S, Jarskog LF, Fleischhacker WW. New therapeutic approaches for treatment-resistant
schizophrenia: a look to the future. J Psychiatr Res. 2014;58:1–6.
Miyamoto S, Jarskog LF, Fleischhacker WW. Schizophrenia: when clozapine fails. Curr Opin
Psychiatry. 2015;28:243–8.
Moore TA, Buchanan RW, Buckley PF, Chiles JA, Conley RR, Crismon ML, Essock SM,
Finnerty M, Marder SR, Miller DD, Mcevoy JP, Robinson DG, Schooler NR, Shon SP,
Stroup TS, Miller AL. The Texas Medication Algorithm Project antipsychotic algorithm for
schizophrenia: 2006 update. J Clin Psychiatry. 2007;68:1751–62.
Morrato EH, Dodd S, Oderda G, Haxby DG, Allen R, Valuck RJ. Prevalence, utilization patterns,
and predictors of antipsychotic polypharmacy: experience in a multistate Medicaid population,
1998–2003. Clin Ther. 2007;29:183–95.
Mustafa FA, Burke JG, Abukmeil SS, Scanlon JJ, Cox M. “Schizophrenia past clozapine”:
reasons for clozapine discontinuation, mortality, and alternative antipsychotic prescribing.
Pharmacopsychiatry. 2015;48:11–4.
Ng CH, Chong SA, Lambert T, Fan A, Hackett LP, Mahendran R, Subramaniam M, Schweitzer I.
An inter-ethnic comparison study of clozapine dosage, clinical response and plasma levels. Int
Clin Psychopharmacol. 2005;20:163–8.
Nielsen J, Dahm M, Lublin H, Taylor D. Psychiatrists’ attitude towards and knowledge of clozapine
treatment. J Psychopharmacol. 2010;24:965–71.
Nielsen J, Damkier P, Lublin H, Taylor D. Optimizing clozapine treatment. Acta Psychiatr Scand.
2011;123:411–22.
Nielsen J, Kane JM, Correll CU. Real-world effectiveness of clozapine in patients with bipolar
disorder: results from a 2-year mirror-image study. Bipolar Disord. 2012a;14:863–9.
Nielsen J, Nielsen RE, Correll CU. Predictors of clozapine response in patients with treatment-
refractory schizophrenia: results from a Danish Register Study. J Clin Psychopharmacol.
2012b;32:678–83.
Nielsen J, Young C, Ifteni P, Kishimoto T, Xiang YT, Schulte PF, Correll CU, Taylor D. Worldwide
differences in regulations of clozapine use. CNS Drugs. 2016;30:149–61.
Percudani M, Barbui C, Fortino I, Petrovich L. Epidemiology of first- and second-generation
antipsychotic agents in Lombardy, Italy. Pharmacopsychiatry. 2005;38:128–31.
Phillips MR, Lu SH, Wang RW. Economic reforms and the acute inpatient care of patients with
schizophrenia: the Chinese experience. Am J Psychiatry. 1997;154:1228–34.
Potter WZ, Ko GN, Zhang LD, Yan WW. Clozapine in China: a review and preview of US/PRC
collaboration. Psychopharmacology (Berl). 1989;99(Suppl):S87–91.
Qiu H, He Y, Zhang Y, He M, Liu J, Chi R, Si T, Wang H, Dong W. Antipsychotic polypharmacy in
the treatment of schizophrenia in China and Japan. Aust N Z J Psychiatry. 2018. https://doi.org/
10.1177/0004867418805559.
Remington G, Saha A, Chong SA, Shammi C. Augmentation strategies in clozapine-resistant
schizophrenia. CNS Drugs. 2005;19:843–72.
Rothbard AB, Kuno E, Foley K. Trends in the rate and type of antipsychotic medications prescribed
to persons with schizophrenia. Schizophr Bull. 2003;29:531–40.
Ruan CJ, De Leon J. Thirty years of both ignorance and clinical experience suggest that clozapine
intoxication during co-occurring infections and inflammation may have higher morbidity and
mortality than is currently believed. Psychosomatics. 2019;60:221–2.
Ruan CJ, Zhang XL, Guo W, Li WB, Zhuang HY, Li YQ, Wang CY, Tang YL, Zhou FC, De Leon J.
Two cases of high serum clozapine concentrations occurring during inflammation in Chinese
patients. Int J Psychiatry Med. 2018;53:292–305.
Neuropsychopharmacotherapy: Differential Dose Regimes in China 867
Schulte P. What is an adequate trial with clozapine?: therapeutic drug monitoring and time to
response in treatment-refractory schizophrenia. Clin Pharmacokinet. 2003;42:607–18.
Shiloh R, Zemishlany Z, Aizenberg D, Radwan M, Schwartz B, Dorfman-Etrog P, Modai I,
Khaikin M, Weizman A. Sulpiride augmentation in people with schizophrenia partially respon-
sive to clozapine. A double-blind, placebo-controlled study. Br J Psychiatry. 1997;171:569–73.
Si TM, Shu L, Yu X, Ma C, Wang GH, Pai PS, Liu XH, Ji LP, Shi JG, Chen XS, Mei QY, Li KQ,
Zhang HY, Ma H. Antipsychotic drug patterns of schizophrenia in China: a cross-sectional study
(in Chinese). Chin J Psychiatry. 2004;37:152–5.
Sim K, Su A, Chan YH, Shinfuku N, Kua EH, Tan CH. Clinical correlates of antipsychotic
polytherapy in patients with schizophrenia in Singapore. Psychiatry Clin Neurosci. 2004;
58:324–9.
Stip E, Yan H, Lee P. Clozapine in a Chinese population. J Psychiatry Neurosci Bull. 1996;
21:283–4.
Tang YL. Pharmacotherapy of schizophrenia. In: Cai ZJ, Weng YZ, Tang YL, editors.
Schizophrenia: from biology to treatments (Chinese). 1st ed. Beijing: The Science Press;
2000. p. 214–27.
Tang YL, Mao PX, Jiang F, Chen Q, Wang CY, Cai ZJ, Mitchell PB. Clozapine in China.
Pharmacopsychiatry. 2008;41:1–9.
Ungvari GS, Chow LY, Chiu HF, Ng FS, Leung T. Modifying psychotropic drug prescription
patterns: a follow-up survey. Psychiatry Clin Neurosci. 1997;51:309–14.
Wang CY, Xiang YT, Cai ZJ, Weng YZ, Bo QJ, Zhao JP, Liu TQ, Wang GH, Weng SM, Zhang HY,
Chen DF, Tang WK, Ungvari GS, Risperidone Maintenance Treatment in Schizophrenia
Investigators. Risperidone maintenance treatment in schizophrenia: a randomized, controlled
trial. Am J Psychiatry. 2010;167:676–85.
Xiang YT, Weng YZ, Leung CM, Tang WK, Ungvari GS. Clinical correlates of clozapine
prescription for schizophrenia in China. Hum Psychopharmacol. 2007;22:17–25.
Xiang YT, Wang CY, Si TM, Lee EH, He YL, Ungvari GS, Chiu HF, Shinfuku N, Yang SY,
Chong MY, Kua EH, Fujii S, Sim K, Yong MK, Trivedi JK, Chung EK, Udomratn P, Chee KY,
Sartorius N, Dixon LB, Kreyenbuhl JA, Tan CH. Clozapine use in schizophrenia: findings
of the Research on Asia Psychotropic Prescription (REAP) studies from 2001 to 2009. Aust
N Z J Psychiatry. 2011;45:968–75.
Xiang YT, Wang CY, Si TM, Lee EH, He YL, Ungvari GS, Chiu HF, Yang SY, Chong MY, Tan CH,
Kua EH, Fujii S, Sim K, Yong KH, Trivedi JK, Chung EK, Udomratn P, Chee KY, Sartorius N,
Shinfuku N. Antipsychotic polypharmacy in inpatients with schizophrenia in Asia (2001–2009).
Pharmacopsychiatry. 2012;45:7–12.
Xiang YT, Buchanan RW, Ungvari GS, Chiu HF, Lai KY, Li YH, Si TM, Wang CY, Lee EH,
He YL, Yang SY, Chong MY, Kua EH, Fujii S, Sim K, Yong MK, Trivedi JK, Chung EK,
Udomratn P, Chee KY, Sartorius N, Tan CH, Shinfuku N. Use of clozapine in older Asian
patients with schizophrenia between 2001 and 2009. PLoS One. 2013;8:e66154.
Xu SW, Dong M, Zhang Q, Yang SY, Chen LY, Sim K, He YL, Chiu HF, Sartorius N, Tan CH,
Chong MY, Shinfuku N, Lin SK, Ng CH, Ungvari GS, Najoan E, Kallivayalil RA,
Jamaluddin R, Javed A, Iida H, Swe T, Zhang B, Xiang YT. Clozapine prescription pattern in
patients with schizophrenia in Asia: the REAP survey (2016). Psychiatry Res. 2019.
Yang HC, Yang KJ, Liu TB, Gao H, Lu YW. The clinical characteristics of patients with treatment-
resistant schizophrenia. Med J Chin People’s Health. 2005;17:257–9.
Ye M, Tang W, Liu L, Zhang F, Liu J, Chen Y, Chen DC, Tan YL, Yang FD, Hong Xiu M, Hui L,
Lv MH, Soares JC, Zhang XY. Prevalence of tardive dyskinesia in chronic male inpatients with
schizophrenia on long-term clozapine versus typical antipsychotics. Int Clin Psychopharmacol.
2014;29:318–21.
Nutrition in Brain Aging: Its Relevance
to Age-Associated Neurodegeneration
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 870
The Molecular Mechanism of Brain Aging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 871
Oxidative Stress and Aging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 871
Impaired Proteolysis Systems for Abnormal Proteins in Aging and PD . . . . . . . . . . . . . . . . . . . 875
Mitochondrial Dysfunction in Aging and PD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 876
How to Intervene the Brain Aging and Age-Associated Neurodegeneration? . . . . . . . . . . . . . . . . . 877
Calorie Restriction (CR) and Small Natural Components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 877
Can Dietary Habits Prevent Aging and Age-Associated Neurodegeneration? . . . . . . . . . . . . . 883
Gut-Brain Axis in Aging and Neurodegeneration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 888
Discussion and Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 888
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 889
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 889
Abstract
In the modern society, the number of the aged population having dementia and
disablement is rapidly increasing, which induces medical, social, and financial
problems in the developed country. The prevalence of Alzheimer and Parkinson
diseases increases according to aging, indicating that brain aging is the most potent
risk factor of these neurodegeneration. The molecular mechanism of aging is still an
enigma. Heterogeneous and complexed factors such as accumulated oxidative stress,
impaired proteolysis system, mitochondrial dysfunction, and gene mutation have
been proposed. Calorie restriction (CR) can prolong the lifespan of various animals,
suggesting that nutrient sensing pathways regulate the aging process. Several natural
compounds called CR mimics, like metformin, rapamycin, and plant food–derived
phytochemicals, were found to target the multiple intracellular signal transduction
systems and delay the aging phenotypes in the animal and cellular models. However,
at present clinical intervention trials have failed to show their effectiveness. Recent
epidemiological studies and clinical trials suggest that lifestyle including diet,
exercise, and sleep affects the risk or progression of brain aging. Especially,
Mediterranean diet and low-carbohydrate diet have shown to ameliorate cognitive
decline and reduce the incidence of age-associated neurodegenerative diseases. This
review presents the vital role of oxidative stress, possible clinical markers, and
beneficial dietary habits in aging and age-related neurodegeneration. Future strategy
to intervene the brain aging by nutrition in human is discussed.
Introduction
Aging is an irreversibly progressive deleterious process and the most potent risk
factor for neurodegenerative diseases, such as Parkinson and Alzheimer diseases
(PD and AD). Aging is caused by accumulation of diverse damages, resulting in the
failure to maintain tissue homeostasis. Gene expression pattern regulates lifespan by
the modulation of several factors in the inter- and intracellular signal pathways. On
the other hand, environmental factors, including diet, exercise, sleep, and exposure
to toxic factors, affect the aging process. Harman proposed “Free radical theory of
ageing” based on that oxidative stress increases the susceptibility to diseases and
death by regulating aging process (Harman 1956). Twin studies presented that the
heredity of the human lifespan is 20–30% (Ekmekciouglu 2020) and that the ratio of
genetic contribution increases with age (vB Hjelmborg et al. 2006). Cellular aging
has been considered to be due to the continuous free radical reactions, leading to
oxidative stress, mitochondrial dysfunction, impaired proteolysis systems, such as
the ubiquitin-proteasome system (UPS) and the autophagy-lysosomal pathway
(ALP), and deficit energy synthesis (Wyss-Coray 2016).
The strategy to slow down the brain aging by dietary manipulations has been
proposed. Caloric restriction (CR) has been shown to prolong the lifespan in exper-
imental models from yeast to the higher animals (Fontana et al. 2010), suggesting that
various molecular effectors regulate nutrient sensing pathways and longevity. In
human, Mediterranean diet (MedDiet) was found to reduce not only the risk of general
geriatric disorders, but also cognitive decline in aging and incidence of AD (van den
Brink et al. 2019). A population-based cohort study in Sweden presented association
of adherence to MedDiet with lower risk for PD (Yin et al. 2021). However, most of
such human intervention trials have not shown sufficient evidence for neuroprotection,
because the nutritional intervention contains complex issues to be solved, such as the
dietary habits regulated by the religions and the histories (Vauzour et al. 2017).
Nutraceuticals are defined as concentrated bioactive compounds from food and used
for enhancing health or medical benefits, as well as prevention and treatment of
diseases, including lipids, vitamins, proteins, glycosides, phenolic compounds, phy-
tochemicals, and others. The role of nutraceuticals in antiaging therapy has been
repeatedly discussed, but the results are still controversial. At present, only a few
nutraceuticals, such as coenzyme Q10 (CoQ1) and n-3 fatty acids, have shown
beneficial effects against various neurodegenerative disorders in clinical trials.
Nutrition in Brain Aging: Its Relevance to Age-Associated Neurodegeneration 871
In this chapter the recent basic and clinical topics of brain aging are reviewed. The
novel strategy to develop the diet and nutrients to ameliorate the human brain aging
is discussed.
condensation product of nitric oxide and superoxide. 3NT and dityrosine were found
in aging-related lipofuscin, composed of lipids, metals, and aggregated proteins in
the neurons (Kato et al. 1998). 3NT significantly increases in the CSF during aging
and patients with AD and amyotrophic lateral sclerosis (ALS) (Tohgi et al. 1999).
3NT is detected in plasma and urine of human, and increases in inflammatory
diseases, suggesting 3NT as a biomarker for systemic nitro-oxidative stress and
inflammation like diabetes mellitus (Feeney and Schöneich 2012). 3NT is identified
in α-synuclein (αSyn), the main composition of Lewy bodies (He et al. 2021) and
mitochondrial complex I (NADH dehydrogenase-ubiqutin oxidoreductase) in the
PD brain (Chinta and Andersen 2011). Decreased complex I activity and accumu-
lation of toxic abnormal αSyn are considered as the causative factors for
idiopathic PD.
Cross-linkage of the amino acids in the proteins destroys its higher structure.
Covalent ortho-ortho coupling of two tyrosine residues forms dityrosine in the
presence of copper, which was detected in the CSF of AD (Ahmed et al. 2005).
Dityrosine cross-linking stabilizes amyloid β (Aβ) oligomer (Al-Hilaly et al. 2013)
and promotes αSyn assemblies and deposition in Lewy bodies of PD brain
(Al-Hilaly et al. 2016). Cysteine residues are also oxidized and nitrated, but the
oxidation is mostly reversible. In the brain of aged humans, S-nitrated cysteine
residues increased in the hippocampus, substantia nigra, and frontal cortex (Finelli
2020). S-Sulfonylation and disulfide bonds were detected in antioxidant enzymes,
SOD and peroxiredoxins (PRDXs).
O
COOH O O
OH
COOH
O
O
4-Hydroxy-2-nonenal (4HNE)
n-6 fatty acid : arachidonic acid derived from n-6 fatty acid Acrolein
+ O2
R1 ・ R2
O
H
N
Lipid radical
HOO O NH
R1 R2
Propanoylation
Propanoylation
Lipid hydroperoxide derived from
derived n-3PUFA
from fatty acid peroxide
peroxide
Acylation to O
lysine residue HOOC
H
N
O NH
Succinylation
Succinylation
derived from
derived n-3 PUFA
from fatty acid peroxide
peroxide
Fig. 1 Oxidative decomposition of n-3 and n-6 polyunsaturated fatty acids (PUFAs). PUFAs rich
in neural membrane are the target of oxyradicals, and hydroxyl radical (.OH) produces intermediate
lipid radicals to form lipid hydroperoxide. Lipid hydroperoxide produces reactive lipid aldehyde,
such as 4-oxo-2-nonenal (4ONE), 4-hydroxy-2-nonenal (4HNE), malon-dialdehyde and acrolein as
shown in the figure. In addition, lipid hydroperoxide and lipid radicals make adducts with the
specific amino acids in proteins
Intracellular
transmission of toxic
oligomers
exosome
Aggregated
Production of modified proteins
membrane-derived Lipid radical-modified proteins
lipid radicals
Toxic oligomer
Autophagosome Autolysosome
Proteolysis by Digestion
ubiquitin-proteosome
Lysosome
Fig. 2 Production and detoxification of toxic oxidized lipid radical–modified proteins in the
neurons. αSyn is stabilized by binding to PUFAs in membrane phospholipids. Oxidation of
membrane PUFA and modification of αSyn by lipid radicals accelerate modified αSyn release to
the cytosol. αSyn modification by lipid aldehyde enhances αSyn oligomerization and aggregation,
and the cell-to-cell transfer in cultured neurons. Toxic αSyn oligomer is degraded by the UPS and
the ALP, or secreted by the exosome
from the cells and declines by aging in human skin fibroblasts. Age-related Drp-1
deficit suppressed fission and mitophagy and was proposed to be associated with
some age-related cognitive decline and hippocampal atrophy (Oettinghaus et al.
2016).
Mitochondrial dysfunctions have been proposed to cause neurodegenerative and
metabolic disorders. After the discovery of 1-methy-4-phenyl-1,2,3,6-teterahydro-
pyridine (MPTP), a synthesized neurotoxin which causes parkinsonism in humans
by inhibiting mitochondrial complex I, numerous papers reported the role of mito-
chondrial dysfunction in PD (Trinh et al. 2021). Complex I activity decreased in the
SN and frontal cortex of PD patients and complex I and IV activities decreased in
platelets and skeletal muscle of patients with sporadic PD (Benecke et al. 1993);
however, the results are not fully reproducible.
In familial (genetic) PD, mitochondrial dysfunction is caused by autosomal dom-
inant mutations of SNCA and leucine-rich repeat kinase 2 (LRRK2), and autosomal
recessive mutations of Parkin, PTEN-induced putative kinase (PINK1) and P5-type
ATPase 13a2 (ATP13a2). PINK1 is a mitochondria-targeted serine/threonine kinase
and phosphorylates Parkin, a cytosolic E3 ubiquitin ligase, which ubiquitinates target
proteins, Mfns (Pickrell and Youle 2015). Parkin is incorporated into mitochondria,
activates peroxisome proliferator-activated receptor γ coactivator-1α (PGC-1α) and
mitogenesis, promotes fission/fusion and mitophagy, and maintains mitochondrial
homeostasis (Poole et al. 2008). Mutated Parkin and PINK1 in autosomal PD
dysregulate mitophagy and mitochondrial dynamics. Damaged mitochondria are
ubiquitinated by activated Parkin and degraded by the UPS. PINK1 or Parkin mutation
causes the accumulation of damaged mitochondria in axons in parkinsonian patients
(Liu et al. 2012).
In conclusion, oxidative stress, impaired proteolysis system, and mitochondrial
dysfunction make a vicious cycle for brain aging and age-associated neurodegeneration,
like PD.
Metformin
Metformin (N,N-dimethylimidodicarbonimidic diamide) is a guanidine derivative
present in French lilac, used as a traditional herb since the seventeenth century in
Europe. Metformin was synthesized for the first time in 1922 and is the first-line
agent for treatment of type 2 diabetes mellitus. It increases the effects of insulin and
suppresses endogenous gluconeogenesis from L-lactate by liver. Metformin is
multifunctional, inhibits mitochondrial complex I activity (Owen et al. 2000),
activates AMPK, and inhibits glucagon-induced elevation of cAMP and mitochon-
drial glycerophosphate dehydrogenase. Recent reports suggest that metformin
reduces human body weight by increasing nonoxidative glucose disposal into
skeletal muscle and gut Escherichia species, which is associated with glycogen-
like peptide-1 (GLP-1) and releasing blood glucose to the intestine (Morita et al.
2020) (Fig. 3).
Recently metformin is proposed as an antiaging drug and a geroprotector
(Piskovatska et al. 2019) from preclinical investigations and beneficial effects for
type 2 diabetics and CVDs (Glossmann and Lutz 2019). Metformin prolongs the
lifespan of lower animals (yeast, fly, and worm) and mammals (mice and dog)
(Martin-Montalvo et al. 2013). Metformin activated mTOR-dependent protein phos-
phatase 2A (PP2A), dephosphorylated tau at Ser202, Ser262, and Ser356, and αSyn
at Ser129, and downregulated their neurotoxicity (Perez-Revuelta et al. 2014). In
Appetite loss
Change of enteric
microbiota
Metformin
Reduced gluconeogenesis
Fig. 3 The multiple functions of metformin. Metformin reduces gluconeogenesis, ameliorates
insulin resistance, and reduces body weight through several target molecules. Metformin prolongs
the lifespan of lower animals and mammals, but the effectiveness in human has not been proved
880 W. Maruyama et al.
Rapamycin
Rapamycin is a macrolide compound, an antibiotic and an inhibitor of mTOR, a
nutrient-sensing signaling. Rapamycin was isolated for the first time in 1972 from
samples of Streptomyces hygroscopicus found on Easter Island. Rapamycin has
immune-suppressant and antiproliferative functions and is proposed to prolong the
lifespan in animals. Rapamycin prolongs the lifespan in yeast, Caenorhabditis
elegans, Drosophila, and mice, improves age-associated decline of physiological
functions in rodents, dogs, and nonhuman primates, and may suppress aging and
age-related diseases, such as cancer, obesity, atherosclerosis, and neurodegeneration
in humans (Blagosklonny 2019). Rapamycin extended the median and maximal
lifespan of old mice (Harrison et al. 2009), suggesting that the pathway for the
longevity effect by rapamycin may be independent of CR response, which needs
early start.
The mTOR pathway is phosphatidylinositol-3-kinase (PI3K)-related serine/thre-
onine kinase, interacts with IGF and AMPK signaling pathways, and regulates
transcription, metabolism, and cytoarchitecture. mTOR drives cellular growth and
functions early in life, and later in life converses quiescence to senescence
(geroconversion); it induces cellular hyperfunction and promotes age-related pathol-
ogies. mTOR signaling pathway regulates autophagy (Kim and Guan 2015). In
postmortem tissue from the inferior parietal lobule of early stage of AD patients,
alteration of mTOR signaling, reduction of autophagy, and hyperactivity of PI3K/
Akt/mTOR pathway were detected (Tramutola et al. 2015). Rapamycin inhibits
mTORC1 (mTOR complex 1) by binding to FKBP12, activates autophagy, degrades
abnormal toxic proteins, and exerts neuroprotection in in vivo models of AD (Singh
et al. 2017).
Rapamycin reduced amyloid plaque deposition and microtubule-associated pro-
tein tau (MAPT/tau) aggregation and improved cognition in transgenic mouse model
of AD (Carosi and Sargeant 2019). In MPTP-induced mouse model of PD,
rapamycin attenuated impaired protein degradation by the UPS and ALP and
protected dopaminergic neurons (Dehay et al. 2010). Rapamycin inhibited
ubiquitination and increased IL-6 with reduced inflammatory cytokines by mTOR/
Akt/NF-κB signal pathway, preserved expression of glutamate transporter in astro-
cytes, and exerted neuroprotection (Zhang et al. 2017). However, the potency of
rapamycin as an immunosuppressant prevents its clinical application as an antiaging
drug in human.
Nutrition in Brain Aging: Its Relevance to Age-Associated Neurodegeneration 881
and modular of distinct cell functions (Joshi and Pratico 2012). In clinical studies in
patients with AD or MCI, vitamin E (2000 IU/day) supplement exerted benefit against
cognitive impairment and progression to AD, but the results are inconsistent (Browne
et al. 2019). The neuroprotective effect of α-tocopherol to early PD patients was
examined but could not show effectiveness (Shoulson 1998). In addition, vitamin E
treatment has potential risk to increase the incidence of malignancy (Nicastro and
Dunn 2013). The effect of fat-soluble antioxidative vitamins may be double edged, as
anti- and pro-oxidants.
Essential water-soluble vitamins are vitamin B1 (thiamine), B2 (riboflavin), B3
(niacin, nicotinic acid), B5 (pantothenic acid), B6 (pyridoxine derivatives), B7
(biotin, vitamin H), B9 (folic acid), B12 (cobalamin), and vitamin C. Declines in
pyridoxine, folic acid, and cobalamin have been reported to be associated with
cognition, depression, and idiopathic fatigue syndrome. These vitamin Bs are
cofactors for neurotransmitter and myelin synthesis, and the deficit increases levels
of homocysteine, a neurotoxic sulfur-containing metabolite of methionine, which is
found to increase in PD and AD. A randomized trial for patients with MCI presented
that vitamin B group (folic acid and vitamins B6 and B12) treatment combined with
n-3 fatty acids decreased plasma homocysteine levels and slowed decline in cogni-
tion by aging (Oulhaj et al. 2016). Vitamin C (ascorbic acid, AA) is a potent water-
soluble antioxidant. The AA concentration in neurons was quite higher (~10 mM)
than in the CSF (~120 μM) and plasma (~40 μM), suggesting the requirement of
constant supplement from diet into neurons. Aging-dependent loss of the BBB
integrity is considered to reduce the ability to retain AA in the brain. However,
epidemiological studies could not confirm the reduced AD risk by high AA intake.
High AA combined with vitamin E intake may reduce incidence of vascular demen-
tia (Masaki et al. 2000), but the effect is not consistent.
The neuroprotective effects of vitamins cannot be denied, but the results are
fluctuated by the dose, duration, and the heterogenity of patients. RCT with more
rational designs and surrogate markers are required considering the cellular mech-
anism of brain aging in future.
efficient energy source than glucose in the brain, and elevated β-hydroxybutyrate
produced from medium-chain triglycerides was considered to improve memory
function. Now several RCTs of ketogenic diet are ongoing in individuals with
subjective memory impairment, mild AD, or healthy control (McDonald and
Cervenka 2018). Also in PD, ketogenic diet exerted moderate to very good
improvement in symptom, a mean of 43% reduction assessed as UPDRS (United
Parkinson’s disease Rating Scale) scores (Vanitallie et al. 2005). In a randomized,
placebo-controlled, crossover study for PD patients, a ketone ester drink enhanced
endurance exercise performance with increased serum β-hydroxybutyrate levels
(Norwitz et al. 2020). However, there are often adverse effects of ketogenic diet
among adult treated for epilepsy and other neurological disorders: weight loss,
gastrointestinal effects, and increase of cholesterol and low-density lipoprotein
(LDL) in serum. In addition, high-fat and low-carbohydrate diet is not tasty and
hard to be digested, which means it may worsen the malnutrition in the aged. Long-
term strategy of ketogenic diet in human has not been established.
1. Increase the intake of fish and seafoods to enhance the contents of PUFA
2. Increase the intake of extra virgin (cold-pressed) olive oil, nuts, and legumes to
enhance the content of monounsaturated fatty acid (MUFA)
3. Increase the intake of leafy green vegetables, fruits, and cereals (whole grain) to
enhance the content of dietary fiber and antioxidative polyphenols
4. Increase the intake of poultry and egg to enhance the content of protein with low
content of saturated fatty acid (SFA)
5. Reduce the intake of red meats and dairy foods to lower the content of SFA
886 W. Maruyama et al.
Neuroprotection ?
Fig. 4 The mechanism of possible neuroprotective effects by low-carbohydrate diet and MedDiet.
Ketone bodies increased by low-carbohydrate diet have multiple biological functions, including
neuroprotection. MedDiet has been confirmed to reduce incidence of metabolic disease, geriatric
disorders, age-related decline in cognition, AD, and PD. Neuroprotection by MedDiet is proposed
to be due to reduced synthesis of lipid radicals, altered gene expression, and change of enteric
microbiota
Nutrition in Brain Aging: Its Relevance to Age-Associated Neurodegeneration 887
MedDiet supplemented with extra virgin oil (1 L/week) or mixed nuts (30 g/day)
improved cognitive composites (memory, attention, and execution and global) in
447 healthy volunteers after average 4.1-year intervention (Valls-Pedret et al. 2015).
Higher adherence to MedDiet and DASH (Dietary Approaches to Stop Hyperten-
sion) diet decreased cognitive decline and incidence of AD (Dominguez and
Barbagallo 2018). MedDiet and DASH accordance score were computed from
food and nutrient components. DASH diet specifies low consumption of saturated
fat, commercial pastries and sweets, and higher consumption of dairy than MedDiet.
In a large-scale prospective, population-based study, the both the MedDiet and
DASH dietary patterns containing high levels of whole grains, nuts, and legumes
were well associated with higher levels of cognitive functions in elderly men and
women over an 11-year period (Wengreen et al. 2013). The effect of dietary habit on
the incidence and progression of AD and PD was investigated in a clinical study
called MIND (Mediterranean DASH Intervention for Neurodegenerative Delay) for
older adults by comparison with the MedDiet. MIND diet slowed cognitive decline
with aging over average 4.7 years among 960 participants and reduced incidence of
AD more remarkedly than MedDiet and DASH (Morris et al. 2015a, b). In a three-
year brain imaging study of 70 cognitively normal participants, high adherence to
MedDiet modulated changes in AD biomarkers, increased 11C-Pittsburg compound
B-PET for brain β-amyloid load, declined 18F-fluorodeoxy-glucose PET for
neurodegeneration, and provided 1.5–3.5 years of protection against AD (Berti
et al. 2018). Higher MIND diet scores were associated with a decreased risk of PD
and a slower rate of the PD progression and MedDiet with a reduced progression,
whereas DASH diet was not associated with either outcome (Agarwal et al. 2018).
MIND diet adherence was strongly associated with later onset of PD in female
subgroup (Metcalfe-Roach et al. 2021) and lower probability of prodromal PD (2%
decrease) in older people (Maraki et al. 2019). In a population-based cohort of
Swedish women (>47,000), higher adherence to MedDiet at the middle age was
associated with lower risk (by a 29%) for PD (Yin et al. 2021). In the European
subject NU-AGE (new dietary strategies addressing the specific needs of elderly
population for a healthy aging in Europe), MedDiet demonstrated beneficial effects
on global cognition and episodic memory (Marseglia et al. 2018). After one-year
adherence to NU-AGE diet–given individually tailored dietary advice, cognition
function showed significant improvement in subjects with higher adherence com-
pared to those with lower adherence.
Recently, the mechanisms behind modulation of aging-related pathology by
MedDiet ingredients have been intensively studied. MedDiet ingredients, such as
oleuropein, oleocanthal, hydroxytyrosol, and resveratrol, activate autophagy pro-
cesses though AMPK, SIRT1/mTOR signaling pathways in in vivo and in vitro
models of AD (Cordero et al. 2018), and modulate aging process, prevent cognition
decline, and promote healthy aging (de Pablos et al. 2019). MedDiet ingredients
have epigenomic effects (DNA methylation and histone modification microRNAs
[miRNAs]) and transcriptomic effects, and modulate pathways including inflamma-
tion, oxidative stress, lipid metabolism, and carcinogenesis, and has potential impact
on human health. In the PREDIMED study, adherence to MedDiet and consumption
888 W. Maruyama et al.
of extra virgin olive oil were correlated with DNA methylation levels of inflamma-
tion- and intermediate metabolism–related genes (Arpon et al. 2016).
The definition of MedDiet is variable according to the studies, e.g., general
descriptive definition, diet pyramids/number of servings of key foods, grams of
key foods/food groups, and nutrient and flavonoid contents. In addition, each
MedDiet receipt depends on the local dietary pattern and food products (Davis
et al. 2015). Global, clinically available scoring to evaluate MedDiet is needed. In
addition, to compare the effect of dietary habit quantitively in individual human,
universal surrogate marker is essential.
Recent research in the gut-brain axis highlights new aspects on the neuroprotective
function of diets and nutraceuticals. Inside the gastrointestinal tract, the gut micro-
biota is composed of more than 100 trillion of microorganisms of numerous species.
Environment of gastrointestinal tract may affect the function of central nervous
system through the neural network and humoral factors. Gut microbiota is associated
with the absorption and metabolism of nutrients, production of short-chain fatty
acids (SCFAs), protection against pathogens, and regulation of host immunity. In
PD, gut microbiota composition changes to Bacrioides enriched from Prevotella-
and Firmicutes-enriched flora. Altered microbiota composition enhances gut inflam-
mation, increases intestinal permeability, and elevates stool inflammatory cytokines,
leading to damages in the gastrointestinal system, which has been proposed to be one
of risk factors of prodromal PD (Heinzel et al. 2020). Dietary habits modulate gut
microbial profiles and exert the disease-modifying effect in PD (Quigley 2020).
Aging affects gut microbial ecology and reduces biodiversity of microbiota in
elderly people, which may change the homeostasis between gut microbiota and
host and cause “immunosenescence.” Decrease of SCFA-producing Eubacterium
rectale and Roseburia reduces the metabolites in gut, then affects various human
physiology. Increased relative population of Bacteroidetes, Streptococcus, and
Blautia was associated with markers of cognitive and nutritional deficiencies
(Claesson et al. 2012). Dietary intervention has been reported to alter microbiota
composition in PD. High MedDiet adherence exerted beneficial effects on micro-
biota profile, smaller presence of E. coli, and increase of Bifidobacteria, thus the
higher levels of SCFAs, suggesting that typical health benefits by MedDiet might be
attributed to specific microbiota composition (Tosti et al. 2018).
Cross-References
Conflict of Interest The authors declare that there are no conflicting financial interests in relation
to the work described.
Source of Funding This work was supported by JSPS KAKENHI Grant Number 18 K07430
(W.M).
References
Agarwal P, Wang Y, Buchman AS, Holland TM, Bennett DA, Morris MC. MIND diet associated
with reduced incidence and delayed progression of Parkinsonism in old age. J Nutr Health
Aging. 2018;22(10):1211–5. https://doi.org/10.1007/s12603-018-1094-5.
Ahmed N, Ahmed U, Thornalley PJ, Hager K, Fleischer G, Münch G. Protein glycation, oxidation
and nitration adduct residues and free adducts of cerebrospinal fluid in Alzheimer’s disease and
link to cognitive impairment. J Neurochem. 2005;92(2):255–63. https://doi.org/10.1111/j.1471-
4159.2004.02864.x.
AL-Hilaly YK, Williams TL, Stewart-Parker M, et al. A central role for dityrosine crosslinking of
amyloid-β in Alzheimer’s disease. Acta Neuropathol Commun. 2013;1:83. https://doi.org/10.
1186/2051-5960-1-83.
890 W. Maruyama et al.
Al-Hilaly Y, Biasetti L, Blakeman BJ, Pollack SJ, Zibaee S, Abdul-Sada A, Thorpe JR, Xue WF,
Seppell LC. The involvement of dityrosine crosslinking in α-synuclein assembly and deposition in
Lewy bodies in Parkinson’s disease. Sci Rep. 2016;16(6):39171. https://doi.org/10.1038/srep39171.
Arpon A, Riezu-Boj JI, Milagro FI, et al. Adherence to Mediterranean diet is associated with
methylation changes in inflammation-related genes in peripheral blood cells. J Physiol Biochem.
2016;73(3):445–55. https://doi.org/10.1007/s13105-017-0552-6.
Ayala A, Munoz MF, Argüelles S. Lipid peroxidation: production, metabolism, and signaling
mechanisms of malondialdehyde and 4-hydroxy-2-nonenal. Oxidative Med Cell Longev.
2014;2014:360438. https://doi.org/10.1155/2014/360438.
Bach-Faig A, Berry EM, Lairon D, et al. Mediterranean diet pyramid today. Science and cultural
updates. Public Health Nutr. 2011;14(12A):2274–84. https://doi.org/10.1017/
S1368980011002515.
Bastianetto S, Menard C, Quirion R. Neuroprotective action of resveratrol. Biochim Biophys Acta.
2015;1852(6):1195–201. https://doi.org/10.1016/j.bbadis.2014.09.011.
Baur JA, Pearson KJ, Price NL, et al. Resveratrol improves health and survival of mice on a high-
calorie diet. Nature. 2006;444(7117):337–42. https://doi.org/10.1038/nature05354.
Bendor JT, Logan TP, Edwards RH. The function of α-synuclein. Neuron. 2013;79(6):1044–66.
https://doi.org/10.1016/j.neuron.2013.09.004.
Benecke R, Strumper P, Weiss H. Electron transfer complexes I and IV of platelets are abnormal in
Parkinson’s disease but normal in Parkinson-plus syndromes. Brain. 1993;116(6):1451–63.
https://doi.org/10.1093/brain/116.6.1451.
Berti V, Walters M, Sterling J, et al. Mediterranean diet and 3-year Alzheimer brain biomarker
changes in middle-aged adults. Neurology. 2018;90(20):e1789–98. https://doi.org/10.1212/
WNL.0000000000005527.
Blagosklonny MV. Rapamycin for longevity: opinion article. Aging. 2019;11(19):8048–67. https://
doi.org/10.18632/aging.102355.
Braak H, Del Tredici K. Neuropathological staging of brain pathology in sporadic Parkinson’s
disease: separating the wheat from the chaff. Parkinsons Dis. 2017;7(s1):S71–85. https://doi.
org/10.3233/JPD-179001.
Browne D, McGuinness B, Woodside JV, McKay GJ. Vitamin E and Alzheimer’s disease: what do
we know so far? Clin Invt Aging. 2019;14:1303–17. https://doi.org/10.2147/CIA.S186760.
Canto C, Auwerx J. Caloric restriction, SIRT1 and longevity. Trends Endocrinol Metab.
2009;20(7):325–31. https://doi.org/10.1016/j.tem.2009.03.008.
Carosi JM, Sargeant TJ. Rapamycin and Alzheimer disease: a double-edged sword? Autophagy.
2019;15(8):1460–2. https://doi.org/10.1080/15548627.2019.1615823.
Chen Y, Zhao S, Fan Z, et al. Metformin attenuates plaque-associated tau pathology and reduced
amyloid-β burden in APP/PS1 mice. Alzheimers Res Ther. 2021;13(1):40. https://doi.org/10.
1186/s13195-020-00761-9.
Chimento A, De Amicis F, Sirianni R, Sinicropi MS, Fuoci F, Casaburi I, Saturnino C, Pezzi
V. Progress to improve oral bioavailability and beneficial effects of resveratrol. Int J Mol Sci.
2019;20(6):1381. https://doi.org/10.3390/ijms20061381.
Chinta SJ, Andersen JK. Nitrosylation and nitration of mitochondrial complex I in Parkinson’s
disease. Free Radic Res. 2011;45(1):53–8. https://doi.org/10.3109/10715762.2010.509398.
Claesson MJ, Jeffery IB, Conde S, et al. Gut microbiota composition correlates with diet and health
in the elderly. Nature. 2012;488(7410):178–84. https://doi.org/10.1038/nature11319.
Colman RJ, Anderson RM, Johnson SC, et al. Caloric restriction delays disease onset and mortality
in rhesus monkeys. Science. 2009;325(5937):201–4. https://doi.org/10.1126/science.1173635.
Cordero JG, Garcia-Escudero R, Avila J, Gargini R, Garcia-Escudero V. Benefit of oleuropein
aglycone for Alzheimer’s disease by promoting autophagy. Oxidative Med Cell Longev.
2018;2018:5010741. https://doi.org/10.1155/2018/5010741.
Cuervo AM, Stefanis L, Fredenburg R, Lansbury PT, Sulzer D. Impaired degradation of mutant
α-synuclein by chaperone-mediated autophagy. Science. 2004;305(5688):1292–5. https://doi.
org/10.1126/science.1101738.
Nutrition in Brain Aging: Its Relevance to Age-Associated Neurodegeneration 891
Dahlin M, Elfving A, Ungerstedt U, Amark P. The ketogenic diet influences the levels of excitatory
and inhibitory amino acids in the CSF in children with refractory epilepsy. Epilepsy Res.
2005;64(3):115–25. https://doi.org/10.1016/j.eplepsyres.2005.03.008.
Das SK, Roberts SB, Bhapkar MV, et al. Body-composition changes in the comprehensive
assessment of long-term effects of reducing intake of energy (CALERIE)-2 study: a 2-y
randomized controlled trials of calorie restriction in nonobese humans. Am J Clin Nutr.
2017;105(4):913–27. https://doi.org/10.3945/ajcn.116.137232.
Davis C, Bryan J, Hodgson J, Murphy K. Definition of the Mediterranean diet; a literature review.
Nutrients. 2015;7(11):9139–53. https://doi.org/10.3390/nu7115459.
De Franceschi G, Fecchio C, Sharon R, Schapira AHV, Proukakis C, Bellotti V, de Laureto
PP. α-Synuclein structural features inhibit harmful polyunsaturated fatty acid oxidation,
suggesting roles in neuroprotection. J Biol Chem. 2017;292(17):6927–37. https://doi.org/10.
1074/jbc.M116.765149.
de Pablos R, Espinosa-Oliva AM, Hornedo-Ortega R, Cano M, Arguelles S. Hydroxytyrosol
protects from aging process via AMPK and autophagy; a review of its effects on cancer,
metabolic syndromes, osteoporosis, immune-mediated and neurodegenerative diseases.
Pharmacol Res. 2019;143:58–72. https://doi.org/10.1016/j.phrs.2019.03.005.
Dehay B, Bove J, Rodriguez-Muela N, Perier C, Recasens A, Boya P, Vila M. Pathogenic lysosomal
depletion in Parkinson’s disease. J Neurosci. 2010;30(37):12535–44. https://doi.org/10.1523/
JNEUROSCI.1920-10.2010.
Dinu M, Pagliai G, Casini A, Sofi F. Mediterranean diet and multiple health outcomes: an umbrella
review of meta-analyses of observational studies and randomized trials. Eur J Clin Nutr.
2018;72(1):30–43. https://doi.org/10.1038/ejcn.2017.58.
Dominguez L, Barbagallo M. Nutritional prevention of cognitive decline and dementia. Acta
Biomed. 2018;89(2):276–90. https://doi.org/10.23750/abm.v89i2.7401.
Ekmekciouglu C. Nutrition and longevity-from mechanisms to uncertainties. Crit Rev Food Sci
Nutr. 2020;60(18):3063–82. https://doi.org/10.1080/10408398.2019.1676698.
Feeney MB, Schöneich C. Tyrosine modifications in aging. Antioxid Redox Signal. 2012;17(11):
1571–9. https://doi.org/10.1089/ars.2012.4595.
Finelli MJ. Redox post-translational modifications of protein thiols in brain aging and neurodegen-
erative conditions – focus on S-nitration. Front Aging Neurosci. 2020;12:254. https://doi.org/
10.3389/fnagi.2020.00254.
Fontana L, Partridge L, Longo VD. Dietary restriction, growth factors and aging: from yeast to
humans. Science. 2010;328(5976):321–6. https://doi.org/10.1126/science.1172539.
Gardener H, Caunca MR. Mediterranean diet in preventing neurodegenerative diseases. Curr Nutr
Rep. 2018;7(1):10–20. https://doi.org/10.1007/s13668-018-0222-5.
GBD 2015 Neurological Disorders Collaborator Group. Global, regional, and national burden of
neurological disorders during 1990–2015: a systematic analysis for the Global Burden of Disease
Study 2015. Lancet Neurol. 2017;16:877–97. https://doi.org/10.1016/S1474-4422(17)30299-5.
Girotti AW. Lipid hydroperoxide generation, turnover, and effector action in biological systems.
J Lipid Res. 1998;39(8):1529–42.
Glossmann HH, Lutz OMD. Metformin and aging: a review. Gerontology. 2019;65(6):581–90.
https://doi.org/10.1159/000502257.
Grimm A, Eckert A. Brain aging and neurodegeneration: from a mitochondrial point of view.
J Neurochem. 2017;143(4):418–31. https://doi.org/10.1111/jnc.14037.
Hager K, Marahrens A, Kenklies M, Riederer P, Münch G. Alpha-lipoic acid as a new treatment
option for Alzheimer type dementia. Arch Gerontol Geriatr. 2001;32(3):275–82. https://doi.org/
10.1016/s0167-4943(01)00104-2.
Han YS, Zheng WH, Bastianetto S, Chabot JG, Quirion R. Neuroprotective effects of resveratrol
against β-amyloid-induced neurotoxicity in rat hippocampal neurons: involvement of protein
kinase C. Br J Pharmacol. 2004;141(6):997–1005. https://doi.org/10.1038/sj.bjp.0705688.
Harman D. Aging: a theory based on free radical and radiation. J Gerontol. 1956;11(3):298–300.
https://doi.org/10.1093/geronj/11.3.298.
892 W. Maruyama et al.
Harman D. The aging process. Proc Natl Acad Sci U S A. 1981;78(11):7124–8. https://doi.org/10.
1073/pnas.78.11.7124.
Harrison DE, Strong R, Sharp ZD, et al. Rapamycin fed late in life extends lifespan in genetically
heterogeneous mice. Nature. 2009;460(7253):392–5. https://doi.org/10.1038/nature08221.
Hawkins CL, Davies MJ. Detection, identification, and quantification of oxidative protein modifi-
cations. J Biol Chem. 2019;194(51):19683–708. https://doi.org/10.1074/jbc.REV119.006217.
He S, Wang F, Yung KKL, Zhang S, Qu S. Effects of α-synuclein-associated post-translational
modifications in Parkinson’s disease. ACS Chem Neurosci. 2021;12(7):1061–71. https://doi.
org/10.1021/acschemneuro.1c00028.
Heinzel S, Aho VTE, Suenkel U, et al. Gut microbiome signatures of risk and prodromal markers of
Parkinson disease. Ann Neurol. 2020;88(2):320–31. https://doi.org/10.1002/ana.25788.
Henderson ST, Vogel JL, Barr LJ, Garvin F, Jones JJ, Costantini LC. Study of the ketogenic agent
AC-1202 in mild to moderate Alzheimer’s disease: a randomized, double-blind, placebo-controlled,
multicenter trial. Nutr Metab (London). 2009;6:31. https://doi.org/10.1186/1743-7075-6-31.
Higashida K, Kim SH, Jung SR, Asaka M, Holloszy JO, Han DH. Effects of resveratrol and Sirt1 on
PGC-1α activity and mitochondrial biogenesis: a reevaluation. PLoS Biol. 2013;11(7):
e1001603. https://doi.org/10.1371/journal.pbio.1001603.
Iside C, Scafuro M, Nebbioso A, Altucci L. SIRT1 activation by natural phytochemicals: and
overview. Front Pharmacol. 2020;11:1225. https://doi.org/10.3389/fphar.2020.01225.
Joshi YB, Pratico D. Vitamin E in aging, dementia, and Alzheimer’s disease. Biofactors.
2012;38(2):90–7. https://doi.org/10.1002/biof.195.
Justice JN, Ferrucci L, Newman AB, et al. A framework for selection of blood-based biomarkers for
geroscience-guided clinical trials: report from the TAME Biomarkers Workgroup. Genoscience.
2018;40(5–6):419–36. https://doi.org/10.1007/s11357-018-0042-y.
Kashiwaya Y, Takeshima T, Mori N, Nakashima K, Clarke K, Veech RL. D-β-hydroxybutyrate
protects neurons in models of Alzheimer’s and Parkinson’s disease. Proc Natl Acad Sci U S
A. 2000;97(19):5440–4. https://doi.org/10.1073/pnas.97.10.5440.
Kato Y, Osawa T. Detection of lipid-lysine amide-type adduct as a marker of PUFA oxidation and
its application. Arch Biochem Biophys. 2010;501(2):182–7. https://doi.org/10.1016/j.abb.2010.
06.010.
Kato Y, Maruyama W, Naoi M, Hashizume Y, Osawa T. Immunohistochemical detection of
dityrosine in lipofuscin pigments in the aged human brain. FEBS Lett. 1998;439(3):231–4.
https://doi.org/10.1016/s0014-5793(98)01372-6.
Kaushik S, Cuervo AM. The coming of age of chaperone-mediated autophagy. Nat Rev Mol Cell
Biol. 2018;19(6):365–81. https://doi.org/10.1038/s41580-018-0001-6.
Kim YC, Guan KL. mTOR: a pharmacologic target for autophagy regulation. J Clin Invest.
2015;125(1):25–32. https://doi.org/10.1172/JCI73939.
Kim DY, Simeone KA, Simeone TA, et al. Ketone bodies mediate antiseizure effects through
mitochondrial permeability transition. Ann Neurol. 2015;78(1):77–87. https://doi.org/10.1002/
ana.24424.
Komatsu M, Wang QJ, Holstein GR, Friedrich VL Jr, Iwata J, Kominami E, Chait BT, Tanaka K,
Yue Z. Essential role for autophagy protein Atg7 in the maintenance of axonal homeostasis and
the prevention of axonal degeneration. Proc Natl Acad Sci U S A. 2007;104(36):14489–94.
https://doi.org/10.1073/pnas.0701311104.
Krikorian R, Shidler MD, Dangelo K, Couch SC, Benoit SC, Clegg DJ. Dietary ketosis enhances
memory in mild cognitive impairment. Neurobiol Aging. 2012;33(2):425.e19–27. https://doi.
org/10.1016/j.neurobiolaging.2010.10.006.
Lee HJ, Choi C, Lee SJ. Membrane-bound α-synuclein has a high aggregation propensity and the
ability to seed the aggregation of the cytosolic form. J Biol Chem. 2002;277(1):671–8. https://
doi.org/10.1074/jbc.M107045200.
Liu S, Sawada T, Lee S, et al. Parkinson’s disease-associated kinase PINK1 regulates Miro protein
level and axonal transport of mitochondria. PLoS Genet. 2012;8(3):e1002537. https://doi.org/
10.1371/journal.pgen.1002537.
Nutrition in Brain Aging: Its Relevance to Age-Associated Neurodegeneration 893
Ludtmann MHR, Angelova PR, Ninkina NN, Gandhi S, Buchman VL, Abramov AY. Monomeric
α-synuclein exerts a physiological role on brain ATP synthase. J Neurosci. 2016;36(41):
10510–21. https://doi.org/10.1523/JNEUROSCI.1659-16.2016.
Ludtmann MHR, Angelova PR, Horrocks MH, et al. α-Synuclein oligomers interact with ATP
synthase and open the permeability transition pore in Parkinson’s disease. Nat Commun.
2018;9(1):2293. https://doi.org/10.1038/s41467-018-04422-2.
Maccariello R, D’Angelo S. Impact of polyphenolic-food on longevity: an elixir of life. An
overview. Antioxidants. 2021;10(4):507. https://doi.org/10.3390/antiox10040507.
Madeo F, Carmona-Gutierrez D, Hofer SJ, Kroemer G. Caloric restriction mimetics against
age-associated disease: targets, mechanisms, and therapeutic potential. Cell Metab.
2019;29(3):592–610. https://doi.org/10.1016/j.cmet.2019.01.018.
Mangialasche F, Polidori MC, Monastero R, Ercolani S, Camarda C, Cecchetti R, Mecocci
P. Biomarkers of oxidative and nitrosative damage in Alzheimer and mild cognitive impairment.
Ageing Res Rev. 2009;8(4):285–305. https://doi.org/10.1016/j.arr.2009.04.002.
Maraki MI, Yannakoulia M, Stamelou M, et al. Mediterranean diet adherence is related to reduced
probability of prodromal Parkinson’s disease. Mov Disord. 2019;34(1):48–57. https://doi.org/
10.1002/mds.27489.
Marseglia A, Xu W, Fratiglioni L, et al. Effect of the NU-AGE diet on cognitive functioning in older
adults: a randomized controlled trial. Front Physiol. 2018;9:349. https://doi.org/10.3389/fphys.
2018.00349.
Martin CK, Bhapkar M, Pittas AG, et al. Effect of caloric restriction on mood, quality of life, sleep
and sexual function in healthy nonobese adults: the CALERIE 2 randomized clinical trial.
JAMA Intern Med. 2016;176(6):743–52. https://doi.org/10.1001/jamainternmed.2016.1189.
Martin-Montalvo A, Mercken EM, Mitchel SJ, et al. Metformin improves healthspan and lifespan
in mice. Nat Commun. 2013;4:2192. https://doi.org/10.1038/ncomms3192.
Masaki KH, Losonczy KG, Izmirlian G, Foley DJ, Ross GW, Petrovitch H, Havlik R, White
LR. Association of vitamin E and C supplement use with cognitive function and dementia in
elderly men. Neurology. 2000;54(6):1265–72. https://doi.org/10.1212/wnl.54.6.1265.
Maswood N, Young J, Tilmont E, et al. Caloric restriction increases neurotrophic factor and
attenuates neurochemical and behavioral deficits in a primate model of Parkinson’s disease.
Proc Natl Acad Sci U S A. 2004;101(52):18171–6. https://doi.org/10.1073/pnas.0405831102.
Mattison JA, Roth GS, Beasley TM, et al. Impact of caloric restriction on health and survival in
rhesus monkeys from the NIA study. Nature. 2012;489(7415):318–21. https://doi.org/10.1038/
nature11432.
McDaniel SS, Rensing NR, Thio LL, Yamada KA, Wong M. The ketogenic diet inhibits the
mammalian target of rapamycin (mTOR) pathway. Epilepsia. 2011;52(3):e7–e11. https://doi.
org/10.1111/j.1528-1167.2011.02981.x.
McDonald TJ, Cervenka MC. Ketogenic diets for adult neurological disorders. Neurotherapeutics.
2018;15(4):1018–31. https://doi.org/10.1007/s13311-018-0666-8.
Metcalfe-Roach A, Yu AC, Golz E, et al. MIND and Mediterranean diets associated with later onset
of Parkinson’s disease. Mov Disord. 2021;36(4):977–84. https://doi.org/10.1002/mds.28464.
Morita Y, Nogami M, Sakaguchi K, Okada Y, Hirota Y, Sugawara K, Tamori Y, Zeng F,
Murakami T, Ogawa W. Enhanced release of glucose into the intraluminal space of the intestine
Associated with metformin treatment as revealed by [18F]Fluorodeoxyglucose PET-MRI. Dia-
betes Care. 2020;43(8):1796–802. https://doi.org/10.2337/dc20-0093.
Morris MC, Tangney CC, Wang Y, Sacks FM, Benett DA, Aggarwal NT. MIND diet associated
with reduced incidence of Alzheimer’s disease. Alzheimers Dement. 2015a;11(9):1007–14.
https://doi.org/10.1016/j.jalz.2014.11.009.
Morris MC, Tangney CC, Wang Y, Sacks FM, Barnes LL, Bennett DA, Aggarwal NT. MIND diet
slow cognitive decline with aging. Alzheimers Dement. 2015b;11(9):1015–22. https://doi.org/
10.1016/j.jalz.2015.04.011.
Muller F, Lustgarten M, Jang Y, Richardson A, Van Remmen H. Trends in oxidative aging theories.
Free Radic Biol Med. 2007;43(4):477–503. https://doi.org/10.1016/j.freeradbiomed.2007.03.034.
894 W. Maruyama et al.
Rochon J, Bales CW, Ravussin E, et al. Design and conduct of the CALERIE study: comprehensive
assessment of long-term effects of reducing intake of energy. J Gerontol A Biol Sci Med Sci.
2011;66A(1):97–108. https://doi.org/10.1093/gerona/glq168.
Ros E, Martinez-Gonzalez MA, Estruch R, Salas-Salvado J, Fito M, Martinez JA, Corella
D. Mediterranean diet and cardiovascular health: teachings of the PREDIMEZ study. Adv
Nutr. 2014;5(3):330S–6S. https://doi.org/10.3945/an.113.005389.
Roth GS, Ingram DK. Manipulation of health span and function by dietary caloric restriction
mimetics. Ann N Y Acad Sci. 2016;1363:5–10. https://doi.org/10.1111/nyas.12834.
Rutjes AW, Denton DA, Di Nisio M, et al. Vitamin and mineral supplement for maintaining
cognitive function in cognitively healthy people in mild and late life. Cochrane Database Syst
Rev. 2018;12(12):CD-11906. https://doi.org/10.1002/14651858.CD011906.pub2.
Sampaio LP. Ketogenic diet for epilepsy treatment. Arq Neuropsiquiatr. 2016;74(10):842–8. https://
doi.org/10.1590/0004-282X20160116.
Schildknecht S, Gerding HR, Karreman C, Drescher M, Lashuel HA, Quteiro TF, Di Monte DA,
Leist M. Oxidative and nitrative alpha-synuclein modifications and proteostatic stress: implica-
tions for disease mechanisms and interventions in synucleinopathies. J Neurochem.
2013;125(4):491–511. https://doi.org/10.1111/jnc.12226.
Schwingshackl L, Morze J, Hoffmann G. Mediterranean diet and health status: active ingredients
and pharmacological mechanisms. Br J Pharmacol. 2020;177(6):1241–57. https://doi.org/10.
1111/bph.14778.
Selley ML. (E)-4-Hydroxy-2-nonenal may be involved in the pathogenesis of Parkinson’s disease.
Free Radic Biol Med. 1998;25(2):169–74. https://doi.org/10.1016/s0891-5849(98)00021-5.
Shamoto-Nagai M, Maruyama W, Hashizume Y, Yoshida M, Osawa T, Riederer P, Naoi M. In
parkinsonian substantia nigra, α-synuclein is modified by acrolein, a lipid peroxidation product,
and accumulates in the dopamine neurons with proteasome activity. J Neural Transm.
2007;114(12):1559–67. https://doi.org/10.1007/s00702-007-0789-2.
Shamoto-Nagai M, Hisaka S, Naoi M, Maruyama W. Modification of α-synuclein by lipid perox-
idation products derived from polyunsaturated fatty acids promotes toxic oligomerization: its
relevance to Parkinson disease. J Clin Biochem Nutr. 2018;62(3):207–12. https://doi.org/10.
3164/jcbn.18-25.
Shoulson I. DATATOP: a decade of neuroprotective inquiry. Parkinson Study Group. Deprenyl and
tocopherol antioxidative therapy of Parkinsonism. Ann Neurol. 1998;44(3 Suppl 1):S160–6.
Singh AK, Kashyap MP, Tripathi VK, Singh S, Garg G, Rizvi SI. Neuroprotection through rapamycin-
induced activation of autophagy and PI3K/Akt1/mTOR/ CREB signaling against amyloid-β-induced
oxidative stress, synaptic/ neurotransmission dysfunction, and neurodegeneration in adult rats. Mol
Neurobiol. 2017;54(8):5815–28. https://doi.org/10.1007/s12035-016-0129-3.
Smith DL Jr, Nagy TR, Allison DB. Calorie restriction: what recent results suggest for the future of
ageing research. Eur J Clin Investig. 2010;40(5):440–50. https://doi.org/10.1111/j.1365-2362.
2010.02276.x.
Snyder H, Mensah K, Theisler C, Lee J, Matouschek A, Wolozin BJ. Aggregated and monomeric
α-synuclein bind to the S6’ proteasomal protein and inhibit proteasomal function. J Biol Chem.
2003;278(14):11753–9. https://doi.org/10.1074/jbc.M208641200.
Taub PR, Ramirez-Sanchez I, Patel M, et al. Beneficial effects of dark chocolate on exercise
capacity in sedentary subjects: underlying mechanisms. A double blind, randomized, placebo
controlled trial. Food Funct. 2016;7(9):3686–93. https://doi.org/10.1039/c6fo00611f.
Tohgi H, Abe T, Yamazaki K, Murata T, Ishizaki E, Isobe C. Alterations of 3-nitrotyroeine
concentration in the cerebrospinal fluid during aging and in patients with Alzheimer’s disease.
Neurosci Lett. 1999;269(1):52–4. https://doi.org/10.1016/s0304-3940(99)00406-1.
Tosti V, Bertozzi B, Fontana L. Health benefits of the Mediterranean diet: metabolic and molecular
mechanisms. J Gerontol A Biol Sci Med Sci. 2018;73(3):318–26. https://doi.org/10.1093/gerona/
glx227.
896 W. Maruyama et al.
Tramutola A, Triplett JC, Di Domenico F, Niedowicz DM, Murphy MP, Coccia R, Perluigi M,
Butterfield DA. Alteration of mTOR signaling occurs early in the progression of Alzheimer disease
(AD): analysis of brain from subjects with pre-clinical AD, amnestic mild cognitive impairment and
late-stage AD. J Neurochem. 2015;133(5):739–49. https://doi.org/10.1111/jnc.13037.
Trichopoulou A, Martinez-Gonzalez MA, Tong TY, Forouhi NG, Khandelwal S, Prabhakaran D,
Mozaffarian D, de Lorgeril M. Definitions and potential health benefits of the Mediterranean
diet: views from experts around the world. BMC Med. 2014;12:112. https://doi.org/10.1186/
1741-7015-12-112.
Trinh D, Israw A, Ararhoon LR, Gleave JA, Nash JE. The multi-faced role of mitochondrial in the
pathology of Parkinson’s disease. J Neurochem. 2021;156(6):715–52. https://doi.org/10.1111/
jnc.15154.
Turner RS, Thomas RG, Craft S, et al. A randomized, double-blind, placebo-controlled trial of
resveratrol for Alzheimer disease. Neurology. 2015;85(16):1383–91. https://doi.org/10.1212/
WNL.0000000000002035.
Valls-Pedret C, Sala-Vila A, Sarra-Mir M, et al. Mediterranean diet and age-related cognitive
decline. A randomized clinical trial. JAMA Intern Med. 2015;175(7):1094–103. https://doi.
org/10.1001/jamainternmed.2015.1668.
van den Brink AC, Brouwer-Brolsma EM, Berendsen AAM, van de Rest O. The Mediterranean,
dietary approaches to stop hypertension (DASH), and Mediterranean-DASH intervention for
neurodegenerative delay (MIND) diets are associated with less cognitive decline and a lower
risk of Alzheimer’s disease – a review. Adv Nutr. 2019;10(6):1040–65. https://doi.org/10.1093/
advances/nmz054.
Van der Auwera I, Wera S, Van Leuven F, Henderson ST. A ketogenic diet reduces amyloid beta
40 and 42 in a mouse model of Alzheimer’s disease. Nutr Metab (Lond). 2005;2:28. https://doi.
org/10.1186/1743-7075-2-28.
Vanitallie TB, Nonas C, Di Rocco A, Boyar K, Hyams K, Heymsfield SB. Treatment of Parkinson
disease with diet-induced hyperketonemia: a feasibility study. Neurology. 2005;64(4):728–30.
https://doi.org/10.1212/01.WNL.0000152046.11390.45.
Varghese N, Werner S, Grimm A, Eckert A. Dietary mitophagy enhancer: a strategy for healthy
brain aging? Antioxidants. 2020;9(10):932. https://doi.org/10.3390/antiox9100932.
Vauzour D, Camprubi-Robles M, Miquel-Kergoat S, et al. Nutrition for the ageing brain: towards
evidence for an optimal diet. Ageing Res Rev. 2017;35:222–40. https://doi.org/10.1016/j.arr.
2016.09.010.
vB Hjelmborg J, Iachine I, Skytthe A, et al. Genetic influence on human lifespan and longevity.
Hum Genet. 2006;119(3):312–21. https://doi.org/10.1007/s00439-006-0144-y.
Wengreen H, Munger RG, Cutler A, et al. Prospective study of dietary approaches to stop
hypertension- and Mediterranean-style dietary patterns and age-related cognitive change: the
Cache Country study on memory, health and aging. Am J Clin Nutr. 2013;98(5):1263–71.
https://doi.org/10.3945/ajcn.112.051276.
Willcox BJ, Willcox DC. Caloric restriction, CR mimetics, and healthy aging in Okinawa: contro-
versies and clinical implications. Curr Opin Clin Nutr Metab Care. 2014;17(1):51–8. https://doi.
org/10.1097/MCO.0000000000000019.
Witte AV, Kerti L, Margulies DS, Flöel A. Effects of resveratrol on memory performance,
hippocampal functional connectivity, and glucose metabolism in healthy older adult.
J Neurosci. 2014;34(23):7862–70. https://doi.org/10.1523/JNEUROSCI.0385-14.2014.
Wyss-Coray T. Ageing, neurodegeneration and brain rejuvenation. Nature. 2016;539(7628):180–6.
https://doi.org/10.1038/nature20411.
Xilouri M, Vogiatzi T, Vekrellis K, Park D, Stefanis L. Abberant α-synuclein confers toxicity to
neurons in part through inhibition of chaperone-mediated autophagy. PLoS One. 2009;4(5):
e5515. https://doi.org/10.1371/journal.pone.0005515.
Nutrition in Brain Aging: Its Relevance to Age-Associated Neurodegeneration 897
Yin JX, Maalouf M, Han P, et al. Ketones block amyloid entry and improve cognition in an
Alzheimer’s model. Neurobiol Aging. 2016;39:25–37. https://doi.org/10.1016/j.
neurobiolaging.2015.11.018.
Yin W, Löf M, Pedersen NL, Sandin S, Fang F. Mediterranean dietary pattern at middle age and risk
of Parkinson’s disease: a Swedish cohort study. Mov Disord. 2021;36(1):255–60. https://doi.
org/10.1002/mds.28314.
Yoritaka A, Kawajiri S, Yamamoto Y, Nakahara T, Ando M, Hashimoto K, Nagase M, Saito Y,
Hattori N. Randomized, double-blind, placebo-controlled pilot trial of reduced coenzyme Q10
for Parkinson’s disease. Parkinsonism Relat Disord. 2015;21(8):911–9. https://doi.org/10.1016/
j.parkreldis.2015.05.022.
Zhang Y, He X, Wu X, et al. Rapamycin upregulates glutamate transporter and IL-6 expression in
astrocytes in a mouse model of Parkinson’s disease. Cell Death Dis. 2017;8(2):e2611. https://
doi.org/10.1038/cddis.2016.491.
Zhang S, Eitan E, Wu TY, Mattson MP. Intercellular transfer of pathogenic α-synuclein by
extracellular vesicles is induced by the lipid peroxidation product 4-hydroxynonenal. Neurobiol
Aging. 2018;61:52–65. https://doi.org/10.1016/j.neurobiolaging.2017.09.016.
Ethnopharmacology and the Development
of Psychoactive Drug: A Critical Overview
Elaine Elisabetsky
Contents
Natural Products, Ethnopharmacology, and Drug Development . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 900
Ethnopharmacology in the Context of Psychopharmacology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 901
Lost in Translation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 902
Like It or Not, You Do Want to See and Know How Homemade Medicines Are Made . . . 903
The Magic Bullet Fairytale . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 904
Too Little About Too Many . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 905
Putting the Cart Before the Horse . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 906
Perspectives for a Translational Ethnopharmacology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 907
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 909
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 909
Abstract
The purpose of this chapter is to provide a critical overview of ethnopharmacology
as a research strategy in the search of psychoactive drugs. Despite the prominent role
of natural products, alkaloids specially, in the development of psychopharmacology,
new plant-derived psychoactive drugs are lacking. Advantages of natural products
over synthetic compounds for interacting with biological systems of therapeutic
interest have been repeatedly documented. Advantages of ethnopharmacology-
driven plant collections over the biodiversity or chemotaxonomy-oriented ones
have also been reported. The main focus of this review is to discuss the advantages
to incorporate ethnopharmacology data into the research design in order to optimize
the identification of innovative drugs for neurological and mental illnesses. For that
matter, traditional medical systems need to be regarded and studied as such, instead
of just the plant species used to prepare the traditional formulas. Studies are used to
illustrate relevant methodological aspects for a proper evaluation of traditional
remedies used for treating the mentally ill, including the collection of clinical data
E. Elisabetsky (*)
Departamento de Bioquímica, Instituto de Ciência Básicas da Saúde, Universidade Federal do Rio
Grande do Sul, Porto Alegre, RS, Brazil
The advantages of screening natural products over synthetic molecules for drug
discovery have been advocated over decades and recently reviewed (Harvey et al.
2015). A pragmatic gauge is that 34% of the new medicines approved by the US Food
and Drug Administration (FDA) between 1981 and 2010 were based on natural
products and/or their direct derivatives. From a chemical point of view, the advantages
of natural products include (Harvey et al. 2015; Skirycz et al. 2016) (a) having a wide
range and high degree of stereochemistry pharmacophores, resulting in not only
bioactivity but likelihood to reach the site of action; (b) they characteristically fit the
molecular chemical descriptors adequate for bioactivity; (c) they probe a different,
wider, and more drug-like chemical space than the scaffolds present among commer-
cial drugs; (d) they cover parts of chemical characteristics of bioactivity targets that
medicinal chemistry compounds do not; and (e) the vast majority of all drugs have a
close match natural product (Harvey et al. 2015). Overcoming snags that marked the
field in the past are the advances in chemical methods for identifying active com-
pounds, better thought off libraries (such as “natural product-derived” and “natural
product-inspired”), improved nuclear magnetic resonance (NMR) techniques to solve
compound structures with minute (microgram range) samples, and new approaches
such as genomics, metabolomics, interactomics, and chemoinformatic (Harvey et al.
2015; Skirycz et al. 2016).
In view of such favorable evidence, it is somewhat surprising that the interest in
natural products as the basis for drug development has been coming in waves at least
since the 1980s. It has been estimated that only 5–15% of higher plants have been
systematically examined as sources of compounds with medicinal properties
(McChesney et al. 2007) nor had been scrutinized the 200,000 plant metabolites
already identified, not to mention the tens of thousands of plant-based homemade
remedies used daily for therapeutic purposes by the majority of mankind. The
explanation may combine the low yield and complex structure of compounds
(Skirycz et al. 2016) and the political nature of drug development, ruled by several
factors other than the medicine qualities/flaws. Specifically for bioprospection
initiatives, the sociopolitical issues, intellectual property rights (IPR), and benefit
sharing issues established by the 1992 Biological Convention are viewed as upfront
hurdles driving pharma interest away (Elisabetsky 1991).
In 1984 Professor Norman Farnsworth (University of Illinois at Chicago) posed
the question, “How can the well be dry when it is filled with water?” (Farnsworth
1984). As considered above, in the nearly four decades since this question was
posed, advances in phytochemistry and synthesis methods can largely overcome the
Ethnopharmacology and the Development of Psychoactive Drug: A Critical Overview 901
of such profound differences, peoples worldwide make use of concepts and practices
from various medical systems in order to maintain or reestablish health (Ceuterick and
Vandebroek 2017; Crandon 1986; Freymann et al. 2006). An informed multi-
disciplinary research team can extract from the description of signs, symptomatology,
and disease course reported by healers and/or users key information to formulate a
sound working hypothesis to foster the necessary hypothesis-driven laboratory exam-
ination (Elisabetsky 2007). Nevertheless, as I hope to illustrate with the following
examples, if ethnopharmacology data is worth using to increase the chances of success
in drug discovery, methods must be optimized. Considering the high rate of dropouts
when translationing preclinical data to successful clinical outcome, the report of a
specific therapeutic outcome (and adverse reactions) from plant-based formulas
ingested orally by men and women is an advantage for drug development that cannot
be underestimated (Mullane et al. 2014).
Lost in Translation
Traditional medical systems are organized as cultural systems (Etkin et al. 1990;
Moerman 2002) and may greatly differ in the meanings of health, disease, disease
etiologies, and cure (Wang 2012), resulting in medicinal practices not easily accom-
modated into the biomechanical paradigm of contemporary Western medicine. On top
of the expected differences in the understanding/beliefs of what causes the condition,
the very description of psychiatric symptomatology varies in different cultures (Gadit
2003; Shepard 1998; Weisman et al. 2000). Nonetheless, the common ground that can
be most effectively used to define the correspondence of emic and Western nosology is
the descriptions of symptomatology and the natural course of the disease. In the same
line, precious information can be obtained about the clinical outcome of therapies, as
when and how amelioration and/or resolution is perceived, adverse effects, etc.
Unfortunately, though, a significant (if not major) part of ethnopharmacology is
based on secondhand information and/or field data collected (and interpreted) by
personnel with no biomedical training. As a result, literature searches and cultural-
driven libraries end up built upon misleading information. Despite cultural beliefs,
informed clinical data collected in the field is highly instrumental for a sound working
hypothesis behind laboratory and/or to designed clinical studies (Graz 2013). Obser-
vational studies have been proposed as an initial step to optimize the choice of species
to be studied, the format of the research design, and to develop improved herbals
(Akubue and Mittal 1982; Graz 2013; Willcox et al. 2011).
Within the context of drug development the interest in traditional medical systems
resides on the assumption that plants processed as traditional medicines have a higher
probability to contain bioactive compounds. Nevertheless, the research designs rarely
(if ever, in industrial settings) consider basic pharmacological hallmarks, such as the
formulation and posology of traditional remedies. Disregarding information on how
any drug is formulated, how much, how often, and for how long it is used is against
basic principles widely admitted as key for a good outcome with any modern drug. As
no investigation on a synthetic drug would disregard the crucial issue of posology, it is
Ethnopharmacology and the Development of Psychoactive Drug: A Critical Overview 903
the lack of respect for traditional knowledge that results in the transformation of a
remedy into an extract or sample, ultimately limiting the quality of the evaluation and
the chances to identify new drugs. As Etkin put it, “ethnopharmacologists of all
backgrounds can enhance their work by projecting pharmacologic data against a
backdrop of medical ethnography” (Etkin 2001).
long-term use may have effects significantly different than acute (single administra-
tion) or sub-chronic (few administrations) challenge in any given molecular target.
Nevertheless, for planning screening or drug development programs, the traditional
posology is rarely taken into consideration. Understandably, the correspondence
between traditional use and its laboratory investigation collides with practical matters,
such as the need for higher amounts of compounds/extracts/fraction for in vivo, orally
administered, and/or repeated administration. Be as it may, the answer to the research
hypothesis of a given biological activity based on users’ claims cannot be conclusive
before a fair assessment is accomplished. Needless to say that for CNS drugs, this issue
is of upmost importance since it has been established that the long-term effects on
initial targets are often required to achieve a new functional state (Hyman and Nestler
1996). Importantly, in vitro methodologies are, unfortunately, more often than not
inadequate for these purposes (Houghton et al. 2007). The use of experimental models
with good predictability validity, usually practical requiring one single dose, is suitable
to initiate the investigation; if these initial results are positive, especially if showing
dose-dependent and sizable effects, a more elaborate phase using models with trans-
lational value and attention for the original posology can be launched.
Even when acute experiments are the only starting possibility (e.g., due to limited
test samples), a realistic notion of the traditional formulation and posology can be
informative. For instance, a hypothesis of antipsychotic properties was raised for
plant-based medicine used by Dr. Chidi, a Nigerian traditional psychiatrist. The
study led to the identification of the active compound, the indole alkaloid alstonine
showing a clear antipsychotic profile (positive, negative, and cognitive symptoms
mice models) and unique mechanism of action (Linck et al. 2015). Alstonine proved
to be active in mice models at 0.5–2.0 mg/kg dose range, far lower than the usual
20–200 mg/kg used in most studies, but compatible with the amount of the alkaloid
expected to be in the traditional preparation as recorded in field work (Costa-Campos
et al. 1998); had we concluded that alstonine had no antipsychotic effect based on the
negative results with 20 mg/kg or higher doses, the entire line of research would
have been abandoned.
Frequently the use of traditional remedies involves mixture of plants, and a single
plant-based remedy may already contain tens of compounds. Nevertheless, the
majority of the studies aim to identify “the” active component, which may be just
a part of what is of interest. For instance, caboclos from the state of Pará in the
Brazilian Amazon use the species Psychotria colorata to prepare a remedy to treat
earache: a handful of cut-up fresh flowers are mushed in milk (preferably “mother’s
milk”), packed in banana leaves, and left for a few minutes over warm ashes; the
resulting mixture is filtered through a piece of cloth and drops applied into the ear; it
is said that abdominal pain also responds to the species, but for that purpose roots
and fruits are mixed with water, left to boil, and the decoction taken orally
(Elisabetsky and Castilhos 1990). Extracts of P. colorata flowers showed marked
Ethnopharmacology and the Development of Psychoactive Drug: A Critical Overview 905
analgesic activity at various pain mice models (Elisabetsky et al. 1995), and
pyrrolidinoindoline alkaloids were identified as the major components in the flower
extract (but absent in the leaves). Several alkaloids with analgesic activity were
isolated and identified: hodgkinsine has a dual mechanism of action acting as opioid
agonist (Amador et al. 2000) and glutamate N-methyl-D-aspartate receptor
(NMDAR) antagonist; psychotridine acts as glutamate NMDAR antagonist
(Amador et al. 2001). It is the combined effects of the alkaloids, equivalent to use
morphine and dizocilpine, that result in the remedy’s potent analgesic effects. Dual
mechanism of action and/or various active components were also identified for
Sceletium tortuosum, where selective serotonin reuptake inhibitor (SSIR) was char-
acterized for mesembrine and mesembrone, whereas mesembrone also acts as
specific phosphodiesterase 4 inhibitor (Harvey et al. 2011). Examples such as
these call attention to the advantage of starting with in vivo analysis, given that
active compounds and/or mechanisms of action could have been lost if a target-
driven screening was the research approach.
Well-documented examples of biological activity resulting from various active
compounds (and lost or diminished when any one in separately assessed) include
plants of relevance in the herb market such as gingko (Liu et al. 2018; Tian et al.
2017), ginseng (Murthy et al. 2018), and kava (Celentano et al. 2019). The same
holds true for the synergic interactions of different active ingredients at complex
formulas, such as in the Ayahuasca beverage, the Japanese Kampö (Satoh 2013), and
traditional Chinese medicine (TCM) (Hong et al. 2017). Unfortunately though, even
with the changes in mainstream pharmacology with regard to multiple mechanisms
of action and network pharmacology (Hopkins 2007, 2008), a significant part of
medicinal plant research still pursue the active compound.
The sample was eventually reduced to 300 studies selected on the basis of the
scientific quality of the reports, from which 61 compounds were identified as being
active at the CNS. The Web of Science survey was repeated using the compound name
and the same CNS activities as keywords. After dismissing reports with scientifically
frail data, the sample was reduced to 34 compounds subjected to thorough revision in
the relevant scientific literature. Of these, 25 compounds were reported to have
anxiolytic activity, 9 with hypnotic properties, 18 as anti-convulsants, 9 as antidepres-
sants, and 10 showed a neuroprotective profile, based on at least one behavior and/or
in vivo electrophysiology assay. Since anxiolytics, hypnotics, and anti-convulsants are
sedatives (psycholetics), with known drugs often sharing the same mechanism of
action and effects (at different doses and efficacy), it was somewhat expected that
the same compounds were active in these three domains. Nevertheless, only six
(phytol, α-asarone, γ-decanolactone, safranal, dehydrofukinone, and oleamide) show
all three properties. Of the 34 reviewed compounds only for linalool, α-asarone, and
thymoquinone a substantial body of data is available, studies were replicated by
different research groups, therefore showing a reliable psychoactive profile
(as antiepileptic/anti-convulsant). Thymol is active in the three most common antide-
pressant experimental models (two with predictive and one with face validity) and
shows mechanisms of action relevant to depression (affecting noradrenaline, serotonin,
and neuroinflammation). Among the compounds claimed to act as neuroprotectors,
data is scattered over compounds and models, except for antioxidant activity assessed
in all but one compound. Several compounds had only one study reported or have been
studied by one group only, which limits the data reliability (Ioannidis 2018). More
often than not, only one experimental model is used, whereas a battery of models
would yield in more reliable picture and rule out false positives. Another potential
limitation is that the mechanism of action is often inferred from the antagonist effect
performed with single-dose experiments, a limitation also noted by Zhu and colleagues
(Zhu et al. 2014) in their review of anti-convulsant compounds isolated from medicinal
plants. Though acceptable in exploratory research (Kenakin et al. 2014), these studies
are unlikely to be selected for drug development before a far more rigorous analysis is
in place (Kimmelman et al. 2014).
Unfortunately, the pattern of a small set of experiments on a wide range of species
is dominant (obviously excluding the best known and/or important commercialized
species). Perhaps the idea of original research is translated into studying a species for
which most data is lacking. Be what may, the lack of accumulated data, including a
battery of in vivo and in vitro tests, phytochemistry, and toxicology, certainly does
not help to make a compelling case for drug development.
(Linde et al. 2008; Ng et al. 2017), though the definition of its active compounds and
mechanisms of action are still unclear (Barnes et al. 2001). The research design often
employed to “validate” traditional uses of medicinal plants by starting with in vitro
tests may gear the search away from truly innovative drugs.
For instance, 34 medicinal species (75 extracts) associated which ethnophar-
macology information interpreted as possible emic description of depressive states
were screened for serotonin transporter affinity. From the five positive hits, two
species were discarded for having compounds deemed inappropriate for long-term
use (tropane alkaloids and cardiac glycosides); the three remaining species were
examined for activity at the forced swimming and tail suspension tests showing
varying positive effects. Bio-guided fractionation led to the identification of
buphanamine-type alkaloids (such as lycorine) and the monoterpene lactone (-)-
loliolide (Jäger 2015). Were the traditional remedies used for long periods? Could
the ones ruled out be fast-acting antidepressants? Did the traditional formulation
contain significant amounts of the potentially toxic alkaloids found in the extracts?
Does the in vitro affinity translate to sizable effect in relevant CNS areas? Do
lycorine and (-)- loliolide administered for a long enough time result in the same
serotoninergic neurotransmitter changes observed with known SSRI? These are
questions that need to be addressed if the data is to be used as incentive for drug
development (trustworthy herbal or new chemical entity). Likewise, a wide range of
species were found to possess MAO-A inhibition properties (Viña et al. 2012), often
due to the presence of quercetin (Chimenti et al. 2006).
The failure to incorporate detailed traditional use (e.g., posology, observed
effects, and side effects) in the research design, the choice of workflow (e.g.,
in vitro before in vivo), and the limited use of experimental models with good
translational value often leads to a somewhat misapplied investment in time and
resources in the study of plant species used for medical purposes. If acceptable for
the varied academic purposes, this modus operandi adds little to the discovery of
innovative CNS drugs. It also may well be part of the reasons to explain the pathetic
yield (and cost/benefit) of high-throughput screenings that were once the hype and
hope in the search for drugs from natural origin (Fox et al. 2006).
Starting the evaluation of medicinal plants with methodologies that gauge the
effect on experimental models with translational value, rather than checking if it
matches a predetermined mechanism of action, may be especially relevant in the
field of mental disorders for which the pathological basis are still elusive (Ledford
2014). A fair assessment requires hypothesis-driven and customized research
designs that consider modes of preparation (types and dose range of components)
and posology (single x fractionated doses) of traditional medicines. As in biomedical
sciences in general, a more rigorous standard is required to afford credibility and
reproducibility to ethnopharmacology studies (Ioannidis 2005), in which the null
hypothesis is often accepted or rejected even when plant part, type of extract, dose
range, and the experimental model might have been inadequate or bear little corre-
spondence to the traditional use.
It is a consensus that the pharmaceutical potential of natural products has yet to be
explored (Cragg and Newman 2013). The preference of combinatorial over natural
product libraries may have contributed to the small number of new chemical entities,
fueling the pharma productivity crisis (Ogbourne and Parsons 2014). Yet, only 5%
out of 2000 studies including ethnopharmacology in the title or keywords include
CNS activities (Heinrich 2015). Collaborative efforts in drug development have
been put in place where consortiums are formed with specific goals such as defining
biomarkers or sharing specific data (Patridge et al. 2016; Wehling 2009). Collabo-
rative efforts, perhaps steered by scientific organizations devoted to specific neuronal
and mental disorders, could collate different academic skills and resources to attain a
substantial enough body of knowledge on promising compounds/species to attract
industrial research.
The extended interpretation of cause-effect relationships in physics calls for the
substitution of “cause” to “determining conditions,” where all the conditions of a
process or state are equally important. For complex diseases such as mental disor-
ders, where genetic, developmental, and ontogenetic factors are increasingly recog-
nized, the concept of determination fits better than causality (Vineis and Porta 1996).
When diseases are understood as processes, where an interplay of multiple factors
has to be considered, it becomes attractive to study complex plant formulas that may
simultaneously modulate more than one target (Rasoanaivo et al. 2011; Roth et al.
2004). In order to truly profit from studying traditional medical systems as such,
rather than medicinal plants as sources of bioactive molecules, ethnopharma-
cologists, pharmacologists, and phytochemists need to be unprejudiced and open-
minded, allowing an unbiased appreciation of potentially new drug paradigms. The
somewhat new paradigm of network pharmacology (Hopkins 2008) bears more
likeness to traditional remedies (more than one active compound, various mecha-
nisms of action) than the now outdated single bullet – yet phytomedicine is often
seen as messy and untrustworthy. Ethnopharmacology combined with a rigorous
hypothesis-driven laboratory scrutiny, by operating a comprehensive understanding
of traditional medical concepts and practices, can be a fruitful strategy to innovative
psychoactive drug development.
Ethnopharmacology and the Development of Psychoactive Drug: A Critical Overview 909
Cross-References
References
Akubue PI, Mittal GC. Clinical evaluation of a traditional herbal practice in Nigeria: a preliminary
report. J Ethnopharmacol. 1982;6:355–9. https://doi.org/10.1016/0378-8741(82)90056-3.
Amador TA, Verotta L, Nunes DS, Elisabetsky E. Antinociceptive profile of Hodgkinsine. Planta
Med. 2000;66:770–2. https://doi.org/10.1055/s-2000-9604.
Amador TA, Verotta L, Nunes DS, Elisabetsky E. Involvement of nmda receptors in the analgesic
properties of psychotridine. Phytomedicine. 2001;8:202–6. https://doi.org/10.1078/0944-7113-
00025.
Anderson LM, Scrimshaw SC, Fullilove MT, Fielding JE, Normand J. Culturally competent
healthcare systems: a systematic review. Am J Prev Med. 2003;24:68–79. https://doi.org/10.
1016/S0749-3797(02)00657-8.
Aponso M, Patti A, Bennett LE. Dose-related effects of inhaled essential oils on behavioural
measures of anxiety and depression and biomarkers of oxidative stress. J Ethnopharmacol.
2020;250:112469. https://doi.org/10.1016/j.jep.2019.112469.
Barnes J, Anderson LA, Phillipson JD. St John’s wort (Hypericum perforatum L.): a review of its
chemistry, pharmacology and clinical properties. J Pharm Pharmacol. 2001;53:583–600. https://
doi.org/10.1211/0022357011775910.
Birari RB, Bhutani KK. Pancreatic lipase inhibitors from natural sources: unexplored potential.
Drug Discov Today. 2007;12:879–89. https://doi.org/10.1016/j.drudis.2007.07.024.
Celentano A, Tran A, Testa C, Thayanantha K, Tan-Orders W, Tan S, Syamal M, McCullough MJ,
Yap T. The protective effects of Kava (Piper Methysticum) constituents in cancers: a systematic
review. J Oral Pathol Med. 2019;48:510–29. https://doi.org/10.1111/jop.12900.
Ceuterick M, Vandebroek I. Identity in a medicine cabinet: discursive positions of Andean migrants
towards their use of herbal remedies in the United Kingdom. Soc Sci Med. 2017;177:43–51.
https://doi.org/10.1016/j.socscimed.2017.01.026.
Chimenti F, Cottiglia F, Bonsignore L, Casu L, Casu M, Floris C, Secci D, Bolasco A, Chimenti P,
Granese A, Befani O, Turini P, Alcaro S, Ortuso F, Trombetta G, Loizzo A, Guarino I. Quercetin
as the active principle of Hypericum hircinum exerts a selective inhibitory activity against
MAO-A: extraction, biological analysis, and computational study. J Nat Prod. 2006;69:945–9.
https://doi.org/10.1021/np060015w.
Costa-Campos L, Lara DR, Nunes DS, Elisabetsky E. Antipsychotic-like profile of alstonine.
Pharmacol Biochem Behav. 1998;60:133–41. https://doi.org/10.1016/s0091-3057(97)00594-7.
Cragg GM, Newman DJ. Natural products: a continuing source of novel drug leads. Biochim
Biophys Acta. 2013;1830:3670–95. https://doi.org/10.1016/j.bbagen.2013.02.008.
910 E. Elisabetsky
Crandon L. medical dialogue and the political economy of medical pluralism: a case from rural
highland Bolivia. Am Ethnol. 1986;13:463–76. https://doi.org/10.1525/ae.1986.13.3.
02a00040.
Elisabetsky E. Sociopolitical, economical and ethical issues in medicinal plant research. J
Ethnopharmacol. 1991;32:235–9. https://doi.org/10.1016/0378-8741(91)90124-V.
Elisabetsky E. Phytotherapy and the new paradigm of drugs mode of action. Scientia et Technica.
2007;33:459–64.
Elisabetsky E, Brum L. Linalool as active component of traditional remedies: anticonvulsant
properties and mechanisms of action. Curare. 2003;26:45–52.
Elisabetsky E, Castilhos ZC. Plants used as analgesics by Amazonian Caboclos as a basis for
selecting plants for investigation. Int J Crude Drug Res. 1990;28:309–20. https://doi.org/10.
3109/13880209009082838.
Elisabetsky E, Amador TA, Albuquerque RR, Nunes DS, do Carvalho Ana CT. Analgesic activity
of Psychotria colorata (Willd. ex R. & S.) Muell. Arg. alkaloids. J Ethnopharmacol. 1995;48:
77–83. https://doi.org/10.1016/0378-8741(95)01287-N.
Elisabetsky E, Brum LF, Souza DO. Anticonvulsant properties of linalool in glutamate-related
seizure models. Phytomed Int J Phytother Phytopharm. 1999;6:107–13. https://doi.org/10.1016/
s0944-7113(99)80044-0.
Elisabetsky, E. Nunes, Domingos S. Central Nervous System Effects of Essential Oil Compounds.
In: K. Husnu Can Baser, Gerhard Buchbauer. (Org.). Handbook of Essential Oils, 3rd edition.
London: Taylor and Francis Group, 2020, v. 1, p. 304–344.
Etkin NL. Perspectives in ethnopharmacology: forging a closer link between bioscience and
traditional empirical knowledge. J Ethnopharmacol. 2001;76:177–82. https://doi.org/10.1016/
s0378-8741(01)00232-x.
Etkin NL, Elisabetsky E. Seeking a transdisciplinary and culturally germane science: the future of
ethnopharmacology. J Ethnopharmacol. 2005;100:23–6. https://doi.org/10.1016/j.jep.2005.05.
025.
Etkin NL, Ross PJ, Muazzamu I. The indigenization of pharmaceuticals: therapeutic transitions in
rural Hausaland. Soc Sci Med. 1990;30:919–28. https://doi.org/10.1016/0277-9536(90)90220-M.
Farnsworth NR. How can the well be dry when it is filled with water? Econ Bot. 1984;38:4–13.
https://doi.org/10.1007/BF02904411.
Fox S, Farr-Jones S, Sopchak L, Boggs A, Nicely HW, Khoury R, Biros M. High-throughput
screening: update on practices and success. J Biomol Screen. 2006;11:864–9. https://doi.org/10.
1177/1087057106292473.
Freymann H, Rennie T, Bates I, Nebel S, Heinrich M. Knowledge and use of complementary and
alternative medicine among British undergraduate pharmacy students. Pharm World Sci.
2006;28:13–8. https://doi.org/10.1007/s11096-005-2221-z.
Gadit AA. Ethnopsychiatry – a review. JPMA. 2003;53:1–6.
Graz B. What is “clinical data”? Why and how can they be collected during field surveys on
medicinal plants? J Ethnopharmacol. 2013;150:775–9. https://doi.org/10.1016/j.jep.2013.08.
036.
Gyllenhaal C, Kadushin MR, Southavong B, Sydara K, Bouamanivong S, Xaiveu M, Xuan LT, Hiep
NT, Hung NV, Loc PK, Dac LX, Bich TQ, Cuong NM, Ly HM, Zhang HJ, Franzblau SG, Xie H,
Riley MC, Elkington BG, Nguyen HT, Waller DP, Ma CY, Tamez P, Tan GT, Pezzuto JM, Soejarto
DD. Ethnobotanical approach versus random approach in the search for new bioactive com-
pounds: support of a hypothesis. Pharm Biol. 2012;50:30–41. https://doi.org/10.3109/13880209.
2011.634424.
Harvey AL, Young LC, Viljoen AM, Gericke NP. Pharmacological actions of the South African
medicinal and functional food plant Sceletium tortuosum and its principal alkaloids. J
Ethnopharmacol. 2011;137:1124–9. https://doi.org/10.1016/j.jep.2011.07.035.
Harvey AL, Edrada-Ebel R, Quinn RJ. The re-emergence of natural products for drug discovery in
the genomics era. Nat Rev Drug Discov. 2015;14:111–29. https://doi.org/10.1038/nrd4510.
Ethnopharmacology and the Development of Psychoactive Drug: A Critical Overview 911
Murthy HN, Dandin VS, Park S-Y, Paek K-Y. Quality, safety and efficacy profiling of ginseng
adventitious roots produced in vitro. Appl Microbiol Biotechnol. 2018;102:7309–17. https://
doi.org/10.1007/s00253-018-9188-x.
Newman DJ, Cragg GM. Natural products as sources of new drugs over the nearly four decades
from 01/1981 to 09/2019. J Nat Prod. 2020;83:770–803. https://doi.org/10.1021/acs.jnatprod.
9b01285.
Ng QX, Venkatanarayanan N, Ho CYX. Clinical use of Hypericum perforatum (St John’s wort) in
depression: a meta-analysis. J Affect Disord. 2017;210:211–21. https://doi.org/10.1016/j.jad.
2016.12.048.
Nunes DS. Chemical approaches to the study of ethnomedicines. In: Medical resources of the
tropical forest. Columbia University Press; 1996. p. 41–7.
Ogbourne SM, Parsons PG. The value of nature’s natural product library for the discovery of New
Chemical Entities: the discovery of ingenol mebutate. Fitoterapia. 2014;98:36–44. https://doi.
org/10.1016/j.fitote.2014.07.002.
Patridge E, Gareiss P, Kinch MS, Hoyer D. An analysis of FDA-approved drugs: natural products
and their derivatives. Drug Discov Today. 2016;21:204–7. https://doi.org/10.1016/j.drudis.
2015.01.009.
Rasoanaivo P, Wright CW, Willcox ML, Gilbert B. Whole plant extracts versus single compounds
for the treatment of malaria: synergy and positive interactions. Malar J. 2011;10(Suppl 1):S4.
https://doi.org/10.1186/1475-2875-10-S1-S4.
Roth BL, Sheffler DJ, Kroeze WK. Magic shotguns versus magic bullets: selectively non-selective
drugs for mood disorders and schizophrenia. Nat Rev Drug Discov. 2004;3:353–9. https://doi.
org/10.1038/nrd1346.
Satoh H. Pharmacological characteristics of Kampo medicine as a mixture of constituents and
ingredients. J Integr Med. 2013;11:11–6. https://doi.org/10.3736/jintegrmed2013003.
Shepard GH. Psychoactive plants and ethnopsychiatric medicines of the Matsigenka. J Psychoac-
tive Drugs. 1998;30:321–32. https://doi.org/10.1080/02791072.1998.10399708.
Silva Brum LF, Emanuelli T, Souza DO, Elisabetsky E. Effects of linalool on glutamate release and
uptake in mouse cortical synaptosomes. Neurochem Res. 2001;26:191–4. https://doi.org/10.
1023/a:1010904214482.
Skirycz A, Kierszniowska S, Méret M, Willmitzer L, Tzotzos G. Medicinal bioprospecting of the
Amazon Rainforest: a modern Eldorado? Trends Biotechnol. 2016;34:781–90. https://doi.org/
10.1016/j.tibtech.2016.03.006.
Tian J, Liu Y, Chen K. Ginkgo biloba extract in vascular protection: molecular mechanisms and
clinical applications. Curr Vasc Pharmacol. 2017;15 https://doi.org/10.2174/
1570161115666170713095545.
Verpoorte R. Good practices: the basis for evidence-based medicines. J Ethnopharmacol. 2012;140:
455–7. https://doi.org/10.1016/j.jep.2012.02.033.
Viña D, Serra S, Lamela M, Delogu G. Herbal natural products as a source of monoamine oxidase
inhibitors: a review. Curr Top Med Chem. 2012;12:2131–44. https://doi.org/10.2174/
156802612805219996.
Vineis P, Porta M. Causal thinking, biomarkers, and mechanisms of carcinogenesis. J Clin
Epidemiol. 1996;49:951–6. https://doi.org/10.1016/0895-4356(96)00118-7.
Wang Q. Individualized medicine, health medicine, and constitutional theory in Chinese medicine.
Front Med. 2012;6:1–7. https://doi.org/10.1007/s11684-012-0173-y.
Wang Z-J, Heinbockel T. Essential oils and their constituents targeting the GABAergic system and
sodium channels as treatment of neurological diseases. Mol Basel Switz. 2018;23 https://doi.
org/10.3390/molecules23051061.
Wehling M. Assessing the translatability of drug projects: what needs to be scored to predict
success? Nat Rev Drug Discov. 2009;8:541–6. https://doi.org/10.1038/nrd2898.
Ethnopharmacology and the Development of Psychoactive Drug: A Critical Overview 913
Weisman AG, López SR, Ventura J, Nuechterlein KH, Goldstein MJ, Hwang S. A comparison of
psychiatric symptoms between Anglo-Americans and Mexican-Americans with Schizophrenia.
Schizophr Bull. 2000;26:817–24. https://doi.org/10.1093/oxfordjournals.schbul.a033496.
Willcox ML, Graz B, Falquet J, Diakite C, Giani S, Diallo D. A “reverse pharmacology” approach
for developing an anti-malarial phytomedicine. Malar J. 2011;10(Suppl 1):S8. https://doi.org/
10.1186/1475-2875-10-S1-S8.
Wu C, Lee S-L, Taylor C, Li J, Chan Y-M, Agarwal R, Temple R, Throckmorton D, Tyner
K. Scientific and regulatory approach to botanical drug development: A U.S. FDA perspective.
J Nat Prod. 2020;83:552–62. https://doi.org/10.1021/acs.jnatprod.9b00949.
Zhu H-L, Wan J-B, Wang Y-T, Li B-C, Xiang C, He J, Li P. Medicinal compounds with
antiepileptic/anticonvulsant activities. Epilepsia. 2014;55:3–16. https://doi.org/10.1111/epi.
12463.
Antidepressants: Molecular Aspects of
SSRIs
Contents
Citalopram Versus S-Citalopram: Molecular Insights into Antidepressant Treatment . . . . . . . . 916
SSRI’s Mode of Action: Targeting the Serotonin Transporter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 917
Citalopram Binding Properties Underlying the Antidepressant Response . . . . . . . . . . . . . . . . . 919
Citalopram-SERT Interaction at the Cellular Level: Regulation of SERT Cell
Surface Exposure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 924
Modulation of SERT Binding Sites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 924
Effect of 5-HT on Citalopram Modulation of SERT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 925
Kinase’s Modulation of SERT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 926
Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 927
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 928
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 928
A. Etievant
Integrative and Clinical Neurosciences EA481, University of Bourgogne Franche-Comté,
Besançon, France
e-mail: adeline.etievant@univ-fcomte.fr; adeline.etievant@gmail.com
N. Haddjeri
Stem Cell and Brain Research Institute, INSERM, U1208, Bron, France
Université de Lyon, Université Lyon 1, Lyon, France
e-mail: nasser.haddjeri@inserm.fr
T. Lau (*)
Central Institute of Mental Health, Department of Translational Brain Research, Medical Faculty
Mannheim; Heidelberg University, Hector Institute for Translational Brain Research, Mannheim,
Germany
German Cancer Research Center (DKFZ), Heidelberg, Germany
e-mail: thorsten.lau@zi-mannheim.de
Abstract
Major depressive disorders are among the most common psychiatric diseases
featuring an insufficient serotonin signaling as the most prominent core patho-
mechanism. Therefore, selective serotonin re-uptake inhibitors are firstly
employed to restore serotonin neurotransmission in depressed patients. One
member of this class of antidepressants, which selectively targets the serotonin
transporter, is citalopram. Citalopram was first synthesized in 1972 and consists
of equal molecular amounts of the isomers R-citalopram and S-citalopram,
generally referred to as escitalopram, which exerts the antidepressant effect.
Underlying citalopram’s mode of action is an allosteric mechanism, by which
citalopram modulates its own binding to the primary binding site at the serotonin
transporter. The following chapter will provide an overview of recent findings on
the molecular binding characteristics determining citalopram’s docking to the
serotonin transporter and the allosteric modulation of the citalopram-induced
antidepressant response. Here, the focus will be on the acute citalopram-induced
effects, which result in dose-dependent internalization of serotonin transporter
molecules and diminished neuronal activity of serotonergic neurons.
Major depressive disorder is one of the most common mental disorders, and according
to a report made by the World Health Organization, in 2015 about 300 million people
suffered from this disease worldwide. Suffering from depression may become a
serious health threat, especially when depression symptoms persist and turn into
long-term burden with moderate or even severe intensities. A more threatening
situation is caused by the lethal outcome of depression: roughly a million people of
patients diagnosed with depression commit suicide each year, which actually is the
second leading cause of death in people aged 15–30 years (World Health Organization
2017). Furthermore, based on the projections of a steadily increasing numbers of
patients, MDD will be the most common disease by the year 2030 (Mathers and
Loncar 2006). Although various treatment options in terms of pharmacological inter-
vention are available, two major problems still remain: (1) there is a significant number
of patients, which does not respond to antidepressant pharmacotherapy and suffer from
chronic and refractory form of depression, and (2) there is a delayed onset of beneficial
antidepressant response in patients that may last up to 2 months (Berton and Nestler
2006; Stahl et al. 2013; Gibbons et al. 2012; Akil et al. 2018).
Up to date, the pathophysiology of depression is poorly understood, and the
interaction of the relation of identified core patho-mechanisms is not well-known.
The most postulated hypotheses focus on dysfunctions of the hypothalamic-pitui-
tary-adrenal axis, immunological disturbances, or altered monoaminergic neuronal
systems. With regard to impaired monoaminergic neurotransmission, diminished
and insufficient serotonin (5-hydroxytryptamine, 5-HT) signaling was identified to
Antidepressants: Molecular Aspects of SSRIs 917
be a main feature of depression (Boku et al. 2018; Dell’Osso et al. 2016; Kraus et al.
2017; Nestler et al. 2002; Robson et al. 2017; Yu et al. 2008). 5-HT signaling is
maintained by serotonergic neurons, which are located in the raphe nuclei.
From here, these neurons give rise to an extensive axonal network, building
projections into different brain regions, such as prefrontal cortex or hippocam-
pus, and providing a constant 5-HT tonus inside the innervated brain regions
(Törk 1990; Hornung 2003; Puig and Gulledge 2011; Quentin et al. 2018; Vizi
2000). One major approach in antidepressant pharmacotherapy is to restore 5-
HT signaling by employing selective serotonin re-uptake inhibitors (SSRI).
These molecules primarily target and functionally inhibit the serotonin trans-
porter (SERT), which plays a key role in serotonergic neurotransmission: its
uptake activity terminates 5-HT signaling and thereby determines duration of
strength of 5-HT receptor activation (Chen et al. 2004, 2005). Upon binding to
their targets, SSRI prevent SERT-dependent 5-HT re-uptake into serotonergic neu-
rons. An acute functional consequence of re-uptake inhibition is the increase of
extracellular 5-HT, which is now able to prolong the activation of 5-HT receptors
and thereby enhance 5-HT signaling (Haenisch and Bonisch 2011; Morrissette and
Stahl 2014). Over the past decades, a number of SSRI were developed as a first
choice antidepressant therapy, including fluoxetine, sertraline, and citalopram, which
was first synthesized by Klaus Bogeso in 1972 (Bogeso and Sánchez 2012; Bigler et
al. 1977). The following chapter will start with a brief overview of SSRI’s
molecular impact on 5-HT signaling and then focus on citalopram-SERT interac-
tion on the cellular level, addressing citalopram-induced SERT internalization and
the allosteric modulation of SERT function.
Fig. 1 Effect of acute citalopram treatment on neuronal activity and SERT availability in 5-
HT neurons. Acute application of citalopram (C) results in binding to and inhibition of SERT
molecules located at the cell surface of serotonergic neurons, thereby elevating extracellular
bioactive 5-HT concentrations (S). 5-HT is now able to enhance stimulation of 5-HT autoreceptors
(R) expressed on serotonergic neurons, which are postulated to differentially modulate 5-HT
neuronal activity. Here, 5-HT 1A autoreceptors (1A) play a central role, as their activation leads
to subsequent inhibition of 5-HT release. Regarding the availability of SERT molecules, citalopram-
SERT interaction triggers the internalization of SERT molecules (+), shifting the kinase-regulated
shuttling of SERT to and from the cell surface toward retrograde trafficking. Extra: extracellular;
intra: intracellular
retrograde SERT trafficking that occurs during the second phase seems to be a time-
and concentration-dependent event. Overall, the redistribution of SERT to internal
compartments during acute SSRI treatment reduces the amount of SERT molecules
that takes up 5-HT after release and thereby further enhances 5-HT signaling
(Benmansour et al. 2002; Kittler et al. 2010; Matthäus et al. 2016). The increased
5-HT levels also act on 5-HT autoreceptors, namely, 5-HT receptor 1A (5-HTR1A)
and putatively 5-HT receptor 2B (5-HTR2B), which are key molecules of the
autoregulatory feedback loop of serotonergic neuronal activity (Blier et al. 1998).
In general, activation of 5-HTR1A autoreceptors inhibits neuronal 5-HT firing
activity inside the dorsal raphe nuclei; indeed, acute antidepressant treatment was
shown to cause a shutdown of 5-HT release (Aghajanian et al. 1970; Piñeyro and
Blier 1999; Courtney and Ford 2016; Mnie-Filali et al. 2016; El Mansari et al. 2005).
Similarly, acute administration of citalopram, as well as its isomer S-citalopram,
suppresses the neuronal activity of serotonergic neurons via activation of 5-HTR1A.
Furthermore, prolonged administration of both substances diminished the spontane-
ous neuronal firing activity of serotonergic neuron. The diminished activity recov-
ered after desensitization of somatodendritic 5-HTR1A, which took 2 weeks in case
of S-citalopram treatment, yet needed one additional week to fully recover in case of
using the racemate citalopram for antidepressant treatment (El Mansari et al. 2005).
In contrast to 5-HTR1A autoreceptors, evidence for a positive modulation of 5-HT
signaling by 5-HTR2B autoreceptors was recently provided in mice (Belmer et al.
2018) but apparently not in rat (Cathala et al. 2019), leaving the question
Antidepressants: Molecular Aspects of SSRIs 919
amino acid residues Asp98, Ile172, and Ser438 were identified for citalopram-SERT
interaction and additionally implied Asn177 to be important for escitalopram-medi-
ated SERT inhibition (Barker et al. 1999, 1998; Henry et al. 2006; Andersen et al.
2009). Regarding its binding to SERT, a study suggested that citalopram adjusts its
docking position with an orientation of the cyano group locating inside the trans-
porter protein to most likely interact with Val343, while the fluorine group is
spatially orientated toward Ile172, Ala173, and Asn177 (Andersen et al. 2010).
Another study addressing the ligand binding characteristics of citalopram-SERT
interaction revealed that the fluorine atom of S-citalopram is located in the proximity
of the amino acid residues Ala173 and Thr439, while the cyano group is proximally
orientated toward the residue of Phe341. For R-citalopram binding to SERT, the
orientation of these groups was found to be in the reversed orientation (Koldso et al.
2010). The recent study by Coleman and colleagues, which analyzed molecular
binding characteristics with the help of X-ray crystallography of human SERT,
identified three subsites regarding the S1 binding site for S-citalopram binding
(Coleman et al. 2016). The residues of Tyr95, Asp98, Tyr176, Ile172, and Phe341
were identified to establish the S1 subsite A, with the latter three residues most likely
providing interaction partners for the hydrophobic groups of S-citalopram and other
antidepressants. The S1 subsite B, a hydrophobic cavity built by the residues of
Ala169, Ala173, Ser439, and Leu443, was identified to be of high importance for S-
citalopram’s high-affinity binding, and to be a signature structure of human SERT,
which is not found in other monoamine transporters. Finally, the residues Phe335,
Thr497, and Val501 form S1 subsite C, which provides a nonpolar/polar region for
interaction of S-citalopram’s fluoro and cyano groups. Compared to the findings
reported using two different docking models based on the leucine transporter
(Andersen et al. 2010), Coleman and colleagues did not identify the residues
Asn177 and Val 343 to be involved in S-citalopram binding to human SERT. Yet
both approaches identified an overlapping set of amino residues to be important
determinants of citalopram docking to SERT. Interestingly, in transgenic mice
carrying a single isoleucine to methionine substitution (I172M), an amino acid
exchange directly affecting the biochemical characteristics of the S1 binding site,
an extremely diminished inhibitory efficacy of multiple antidepressant agents was
observed, including citalopram (Thompson et al. 2011; Henry et al. 2006).
the transmembrane domains 1, 6, 10, and 11, regions identical to those carrying the
amino acid exchanges for the transgene allosteric site-missing animal models.
However, the binding studies have revealed different interacting residues for S-
citalopram docking to the allosteric site. The residues of Ala331, Gln332, and
Phe556 most likely establish a groove and interaction domains for S-citalopram’s
fluoro group, while the extracellularly located amino acids Arg104, Asp308, and
Glu494 provide interaction partners for the cyano group. Although the amino acid
exchanges in the animal models affect different amino acid residues, the exchanges
may most likely alter the biochemical surroundings for efficient citalopram-SERT
interaction at the allosteric S2 binding site. Interestingly, Anderson and colleagues
were not able to detect robust docking kinetics of S-citalopram to the allosteric
binding site in both kinetic models applied in their study (Andersen et al. 2010).
Recently, Zhu and colleagues employed atomic force microscopy to gain further
insight on citalopram’s binding characteristics on the molecular level. In their study,
two SERT binding sites for citalopram were identified, providing physical evidence
for the existence of the S1 and S2 binding sites (Zhu et al. 2018).
The binding of citalopram to the allosteric site is thought to affect citalopram’s
affinity for binding to the primary S1 site (Fig. 2; Zhong et al. 2012a, b). The
structural-conformational analysis of citalopram binding to the allosteric S2 site
revealed that occupation of the allosteric site results in a spatial barrier that prevents
ligands to dissociate from the primary S1 site. The interaction with both binding sites
is thought to cause the superior antidepressant effect of S-citalopram compared to
racemic citalopram. For the racemic citalopram, R-citalopram’s binding to the
allosteric binding site of SERT is thought to antagonize S-citalopram binding to
the primary binding site SERT (Coleman et al. 2016; Sanchez 2006; Zhong et al.
2009). There are a number of studies that provide evidence for the hypothesis that R-
citalopram negatively influencing S-citalopram-SERT binding efficacy by an allo-
steric mechanism (Zhong et al. 2012a). On the molecular level, R-citalopram’s
antagonistic effect is assumed to depend on the binding to the low-affinity allosteric
S2 site which affects escitalopram binding to the high-affinity orthosteric S1 site. In
the absence of R-citalopram, escitalopram binds to the allosteric binding site and
thereby enhances its own binding to the orthosteric binding site. The latter would
result in a stronger inhibition of 5-HT uptake and enhanced 5-HT signaling, while R-
citalopram interference via S2 binding would diminish S-citalopram binding to S1
and thereby resulting in an incomplete block of the 5-HT binding pocket (Plenge et
al. 2007; Sanchez 2006; Zhong et al. 2012a). Experimental animal studies success-
fully demonstrated that R-citalopram is able to antagonize escitalopram’s antide-
pressant action using chronic mild stress paradigms (Sánchez et al. 2003), to abolish
escitalopram-enhanced 5-HT signaling (Mansari et al. 2007; Mork et al. 2003), and
to suppress acute escitalopram-mediated neuronal firing of dorsal raphe serotonergic
neurons (El Mansari et al. 2005; Mansari et al. 2007; Mnie-Filali et al. 2007). In
addition, tomography studies demonstrated R-citalopram’s ability to bind human
SERT (Klein et al. 2006; Lundberg et al. 2007), most likely also contributing to
initial 5-HT re-uptake inhibition, and that R-citalopram’s antagonizing effect is most
likely caused by chronic treatment with the racemate citalopram (Klein et al. 2007).
Antidepressants: Molecular Aspects of SSRIs 923
(Sanchez 2006; Zhong et al. 2009). Similarly, diminished SERT availability may
also result in enhanced 5-HTR1A activity and subsequent reduced 5-HT release.
In conclusion, the discrepant results observed in rodent models may be based on
species-specific differential functional regulation of rodent and human SERT.
Research approaches may overcome this obstacle by employing humanized rodent
models, for example, as in the studies by Jacobsen or Matthäus and colleagues
(Jacobsen et al. 2014; Matthäus et al. 2016), which provides the cornerstone for
pharmacological analysis of human SERT by means of experimental approaches
normally restricted to animal models. Furthermore, such experimental approaches
may be supported by complementary murine or human stem cell-derived serotonergic
neurons that contribute to understand citalopram’s mode of action on a cellular level.
voltammetry technic, it has been shown that acute citalopram treatment resulted in
frequency-dependent increases of evoked 5-HT release that required SERT expression
(Dankoski et al. 2016). Therefore, if protein kinase signaling is involved in citalopram-
SERT interaction by regulating SERT trafficking to and from the cell surface, one also
has to consider SERT expression levels, which represent the very beginning of SERT
availability. Indeed, a study using transgenic mice that feature either a heterozygous or
homozygous SERT deletion showed an age-dependent response to S-citalopram
treatment linked to SERT expression levels (Mitchell et al. 2016). SERT expression
changed during the age of the mice, steadily increasing from postnatal to adult stages.
However, the binding kinetics remained unchanged, indicating that SERT expressed in
postnatal and adolescent brain stages displays similar pharmacological characteristics
compared to SERT expressed in adult brain stages. Therefore, at constant S-citalopram
concentrations, the strongest antidepressant effect in wild-type and heterozygous
SERT mice was observed during early stages, up to 4 weeks after birth, while
SERT-depleted mice did not show antidepressant responses.
Summary
Citalopram’s binding kinetics to SERT depend on the interaction to the amino acid
residues establishing the orthosteric (primary) S1 binding site and the allosteric S2
binding site. The superior antidepressant efficacy of citalopram’s isomer S-
citalopram compared to the racemate itself is attributed to the lacking allosteric
modulation by the isomer R-citalopram. Upon acute exposure, citalopram and S-
citalopram both result in elevation of bioactive 5-HT concentrations. On one side,
this affects the neuronal activity of serotonergic neurons, suppressing 5-HT release
predominantly due to enhanced activation of 5-HT1A autoreceptors. On the other
side, protein kinase-dependent SERT shuttling to and from the cell surface is
shifted toward internalization of SERT molecules, enhancing SERT’s retrograde
trafficking. Both events present the first steps in citalopram-induced antidepressant
response, similar to molecular events triggered by acute exposure to other SSRI.
Especially, the interference with 5-HT neuronal activity is considered to add to the
already dysfunctional 5-HT signaling under disease conditions and provide the
cornerstone for the hypothesis of autoreceptor-dependent delay of clinical efficacy
shown by established antidepressant pharmacotherapy. Up to date, many interac-
tions driving 5-HT autoregulatory feedback loops maintained by 5-HT auto-
receptor and SERT activity are not well understood, particularly under complex
pathological conditions. In order to identify new and more efficient antidepres-
sants, a more detailed insight into SSRI’s potential to interfere with physiological
molecular and cellular processes is fundamental. Major depressive disorders were
too long to be considered as synaptopathies, addressing how 5-HT dysfunction
affects synaptic plasticity, while neglecting the putative role of 5-HT signaling, and
SSRI-dependent modulation of 5-HT signaling, in intrinsic, cellular, and circuit
changes.
928 A. Etievant et al.
Cross-References
References
Adkins EM, Barker EL, Blakely RD. Interactions of tryptamine derivatives with serotonin trans-
porter species variants implicate transmembrane domain I in substrate recognition. Mol
Pharmacol. 2001;59(3):514–23.
Aghajanian GK, Graham AW, Sheard MH. Serotonin-containing neurons in brain: depression of
firing by monoamine oxidase inhibitors. Science. 1970;169(3950):1100–2.
Akil H, Gordon J, Hen R, Javitch J, Mayberg H, McEwen B, Meaney MJ, Nestler EJ. Treatment
resistant depression: a multi-scale, systems biology approach. Neurosci Biobehav Rev.
2018;84:272–88.
Andersen J, Taboureau O, Hansen KB, Olsen L, Egebjerg J, Stromgaard K, Kristensen AS.
Location of the antidepressant binding site in the serotonin transporter: importance of Ser-438
in recognition of citalopram and tricyclic antidepressants. J Biol Chem. 2009;284
(15):10276–84.
Andersen J, Olsen L, Hansen KB, Taboureau O, Jorgensen FS, Jorgensen AM, Bang-Andersen B,
Egebjerg J, Stromgaard K, Kristensen AS. Mutational mapping and modeling of the binding site
for (S)-citalopram in the human serotonin transporter. J Biol Chem. 2010;285(3):2051–63.
Barker EL, Perlman MA, Adkins EM, Houlihan WJ, Pristupa ZB, Niznik HB, Blakely RD. High
affinity recognition of serotonin transporter antagonists defined by species-scanning mutagen-
esis. An aromatic residue in transmembrane domain I dictates species-selective recognition of
citalopram and mazindol. J Biol Chem. 1998;273(31):19459–68.
Barker EL, Moore KR, Rakhshan F, Blakely RD. Transmembrane domain I contributes to the
permeation pathway for serotonin and ions in the serotonin transporter. J Neurosci. 1999;19
(12):4705–17.
Belmer A, Quentin E, Diaz SL, Guiard BP, Fernandez SP, Doly S, Banas SM, Pitychoutis PM,
Moutkine I, Muzerelle A, Tchenio A, Roumier A, Mameli M, Maroteaux L. Positive regulation
of raphe serotonin neurons by serotonin 2B receptors. Neuropsychopharmacology. 2018;43
(7):1623–32.
Benmansour S, Owens WA, Cecchi M, Morilak DA, Frazer A. Serotonin clearance in vivo is altered
to a greater extent by antidepressant-induced downregulation of the serotonin transporter than
by acute blockade of this transporter. J Neurosci. 2002;22(15):6766–72.
Berton O, Nestler EJ. New approaches to antidepressant drug discovery: beyond monoamines. Nat
Rev Neurosci. 2006;7(2):137–51.
Bigler AJ, Bøgesø KP, Toft A, Hansen V. Quantitative structure–activity relationships in a series of
selective 5-HT uptake inhibitors. Eur J Med Chem. 1977;12:289–95.
Blakely RD, Bauman AL. Biogenic amine transporters: regulation in flux. Curr Opin Neurobiol.
2000;10:328–36.
Blier P, de Montigny C. Current advances and trends in the treatment of depression. Trends
Pharmacol Sci. 1994;15(7):220–6.
Blier P, Piñeyro G, el Mansari M, Bergeron R, de Montigny C. Role of somatodendritic 5-HT
autoreceptors in modulating 5-HT neurotransmission. Ann N Y Acad Sci. 1998;861:204–16.
Antidepressants: Molecular Aspects of SSRIs 929
Bogeso K, Sánchez C. The discovery of citalopram and its refinement to escitalopram. In: Fischer J,
Ganellin CR, Rotella DP, editors. Analogue-based drug design III. Weinheim: Wiley-VCH
Verlag GmbH & Co. KGaA; 2012.
Boku S, Nakagawa S, Toda H, Hishimoto A. Neural basis of major depressive disorder: beyond
monoamine hypothesis. Psychiatry Clin Neurosci. 2018;72(1):3–12.
Canli T, Lesch KP. Long story short: the serotonin transporter in emotion regulation and social
cognition. Nat Neurosci. 2007;10(9):1103–9.
Carneiro AM, Blakely RD. Serotonin-, protein kinase C-, and Hic-5-associated redistribution of the
platelet serotonin transporter. J Biol Chem. 2006;281(34):24769–80.
Cathala A, Devroye C, Drutel G, Revest JM, Artigas F, Spampinato U. Serotonin(2B) receptors in
the rat dorsal raphe nucleus exert a GABA-mediated tonic inhibitory control on serotonin
neurons. Exp Neurol. 2019;311:57–66.
Celada P, Bortolozzi A, Artigas F. Serotonin 5-HT1A receptors as targets for agents to treat
psychiatric disorders: rationale and current status of research. CNS Drugs. 2013;27(9):703–16.
Celik L, Sinning S, Severinsen K, Hansen CG, Moller MS, Bols M, Wiborg O, Schiott B. Binding
of serotonin to the human serotonin transporter. Molecular modeling and experimental valida-
tion. J Am Chem Soc. 2008;130(12):3853–65.
Chen NH, Reith ME, Quick MW. Synaptic uptake and beyond: the sodium- and chloride-dependent
neurotransmitter transporter family SLC6. Pflugers Arch. 2004;447(5):519–31.
Chen F, Larsen MB, Sanchez C, Wiborg O. The S-enantiomer of R,S-citalopram, increases inhibitor
binding to the human serotonin transporter by an allosteric mechanism. Comparison with other
serotonin transporter inhibitors. Eur Neuropsychopharmacol. 2005;15:193–8.
Cipriani A, Furukawa TA, Salanti G, Geddes JR, Higgins JP, Churchill R, Watanabe N, Nakagawa
A, Omori IM, McGuire H, Tansella M, Barbui C. Comparative efficacy and acceptability of 12
new-generation antidepressants: a multiple-treatments meta-analysis. Lancet. 2009;373:746–58.
Coleman JA, Green EM, Gouaux E. X-ray structures and mechanism of the human serotonin
transporter. Nature. 2016;532(7599):334–9.
Courtney NA, Ford CP. Mechanisms of 5-HT1A receptor-mediated transmission in dorsal raphe
serotonin neurons. J Physiol. 2016;594(4):953–65.
Coyle JT, Duman RS. Finding the intracellular signaling pathways affected by mood disorder
treatments. Neuron. 2003;38(2):157–60.
Dankoski EC, Carroll S, Wightman RM. Acute selective serotonin reuptake inhibitors regulate the
dorsal raphe nucleus causing amplification of terminal serotonin release. J Neurochem.
2016;136(6):1131–41.
Dell’Osso L, Carmassi C, Mucci F, Marazziti D. Depression, serotonin and tryptophan. Curr Pharm
Des. 2016;22(8):949–54.
Duric V, Duman RS. Depression and treatment response: dynamic interplay of signaling pathways
and altered neural processes. Cell Mol Life Sci. 2013;70(1):39–53.
El Mansari M, Sánchez C, Chouvet G, Renaud B, Haddjeri N. Effects of acute and long-term
administration of escitalopram and citalopram on serotonin neurotransmission: an in vivo
electrophysiological study in rat brain. Neuropsychopharmacology. 2005;30(7):1269–77.
Gabrielsen M, Ravna AW, Kristiansen K, Sylte I. Substrate binding and translocation of the
serotonin transporter studied by docking and molecular dynamics simulations. J Mol Model.
2012;18(3):1073–85.
Gibbons RD, Hur K, Brown CH, Davis JM, Mann JJ. Benefits from antidepressants: synthesis of 6-
week patient-level outcomes from double-blind placebo-controlled randomized trials of fluox-
etine and venlafaxine. Arch Gen Psychiatry. 2012;69:572–9.
Guiard BP, Mansari ME, Murphy DL, Blier P. Altered response to the selective serotonin reuptake
inhibitor escitalopram in mice heterozygous for the serotonin transporter: an electrophysiolog-
ical and neurochemical study. Int J Neuropsychopharmacol. 2012;15(3):349–61.
Haase J, Killian AM, Magnani F, Williams C. Regulation of the serotonin transporter by interacting
proteins. Biochem Soc Trans. 2001;29(Pt 6):722–8.
Haenisch B, Bonisch H. Depression and antidepressants: insights from knockout of dopamine,
serotonin or noradrenaline re-uptake transporters. Pharmacol Ther. 2011;129:352–68.
930 A. Etievant et al.
Henry LK, Field JR, Adkins EM, Parnas ML, Vaughan RA, Zou MF, Newman AH, Blakely RD.
Tyr-95 and Ile-172 in transmembrane segments 1 and 3 of human serotonin transporters interact
to establish high affinity recognition of antidepressants. J Biol Chem. 2006;281(4):2012–23.
Hornung JP. The human raphe nuclei and the serotonergic system. J Chem Neuroanat. 2003;26
(4):331–43.
Iceta R, Aramayona JJ, Mesonero JE, Alcalde AI. Regulation of the human serotonin trans-
porter mediated by long-term action of serotonin in Caco-2 cells. Acta Physiol. 2008;193
(1):57–65.
Jacobsen JP, Plenge P, Sachs BD, Pehrson AL, Cajina M, Du Y, Roberts W, Rudder ML, Dalvi P,
Robinson TJ, O’Neill SP, Khoo KS, Morillo CS, Zhang X, Caron MG. The interaction of
escitalopram and R-citalopram at the human serotonin transporter investigated in the mouse.
Psychopharmacology. 2014;231(23):4527–40.
Jayanthi LD, Samuvel DJ, Blakely RD, Ramamoorthy S. Evidence for biphasic effects of protein
kinase C on serotonin transporter function, endocytosis, and phosphorylation. Mol Pharmacol.
2005;67(6):2077–87.
Jørgensen TN, Christensen PM, Gether U. Serotonin-induced down-regulation of cell surface
serotonin transporter. Neurochem Int. 2014;73:107–12.
Kasper S, Sacher J, Klein N, Mossaheb N, Attarbaschi-Steiner T, Lanzenberger R, Spindelegger C,
Asenbaum S, Holik A, Dudczak R. Differences in the dynamics of serotonin reuptake trans-
porter occupancy may explain superior clinical efficacy of escitalopram versus citalopram. Int
Clin Psychopharmacol. 2009;24(3):119–25.
Kim HS, Park IS, Park WK. NMDA receptor antagonists enhance 5-HT2 receptor-mediated
behavior, head-twitch response, in mice. Life Sci. 1998;63:2305–11.
Kittler K, Lau T, Schloss P. Antagonists and substrates differentially regulate serotonin transporter
cell surface expression in serotonergic neurons. Eur J Pharmacol. 2010;629(1–3):63–7.
Klein N, Sacher J, Geiss-Granadia T, Attarbaschi T, Mossaheb N, Lanzenberger R, Potzi C, Holik
A, Spindelegger C, Asenbaum S, Dudczak R, Tauscher J, Kasper S. In vivo imaging of
serotonin transporter occupancy by means of SPECT and [123I]ADAM in healthy subjects
administered different doses of escitalopram or citalopram. Psychopharmacology.
2006;188:263–72.
Klein N, Sacher J, Geiss-Granadia T, Mossaheb N, Attarbaschi T, Lanzenberger R, Spindelegger C,
Holik A, Asenbaum S, Dudczak R, Tauscher J, Kapser S. Higher serotonin transporter occu-
pancy after multiple dose administration of escitalopram compared to citalopram: an [123I]
ADAM SPECT study. Psychopharmacology. 2007;191:333–9.
Koldso H, Severinsen K, Tran TT, Celik L, Jensen HH, Wiborg O, Schiott B, Sinning S. The two
enantiomers of citalopram bind to the human serotonin transporter in reversed orientations. J
Am Chem Soc. 2010;132(4):1311–22.
Kraus C, Castrén E, Kasper S, Lanzenberger R. Serotonin and neuroplasticity – links between
molecular, functional and structural pathophysiology in depression. Neurosci Biobehav Rev.
2017;77:317–6.
Krout D, Rodriquez M, Brose SA, Golovko MY, Henry LK, Thompson BJ. Inhibition of the
serotonin transporter is altered by metabolites of selective serotonin and norepinephrine reup-
take inhibitors and represents a caution to acute or chronic treatment paradigms. ACS Chem
Neurosci. 2017;8(5):1011–8.
Lau T, Horschitz S, Bartsch D, Schloss P. Monitoring mouse serotonin transporter internalization in
stem cell-derived serotonergic neurons by confocal laser scanning microscopy. Neurochem Int.
2009;54(3–4):271–6.
Lundberg J, Christophersen JS, Petersen KB, Loft H, Halldin C, Farde L. PET measurement of
serotonin transporter occupancy: a comparison of escitalopram and citalopram. Int J Neuropsy-
chopharmacol. 2007;10:777–85.
Mansari ME, Wiborg O, Mnie-Filali O, Benturquia N, Sánchez C, Haddjeri N. Allosteric modula-
tion of the effect of escitalopram, paroxetine and fluoxetine: in-vitro and in-vivo studies. Int J
Neuropsychopharmacol. 2007;10(1):31–40.
Mathers CD, Loncar D. Projections of global mortality and burden of disease from 2002 to 2030.
PLoS Med. 2006;3(11):e442.
Antidepressants: Molecular Aspects of SSRIs 931
Matthäus F, Haddjeri N, Sánchez C, Martí Y, Bahri S, Rovera R, Schloss P, Lau T. The allosteric
citalopram binding site differentially interferes with neuronal firing rate and SERT trafficking in
serotonergic neurons. Eur Neuropsychopharmacol. 2016;26(11):1806–17.
Mitchell NC, Gould GG, Koek W, Daws LC. Ontogeny of SERT expression and antidepressant-like
response to escitalopram in wild-type and SERT mutant mice. J Pharmacol Exp Ther. 2016;358
(2):271–81.
Mnie-Filali O, Faure C, Mansari ME, Lambas-Senas L, Berod A, Zimmer L, Sanchez C, Haddjeri
N. R-citalopram prevents the neuronal adaptive changes induced by escitalopram. Neuroreport.
2007;18:1553–6.
Mnie-Filali O, Lau T, Matthaeus F, Abrial E, Delcourte S, El Mansari M, Pershon A, Schloss P,
Sánchez C, Haddjeri N. Protein kinases alter the allosteric modulation of the serotonin trans-
porter in vivo and in vitro. CNS Neurosci Ther. 2016;22(8):691–9.
Montgomery S, Hansen T, Kasper S. Efficacy of escitalopram compared to citalopram: a meta-
analysis. Int J Neuropsychopharmacol. 2011;14(2):261–8.
Mork A, Kreilgaard M, Sanchez C. The R-enantiomer of citalopram counteracts escitalopram-
induced increase in extracellular 5-HT in the frontal cortex of freely moving rats. Neurophar-
macology. 2003;45:167–73.
Morrissette DA, Stahl SM. Modulating the serotonin system in the treatment of major depressive
disorder. CNS Spectr. 2014;19(Suppl 1):57–67; quiz 54–7, 68
Murphy DL, Lesch KP. Targeting the murine serotonin transporter: insights into human neurobi-
ology. Nat Rev Neurosci. 2008;9(2):85–96.
Murray KE, Ressler KJ, Owens MJ. In vivo investigation of escitalopram’s allosteric site on the
serotonin transporter. Pharmacol Biochem Behav. 2016;141:50–7.
Nemeroff CB. Psychopharmacology of affective disorders in the 21st century. Biol Psychiatry.
1998;44(7):517–25.
Nestler EJ, Barrot M, DiLeone RJ, Eisch AJ, Gold SJ, Monteggia LM. Neurobiology of depression.
Neuron. 2002;34(1):13–25.
Oz M, Libby T, Kivell B, Jaligam V, Ramamoorthy S, Shippenberg TS. Real-time, spatially
resolved analysis of serotonin transporter activity and regulation using the fluorescent substrate,
ASP+. J Neurochem. 2010;114(4):1019–29.
Penmatsa A, Wang KH, Gouaux E. X-ray structure of dopamine transporter elucidates antidepres-
sant mechanism. Nature. 2013;503(7474):85–90.
Piñeyro G, Blier P. Autoregulation of serotonin neurons: role in antidepressant drug action.
Pharmacol Rev. 1999;51(3):533–91.
Plenge P, Gether U, Rasmussen SG. Allosteric effects of R- and S-citalopram on the human 5-HT
transporter: evidence for distinct high- and low-affinity binding sites. Eur J Pharmacol.
2007;567:1–9.
Prasad HC, Zhu CB, McCauley JL, Samuvel DJ, Ramamoorthy S, Shelton RC, Hewlett WA,
Sutcliffe JS, Blakely RD. Human serotonin transporter variants display altered sensitivity to
protein kinase G and p38 mitogen-activated protein kinase. Proc Natl Acad Sci U S A. 2005;102
(32):11545–50.
Puig MV, Gulledge AT. Serotonin and prefrontal cortex function: neurons, networks, and circuits.
Mol Neurobiol. 2011;44(3):449–64.
Qian Y, Galli A, Ramamoorthy S, Risso S, DeFelice LJ, Blakely RD. Protein kinase C activation
regulates human serotonin transporters in HEK-293 cells via altered cell surface expression. J
Neurosci. 1997;17(1):45–57.
Quentin E, Belmer A, Maroteaux L. Somato-dendritic regulation of raphe serotonin neurons; a key
to antidepressant action. Front Neurosci. 2018;12:982.
Ramamoorthy S, Giovanetti E, Qian Y, Blakely RD. Phosphorylation and regulation of antidepres-
sant-sensitive serotonin transporters. J Biol Chem. 1998;273(4):2458–66.
Richardson-Jones JW, Craige CP, Guiard BP, Stephen A, Metzger KL, Kung HF, Gardier AM,
Dranovsky A, David DJ, Beck SG, Hen R, Leonardo ED. 5-HT1A autoreceptor levels deter-
mine vulnerability to stress and response to antidepressants. Neuron. 2010;65(1):40–52.
Robson MJ, Quinlan MA, Blakely RD. Immune system activation and depression: roles of
serotonin in the central nervous system and periphery. ACS Chem Neurosci. 2017;8(5):932–42.
932 A. Etievant et al.
Felix Tretter
Contents
Quest for Theoretical Modeling in Neuropsychiatry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 934
Empiricism, “Methodism,” and “Dataism”: How to Understand “Complexity”? . . . . . . . . . . 934
“Theoretical Neurobiology”: Ways to the Big Picture? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 935
The Brain as a Systemic Symptom Generator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 936
Systems Pathology of the Brain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 936
Systems Neurobiology and Addiction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 937
The Basic Whole-Brain Macro-circuit Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 938
The View of Neurochemistry of the Circuitry: The Cellular and Subcellular Level . . . . . . . 939
Clinical Foundations of a Systems Pathology of Addiction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 940
Symptoms of Alcohol Withdrawal: Clinical Aspects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 940
Quantification of Withdrawal Course: Assessment Scales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 941
Treatment of Alcohol Withdrawal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 942
Neurobiological Theory and Modeling of Addiction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 943
Alcohol Effects on the Main Neurotransmitter Systems (NTMS) . . . . . . . . . . . . . . . . . . . . . . . . . . 944
The Basic Neurochemical Systems Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 945
The “Neurochemical Matrix”: A Heuristic Structure Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 945
The “Neurochemical Mobile”: A Metaphorical Process Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 947
Alcohol and the Neurochemical Mobile . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 948
Conclusion and Perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 950
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 952
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 952
Abstract
This chapter initially criticizes the lack of a “theoretical” (neuro)psychiatry that
helps to understand mental disorders and their pharmaceutical treatment. In the
chapter, the complex pathophysiology of alcohol withdrawal is used as an
example to demonstrate the utility of systemic modeling of the brain: pharma-
ceutical treatment of alcohol withdrawal syndrome is usually based on
F. Tretter (*)
Bertalanffy Center for the Study of Systems Science, Vienna, Austria
e-mail: felix.tretter@bcsss.org; felix.tretter@yahoo.com
The “explanation” of mental disorders by neurobiological data has some basic short-
comings: these are the explanatory brain-mind gap, the choice of the appropriate level
of neural organization, knowledge integration of animal data and human data, under-
standing circular causality, conceptually coping with emergence, etc. (Craver 2006;
Kotchoubey et al. 2016). Moreover, neuropsychiatry has no elaborated field of
“theoretical psychiatry” (comp. Wang and Krystal 2014, Jakovljevic and Jakovljevic
2019). Instead of theoretical efforts, neuropsychiatry promotes further detailed exper-
imental research on neurobiological correlates.
In contrast, it is suggested here that a co-evolution of empirical research and
theoretical research should be practiced, as it can be seen in the history of physics,
which shows a balanced interplay between these two fields. The basic idea proposed
here is that a principle-oriented interpretation of different observations and data (e.g.,
self-organized re-balancing) could provide a heuristic framework that can “explain”
hypothetically and even “forecast” psychopharmaceutical strategies.
The prevailing roadmap for biological psychiatry has been a top-down and structure-
oriented experimental approach aiming to identify the genes responsible for the
occurrence of mental disorders. After the “systemic turn” in molecular biology
around 2000 (Kitano 2002), management of data complexity by computational
Neurochemical Mobile: A Heuristic Tool for Understanding Dynamic. . . 935
science was prioritized, and the idea of reconstructing “the whole” (brain) from the
bottom-up through computational tools guided research strategies (Tretter et al.
2010a).
Because the National Institute of Mental Health was aware of the difficulties to
integrate the plenty of methodologically heterogeneous data gathered by biological
psychiatry, the Research Domain Criteria (RDoC) were established (Insel et al.
2010). RDoC essentially provides a matrix of five domains of functions specified
as basic regulatory functions, positive valence functions, negative valence functions,
cognitive functions, and social functions. These domains – for addiction, mainly the
valence systems – are described through various methodologies including genetics,
molecules, cells, circuitry, physiology, behavior, and even self-reports.
In a similar way, the Human Brain Project, funded by the EU as a flagship project,
intends to gather a broad range of basic biological knowledge about the brain (HBP
2021).
Lately, promising connectomics brain research emerged (Sporns 2013; Morgan
and Lichtman 2013). In this context, also modeling strategies based on system theory
are used, such as the coupled neuronal-neurotransmitter whole-brain model
(Kringelbach and Deco 2020). However, there are several severe problems to
understand structural and functional connectomics already on the ultramicro-level
of “synapses” which are the connecting devices in the nervous system. Functional
understanding of synapses still is insufficient, even in the context of “systems
biology of the brain” (Tretter et al. 2010b). This raises the fundamental question of
how to integrate the complexity of data within an appropriate conceptual framework.
Also Harvard brain researchers Joshua Morgan and Jeff Lichtman summarize
connectomics research: “Both cellular and systems neuroscience are making steady
progress, but the critical bridge between them that understands how many neurons
organize into functional networks is still unbuilt” (Morgan and Lichtman 2017). This
is also true for the brain-mind gap (Kotchoubey et al. 2016).
The coincidence of neural activity with mental states and processes could be
explained by the high structural connectivity of the nervous system and the enor-
mous number of possible states of the brain. This is mainly due to the about
100 trillion synapses and their information transmitting sites for transmitter sub-
stances. Also each neuron has receptors for different transmitter substances, and
therefore, for example, benzodiazepines via GABA-A receptors can act in an
inhibitory way – so to say as a master molecule – on all cellular symptom generators:
inhibition of noradrenaline (norepinephrine) transmitting neurons results in a reduc-
tion of sympathetic activation, inhibition of glutamate-transmitting neurons reduces
cerebral excitation, etc. The effects on organs as final effectors vary in terms of
latency, intensity, and duration of responses, depending on the temporal pattern of
the activity of the neuron assemblies in different brain areas.
The high degree of connectivity of brain cells and cells of other organs justifies to
propose a general systems pathology that is embedded into integrative systems
medicine (Tretter 2019). This approach understands syndromes (e.g., withdrawal
syndrome) as the emergence of a causally effective network of symptoms that are
based on physiological/psychological dysfunctions. Systems pathology is rooted in
concepts of general systems theory (GST) as outlined by Ludwig von Bertalanffy
(von Bertalanffy 1968). One central theme of GST is the assumption that diseases
can be understood basically as a persistent imbalance of opposing forces that, under
normal conditions, are in a fluctuating dynamic balance (Bertalanffy: “Fließgle-
ichgewicht”): stress syndromes can be understood by the persistent dominance of
ergotropic mechanisms over trophotropic mechanisms, several endocrine dysfunc-
tions are the result of disbalances of upregulating and downregulating organ and/or
tissue functions, and immune disorders in some cases are a result of dominance of
pro-inflammatory versus anti-inflammatory mechanisms and vice versa. According
to this general theoretical principle, every disease is conceptually framed by the basic
notion of dynamics in terms of spatio-temporal on-off patterns of symptom inten-
sity and symptom generators. This principle will be applied later to justify a
neurochemical systems model.
In this context, it has to be highlighted that building systemic models in medicine
should be based on an explicit systems methodology that is “transdisciplinary”
(researchers and clinicians). It starts with a list of components which are supposed
to constitute a system and that are combined to a comprehensive network model
depicting all possible bidirectional connections. In a next step, a circuit model with
Neurochemical Mobile: A Heuristic Tool for Understanding Dynamic. . . 937
A B
ENVIRONMENT COR
MIND, + +
BEHAVIOR
ORGANISM
STRIA
_ _
+
ORGANSYSTEM D1R D2R GPe
_
ORGAN
SN/
VTA
STN
Kortex
+
TISSUE
GPi
CELL
2 in 3 in
GENES,
PROTEINES
THAL
Fig. 1 Multi-level view of systems neurobiology and selective modeling. (a) Top-down analysis to
molecular biology and bottom-up reconstruction of the whole by computational science. (b)
Qualitative central circuit model with imbalances accompanied with Parkinson syndrome or
schizophrenia. D2R-mediated indirect and D1R-mediated direct pathways that modulate the
thalamus-cortex interaction via 3 and 2 inhibitions that are exerted on the thalamus. Legend: Cor,
cortex; Stria, striatum; GPe, globus pallidus externus; SN, substantia nigra; VTA, ventral tegmental
area; STN, subthalamic nucleus; GPi, globus pallidus internus; Thal, thalamus; SN, substantia
nigra; VTA, ventral tegmental area; D1R, dopamine D1 receptors; D2R, dopamine D2 receptors; in,
inhibitory component; +, activating effect; , inhibitory effect
Frequency/
intensity
ES (Glu) DT (DA, 5HT)
SHA
(NA)
0 5 10 15
Time (days)
Fig. 2 Schematic of clinical observations and study results of time course of symptoms of severe
alcohol withdrawal syndromes (and their neurochemical “drivers”). SHA ¼ sympathetic hyperac-
tivity (NA ¼ noradrenaline, norepinephrine). ES ¼ epileptic seizures (Glu ¼ glutamate). DT ¼
delirium tremens (DA ¼ dopamine, 5-HT ¼ serotonin)
The basic macro-neuroanatomy of the whole brain shows the structure of long-
range circuits that exist between different brain areas such as cortex, striatum,
globus pallidus, hippocampus, amygdala, nucleus accumbens, brain stem/mid-
brain, and so on (compare Fig. 1b). The functions of macro-level circuits of
these components and the conditions of their local neural networks and their
synapse-based transmission systems, driven by glutamate (Glu), GABA, and
dopamine (DA), are relevant for neurology and psychiatry (Carlsson 1959).
They are called here simply as “Glu system,” etc. instead of “glutamatergic”
systems, etc. These systems were first discovered by pathology of neuroanatomy
and neurochemistry of motor system disorders such as Parkinson’s disease (PD;
Birkmayer and Hornykiewicz 1961) but also of schizophrenia (SZ; Carlsson
2006). The “standard circuit model” of this core system of the brain encompasses
the basal ganglia, consisting essentially of the striatum (Stria), globus pallidus pars
interna (GPi), globus pallidum pars externa (GPe), subthalamic nucleus (STN), and
substantia nigra (SN; see Fig. 2b). Although structurally well known (Albin et al.
1989; DeLong 1990; Sian et al. 1999; Foley and Riederer 2000), this system is still
not completely understood functionally, although even sophisticated systems ana-
lyses and computer simulations were conducted (Porenta and Riederer 1986;
Schroll and Hamker 2013; Berns and Sejnowski 2021). There is general agreement
however that – in the case of neuropathological conditions – two dynamic imbal-
ances, between the direct subsystem (accelerator) and the indirect subsystem
(brake) and in addition between the D1 and D2 receptor-based DA subsystem,
can result in distinct clinical syndromes (Fig. 2b):
Neurochemical Mobile: A Heuristic Tool for Understanding Dynamic. . . 939
– Normally, the activation of the signal flow from Cor to Stria and from there with
inhibitory effects to GPi that inhibits Thal corresponds to a so-called “direct”
system. With its double serial inhibitory function (i.e., activating function) on
Thal itself, it is activated by the dopamine system on the level of Stria via D1
receptors.
– The (long) “indirect” system starts also from Cor to Stria and from there runs over
GABA-based inhibitory connections to the GPe that inhibits the STN which
activates the GPi that inhibits Thal. Also this subsystem with triple serial inhibi-
tions on Thal causing an inhibition of Thal is modulated by DA, namely, by
inhibitory D2 receptors on the level of Stria.
– In addition a “short indirect pathway” has been identified, extending from Cor to
Stria to GPe and directly to GPi and Thal, and also a “hyperdirect pathway” from
Cor to STN and GPi was found.
– In the pathological case of a low-level dopamine function, PD occurs, where the
D1 receptors could dominate.
– In the pathological case with a high dopamine level, the D2 receptors could play a
strong role and psychotic states of SZ can emerge.
– Interestingly, in addicts, which are the focus of this chapter, a low level of D2
receptors is observed (Volkow et al. 2006).
It must be added here to this brain-global model that the various cell types of the
different involved brain areas are not considered explicitly in this so-called standard
model of basal ganglia. Furthermore, at the level of opposing receptor types of the
DA system where pharmaceuticals can act upon, the “behavioral function” is still
highly hypothetical: maybe the D1 receptors are involved in a behavioral “prepare”
function, whereas the D2 receptors enable a behavioral “select” function (Keeler
et al. 2014). For a deeper functional understanding of this system, it must be seen that
on the meso-level of local networks (e.g., Stria), the peculiarity of the nervous
system, especially of the brain, to have multiply connected elements (neurons)
implies that, on average, feedback occurs after about three or four connections
(synapses) mixed with convergences, divergences, and feedforwards: striatal
GABA output neurons, in addition to dopaminergic input, receive glutamatergic
corticofugal input and inhibitory input from local cholinergic interneurons which
already constitute a complicated system, as these components of the system also
have several interactions. Looking at the local networks in the cortex, mainly
940 F. Tretter
neurons are present that operate via excitatory Glu transmission (e.g., pyramidal
cells) and via inhibitory GABA transmission (e.g., basket cells). Through their
interaction with excitation and inhibitory returns, they could work like a
two-component oscillator, where the relative strength of inhibitory GABA deter-
mines the intensity and frequency of the oscillations as is known by simulation
studies (Buzsáki 2006; Liljenström 2010).
At the micro-level of the cells, in addition to the synapses at the cell body,
dendritic synapses as input sites and synaptic sites at axon terminals as output
structures are of particular interest.
Studies at the ultramicro-level of the intracellular molecular world are concerned
with omics studies of the genome, epigenome, transcriptome, proteome, meta-
bolome, etc. and with the analysis of various signal transduction pathways (Nestler
and Lüscher 2019).
All these areas of neurobiology provide information about the mechanisms of
addiction. However, we still lack heuristically valuable models. Here a sketch of
such a model will be presented, of course ignoring much of the data of neurobiology,
but with a high ecological validity in relation to clinical issues. Also a lot of data is
missing at constituting the model, but this was already mentioned in the section on
“theory.” As a next step and as a concrete example, some of the clinical knowledge
of alcoholism is mentioned here.
With regard to addiction as a form of alcohol use disorders (AUD), it is assumed here
that the global time course of this disorder can be divided – with overlaps – into an
acute phase, a chronic phase, a withdrawal phase, a restitutional phase, and a
relapse phase (Tretter 2000). The clinically relevant symptoms for the diagnosis
are listed according to DSM 5 with dimensional categories (APA 2021). For basic
function analytic considerations, a reduction of these features to symptoms that are
significant for the disease dynamics to neurobiological correlates seems to be
helpful.
Interestingly, the main knowledge about the pathophysiology of alcohol with-
drawal is known from “cold withdrawal” without medication in the context of prison
Neurochemical Mobile: A Heuristic Tool for Understanding Dynamic. . . 941
The sequence of these syndromes is not always in this way. In some cases an
initial phase of sympathetic hyperactivation is followed by a delirium, but very
seldom a delirium occurs without clinical precursors. In some cases withdrawal
seizures appear first, sometimes deliria follow seizures, etc. Interestingly, in terms of
biological rhythmicity, deliria very often occur in evening hours.
Several rating scales are used to quantify withdrawal syndromes, with at least daily
applications: the Clinical Institute Withdrawal Assessment for Alcohol (CIWA and
CIWA-Ar; Sullivan et al. 1989) or the Short Alcohol Withdrawal Scale (SAWS;
Gossop et al. 2002) are widely used scales for description of the course of the
severity of withdrawal symptoms and to control the efficacy of pharmaceutical
942 F. Tretter
treatment. Other scales have been designed to make predictions about the prospec-
tive time course of the withdrawal: the Prediction of Alcohol Withdrawal Severity
Scale (PAWSS; Maldonado et al. 2014) or the Luebeck Alcohol-Withdrawal Risk
Scale (LARS; Junghanns and Wetterling 2017). They use levels of blood alcohol
concentration at clinical admission, level of liver enzymes, electrolytes, blood
pressure, history of addiction and withdrawals, etc. as predictors.
Summarizing the frequently observed time courses, sympathetic arousal in most
cases is first, then a risk for probably glutamatergic seizures occurs, and a few hours
later a dopamine-based delirium can develop (Fig. 2).
The most critical phase in the treatment of alcohol dependence is the withdrawal
phase. Depending on the degree of dependence, an in-patient treatment is
recommended and also because of uncertainties of predictive power of assessment
scales in individual cases.
The strategy of “cold withdrawal” was continued in general psychiatry until
the 1980s. The rationale was that an addicted person should not be given
medications that themselves have an addictive potential, such as what benzodi-
azepines are supposed to have. On the other hand, these withdrawal procedures
were accompanied with several severe complications, mainly due to generalized
seizures and severe courses of delirium tremens, even with fatal outcome. Today,
in withdrawal management mainly the reduction of risks for severe complica-
tions such as generalized withdrawal seizures and delirium tremens is aimed. In
consequence, the benefit-risk assessment to apply psychopharmaceuticals or not
eventually led to the low-threshold indication of benzodiazepines although it is
still observed that later on some of these patients substitute alcohol with benzo-
diazepines or take both substances together. From the practical standpoint,
individuals with AUD are recommended to reduce or to discontinue their daily
alcohol dose. In addition, benzodiazepines are used as an effective and safe
withdrawal suppressant, but also other medications could be applied, although
satisfying randomized controlled studies still are lacking to make strong
recommendations.
Benzodiazepines
These drugs act on the GABA receptor. They are effective and recommended:
“Benzodiazepines are first-line treatment because of their well-documented effec-
tiveness in reducing the signs and symptoms of withdrawal including the incidence
of seizure and delirium” (ASAM 2021, p. 7). Benzodiazepines can be antagonized
by anexate.
Clomethiazole
It operates on the GABA receptor, but there is no antidote known. Treatment with
this substance is very effective but the drug has several side effects.
Neurochemical Mobile: A Heuristic Tool for Understanding Dynamic. . . 943
Clonidine
Clonidine acts as an alpha-2 adrenergic receptor agonist that enhances the inhibitory
action of the presynaptic receptor: as a result, the release of noradrenaline (norepi-
nephrine) is reduced and the sympathetic symptoms are attenuated.
Physostigmine
The utility of physostigmine as an acetylcholine agonist in treatment of delirium
tremens in intensive care units is reappearing (Rasimas et al. 2018).
Phenobarbital
This substance acts agonistically on the GABA receptors and is widely used as an
alternative to benzodiazepines in anglophonic countries. In these countries, they are
also used as an adjunct to benzodiazepines to control resistant alcohol withdrawal
syndrome in settings with close monitoring. These drugs cannot be antagonized.
Propofol
Acting on the GABA receptor it may be used in patients in intensive care units
experiencing resistant alcohol withdrawal and who already require mechanical
ventilation (APA 2013).
Anti-convulsants
Valproate and carbamazepine supposedly act on voltage-gated sodium channels.
They are frequently used, especially for prevention of withdrawal seizures.
Antipsychotics
If the patient develops productive psychotic symptoms, haloperidol with anti-
dopaminergic effects is a widely used drug in different medical disciplines. How-
ever, risperidone and quetiapine with anti-serotonergic effects are preferred because
they comparably do not reduce the seizure threshold.
Summarizing this section, from a clinical point of view, the two main risks –
seizures and delirium – can be prevented and treated at best if benzodiazepines are
applied from the very beginning of the withdrawal. It is also useful if clonidine as an
anti-noradrenergic drug is applied early in treatment.
Focusing on the neurochemical level of the addicted brain is useful as the drug and
the medications act on the molecular sites of the neurons. But it should be minded
that only recently, George Koob, Michel Le Moal, and Nora Volkow have developed
what is currently the leading comprehensive neurobiological systems model of
addiction that also bridges some gaps that exist to psychotherapy (Koob and Le
Moal 2001; Koob and Volkow 2016). For instance, these authors designed a stage-
based concept that ranges from a brain center-oriented paradigm to a (“systemic”)
944 F. Tretter
At first, it has to be considered that alcohol as a very small molecule acts at the
molecular level of cells, namely, at sites of synapses that provide intercellular
communication. These sites are mainly receptors, which enforce or inhibit signal
transmission through ligand-gated ion channels or by other gating mechanism and
have direct effects on intracellular signal transduction. Generally saying, the most
important neurotransmitter systems (NTMSs) for both, psychiatry and neurology,
are as follows (Birkmayer and Riederer 1989; Stahl 2013): noradrenaline (NA;
norepinephrine), dopamine (DA), serotonin (5-HT), acetylcholine (Ach), glutamate
(Glu), and gamma-aminobutyric acid (GABA). These TMSs have topographical
features. For instance, in the cortex Glu and GABA are predominant, whereas in
brain stem areas NA, DA, 5-HT, and Ach exhibit cellular clusters.
Theoretical modeling requires proper summarizing of empirical observations and
data. However, regarding alcohol effects many neurobiological studies are still
contradictory, and they also use different subjects (e.g., healthy individuals, animals
of different species, in vitro tissue studies), different substrates (genes, cell tissue),
and different methods (intracerebral concentration measures, spinal fluid concentra-
tions of metabolites, etc.). For these reasons, generalizations such as those made here
are not as consistent as they should be. Here, a selection of data of six basic
transmitter systems is made:
Glutamate (Glu): Alcohol inhibits N-methyl-D-aspartate (NMDA) receptor func-
tion and reduces calcium influx (Tabakoff and Hoffman 1993; Lovinger et al. 1989;
Lovinger et al. 1990). An upregulation of NMDA receptors is found with chronic use
(Hoffman and Tabakoff 1994). Alcohol withdrawal is associated with an increase in
glutamate in the striatum, the nucleus accumbens, and the hippocampus (Tsai and
Coyle 1998).
GABA: Low doses of alcohol increase tonic inhibition that is mediated by GABA-
A receptors, particularly those containing δ-subunits and located extrasynaptically
(Wei et al. 2004). At higher concentrations, it leads to acute intoxication by poten-
tiating the gamma-aminobutyric acid (GABA) effects. Under conditions of chronic
alcohol consumption, downregulation of GABA-A receptors occurs (Davies 2003).
Noradrenaline (NA): Interestingly, cross-species findings of suppressive effects
of alcohol on NA are by no means consistent, but chronic alcohol use suppresses
noradrenergic signaling, and there is a rebound of the system during acute with-
drawal, resulting in increased NA release (Fitzgerald 2013).
Dopamine (DA): Alcohol stimulates dopamine release in the nucleus accumbens
(Weiss et al. 1995) and is suppressed in chronic alcohol use (Di Chiara 1997) and in
Neurochemical Mobile: A Heuristic Tool for Understanding Dynamic. . . 945
The design of systems models in biology can be started from a list of described
elements, compiled in an interaction matrix to make an overview about the kind and
strength of effective interrelations. In addition to this structure model, a process model
can be constructed that depicts the imbalances and the process dynamics of the
components. If enough data are available, “exploratory” and even “explanatory”
computer simulations can be realized to test the model and to put some new empirical
questions for improving the data basis of the model (Tretter 2001; Qi et al. 2014).
Theories of health and disease can be reduced for heuristic reasons to the interaction
of opposing forces or by adaptive counter-actions: growth and apoptosis, activation
and inhibition of functions, etc. A balance concept also was successfully applied at
the level of transmitter systems (Birkmayer et al. 1972). This principle was already
used in addiction to explain drug effects and complementary withdrawal symptoms
within one conceptual framework proposed by Himmelsbach (Himmelsbach 1941).
Similarly, the allostasis concept by Solomon and Corbit has been a fruitful heuristic
(Solomon and Corbit 1974). This idea of a general opponent principle is extended in
this sketch of a neurochemical systems model of alcohol addiction.
In this view, summarizing roughly the effects of acute and chronic alcohol
consumption on the NTMSs mentioned in the section above, in a first step of
modeling, an effect table of alcohol on NTMSs can be constructed that shows
opposite effects of acute versus chronic alcohol application (Table 1). This table
Table 1 Effects of acute and chronic alcohol consumption. 5-HT ¼ serotonin, DA ¼ dopamine,
NA ¼ noradrenaline, ACh ¼ acetylcholine, Glu ¼ glutamate, GABA ¼ gamma-aminobutyric acid
5-HT DA NA ACh Glu GABA
Acute alcohol consumption " / (#) " # "/0? # "
Chronic alcohol consumption and withdrawal #? # " # " #
946 F. Tretter
needs further confirmation. However, it could be already extended with regard to the
opioid system, the cortisol system, etc. if it is important for further explanations.
Also brain-local peculiarities could be depicted, and of course the local/regional
presence of receptor subtypes is essential.
In the framework of the basic six-component model, it is assumed that the GABA
system and the DA system are enhanced during acute alcohol consumption, whereas
Glu and NA are inhibited. By chronic alcohol consumption, the GABA system and
DA are downregulated and Glu and NA are upregulated. In the situation of with-
drawal, this constellation characterizes the clinical syndrome: hyperexcitation, gen-
eralized withdrawal seizures, and delirious states can occur.
The NTMSs are highly connected neural networks in the brain stem. To under-
stand this dynamics, a heuristic network model can be derived from the tabular
description of the NTMSs and transformed into a network model for exploratory
purposes. Such a model that represents only six relevant transmission systems
already exerts a remarkable dynamics which indicates that more complex multi-
level models with hundreds or thousands of differential equations might not have
much more explanatory value (Fig. 3).
Spontaneous and circadian fluctuations of the activity of the brain interfere with
the on-off pattern of the activity of the various NTMSs during alcohol withdrawal as
it was mentioned above by the time course of significant symptoms (Fig. 1). Also
circadian oscillations are of importance with the dominance of the noradrenaline
system in the morning and the dominance of the acetylcholine system in the
afternoon and evening/night. This is probably relevant as heavy dreaming (REM
phases) is a predictor for delirious states. In consequence, the time structure of the
effects of the various NTMSs is of importance for an appropriate medication.
However, there is not enough known to draw a strong evidence-based picture of
NA ACh
ALCOHOL
DA Glu
ALCOHOL
5-HT GABA
Fig. 3 Hypothetical neurochemical network model. Hypothetical interactions and observed main
effects of acute alcohol consumption on DA, NA, GABA, and Glu. Legend: Lines with tran-
soms ¼ inhibitory action; for abbreviations see Table 1
Neurochemical Mobile: A Heuristic Tool for Understanding Dynamic. . . 947
the situation. For this reason, only a preliminary schematic for the dynamics of
glutamate and GABA can be suggested. It could serve as a theoretical guide for
further empirical research. In spite of these flaws, the network model was explored
validly in our computer simulation of alcoholism (62; see Fig. 4).
The model proposed here is a heuristic and can be used for further developments of
hypotheses. The construction of the mobile relies on the view that each component
of the neurochemical mobile represents the “relative functional weight” of a neuro-
transmitter, a term which summarily captures its physiological or behavioral impor-
tance and will be discussed later. The functional weight determines how the
respective neurotransmitter system is balanced during health or is off balance if it
is perturbed by substance abuse and/or by addiction.
The mobile can be differentiated according to other subclasses of receptors
(Fig. 3b): alpha 1 (a1) and alpha 2 (a2) receptors in the noradrenaline system, D1
and D2 receptors of the dopamine system, 5-HT1 and 5-HT2 receptors in the
serotonin system, nicotinergic (nACh) and muscarinergic (mACh) receptors in the
acetylcholine system, GABA-A and GABA-B receptors of the GABA system,
NMDA or AMPA (α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid) recep-
tors, and metabotropic glutamate receptors (mGlu). By this diversification already
12 subsystems are depicted in the model and the formalized and numerical calcula-
tion becomes difficult. For such reasons of complexity reduction, a qualitative
(semiquantitative) modeling strategy with only six components is a heuristically
useful first step (Fig. 5).
948 F. Tretter
NORADRENALINE ACETYLCHOLINE
NA ACh
+ _ + _
a1 a2 nACh mACh
Fig. 5 The neurochemical mobile as a system of scales. (a) Basic model. (b) With paired subtypes
of receptors that have opposing effects. Abbreviations: see Table 1 and text
When acute alcohol is consumed, activation of the dopamine system and the
GABA system and inhibition of the noradrenaline system and the glutamate
system occur (Fig. 6). If the person consumes alcohol chronically, these
Neurochemical Mobile: A Heuristic Tool for Understanding Dynamic. . . 949
Acute
A alcohol
NORADRENALINE ACETYLCHOLINE
SEROTONIN
GLUTAMATE
GABA
DOPAMINE
B
Chronic
alcohol
ACETYLCHOLINE
NORADRENALINE
Fig. 6 Constellation of the neurochemical mobile in acute alcohol consumption (a) and in chronic
alcohol consumption (b). (a) The activity of several neurochemical transmitter systems is altered, for
example, GABA is enforced and glutamate is inhibited. This correlates with the pleasant feeling of
alcoholization. (b) Adaptation with recovery of balance of the scales. There is dominance of noradren-
aline, serotonin, and glutamate, whereas mainly GABA is subdominating. GABA downregulates to a
lower level of function, whereas glutamate transmission exhibits enhancement of function
950 F. Tretter
A
Withdrawal
ACETYLCHOLINE
NORADRENALINE
GABA
DOPAMINE GLUTAMATE
SEROTONIN
B Withdrawal
Anti-noradrenergic
drugs
ACETYLCHOLINE
NORADRENALINE Cholinergic
drugs ?
Glutamate
Serotonin antagonists ?
antagonists ?
GABA
DOPAMINE GLUTAMATE
SEROTONIN
Benzodiazepines
Dopmain antagonists
Fig. 7 Constellation of the neurochemical mobile in alcohol withdrawal (a) and options for
treatment (b). Substances with question marks are possible candidates for pharmaceutical treatment
952 F. Tretter
Cross-References
References
Albin RL, Young AB, Penney JB. The functional anatomy of basal ganglia disorders. Trends
Neurosci. 1989;12:366–75. https://doi.org/10.1016/0166-2236(89)90074-X.
APA, American Psychiatric Association. Diagnostic and statistical manual of mental disorders. 5th
ed. Arlington: American Psychiatric Publishing; 2013.
APA, American Psychiatric Association. Clinical practice guidelines. APA; 2021. https://www.
psychiatry.org/psychiatrists/practice/clinical-practice-guidelines
ASAM, American Society of Addiction Medicine. The Asam clinical practice guideline on alcohol.
ASAM 2021. https://www.Asam.Org/Docs/Default-Source/Quality-Science/The_Asam_Clini
cal_Practice_Guideline_On_Alcohol-1.pdf?sfvrsn¼ba255c2_2
Berns GS, Sejnowski TJ. A computational model of how the basal ganglia produce sequences. J
Cogn Neurosci. 2021;10:108–21. https://doi.org/10.1162/089892998563815.
Bertalanffy L. General system theory. New York: Braziller; 1968.
Birkmayer W, Hornykiewicz O. Der L-Dioxyphenylalanin (¼ L-Dopa)-Effekt bei der Parkinson-
Akinese. Wien Klien Wschr. 1961;73:787–78.
Birkmayer W, Riederer P, editors. Understanding the neurotransmitters: key to the workings of the
brain. Vienna, New York: Springer; 1989.
Birkmayer W, Danielczyk W, Neumayer E, et al. The balance of biogenic amines as condition for
normal behaviour. J Neural Transm. 1972;33:163–78. https://doi.org/10.1007/BF01260902.
Bonhoeffer K. Die akuten Geisteskrankheiten der Gewohnheitstrinker. Eine klinische Studie. Jena:
Fischer Verlag; 1901.
Buzsáki G. Rhythms of the brain. Oxford: Oxford University Press; 2006.
Carlsson A. The occurrence, distribution and physiological role of catecholamines in the nervous
system. Pharmacol Rev. 1959;11(2, Part 2):490–3. [PMID: 13667431]
Carlsson A. The neurochemical circuitry of schizophrenia. Pharmacopsychiatry. 2006;39(S1):10–4.
Craver C. Explaining the brain. Oxford: Oxford University Press; 2006.
Davies M. The role of GABAA receptors in mediating the effects of alcohol in the central nervous
system. J Psychiatry Neurosci. 2003;28:263–74.
Dayan P, Abbott L. Theoretical neuroscience. Computational and mathematical modeling of neural
systems. Cambridge: MIT Press; 2001.
DeLong MR. Primate models of movement disorders of basal ganglia origin. Trends Neurosci.
1990;13:281–5. https://doi.org/10.1016/0166-2236(90)90110-V.
Di Chiara G. Alcohol and dopamine. Alcohol Health Res World. 1997;21(2):108–14.
Fitzgerald PJ. Elevated norepinephrine may be a unifying etiological factor in the abuse of a broad
range of substances: alcohol, nicotine, marijuana, heroin, cocaine, and caffeine. Substance
Abuse Res Treat. 2013;7:171–83. https://doi.org/10.4137/SART.S13019.
Foley P, Riederer P. The motor circuit of the human basal ganglia reconsidered. In: Mizuno Y, Calne
DB, Horowski R, Poewe W, Riederer P, Youdim MBH, editors. Advances in research on
neurodegeneration. Vienna: Springer; 2000. https://doi.org/10.1007/978-3-7091-6284-2_8.
Gossop M, Keaney F, Stewart D, Marshall EJ, Strang J. A short alcohol withdrawal scale (SAWS):
development and psychometric properties. Addict Biol. 2002;7(1):37–43. https://doi.org/10.
1080/135562101200100571. PMID: 11900621
HBP, Human Brian Project. The human Brian project. 2021. https://www.humanbrainproject.eu/en/
about/project-structure/work-packages/work-package-1/
Neurochemical Mobile: A Heuristic Tool for Understanding Dynamic. . . 953
Himmelsbach CK. The morphine abstinence syndrome, its nature and treatment. Ann Intern Med.
1941;15:829–43.
Hoffman PL, Tabakoff B. The role of the NMDA receptor in ethanol withdrawal. EXS. 1994;71:
61–70. https://doi.org/10.1007/978-3-0348-7330-7_7. PMID: 8032173
Insel T, Cuthbert B, Garvey M, Heinssen R, Pine DS, Quinn K, Sanislow C, Wang P. Research
domain criteria (RDoC): toward a new classification framework for research on mental disor-
ders. Am J Psychiatry. 2010;167(7):748–51.
Jakovljevic M, Jakovljevic I. Theoretical psychiatry as a link between academic and clinical psychi-
atry. In: Kim YK, editor. Frontiers in psychiatry. Advances in experimental medicine and biology,
vol. 1192. Singapore: Springer; 2019. https://doi.org/10.1007/978-981-32-9721-0_19.
Junghanns K, Wetterling T. Der komplizierte Alkoholentzug: Grand-Mal-Anfälle, Delir und
Wernicke-Enzephalopathie [Alcohol withdrawal and its major complications]. Fortschr Neurol
Psychiatr. 2017;85(3):163–77. https://doi.org/10.1055/s-0043-103052. Epub 2017 Mar 20.
German. PMID: 28320026
Keeler JF, Pretsell DO, Robbins TW. Functional implications of dopamine D1 vs. D2 receptors: a
‘prepare and select’ model of the striatal direct vs. indirect pathways. Neuroscience. 2014;282:
156–75.
Kitano H. Systems biology: a brief overview. Science. 2002;295(5560):1662–4. https://doi.org/10.
1126/science.1069492. PMID: 11872829
Klein JT. Interdisciplinarity and transdisciplinarity: keyword meanings for collaboration science
and translational medicine. J Transl Med Epidemiol. 2014;2(2):1024.
Koob GF, Le Moal M. Drug addiction, dysregulation of reward, and allostasis. Neuropsychophar-
macology. 2001;24(2):97–129. https://doi.org/10.1016/S0893-133X(00)00195-0. PMID:
11120394
Koob GF, Volkow ND. Neurobiology of addiction: a neurocircuitry analysis. Lancet Psychiatry.
2016;3(8):760–73. https://doi.org/10.1016/S2215-0366(16)00104-8.
Kotchoubey B, Tretter F, Braun HA, Buchheim T, Draguhn A, Fuchs T, Hasler F, Hastedt H,
Hinterberger T, Northoff G, Rentschler I, Schleim S, Sellmaier S, Tebartz van Elst L, Tschacher
W. Methodological problems on the way to integrative human neuroscience. Front Integr
Neurosci. 2016;10:41.
Kringelbach ML, Deco G. Brain states and transitions: insights from computational neuroscience.
Cell Rep. 2020;32(10):108128. https://doi.org/10.1016/j.celrep.2020.108128. PMID: 32905760
Larsson A, Edström L, Svensson L, Söderpalm B, Engel JAV. Voluntary ethanol intake increases
extracellular acetylcholine levels in the ventral tegmental area in the rat. Alcohol Alcohol.
2005;40(5):349–58. https://doi.org/10.1093/alcalc/agh180.
Liljenström H. Network effects of synaptic modifications. Pharmacopsychiatry. 2010;43(Suppl. 1):
S67–81.
Lovinger DM, Zhou Q. Alcohols potentiate ion current mediated by recombinant 5-HT3RA
receptors expressed in a mammalian cell line. Neuropharmacology. 1994;33:1567–72.
Lovinger DM, White G, Weight FF. Ethanol inhibits NMDA-activated ion current in hippocampal
neurons. Science. 1989;243:1721–4.
Lovinger DM, White G, Weight FF. NMDA receptor-mediated synaptic excitation selectively
inhibited by ethanol in hippocampal slice from adult rat. J Neurosci. 1990;10:1372–9.
Maldonado JR, Sher Y, Ashouri JF, Hills-Evans K, Swendsen H, Lolak S, Miller AC. The
“prediction of alcohol withdrawal severity scale” (PAWSS): systematic literature review and
pilot study of a new scale for the prediction of complicated alcohol withdrawal syndrome.
Alcohol. 2014;48(4):375–90. https://doi.org/10.1016/j.alcohol.2014.01.004. Epub 2014 Feb 19.
PMID: 24657098
McBride WJ, Murphy JM, Yoshimoto K, Lumeng L, Li T-K. Serotonin mechanisms in alcohol
drinking behavior. Drug Dev Res. 1993;30:170–7.
Morgan JL, Lichtman JW. Why not connectomics? Nat Methods. 2013;10(6):494–500. https://doi.
org/10.1038/nmeth.2480.
Morgan JL, Lichtman JW. Digital tissue and what it may reveal about the brain. BMC Biol.
2017;15:101. https://doi.org/10.1186/s12915-017-0436-9.
954 F. Tretter
Nestler EJ, Lüscher C. The molecular basis of drug addiction: linking epigenetic to synaptic and
circuit mechanisms. Neuron. 2019;102(1):48–59. https://doi.org/10.1016/j.neuron.2019.01.
016. PMID: 30946825; PMCID: PMC6587180
Porenta G, Riederer P. Some aspects of modeling neuronal interactions in the basal ganglia. In:
Trappl R, editor. Cybernetics and systems. Vienna: Austrian Research Institute for Artificial
Intelligence; 1986. p. 351–8.
Qi Z, Tretter F, Voit EO. A heuristic model of alcohol dependence. PLoS One. 2014;9(3):e92221.
https://doi.org/10.1371/journal.pone.0092221. Published online 2014 Mar 21
Rasimas JJ, Sachdeva KK, Donovan IW. Revival of an antidote: bedside experience with physo-
stigmine. Toxicol Commun. 2018;2(1):85–101. https://doi.org/10.1080/24734306.2018.
1535538.
Schroll H, Hamker FH. Computational models of basal-ganglia pathway functions: focus on
functional neuroanatomy. Front Syst Neurosci. 2013;7:122. https://doi.org/10.3389/fnsys.
2013.00122.
Sian J, Youdim MBH, Riederer P, et al. Biochemical anatomy of the basal ganglia and associated
neural systems. In: Siegel GJ, Agranoff BW, Albers RW, et al., editors. Basic neurochemistry:
molecular, cellular and medical aspects. 6th ed. Philadelphia: Lippincott-Raven; 1999. Available
from: https://www.ncbi.nlm.nih.gov/books/NBK27905/.
Solomon RL, Corbit JD. An opponent-process theory of motivation: I. Temporal Dynam Affect
Psychol Rev. 1974;81:119–45.
Sporns O. The human connectome: origins and challenges. NeuroImage. 2013;80:53–61.
Stahl SM. Stahl’s essential psychopharmacology: neuroscientific basis and practical applications.
Cambridge: Cambridge University Press; 2013.
Sullivan JT, Sykora K, Schneiderman J, Naranjo CA, Sellers EM. Assessment of alcohol withdrawal:
the revised clinical institute withdrawal assessment for alcohol scale (CIWA-Ar). Br J Addict.
1989;84(11):1353–7. https://doi.org/10.1111/j.1360-0443.1989.tb00737.x. PMID: 2597811
Tabakoff B, Hoffman PL. Ethanol, sedative hypnotics, and glutamate receptor function in brain and
cultured cells. Behav Genet. 1993;23:231–6. https://doi.org/10.1007/BF01067428.
Tretter F. Suchtmedizin. Stuttgart: Schattauer; 2000.
Tretter F. Systemisch-kybernetische Modellansätze der Psychologie der Sucht. In: Tretter F,
Müller A, editors. Psychologie der Sucht. Göttingen: Hogrefe; 2001. p. 165–99.
Tretter F. “Systems medicine” in the view of von Bertalanffy’s “organismic biology” and systems
theory. Res Behav Sci. 2019;36:346–62.
Tretter F, Winterer G, Gebicke-Haerter PJ, Mendoza ER, editors. Systems biology in psychiatric
research: from high-throughput data to mathematical modeling. Weinheim: Wiley-Blackwell;
2010a.
Tretter F, Rujescu D, Pogarell O, Mendoza E, editors. Systems biology of the synapse. Pharmacop-
sychiatry. 2010b;43:S1–S98.
Tsai G, Coyle JT. The role of glutamatergic neurotransmission in the pathophysiology of alcohol-
ism. Annu Rev Med. 1998;49:173–84.
Volkow ND, Wang G, Begleiter H, et al. High levels of dopamine D2 receptors in unaffected
members of alcoholic families: possible protective factors. Arch Gen Psychiatry. 2006;63(9):
999–1008. https://doi.org/10.1001/archpsyc.63.9.999.
Wang X-J, Krystal JH. Computational psychiatry. Neuron. 2014;84:638–54.
Wei W, Faria LC, Mody I. Low ethanol concentrations selectively augment the tonic inhibition
mediated by δ subunit- containing GABAA receptors in hippocampal neurons. J Neurosci.
2004;24:8379–82.
Weiss F, Lorang MT, Bloom FE, Koob GF. Oral alcohol self-administration stimulates dopamine
release in the rat nucleus accumbens: genetic and motivational determinants. J Pharmacol Exp
Ther. 1995;267:250–8.
Weiss F, Parsons LH, Schulteis G, Hyytia P, Lorang MT, et al. Ethanol self-administration restores
withdrawal-associated deficiencies in accumbal dopamine and 5-hydroxytryptamine release in
dependent rats. J Neurosci. 1996;16:3474–85.
Delusion and Dopamine: Neuronal Insights
in Psychotropic Drug Therapy
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 956
Delusions and Predictions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 958
The Role of Dopamine in Delusion Formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 960
Moving Current Models Forward . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 961
Understanding Delusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 963
Delusional Infestation: Interface Between Hallucination and Delusion . . . . . . . . . . . . . . . . . . . . . . . 964
Psychotropic Drug Therapy in Delusional Disorder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 967
Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 969
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 970
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 970
Abstract
Delusions are a central topic for neuropsychiatric research given their insidious
psychotic nature. According to the Bayesian brain hypothesis (predictive coding
theory), delusions are regarded as an aberrant inference process. It is believed that
an excess of dopamine contributes to abnormal salience attribution, which is
considered to be the basis of delusional formation. This is explained in detail by
the neuromodulatory effect of dopamine on the integration of prediction errors
and aberrant prediction errors respectively, into a person’s thinking. All currently
Introduction
Delusions represent a central topic for neuropsychiatric research, given their insid-
ious psychotic nature (Corlett et al. 2010; Feeney et al. 2017). Delusions are key
symptoms of mental illness, and physicians have been trying for more than a
hundred years to understand deluded patients and offer them treatment. Nosologi-
cally, delusions are a nonspecific symptom, and delusions occur in a wide range of
mental, medical, neurological, and substance-induced disorders. For example, there
is a core feature of schizophrenia, occurring in around 70% of patients, most
commonly with concomitant hallucinations, and grouped as “positive symptoms”
(Andreasen and Olsen 1982; Fenton et al. 1997). Delusions present as a primary
form of mental disorder, when occurring as a purely delusional disorder (ICD-11
delusional disorder/DSM-5:297.1). As such, they are isolated as a circumscribed
monothematic disorder without identifiable psychological or physical triggers
(Huber et al. 2020). Secondary forms occur as a result of another medical, neuro-
logical, or psychiatric condition. The distinction between primary and secondary
forms of delusions is of great importance because it focuses the treatment on the
underlying cause whenever possible. The term “primary” does not exclude neuro-
biological backgrounds (Keshavan and Kaneko 2013), but the emphasis of treatment
will center around antipsychotic medication and specific psychotherapies, rather
than the treatment of an underlying illness.
A delusion is usually defined as a “fixed false belief out of keeping with the
patient’s cultural background,” but recently in clinical psychiatry the aspect of being
“false” is considered less important than the aspect of being “fixed” (Freudenmann
and Lepping 2009). According to Karl Jaspers (1883–1969), one of the founders of
modern psychiatry, delusions can be recognized by (i) an extraordinary conviction
and an incomparable, subjective certainty, (ii) which cannot be influenced by
experience or logical conclusions, although (iii) their content is impossible. Because
ultimately, it can be difficult to prove or disprove certain beliefs, psychiatrists
identify delusions not primarily by judging the reality or falsity of the content of
the belief, although this might seem the most obvious. According to Freudenmann
and Lepping, the “best practice in diagnosing delusions is to look at the form of
reasoning, not the content, because the third Jasperian criterion of delusions (impos-
sibility) can be a pitfall” (Freudenmann and Lepping 2009). They use the example of
a patient with schizophrenia who may believe him- or herself to be persecuted by the
Mafia (possible; it is difficult, if not impossible, to prove the contrary): Asked to
substantiate this belief, the patient refers to strikingly many black Italian cars in front
of his or her house (obviously insufficient explanation or proof). Unlike a
Delusion and Dopamine: Neuronal Insights in Psychotropic Drug Therapy 957
Karl John Friston (born 1959) represents the doctrine that one of the main tasks of
our brain is to make predictions about our environment (Friston 2010; Friston et al.
2016). According to Friston and other scientists who have worked intensively on
models of delusions in recent years, our brain represents a “prediction organ” that
designs mental life as a constant stream of predictions and error corrections (predic-
tive-coding theory) (Corlett et al. 2010; Sterzer et al. 2018). With regard to delusions,
this means in a simplified form: Cogito, ergo praedico. I think, therefore I make
predictions; if these predictions are wrong, and yet I am uncorrectably convinced of
the opposite, then I have a delusion. The French philosopher Jean-Jacques Rousseau
(1712–1778) claimed that human foresight is the beginning of all madness (Les
Confessions). The Prussian philosopher Immanuel Kant (1724–1804) also saw
much adversity in the human foresight and describes the madman as the figure
who is used to confusing imagination and object (Versuch über die Krankheiten des
Kopfes). The British clergyman and mathematician Reverend Thomas Bayes (1701–
1761) is regarded as the mastermind of the “Bayesian statistics,” named after him,
which provides the mathematical (statistical) basis of the predictive-coding theory
(Fletcher and Frith 2009; Aitchison and Lengyel 2017). It is unlikely that the three
contemporaries (Rousseau, Kant, and Bayes) ever met, but their considerations
exemplify how early philosophers discussed the subject of delusions.
More recent hypotheses about the origins of delusions assert that humans and
animals undeniably have a brain that allows them to anticipate future situations by
pattern recognition. This enables survival by maximizing rewards and minimizing
punishments. The brain accomplishes this mathematically by making predictions
and minimizing prediction errors because of its functional-hierarchical organized
brain anatomy (neuronal networks). Predictions are communicated in a top-down
fashion from higher to lower cerebral layers. When predictions encounter bottom-up
sensory information that does not match the predictions, errors are generated (pre-
diction error), which are either accommodated (ignored) or assimilated (integrated
into future predictions). Such future predictions, named priors, provide the source of
top-down predictions (Friston and Kiebel 2009).
Delusion and Dopamine: Neuronal Insights in Psychotropic Drug Therapy 959
encoding of the prediction error in frontal cortical regions under ketamine action
(Corlett et al. 2006). A functional MRI (fMRI) study with ketamine confirmed the
relationship between a disturbed encoding of a prediction error and the development
of delusions (the extent of the disturbance correlated with the tendency to form
delusions) (Corlett et al. 2007b). Importantly, these abnormal responses under
ketamine correlated with the severity of altered beliefs across all subjects. While
the primary action of ketamine is the blockade of the glutamatergic NMDA receptor,
ketamine has important direct and indirect actions on both glutamate and dopamine.
Indeed, it has been argued that the cognitive and behavioral effects of ketamine
might be attributable to stimulation of D2 receptors rather than to the blockade of
NMDA receptors (Kapur and Seeman 2001). It has been ascertained that acute
treatment with NMDA receptor antagonists increases dopamine release in the
striatum, nucleus accumbens, and prefrontal cortex of experimental animals and
enhances the firing rate of dopamine in the midbrain (Svensson et al. 1998; Sitges et
al. 2000). In human subjects, positron emission tomography (PET) imaging of 11C-
raclopride binding following ketamine administration revealed that ketamine
induces striatal dopamine release. The magnitude of this release correlated with
the intensity of the ketamine-induced psychosis (Smith et al. 1998; Vollenweider et
al. 2000). Lawrence Kegeles was the first to suggest that the psychotomimetic effects
of NMDA receptor antagonism may be due to a disruption in the glutamatergic
control of dopamine function (Kegeles et al. 2000). Ketamine-induced disruption of
fronto-striatal dopamine/glutamate function leads to characteristic psychopathology
by aberrant prediction error-based associative causal learning.
Understanding Delusions
we are not actively concentrating (e.g., during daydreaming) (Greicius et al. 2003).
The central executive network is involved in our decisions by recognizing what is
important. It can therefore take complex information into account, process it, and
react to it. The central executive network is involved in our decisions to understand
what is important in order to respond to diversity (environmental complexity)
(Talpos and Shoaib 2015). Individuals with psychosis have structural, functional,
and neurochemical changes within these networks and how they are activated. While
Karl Jaspers wrote about the phenomenology and subjective experiences of delu-
sional patients, there is now a wealth of information that allows us to see our
understanding of delusional beliefs more in the context of altered brain structures
and functions. However, it should be noted that these changes are not necessarily
specific to delusions and go beyond positive symptoms. But this is not a contradic-
tion, because the theory of predictive coding is comprehensive and provides a
unified brain theory. Prediction errors or abnormal prediction errors occur at different
hierarchical levels of the brain, with different consequences, whether as learning
disturbances, hallucinations, or delusions. One and the same neurotransmitter (e.g.,
dopamine) can have different modes of action at different hierarchical levels of the
brain, and thus cause different functional disorders. Hallucinations could also be due
to cue-independent dopamine activation and be experienced as heard, seen, or felt
(false perception), but at a lower level of the brain hierarchy as delusions (false
beliefs) (Upthegrove et al. 2016). The concepts of salience dysregulation, self-
reference (via the so-called default mode network), and the prediction error model
have made a fundamental contribution to the neurobiological understanding of the
formation of delusional beliefs (overactive dopaminergic neurotransmission, abnor-
mal prediction errors, and increased influence of cognitive predictions/top-down).
The need to understand delusions and hallucinations together is still of great psychi-
atric and neuroscientific interest. Since primary delusions are rarely recorded, the
current understanding of delusional beliefs is largely limited to persecutory beliefs,
secondary delusions (e.g., in substance abuse), and investigations within the frame-
work of schizophrenia research. Within the category of delusional disorders, delu-
sional infestation (previously known as delusional parasitosis or Ekbom syndrome) is
a unique subtype of somatic delusion (Delusional Disorder in ICD-11; DSM-5:297.1).
Individuals with delusional infestation are convinced that they are infested by small
animals (insects, mites, and bugs) or inanimate pathogens that effect the patient’s skin,
other organs, or immediate environment (Lepping et al. 2015). Usually, patients with
delusional infestation complain of itching and may have secondary skin damage
(scratching effects or other self-damage with, e.g., disinfectants). While they often
complain of multiple systemic symptoms, dermatological symptoms are typical. The
alleged pathogen can be felt (tactile hallucinations), and often seen (visual
Delusion and Dopamine: Neuronal Insights in Psychotropic Drug Therapy 965
biasing perception by selecting the expected sensation and initiating the top-down
transmission of those expectations in predictive coding (Colder 2015).
The evidence of structural abnormalities in frontal and body perception-related
brain areas in delusional infestation is consistent with current concepts of delusion
formation. Frontal dysfunction explains the impaired judgment with resulting false
beliefs (delusions), while dysfunctional processing in the dorsal striato-subcortical
loop explains false (visuo-tactile) perceptions (hallucinations). Such a distinction
between perception and belief is consistent with Max Coltheart’s two-factor theory
of monothematic delusions, postulating that the perception is disturbed (factor 1),
and then linked to an irrational explanation, due to a nonconforming probabilistic
frontal reasoning (factor 2) (Coltheart 2010; Coltheart et al. 2018). Moreover, the
structural data are compatible with the current predictive coding model of delusions,
postulating that perception is mainly colored by prior expectations (Corlett 2020).
Subsequently, our findings support the notion that the delusional belief to be infested
with pathogens is likely to be associated with disrupted prefrontal control over
somato-sensory representations. This may indicate that delusions originate from a
neuronal network failure between top-down and bottom-up processing and represent
a mismatch between expectations and experience (aberrant prediction errors).
On a neuronal level, the dopamine system has been explored as a main target in
which biological, psychological, and social stressors have their impact to induce
psychotic symptoms (McCutcheon et al. 2019). In particular, neuroimaging studies
(PET) revealed increased striatal dopamine in the synaptic cleft in people with
positive symptoms and delusions, respectively (Fusar-Poli and Meyer-Lindenberg
2012). Additional dopamine agonists trigger psychosis, of which paranoid delusions
are the most common symptom (Voce et al. 2019). It is believed that an excess of
dopamine contributes to abnormal salience attribution, which is considered to be the
basis of delusional formation. This is explained in detail by the neuromodulatory
effect of dopamine on the integration of prediction errors within the predictive-
coding theory (Bayesian brain) (Corlett et al. 2009; Sterzer et al. 2018; Corlett 2020).
It is presumed that altered hyperdopaminergic transmission leads to abnormal
weighting of perceptual prediction errors. Such abnormal weighting leads to aberrant
prediction errors, inappropriate learning, and consequently to misinterpretations
(delusions) in “higher thinking” brain structures. Chronically, overstimulated dopa-
mine receptors lose their ability to block out irrelevant signals (stimuli). Therefore,
inappropriate associations arise and delusions are formed.
Importantly, almost all currently approved antipsychotics (neuroleptics) block
dopamine receptors, traditionally by D2 dopamine receptor antagonism (Kapur and
Seeman 2001; Kapur 2003). In addition, stimulant drugs (e.g., methamphetamine),
which increase activity at mesolimbic D2 dopamine receptors, raise the risk of
psychosis and delusions (Voce et al. 2019). Furthermore, PET studies (in vivo)
have shown that for most antipsychotic drugs, a 65–70% occupancy of the D2-
968 M. K. Huber et al.
Conclusion
Delusions and dopamine belong together like Pythagoras and the right angle.
Prediction error signaling and stimulus-reward learning primarily implicate dopa-
mine as the major neurotransmitter in delusion formation. Similarly, the dopamine
hypothesis is the dominant pathophysiological model of psychosis and delusions.
This is also supported by the neurochemical actions of antipsychotics (dopamine
antagonism). Inappropriate prediction error signaling explains the formation of
delusions and provides the basis for inappropriate processing of stimuli, perception,
attention, associations, thoughts, and formation of beliefs. A chronically inappropri-
ate excess of dopamine leads to irrelevant associations and beliefs being imbued
with inappropriate significance. The brain’s attempt to integrate these inappropriate
experiences leads to inappropriate priors, the beginning, and the maintaining of
delusions. Such processes are driven by prediction error signals, i.e., discrepancies
between what an organism expects in a given environment and what actually
happens. These neuronal processes are mainly signaled by dopaminergic and
glutamatergic mechanisms. Disruption of these neurobiological systems may under-
lie delusional formation through aberrant prediction error firing. This leads to
attention being drawn to inappropriate stimuli resulting in inappropriate priors, the
source of predictions. On a practical clinical level, the long-standing clinical
970 M. K. Huber et al.
experience of the authors confirms the importance of a timely and adequate treatment
of patients with delusions with antipsychotics, accompanied by a stable and trusting
doctor-patient relationship.
Cross-References
References
Abi-Dargham A, Rodenhiser J, Printz D, Zea-Ponce Y, Gil R, Kegeles LS, Weiss R, Cooper TB,
Mann JJ, Van Heertum RL, Gorman JM, Laruelle M. Increased baseline occupancy of D2
receptors by dopamine in schizophrenia. Proc Natl Acad Sci USA. 2000;97:8104–9.
Aitchison L, Lengyel M. With or without you: predictive coding and Bayesian inference in the
brain. Curr Opin Neurobiol. 2017;46:219–27.
Andreasen NC, Olsen S. Negative v positive schizophrenia. Definition and validation. Arch Gen
Psychiatry. 1982;39:789–94.
Baker PB, Cook BL, Winokur G. Delusional infestation: the interface of delusions and hallucina-
tions. Psychiatr Clin North Am. 1995;18:345–61.
Berrios GE. Tactile hallucinations: conceptual and historical aspects. J Neurol Neurosurg Psychi-
atry. 1982;45:285–93.
Colder B. The basal ganglia select the expected sensory input used for predictive coding. Front
Comput Neurosci. 2015;23(9):119.
Coltheart M. The neuropsychology of delusions. Ann N Y Acad Sci. 2010;1191:16–26.
Coltheart M, Cox R, Sowman P, Morgan H, Barnier A, Langdon R, Connaughton E, Teichmann L,
Williams N, Polito V. Belief, delusion, hypnosis, and the right dorsolateral prefrontal cortex: a
transcranial magnetic stimulation study. Cortex. 2018;101:234–48.
Corlett PR. Predicting to perceive and learning when to learn. Trends Cogn Sci. 2020;24(4):259–60.
Corlett PR, Fletcher PC. Delusions and prediction error: clarifying the roles of behavioural and
brain responses. Cogn Neuropsychiatry. 2015;20:95–105.
Corlett PR, Honey GD, Aitken MR, Dickinson A, Shanks DR, Absalom AR, Lee M, Pomarol-
Clotet E, Murray GK, McKenna PJ, Robbins TW, Bullmore ET, Fletcher PC. Frontal responses
during learning predict vulnerability to the psychotogenic effects of ketamine: linking cognition,
brain activity, and psychosis. Arch Gen Psychiatry. 2006;63:611–21.
Corlett PR, Murray GK, Honey GD, Aitken MRF, Shanks DR, Robbins TW, Bullmore ET,
Dickinson A, Fletcher PC. Disrupted prediction-error signal in psychosis: evidence for an
associative account of delusions. Brain. 2007a;130:2387–400.
Corlett PR, Honey GD, Fletcher PC. From prediction error to psychosis: ketamine as a pharmaco-
logical model of delusions. J Psychopharmacol. 2007b;21:238–52.
Corlett PR, Krystal JH, Taylor JR, Fletcher PC. Why do delusions persist? Front Hum Neurosci.
2009;3:12.
Corlett PR, Taylor JR, Wang XJ, Fletcher PC, Krystal JH. Toward a neurobiology of delusions. Prog
Neurobiol. 2010;92(3):345–69.
Delusion and Dopamine: Neuronal Insights in Psychotropic Drug Therapy 971
Corlett PR, Cambridge V, Gardner JM, Piggot JS, Turner DC, Everitt JC, Arana FS, Morgan HL,
Milton AL, Lee JL, Aitken MR, Dickinson A, Everitt BJ, Absalom AR, Adapa R, Subramanian
N, Taylor JR, Krystal JH, Fletcher PC. Ketamine effects on memory reconsolidation favor a
learning model of delusions. PLoS One. 2013;8(6):e65088.
Di Pellegrino G, Ladavas E. Peripersonal space in the brain. Neuropsychologia. 2015;66:126–33.
Dickinson A. The 28th Bartlett memorial lecture causal learning: an associative analysis. Q J Exp
Psychol. 2001;54(1):3–25.
Dudley R, Taylor P, Wickham S, Hutton P. Psychosis, delusions and the “Jumping to conclu-
sions” reasoning Bias: a systematic review and meta-analysis. Schizophr Bull. 2016;42
(3):652–65.
Eccles JA, Garfinkel SN, Harrison NA, Ward J, Taylor RE, Bewley A, Critchley HD. Sensations of
skin infestation linked to abnormal frontolimbic brain reactivity and differences in self-repre-
sentation. Neuropsychologia. 2015;77:90–6.
Feeney EJ, Groman SM, Taylor JR, Corlett PR. Explaining delusions: reducing uncertainty through
basic and computational neuroscience. Schizophr Bull. 2017;43(2):263–72.
Fenton WS, McGlashan TH, Victor BJ, Blyler CR. Symptoms, subtype, and suicidality in patients
with schizophrenia spectrum disorders. Am J Psychiatry. 1997;154(2):199–204.
Fiorillo CD. Towards a general theory of neural computation based on prediction by single neurons.
PLoS One. 2008;3:e3298.
Fiorillo CD, Tobler PN, Schultz W. Evidence that the delay-period activity of dopamine neurons
corresponds to reward uncertainty rather than backpropagating TD errors. Behav Brain Funct.
2005;1:7.
Fletcher PC, Frith CD. Perceiving is believing: a Bayesian approach to explaining the positive
symptoms of schizophrenia. Nat Rev Neurosci. 2009;10:48–58.
Freudenmann RW, Lepping P. Delusional infestation. Clin Microbiol Rev. 2009;22(4):690–732.
Friston K. A theory of cortical responses. Philos Trans R Soc Lond B Biol Sci. 2005;360:815–36.
Friston K. The free energy principle: a unified brain theory? Nat Rev Neurosci. 2010;11(2):127–38.
Friston K, Kiebel S. Predictive coding under the free-energy principle. Phil Trans R Soc B.
2009;364:1211–21.
Friston K, Schwartenbeck P, FitzGerald T, Moutoussis M, Behrens T, Dolan RJ. The anatomy of
choice: dopamine and decision-making. Philos Trans R Soc Lond Ser B Biol Sci. 2014;5:369.
Friston K, FitzGerald T, Rigoli F, Schwartenbeck P, Pezzulo G. Active inference. A process theory.
Neural Comput. 2016;21:1–49.
Fusar-Poli P, Meyer-Lindenberg A. Striatal presynaptic dopamine in schizophrenia, Part II: meta-
analysis of [18F/11C]-DOPA PET studies. Schizophr Bull. 2012;39(1):33–42.
González-Rodríguez A, Estrada F, Monreal JA, Palao D, Labad J. A systematic review of the
operational definitions for antipsychotic response in delusional disorder. Int Clin
Psychopharmacol. 2018;33(5):261–7.
Gordon N, Tsuchiya N, Koenig-Robert R, Hohwy J. Expectation and attention increase the
integration of top-down and bottom-up signals in perception through different pathways.
PLoS Biol. 2019;17(4):e3000233.
Graziano MS, Gross CG. A bimodal map of space: somatosensory receptive fields in the macaque
putamen with corresponding visual receptive fields. Exp Brain Res. 1993;97(1):96–109.
Greicius MD, Krasnow B, Reiss AL, Menon V. Functional connectivity in the resting brain: a
network analysis of the default mode hypothesis. Proc Natl Acad Sci USA. 2003;100
(1):253–8.
Gunduz-Bruce H, McMeniman M, Robinson DG, Woerner MG, Kane JM, Schooler NR,
Lieberman JA. Duration of untreated psychosis and time to treatment response for delusions
and hallucinations. Am J Psychiatry. 2005;162(10):1966–9.
Heinz A, Murray GK, Schlagenhauf F, Sterzer P, Grace AA, Waltz JA. Towards a unifying
cognitive, neurophysiological, and computational neuroscience account of schizophrenia.
Schizophr Bull. 2019;45(5):1092–100.
Herrero JL, Roberts MJ, Delicato LS, Gieselmann MA, Dayan P, Thiele A. Acetylcholine contrib-
utes through muscarinic receptors to attentional modulation in V1. Nature. 2008;454:1110–4.
972 M. K. Huber et al.
Howes OD, Montgomery AJ, Asselin MC, Murray RM, Valli I, Tabraham P, Bramon-Bosch E,
Valmaggia L, Johns L, Broome M, McGuire PK, Grasby PM. Elevated striatal dopamine
function linked to prodromal signs of schizophrenia. Arch Gen Psychiatry. 2009;66(1):13–20.
Howes OD, Kambeitz J, Kim E, Stahl D, Slifstein M, Abi-Dargham A, Kapur S. The nature of
dopamine dysfunction in schizophrenia and what this means for treatment: meta-analysis of
imaging studies. Arch Gen Psychiatry. 2012;69(8):776–86.
Huber M, Kirchler E, Karner M, Pycha R. Delusional parasitosis and the dopamine transporter. A
new insight of etiology? Med Hypotheses. 2007;68(6):1351–8.
Huber M, Karner M, Kirchler E, Lepping P, Freudenmann RW. Striatal lesions in delusional
parasitosis revealed by magnetic resonance imaging. Prog Neuro-Psychopharmacol Biol Psy-
chiatry. 2008;32(8):1967–71.
Huber M, Lepping P, Pycha R, Karner M, Schwitzer J, Freudenmann RW. Delusional infestation:
treatment outcome with antipsychotics in 17 consecutive patients (using standardized reporting
criteria). Gen Hosp Psychiatry. 2011;33(6):604–11.
Huber M, Wolf RC, Lepping P, Kirchler E, Karner M, Sambataro F, Herrnberger B, Corlett PR,
Freudenmann RW. Regional gray matter volume and structural network strength in somatic vs.
non-somatic delusional disorders. Prog Neuro-Psychopharmacol Biol Psychiatry. 2018;82:115–22.
Huber M, Kirchler E, Wolf RC. Pharmakotherapie der wahnhaften Störung. NeuroTransmitter.
2020;31(6):46–53. (SpringerMedizin).
Kapur S. Psychosis as a state of aberrant salience: a framework linking biology, phenomenology,
and pharmacology in schizophrenia. Am J Psychiatry. 2003;160:13–23.
Kapur S, Seeman P. Ketamine has equal affinity for NMDA receptors and the high-affinity state of
the dopamine D2 receptor. Biol Psychiatry. 2001;49:954–7.
Kegeles LS, Abi-Dargham A, Zea-Ponce Y, Rodenhiser-Hill J, Mann JJ, Van Heertum RL, Cooper
TB, Carlsson A, Laruelle M. Modulation of amphetamine-induced striatal dopamine release by
ketamine in humans: implications for schizophrenia. Biol Psychiatry. 2000;48(7):627–40.
Kendler KS, Hays P. Paranoid psychosis (delusional disorders) and schizophrenia. Arch Gen
Psychiatry. 1981;38:547–51.
Keshavan MS, Kaneko Y. Secondary psychoses: an update. World Psychiatry. 2013;12(1):4–15.
Keysers C, Wicker B, Gazzola V, Anton JL, Fogassi L, Gallese V. A touching sight: SII/PV
activation during the observation and experience of touch. Neuron. 2004;42(2):335–46.
Kienast T, Heinz A. Dopamine and the diseased brain. CNS Neurol Disord Drug Targets.
2006;5:109–31.
Kokkinou M, Ashok AH, Howes OD. The effects of ketamine on dopaminergic function: meta-
analysis and review of the implications for neuropsychiatric disorders. Mol Psychiatry.
2018;23:59–69.
Ladavas E, Zeloni G, Farne A. Visual peripersonal space centered on the face in humans. Brain.
1998;121(Pt12):2317–26.
Lavin A, Nogueira L, Lapish CC, Wightman RM, Phillips PE, Seamans JK. Mesocortical dopamine
neurons operate in distinct temporal domains using multimodal signaling. J Neurosci. 2005;25
(20):5013–23.
Lepping P, Russell I, Freudenmann RW. Antipsychotic treatment of delusional parasitosis–a
systematic review. Br J Psychiatry. 2007;191:198–205.
Lepping P, Baker C, Freudenmann RW. Delusional infestation in dermatology in the UK: preva-
lence, treatment strategies, and feasibility of a RCT. Clin Exp Dermatol. 2010;35(8):841–4.
Lepping P, Huber M, Freudenmann RW. How to approach delusional infestation. BMJ. 2015;350:
h1328.
Lepping P, Aboalkaz S, Squire SB, Romanov DV, Bewley A, Huber M, Noorthoorn EO. Later age
of onset and longer duration of untreated psychosis are associated with poorer outcome in
delusional infestation. Acta Derm Venereol Acta Ven. 2020. in print.
McCutcheon RA, Abi-Dargham A, Howes OD. Schizophrenia, dopamine and the striatum: from
biology to symptoms. Trends Neurosci. 2019;42:205–20.
Delusion and Dopamine: Neuronal Insights in Psychotropic Drug Therapy 973
Sterzer P, Adams RA, Fletcher P, Frith C, Lawrie SM, Muckli L, Petrovic P, Uhlhaas P, Voss M,
Corlett PR. The predictive coding account of psychosis. Biol Psychiatry. 2018;84(9):634–43.
Svensson TH, Mathe JM, Nomikos GG, Schilstrom B. Role of excitatory amino acids in the ventral
tegmental area for central actions of non-competitive NMDA-receptor antagonists and nicotine.
Amino Acids. 1998;14:51–6.
Talpos J, Shoaib M. Executive function. Handb Exp Pharmacol. 2015;228:191–213.
Tauscher J, Hussain T, Agid O, Verhoeff NPLG, Wilson AA, Houle S, Remington G, Zipursky RB,
Kapur S. Equivalent occupancy of dopamine D1 and D2 receptors with clozapine: differentia-
tion from other atypical antipsychotics. Am J Psychiatry. 2004;161:1620–5.
Tsakiris M, Haggard P. The rubber hand illusion revisited: visuotactile integration and self-attribu-
tion. J Exp Psychol Hum Percept Perform. 2005;31:80–91.
Upthegrove R, Ives J, Broome MR, Caldwell K, Wood SJ, Oyebode F. Auditory verbal hallucina-
tions in first-episode psychosis: a phenomenological investigation. Br J Psychiatr Open. 2016;2
(1):88–95.
Voce A, Calabria B, Burns R, Castle D, McKetin R. A systematic review of the symptom profile and
course of methamphetamine-associated psychosis substance use and misuse. Subst Use Misuse.
2019;54(4):549–59.
Vollenweider FX, Vontobel P, Oye I, Hell D, Leenders KL. Effects of (S)-ketamine on striatal
dopamine: a [11C] raclopride PET study of a model psychosis in humans. J Psychiatr Res.
2000;34:35–43.
Winton-Brown TT, Fusar-Poli P, Ungless MA, Howes OD. Dopaminergic basis of salience
dysregulation in psychosis. Trends Neurosci. 2014;37:85–94.
Wolf RC, Huber M, Depping MS, Thomann PA, Karner M, Lepping P, Freudenmann RW.
Abnormal gray matter and white matter volume in delusional infestation. Prog Neuro-
Psychopharmacol Biol Psychiatry. 2013;46(1):19–24.
Hallucinogens as Therapeutic Agents: Past,
Present, and Future
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 976
History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 976
Chemistry and Mechanism of Action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 978
Pharmacokinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 980
Clinical Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 980
Indications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 982
Adverse Effects and Toxicology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 982
Combination Therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 983
Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 984
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 984
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 984
Abstract
Hallucinogens have a long history as therapeutic agents. After the synthesis of
lysergic acid diethylamide (LSD) in 1938 by Albert Hofmann, the popularity of
classical hallucinogens with psychedelic properties increased among scientists,
psychiatrists, neurologists, and psychotherapists. Research in the 1950s and
1960s showed great promise for the use of psychedelics in medical research
and treatment. Psychedelics are characterized by their mind revealing effects,
which are valuable in the treatment of major depressive disorder (MDD) and of
stress- and trauma-related mental disorders (PTSD). However, research into the
E. Vos
University of Amsterdam, Amsterdam, The Netherlands
e-mail: elena.vos@student.uva.nl
S. Snelders
Freudenthal Institute, Utrecht University, Utrecht, The Netherlands
e-mail: s.a.m.snelders@uu.nl
T. Pieters (*)
Department of Pharmaceutical Sciences, Utrecht University, Utrecht, The Netherlands
e-mail: t.pieters@uu.nl
Introduction
For thousands of years, people have been using mind-altering drugs for religious and
healing purposes. The earliest evidence of the use of classical hallucinogens or
psychedelic substances dates back 5700 years to northeastern Mexico, where the
remains of peyote cacti and red beans containing mescaline were found in inhabited
caves (Rucker et al. 2018). The term “psychedelic” is a neologism derived from the
ancient Greek words “psychē” (“mind” or “soul”) and dēlein (“to reveal” or “to
open”). The mind-opening effects of the drugs make psychedelics valuable in
psychotherapy. This therapeutic value was widely researched in the 1950s and
1960s. Psychedelics showed great promise for medical use in combination with
psychotherapy and were considered breakthrough treatments for patients with major
depressive disorder (MDD), addiction, end-of-life anxiety, chronic pain, obsessive–
compulsive disorder, and anorexia (Nutt et al. 2020). Use was advocated by many
psychiatrists and psychotherapists from diverse cultural backgrounds (Snelders and
Kaplan 2002). Although the therapeutic effects seemed promising, research became
restricted after the 1960s because of the popularity of psychedelics for recreational
use (especially among young people), which led to most classical hallucinogens
being classified as drugs with abuse potential. In recent years, research on these
substances has been reestablished by different groups around the world (Snelders
and Pieters 2020).
History
LSD available to qualified researchers in Europe and the USA under the brand
name Delysid (Geiger et al. 2018). The goal of this distribution was to gain insight
into the safety risks and potential clinical benefits of LSD (Belouin and
Henningfield 2018). The timing of this introduction turned out to be ideal when,
as part of a psychopharmacologic revolution, psychiatry changed during the
1950s. New stimulating, calming, and narcotic agents became available. LSD
and psilocybin were two such agents, fitting perfectly within this psychopharma-
cological revolution (Snelders and Pieters 2020). LSD research also played a
crucial role in the emerging field of modern neuropsychopharmacology. Scientists
noticed close similarities between the chemical structures of serotonin and LSD in
1954 (Woolley and Shaw 1954). In 1957, the influential American journal Life
published an article by Robert Gordon Wasson about the Mexican psilocybin
mushroom. One year later, Hofmann isolated the psychoactive compounds in the
mushroom: psilocybin and its most prominent active metabolite psilocin. Subse-
quently, Sandoz promoted the use of these psychedelic compounds for both basic
pharmaceutical and therapeutic clinical research (Geiger et al. 2018). The popu-
larity of psychedelics among scientists and psychiatrists increased. From the end
of the 1940s onwards, clinical studies of psychedelics were performed on patients
with psychoses, MDD, addiction, anxiety, or compulsive disorders (Rucker et al.
2018). Psychedelics share the remarkable ability of enhancing both perception of
meaning and suggestibility. Experiences of ego dissolution and spiritually mean-
ingful experiences are commonly mentioned responses, and are correlated with
the success of therapy (Breeksema et al. 2020). Between 1950 and 1965, over
1000 scientific papers on psychedelics were published and LSD alone was
prescribed as a treatment to over 40,000 patients. Unfortunately, most of these
studies were observational, uncontrolled, and with small number of participants.
When research continued, concern arose among scientists about the long-term
effects of psychedelics. The “dark side” of psychedelics (so-called “bad trips,”
psychoses, “flashbacks,” and suicides) became more visible (Snelders and Pieters
2020). In the sociopolitical context, psychedelics became associated with the
counterculture and political activism in the USA and Europe during the 1960s
and 1970s. These social developments led to psychedelics earning an unfavorable
reputation among many political and medical leaders. In 1966, hallucinogens
were placed in Schedule I of the UN Single Convention on Narcotic Drugs.
Manufacture, possession, and use of most psychedelics were regulated in national
legislations (Belouin and Henningfield 2018; Rucker et al. 2018). Thus, reduced
interest in psychedelics in medicine was related to a complex interplay of these
internal and external factors. It seemed that psychedelics, mainly LSD and
psilocybin, had followed the typical cyclic movement of psychotropic drugs:
from popularity to criticism to oblivion (Snelders et al. 2006). Still, psychoactive
substances have never completely disappeared and were always propagated by
psychedelic enthusiasts, both outside and inside of medicine (Snelders and Pieters
2020). Nowadays, there is growing anticipation, awareness, and hope for the
potential of psychedelic compounds to become medically approved as psychoac-
tive treatments (Belouin and Henningfield 2018).
978 E. Vos et al.
(van den Brink et al. (van den Brink et al. 2020b) contact (van den Brink experience (van den Brink
2020b) 2020b) et al. 2020b) et al. 2020b)
Risks Psychosis (when Psychosis, HPPD (van Addiction, bladder Hyperthermia, liver Heart rhythm disturbance,
vulnerable), hallucinogen den Brink et al. 2020b) problems, HPPD (van den damage, serotonin psychosis (when
persisting perception Brink et al. 2020b) syndrome, HPPD (van vulnerable), HPPD (van
disorder (HPPD) (van den den Brink et al. 2020b), den Brink et al. 2020b)
Brink et al. 2020b) cardiac arrhythmias
(Ionescu et al. 2019)
Adapted from van den Brink et al. (2020b)
Prim. primarily
979
980 E. Vos et al.
Pharmacokinetics
Clinical Studies
(2018)
RCT; N ¼ 26, US Patients with treatment-resistant No significant difference in N¼2 No adverse events described
Ionescu depression, 6 0.5 mg/kg ketamine depression severity between placebo
et al. (2019) or saline placebo and ketamine
RCT; N ¼ 26, US Patients with post-traumatic stress Active doses of MDMA with N¼2 Anxiety, headache, muscle
Mithoefer disorder (PTSD), 30, 75, or 125 mg adjunctive psychotherapy were tension, and fatigue
et al. (2018) MDMA effective and well tolerated in
reducing PTSD symptoms
981
982 E. Vos et al.
Indications
Combination Therapy
Conclusion
Cross-References
References
Agin-Liebes GI, Malone T, Yalch MM, Mennenga SE, Ponté KL, Guss J, Bossis AP, Grigsby J,
Fischer S, Ross S. Long-term follow-up of psilocybin-assisted psychotherapy for psychiatric
and existential distress in patients with life-threatening cancer. J Psychopharmacol. 2020;34
(2):155–66.
Belouin SJ, Henningfield JE. Psychedelics: where we are now, why we got here, what we must do.
Neuropharmacology. 2018;142:7–19.
Breeksema JJ, van den Brink W, Veraart JKE, Smith-Apeldoorn SY, Vermetten E, Schoevers RA.
Psychedelica bij de behandeling van depressie, angst en obsessieve-compulsieve stoornis.
Tijdschr Psychiatr. 2020;62(8):618–28.
Brown RT, Nicholas CR, Cozzi NV, Gassman MC, Cooper KM, Muller D, Thomas CD, Hetzel SJ,
Henriquez KM, Ribaudo AS, Hutson PR. Pharmacokinetics of escalating doses of oral psilo-
cybin in healthy adults. Clin Pharmacokinet. 2017;56(12):1543–54.
Carhart-Harris RL. How do psychedelics work? Curr Opin Psychiatry. 2019;32(1):16–21.
Carhart-Harris RL, Leech R, Williams TM, Erritzoe D, Abbasi N, Bargiotas T, Hobden P, Sharp DJ,
Evans J, Feilding A, Wise RG, Nutt DJ. Implications for psychedelic-assisted psychotherapy:
functional magnetic resonance imaging study with psilocybin. Br J Psychiatry. 2012;200
(3):238–44.
Carhart-Harris RL, Bolstridge M, Day CMJ, Rucker J, Watts R, Erritzoe DE, Kaelen M, Giribaldi
B, Bloomfield M, Pilling S, Rickard JA, Forbes B, Feilding A, Taylor D, Curran HV, Nutt DJ.
Hallucinogens as Therapeutic Agents: Past, Present, and Future 985
Smith-Apeldoorn SY, Veraart JKE, Kamphuis J, Breeksema JJ, van den Brink W, Aan Het Rot M,
Schoevers RA. Ketamine als anestheticum, analgeticum en als antidepressivum. Tijdschr
Psychiatr. 2020;62(8):629–39.
Snelders S, Kaplan C. LSD therapy in Dutch psychiatry: changing socio-political settings and
medical sets. Med Hist. 2002;46(2):221–40.
Snelders S, Pieters T. De opkomst, val en opkomst van LSD. Tijdschr Psychiatr. 2020;62(8):707–12.
Snelders S, Kaplan C, Pieters T. On cannabis, chloral hydrate, and career cycles of psychotropic
drugs in medicine. Bull Hist Med. 2006;80(1):95–114.
van Amsterdam J, van den Brink W. Recreatief ecstasygebruik in Nederland: gebruikskenmerken,
gezondheidsschade en criminaliteit. Tijdschr Psychiatr. 2020;62(8):693–701.
van den Brink W, Breeksema JJ, Vermetten E, Schoevers RA. Psychedelica bij de behandeling van
verslaving en psychose. Tijdschr Psychiatr. 2020a;62(8):650–8.
van den Brink W, Schoevers RA, Vermetten E, Van R, Breeksema JJ. Effectiviteit van psychedelica
bij de behandeling van psychiatrische aandoeningen: inleiding. Tijdschr Psychiatr. 2020b;62
(8):613–7.
van der Gronde T, Los L, Herremans A, Oosting R, Zorzanelli R, Pieters T. Toward a new model of
understanding, preventing, and treating adolescent depression focusing on exhaustion and
stress. Front Psychiatry. 2020;11:412.
van Elk M. Neurowetenschappelijke en psychologische verklaringen voor de therapeutische
effecten van psychedelica. Tijdschr Psychiatr. 2020;62(8):677–83.
Vermetten E, Yehuda R. MDMA-assisted psychotherapy for posttraumatic stress disorder: a
promising novel approach to treatment. Neuropsychopharmacology. 2020;45(1):231–2.
Woolley DW, Shaw E. A biochemical and pharmacological suggestion about certain mental
disorders. Proc Natl Acad Sci U S A. 1954;40(4):228–31.
World Health Organization. Ketamine (INN): update review report. Geneva: World Health Orga-
nization; 2015.
Zanos P, Moaddel R, Morris PJ, Riggs LM, Highland JN, Georgiou P, Pereira EFR, Albuquerque
EX, Thomas CJ, Zarate CA Jr, Gould TD. Ketamine and ketamine metabolite pharmacology:
insights into therapeutic mechanisms. Pharmacol Rev. 2018;70(3):621–60.
Lipid Metabolism Disturbances During
Antipsychotic Treatment for Schizophrenia
Contents
Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 988
Lipid Metabolism and Schizophrenia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 989
The Risk of Hyperlipidemia and Antipsychotics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 990
Antipsychotic Drugs and Lipid Disturbances . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 992
Meta-analysis of Lipid Parameters Among Antipychotics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 994
Effect of Smoking Status and Gender Differences on Lipid Parameters . . . . . . . . . . . . . . . . . . . . . . 994
Race/Ethnicity Differences in the Risk of Hyperlipidemia Among Schizophrenia . . . . . . . . . . . . 995
Hypotheses Regarding the Mechanisms for the Association Between Antipsychotics
and Lipids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 996
Guidelines and Interventions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 996
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 999
Abstract
Patients with schizophrenia experience increased morbidity and mortality and have
an approximately 20% shorter lifespan. Patients with schizophrenia also have a
relatively increased prevalence of cardiometabolic risk factors, such as obesity, type
2 diabetes, hypertension, hyperglycemia, and dyslipidemia. The odds ratio (OR) for
metabolic syndrome in patients with chronic schizophrenia compared to the general
population was 2.35. Although the individual differences of these metabolic
disturbances have been observed, an understanding of the underlying mechanisms
is still uncertain. A meta-analysis demonstrated that approximately 40% of patients
with schizophrenia and related disorders have abnormalities in lipid metabolism.
S. Ono (*)
Department of Community Psychiatric Medicine, Niigata University Graduate School of Medical
and Dental Sciences, Niigata, Japan
e-mail: onoshin@med.niigata-u.ac.jp
T. Someya
Department of Psychiatry, Niigata University Graduates School of Medicine and Dental Science,
Niigata, Japan
e-mail: psy@med.niigata-u.ac.jp
Background
Patients with schizophrenia experience increased morbidity and mortality and have
an approximately 20% shorter life expectancy (Harris and Barraclough 1998),
manifesting as a premature mortality gap of 10–20 years compared with the
general population (Walker et al. 2015). Patients with schizophrenia also have a
relatively increased prevalence of cardiometabolic risk factors, such as obesity,
type 2 diabetes, hypertension, hyperglycemia, and dyslipidemia (Dixon et al. 2000;
McEvoy et al. 2005; Meyer and Stahl 2009). A retrospective cohort study of adult
patients with schizophrenia aged 20–64 years in the United States showed mortal-
ity rate was >3.5 times greater than that for general population adults for all-cause
standardized mortality ratios (3.7; 95% confidence interval [CI], 3.7–3.7) (Olfson
et al. 2015). In this cohort, schizophrenic patients with cardiovascular disease had
the highest mortality rate (403.2 per 100,000 person-years) and a standardized
mortality ratio of 3.6 (95% CI, 3.5%–3.6%) (Olfson et al. 2015). Furthermore,
meta-analysis revealed that 33% of people with schizophrenia have metabolic
syndrome, with a rate as high as 69% in schizophrenic patients with chronic illness
(Vancampfort et al. 2015). The odds ratio (OR) for metabolic syndrome in patients
with chronic schizophrenia compared to the general population was 2.35
(Vancampfort et al. 2013).
In the general population, lipid metabolism disorders have been associated
with an increased risk for cardiovascular events, such as myocardial infarction and
stroke (Kannel et al. 1971; Gordon et al. 1981; Castelli et al. 1986). A decreased
high-density lipoprotein (HDL) level, a cardiometabolic risk factor, is strongly and
inversely associated with the risk of coronary heart disease (Chirovsky et al. 2009).
Some studies in patients with schizophrenia attributed their metabolic disturbance
to poor lifestyle choices, such as a sedentary lifestyle, smoking, poor diet, and
prolonged stress, as well as use of antipsychotic drugs, and genetic susceptibility
(De Hert et al. 2009; Correll et al. 2011). In addition, significant elevations in
triglyceride and cholesterol levels were reported in association with second-
generation antipsychotic (SGA) treatment (Meyer 2002).
Although the individual differences of these metabolic disturbances have been
observed, an understanding of the underlying mechanisms is still uncertain (De Hert
Lipid Metabolism Disturbances During Antipsychotic Treatment for. . . 989
Metabolic risk factors are often present in patients with a first episode of psychosis or
drug-naive schizophrenia and are thought to be associated with sedentary lifestyle,
poor diet, smoking, and prolonged stress. Regarding sedentary lifestyle, the degree
of moderate and vigorous physical activity for the current population with schizo-
phrenia remains significantly less compared with healthy controls (Stubbs et al.
2016). Regarding HDL, levels were lower in both the drug-naive and drug-free
groups of patients with schizophrenia compared with healthy controls (Enez Darcin
et al. 2015). A case control study that compared differences in lipid metabolism
parameters between drug-free patients with a first episode of schizophrenia and
healthy individuals matched for age, race, and sex showed lower total cholesterol
and higher HDL levels for the schizophrenia patients (Wu et al. 2013). In contrast,
another study conducted in drug-free psychotic patients diagnosed with schizophre-
nia spectrum disorder after their first episode reported no significant difference in
lipid profile compared with healthy controls (Sengupta et al. 2008). Studies with
drug-naive patients with a first episode of schizophrenia found no significant differ-
ences in a range of metabolic measures, including blood glucose, lipids, insulin, and
measures of fat deposits (Arranz et al. 2004; Zhang et al. 2004). This no significant
metabolic difference from the healthy population has also been reported for drug-
naive participants in the European First Episode schizophrenia Trial (the EUFEST
study) (Fleischhacker et al. 2013). This range of findings illustrates the controversy
over the presence of dyslipidemia in unmedicated schizophrenia.
A meta-analysis investigated lipid metabolism in which fasting total cholesterol,
low-density lipoprotein (LDL), HDL, triglyceride, and adipocytokine parameters
were compared between individuals with a first episode of schizophrenia and no or
minimal antipsychotic drug exposure and healthy controls. The 20 case-control
studies that met inclusion criteria included 1167 patients and 1184 controls. Total
cholesterol and LDL levels were significantly decreased in schizophrenic patients
versus controls, corresponding to an absolute reduction of 0.26 mmol/L and
0.15 mmol/L, respectively. Triglyceride levels were also significantly increased in
the schizophrenia group, corresponding to an absolute increase of 0.08 mmol/L.
However, HDL and leptin levels were not altered in patients with schizophrenia
compared with controls (Pillinger et al. 2017).
990 S. Ono and T. Someya
It is well established that antipsychotic drug treatment can induce substantial weight
gain, with different drugs having greater or lesser effects and the greatest increases
occurring with SGA (Allison et al. 1999; Taylor and McAskill 2000). The mecha-
nisms underlying this antipsychotic drug-induced weight gain are likely multifacto-
rial. Antagonism of the serotonin receptor 5-HT2C, for which several antipsychotic
drugs – particularly olanzapine and clozapine – have high affinity, is a strong
candidate because the receptor is known to influence appetite and thereby weight
gain (Nonogaki et al. 1998). Association of a polymorphism of the 5-HT2C receptor
gene with antipsychotic drug-induced weight gain provides further evidence of the
importance of the receptor in this process (Reynolds et al. 2002). Dyslipidemia and
impaired glucose tolerance may be secondary to obesity and lifestyle habits or may
be caused by oral antipsychotics themselves.
Several database studies have estimated the relative risk of developing hyper-
lipidemia associated with SGAs. A nested case-control study of 1268 incident
cases of hyperlipidemia among 18,309 schizophrenia patients was conducted in
England and Wales (Koro et al. 2002). Olanzapine use was associated with nearly a
five-fold increase in the odds of developing hyperlipidemia compared with no
antipsychotic exposure, controlling for sex, age, and other medications and disease
conditions influencing lipid levels (OR, 4.65; 95% CI, 2.44%–8.85%). Risperi-
done was not associated with an increased OR for hyperlipidemia compared with
no antipsychotic exposure (1.12; 95% CI, 0.60%–2.11%), whereas the OR of
olanzapine versus first-generation antipsychotics (FGAs) was 1.4 (95% CI,
0.92%–2.22%).
In a study using a medical claims database in California and controlling for age,
ethnicity, type 2 diabetes, or hypothyroidism, the reported risk of developing
hyperlipidemia was increased for olanzapine (OR, 1.20; 95% CI, 1.08%–1.33%)
Lipid Metabolism Disturbances During Antipsychotic Treatment for. . . 991
compared with FGAs. Exposure to clozapine (OR, 1.16; 95% CI, 0.99%–1.37%),
risperidone (OR, 1.00; 95% CI, 0.90%–1.12%), and quetiapine (OR, 1.01; 95% CI,
0.78%–1.32%) was not associated with an increased risk during the 12-week
exposure window (Lambert et al. 2005).
A case-control analysis using pharmacy and claims in the United States compared
patients managed without antipsychotics to those who received antipsychotics
and showed that treatment was associated with a significant increase in risk of
incident hyperlipidemia with the following drugs: clozapine (OR, 1.82; 95% CI,
1.61%–2.05%), risperidone (OR, 1.53; 95% CI, 1.43%–1.64%), quetiapine (OR,
1.52; 95% CI, 1.40%–1.65%), olanzapine (OR, 1.56; 95% CI, 1.47%–1.67%),
ziprasidone (OR, 1.40; 95% CI, 1.19%–1.65%), and FGAs (OR, 1.26; 95% CI,
1.14%–1.39%). However, aripiprazole (OR, 1.19; 95% CI, 0.94%–1.52%) was not
associated with an increased risk (Olfson et al. 2006).
In Japan, a sequence symmetry analysis using health insurance claims data was
conducted to analyze the risks for hyperlipidemia associated with these drugs:
risperidone, paliperidone, perospirone hydrochloride hydrate, blonanserin, cloza-
pine, olanzapine, quetiapine fumarate, aripiprazole, and zotepine. Olanzapine was
significantly associated with increased hyperlipidemia (adjusted sequence ratio
[ASR], 1.56; 95% CI, 1.25%–1.95%). The ASRs obtained for risperidone (1.01;
95% CI, 0.80%–1.27%), perospirone (0.93; 95% CI, 0.63–1.39), blonanserin
(0.83; 95% CI, 0.52–1.33), quetiapine fumarate (0.93; 95% CI, 0.73%–1.18%),
and aripiprazole (1.02; 95% CI, 0.82%–1.26%) were all approximately 1.0. Unsta-
ble estimates, defined as wide CIs, were obtained for paliperidone and zotepine
because of the small sample size (Takeuchi et al. 2015). These database studies
show that exposure to olanzapine or clozapine results in a high OR for
hyperlipidemia.
SGAs are thought to be more commonly associated with dyslipidemia than
FGAs. Lund et al. investigated the risk of dyslipidemia in a comparison between
clozapine and FGAs in a retrospective cohort using medical and pharmacy claims
(Lund et al. 2001). They reported no significant differences in overall incidence
rates for hyperlipidemia between the two patient groups, adjusting for age, sex, and
duration of available follow-up. Among younger patients (age 20–34 years),
clozapine was associated with a significantly increased relative risk of hyperlipid-
emia (2.4 [95% CI, 1.1–5.2]). A systematic review and meta-analysis evaluated
SGAs and FGAs for the risk of dyslipidemia in patients with severe mental illness.
The analysis included a cross-sectional study, a cohort study, a case-control study,
and an intervention study that directly compared SGAs and FGAs in patients with
severe psychiatric disorders and evaluated dyslipidemia as the primary or second-
ary outcome. The reported associations between SGAs versus FGAs with
dyslipidemia were inconsistent, with high variability among studies, and ulti-
mately only a full qualitative synthesis was feasible. Determining the relative
risk of SGAs and FGAs in the effects of antipsychotics on dyslipidemia can be
difficult to clarify clinically because each drug can cause a variety of adverse
effects (Buhagiar and Jabbar 2019).
992 S. Ono and T. Someya
patients who gained >20% or <20% of their baseline body mass index, and no
significant differences in HDL levels between olanzapine- and risperidone-treated
patients were shown (Perez-Iglesias et al. 2009). In a randomized comparison study
of aripiprazole and olanzapine, no group differences were observed in the propor-
tions of patients with potentially clinically significant fasting HDL levels because the
weight gain at the study endpoint was significantly greater with olanzapine than with
aripiprazole (Fleischhacker et al. 2009).
In a 28-week RCT of olanzapine and ziprasidone, lipid parameters of
olanzapine versus ziprasidone were shown: increased total cholesterol (0.08 mmol/L
(SD 0.95) vs. mean, 0.33 mmol/L (SD 0.81)); increased LDL (0.02 mmol/L
(SD 0.77) vs. –0.27 mmol/L (SD 0.67)), and increased triglyceride (0.39 mmol/L
(SD 1.99) vs. –0.24 mmol/L (SD 1.09)). However, a significant decrease in HDL levels
was shown for the olanzapine (0.06 mmol/L (SD 0.26)) versus ziprasidone
(0.02 mmol/L (SD 0.25)) (Breier et al. 2005).
In a 6-week RCT of paliperidone extended release (ER), mean changes from
baseline in LDL, HDL, total cholesterol, and triglyceride levels were not different
among the lower-dose, higher-dose paliperidone ER and placebo (Canuso et al.
2010). In another study, paliperidone ER at three doses (3 mg, 9 mg, and 15 mg) and
olanzapine were compared with placebo in a 6-week RCT. No significant changes
from baseline to endpoint in LDL, HDL, total cholesterol, or triglyceride levels were
found (Davidson et al. 2007).
In a 6-week RCT of brexpiprazole, slight increases occurred in total cholesterol
and HDL levels from baseline to the last visit with 2 mg (0.63 mg/dL (SD 31.74)
and 1.33 mg/dL (SD 9.11)) and 4 mg (4.24 mg/dL (SD 30.33) and 0.48 mg/dL
(SD 7.26)) for the brexpiprazole groups compared with slight decreases in the
placebo group; however, these changes were neither clinically relevant nor statis-
tically significant. Slight increases were also shown in LDL and triglyceride levels
in the 4-mg brexpiprazole group (2.57 mg/dL (SD 26.79) and 6.76 mg/dL
(SD 69.54)); however, these differences were also not clinically relevant (Correll
et al. 2015).
In a phase III trial of cariprazine, mean changes from baseline of total cholesterol
and LDL levels were significantly decreased in the cariprazine 6- mg/day group
(4.5 md/dL (SD 32.4) and 3.5 md/dL (SD 30.7)) versus placebo. No other
differences among groups were shown for changes in total cholesterol, LDL,
HDL, and triglyceride levels (Durgam et al. 2015). Subsequent placebo-controlled
trials also reported similar results for parameters such as total cholesterol, LDL,
HDL, and triglyceride levels (Earley et al. 2017).
In a 6-week RCT of lurasidone, endpoint changes in levels of total cholesterol,
triglyceride, and LDL were comparable for both the lurasidone (40 mg/day and
120 mg/day) and placebo groups, whereas the olanzapine group showed a signif-
icant median increase compared with the placebo group in cholesterol levels,
LDL, and triglyceride (Meltzer et al. 2011). In another 6-week RCT in Asia, no
clinically meaningful differences were shown between lurasidone and placebo in
terms of the effect on total cholesterol, LDL, HDL, and triglyceride levels
(Higuchi et al. 2019).
994 S. Ono and T. Someya
known that smoking alters lipid profiles, as characterized by increased total choles-
terol, triglyceride, and LDL levels, along with decreased HDL levels (Maeda et al.
2003); however, the mechanisms by which smoking changes lipid levels are still
unknown. A few studies have investigated the smoking versus nonsmoking and lipid
profiles in schizophrenia, and one such study showed no significant difference
in total cholesterol, triglyceride, HDL, and LDL levels between the smokers and
nonsmokers (Hui-Mei et al. 2016).
In the general population, cholesterol and triglyceride levels are known to be
affected by age and sex. It was suggested that women have a greater vulnerability to
develop psychotropic-drug–induced metabolic disturbances than men (Correll et al.
2015; Seeman 2009). The risk of developing hypercholesterolemia was significantly
greater for female versus male patients in a retrospective adolescent cohort
(McIntyre and Jerrell 2008). A study that investigated the sex differences in the
effects of antipsychotics on glucose and lipid metabolism in first-episode schizo-
phrenia randomized 112 patients with schizophrenia to receive clozapine,
olanzapine, risperidone, or sulpiride for 8 weeks. After treatment, the waist-to-hip
ratio, triglyceride levels, and insulin resistance index for men were higher than that
for women in the clozapine and olanzapine groups. In the sulpiride group, body mass
index and triglyceride levels, insulin levels, and insulin resistance index for women
was higher than that for men. No significant sex difference was shown for all
assessments in the risperidone group (Wu et al. 2007).
Body mass index and triglyceride levels are also positively correlated in the
general population. A study on the prevalence of obesity in chronic schizophrenia
reported an approximately two times rate higher for female versus male patients.
This study used a stepwise multiple regression to determine which factors contrib-
uted to obesity and found that type 2 diabetes and triglyceride levels were signifi-
cantly associated with obesity in female patients but that only triglyceride levels
were associated with obesity in male patients (Li et al. 2016). It is thought that lipid
metabolism is largely regulated by sex hormones. Indeed, in another study plasma
triglyceride concentrations during both fasting and fed conditions were reported to
be lower in premenopausal women than in age-matched men in the general popula-
tion (Wang et al. 2011).
metabolic parameters except HDL and fasting glucose, which was significantly
different than aripiprazole for every outcome except fasting glucose. In the black/
Hispanic cohort of 137 patients, olanzapine resulted in adverse metabolic outcomes,
and these changes were significantly different from aripiprazole for adiposity, total
cholesterol, and non- HDL cholesterol. Aripiprazole decreased the odds of metabolic
syndrome at the study endpoint compared with olanzapine for all study participants
(OR, 0.33; 95% CI, 0.19–0.55), for the white cohort (OR, 0.20; 95% CI, 0.10–0.41),
and for the African American/Hispanic cohort (OR, 0.53; 95% CI, 0.25–1.12);
however, the results for the African American/Hispanic were not statistically signifi-
cant ( p ¼ 0.096). Within the aripiprazole group, white patients had significantly lower
risk for metabolic syndrome, whereas no significant difference in metabolic syndrome
was evident between white and African American/Hispanic patients exposed to
olanzapine. These findings suggest that race/ethnicity may be an important moderator
of metabolic risk during SGA therapy (Meyer et al. 2009).
References
Allison DB, Mentore JL, Heo M, et al. Antipsychotic-induced weight gain: a comprehensive
research synthesis. Am J Psychiatr. 1999;156:1686–96.
Arranz B, Rosel P, Ramírez N, Dueñas R, Fernández P, Sanchez JM, Navarro MA,
San L. Insulin resistance and increased leptin concentrations in noncompliant schizophrenia
patients but not in antipsychotic-naive first-episode schizophrenia patients. J Clin Psychiatry.
2004;65(10):1335–42.
1000 S. Ono and T. Someya
Baller JB, McGinty EE, Azrin ST, Juliano-Bult D, Daumit GL. Screening for car- diovascular
risk factors in adults with serious mental illness: a review of the evidence. BMC Psychiatry.
2015;15:55.
Breier A, Berg PH, Thakore JH, Naber D, Gattaz WF, Cavazzoni P, Walker DJ, Roychowdhury SM,
Kane JM. Olanzapine versus ziprasidone: results of a 28-week double-blind study in patients
with schizophrenia. Am J Psychiatry. 2005;162(10):1879–87.
Buhagiar K, Jabbar F. Association of first- vs. second-generation antipsychotics with lipid
abnormalities in individuals with severe mental illness: a systematic review and meta-analysis.
Clin Drug Investig. 2019;39(3):253–73.
Canuso CM, Lindenmayer JP, Kosik-Gonzalez C, Turkoz I, Carothers J, Bossie CA,
Schooler NR. A randomized, double-blind, placebo-controlled study of 2 dose ranges of
paliperidone extended-release in the treatment of subjects with schizoaffective disorder. J Clin
Psychiatry. 2010;71(5):587–98.
Castelli WP, Garrison RJ, Wilson PW, Abbott RD, Kalousdian S, Kannel WB. Incidence of
coronary heart disease and lipoprotein cholesterol levels: the Framingham study. JAMA.
1986;256:2835–8.
Chirovsky DR, Fedirko V, Cui Y, Sazonov V, Barter P. Prospective studies on the relationship
between high-density lipoprotein cholesterol and cardiovascular risk: a systematic review. Eur
J Cardiovasc Prev Rehabil. 2009;16:404–23.
Chrzanowski WK, Marcus RN, Torbeyns A, Nyilas M, McQuade RD. Effectiveness of long-term
aripiprazole therapy in patients with acutely relapsing or chronic, stable schizophrenia: a
52-week, open-label comparison with olanzapine. Psychopharmacology. 2006;189:259–66.
Cooper SJ, Reynolds GP, Barnes T, With expert co-authors (in alphabetical order), et al.
BAP guidelines on the management of weight gain, metabolic disturbances and cardiovascular
risk associated with psychosis and antipsychotic drug treatment. J Psychopharmacol.
2016;30(8):717–48.
Correll CU, Lencz T, Malhotra AK. Antipsychotic drugs and obesity. Trends Mol Med. 2011;17(2):
97–107.
Correll CU, Skuban A, Ouyang J, Hobart M, Pfister S, McQuade RD, Nyilas M, Carson WH,
Sanchez R, Eriksson H. Efficacy and safety of brexpiprazole for the treatment of acute
schizophrenia: a 6-week randomized, double-blind, placebo-controlled trial. Am J Psychiatry.
2015;172(9):870–80.
Daumit GL, Goff DC, Meyer JM, Davis VG, Nasrallah HA, McEvoy JP, Rosenheck R, Davis SM,
Hsiao JK, Stroup TS, Lieberman JA. Antipsychotic effects on estimated 10-year coronary heart
disease risk in the CATIE schizophrenia study. Schizophr Res. 2008;105(1–3):175–87.
Davidson M, Emsley R, Kramer M, Ford L, Pan G, Lim P, Eerdekens M. Efficacy, safety and early
response of paliperidone extended-release tablets (paliperidone ER): results of a 6-week,
randomized, placebo-controlled study. Schizophr Res. 2007;93(1–3):117–30.
De Hert M, Schreurs V, Vancampfort D, VAN Winkel R. Metabolic syndrome in people with
schizophrenia: a review. World Psychiatry. 2009;8:15–22.
De Hert M, Vancampfort D, Correll CU, Mercken V, Peuskens J, Sweers K, van Winkel R,
Mitchell AJ. Guidelines for screening and monitoring of cardiometabolic risk in schizophrenia:
systematic evaluation. Br J Psychiatry. 2011;199(2):99–105.
De Hert M, Detraux J, van Winkel R, Yu W, Correll CU. Metabolic and cardiovascular adverse
effects associated with antipsychotic drugs. Nat Rev Endocrinol. 2012;8(2):114–26.
de Leon J, Diaz F. A meta-analysis of worldwide studies demonstrates an association between
schizophrenia and tobacco smoking behaviors. Schizophr Res. 2005;7(6):135–57.
Del Valle MC, Loebel AD, Murray S, Yang R, Harrison DJ, Cuffel BJ. Change in framingham risk
score in patients with schizophrenia: a post hoc analysis of a randomized, double-blind, 6-week
trial of ziprasidone and olanzapine. Prim Care Companion J Clin Psychiatry. 2006;8(6):329–33.
Delacretaz A, Vandenberghe F, Gholam-Rezaee M, et al. Early changes of blood lipid levels during
psychotropic drug treatment as predictors of long-term lipid changes and of new onset
dyslipidemia. J Clin Lipidol. 2018;12(1):219–29.
Lipid Metabolism Disturbances During Antipsychotic Treatment for. . . 1001
Lambert BL, et al. Association between antipsychotic treatment and hyperlipidemia among
California Medicaid patients with schizophrenia. J Clin Psychopharmacol. 2005;25(1):12–8.
Li Q, Chen D, Liu T, et al. Sex differences in body mass index and obesity in chinese patients with
chronic schizophrenia. J Clin Psychopharmacol. 2016;36:643–8.
Lindenmayer JP, Czobor P, Volavka J, Citrome L, Sheitman B, McEvoy JP, Cooper TB, Chakos M,
Lieberman JA. Changes in glucose and cholesterol levels in patients with schizophrenia treated
with typical or atypical antipsychotics. Am J Psychiatry. 2003;160(2):290–6.
Lund BC, Perry PJ, Brooks JM, et al. Clozapine use in patients with schizophrenia and the risk of
diabetes, hyperlipidemia, and hypertension: a claims-based approach. Arch Gen Psychiatry.
2001;58:1172–6.
Maeda K, Noguchi Y, Fukui T. The effects of cessation from cigarette smoking on the lipid and
lipoprotein profiles: a meta-analysis. Prev Med. 2003;37:283–90.
Manu P, Dima L, Shulman M, et al. Weight gain and obesity in schizophrenia: epidemiology,
pathobiology, and management. Acta Psychiatr Scand. 2015;132(2):97–108.
McEvoy JP, Meyer JM, Goff DC, Nasrallah HA, Davis SM, Sullivan L, et al. Prevalence of
the metabolic syndrome in patients with schizophrenia: baseline results from the Clinical
Antipsychotic Trials of Intervention Effectiveness (CATIE) Schizophrenia Trial and comparison
with national estimates from NHANES III. Schizophr Res. 2005;80:19–32.
McIntyre RS, Jerrell JM. Metabolic and cardiovascular adverse events associated with antipsy-
chotic treatment in children and adolescents. Arch Pediatr Adolesc Med. 2008;162(10):929–35.
McQuade RD, Stock E, Marcus R, Jody D, Gharbia NA, Vanveggel S, Archibald D,
Carson WH. A comparison of weight change during treatment with olanzapine or aripiprazole:
results from a randomized, double-blind study. J Clin Psychiatry. 2004;65(Suppl 18):47–56.
Meltzer HY, Cucchiaro J, Silva R, Ogasa M, Phillips D, Xu J, Kalali AH, Schweizer E, Pikalov A,
Loebel A. Lurasidone in the treatment of schizophrenia: a randomized, double-blind, placebo-
and olanzapine-controlled study. Am J Psychiatry. 2011;168(9):957–67.
Meyer JM. A retrospective comparison of weight, lipid, and glucose changes between risperidone-
and olanzapine-treated inpatients: metabolic outcomes after 1 year. J Clin Psychiatry.
2002;63(5):425–33.
Meyer JM, Stahl SM. The metabolic syndrome and schizophrenia. Acta Psychiatr Scand. 2009;119:
4–14.
Meyer JM, Rosenblatt LC, Kim E, Baker RA, Whitehead R. The moderating impact of ethnicity
on metabolic outcomes during treatment with olanzapine and aripiprazole in patients with
schizophrenia. J Clin Psychiatry. 2009;70(3):318–25.
Misiak B, Stanczykiewicz B, Laczmanski L, Frydecka D. Lipid profile disturbances in
antipsychotic-naive patients with first-episode non-affective psychosis: a systematic review
and meta-analysis. Schizophr Res. 2017;190:18–27.
Mitchell AJ, Vancampfort D, Sweers K, van Winkel R, Yu W, De Hert M. Prevalence of metabolic
syndrome and metabolic abnormalities in schizophrenia and related disorders – a systematic
review and meta-analysis. Schizophr Bull. 2013;39:306–18.
Mizuno Y, Suzuki T, Nakagawa A, Yoshida K, Mimura M, Fleischhacker WW,
Uchida H. Pharmacological strategies to counteract antipsychotic-induced weight gain and
metabolic adverse effects in schizophrenia: a systematic review and meta-analysis. Schizophr
Bull. 2014;40(6):1385–403.
Nonogaki K, Strack AM, Dallman MF, et al. Leptin-independent hyperphagia and type 2 diabetes in
mice with a mutated serotonin 5-HT2C receptor gene. Nat Med. 1998;4:1152–6.
Olfson M, Marcus SC, Corey-Lisle P, et al. Hyperlipidemia following treatment with antipsychotic
medications. Am J Psychiatry. 2006;163:1821–5.
Olfson M, Gerhard T, Huang C, Crystal S, Stroup TS. Premature mortality among adults with
schizophrenia in the United States. JAMA Psychiat. 2015;72(12):1172–81.
Paramsothy P, Knopp R, Bertoni AG, Tsai MY, Rue T, Heckbert SR. Combined hyperlipidemia
in relation to race/ethnicity, obesity, and insulin resistance in the Multi-Ethnic Study of
Atherosclerosis. Metabolism. 2009;58:212–9.
Lipid Metabolism Disturbances During Antipsychotic Treatment for. . . 1003
Perez-Iglesias R, Mata I, Pelayo-Teran JM, Amado JA, Garcia-Unzueta MT, Berja A, et al. Glucose
and lipid disturbances after 1 year of antipsychotic treatment in a drug-naïve population.
Schizophr Res. 2009;107:115–21.
Peter WFW, Ralph BD’A, Daniel L, et al. Prediction of coronary heart disease using risk factor
categories. Circulation. 1998;97:1837–1847. https://www.ahajournals.org/doi/full/10.1161/01.
CIR.97.18.1837.
Pillinger T, Beck K, Stubbs B, Howes OD. Cholesterol and triglyceride levels in first-episode
psychosis: systematicreview and meta-analysis. Br J Psychiatry. 2017;211(6):339–49.
Pillinger T, McCutcheon RA, Vano L, Mizuno Y, Arumuham A, Hindley G, Beck K, Natesan S,
Efthimiou O, Cipriani A, Howes OD. Comparative effects of 18 antipsychotics on metabolic
function in patients with schizophrenia, predictors of metabolic dysregulation, and association
with psychopathology: a systematic review and network meta-analysis. Lancet Psychiatry.
2020;7(1):64–77.
Reynolds GP, Zhang ZJ, Zhang XB. Association of antipsychotic drug-induced weight gain with a
polymorphism of the promoter region of the 5-HT2C receptor gene. Lancet. 2002;359:2086–7.
Rummel-Kluge C, Komossa K, Schwarz S, Hunger H, Schmid F, Lobos CA, Kissling W, Davis JM,
Leucht S. Head-to-head comparisons of metabolic side effects of second generation antipsy-
chotics in the treatment of schizophrenia: a systematic review and meta-analysis. Schizophr Res.
2010;123(2–3):225–33.
Seeman MV. Secondary effects of antipsychotics: women at greater risk than men. Schizophr Bull.
2009;35(5):937–48.
Sengupta S, Parrilla-Escobar MA, Klink R, Fathalli F, Ying KN, Stip E, Baptista T, Malla A,
Joober R. Are metabolic indices different between drug-naive first-episode psychosis patients
and healthy controls? Schizophr Res. 2008;102(1–3):329–36.
Siafis S, Tzachanis D, Samara M, Papazisis G. Antipsychotic drugs: from receptor-binding profiles
to metabolic side effects. Curr Neuropharmacol. 2018;16(8):1210–23.
Smith RC, Lindenmayer JP, Bark N, Warner-Cohen J, Vaidhyanathaswamy S,
Khandat A. Clozapine, risperidone, olanzapine, and conventional antipsychotic drug effects
on glucose, lipids, and leptin in schizophrenic patients. Int J Neuropsychopharmacol. 2005;8(2):
183–94.
Smith RC, Lindenmayer JP, Hu Q, Kelly E, Viviano TF, Cornwell J, Vaidhyanathaswamy S,
Marcovina S, Davis JM. Effects of olanzapine and risperidone on lipid metabolism in chronic
schizophrenic patients with long-term antipsychotic treatment: a randomized five month study.
Schizophr Res. 2010;120(1–3):204–9.
Spivak B, Lamschtein C, Talmon Y, Guy N, Mester R, Feinberg I, Kotler M,
Weizman A. The impact of clozapine treatment on serum lipids in chronic schizophrenic
patients. Clin Neuropharmacol. 1999;22:98–101.
Stroup TS, McEvoy JP, Ring KD, Hamer RH, LaVange LM, Swartz MS, Rosenheck RA,
Perkins DO, Nussbaum AM, Lieberman JA, Schizophrenia Trials Network. A randomized
trial examining the effectiveness of switching from olanzapine, quetiapine, or risperidone to
aripiprazole to reduce metabolic risk: comparison of antipsychotics for metabolic problems
(CAMP). Am J Psychiatry. 2011;168(9):947–56.
Stroup TS, Byerly MJ, Nasrallah HA, Ray N, Khan AY, Lamberti JS, Glick ID, Steinbook RM,
McEvoy JP, Hamer RM. Effects of switching from olanzapine, quetiapine, and risperidone to
aripiprazole on 10-year coronary heart disease risk and metabolic syndrome status: results from
a randomized controlled trial. Schizophr Res. 2013;146(1–3):190–5.
Stubbs B, Firth J, Berry A, Schuch FB, Rosenbaum S, Gaughran F, et al. How much physical
activity do people with schizophrenia engage in? A systematic review, comparative meta-
analysis and meta-regression. Schizophr Res. 2016;176:431–40.
Takeuchi Y, Kajiyama K, Ishiguro C, Uyama Y. Atypical antipsychotics and the risk of hyperlip-
idemia: a sequence symmetry analysis. Drug Saf. 2015;38(7):641–50.
Taylor DM, McAskill R. Atypical antipsychotics and weight gain – a systematic review. Acta
Psychiatr Scand. 2000;101:416–32.
1004 S. Ono and T. Someya
Vancampfort D, Wampers M, Mitchell AJ, Correll CU, De Herdt A, Probst M, et al. A meta-analysis
of cardio-metabolic abnormalities in drug naive, first-episode and multi-episode patients with
schizophrenia versus general population controls. World Psychiatry. 2013;12:240–50.
Vancampfort D, Stubbs B, Mitchell AJ, De Hert M, Wampers M, Ward PB, et al. Risk of metabolic
syndrome and its components in people with schizophrenia and related psychotic disorders,
bipolar disorder and major depressive disorder: a systematic review and meta-analysis. World
Psychiatry. 2015;14:339–47.
Vancampfort D, Firth J, Correll CU. The impact of pharmacological and non-pharmacological
interventions to improve physical health outcomes in people with schizophrenia: a meta-review
of meta-analyses of randomized controlled trials. World Psychiatry. 2019;18(1):53–66.
Walker ER, McGee RE, Druss BG. Druss. Mortality in mental disorders and global disease burden
implications: a systematic review and meta-analysis. JAMA Psychiat. 2015;72(4):334–41.
Wang X, Magkos F, Mittendorfer B. Sex differences in lipid and lipoprotein metabolism: it’s not
just about sex hormones. J Clin Endocrinol Metab. 2011;96:885–93.
Weiden PJ, Buckley PF. Reducing the burden of side effects during long-term antipsychotic
therapy: the role of “switching” medications. J Clin Psychiatry. 2007;68(Suppl 6):14–23.
Wilson PW, D’Agostino RB, Levy D, Balanger AM, Silbershatz H, Kannel WB. Prediction of
coronary heart disease using risk factor categories. Circulation. 1998;97:1837–47.
Wu RR, Zhao JP, Zhai JG, Guo XF, Guo WB. Sex difference in effects of typical and atypical
antipsychotics on glucose-insulin homeostasis and lipid metabolism in first-episode schizophre-
nia. J Clin Psychopharmacol. 2007;27(4):374–9.
Wu X, Huang Z, Wu R, Zhong Z, Wei Q, Wang H, Diao F, Wang J, Zheng L, Zhao J,
Zhang J. The comparison of glycometabolism parameters and lipid profiles between drug-
naïve, first-episode schizophrenia patients and healthy controls. Schizophr Res. 2013;150(1):
157–62.
Wulffele MG, Kooy A, de Zeeuw D, Stehouwer CD, Gansevoort RT. The effect of metformin on
blood pressure, plasma cholesterol and triglycerides in type 2 diabetes mellitus: a systematic
review. J Intern Med. 2004;256(1):1–14.
Yan H, Chen JD, Zheng XY. Potential mechanisms of atypical antipsychotic-induced hyper-
triglyceridemia. Psychopharmacology. 2013;229(1):1–7.
Zhang ZJ, Yao ZJ, Liu W, Fang Q, Reynolds GP. Effects of antipsychotics on fat deposition and
changes in leptin and insulin levels. Magnetic resonance imaging study of previously untreated
people with schizophrenia. Br J Psychiatry. 2004;184:58–62.
Zhou T, Xu X, Du M, Zhao T, Wang J. A preclinical overview of metformin for the treatment of type
2 diabetes. Biomed Pharmacother. 2018;106:1227–35.
Cannabinoid Drugs in Mental Health
Disorders
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1006
Search Strategy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1007
Cannabinoids in Treatment of Mental Health Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1008
Anxiety Disorders / OCD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1009
Post-Traumatic Stress Disorder (PTSD) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1016
Psychotic Disorders . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1017
Tourette Syndrome . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1017
Dementia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1017
Depression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1018
Other Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1019
Cannabinoids in Treatment of Substance Use Disorders . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1019
Cannabis Use Disorder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1019
Nicotine Dependence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1027
Opioid Use Disorder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1028
Alcohol Use Disorder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1029
S. Kloiber (*)
Campbell Family Mental Health Research Institute, Centre for Addiction and Mental Health
(CAMH), Department of Psychiatry, University of Toronto, Toronto, ON, Canada
e-mail: Stefan.Kloiber@camh.ca
J. Matheson
Translational Addiction Research Laboratory, CAMH, Toronto, ON, Canada
e-mail: Justin.Matheson@camh.ca
H. K. Kim
Department of Psychiatry, University of Toronto, Toronto, ON, Canada
e-mail: helena.kim@camh.ca
B. Le Foll
Translational Addiction Research Laboratory, Acute Care Program, and Campbell Family Mental
Health Research Institute, CAMH, Departments of Family and Community Medicine,
Pharmacology and Toxicology, Psychiatry, and Institute of Medical Science, University of Toronto,
Toronto, ON, Canada
e-mail: Bernard.LeFoll@camh.ca
Abstract
Endocannabinoid signalling plays an important role in affect, anxiety, and
reward, and thus targeting this system could potentially be used to treat mental
health and substance use disorders. This chapter discusses the current state of the
clinical evidence evaluating the potential of various cannabinoid drugs as mental
health and addiction pharmacotherapy, with a focus on randomized controlled
trials. A few small clinical trials have found preliminary evidence for different
cannabinoids in mental health disorders, including cannabidiol (CBD) for anxiety
disorders, and Δ9-tetrahydrocannabinol (THC) for post-traumatic stress disorder
(PTSD) and Tourette syndrome. The evidence for cannabinoids in other mental
health conditions (e.g., psychotic disorders) is mixed. The clinical evidence also
suggests that a combination of CBD and THC may be useful in reducing cannabis
craving and/or withdrawal in patients with cannabis use disorder (CUD), with
possible longer-term effects on reducing cannabis use and promoting abstinence.
CBD alone may have the potential to reduce cannabis use, promote tobacco
smoking cessation, and attenuate opioid craving, though these findings need
replication. Finally, the cannabinoid type-1 (CB1) receptor antagonist rimonabant
showed promise as a smoking cessation pharmacotherapy, yet this drug was
withdrawn from the market due to serious psychiatric adverse effects. The current
evidence demonstrates the potential for cannabinoid drugs in the treatment of
mental health and substance use disorders, yet this evidence is clearly in its early
stages. Future directions for the field are discussed.
Introduction
very little activity at CB2, while 2-AG is a low-affinity, high-efficacy agonist for both
receptors (Lu and Mackie 2016). Many enzymes are involved in the synthesis and
degradation of AEA and 2-AG, including fatty acid amide hydrolase (FAAH), which
is responsible for the majority of AEA metabolism, and monoacylglycerol lipase
(MAGL), which is often responsible for the degradation of 2-AG (Pamplona and
Takahashi 2012).
The wide expression of CB1 receptors in the brain, especially in the frontal cortex,
basal ganglia, and medial temporal lobes (Hu and Mackie 2015), positions the ECS
as an important mediator of cognitive and affective processes. Research has consis-
tently shown an important role of the ECS in symptoms and models of anxiety,
depression, and psychosis in both animals and humans (Ibarra-Lecue et al. 2018;
Connor et al. 2021). In humans, there is a strong link between the use of cannabis/
cannabinoids and mood and anxiety disorders (Botsford et al. 2020; Lev-Ran et al.
2012; Mammen et al. 2018), and alleviation of mood/anxiety symptoms is a com-
mon motivation for using cannabis (Wycoff et al. 2018). Similarly, there is a wealth
of evidence demonstrating the role of the ECS in drug reward and addiction (Parsons
and Hurd 2015). For example, CB1 receptor signalling plays a role in nicotine,
cannabis, alcohol, and opioid behaviours in animals and humans (Sloan et al. 2017;
Le Foll and Goldberg 2005).
To date, several different ECS-targeting drugs have been evaluated for their
potential to treat mental health disorders and substance use disorders (SUDs).
These include Δ9-tetrahydrocannabinol (THC), the primary psychoactive compound
in the cannabis plant, and its encapsulated oral formulation dronabinol (Marinol ®);
cannabidiol (CBD), a non-intoxicating cannabinoid; nabiximols (Sativex ®), an
oromucosal spray containing both THC and CBD (2.7 mg THC and 2.5 mg CBD
per spray); nabilone (Cesamet ®), a synthetic analogue of THC; and various CB1
receptor antagonists/inverse agonists, most notably rimonabant, among other drugs
targeting the ECS.
The goal of this chapter is to provide an overview of the existing clinical
evidence, with a focus on controlled clinical trials evaluating the feasibility and/or
efficacy of ECS-targeting drugs as treatment for mental health and addictive disor-
ders. Preclinical and human laboratory evidence will be discussed to support the
clinical evidence where relevant.
Search Strategy
PubMed was searched in February 2021 with the following terms: cannabinoid or
THC or CBD or “FAAH inhibitor” or “MAGL inhibitor” and addiction or “use
disorder” or “mental health” or psychiatry or depression or anxiety or schizophrenia
or “mood disorder” or GAD or OCD or “social phobia” or SAD or PTSD), and the
“Clinical Trial” filter was applied. This resulted in a total of 424 articles. Eligible
studies were clinical trials or case series in humans, involved the administration of a
drug targeting the ECS, and included a primary endpoint related to mental health or
substance use. Studies were excluded if they recruited only healthy participants.
1008 S. Kloiber et al.
Fig. 1 Clinical trials assessing the effects of cannabinoid drugs on mental health symptoms and
psychiatric disorders
cannabinoids, primarily THC, on mental health are also well studied, where
cannabinoids were associated with early age of onset and severity of bipolar
disorder, risk for schizophrenia and worsening of psychotic symptoms, and neg-
ative effects on anxiety and mood (reviewed in Suryadevara et al. 2017). Canna-
binoids examined for their potential effects in mental health disorders to date
include CBD, THC, nabiximols, dronabinol, rimonabant, and nabilone (Black
et al. 2019; Bonaccorso et al. 2019). The following sections will discuss the
potential effect of these compounds in anxiety disorders and OCD, PTSD,
psychotic disorders, Tourette syndrome, psychiatric symptoms of dementia,
non-motor symptoms of Parkinson’s disease, depression, and other conditions.
See Table 1 for an overview of controlled clinical studies that met our eligibility
criteria.
Table 1 Clinical trials assessing the effects of cannabinoid drugs on mental health symptoms in
patients
Reference Sample Study design Intervention Relevant findings
Anxiety disorders / OCD
Bergamaschi 24 treatment- Double-blind, CBD 600 mg, CBD significantly
et al. (2011) naïve adult placebo- single dose reduced anxiety,
patients with controlled, cognitive
social anxiety randomized impairment, and
disorder study discomfort during a
(simulated public simulated public
speaking test) speaking test as
measured by the
Visual Analogue
Mood Scale
compared to placebo
Crippa et al. 10 male adult Double-blind, CBD 400 mg, CBD significantly
(2011) patients with placebo- single dose decreased subjective
social anxiety controlled, anxiety measured by
disorder cross-over the Visual Analogue
study Mood Scale
compared to placebo
Fabre and 20 adult patients Double-blind, Nabilone 1 mg Nabilone
McLendon with placebo- three times a day significantly
(1981) psychoneurotic controlled for 28 days decreased anxiety
anxiety study as measured by the
Hamilton
Anxiety Rating
Scale compared to
placebo
Kayser et al. 12 adult patients Placebo- Single dose, 50% Both THC and
(2020) with obsessive- controlled, of 800 mg CBD did not change
compulsive randomized mixture OCD symptoms as
disorder study administered in measured by the
cigarette form of Yale-Brown
THC and CBD Obsessive-
combined, Compulsive Scale
ranging from 7% and the Obsessive-
to 0% THC and Compulsive Visual
10.4% to 0% Analog Scale
CBD. compared to
placebo
Masataka 37 teenage Double-blind, CBD oil 300 mg CBD significantly
(2019) patients with placebo- daily for 4 weeks lowered symptoms
social anxiety controlled of social anxiety
disorder and study disorder as
avoidant measured by the
personality Fear of Negative
disorder Evaluation
Questionnaire and
the Liebowitz
Social Anxiety
Scale compared to
placebo
(continued)
Cannabinoid Drugs in Mental Health Disorders 1011
Table 1 (continued)
Reference Sample Study design Intervention Relevant findings
Post-traumatic stress disorder
Jetly et al. 10 Canadian Double-blind, Nabilone Nabilone produced a
(2015) male military placebo- 0.5–3 mg for significant decrease
personnel with controlled, 7 weeks in nightmares as
PTSD randomized, measured by the
cross-over Clinician-
study Administered PTSD
Scale, Recurring and
Distressing Dream
Scale, and increased
general well-being
measured by the
General Well Being
Questionnaire
compared to placebo
Rabinak et al. 19 adult patients Randomized, Dronabinol Response time to
(2020) with PTSD and double-blind, 7.5 mg single threatening faces
27 trauma- and placebo- dose was used as a
exposed adults controlled marker of fear
without PTSD study response in patients
combined with PTSD, where
slower response
time would indicate
greater fear and
anxiety. THC
resulted in faster
response time to
threatening faces
compared to
placebo treatment
Psychotic disorders
Boggs et al. 17 adult patients Double-blind, Adjunctive Rimonabant did not
(2012) with placebo- rimonabant produce
schizophrenia or controlled, 20 mg daily for improvement in the
schizoaffective randomized 16 weeks Repeatable Battery
disorder being study for the
treated with Assessment of
second- Neuropsychological
generation Status total score
antipsychotics compared to placebo
Boggs et al. 41 adult patients Double- Adjunctive CBD CBD augmentation
(2018) with blinded, 600 mg daily for was not associated
schizophrenia randomized, 6 weeks with improvement in
treated with placebo- Positive And
antipsychotics controlled Negative Syndrome
study Scale (PANSS)
scores or MATRICS
Consensus
Cognitive Battery
scores compared to
placebo
(continued)
1012 S. Kloiber et al.
Table 1 (continued)
Reference Sample Study design Intervention Relevant findings
Kelly et al. 15 adult patients Double-blind, Adjunctive Rimonabant
(2011) with placebo- rimonabant significantly
schizophrenia or controlled 20 mg daily for decreased Brief
schizoaffective study 16 weeks Psychiatric Rating
disorder taking Scale (BPRS)
second- scores compared to
generation placebo
antipsychotics
Leweke et al. 39 adult patients Double-blind, Cannabidiol Cannabidiol
(2012) with acute placebo- 600–800 mg produced clinical
paranoid controlled, daily for 28 days improvement as
schizophrenia and measured by the
randomized Positive And
study Negative Syndrome
Scale (PANSS)
compared to placebo
McGuire 88 adult patients Double-blind, Adjunct CBD CBD lowered
et al. (2018) with randomized, 1000 mg/day for positive symptoms
schizophrenia placebo- 6 weeks as measured
treated with controlled Positive And
antipsychotics study Negative Syndrome
Scale (PANSS),
produced
significant global
improvement as
measured by the
Clinical Global
Impression
scale (CGI), but did
not produce
significant changes
in overall
functioning or
cognitive
performance
compared to
placebo
Meltzer et al. 72 adult patients Double-blind, SR141716 (CB1 SR141716 did not
(2004) with randomized, antagonist) for change any
schizophrenia or placebo- 6 weeks outcomes compared
schizoaffective controlled to placebo as
disorder study measured by the
Positive and
Negative Syndrome
Scale (PANSS),
Brief Psychiatric
Rating
Scale (BPRS), and
Clinical Global
Impression Scale
(CGI)
(continued)
Cannabinoid Drugs in Mental Health Disorders 1013
Table 1 (continued)
Reference Sample Study design Intervention Relevant findings
Schwarcz 6 adult patients Clinical case Dronabinol Four out of 6
et al. (2009) with severe series up-titrated from patients improved
chronic 2.5 mg twice a with dronabinol as
schizophrenia day to 10 mg measured by the
treated with twice a day in Clinical Global
antipsychotics 3 weeks, then Impression Scale
maintained at (CGI) and Brief
10 mg for up to Psychiatric Rating
5 more weeks. Scale (BPRS)
compared to
baseline.
Zuardi et al. 3 male inpatients Cross-over CBD up to One out of 3 patients
(2006) with treatment- design, 1280 mg daily showed mild
resistant placebo- for 29 days improvement as
schizophrenia controlled measured by the
Brief Psychiatric
Rating Scale
(BPRS) and Positive
and Negative
Syndrome Scale
(PANSS)
Tourette syndrome
Muller-Vahl 12 adult patients Double-blind, THC 5–10 mg THC significantly
et al. (2002) with Tourette placebo- per day for improved tics and
syndrome controlled, 2 days obsessive-
randomized- compulsive
cross-over behaviour, motor
design tics, but not vocal
tics as measured by
the Tourette’s
Syndrome Symptom
List, Yale Global Tic
Severity Scale,
Shapiro Tourette’s
Syndrome Severity
Scale, and Tourette’s
Syndrome Global
Scale compared to
placebo
Muller-Vahl 24 adult patients Double-blind, THC 10 mg daily THC decreased
et al. (2003) with Tourette placebo- for 6 weeks symptoms as
syndrome controlled, measured by the
randomized Tourette Syndrome
study Clinical Global
Impression Scale,
Tourette-Syndrome
Severity Scale, Yale
Global Tic Severity
Scale, and Video
Rating Scale
compared to placebo
(continued)
1014 S. Kloiber et al.
Table 1 (continued)
Reference Sample Study design Intervention Relevant findings
Dementia (psychiatric symptoms)
de Faria et al. 24 patients with Double-blind, CBD 300 mg, CBD significantly
(2020) Parkinson’s placebo- single dose decreased simulated
disease controlled, public speaking test-
(simulated public randomized, induced anxiety as
speaking test) cross-over measured by the
study Visual Analog
Mood Scale
compared to placebo
Herrmann 39 adult patients Double-blind, Nabilone 1–2 mg Nabilone decreased
et al. (2019) with Alzheimer’s randomized, daily for 6 weeks agitation as
disease and placebo- measured by the
agitation controlled, Coben Mansfield
cross-over agitation inventory,
design Neuropsychiatric
Inventory – Nursing
home total score and
caregiver distress
score, and improved
standardized Mini-
Mental State
Examination scores
compared to placebo
Peball et al. 47 patients with Placebo- Nabilone 0.25 Nabilone
(2020) Parkinson’s controlled, once daily to significantly
disease with double-blind, 1 mg twice daily decreased
disturbing randomized treatment for symptoms
non-motor study 4 weeks compared to
symptoms, placebo as
including measured by the
sensory changes, Movement Disorder
sleep Society – Unified
disturbances, and PD Rating Scale – I
autonomic
dysfunction
van den 24 patients Randomized, THC 1.5 mg Neuropsychiatric
Elsen et al. diagnosed with double-blind, 3 times daily for symptoms as
(2015a) Alzheimer’s placebo- 3 weeks measured by the
disease/vascular controlled Neuropsychiatric
dementia/mixed study Inventory did not
dementia significantly change
with THC compared
to placebo
van den 22 adult patients Randomized, THC 0.75 mg THC did not reduce
Elsen et al. with dementia double-blind, twice daily for neuropsychiatric
(2015b) with placebo- 6 weeks symptoms as
neuropsychiatric controlled, measured by the
symptoms cross over Neuropsychiatric
design Inventory compared
to placebo
(continued)
Cannabinoid Drugs in Mental Health Disorders 1015
Table 1 (continued)
Reference Sample Study design Intervention Relevant findings
Volicer et al. 11 hospitalized Randomized, Dronabinol Dronabinol
(1997) patients in the placebo- 2.5 mg for significantly
dementia Study controlled, 6 weeks decreased severity
unit with cross over of disturbed
Alzheimer’s design behaviour as
disease who measured by Cohen-
were Mansfield Agitation
refusing food. Inventory compared
to placebo
Walther et al. 6 patients with Open-label, Dronabinol Dronabinol caused
(2006) late-stage not placebo- 2.5 mg treatment a significant
dementia controlled for 2 weeks reduction in
(Alzheimer’s or nocturnal motor
vascular) activity compared
suffering from to placebo as
circadian and measured by
behavioural actigraphy and
disturbances improvement in
Neuropsychiatric
Inventory total
score and agitation,
aberrant motor, and
night-time
behaviour subscales
Depression
Kotin et al. 8 adult inpatients Double-blind, THC at 0.3 mg/ THC did not show
(1973) with moderate to placebo- kg twice a day an antidepressant
severe controlled for 7 days effect measured
depression study using a 15-point
nurse-rated scale
compared to placebo
Other conditions
Cooper et al. 30 adult patients Double-blind, Sativex (1:1 Sativex spray did
(2017) with ADHD randomized, THC and CBD; not produce a
placebo- 2.7 mg and significant change in
controlled 2.5 mg per cognitive
study 100 microlitres, performance and
respectively) activity level
oromucosal measured using the
spray up to QbTest compared to
14 sprays per day placebo;
Sativex did produce
a significant
improvement in
hyperactivity/
impulsivity
measured with the
Conners’ Adult
ADHD rating scale
compared to placebo
(continued)
1016 S. Kloiber et al.
Table 1 (continued)
Reference Sample Study design Intervention Relevant findings
Grant et al. 14 female adult Open-label Dronabinol 2.5 – Dronabinol
(2011) patients with study, no 15 mg daily for significantly
trichotillomania control group 12 weeks decreased hair
pulling as
measured by the
Massachusetts
General Hospital
Hair Pulling Scale
compared to
baseline
Heussler 20 children and Open-label ZYN002 ZYN002
et al. (2019) adolescents clinical trial, (transdermal significantly
(6–17 yo) with no control CBD gel) decreased Anxiety,
fragile X group 50–250 mg twice Depression, and
syndrome daily for Mood Scale
12 weeks score, Aberrant
Behavior
Checklist –
Community for
FXS score,
Pediatric Quality of
Life Inventory
score, and Pediatric
Anxiety Rating
Scale score
compared to
baseline scores
a
Abbreviations: CBD (cannabinol), THC (Δ9 – tetrahydrocannabinol)
An emerging body of studies has also evaluated a potential role for cannabinoids in
the treatment of intrusive and arousal symptoms of PTSD (Walsh et al. 2017). Two
randomized, double-blind, and placebo-controlled clinical trials have examined the
effect of cannabinoids in PTSD (Jetly et al. 2015; Rabinak et al. 2020). The sample
sizes were small, with one study having 10 participants (Jetly et al. 2015) and the
other 19 (Rabinak et al. 2020). Both studies reported a significant effect of
cannabinoids, where nabilone ranging from 0.5 to 3 mg administered over
7 weeks (Jetly et al. 2015) was found to decrease PTSD symptoms and a single
dose of dronabinol 7.5 mg (Rabinak et al. 2020) using a behavioural/fMRI
paradigm resulted in the attenuation of amygdala activation and increase of medial
prefrontal cortex and adjacent rostral anterior cingulate cortex activation and
corticolimbic functional connectivity to threat compared to placebo, as well as
reduction of slowing of response to threatening vs. non-threatening cues. This
finding may reflect a normalization of behavioural responses to the threat that
accompanies the observed THC-related modulation of underlying corticolimbic
circuitry.
Cannabinoid Drugs in Mental Health Disorders 1017
Psychotic Disorders
Tourette Syndrome
Dementia
disorders (Bonaccorso et al. 2019; Aso et al. 2013; Iuvone et al. 2009). Synthetic
agonists at both CB1 and CB2 receptors have been demonstrated to reduce the
harmful effects of beta-amyloid peptide action in preclinical studies and prevent
memory deficits in beta-amyloid treated rodents (Aso and Ferrer 2014). Further-
more, CBD was reported to reduce tau protein hyperphosphorylation in a mouse
model of Alzheimer’s dementia and decrease the production of nitric oxide and
cytokines (Aso et al. 2013, 2016; Casarejos et al. 2013; Iuvone et al. 2009).
Our search identified seven clinical studies examining the effect of cannabinoids
on psychiatric symptoms in patients with dementia, including Alzheimer’s dementia,
vascular dementia, and Parkinson’s disease, where six studies were blinded, placebo-
controlled, and randomized (de Faria et al. 2020; Herrmann et al. 2019; Peball et al.
2020; van den Elsen et al. 2015a, b; Volicer et al. 1997) and one was an open-label
study (Walther et al. 2006). Sample sizes were small, ranging from 6 to 47 partici-
pants. Five studies reported positive findings, where nabilone 1–2 mg for 6 weeks
was found to decrease agitation (Herrmann et al. 2019), nabilone 0.25–1 mg for
4 weeks improved non-motor symptoms of Parkinson’s disease (Peball et al. 2020),
dronabinol 2.5 mg for 6 weeks or 2 weeks decreased severity of disturbed behaviour
and nocturnal motor activity, respectively (Volicer et al. 1997; Walther et al. 2006),
and CBD 300 mg decreased anxiety (de Faria et al. 2020). However, 2 studies
reported no effect of THC 1.5 or 4.5 mg for 3 or 6 weeks, respectively, in patients
with dementia (van den Elsen et al. 2015a, b).
Depression
Other Conditions
The effect of cannabinoids was also examined in adult patients with ADHD in a
double-blind placebo-controlled study, where nabiximols was not found to be
effective in decreasing symptoms of ADHD (Cooper et al. 2017). One study
examined the effect of dronabinol 2.5–15 mg administered over 12 weeks in patients
with trichotillomania, where it was found to be effective in decreasing hair pulling
(Grant et al. 2011). Lastly, one open-label clinical trial examined the effect of a
transdermal CBD gel (ZYN002) administered daily for 12 weeks on children and
adolescents with fragile X syndrome, where ZYN002 was found to significantly
improve mood and anxiety (Heussler et al. 2019).
Given the important role of the ECS in mediating reward and learning, there has been
considerable interest in investigating cannabinoids as potential treatments for SUDs
(Chye et al. 2019; Sloan et al. 2017). It is important to note that pharmacotherapeutic
strategies for treating SUDs with ECS-targeting drugs are likely to be substance-
specific. For example, CB1 receptor agonists may be useful as a substitution strategy
to treat cannabis use disorder (CUD), while CB1 receptor antagonists may be useful
to mitigate some of the reinforcing effects of other psychoactive drugs (Sloan et al.
2017; Le Foll and Goldberg 2005). The following section will review the clinical
evidence for the role of cannabinoid drugs in reducing symptoms of SUDs, including
CUD, nicotine dependence, opioid use disorder (OUD), and alcohol use disorder
(AUD). All studies administered either CBD or THC (alone, or in combination, i.e.,
nabiximols), nabilone (the synthetic analogue of THC), or a CB1 receptor antagonist/
inverse agonist (rimonabant, taranabant, or surinabant). See Table 2 for an overview
of included clinical studies that met our eligibility criteria.
Table 2 Clinical trials assessing the effects of cannabinoid drugs on substance-related outcomes in
patients with substance use disorders
Reference Sample Study design Intervention Relevant findings
Cannabis
Allsop 51 participants Two-site Six-day Nabiximols significantly
et al. (2014) with DSM-IV parallel- treatment with reduced cannabis
cannabis groups, nabiximols withdrawal scores
dependence double-blind, (maximum compared to placebo
randomized, daily dose No significant difference
inpatient trial 86.4 mg in reductions in cannabis
THC/80 mg use at follow-up
CBD) (n ¼ 27)
or placebo
(n ¼ 24) with
standardized
psychosocial
intervention
Freeman 82 participants Phase 2a, Four weeks of In the first treatment
et al. (2020) with DSM-5 randomised, treatment in phase, CBD 200 mg was
cannabis use double-blind, two phases: eliminated as an
disorder placebo (I) placebo ineffective dose
(at least controlled, (n ¼ 12), oral At the final analysis, both
moderate parallel-group CBD 200 mg/ CBD 400 mg and 800 mg
severity) trial day (n ¼ 12), exceeded primary
CBD 400 mg endpoint criteria (i.e. had
(n ¼ 12), CBD a probability of greater
800 mg than 0.9 of being superior
(n ¼ 12); to placebo) for both
(II) placebo urinary THC-COOH-
(n ¼ 11), CBD creatinine ratio and days
400 mg of abstinence from
(n ¼ 12), CBD cannabis
800 mg
(n ¼ 11) with
motivational
interviewing
Hill et al. 12 participants Parallel- Ten weeks of No difference between
(2017) with DSM-IV groups, treatment with nabilone and placebo in
cannabis double-blind, oral nabilone cannabis-related
dependence randomized (maximum outcomes (as measured
trial 2 mg/day) by Timeline Follow Back
(n ¼ 6) or and daily diaries)
placebo
(n ¼ 6) with
medical
management
behavioural
intervention
Levin et al. 156 participants Randomized, Twelve weeks No difference between
(2011) with DSM-IV- double-blind, of treatment dronabinol and placebo in
with oral 2-week continuous
(continued)
Cannabinoid Drugs in Mental Health Disorders 1021
Table 2 (continued)
Reference Sample Study design Intervention Relevant findings
TR cannabis parallel-group dronabinol abstinence from cannabis.
dependence trial (maximum Greater reduction in
40 mg/day) withdrawal symptoms in
(n ¼ 79) or dronabinol
placebo group vs. placebo
(n ¼ 77) with
psychosocial
intervention
(coping skills
plus
motivational
enhancement
therapy)
Levin et al. 122 participants Randomized, Eleven weeks No difference between
(2016) with DSM-IV- double-blind, of treatment dronabinol-lofexidine
TR cannabis parallel-group with oral and placebo in achieving
dependence trial dronabinol abstinence from cannabis
(maximum No group difference in
60 mg/day) withdrawal symptoms
and lofexidine
(maximum
1.8 mg/day)
(n ¼ 61) or
placebo
(n ¼ 61) plus
motivational
enhancement
and cognitive
behavioural/
relapse
prevention
therapy
Lintzeris 128 participants Phase Twelve weeks Outcomes during
et al. (2019), with ICD-10 3, multisite, of treatment treatment (Lintzeris et al.
Lintzeris cannabis outpatient with 2019): nabiximols group
et al. (2020) dependence randomized, nabiximols reported significantly
double-blind, (maximum fewer days of cannabis
Parallel-design daily dose use over the course of the
trial 86.4 mg 12 weeks of treatment
THC/80 mg compared to the placebo
CBD) (n ¼ 61) group
or placebo Three-month follow up
(n ¼ 67) plus (Lintzeris et al. 2020): the
CBT nabiximols groups
psychosocial reported significantly
intervention fewer days of cannabis
and had a higher
proportion of abstinent
participants in the 28 days
prior to the 3-month
(continued)
1022 S. Kloiber et al.
Table 2 (continued)
Reference Sample Study design Intervention Relevant findings
follow up, as compared to
placebo
Lundahl and 14 participants Placebo- Three single- Both doses of dronabinol
Greenwald with DSM-IV controlled, dose (10 mg and
(2015) cannabis within- experimental 20 mg vs. placebo)
dependence subjects, sessions: attenuated cannabis-
crossover trial Placebo, 10 mg induced increases in
oral craving and anxiety
dronabinol,
20 mg
dronabinol
Trigo et al. 4 participants Open-label Twelve weeks Decrease in cannabis use
(2016b) with DSM-IV pilot trial of treatment over the course of
cannabis with treatment, with
dependence nabiximols accompanying decrease
(maximum in craving and
daily dose withdrawal
113.4 mg
THC/105 mg
CBD) plus
motivational
enhancement
therapy and
cognitive
behavioural
therapy
Trigo et al. 9 participants Double-blind, Five days of Nabiximols (compared to
(2016a) with current placebo- treatment with placebo) significantly
DSM-IV controlled, self-titration of reduced withdrawal, but
cannabis crossover trial placebo, fixed not craving, during
dependence (ABACADAE dose of abstinence
study design, placebo, self-
where B-E are titration of
abstinence nabiximols
phases with (up to
medication maximum
intervention 108 mg
and A is THC/100 mg
smoking as CBD), or fixed
usual phase) dose of
nabiximols
(108 mg
THC/100 mg
CBD) or
placebo
(medication
phases B-D),
each followed
by smoking as
usual period
(A)
(continued)
Cannabinoid Drugs in Mental Health Disorders 1023
Table 2 (continued)
Reference Sample Study design Intervention Relevant findings
Trigo et al. 27 participants Double-blind, Twelve weeks No significant difference
(2018) with current placebo- of treatment in abstinence rates
DSM-IV controlled, with between nabiximols and
cannabis parallel-group nabiximols placebo groups
dependence randomized (maximum Significant group by time
trial daily dose interaction for craving
113.4 mg (greater reduction in
THC/105 mg craving in nabiximols
CBD) (n ¼ 13) group vs. placebo at week
or placebo 7), but no effect on
(n ¼ 14) with withdrawal
motivational
enhancement
therapy and
cognitive
behavioural
therapy
Nicotine
Morgan et al. 24 daily Double-blind, One week of Significant reduction in
(2013) tobacco placebo- treatment with the number of cigarettes
smokers controlled, a CBD inhaler smoked in the CBD
intending to parallel-group (participants group, but not in the
quit randomized instructed to placebo group
trial use the inhaler No effect of CBD on
when they had nicotine craving
an urge to
smoke;
estimated
400μg CBD
per use)
(n ¼ 12) or
placebo
(n ¼ 12)
Morrison 317 daily Multicentre Eight weeks of No significant difference
et al. (2010) tobacco (11-site), treatment with (taranabant vs. placebo)
smokers who double-blind, oral taranabant in continuous abstinence
had at least randomized, (CB1 receptor
three previous placebo- inverse
quit attempts controlled, agonist;
outpatient trial maximum
daily dose
8 mg)
(n ¼ 159) or
placebo
(n ¼ 158) with
brief,
individualized
counselling
intervention
(continued)
1024 S. Kloiber et al.
Table 2 (continued)
Reference Sample Study design Intervention Relevant findings
Rigotti 735 daily Multicentre Rimonabant Significantly greater odds
et al. (2009) tobacco (15-site) (20 mg daily, of achieving smoking
smokers with randomized, oral) given cessation in the nicotine
intent to quit double-blind, open-label for patch + rimonabant
placebo- 9 weeks; compared to placebo
controlled participants patch + rimonabant
parallel-group randomized to
trial nicotine patch
(n ¼ 369) or
placebo patch
(n ¼ 366) plus
weekly
cognitive-
behavioural
smoking
counselling
Robinson 2097 daily Pooled Ten weeks of Continuous abstinence
et al. (2018) tobacco analysis of treatment with rates were significantly
smokers with three oral greater in the 20 mg (but
intent to quit multicentre, rimonabant not 5 mg) rimonabant
randomized, (5 mg group vs. placebo (when
double-blind, (n ¼ 518) or assessed at end of
placebo- 20 mg treatment, and at 48-week
controlled (n ¼ 790) follow-up)
trials daily) or
placebo
(n ¼ 789) plus
behavioural
counselling
Tonstad and 810 daily Randomized, Eight weeks of No difference in
Aubin (2012) tobacco multicentre treatment with continuous abstinence
smokers (35 site), oral surinabant rates between surinabant
motivated to double-blind, (CB1 receptor (at any dose tested) and
quit 4-arm parallel- antagonist) at placebo
group placebo- 2.5 mg/day
controlled trial (n ¼ 199),
5 mg/day
(n ¼ 204),
10 mg/day
(n ¼ 205), or
placebo
(n ¼ 202) plus
brief smoking
cessation
counselling
Opioids
Bisaga et al. 60 treatment- Randomized, Treatment with Dronabinol (vs. placebo)
(2015) seeking double-blind, oral dronabinol reduced withdrawal
participants placebo- (maximum during the initial acute
with current controlled daily dose inpatient phase, prior to
30 mg) naltrexone induction
(continued)
Cannabinoid Drugs in Mental Health Disorders 1025
Table 2 (continued)
Reference Sample Study design Intervention Relevant findings
DSM-IV opioid parallel-group (n ¼ 40) or No difference between
dependence trial placebo dronabinol and placebo
(n ¼ 20) groups in withdrawal
during symptoms during the
inpatient outpatient phase or
detoxification proportion of patients
and naltrexone inducted onto and
induction maintained on injection
(8 days total) naltrexone
and then
subsequent
5 weeks of
outpatient care
Hurd et al. 42 participants Double-blind, Three days of Both doses of CBD
(2019) with current placebo- treatment with reduced heroin
DSM-IV opioid controlled, oral CBD at cue-induced craving and
dependence, randomized, 400 mg anxiety compared to
recently parallel-group (n ¼ 14) and placebo, both when
abstinent from trial 800 mg assessed acutely (i.e.,
heroin (n ¼ 13) or after the first dose of
placebo CBD/placebo during the
(n ¼ 15) first experimental
session) and when
assessed 1 week after the
final CBD exposure
Alcohol
Soyka et al. 258 participants Phase 2a Twelve weeks No significant difference
(2008) with current (proof-of- of treatment (rimonabant vs. placebo)
DSM-IV concept), with oral in time to first drink or
alcohol double-blind, rimonabant heavy drinking
dependence placebo- (20 mg/day)
who had controlled (n ¼ 131) or
recently parallel-group placebo
detoxified from trial (n ¼ 127)
alcohol
a
Abbreviations: CBD (cannabinol), THC (Δ9 – tetrahydrocannabinol)
human laboratory studies have found that oral THC or its synthetic analogue
nabilone reduce cannabis withdrawal when compared to placebo (Panlilio et al.
2016). CBD has also shown promise in reducing cannabis-related outcomes (e.g.,
self-administration, withdrawal) in animal models and in humans (Chye et al. 2019).
Combining THC and CBD (e.g., nabiximols) has received particular attention given
the extant evidence from human studies that CBD can at least partly attenuate the
acute effects of THC (Freeman et al. 2019).
Eleven studies that met our eligibility criteria assessed the impact of cannabinoid
drugs on cannabis-related outcomes. All included studies administered either oral
1026 S. Kloiber et al.
CBD alone (Freeman et al. 2020), oral THC (i.e., dronabinol) alone (Levin et al.
2011, 2016; Lundahl and Greenwald 2015), nabiximols (Allsop et al. 2014; Trigo
et al. 2016a, b, 2018; Lintzeris et al. 2019, 2020), or oral nabilone (Hill et al. 2017).
Freeman et al. (2020) conducted a phase 2a, double-blind, placebo-controlled
trial that randomized 82 participants with DSM-5 CUD (at least moderate severity)
to receive 4 weeks of treatment with oral CBD (200, 400, or 800 mg/day) or placebo
in two phases. In the first phase, CBD 200 mg was eliminated as an ineffective dose.
In the second phase, at the final analysis, both CBD 400 mg and 800 mg exceeded
primary endpoint criteria (i.e., had a probability greater than 0.9 of being superior to
placebo) for both urinary cannabinoid concentrations and days of abstinence from
cannabis (Freeman et al. 2020).
Levin and colleagues conducted two RCTs evaluating the efficacy of dronabinol
in promoting abstinence from cannabis use. In the first trial, 156 participants with
DSM-IV-TR cannabis dependence were randomized to 12 weeks of treatment with
oral dronabinol (maximum 40 mg/day, n ¼ 79) or placebo (n ¼ 77). There was no
difference between dronabinol and placebo in 2-week continuous abstinence from
cannabis, but there was a greater effect of dronabinol on withdrawal symptoms when
compared to placebo (Levin et al. 2011). In a follow-up trial, 122 participants with
DSM-IV-TR cannabis dependence were randomized to 11 weeks of treatment with
combined oral dronabinol (maximum 60 mg/day) and lofexidine (an α2A adrenergic
receptor agonist; maximum 1.8 mg/day) (n ¼ 61) or placebo (n ¼ 61). There was no
difference in continuous abstinence or withdrawal symptoms between treatment
groups; however, dronabinol was not administered alone (only in combination
with lofexidine) (Levin et al. 2016). Lundahl and Greenwald (2015) conducted a
single-dose crossover trial with 14 participants meeting DSM-IV criteria for canna-
bis dependence. Dronabinol (at 10 mg and 20 mg) significantly attenuated cannabis-
induced increases in craving and anxiety (compared to placebo) (Lundahl and
Greenwald 2015).
Six clinical studies have found evidence that nabiximols may be an effective
treatment for cannabis dependence. Allsop et al. (2014) randomized 51 participants
with DSM-IV cannabis dependence to receive 6 days of treatment with nabiximols
(maximum daily dose 86.4 mg THC/80 mg CBD, n ¼ 27) or placebo (n ¼ 24) and
found that nabiximols significantly reduced cannabis withdrawal compared to pla-
cebo, though there was no difference in cannabis use at 28-day follow up. Trigo and
colleagues provided additional clinical evidence suggesting nabiximols as an effec-
tive therapy for CUD. In an open-label pilot study, they found that 12 weeks of
treatment with nabiximols (maximum daily dose 113.4 mg THC/105 mg CBD) led
to a significant decrease in cannabis use, craving, and withdrawal (Trigo et al.
2016b). Trigo and colleagues also conducted a crossover trial in nine participants
with current DSM-IV cannabis dependence, who underwent 5 days of treatment with
self-titration of a placebo, self-titration of nabiximols, fixed dose of placebo, and
fixed dose of nabiximols (108 mg THC/100 mg CBD) during a period of abstinence,
each followed by smoking as usual period. This study found that nabiximols
significantly reduced craving but not withdrawal, as compared to placebo (Trigo
et al. 2016a). Finally, the authors conducted a double-blind, placebo-controlled trial
Cannabinoid Drugs in Mental Health Disorders 1027
Nicotine Dependence
Converging lines of evidence from preclinical and human studies have strongly
implicated a role of the CB1 receptor in nicotine reward and reinforcement (Gamaleddin
et al. 2015). In animal models, rimonabant (a CB1 receptor antagonist/inverse agonist)
has been shown to decrease nicotine self-administration, block the development of
nicotine-induced conditioned place preference, and block reinstatement of previously
extinguished nicotine-seeking behaviour (Gamaleddin et al. 2015; Le Foll and Gold-
berg 2004). Rimonabant emerged as a promising candidate for smoking cessation
pharmacotherapy based on converging evidence from animal models and human trials,
yet significant psychiatric adverse effects ultimately led to rimonabant being withdrawn
from the market (Sloan et al. 2017). Similarly, the development of two other CB1
receptor antagonists (surinabant and taranabant) was halted due to the risk of serious
adverse effects. CBD has been another promising candidate for a smoking cessation
drug, in large part due to its non-competitive antagonism of CB1 receptor signalling,
likely through negative allosteric modulation, which would mean it lacks the adverse
effects caused by inverse agonists such as rimonabant (Sloan et al. 2017; Chye et al.
2019). Human laboratory evidence has suggested that CBD can reduce attentional bias
to tobacco cues (e.g., Hindocha et al. 2018).
Five studies that met our eligibility criteria assessed the impact of cannabinoid
drugs on nicotine-related outcomes. One study administered inhaled CBD (Morgan
et al. 2013), while the remaining four administered one of three different CB1 receptor
1028 S. Kloiber et al.
induction, but did not have sustained effects during the subsequent 5 weeks of
outpatient care (Bisaga et al. 2015). Hurd et al. (2019) recruited 42 participants
with current DSM-IV opioid dependence who were recently abstinent from heroin
and randomized them to receive three consecutive days of treatment with oral CBD
400 mg (n ¼ 14), CBD 800 mg (n ¼ 13), or placebo (n ¼ 15). The authors found that
both doses of CBD effectively reduced cue-induced craving and anxiety compared
to placebo, both when measured after the first dose of CBD/placebo and when
assessed 1 week after the final day of CBD/placebo treatment (Hurd et al. 2019).
The bidirectional relationship between alcohol exposure and the ECS has been
extensively characterized in animal models, with studies showing that alcohol intake
leads to changes in CB1 receptors and endocannabinoid levels and that infusion of
CB1 receptor agonists can enhance alcohol self-administration (Basavarajappa et al.
2019). Preclinical evidence has also suggested that CB1 receptor antagonism can
reduce alcohol self-administration (Sloan et al. 2017). Only one relevant clinical trial
was identified for the present review. A single phase 2a (proof-of-concept) trial
found no significant evidence of an effect of rimonabant (20 mg daily for 12 weeks)
compared to placebo on time to first drink or time to first heavy drinking in a sample
of 258 participants with DSM-IV alcohol dependence who had recently undergone
detoxification from alcohol (Soyka et al. 2008). Emerging evidence has also
supported the role of CBD in reducing alcohol-related outcomes (Nona et al.
2019), though no published clinical trials were identified for inclusion in this chapter.
The evidence reviewed supports the role of cannabinoid drugs in treating SUDs,
though most findings have yet to be replicated in large trials. The included clinical
studies seem to support a role for CBD in reducing cannabis use (Freeman et al. 2020),
promoting tobacco smoking cessation (Morgan et al. 2013), and attenuating opioid
craving (Hurd et al. 2019). Cannabinoid substitution therapy (with dronabinol or
nabiximols) shows promise as a strategy to reduce cannabis craving and/or withdrawal
in patients with CUD (Allsop et al. 2014; Levin et al. 2011, 2016; Trigo et al. 2016a,
2018), and the most recent evidence suggests that nabiximols has potential to reduce
cannabis use and promote abstinence (Lintzeris et al. 2019, 2020). Preliminary
evidence suggests that dronabinol may also be useful for reducing opioid withdrawal
(Bisaga et al. 2015), though this finding also needs replication in a larger RCT.
A pooled analysis of over 2000 participants of smoking cessation trials found
compelling evidence that rimonabant is effective in promoting abstinence from
tobacco smoking (Robinson et al. 2018), yet rimonabant has been withdrawn due
to its serious psychiatric adverse effects, including depressive symptoms and
Cannabinoid Drugs in Mental Health Disorders 1031
suicidality (Sloan et al. 2017). So far, the use of other CB1 receptor antagonists – i.e.,
surinabant (Tonstad and Aubin 2012) and taranabant (Morrison et al. 2010) – have
not shown efficacy in promoting smoking cessation in clinical trials, and develop-
ment of these drugs ceased after the withdrawal of rimonabant. Given the success of
rimonabant, there has been interest in developing alternative strategies for antago-
nizing the CB1 receptor without the serious psychiatric adverse effects. Such strat-
egies include peripherally restricted CB1 antagonists (e.g., JD5037), neutral CB1
receptor antagonists that lack the inverse agonism of rimonabant (e.g., AM4113),
and negative allosteric modulators of CB1 receptors (e.g., Pepcan-12) (Galaj and Xi
2019). So far, only the neutral CB1 receptor antagonist AM4113 has demonstrated
the potential to reduce nicotine self-administration, yet all three types of CB1
receptor-targeting compounds have demonstrated the potential to reduce alcohol
intake, suggesting that these alternative CB1 receptor antagonism strategies may
have the potential for AUD (Galaj and Xi 2019).
Another strategy for modulating the ECS to treat SUDs and mental health conditions
that has received increasing attention is the use of enzyme inhibitors such as FAAH or
MAGL inhibitors (Galaj and Xi 2019; Sloan et al. 2017). Inhibiting FAAH increases
levels of AEA, which has shown promise in preclinical studies to modulate nicotine and
alcohol intake and in preliminary human studies to reduce cannabis use and withdrawal
in CUD (Galaj and Xi 2019). Similarly, the inhibition of MAGL leads to increased
levels of 2-AG, though currently, there is less evidence suggesting that MAGL
inhibitors can reduce drug-related behaviours (Galaj and Xi 2019). For example, in
rodents, FAAH inhibition was shown to attenuate cue-induced reinstatement of
nicotine-seeking (Forget et al. 2016), while MAGL inhibition actually enhanced
cue-induced reinstatement (Trigo and Le Foll 2016), so more work is clearly needed
to delineate the effect of MAGL inhibition. Interestingly, dual inhibition of FAAH and
MAGL with the compound SA-57 has been shown to reduce heroin self-administration
in rodents (Wilkerson et al. 2017). More work is clearly needed to evaluate the potential
of FAAH and MAGL inhibitors to reduce drug self-administration and reinforcement,
but this remains a promising avenue for future human trials. Preclinical studies of
FAAH or MAGL inhibitors using stress paradigms indicated potential effects on
depression- and anxiety-like behaviours, as well as social behaviours, potentially
mediated through the restoration of AEA and brain-derived neurotrophic factor
(BDNF) (Alteba et al. 2020; Dong et al. 2020; Carnevali et al. 2020). A recently
published clinical trial of an FAAH inhibitor in social anxiety disorder (Schmidt et al.
2021) showed a negative result of the primary outcome, though the authors indicated a
small to modest anxiolytic effect in severe social anxiety disorder and that measured
FAAH and drug levels may warrant exploration of higher doses.
Conclusion
Cross-References
References
Allsop DJ, Copeland J, Lintzeris N, Dunlop AJ, Montebello M, Sadler C, Rivas GR, Holland RM,
Muhleisen P, Norberg MM, Booth J, McGregor IS. Nabiximols as an agonist replacement
therapy during cannabis withdrawal: a randomized clinical trial. JAMA Psychiat. 2014;71:281–
91.
Allsop DJ, Lintzeris N, Copeland J, Dunlop A, McGregor IS. Cannabinoid replacement therapy
(CRT): Nabiximols (Sativex) as a novel treatment for cannabis withdrawal. Clin Pharmacol
Ther. 2015;97:571–4.
Alteba S, Mizrachi Zer-Aviv T, Tenenhaus A, Ben David G, Adelman J, Hillard CJ, Doron R,
Akirav I. Antidepressant-like effects of URB597 and JZL184 in male and female rats exposed to
early life stress. Eur Neuropsychopharmacol. 2020;39:70–86.
Aso E, Ferrer I. Cannabinoids for treatment of Alzheimer’s disease: moving toward the clinic. Front
Pharmacol. 2014;5:37.
Aso E, Juves S, Maldonado R, Ferrer I. CB2 cannabinoid receptor agonist ameliorates Alzheimer-
like phenotype in AbetaPP/PS1 mice. J Alzheimers Dis. 2013;35:847–58.
Aso E, Andres-Benito P, Ferrer I. Delineating the efficacy of a Cannabis-based medicine at
advanced stages of dementia in a murine model. J Alzheimers Dis. 2016;54:903–12.
Basavarajappa BS, Joshi V, Shivakumar M, Subbanna S. Distinct functions of endogenous canna-
binoid system in alcohol abuse disorders. Br J Pharmacol. 2019;176:3085–109.
Bergamaschi MM, Queiroz RH, Chagas MH, de Oliveira DC, De Martinis BS, Kapczinski F,
Quevedo J, Roesler R, Schroder N, Nardi AE, Martin-Santos R, Hallak JE, Zuardi AW, Crippa
JA. Cannabidiol reduces the anxiety induced by simulated public speaking in treatment-naive
social phobia patients. Neuropsychopharmacology. 2011;36:1219–26.
Cannabinoid Drugs in Mental Health Disorders 1033
Bisaga A, Sullivan MA, Glass A, Mishlen K, Pavlicova M, Haney M, Raby WN, Levin FR, Carpenter
KM, Mariani JJ, Nunes EV. The effects of dronabinol during detoxification and the initiation of
treatment with extended release naltrexone. Drug Alcohol Depend. 2015;154:38–45.
Black N, Stockings E, Campbell G, Tran LT, Zagic D, Hall WD, Farrell M, Degenhardt
L. Cannabinoids for the treatment of mental disorders and symptoms of mental disorders: a
systematic review and meta-analysis. Lancet Psychiatry. 2019;6:995–1010.
Blessing EM, Steenkamp MM, Manzanares J, Marmar CR. Cannabidiol as a potential treatment for
anxiety disorders. Neurotherapeutics. 2015;12:825–36.
Boggs DL, Kelly DL, McMahon RP, Gold JM, Gorelick DA, Linthicum J, Conley RR, Liu F,
Waltz J, Huestis MA, Buchanan RW. Rimonabant for neurocognition in schizophrenia: a
16-week double blind randomized placebo controlled trial. Schizophr Res. 2012;134:207–10.
Boggs DL, Surti T, Gupta A, Gupta S, Niciu M, Pittman B, Schnakenberg Martin AM,
Thurnauer H, Davies A, D’Souza DC, Ranganathan M. The effects of cannabidiol (CBD) on
cognition and symptoms in outpatients with chronic schizophrenia a randomized placebo
controlled trial. Psychopharmacology. 2018;235:1923–32.
Bonaccorso S, Ricciardi A, Zangani C, Chiappini S, Schifano F. Cannabidiol (CBD) use in
psychiatric disorders: a systematic review. Neurotoxicology. 2019;74:282–98.
Bonn-Miller MO, Boden MT, Bucossi MM, Babson KA. Self-reported cannabis use characteristics,
patterns and helpfulness among medical cannabis users. Am J Drug Alcohol Abuse. 2014;40:
23–30.
Botsford SL, Yang S, George TP. Cannabis and cannabinoids in mood and anxiety disorders: impact
on illness onset and course, and assessment of therapeutic potential. Am J Addict. 2020;29:9–26.
Campos AC, Ortega Z, Palazuelos J, Fogaca MV, Aguiar DC, Diaz-Alonso J, Ortega-Gutierrez S,
Vazquez-Villa H, Moreira FA, Guzman M, Galve-Roperh I, Guimaraes FS. The anxiolytic effect
of cannabidiol on chronically stressed mice depends on hippocampal neurogenesis: involvement
of the endocannabinoid system. Int J Neuropsychopharmacol. 2013;16:1407–19.
Carnevali L, Statello R, Vacondio F, Ferlenghi F, Spadoni G, Rivara S, Mor M, Sgoifo
A. Antidepressant-like effects of pharmacological inhibition of FAAH activity in socially
isolated female rats. Eur Neuropsychopharmacol. 2020;32:77–87.
Casarejos MJ, Perucho J, Gomez A, Munoz MP, Fernandez-Estevez M, Sagredo O, Fernandez
Ruiz J, Guzman M, de Yebenes JG, Mena MA. Natural cannabinoids improve dopamine
neurotransmission and tau and amyloid pathology in a mouse model of tauopathy. J Alzheimers
Dis. 2013;35:525–39.
Chye Y, Christensen E, Solowij N, Yücel M. The endocannabinoid system and Cannabidiol’s
promise for the treatment of substance use disorder. Front Psych. 2019;10:63.
Connor JP, Stjepanović D, Le Foll B, Hoch E, Budney AJ, Hall WD. Cannabis use and cannabis use
disorder. Nat Rev Dis Primers. 2021;7:16.
Cooper RE, Williams E, Seegobin S, Tye C, Kuntsi J, Asherson P. Cannabinoids in attention-deficit/
hyperactivity disorder: a randomised-controlled trial. Eur Neuropsychopharmacol. 2017;27:
795–808.
Crippa JA, Zuardi AW, Martín-Santos R, Bhattacharyya S, Atakan Z, McGuire P, Fusar-Poli
P. Cannabis and anxiety: a critical review of the evidence. Hum Psychopharmacol. 2009;24:
515–23.
Crippa JA, Derenusson GN, Ferrari TB, Wichert-Ana L, Duran FL, Martin-Santos R, Simoes MV,
Bhattacharyya S, Fusar-Poli P, Atakan Z, Santos Filho A, Freitas-Ferrari MC, McGuire PK,
Zuardi AW, Busatto GF, Hallak JE. Neural basis of anxiolytic effects of cannabidiol (CBD) in
generalized social anxiety disorder: a preliminary report. J Psychopharmacol. 2011;25:121–30.
D’Souza DC, Perry E, MacDougall L, Ammerman Y, Cooper T, Wu YT, Braley G, Gueorguieva R,
Krystal JH. The psychotomimetic effects of intravenous delta-9-tetrahydrocannabinol in healthy
individuals: implications for psychosis. Neuropsychopharmacology. 2004;29:1558–72.
de Faria SM, de Morais Fabricio D, Tumas V, Castro PC, Ponti MA, Hallak JE, Zuardi AW, Crippa
JAS, Chagas MHN. Effects of acute cannabidiol administration on anxiety and tremors induced
by a simulated public speaking test in patients with Parkinson’s disease. J Psychopharmacol.
2020;34:189–96.
1034 S. Kloiber et al.
Dong B, Shilpa BM, Shah R, Goyal A, Xie S, Bakalian MJ, Suckow RF, Cooper TB, Mann JJ,
Arango V, Vinod KY. Dual pharmacological inhibitor of endocannabinoid degrading enzymes
reduces depressive-like behavior in female rats. J Psychiatr Res. 2020;120:103–12.
Elphick MR, Egertova M. The phylogenetic distribution and evolutionary origins of endo-
cannabinoid signalling. Handb Exp Pharmacol. 2005;168:283–97.
Fabre LF, McLendon D. The efficacy and safety of nabilone (a synthetic cannabinoid) in the
treatment of anxiety. J Clin Pharmacol. 1981;21:377S–82S.
Feingold D, Weinstein A. Cannabis and depression. Adv Exp Med Biol. 2021;1264:67–80.
Forget B, Guranda M, Gamaleddin I, Goldberg SR, Le Foll B. Attenuation of cue-induced
reinstatement of nicotine seeking by URB597 through cannabinoid CB1 receptor in rats.
Psychopharmacology. 2016;233:1823–8.
Freeman AM, Petrilli K, Lees R, Hindocha C, Mokrysz C, Curran HV, Saunders R, Freeman
TP. How does cannabidiol (CBD) influence the acute effects of delta-9-tetrahydrocannabinol
(THC) in humans? A systematic review. Neurosci Biobehav Rev. 2019;107:696–712.
Freeman TP, Hindocha C, Baio G, Shaban NDC, Thomas EM, Astbury D, Freeman AM, Lees R,
Craft S, Morrison PD, Bloomfield MAP, O’Ryan D, Kinghorn J, Morgan CJA, Mofeez A,
Curran HV. Cannabidiol for the treatment of cannabis use disorder: a phase 2a, double-blind,
placebo-controlled, randomised, adaptive Bayesian trial. Lancet Psychiatry. 2020;7:865–74.
Galaj E, Xi ZX. Potential of cannabinoid receptor ligands as treatment for substance use disorders.
CNS Drugs. 2019;33:1001–30.
Gamaleddin IH, Trigo JM, Gueye AB, Zvonok A, Makriyannis A, Goldberg SR, Le Foll B. Role of
the endogenous cannabinoid system in nicotine addiction: novel insights. Front Psych.
2015;6:41.
Gomes FV, Llorente R, Del Bel EA, Viveros MP, Lopez-Gallardo M, Guimaraes FS. Decreased
glial reactivity could be involved in the antipsychotic-like effect of cannabidiol. Schizophr Res.
2015;164:155–63.
Grant JE, Odlaug BL, Chamberlain SR, Kim SW. Dronabinol, a cannabinoid agonist, reduces hair
pulling in trichotillomania: a pilot study. Psychopharmacology. 2011;218:493–502.
Gueye AB, Pryslawsky Y, Trigo JM, Poulia N, Delis F, Antoniou K, Loureiro M, Laviolette SR,
Vemuri K, Makriyannis A, Le Foll B. The CB1 neutral antagonist AM4113 retains the
therapeutic efficacy of the inverse agonist Rimonabant for nicotine dependence and weight
loss with better psychiatric tolerability. Int J Neuropsychopharmacol. 2016;19
Hahn B. The potential of Cannabidiol treatment for Cannabis users with recent-onset psychosis.
Schizophr Bull. 2018;44:46–53.
Herrmann N, Ruthirakuhan M, Gallagher D, Verhoeff N, Kiss A, Black SE, Lanctot
KL. Randomized placebo-controlled trial of Nabilone for agitation in Alzheimer’s disease.
Am J Geriatr Psychiatry. 2019;27:1161–73.
Heussler H, Cohen J, Silove N, Tich N, Bonn-Miller MO, Du W, O’Neill C, Sebree T. A phase 1/2,
open-label assessment of the safety, tolerability, and efficacy of transdermal cannabidiol
(ZYN002) for the treatment of pediatric fragile X syndrome. J Neurodev Disord. 2019;11:16.
Hill AJ, Williams CM, Whalley BJ, Stephens GJ. Phytocannabinoids as novel therapeutic agents in
CNS disorders. Pharmacol Ther. 2012;133:79–97.
Hill KP, Palastro MD, Gruber SA, Fitzmaurice GM, Greenfield SF, Lukas SE, Weiss RD. Nabilone
pharmacotherapy for cannabis dependence: a randomized, controlled pilot study. Am J Addict.
2017;26:795–801.
Hindocha C, Freeman TP, Grabski M, Stroud JB, Crudgington H, Davies AC, Das RK, Lawn W,
Morgan CJA, Curran HV. Cannabidiol reverses attentional bias to cigarette cues in a human
experimental model of tobacco withdrawal. Addiction. 2018;113:1696–705.
Howlett AC, Barth F, Bonner TI, Cabral G, Casellas P, Devane WA, Felder CC, Herkenham M,
Mackie K, Martin BR, Mechoulam R, Pertwee RG. International Union of Pharmacology.
XXVII. Classification of cannabinoid receptors. Pharmacol Rev. 2002;54:161–202.
Hu SS-J, Mackie K. Distribution of the endocannabinoid system in the central nervous system. In:
Pertwee RG, editor. Endocannabinoids. Cham: Springer International Publishing; 2015.
Cannabinoid Drugs in Mental Health Disorders 1035
Hurd YL, Yoon M, Manini AF, Hernandez S, Olmedo R, Ostman M, Jutras-Aswad D. Early phase
in the development of Cannabidiol as a treatment for addiction: opioid relapse takes initial center
stage. Neurotherapeutics. 2015;12:807–15.
Hurd YL, Spriggs S, Alishayev J, Winkel G, Gurgov K, Kudrich C, Oprescu AM, Salsitz
E. Cannabidiol for the reduction of Cue-induced craving and anxiety in drug-abstinent individ-
uals with heroin use disorder: a double-blind randomized placebo-controlled trial. Am J
Psychiatry. 2019;176:911–22.
Ibarra-Lecue I, Pilar-Cuéllar F, Muguruza C, Florensa-Zanuy E, Díaz Á, Urigüen L, Castro E,
Pazos A, Callado LF. The endocannabinoid system in mental disorders: evidence from human
brain studies. Biochem Pharmacol. 2018;157:97–107.
Iuvone T, Esposito G, De Filippis D, Scuderi C, Steardo L. Cannabidiol: a promising drug for
neurodegenerative disorders? CNS Neurosci Ther. 2009;15:65–75.
Jenny M, Schrocksnadel S, Uberall F, Fuchs D. The potential role of cannabinoids in modulating
serotonergic signaling by their influence on tryptophan metabolism. Pharmaceuticals (Basel).
2010;3:2647–60.
Jetly R, Heber A, Fraser G, Boisvert D. The efficacy of nabilone, a synthetic cannabinoid, in the
treatment of PTSD-associated nightmares: a preliminary randomized, double-blind, placebo-
controlled cross-over design study. Psychoneuroendocrinology. 2015;51:585–8.
Karniol IG, Shirakawa I, Kasinski N, Pfeferman A, Carlini EA. Cannabidiol interferes with the
effects of delta 9 - tetrahydrocannabinol in man. Eur J Pharmacol. 1974;28:172–7.
Kayser RR, Haney M, Raskin M, Arout C, Simpson HB. Acute effects of cannabinoids on
symptoms of obsessive-compulsive disorder: a human laboratory study. Depress Anxiety.
2020;37:801–11.
Kelly DL, Gorelick DA, Conley RR, Boggs DL, Linthicum J, Liu F, Feldman S, Ball MP, Wehring HJ,
McMahon RP, Huestis MA, Heishman SJ, Warren KR, Buchanan RW. Effects of the cannabinoid-
1 receptor antagonist rimonabant on psychiatric symptoms in overweight people with schizophre-
nia: a randomized, double-blind, pilot study. J Clin Psychopharmacol. 2011;31:86–91.
Kotin J, Post RM, Goodwin FK. 9 -tetrahydrocannabinol in depressed patients. Arch Gen Psychi-
atry. 1973;28:345–8.
Le Foll B, Goldberg SR. Rimonabant, a CB1 antagonist, blocks nicotine-conditioned place
preferences. Neuroreport. 2004;15:2139–43.
Le Foll B, Goldberg SR. Cannabinoid CB1 receptor antagonists as promising new medications for
drug dependence. J Pharmacol Exp Ther. 2005;312:875–83.
Levin FR, Mariani JJ, Brooks DJ, Pavlicova M, Cheng W, Nunes EV. Dronabinol for the treatment
of cannabis dependence: a randomized, double-blind, placebo-controlled trial. Drug Alcohol
Depend. 2011;116:142–50.
Levin FR, Mariani JJ, Pavlicova M, Brooks D, Glass A, Mahony A, Nunes EV, Bisaga A,
Dakwar E, Carpenter KM, Sullivan MA, Choi JC. Dronabinol and lofexidine for cannabis use
disorder: a randomized, double-blind, placebo-controlled trial. Drug Alcohol Depend.
2016;159:53–60.
Lev-Ran S, Le Foll B, McKenzie K, Rehm J. Cannabis use and mental health-related quality of life
among individuals with anxiety disorders. J Anxiety Disord. 2012;26:799–810.
Leweke FM, Piomelli D, Pahlisch F, Muhl D, Gerth CW, Hoyer C, Klosterkotter J, Hellmich M,
Koethe D. Cannabidiol enhances anandamide signaling and alleviates psychotic symptoms of
schizophrenia. Transl Psychiatry. 2012;2:e94.
Lintzeris N, Bhardwaj A, Mills L, Dunlop A, Copeland J, McGregor I, Bruno R, Gugusheff J,
Phung N, Montebello M, Chan T, Kirby A, Hall M, Jefferies M, Luksza J, Shanahan M,
Kevin R, Allsop D. Nabiximols for the treatment of Cannabis dependence: a randomized
clinical trial. JAMA Intern Med. 2019;179:1242–53.
Lintzeris N, Mills L, Dunlop A, Copeland J, McGregor I, Bruno R, Kirby A, Montebello M, Hall M,
Jefferies M, Kevin R, Bhardwaj A. Cannabis use in patients 3 months after ceasing nabiximols
for the treatment of cannabis dependence: results from a placebo-controlled randomised trial.
Drug Alcohol Depend. 2020;215:108220.
1036 S. Kloiber et al.
Raby WN, Carpenter KM, Rothenberg J, Brooks AC, Jiang H, Sullivan M, Bisaga A, Comer S,
Nunes EV. Intermittent marijuana use is associated with improved retention in naltrexone
treatment for opiate-dependence. Am J Addict. 2009;18:301–8.
Rigotti NA, Gonzales D, Dale LC, Lawrence D, Chang Y. A randomized controlled trial of adding
the nicotine patch to rimonabant for smoking cessation: efficacy, safety and weight gain.
Addiction. 2009;104:266–76.
Robinson JD, Cinciripini PM, Karam-Hage M, Aubin HJ, Dale LC, Niaura R, Anthenelli
RM. Pooled analysis of three randomized, double-blind, placebo controlled trials with
rimonabant for smoking cessation. Addict Biol. 2018;23:291–303.
Sarris J, Sinclair J, Karamacoska D, Davidson M, Firth J. Medicinal cannabis for psychiatric
disorders: a clinically-focused systematic review. BMC Psychiatry. 2020;20:24.
Sartim AG, Guimaraes FS, Joca SR. Antidepressant-like effect of cannabidiol injection into the
ventral medial prefrontal cortex-possible involvement of 5-HT1A and CB1 receptors. Behav
Brain Res. 2016;303:218–27.
Sarvet AL, Wall MM, Keyes KM, Olfson M, Cerda M, Hasin DS. Self-medication of mood and
anxiety disorders with marijuana: higher in states with medical marijuana laws. Drug Alcohol
Depend. 2018;186:10–5.
Schmidt ME, Liebowitz MR, Stein MB, Grunfeld J, Van Hove I, Simmons WK, Van Der Ark P,
Palmer JA, Saad ZS, Pemberton DJ, Van Nueten L, Drevets WC. The effects of inhibition of fatty
acid amide hydrolase (FAAH) by JNJ-42165279 in social anxiety disorder: a double-blind,
randomized, placebo-controlled proof-of-concept study. Neuropsychopharmacology. 2021;46:
1004–10.
Schwarcz G, Karajgi B, McCarthy R. Synthetic delta-9-tetrahydrocannabinol (dronabinol) can
improve the symptoms of schizophrenia. J Clin Psychopharmacol. 2009;29:255–8.
Seeman P. Cannabidiol is a partial agonist at dopamine D2High receptors, predicting its antipsy-
chotic clinical dose. Transl Psychiatry. 2016;6:e920.
Shalit N, Lev-Ran S. Does cannabis use increase anxiety disorders? A literature review. Curr Opin
Psychiatry. 2020;33:8–13.
Sloan ME, Gowin JL, Ramchandani VA, Hurd YL, Le Foll B. The endocannabinoid system as a
target for addiction treatment: trials and tribulations. Neuropharmacology. 2017;124:73–83.
Socías ME, Wood E, Lake S, Nolan S, Fairbairn N, Hayashi K, Shulha HP, Liu S, Kerr T, Milloy
MJ. High-intensity cannabis use is associated with retention in opioid agonist treatment: a
longitudinal analysis. Addiction. 2018;113:2250–8.
Soyka M, Koller G, Schmidt P, Lesch OM, Leweke M, Fehr C, Gann H, Mann KF. Cannabinoid
receptor 1 blocker rimonabant (SR 141716) for treatment of alcohol dependence: results from a
placebo-controlled, double-blind trial. J Clin Psychopharmacol. 2008;28:317–24.
Stanciu CN, Brunette MF, Teja N, Budney AJ. Evidence for use of cannabinoids in mood disorders,
anxiety disorders, and PTSD: a systematic review. Psychiatr Serv. 2021;72:appips202000189.
Suryadevara U, Bruijnzeel DM, Nuthi M, Jagnarine DA, Tandon R, Bruijnzeel AW. Pros and cons
of medical Cannabis use by people with chronic brain disorders. Curr Neuropharmacol.
2017;15:800–14.
Tonstad S, Aubin HJ. Efficacy of a dose range of surinabant, a cannabinoid receptor blocker, for
smoking cessation: a randomized controlled clinical trial. J Psychopharmacol. 2012;26:1003–9.
Trigo JM, Le Foll B. Inhibition of monoacylglycerol lipase (MAGL) enhances cue-induced
reinstatement of nicotine-seeking behavior in mice. Psychopharmacology. 2016;233:1815–22.
Trigo JM, Lagzdins D, Rehm J, Selby P, Gamaleddin I, Fischer B, Barnes AJ, Huestis MA, Le Foll
B. Effects of fixed or self-titrated dosages of Sativex on cannabis withdrawal and cravings. Drug
Alcohol Depend. 2016a;161:298–306.
Trigo JM, Soliman A, Staios G, Quilty L, Fischer B, George TP, Rehm J, Selby P, Barnes AJ,
Huestis MA, Le Foll B. Sativex associated with behavioral-relapse prevention strategy as
treatment for Cannabis dependence: a case series. J Addict Med. 2016b;10:274–9.
Trigo JM, Soliman A, Quilty LC, Fischer B, Rehm J, Selby P, Barnes AJ, Huestis MA, George TP,
Streiner DL, Staios G, Le Foll B. Nabiximols combined with motivational enhancement/
1038 S. Kloiber et al.
cognitive behavioral therapy for the treatment of cannabis dependence: a pilot randomized
clinical trial. PLoS One. 2018;13:e0190768.
Vadhan NP, Corcoran CM, Bedi G, Keilp JG, Haney M. Acute effects of smoked marijuana in
marijuana smokers at clinical high-risk for psychosis: a preliminary study. Psychiatry Res.
2017;257:372–4.
van den Elsen GA, Ahmed AI, Verkes RJ, Kramers C, Feuth T, Rosenberg PB, van der Marck MA,
Olde Rikkert MG. Tetrahydrocannabinol for neuropsychiatric symptoms in dementia: a ran-
domized controlled trial. Neurology. 2015a;84:2338–46.
van den Elsen GAH, Ahmed AIA, Verkes RJ, Feuth T, van der Marck MA, Olde Rikkert MGM.
Tetrahydrocannabinol in behavioral disturbances in dementia: a crossover randomized con-
trolled Trial. Am J Geriatr Psychiatry. 2015b;23:1214–24.
Volicer L, Stelly M, Morris J, McLaughlin J, Volicer BJ. Effects of dronabinol on anorexia and
disturbed behavior in patients with Alzheimer’s disease. Int J Geriatr Psychiatry. 1997;12:913–9.
Walsh Z, Gonzalez R, Crosby K, Thiessen MS, Carroll C, Bonn-Miller MO. Medical cannabis and
mental health: a guided systematic review. Clin Psychol Rev. 2017;51:15–29.
Walther S, Mahlberg R, Eichmann U, Kunz D. Delta-9-tetrahydrocannabinol for nighttime agitation
in severe dementia. Psychopharmacology. 2006;185:524–8.
Whiting PF, Wolff RF, Deshpande S, Di Nisio M, Duffy S, Hernandez AV, Keurentjes JC, Lang S,
Misso K, Ryder S, Schmidlkofer S, Westwood M, Kleijnen J. Cannabinoids for medical use: a
systematic review and meta-analysis. JAMA. 2015;313:2456–73.
Wilkerson JL, Ghosh S, Mustafa M, Abdullah RA, Niphakis MJ, Cabrera R, Maldonado RL,
Cravatt BF, Lichtman AH. The endocannabinoid hydrolysis inhibitor SA-57: intrinsic anti-
nociceptive effects, augmented morphine-induced antinociception, and attenuated heroin seek-
ing behavior in mice. Neuropharmacology. 2017;114:156–67.
Wycoff AM, Metrik J, Trull TJ. Affect and cannabis use in daily life: a review and recommenda-
tions for future research. Drug Alcohol Depend. 2018;191:223–33.
Zamberletti E, Rubino T, Parolaro D. The endocannabinoid system and schizophrenia: integration
of evidence. Curr Pharm Des. 2012;18:4980–90.
Zanelati TV, Biojone C, Moreira FA, Guimaraes FS, Joca SR. Antidepressant-like effects of
cannabidiol in mice: possible involvement of 5-HT1A receptors. Br J Pharmacol. 2010;159:
122–8.
Zuardi AW, Hallak JE, Dursun SM, Morais SL, Sanches RF, Musty RE, Crippa JA. Cannabidiol
monotherapy for treatment-resistant schizophrenia. J Psychopharmacol. 2006;20:683–6.
Cognitive Enhancement and American
Constitutional Law
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1040
Government Safety Interests and Limits on a Constitutional Right to Cognitive Liberty . . . . 1046
Tradition, Social Convention, and Limits on a Constitutional Right to Cognitive Liberty . . . 1050
Quasi-Constitutional Rights to Cognitive Liberty . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1055
Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1056
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1057
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1058
Abstract
In recent years, there has been significant research and debate about whether
individuals can – and should be able to – improve their memory or other aspects
of their cognition with cognitive enhancement drugs.
This debate has largely been about ethics, but a debate has also emerged in legal
scholarship about whether, if constitutional systems give individuals the freedom
to shape their thinking in other ways (with books, conversations, software, or
games), it should also give them freedom to safely do so with cognitive enhance-
ment drugs. In the American context (which is the focus of this chapter), certain
scholars argue that individuals’ long-recognized right to “freedom of thought,”
under the First Amendment of the US Constitution, should, in the twenty-first
century, be understood as a broad right to “cognitive liberty,” allowing us to shape
our own minds medically as well as culturally.
This chapter provides an overview of this argument and then examines two
major reasons why American courts might resist it. One concerns the risks and
uncertainties about the safety of pharmacological cognitive enhancement. Indi-
viduals have constitutional freedom to take action that comes with at least some
risks to safety: Many forms of protest can involve physical action government has
M. J. Blitz (*)
Oklahoma City University School of Law, Okahoma City, OK, USA
e-mail: mblitz@okcu.edu
cause to regulate. So it is not safety risks by themselves that weigh against finding
constitutional rights. It is rather because courts may well find, in analyzing laws
about cognition enhancement drugs, that it is harder to disentangle safety issues
(that government does have grounds to regulate) from liberty interests (that
government does not). A second reason courts may continue to treat different
forms of cognition enhancement differently is that historical tradition does so –
and traditional differences may, in rights jurisprudence, often override functional
similarities. The chapter ends by considering how, even when individuals lack
constitutional rights to use pharmacological means of cognitive enhancement,
legislatures can give them “quasi-constitutional” rights to do so.
Introduction
If I wish to improve my memory, there are a number of ways I might do so. I might
read a book on how to construct a “memory palace” or other mnemonic devices. I
can obtain a computer or Smartphone applications that let me tackle exercises that
challenge me to remember information – and perhaps build my capacity for doing so.
I might do a crossword puzzle. Each of these has been recommended as a way to
improve memory.
I also might conceivably do so by taking drugs – so-called “nootropics” substances
that can purportedly “boost creativity, memory, decision-making or other high-level
brain functions” (Heid 2019). These include substances individuals have long ingested
in common foods and beverages – such as caffeine (Dressler et al. 2019) or omega-3
oils in fish (Dressler et al. 2019; Luchtman and Song 2013). But they can also include
psychoactive drugs developed to treat illness – and repurposed to improve memory or
other cognitive functions in the “healthy.” To be sure, the benefits of such drugs remain
uncertain. Many have touted substances such as methylphenidate (or Ritalin) and
modafinil (or Provigil) as “smart pills.” Others have noted that acetylcholinesterase
inhibitors like those used to treat some Alzheimer’s patients may improve memory in
healthy adults as well – as might other drugs that increase levels of acetylcholine
(Farah et al. 2014). Others have argued that taking “microdoses” of LSD or mescaline
– doses too small to produce the hallucinatory experiences associated with those drugs
– improves individuals’ focus and creativity (Rose 2019).
But the research has been mixed. In a 2015 review of the literature on the
cognition-enhancing effects of these two substances, Veljko Dubljevic and Christo-
pher James Ryan found that there was as yet “little evidence that either methylphe-
nidate or modafinil are likely to provide any useful cognitive enhancement to those
who choose to use them” (Dubljevic and Ryan 2015). Dressler and his colleagues
similarly stress that “evidence for their efficacy for augmenting brain function and
cognition in healthy subjects is often markedly lower than assumed in theoretical
discussions” (Dressler et al. 2019; Luchtman and Song 2013). As Farah and her
colleagues stress, in a 2014 review of cognitive enhancement studies, “there are
surprisingly few generalizations about cognitive enhancement that can be stated with
confidence at present” (Farah et al. 2014). And Rose notes that when it comes to
Cognitive Enhancement and American Constitutional Law 1041
1
Other scholars have explored the implications of cognitive liberty in European law. Charlotte
Walsh argues that right to freedom of thought and cognitive liberty in the European Convention on
Human Rights justify decriminalization of psychedelics, and also argues that “should inform not
only defences raised in court but also the discourse of drug policy activism more broadly.” (Walsh
2016). Understanding the constitutional status of cognitive enhancement might also be informed by
a deeper exploration of scholarship on the ethics of cognitive enhancement in the United States and
Europe. (See e.g., Jotterand and Dubljevic 2016; Hildt and Franke 2013; Helmchen 2021; chapter
“▶ Ethical Issues in Neuropsychopharmacotherapy: US Perspective”).
Cognitive Enhancement and American Constitutional Law 1043
greater threat to safety. In fact, even those who have advanced ethical arguments in
favor of permitting healthy individuals to use cognitive enhancement drugs have
emphasized they continue to believe that government regulation of such drugs is
necessary to assure safety. For example, in an editorial in Nature calling for the FDA
to permit use of cognition-enhancing drugs not only to treat sick patients but to
improve the healthy, Henry Greeley and other experts on law, ethics, or neuroscience
emphasize that while they believe cognitive enhancement drugs should be available
to mentally healthy individuals, this does not mean they should be unregulated:
“Cognitive enhancements,” they noted, “affect the most complex and important
human organ [the brain], and the risk of unintended side effects is therefore both
high and consequential” (Greely et al. 2008). Neil Levy likewise notes, before
examining the difference between internal and external interventions to the mind,
that “there are understandable safety concerns” linked to new technologies for
cognitive enhancements (Levy 2007).
But pointing to the safety risks of pharmacology is at best a partial answer to the
challenge raised by defenders of cognitive liberty. It explains that government has to
be left with power to protect individuals’ physical safety, but not why, once govern-
ment has imposed the regulation it needs to do so, individuals cannot then be left free
to engage in any kind of cognitive enhancement they wish to engage in, with drugs
or through other means, as long as it is consistent with such safety interests. Once
government has relied on its duty of protecting health and safety to claim its own
share of the “territory” encompassed by pharmacological enhancement, why isn’t
the rest of that “territory” constitutionally reserved for individuals?
After all, this is what already happens in free speech law and other areas of
American law where courts have to balance liberty and the need for government
measures to assure public safety: In free speech jurisprudence, for example, Amer-
ican courts allow the government to prevent protestors from blocking traffic or
inciting violence, but as long as individuals steer clear of preventing others’ freedom
of movement or threatening their safety, the First Amendment lets them say what
they wish. It does not allow government to restrict a form of protest that raises no
safety threat on the ground that government wishes to suppress the expression in that
protest (United States v. O’Brien 1968). Government, in other words, cannot use
traffic- or violence-prevention as a basis for suppressing certain ideas it doesn’t like.
Why then, one might ask, can courts not apply a variant of the same framework when
it addresses the exercise of cognitive liberty by changing mental states – and find that
government may regulate drug safety, but may not constitutionally use it as an
excuse to suppress individuals’ decisions about how to shape their consciousness?
In the remainder of this chapter, I will explore two possible answers to this
question. The first is that, when it comes to pharmacology, it is far harder (if not
impossible) for courts – or for government – to bracket safety considerations. To
continue using the spatial metaphor I discussed above: It’s not the case that safety
considerations only fill some of the territory of pharmacological enhancement or
alteration of cognition, leaving the rest as a sphere of individual liberty. Rather, one
might argue, safety considerations permeate all of this territory in a way they do not
in the realm of speech or cultural expression.
1046 M. J. Blitz
The second possible answer is that protecting physical health and safety is not the
sole reason that government receives the significant power it does to regulate and
restrict individuals’ use of drugs – whether for psychopharmacology or for other
purposes. Outside the realms of human behavior that are constitutionally reserved for
individual liberty, after all, government regulates myriad activities for many reasons
other than assuring safety: It enacts laws that promote education, economic well-
being, the aesthetic quality of our shared public spaces, and many other public
benefits. So one might argue that if the government is left unconstrained in the
territory that constitutes the realm of psychopharmacology, this is not necessarily
because safety concerns permeate all of this territory. It is rather because there are
other reasons that our constitutional tradition reserves a realm of speech and expres-
sion for cognitive liberty while failing to similarly insulate other spheres where such
cognitive liberty could potentially be at stake. To some extent, this division of
territory between a sphere of constitutional rights and a sphere of government leeway
to regulate may be as much a matter of tradition and social convention as of
principle. Government may face constitutional limitations in regulating one means
of cognitive enhancement (through speech) and not another (with pharmacology) not
because of any inherent difference between these two means of exercising control
over mental autonomy – but rather because the social and constitutional traditions of
the United States (and of other jurisdictions) have historically divided up the world
that way, and the division leaves us with sufficient freedom for mental autonomy.
Below, I examine each of these two possibilities more closely. Even if either or
both of these frameworks succeeds in justifying why the constitution does not
insulate psychopharmacology from government, this still leaves open the possibility
of what we might call “quasi-constitutional limitations” on regulations that limit
cognitive liberty. In other words, even if our liberty to shape our consciousness is not
fully protected by the United States Constitution, as Boire and others argue it should
be, the government can still deem itself bound to refrain from unnecessarily inter-
fering in such liberty even when it is free to do so. It can enact statutes that constrain
itself even when the Constitution doesn’t do so.
As noted above, in Stanley v. Georgia, the Supreme Court writes that government
“may not constitutionally premise legislation on the desirability of controlling
a person’s private thoughts” (Stanley v. Georgia 1969). Why does this not present
a barrier to drug legislation that is premised on the “desirability” of preventing a
person from using such drugs to generate certain mental states?
One potential response by the courts is that – although such drug legislation
would raise constitutional concerns – this is not a possibility that arises, because all
restrictions on drugs in American law are premised at least in part on some
justification other than thought control – that is protecting patients’ health and safety.
The criteria that the Controlled Substance Act requires the FDA to use in classifying
Cognitive Enhancement and American Constitutional Law 1047
a drug as a Schedule I drug, after all, are safety-based: Such drugs have, among other
things, a “high potential for abuse,” and “a lack of accepted safety for use of the drug
or other substance under medical supervision.” 21 United States Code § 812. Since
concern for safety is always available as a justification for drug regulation, there are
never any cases where government is forced to concede that its true justification is to
limit individuals’ cognitive liberty.
But this isn’t a decisive objection. First, in both First and Fourteenth Amend-
ment law, courts are sometimes willing to show skepticism towards government’s
asserted justifications for enacting a law. They are willing to look behind govern-
ment claims that invoke justifiable grounds for government – to see if there are
constitutionally impermissible designs underlying the government’s action. Con-
sider the case of Sorrell v. IMS Health. That was a case not about drugs, but rather
speech about drugs: The State of Vermont regulated the practice of “detailing,”
wherein representatives of pharmaceutical companies attempt to convince physi-
cians to recommend and prescribe the medicines sold by those companies. When
making their sales pitch to physicians, these pharmaceutical company representa-
tives benefit from having information about the prescribing history of specific
physicians. The more they know about their audience, the better able they are to
convince them. Vermont was concerned that these sale pitches were convincing
physicians to prescribe expensive brand name drugs rather than generic drugs that
were just as effective. So they barred data-mining companies from sharing data on
physician prescribing practices with pharmaceutical company representatives.
Vermont said that their justification for doing so was to protect the privacy of
patients and doctors. But the Supreme Court was skeptical. It pointed out that
Vermont banned the sale of prescription data only to pharmaceutical company
representatives, but continued to let data-mining companies make it available to
numerous other entities, who could also use it in ways that threatened doctors’ or
patients’ privacy. The Court thus found that Vermont’s law seemed designed not to
protect patient safety but rather “to tilt public debate [about drugs] in a preferred
direction” – namely against argument for using brand name drugs and in favor of
using their generic equivalents (Sorrell v. IMS Health 2011).
The Court exhibited a similar kind of skepticism in a well-known case on freedom
of religion. In Lukumi Babalu Aye v City of Hialeah, Florida, the city of Hialeah had
banned the sacrifice of animals ostensibly in order to prevent cruelty to animals. But
the Court found that this justification was implausible because the ordinance
contained numerous exceptions that allowed the slaughter of animals (e.g., by
butchers). The only kind of animal slaughter occurring in the city that the exceptions
did not cover was that performed by the Santeria religious group as a part of their
religious ritual. Moreover, there had been numerous statements in city council
meetings expressing disdain for the Santeria. With the benefit of this background
information, the Court concluded that the only plausible explanation for the legal
ban on animal sacrifice was to prevent the Santeria from practicing their religion.
This purpose, to prevent members of a religious group from practicing their religion,
is an impermissible one under the US Constitution’s First Amendment protection for
religious liberty (Church of Lukumi Babalu Aye v. City of Hialeah, 1993).
1048 M. J. Blitz
So why shouldn’t courts engage in a similarly searching inquiry of laws that bar
pharmacological cognitive enhancement or other use of drugs to shape one’s mental
states? There are at least three different reasons one might offer.
One is that physical safety is always at stake when individuals ingest psychoac-
tive substances – and the state can thus always justify such regulation even when
certain government officials may have other motivations for preventing cognitive
enhancement. To quote the Nature editorial discussed earlier, psychoactive drugs all
affect “most complex and important human organ [the brain],” (Greely et al. 2008,
703) so, one might argue, the government’s safety justifications will never be
implausible in the way that the rejected justifications were in the above cases on
pharmaceutical company sales or animal slaughter. As Ricci notes, “the use of
methylphenidate, modafinil, amphetamine, and other prescription drugs involves
health risks, including dependence, tolerance, and cardiovascular, neurologic, and
psychological disorders” (Ricci 2020). Even when use of such drugs has a legitimate
medical purpose, the US government is not constitutionally bound to permit their
use. As Glannon points out, in a discussion of the ethics of cognitive enhancement
“adverse effects from using methylphenidate for cognitive enhancement intuitively
would be less acceptable than using it therapeutically” (Glannon 2015). As Bruhl et
al. point out, there are also safety risks associated with microdosing with LSD or
other hallucinogens: “Those who microdose incorrectly risk having unwanted, full-
blown trips or even experience unpleasant trips. There are some reports of psychosis-
like symptoms in certain vulnerable individuals who use LSD recreationally” (Bruhl
et al. 2019). Court may decide they are far less well-placed to second guess a safety
assessment by government than they are to second guess asserted justifications of
limiting speech or restrictions that burden religious activity.
In fact, courts may take the position that when chemicals affect brain structure
and function, it is permissible for government to begin by assuming that use of
cognitive enhancement drugs raises significant safety concerns – and then make their
use illegal until safety is proven. This, after all, is the stance that drug regulators
often take. When government claims it needs to restrict speech or religious practice
in order to promote safety, it has to provide powerful evidence of the safety
justification (and show it is real). But this approach is a poor fit for a system, like
that used by the US Food and Drug Administration that always begins with a
presumption that drugs that a company wishes to manufacture and market unsafe
until sufficient data show that it is safe and effective (21 United States Code
Annotated § 355; 21 Code of Federal Regulations § 312.23; Abigail All. for Better
Access to Developmental Drugs v. von Eschenbach, 2007). Numerous articles
emphasize how little is known about the long-term effects of many cognitive
enhancement drugs, so given this uncertainty, courts may simply refuse to scrutinize
government restrictions of pharmacological cognition enhancement. As Bruhl and
her co-authors stress, “there are no studies investigating the long-term effects of
cognitive enhancers in healthy people, where effects such as habituation or even a
potentiation of the beneficial effects, as well as different side effects (mental,
somatic), compared to a single dose could be detected” (Bruhl et al. 2019). And
past experiences with drugs “enthusiastically acclaimed for enhancing cognition and
Cognitive Enhancement and American Constitutional Law 1049
mood” provide reason for caution: “initial Golden Periods were followed by a
painful sobering, with the social and health–economic consequences still visible
today” (Bruhl et al. 2019).
To be sure, in debates over the ethics of cognitive enhancement, scholars have
pointed out that the mere existence of a health or safety risk is often not treated as
sufficient justification to ban an activity. Individuals are allowed to ski, climb
mountains, and participate in boxing matches and football games despite the risks
of these activities. They are allowed to eat high-cholesterol food that can increase
their risk of having a stroke or a heart attack. As Greely notes, “outside the world of
medicine we allow people to take non-trivial risks without requiring the government
to agree with consumers’ cost-benefit assessments” (Greely 2006). Of course,
an ethical or policy calculation is different from a constitutional rule: That govern-
ment tolerates certain risks when it leaves people free to engage in certain activities
(such as playing football) does not mean that government is constitutionally required
to tolerate such risks. Government is not constitutionally prohibited from banning
skiing or football, even though it is likely such a ban would be deeply unpopular in
the United States. One might thus argue that where shaping one’s mind with drugs
generates safety risks of any kind, it for democratic officials, not courts applying
constitutional law, to weigh the costs and benefits of each activity and decide
whether individuals should be left free to undertake it.
Second, one might argue that courts will be unable to distinguish permissible
safety-related justifications from (what one might argue are) impermissible designs
to constrain mental autonomy because these two interests are not entirely distinct.
Some of the worrisome side effects that patients, doctors, and government regulators
worry about are psychological. The FDA, for example, has issued warnings about
drugs that can cause depression as a side effect.2
If it is constitutionally permissible for government to prevent individuals from
ingesting chemicals on the grounds that doing so comes with a risk of severe
depression then it may be permissible for government to invoke health and safety
interests to prevent individuals from inducing other mental states, such as extreme
anxiety or paranoia. In fact, the Court has even found it constitutionally permissible
for prison officials to require that certain prisoners take antipsychotic medications in
order to treat certain psychological conditions that might be dangerous (Washington
v. Harper 1990, Riggins v. Nevada 1992). It has also suggested that authorities can
government may be able to forcibly medicate an inmate of a prison or psychiatric
facility in order to make that person “competent to stand trial,” where the govern-
ment can make certain showings. In Sell v. United States, the Court made clear that
the hurdle government has to meet to do so is a rather high one: Government must
show that “the treatment is medically appropriate, is substantially unlikely to have
side effects that may undermine the fairness of the trial, and, taking account of less
2
It did so, for example, with respect to smoking cessation drug, varenicline. See Food and Drug
Administration. FDA Drug Safety Communication: Safety review update of Chantix (varenicline)
and risk of neuropsychiatric adverse events.
1050 M. J. Blitz
In the preceding paragraphs, I have been assuming that – with respect to decisions
about psychopharmacology – where the government’s interest in guaranteeing safety
ends, constitutional cognitive liberty interests begin. But this isn’t the only possibil-
ity. It is also possible that constitutional freedom of thought protects not all the things
we do to shape our own thinking processes or mental, but only a subset of them. If
that is the case, our First Amendment freedom of thought – or Fifth and Fourteenth
Amendment liberty interests – may fall short of covering psychopharmacology not
because these protections have been pushed out of that territory to make room for
safety regulation, but because constitutional liberty never covered it in the first place.
Consider an analogy with another doctrine in First Amendment law: The right to
“receive information and ideas.” The Supreme Court has held that that the US
Constitution’s guarantee of free speech not only gives individuals a right to express
Cognitive Enhancement and American Constitutional Law 1051
ideas, but also a right to receive them from others – the right that is to be an audience,
and learn information (which they can then, of course, make the basis of their own
expression). (Martin v. City of Struthers 1943). But this right to receive information
has to have limits, as the Supreme Court emphasized in the 1965 case, Zemel v. Rusk.
This case arose when the United States State Department refused to validate the
passport of Louis Zemel for travel to Cuba. Zemel insisted that this violated his First
Amendment right to gather information: It was, he said, a “direct interference with
the First Amendment rights of citizens to travel abroad so that they might acquaint
themselves at first hand with the effects abroad of [the US] Government’s policies,
foreign and domestic, and with conditions abroad which might affect such policies.”
But the Court rejected Zemel’s claim. Treating the First Amendment as containing an
“unrestrained right to gather information,” said the Court, would stretch it to cover –
and shield from government regulation – virtually all human activity: “There are few
restrictions on action which could not be clothed by ingenious argument in the garb
of decreased data flow” (Zemel v. Rusk 1965).
The same argument might be made about freedom of thought. Every time we are
denied the chance to perceive a certain environment, we are denied the raw material
for a certain type of mental experience: Preventing Louis Zemel from visiting Cuba
not only prevented him from gathering information about Cuba. It prevented him
from having the kind of mental experiences that is difficult to have except when one
is in Cuba, perceiving the sights and sounds there. Freedom of thought, or cognitive
liberty, cannot plausibly entail a freedom to do anything one wishes to do in the
outside world – so that it will generate a particular kind of mental experience (See
Blitz 2010). There have to be limits on freedom of thought.
The Court of Appeals for the Seventh Circuit has attempted to identify such a
limit. In Doe v. City of Lafayette, it said that “only “governmental regulations
aimed at mere thought, and not thought plus conduct, trigger freedom of thought
protection” (Doe v. City of Lafayette, Indiana 2004). The problem with this
standard is that government regulations generally haven’t been aimed at “mere
thought” because, as Justice Murphy wrote in Jones v. Opelika, “the most tyran-
nical government” has generally been “powerless to control the inward workings
of the mind” (Jones v. Opelika 1942). One might argue that, with advances in
psychosurgery and greater understanding of pharmacological interventions, Mur-
phy’s generalization is no longer true: The government can control the inward
workings of the mind by controlling the functioning of the brain. But as a general
matter, when courts have raised worries that government action was violating or
threatening freedom of thought, they have focused on ways government does so
from the outside. For example, in Stanley v. Georgia, the Court worried that
government wished to control Stanley’s thoughts by restricting the films he
could possess and watch in his own home. One might argue that even here, the
government justified its limits on external conduct (watching a film) solely on the
basis of its effect on thinking (watching the film allowed the generation of a certain
mental state). But this begs the question of how courts can know a limit like that
imposed on Stanley is solely about targeting mental experiences rather than the
effects the government believes to flow from them.
1052 M. J. Blitz
Moreover, if Chalmers and Clark are right in arguing that the physical substrate
underlying our thinking includes not only our brain processes, but also the tools we
use outside of our bodies to store memories – from notebooks to Smartphones – then
many government regulations of these tools can potentially count as a restriction on
freedom of thought and at least incidentally burden it. The line between regulating
thought and regulating “conduct,” then, will not be entirely clear.
Perhaps for this reason, one might argue, the law’s answer to such a challenge
must often be based not only on principled line-drawing between categories, but on
history and on social conventions. More specifically, the reason that the constitu-
tional protections for freedom of thought cover only certain kinds of cognitive
enhancements and not others might be, in part, that (1) we need some protected
realm where we can exercise mental autonomy, even if we cannot exercise it in all
circumstances where it is possible and (2) the legal history of the United States has
marked out the realm of expression, and self-formation with cultural resources, as a
realm where individuals are generally free from state restriction as they shape and
reshape their own consciousness. Even if pharmacology is another tool that allows
us to similarly reshape our minds, it might lie outside of the sphere of tools that our
legal history has set aside for the exercise of mental autonomy. One might not
establish a constitutional right to pharmacological cognitive enhancement, then,
simply by showing that it is functionally analogous to cultural enhancement that is
already shielded by constitutional liberties. The parity principle that Levy defense
for ethics, in other words, may not translate to constitutional law – because even
when there is a functional parity between two forms of enhancement, social and
historical tradition may have classified them differently. Courts may treat this as
reason enough to continue treating them differently, so long as it leaves us with a
satisfactory amount of mental liberty, even if not the maximum amount conceivable.
In analyzing American doctrines of constitutional liberty, this account is best
elaborated in two steps. First, one might argue, to the extent the First Amendment’s
freedom of speech protects the thought underlying expression, this doesn’t necessarily
mean it protects all action we might take to shape the mental process that produce such
thought. Rodney Deaton has made an argument like this: The Supreme Court’s First
Amendment jurisprudence in Stanley v. Georgia, he argues, prevents government from
denying individuals the ability to generate particular beliefs or experiences. It bars
government from preventing individuals from obtaining the information or visual
experience that comes from watching a film (Deaton 2006). But when we ingest a
drug psilocybin or methylphenidate, this doesn’t generate a particular experience.
Outside of science fiction, an enhancement drug of that kind generally isn’t pro-
grammed to deliver us a custom-made dream or mental experience. Rather, it produces
certain changes in the way our brain operates (and, perhaps, in doing so, enhances
certain mental capacities). But that is different from delivering us particular informa-
tional or experiential content, and thus, it is different from the kind of thinking that is
arguably most closely to speech: namely, the mental formation of content that can then
be expressed (or that we find in others’ expression).
That a certain exercise of mental autonomy lies outside the realm of First
Amendment freedom of expression and the freedom of thought that makes it
Cognitive Enhancement and American Constitutional Law 1053
possible doesn’t mean that it lies outside the realm of constitutional liberties alto-
gether. In a series of cases, the Supreme Court has made it clear that pharmacological
alteration of a person’s brain function can implicate their constitutional rights – at
least when government forces such alteration on people against their will (Washing-
ton v. Harper 1990; Nevada v Riggins 1992; Sell v. United States 2003). The Court
found that government is subject to constitutional constraints when it requires
prisoners or inmates of psychiatric facilities to take antipsychotic medications to
make them less dangerous to themselves or others, or to make them competent to
stand trial. But the Court did so not on the grounds that compelled use of psycho-
active substances would violate the First Amendment’s right to “freedom of
thought,” but rather that it would violate the right that the Fifth and Fourteenth
Amendment “due process” clauses give each individuals to be free from government
coercion that invades their bodily autonomy or freedom to determine their own
medical treatment. The Court has never said that government might violate the
Constitution if, instead of compelling a person to take a psychoactive drug against
her will, it denied her the freedom to take a psychoactive drug voluntarily.
Even if it did, however, it may well conclude that if the right to reshape one’s
brain operations with a psychiatric medication or cognitive enhancement interven-
tion is a constitutional right – it is not a First Amendment right (to be analyzed under
doctrines similar to those that protect freedom of expression) but rather a Fifth or
Fourteenth Amendment right (to be analyzed under doctrines the protect our bodily
autonomy). This difference could have significance because, in certain respects,
constitutional protections for our right to bodily autonomy and medical treatment are
more limited than the First Amendment rights to advance these interests through the
adoption of a “clear and convincing” standard of proof to govern such autonomy in
the realm of expression and belief. The Supreme Court has suggested, for example,
that when individuals make decisions about whether to continue life-saving treat-
ment, such decisions are not completely insulated from government legal measures
furthering the state’s interest in “the preservation of human life” (Cruzan by Cruzan
v. Dir., Missouri Dep’t of Health 1990).
Second, courts might well not only reject arguments for shielding pharmacological
cognitive enhancement with First Amendment protections for freedom of thought, but
also arguments for shielding it with Fifth or Fourteenth Amendment protections. There
are a variety of reasons they might do so. But one form such a rejection is likely to take
in some legal argument is one that emphasizes the importance of social convention
and tradition in marking the boundaries of our right to physical liberty or bodily
autonomy. Many judges have argued that these boundaries are inherently vague –
because the United States Constitution, at least, says little about them. Courts have
rather implied them from the mention of the word “liberty” in the text of the due
process clauses of the Fifth and Fourteenth Amendments.
They have, primarily in the latter part of the twentieth century, found that certain
aspects of individuals’ “intimate conduct” occur within a constitutionally shielded
“zone of privacy” where state interference is generally out of place (Griswold v.
Connecticut 1965; Lawrence v. Texas 2003). This is the basis, for example, of the
decisions in which the Court has found unconstitutional laws that restrict
1054 M. J. Blitz
intuitively count as drugs and have long been available to individuals without the
need for any prescription: It is conceivable, for example, the courts might be more
receptive to constitutional arguments against laws that which bar individuals from
using omega-3 fish oil for cognitive enhancement.
What implications does the foregoing analysis have for the constitutional status of
pharmacological cognitive enhancement? My central point is that when courts and
legal scholars analyzing American law confront the question of whether there is, or
should be, a constitutional right to shape our minds medically as well as with words,
they are not constrained to reach only one answer. Rather, they come to a fork in the
road. One can certainly make the argument that the logic of the First Amendment’s
cases on freedom of thought, or that on Fifth and Fourteenth Amendment liberty to
bodily autonomy and to freedom from state interference in one’s brain processes,
should prevent the state from restricting individuals’ decisions to use pharmacolog-
ical methods to enhance their cognition or otherwise alter their thinking patterns.
More specifically, one might argue that, except where the state has a valid safety-
based reason for limiting such activity, it should respect individuals’ right to
determine the content of their own consciousness. As noted above, Sententia and
Boire make an argument of this kind.
On the other hand, for the reasons I’ve given above, it’s also quite plausible for
courts to conclude that, when it comes to constitutional law at least, there isn’t parity
between cultural and pharmacological modes of cognitive enhancement. This could
be because concerns about physical health and safety permeate the latter and make it
impossible to shut government regulators out of this realm to the extent that
constitutional rights generally do shut government out of the spheres they protect.
Or it could be because the constitutional status of an activity (like enhancement)
depends not merely on what it does (and whether it is functionally analogous to those
covered by existing rights) but on how American tradition has classified it.
Even if pharmacological enhancement is not protected by a constitutional right, it
could still be protected by a quasi-constitutional right. That is, even if the US
Constitution does not establish a right of cognitive liberty that courts might invoke
to strike down legislative limits on certain types of cognitive enhancement, members
of the US Congress, or of state legislators, may still limit themselves – and those in
the federal and state government bound by their laws. This is what the US Congress
did with respect to religious liberty. The 2006 Supreme Court case, Gonzales v. O
Centro Espírita Beneficente União do Vegetal, provides an example of this. In that
case, the Supreme Court found that the federal government did not have a “compel-
ling interest” and thus could not legally seize a shipment of ayahuasca that the União
do Vegetal Church intended to use in its religious ceremonies. This was not,
however, because the religious liberty protection of the US Constitution itself barred
government from restricting religious use of a psychoactive substance (Gonzales v.
O Centro Espírita Beneficente União do Vegetal, 2006). On the contrary, the 1990
1056 M. J. Blitz
Supreme Court case of Employment Division, Oregon v. Smith had made it clear that
a “neutral law of general applicability” – for example, a law that bans the use of
mescaline or ayahuasca by all individuals – is constitutional even if, in imposing
such a ban, it incidentally burdens individuals who wish to use such practices as part
of their religious worship. But that wasn’t the end of the story: Although Smith gave
government virtually unlimited room to enact drug laws (or other laws) that inci-
dentally burdened religion, Congress then enacted a statute holding the federal
government to a higher standard: In the Religious Freedom Restoration Act
(RFRA), it barred any part of the federal government from burdening religion,
even incidentally, unless doing so was justified by a “compelling government
interest” (42 U.S.C. § 2000bb-bb4). This standard (known in American constitu-
tional law, as “strict scrutiny”) is a very high hurdle for government – so it is not
surprising that the government was unable to meet it when it tried to seize the
ayahuasca used by the União do Vegetal Church. In short, the religious liberty
protection that Supreme Court refused to impose on government in Employment
Division, Oregon v. Smith as a matter of constitutional law was imposed by the
federal government on itself in federal legislation (Employment Division v. Smith,
1990). Many states have followed this example and impose state law constraints on
their own government that mirror the religious liberty protections in RFRA. I call
such law “quasi-constitutional” here because it imposes rights-based restraints of a
kind courts often find in the Constitution, but locates those constraints in statutes
rather than in amendments to the Constitution (which are, in the US system, much
harder to enact, especially at the level of federal rather than state government).
RFRA only protects the use of psychoactive substances in religious worship. But
it is possible for Congress or state legislatures to enact other statutes that provide
broader quasi-constitutional protections for cognitive liberty. Even if government
may restrict even generally safe uses of cognitive enhancement drugs (for purposes
other than treating illness), Congress or state legislatures could adopt laws barring
any restriction that does have a justification rooted in health or physical security. As
Veljko Dubljevic writes, a state might instead institute a “gatekeeper” system in
which patients can only obtain and use cognition-enhancement drugs under the
supervision of a physician, and with educational programs that require them to
learn about safety risks and the safest methods of using such drugs (Dubljevic 2013).
Even if arguments for parity between pharmacological and cultural cognitive
enhancement don’t justify giving the two the same status as a matter of constitutional
law, they may justify – as a policy matter – legislation that results in both of them
being treated equally, that is, in such a way that the government does not unneces-
sarily restrict cognitive liberty.
Conclusion
In the past 20 years, the possibility of enhancing cognition – with drugs or medical
devices, such as transcranial direct current stimulation (tDCS) – has generated
debates among ethicists about whether it is ethical for a society to prohibit
Cognitive Enhancement and American Constitutional Law 1057
individuals from enhancing themselves in this way, or to bar psychiatrists and other
physicians from aiding in that enhancement.
Such debates are, in part, debates about liberty. They ask if individuals should
have the freedom to use this kind of cognition enhancement. This chapter has
considered whether, if such liberty exists, it should come in the form of a constitu-
tional right to “cognitive liberty.” Such questions are likely to arise in any jurisdic-
tion where there are constitutional protections for freedom of thought or belief. My
focus here, however, has been on American constitutional law. More specifically,
this chapter focuses specifically on whether there is a right to cognitive enhancement
included within the “freedom of thought” courts have found in the First Amend-
ment’s free speech guarantees or among the liberty rights they have found in the
Fifth and Fourteenth Amendment due process clauses.
Richard Glen Boire points out that the right to cognitive liberty “remains only
obliquely defined within the US legal system” (Boire 2000b). One reason for this is
simply that in the past, government has not easily been able to – in Justice
Murphy’s words – invade and control “the inward workings of the mind” (Jones
v. Opelika, 1942). It can’t know individuals’ unexpressed thoughts. Nor can it
control their thoughts as effectively as it can coerce physical behavior. But there are
two other reasons that the cognitive liberty Boire, Sententia, and other describe
might have remained ill-defined. One is that once our freedom to shape our own
consciousness extends beyond the realm of speech and expression, it enters a realm
where health and safety interests normally justify extensive government monitoring
and regulation of a kind that often has an uneasy coexistence with constitutional
rights (which push government power back). Another is that although there is a
long tradition in American law of limiting the government’s power to restrict
individuals’ speech, and the public discourse and shaping of culture to which it
contributes, there is not a similar tradition of insulating, against government
regulation, the chemical inputs to our consciousness. In American law, courts that
announce new constitutional rights – or announce that they are extending familiar
constitutional rights (such as freedom of speech or due process rights) into new
territory – often present themselves as formally endorsing, in law, constitutional
understandings that have already found in a place in America’s traditions or cultural
understandings about fundamental liberties. Where such a path for extending
constitutional rights has not already been carved by tradition or social change,
courts often leave it to democratic processes to do the carving: They might, for
example, refuse to find any general constitutional right to cognitive liberty and
leave, to Congress and state legislatures, the decision of whether to legislatively
establish quasi-constitutional rights.
Cross-References
References
Abigail All. for Better Access to Developmental Drugs v. von Eschenbach, 469 F.3d 129 (DC Cir.
2007) (Federal Reporter 3d, Vol. 469, 129–138).
Church of Lukumi Babalu Aye v. City of Hialeah, 508 U.S. 520 (1993) (United States Reports, Vol.
508, 520–580).
Clark, A. and Chalmers. D, The Extended Mind, in Clark. A., Supersizing the Mind: Embodiment,
Action, and Cognitive Experience (Oxford University Press 2008).
Cruzan by Cruzan v. Dir., Missouri Dep’t of Health, 497 U.S. 261 (1990) (United States Reports,
Vol. 497, 261–357).
Doe v. City of Lafayette, 377 F.3d 757 (7th Cir. 2004) (Federal Reporter 3d, Vol. 377, 757–785).
Gonzales v. O Centro Espírita Beneficente União do Vegetal, 546 U.S. 418 (2006) (United States
Reports, Vol. 564, 418–439).
Griswold v. Connecticut, 381 U.S. 479 (1965) (United States Reports, Vol. 381, 479–531).
Jones v. Opelika, 316 U.S. 584 (1942) (Murphy, J., dissenting) (United States Reports, Vol. 316,
584–624).
Lawrence v. Texas, 539 U.S. 558 (2003) (United States Reports, Vol. 539, 558–606).
Martin v. Struthers, 319 U.S. 141 (1943) (United States Reports, Vol. 319, 141–157).
Michael H. v. Gerald D., 491 U.S. 110 (1989) (United States Reports, Vol. 491, 110–163).
Moore v. City of East Cleveland, 431 U.S. 494 (1977) (United States Reports, Vol. 494, 431–552)
Planned Parenthood v. Casey, 505 U.S. 833 (1992) (United States Reports, Vol. 505, 883–1002).
Riggins v. Nevada, 504 U.S. 127 (1992) (United States Reports, Vol. 504, 127–157).
Roe v. Wade, 410 U.S. 113 (1973) (United States Reports, Vol. 410, 113–178).
Sell v. United States, 539 U.S. 166 (2003) (United States Reports, Vol. 539, 166–193) Skiadas v.
Acer Therapeutics Inc., No. 1:19-CV-6137-GHW, 2020 WL 3268495, *1-*17 (S.D.N.Y. June
16, 2020) Fed. Sec. L. Rep. P 100,848.
Sorrell v. IMS Health, 564 U.S. 552 (2011) (United States Reports, Vol. 564, 552–603).
Stanley v. Georgia, 394 U.S 557 (1969) (United States Reports, Vol. 394, 557–572).
United States v. O’Brien, 391 U.S. 367 (1968) (United States Reports, Vol. 391, 367–391).
Video Software Deals Ass’n v. Schwarzenegger, 556 F.3d 950 (2009). (Federal Reporter 3d, Vol.
556, 950–967).
Washington v. Glucksberg, 521 U.S. 702 (1997) (United States Reports, Vol. 521, 702–789).
Washington v. Harper, 494 U.S. 210 (1990) (United States Reports, Vol. 494, 210–258).
Zemel v. Rusk, 381 U.S. 1 (1965) (United States Reports, Vol. 381, 1–39).
Statutes
21 United States Code § 812.
21 United State Code §§ 821–831.
Bruhl AB, Angelo CD, Sahakian B. Neuroethical issues in cognitive enhancement: modafinil as the
example of a workplace drug? Brain Neurosci Adv. 2019;3
Bublitz JC, Merkel R. Crimes against minds: on mental manipulations, harms, and human right to
mental self-determination. Crim L Phil. 2014;8:51–77.
Carter JA, Palermos SO. Is having your computer compromised a personal assault? The ethics of
extended cognition. J Am Phil Assoc. 2016;2(4):542–60.
Deaton RJS. Neuroscience and the in Corpore-ted first amendment. First Amendment L Rev.
2006;4:181–221.
Dressler M, Sandberg A, Bublitz C, Ohla K, Trenado C, Mroczko-Wąsowicz A, Kühn S, Repantis
D. Hacking the brain. ACS Chem Neurosci. 2019;10(3):1137–48.
Dubljevic V. Cognitive enhancement, rational choice and justification. Neuroethics. 2013;6:179–87.
Dubljevic V, Ryan C. Cognitive enhancement with methylphenidate and modafinil: conceptual
advances and societal implications. Neurosci Neuroecono. 2015;4:25–33.
Farah MJ, Smith M, Ilieva I, Hamilton R. Cognitive enhancement. Wiley Interdiscip Rev Cogn Sci.
2014;5(1):95–103.
Glannon W. Neuroethics: cognitive enhancement. In: Goldberg S, editor. Oxford philosophy
handbooks. New York: Oxford University Press; 2015. p. 1–13.
Greely HT. The social effects of advances in neuroscience: legal problems, legal perspectives. In:
Neuroethics: defining the issues in theory, practice, and policy, vol. 245. Judy Illes ed. Oxford:
Oxford University Press; 2006.
Greely HT, et al. Towards responsible use of cognitive-enhancing drugs by the healthy. Nature.
2008;456:702–5.
Griffith RR, et al. Mystical-type experiences occasioned by psilocybin mediate the attribution of
personal meaning and spiritual significance 14 months later. J Psychopharmacol. 2008;22
(6):621–32.
Heid M. Nootropics, or ‘Smart Drugs,’ are gaining popularity. But should you take them? Time.
2019;23
Helmchen H. Some remarks on the ethics of Psychopharmacotherapy from a European viewpoint.
In: Riederer P, Laux G, Le W, Nagatsu T, editors. NeuroPsychopharmacotherapy; 2021.
Hildt E, Franke AG, editors. Cognitive enhancement. An interdisciplinary perspective. Dordrecht:
Springer; 2013.
Jotterand F, Dubljevic V, editors. Cognitive enhancement: ethical and policy implications in
international perspectives. New York: Oxford University Press; 2016.
Levy N. Neuroethics: challenges for the 21st century. Cambridge: Cambridge University Press; 2007.
Luchtman DW, Song C. Cognitive enhancement by omega-3 fatty acids from child-hood to old age:
findings from animal and clinical studies. Neuropharmacology. 2013;64:550–65.
Mehlman MJ. Cognition enhancing drugs. Millbank Q. 2004;82(3):483–506.
Ricci G. Pharmacological human enhancement: an overview of the looming bioethical and regu-
latory challenges. Front Psych. 2020;11:–53.
Rose S. Do microdoses of LSD change your mind? Sci Am. 2019;161
Sabet KA. Much Ado about nothing: why rescheduling won’t solve advocates’ medical Marijuana
problem. Wayne L Rev. 2012;58:81–101.
Sententia W. Neuroethical considerations: cognitive liberty and converging technologies for
improving human cognition. Ann N Y Acad Sci. 2004;1013(1)
Tribe LH. American constitutional law. 2d ed. New York: Foundation Press; 1988.
Walsh C. Psychedelics and cognitive liberty: reimagining drug policy through the prism of human
rights. Int J Drug Policy. 2016;29:80–7.
Challenges of Parkinson’s Disease Care
in Southeast Asia
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1062
Geography of Southeast Asia and Regional Epidemiology of Parkinson’s Disease . . . . . . . . . . 1063
Burden of Parkinson’s Disease in Southeast Asia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1066
Diagnostic Delay and Misdiagnosis of Parkinson’s Disease in Southeast Asia . . . . . . . . . . . . . . 1069
Knowledge Gaps of Parkinson’s Disease in Southeast Asia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1070
Parkinson’s Disease Treatment: Availabilities and Accessibilities in Southeast Asia . . . . . . . . 1071
Telehealth as a Solution for Specialist Parkinson’s Care in Southeast Asia . . . . . . . . . . . . . . . . . . 1075
Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1076
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1077
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1077
R. Bhidayasiri (*)
Chulalongkorn Centre of Excellence for Parkinson’s Disease and Related Disorders, Department of
Medicine, Faculty of Medicine, Chulalongkorn University and King Chulalongkorn Memorial
Hospital, Thai Red Cross Society, Bangkok, Thailand
The Academy of Science, The Royal Society of Thailand, Bangkok, Thailand
e-mail: rbh@chulapd.org
S. Virameteekul
Chulalongkorn Centre of Excellence for Parkinson’s Disease and Related Disorders, Department of
Medicine, Faculty of Medicine, Chulalongkorn University and King Chulalongkorn Memorial
Hospital, Thai Red Cross Society, Bangkok, Thailand
B. Sukoandari
Red Cross Hospital, Bogor, West Java, Indonesia
T. N. Tran
Movement Disorder Unit, Department of Neurology, University Medical Center, University of
Medicine and Pharmacy, Ho Chi Minh, Vietnam
T. T. Lim
Island Hospital, Penang, Malaysia
Abstract
When reviewing literature on Parkinson’s disease (PD), the focus tends to be on
Western studies where as this, historically, is where the majority of studies have
been undertaken. However, this bias is gradually changing as the prevalence of
PD is rising steeply in highly populous nations, which are mainly located in
Eastern hemisphere and including some within Southeast Asia (SEA). With this
rising demand for PD care in SEA, it is becoming apparent that there is a
mismatch of what healthcare resources are available across the region including
access to facilities, qualified personnel, approved and reimbursed therapeutic
options, or even knowledge of those who treat and those who receive the
treatments. This chapter looks at these challenges through the lens of a typical
case of PD patient in SEA, where various challenges are faced along the case
history, from diagnosis to advanced disease stage, with the intention of providing
real regional perspectives, supported by data or published literature when avail-
able. In addition, further insights on the practicalities of each challenge are sought
using individual authors’ clinical experience from within various countries of
SEA. It is unlikely that this chapter includes all challenges that PD patients face in
this region, but the authors have attempted to outline all those that are relevant in
clinical situations and provide rationales or solutions for some when available. It
is not the intention of the authors to fill this chapter with statistics or theoretical
approaches that, though seemingly ideal, have never been implemented in real
clinical scenarios. Rather, the authors want readers to be able to visualize the
challenges from a real clinical vignette typical of this region and to compare and
contrast these challenges with what is happening in the reader’s own environ-
ment. With the era of information technology and the “new” normal that we all
are facing, we all can benefit from information gained from any part of the world
and apply it appropriately in our clinical practice.
Introduction
Fig. 1 Map of Southeast Asia depicting statistical data on the number of Parkinson’s disease patients and neurologists
1066 R. Bhidayasiri et al.
Clinical Vignette (Part 1) At the age of 63, the patient began to develop left
shoulder pain associated with dragging of her left foot during fast walking. She
Challenges of Parkinson’s Disease Care in Southeast Asia 1069
was initially treated for tendinitis by a general practitioner, internist, and orthopedic
surgeon for 2 years in her remote hometown without any improvement before being
referred to a neurologist in the university hospital located 250 km away where the
diagnosis of PD was made. (To be continued)
was identified in a systematic review as one of the common conditions that are
associated with diagnostic errors (overdiagnosis vs. underdiagnosis) in older patients
(Skinner et al. 2016). Specialists were found to less overdiagnose with a 1-PPV of
1.4–25% while overdiagnosis was more prevalent in patients with no bradykinesia,
rigidity, resting tremor, hypomimia, less severe disease, and higher scores on
cognitive testing (Schrag et al. 2002; Hughes et al. 2002; Newman et al. 2009).
Red flags for non-PD conditions include patients receiving low doses of anti-
parkinsonian medications, whose dose had not increased in 3 years, who had been
on monotherapy for 5 years, and who had neither a documented response to
dopaminergic medications nor clinical progression were more commonly over-
diagnosed (Newman et al. 2009; Schrag et al. 2002). In contrast, factors associated
with underdiagnosis included lower disease severity, shorter disease duration, living
in a nursing home, and absence of typical complications of PD and/or its treatment
(e.g., dyskinesia) (Schrag et al. 2002). Young-onset (defined in this study as the onset
between 21 and 45 years) was also associated with a delayed diagnosis and a
substantial proportion of young-onset patients were misdiagnosed with functional
disorders (Rana et al. 2012). Even in the hands of specialists, misdiagnosis can still
occur in the early stages of the disease due to subtle clinical features, absence of
typical presentations, and a lack of red flags for alternative diagnoses (Rizzo et al.
2016; Litvan et al. 2003; Bhidayasiri et al. 2019a). Conditions that were found to
mimic PD were similar across studies, including ET, drug-induced parkinsonism,
dementia with Lewy bodies, dystonic tremor, multiple system atrophy (MSA),
Alzheimer’s disease, and vascular parkinsonism (Breen et al. 2013; Bhidayasiri
et al. 2011a; Hughes et al. 1993). Symptom characteristics also influence the timing
of diagnosis in which gait disturbances were associated with the longest delay, while
tremor had the shortest (Breen et al. 2013).
Recent clinical diagnostic criteria proposed by the International Parkinson
and Movement Disorder Society (MDS) has incorporated normal functional
neuroimaging of the presynaptic dopaminergic system (e.g., 123I-2β-carbomethoxy-
3β-(4-iodophenyl)-N-(3-fluoropropyl) nortropane dopamine transporter single photon
emission computed tomography (123I-FP-CIT DAT SPECT), or L-3,4-dihydroxy-6-
(18F) fluoro-phenylalanine positron emission tomography (18FDOPA PET)) as one of
the absolute exclusion criteria for PD (Postuma et al. 2015). However, despite the
evidence supporting its use in the differential diagnosis of parkinsonism, its availability
and cost limits any widespread adoption in clinical practice in SEA, with its use here
mainly restricted to a few centers for research purposes (Mirpour et al. 2018; Bega et al.
2015; Bhidayasiri et al. 2012).
Several factors are likely to contribute to diagnostic delay and misdiagnosis of PD,
encompassing disease characteristics, knowledge of both patients and physicians,
socioeconomic situations, availabilities of investigations (to support or refute the
diagnosis), and even misperceptions that PD was attributed to other nonmedical
Challenges of Parkinson’s Disease Care in Southeast Asia 1071
causes (e.g., cold weather, witchcraft), resulting in patients seeking treatment from
traditional healers, instead of healthcare professionals. One study in Thailand
documented significant knowledge gaps in PD among medical professionals, with
internists and general practitioners’ knowledge of pharmacotherapy and treatment
guidelines limited compared to neurologists (Bhidayasiri et al. 2014). However, an
intensive training program among healthcare professionals in Laos has been shown
to improve knowledge in PD even among those with limited existing knowledge
(Phokaewvarangkul et al. 2020). As for general public knowledge on PD, one large
study involving more than 1,000 responders in Malaysia found they still incorrectly
believe that PD is always associated with tremor and it can be cured (Tan et al. 2015).
Although patients were found to have superior knowledge to that of the general
public, less than half of PD patients in Malaysia recognized NMS as part of the
disease manifestations, with the least recognized reduced sense of smell, followed by
weight loss, and pain (Tan et al. 2015). Among Thai PD patients, younger age,
female gender, and higher disease duration were identified as predictors of a higher
level of PD knowledge and this group could potentially serve as patient advocates
for longitudinal patient educational programs within the country (Jitkritsadakul et al.
2017). The issue of stigma/shame was also observed in one-third of PD patients and
is recognized by several patient’s support groups within SEA where a recent public
education to break this stigma using integrated media has been shown to be effective
(Choo et al. 2020; Jagota et al. 2020).
Clinical Vignette (Part 2) The patient responded well to levodopa, taken four times
daily, to the point that she could return to her usual, pre-illness, activities. Due to a
long duration response, she cut back the frequency of levodopa to twice daily, but at
a higher dosage, to make her treatment regime more convenient. After only a year,
she started to develop motor fluctuations of predominantly wearing-off and peak-
dose dyskinesia. She required additional classes of dopaminergic medications to
improve her fluctuations, but these medications were not reimbursed and also not
available from her local health authority. Moreover, she did want to pursue deep
brain stimulation as this procedure was not available locally and she wished to stay
close to her family. As a result, she self-adjusted her own medications by increasing
the frequency of levodopa, leading to worsening dyskinesias. In addition, she
followed her neighbors’ suggestions to experiment with several herbal medicines
with the hope that her disease would be cured.
While levodopa remains a gold standard for the treatment of PD, recent evidence has
indicated that there are subgroups of PD patients whose optimal treatment in the early
stage may not be necessarily levodopa, but other classes of dopaminergic medications
(De Bie et al. 2020). Moreover, as the disease progresses, it is likely that most PD
patients will require additional classes of dopaminergic medications alongside
1072 R. Bhidayasiri et al.
SAI N N N N N N Y N
MAOBIs
Selegiline Y N N Y Y Y Y N
Rasagiline Y N N N Y N Y N
Amantadine Y N N Y Y Y N N
Zonisamide N Y N Y Y N Y N
LCIG N N N N N N Y N
DBS Y Y N Y Y Y Y Y
SR: standard-release; CR: controlled-release; COMTI: catechol-O-methyltransferase inhibitor; LCE: levodopa carbidopa entacapone; DAs: dopamine agonists;
SAI: subcutaneous apomorphine infusion; MAOBIs: monoamine-oxidase B inhibitors; LCIG: levodopa-carbidopa intestinal gel; DBS; deep brain stimulation;
N: Not available; Y; available
1073
1074 R. Bhidayasiri et al.
relationship (Kim and Jeon 2012; Tan et al. 2006). While approximately 40% of
Western PD patients reported using at least one type of CAM methods, the rate is
probably much higher in Asian populations, as documented in 61% and 76% of
Singaporean and Korean PD patients, respectively (Kim et al. 2009; Tan et al. 2006).
CAM seems to be common practice among Asian population who believe that
Western medicine is toxic and CAM can help balance “Gi” for the promotion of
health. Evidence for CAM users in SEA is strong. In recent surveys, 6.4% of
Singaporeans and 32.9% of Indonesians reported using at least one CAM method
for their chronic conditions, ranging from mental illness, sleep disturbance, and
depression (Seet et al. 2020, Pengpid and Peltzer 2018b). In PD, patients with severe
motor dysfunction at onset are more likely to use CAM and it is often difficult to
determine a definite conclusion on the efficacy as most studies are limited by study
designs, inappropriate controlled settings, and inaccessible methods (Tan et al.
2006). However, an evidence-based review of CAM in PD ranging from natural
products (herbals, vitamins, minerals, and probiotics), mind and body practices
(acupuncture, massage, meditation, movement therapies, relaxation techniques, tai
chi, and yoga), and alternative systems (Korean and Chinese medicine, Ayurvedic,
and homeopathy) has shown some promising results that mind–body interventions
are effective forms of physical activity that are likely to foster good adherence and
may reduce disability related to PD (Bega and Zadikoff 2014). Another more recent
review also demonstrated beneficial effects of massage therapy on both motor and
NMS of PD along with safe techniques (Angelopoulou et al. 2020). Among the
studies included in this systematic review is a study from Thailand on the efficacy of
therapeutic Thai massage in improving upper limb motor strength in PD patients
(Miyahara et al. 2018).
Clinical Vignette (Part 3) The patient’s condition progressed to the point that
ambulation was only possible with a wheelchair. This protracted mobility further
limiting visits to her neurologist. At this stage, her husband acted as a 24-h full-time
carer so that medications were given to the patient at the right time and right dosage,
lessening the severity of dyskinesia and “off” episodes. However, he lost his job as a
mechanic, causing financial constraints and stress to himself and the family. Due to
her present condition and COVID-19 pandemic, the patient was evaluated by her
neurologist via teleclinic, providing convenient, specialist care, and a reduced risk
of contagion to the patient.
In the past, specialist consultations were only achievable by face-to-face visits. This
was particularly troublesome for PD patients who have impaired mobility and live
far from those practices, resulting in inconvenience, discomfort, and inadequate care
to patients. In SEA, most neurologists tend to congregate in large cities, providing
significant disadvantages to nonurban patients who lack access to specialist care
1076 R. Bhidayasiri et al.
Conclusion
also one global society, should be aware of the impact PD has on all levels of society.
This chapter examines various challenges, outlining the problems, statistics, litera-
ture, and even solutions for some that are prevalent in SEA in relation to the rest of
the World. Although challenges have been evident in SEA, we must also recognize
that healthcare professionals in our region do their utmost to overcome these
significant challenges in order to deliver to their patients their best diagnostic
abilities, treatment decisions, and multidisciplinary, interdisciplinary, or transdisci-
plinary management wherever possible. These efforts should not be forgotten, and
the authors believe that our community has made significant and positive changes to
the care of PD in SEA, but should still continue working together to bridge the gaps
to improve care for our patients.
Cross-References
Financial Disclosure Roongroj Bhidayasiri: receives salary from Chulalongkorn University and
stipend from the Royal Society of Thailand, has received consultancy and/or honoraria/lecture
fees from Abbott, Boehringer-Ingelheim, Britannia, Ipsen, Novartis, Teva-Lundbeck, Takeda,
and Otsuka pharmaceuticals; he has received research funding from the Newton Fund, the UK
Government, Thailand Science and Research Innovation Bureau, Thailand Research Fund,
Crown Property Bureau, Chulalongkorn University, and the National Science and Technology
Development Agency; he holds patents for laser-guided walking stick, portable tremor device,
nocturnal monitoring, and electronic Parkinson’s disease symptom diary as well as copyright on
Parkinson’s mascot, dopamine lyrics and teaching video clips for common nocturnal and gastroin-
testinal symptoms for Parkinson’s disease.
Other authors report no financial disclosure.
References
2021a. Global Health Observatory Data Repository (South-East Asia Region) [Online]. World Health
Organization. Available: https://apps.who.int/gho/data/view.main-searo.SDG2016LEXREGv?
lang¼en. Accessed 22 Jan 2021.
2021b. South Eastern Asian Population [Online]. Available: https://www.worldometers.info/world-
population/south-eastern-asia-population/#:~:text¼The%20current%20population%20of%
20South,of%20the%20total%20world%20population. Accessed.
Abbas MM, et al. Epidemiology of Parkinson’s disease-east versus west. Mov Disord Clin Pract.
2018;5:14–28.
Angelopoulou E, et al. Massage therapy as a complementary treatment for Parkinson’s disease: a
systematic literature review. Complement Ther Med. 2020;49:102340.
1078 R. Bhidayasiri et al.
Gavidia M. Poll finds 1 in 4 people with Parkinson disease misdiagnosed [Online]. Am J Managed
Care. 2020. Available: https://www.ajmc.com/view/poll-finds-1-in-4-people-with-parkinson-
disease-misdiagnosed. Accessed 28 Jan 2021.
GBD 2015 Neurological Disorders Collaborator Group. Global, regional, and national burden of
neurological disorders during 1990–2015: a systematic analysis for the Global Burden of
Disease Study 2015. Lancet Neurol. 2017;16:877–97.
GBD 2016 Neurology Collaborators. Global, regional, and national burden of neurological disor-
ders, 1990–2016: a systematic analysis for the Global Burden of Disease Study 2016. Lancet
Neurol. 2019;18:459–80.
GBD 2016 Parkinson’s Disease Collaborators. Global, regional, and national burden of Parkinson’s
disease, 1990–2016: a systematic analysis for the Global Burden of Disease Study 2016. Lancet
Neurol. 2018;17:939–53.
GBD 2019 Diseases and Injuries Collaborators. Global burden of 369 diseases and injuries in 204
countries and territories, 1990–2019: a systematic analysis for the Global Burden of Disease
Study 2019. Lancet. 2020;396:1204–22.
Hessel F. Burden of disease. In: Kirch W, editor. Encyclopedia of public health. Dordrecht:
Springer; 2008.
Hughes AJ, et al. A clinicopathologic study of 100 cases of Parkinson’s disease. Arch Neurol.
1993;50:140–8.
Hughes AJ, et al. The accuracy of diagnosis of parkinsonian syndromes in a specialist movement
disorder service. Brain. 2002;125:861–70.
Jagota P, et al. If your patients were too embarrassed to go out in public, what would you do? –
public education to break the stigma on Parkinson’s disease using integrated media. Patient
Relat Outcome Meas. 2020;11:143–8.
Jamora RDG, Miyasaki JM. Treatment gaps in Parkinson’s disease care in the Philippines.
Neurodegener Dis Manag. 2017;7:245–51.
Jitkritsadakul O, et al. Knowledge, attitudes and perceptions of Parkinson’s disease: a cross-
sectional survey of Asian patients. J Neurol Sci. 2017;374:69–74.
Kim JY, Jeon BS. Complementary and alternative medicine in Parkinson’s disease patients in
Korea. Curr Neurol Neurosci Rep. 2012;12:631–2.
Kim SR, et al. Use of complementary and alternative medicine by Korean patients with Parkinson’s
disease. Clin Neurol Neurosurg. 2009;111:156–60.
Kowal SL, et al. The current and projected economic burden of Parkinson’s disease in the United
States. Mov Disord. 2013;28:311–8.
Kwok JYY, et al. Symptom burden and unmet support needs of patients with Parkinson’s disease: a
cross-sectional study in Asia-Pacific regions. J Am Med Dir Assoc. 2020;S1525–8610(20)
30795-7. https://doi.org/10.1016/j.jamda.2020.09.012. Online ahead of print.
Leochico CFD, et al. Challenges to the emergence of telerehabilitation in a developing country: a
systematic review. Front Neurol. 2020;11:1007.
Lim SY, et al. Parkinson’s disease in the Western Pacific region. Lancet Neurol. 2019;18:865–79.
Litvan I, et al. Movement disorders society scientific issues committee report: SIC
task force appraisal of clinical diagnostic criteria for parkinsonian disorders. Mov Disord.
2003;18:467–86.
Maetzler W, et al. Modernizing daily function assessment in Parkinson’s disease using capacity,
perception, and performance measures. Mov Disord. 2020;36(1):76–82. https://doi.org/10.
1002/mds.28377. Epub 2020 Nov 15.
Mirpour S, et al. Impact of DAT-SPECT on management of patients suspected of parkinsonism.
Clin Nucl Med. 2018;43:710–4.
Miyahara Y, et al. Can therapeutic Thai massage improve upper limb muscle strength in Parkinson’s
disease? An objective randomized-controlled trial. J Tradit Complement Med. 2018;8:261–6.
Mohammadi D. Neurology in resource-poor countries: fighting for funding. Lancet Neurol.
2011;10:953–4.
1080 R. Bhidayasiri et al.
Muangpaisan W, et al. Systematic review of the prevalence and incidence of Parkinson’s disease in
Asia. J Epidemiol. 2009;19:281–93.
Murray CJ, Lopez AD. Measuring the global burden of disease. N Engl J Med. 2013;369:448–57.
National Institute for Health and Care Excellence (NICE). Parkinson’s disease in adults. 2017.
https://www.nice.org.uk/guidance/ng71.
Newman EJ, et al. Accuracy of Parkinson’s disease diagnosis in 610 general practice patients in the
West of Scotland. Mov Disord. 2009;24:2379–85.
Pengpid S, Peltzer K. Multimorbidity in chronic conditions: public primary care patients in four
greater Mekong countries. Int J Environ Res Public Health. 2017;14:1019.
Pengpid S, Peltzer K. The impact of chronic diseases on the quality of life of primary care patients in
Cambodia, Myanmar and Vietnam. Iran J Public Health. 2018a;47:1308–16.
Pengpid S, Peltzer K. Utilization of traditional and complementary medicine in Indonesia: results of
a national survey in 2014–15. Complement Ther Clin Pract. 2018b;33:156–63.
Phokaewvarangkul O, et al. Addressing knowledge gaps in Parkinson’s disease: a report on
the Movement Disorder Society’s Centre-to-Centre initiative to improve Parkinson’s disease
services in Lao People’s Democratic Republic. BMC Med Educ. 2020;20:239.
Postuma RB, et al. MDS clinical diagnostic criteria for Parkinson’s disease. Mov Disord.
2015;30:1591–601.
Poungvarin N. Resources and organization of neurology care in South East Asia. Neurol Asia.
2007;12:41–6.
Prado M Jr, Jamora RD. Cost of Parkinson’s disease among Filipino patients seen at a public tertiary
hospital in Metro Manila. J Clin Neurosci. 2020;74:41–6.
Pringsheim T, et al. The prevalence of Parkinson’s disease: a systematic review and meta-analysis.
Mov Disord. 2014;29:1583–90.
Rana AQ, et al. Challenges in diagnosis of young onset Parkinson’s disease. J Neurol Sci. 2012;323:
113–6.
Razali R, et al. Burden of care among caregivers of patients with Parkinson disease: a cross-
sectional study. Clin Neurol Neurosurg. 2011;113:639–43.
Rizzo G, et al. Accuracy of clinical diagnosis of Parkinson disease: a systematic review and meta-
analysis. Neurology. 2016;86:566–76.
Sakdisornchai K, et al. Availability of anti-parkinsonian drugs in Thailand. Mov Disord. 2017;32:
S512–4.
Schrag A, et al. How valid is the clinical diagnosis of Parkinson’s disease in the community? J
Neurol Neurosurg Psychiatry. 2002;73:529–34.
Seet V, et al. The use of complementary and alternative medicine in a multi-ethnic Asian population:
results from the 2016 Singapore Mental Health Study. BMC Complement Med Ther.
2020;20:52.
Singhal BS, Khadilkar SV. Neurology in the developing world. Handb Clin Neurol. 2014;121:
1773–82.
Skinner TR, et al. Diagnostic errors in older patients: a systematic review of incidence and potential
causes in seven prevalent diseases. Int J Gen Med. 2016;9:137–46.
Tan CT. Neurology in Asia. Neurology. 2015;84:623–5.
Tan LC, et al. Prevalence of Parkinson disease in Singapore: Chinese vs Malays vs Indians.
Neurology. 2004;62:1999–2004.
Tan EK, et al. Prescribing pattern in Parkinson’s disease: are cost and efficacy overriding factors?
Int J Clin Pract. 2005;59:511–4.
Tan LC, et al. Use of complementary therapies in patients with Parkinson’s disease in Singapore.
Mov Disord. 2006;21:86–9.
Tan AH, et al. Knowledge of Parkinson’s disease in a multiethnic urban Asian setting. J Parkinsons
Dis. 2015;5:865–79.
Tan SB, et al. Parkinson’s disease caregiver strain in Singapore. Front Neurol. 2020;11:455.
Viwattanakulvanid P, et al. The impact of the nocturnal disabilities of Parkinson’s disease on
caregivers’ burden: implications for interventions. J Neural Transm. 2014;121(Suppl 1):15–24.
Challenges of Parkinson’s Disease Care in Southeast Asia 1081
Willis AW, et al. Neurologist care in Parkinson disease: a utilization, outcomes, and survival study.
Neurology. 2011;77:851–7.
Yang W, et al. Current and projected future economic burden of Parkinson’s disease in the U.S. NPJ
Parkinsons Dis. 2020;6:15.
Zhao YJ, et al. Economic burden of Parkinson’s disease in Singapore. Eur J Neurol. 2011;18:519–26.
Zhou X, et al. Burnout, psychological morbidity, job stress, and job satisfaction in Chinese
neurologists. Neurology. 2017;88:1727–35.
Neuropsychopharmacotherapy:
Complementary Treatments
Hans Moises
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1084
Psychotherapies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1085
Behavioral Activation Therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1085
Cognitive Behavioral Therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1085
Problem-Solving Therapies (PST) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1086
First-Wave Therapies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1086
Second-Wave Therapies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1086
Third-Wave Therapies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1087
Interpersonal Psychotherapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1088
Psychodynamic Therapies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1088
Non-directive Supportive Therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1088
Life Review Therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1089
Exposure Therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1089
Eye Movement Desensitization and Reprocessing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1089
The Dodo Bird Verdict . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1090
Common Factors Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1090
Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1091
Problems with Meta-Analyses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1093
Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1093
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1094
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1094
Abstract
Psychotherapies are popular complementary treatments. These treatments include
behavioral activation therapy (BA), cognitive behavior therapy (CBT), problem-
solving therapies (PST), and self-examination therapy (SET); “third-wave” therapies
H. Moises (*)
Department of Psychiatry and Psychotherapy, Kiel University Hospital, Kiel, Germany
e-mail: moises@icloud.com; hans.moises@alumni.uni-kiel.de
Introduction
0.54
0.5
adjusted)
0.41
0.36 0.35
0.4 0.28
0.23 0.23 0.23 0.21 Clinically Irrelevant
0.2
Behavioral Activation Therapy
CBT Generic
Psychodynamic - Generic
Problem-Solving Therapy
Fig. 1 Effects of psychotherapies compared to control groups in RCTs with low risk of bias. Data
from “Table V. The effects of fifteen evidence-supported therapies for adult depression: A meta-
analytic review.” (Cuijpers et al. 2019). The result for third-wave therapies is only based on RCTs of
ACT and MBCT. Insufficient numbers of studies were available for other types of third-wave therapies
Psychotherapies
Cognitive behavioral therapy (CBT) is based on the insight of ancient Stoic philos-
ophers that an individual’s emotions are strongly influenced by their thoughts
1086 H. Moises
(Evans, 2013). The focus of the cognitive behavioral therapist is on the pessimistic
and negative thoughts of the depressive patient, the so-called thinking errors. The
aim is to modify the cognitive distortions of the patient by challenging them to
question and investigate their reality (Cuijpers et al. 2019). The results of a recent
meta-analysis are shown in Fig. 1.
Self-Examination Therapy
In self-examination therapy (SET), patients determine their major goals in life and
subsequently focus on employing problem-solving methods to achieve their prior-
ities. Patients learn to accept problems, which cannot be solved (Bowman et al.
1996; Cuijpers et al. 2019). The 2019 meta-analysis by Cuijpers et al. reveals a low
ES of 0.28 of self-examination therapy in depression (see Fig. 1).
First-Wave Therapies
The first generation of behavioral therapy was a reaction to the prevailing psycho-
analysis by Sigmund Freud, which was thought to have weak scientific support
(Hayes 2004). By contrast, behavioral therapy was based on the experimentally well-
established principles of learning theory. In the first wave of behavioral
therapy, anxiety was mainly treated by exposure to the feared stimuli. Cognition
was not taken into account. The primary purpose was to eliminate specific behaviors
(Wolpe 1958; Wolpe and Rachmann 1960; Hayes 2004, 2017).
Second-Wave Therapies
In the second wave, Aaron Beck developed CBT by targeting the content of thoughts
(Beck 1979; Hunot et al. 2013; Hayes 2017; Cuijpers and Cristea 2016).
Neuropsychopharmacotherapy: Complementary Treatments 1087
Third-Wave Therapies
Mindfulness-Based CBT
MBCT combines meditation or mindfulness exercises with CBT. It was developed for
the prevention of relapse in people with recurrent depression by Mark Williams
(Oxford) together with John Teasdale (Cambridge) and Zindel Segal (Toronto)
based on the work by Jon Kabat-Zinn (Worcester, USA) on mindfulness-based stress
reduction (Segal et al. 2002, 2012). Even after remission, depressed patients tend to
react to negative mood by ruminating about their life situation. In MBCT, patients
learn to change their reaction to unwanted thoughts and feelings by no longer trying to
avoid them but to responding in a more detached way. MBCT is an 8-week group
program. A considerable number of studies have now examined the effects in patients
with acute depression and reported a significant effect (Cuijpers et al. 2019) (in Fig. 1
as third-wave therapies).
exposure therapy, patients learn a new skill to confront the feared stimuli to achieve
their interpersonal goals (McCullough et al. 2014, 2015). Meta-analyses by Negt
et al. (2016) reported a moderate to high ES for CBASP when compared to treatment
as usual and interpersonal psychotherapy (ES ¼ 0.64–0.75, p < 0.05).
Interpersonal Psychotherapy
Psychodynamic Therapies
Life review therapy (LRT) employs reminiscence of the patient’s life to treat
depression in older adults. Patients are encouraged to systematically re-evaluate all
phases of their life and to try to resolve conflicts and to assess adaptive coping
responses (Lewis and Butler 1974; Cuijpers et al. 2019). A meta-analysis by
Bohlmeijer et al. (2003) obtained an ES of 0.84 for the treatment of depression in
the elderly following life review therapy. The ES was stronger in more severely
depressed patients (ES ¼ 1.23).
Exposure Therapy
Exposure treatment for phobia and other anxiety disorders was developed by Joseph
Wolpe (Wolpe 1958). Its foundation is the learning theory of Pavlov, Hull, and others
(Pavlov 1963; Hull 1943). Neuroses (as anxiety disorders were called at this time)
were thought to consist of learned behavior which can be unlearned by being
exposed to the feared stimuli (Wolpe 1954, 1958).
Exposure has become an essential part of many psychotherapies including CBT
(Sisemore 2012), DBT (Linehan 2018), ACT (Hayes 2004), prolonged exposure
therapy (PE) (Foa et al. 2007), and narrative exposure therapy (NE) for posttraumatic
stress disorder (PTSD), emotional exposure, or processing (Sisemore 2012).
In a meta-analysis, exposure therapy was directly compared to non-exposure
treatments. Exposure treatment led to significantly greater improvement at both
post-treatment (ES ¼ 0.44, p < 0.001) and follow-up (ES ¼ 0.35, p < 0.001),
contradicting the contention of the Dodo bird verdict that all psychotherapies are
equally effective (Wolitzky-Taylor et al. 2008).
Shapiro claims to have created an entirely novel paradigm, the “accelerated infor-
mation processing model,” to explain the effect of EMDR.
Scientific-sounding terms are employed; for example:
[The] valences of the neural receptors (synaptic potential) of the respective neuro networks,
which separately store various information plateaus and levels of adaptive information, are
represented by the letters Z through A. It is hypothesized that the high-valence target
network (Z) cannot link up with the more adaptive information, which is stored in networks
with a lower valence. That is, the synaptic potential is different for each level of effect held in
the various neuro networks. . .. (Shapiro 1995 as cited in Herbert et al. 2000)
The Dodo bird is a fictional character in Lewis Carroll’s novel Alice’s Adventures in
Wonderland. The Dodo proposes a race where no measurement is taken of how far or
for how long the competitors run, and at the end, he judges that “everyone has won
and all must have prizes.” This quote has come to represent one of the best-replicated
results of psychotherapy research: all psychotherapies are equally effective
(Luborsky et al. 1975). Despite numerous replications, this conclusion is controver-
sial among psychotherapists, and the possible causes of the Dodo bird verdict are
hotly debated (Budd and Hughes, 2009). Discussions of the controversy can be
found in “The dodo bird is extinct” (Beutler 2002) and “The Dodo Bird Verdict is
alive and well—mostly” (Luborsky et al. 2002). One of the possible explanations for
the Dodo bird verdict is that common factors across different psychotherapies may
be responsible for the equivalent efficacy of therapies (Luborsky 1995).
The common factors theory is based on meta-analyses that have been unable to find a
significant difference in effectiveness between different psychotherapeutic methods
(Luborsky 1995; Cuijpers 2019). Figure 2 shows the effect sizes of meta-analyses of
psychotherapies investigating common and specific factors. Interestingly, neither
specific ingredients, protocol adherence, the competence of the therapist, nor treat-
ment differences have a clinically relevant influence of the effect of psychotherapies.
In contrast, the characteristics of the therapist – their genuineness, their positive
regard for and affirmation of the patient, and their empathy – and factors such as the
strength of the alliance between patient and therapist as well as the consensus of the
Neuropsychopharmacotherapy: Complementary Treatments 1091
0.49
0.4
0.35
0.32
0.24
0.20 Clinially irrelevant
0.2
0.14
0.04
0.01
Therapist in RCTs
Treatment Di erences
Psychotherapy
Empathy
Alliance
Therapist’s Competence
Goal Consensus / Collaboration
Genuineness
Therapists in Naturalistic Setting
Cultural Adaptation
Expectations
Protocol Adherence
Fig. 2 Effect sizes of common factors in different psychotherapies obtained from meta-analyses.
Data from Table 9.1 of “The Great Psychotherapy Debate – Evidence for What Makes Psychother-
apy Work” (Wampold and Imel 2015)
therapeutic goal and the collaboration to achieve that goal appear to be of major
importance for the outcome of the psychotherapy.
Even in antidepressant studies, there was significant variability in treatment
outcome between psychiatrists. In one study, psychiatrists accounted for 9.1% of
variability on the Beck Depression Inventory as compared to 3.4% accounted for by
the treatment. The best psychiatrists had better outcomes giving a placebo than the
worst psychiatrists had when giving the antidepressant (Wampold and Imel 2015).
Some psychiatrists and therapists consistently achieve better results than others.
For example, in one study the best therapists had a response rate of 80% compared to
a response rate of 20% for the lowest-performing therapists.
Discussion
Psychotherapy studies are criticized for being – on average – of low quality (Eysenck
1984; Coyne and Kok 2014; Cuijpers et al. 2019a) although there is some evidence
of improvement (Chen et al. 2014). For example, 65% of psychotherapy trials in
1092 H. Moises
depression employed improper randomization (Chen et al. 2014), and only 20% of
trials were found to have a low risk of bias (Cuijpers et al. 2019a). Typically,
psychotherapy studies are small and methodologically flawed, and non-blinded
raters with strong allegiances to one of the treatments under investigation evaluate
the results (Coyne and Kok 2014).
Studies employing proper blinding result in smaller effect sizes for psychotherapy
for depression (Cuijpers et al. 2010). Allegiance to the psychotherapy under inves-
tigation is a very strong predictor of outcome in RCTs (Miller et al. 2008; Munder
et al. 2013; Wampold and Imel 2015; Cuijpers and Cristea 2016).
There are more than 250 different psychotherapies (Herink 2012). All reported
improvements in patients treated by these different methods might be due to placebo
effects (Cuijpers and Cristea 2016). Placebos are thought to work by an increased
expectation of improvement or cure on the part of the patient (Colagiuri et al. 2015).
As long as patients expect the therapy to work, they will get better (Cuijpers and
Cristea 2016). The placebo effect might explain why some exotic treatments, such as
Mesmer’s animal magnetism (Lanska and Lanska 2007), acupuncture (Derry et al.
2006), homeopathy (Grams 2019), swimming with dolphins, or dancing the Argen-
tine tango, appear to be effective (Cuijpers and Cristea 2016).
Improper randomization can lead to larger reported effect sizes. In a meta-
analysis of psychotherapies for depression, Cuijpers and Cristea (2016) found
significant smaller effect sizes in studies using concealed assignment of patients to
the treatments compared to studies which did not use such a procedure.
In expectation of receiving psychotherapy, patients on waiting lists appear to do
nothing to solve their problems. (Cuijpers et al. 2010). In consequence, the outcome
of patients on waiting lists is worse than those receiving no treatment. For this
reason, waiting lists can be termed nocebos (Furukawa et al. 2014). Indeed, several
meta-analyses have found much larger effect sizes in studies using waiting list
control groups as compared to studies employing other control groups (Mohr et al.
2009, 2014; Barth et al. 2013).
The exclusion of drop-outs from the analysis of outcome results in larger effect
sizes (Cuijpers and Cristea 2016). Usually, multiple outcome measures are employed
in psychotherapeutic studies. Selective outcome reporting introduces a bias (Cuijpers
and Cristea 2016).
Small samples can produce better results than large samples due to an increased
chance of systematic differences between the groups under investigation (Cuijpers
et al. 2010; Gibertini et al. 2012; Cuijpers and Cristea 2016).
Trials involving hundreds or even thousands patients are required to find a small
difference between psychotherapies, and consequently, a large number of psycho-
therapeutic studies have insufficient power to report statistically robust outcomes
(Cuijpers and Cristea 2016).
A considerable publication bias exists in the field of psychotherapeutic research, as
demonstrated by the negative correlation between effect size and sample size, the
biased distribution of p values, an excess of significant findings, and the fact that a
considerable number of trials are never published (Kühberger et al. 2014; Flint et al.
2015; Cuijpers and Cristea 2016). Similar problems exist in antidepressant research
(Turner et al. 2008).
Neuropsychopharmacotherapy: Complementary Treatments 1093
The results of meta-analyses have become a munition in the battle between psycho-
therapists advocating the Freudian and those favoring the Pavlovian paradigm. In
1977, the first meta-analysis of psychotherapy outcome studies was published by
Smith and Glass (Smith and Glass 1977; Smith et al. 1980). Combining the results of
nearly 400 controlled studies of different diagnoses, psychotherapies, and counselling,
they obtained an average ES of 0.68. Favoring the Pavlovian paradigm, Hans Eysenck
criticized such a meta-analysis as a meaningless exercise in adding apples and oranges.
He further claimed that certain excellent studies by prolific and prominent behavioral
researchers had been omitted, while others were included that were so flawed as to put
them beyond the pale of acceptability (Eysenck 1984, 1994). In his view, meta-
analysis muddies the waters in the majority of cases where it is being used and leads
to meaningless conclusions that are likely to hamper proper scientific research.
Thirty years later, Coyne and Kok observed that:
meta-analyses of psychotherapies are strongly biased toward concluding that treatments work,
especially when conducted by those who have undeclared conflicts of interest and investigator
allegiances. This includes developers and promoters of treatments that stand to gain financially
or otherwise from their branding as ‘evidence-supported’. (Coyne and Kok 2014)
that meta-analyses that include all published trials, without taking into account the problems
of meta-analyses, heavily overestimate the effects of psychotherapies. This makes it highly
probable that Smith and Glass (1977) in their early meta-analysis considerably over-
estimated the effects of psychotherapy. . . . we found the overall effect of psychotherapy
was small but significant. (Cuijpers et al. 2019)
However,
because the effect sizes are relatively small, only a small change in the overall estimate,
caused by an unknown source of bias, can make this non-significant. . . The possibility that
psychotherapies do not have effects that are larger than spontaneous recovery cannot be
excluded. (Cuijpers et al. 2019b)
Conclusions
Psychotherapy is well accepted by patients and the general public alike. Having seen
recovery from anxiety and depression in patients following behavior therapy or other
forms of psychotherapy as well as antidepressants, it is difficult to imagine that these
improvements are purely the result of spontaneous remission. The significant role of
the therapist’s personality including their empathy and genuineness (see Fig. 2) is in
contradiction to the thesis of spontaneous recovery.
1094 H. Moises
Currently, the best evidence for the efficiency of psychotherapies is obtained from
meta-analyses of studies with low risk of bias. The results show that BA, the third
wave’s ACT and MBCT, and Beck’s CBT have the largest effect sizes (see Fig. 1).
In summary, meta-analyses suggest that all psychotherapies and antidepressants
are equivalently effective (Cuijpers 2016). For future psychotherapy research, the
reduction of biases and a considerable increase in sample sizes are of major impor-
tance to achieve further progress in this field. There are achievable strategies
available to researchers to reduce bias. As an example, blinding of raters is not
widely used, yet it can be achieved by recording video of patients during interviews,
editing the records to remove any hints to the psychotherapeutic treatment obtained,
and then presenting the videos to blinded raters in a randomized sequence.
Cross-References
References
Barth J, Munder T, Gerger H, et al. Comparative efficacy of seven psychotherapeutic interventions
for patients with depression: a network meta-analysis. PLoS Med. 2013;10:e1001454.
Beck AT. Cognitive therapy of depression. New York: Guilford Press; 1979.
Beutler LE. The dodo bird is extinct. Clin Psychol Sci Pract. 2002;9:30–4.
Bisson JI, Roberts NP, Andrew M, Cooper R, Lewis C. Psychological therapies for chronic post-
traumatic stress disorder (PTSD) in adults. Cochrane Database Syst Rev. 2013. CD003388.
Bohlmeijer E, Smit F, Cuijpers P. Effects of reminiscence and life review on late-life depression: a
meta-analysis. Int J Geriatr Psychiatry. 2003;18:1088–94.
Bowman V, Ward LC, Bowman D, Scogin F. Self-examination therapy as an adjunct treatment for
depressive symptoms in substance abusing patients. Addict Behav. 1996;21:129–33.
Budd R, Hughes I. The Dodo Bird Verdict – controversial, inevitable and important: a commentary
on 30 years of meta-analyses. Clin Psychol Psychother. 2009;16:510–22.
Chen P, Furukawa TA, Shinohara K, et al. Quantity and quality of psychotherapy trials for
depression in the past five decades. J Affect Disord. 2014;165:190–5.
Colagiuri B, Schenk LA, Kessler MD, Dorsey SG, Colloca L. The placebo effect: from concepts to
genes. Neuroscience. 2015;307:171–90.
Coyne JC, Kok RN. Salvaging psychotherapy research: a manifesto. J Evid Based Psychother.
2014;14:105–24.
Cristea IA, Gentili C, Cotet CD, Palomba D, Barbui C, Cuijpers P. Efficacy of psychotherapies for
borderline personality disorder: a systematic review and meta-analysis. JAMA Psychiat.
2017;74:319–28.
Cuijpers P. The future of psychotherapy research: stop the waste and focus on issues that matter.
Epidemiol Psychiatr Sci. 2016;25:291–4.
Cuijpers P. The role of common factors in psychotherapy outcomes. Annu Rev Clin Psychol.
2019;15:207.
Cuijpers P, Cristea IA. How to prove that your therapy is effective, even when it is not: a guideline.
Epidemiol Psychiatr Sci. 2016;25:428–35.
Cuijpers P, van Straten A, Bohlmeijer E, Hollon SD, Andersson G. The effects of psychotherapy for
adult depression are overestimated: a meta-analysis of study quality and effect size. Psychol
Med. 2010;40:211–23.
Neuropsychopharmacotherapy: Complementary Treatments 1095
Cuijpers P, Geraedts AS, van Oppen P, et al. Interpersonal psychotherapy for depression: a meta-
analysis. Am J Psychiatry. 2011;168:581–92.
Cuijpers P, Driessen E, Hollon SD, et al. The efficacy of non-directive supportive therapy for adult
depression: a meta-analysis. Clin Psychol Rev. 2012;32:280–91.
Cuijpers P, Karyotaki E, de Wit L, Ebert DD. The effects of fifteen evidence-supported therapies for
adult depression: a meta-analytic review. Psychother Res. 2019. p. 1–15.
Cuijpers P, Karyotaki E, Reijnders M, Ebert DD. Is psychotherapy effective? Pretending everything
is fine will not help the field forward. Epidemiol Psychiatr Sci. 2019a;28:56–357.
Cuijpers P, Karyotaki E, Reijnders M, Ebert DD. Was Eysenck right after all? A reassessment of the
effects of psychotherapy for adult depression. Epidemiol Psychiatr Sci. 2019b;28:21–30.
Davidson PR, Parker KCH. Eye movement desensitization and reprocessing (EMDR): a meta-
analysis. J Consult Clin Psychol. 2001;69:305–16.
Derry CJ, Derry S, McQuay HJ, Moore RA. Systematic review of systematic reviews of acupunc-
ture published 1996–2005. Clin Med (Lond). 2006;6:381–6.
Dimidjian S, Hollon SD, Dobson KS, et al. Randomized trial of behavioral activation, cognitive
therapy, and antidepressant medication in the acute treatment of adults with major depression.
J Consult Clin Psychol. 2006;74:658–70.
D’Zurilla TJ, Nezu AM, Maydeu-Olivares A. Social problem solving: theory and assessment. In:
Chang EC, D’Zurilla TJ, Sanna LJ, editors. Social problem solving: theory, research, and
training, American Psychological Association, vol. 2004; 2004. p. 11–27.
Evans J. Philosophy for life: and other dangerous situations. London: Rider; 2013.
Eysenck HJ. Meta-analysis: an abuse of research integration. J Spec Educ. 1984;18:41–59.
Eysenck HJ. Meta-analysis and its problems. BMJ. 1994;309:789–92.
Flint J, Cuijpers P, Horder J, Koole SL, Munafò MR. Is there an excess of significant findings in
published studies of psychotherapy for depression. Psychol Med. 2015;45:439–46.
Foa E, Hembree E, Rothbaum BO. Prolonged exposure therapy for PTSD. New York:
Oxford University Press; 2007.
Furukawa TA, Noma H, Caldwell DM, et al. Waiting list may be a nocebo condition in psychotherapy
trials: a contribution from network meta-analysis. Acta Psychiatr Scand. 2014;130:181–92.
Gibertini M, Nations KR, Whitaker JA. Obtained effect size as a function of sample size in
approved antidepressants: a real-world illustration in support of better trial design. Int Clin
Psychopharmacol. 2012;27:100–6.
Grams N. Homeopathy-where is the science? A current inventory on a pre-scientific artifact. EMBO
Rep. 2019. p. 20.
Hayes SC. Acceptance and commitment therapy, relational frame theory, and the third wave of
behavioral and cognitive therapies–republished article. Behav Ther. 2004;35:639–65.
Hayes SC. Acceptance and commitment therapy, relational frame theory, and the third wave of
behavioral and cognitive therapies – republished article. Behav Ther. 2016;47:869–85.
Hayes SC. The third wave of cognitive behavioral therapy and the rise of process-based care.
[editorial]. World Psychiatry. 2017;16(3):245.
Hayes SC, Wilson KG. Acceptance and commitment therapy: altering the verbal support for
experiential avoidance. Behav Anal. 1994;17:289–303.
Hayes SC, Strosahl K, Wilson KG. Acceptance and commitment therapy: understanding and
treating human suffering. New York: Guilford Press; 1999.
Herbert JD, Lilienfeld SO, Lohr JM, et al. Science and pseudoscience in the development of eye
movement desensitization and reprocessing: implications for clinical psychology. Clin Psychol
Rev. 2000;20:945–71.
Herink R. The psychotherapy guidebook. Martinsville: Fideli Publishing; 2012.
Hull CL. Principles of behavior. New York: Appleton; 1943.
Hunot V, Moore TH, Caldwell DM, et al. ‘Third wave’ cognitive and behavioural therapies versus
other psychological therapies for depression. Cochrane Database Syst Rev. 2013. CD008704.
Jacobson NS, Martell CR, Dimidjian S. Behavioral activation treatment for depression: returning to
contextual roots. Clin Psychol Sci Pract. 2001;8:255–70.
Kanter J, Busch AM, Rusch LC. Behavioral activation. Routledge; 2009.
1096 H. Moises
Klerman GL, Weissman MM. Interpersonal psychotherapy of depression. Lanham: Jason Aronson;
1994.
Kühberger A, Fritz A, Scherndl T. Publication bias in psychology: a diagnosis based on the
correlation between effect size and sample size. PLoS One. 2014;9:e105825.
Lanska DJ, Lanska JT. Franz Anton Mesmer and the rise and fall of animal magnetism: dramatic
cures, controversy, and ultimately a triumph for the scientific method. In: Whitaker H, Smith
CUM, Finger S, editors. Brain, mind and medicine: essays in eighteenth-century neuroscience.
Boston: Springer; 2007.
Lewinsohn PM. A behavioral approach to depression. Essential papers on depression. 1974.
p. 150–172.
Lewis MI, Butler RN. Life-review therapy. Putting memories to work in individual and group
psychotherapy. Geriatrics. 1974;29:165–73.
Linehan MM. Cognitive-behavioral treatment of borderline personality disorder. New York:
Guilford Publications; 2018.
Linehan MM, Armstrong HE, Suarez A, Allmon D, Heard HL. Cognitive-behavioral treatment of
chronically parasuicidal borderline patients. Arch Gen Psychiatry. 1991;48:1060–4.
Luborsky L. Are common factors across different psychotherapies the main explanation for the
dodo bird verdict that “everyone has won so all shall have prizes”. Clin Psychol Sci Pract.
1995;2:106–9.
Luborsky L, Singer B, Luborsky L. Comparative studies of psychotherapies. Is it true that
“everyone has won and all must have prizes”. Arch Gen Psychiatry. 1975;32:995–1008.
Luborsky L, Rosenthal R, Diguer L, et al. The Dodo Bird Verdict is alive and well – mostly. Clin
Psychol Sci Pract. 2002;9:2–12.
Martell CR, Addis ME, Jacobson NS. Depression in context: strategies for guided action.
New York: WW Norton & Co; 2001.
McCullough JP Jr, Schramm E, Penberthy JK. CBASP as a distinctive treatment for persistent
depressive disorder. London: Routledge; 2014. p. 152.
McCullough JP Jr, Schramm E, Penberthy K. CBASP – cognitive behavioral analysis system of
psychotherapy: Chronische Depressionen effektiv behandeln. Paderborn: Junfermann Verlag;
2015.
Miller S, Wampold B, Varhely K. Direct comparisons of treatment modalities for youth disorders: a
meta-analysis. Psychother Res. 2008;18:5–14.
Mohr DC, Spring B, Freedland KE, et al. The selection and design of control conditions for
randomized controlled trials of psychological interventions. Psychother Psychosom.
2009;78:275–84.
Mohr DC, Ho J, Hart TL, et al. Control condition design and implementation features in controlled
trials: a meta-analysis of trials evaluating psychotherapy for depression. Transl Behav Med.
2014;4:407–23.
Munder T, Brütsch O, Leonhart R, Gerger H, Barth J. Researcher allegiance in psychotherapy
outcome research: an overview of reviews. Clin Psychol Rev. 2013;33:501–11.
Mynors-Wallis LM, Gath DH, Day A, Baker F. Randomised controlled trial of problem solving
treatment, antidepressant medication, and combined treatment for major depression in primary
care. BMJ. 2000;320:26–30.
Negt P, Brakemeier EL, Michalak J, et al. The treatment of chronic depression with cognitive
behavioral analysis system of psychotherapy: a systematic review and meta-analysis of
randomized-controlled clinical trials. Brain Behav. 2016;6:e00486.
Nezu AM. Efficacy of a social problem-solving therapy approach for unipolar depression. J Consult
Clin Psychol. 1986;54:196.
Öst LG. The efficacy of acceptance and commitment therapy: an updated systematic review and
meta-analysis. Behav Res Ther. 2014;61:105–21.
Pavlov IP. Conditioned reflexes and psychiatry. Translation by W.H. Gantt. New York:
International Publishers; 1963.
Neuropsychopharmacotherapy: Complementary Treatments 1097
Segal ZV, Williams JMG, Teasdale JD. Mindfulness-based cognitive therapy for depression. 1st ed. -
New York: Guilford Press; 2002. p. 351.
Segal ZV, Williams JMG, Teasdale JD. Mindfulness-based cognitive therapy for depression.
New York: Guilford Press; 2012. p. 451.
Shapiro F. Eye movement desensitization: a new treatment for post-traumatic stress disorder.
J Behav Ther Exp Psychiatry. 1989;20:211–7.
Shapiro F. Eye movement desensitization and reprocessing. Guilford Publications; 1995. p. 398.
Sisemore TA. The clinician’s guide to exposure therapies for anxiety spectrum disorders. Oakland:
New Harbinger Publications; 2012.
Smith ML, Glass GV. Meta-analysis of psychotherapy outcome studies. Am Psychol.
1977;32:752–60.
Smith ML, Glass GV, Miller TI. The benefits of psychotherapy. Johns Hopkins University Press,
Baltimore; 1980. p. 269.
Torrey EF. Witchdoctors and psychiatrists: the common roots of psychotherapy and its future.
Northvale/London: Jason Aronson Inc.; 1986. p. 316.
Turner EH, Matthews AM, Linardatos E, Tell RA, Rosenthal R. Selective publication of antide-
pressant trials and its influence on apparent efficacy. N Engl J Med. 2008;358:252–60.
Wampold BE, Imel ZE. The great psychotherapy debate. Routledge; 2015. p. 334.
Weissman MM, Klerman GL. Interpersonal psychotherapy and tricyclics for depression. In:
Psychiatry the state of the art. Heidelberg/New York: Springer; 1985. p. 99–104.
Wolitzky-Taylor KB, Horowitz JD, Powers MB, Telch MJ. Psychological approaches in the
treatment of specific phobias: a meta-analysis. Clin Psychol Rev. 2008;28:1021–37.
Wolpe J. Reciprocal inhibition as the main basis of psychotherapeutic effects. AMA Arch Neurol
Psychiatry. 1954;72:205–26.
Wolpe J. Psychotherapy by reciprocal inhibition. Stanford: Stanford University Press; 1958.
Wolpe J, Rachmann S. Psychoanalytic “evidence”: a critique based on Freud’s case of Little Hans.
J Nerv Ment Dis. 1960;131:135–48.
Part III
Antidepressants
Gerd Laux
Contents
History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1102
Definition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1102
Classification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1102
New Classification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1103
Indications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1106
Guidelines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1106
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1107
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1107
Abstract
Antidepressant effects of the monoamine oxidase inhibitor iproniazid and
imipramine have been discovered serendipitously 1957/1958. Their effects of sero-
tonergic and noradrenergic transmission led to hypotheses of the origin of depression.
Various tri- and tetracyclic substances have been synthesized; in the 1980s, seroto-
nin-selective inhibitors like fluoxetine got first-line antidepressants. Dual acting,
combined serotonin-noradrenaline-reuptake inhibitors like venlafaxine have been
developed, so classification of antidepressants nowadays includes these main groups.
Recently, a new nomenclature for neuropsychopharmacological CNS medications
(NbN), based on neuroscience regarding pharmacology and mode of action, has been
proposed. Guidelines have been proposed by international and national societies like
National Institute for Health and Clinical Excellence (NICE), World Federation
of Societies of Biological Psychiatry (WFSBP), The International College of
G. Laux (*)
Institute of Psychological Medicine (IPM), Soyen, Germany
MVZ Waldkraiburg, Center of Neuropsychiatry, Waldkraiburg, Germany
Department of Psychiatry and Psychotherapy, Ludwig-Maximilians-University (LMU), Munich,
Germany
e-mail: ipm@ipm-laux.de
History
Definition
Classification
New Classification
Vilazodone ++ +
Vortioxetine +++ + +
5-HT serotonin receptors, α α-adreno receptors, ACh muscarinic acetylcholine receptors, H1 histamine receptors, DA dopamine receptors
1105
1106 G. Laux
Indications
Some antidepressants are also effective and recommended for the treatment of
• Anxiety disorders (generalized anxiety disorder, panic disorder, and social anx-
iety disorder)
• Obsessive-compulsive disorder
• Chronic pain syndromes.
Guidelines
Guidelines have been developed by international and national societies like the
British Association for Psychopharmacology, World Federation of Societies of
Biological Psychiatry (WFSBP), National Institute for Health and Clinical Excel-
lence (NICE), American Psychiatric Association (APA), Canadian Network for
Antidepressants: Definition, Classification, Guidelines 1107
Mood and Anxiety Treatments (CANMAT), the Australian and New Zealand
Guidelines, and the German S3 Guideline (Bauer et al. 2013, 2015; Cleare et al.
2015; DGPPN 2015).
Today SSRIs are seen as first-line antidepressants. Acute treatment is
recommended for 6–12 months, maintenance and relapse prevention for 2 years,
and longer depending from history of course (genetic load, and frequency of relapses
and recurrences).
Cross-References
References
Baldwin DS, Anderson IM, Nutt DJ, et al. Evidence-based pharmacological treatment of anxiety
disorders, post-traumatic stress disorder and obsessive-compulsive disorder: a revision of the
2005 guidelines from the British Association for Psychopharmacology. J Psychopharmacol.
2014;28:403–39.
Bauer M, Pfennig A, Severus E, et al. World Federation of Societies of Biological Psychiatry
(WFSBP) guidelines for biological treatment of unipolar depressive disorders, part 1: update
2013 on the acute and continuation treatment of unipolar depressive disorders. World J Biol
Psychiatry. 2013;14:334–85.
1108 G. Laux
Contents
Biochemical Mechanisms of Action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1112
Effect on the Monoaminergic Synapse . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1112
Effects in the Entire Neuronal System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1113
Reorganization Instead of Deficit Regulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1113
Effect by Inhibition of Reuptake . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1113
Effect on Receptors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1114
Effect by Delayed Decomposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1117
Clinical Efficacy Profiles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1117
Effect on the Noradrenergic System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1118
Effect on the Serotonergic System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1119
Effect on the Dopaminergic System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1120
Effect on the Glutamatergic System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1121
Adaptive Changes with Prolonged Use of Antidepressants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1124
Neuroplasticity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1124
Behavioural Pharmacological Effects of Antidepressants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1128
Cross-references . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1129
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1130
Abstract
The first antidepressants, discovered in the 1950s, and any successors directly
target the monoaminergic brain systems. Problems in this group, however, are
low effectiveness (only two-thirds of those treated responded to the therapy) and a
Abbreviations
5-HT 5-hydroxytryptamine/
serotonin
5-HT1A/5-HT1D/5-HT2/5-HT2A/5-HT2C/5-HT3 Serotonin receptors; subtypes
1A/1D/2/2A/2C/3
α1/α2/β Adrenergic receptors; subtypes
α1/α2/β
AC Adenylate cyclase
Akt Protein kinase B
AMP Adenosine monophosphate
AMPA α-amino-3-hydroxy-5-methyl-
4-isoxazolepropionic acid
ATP Adenosine Triphosphate
BDNF Brain-derived neurotrophic
factor
CA Cornu ammonis
Ca2+ Calcium
CaMK Ca2+/calmodulin-dependent
protein kinase
cAMP Cyclic adenosine monophos
phate
Cl Chloride
CNS Central Nervous System
COMT Catechol-O-methyltransferase
CREB cAMP response element-
binding protein
D1/D2 Dopamine receptors; subtypes
1/subtype 2
DA Dopamine
Antidepressants: Pharmacology and Biochemistry 1111
The mechanism of action of antidepressants has not yet been completely resolved
despite intensive research over the last 60 years. The neurobiochemical effects of
antidepressant substances are relatively well understood. However, there is still a
lack of clarity about the importance of these actions for the impact on depression,
since the biochemical basis of the disease itself has not yet been clarified due to the
lack of valid models of depression (Müller and Eckert 2017; Frazer 1997; Leonard
1995, 1996; Ebmeier et al. 2006).
Since the discovery of the thymoleptic effect of imipramine over 50 years ago,
the transmission at monoaminergic synapses of the CNS is the centre of atten-
tion in the research on mechanisms of action of antidepressants. Initially, the
biochemical hypotheses about the causes of depression related to a lack of
transmitters in the synaptic cleft, later also to a reduced sensitivity of postsyn-
aptic receptors. The mode of action of antidepressants was considered to be a
specific effect (see below) on these hypothetical deficits. To this day, there is
virtually no effective antidepressant that does not also affect monoaminergic
synapses (Berton and Nestler 2006). An exception is the fast-acting substance
ketamine.
Antidepressants: Pharmacology and Biochemistry 1113
The possibility that the effect of antidepressants cannot be ascribed to such specific
processes has lately been discussed increasingly. In 1983, Paioni et al. already
suspected that monoaminergic synapses are to be regarded as particularly favourable
intervention points for pharmacological influence on neuronal systems, namely, in
the sense of an impulse with the consequence of a slow normalization of the
previously disturbed overall regulatory system. This is supported by the detectable
adaptive changes in many neurotransmitter systems (see above) among all antide-
pressants and the fact that the central transmitters (norepinephrine, serotonin, dopa-
mine) triggered primarily by the antidepressants have a modulating action on many
different anatomical structures or many different functional processes of the CNS.
Stassen et al. (1996), likewise, conclude, based on the evaluation of the time courses
of antidepressants and placebo therapy, that the antidepressants tend to only accel-
erate a physiological normalization process.
This would mean that neither the acute pharmacological effects (amine reuptake
inhibition, MAO-inhibition) nor the latency-appearing adaptive changes in various
signal transduction mechanisms correct neurochemical deficits in depression but
rather are an expression of a rearrangement of certain functions of the central
neurotransmission, which ultimately lead to the resolution of the depression in the
patient. Since not only a lot of very different classes of substances (tricyclic drugs,
SSRIs, MAO-inhibitors, St. John’s wort extract) but also therapeutic measures like
electroconvulsive therapy (ECT) and rapidly effective interventions like sleep
deprivation and ketamine infusion converge on the level of adaptive changes,
this “reorganization hypothesis” could explain why many pharmacologically
very different antidepressant therapy options can show very similar clinical effects.
An important argument for this hypothesis is the fact that these adaptive changes
also exhibit a certain latency (1–2 weeks) in animal experiments, which in turn
better correlates with the delayed antidepressant action on the patient than the acute
effects.
The biochemical profile of the antidepressants is mainly derived from the acute
effects on the inhibition of NE and serotonin reuptake and the latency appearing
adaptive changes in certain central signal transduction mechanisms (Fig. 1 and
Table 1). We distinguish
Fig. 1 Effects of different classes of antidepressants in (a) central noradrenergic and (b) central
serotonergic synapses (Müller and Eckert 2017)
Bupropion, which is also approved for the treatment of depression, is the only
antidepressant that inhibits the dopamine reuptake to a relevant extent under
therapeutic conditions (Stahl et al. 2004). Another exception is St. John’s wort
extract, which inhibits the synaptosomal uptake of serotonin, norepinephrine and
dopamine to about the same extent via the most important ingredient hyperforin
(Müller 2003).
Effect on Receptors
Besides this characteristic, the action of the individual antidepressants on the pre-
and postsynaptic receptors needs to be considered for the profile. However, those are
less responsible for the antidepressant effect (exception α2-antagonsim, primary
mechanisms in mirtazapine) than for the desired – but especially for the many
undesired vegetative – side effects (sedation, anxiolysis) of the respective antide-
pressant (Table 2).
Agomelatine is the first melatonin receptor (MT1, MT2) agonist and simulta-
neously antagonist on the serotonin receptor subtype 5-HT2C, so that the substance at
least intervenes in the serotonergic neurotransmission like most other antidepres-
sants (Fuchs et al. 2006).
Table 1 Inhibition constants (Ki-values in nmol) and receptor profiles of typical antidepressants for the inhibition of neuronal reuptake of norepinephrine and
serotonin (Müller and Eckert 2017)
Drug NE uptake 5-HT uptake 5-HT selectivitya H1 receptor M receptor α1 receptor α2 receptor 5-HT2 receptor
TCA
Amitriptyline 14 84 0,17 1 10 24 940 18
Clomipramine 28 5 5,6 31 37 38 >1000 54
Desipramine 0,6 180 0,003 60 66 100 >1000 350
Doxepin 18 220 0,08 0,2 23 24 >1000 27
Imipramine 14 41 0,3 37 46 32 >1000 150
Mirtazapine >1000 >1000 – 0,5 500 500 10 5
Nortriptyline 2 154 0,01 6 37 55 >1000 41
Trimipramine 510 >1000 0,02 0,3 58 24 680 32
Viloxazine 170 >1000 0,01 >1000 1000 >1000 1000 >1000
SSRI
Antidepressants: Pharmacology and Biochemistry
Table 2 Possible adverse drug reactions of inhibition of neuronal reuptake systems and of
blockade of neuroreceptors (Müller and Eckert 2017)
Reuptake systems Adverse drug reactions
NE reuptake Enhancement of the effects of sympathomimetics
Tachycardia
RR "
Restlessness, tremor
Erectile and ejaculation dysfunctions
5-HT reuptake Gastrointestinal dysfunction, nausea, vomiting
Restlessness, sleep disturbances
Extrapyramidal motor symptoms (?)
Reduced appetite, weight loss
Headache
Sexual dysfunctions
DA reuptake Psychomotor activation
Triggering resp. reinforcement psychosis
Anti-Parkinson effect
Neuroreceptors Adverse drug reactions
M Dry mouth
Blurred vision, accommodative dysfunction
Sinus tachycardia
Obstipation
Urinary retention, voiding disorders
Disturbances of memory
H1 Sedation, fatigue, sleepiness
Enhancement of other central damping substances
Weight gain (?)
α1 Orthostasis, RR #
Dizziness, drowsiness, sedation
Reflex tachycardia (+ α2 blocking?)
Enhancement of the effect of other α1 inhibitors
D2 Extrapyramidal motor symptoms
Prolactin "
Sexual dysfunctions
5-HT2 Increased appetite, weight gain
RR #
5-HT3 Antiemetic effect
Anxiolysis (?)
Other substances occupying a special position in this concept are tianeptine, the
new substance vortioxetine, as well as ketamine (McEwen et al. 2010; Otte 2014;
Müller 2015). Paradoxically, tianeptine does not lead to inhibition but rather to an
increase in serotonin uptake. However, like other serotonergic antidepressants,
tianeptine has similar adaptive effects at the serotonergic synapse and shows very
distinct actions on neuroplasticity mechanisms. In addition, it directly modulates
glutamatergic neurotransmission, especially at the AMPA receptor (McEwen et al.
2010).
Antidepressants: Pharmacology and Biochemistry 1117
Vortioxetine inhibits serotonin reuptake and has agonistic and antagonistic effects
on various serotonin receptors, which synergistically contribute to the antidepressant
effect, explaining the positive effect on cognitive dysfunction in depressed patients.
This improvement of cognition distinguishes the substance from other antidepressants.
Ketamine, a drug from the class of narcotics, has an antagonistic effect on the
glutamatergic NMDA receptor complex. It inhibits the NMDA-dependent release of
acetylcholine and appears to possess an additional opioid-like active component.
In the human brain, serotonin, norepinephrine and dopamine are degraded by the
mitochondrial enzyme monoamine oxidase (MAO) via oxidative deamination. The
inhibition of this enzyme delays the decomposition of the above-mentioned neuro-
transmitters and thus leads to an increased synaptic availability (Figs. 1 and 3).
Currently, the substances tranylcypromine, an irreversible, non-selective inhibitor,
and moclobemide, a reversible, MAO-A-selective inhibitor, are of clinical relevance
in the treatment of depression. Because of the dietary restrictions required and the
significant risk of interaction, classical MAO inhibitors have been used mainly to
treat patients with atypical depression and treatment-resistant patients. Reversible
MAO inhibitors do not have these restrictions. They are therefore used for all forms
of depression (Riederer et al. 2002).
severe
medium
weak
missing
Early hypotheses, now more than 50 years old, also assumed that serotonin defi-
ciency in the synaptic cleft in the CNS is part of the pathophysiology of depression.
Indeed, there is evidence that lowering the brain 5-HT concentration, by the so-called
tryptophan depletion, can cause clinical depressive symptomatology in those with
risk factors for depression (Cowen 2008). Moreover, post-mortem analysis and
positron-emission tomography showed findings of a reduction in serotonin trans-
porter binding density in the raphe nuclei, amygdala, and other brain areas in
depressed patients, especially suicidal individuals (Nemeroff and Owens 2009).
But it is still not clear whether the above-mentioned changes in the brain’s 5-HT
system is the cause of patients becoming depressed, for example, when they are
exposed to psychosocial stress, or whether these changes are remnants of earlier
acute depressive episodes and their treatment (Cowen 2008). Nevertheless, the
increased availability of synaptic serotonin seems to also be an important initial
mechanism of action of many antidepressants. In today’s often used group of SSRIs,
it is the sole initial effect. However, it is still important to bear in mind that the
increase in serotonin concentration is rapid, but the antidepressant effect only occurs
after some time. One assumption is therefore that in the successful therapy of
depressive patients, longer-lasting plastic processes are initiated or at least
influenced by serotonergic or noradrenergic signalling pathways that are associated
with the alteration of existing or the formation of new synaptic contacts and new
neurons (Riederer et al. 2002).
Reuptake inhibition. Various antidepressants selectively (e.g. SSRI) or
non-selectively (e.g. amitriptyline, clomipramine, imipramine and venlafaxine)
inhibit the reuptake of serotonin into the presynaptic nerve ending.
1120 V. Efinger et al.
5-HT1A receptor activation. There are also 5-HT1A autoreceptors in the serotoner-
gic system, which regulate the transmitter release analogously to the α2 receptors in the
noradrenergic system. In addition to the anxiolytic buspirone, there are some devel-
opmental substances (e.g. gepirone) that, as partial agonists, reduce the activity of the
serotonergic neurons by activating the 5-HT1A autoreceptors but directly activate
postsynaptic 5-HT1A receptors and have therefore been tested as antidepressants.
Inhibition of decomposition. The intra- and extraneuronal decomposition of
serotonin also occurs via MAO-A, whose inhibition by selective (moclobemide)
and non-selective (tranylcypromine) inhibitors leads to an increase in the concen-
tration of serotonin in the synapse respectively in the synaptic cleft.
Central α2 antagonism. These antagonists, e.g. mirtazapine, block the noradren-
ergic inhibition of serotonergic neurons and thus lead to an increased synaptic
activity of the serotonergic system.
5-HT2 antagonism. This antagonism, e.g. also mediated by mirtazapine, can lead
to an increase in the neuronal serotonin release and thus to an increased 5-HT1A
activation via interconnection mechanisms that have not yet been completely clar-
ified. This mechanism also applies to some atypical antipsychotics.
Reuptake activation. In contrast to most of the other antidepressants, which
influences the serotonergic system, tianeptine seems to enhance the reuptake of
serotonin into the presynapse, thus decreasing the serotonin concentration in the
synaptic cleft (Simoni et al. 1992). Nevertheless, tianeptine shows very good
antidepressant efficacy comparable to other substances (Kasper and McEwen 2008).
The dopaminergic system (Fig. 3) is very similar to the noradrenergic system (dopa-
mine is the precursor of norepinephrine; the synthesis pathway of the two neurotrans-
mitters is the same up to this stage!). The effects of antidepressants on this system and
their importance in influencing depression have not been well examined. Most of the
antidepressants have no relevant action on the neuronal DA reuptake. Two exceptions
are nomifensine and amineptine, which have been withdrawn from the market because
of serious side effects, and one newer substance: bupropion. St. John’s wort extract
also inhibits neuronal DA uptake to about the same extent as NE uptake (Müller 2003).
The following dopaminergic mechanisms are relevant for an antidepressant effect:
reuptake inhibition, inhibition of dopamine decomposition and presynaptic receptors.
Reuptake inhibition. Bupropion selectively inhibits the reuptake of dopamine and
norepinephrine. It has no effects on other reuptake systems or neurotransmitter
receptors; it is thus not associated with significant sedation, cognitive or anticholin-
ergic side effects. This pharmacologic profile is unique to bupropion, which is
currently the only available NDRI shown to increase the neurotransmission of
dopamine in humans at clinically relevant doses (Stahl et al. 2004).
Inhibition of decomposition. In humans, dopamine is degraded intra- and extra-
neuronally by MAO-A but mainly by MAO-B over oxidative deamination (Fritze
et al. 1989). Therefore, especially the older, non-selective MAO inhibitor
tranylcypromine leads to increased synaptic availability of dopamine.
Antidepressants: Pharmacology and Biochemistry 1121
occur within a few hours and even in patients considered treatment-resistant lasting
up to one week. These findings were later confirmed in a double-blind, placebo-
controlled study by Zarate et al. (2006) and many other trials by different research
groups (Sanacora et al. 2017; Wilkinson et al. 2017; Newport et al. 2015). Thus,
ketamine represents a new chance of recovery, especially for therapy-resistant
depression patients. However, the widespread use of ketamine poses a problem
due to its side effects. Besides psychomimetic and dissociative effects, ketamine
has a high potential for abuse. Nevertheless, the findings can be used for the
development of ketamine-like drugs with fewer adverse reactions and a sustained
action, as well as agents that act at other sites within the glutamatergic system.
Examples of new developments are NMDA allosteric modulators (e.g. brexanolone
and SAGE-217) or mGluR2/3 autoreceptor inhibitors (notably LY341,495 and
MGS0039) (Duman 2018). It is believed that the rapid effect of ketamine is due to
the release of growth factors which promote the formation and function of synapses
in the hippocampus and layer V pyramidal neurons in the medial prefrontal cortex
(mPFC) (Björkholm and Monteggia 2016; Duman 2018). This restores connections
that have been broken down by stress mechanisms due to disease. The mechanism of
action is not yet fully understood. The assumption is that a rapid, transient increase in
glutamate (“glutamate burst”) occurs via disinhibition of the glutamatergic trans-
mission that is disturbed during depression. This disinhibition is achieved by
blocking NMDA receptors on GABAergic interneurons by ketamine. The increase
in glutamate in the synaptic cleft activates postsynaptic AMPA receptors, which in
turn leads to increased formation and release of BDNF. BDNF for its part activates
the protein mTORC1 via TrkB and Akt, which ultimately results in the effects
described above ((Duman 2014), Fig. 4). Another assumption suggests that
NMDA receptors on glutamatergic post-synapses are also inhibited by ketamine.
This inhibition leads to a deactivation of eEF2, which normally suppresses BDNF
levels (Krystal et al. 2019). In addition to its influence on the hippocampus and
mPFC, there is evidence that ketamine also has an effect on the lateral habenula
(LHb), a brain region that inhibits major reward centres (Yang et al. 2018). Blocking
the LHb with ketamine cancels out the anti-reward effect. Overall, it is postulated
that the actions on the lateral habenula lead to the rapid onset of antidepressant
effects, whereas the effects on the glutamatergic synapses in the hippocampus and
mPFC lead to the sustained actions of ketamine (Duman 2018). This postulate is
corroborated by the fact that ketamine has a half-life of only about 3 h, suggesting
that the long-lasting antidepressant effects are not mediated by NMDA receptor
blockade per se but rather by synaptic plasticity mechanisms (Björkholm and
Monteggia 2016).
Esketamine, the S-enantiomer of ketamine, has a three- to fourfold higher affinity
to NMDA receptors than R-ketamine. Two phase II RCTs were able to show that
intranasally administered esketamine diminished depressive symptoms already 4 h
post-dose and suicidal ideation in depressed patients. Based on the positive results,
the FDA granted a breakthrough therapy designation for intranasal esketamine, with
the result that it entered phase III development (Garay et al. 2017). In March 2019,
an esketamine nasal spray was approved by the FDA (U.S. Food and Drug
Antidepressants: Pharmacology and Biochemistry 1123
Fig. 4 Glutamatergic synapse and the therapeutic mechanisms of action of the rapid-acting
antidepressant ketamine. (Modified after Duman 2018; Krystal et al. 2019)
Administration 2019). Since the doses of esketamine required for depression may
cause dissociation and delirium, the approval is restricted for the therapy of
treatment-resistant patients with MDD in combination with an oral antidepressant
(Okada et al. 2020).
Scopolamine, a non-selective acetylcholine muscarinic receptor antagonist, also
exhibits promising antidepressant effects. Two clinical trials from 2006 and 2010
reported that a single low dose of scopolamine (4 μg/kg i.v.) already produces
antidepressant responses within the first 3 days (Furey and Drevets 2006; Drevets
and Furey 2010). A much earlier study from 1991 using other administration
(e.g. intramuscular) and dosing routes already suggested that scopolamine can
produce his antidepressant effect within several hours after treatment (Gillin et al.
1991). Moreover, it seems that the antidepressant actions of scopolamine are
increased with repeated administrations (Ellis et al. 2014). Interestingly, it seems
that scopolamine appears to mediate its rapid therapeutic action also via an increased
glutamatergic transmission (Duman 2014). Studies with rodent models were able to
show that scopolamine, just like ketamine, causes a burst of glutamate, with the
result of an increased release of BDNF and an increased mTORC1 signalling, so that
the synapse formation is ameliorated (Voleti et al. 2013; Ghosal et al. 2018).
Furthermore, there are findings that the effect of scopolamine also occurs via
disinhibition of the glutamatergic transmission through the blockage of M1 receptors
on GABAergic interneurons (Wohleb et al. 2016).
Besides (es)ketamine and scopolamine, rapastinel is a new substance of interest.
Rapastinel acts as a partial agonist at the allosteric glycine site of the NMDA
1124 V. Efinger et al.
receptor complex (Moskal et al. 2017). The consequences are an increase in synapse
number and functionality in the mFPC, as well as an increased BDNF release and
mTORC1 activation. In addition to preclinical studies in different rodent models
(including FST, NSFT, LH, chronic mild and social defeat stress), a double-blind
randomised trial showed that a single dose of rapastinel leads to antidepressant
effects lasting for about 7 days (Burgdorf et al. 2013; Kato et al. 2018; Preskorn
et al. 2015). Currently, it is in phase III clinical trials and has received breakthrough
classification.
Tianeptine also seems to influence the glutamatergic system, but in a modulating,
not in a direct, way. It has been shown that tianeptine leads, over various kinases, to
increased phosphorylation of the AMPA receptor subunit GInR1, which is partly
responsible for its antidepressant effect (Müller 2016).
Antidepressants have a number of different primary effects on the CNS. What all
these acute mechanisms of action have in common is that they occur immediately
after application and therefore do not correspond to the delayed development of the
antidepressant action in the patient. It is therefore assumed today that secondary to
these acute influences on the central neurotransmission, adaptive changes occur in
response to the acute intervention, especially at the level of receptors and receptor-
coupled transduction mechanisms. Here, a downregulation of the density and sen-
sitivity of central β-receptors has been best investigated (β-downregulation) but has
neither been found for all substances.
GABAergic mechanisms, glutamatergic mechanisms, the sensitivity of glucocor-
ticoid receptors and the regulation of transcription factors may also be affected by
such adaptive changes. To date, it is not possible to associate one of these adaptive
changes exclusively with the antidepressant efficacy, but rather see these changes,
which can be determined in animal experiments as an expression of an adaptation or
functional plasticity, which is possibly a direct correlate of the antidepressant effect
(Müller and Eckert 2017).
Neuroplasticity
Finally, the formation of new nerve cells, the neurogenesis, which in adult humans can
only be found in the olfactory bulb, the dentate gyrus (DG) of the hippocampus and the
subventricular zone, is also part of structural plasticity. CREB and other inducible
transcription factors induce effector genes that contribute to the stabilization of struc-
tural plasticity. The growth factor BDNF (“brain-derived neurotrophic factor”) is one
of the target genes or target proteins of CREB. Thus, BDNF not only can induce CREB
(Fig. 5) but also is itself a target of CREB induction. BDNF represents an important
growth factor for neuronal function in the CNS. Under the influence of BDNF, there is
dendrite and synapse growth of neuronal cells, whereas without the stimulating effect
of BDNF, atrophy up to the risk of cell death occurs. The fact that BDNF is also a target
gene of the transcription factor CREB has now led to the speculation that chronic
treatment with antidepressants may alter the concentration of BDNF. Interestingly, this
could be confirmed; various studies were able to show that under subchronic treatment
of different antidepressants, the BDNF-m-RNA is upregulated in various brain areas,
but mainly in the hippocampus (Duman et al. 1997, 1999; Duman 2014). Furthermore,
there are data which suggest that tianeptine increases the expression of BDNF and
other nerve growth factors in the hippocampus and amygdala and thus promotes
neuroplasticity (Alfonso et al. 2006; Reagan et al. 2007).
Neurodegenerative hypothesis of depression. Under normal conditions,
neurogenesis and normal growth of dendrites and synapses can be seen in the
adult brain. For many years now, there have been indications that chronic stress,
combined with upregulation of glucocorticoids, can lead along with other biochem-
ical changes to atrophic or degenerative changes, especially in the CA3 regions of the
hippocampus (Duman et al. 1999; Sapolsky et al. 1985). Seeing that the hippocam-
pus has been established to play an important role in processing psychological stress,
it immediately suggests itself to suppose that structural hippocampal changes under-
lie the pathophysiology of major depression (Dranovsky and Hen 2006). Together
with current findings from modern imaging clinical research, in which certain
indications of neurodegenerative changes, notably a decreased volume, in the
hippocampus of depressed patients are described, the finding mentioned above has
led to the so-called neurodegenerative hypothesis of depression, respectively, the
neurotrophic hypothesis of antidepressant effects (Rajkowska et al. 1999; Soares and
Mann 1997; Ebmeier et al. 2006). Interestingly – and this brings us back to the
growth factor BDNF – the hippocampal concentration of BDNF also tends to be
reduced under chronic stress. Other factors, especially genetic ones, which can be
relevant for depression as well, also seem to have a negative influence on the growth
of CA3 neurons. For example, the Met-allele, the result of the genetic polymorphism
Val/Met at codon 66, blocks the release of BDNF, with the consequence that the
carriers of this allele have an increased vulnerability to stress and depression
(Gerhard et al. 2016). But overall, it is still largely a matter of speculation as to
which factors interact in depressive patients.
Neurotrophic hypothesis of antidepressant effects. Under certain conditions, an
improved survival rate of hippocampal neurons with improved dendrite growth and
synapse formation and, in addition, a neoformation of nerve cells (neurogenesis)
have been seen under biological antidepressant therapies (Malberg et al. 2000).
Antidepressants: Pharmacology and Biochemistry 1127
It is known that serotonin over the 5-HT1A receptor and norepinephrine have
positive effects on hippocampal neurogenesis ((Santarelli et al. 2003; Dranovsky and
Hen 2006), Fig. 6d). It is therefore not surprising that SSRIs (e.g. fluoxetine)
increase neurogenesis in the ventral dentate gyrus (Duman 2004). A longitudinal
MRI study from 2013 could show that SSRI treatment over 8 weeks of patients with
major depression led to hippocampal growth, which is in consistency with the results
of a study using high-resolution MRI, where medicated patients with depression had
bigger dentate gyri than unmedicated patients (Huang et al. 2013; Arnone et al.
2013). There is also evidence from various other studies that, next to SSRIs, further
antidepressants (SNRIs, TCAs, MAO inhibitors and ketamine) increase
neurogenesis (Levy et al. 2018). In addition to their effects on neurogenesis,
antidepressants can also regulate other types of neuroplasticity (Fig. 6a–c). Chronic
mild stress has been shown to reduce volume as well as length and branching of
dendrites within the DG, CA3 region and the PFC in rats (Bessa et al. 2009). But
there is evidence that antidepressant treatment increases the variability and turnover
in the branches of dendrites and axons (Castrén and Hen 2013). The chronic use of
fluoxetine, for example, simultaneously increases the elongation and retraction of
branch tips in the mouse visual cortex (Chen et al. 2011). Moreover, imipramine and
tianeptine are able to restore dendritic arborisation to prestress levels (Patrício et al.
2015). Chronic stress also leads to the loss of dendritic spines and synaptic contacts
in the hippocampus and the PFC (McEwen 1999; Cerqueira et al. 2007; Hajszan
et al. 2010; Radley et al. 2006). Fluoxetine treatment, however, increased spine
number back to the baseline level (Chen et al. 2011). In addition to synapse number,
synaptic strength is also regulated by antidepressant drugs. Notably, fluoxetine and
tianeptine are able to modulate long-term potentiation (LTP) within the hippocampus
Fig. 6 Neuronal plasticity. In addition to synaptic plasticity at the level of signal transduction
(receptors, transporters, enzymes, etc.), neuronal plasticity also includes functional aspects of the
synapse such as (a) long-term potentiation (a mechanism of signal impregnation) and (b) histolog-
ically tangible changes (such as density and structure of postsynaptic spines), (c) length and
branching of neurites and (d) neoformation of neurons (neurogenesis), which is, however, mainly
reserved for two small areas of the hippocampus (Müller 2016)
1128 V. Efinger et al.
(Rocher et al. 2004; Wang et al. 2008). The presumption is that this finding may be
related to the “dematuration” process observed in the dentate granule neurons
(Kobayashi et al. 2010), indicating that antidepressant treatment reactivated
juvenile-like plasticity in the brain (Maya Vetencourt et al. 2008; Karpova et al.
2011).
Up to now, the effects of antidepressants on several mechanisms of neuroplasticity
were mainly seen as a consequence of the elevated synaptic concentrations of the amine
neurotransmitters leading to enhanced activation of postsynaptic receptors of the respec-
tive amine neurotransmitters (Fig. 5). However, there is increasing evidence that some
antidepressants may additionally affect neuroplasticity by directly activating the signal
cascades finally leading to enhanced levels of neurotrophic factors like BDNF (Fig. 5)
and NGF as shown for tianeptine (McEwen et al. 2010), fluvoxamine (Ishima et al.
2014) and fluoxetine (Levy et al. 2019). Similarly, several other drugs with antidepres-
sant properties have been shown to possess direct effects on neuroplasticity like
hyperforin, the main active ingredient of hypericum extract (Heiser et al. 2013) and
the special lavender oil Silexan (Friedland et al. 2021).
The results of all the studies mentioned above point towards that the adaptive
effects of antidepressants may repose on influencing the neuroplasticity in the
dentate gyrus and several other brain regions. Nevertheless, this hypothesis is still
not proven beyond doubt, and there are still a lot of unanswered questions, so further
experiments and clinical trials are needed. But the findings can help to better
understand the pathophysiology of major depression and to develop new
antidepressants.
Just like the chemical structure, the pharmacological properties of the preparations
used as antidepressants are quite different. The basic problem of research in this area
is the lack of an adequate animal model of depression. After about 30 years of
research, the following animal models are mainly used to assess the antidepressant
properties (Willner 1984). However, only some of these are actual “animal models of
depression,” while others are based on the pharmacological properties of tricyclics
and primarily record their NE- or serotonin-enhancing effects.
Spontaneous behaviour. Antidepressants, especially tricyclics, inhibit the spon-
taneous behaviour in animals, and they show centrally depressing actions at medium
and higher doses. So, this model tests sedative, but not antidepressant properties. In
contrast to neuroleptics, which produce similar effects, antidepressants cause a slight
to strong increase in excitability after increasing doses.
Reserpine antagonism. Antidepressants, particularly tricyclics, cancel out the
effects caused by reserpine (psychomotor inhibition, reduced autonomic reactions).
Potentiation of various catecholamine effects. Tricyclic antidepressants intensify,
probably by inhibiting the reabsorption of norepinephrine and the resulting increase
in the concentration at the receptor, the increases in blood pressure caused by this
transmitter.
Antidepressants: Pharmacology and Biochemistry 1129
Separation model. If young animals are socially isolated (separation from parent
animals), after some time, there is a considerable loss of activity and significant
changes in posture. These behaviours are reversed by antidepressants. However, this
model is not specific, as similar effects can also be achieved with alcohol, benzodi-
azepines and opioids.
Behavioural despair test. This “swimming test” determines how long the animals
swim after being immersed in a small water-filled container before they adopt an
immobile position. Antidepressants prolong the swimming phase, but this is also
achieved with antihistamines and anticholinergics.
Chronic stress test. In this experiment, rats or mice are exposed to chronic stress
(fasting, electric shocks, isolation, immersion in cold water) for a longer period of
time. The resulting reduced exploratory behaviour is increased again by antidepres-
sants, especially tricyclics.
Learned helplessness test. In this test, the animals learn “helplessness” through
unavoidable stimuli, which even after the experimental situation is ceased, no longer
enables them to influence tasks through their own behavioural reactions. This
helplessness is reversed by antidepressants, but not by neuroleptics and
tranquillizers.
Bulb ectomised rats. After surgical removal of the olfactory bulb, rats show
various behavioural changes that are similar to depression and can be corrected by
antidepressants.
None of these models is sufficient on their own to reliably predict an antidepres-
sant effect in humans. Though the hit rate can be significantly increased by a
combination of several of the tests mentioned above, all of which show a certain
“depression analogy.” However, to date, with all of these models, it has not yet been
possible to develop substances that exceed the usual approx. 70% response rate at
initial use.
Cross-references
References
Alfonso J, Frick LR, Silberman DM, Palumbo ML, Genaro AM, Frasch AC. Regulation of
hippocampal gene expression is conserved in two species subjected to different stressors and
antidepressant treatments. Biol Psychiatry. 2006;59(3):244–51.
Arnone D, McKie S, Elliott R, Juhasz G, Thomas EJ, Downey D, et al. State-dependent changes in
hippocampal grey matter in depression. Mol Psychiatry. 2013;18(12):1265–72.
Berman RM, Cappiello A, Anand A, Oren DA, Heninger GR, Charney DS, Krystal
JH. Antidepressant effects of ketamine in depressed patients. Biol Psychiatry. 2000;47(4):351–4.
Berton O, Nestler EJ. New approaches to antidepressant drug discovery: beyond monoamines. Nat
Rev Neurosci. 2006;7:137–51.
Bessa JM, Ferreira D, Melo I, Marques F, Cerqueira JJ, Palha JA, et al. The mood-improving
actions of antidepressants do not depend on neurogenesis but are associated with neuronal
remodeling. Mol Psychiatry. 2009;14(8):764–73, 739.
Björkholm C, Monteggia LM. BDNF – a key transducer of antidepressant effects. Neuropharma-
cology. 2016;102:72–9.
Burgdorf J, Zhang X, Nicholson KL, Balster RL, Leander JD, Stanton PK, et al. GLYX-13, a
NMDA receptor glycine-site functional partial agonist, induces antidepressant-like effects
without ketamine-like side effects. Neuropsychopharmacology. 2013;38(5):729–42.
Castrén E, Hen R. Neuronal plasticity and antidepressant actions. Trends Neurosci. 2013;36(5):
259–67.
Cerqueira JJ, Mailliet F, Almeida OFX, Jay TM, Sousa N. The prefrontal cortex as a key target of
the maladaptive response to stress. J Neurosci. 2007;27(11):2781–7.
Chen AC, Shirayama Y, Shin KH, Neve RL, Duman RS. Expression of the cAMP response element
binding protein (CREB) in hippocampus produces an antidepressant effect. Biol Psychiatry.
2001;49:753–62.
Chen JL, Lin WC, Cha JW, So PT, Kubota Y, Nedivi E. Structural basis for the role of inhibition in
facilitating adult brain plasticity. Nat Neurosci. 2011;14(5):587–94.
Cowen PJ. Serotonin and depression: pathophysiological mechanism or marketing myth? Trends
Pharmacol Sci. 2008;29(9):433–6.
de Simoni MG, de Luigi A, Clavenna A, Manfridi A. In vivo studies on the enhancement of
serotonin reuptake by tianeptine. Brain Res. 1992;574(1–2):93–7.
Dranovsky A, Hen R. Hippocampal neurogenesis: regulation by stress and antidepressants. Biol
Psychiatry. 2006;59(12):1136–43.
Drevets WC, Furey ML. Replication of scopolamine’s antidepressant efficacy in major depressive
disorder: a randomized, placebo-controlled clinical trial. Biol Psychiatry. 2010;67(5):432–8.
Duman RS. Depression: a case of neuronal life and death? Biol Psychiatry. 2004;56(3):140–5.
Duman RS. Pathophysiology of depression and innovative treatments: remodeling glutamatergic
synaptic connections. Dialogues Clin Neurosci. 2014;16:11–27.
Duman RS. Ketamine and rapid-acting antidepressants: a new era in the battle against depression
and suicide. F1000Res. 2018;7:F1000.
Antidepressants: Pharmacology and Biochemistry 1131
Duman RS, Heninger GR, Nestler EJ. A molecular and cellular theory of depression. Arch Gen
Psychiatry. 1997;54:597–606.
Duman RS, Malberg J, Thorne J. Neural plasticity to stress and antidepressant treatment. Biol
Psychiatry. 1999;46:1181–91.
Duncan GE, Johnson KB, Breese GR. Topographic patterns of brain activity in response to swim
stress assessment by 2-deoxyglucose uptake and expression of fos-like immunreacivity. J
Neurosci. 1993;13:3932–4.
Ebmeier KP, Donaghex C, Steele JD. Recent developments and current controversies in depression.
Lancet. 2006;367:153–67.
Ellis JS, Zarate CA, Luckenbaugh DA, Furey ML. Antidepressant treatment history as a predictor of
response to scopolamine: clinical implications. J Affect Disord. 2014;162:39–42.
Frazer A. Antidepressants. J Clin Psychiatry. 1997;58(Suppl 6):9–25.
Friedland K, Silani G, Schuwald A, Stockburger C, Koch E, Nöldner M, Müller WE. Neurotrophic
properties of Silexan, an essential oil from the flowers of lavender-preclinical evidence for
antidepressant-like properties. Pharmacopsychiatry. 2021;54(1):37–46.
Fritze J, Koronakis P, Konradi C, Riederer P. Isoelectric focusing of monoamine oxidase subtypes
as identified by MAO inhibitors. Eur J Pharmacol Mol Pharmacol. 1989;172(2):147–54.
Fuchs E, Simon M, Schmelting B. Pharmacology of a new antidepressant: benefit of the implication
of the melatonergic system. Int Clin Psychopharmacol. 2006;21(Suppl 1):17–20.
Furey ML, Drevets WC. Antidepressant efficacy of the antimuscarinic drug scopolamine: a
randomized, placebo-controlled clinical trial. Arch Gen Psychiatry. 2006;63(10):1121–9.
Garay RP, Zarate CA, Charpeaud T, Citrome L, Correll CU, Hameg A, Llorca P-M. Investigational
drugs in recent clinical trials for treatment-resistant depression. Expert Rev Neurother. 2017;17
(6):593–609.
Gerhard DM, Wohleb ES, Duman RS. Emerging treatment mechanisms for depression: focus on
glutamate and synaptic plasticity. Drug Discov Today. 2016;21(3):454–64.
Ghosal S, Bang E, Yue W, Hare BD, Lepack AE, Girgenti MJ, Duman RS. Activity-dependent
brain-derived neurotrophic factor release is required for the rapid antidepressant actions of
scopolamine. Biol Psychiatry. 2018;83(1):29–37.
Gillin JC, Sutton L, Ruiz C, Darko D, Golshan S, Risch SC, Janowsky D. The effects of
scopolamine on sleep and mood in depressed patients with a history of alcoholism and a normal
comparison group. Biol Psychiatry. 1991;30(2):157–69.
Hajszan T, Szigeti-Buck K, Sallam NL, Bober J, Parducz A, Maclusky NJ, et al. Effects of estradiol
on learned helplessness and associated remodeling of hippocampal spine synapses in female
rats. Biol Psychiatry. 2010;67(2):168–74.
Heiser JH, Schuwald AM, Sillani G, Ye L, Müller WE, Leuner K. TRPC6 channel-mediated neurite
outgrowth in PC12 cells and hippocampal neurons involves activation of RAS/MEK/ERK,
PI3K, and CAMKIV signaling. J Neurochem. 2013;127(3):303–13.
Hope BT, Kelz MB, Duman RS, Nestler EJ. Chronic electroconvulsive seizure (ECS) treatment
results in expression of a long-lasting AP-1 complex in brain with altered composition and
characteristics. J Neurosci. 1994;14(7):4318–28.
Huang Y, Coupland NJ, Lebel RM, Carter R, Seres P, Wilman AH, Malykhin NV. Structural
changes in hippocampal subfields in major depressive disorder: a high-field magnetic resonance
imaging study. Biol Psychiatry. 2013;74(1):62–8.
Ishima T, Fujita Y, Hashimoto K. Interaction of new antidepressants with sigma-1 receptor
chaperones and their potentiation of neurite outgrowth in PC12 cells. Eur J Pharmacol.
2014;727:167–73.
Karpova NN, Pickenhagen A, Lindholm J, Tiraboschi E, Kulesskaya N, Agústsdóttir A, et al. Fear
erasure in mice requires synergy between antidepressant drugs and extinction training. Science.
2011;334(6063):1731–4.
Kasper S, McEwen BS. Neurobiological and clinical effects of the antidepressant tianeptine. CNS
Drugs. 2008;22(1):15–26.
Kato T, Fogaça MV, Deyama S, Li X-Y, Fukumoto K, Duman RS. BDNF release and signaling are
required for the antidepressant actions of GLYX-13. Mol Psychiatry. 2018;23(10):2007–17.
1132 V. Efinger et al.
Kielholz P. Diagnose und Therapie der Depression für den Praktiker. 3rd ed. München: J. F.
Lehmanns Verlag; 1971.
Kobayashi K, Ikeda Y, Sakai A, Yamasaki N, Haneda E, Miyakawa T, Suzuki H. Reversal of
hippocampal neuronal maturation by serotonergic antidepressants. Proc Natl Acad Sci U S
A. 2010;107(18):8434–9.
Krystal JH, Abdallah CG, Sanacora G, Charney DS, Duman RS. Ketamine: a paradigm shift for
depression research and treatment. Neuron. 2019;101(5):774–8.
Leonard BE. Mechanisms of action of antidepressants. CNS Drugs. 1995;4(Suppl 1):1–12.
Leonard BE. New approaches to the treatment of depression. J Clin Psychiatry. 1996;57(Suppl
4):26–33.
Levy MJF, Boulle F, Steinbusch HW, van den Hove DLA, Kenis G, Lanfumey L. Neurotrophic
factors and neuroplasticity pathways in the pathophysiology and treatment of depression.
Psychopharmacology. 2018;235(8):2195–220.
Levy MJF, Boulle F, Emerit MB, Poilbout C, Steinbusch HWM, van den Hove DLA, et al. 5-HTT
independent effects of fluoxetine on neuroplasticity. Sci Rep. 2019;9(1):6311.
Malberg JE, Blendy JA. Antidepressant action: to the nucleus and beyond. Trends Pharmacol Sci.
2005;26(12):631–8.
Malberg JE, Eisch AJ, Nestler EJ, Duman RS. Chronic antidepressant treatment increases
neurogenesis in adult rat hippocampus. J Neurosci. 2000;20(24):9104–10.
Maya Vetencourt JF, Sale A, Viegi A, Baroncelli L, de Pasquale R, O’Leary OF, et al. The
antidepressant fluoxetine restores plasticity in the adult visual cortex. Science. 2008;320
(5874):385–8.
McEwen BS. Stress and hippocampal plasticity. Annu Rev Neurosci. 1999;22:105–22.
McEwen BS, Chattarji S, Diamond DM, Jay TM, Reagan LP, Svenningsson P, Fuchs E. The
neurobiological properties of tianeptine (Stablon): from monoamine hypothesis to glutamatergic
modulation. Mol Psychiatry. 2010;15(3):237–49.
Moret C, Briley M. The importance of norepinephrine in depression. Neuropsychiatr Dis Treat.
2011;7(Suppl 1):9–13.
Morinobu S, Nibuya M, Duman RS. Chronic antidepressant treatment down-regulates the induction
of c-fos mRNA in response to acute stress in rat frontal cortex. Neuropsychopharmacology.
1995;12(3):221–8.
Moskal JR, Burgdorf JS, Stanton PK, Kroes RA, Disterhoft JF, Burch RM, Khan MA. The
development of Rapastinel (formerly GLYX-13); a rapid acting and long lasting antidepressant.
Curr Neuropharmacol. 2017;15(1):47–56.
Müller WE. Wirkungsmechanismus niedrigdosierter Neuroleptika bei Angst und Depression. In:
Pöldinger W, editor. Niedrigdosierte Neuroleptika bei ängstlich-depressiven Zustandsbildern
und psychosomatischen Erkrankungen. Karlsruhe: Braun; 1991. p. 24–38.
Müller WE. Current St John’s wort research from mode of action to clinical efficacy. Pharmacol
Res. 2003;47(2):101–9.
Müller WE. Antidepressiva und kognitive Dysfunktion: die Rolle von Vortioxetin. Psychopharma-
kotherapie. 2015;22:177–88.
Müller WE. Normalisierung gestörter Neuroplastizitätsmechanismen als gemeinsame Endstrecke
im Wirkungsmechanismus von Antidepressiva: Die besondere Rolle von Tianeptin.
Psychopharmakotherapie. 2016;23(6):230–8.
Müller WE, Eckert A. Psychopharmakotherapie – pharmakologische Grundlagen. In: Möller H-J,
Laux G, Kapfhammer H-P, editors. Psychiatrie, Psychosomatik, Psychotherapie. 5th
ed. Springer; 2017. p. 749–93.
Nemeroff CB, Owens MJ. The role of serotonin in the pathophysiology of depression: as important
as ever. Clin Chem. 2009;55(8):1578–9.
Newport DJ, Carpenter LL, McDonald WM, Potash JB, Tohen M, Nemeroff CB. Ketamine and
other NMDA antagonists: early clinical trials and possible mechanisms in depression. Am J
Psychiatry. 2015;172(10):950–66.
Antidepressants: Pharmacology and Biochemistry 1133
Wilkinson ST, Toprak M, Turner MS, Levine SP, Katz RB, Sanacora G. A survey of the clinical,
off-label use of ketamine as a treatment for psychiatric disorders. Am J Psychiatry. 2017;174(7):
695–6.
Willner P. The validity of animal models of depression. Psychopharmacology. 1984;83(1):1–16.
Wohleb ES, Wu M, Gerhard DM, Taylor SR, Picciotto MR, Alreja M, Duman RS. GABA
interneurons mediate the rapid antidepressant-like effects of scopolamine. J Clin Invest.
2016;126(7):2482–94.
Yang Y, Cui Y, Sang K, Dong Y, Ni Z, Ma S, Hu H. Ketamine blocks bursting in the lateral habenula
to rapidly relieve depression. Nature. 2018;554(7692):317–22.
Zarate CA, Singh JB, Carlson PJ, Brutsche NE, Ameli R, Luckenbaugh DA, et al. A randomized
trial of an N-methyl-D-aspartate antagonist in treatment-resistant major depression. Arch Gen
Psychiatry. 2006;63(8):856–64.
Antidepressants: Indications,
Contraindications, Interactions, and Side
Effects
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1137
Indications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1138
Major Depressive Disorder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1138
Anxiety Disorders, Obsessive-Compulsive Disorder, Trauma- and Stressor-Related
Disorders . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1143
Other Approved Indications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1144
Contraindications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1145
Bipolar Disorder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1145
Pregnancy and Breastfeeding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1146
Liver and Kidney Impairment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1147
Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1148
MAOIs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1148
TCAs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1150
SSRIs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1150
SNRIs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1151
Mirtazapine and Mianserin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1152
Reboxetine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1152
Bupropion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1153
Agomelatine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1153
Vortioxetine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1154
Side Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1154
Gastrointestinal System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1154
Genitourinary System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1157
Cardiovascular System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1158
Abstract
Antidepressants are among the most largely prescribed medications worldwide.
They are the mainstay of depression treatment but are used for other conditions
too. As of today, over 40 medications are available. The history of new antide-
pressant development reflects an endeavor to reduce side effects while preserving
efficacy. In this perspective, newer agents have largely replaced monoamine
oxidase inhibitors and tricyclic antidepressants on the ground of higher tolerabil-
ity and similar efficacy. Considerations of safety, tolerability, comorbidities, and
possible drug interactions rather than efficacy generally guide antidepressant
choice. Personalized antidepressant choice should be implemented by matching
each patient’s symptom profile with the pharmacodynamic/pharmacokinetic pro-
file of available antidepressants. However, the use of clinical criteria to guide drug
choice does not prevent the occurrence of appreciable side effects, poor treatment
adherence, and insufficient response in a relevant proportion of patients. The need
of slow dose titration and the delayed development of the full antidepressant
action represent further issues. Several approaches can address these issues and
improve the outcomes of antidepressant treatment. These include the use of
genetic variants modulating antidepressant metabolism to guide drug choice
and dosing, as recommended by available guidelines. The validation of further
genetic and nongenetic biomarkers is expected in the next future. However, such
strategies merely capitalize on currently available medications which act on
substantially similar pathways and do not reflect the heterogeneity in depression
pathogenesis. The development of antidepressants with new mechanisms may
involve the modulation of endocrinological and immunological pathways, as well
as non-monoaminergic neurotransmission and metabolic regulation.
List of Abbreviations
5-HT Serotonin
ACT Anatomical Therapeutic Chemical
ADHH Attention deficit hyperactivity disorder
AIJ Antidepressant-induced jitteriness
BD Bipolar disorder
Antidepressants: Indications, Contraindications, Interactions, and Side Effects 1137
Introduction
Depression is a highly recurrent disorder taking a heavy toll on both personal and
public health. It is the leading cause of disability worldwide and has been substan-
tially increasing over the past few decades, mainly driven by population growth and
ageing (Ahmad Kiadaliri 2016). With an estimated 350 million people affected
globally, the economic burden of depressive disorders in the USA alone has been
estimated to be more than US$210 billion (WHO 2017). Antidepressants are the
mainstay of depression treatment and rank among the top three most commonly used
therapeutic drug classes in the USA – in the 2011–2014 period, 12.7% of persons in
the USA aged 12 and over reported antidepressant use in the past month (Pratt et al.
2017). Antidepressants are prescribed for a wide range of conditions other than
major depression. It has been over 60 years since the serendipitous discovery of the
first antidepressant, when an antituberculosis drug, iproniazid, was observed to
improve mood in patients treated for tuberculosis. As of today, over 40 antidepres-
sants have been approved by regulatory agencies.
There are a number of non-pharmacological medical treatments for depression,
but it is beyond the scope of the present chapter to discuss them, and the reader can
refer to the relevant literature (Farah et al. 2016).
1138 F. Corponi et al.
Indications
While depression accounts for the majority of antidepressant prescription, this class
of drugs is also used to treat other conditions. Indications are established through
randomized clinical trials (RCTs), and being depression the most relevant and
prevalent disease, pharmaceutical companies tend to set up RCTs aimed at gaining
approval for this disorder in the first place. The relatively high rates of placebo
response (20–40%) have often represented a concern and contributed to failed trials
and delayed delivery of new medications to market. A number of factors purportedly
drive placebo response, including the randomization ratio, the expectation of receiv-
ing an active treatment, the therapeutic setting, the frequency of study visits, and the
entry criteria (e.g., milder forms of depression). These factors are potential causes for
an inflation of placebo response (Furukawa et al. 2016). However, antidepressants
are more efficacious than placebo in adults with major depressive disorder (MDD),
and although the overall effect size is statistically modest (standardized mean
difference (SMD) of 0.30 in acute treatment), statistical significance does not
equal to clinical significance (Cipriani et al. 2018).
The approval for indications other than MDD requires additional RCTs that
pharmaceutical companies may decide not to fund, resulting in common “off-
label” use of medications. By the same token, while not all drugs appear to be
equally effective in each condition, limited data on comparability across drugs is
available for cost reasons. At present, due to the absence of validated biomarkers to
guide treatment choice and dosing, treatment decisions rely to a great extent on
clinical factors. A comprehensive list of antidepressants licensed by the US Food and
Drug Administration (FDA) and European Medicines Agency (EMA) is reported in
Table 1.
Table 1 Antidepressants indications approved by the US Food and Drug Administration (FDA)
and European Medicines Agency (EMA)
Drug FDA EMA
Fluoxetine MDD, OCD, BN, PD MDD, OCD, BN
Sertraline MDD, OCD, PD, PTSD, SAD, MDD, OCD, PD, PTSD, SAD
PMD
Paroxetine MDD, OCD, PD, SAD, GAD, MDD, OCD, PD, SAG, GAD,
PTSD PTSD
Fluvoxamine OCD MDD, OCD
Citalopram MDD MDD, PD
Escitaloprama MDD, GAD MDD, PD, SAD, OCD, GAD
Venlafaxine MDD, GAD, SAD, PD MDD
Duloxetine MDD, GAD, DPNP, FM, CMP MDD, GAD, DPNP
Milnacipram FM \
Vilazodone MDD \
Bupropiona MDD, SAD, SC MDD, SC
Reboxetinea \ MDD
Atomoxetine ADHD ADHD
Agomelatine \ MDD
Mirtazapine MDD MDD
Mianserina MDD
Trazodonea MDD MDD
Nefazodone Withdrawn due to incidence of Withdrawn due to incidence of
severe liver damage (1 in every severe liver damage (1 in every
250,000 to 300,000 patient-years) 250,000 to 300,000 patient-years)
Vortioxetine MDD MDD
Clomipraminea OCD MDD, OCD, PD, PTSD, SAD,
GAD
Nortriptylinea MDD MDD
Amitriptyline MDD MDD, NP, CTTH (prophylactic),
migraine (prophylactic), NE
Desipraminea MDD MDD
Tranylcyprominea MDD MDD
Doxepina MDD, insomnia characterized by MDD
difficulties with sleep maintenance
Imipraminea MDD, CE MDD, CE
ADHD attention-deficit/hyperactivity disorder, BN bulimia nervosa, CE childhood enuresis, CMP
chronic musculoskeletal pain, CTTH chronic tension-type headache, DPNP diabetic peripheral
neuropathic pain, FB fibromyalgia, GAD generalized anxiety disorder, MDD major depressive
disorder, NE nocturnal enuresis, NP neuropathic pain, OCD obsessive-compulsive disorder, PD
panic disorder, PMD premenstrual dysphoric disorder, PTSD post-traumatic stress disorder, SAD
social anxiety disorder, SC smoking cessation, VSM vasomotor symptoms associated with
menopause
a
Approved by individual European countries through a decentralized procedure – indications
licensed by the following national agencies are reported: Medicines and Healthcare Products
Regulatory Agency (MHRA); Federal Institute for Drugs and Medical Devices (BfArM);
Italian Medicines Agency (AIFA); Spanish Agency for Medicines and Health Products
(AEMPS); National Agency for the Safety of Medicine and Health Products (ANSM)
1140 F. Corponi et al.
first-line treatment, but evidence for superiority of one antidepressant over the other
is flimsy, making the identification of the most appropriate drug difficult. By and
large, effectiveness is comparable between classes and within classes of medications
and does not therefore stand as the main criterion of choice. However, differences in
the risk of causing particular side effects and drug interactions are better known.
Thus, considerations such as safety, tolerability, co-occurring medical conditions,
and potential for drugs interactions are usually recommended to guide the initial
selection of the antidepressant medication. Personal and familial history of prior
treatments should be investigated since an individual is more likely to respond to an
antidepressant to which himself/herself or a next of kin have responded in the past. In
addition, when choosing which antidepressant to prescribe, the specific pharmaco-
dynamic profile (i.e., pharmacological targets and binding affinity) should be con-
sidered in order to match the observed clinical profile (Serretti 2018). On these
premises, selective serotonin reuptake inhibitors (SSRIs), serotonin and noradrena-
line reuptake inhibitors (SNRIs), mirtazapine, bupropion, and vortioxetine are con-
sidered the first-line medications in most cases (Gelenberg et al. 2010; Kennedy et al.
2016).
The elderly on the one hand, children and adolescents on the other represent
special populations. The elderly, particularly patients aged 65 and over, seem to
benefit less from antidepressants, in terms of both smaller drug-placebo difference
and later onset of antidepressant action. As comorbidity is frequent in this age range,
patients should be thoroughly evaluated for the presence of co-occurring medical
conditions which can precipitate depressive episodes and hamper antidepressant
action. There is no ideal antidepressant in the elderly, and they should be closely
monitored for potential side effects and drug interactions. SSRIs are generally better
tolerated than tricyclic antidepressants (TCAs). Agomelatine, sertraline,
vortioxetine, and duloxetine as well are effective and generally well-tolerated (Tay-
lor et al. 2018; Tedeschini et al. 2011). At the other end of the age spectrum,
fluoxetine is the only both FDA- and EMA-approved medication for the treatment
of depression in children and adolescents. Escitalopram carries FDA approval solely
for adolescent depression. Psychological interventions are considered first-line treat-
ment for child and adolescent depression, and pharmacotherapy should be used for
more severe episodes of disease in combination with psychological treatment. In
younger children (<13 years), antidepressant efficacy seems smaller (Birmaher et al.
2007).
Over the past decades, there has been an endeavor in psychiatric research to boost
precision medicine (Serretti 2018), which ultimately aims at therapies tailored to the
specific patient, on the basis of a deep genetic, environmental and psychological
characterization. Despite some initial promising results, no pharmacogenetic test has
become of common use in non-research clinical settings because of the lack of
definitive evidence of favorable cost/effectiveness ratio. Current clinical guidelines
and drug labels recommend genotyping of variants in two cytochrome genes
(CYP2D6 and CYP2C19) to guide choice and dosing of a number of antidepressants
(Consortium CPI 2017; FDA 2017). Indeed genetic polymorphisms in cytochrome
P450 (CYP450) genes, alongside other factors, affect drug plasma concentration and
Antidepressants: Indications, Contraindications, Interactions, and Side Effects 1141
options are available: 1) maximizing the dose of medication, especially if the upper
limit dosage has not been reached; 2) switching to a different compound, considering
that the appropriateness of changing between or within classes is controversial; and
3) adding an adjunctive agent from an armamentarium comprising lithium, thyroid
hormones, another antidepressant, some anti-convulsants, some atypical antipsy-
chotics, and psychostimulants. Non-response to two or more antidepressants pre-
scribed at an adequate dose and duration is usually defined treatment-resistant
depression (TRD). A discussion of TRD and its management is beyond the scope
of this chapter. Overall, one-third of MDD patients will experience TRD. Approx-
imately one-third of patients on their first MDD episode will remit on their first
antidepressant trial after the acute phase of treatment. Unfortunately, for those who
fail to remit, chances of remission decrease with each successive trial.
In the continuation phase, antidepressant treatment should be continued at the
same therapeutic dose at which remission was reached for 6–9 months, unless
tolerability issues require a dose reduction. The goal of this phase is to forestall
relapse, which is defined as a return of symptoms to the full syndrome criteria for an
episode during remission but before recovery. Recovery is defined as remission
maintained for over 6 months, thus it is distinguished from remission by a temporal
criterion. Average rate of relapse in patients not going on continuation phase is 41%
compared with 18% on active treatment (Geddes et al. 2003), and thus continuation
should always be pursued. Even when antidepressant therapy is aptly administered
during the continuation phase, relapse may occur requiring a return to the acute
phase of treatment.
At the end of the continuation phase, a careful evaluation should be made as to
whether to maintain treatment or gradually go off medication(s). The aim of main-
tenance is to protect susceptible patients against recurrence, which is defined as the
appearance of a new episode after a period of recovery. A prior history of three or
more MDD episodes was associated with an above 90% recurrence rate (Angst et al.
1973; Lavori et al. 1994), thus it is the main condition that requires maintenance.
Presence of residual symptoms, ongoing psychosocial stressors, severity of prior
episodes, presence of side effects, and patient’s preference should be taken into
account when choosing between maintenance and discontinuation. In general, the
same medication dose of the continuation phase should be kept during maintenance,
especially if it is well-tolerated. Recurrence however will still happen in as many as
25% of individuals with adequate maintenance therapy.
On the other hand, in patients discontinuing antidepressants, it is important to
taper the medication over the course of some weeks in order to prevent discontin-
uation syndrome. The highest risk of relapse in these patients is in the first 2 months
after therapy cessation, thus a careful monitoring is recommended in this period
(Taylor et al. 2018; Gelenberg et al. 2010; Kennedy et al. 2016; Baldwin et al. 2015).
Of notice, as many as 50% of patients from psychiatric settings and primary care
prematurely discontinue antidepressant therapy on their own initiative for a number
of reasons, some of which are patient-related (e.g., side effects, misperceptions about
the medication), while others are clinician-related (e.g., poor instruction by the
clinician about the medication, lack of follow-up care). Poor treatment compliance
Antidepressants: Indications, Contraindications, Interactions, and Side Effects 1143
phobia, and OCD and 6–8 months for panic disorder and PTSD. When discontinued,
antidepressants should be reduced slowly over several weeks to months to avoid
discontinuation syndrome.
Contraindications
Bipolar Disorder
Much of the concern about antidepressant safety in bipolar disorder (BD) regards the
possibility of affective switch to mania/hypomania. Although the empirical data are
arguably inconclusive (evidence level “C,” according to the the International Society
for Bipolar Disorders (ISBD)) (Pacchiarotti et al. 2013), current treatment guidelines
for BD (Taylor et al. 2018; Yatham et al. 2018) espouse this view and discourage the
use of antidepressant monotherapy. In fact, the risk of switch is significantly reduced
with a concurrent mood stabilizer compared to patients on antidepressant mono-
therapy: among patients treated with a concurrent mood stabilizer, no change in the
risk of (hypo)manic switch was observed during the 3 months after the start of an
antidepressant treatment, and a decreased risk was observed during the period
3–9 months after treatment initiation (hazard ratio = 0.63, 95% CI = 0.42; 0.93)
(Viktorin et al. 2014). Since a major depressive episode may be the initial presen-
tation of BD, clinicians are encouraged to screen for personal and familial history of
mania/hypomania. Patients should be monitored, particularly during the initial phase
of treatment, for activation indicators (Taylor et al. 2018; Yatham et al. 2018). The
highest risk of affective switch was associated with tricyclic antidepressants (TCAs)
and venlafaxine, while the lowest risk was reported for bupropion. Monoamine
oxidase inhibitors (MAOIs) may have a relatively low risk too. For BDII patients,
switch rate is lower, and when switch does occur, it is almost exclusively into
hypomania. Another controversial question is whether antidepressants can induce
rapid cycling (i.e., 4 recurrences/year) in BD. The ISBD concluded that the overall
quality rating of the evidence on this topic is poor (evidence level “D”). On the flip
side, there is strong evidence that antidepressants are harmful in terms of switch rates
in patients with a history of rapid cycling, and thus their use should be discouraged in
this case. One last issue is whether antidepressant may alter the risk of suicidal
behavior in BD. Once again, quality of evidence was assessed as poor by the ISBD
(evidence level “D”) (Pacchiarotti et al. 2013).
Besides the aforementioned issues, evidence of antidepressant efficacy in bipolar
depression is poor. Consequently, treatment guidelines (Taylor et al. 2018; Yatham
et al. 2018) recommend the use of these medications as second/third-line option. In
particular, in individuals with a personal history of response to an antidepressant, the
1146 F. Corponi et al.
The use of medications during pregnancy and breastfeeding raises safety concerns
about both the fetus/newborn and the mother. Antidepressants safety in these patient
groups cannot be clearly established since robust; prospective studies are lacking as
obviously unethical. Data are thus largely sparse and methodologically limited. If
antidepressants carry some known risks, untreated depression does too, for both the
fetus/newborn and the mother. Although minor forms of depression may respond
well to support or psychotherapy, major depression may benefit from the use of
antidepressants. Furthermore, pregnancy is not protective with respect to the risk of
depression relapse, contrary to what was once though, and relapse rates are high in
those with a history of depression who discontinue medications. In one study 68% of
women who were euthymic on antidepressant therapy and stopped treatment during
pregnancy relapsed, compared with 26% who continued antidepressants (Cohen et
al. 2006). Consequently, despite pregnancy and breastfeeding being relative contra-
indications, the use of antidepressants in such conditions is common in the clinical
practice, with around 8% of women in the USA being prescribed antidepressants at
some point during their pregnancy; this rate has shown a trend of increase over the
years, and SSRIs are the drugs class most commonly prescribed (Alwan et al. 2011).
Overall, antidepressants appear not to be major teratogens. Some rare birth
defects however have been observed to occur at higher rates in case of first trimester
exposure to specific compounds. Paroxetine in particular was specifically associated
with cardiac malformations. This resulted in labelling changes and treatment guide-
lines recommending that paroxetine should not be the first-line option when
selecting a new antidepressant for a pregnant patient (Berard et al. 2016). Sertraline
has the least degree of placental exposure and is recommended as first-line choice
during pregnancy (Hendrick et al. 2003).
Antidepressants were associated with spontaneous abortion, decreased gesta-
tional age, and decreased birth weight. It is possible however that any increased
risk may be related to depression, not the medication used to treat it. As regards
persistent pulmonary hypertension of the newborn, available studies make it
likewise difficult to differentiate the contribution of antidepressants from that of
Antidepressants: Indications, Contraindications, Interactions, and Side Effects 1147
depression itself, and if any risk exists, it was estimated to be less than 1% (t Jong
et al. 2012).
With third trimester antidepressant use, neonatal abstinence syndrome can occur.
This condition is the result of medication withdrawal from intra-uterus exposure and
entails transient symptoms such as jitteriness, tremor, sleep, and feeding distur-
bances. It may occur in as many as 30% of neonates exposed to SSRIs in utero,
but it is usually a mild and self-limiting condition. Paroxetine may be the SSRI most
likely to cause abstinence syndrome, allegedly due to its short half-life compara-
tively to other SSRIs and lack of active metabolites. Conversely, the long half-life of
fluoxetine and its active metabolite norfluoxetine, 2–4 days and 7–15 days, respec-
tively, make neonatal abstinence syndrome less likely with this medication
(Levinson-Castiel et al. 2006).
Women on antidepressants during pregnancy may be at increased risk of devel-
oping hypertension, preeclampsia, and postpartum hemorrhage. All these three
conditions are purportedly driven by serotonin (5-HT) and NE interference with
vascular tone and hemostasis.
Exposure to antidepressants via breast milk is substantially lower than in utero
exposure. Comparatively greater experience has been acquired for SSRIs, being the
class of drugs most commonly prescribed to breastfeeding women. Paroxetine and
sertraline should be considered first-line choices when needing to start an antide-
pressant treatment in nursing women. Contrariwise fluoxetine should be avoided
since accumulation may occur in the infant in consideration of the long elimination
half-life of fluoxetine and its primary active metabolite (Taylor et al. 2018).
Antidepressants are highly protein bound, as lipophilic compounds, and are exten-
sively metabolized through the liver, primarily by CYP450 isoenzymes. Given the
crucial role the liver plays in the pharmacokinetics of antidepressants, it is unsur-
prising that various degrees of liver impairment have important consequences on
drug metabolism via a number of routes: decreased blood/plasma clearance by
hepatic metabolism or biliary excretion, reduced synthesis of transport proteins,
augmented distribution volume due to peripheral edemas and ascites, and reduced
first-pass metabolism as a result of portal-systemic shunts. Although kidneys are less
relevant to antidepressant metabolism, many active metabolites are excreted via this
route. Besides, patients with liver and/or kidney impairment, especially moderate to
severe forms, are prone to a number of comorbidities and are likely to receive
various medications. Thus the use of antidepressants in such patients requires
particular caution. There is a lack of clinical trials supporting effectiveness and
safety of antidepressants in liver and/or kidney impairment, since these conditions
usually stand as exclusion criteria. For the same reason, many drugs lack information
regarding pharmacokinetics and dosing adjustments in liver and/or kidney impair-
ment. The current consensus is that SSRIs are the safest antidepressants in both liver
and kidney impairment, principally based upon their generally high tolerability and
1148 F. Corponi et al.
wider therapeutic index. Lower starting/maintenance doses and slower titration are
recommended. Assessing hepatic and/or renal function is necessary so that dose
adjustment can be made. Creatinine clearance calculated with the Cockcroft-Gault
equation provides an estimation of the renal function, while the Child-Pugh score is
commonly used to establish the severity of liver impairment (Mullish et al. 2014;
Hedayati et al. 2012).
Interactions
Considering their large use (Pratt et al. 2017), antidepressants are very likely to be
co-prescribed with other drugs and to be encountered not only by mental health
specialists. Hence the evaluation of potential drug-drug interactions involving anti-
depressants is particularly relevant as well as the evaluation of the clinical signifi-
cance of such interactions. Drug-drug interactions are broadly classified as either
pharmacokinetic or pharmacodynamic. The former is the case of one drug interfering
with the absorption, distribution, metabolism, or elimination of another. This may
cause subtherapeutic effect or toxicity. The majority of the significant pharmacoki-
netic interactions with antidepressants involve druginduced changes in hepatic
metabolism, predominantly through cytochrome P450 (CYP450) isoenzymes.
While enzyme inhibition occurs rapidly, within 2–3 days, due to the offending
drug binding to the metabolizing enzyme, enzyme induction is generally a slower
process, hinging on protein expression reprogramming. Pharmacodynamic interac-
tions happen when one drug modulates the pharmacologic effect of another one
resulting in additive, synergistic, or antagonistic effects through action at the same
biochemical or molecular site, on the same target organ or on a different target that is
associated with the same physiological process. When considering the relevance and
significance of potential drug-drug interactions, taking into account patient’s char-
acteristics is of upmost importance: age, comorbidities, and genetic variability. In
particular, polymorphisms in CYP450 genes underlie differences in enzyme activity
that range from no enzyme activity (poor metabolizers), decreased (intermediate
metabolizers), normal (extensive metabolizers/wild type), to increased (ultra-rapid
metabolizers) enzyme activity. A common consequence of these different pheno-
types is therapeutic resistance or inefficacy due to insufficient plasma levels in ultra-
rapid metabolizers. In contrast, slow metabolism (poor metabolizers) may lead to
elevated plasma levels and side effects/toxicity. Of notice, most drug interactions are
detected after drug marketing (phase IV) and published either as single case reports
or case series (Taylor et al. 2018).
MAOIs
MAOIs are associated with the most clinically significant drug interactions among
all antidepressant classes. Despite decades of use, the pharmacokinetic interactions
of MAOIs are still poorly characterized but are certainly less problematic than
Antidepressants: Indications, Contraindications, Interactions, and Side Effects 1149
serotonergic agent and vice versa. Because fluoxetine has a particularly long
dynamic half-life, a washout period of more than 5 weeks may be needed
(Flockhart 2012).
TCAs
TCAs are primarily metabolized through the hepatic isoenzyme CYP2D6, with 1A2
and 3A4 as secondary routes; on the other hand, they do not cause enzyme inhibition
or induction. As a result, this class of drugs is not problematic itself in terms of
pharmacokinetic interactions, but, because of their narrow therapeutic index, TCAs
are susceptible to interactions when combined with inhibitors or inducers of the
aforementioned CYP450 enzymes.
Pharmacodynamic interactions on their part can exacerbate TCAs side effects.
Their sedative action due to H1 receptor agonism can be aggravated by sedative
drugs or alcohol, leading to respiratory depression in the worst case scenario.
Antihistamines and antipsychotics can precipitate anticholinergic effects, such as
blurred vision, dry mouth, constipation, urinary retention, and impaired cognition.
Adrenergic α1 blockers and antihypertensive drugs in general increase the risk of
orthostatic hypotension and subsequently of falls. Since TCAs lower the seizure
threshold, caution is needed with other proconvulsant drugs, e.g., antipsychotics,
particularly if the patient is being treated for epilepsy (higher doses of anti-convul-
sants may be required). Drugs that alter cardiac conduction either directly (e.g.,
antiarrhythmics or phenothiazines) or indirectly (e.g., diuretics via electrolyte dis-
turbances) should be used with caution in patients taking TCAs, in consideration of
the increased risk of QTc interval prolongation and arrhythmias. Lastly, combination
with drugs enhancing serotonergic transmission can result in serotonin syndrome
(Gillman 2007).
SSRIs
should thus be avoided in women taking tamoxifen for the treatment or prevention of
recurrence of breast cancer (Spina et al. 2012).
SSRIs are liable to a number of potential pharmacodynamics interactions which
make their side effects more likely to occur or more severe. Co-prescription with
other drugs raising serotonergic transmission can result in serotonin syndrome.
These include other 5-HT enhancing antidepressants (MAOIs and TCAs in partic-
ular), buspirone, linezolid, lithium, fentanyl, tramadol, and pethidine. Combination
with nonsteroidal anti-inflammatory drugs (NSAIDs), oral anticoagulants, or anti-
platelet drugs can lead to hemorrhagic risk, especially in patients with a pre-existing
hemorrhagic diathesis or undergoing major surgery. Antipsychotics can worsen
sexual dysfunction associated with SSRIs. Thiazide diuretics increase the risk of
hyponatremia. Medicines known to prolong QTc interval, such as most antipsy-
chotics and TCAs, warrant caution when used with citalopram in particular, for
which the risk of QTc prolongation seems the highest among SSRIs (Taylor et al.
2018; Gelenberg et al. 2010).
SNRIs
Mirtazapine and mianserin have three main metabolic pathways, via CYP2D6, 1A2,
and 3A4, and have no appreciable inhibitory or inducing effects itself on the various
CYP450 isoforms. Thus, they seemingly carry a low risk of pharmacokinetic
interactions. However, the few interaction studies available showed that cimetidine,
fluvoxamine, and ketoconazole markedly increase plasma concentrations of both
drugs while carbamazepine and phenytoin (CYP3A4 inducers) reduce their plasma
concentrations. Smokers have significantly lower plasma mirtazapine and mianserin
concentration than nonsmokers, confirming the role of CYP1A2, which is induced
by cigarette smoking, in the metabolism of both drugs.
An additive sedative effect was documented when mirtazapine was co-adminis-
tered with diazepam or alcohol. The combinations likely to potentiate mirtazapine
sedative effect warrant caution, especially in the elderly. Likewise, mianserin may
potentiate the central nervous depressant action of alcohol, anxiolytics, and hyp-
notics. Weight gain is another common adverse effect of mirtazapine, which can
escalate when the drug is combined with antipsychotics (Spina et al. 2012). As
compared to both SSRIs and SNRIs, mirtazapine probably carries a lower risk of
serotonin syndrome and bleeding (Hauta-Aho et al. 2009). Lastly, the combined use
of mirtazapine with tramadol or dopamine-blocking agents could enhance the risk of
restless legs syndrome, possibly through a pharmacodynamic mechanism involving
5-HT2 receptor-mediated suppression of dopamine release and H1 receptor antag-
onism (Kim et al. 2008). Clinical experience has shown that mianserin does not
interact with the antihypertensives bethanidine, clonidine, guanethidine, or propran-
olol. There may be an enhanced hypotensive effect if mianserin is taken with
diazoxide, hydralazine, or nitroprusside (Kopera 1978).
Reboxetine
on CYP2D6 and CYP3A4, unlikely to be relevant at the doses usually reached in the
clinical practice (Wienkers et al. 1999). Triiodothyronine, when co-administered as
an adjunctive treatment to reboxetine in MDD, was recorded to precipitate norad-
renergic effects such as anxiety, irritability, increased sweating, tremor, and sleep
disturbances (Cooper-Kazaz et al. 2010).
Bupropion
Agomelatine
Vortioxetine
Side Effects
Side effects of antidepressants limit compliance and are often the primary cause of
medication discontinuation or dose reduction. Roughly one-third to one-half of
patients discontinue pharmacological treatment, and a quarter of those patients
reported side effects as the reason for discontinuation (Lader et al. 2004). The history
of new antidepressant development reflects an effort by the pharmaceutical industry
to reduce side effects while preserving efficacy. In this perspective, TCAs and
MAOIs have come to be partially replaced by new generations of antidepressants
as these proved to be generally safer and better tolerated. The likelihood of different
side effects varies among drug classes, among individual agents, and, on the other
hand, among different subjects. Overall, side effects tend to manifest earlier than
therapeutic effects but they diminish over time with receptor desensitization. In
efficacy trials, an average of 63% of patients experienced at least one adverse
event during treatment (Gartlehner et al. 2011). Table 2 lists antidepressant side
effects and specifies more frequently associated and safer agents for each side effect.
Gastrointestinal System
Table 2 Antidepressant side effects with more frequently associated drugs and, where applicable,
drugs recognized as relatively safer
Side effect Drug(s) more frequently associated Safer drug(s)
Nausea Fluvoxamine, fluoxetine, venlafaxine, Escitalopram
vortioxetine, vilazodone
Vomiting Fluoxetine, venlafaxine Escitalopram
Diarrhea Sertraline, vilazodone \
Anorexia Fluoxetine Escitalopram
Xerostomia TCAs, SSRIs, SNRIs, bupropion Fluvoxamine,
vortioxetine
Constipation TCAs \
Drug-induced liver MAOIs, TCAs, nefazodone, bupropion, Citalopram,
injury duloxetine, agomelatine escitalopram, paroxetine,
fluvoxamine
Sexual dysfunction Paroxetine, venlafaxine Mirtazapine,
agomelatine,
vortioxetine, vilazodone
Urinary retention TCAs (particularly imipramine) \
QTc interval TCAs, citalopram Lofepramine
prolongation
Hypertension Venlafaxine, MAOIs
Orthostatic TCAs, mirtazapine
hypotension
" resting heart rate, # TCAs, SNRIs \
heart rate variability
Seizures TCAs, bupropion \
Extrapyramidal SSRIs \
symptoms
Blurred vision TCAs \
Angle-closure TCAs, paroxetine
glaucoma
Weight gain Paroxetine, amitriptyline Venlafaxine,
vortioxetine, vilazodone
Weight loss Fluoxetine, bupropion Escitalopram
Bleeding SSRIs, venlafaxine \
Hyponatremia SSRIs, venlafaxine TCAs, bupropion,
trazodone
Hyperhidrosis TCAs, sertraline, paroxetine Fluvoxamine, bupropion,
vortioxetine
Insomnia Bupropion, desvenlafaxine, duloxetine, Agomelatine, citalopram,
fluoxetine, fluvoxamine, reboxetine, escitalopram
sertraline, venlafaxine
Somnolence Fluvoxamine, mirtazapine, paroxetine, Bupropion
citalopram
Restless leg Mianserin, mirtazapine \
syndrome
Discontinuation Paroxetine, venlafaxine Agomelatine,
syndrome vortioxetine
(continued)
1156 F. Corponi et al.
Table 2 (continued)
Side effect Drug(s) more frequently associated Safer drug(s)
SSRI-induced Paroxetine Fluvoxamine
indifference
Switch to mania/ TCAs, venlafaxine Bupropion
hypomania
MAOIs monoamine oxidase inhibitors, REM rapid eye movement, SNRIs serotonin and noradren-
aline reuptake inhibitors, SSRIs selective serotonin reuptake inhibitors, TCAs tricyclic
antidepressants
Among SSRIs, nausea, vomiting, weight loss, and anorexia are more frequent in
fluvoxamine- and fluoxetine-treated patients, whereas they are less likely with
escitalopram. Venlafaxine consistently showed a higher rate of nausea and vomiting
than SSRIs, while sertraline has a higher incidence of diarrhea than other SSRIs and
venlafaxine. The data for the most recently approved antidepressants, i.e.,
vortioxetine and vilazodone, are comparatively limited. Nevertheless, nausea is
one of the most frequently reported side effect for both of them, and, as regards
vilazodone, diarrhea seems a common occurrence, reported by over 10% of patients,
with most cases being mild to moderate and only a very few leading to premature
discontinuation (Wang et al. 2018).
Xerostomia
Xerostomia, an unpleasant dry mouth sensation, is a common complaint with TCAs
use experienced by up to 27% patients on these medications. It is one of the
anticholinergic side effects of TCAs. In fact, antagonism of muscarinic M3 receptors
is responsible for decreasing salivary flow rate. Other antidepressants as well can
produce xerostomia, arguably through different pathogenetic mechanisms. In particu-
lar, significant risk of dry mouth is associated with SSRIs (relative risk = 1.64), SNRIs
(relative risk = 2.24), and bupropion (relative risk = 2.0). With regard to SNRIs, it is
hypothesized that xerostomia results from central accumulation of NE, which causes
activation of α2-receptors and concurrent inhibition of parasympathetic salivary
neurons in the brainstem leading to decreased salivary flow. On the other hand,
SSRIs, having less affinity for muscarinic cholinergic receptors, α-receptors, and NE
transporter, have less impact on salivary flow rate. Fluvoxamine and vortioxetine are
considered unencumbered by this adverse event (Cappetta et al. 2018).
Constipation
Constipation is another common anticholinergic side effect of TCAs. Cholinergic
receptors antagonism reduces the contractions propelling waste matter through the
digestive tract and the intestinal secretions lubricating. Incidence rate is generally
higher with TCAs compared to most SSRIs (Trindade et al. 1998).
presentations but in most cases patients are clinically asymptomatic, and the alter-
ations identified in liver function tests are the only suggestive element of DILI.
Overall, the incidence of antidepressant-induced liver toxicity requiring hospitaliza-
tion is only 1.28–4 cases per 100,000 patient-years. The mechanism of DILI
associated with antidepressants can be immuno-allergic or metabolic. A hypersen-
sitivity syndrome (fever, rash, eosinophilia, and auto-antibodies) and a short latency
period (1 to 6 weeks) are earmarks of the former pathogenetic mechanism. On the
contrary, the absence of any hypersensitivity syndrome and a delayed onset (1 month
to 1 year) point to an immuno-allergic reaction. In most cases, liver damage is
unpredictable, dose-independent, and of the hepatocellular type rather than of the
cholestatic type. Old age and polytherapy with more than one drug targeting the
same CYP450 isoenzyme pathway are the main risk factors. Of notice, symptoms of
DILI can mimic neurovegetative manifestations of MDD. As a consequence of
misdiagnosis, clinicians may be tricked into increasing the antidepressant dosage
or into co-prescription. The drugs for which the frequency of reported DILI appears
to be highest are MAOIs, TCAs, nefazodone, bupropion, duloxetine, and
agomelatine. Those with apparently lower risk are citalopram, escitalopram, parox-
etine, and fluvoxamine. For nefazodone the risk of liver toxicity was considered
unacceptably high, and it was subsequently withdrawn from the market (Voican et al.
2014).
Genitourinary System
Sexual Dysfunction
The incidence of sexual dysfunction during antidepressant treatment can be high
(50–70%), mainly when the mechanism of antidepressant action is highly related to
5-HT reuptake blockade. On the contrary, drugs that predominantly decrease NE or
dopamine uptake have either fewer or null negative effects on sexual functioning.
5-HT2A receptor agonism at descending spinal cord synapses on sympathetic/
parasympathetic neurons is purportedly the main mechanism of sexual dysfunction,
and the side effect is likely to be dose-dependent. Other factors may play a role too:
sedation, hormonal changes, disturbance of cholinergic/adrenergic balance, periph-
eral α–adrenergic antagonism, and inhibition of nitric oxide. All phases of sexual
activity can be affected: sexual desire, arousal (including clitoral engorgement and
lubrication in women and erectile function in men), and orgasm with possible
differences between men and women. However, all these effects are reversible.
Despite the high frequency of sexual dysfunction, patients tend not to report it,
unless directly questioned. Moreover, impaired sexual functioning can be a result of
depression itself or a concurrent medical/psychological condition (Serretti and
Chiesa 2009).
Compared with other second-generation antidepressants, paroxetine has higher
rates of sexual dysfunction, up to 80% (Montejo et al. 2010), particularly ejaculatory
dysfunction. Bupropion has large evidence of showing no influence on sexual
dysfunction or slight improvement of sexual function both in depressed populations
and in healthy volunteers. Mirtazapine and agomelatine have a good sexual
1158 F. Corponi et al.
tolerability, similarly to the newer agents vortioxetine, at least in the lower dose
range, and vilazodone (Reichenpfader et al. 2014).
A relatively rare but important side effect with trazodone is priapism. It may be
most likely to occur within the first 28 days of treatment and that the majority of
cases occur with doses of 150 mg/day or less. All age groups appear to be vulnerable
to this adverse effect. Patients should be informed to discontinue the medication on
occurrence of any unusual erectile problem (Warner et al. 1987).
Urinary Disturbances
Urination is controlled by a complex mechanism that coordinates bladder storage,
emptying, and urinary sphincter activity by regulation of smooth muscle in the
bladder and urethra. TCAs have a significant association with urinary retention,
especially imipramine. However, when all TCAs are considered together, the risk is
very low (0.1%). TCAs may cause urinary retention through muscarinic cholinergic
receptors antagonism and, secondly, 5-HT and NE reuptake inhibition. Urinary
disturbances are rather rare events with newer antidepressants. Urinary retention
secondary to the use of SSRIs is supported only by case reports, and in most cases
SSRIs were used in combination with benzodiazepine and/or antipsychotics. 5-HT
may increase the central sympathetic outflow leading to urinary storage, and, at the
same time, it inhibits the parasympathetic flow that promotes voiding. SNRIs may
produce urinary retention via α1–adrenoreceptors stimulation as an additional mech-
anism. As with SSRIs, the association for SNRIs is likewise anecdotal and
documented only in case reports. Venlafaxine was associated with urinary inconti-
nence as well, the action on 5-HT4 receptors being allegedly the underlying mech-
anism (Dane et al. 2016; Faure Walker et al. 2016).
Cardiovascular System
Heart Rate
Lastly, antidepressants unfavorably impact on resting-state heart rate and its vari-
ability. Increases in the former and decreases in the latter are associated with
substantial morbidity and mortality. All antidepressants, except possibly SSRIs,
were associated with increases in resting heart rate and decreased heart rate vari-
ability. These negative effects are more prominent in people treated with TCAs
(mean resting heart rate = 73.94 bpm) followed by SNRIs (mean resting heart
rate = 71.00 bpm) compared to those not on antidepressants (mean resting heart
rate = 66.87 bpm) (Kemp et al. 2014).
Nervous System
Serotonin Syndrome
The serotonin syndrome is a potentially life-threatening adverse reaction that can
result from therapeutic drug use, intentional self-poisoning, or inadvertent interac-
tion between drugs. This condition importantly is not an idiopathic drug reaction but
a predictable consequence of excessive serotonergic agonism at the central and
peripheral serotonergic receptors. The classic presentation is described as a clinical
triad of mental status changes, autonomic hyperactivity, and neuromuscular abnor-
malities, but clinical manifestations are highly variable and range from benign to
1160 F. Corponi et al.
Seizures
Antidepressants can lower the seizure threshold. Epileptogenic risk is particularly
high for TCAs and bupropion which are indeed contraindicated in individuals with
seizure disorders. As regards bupropion, the use of immediate release formulation in
dosages over 450 mg may produce a tenfold increase of estimated seizure incidence.
However, with the development of sustained release, the incidence of seizure
dropped to 0.1% in a study including 3.094 patients (50–300 mg) (Dunner et al.
1998).
Extrapyramidal Symptoms
Case reports have associated the use of antidepressants with extrapyramidal symp-
toms (EPS). The whole spectrum of EPS has been reported over the years, but
dystonia appears to be the most common presentation. SSRIs are the most implicated
agents (80.2% of all cases). The putative pathogenetic mechanism would be a
disruption in dopaminergic neurons in the nigrostriatal and tuberoinfundibular
pathways caused by excessive levels of 5-HT. EPS usually arise within 30 days of
either treatment initiation or dose increase. Age, gender, dosing, and concurrent
antipsychotic treatment seemingly do not impact substantially on liability to EPS
during treatment with SSRIs (Hawthorne and Caley 2015).
Headache
Although a commonly listed side effect reported for second-generation antidepres-
sants, headache is much more likely to be coincidental than associated with medi-
cation use. Second-generation antidepressants emerged as either not associated with
an increased risk of headache or associated with a clinically inconsequential, albeit
statistically significant, risk in comparison to placebo. The only antidepressants that
were found to be significantly associated with increased risk of headache compared
to placebo were bupropion (RR = 1.22, 95% CI = 1.06–1.41) and escitalopram
Antidepressants: Indications, Contraindications, Interactions, and Side Effects 1161
Ophthalmic Effects
Ophthalmic effects are more prominent with TCAs due to their anticholinergic
activity. Blurred vision occurs in roughly one-third of patients in the initial stages
of treatment with TCAs. Patients usually develop tolerance to this effect and vision
spontaneously return to normal after some time elapses. A more dangerous event is
angle-closure glaucoma, which TCAs may trigger through mydriasis and
cycloplegia. Pre-existing risk factors for angle-closure glaucoma indeed contraindi-
cate TCAs. SSRIs have been associated with mydriasis, increased intraocular pres-
sure, and angle-closure glaucoma via anticholinergic or noradrenergic effects (albeit
weak for the majority of SSRIs, especially in comparison to TCAs) and serotonergic
effects. Most of the reported glaucoma cases involved paroxetine, which compara-
tively to other SSRIs produces a relatively strong NE reuptake inhibition (Richa and
Yazbek 2010). Case reports have also associated venlafaxine with angle-closure
glaucoma (Richa and Yazbek 2010; Zhou et al. 2018; Ezra et al. 2006). In addition, a
nested case-control study found that that the risk of angle-closure glaucoma was
doubled among patients younger than 50 years taking bupropion (Symes et al. 2015).
New antidepressants thus do not appear to be completely safe from the risk of
glaucomatous attacks, even though well below the level of risk associated with
TCAs; baseline and follow-up ophthalmic consultations may be warranted in
patients with a higher risk for glaucoma (e.g., the elderly and those with a familial
high risk of glaucomatous diseases).
Weight Gain
Although weight loss can occur in the early stage of antidepressant treatment
(4–12 weeks), weight is often regained after 6 months and can be followed by
additional gain with long-term use. The interaction of a number of mechanisms is
likely to drive this phenomenon: 1) antagonism to histaminergic H1 receptors and
5-HT2C receptors; 2) diminished caloric expenditure due to sedative effects of
some antidepressants; 3) a shift in food preference; and 4) dry mouth, caused by
some antidepressants, may lead to an increased intake of caloric beverages
(Carvalho et al. 2016). Antidepressants markedly differ in their propensity to
induce changes in body weight, and overall the magnitude of effect is mild with
some notable exceptions. Paroxetine and amitriptyline are associated with a mean
weight gain after 4 months of 2.73 kg (95% CI = 0.78; 4.68) and 2.24 kg (95%
CI = 1.82; 2.66), respectively. Mirtazapine is the newer-generation antidepressant
most consistently associated with significant weight gain in the initial phases of
treatment with a mean weight gain of 1.74 kg between 4 and 12 weeks (95%
CI = 1.28; 2.20). Contrariwise, some weight loss occurs with fluoxetine and
bupropion, although the effect of fluoxetine appears to be limited to the acute
1162 F. Corponi et al.
Bleeding
Hyponatremia
Hyponatremia is defined as a serum sodium level of less than 135 mEq/L by some
sources or of less than 130 mEq/L by others, the normal range being 135–145 mEq/L.
Clinical manifestations depend on the severity of hyponatremia. Minor reductions in
serum sodium level can remain clinically undetected, whereas, in case of a more
important drop in serum sodium level, symptoms range from nausea and malaise to
lethargy, decreased level of consciousness, headache, and, in severe cases, seizures
and coma. Overt neurologic symptoms are generally due to very low serum sodium
levels (usually <115 mEq/L) and result from intracerebral osmotic fluid shifts and
Antidepressants: Indications, Contraindications, Interactions, and Side Effects 1163
Hyperhidrosis
Sleep Disturbances
Discontinuation Syndrome
Suicidality
Patients on SSRIs may develop diminished motivation and reduction in the intensity
and/or experience of all emotions, both positive and negative, with a subsequent
sense of emotional detachment. Since apathy is a feature of depression itself, SSRI-
induced indifference can be mistaken for a residual symptom of the illness. In SSRIs-
induced indifference, however, patients do not report emotional distress as with
depression, but, on the contrary, they report blunted emotional responsiveness; they
usually describe themselves as “just not caring.” While such a state of emotional
flattening can be in part desirable with regard to negative life events, it also limits
responsiveness to positive events and impairs interpersonal relationships, hindering
emotional resonance with beloved ones. SSRIs-induced indifference is dose-depen-
dent and completely reversible on drug discontinuation; usually a dose reduction is
sufficient. Onset is often insidious and delayed. Two putative mechanisms for
development of this side effect have been hypothesized: serotonergic effects on the
frontal lobes and/or serotonergic modulation of mid-brain dopaminergic systems,
which project to the prefrontal cortex. At present, data on epidemiology, effect of
gender, age and ethnicity, differential rate (if any) among SSRIs, and adjunctive risk
factors are lacking (Sansone and Sansone 2010). According to some findings
including animal models and healthy volunteers, SSRI-induced indifference could
be most pronounced with paroxetine, while fluvoxamine may be the least subject to
this effect (Serretti et al. 2010).
Antidepressant-Induced Jitteriness
Conclusions
Antidepressants are one of the most prescribed drug classes worldwide. Depression
accounts for the majority of antidepressant prescriptions; therefore for new drugs
pharmaceutical industries tend to set up RCTs aimed at gaining approval for this
condition in the first place. Additional RCTs to license drugs for other conditions
may not be funded, leaving a gray area of off-label prescriptions. Besides depression,
approved indications mainly comprise other psychiatric conditions in the realm of
anxiety spectrum disorders, obsessive-compulsive disorders, stressors-related disor-
ders, and medical painful syndromes. Whatever the illness, undertaking an antide-
pressant treatment is appropriate when some severity indicators are present.
Monotherapy is generally the first-line treatment, but evidence of superior efficacy
of one agent over another is scarce. Therefore, risk of different side effects and drug
interactions are usually recommended to guide antidepressant choice. Consequently,
considerations of safety, tolerability, co-occurring morbid conditions, and co-pre-
scribed medications rather than efficacy are the main criteria for antidepressant
choice.
The delayed development of the full antidepressant effect and the limited remis-
sion rate, with around 20–30% of patients considered treatment-resistant depression
(TRD), remain the most problematic issues of antidepressant therapy. Secondarily,
while newer antidepressants are safer and better tolerated, side effects are still a
major limit to compliance. A number of strategies can be applied in a bid to address
these still unmet needs. First, a clinical approach that matches clinical features (i.e.,
disease characteristics and vulnerability to side effects) with antidepressant specific-
ity (pharmacodynamic and pharmacokinetic profile) is a good option to personalize
treatment. This line of action can optimize efficacy and minimize side effects
(Serretti 2018), and it can be implemented more easily by using a recently proposed
new classification of psychotropic drugs based on pharmacological domains and
modes of action, the Neuroscience-based Nomenclature (NbN). Indeed the currently
applied Anatomical Therapeutic Chemical (ATC) classification classifies psychotro-
pic drugs by main indication, not including all potential clinical uses of a particular
agent. NbN aims to avoid confusion about drug indication, since patients and
Antidepressants: Indications, Contraindications, Interactions, and Side Effects 1167
References
Ables AZ, Nagubilli R. Prevention, recognition, and management of serotonin syndrome. Am Fam
Physician. 2010;81(9):1139–42.
Ahmad Kiadaliri A. Global, regional, and national disability-adjusted life-years (DALYs) for 315
diseases and injuries and healthy life expectancy (HALE), 1990-2015: a systematic analysis for
the Global Burden of Disease Study 2015. Lancet. 2016;388(10053):1603–58.
Alberti S, Chiesa A, Andrisano C, Serretti A. Insomnia and somnolence associated with second-
generation antidepressants during the treatment of major depression: a meta-analysis. J Clin
Psychopharmacol. 2015;35(3):296–303.
Alwan S, Reefhuis J, Rasmussen SA, Friedman JM, National Birth Defects Prevention S. Patterns
of antidepressant medication use among pregnant women in a United States population. J Clin
Pharmacol. 2011;51(2):264–70.
American Psychiatric A. Diagnostic and statistical manual of mental disorders. Washington, DC:
American Psychiatric Association; 2013.
Andrade C, Sharma E. Serotonin reuptake inhibitors and risk of abnormal bleeding. Psychiatr Clin
North Am. 2016;39(3):413–26.
Angst J, Baastrup P, Grof P, Hippius H, Poldinger W, Weis P. The course of monopolar depression
and bipolar psychoses. Psychiatr Neurol Neurochir. 1973;76(6):489–500.
Baldwin DS, Anderson IM, Nutt DJ, Allgulander C, Bandelow B, den Boer JA, et al. Evidence-
based pharmacological treatment of anxiety disorders, post-traumatic stress disorder and obses-
sive-compulsive disorder: a revision of the 2005 guidelines from the British Association for
Psychopharmacology. J Psychopharmacol. 2014;28(5):403–39.
Baldwin D, Barnes T, Coghill D, Goodwin G, Hale T, Howard L, et al. Evidence-based guidelines
for treating depressive disorders with antidepressants: a revision of the 2008 British Association
for Psychopharmacology guidelines. J Psychopharmacol. 2015;29(295):459–525.
Beach SR, Kostis WJ, Celano CM, Januzzi JL, Ruskin JN, Noseworthy PA, et al. Meta-analysis of
selective serotonin reuptake inhibitor-associated QTc prolongation. J Clin Psychiatry.
2014;75(5):e441–9.
Berard A, Iessa N, Chaabane S, Muanda FT, Boukhris T, Zhao JP. The risk of major cardiac
malformations associated with paroxetine use during the first trimester of pregnancy: a system-
atic review and meta-analysis. Br J Clin Pharmacol. 2016;81(4):589–604.
Beyer C, Cappetta K, Johnson JA, Bloch MH. Meta-analysis: risk of hyperhidrosis with second-
generation antidepressants. Depress Anxiety. 2017;34(12):1134–46.
Birmaher B, Brent D, Issues AWGoQ, Bernet W, Bukstein O, Walter H, et al. Practice parameter for
the assessment and treatment of children and adolescents with depressive disorders. J Am Acad
Child Adolesc Psychiatry. 2007;46(11):1503–26.
Boyer EW, Shannon M. The serotonin syndrome. N Engl J Med. 2005;352(11):1112–20.
Brintellix (vortioxetine) Summary of product characteristics (online) Lundbeck Limited 2015.
www.medicines.org.uk/emc/medicine/30904. Accessed 15 Aug 2018.
Cappetta K, Beyer C, Johnson JA, Bloch MH. Meta-analysis: risk of dry mouth with second generation
antidepressants. Prog Neuro-Psychopharmacol Biol Psychiatry. 2018;84(Pt A):282–93.
Carvalho AF, Sharma MS, Brunoni AR, Vieta E, Fava GA. The safety, tolerability and risks
associated with the use of newer generation antidepressant drugs: a critical review of the
literature. Psychother Psychosom. 2016;85(5):270–88.
Chappell J, He J, Knadler MP, Mitchell M, Lee D, Lobo E. Effects of duloxetine on the pharma-
codynamics and pharmacokinetics of warfarin at steady state in healthy subjects. J Clin
Pharmacol. 2009;49(12):1456–66.
Chen G, Lee R, Hojer AM, Buchbjerg JK, Serenko M, Zhao Z. Pharmacokinetic drug interactions
involving vortioxetine (Lu AA21004), a multimodal antidepressant. Clin Drug Investig.
2013;33(10):727–36.
Cipriani A, Furukawa TA, Salanti G, Chaimani A, Atkinson LZ, Ogawa Y, et al. Comparative efficacy
and acceptability of 21 antidepressant drugs for the acute treatment of adults with major depressive
disorder: a systematic review and network meta-analysis. Lancet. 2018;391(10128):1357–66.
Antidepressants: Indications, Contraindications, Interactions, and Side Effects 1169
Cohen LS, Altshuler LL, Harlow BL, Nonacs R, Newport DJ, Viguera AC, et al. Relapse of major
depression during pregnancy in women who maintain or discontinue antidepressant treatment.
JAMA. 2006;295(5):499–507.
Consortium CPI. CPIC guidelines. 2017. Available at: https://cpicpgx.org/guidelines/. Accessed 24
July 2018.
Cooper-Kazaz R, Cohen A, Lerer B. Noradrenergic adverse effects due to combined treatment with
reboxetine and triiodothyronine. J Clin Psychopharmacol. 2010;30(2):211–2.
Dane KE, Gatewood SB, Peron EP. Antidepressant use and incident urinary incontinence: a
literature review. Consult Pharm. 2016;31(3):151–60.
Dunner DL, Zisook S, Billow AA, Batey SR, Johnston JA, Ascher JA. A prospective safety
surveillance study for bupropion sustained-release in the treatment of depression. J Clin
Psychiatry. 1998;59(7):366–73.
Ezra DG, Storoni M, Whitefield LA. Simultaneous bilateral acute angle closure glaucoma following
venlafaxine treatment. Eye (Lond). 2006;20(1):128–9.
Farah WH, Alsawas M, Mainou M, Alahdab F, Farah MH, Ahmed AT, et al. Non-pharmacological
treatment of depression: a systematic review and evidence map. Evid Based Med. 2016;21(6):214–21.
Faure Walker N, Brinchmann K, Batura D. Linking the evidence between urinary retention and
antipsychotic or antidepressant drugs: a systematic review. Neurourol Urodyn. 2016;35(8):866–74.
FDA. Table of pharmacogenomic biomarkers in drug labeling. 2017. Available at: https://www.fda.
gov/Drugs/ScienceResearch/ucm572698htm. Accessed 24 July 2018.
Flockhart DA. Dietary restrictions and drug interactions with monoamine oxidase inhibitors: an
update. J Clin Psychiatry. 2012;73(Suppl 1):17–24.
Furukawa TA, Cipriani A, Atkinson LZ, Leucht S, Ogawa Y, Takeshima N, et al. Placebo response
rates in antidepressant trials: a systematic review of published and unpublished double-blind
randomised controlled studies. Lancet Psychiatry. 2016;3(11):1059–66.
Gabriel M, Sharma V. Antidepressant discontinuation syndrome. CMAJ. 2017;189(21):E747.
Garnock-Jones KP, Keating GM. Atomoxetine: a review of its use in attention-deficit hyperactivity
disorder in children and adolescents. Paediatr Drugs. 2009;11(3):203–26.
Gartlehner G, Hansen RA, Morgan LC, Thaler K, Lux L, Van Noord M, et al. Comparative benefits
and harms of second-generation antidepressants for treating major depressive disorder: an
updated meta-analysis. Ann Intern Med. 2011;155(11):772–85.
Geddes JR, Carney SM, Davies C, Furukawa TA, Kupfer DJ, Frank E, et al. Relapse prevention
with antidepressant drug treatment in depressive disorders: a systematic review. Lancet.
2003;361(9358):653–61.
Gelenberg AJ, Marlene Freeman CP, Markowitz JC, Rosenbaum JF, Thase ME, Trivedi MH, et al.
Practice guideline for the treatment of patients with major depressive disorder. 3rd ed. Work
group on major depressive disorder. Virginia: American Psychiatric Association. 2010.
Gillman PK. Tricyclic antidepressant pharmacology and therapeutic drug interactions updated. Br J
Pharmacol. 2007;151(6):737–48.
Goldstein DJ, Wilson MG, Thompson VL, Potvin JH, Rampey AH Jr. Long-term fluoxetine
treatment of bulimia nervosa. Fluoxetine Bulimia Nervosa Research Group. Br J Psychiatry.
1995;166(5):660–6.
Hamilton M. Rating depressive patients. J Clin Psychiatry. 1980;41(12 Pt 2):21–4.
Harada E, Tokuoka H, Fujikoshi S, Funai J, Wohlreich MM, Ossipov MH, et al. Is duloxetine's effect
on painful physical symptoms in depression an indirect result of improvement of depressive
symptoms? Pooled analyses of three randomized controlled trials. Pain. 2016;157(3):577–84.
Harmer CJ, Duman RS, Cowen PJ. How do antidepressants work? New perspectives for refining
future treatment approaches. Lancet Psychiatry. 2017;4(5):409–18.
Hauta-Aho M, Tirkkonen T, Vahlberg T, Laine K. The effect of drug interactions on bleeding risk
associated with warfarin therapy in hospitalized patients. Ann Med. 2009;41(8):619–28.
Hawthorne JM, Caley CF. Extrapyramidal reactions associated with serotonergic antidepressants.
Ann Pharmacother. 2015;49(10):1136–52.
Hedayati SS, Yalamanchili V, Finkelstein FO. A practical approach to the treatment of depression in
patients with chronic kidney disease and end-stage renal disease. Kidney Int. 2012;81(3):247–55.
1170 F. Corponi et al.
Hendrick V, Stowe ZN, Altshuler LL, Hwang S, Lee E, Haynes D. Placental passage of antide-
pressant medications. Am J Psychiatry. 2003;160(5):993–6.
Hiemke C, Bergemann N, Clement HW, Conca A, Deckert J, Domschke K, et al. Consensus
guidelines for therapeutic drug monitoring in neuropsychopharmacology: update 2017.
Pharmacopsychiatry. 2018;51(1–02):9–62.
Hilbert A, Hoek HW, Schmidt R. Evidence-based clinical guidelines for eating disorders: interna-
tional comparison. Curr Opin Psychiatry. 2017;30(6):423.
Jiang HY, Chen HZ, Hu XJ, Yu ZH, Yang W, Deng M, et al. Use of selective serotonin reuptake
inhibitors and risk of upper gastrointestinal bleeding: a systematic review and meta-analysis.
Clin Gastroenterol Hepatol. 2015;13(1):42–50.e3.
Katzman MA, Bleau P, Blier P, Chokka P, Kjernisted K, Van Ameringen M, et al. Canadian clinical
practice guidelines for the management of anxiety, posttraumatic stress and obsessive-compul-
sive disorders. BMC Psychiatry. 2014;14(Suppl 1):S1.
Katzman MA, Nierenberg AA, Wajsbrot DB, Meier E, Prieto R, Pappadopulos E, et al. Speed of
improvement in symptoms of depression with desvenlafaxine 50 mg and 100 mg compared with
placebo in patients with major depressive disorder. J Clin Psychopharmacol. 2017;37(5):555–61.
Kemp AH, Brunoni AR, Santos IS, Nunes MA, Dantas EM, Carvalho de Figueiredo R, et al. Effects
of depression, anxiety, comorbidity, and antidepressants on resting-state heart rate and its
variability: an ELSA-Brasil cohort baseline study. Am J Psychiatry. 2014;171(12):1328–34.
Kennedy SH, Lam RW, McIntyre RS, Tourjman SV, Bhat V, Blier P, et al. Canadian Network for
Mood and Anxiety Treatments (CANMAT) 2016 clinical guidelines for the management of
adults with major depressive disorder: section 3. Pharmacological treatments. Can J Psychiatr
Rev Can Psychiatr. 2016;61(9):540–60.
Kim SW, Shin IS, Kim JM, Park KH, Youn T, Yoon JS. Factors potentiating the risk of mirtazapine-
associated restless legs syndrome. Hum Psychopharmacol. 2008;23(7):615–20.
Kopera H. Cardiovascular tolerance of mianserin and interactions of mianserin with other drugs.
Acta Psychiatr Belg. 1978;78(5):787–97.
Lader M, Stender K, Burger V, Nil R. Efficacy and tolerability of escitalopram in 12- and 24-week
treatment of social anxiety disorder: randomised, double-blind, placebo-controlled, fixed-dose
study. Depress Anxiety. 2004;19(4):241–8.
Lavori PW, Keller MB, Mueller TI, Scheftner W. Recurrence after recovery in unipolar MDD: an
observational follow-up study of clinical predictors and somatic treatment as a mediating factor.
Int J Methods Psychiatr Res. 1994;4(4):211–29.
Levinson-Castiel R, Merlob P, Linder N, Sirota L, Klinger G. Neonatal abstinence syndrome after in
utero exposure to selective serotonin reuptake inhibitors in term infants. Arch Pediatr Adolesc
Med. 2006;160(2):173–6.
Montejo AL, Prieto N, Terleira A, Matias J, Alonso S, Paniagua G, et al. Better sexual acceptability
of agomelatine (25 and 50 mg) compared with paroxetine (20 mg) in healthy male volunteers.
An 8-week, placebo-controlled study using the PRSEXDQ-SALSEX scale. J Psychopharmacol.
2010;24(1):111–20.
Montgomery SA, Asberg M. A new depression scale designed to be sensitive to change. Br J
Psychiatry. 1979;134:382–9.
Moore RA, Derry S, Aldington D, Cole P, Wiffen PJ. Amitriptyline for neuropathic pain in adults.
Cochrane Database Syst Rev. 2015;7:CD008242.
Mullish BH, Kabir MS, Thursz MR, Dhar A. Review article: depression and the use of antidepres-
sants in patients with chronic liver disease or liver transplantation. Aliment Pharmacol Ther.
2014;40(8):880–92.
Nierenberg AA, Husain MM, Trivedi MH, Fava M, Warden D, Wisniewski SR, et al. Residual
symptoms after remission of major depressive disorder with citalopram and risk of relapse: a
STARD report. Psychol Med. 2010;40(1):41–50.
Ostad Haji E, Hiemke C, Pfuhlmann B. Therapeutic drug monitoring for antidepressant drug
treatment. Curr Pharm Des. 2012;18(36):5818–27.
Antidepressants: Indications, Contraindications, Interactions, and Side Effects 1171
Pacchiarotti I, Bond DJ, Baldessarini RJ, Nolen WA, Grunze H, Licht RW, et al. The International
Society for Bipolar Disorders (ISBD) task force report on antidepressant use in bipolar disor-
ders. Am J Psychiatry. 2013;170(11):1249–62.
Pratt LA, Brody DJ, Gu Q. Antidepressant use among persons aged 12 and over: United States,
2011–2014: US Department of Health and Human Services, Centers for Disease Control and
Prevention, Atlanta: National Center for Health Statistics; 2017.
Reichenpfader U, Gartlehner G, Morgan LC, Greenblatt A, Nussbaumer B, Hansen RA, et al.
Sexual dysfunction associated with second-generation antidepressants in patients with major
depressive disorder: results from a systematic review with network meta-analysis. Drug Saf.
2014;37(1):19–31.
Richa S, Yazbek JC. Ocular adverse effects of common psychotropic agents: a review. CNS Drugs.
2010;24(6):501–26.
Romano SJ, Halmi KA, Sarkar NP, Koke SC, Lee JS. A placebo-controlled study of fluoxetine in
continued treatment of bulimia nervosa after successful acute fluoxetine treatment. Am J
Psychiatry. 2002;159(1):96–102.
Sansone RA, Sansone LA. SSRI-induced indifference. Psychiatry (Edgmont). 2010;7(10):14–8.
Sansone RA, Sansone LA. Antidepressant adherence: are patients taking their medications? Innov
Clin Neurosci. 2012;9(5–6):41–6.
Serretti A. The present and future of precision medicine in psychiatry: focus on clinical psycho-
pharmacology of antidepressants. Clin Psychopharmacol Neurosci. 2018;16(1):1–6.
Serretti A, Chiesa A. Treatment-emergent sexual dysfunction related to antidepressants: a meta-
analysis. J Clin Psychopharmacol. 2009;29(3):259–66.
Serretti A, Mandelli L. Antidepressants and body weight: a comprehensive review and meta-
analysis. J Clin Psychiatry. 2010;71(10):1259–72.
Serretti A, Calati R, Goracci A, Di Simplicio M, Castrogiovanni P, De Ronchi D. Antidepressants in
healthy subjects: what are the psychotropic/psychological effects? Eur Neuropsychopharmacol.
2010;20(7):433–53.
Sinclair LI, Christmas DM, Hood SD, Potokar JP, Robertson A, Isaac A, et al. Antidepressant-
induced jitteriness/anxiety syndrome: systematic review. Br J Psychiatry. 2009;194(6):483–90.
Spina E, Trifiro G, Caraci F. Clinically significant drug interactions with newer antidepressants.
CNS Drugs. 2012;26(1):39–67.
Strawbridge R, Young AH, Cleare AJ. Biomarkers for depression: recent insights, current chal-
lenges and future prospects. Neuropsychiatr Dis Treat. 2017;13:1245–62.
Symes RJ, Etminan M, Mikelberg FS. Risk of angle-closure glaucoma with bupropion and
topiramate. JAMA Ophthalmol. 2015;133(10):1187–9.
Sysko R, Sha N, Wang Y, Duan N, Walsh BT. Early response to antidepressant treatment in bulimia
nervosa. Psychol Med. 2010;40(6):999–1005.
Szegedi A, Jansen WT, van Willigenburg AP, van der Meulen E, Stassen HH, Thase ME. Early
improvement in the first 2 weeks as a predictor of treatment outcome in patients with major
depressive disorder: a meta-analysis including 6562 patients. J Clin Psychiatry. 2009;70(3):344–53.
t Jong GW, Einarson T, Koren G, Einarson A. Antidepressant use in pregnancy and persistent
pulmonary hypertension of the newborn (PPHN): a systematic review. Reprod Toxicol.
2012;34(3):293–7.
Tampi RR, Balderas M, Carter KV, Tampi DJ, Moca M, Knudsen A, et al. Citalopram, QTc
prolongation, and Torsades de pointes. Psychosomatics. 2015;56(1):36–43.
Taylor D, Barnes TE, Young A. The Maudsley prescribing guidelines in psychiatry 2018. Newark:
John Wiley & Sons. 2018.
Tedeschini E, Levkovitz Y, Iovieno N, Ameral VE, Nelson JC, Papakostas GI. Efficacy of
antidepressants for late-life depression: a meta-analysis and meta-regression of placebo-con-
trolled randomized trials. J Clin Psychiatry. 2011;72(12):1660–8.
Telang S, Walton C, Olten B, Bloch MH. Meta-analysis: second generation antidepressants and
headache. J Affect Disord. 2018;236:60–8.
1172 F. Corponi et al.
Trindade E, Menon D, Topfer LA, Coloma C. Adverse effects associated with selective serotonin
reuptake inhibitors and tricyclic antidepressants: a meta-analysis. CMAJ. 1998;159(10):1245–52.
Uher R, Farmer A, Henigsberg N, Rietschel M, Mors O, Maier W, et al. Adverse reactions to
antidepressants. Br J Psychiatry. 2009;195(3):202–10.
Viktorin A, Lichtenstein P, Thase ME, Larsson H, Lundholm C, Magnusson PK, et al. The risk of
switch to mania in patients with bipolar disorder during treatment with an antidepressant alone
and in combination with a mood stabilizer. Am J Psychiatry. 2014;171(10):1067–73.
Viramontes TS, Truong H, Linnebur SA. Antidepressant-induced hyponatremia in older adults.
Consult Pharm. 2016;31(3):139–50.
Voican CS, Corruble E, Naveau S, Perlemuter G. Antidepressant-induced liver injury: a review for
clinicians. Am J Psychiatry. 2014;171(4):404–15.
Wang SM, Han C, Bahk WM, Lee SJ, Patkar AA, Masand PS, et al. Addressing the side effects of
contemporary antidepressant drugs: a comprehensive review. Chonnam Med J. 2018;54(2):101–12.
Warner MD, Peabody CA, Whiteford HA, Hollister LE. Trazodone and priapism. J Clin Psychiatry.
1987;48(6):244–5.
Wasserman D, Rihmer Z, Rujescu D, Sarchiapone M, Sokolowski M, Titelman D, et al. The
European Psychiatric Association (EPA) guidance on suicide treatment and prevention. Eur
Psychiatry. 2012;27(2):129–41.
WHO. Depression: fact sheet. Geneva: World Health Organisation. Available at: http://www.
whoint/mediacentre/factsheets/fs369/en/; 2017.
Wichniak A, Wierzbicka A, Walecka M, Jernajczyk W. Effects of antidepressants on sleep. Curr
Psychiatry Rep. 2017;19(9):63.
Wienkers LC, Allievi C, Hauer MJ, Wynalda MA. Cytochrome P-450-mediated metabolism of the
individual enantiomers of the antidepressant agent reboxetine in human liver microsomes. Drug
Metab Dispos. 1999;27(11):1334–40.
Wilkes S. The use of bupropion SR in cigarette smoking cessation. Int J Chron Obstruct Pulmon
Dis. 2008;3(1):45–53.
Yatham LN, Kennedy SH, Parikh SV, Schaffer A, Bond DJ, Frey BN, et al. Canadian Network for
Mood and Anxiety Treatments (CANMAT) and International Society for Bipolar Disorders
(ISBD) 2018 guidelines for the management of patients with bipolar disorder. Bipolar Disord.
2018;20(2):97–170.
Zhou N, Zhao JX, Zhu YN, Zhang P, Zuo Y. Acute angle-closure glaucoma caused by venlafaxine.
Chin Med J. 2018;131(12):1502–3.
Zohar J, Stahl S, Moller HJ, Blier P, Kupfer D, Yamawaki S, et al. A review of the current
nomenclature for psychotropic agents and an introduction to the neuroscience-based nomencla-
ture. Eur Neuropsychopharmacol. 2015;25(12):2318–25.
Antidepressants: Course and Duration of
Therapy, Withdrawal Syndromes, and
Resistance to Therapy
Contents
Course and Duration of Pharmacological Therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1174
Acute Treatment Phase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1174
Continuation Phase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1175
Recurrence Prevention Phase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1175
Withdrawal Syndromes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1176
Acute Discontinuation Syndrome/Withdrawal Syndrome . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1177
Severe Cases of ADS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1180
Persistent Post-discontinuation Syndromes and Rebound Phenomena . . . . . . . . . . . . . . . . . . . . 1180
Basic Principles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1181
Treatment and Prevention . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1181
Resistance to Therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1182
Clinical Assessment of Resistance to Therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1183
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1185
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1185
Abstract
The course and duration of therapy, withdrawal syndromes, and resistance to
therapy are important clinical topics in antidepressant therapy. The reviewed
clinical evidence may help to guide clinicians and patients when planning the
pharmacological part of the therapy.
The treatment with antidepressants is typically divided into different phases
that each require specific considerations.
The recommended course and duration of treatment may vary depending on
the respective treatment phase, as well as on individual patient characteristics.
Insufficient treatment response occurs frequently and a series of therapeutic
options are available such as increase in dosage, combining the current antide-
pressant with an additional antidepressant, augmentation with another pharma-
cological substance, and switching to a different antidepressant.
When antidepressant therapy is discontinued, patients may experience with-
drawal symptoms/syndromes. Symptoms are generally mild and transient, but,
depending on the individual drug, can be severe and impairing for individual
patients. Recognizing discontinuation syndromes and differentiating those from
re-emergence of the primary disorder is of crucial importance and will have major
consequences for further treatment. Indications and guidance for gradual tapering
of antidepressants will be discussed.
Generally, medication can be used for reducing symptom severity in clinical depres-
sion, particularly for severe forms of major depression (Cipriani et al. 2018a).
Medication, if necessary, should be combined with psychotherapeutic interventions
and social support.
Important aims of the pharmacological treatment in the acute phase are to reduce
symptoms of depressions that are associated with a burden of distress for the
individual person, to reduce mortality, and to increase psychosocial functioning.
The acute phase typically lasts for 6–12 weeks until individual treatment aims have
been fully or partially reached.
The degree of symptom reduction, also known as treatment response, can be
measured quantitatively by standardized rating scales (e.g., Beck’s Depression
Inventory, Hamilton Depression Scale, Hospital Anxiety and Depression Scale, or
Montgomery–Åsberg Depression Rating Scale). “Response” to treatment is typi-
cally defined as a 50% reduction in depression severity, as measured by standardized
rating scales. A broadly implemented four-tier definition of treatment response
categorizes the symptom reduction in percentage according to standardized rating
scale: <20% ¼ no significant effect; 20–50% ¼ minimal effect; 51–99% partial
remission; 100% ¼ full remission (Crismon et al. 1999).
In the acute treatment phase, medication will be increased in dosage until a low
start dosage has been reached (Hiemke et al. 2011). Current treatment guidelines for
unipolar depression recommend to increase the dosage stepwise and keep this initial
phase of the treatment as long as needed with regard to tolerability of the medication
but as short as possible to prevent delays in reaching the low start dosage (DGPPN
et al. 2015). Older persons or persons with relevant comorbidities such as
Antidepressants: Course and Duration of Therapy, Withdrawal Syndromes, and . . . 1175
cardiovascular illnesses or epilepsy may require smaller dosages (e.g., less than half
dosages of tricyclic antidepressants such as amitriptyline) and slower increases in
dosage to reduce the risk of adverse events (Arroll 2005; Leucht et al. 2012; Roose
and Spatz 1999). Patients should be well informed about critical factors such as
potential adverse events, planned course and duration of treatment, and potential
withdrawal phenomena prior to the start of the treatment. Regular clinical monitor-
ing with regard to occurrence of adverse events is of particular importance in this
first phase.
The time to treatment response may vary between individuals and according to
the specific antidepressant medication. In clinical routine, treatment response is often
assessed approximately four weeks after the beginning of the acute treatment phase
(DGPPN et al. 2015; NICE 2009). While a statistically significant and sustained
effect of antidepressant treatment can be found in clinical depression (Cipriani et al.
2018a; Henssler et al. 2018a), clinical significance of improvements are currently
under discussion (Cipriani et al. 2018b; Kirsch and Jakobsen 2018). Likewise, the
assumption that antidepressant efficacy is dependent on the severity of depression is
challenged by recent findings from patient-level data analyses (Rabinowitz et al.
2016).
Continuation Phase
The acute treatment phase is followed by the continuation phase that generally lasts
for 6–9 months (American Psychiatric Association (APA) 2000; Bauer et al. 2015):
As the name implies, the continuation phase is intended as a continuation of the
acute treatment with the same medication and dosage and the main aim is to
“stabilize” the response (i.e., avoid relapse of the depressive episode) that has been
reached in the acute treatment. It has been shown that the initial phase following the
acute treatment is characterized by an increased risk of relapse of the depressive
symptoms and withdrawal phenomena, and treatment during the continuation
phase may reduce the odds of relapse by 70% (Cipriani et al. 2018a; Geddes et al.
2003).
Recurrence prevention is the third phase that may start after the continuation phase
and is generally considered to be indicated for persons with at least two recent
depressive episodes that caused functional deficits (Frank et al. 1990; Geddes et al.
2003). Pharmacological recurrence prevention is not indicated for all patients and
limited to persons with an increased risk of recurrence such as persons with
sustaining factors (e.g., residual symptoms from the last depressive episode, low
coping resources, increasing number of depressive episodes, decreasing time in
between episodes, severity of recent depressive episodes, a history of suicidality,
onset early in life, a family history of major depression in first degree relatives,
1176 L. Brandt et al.
Withdrawal Syndromes
(Van Geffen et al. 2007). It seems essential to provide patients with relevant
information on possible withdrawal phenomena already at treatment initiation
(Brandt et al. 2020b).
In clinical practice, when facing a patient who experiences newly emergent
symptoms after discontinuation of dose reduction of an antidepressant, the following
possible differential diagnoses have to be distinguished:
• Flu-like symptoms
• Insomnia (disturbed sleep, vivid dreams/nightmares)
1178 L. Brandt et al.
• Nausea
• Imbalance (vertigo, light-headedness)
• Sensory disturbances (electric shock-like sensations, dysesthesia)
• Hyperarousal (anxiety, agitation, irritability, etc.)
It has been proposed to define withdrawal symptoms that persist for over more than 6
weeks as “persistent post-discontinuation syndromes” (Chouinard and Chouinard
2015). Also, mood and anxiety symptoms have been reported to re-emerge after
discontinuation to a greater extent than prior to initiation of the antidepressant (Fava
et al. 2007). These rebound phenomena are thought to represent an organism’s
increased susceptibility and vulnerability following drug discontinuation, most
Antidepressants: Course and Duration of Therapy, Withdrawal Syndromes, and . . . 1181
Basic Principles
Generally, the risk of ADS seems to appear with a minimum treatment duration of
4 weeks (Hohagen et al. 1994); robust data exists after 8 weeks. Afterwards, it
appears that the risk of ADS does not further increase with longer treatment
duration. ADS has been reported irrespective of the primary disorder, for which
the antidepressant treatment was initiated (Baldwin et al. 2007; Bogetto et al.
2002). As described above, incidence and severity of ADS are different between
single agents and classes of ADs. Additionally, plasma elimination time of the AD
plays a major role. In this vein, ADs with shorter elimination time inherit a
relatively higher risk of severe and frequent discontinuation symptoms. It has
been reported that ADS occurs after approximately 3–5 half-lives following
discontinuation or dose-reduction (Montgomery 2002). Discontinuation symp-
toms occur independent of medication dose within the normal therapeutic dosing
range, but have been reported to appear more frequently after high-dose treatment
(e.g., duloxetine 120 mg/day; escitalopram 20 mg/day) (Baldwin et al. 2007;
Markowitz et al. 2000; Perahia et al. 2005).
Resistance to Therapy
Cross-References
References
Adli M, Baethge C, Heinz A, Langlitz N, Bauer M. Is dose escalation of antidepressants a rational
strategy after a medium–dose treatment has failed? Eur Arch Psychiatry Clin Neurosci.
2005;255(6):387–400.
Ainsworth K, Smith SE, Zetterström TSC, Pei Q, Franklin M, Sharp T. Effect of antidepressant
drugs on dopamine D1 and D2 receptor expression and dopamine release in the nucleus
accumbens of the rat. Psychopharmacology. 1998;140(4):470–7.
Allgulander C, Florea I, Huusom A. Prevention of relapse in generalized anxiety disorder by
escitalopram treatment. Int J Neuropsychopharmacol. 2006;9(5):495–505.
American Psychiatric Association (APA). Practice guideline for the treatment of patients with major
depressive disorder. In: American Psychiatric Association, editor. Practice guidelines for the
treatment of people with psychiatric disorders. Washington, DC: American Psychiatric Associ-
ation (APA). Guideline Central; 2000. p. 413–96.
Arroll B. Efficacy and tolerability of tricyclic antidepressants and SSRIs compared with placebo for
treatment of depression in primary care: a meta-analysis. Ann Fam Med. 2005;3(5):449–56.
Baethge C, Gruschka P, Smolka MN, Berghöfer A, Bschor T, Müller-Oerlinghausen B, et al.
Effectiveness and outcome predictors of long-term lithium prophylaxis in unipolar major
depressive disorder. J Psychiatry Neurosci. 2003;28(5):355–61.
Baldwin DS, Cooper JA, Huusom AKT, Hindmarch I. A double-blind, randomized, parallel-group,
flexible-dose study to evaluate the tolerability, efficacy and effects of treatment discontinuation
1186 L. Brandt et al.
with escitalopram and paroxetine in patients with major depressive disorder. Int Clin
Psychopharmacol. 2006;21(3):159–69.
Baldwin DS, Montgomery SA, Nil R, Lader M. Discontinuation symptoms in depression and
anxiety disorders. Int J Neuropsychopharmacol. 2007;10(1):73–84.
Bartova L, Dold M, Kautzky A, Fabbri C, Spies M, Serretti A, et al. Results of the European Group
for the Study of Resistant Depression (GSRD) – basis for further research and clinical practice.
The World. J Biol Psychiatry. Taylor & Francis. 2019;20(6):427–48.
Bauer M, Döpfmer S. Lithium augmentation in treatment-resistant depression: meta-analysis of
placebo-controlled studies. J Clin Psychopharmacol. 1999;19(5):427–34.
Bauer M, Adli M, Baethge C, Berghöfer A, Sasse J, Heinz A, et al. Lithium augmentation therapy in
refractory depression: clinical evidence and neurobiological mechanisms. Can J Psychiatr.
2003;48(7):440–8.
Bauer M, Whybrow PC, Angst J, Versiani M, Möller HJ. Biologische Behandlung unipolarer
depressiver Störungen: Behandlungslinien der World of Federation of Societies of Biological
Psychiatry. 1st ed. Stuttgart: Wissenschaftliche Verlagsgesellschaft; 2004.
Bauer M, Pfennig A, Severus E, Whybrow PC, Angst J, Möller H-J, et al. World Federation of
Societies of Biological Psychiatry (WFSBP) guidelines for biological treatment of unipolar
depressive disorders, part 1: update 2013 on the acute and continuation treatment of unipolar
depressive disorders. World J Biol Psychiatry. 2013;14(5):334–85.
Bauer M, Severus E, Köhler S, Whybrow PC, Angst J, Möller H-J, et al. World Federation of
Societies of Biological Psychiatry (WFSBP) guidelines for biological treatment of unipolar
depressive disorders. Part 2: maintenance treatment of major depressive disorder-update 2015.
World J Biol Psychiatry. Taylor & Francis. 2015;16(2):76–95.
Baumann P, Hiemke C, Ulrich S, Eckermann G, Gaertner I, Gerlach M, et al. The AGNP-TDM
expert group consensus guidelines: therapeutic drug monitoring in psychiatry. Pharmacop-
sychiatry. 2004;37(6):243–65.
Berber MJ. FINISH: remembering the discontinuation syndrome. Flu-like symptoms, Insomnia,
Nausea, Imbalance, Sensory disturbances, and Hyperarousal (anxiety/agitation). J Clin Psychi-
atry. 1998;59(5):255.
Berigan TR. Bupropion-associated withdrawal symptoms revisited: a case report. Prim Care
Companion J Clin Psychiatry. 2002;4(2):78.
Bogetto F, Bellino S, Revello RB, Patria L. Discontinuation syndrome in dysthymic patients treated with
selective serotonin reuptake inhibitors: a clinical investigation. CNS Drugs. 2002;16(4):273–83.
Boschloo L, Spijker AT, Hoencamp E, Kupka R, Nolen WA, Schoevers RA, et al. Predictors of the
onset of manic symptoms and a (hypo) manic episode in patients with major depressive disorder.
PLoS One. 2014;9(9):e106871.
Brandt L, Bschor T, Henssler J, Müller M, Hasan A, Heinz A, et al. Antipsychotic withdrawal
symptoms: a systematic review and meta-analysis. Front Psych. 2020a;11:569912.
Brandt L, Henssler J, Gutwinski S. Entstehung, Merkmale, Prävention und Behandlung von
Absetzphänomenen. InFo Neurologie. 2020b;22(3):26–35.
Bschor T, Kern H, Henssler J, Baethge C. Switching the antidepressant after nonresponse in adults
with major depression: a systematic literature search and meta-analysis. J Clin Psychiatry.
2018;79(1):16r10749.
Burgess SS, Geddes J, Hawton KK, Taylor MJ, Townsend E, Jamison K, et al. Lithium for
maintenance treatment of mood disorders. Cochrane Database Syst Rev [Internet]. 2001[cited
2020 Nov 10];2001(3). Available from https://www.ncbi.nlm.nih.gov/pmc/articles/
PMC7005360/
Carter B, Strawbridge R, Husain MI, Jones BDM, Short R, Cleare AJ, et al. Relative effectiveness
of augmentation treatments for treatment-resistant depression: a systematic review and network
meta-analysis. Int Rev Psychiatry. Taylor & Francis; 2020;32(5–6):477–90.
Charney DS, Heninger GR, Sternberg DE, Landis H. Abrupt discontinuation of tricyclic antide-
pressant drugs: evidence for noradrenergic hyperactivity. Br J Psychiatry J Ment Sci. 1982;141:
377–86.
Antidepressants: Course and Duration of Therapy, Withdrawal Syndromes, and . . . 1187
Chouinard G, Chouinard V-A. New classification of selective serotonin reuptake inhibitor with-
drawal. Psychother Psychosom. 2015;84(2):63–71.
Cipriani A, Pretty H, Hawton K, Geddes JR. Lithium in the prevention of suicidal behavior and all-
cause mortality in patients with mood disorders: a systematic review of randomized trials. Am J
Psychiatry. 2005;15:1805–1819.
Cipriani A, Hawton K, Stockton S, Geddes JR. Lithium in the prevention of suicide in mood
disorders: updated systematic review and meta-analysis. BMJ. 2013;346(4):f3646.
Cipriani A, Furukawa TA, Salanti G, Chaimani A, Atkinson LZ, Ogawa Y, et al. Comparative efficacy
and acceptability of 21 antidepressant drugs for the acute treatment of adults with major depressive
disorder: a systematic review and network meta-analysis. Lancet. 2018a;391(10128):1357–66.
Cipriani A, Salanti G, Furukawa TA, Turner E, Ioannidis JPA, Geddes JR. Network meta-analysis
of antidepressants – authors’ reply. Lancet. 2018b;392(10152):1012–3.
Cowen PJ. Pharmacological management of treatment-resistant depression. Adv Psychiatr Treat.
1998;4:320–7.
Crismon ML, Trivedi M, Pigott T, Rush J, Hirschfeld R, Kahn D, et al. The Texas Medication
Algorithm Project: report of the Texas Consensus Conference Panel on medication treatment of
major depressive disorder. J Clin Psychiatry. 1999;15:142–156.
Davies J, Read J. A systematic review into the incidence, severity and duration of antidepressant
withdrawal effects: are guidelines evidence-based? Addict Behav. 2018;97:111–121.
Davies P, Ijaz S, Williams CJ, Kessler D, Lewis G, Wiles N. Pharmacological interventions for
treatment-resistant depression in adults. Cochrane Database Syst Rev. 2019;12:CD010557.
DGPPN, BÄK, KBV, AWMF. S3-Leitlinie/Nationale VersorgungsLeitlinie Unipolare Depression –
Langfassung, 2. Auflage [Internet]. Deutsche Gesellschaft für Psychiatrie, Psychotherapie und
Nervenheilkunde (DGPPN); Bundesärztekammer (BÄK); Kassenärztliche Bundesvereinigung
(KBV); Arbeitsgemeinschaft der Wissenschaftlichen Medizinischen Fachgesellschaften
(AWMF); 2015 [cited 2020 Nov 3]. Available from http://www.leitlinien.de/mdb/downloads/
nvl/depression/depression-2aufl-vers5-lang.pdf
Dilsaver SC, Greden JF. Antidepressant withdrawal phenomena. Biol Psychiatry. 1984;19(2):
237–56.
Dilsaver SC, Feinberg M, Greden JF. Antidepressant withdrawal symptoms treated with anticho-
linergic agents. Am J Psychiatry. 1983;140(2):249–51.
Edwards S, Hamilton V, Nherera L, Trevor N. Lithium or an atypical antipsychotic drug in the
management of treatment-resistant depression: a systematic review and economic evaluation.
Health Technol Assess. 2013;17(54):1.
Ellison JM. SSRI withdrawal buzz. J Clin Psychiatry. 1994;55(12):544–5.
Fauchere PA. Recurrent, persisting panic attacks after sudden discontinuation of mirtazapine
treatment: a case report. Int J Psychiatry Clin Pract. 2004;8(2):127–9.
Fava GA, Bernardi M, Tomba E, Rafanelli C. Effects of gradual discontinuation of selective
serotonin reuptake inhibitors in panic disorder with agoraphobia. Int J Neuropsychopharmacol.
2007;10(6):835–8.
Fava GA, Gatti A, Belaise C, Guidi J, Offidani E. Withdrawal symptoms after selective serotonin
reuptake inhibitor discontinuation: a systematic review. Psychother Psychosom. 2015;84(2):
72–81.
Fekadu A, Wooderson S, Donaldson C, Markopoulou K, Masterson B, Poon L, et al. A multi-
dimensional tool to quantify treatment resistance in depression: the Maudsley staging method. J
Clin Psychiatry. 2009;70(2):177–84.
Fekadu A, Donocik JG, Cleare AJ. Standardisation framework for the Maudsley staging method for
treatment resistance in depression. BMC Psychiatry. 2018;18(1):100.
Frank E, Kupfer D, Perel J. Three-year outcomes for maintenance therapies in recurrent depression.
Arch Gen Psychiatry. 1990;47(12):1093.
Gahr M, Schonfeldt-Lecuona C, Kolle MA, Freudenmann RW. Withdrawal and discontinuation
phenomena associated with tranylcypromine: a systematic review. Pharmacopsychiatry.
2013;46(4):123–9.
1188 L. Brandt et al.
Geddes JR, Carney S, Davies C, Furukawa T, Kupfer D, Frank E, et al. Relapse prevention with
antidepressant drug treatment in depressive disorders: a systematic review. Lancet. 2003;361:
653–1.
Giakas WJ, Davis JM. Intractable withdrawal from venlafaxine treated with fluoxetine. Psychiatr
Ann. 1997;27(2):85–92.
Giller E, Bialos D, Harkness L, Jatlow P, Waldo M. Long-term amitriptyline in chronic depression.
Hillside J Clin Psychiatry. 1985;7(1):16–33.
GlaxoSmithKline. A double-blind comparative study of withdrawal effects following abrupt dis-
continuation of treatment with paroxetine in low or high dose or imipramine. GSK – clinical
study register. https://www.gsk-clinicalstudyregister.com. 1992.
Goodwin G, Emsley R, Rembry S, Rouillon F. Agomelatine prevents relapse in patients with major
depressive disorder without evidence of a discontinuation syndrome: a 24-week randomized,
double-blind, placebo-controlled trial. J Clin Psychiatry. 2009;70(8):1128–37.
Groot PC, van Os J. Outcome of antidepressant drug discontinuation with tapering strips after 1–5
years. Ther Adv Psychopharmacol. 2020;10:2045125320954609.
Guzzetta F, Tondo L, Centorrino F, Baldessarini R. Lithium treatment reduces suicide risk in
recurrent major depressive disorder. J Clin Psychiatry. 2007;4:380–383.
Hardeveld F, Spijker J, De Graaf R, Nolen WA, Beekman ATF. Prevalence and predictors of
recurrence of major depressive disorder in the adult population. Acta Psychiatr Scand.
2010;122(3):184–91.
Hartford J, Kornstein S, Liebowitz M, Pigott T, Russell J, Detke M, et al. Duloxetine as an SNRI
treatment for generalized anxiety disorder: results from a placebo and active-controlled trial. Int
Clin Psychopharmacol. 2007;22(3):167–74.
Harvey BH, Slabbert FN. New insights on the antidepressant discontinuation syndrome. Human
Psychopharmacol. 2014;29(6):503–16.
Heinz AJ, Beck A, Meyer-Lindenberg A, Sterzer P, Heinz A. Cognitive and neurobiological
mechanisms of alcohol-related aggression. Nat Rev Neurosci. Nature Publishing Group;
2011;12(7):400–13.
Heinz, A., Daedelow, L., Wackerhagen, C., Di Chiara, G. Addiction theory matters – why there is
no dependence on caffeine or antidepressant medication. Addict Biol. 2019;25(2):e12735. (in
press).
Henssler J, Bschor T, Baethge C. Combining antidepressants in acute treatment of depression: a
meta-analysis of 38 studies including 4511 patients. Can J Psychiatr. 2016;61(1):29–43.
Henssler J, Kurschus M, Franklin J, Bschor T, Baethge C. Long-term acute-phase treatment with
antidepressants, 8 weeks and beyond: a systematic review and meta-analysis of randomized,
placebo-controlled trials. J Clin Psychiatry. 2018a;79(1):15r10545.
Henssler J, Kurschus M, Franklin J, Bschor T, Baethge C. Trajectories of acute antidepressant
efficacy: how long to wait for response? A systematic review and meta-analysis of long-term,
placebo-controlled acute treatment trials. J Clin Psychiatry. 2018b;79(3):17r11470.
Henssler J, Heinz A, Brandt L, Bschor T. Antidepressant withdrawal and rebound phenomena.
Dtsch Arztebl Int. 2019;116(20):355–61.
Hiemke C, Baumann P, Bergemann N, Conca A, Dietmaier O, Egberts K, et al. AGNP consensus
guidelines for therapeutic drug monitoring in psychiatry: update 2011. Pharmacopsychiatry.
2011;44(06):195–235.
Hohagen F, Montero RF, Weiss E, Lis S, Schonbrunn E, Dressing H, et al. Treatment of primary
insomnia with trimipramine: An alternative to benzodiazepine hypnotics? Eur Arch Psychiatry
Clin Neurosci. 1994;244(2):65–72.
Horowitz MA, Taylor D. Tapering of SSRI treatment to mitigate withdrawal symptoms. Lancet
Psychiatry. 2019;6(6):538–46.
Hotopf M, Hardy R, Lewis G. Discontinuation rates of SSRIs and tricyclic antidepressants: a meta-
analysis and investigation of heterogeneity. Br J Psychiatry. 1997;170(2):120–7.
Kaufman MJ, Henry ME, Frederick B, Hennen J, Villafuerte RA, Stoddard EP, et al. Selective
serotonin reuptake inhibitor discontinuation syndrome is associated with a rostral anterior
Antidepressants: Course and Duration of Therapy, Withdrawal Syndromes, and . . . 1189
Spijker J, de Graaf R, Bijl RV, Beekman ATF, Ormel J, Nolen WA. Duration of major depressive
episodes in the general population: results from The Netherlands Mental Health Survey and
Incidence Study (NEMESIS). Br J Psychiatry. 2002;181:208–13.
Stein DJ, Ahokas A, Albarran C, Olivier V, Allgulander C. Agomelatine prevents relapse in
generalized anxiety disorder: a 6-month randomized, double-blind, placebo-controlled discon-
tinuation study. J Clin Psychiatry. 2012;73(7):1002–8.
Stoukides JA, Stoukides CA. Extrapyramidal symptoms upon discontinuation of fluoxetine. Am J
Psychiatry. 1991;148(9):1263.
Tint A, Haddad PM, Anderson IM. The effect of rate of antidepressant tapering on the incidence of
discontinuation symptoms: a randomised study. J Psychopharmacol (Oxford, England). 2008;22
(3):330–2.
Tondo L, Hennen J, Baldessarini RJ. Lower suicide risk with long-term lithium treatment in major
affective illness: a meta-analysis. Acta Psychiatr Scand. 2001;104(3):163–72.
Tyrer P. Clinical effects of abrupt withdrawal from tri-cyclic antidepressants and monoamine
oxidase inhibitors after long-term treatment. J Affect Disord. 1984;6(1):1–7.
Van Geffen ECG, Brugman M, Van Hulten R, Bouvy ML, Egberts ACG, Heerdink ER. Patients’
concerns about and problems experienced with discontinuation of antidepressants. Int J Pharm
Pract. 2007;15(4):291–3.
Vandel P, Sechter D, Weiller E, Pezous N, Cabanac F, Tournoux A, et al. Post-treatment emergent
adverse events in depressed patients following treatment with milnacipran and paroxetine.
Human Psychopharmacol. 2004;19(8):585–6.
Westermeyer J. Addiction to tranylcypromine (Parnate): a case report. Am J Drug Alcohol Abuse.
Taylor & Francis;. 1989;15(3):345–50.
Wittchen H-U, Beesdo K, Bittner A, Goodwin RD. Depressive episodes–evidence for a causal role
of primary anxiety disorders? Eur Psychiatry. 2003;18(8):384–93.
Yiend J, Paykel E, Merritt R, Lester K, Doll H, Burns T. Long term outcome of primary care
depression. J Affect Disord. 2009;118(1–3):79–86.
Zajecka J, Fawcett J, Amsterdam J, Quitkin F, Reimherr F, Rosenbaum J, et al. Safety of abrupt
discontinuation of fluoxetine: a randomized, placebo-controlled study. J Clin Psychopharmacol.
1998;18(3):193–7.
Tricyclics: Imipramine, Clomipramine,
Trimipramine (Dibenzazepines)
Contents
Chemistry, Developmental History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1194
Pharmacology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1195
Pharmacokinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1195
Pharmacodynamics/Mechanism of Action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1195
Indications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1197
Clinical Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1197
Side Effects/Adverse Reactions and Toxicology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1198
Toxicity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1200
Combination Therapies, Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1201
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1201
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1201
Abstract
Imipramine was the first tricyclic antidepressant (TCA), discovered serendipi-
tously being the prototype of antidepressants for long time. Clomipramine and
trimipramine are further developed dibenzazepine TCAs belonging to the
tertiary-amine tricyclics. Numerous randomized clinical trials (RCTs) have pro-
ved the antidepressive efficacy for acute and long-term treatment of depressions.
For clomipramine some studies and clinical experience claim higher efficacy in
H.-P. Volz
Krankenhaus für Psychiatrie, Psychotherapie und Psychosomatische Medizin Schloss Werneck,
Werneck, Germany
G. Laux (*)
Institute of Psychological Medicine (IPM), Soyen, Germany
MVZ Waldkraiburg, Center of Neuropsychiatry, Waldkraiburg, Germany
Department of Psychiatry and Psychotherapy, Ludwig-Maximilians-University (LMU), Munich,
Germany
e-mail: ipm@ipm-laux.de
Tricyclics have a central three-ring structure, hence the name tertiary-amine tricyclics
being representatives of a chemical group in which two benzene rings are fused to a
central, nitrogen-containing seven-membered ring, the 1H-azepine (Dibenzoazepines).
Examples are imipramine, clomipramine, and trimipramine, but also the anxio-
lytic opipramol or the anti-convulsant carbamazepine. Figure 1 shows the chemical
structure of imipramine as an example.
Imipramine was the first TCA to be developed by Ciba in 1951, as an antihista-
mine. In 1957 the Swiss psychiatrist Roland Kuhn observed that imipramine ame-
liorated symptoms in depressed patients, so serendipitously antidepressant effects
have been discovered. Subsequently, imipramine was introduced for the treatment of
depression in Europe in 1958 and in the United States in 1959 being the prototypical
drug for the development of the later-released TCAs. The discoveries of imipramine,
the first TCA, and iproniazid, the first monoamine oxidase inhibitor (MAOI), were
results of astute and rather fortuitous clinical observations and did not emerge form
systematic pharmacological development (Humble 2000).
Clomipramine was developed by Geigy as a chlorinated derivative of Imipramine.
It was first referenced in the literature in 1961 and was patented in 1963, approved for
medical use in Europe in the treatment of depression in 1970. Clomipramine remains
the only TCA in the United States that is not approved for the treatment of depression,
in spite of the fact that it is a highly effective antidepressant. It has been approved in
the United States for the treatment of OCD in 1989 and became available in 1990. It
was the first drug to be investigated and found effective in the treatment of OCD. It is
on the World Health Organization’s List of Essential Medicines.
Trimipramine was developed by Rhône-Poulenc, patented in 1959 and first
introduced for medical use in 1966 in Europe. It was not introduced in the United
States, is available worldwide except Australia (Heal 1997).
Pharmacology
Pharmacokinetics
Pharmacodynamics/Mechanism of Action
The most important pharmacodynamic property of tricyclics are the inhibition of the
reuptake of noradrenaline or serotonin at the presynaptic site. However, this is only
true for clomipramine and imipramine, whereas trimipramine exerts nearly no
inhibiting effect on the presynaptic noradrenaline and serotonin transporter. Clomip-
ramine acts mainly on the serotonin transporter, imipramine has a nearly balanced
inhibiting potential (Richelson and Nelson 1984; Owens et al. 1997; Humble 2000).
However, the compounds do not only exhibit activities on the presynaptically
located noradrenaline and serotonin transporter but also at different receptor sites
located in the CNS but also in the periphery. The most important receptors which
those compounds usually block are the muscarinergic-acetylcholine receptor
(AchR), the histamine-1-receptor (H1R), the alpha-1-adrenoreceptor (α1R), and the
Serotonin-2(A)-Receptor (5HT2[A] (Gillman 2007, Haenisch et al. 2011).).
Trimipramine is described as an atypical or “second-generation” TCA because,
unlike other TCAs, it seems to be a fairly weak monoamine reuptake inhibitor
Similarly to other TCAs however, trimipramine does have antihistamine, anti-
serotonergic, antiadrenergic, antidopaminergic, and anticholinergic activities. Its
well-documented antidepressant action cannot be explained by noradrenaline or
serotonin reuptake inhibition or by a down-regulation of beta-adrenoceptors. Fur-
thermore, its receptor affinity profile – trimipramine is the only antidepressant
showing also activity at the Dopamine-2-receptor (D2) – resembles more that of
clozapine, a neuroleptic/antipsychotic drug, than that of tricyclic antidepressants
(Gastpar 1989). Trimipramine does not reduce, but rather increases, rapid eye
movement sleep. It stimulates nocturnal prolactin secretion and inhibits nocturnal
1196 H.-P. Volz and G. Laux
cortisol secretion and may act at the level of the hypothalamus on corticotropin-
releasing hormone secretion (Berger and Gastpar 1996).
In Table 2 the respective activities of imipramine, clomipramine, and tri-
mipramine are summarized qualitatively.
The different receptor profiles are shown in Fig. 2:
Legend Imipramine
5HT2
5HT2 9%
NE-receptor NE-receptor
5HT1 5HT-receptor 17%
AChM
29%
5HT-receptor
D2 DA-receptor 11%
α1 2%
AChM α1
H2 α2 H1 32%
Clomipramine Trimipramine
5HT2
D2 NE-receptor
7%
1% 5%
AChM
13%
H1
10% H1
α1 5HT-receptor
3%
61%
Fig. 2 Receptor binding profiles of imipramine, clomipramine, and trimipramine. (Adapted from
Fritze 2002)
Tricyclics: Imipramine, Clomipramine, Trimipramine (Dibenzazepines) 1197
Indications
Clinical Studies
positive results have been reported in chronic pain syndromes (Nilsson and von
Knorring 1989).
According to his atypical pharmacological properties trimipramine is effective in
the therapy of insomnia and psychotic, delusional depression (Pecknold and Luthe
1989; Gross et al. 1991; Sonntag et al. 1996; Berger and Gastpar 1996; Riemann
et al. 2002; Künzel et al. 2009).
Table 3 gives a summary of important controlled clinical studies:
Thus, overall, there seems to be no major efficacy difference of the tricyclics
compared to other antidepressants.
Table 3 Surveys Review RCTs of imipramine (I), clomipramine (C) and trimipramine (T) in
depression
Study arms, Results Relevant and
N, Patients comparators, (Response-rate Dropout severe side
Study Duration Placebo RR) rates effects
Prien et al. N ¼ 77 I N ¼ 38 Relapse rates
(1973) 20 months Pl N ¼ 39 I 37%
RCT Pl 67%
Frank et al. N ¼ 128 I N ¼ 28 Relapse rates
(1990) 3 years IPT þ I N ¼ 25 21%
RCT IPT þ Pl N ¼ 26 24%
IPT N ¼ 26 66%
Pl N ¼ 23 62%
78%
Effective
relapse
prevention of I
at 200 mg/d
Wehmeier N ¼ 41 T N ¼ 21,150 mg/d T ¼ F T¼F
et al. (2005) geriatric Fluoxetine (F)
RCT MDD pts N ¼ 20 20 mg/d
6 weeks
Pecknold et al. N ¼ 39 T N ¼ 20, T ¼ M in M>T
(1985) Anxiety- 150 mg/d depression and M weight gain, T
RCT depression Maprotiline anxiety scales # RR, lower
5 weeks (M) N ¼ 19, anticholinergic,
150 mg/d neurologic and
cardiovascular
effects
Greist et al. N ¼ 1520 C N ¼ 520 C > F, Fl,
(1995) Fluoxetine (F) N ¼ S > Pl
Meta-analysis 355
4 multicenter, Fluvoxamine
placebo- (Fl) N ¼ 320
controlled Sertraline (S) N ¼
studies 325
Arroll et al. Primary TCAs vs Pl RR TCA 60% NNH TCA 5–11
(2005) care pts. Pl 45%
Meta-analysis 6–8 weeks NNT TCA 4
12 RCT Low dose also
studies effective
Sonntag et al. N ¼ 20 T vs I T enhanced
(1996) male inpts REM and slow
RCT Sleep wave sleep,
EEG, total sleep time
nocturnal increased
hormone with T, cortisol
secretion secretion
decreased
with T
(continued)
1200 H.-P. Volz and G. Laux
Table 3 (continued)
Study arms, Results Relevant and
N, Patients comparators, (Response-rate Dropout severe side
Study Duration Placebo RR) rates effects
Künzel et al. N ¼ 94 T N ¼ 33 Ø RR: T 84.8% Not reported
(2009) Delusional 356 mg/d A/H 70.8%
RCT depression Amitriptyline/ Decrease of
haloperidol (A/H) cortisol and
N ¼ 24 Ø ACTH
184/6 mg/d response in the
Dex/CRH test
in both groups
Note > more effective or more frequent than
NNT number needed to treat, NNH number needed to harm, Randomized Controlled Trial (RCT)
Toxicity
TCAs have greater toxicity than SSRIs, SNRIs, or mirtazapine (Hawton et al. 2010).
Clomipramine was associated with a significant degree of toxicity in overdose,
trimipramine is relatively safe in overdose (White et al. 2008).
Contraindications for tricyclics include recent myocardial infarction, heart block
or other cardiac arrhythmias, severe liver disease, narrow angle glaucoma, and
urinary retention.
Pregnancy: For imipramine positive evidence of risk to human fetus (risk cate-
gory D), for clomipramine and trimipramine risk category C (no controlled studies in
humans, some animal studies show adverse effects).
Withdrawal symptoms may occur during gradual or particularly abrupt with-
drawal of tricyclic antidepressant drugs. Possible symptoms include: nausea,
vomiting, abdominal pain, diarrhea, insomnia, headache, nervousness, anxiety,
dizziness and worsening of psychiatric status.
Tricyclics: Imipramine, Clomipramine, Trimipramine (Dibenzazepines) 1201
Cross-References
References
Anderson I. Selective serotonin reuptake inhibitors versus tricyclic antidepressants; a meta-analysis
of efficacy and tolerability. J Affect Disord. 2000;58:19–36.
Arroll B, Macgillivray S, Ogston S, et al. Efficacy and tolerability of tricyclic antidepressants and
SSRIs compared with placebo for treatment of depression in primary care: a meta-analysis. Ann
Fam Med. 2005;3:449–56.
Balant-Gorgia AE, Gex-Fabry M, Balant LP. Clinical pharmacokinetics of clomipramine. Clin
Pharmacokinet. 1991;20:447–62.
Berger M, Gastpar M. Trimipramine: a challenge to current concepts on antidepressives. Eur Arch
Psychiatry Clin Neurosci. 1996;246:235–9.
Danish University Antidepressant Group. Citalopram: clinical effectprofile in comparison with
clomipramine: a controlled multicenter study. Psychopharmacology. 1986;90:131–8.
Del Casale A, Sorice S, Padovano A, et al. Psychopharmacological treatment of obsessive-
compulsive disorder (OCD). Curr Neuropharmacol. 2019;17:710–36.
Fayez R, Gupta V. Imipramine. StatPearls [Internet]. Treasure Island: StatPearls Publishing; 2021.
Frank E, Kupfer DJ, Perel JM, et al. Three-year outcomes for maintenance therapies in recurrent
depression. Arch Gen Psychiatry. 1990;47:1093–9.
Fritze J. Unerwünschte Wirkungen, Kontraindikationen, Überdosierung, Intoxikation. In:
Riederer P, Laux G, Pöldinger W, editors. Neuropsychopharmaka. Ein Therapiehandbuch.
Band 3: Antidepressiva, Phasenprophylaktika und Stimmungsstabilisierer. Wien: Springer;
2002. p. 162–82.
1202 H.-P. Volz and G. Laux
Gastpar M. Clinical originality and new biology of trimipramine. Drugs. 1989;38(Suppl 1):43–8.
Gerson SC, Plotkin DA, Jarvik LF. Antidepressant drug studies 1964 to 1986: empirical evidence
for aging patients. J Clin Psychopharmacol. 1988;8:311–22.
Gillman PK. Tricyclic antidepressant pharmacology and therapeutic drug interactions updated. Br
J Pharmacol. 2007;151:737–48.
Greist JH, Jefferson JW, Kobak KA, et al. Efficacy and tolerability of serotonin transport inhibitors
in obsessive-compulsive disorder. A meta-analysis. Arch Gen Psychiatr. 1995;52:53–60.
Gross G, Xin X, Gastpar M. Trimipramine: pharmacological reevaluation and comparison with
clozapine. Neuropharmacology. 1991;30:1159–6.
Haenisch B, Hiemke C, Bönisch H. Inhibitory potencies of trimipramine and its main metabolites at
human monoamine and organic cation transporters. Psychopharmacology. 2011;217:289–95.
Hawton K, Bergen H, Simkin S, Cooper J, et al. Toxicity of antidepressants: rates of suicide relative
to prescribing and non-fatal overdose. Br J Psychiatry. 2010;196:354–8.
Heal D. The antidepressant era. Harvard University Press; 1997. ISBN 9780674039575
Hiemke C, Bergemann N, Clement HW, Conca A, et al. Consensus guidelines for therapeutic drug
monitoring in neuropsychopharmacology: update 2017. Pharmacopsychiatry. 2018;51:9–62.
Humble M. Noradrenaline and serotonin reuptake inhibition as clinical principles: a review of
antidepressant efficacy. Acta Psychiatr Scand Suppl. 2000;402:28–36.
Künzel HE, Ackl N, Hatzinger M, et al. Outcome in delusional depression comparing trimipramine
monotherapy with a combination of amitriptyline and haloperidol – a double-blind multicentre
trial. L Psyhiatr Res. 2009;43:702–10.
Lapierre YD. A review of trimipramine. 30 years of clinical use. Drugs. 1989;38(suppl 1):17–24.
McTavish D, Benfield P. Clomipramine. An overview of its pharmacological properties and a
review of its therapeutic use in obsessive compulsive disorder and panic disorder. Drugs.
1990;39:136–53.
Mediq.ch. Accessed 15 Feb 2020.
Möller HJ, Fuger J, Kasper S. Efficacy of new generation antidepressants: meta-analysis of
imipramine-controlled studies. Pharmacopsychiatry. 1994;27:215–23.
Montgomery SA, Baldwin DS, Blier P, et al. Which antidepressants have demonstrated superior
efficacy? A review of the evidence. Int Clin Psychopharmacol. 2007;22:323–9.
Morris JB, Beck AT. The efficacy of antidepressant drugs. A review of research (1958-1972). Arch
Gen Psychiatry. 1974;30:667–74.
Nilsson HL, von Knorring L. Review. Clomipramine in acute and chronic pain syndrome. Nord
Psykiatr Tidsskr. 1989;43:101–13.
Owens MJ, Morgan WN, Plott SJ, Nemeroff CB. Neurotransmitter receptor and transporter binding
profile of antidepressants and their metabolites. J Pharmacol Exp Ther. 1997;283:1305–22.
Papp LA, Schneier FR, Fyer AJ, et al. Clomipramine treatment of panic disorder: pros and cons.
J Clin Psychiatry. 1997;58:423–5.
Pecknold JC, Luthe L. Trimipramine, anxiety, depression and sleep. Drugs. 1989;38(Suppl
1):25–31.
Pecknold JC, Familamiri P, McClure DJ, Elie R, Chang H. Trimipramine and maprotiline: antide-
pressant, anxiolytic, and cardiotoxic comparison. J Clin Psychiatry. 1985;46:166–71.
Peters MD, Davis SK, Austin LS. Clomipramine: an antiobsessional tricyclic antidepressant. Clin
Pharm. 1990;9:165–78.
Prien RF, Klett CJ, Caffey EM Jr. Lithium carbonate and imipramine in prevention of affective
episodes. A comparison in recurrent affective illness. Arch Gen Psychiatry. 1973;29:420–5.
Psiac.de. Accessed 15 Feb 2020.
Richelson E, Nelson A. Antagonism by antidepressants of neurotransmitter receptors of normal
human brain in vitro. J Pharmacol Exp Ther. 1984;230:94–102.
Riemann D, Voderholzer U, Cohrs S, et al. Trimipramine in primary insomnia: results of a
polysomnographic double-blind controlled study. Pharmacopsychiatry. 2002;35:165–74.
Sallee FR, Pollock BG. Clinical pharmacokinetics of imipramine and desipramine. Clin
Pharmacokinet. 1990;18:346–64.
Tricyclics: Imipramine, Clomipramine, Trimipramine (Dibenzazepines) 1203
Sonntag A, Rothe B, Guldner J, et al. Trimipramine and imipramine exert different effects on the
sleep EEG and on nocturnal hormone secretion during treatment of major depression. Depres-
sion. 1996;4:1–13.
Stahl SM. Stahl’s essential psychopharmacology. In: The prescriber’s guide. Antidepressants. 6th
ed. Cambridge University Press; 2017.
Wehmeier PM, Kluge M, Maras A, Riemann D, et al. Fluoxetine versus trimipramine in the
treatment of depression in geriatric patients. Pharmacopsychiatry. 2005;38:13–6.
White N, Litovitz T, Clancy C. Suicidal antidepressant overdoses: a comparative analysis by
antidepressant type. J Med Toxicol. 2008;4(4):238–50.
Amitriptyline and Depressions
Mellar P. Davis
Contents
Chemistry, Developmental History, and Chemical Graphics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1206
Pharmacology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1206
Pharmacodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1206
Pharmacokinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1208
Indication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1209
Clinical Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1210
Depression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1210
Other Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1212
Amitriptyline and Migraines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1214
Amitriptyline and Pain: Recent Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1214
Amitriptyline and Insomnia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1215
Adverse Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1215
Drug Interactions and Drug Combinations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1217
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1218
Abstract
Amitriptyline was the second tricyclic antidepressant to be introduced into
practice in the United States. In a network meta-analysis, amitriptyline is equiv-
alent to selective serotonin reuptake inhibitors (SSRIs) and is superior to trazo-
done, fluoxetine, and reboxetine in efficacy. Dropouts for reasons of adverse
effects were higher than for most antidepressants particularly in the outpatient
setting. Therapeutic drug monitoring improves clinical utility since there are
wide individual differences in serum drug levels per dose. Drug interactions are
a result of competitive inhibition with the mixed function oxidases CYP2D6 and
M. P. Davis (*)
Department of Palliative Care, Geisinger Medical Center, Danville, PA, USA
e-mail: mdavis2@geisinger.edu
Amitriptyline was synthesized in 1960 and introduced into the United States in 1961
and in the United Kingdom in 1962 (Pressman and Weiss 1961). It was the second
tricyclic antidepressant to be introduced for the treatment of depression. Imipramine
was the first which was available in 1957. Amitriptyline is a dibenzocycloheptadiene
(Fig. 1). The chemical name is 3-(10,11)-dihydro-5H-dibenzo [a, d] cycloheptene-5-
y(diene)-N, N-dimethyl-propan-1-amine.
Pharmacology
Pharmacodynamics
Amitriptyline has a high affinity for norepinephrine and serotonin transporters but
very low affinity for dopamine transporters. It binds to and inhibits adrenergic,
histaminergic, and muscarinic receptors which are largely responsible for side effects
associated with amitriptyline. Amitriptyline also binds to certain serotonin receptors
(5HT1B, 5HT2A, 5HT2B, and 5HT2C) (Vaishnavi et al. 2004; Mestres et al. 2011;
Giros et al. 1992) (Table 1). The hypnotic effects of amitriptyline are thought to be
due to histamine receptors (H1, H2) and certain serotonin receptors (5HT2A and
5HT2C) (Atkin et al. 2018).
Several recent studies demonstrated amitriptyline effects beyond monoamine
receptor and transporter interactions. The hippocampus plays a major role in depres-
sion. Reductions in neurogenesis within the hippocampus are associated with major
depression disorders. The Janus kinase 3 (Jak-3) is induced in stress-related major
depression which causes inhibition of neurogenesis within the hippocampus as
described in animal models. Amitriptyline reduces Jak-3 expression through the
Amitriptyline and Depressions 1207
Table 1 Affinity of amitriptyline for transporters and receptors (Vaishnavi et al. 2004; Mestres
et al. 2011)
Transporter/receptor K1 (nM) IC50(nM)
Norepinephrine transporter 13.3 63
Serotonin transporter 3.13 47
Dopamine transporter 2580 7500
Norepinephrine receptor 24 690
Muscarinic receptor 7.2 –
Histamine receptor – 26(H1)
Serotonin-5HT1B – 40
Serotonin-5HT2A – 4
Sertonin-5HT2B – 40
Serotonin-5HT2C – 6
Ki-binding constant of the protein in concentration
IC50-concentration at which 50% of the protein function is inhibited
Pharmacokinetics
Indication
Clinical Studies
Depression
AMT, amitriptyline; RCT, randomized control trials; OR, odds ratio; RR, relative risk; SSRIs, selective serotonin reuptake inhibitors; SMD, standard mean
difference
1212 M. P. Davis
Other Studies
Other studies have bearing on amitriptyline utility and are worth reviewing. These
studies include patients with comorbidities as well as major depression disorders.
Patients with melancholic depression, depression, and panic disorders or psychotic
depression have not been included in randomized trials reviewed in the previous
section. An important study has analyzed whether lower doses of tricyclic antide-
pressants can be used and still maintain a response but have fewer side effects.
In a Cochrane Collaboration review, low doses (<100 mg/d) of tricyclic antide-
pressants as a drug class were compared to standard doses using response (>50%
decrease) in the Hamilton Assessment of Major Depression Rating Scale (HAMD)
(Furukawa et al. 2003). Low-dose tricyclic antidepressants as a class were superior
to placebo at 4 weeks (OR 1.65, 95% CI 1.36–2.0) and at 8 weeks (OR 1.47, 95%CI
1.12–1.94). Adverse effects were 63% higher than placebo, and dropout were 111%
higher than placebo. However, low-dose tricyclics were 55% less likely to produce
dropouts than standard doses.
A systematic review and meta-analysis of 41 trials involving antidepressants in
the treatment of panic disorders found that tricyclic antidepressants as a class were
superior to SSRIs and serotonin norepinephrine reuptake inhibitors (SNRIs)
(Bighelli et al. 2018). However, tricyclics also produced greater numbers of dropout.
Definitions of response in the study was “very much or much improved” on the
Clinical Global Impression by investigators or more than 40% reduction in Panic
Disorder Severity scale or 50% decrease in Fear Questionnaire Agoraphobia sub-
scale. The tricyclic antidepressants involved in trials were imipramine, clomipra-
mine, and desipramine but not amitriptyline.
Though tricyclic antidepressants have been used frequently in certain countries to
treat childhood depression, a meta-analysis of 14 studies found that the tricyclics as a
class did not improve depression (RR for response 1.07, 95%CI 0.91–1.29) (Hazell
and Mirzaie 2013). Some reduction in depression symptoms did occur (standard
mean differences [SMD], 0.32,95%CI 0.59 to 0.04). Adverse effects were
significantly worse with tricyclics relative to placebo in children. The risk of vertigo
(RR 2.76, 95%CI 1.73–4.43), orthostatic hypotension (RR 4.86, 95%CI1.69–13.97),
tremors (RR 5.53, 95%CI 1.64–17.98), and dry mouth (RR 3.35, 95%CI 1.98–5.64).
SSRIs were deemed safer due to potential overdose deaths associated with tricyclic
antidepressants. The wide confidence intervals demonstrate the imprecision in the
analysis secondary to low numbers of participants in the studies.
Amitriptyline and Depressions 1213
In adults with chronic, nonspecific, low back pain, low-dose amitriptyline (25 mg/d)
or an active comparator (benztropine mesylate, 1 mg/d) for 6 months. There were no
significant improvements in outcomes including pain intensity at 6 months, but there
was a reduction in disability at 3 months (Urquhart et al. 2018). For masticatory
myofascial pain, nortriptyline and amitriptyline were compared in a randomized
trial. The 50% improvement rate with nortriptyline was better (P = 0.036) (Haviv
et al. 2019). Depression and chronic pain often coexist. Of 1166 people prescribed
opioids for chronic non-cancer pain, half the cohort (N = 637, 54.6%) used antide-
pressants. Three hundred twenty-nine (51.7%) reported moderate to severe depres-
sion. Amitriptyline was the most commonly used antidepressant (17.3%) (Gisev
et al. 2019). Patients with chronic neck pain were randomized between amitriptyline
at night and placebo. Amitriptyline significantly lowers pain scores greater than
in the placebo group (3.34 +/ 1.45 vs. 6.12 +/ 0.92; p < 0.0001), which
Amitriptyline and Depressions 1215
Many drugs are used off-label for treatment of insomnia. The most common are
antidepressants such as trazodone, mirtazapine, and amitriptyline. There is little
evidence that these drugs are effective in treating insomnia not associated with
depression (2018). There was no evidence for amitriptyline use for primary insomnia
(Everitt et al. 2018). Sertraline, fluoxetine, and amitriptyline appear to increase
periodic limb movements that do not disrupt sleep (Kolla et al. 2018).
Adverse Effects
One of the difficulties in measuring side effects is that antidepressant side effects can
be learned using a conditioning paradigm and evoked via a placebo pill. This may
result in a crossover effect with placebo side effects looking very similar to active
treatment. This should inform us when we read about side effects (Rheker et al.
2017). In systematic reviews of randomized trials comparing amitriptyline to pla-
cebo, side effects experienced significantly more often were largely related to
anticholinergic toxicities: tachycardia, nervousness, sedation, tremor, dyspepsia,
sexual dysfunction, and weight gain (Leucht et al. 2012). Anticholinergic adverse
effects may be worse at the time of peak plasma concentrations. In a study dividing
the dose of amitriptyline to every 8 h appears to reduce dry mouth and sedation
compared to a single-dose bedtime dose which is the usual dosing strategy (Gupta
et al. 1999). Individuals starting out on amitriptyline are better at recognizing
treatment-emerging side effects. However, individuals on long-term amitriptyline
may eventually complain of anticholinergic side effects they did not recognize with
initial therapy. Patients may initially tolerate amitriptyline anticholinergic effects and
not recognize symptoms as side effects, and yet over time tolerance diminishes at a
threshold at which the symptom is recognized as a side effect (Bryant et al. 1987).
In early studies, the incidence of side effects recognized by patients occurred in
15.4%. Major adverse effects described in an early study were psychosis, halluci-
nations, disorientation, and agitation in 4.6% of a group of patients treated in the
Boston Collaborative Drug Surveillance Program (Jick et al. 1972). However, these
severe adverse effects have not been described in recent trials and may have been
related to treating psychotic depression or bipolar disorders with amitriptyline; both
of which are poorly responsive to amitriptyline alone. Amitriptyline converts depres-
sion to mania in 42% of individuals treated for bipolar depression (Koszewska and
Rybakowski 2009).
1216 M. P. Davis
There has been concern that hormonal contraceptives may increase amitriptyline
exposure through pharmacokinetic interactions (Berry-Bibee et al. 2016). The com-
bined use of carbamazepine with amitriptyline reduces mean amitriptyline levels by
42% and nortriptyline by 40% (Leinonen et al. 1991). Valproic acid has been
combined with amitriptyline to treat refractory depression. Valproic acid signifi-
cantly increases amitriptyline serum levels (Unterecker et al. 2013). On the other
hand, the combination of mirtazapine (30 mg) and amitriptyline (75 mg) has minor
and clinically insignificant interactions. Mirtazapine may in fact prevent this sero-
tonin syndrome by blocking postsynaptic serotonin receptors (Sennef et al. 2003).
The combination of citalopram with amitriptyline has been used for individuals with
depression and either migraine or tension headaches who have not responded to
either drug alone. There are no increases in major side effects or serotonergic adverse
effects with the combination (Rampello et al. 2004). However, citalopram and
escitalopram like amitriptyline prolongs the QTC interval and may cause torsades
de pointes. The combination, if used, should be monitored by ECG (Tampi et al.
2015). The risk appears to be related to the CYP2C19 genotype and not citalopram
blood levels (Kumar et al. 2014). CYP2C19 poor metabolizers will be at particular
risk with the combination (Chang et al. 2014).
Antidepressants have been augmented with coadministration of atypical antipsy-
chotics, most commonly aripiprazole, quetiapine, and olanzapine, for treatment of
refractory depression, psychotic depression, and depression associated with schizo-
phrenia or anxiety (Katzman et al. 2008; Brawman-Mintzer et al. 2005). The most
common combination consists of a SSRI and an atypical antipsychotic. Drug-drug
interactions are a risk when combining atypical antipsychotics with amitriptyline.
For instance, risperidone and aripiprazole are both metabolized through CYP2D6
and could competitively inhibit amitriptyline clearance. Delayed clearance of both
drugs increases the risk for extrapyramidal side effects (Berman et al. 2009; Wright
et al. 2013). One systematic review of augmentation therapy with atypical antipsy-
chotics for depression included 35 studies, and no combinations included amitrip-
tyline (Wright et al. 2013).
A small study found the combination of olanzapine in average dose 6.5 mg plus
amitriptyline effectively treated melancholic depression which had failed to respond
to amitriptyline alone. The response (50% reduction in score) based on the
Montgomery-Asberg Depression Rating Scale (MADRS) occurred in 10 of 23 avail-
able patients. The combination appeared to be effective and safe; however, the study
was not a randomized trial and was of short duration (Takahashi et al. 2008).
Amitriptyline has also been combined with another antipsychotic, haloperidol (aver-
age dose 9 mg) and compared with risperidone for patients with schizoaffective
1218 M. P. Davis
References
Achar E, Achar RA, Paiva TB, Campos AH, Schor N. Amitriptyline eliminates calculi through
urinary tract smooth muscle relaxation. Kidney Int. 2003;64:1356–64.
Agabio R, Trogu E, Pani PP. Antidepressants for the treatment of people with co-occurring
depression and alcohol dependence. Cochrane Database Syst Rev. 2018;4:CD008581.
Alcocer-Gomez E, De Miguel M, Casas-Barquero N, Nunez-Vasco J, Sanchez-Alcazar JA,
Fernandez-Rodriguez A, Cordero MD. NLRP3 inflammasome is activated in mononuclear
blood cells from patients with major depressive disorder. Brain Behav Immun. 2014;36:111–7.
Amitriptyline and Depressions 1219
Dean L. Amitriptyline therapy and CYP2D6 and CYP2C19 Genotype. In: Pratt V, McLeod H,
Rubinstein W, Dean L, Malheiro A, editors. Medical genetics summaries [Internet]. Bethesda:
National Center for Biotechnology Information (US); 2017; 2012-.2017 Mar 23.
Drugs for chronic insomnia. Med Lett Drugs Ther. 2018;60:201–205.
Eshun-Wilson I, Siegfried N, Akena DH, Stein DJ, Obuku EA, Joska JA. Antidepressants for
depression in adults with HIV infection. Cochrane Database Syst Rev. 2018;1:CD008525.
Everitt H, Baldwin DS, Stuart B, Lipinska G, Mayers A, Malizia AL, Manson CC, Wilson
S. Antidepressants for insomnia in adults. Cochrane Database Syst Rev. 2018;5:CD010753.
Franke L, Schewe HJ, Uebelhack R, Muller-Oerlinghausen B. Predictors of therapeutic effects in
amitriptyline treatment – 1. Plasma drug levels. Pharmacopsychiatry. 2003;36:134–42.
Furlanut M, Benetello P. The pharmacokinetics of tricyclic antidepressant drugs in the elderly.
Pharmacol Res. 1990;22:15–25.
Furukawa T, McGuire H Barbui C. Low dosage tricyclic antidepressants for depression. Cochrane
Database Syst Rev. 2003;(3):CD003197.
Gassen NC, Hartmann J, Zschocke J, Stepan J, Hafner K, Zellner A, Kirmeier T,
Kollmannsberger L, Wagner KV, Dedic N, Balsevich G, Deussing JM, Kloiber S, Lucae S,
Holsboer F, Eder M, Uhr M, Ising M, Schmidt MV, Rein T. Association of FKBP51 with
priming of autophagy pathways and mediation of antidepressant treatment response: evidence in
cells, mice, and humans. PLoS Med. 2014;11:e1001755.
Gassen NC, Hartmann J, Schmidt MV, Rein T. FKBP5/FKBP51 enhances autophagy to synergize
with antidepressant action. Autophagy. 2015;11:578–80.
Giros B, El Mestikawy S, Godinot N, Zheng K, Han H, Yang-Feng T, Caron MG. Cloning,
pharmacological characterization, and chromosome assignment of the human dopamine trans-
porter. Mol Pharmacol. 1992;42:383–90.
Gisev N, Nielsen S, Campbell G, Santo T, Mant A, Bruno R, Cohen M, Hall WD, Larance B,
Lintzeris N, Farrell M, Degenhardt L. Antidepressant use among people prescribed opioids for
chronic noncancer pain. Pain Med. 2019;20:2450–8.
Gomolin IH, Melmed CA. Prolonged delirium without anticholinergic signs following amitriptyline
overdose. Can Med Assoc J. 1983;129:1203–4.
Guaiana G, Barbui C, Hotopf M. Amitriptyline for depression. Cochrane Database Syst Rev.
2007;18:CD004186.
Gulbins A, Grassme H, Hoehn R, Kohnen M, Edwards MJ, Kornhuber J, Gulbins E. Role of Janus-
kinases in major depressive disorder. Neurosignals. 2016;24:71–80.
Gupta SK, Shah JC, Hwang SS. Pharmacokinetic and pharmacodynamic characterization of OROS
and immediate-release amitriptyline. Br J Clin Pharmacol. 1999;48:71–8.
Haviv Y, Zini A, Sharav Y, Almoznino G, Benoliel R. Nortriptyline compared to amitriptyline for
the treatment of persistent masticatory myofascial pain. J Oral Facial Pain Headache.
2019;33:7–13.
Hazell P, Mirzaie M. Tricyclic drugs for depression in children and adolescents. Cochrane Database
Syst Rev. 2013;CD002317.
Hefner G, Hahn M, Hohner M, Roll SC, Klimke A, Hiemke C. QTc time correlates
with amitriptyline and venlafaxine serum levels in elderly psychiatric inpatients.
Pharmacopsychiatry. 2018;52:38–43.
Hiemke C, Bergemann N, Clement HW, Conca A, Deckert J, Domschke K, Eckermann G,
Egberts K, Gerlach M, Greiner C, Grunder G, Haen E, Havemann-Reinecke U, Hefner G,
Helmer R, Janssen G, Jaquenoud E, Laux G, Messer T, Mossner R, Muller MJ, Paulzen M,
Pfuhlmann B, Riederer P, Saria A, Schoppek B, Schoretsanitis G, Schwarz M, Gracia MS,
Stegmann B, Steimer W, Stingl JC, Uhr M, Ulrich S, Unterecker S, Waschgler R, Zernig G,
Zurek G, Baumann P. Consensus guidelines for therapeutic drug monitoring in neuropsycho-
pharmacology: update 2017. Pharmacopsychiatry. 2018;51:9–62.
Jick H, et al. Adverse reactions to the tricyclic-antidepressant drugs. Report from Boston Collab-
orative Drug Surveillance Program. Lancet. 1972;1:529–31.
Katzman MA, Vermani M, Jacobs L, Marcus M, Kong B, Lessard S, Galarraga W, Struzik L,
Gendron A. Quetiapine as an adjunctive pharmacotherapy for the treatment of non-remitting
Amitriptyline and Depressions 1221
Contents
Chemistry, Developmental History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1225
Pharmacology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1226
Pharmacokinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1226
Mechanism of Action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1226
Indications (of Marketed Products) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1226
Clinical Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1226
Side Effects/Adverse Reactions and Toxicology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1227
Combination Therapy: Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1231
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1231
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1232
Abstract
Amoxapine is a tricyclic antidepressant (TCA) and used in the treatment of major
depressive disorder. It is a norepinephrine and serotonin reuptake inhibitor. A
more rapid onset of action is discussed in comparison to other antidepressants.
There are notable, especially anticholinergic, side effects such as constipation and
paralytic ileus, urinary retention, and ECG changes. Amoxapine is potentially
lethal when overdosed.
Pharmacology
Pharmacokinetics
The absorption of amoxapine is rapid and almost complete from the GI tract. It is
extensively metabolized by hepatic hydroxylation resulting in the active metabolites
7-hydroxyamoxapine and 8-hydroxyamoxapine. These metabolites are conjugated
to form glucuronides (Uptodate.com, 12/12/2018). The time to peak serum levels is
1.5 h. Half-life of amoxapine is 8 h and of the metabolite 8-hydroxyamoxapine 30 h
(Jue et al. 1982). It is excreted predominantly in urine as conjugated metabolites
(drugs.com, 12/12/2018).
Mechanism of Action
Clinical Studies
Already in 1978, Hekimian et al. compared the therapeutic efficacy and the onset of
action of amoxapine and amitriptyline. This double-blind study showed for a smaller
number of patients that amoxapine was as effective as amitriptyline and had an
earlier onset of action (Hekimian et al. 1978).
Amoxapine and Depressions 1227
Table 1 (continued)
Study (RCT, meta-
analysis, NIS; Study arms,
name/acronym, N, patients comparators, Results (scores, response/ Relevant and severe side
authors, year Country(ies) placebo remissions rates) Dropout rates effects
Amoxapine as an 54 Amoxapine PANSS: Total score mean 18 of 54 No fewer extrapyramidal
antipsychotic: 150–250 mg, difference amoxapine side effects in amoxapine
comparative study haloperidol 15.74, haloperidol 13.41 group
versus haloperidol; 5–15 mg CGI
Chaudhry et al. CDSS
2007.
BAS barnes akathisia rating scale, BDI beck depression inventory, BPRS brief psychiatric rating scale, CDSS calgary depression scale for schizophrenia, CGI
clinical global impression scale, HAM-D hamilton rating scale for depression, IMPS inpatient multidimensional psychiatric rating scale, MADRS montgomery-
asberg depression rating scale, NOSIE nurses’ observation scale for inpatient evaluation, PANSS positive and negative syndrome scale, SAS simpson-angus scale
for extrapyramidal symptoms, ZSDS zung self-rating depression scale
D. Rujescu et al.
Amoxapine and Depressions 1231
Cross-References
References
Apiquian R, Fresan A, Ulloa RE, de la Fuente-Sandoval C, Herrera-Estrella M, Vazquez A, Nicolini
H, Kapur S. Amoxapine as an atypical antipsychotic: a comparative study vs risperidone.
Neuropsychopharmacology. 2005;30:2236–44.
Ban TA, Wilson WH, McEvoy JP. Amoxapine: a review of literature. Int Pharmacopsychiatry.
1980;15:166–70.
Buckley NA, McManus PR. Can the fatal toxicity of antidepressant drugs be predicted with
pharmacological and toxicological data? Drug Saf. 1998;18:369–81.
Chaudhry IB, Husain N, Khan S, Badshah S, Deakin B, Kapur S. Amoxapine as an antipsychotic:
comparative study versus haloperidol. J Clin Psychopharmacol. 2007;27:575–81.
Davis JM, Wang Z, Janicak PG. A quantitative analysis of clinical drug trials for the treatment of
affective disorders. Psychopharmacol Bull. 1993;29:175–81.
Drugbank.ca. https://www.drugbank.ca/drugs/DB00543. Accessed 12 Nov 2019.
Fitzgerald PB, Filia S, De Castella A, McBain N, Kapur S, Kulkarni J. Amoxapine in schizophrenia:
a negative double-blind controlled trial. J Clin Psychopharmacol. 2004;24:448–50.
Food and Drug Administration (FDA). https://www.accessdata.fda.gov/drugsatfda_docs/label/
2014/072691s036lbl.pdf. Accessed 16 Dec 2018.
Fruensgaard K, Hansen CE, Korsgaard S, Nymgaard K, Vaag UH. Amoxapine versus amitriptyline
in endogenous depression. A double-blind study. Acta Psychiatr Scand. 1979;59:502–8.
Hekimian LJ, Friedhoff AJ, Deever E. A comparison of the onset of action and therapeutic efficacy
of amoxapine and amitriptyline. J Clin Psychiatry. 1978;39:633–7.
Jue SG, Dawson GW, Brogden RN. Amoxapine: a review of its pharmacology and efficacy in
depressed states. Drugs. 1982;24:1–23.
Kapur S, Cho R, Jones C, McKay G, Zipursky RB. Is amoxapine an atypical antipsychotic?
Positron-emission tomography investigation of its dopamine2 and serotonin2 occupancy. Biol
Psychiatry. 1999;45:1217–20.
Kim S, Chen J, Cheng T, Gindulyte A, He J, He S, Li Q, Shoemaker BA, Thiessen PA, Yu B,
Zaslavsky L, Zhang J, Bolton EE. PubChem 2019 update: improved access to chemical data.
Nucleic Acids Res. 2019;47(D1):D1102–9.
Kinney JL, Evans RL Jr. Evaluation of amoxapine. Clin Pharm. 1982;1:417–24.
Neuroscience based Nomenclature. http://www.nbn2.com/search#drug/6. Accessed 14 Dec 2018.
Stahl SM. Prescriber’s guide. 6th ed. Cambridge, MA: Cambridge University Press; 2017.
Tatsumi M, Groshan K, Blakely RD, Richelson E. Pharmacological profile of antidepressants and
related compounds at human monoamine transporters. Eur J Pharmacol. 1997;11:249–58.
Uptodate.com. Accessed 14 Dec 2018.
White N, Litovitz T, Clancy C. Suicidal antidepressant overdoses: a comparative analysis by
antidepressant type. J Med Toxicol. 2008;4:238–50.
Winek CL, Wahba WW, Rozin L. Amoxapine fatalities: three case studies. Forensic Sci Int.
1984;26:33–8.
Nortriptyline and Maprotiline
for Depressions
Gerd Laux
Contents
Chemistry and Developmental History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1234
Pharmacology and Pharmacokinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1234
Mechanism of Action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1235
Indications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1235
Clinical Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1236
Side Effects, Adverse Reactions, and Toxicology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1237
Combination Therapy: Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1237
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1238
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1238
Abstract
Nortriptyline is a tricyclic antidepressant compound (TCA) methylated of the
parent amitriptyline. It is a serotonin and norepinephrine reuptake inhibitor. The
drug has a “therapeutic window,” that is, a curvilinear relationship between
therapeutic response and plasma levels.
Maprotiline is a tetracyclic compound closely related to nortriptyline
exhibiting strong effects as a norepinephrine reuptake inhibitor showing a high
degree of selectivity.
Both are released for treatment of major depressions, they are also used off-label
for the treatment of panic disorder, chronic pain/neuralgia, and neuropathic pain.
The most common side effects of nortriptyline include dry mouth, sedation,
constipation, increased appetite, blurred vision, and tinnitus. Maprotiline causes
G. Laux (*)
Institute of Psychological Medicine (IPM), Soyen, Germany
MVZ Waldkraiburg, Center of Neuropsychiatry, Waldkraiburg, Germany
Department of Psychiatry and Psychotherapy, Ludwig-Maximilians-University (LMU), Munich,
Germany
e-mail: ipm@ipm-laux.de
Nortriptyline
Maprotiline
NE-Uptake
alpha-adrenerg
histaminerg
anticholinerg
Nortriptyline Maprotiline
Mechanism of Action
Indications
Clinical Studies
The most common side effects of nortriptyline are central and peripheral
anticholinergic effects like dry mouth, sedation, constipation, urinary hesitancy,
increased appetite, blurred vision, and tinnitus. Cardiovascular effects like ortho-
static hypotension are less pronounced compared to other TCAs; however,
quinide-like effects on cardiac conduction are possible. From a clinical point of
view the substance was well tolerated also in elderly patients (Rosen et al. 1993).
Maprotiline causes a strong initial sedation (first 2–3 weeks of therapy) and is
therefore indicated to treat agitated patients. It can cause anticholinergic side effects
(dry mouth, constipation, confusion, tachycardia) with a lower incidence than ami-
triptyline. Most often seen are antihistamine effects like sedation, dizziness, drowsi-
ness, increased appetite, and weight gain. Leukopenia and skin reactions occur more
often with maprotiline than with comparable drugs like amitriptyline. Maprotiline
caused a significant prolongation of atrioventricular and intraventricular conduction
and a rise in heart rate. Although these effects were not clinically relevant in samples of
patients without overt heart disease, they should be taken into account when treating
depressed patients with concomitant cardiac disease (Hewer et al. 1995). Under higher
doses (>200 mg/d) the risk of seizures is increased (Molnar 1993).
Nortriptyline should not be used in the acute recovery phase after myocardial
infarction, hypertrophy of the prostate gland with urine hesitancy, closed angle
glaucoma, and along with a monoamine oxidase (MAO) inhibitor.
A nortriptyline overdose is considered a medical emergency and frequently
results in death. Symptoms of overdose include irregular heartbeat, seizures, coma,
confusion, hallucination, widened pupils, drowsiness, agitation, fever, low body
temperature, stiff muscles, and vomiting.
Nortriptyline may cause problems if taken during pregnancy. Use during
breastfeeding appears to be relatively safe.
Contraindications for maprotiline are hypertrophy of the prostate gland with urine
hesitancy, closed angle glaucoma, serious cardiovascular conditions, and epilepsy.
During pregnancy it should be used only if clearly needed, and it is excreted in breast
milk corresponding closely to blood concentrations.
Overdose (>1 g) can be toxic and fatal due to cardiac arrhythmias and CNS
depression.
Cross-References
References
Asberg M, Cronholm B, Sjöqvist F, Tuck D. Relationship between plasma level and therapeutic
effect of nortriptyline. Br Med J. 1971;3:331–4.
Baumann PA, Maitre L. Neurobiochemical aspects of maprotiline (Ludiomil). J Int Med Res.
1979;7:391–400.
Cusack B, Nelson A, Richelson E. Binding of antidepressants to human brain receptors: focus on
newer generation compounds. Psychopharmacology. 1994;114:559–65.
Derry SW, Moore AD. Nortriptyline for neuropathic pains in adults. Cochrane Database Syst Rev.
2015; https://doi.org/10.1002/14651858.CD011605.
Flint A, Rifat S. The effect of sequential antidepressant treatment on geriatric depression. J Affect
Disord. 1996;36:95–105.
Georgotas A, McCue RE, Hapworth W, et al. Comparative efficacy and safety of MAOIs versus
TCAs in treating depression in the elderly. Biol Psychiatry. 1986;21:1155–66.
Ghaemi NS. Clinical psychopharmacology. Oxford: Oxford University Press; 2019.
Grüter W, Pöldinger W. Maprotiline. Mod Probl Pharmacopsychiatry. 1982;18:1–30.
Gwirtsman HE, Ahles S, Halaris A, et al. Therapeutic superiority of maprotiline versus doxepin in
geriatric depression. J Clin Psychiatry. 1983;44:449–53.
Hewer W, Rost W, Gattaz F. Cardiovascular effects of fluvoxamine and maprotiline in depressed
patients. Eur Arch Psychiatry Clin Neurosci. 1995;246:1–6.
Kuhn R. Die Therapie der larvierten Depression. In: Kielholz P, editor. Die larvierte Depression.
Bern: Huber; 1973. p. 194–203.
Lindsay PG, Olsen RB. Maprotiline in pain depression. J Clin Psychiatry. 1985;46:226–8.
McCue RE. Using tricyclic antidepressants in the elderly. Clin Geriatr Med. 1992;8:323–34.
Molnar G. Seizures associated with high maprotiline serum concentrations. Can J Psychiatr.
1993;28:555–6.
Mulsant BH, Blumberger DM, Ismail Z, et al. A systematic approach to the pharmacotherapy of
geriatric major depression. Clin Geriatr Med. 2014;30:517–34.
Nelson JC. Tricyclic and tetracyclic drugs. In: Schatzberg AF, Nemeroff CB, editors. Textbook of
psychopharmacology. 5th ed. Washington, DC: American Psychiatric Publishing; 2017.
Perry PJ, Zeilmann C, Arndt S. Tricyclic antidepressant concentrations in plasma: an estimate of
their sensitivity and specificity as a predictor of response. J Clin Psychopharmacol. 1994;14:
230–40.
Nortriptyline and Maprotiline for Depressions 1239
Pinder RM, Brogden RN, Speight TM, Avery GS. Maprotiline: a review of its pharmacological
properties and therapeutic efficacy in mental depressive states. Drugs. 1977;13:321–52.
Power RA, Muthen B, Henigsberg N, et al. Non-random drop-out and the relative efficacy of
escitalopram and nortriptyline in treating major depressive disorder. Psychiatry Res. 2012;46:
1333–8.
Reynolds CF. Nortriptyline and interpersonal psychotherapy as maintenance therapies for recurrent
major depression. JAMA. 1999;281:39–45.
Richelson E, Nelson A. Antagonism of antidepressants of neurotransmitter receptors of normal
human brain in vitro. J Pharmacol Exp Ther. 1984;230:94–102.
Rosen J, Sweet R, Pollock BG, et al. Nortriptyline in the hospitalized elderly: tolerance and side
effect reduction. Psychopharmacol Bull. 1993;29:327–31.
Rouillon F, Serrurier PR, et al. Recurrence of unipolar depression and efficacy of maprotiline.
L'Encéphale. 1989;15:527–34.
Seppi K, Weintraub D, Coelho M, et al. The movement disorder society evidence-based medicine
review update: treatments for the non-motor symptoms of Parkinson’s disease. Mov Disord.
2011;26:S42–80.
Stahl S. Essential psychopharmacology. The prescriber's guide. Antidepressants. 6th
ed. Cambridge: Cambridge University Press; 2017.
Tatsumi M, GroshanK BRD, Richelson E. Pharmacological profile of antidepressants and related
compounds at human monoamine transporters. Eur J Pharmacol. 1997;340:249–58.
Van der Velde CD. Maprotiline versus imipramine and placebo in neurotic depression. J Clin
Psychiatry. 1981;42:138–41.
Weissman MM, Lieb J, Prusoff B, Bothwell S. A double-blind trial of maprotiline (Ludiomil) and
amitriptyline in depressed outpatients. Acta Psychiatr Scand. 1975;52:225–36.
Wells BG, Gelenberg AJ. Chemistry, pharmacology, pharmacokinetics, adverse effects, and effi-
cacy of the antidepressant maprotiline hydrochloride. Pharmacotherapy. 1981;1:121–39.
Mianserine and Depressions
Contents
Chemistry, Developmental History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1241
Pharmacology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1242
Indications (of Marketed Products) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1242
Clinical Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1243
Side Effects/Adverse Reactions and Toxicology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1243
Combination Therapy: Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1246
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1246
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1246
Abstract
Mianserine is a mood-enhancing and anxiety-relieving tetracyclic antidepressant,
which enhances noradrenergic neurotransmission and interacts with serotonin
receptors.
Pharmacology
Mianserine is absorbed after oral administration with peak plasma concentrations being
attained 2–3 h after ingestion. Plasma concentration increases progressively during
continued treatment, reaching possible steady state after 2 weeks. The maximum
antidepressant effect occurs after a delay within 2–4 weeks (Brogden et al. 1978).
The profile of action differs from that of tricyclic antidepressants. Mianserine
enhances noradrenergic neurotransmission by presynaptic α-adrenoceptor-blocking
activity and interacts with serotonin receptors but has no central anticholinergic
activity. It is a histamine, alpha1 and alpha2 antagonist (Brogden et al. 1978).
doses or as a single bedtime dose of 60 mg. The treatment is started gradually, and
the withdrawal should take place gradually as well to prevent withdrawal symptoms
(Brogden et al. 1978).
Mianserine should not be used in case of hypersensitivity, mania, and severe liver
disease as well as in combination with monoaminooxidase (MAO) inhibitors. It was
approved in Switzerland in 1981 and is commercially available as film-coated tablets
(Generika). The original Tolvon ® is no longer marketed.
Clinical Studies
Mianserine should not be combined with MAO inhibitors. Blood pressure should be
monitored if antihypertensive drugs are administered simultaneously. Mianserine has
an EEG and clinical activity profile similar to that of amitriptyline. It is a substrate of
CYP3A4, and corresponding drug interactions with inducers and inhibitors are pos-
sible. Interactions with vitamin K antagonists are also described (Brogden et al. 1978).
Cross-References
References
Berilgen MS. Delirium associated with mianserine in demented patients. J Clin Psychopharmacol.
2010;30(4):467–8.
Bonne O, Shalev AY, Bloch M. Delirium associated with mianserine. Eur Neuropsychopharmacol.
1995;5(2):147–9.
Brogden RN, Heel RC, Speight TM, Avery GS. Mianserine: a review of its pharmacological
properties and therapeutic efficacy in depressive illness. Drugs. 1978;16(4):273–301.
Haine SE, Miljoen HP, Blankoff I, Vrints CJ. Mianserine and ventricular tachycardia: case report
and review of the literature. Cardiology. 2006;106(4):195–8.
Itil TM, Polvan N, Hsu W. Clinical and EEG effects of GB-94, a “tetracyclic” antidepressant (EEG
model in discovery of a new psychotropic drug). Curr Ther Res Clin Exp. 1972;14(7):395–413.
Karlsson I, Godderis J, Augusto De Mendonca Lima C, Nygaard H, Simanyi M, Taal M, et al. A
randomised, double-blind comparison of the efficacy and safety of citalopram compared to
mianserine in elderly, depressed patients with or without mild to moderate dementia. Int
J Geriatr Psychiatry. 2000;15(4):295–305.
Kleemann A, Engel J, Kutscher B, Reichert D. Pharmaceutical substances, 2 Bände. 4th
ed. Stuttgart: Thieme-Verlag; 2000. ISBN 978-1-58890-031-9
Mizuki Y, Kajimura N, Kai S, Suetsugi M, Ushijima I, Yamada M. Effects of mianserine in chronic
schizophrenia. Prog Neuro-Psychopharmacol Biol Psychiatry. 1992;16(4):517–28.
Van der Burg WJ, Bonta IL, Delobelle J, Ramon C, Vargaftig B. Novel type of substituted
piperazine with high antiserotonin potency. J Med Chem. 1970;13:35. https://doi.org/10.1021/
jm00295a010.
van Heeringen K, Zivkov M. Pharmacological treatment of depression in cancer patients. A
placebo-controlled study of mianserine. Br J Psychiatry. 1996;169(4):440–3.
Wakeling A. Efficacy and side effects of mianserine, a tetracyclic antidepressant. Postgrad Med
J. 1983;59(690):229–31.
Trazodone and Depression
Gerd Laux
Contents
Chemistry, Developmental History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1248
Pharmacology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1248
Pharmacokinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1248
Mechanism of Action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1249
Indications (of Marketed Products) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1249
Clinical Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1249
Side Effects, Adverse Reactions, and Toxicology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1252
Combination Therapy: Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1253
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1254
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1254
Abstract
Trazodone is a multimodal, multifunctional antidepressant working as an antag-
onist of 5-HT2A and 5-HT2B receptors, as a weak 5-HT-reuptake inhibitor and a
partial agonist of the 5-HT1A receptor, and an antagonist of the α1- and α2-
adrenergic and histamine receptors. Trazodone has shown antidepressant efficacy
and effectiveness in randomized placebo-controlled clinical studies (RCTs) and
noninterventional studies; long-term studies are missing, however. Frequently
low-dose trazodone is used off-label in the treatment of insomnia. Common side
effects include sedation, dizziness, somnolence, and orthostatic hypotension; a
relatively rare side effect associated with trazodone is priapism.
G. Laux (*)
Institute of Psychological Medicine (IPM), Soyen, Germany
MVZ Waldkraiburg, Center of Neuropsychiatry, Waldkraiburg, Germany
Department of Psychiatry and Psychotherapy, Ludwig-Maximilians-University (LMU), Munich,
Germany
e-mail: ipm@ipm-laux.de
Pharmacology
Trazodone is the first member of the serotonin antagonist and reuptake inhibitor
(SARI) class, described as a multimodal, multifunctional antidepressant (Stahl
2009). Trazodone is an antagonist of 5-HT2A and 5-HT2B receptors, a weak
5-HT-reuptake inhibitor and a partial agonist of the 5-HT1A receptor, and an antag-
onist of the α1- and α2-adrenergic and histamine receptors (Cusack et al. 1994;
Tatsumi et al. 1997; Richelson 2002; Stahl 2009).
Pharmacokinetics
Mechanism of Action
Clinical Studies
Data from open and double-blind trials suggest the antidepressant efficacy of
trazodone is comparable to that of TCAs (amitriptyline, doxepin, and imipramine),
mianserin, and SSRIs (Fagiolini et al. 2012). Three double-blind studies reported
trazodone has antidepressant efficacy similar to that of other antidepressants in
geriatric patients being helpful for severe agitation and insomnia mainly (Gerner
1987). Comparative studies in a total of 320 evaluable elderly patients with major
depression suggest that trazodone at therapeutic doses is superior to placebo and as
effective as amitriptyline, imipramine, fluoxetine, and mianserin in relieving depres-
sive symptoms (Haria et al. 1994). A survey review of important studies is given in
Table 1.
1250
Common side effects include sedation, dizziness, nausea, feeling faint, vomiting, dry
mouth, and headache. More serious side effects may include suicide, mania, irregular
heart rate (arrhythmogenic), pathologically prolonged erections, and orthostatic hypo-
tension, which may cause dizziness and increase the risk of falling, and can have
devastating consequences for elderly patients mainly. Most frequent side effects in a
recent study were dizziness (11,2%) and somnolence (8,7%) (Fagiolini et al. 2020).
A relatively rare side effect associated with trazodone is priapism, likely due to its
antagonism at α-adrenergic receptors. More than 200 cases have been reported, and
the manufacturer estimated that the incidence of any abnormal erectile function is
about one in 6000 male patients treated with trazodone. The risk for this side effect
appears to be greatest during the first month of treatment at low dosages
(i.e., <150 mg/day). Early recognition of any abnormal erectile function is impor-
tant, including prolonged or inappropriate erections, and should prompt discontin-
uation of trazodone. In females, increased libido has been reported (Abber et al.
1987; Clayton et al. 2002).
Trazodone and Depression 1253
Trazodone should not be taken with serotonergic agents like SSRIs, tramadol,
fentanyl, dextro-methorphane, pethidine, St John’s wort, tryptophan, or 5-HTP as
the resulting drug interaction could lead to serotonin syndrome. Use with caution in
patients with history of seizures.
CYP3A4 inhibitors such as clarithromycin, erythromycin, fluvoxamine,
grapefruit juice, ketoconazole, and ritonavir may lead to increased concentrations
of trazodone, CYP3A4 inducers like carbamazepine, phenytoin, and St. John’s
wort may result in decreased trazodone concentrations. CYP2D6 inhibitors
(bupropion, cannabidiol, and metoclopramide) may result in increased concen-
trations of both trazodone and mCPP. CYP1A2 inhibitors (ciprofloxacin) may
increase trazodone concentrations while CYP1A2 inducers (rifampin, ritonavir,
and tobacco) may decrease trazodone concentrations (Fagiolini et al. 2012; Stahl
2017; Ghaemi 2019).
1254 G. Laux
Cross-References
References
Abber JC, Lue TF, Luo JA, Juenemann KP, Tanagho EA. Priapism induced by chlorpromazine and
trazodone: mechanism of action. J Urol. 1987;137:1039–42.
Beasley CM Jr, Dornseif BE, Pultz JA, Bosomworth JC, Sayler ME. Fluoxetine versus trazodone:
efficacy and activating-sedating effects. J Clin Psychiatry. 1991;52:294–9.
Brent DA. Antidepressants and suicidality. Psychiatr Clin North Am. 2016;39:503–12.
Brogden RN, Heel RC, Speight TM, Avery GS. Trazodone: a review of its pharmacological
properties and therapeutic use in depression and anxiety. Drugs. 1981;21:401–29. https://doi.
org/10.2165/00003495-198121060-00001.
Buckley NA, McManus PR. Fatal toxicity of serotonergic and other antidepressant drugs: analysis
of United Kingdom mortality data. Br Med J. 2002;325:1332–3.
Cipriani A, Furukawa TA, Salanti G, Chaimani A, et al. Comparative efficacy and acceptability of
21 antidepressant drugs for the acute treatment of adults with major depressive disorder: a
systematic review and network meta-analysis. Lancet. 2018;391:1357–66.
Clayton AH, Pradko JF, Croft HA, et al. Prevalence of sexual dysfunction among newer antide-
pressants. J Clin Psychiatry. 2002;63:357–66.
Cunningham LA, Borison RL, Carman JS, Chouinard G, et al. A comparison of venlafaxine,
trazodone, and placebo in major depression. J Clin Psychopharmacol. 1994;14:99–106.
Cusack B, Nelson A, Richelson E. Binding of antidepressants to human brain receptors: focus on
newer generation compounds. Psychopharmacology. 1994;114:559–65.
Davies J, Read J. A systematic review into the incidence, severity and duration of antidepressant
withdrawal effects: are guidelines evidence-based? Addict Behav. 2019;97:111–21.
Einarson A, Bonari L, Voyer-Lavigne S, et al. A multicentre prospective controlled study to
determine the safety of trazodone and nefazodone use during pregnancy. Can J Psychiatr.
2003;48:106–10.
Entsuah AR, Rudolph RI, Hackett D et al. Efficacy of venlafaxine and placebo during long-term
treatment of depression: a pooled analysis of relapse rates. Int Clin Psychopharmacol.
1996;11:137–145.
Fagiolini A, Comandini A, Catena Dell’Osso M, Kasper S. Rediscovering trazodone for the
treatment of major depressive disorder. CNS Drugs. 2012;26:1033–49.
Fagiolini A, Albert U, Ferrando L, Herman E, et al. A randomized, double-blind study comparing
the efficacy and safety of trazodone once-a-day and venlafaxine extended-release for the
treatment of patients with major depressive disorder. Int Clin Psychopharmacol. 2020;35:
137–46. https://doi.org/10.1097/YIC.0000000000000304.
Gabriel M, Sharma V. Antidepressant discontinuation syndrome. CMAJ. 2017;189:E747. https://
doi.org/10.1503/cmaj.160991.
Gerner RH. Geriatric depression and treatment with trazodone. Psychopathology. 1987;20:82–91.
Trazodone and Depression 1255
Gershon S, Mann J, Newton R, Gunther BJ. Evaluation of trazodone in the treatment of endogenous
depression: results of a multicenter double-blind study. J Clin Psychopharmacol. 1981;Suppl 6:
39–44.
Ghaemi NS. Clinical psychopharmacology. Oxford, 2019.
Goldberg HL, Finnerty RJ. Trazodone in the treatment of neurotic depression. J Clin Psychiatry.
1980;41:430–4.
Haria M, Fitton A, Trazodone MTD. A review of its pharmacology, therapeutic use in depression
and therapeutic potential in other disorders. Drugs Aging. 1994;4:331–55.
Hiemke C, Bergemann N, Clement HW, Conca A, et al. Consensus guidelines for therapeutic drug
monitoring in neuropsychopharmacology: update 2017. Pharmacopsychiatry. 2018;51:9–62.
Jaffer KY, Chang T, Vanle B, et al. Trazodone for insomnia: A systematic review. Innov Clin
Neurosci. 2017;14:24–34.
Jauch R, Kopitar Z, Prox A, Zimmer A. Pharmacokinetics and metabolism of trazodone in man
(author’s transl). Arzneimittel-Forsch. 1976;26:2084–9.
Kerr TA, McClelland HA, Stephens DA, Trazodone ASI. A comparative clinical and predictive study.
Acta Psychiatr Scand. 1984;70:573–7. https://doi.org/10.1111/j.1600-0447.1984.tb01251.x.
Khouzam HR. A review of trazodone use in psychiatric and medical conditions. Postgrad Med.
2017;129:140–8.
Mendelson WB. A review of the evidence for the efficacy and safety of trazodone in insomnia.
J Clin Psychiatry. 2005;66:469–76. https://doi.org/10.4088/jcp.v66n0409.
Moises HW, Kasper S, Beckmann H. Trazodone and amitriptyline in treatment of depressed
inpatients. A double-blind study. Pharmacopsychiatria. 1981;14:167–71. https://doi.org/10.
1055/s-2007-1019592.
Murasaki M, Kurihara M, Takahashi R, Mori A, Hasegawa K, et al. Clinical effects of trazodone
(KB-831) on depression – a double-blind comparative study using amitriptyline as a standard
drug. Rinsho Hiyoka (Jpn). 1990;18:279–313.
Nierenberg AA, Adler LA, Peselow E, Zornberg G, Rosenthal M. Trazodone for antidepressant-
associated insomnia. Amer J Psychiatry. 1994;151:1069–72.
Richelson E. The clinical relevance of antidepressant interaction with neurotransmitter transporters
and receptors. Psychopharmacol Bull. 2002;36:133–50.
Rickels K, Downing R, Schweizer E, Hassman H. Antidepressants for the treatment of generalized
anxiety disorder. A placebo-controlled comparison of imipramine, trazodone, and diazepam.
Arch Gen Psychiatry. 1993;50:884–95. https://doi.org/10.1001/archpsyc.1993.
01820230054005.
Stahl SM. Mechanism of action of trazodone: a multifunctional drug. CNS Spectr. 2009;14:536–46.
Stahl S. Essential psychopharmacology. The Prescriber’s guide. Antidepressants. 6th
ed. Cambridge: Cambridge University Press; 2017.
Tatsumi M, GroshanK BRD, Richelson E. Pharmacological profile of antidepressants and related
compounds at human monoamine transporters. Eur J Pharmacol. 1997;340:249–58.
Tsutsui S, Nakano K, Tsuboi K, Tahara K, Igarashi M, et al. Clinical evaluation of trazodone
hydrochloride, a new antidepressant, in the field of medicine – a double-blind controlled study
in comparison with maprotiline hydrochloride. Rinsho Iyaku (Jpn). 1990;6:1193–214.
van Moffaert M, de Wilde J, Vereecken A, Dierick M, et al. Mirtazapine is more effective than
trazodone: a double-blind controlled study in hospitalized patients with major depression. Int
Clin Psychopharmacol. 1995;10:3–9.
Weisler RH, Johnston JA, Lineberry CG, Samara B, et al. Comparison of bupropion and trazodone
for the treatment of major depression. J Clin Psychopharmacol. 1994;14:170–9.
White N, Litovitz T, Clancy C. Suicidal antidepressant overdoses: a comparative analysis by
antidepressant type. J Med Toxicol. 2008;4:238–50.
Yi XY, Ni SF, Ghadami MR, et al. Trazodone for the treatment of insomnia: a meta-analysis of
randomized placebo-controlled trials. Sleep Med. 2018;45:25–32.
Serotonin Reuptake Inhibitors: Citalopram,
Escitalopram, Fluoxetine, Fluvoxamine,
Paroxetine, and Sertraline
Gerd Laux
Contents
Chemistry, Developmental History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1258
Pharmacology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1259
Pharmacokinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1259
Mechanism of Action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1259
Indications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1260
Clinical Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1260
Side Effects, Adverse Reactions, and Toxicology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1265
Combination Therapy: Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1266
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1266
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1267
Abstract
Selective serotonin reuptake inhibitors (SSRIs) are the first-line antidepressants today
showing similar efficacy as nonselective monoamine reuptake inhibitors (NSMRIs,
“classical” tricyclic antidepressants) with lower side effects, however. So long, six
substances are approved and marketed – citalopram, escitalopram, fluoxetine,
fluvoxamine, paroxetine, and sertraline. Indications are major depressive disorder,
generalized anxiety disorder, panic disorder, social anxiety disorder, and obsessive-
compulsive disorder mainly. They differ regarding pharmacokinetic properties (half-
life, metabolism) mainly. Numerous randomized placebo-controlled clinical studies
(RCTs) and noninterventional studies have shown efficacy and effectiveness. Typical
side effects are nausea, troubled sleeping, and sexual dysfunction.
G. Laux (*)
Institute of Psychological Medicine (IPM), Soyen, Germany
MVZ Waldkraiburg, Center of Neuropsychiatry, Waldkraiburg, Germany
Department of Psychiatry and Psychotherapy, Ludwig-Maximilians-University (LMU), Munich,
Germany
e-mail: ipm@ipm-laux.de
Selective serotonin reuptake inhibitors (SSRIs) have been developed and introduced
as antidepressants between 1972 and 2003. Zimelidine (INN) developed by Arvid
Carlsson was the first to be marketed; rare case reports of immunotoxic effects
(Guillain–Barré syndrome) appeared to be caused by the drug, prompting its man-
ufacturer to withdraw it from the market in 1983. After its withdrawal, it was
succeeded by fluvoxamine and fluoxetine (derived from the antihistamine diphen-
hydramine). Fluvoxamine was introduced in Switzerland in 1983, in the USA in
1994, and in Japan in 1999. Fluoxetine was discovered by Eli Lilly in 1972; in 1975,
it was given the official chemical name fluoxetine and was marketed in Belgium in
1986; Lilly began marketing it in the USA in 1988 with the trade name Prozac, a
cultural icon and symbol of blockbuster and medicalization problem. It is on the
World Health Organization’s List of Essential Medicines. Paroxetine was approved
for medical use in the United States in 1992. Citalopram was first synthesized in
1972, first marketed in 1989 in Denmark, and approved for medical use in the United
States in 1998. Lundbeck now markets the (S)-(+) enantiomer escitalopram. FDA
issued the approval of escitalopram for major depression in 2002 and for generalized
anxiety disorder in 2003. Sertraline was approved by the FDA for medical use in the
United States in 1991 (Ciraulo and Shader 2011; Ghaemi 2019).
The chemical formulas of the SSRIs available are shown in Fig. 1.
Pharmacology
Pharmacokinetics
The six SSRIs differ mainly regarding elimination half-life and liver metabolism
via cytochrome P450 isoenzymes shown in Table 1. Great genetic variability in
the function of 2D6 enzyme among people should be mentioned. Range of
therapeutic plasma levels also vary. Paroxetine is the most potent 2D6 inhibitor
(DeVane et al. 2002; Hiemke and Härtter 2000; Kaye et al. 1989, Sanchez et al.
2014).
Mechanism of Action
SSRIs share common features, increase 5-HT, and enhance serotonergic neuro-
transmission through blockade of the serotonin transporter so blocking 5-HT
reuptake. 5-HT1A and 5-HT2 receptor density among patients with depression is
normalized (Cusack et al. 1994; Richelson and Nelson 1984; Tatsumi et al.
1997).
Fluvoxamine is a potent selective serotonin reuptake inhibitor with around
100-fold affinity for the serotonin transporter. Additionally, it behaves as a
potent σ1 receptor-agonist (Wilde et al. 1993). Fluoxetine delays the reuptake
of serotonin, resulting in serotonin persisting longer when it is released;
norfluoxetine is the primary active metabolite. Fluoxetine has been shown to
inhibit acid sphingomyelinase, a key regulator of ceramide levels which derives
ceramide from sphingomyelin (Benfield et al. 1986). Escitalopram has the
highest selectivity for the serotonin transporter (SERT) (Sanchez et al. 2014).
Paroxetine has mild anticholinergic effects and is also a relatively potent nor-
epinephrine reuptake inhibitor (Kaye et al. 1989). Sertraline also shows rela-
tively high activity as an inhibitor of the dopamine transporter (DAT) and
antagonist of the sigma σ1 receptor (DeVane et al. 2002).The unique effect of
sertraline on dopaminergic neurotransmission may be related to these effects on
cognition and vigilance.
Indications
SSRIs have a broad spectrum of indications. They are used and in many countries
released for the treatment of major depressive disorder, anxiety disorders, such as
panic disorder, social anxiety disorder, obsessive-compulsive disorder (OCD),
bulimia nervosa, and premenstrual dysphoric disorder (PMDD). Fluoxetine is also
approved for treatment of major depressive disorder in adolescents and children (Lee
et al. 2016; Stahl 2017; Ghaemi 2019).
For subtypes of depressive disorders such as poststroke depression and postpar-
tum/postnatal depressions, SSRIs are also first-line medications (Villa et al. 2018;
Deng et al. 2018).
Efficacy in the treatment of post-traumatic stress disorder (PTSD) is also evidence
based (Ipser and Stein 2012), e.g., sertraline is approved for the treatment of PTSD.
There is evidence that paroxetine may be effective in the treatment of compulsive
gambling, and fluoxetine in impulsive aggression (Kim et al. 2002; Felthous and
Stanford 2021). Escitalopram has shown positive effects regarding cognitive func-
tions including driving ability (Montgomery and Möller 2009). Anti-inflammatory
effects of fluvoxamine are being researched for potential use in Covid-19.
Clinical Studies
Multiple RCTs proved SSRIs superior to placebo required for approval; they are as
effective as tricyclic antidepressants, but better tolerated (Undurraga and
Baldessarini 2017; Ciraulo and Shader 2011; Ghaemi 2019). Table 2 gives a
selection of RCTs between SSRIs and TCAs and within SSRIs as well as some
major noninterventional/surveillance/observational studies (NIS). In conclusion,
efficacy and effectiveness of SSRIs is established.
A review concluded that second-generation antidepressants appear equally effec-
tive, although they may differ in efficacy (response rates) and acceptance (dropout
rates) (Cipriani et al. 2018). Treatment guidelines issued by the National Institute of
Health and Clinical Excellence, the American Psychiatric Association, and the
German Association of Psychiatry, Psychotherapy and Psychosomatics reflect the
viewpoint of equal efficacy.
Efficacy of fluoxetine for acute and maintenance treatment of major depressive
disorder in adults as well as children and adolescents (8–18 years) was established in
multiple clinical trials (Rossi et al. 2004; Magni et al. 2013). Efficacy for geriatric
depression was also demonstrated in placebo-controlled trials. It seems less effective
than sertraline, mirtazapine, and venlafaxine. The efficacy in the treatment of
obsessive-compulsive disorder (OCD) was demonstrated in two randomized multi-
center phase III clinical trials. The pooled results of these trials demonstrated that
47% of completers treated with the highest dose were “much improved” or “very
much improved” after 13 weeks of treatment, compared to 11% in the placebo arm of
the trial. The efficacy in panic disorder was demonstrated in two 12-week-random-
ized multicenter phase III clinical trials that enrolled patients diagnosed with panic
Table 2 Survey review SSRIs studies in MDD
Study arms, comparators, and Results Dropout
Study N patients placebo response-/remission-rate rates Relevant side effects
Citalopram (C) 316 pts C 103, 20–60 mg/d 62%
Stahl (2000) S 106, 50–150 mg/d 57%
RCT 24 weeks Pl 107 40%
Bougerol et al. 296 in-/out-pts C 147, 40 mg/d 69%
(1997) Flu 149 20 mg/d 73%
RCT
Escitalopram (E) 1282 pts E 12,6 mg/d 55,5%
Auquier et al. (2003) C 26,4 mg/d 50,8%
Meta-analysis
4 RCTs
Moore et al. (2005) 294 pts E versus C E 76% /56% C>E C 2,8%
RCT C 61%/44% E 0,7%
Holsboer-Trachsler 5.649 pts 52% Nausea 7%
et al. (2011) Depression
NIS and anxiety
Möller et al. (2007) 11.760 pts MADRS 6,7%
NIS initial 31,8 Nausea 1,7%,
week 8 12, and 4 anxiety 0,7%,
two suicides
Fluoxetine (FLU) 65 inpts Flu 40–80 mg/d versus maprotiline Flu ¼ M Four each Flu nausea
Serotonin Reuptake Inhibitors: Citalopram, Escitalopram, Fluoxetine,. . .
Table 2 (continued)
Study arms, comparators, and Results Dropout
Study N patients placebo response-/remission-rate rates Relevant side effects
Newhouse et al. 236 older pts Flu 20–40 mg/d Flu 71% S lower cognitive
(2000) S 50–100 mg/d S 73% symptoms
RCT
Magni et al. (2013) 171 studies Flu ¼ TCAs Flu better tolerated than
Review RCTs 24.868 pts Flu < sertraline, mirtazapine, TCAs
and venlafaxine
Fluvoxamine (F) 84 inpts F 100–300 mg/d F¼C F>C F nausea
Coleman and Block clomipramine (C) 100–300 mg/d C tremor, and
(1982) anticholinergic symptoms
RCT
Nemeroff et al. 95 pts F 124 mg/d F¼S F 31% nausea
(1995) S 137 mg/d S 35% insomnia
RCT
Paroxetine (P) 258 pts P 125, 20–30 mg/d P 60,0%
Bignamini and Ami 133, 75–150 mg/d Ami 65,4%
Rapisarda (1992)
RCT
Möller et al. (1993) 160 inpts P 84, 30 mg/d P 74,0%
RCT Ami 76, 150 mg/d Ami 87,0%
G. Laux
Ravindran et al. 953 pts P 479, 20–40 mg/d P 68,5%
(1997) 8–12 weeks Clo 474, 75–150 mg/d Clo 66,9%
RCT
Sertraline (S) 448 pts S 149, 50–200 mg/d S 61,1%
Reimherr et al. Ami 149, 50–150 mg/d Ami 70,5%
(1990) Pl 150 Pl 38,3%
RCT S ¼ Ami > Pl
Lydiard et al. (1997) 392 pts S 132, 50–200 mg/d S 55%
RCT Ami 131, 50–150 mg/d Ami 53%
Pl 129 Pl 37%
Sechter et al. (1999) 238 pts S 118, 50–150 mg/d S 74%
RCT 24 weeks Flu 120, 20–60 mg/d Flu 64%
Möller et al. (2000) 205 pts S 100, 50–100 mg/d S 66,3%
RCT Ami 105, 75–150 mg/d Ami 65,9%
Note > ¼ significantly more effective or better as, ¼ not significant different, MADRS¼ Montgomery Asberg Depression Rating Scale, Pl¼ Placebo, TCA¼
Tricyclic Antidepressants
Serotonin Reuptake Inhibitors: Citalopram, Escitalopram, Fluoxetine,. . .
1263
1264 G. Laux
SSRIs have much lower rates of adverse effects than TCAs (Undurraga et al. 2017;
Stahl 2017; Ghaemi 2019). Common side effects mainly include gastrointestinal
side effects such as nausea, loss of appetite, vomiting, diarrhea, and CNS effects
such as trouble sleeping, sexual problems, shakiness, feeling tired, excessive yawn-
ing, severe tinnitus, and sexual dysfunction. Rates vary depending if spontaneous
reporting or direct questioning is applied.
SSRIs are associated with a modest increase in the QTc interval although to a
lesser extent than TCAs.
Citalopram and escitalopram are associated with more dose-dependent QT inter-
val prolongation, so maximum daily doses are limited at 40 resp.20 mg for adults
and 20/10 mg for those older than 65 years or pts. with liver impairment (Beach et al.
2014).
Rare are hyponatremia and SIADH (syndrome of inappropriate antidiuretic
hormone secretion).
Serious side effects include serotonin syndrome and blood-thinning effects with
increased risk of bleeding especially when combined with anticoagulants or NSAIDs
(Andrade and Sharma 2016).
An increased risk of suicidal behavior in people under 25 years has been claimed.
Newer analyses conclude among adults a reduction in mean rating of suicidality in
patients receiving an SSRI; in young adults, SSRI treatment neither reduced nor
increased suicidality ratings relative to placebo (Brent 2016; Näslund et al. 2018).
Fluoxetine is considered the most stimulating of the SSRIs (that is, it is most
prone to causing insomnia and agitation). It also appears to be the most prone of the
SSRIs for producing dermatologic reactions (e.g., urticaria, rash, and itchiness).
Fluvoxamine is associated with nausea and sedation mainly. Paroxetine has a higher
incidence of anticholinergic effects (e.g., dry mouth, constipation, and blurred
vision) and sedation. Sertraline has higher incidence of diarrhea and can have
activating effects.
Regarding safety of overdose, SSRIs have low-hazard indices compared to
TCAs and MAOIs (White et al. 2008). Acute overdosage is often manifested by
emesis, lethargy, ataxia, disturbances in heart rhythm, tremor, amnesia, confusion,
coma, or convulsions.
Pregnancy. Data regarding teratogenic effects of SSRIs are controversial. Risk of
treatment (first trimester fetal development, third trimester newborn delivery) must
be weighed against risk of recurrence of depression (Fischer et al. 2019). Increased
risk of pulmonary hypertension has been associated with paroxetine mainly, so it is
not recommended for use in pregnancy (Berard et al. 2017; De Vries et al. 2021). In
sum, evidence suggests a generally small risk of congenital malformations and
argues against a substantial teratogenic effect of SSRIs (Gao et al. 2018).
Contraindications are MAOIs due to the potential for serotonin syndrome.
Use of citalopram, escitalopram, fluoxetine, or sertraline is permitted for pilots
with authorization from an aviation medical examiner by Federal Aviation
1266 G. Laux
Concomitant use of serotonergic drugs such as triptans, tramadol, and buspiron must
be monitored, and dose must be reduced by half when strong inhibitors of CYP2D6
are coadministered. Concomitant use of aspirin, nonsteroidal anti-inflammatory
drugs (NSAIDs), and anticoagulants can result in gastrointestinal and other bleeding.
SSRIs should not be taken with St John’s wort, tryptophan, or 5-HTP as the resulting
drug interaction could lead to serotonin syndrome.
Fluvoxamine has been observed to increase serum concentrations of mirtazapine,
which is mainly metabolized by CYP1A2, CYP2D6, and CYP3A4 (Wilde et al.
1993).
Paroxetine interacts with CYP2D6CYP2B6 and reaches a steady state in 7–14 days
(Kaye et al. 1989).
Taking citalopram with omeprazole may cause higher blood levels of citalopram
(Hiemke et al. 2018).
Cross-References
References
Andrade C, Sharma E. Serotonin reuptake inhibitors and risk of abnormal bleeding. Psychiatr Clin
North Am. 2016;39:413–26.
Auquier P, Robitail S, Llorca PM, Rive B. Comparison of escitalopram and citalopram efficacy: a
meta-analysis. Int J Psych Clin Pract. 2003;7:259–68.
Baldwin DS, Asakura S, Koyama T, Hayano T, Hagino A, Reines E, Larsen K. Efficacy of
escitalopram in the treatment of social anxiety disorder: a meta-analysis versus placebo. Eur
Neuropsychopharmacol. 2016;26:1062–9.
Beach SR, Kostis WJ, Celano CM, et al. Meta-analysis of selective serotonin reuptake inhibitor-
associated QTc prolongation. J Clin Psychiatry. 2014;75:441–9.
Benfield P, Heel RC, Lewis SP. Fluoxetine. A review of its pharmacodynamic and pharmacokinetic
properties, and therapeutic efficacy in depressive illness. Drugs. 1986;32:481–508.
Bérard A, Sheehy O, Zhao JP, Vinet É, Bernatsky S, Abrahamowicz M. SSRI and SNRI use during
pregnancy and the risk of persistent pulmonary hypertension of the newborn. Br J Clin
Pharmacol. 2017;83:1126–33.
Bignamini A, Rapisarda V. A double-blind multicenter study of paroxetine and amitriptyline in
depressed outpatients. Int Clin Psychopharmacol. 1992;6(Suppl 4):37–41.
Bougerol T, Scotto JC, Patris FM, et al. Citalopram and fluoxetine in major depression: comparison
of two clinical trials in a psychiatrist setting and in general practice. Clin Drug Invest. 1997;14:
77–89.
Brent DA. Antidepressants and suicidality. Psychiatr Clin North Am. 2016;39:503–12.
Cipriani A, La Ferla T, Furukawa TA, Signoretti A, et al. Sertraline versus other antidepressive
agents for depression. Cochrane Database Syst Rev. 2010;2:CD006117.
Cipriani A, Furukawa TA, Salanti G, Chaimani A, et al. Comparative efficacy and acceptability of
21 antidepressant drugs for the acute treatment of adults with major depressive disorder: a
systematic review and network meta-analysis. Lancet. 2018;391:1357–66.
Ciraulo DA, Shader RI, editors. Pharmacotherapy of depression. 2nd ed. New York: Humana Press/
Springer; 2011.
Coleman BS, Block BA. Fluvoxamine maleate, a serotonergic antidepressant: a comparison with
chlorimipramine. Prog Neuro-Psychopharmacol Biol Psychiatry. 1982;6:475–8.
Cusack B, Nelson A, Richelson E. Binding of antidepressants to human brain receptors: focus on
newer generation compounds. Psychopharmacology. 1994;114:559–65.
Davies J, Read J. A systematic review into the incidence, severity and duration of antidepressant
withdrawal effects: are guidelines evidence-based? Addict Behav. 2019;97:111–21.
De Vries C, Gadzhanova S, Sykes MJ, Ward M, Roughead E. A systematic review and meta-
analysis considering the risk for congenital heart defects of antidepressant classes and individual
antidepressants. Drug Saf. 2021;44:291–312.
DeJonghe F, Swinkels J, Tuyman-Qua H. Randomized double-blind study of fluvoxamine and
maprotiline in treatment of depression. Pharmacopsychiatry. 1991;24:21–7.
Deng L, Qiu S, Yang Y, et al. Efficacy and tolerability of pharmacotherapy for post-stroke
depression: a network meta-analysis. Oncotarget. 2018;9:23718–28.
DeVane CL, Liston HL, Markowitz JS. Clinical pharmacokinetics of sertraline. Clin
Pharmacokinet. 2002;41:1247–66.
Fava GA, Gatti A, Belaise C, Guidi J, Offidani E. Withdrawal symptoms after selective serotonin
reuptake inhibitor discontinuation: a systematic review. Psychother Psychosom. 2015;84:72–81.
Favré P. Clinical efficacy and achievement of a complete remission in depression: increasing
interest in treatment with escitalopram. L’Encephale. 2012;38:86–96.
Feighner J, Cohn JB. Double-blind comparative trials of fluoxetine and doxepine in geriatric
patients with major depressive disorder. J Clin Psychiatry. 1985;46:20–5.
Felthous A, Stanford M. The pharmacotherapy of impulsive aggression in psychopathic disorders.
In: Felthous A, Sass H, editors. The Wiley international handbook on psychopathic disorders
and the law. 2nd ed. Wiley; 2021. p. 810–3.
1268 G. Laux
Fischer Fumeaux CJ, Morisod Harari M, Weisskopf E, Eap CB, Epiney M, Vial Y, Csajka C, Bickle
Graz M, Panchaud A. Risk-benefit balance assessment of SSRI antidepressant use during
pregnancy and lactation based on best available evidence – an update. Expert Opin Drug Saf.
2019;18:949–63.
Gabriel M, Sharma V. Antidepressant discontinuation syndrome. Can Med Assoc J. 2017;189:747.
Gao SY, Wu QJ, Sun C, Zhang TN, Shen ZQ, et al. Selective serotonin reuptake inhibitor use during
early pregnancy and congenital malformations: a systematic review and meta-analysis of cohort
studies of more than 9 million births. BMC Med. 2018;16:205.
Ghaemi NS. Clinical psychopharmacology. Oxford: Oxford University Press; 2019.
Hiemke C, Härtter S. Pharmacokinetics of selective serotonin reuptake inhibitors. Pharmacol Ther.
2000;85:11–28.
Hiemke C, Bergemann N, Clement HW, Conca A, et al. Consensus guidelines for therapeutic drug
monitoring in neuropsychopharmacology: update 2017. Pharmacopsychiatry. 2018;51:9–62.
Hirschfeld RM. Sertraline in the treatment of anxiety disorders. Depress Anxiety. 2000;11:139–57.
Holsboer-Trachsler E, Baumann P, Höck P, et al. Escitalopram in clinical practice: results of an
observational study in patients with depression and anxiety. Psychopharmakotherapie. 2011;18:
59–65.
Ipser JC, Stein DJ. Evidence-based pharmacotherapy of post-traumatic stress disorder (PTSD). Int J
Neuropsychopharmacol. 2012;15:825–40.
Kaye CM, Haddock RE, Langley PF, Mellows G, et al. A review of the metabolism and pharma-
cokinetics of paroxetine in man. Acta Psychiatr Scand Suppl. 1989;350:60–75.
Keller MB. Citalopram therapy for depression: a review of 10 years of European experience and
data from U.S. clinical trials. J Clin Psychiatry. 2000;61:896–908.
Kennedy SH, Andersen HF, Lam RW. Efficacy of escitalopram in the treatment of major depressive
disorder compared with conventional selective serotonin reuptake inhibitors and venlafaxine
XR: a meta-analysis. J Psychiatry Neurosci. 2006;31:122–31.
Kim SW, Grant JE, Adson DE, Shin YC, Zaninelli R. A double-blind placebo-controlled study of
the efficacy and safety of paroxetine in the treatment of pathological gambling. J Clin Psychi-
atry. 2002;63:501–7.
Lee DJ, Schnitzlein CW, Wolf JP, Vythilingam M, et al. Psychotherapy versus pharmacotherapy for
posttraumatic stress disorder: systemic review and meta-analyses to determine first-line treat-
ments. Depress Anxiety. 2016;33:792–806.
Lydiard RB, Stahl SM, Hertzman M, Harrison WM. A double-blind, placebo-controlled study
comparing the effects of sertraline versus amitriptyline in the treatment of major depression. J
Clin Psychiatry. 1997;58:484–91.
Magni LR, Purgato M, Gastaldon C, Papola D, Furukawa TA, Cipriani A, Barbui C. Fluoxetine
versus other types of pharmacotherapy for depression. Cochrane Database Syst Rev. 2013;7:
CD004185.
Möller HJ, Berzewski H, Eckmann F, Gonzalves N, Kissling W, Knorr W, Ressler P, Rudolf GA,
Steinmeyer EM, Magyar I, et al. Double-blind multicenter study of paroxetine and amitriptyline
in depressed inpatients. Pharmacopsychiatry. 1993;26:75–8.
Möller HJ, Glaser K, Leverkus F, Göbel C. Double-blind, multicenter comparative study of
sertraline versus amitriptyline in outpatients with major depression. Pharmacopsychiatry.
2000;33:206–12.
Möller HJ, Friede M, Schmauß M. Treatment of depression with escitalopram: results of a large
post-marketing surveillance study. Psychopharmakotherapie. 2007;14:149–56.
Montgomery SA, Möller HJ. Is the significant superiority of escitalopram compared with other
antidepressants clinically relevant? Int Clin Psychopharmacol. 2009;24:111–8.
Moore N, Verdoux H, Fantino B. Prospective, multicenter, randomized, double-blind study of the
efficacy of escitalopram versus citalopram in outpatient treatment of major depressive disorder.
Int Clin Psychopharmacol. 2005;20:131–7.
Muijsers RB, Plosker GL, Noble S. Sertraline: a review of its use in the management of major
depressive disorder in elderly patients. Drugs Aging. 2002;19:377–92.
Serotonin Reuptake Inhibitors: Citalopram, Escitalopram, Fluoxetine,. . . 1269
Contents
Chemistry, Developmental History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1272
Pharmacology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1272
Pharmacokinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1272
Mechanism of Action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1272
Indications (of Marketed Products) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1272
Clinical Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1272
Side Effects/Adverse Reactions and Toxicology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1273
Combination Therapy: Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1278
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1279
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1279
Abstract
Bupropion is a norepinephrine and dopamine reuptake inhibitor. It is indi-
cated for the treatment of major depressive disorder, for the prevention of
major depressive episodes in patients with seasonal affective disorder, and for
the treatment of nicotine addiction (Food and Drug Administration, FDA).
Positive effects on fatigue and hypersomnia have been described. Compared
to other antidepressants, it has a positive side effect profile especially
concerning the risk of weight gain, sexual dysfunction, and QT prolongation.
At higher doses, there is an increased risk of seizures. Due to its unique
pharmacology, it might be used for pharmacological augmentation of seroto-
nergic antidepressants.
Pharmacology
Pharmacokinetics
Mechanism of Action
The mechanism of action is not fully understood. Bupropion blocks the norepineph-
rine and the dopamine reuptake pump. That is why it presumably increases norepi-
nephrine and dopaminergic neurotransmission (Stahl 2017).
The indications for bupropion are the treatment of major depressive disorder, the
prevention of major depressive episodes in patients with seasonal affective disorder,
and the treatment of nicotine addiction.
Positive effects on fatigue and hypersomnia are described. In clinical practice
bupropion is also used as augmentation of serotonergic antidepressants like selective
serotonergic reuptake inhibitors (SSRIs).
Clinical Studies
Most side effects seem to be linked to norepinephrine and dopamine. Most common
are headache, dry mouth, and transient insomnia. Increased restlessness, anxiety, and
agitation were repeatedly reported shortly after initiation of treatment. The rate of
insomnia for bupropion is higher than for placebo but similar to the rate associated
with SSRIs (Fava et al. 2005).
Table 1 “Surveys review”
1274
(continued)
1276
Table 1 (continued)
Study (RCT, meta-
analysis, NIS; Relevant and
name/acronym, N, patients Results (scores, response/remission Dropout severe side
authors, year Country(ies) Study arms, comparators, placebo rates) rates effects
double-blind, Bupropion 300 mg: 29.8%
placebo-controlled Placebo: 28.5%
study in Asian
patients, Koshino
et al. 2013
Systematic review 14,372 records Meta-analysis bupropion and placebo Hedge’s g < 0: Favoring bupropion
and meta-analysis identified by Hedge’s g > 0: Favoring placebo
Bupropion: a database Overall:
systematic review searching, five Hedge’s g = 2.02, df = 4,
and meta-analysis studies eligible for p < 0.001, CI -2.93 – 1.11
of effectiveness as inclusion in High heterogeneity of study findings
an Antidepressant meta-analysis ( p =0.001, I2 = 79.4%,
Patel et al. 2016 τ2 = 0.832), prediction interval
(PI 5.28 to 1.24).
Egger’s test p = 0.043 and Begg’s
test p = 0.027 for publication bias
D. Rujescu et al.
Comparative Review of 117 Bupropion 150–450 mg compared Efficacy (response rate) OR (95%
efficacy and randomized with new-generation antidepressants CI) 0,93 (0,77–1,11)
acceptability of 12 controlled trials Acceptability (dropout rate) OR
new-generation (25,928 (95% CI) 1,12 (0,92–1,36)
antidepressants: a participants)
multiple-
treatments meta-
analysis, Cipriani
et al. 2009
Bupropion and Depressions
Comparative Review of 522 Bupropion and placebo Efficacy (response rate) compared to
efficacy and trials comprising Bupropion and 20 antidepressants placebo (95% CrI) 1,58 (1,35–1,86)
acceptability of 21 116,477 Acceptability (dropout rate)
antidepressant participants, 33 compared to placebo (95% CrI) 0,96
drugs for the acute trials including (0,81–1,14)
treatment of adults bupropion
with major
depressive
disorder: a
systematic review
and network meta-
analysis, Cipriani
et al. 2018
HAM-D Hamilton rating scale for depression, IDS-IVR Inventory of depressive symptomatology self-reported, MADRS Montgomery-Asberg depression rating
scale
1277
1278 D. Rujescu et al.
Cross-References
References
Cipriani A, Furukawa TA, Salanti G, Geddes JR, Higgins JP, Churchill R, Watanabe N, Nakagawa
A, Omori IM, McGuire H, Tansella M, Barbui C. Comparative efficacy and acceptability of 12
new-generation antidepressants: a multiple-treatments meta-analysis. Lancet. 2009;373:746–58.
Cipriani A, Furukawa TA, Salanti G, Chaimani A, Atkinson LZ, Ogawa Y, Leucht S, Ruhe HG,
Turner EH, Higgins JPT, Egger M, Takeshima N, Hayasaka Y, Imai H, Shinohara K, Tajika A,
Ioannidis JPA, Geddes JR. Comparative efficacy and acceptability of 21 antidepressant drugs for
the acute treatment of adults with major depressive disorder: a systematic review and network
meta-analysis. Lancet. 2018;391:1357–66.
Clayton AH, Croft HA, Horrigan JP, Wightman DS, Krishen A, Richard NE, Modell JG. Bupropion
extended release compared with escitalopram: effects on sexual functioning and antidepressant
efficacy in 2 randomized, double-blind, placebo-controlled studies. J Clin Psychiatry.
2006;67:736–46.
Drugbank.ca. 2018. https://www.drugbank.ca/drugs/DB01156. Accessed 03 Dec 2018.
Embroytox.de. 2019. https://www.embryotox.de/arzneimittel/details/bupropion/. Accessed 11 Nov
2019.
Fava M, Rush AJ, Thase ME, Clayton A, Stahl SM, Pradko JF, Johnston JA. 15 years of clinical
experience with bupropion HCl: from bupropion to bupropion SR to bupropion XL. Prim Care
Companion J Clin Psychiatry. 2005;7:106–13.
Food and Drug Administration (FDA). 2018. https://www.accessdata.fda.gov/drugsatfda_docs/
label/2011/021515s026s027lbl.pdf. Accessed 03 Dec 2018.
Hasnain M, Vieweg WV. Weight considerations in psychotropic drug prescribing and switching.
Postgrad Med. 2013;125:117–29.
Hendrick V, Suri R, Gitlin MJ, Ortiz-Portillo E. Bupropion use during pregnancy: a systematic
review. Prim Care Companion CNS Disord. 2017;19(5) https://doi.org/10.4088/PCC.17r02160.
1280 D. Rujescu et al.
Hewett K, Chrzanowski W, Jokinen R, Felgentreff R, Shrivastava RK, Gee MD, Wightman DS,
O'Leary MC, Millen LS, Leon MC, Briggs MA, Krishen A, Modell JG. Double-blind, placebo-
controlled evaluation of extended-release bupropion in elderly patients with major depressive
disorder. J Psychopharmacol. 2010a;24:521–9.
Hewett K, Gee MD, Krishen A, Wunderlich HP, Le Clus A, Evoniuk G, Modell JG. Double-blind,
placebo-controlled comparison of the antidepressant efficacy and tolerability of bupropion XR
and venlafaxine XR. J Psychopharmacol. 2010b;24:1209–16.
Jefferson JW, Rush AJ, Nelson JC, VanMeter SA, Krishen A, Hampton KD, Wightman DS, Modell
JG. Extended-release bupropion for patients with major depressive disorder presenting with
symptoms of reduced energy, pleasure, and interest: findings from a randomized, double-blind,
placebo-controlled study. J Clin Psychiatry. 2006;67:865–73.
Kim S, Chen J, Cheng T, Gindulyte A, He J, He S, Li Q, Shoemaker BA, Thiessen PA, Yu B,
Zaslavsky L, Zhang J, Bolton EE. PubChem 2019 update: improved access to chemical data.
Nucleic Acids Res. 2019;47(D1):D1102–9. https://doi.org/10.1093/nar/gky1033. [PubMed
PMID: 30371825]
Koshino Y, Bahk W, Sakai H, Kobayashi T. The efficacy and safety of bupropion sustained-release
formulation for the treatment of major depressive disorder: a multi-center, randomized, double-
blind, placebo-controlled study in Asian patients. Neuropsychiat Dis Treat. 2013;9:1273–80.
Monden R, Roest AM, van Ravenzwaaij D, Wagenmakers EJ, Morey R, Wardenaar KJ, de Jonge P.
The comparative evidence basis for the efficacy of second-generation antidepressants in the
treatment of depression in the US: a Bayesian meta-analysis of food and drug administration
reviews. J Affect Disord. 2018;235:393–8.
Neuroscience based Nomenclature. 2018. http://www.nbn2.org/search#drug/15. Accessed 03 Dec
2018.
Patel K, Allen S, Hague MN, Angelescu I, Baumeister D, Tracy DK. Bupropion: a systematic
review and meta-analysis of effectiveness as an antidepressant. Ther Adv Psychopharmacol.
2016;6:99–144.
Post RM, Altshuler LL, Leverich GS, Frye MA, Nolen WA, Kupka RW, Suppes T, McElroy S,
Keck PE, Denicoff KD, Grunze H, Walden J, Kitchen CM, Mintz J. Mood switch in bipolar
depression: comparison of adjunctive venlafaxine, bupropion and sertraline. Br J Psychiatry.
2006;189:124–31.
Stahl SM. Prescriber’s guide. 6th ed. Cambridge, MA: Cambridge University Press; 2017.
Steinert T, Fröscher W. Epileptic seizures under antidepressive drug treatment: systematic review.
Pharmacopsychiatry. 2018;51:121–35.
Valenstein M, Kim HM, Ganoczy D, Eisenberg D, Pfeiffer PN, Downing K, Hoggatt K, Ilgen M,
Austin KL, Zivin K, Blow FC, McCarthy JF. Antidepressant agents and suicide death among US
Department of Veterans Affairs patients in depression treatment. J Clin Psychopharmacol.
2012;32:346–53.
White N, Litovitz T, Clancy C. Suicidal antidepressant overdoses: a comparative analysis by
antidepressant type. J Med Toxicol. 2008;4:238–50.
Mirtazapine and Depressions
Mellar P. Davis
Contents
Chemistry and Developmental History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1282
Pharmacology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1282
Pharmacokinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1282
Special Populations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1284
Elderly . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1284
Renal Failure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1284
Hepatic Failure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1284
Pharmacodynamics: Mechanisms of Action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1284
Indications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1286
Clinical Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1286
Mirtazapine Compared with Placebo and Amitriptyline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1286
Other Antidepressants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1290
Other Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1291
General Anxiety Disorders . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1294
Pain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1294
Side Effects and Adverse Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1294
Drug Interactions and Combinations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1297
Drug Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1297
Drug Combinations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1298
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1300
Abstract
Mirtazapine was introduced into the United States in 1996 as a norepinephrine
and specific serotonergic antidepressant which would have a broad spectrum of
activity and hopefully fewer side effects compared with selective serotonin
reuptake inhibitors. Mirtazapine consists of S- and R-enantiomers which are
metabolized by cytochromes CYP1A2, CYP2D6, and CYP3A4 which are then
M. P. Davis (*)
Department of Palliative Care, Geisinger Medical Center, Danville, PA, USA
e-mail: mdavis2@geisinger.edu
Fig. 1 Mirtazapine
Pharmacology
Pharmacokinetics
Oral by availability is 49.7% 9.8% which does not change with chronic dosing
and is not significantly altered by food (Fawcett and Barkin 1998b). The mean
time to maximum concentration is 1.8 0.7 h for single dose and a bit shorter
Mirtazapine and Depressions 1283
(1.5 h) with multiple doses (Timmer et al. 2000). Peak plasma levels for a single
15 mg dose are 31.6 12.8 ng/mL and for multiple doses 41.8 7.7 ng/mL.
Dose proportional increases in plasma levels occur with doses between 5 and
75 mg. Serum half-life is 16.4 4.6 h which remains the same at steady state
(Timmer et al. 1996, 1997, Timmer et al. 2000). Mirtazapine is 85% protein
bound, a central volume of distribution of 107 42 L and at steady state a total
volume of distribution of 339 125 L or about 4.5 l/kg (Fawcett and Barkin
1998b).
Mirtazapine is metabolized in the gut wall and then hepatic first-pass clearance
through multiple mixed function oxidases (cytochromes). Cytochromes
CYP1A2, CYP2D6, and CYP3A4 are largely responsible for metabolism with
CYP2D6 and CYP1A2 involved in the formation of the 8-hydroxy metabolite
and CYP3A4 responsible for the formation of the N-desmethyl and N-oxide
metabolite (Dodd et al. 2001). Mirtazapine is also glucuronidated. Metabolites
are active but contribute only 3–10% to antidepressant efficacy (Timmer et al.
2000; Sandker et al. 1994; Delbressine et al. 1998; Kasper 1995, 1997). The
R-enantiomer has higher plasma levels than the S-enantiomer and a longer half-
life. However, the S-enantiomer serum levels are increased in CYP2D6 poor
metabolizers, while the R-enantiomer levels are unchanged (Delbressine et al.
1998; Hayashi et al. 2015). A multicenter study compared clinical effectiveness
of mirtazapine with pharmacokinetics and pharmacokinetics (Jaquenoud Sirot et
al. 2012). Smokers had lower mirtazapine and metabolites serum levels, while
women had higher levels. And although CYP1A2 and CYP2D6 genotype influ-
ences the parent drug and metabolite level, neither cytochrome influences effi-
cacy (Okubo et al. 2015; Kirchheiner et al. 2004). However, CYP2B6 genotype
influences R-hydroxy metabolites which have some correlation with depression
responses. The authors found that a S-mirtazapine levels greater than 5 ng/mL
had a 77% probability of response with a specificity of 50% and a sensitivity of
91%. On the other hand, there is no relationship between pharmacokinetic
parameters and adverse effects. At low drug concentrations, the three major
cytochromes each contribute approximately 25–45% to mirtazapine metabolism,
whereas at high drug concentrations, CYP3A4 contributes 70% to metabolism.
However, complete inhibition of one cytochrome enzyme is unlikely to influence
mirtazapine clearance to significant degree (Stormer et al. 2000). Unlike tricyclic
antidepressants, CYP2D6 genotype does not influence drug levels (Lind et al.
2009). In general, despite the abovementioned study, serum levels do not corre-
late with antidepressant efficacy (Timmer et al. 2000). Therapeutic drug moni-
toring is not recommended as it is with amitriptyline. Mirtazapine does not inhibit
cytochromes to any significant degree and is less subject to drug-drug interac-
tions than SSRIs and tricyclic antidepressants (Barkin et al. 1999; Stormer et al.
2000).
Mirtazapine and metabolites are mainly eliminated in the urine (75%) with 15%
eliminated in stool. Mirtazapine is nearly completely eliminated within 3–4 days.
The eliminations half-life is 20–40 h. It is longer in women (37 h) than men (26 h)
(Timmer et al. 2000).
1284 M. P. Davis
Special Populations
Elderly
Renal Failure
Hepatic Failure
autoreceptors is 30-fold greater than for central and peripheral alpha-1 adrenergic
receptors (de Boer 1995). Norepinephrine inhibits hippocampal serotonergic neuron
release of serotonin by binding to presynaptic alpha-2 heteroreceptors which is
prevented by mirtazapine. At the same time, norepinephrine release is amplified
by mirtazapine binding to alpha-2 autoreceptors. Subsequently norepinephrine binds
to alpha-1 adrenergic receptors on brainstem raphe neurons which facilitate seroto-
nin release (Fawcett and Barkin 1998b). By blocking postsynaptic 5HT-2 and 5HT-3
receptors, mirtazapine reduces nausea and anorexia commonly seen with SSRIs and
reduces the risk of a serotonin syndrome (Fawcett and Barkin 1998b). Serotonin
released by raphe neurons is directed to 5HT-1 receptors and away from 5HT-2
receptors which contribute to mirtazapine’s anti-anxiolytic activity (Fawcett and
Barkin 1998b). Blockade of 5 HT-3 receptors in the limbic system may be important
in treating clinical states associated with increased dopaminergic tone such as
emesis, schizophrenia, manic, and substance abuse (Dubovsky 1994). This may be
the reason for reduced extrapyramidal effects when mirtazapine is combined with an
antipsychotic.
Mirtazapine has low affinity for muscarinic and dopaminergic receptors such that
anticholinergic side effects are significantly less than that seen with tricyclic antide-
pressants (Nutt 1997; de Boer et al. 1988b). The high affinity for histamine (H-1)
receptors is responsible for sedation and improves sleep and weight gain commonly
seen with mirtazapine (Salvi et al. 2016; Holm and Markham 1999).
Mirtazapine has very weak and clinically insignificant norepinephrine reuptake
inhibition. It is a 1000-fold less avid for transporters compared with desipramine and
100-fold less than imipramine (de Boer et al. 1988a).
Mirtazapine responses are associated with plasma brain-derived neurotrophic
factor (BDNF) and reduced plasma tumor necrosis factor-alpha (TNF-alpha)
suggesting that mirtazapine has some anti-neuroinflammatory activity which con-
tributes to its antidepressant efficacy (Gupta et al. 2016; Deuschle et al. 2013).
Weight gain is common in the first 4 weeks of mirtazapine therapy. Weight gain is
associated with an increase leptin serum level. Tricyclic antidepressants are also
associated with weight gain, but this is not related to serum leptin levels unlike
mirtazapine (Kraus et al. 2002; Goodnick et al. 1999; Fava 2000).
Mirtazapine has less sexual dysfunction associated with it compared with SSRIs
and, in fact, reverses SSRI sexual dysfunction. This may be related to reduced effects
on the hypothalamic pituitary axis. Mirtazapine does not increase prolactin levels or
reduced growth hormone levels; although mirtazapine transiently reduces cortisol
levels, the clinical significance of this is unknown (Farah 1998, 1999; Laakmann
et al. 1999, 2000).
Several studies have attempted to use pharmacogenetics to predict mirtazapine
responders. The short form of the monoamine oxidase A promoter region has been
reported to predict mirtazapine responses (Tzeng et al. 2009). Monoamine oxidase A
polymorphism T941G is reported to predict antidepressant responses only in females
(Tadic et al. 2007). The catechol-o-methyl transferase gene Val/Val and Val/Met
predicts responses with mirtazapine but not paroxetine (Szegedi et al. 2005). Certain
polymorphisms of the 5HT-2A receptor predict improved insomnia but not
1286 M. P. Davis
Indications
Clinical Studies
(continued)
Table 1 (continued)
1288
Placebo 10
MIR response 51%
Kasper N = 732 MIR HAMD change 6 weeks
1995 RCT = 5 Amitriptyline MIR 15.7
Amitriptyline 16.1
Responses MIR 72%
Amitriptyline 71%
Kasper N = 265 MIR HAMD change 6 weeks
RCT = 3 Doxepin MIR 18.1
Trazodone Clomipramine 18.1
Clomipramine MIR 13.6
Doxepin 14.7
MIR 18.7
Trazodone 15.7
MIR = clomipramine,
doxepin > trazodone
HAMD Hamilton Depression Rating Scale, MIR mirtazapine, n number, NNT number needed to treat, OR odds ratio, RR relative risks, RCT randomized
controlled trial, SMD standard mean difference
1289
1290 M. P. Davis
which differed significantly from placebo and SSRIs but not amitriptyline. At
6 weeks the magnitude of change was 14 points for mirtazapine and 10 points for
placebo (Kasper 1995). Responders overall in this review were 51% for mirtazapine
and 33% for placebo. Overall dropouts were 13.7% for mirtazapine and 21.5% for
placebo. Mirtazapine maintenance therapy after remission-maintained remissions
was better than placebo at 2 years (77% versus 43.9%, p < 0.001) (Montgomery
1995). One of the original studies found that difference in the HAMD between
mirtazapine and placebo to be 3.5 at 6 weeks favoring mirtazapine. Responses
were 63% for mirtazapine and 30% for placebo (Smith et al. 1990). A fourth review
found that mirtazapine decreased the HAMD by 11.1 points versus 6.9 points for
placebo at 6 weeks (Stahl et al. 1997). A fifth review found that the SMD in the
HAMD between mirtazapine and placebo was 0.49 (95% confidence interval (CI),
0.33–0.64) and reduction in the HAMD was 11.08 for mirtazapine versus 7.21
for placebo (Bech 2001). If only the depression subscale was compared, the differ-
ence was 1.78 in favor of mirtazapine with a SMD of 0.42 (95% CI, 0.28–0.57).
Dropouts were not different between placebo and mirtazapine. Reasons for dropouts
were adverse effects for mirtazapine and lack of efficacy for placebo (Cipriani et al.
2018a). In a network meta-analysis, mirtazapine was superior to placebo with an
odds ratio (OR) of 1.89 (95% CI, 1.64–2.2) (Cipriani et al. 2018a). The overall
number needed to treat to benefit a single patient (NNT) with mirtazapine is 4.
In all the comparisons of mirtazapine with amitriptyline in particular and with
other classes of antidepressants in general, mirtazapine was not inferior in the
primary endpoints which included improvement, responses, and remissions
(Cipriani et al. 2018a; Montgomery et al. 1998; Stahl et al. 1997; Smith et al.
1990; Fawcett and Barkin 1998a; Bech 2001; Kasper 1995; Watanabe et al. 2008,
2011; Papakostas et al. 2008; Thase et al. 2010; Khoo et al. 2015). In fact,
mirtazapine was better tolerated than amitriptyline with fewer side effects and was
found to be the most cost-effective antidepressant of any drug class. In the original
trials, there tended to be less dropout with mirtazapine compared with amitriptyline
(Montgomery et al. 1998; Montgomery 1995).
Other Antidepressants
There have been several systematic reviews and meta-analysis which have compared
mirtazapine to other antidepressants. Cipriani and colleagues in their network meta-
analysis found that mirtazapine responses ranged between 57.4% and 59.4% and
remissions ranged between 41.8% and 46.6% at 7 months (Cipriani et al. 2018a).
Mirtazapine produced similar responses as SSRIs but was superior to reboxetine and
trazodone. Clomipramine was less well tolerated, and escitalopram was better
tolerated than mirtazapine. In the two meta-analyses published by Watanabe and
colleagues, mirtazapine produced equivalent responses and remissions as tricyclic
antidepressants within the first 2 weeks of therapy, whereas SSRI responses did not
occur until later and matched mirtazapine efficacy at only at 6 weeks (Watanabe et al.
2008, 2011). In addition, mirtazapine responses were superior to venlafaxine and
Mirtazapine and Depressions 1291
Other Studies
Mirtazapine has been used to treat refractory depression defined as failure to respond
to two consecutive medical treatments. Remissions by the HAMD were 12% (Fava
et al. 2006). However, patients who have not responded to a SSRI alone will respond
to mirtazapine nearly 50% of the time (48%) (Fava et al. 2001). Patients who have
impaired sexual function on an SSRI will have a return of sexual function when
switched to mirtazapine (Gelenberg et al. 2000; Koutouvidis et al. 1999).
Mirtazapine is particularly effective in patients with depression and anxiety or
those with a general anxiety disorder (Falkai 1999; Nutt 1999; Croom et al. 2009).
Mirtazapine is equally effective as amitriptyline and clomipramine when treating
melancholic depression (Kasper 1997). In a randomized trial, mirtazapine produced
a higher remission rate than venlafaxine when treating melancholic depression and a
higher completion percentage (76.9% vs. 64.6%) with fewer dropouts (5.1% vs.
15.3%). Sleep disturbances were also improved with mirtazapine in this group of
patients (Guelfi et al. 2001).
In a prospective study which compared doses of sertraline and mirtazapine as an
“add-on” or “switch” for patients not responding to sertraline, there was no statis-
tically significant difference in the adjusted PHQ-9 score at week 9 between the
1292 M. P. Davis
50 mg/day arm and the 100 mg/day arm (0.25 point, 95% confidence interval (CI), –
0.58 to 1.07, P = 0.55). In step 2, participants not remitted by week 3 were
randomized to continue sertraline and add mirtazapine) or to switch to mirtazapine.
At week 9, adding mirtazapine reduced the PHQ-9 scores by 0.99 point (0.43–1.55,
P = 0.0012); switching reduced it by 1.01 points (0.46–1.56, P = 0.0012), relative to
continuing sertraline (Kato et al. 2018). It appears that if patients have no response
to sertraline using 50 mg a day for 3 weeks, there may be a modest advantage to
switching to mirtazapine.
Mirtazapine prevents poststroke depression. In a randomized trial, 2 of 35
patients developed depression on mirtazapine, whereas 14 of 35 placebo-treated
patients developed depression. Mirtazapine was effective treating depression in
those who developed depression on placebo in the crossover (Niedermaier et al.
2004). Mirtazapine has been used to treat post-traumatic stress disorders (PTSD).
However, mirtazapine is less effective than SSRIs as a drug class and venlafaxine for
PTSD (Cipriani et al. 2018b).
Mirtazapine has been used to treat patients with substance use and depression.
Patients who use cocaine have improved depression but not reduced cocaine use
(Afshar et al. 2012). The benefits of mirtazapine in reducing depression in patients
with alcohol abuse have been confirmed in two studies. However the evidence for
reducing alcohol use in this population is mixed (Cornelius et al. 2016; Altintoprak
et al. 2008).
Chronic pain and depression co-occur with a high degree of frequency. In an
observational study of 594 patients, mirtazapine improved depression and pain
severity independent of age and type of pain syndrome (Freynhagen et al. 2006).
Sleep disturbances due to pain were also improved. Dropouts (18%) were largely
due to lack of efficacy. Eighty percent elected to continue mirtazapine beyond the 6-
week trial period. Adverse effects were fatigue and weight gain.
Responses in post-traumatic stress syndrome (PTSD) to antidepressants are
mixed (Mcgrane and Shuman 2018). While placebo-controlled trials are few, avail-
able data suggest that mirtazapine produces only a small benefit in civilians with
PTSD as either monotherapy or as adjunct to sertraline (Davidson et al. 2003). As
one would anticipate, open-label studies demonstrate improved outcomes in combat-
related PTSD populations relative to placebo-controlled studies (Chung et al. 2004).
Open-label trials bias responses toward favorable outcomes which may not be
validated or confirmed by double-blind randomized trials. Though uncommon,
worsening anxiety has been experienced secondary to mirtazapine; thus, mirtazapine
should be used with caution in patients with PTSD (Davidson et al. 2003; Connor et
al. 1999; Alderman et al. 2009).
Depression in Alzheimer’s disease is much less responsive to antidepressants
than major depression disorder (Orgeta et al. 2017). Mirtazapine, imipramine clo-
mipramine, and sertraline responses are the same as placebo (OR 1.08, 95% CI
0.69–1.69). Mirtazapine improvement in depression by the Cornell Scale for Depres-
sion in Dementia was discouraging (SMD of 0.01, 95% CI 1.37 to +1.38)
(Banerjee et al. 2013). Mirtazapine worsened daytime sleeping patterns without
improving night time sleep (Scoralick et al. 2017). In general, classical
Mirtazapine and Depressions 1293
recent survey, mirtazapine was the most common in a depressant prescribed for
depression in cancer patients because of its multiple benefits (appetite, weight gain,
sleep, pain, and depression) (Sanjida et al. 2018).
Pain
Mirtazapine has been used in patients with pain, depressed mood, and/or insomnia.
The beneficial effects are present, but the mechanism for analgesia is unknown. It is
unclear whether the reduction of pain, the enhancement of the depressed mood, or
the combination of both effects leads to analgesia (Freynhagen et al. 2006). In animal
models there is evidence that mirtazapine has direct analgesic benefits (Bomholt et
al. 2005). Analgesia appears to be at the supraspinal level and not spinal (Kilic et al.
2011). In an open-label crossover study, mirtazapine analgesia was assessed by the
Memorial Pain Assessment Scale items for pain, pain relief (Theobald et al. 2002).
Low-dose pregabalin and mirtazapine improved pain from bone metastases better
than pregabalin alone in a randomized trial. The rationale behind the combination is
that a subset with bone metastases suffer from neuropathic pain from damage to
trabecular nerve-free endings (Nishihara et al. 2013). A meta-analysis of fibromyal-
gia pain found that mirtazapine potential benefits were outweighed by its potential
harms (Welsch et al. 2018). A systematic review published the same year as the
meta-analyses found that patients with fibromyalgia, treated with mirtazapine, had
improved pain, sleep, and quality of life. Study durations ranged from 6 to 13 weeks,
and studies used varying dosing strategies for mirtazapine (Ottman et al. 2018).
A summary of side effects relative to placebo are listed on Table 2, and the side
effects relative to SSRIs are summarized on Table 3.
Mirtazapine and Depressions 1295
significant impairment, and individuals with marginal driving ability may become
dangerously impaired on mirtazapine. Trazodone which is also given at night for
sleep does not impair driving skills (Sasada et al. 2013).
Antidepressants as a drug class can lower seizure threshold. However, seizures
are very rare with mirtazapine (less than 0.1%). In one study incidence of seizures
with mirtazapine was 0.04% compared with tricyclic antidepressants which was as
high as 4% (Steinert and Froscher 2018; Kasper 1997). Mirtazapine occasionally
produces nightmares and has been associated with the restless leg syndrome and
periodic limb movements (Menon and Madhavapuri 2017; Kolla et al. 2018).
Mirtazapine does not adversely affect blood pressure or heart rate (Montgomery
1995). However SSRIs are significantly associated with lower all-cause mortality in
patients with cardiovascular disease (OR 0.37, 95% CI 0.19–0.71, p = 0.0028) in a
group of veterans; mirtazapine was significantly associated with higher prevalence
of heart failure (OR 3.26, 95% CI 1.029–10.38, p = 0.0445) (Acharya et al. 2013).
This may be related to weight gain and increase fat on mirtazapine (Lee et al. 2016;
Laimer et al. 2006; Kraus et al. 2002). Depression has been associated with increased
heart rate and reduced respiratory sinus arrhythmia (a sign of reduced vagal tone)
(Hage et al. 2017). Mirtazapine reduces heart rate variability further. The
sympathovagal imbalance of depression is not reversed by mirtazapine (Terhardt
et al. 2013).
Mirtazapine reduces nausea, retching, and reduced appetite and can increase
weight in patients with functional dyspepsia and gastroparesis (Tack and Carbone
2017; Malamood et al. 2017; Yin et al. 2014; Kim et al. 2006). Topical mirtazapine is
well absorbed and has been associated with weight gain in cats with unintended
weight loss (Poole et al. 2019). This may be an attractive option in those suffering
from cancer-related anorexia, thus seen more as a benefit then a side effect. The
benefits may be related to neuroendocrine changes. Serum neuropeptide Y, ghrelin,
and motilin are increased, while serotonin and cholecystokinin are reduced (Jiang et
al. 2016). The downside into mirtazapine like other antidepressants is the increased
risk of gastrointestinal bleeding (OR 1.17, 95% CI 1.01–1.38) (Na et al. 2018).
Mirtazapine has been associated with the serotonin syndrome. In one case report,
the combination of metoclopramide and mirtazapine leads to serotonin-related
toxicity (Harada et al. 2017). This has also been reported with olanzapine plus
mirtazapine and escitalopram plus mirtazapine and mirtazapine plus venlafaxine
(Wu et al. 2015; Ansermot et al. 2014; Decoutere et al. 2012). On the other hand,
mirtazapine has been used to treat the serotonin syndrome since it blocks postsyn-
aptic serotonin receptors (Hoes and Zeijpveld 1996).
Antidepressants have been associated with sleep bruxism. Duloxetine (odds ratio
[OR] = 2.16; 95% confidence interval [95% CI] = 1.12–4.17), paroxetine
(OR = 3.63; 95% CI = 2.15–6.13), and venlafaxine (OR = 2.28; 95%
CI = 1.34–3.86) are associated with sleep bruxism, whereas mirtazapine is not
(Melo et al. 2018).
Mirtazapine treats pruritus of various causes, those resulting from spinal mor-
phine, related to cancer and dermatological diseases (Kouwenhoven et al. 2017;
Yosipovitch and Bernhard 2013; Hundley and Yosipovitch 2004; Davis et al. 2003).
Mirtazapine and Depressions 1297
Drug Interactions
Drug Combinations
the imipramine subgroup (Navarro et al. 2019). A second study did not find
evidence of important benefits to adding mirtazapine to an SSRI or SNRI over
placebo in a treatment-resistant group of primary care patients with depression
(Kessler et al. 2018).
Mirtazapine has been combined with methylphenidate to treat depression in
cancer patients. Depression was significantly better by day 3 as measured by the
MADRS ( 5.29 versus 1.02), and responses to the combination were greater at
day 9 (20.5% versus 6.8%) and day 14 (29.5% versus 6.8%) (Zaini et al. 2018).
Adverse effects were the same as mirtazapine alone. In contrast, aripiprazole aug-
mentation therapy appears to be effective and safe in patients who do not respond to
SSRIs or SNRIs (Kamijima et al. 2018).
Early studies reported that mirtazapine did augment monotherapy activity in
those with resistant depression. Mirtazapine has been used to augment antidepres-
sant responses in patients not responding to monotherapy. Patients not responding at
4 weeks to a standard antidepressant were randomized to continued monotherapy or
mirtazapine augmentation using 15–30 mg/d. Responses at 4 weeks were 64% for
the combination versus 20% for monotherapy and remissions 45.4% versus 13.3%
(Carpenter et al. 1999, 2002). A study which compared mirtazapine augmentation of
SSRIs (paroxetine, sertraline) versus SSRI augmentation of mirtazapine in patients
failing to respond to 4 weeks of monotherapy found that mirtazapine improved
responses but adding SSRIs to mirtazapine failed to improve responses. When
mirtazapine was added to SSRIs non-responders after 4 weeks, the remission rate
increased by 5% and HAMD score improved by 4 points. While for mirtazapine
non-responders, the addition of a SSRIs was not effective (Kato et al. 2017). This
begs the question of whether mirtazapine should be used first line or second line or as
augmentation to SSRI therapy.
Two systematic reviews shed some light on the question. Upfront combinations
of mirtazapine plus a SSRI are superior to SSRI monotherapy for remissions (RR
1.88, 95% CI 1.06–3.33) but not responses. Weight gain was the major side effect
with the combination (Rocha et al. 2012). The second systematic review had safety
as the primary outcome. The combination of mirtazapine plus an SSRI was associ-
ated with greater overall side effects than SSRI monotherapy (RR 1.65, 95% CI
1.19–2.33). Sedation (RR 3.22, 95% CI 2.16–4.08) and weight gain (RR 3.81, 95%
CI 1.37–10.55) were the most prominent additional side effects with the combination
(Galling et al. 2015).
Citalopram has been used to treat obsessive-compulsive disorders. In a ran-
domized trial, the combination of mirtazapine plus citalopram produced
responses at 4 weeks which took 8 weeks for citalopram (Pallanti et al. 2004).
The combination of mirtazapine plus sertraline produced greater PTSD remis-
sions than sertraline alone (OR 4.7, 95% CI 1.1–19.9, NNT 3.5). Depression was
significantly better with the combination at 24 weeks ( p = 0.023) (Schneier et al.
2015).
Prolonged release melatonin (2 mg) has been added to mirtazapine as it was being
withdrawn for a group of premenopausal women treated for insomnia but suffered
from weight gain. Mirtazapine was withdrawn over 1–3 months, and so there was a
1300 M. P. Davis
1–3-month overlap. Sleep quality improved 103% with the combination and further
improved 180% with the withdrawal of mirtazapine. Weight decreased with the
withdrawal of mirtazapine (Dolev 2011).
References
Abo-Zena RA, Bobek MB, Dweik RA. Hypertensive urgency induced by an interaction of
mirtazapine and clonidine. Pharmacotherapy. 2000;20:476–8.
Acharya T, Acharya S, Tringali S, Huang J. Association of antidepressant and atypical antipsychotic
use with cardiovascular events and mortality in a veteran population. Pharmacotherapy.
2013;33:1053–61.
Afshar M, Knapp CM, Sarid-Segal O, Devine E, Colaneri LS, Tozier L, Waters ME, Putnam MA,
Ciraulo DA. The efficacy of mirtazapine in the treatment of cocaine dependence with comorbid
depression. Am J Drug Alcohol Abuse. 2012;38:181–6.
Alderman CP, Condon JT, Gilbert AL. An open-label study of mirtazapine as treatment for combat-
related PTSD. Ann Pharmacother. 2009;43:1220–6.
Altintoprak AE, Zorlu N, Coskunol H, Akdeniz F, Kitapcioglu G. Effectiveness and tolerability of
mirtazapine and amitriptyline in alcoholic patients with co-morbid depressive disorder: a
randomized, double-blind study. Hum Psychopharmacol. 2008;23:313–9.
Ansermot N, Hodel PF, Eap CB. Serotonin toxicity after addition of mirtazapine to escitalopram. J
Clin Psychopharmacol. 2014;34:540–1.
Anttila SA, Leinonen EV. A review of the pharmacological and clinical profile of mirtazapine. CNS
Drug Rev. 2001;7:249–64.
Anttila AK, Rasanen L, Leinonen EV. Fluvoxamine augmentation increases serum mirtazapine
concentrations three- to fourfold. Ann Pharmacother. 2001;35:1221–3.
Arora S, Vohora D. Comparative evaluation of partial alpha2 -adrenoceptor agonist and pure alpha2
-adrenoceptor antagonist on the behavioural symptoms of withdrawal after chronic alcohol
administration in mice. Basic Clin Pharmacol Toxicol. 2016;119:202–9.
Banerjee S, Hellier J, Romeo R, Dewey M, Knapp M, Ballard C, Baldwin R, Bentham P, Fox C,
Holmes C, Katona C, Lawton C, Lindesay J, Livingston G, Mccrae N, Moniz-Cook E, Murray J,
Nurock S, Orrell M, O’brien J, Poppe M, Thomas A, Walwyn R, Wilson K, Burns A. Study of
the use of antidepressants for depression in dementia: the HTA-SADD trial – a multicentre,
randomised, double-blind, placebo-controlled trial of the clinical effectiveness and cost-effec-
tiveness of sertraline and mirtazapine. Health Technol Assess. 2013;17:1–166.
Barkin RL, Chor PN, Braun BG, Schwer WA. A trilogy case review highlighting the clinical and
pharmacologic applications of mirtazapine in reducing polypharmacy for anxiety, agitation,
insomnia, depression, and sexual dysfunction. Prim Care Companion J Clin Psychiatry.
1999;1:142–5.
Bech P. Meta-analysis of placebo-controlled trials with mirtazapine using the core items of the
Hamilton Depression Scale as evidence of a pure antidepressive effect in the short-term
treatment of major depression. Int J Neuropsychopharmacol. 2001;4:337–45.
Benazzi F. Mirtazapine withdrawal symptoms. Can J Psychiatr. 1998a;43:525.
Benazzi F. Serotonin syndrome with mirtazapine-fluoxetine combination. Int J Geriatr Psychiatry.
1998b;13:495–6.
Bomholt SF, Mikkelsen JD, Blackburn-Munro G. Antinociceptive effects of the antidepressants
amitriptyline, duloxetine, mirtazapine and citalopram in animal models of acute, persistent and
neuropathic pain. Neuropharmacology. 2005;48:252–63.
Botts S, Diaz FJ, Santoro V, Spina E, Muscatello MR, Cogollo M, Castro FE, De Leon J. Estimating
the effects of co-medications on plasma olanzapine concentrations by using a mixed model.
Prog Neuro-Psychopharmacol Biol Psychiatry. 2008;32:1453–8.
Mirtazapine and Depressions 1301
Bremner JD, Wingard P, Walshe TA. Safety of mirtazapine in overdose. J Clin Psychiatry.
1998;59:233–5.
Cankurtaran ES, Ozalp E, Soygur H, Akbiyik DI, Turhan L, Alkis N. Mirtazapine improves sleep
and lowers anxiety and depression in cancer patients: superiority over imipramine. Support Care
Cancer. 2008;16:1291–8.
Carpenter LL, Jocic Z, Hall JM, Rasmussen SA, Price LH. Mirtazapine augmentation in the
treatment of refractory depression. J Clin Psychiatry. 1999;60:45–9.
Carpenter LL, Yasmin S, Price LH. A double-blind, placebo-controlled study of antidepressant
augmentation with mirtazapine. Biol Psychiatry. 2002;51:183–8.
Chang HS, Won ES, Lee HY, Ham BJ, Kim YG, Lee MS. Association of ARRB1 polymorphisms
with the risk of major depressive disorder and with treatment response to mirtazapine. J
Psychopharmacol. 2015;29:615–22.
Chew-Graham CA, Shepherd T, Burroughs H, Dixon K, Kessler D. The value of an embedded
qualitative study in a trial of a second antidepressant for people who had not responded to one
antidepressant: understanding the perspectives of patients and general practitioners. BMC Fam
Pract. 2018;19:197.
Chung MY, Min KH, Jun YJ, Kim SS, Kim WC, Jun EM. Efficacy and tolerability of mirtazapine
and sertraline in Korean veterans with posttraumatic stress disorder: a randomized open label
trial. Hum Psychopharmacol. 2004;19:489–94.
Cipriani A, Furukawa TA, Salanti G, Chaimani A, Atkinson LZ, Ogawa Y, Leucht S, Ruhe HG,
Turner EH, Higgins JPT, Egger M, Takeshima N, Hayasaka Y, Imai H, Shinohara K, Tajika A,
Ioannidis JPA, Geddes JR. Comparative efficacy and acceptability of 21 antidepressant drugs for
the acute treatment of adults with major depressive disorder: a systematic review and network
meta-analysis. Lancet. 2018a;391:1357–66.
Cipriani A, Williams T, Nikolakopoulou A, Salanti G, Chaimani A, Ipser J, Cowen PJ, Geddes JR,
Stein DJ. Comparative efficacy and acceptability of pharmacological treatments for post-
traumatic stress disorder in adults: a network meta-analysis. Psychol Med. 2018b;48:1975–84.
Connor KM, Davidson JR, Weisler RH, Ahearn E. A pilot study of mirtazapine in post-traumatic
stress disorder. Int Clin Psychopharmacol. 1999;14:29–31.
Cornelius JR, Chung TA, Douaihy AB, Kirisci L, Glance J, Kmiec J, Wesesky MA, Fitzgerald D,
Salloum I. A review of the literature of mirtazapine in co-occurring depression and an alcohol use
disorder. J Addict Behav Ther Rehabil. 2016;5. https://doi.org/10.4172/2324-9005.1000159.
Croom KF, Perry CM, Plosker GL. Mirtazapine: a review of its use in major depression and other
psychiatric disorders. CNS Drugs. 2009;23:427–52.
Davidson JR, Weisler RH, Butterfield MI, Casat CD, Connor KM, Barnett S, Van Meter S.
Mirtazapine vs. placebo in posttraumatic stress disorder: a pilot trial. Biol Psychiatry.
2003;53:188–91.
Davis MP, Frandsen JL, Walsh D, Andresen S, Taylor S. Mirtazapine for pruritus. J Pain Symptom
Manag. 2003;25:288–91.
De Boer T. The effects of mirtazapine on central noradrenergic and serotonergic neurotransmission.
Int Clin Psychopharmacol. 1995;10(Suppl 4):19–23.
De Boer T. The pharmacologic profile of mirtazapine. J Clin Psychiatry. 1996;57(Suppl 4):19–25.
De Boer T, Broekkamp CL, Gower A, De Graaf JS, De Vos CJ, Rae D, Van Delft AM. The
pharmacological profile of Org 6906, a potential non-sedative antidepressant that combines
monoamine uptake inhibition with alpha 2-adrenolytic activity. Neuropharmacology.
1988a;27:251–60.
De Boer TH, Maura G, Raiteri M, De Vos CJ, Wieringa J, Pinder RM. Neurochemical and
autonomic pharmacological profiles of the 6-aza-analogue of mianserin, Org 3770 and its
enantiomers. Neuropharmacology. 1988b;27:399–408.
De Boer T, Nefkens F, Van Helvoirt A. The alpha 2-adrenoceptor antagonist Org 3770 enhances
serotonin transmission in vivo. Eur J Pharmacol. 1994;253:R5–6.
De Boer TH, Nefkens F, Van Helvoirt A, Van Delft AM. Differences in modulation of noradrenergic
and serotonergic transmission by the alpha-2 adrenoceptor antagonists, mirtazapine, mianserin
and idazoxan. J Pharmacol Exp Ther. 1996;277:852–60.
1302 M. P. Davis
other than anti-epileptic drugs and antibiotics: a population-based case-control study. Drug Saf.
2019;42:55–66.
Freynhagen R, Muth-Selbach U, Lipfert P, Stevens MF, Zacharowski K, Tolle TR, Von Giesen HJ.
The effect of mirtazapine in patients with chronic pain and concomitant depression. Curr Med
Res Opin. 2006;22:257–64.
Galling B, Calsina Ferrer A, Abi Zeid Daou M, Sangroula D, Hagi K, Correll CU. Safety and
tolerability of antidepressant co-treatment in acute major depressive disorder: results from a
systematic review and exploratory meta-analysis. Expert Opin Drug Saf. 2015;14:1587–608.
Gandhi S, Shariff SZ, Al-Jaishi A, Reiss JP, Mamdani MM, Hackam DG, Li L, Mcarthur E, Weir
MA, Garg AX. Second-generation antidepressants and hyponatremia risk: a population-based
cohort study of older adults. Am J Kidney Dis. 2017;69:87–96.
Gandotra K, Chen P, Jaskiw GE, Konicki PE, Strohl KP. Effective treatment of insomnia with
mirtazapine attenuates concomitant suicidal ideation. J Clin Sleep Med. 2018;14:901–2.
Gelenberg AJ, Mcgahuey C, Laukes C, Okayli G, Moreno F, Zentner L, Delgado P. Mirtazapine
substitution in SSRI-induced sexual dysfunction. J Clin Psychiatry. 2000;61:356–60.
Gnanadesigan N, Espinoza RT, Smith R, Israel M, Reuben DB. Interaction of serotonergic
antidepressants and opioid analgesics: is serotonin syndrome going undetected? J Am Med
Dir Assoc. 2005;6:265–9.
Goodnick PJ, Puig A, Devane CL, Freund BV. Mirtazapine in major depression with comorbid
generalized anxiety disorder. J Clin Psychiatry. 1999;60:446–8.
Guelfi JD, Ansseau M, Timmerman L, Korsgaard S, Mirtazapine-Venlafaxine Study G. Mirtazapine
versus venlafaxine in hospitalized severely depressed patients with melancholic features. J Clin
Psychopharmacol. 2001;21:425–31.
Gupta R, Gupta K, Tripathi AK, Bhatia MS, Gupta LK. Effect of mirtazapine treatment on serum
levels of brain-derived neurotrophic factor and tumor necrosis factor-alpha in patients of major
depressive disorder with severe depression. Pharmacology. 2016;97:184–8.
Hage B, Britton B, Daniels D, Heilman K, Porges SW, Halaris A. Low cardiac vagal tone index by
heart rate variability differentiates bipolar from major depression. World J Biol Psychiatry.
2019;20(5):359–367. https://doi.org/10.1080/15622975.2017.1376113. Epub 2017 Oct 5.
Harada T, Hirosawa T, Morinaga K, Shimizu T. Metoclopramide-induced Serotonin Syndrome.
Intern Med. 2017;56:737–9.
Hayashi Y, Watanabe T, Aoki A, Ishiguro S, Ueda M, Akiyama K, Kato K, Inoue Y, Tsuchimine S,
Yasui-Furukori N, Shimoda K. Factors affecting steady-state plasma concentrations of enantio-
meric mirtazapine and its desmethylated metabolites in Japanese psychiatric patients.
Pharmacopsychiatry. 2015;48:279–85.
Hoes MJ, Zeijpveld JH. Mirtazapine as treatment for serotonin syndrome. Pharmacopsychiatry.
1996;29:81.
Holm KJ, Markham A. Mirtazapine: a review of its use in major depression. Drugs.
1999;57:607–31.
Honig A, Kuyper AM, Schene AH, Van Melle JP, De Jonge P, Tulner DM, Schins A, Crijns HJ,
Kuijpers PM, Vossen H, Lousberg R, Ormel J, MIND-IT Investigators. Treatment of post-
myocardial infarction depressive disorder: a randomized, placebo-controlled trial with
mirtazapine. Psychosom Med. 2007;69:606–13.
Horstmann S, Dose T, Lucae S, Kloiber S, Menke A, Hennings J, Spieler D, Uhr M, Holsboer F, Ising
M. Suppressive effect of mirtazapine on the HPA system in acutely depressed women seems to be
transient and not related to antidepressant action. Psychoneuroendocrinology. 2009;34:238–48.
Houlihan DJ. Serotonin syndrome resulting from coadministration of tramadol, venlafaxine, and
mirtazapine. Ann Pharmacother. 2004;38:411–3.
Hundley JL, Yosipovitch G. Mirtazapine for reducing nocturnal itch in patients with chronic
pruritus: a pilot study. J Am Acad Dermatol. 2004;50:889–91.
Iwamoto K, Kawano N, Sasada K, Kohmura K, Yamamoto M, Ebe K, Noda Y, Ozaki N. Effects of
low-dose mirtazapine on driving performance in healthy volunteers. Hum Psychopharmacol.
2013;28:523–8.
1304 M. P. Davis
Jaquenoud Sirot E, Harenberg S, Vandel P, Lima CA, Perrenoud P, Kemmerling K, Zullino DF,
Hilleret H, Crettol S, Jonzier-Perey M, Golay KP, Brocard M, Eap CB, Baumann P. Multicenter
study on the clinical effectiveness, pharmacokinetics, and pharmacogenetics of mirtazapine in
depression. J Clin Psychopharmacol. 2012;32:622–9.
Jha MK, Malchow AL, Grannemann BD, Rush AJ, Trivedi MH. Do baseline sub-threshold
hypomanic symptoms affect acute-phase antidepressant outcome in outpatients with major
depressive disorder? Preliminary findings from the randomized CO-MED trial. Neuropsycho-
pharmacology. 2018;43:2197–203.
Jiang SM, Jia L, Liu J, Shi MM, Xu MZ. Beneficial effects of antidepressant mirtazapine in
functional dyspepsia patients with weight loss. World J Gastroenterol. 2016;22:5260–6.
Kamijima K, Yasuda M, Yamamura K, Fukuta Y. Real-world effectiveness and safety of
aripiprazole augmentation therapy in patients with major depressive disorder. Curr Med Res
Opin. 2018;34:2105–12.
Kang RH, Choi MJ, Paik JW, Hahn SW, Lee MS. Effect of serotonin receptor 2A gene polymor-
phism on mirtazapine response in major depression. Int J Psychiatry Med. 2007;37:315–29.
Karsten J, Hagenauw LA, Kamphuis J, Lancel M. Low doses of mirtazapine or quetiapine for
transient insomnia: a randomised, double-blind, cross-over, placebo-controlled trial. J
Psychopharmacol. 2017;31:327–37.
Kasper S. Clinical efficacy of mirtazapine: a review of meta-analyses of pooled data. Int Clin
Psychopharmacol. 1995;10(Suppl 4):25–35.
Kasper S. Efficacy of antidepressants in the treatment of severe depression: the place of mirtazapine.
J Clin Psychopharmacol. 1997;17(Suppl 1):19S–28S.
Kato M, Takekita Y, Koshikawa Y, Sakai S, Bandou H, Nishida K, Sunada N, Onohara A, Hatashita
Y, Serretti A, Kinoshita T. Non response at week 4 as clinically useful indicator for antidepressant
combination in major depressive disorder. A sequential RCT. J Psychiatr Res. 2017;89:97–104.
Kato T, Furukawa TA, Mantani A, Kurata K, Kubouchi H, Hirota S, Sato H, Sugishita K, Chino B,
Itoh K, Ikeda Y, Shinagawa Y, Kondo M, Okamoto Y, Fujita H, Suga M, Yasumoto S, Tsujino N,
Inoue T, Fujise N, Akechi T, Yamada M, Shimodera S, Watanabe N, Inagaki M, Miki K, Ogawa
Y, Takeshima N, Hayasaka Y, Tajika A, Shinohara K, Yonemoto N, Tanaka S, Zhou Q, Guyatt
GH, SUND Investigators. Optimising first- and second-line treatment strategies for untreated
major depressive disorder – the SUND study: a pragmatic, multi-centre, assessor-blinded
randomised controlled trial. BMC Med. 2018;16:103.
Kessler DS, Macneill SJ, Tallon D, Lewis G, Peters TJ, Hollingworth W, Round J, Burns A, Chew-
Graham CA, Anderson IM, Shepherd T, Campbell J, Dickens CM, Carter M, Jenkinson C,
Macleod U, Gibson H, Davies S, Wiles NJ. Mirtazapine added to SSRIs or SNRIs for treatment
resistant depression in primary care: phase III randomised placebo controlled trial (MIR). BMJ.
2018;363:k4218.
Khoo AL, Zhou HJ, Teng M, Lin L, Zhao YJ, Soh LB, Mok YM, Lim BP, Gwee KP. Network meta-
analysis and cost-effectiveness analysis of new generation antidepressants. CNS Drugs.
2015;29:695–712.
Kilic FS, Dogan AE, Baydemir C, Erol K. The acute effects of mirtazapine on pain related behavior
in healthy animals. Neurosciences (Riyadh). 2011;16:217–23.
Kim SW, Shin IS, Kim JM, Kang HC, Mun JU, Yang SJ, Yoon JS. Mirtazapine for severe
gastroparesis unresponsive to conventional prokinetic treatment. Psychosomatics.
2006;47:440–2.
Kirchheiner J, Henckel HB, Meineke I, Roots I, Brockmoller J. Impact of the CYP2D6 ultrarapid
metabolizer genotype on mirtazapine pharmacokinetics and adverse events in healthy volun-
teers. J Clin Psychopharmacol. 2004;24:647–52.
Kolla BP, Mansukhani MP, Bostwick JM. The influence of antidepressants on restless legs
syndrome and periodic limb movements: a systematic review. Sleep Med Rev. 2018;38:131–40.
Koutouvidis N, Pratikakis M, Fotiadou A. The use of mirtazapine in a group of 11 patients
following poor compliance to selective serotonin reuptake inhibitor treatment due to sexual
dysfunction. Int Clin Psychopharmacol. 1999;14:253–5.
Mirtazapine and Depressions 1305
Kouwenhoven TA, Van De Kerkhof PCM, Kamsteeg M. Use of oral antidepressants in patients with
chronic pruritus: a systematic review. J Am Acad Dermatol. 2017;77:1068–1073.e7.
Kraus T, Haack M, Schuld A, Hinze-Selch D, Koethe D, Pollmacher T. Body weight, the tumor
necrosis factor system, and leptin production during treatment with mirtazapine or venlafaxine.
Pharmacopsychiatry. 2002;35:220–5.
Laakmann G, Schule C, Baghai T, Waldvogel E. Effects of mirtazapine on growth hormone,
prolactin, and cortisol secretion in healthy male subjects. Psychoneuroendocrinology.
1999;24:769–84.
Laakmann G, Schule C, Baghai T, Waldvogel E, Bidlingmaier M, Strasburger C. Mirtazapine: an
inhibitor of cortisol secretion that does not influence growth hormone and prolactin secretion. J
Clin Psychopharmacol. 2000;20:101–3.
Laakmann G, Hennig J, Baghai T, Schule C. Influence of mirtazapine on salivary cortisol in
depressed patients. Neuropsychobiology. 2003;47:31–6.
Ladino M, Guardiola VD, Paniagua M. Mirtazapine-induced hyponatremia in an elderly hospice
patient. J Palliat Med. 2006;9:258–60.
Laimer M, Kramer-Reinstadler K, Rauchenzauner M, Lechner-Schoner T, Strauss R, Engl J,
Deisenhammer EA, Hinterhuber H, Patsch JR, Ebenbichler CF. Effect of mirtazapine treatment
on body composition and metabolism. J Clin Psychiatry. 2006;67:421–4.
Lee HY, Kang RH, Paik JW, Jeong YJ, Chang HS, Han SW, Lee MS. Association of the adrenergic
alpha 2a receptor – 1291C/G polymorphism with weight change and treatment response to
mirtazapine in patients with major depressive disorder. Brain Res. 2009;1262:1–6.
Lee SH, Paz-Filho G, Mastronardi C, Licinio J, Wong ML. Is increased antidepressant exposure a
contributory factor to the obesity pandemic? Transl Psychiatry. 2016;6:e759.
Leth-Moller KB, Hansen AH, Torstensson M, Andersen SE, Odum L, Gislasson G, Torp-Pedersen
C, Holm EA. Antidepressants and the risk of hyponatremia: a Danish register-based population
study. BMJ Open. 2016;6:e011200.
Lind AB, Reis M, Bengtsson F, Jonzier-Perey M, Powell Golay K, Ahlner J, Baumann P, Dahl ML.
Steady-state concentrations of mirtazapine, N-desmethylmirtazapine, 8-hydroxymirtazapine
and their enantiomers in relation to cytochrome P450 2D6 genotype, age and smoking behav-
iour. Clin Pharmacokinet. 2009;48:63–70.
Loonen AJ, Doorschot CH, Oostelbos MC, Sitsen JM. Lack of drug interactions between
mirtazapine and risperidone in psychiatric patients: a pilot study. Eur Neuropsychopharmacol.
1999;10:51–7.
Lovecchio F, Riley B, Pizon A, Brown M. Outcomes after isolated mirtazapine (Remeron)
supratherapeutic ingestions. J Emerg Med. 2008;34:77–8.
Lozupone M, La Montagna M, D’urso F, Piccininni C, Sardone R, Dibello V, Giannelli G, Solfrizzi
V, Greco A, Daniele A, Quaranta N, Seripa D, Bellomo A, Logroscino G, Panza F. Pharmaco-
therapy for the treatment of depression in patients with Alzheimer’s disease: a treatment-
resistant depressive disorder. Expert Opin Pharmacother. 2018;19:823–42.
Malamood M, Roberts A, Kataria R, Parkman HP, Schey R. Mirtazapine for symptom control in
refractory gastroparesis. Drug Des Devel Ther. 2017;11:1035–41.
Mattila M, Mattila MJ, Vrijmoed-De Vries M, Kuitunen T. Actions and interactions of psychotropic
drugs on human performance and mood: single doses of ORG 3770, amitriptyline, and
diazepam. Pharmacol Toxicol. 1989;65:81–8.
Mcgrane IR, Shuman MD. Mirtazapine therapy for posttraumatic stress disorder: implications of
alpha-adrenergic pharmacology on the startle response. Harv Rev Psychiatry. 2018;26:36–41.
Melo G, Dutra KL, Rodrigues Filho R, Ortega AOL, Porporatti AL, Dick B, Flores-Mir C, De Luca
Canto G. Association between psychotropic medications and presence of sleep bruxism: a
systematic review. J Oral Rehabil. 2018;45:545–54.
Menon V, Madhavapuri P. Low-dose mirtazapine-induced nightmares necessitating its discontinu-
ation in a young adult female. J Pharmacol Pharmacother. 2017;8:182–4.
Meyboom RH, Edwards IR, Egberts AC. Mirtazapine and the granulocytes – so far so good. N Z
Med J. 1999;112:104.
1306 M. P. Davis
Ramaekers JG, Conen S, De Kam PJ, Braat S, Peeters P, Theunissen EL, Ivgy-May N. Residual
effects of mirtazapine on actual driving performance: overall findings and an exploratory
analysis into the role of CYP2D6 phenotype. Psychopharmacology (Berl). 2011;215:321–32.
Rocha FL, Fuzikawa C, Riera R, Hara C. Combination of antidepressants in the treatment of major
depressive disorder: a systematic review and meta-analysis. J Clin Psychopharmacol.
2012;32:278–81.
Rush AJ, Trivedi MH, Stewart JW, Nierenberg AA, Fava M, Kurian BT, Warden D, Morris DW,
Luther JF, Husain MM, Cook IA, Shelton RC, Lesser IM, Kornstein SG, Wisniewski SR.
Combining medications to enhance depression outcomes (CO-MED): acute and long-term
outcomes of a single-blind randomized study. Am J Psychiatry. 2011;168:689–701.
Salvi V, Mencacci C, Barone-Adesi F. H1-histamine receptor affinity predicts weight gain with
antidepressants. Eur Neuropsychopharmacol. 2016;26:1673–7.
Sandker GW, Vos RM, Delbressine LP, Slooff MJ, Meijer DK, Groothuis GM. Metabolism of three
pharmacologically active drugs in isolated human and rat hepatocytes: analysis of interspecies
variability and comparison with metabolism in vivo. Xenobiotica. 1994;24:143–55.
Sanjida S, Mulvogue K, Shaw J, Couper J, Kissane D, Pearson SA, Price MA, Janda M. What type
and dose of antidepressants are cancer and non-cancer inpatients being prescribed: a retrospec-
tive case-control study at an Australian tertiary hospital. Support Care Cancer. 2018;26:625–34.
Sasada K, Iwamoto K, Kawano N, Kohmura K, Yamamoto M, Aleksic B, Ebe K, Noda Y, Ozaki N.
Effects of repeated dosing with mirtazapine, trazodone, or placebo on driving performance and
cognitive function in healthy volunteers. Hum Psychopharmacol. 2013;28:281–6.
Schneier FR, Campeas R, Carcamo J, Glass A, Lewis-Fernandez R, Neria Y, Sanchez-Lacay A,
Vermes D, Wall MM. Combined mirtazapine and SSRI treatment of PTSD: a placebo-controlled
trial. Depress Anxiety. 2015;32:570–9.
Schule C, Baghai T, Rackwitz C, Laakmann G. Influence of mirtazapine on urinary free cortisol
excretion in depressed patients. Psychiatry Res. 2003;120:257–64.
Schule C, Baghai TC, Eser D, Hecht S, Hermisson I, Born C, Hafner S, Nothdurfter C, Rupprecht
R. Mirtazapine monotherapy versus combination therapy with mirtazapine and aripiprazole in
depressed patients without psychotic features: a 4-week open-label parallel-group study. World J
Biol Psychiatry. 2007;8:112–22.
Schule C, Baghai TC, Eser D, Nothdurfter C, Rupprecht R. Lithium but not carbamazepine
augments antidepressant efficacy of mirtazapine in unipolar depression: an open-label study.
World J Biol Psychiatry. 2009;10:390–9.
Scoralick FM, Louzada LL, Quintas JL, Naves JO, Camargos EF, Nobrega OT. Mirtazapine does
not improve sleep disorders in Alzheimer’s disease: results from a double-blind, placebo-
controlled pilot study. Psychogeriatrics. 2017;17:89–96.
Sennef C, Timmer CJ, Sitsen JM. Mirtazapine in combination with amitriptyline: a drug-drug
interaction study in healthy subjects. Hum Psychopharmacol. 2003;18:91–101.
Seo HJ, Jung YE, Woo YS, Jun TY, Chae JH, Bahk WM. Effect of augmented atypical antipsy-
chotics on weight change in patients with major depressive disorder in a naturalistic setting.
Hum Psychopharmacol. 2009;24:135–43.
Sitsen J, Maris F, Timmer C. Drug-drug interaction studies with mirtazapine and carbamazepine in
healthy male subjects. Eur J Drug Metab Pharmacokinet. 2001;26:109–21.
Slee A, Nazareth I, Bondaronek P, Liu Y, Cheng Z, Freemantle N. Pharmacological treatments for
generalised anxiety disorder: a systematic review and network meta-analysis. Lancet.
2019;393:768–77.
Smit M, Dolman KM, Honig A. Mirtazapine in pregnancy and lactation – a systematic review. Eur
Neuropsychopharmacol. 2016;26:126–35.
Smit M, Wennink H, Heres M, Dolman KM, Honig A. Mirtazapine in pregnancy and lactation: data
from a case series. J Clin Psychopharmacol. 2015;35:163–7.
Smith WT, Glaudin V, Panagides J, Gilvary E. Mirtazapine vs. amitriptyline vs. placebo in the
treatment of major depressive disorder. Psychopharmacol Bull. 1990;26:191–6.
1308 M. P. Davis
Spaans E, Van Den Heuvel MW, Schnabel PG, Peeters PA, Chin-Kon-Sung UG, Colbers EP, Sitsen
JM. Concomitant use of mirtazapine and phenytoin: a drug-drug interaction study in healthy
male subjects. Eur J Clin Pharmacol. 2002;58:423–9.
Spyridi S, Sokolaki S, Nimatoudis J, Iacovides A, Kaprinis G. Status epilepticus in a patient treated
with olanzapine and mirtazapine. Int J Clin Pharmacol Ther. 2009;47:120–3.
Stahl S, Zivkov M, Reimitz PE, Panagides J, Hoff W. Meta-analysis of randomized, double-blind,
placebo-controlled, efficacy and safety studies of mirtazapine versus amitriptyline in major
depression. Acta Psychiatr Scand Suppl. 1997;391:22–30.
Steinert T, Froscher W. Epileptic seizures under antidepressive drug treatment: systematic review.
Pharmacopsychiatry. 2018;51:121–35.
Stimmel GL, Dopheide JA, Stahl SM. Mirtazapine: an antidepressant with noradrenergic and
specific serotonergic effects. Pharmacotherapy. 1997;17:10–21.
Stormer E, Von Moltke LL, Shader RI, Greenblatt DJ. Metabolism of the antidepressant mirtazapine
in vitro: contribution of cytochromes P-450 1A2, 2D6, and 3A4. Drug Metab Dispos.
2000;28:1168–75.
Szegedi A, Rujescu D, Tadic A, Muller MJ, Kohnen R, Stassen HH, Dahmen N. The catechol-O-
methyltransferase Val108/158Met polymorphism affects short-term treatment response to
mirtazapine, but not to paroxetine in major depression. Pharmacogenomics J. 2005;5:49–53.
Tack J, Carbone F. Functional dyspepsia and gastroparesis. Curr Opin Gastroenterol.
2017;33:446–54.
Tadic A, Muller MJ, Rujescu D, Kohnen R, Stassen HH, Dahmen N, Szegedi A. The MAOA
T941G polymorphism and short-term treatment response to mirtazapine and paroxetine in major
depression. Am J Med Genet B Neuropsychiatr Genet. 2007;144B:325–31.
Terhardt J, Lederbogen F, Feuerhack A, Hamann-Weber B, Gilles M, Schilling C, Lecei O,
Deuschle M. Heart rate variability during antidepressant treatment with venlafaxine and
mirtazapine. Clin Neuropharmacol. 2013;36:198–202.
Thase ME, Nierenberg AA, Vrijland P, Van Oers HJ, Schutte AJ, Simmons JH. Remission with
mirtazapine and selective serotonin reuptake inhibitors: a meta-analysis of individual patient
data from 15 controlled trials of acute phase treatment of major depression. Int Clin
Psychopharmacol. 2010;25:189–98.
Theobald DE, Kirsh KL, Holtsclaw E, Donaghy K, Passik SD. An open-label, crossover trial of
mirtazapine (15 and 30 mg) in cancer patients with pain and other distressing symptoms. J Pain
Symptom Manag. 2002;23:442–7.
Timmer CJ, Paanakker JE, Vrijmoed-De Vries M. Mirtazapine pharmacokinetics with two dosage
regimens and two pharmaceutical formulations. Pharm Res. 1997;14:98–102.
Timmer CJ, Sitsen JM, Delbressine LP. Clinical pharmacokinetics of mirtazapine. Clin
Pharmacokinet. 2000;38:461–74.
Tzeng DS, Chien CC, Lung FW, Yang CY. MAOA gene polymorphisms and response to
mirtazapine in major depression. Hum Psychopharmacol. 2009;24:293–300.
Varia I, Venkataraman S, Hellegers C, Gersing K, Doraiswamy PM. Effect of mirtazapine orally
disintegrating tablets on health-related quality of life in elderly depressed patients with comor-
bid medical disorders: a pilot study. Psychopharmacol Bull. 2007;40:47–56.
Verster JC, Van De Loo AJ, Roth T. Mirtazapine as positive control drug in studies examining the
effects of antidepressants on driving ability. Eur J Pharmacol. 2015;753:252–6.
Vidal C, Reese C, Fischer BA, Chiapelli J, Himelhoch S. Meta-analysis of efficacy of mirtazapine as
an adjunctive treatment of negative symptoms in schizophrenia. Clin Schizophr Relat Psycho-
ses. 2015;9:88–95.
Warden D, Trivedi MH, Carmody T, Toups M, Zisook S, Lesser I, Myers A, Kurian KR, Morris D,
Rush AJ. Adherence to antidepressant combinations and monotherapy for major depressive
disorder: a CO-MED report of measurement-based care. J Psychiatr Pract. 2014;20:118–32.
Waring WS, Good AM, Bateman DN. Lack of significant toxicity after mirtazapine overdose: a
five-year review of cases admitted to a regional toxicology unit. Clin Toxicol (Phila).
2007;45:45–50.
Mirtazapine and Depressions 1309
Contents
Chemistry, Developmental History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1311
Pharmacology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1312
Indications (of Marketed Products) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1312
Clinical Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1316
Side Effects/Adverse Reactions and Toxicology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1316
Combination Therapy: Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1316
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1316
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1317
Abstract
Vilazodone is an antidepressant, which enhances the serotonin concentration in
the synapse and also is a partial agonist for 5-HT1A-receptors.
The synthesis of vilazodone proceeds via two synthetic pathways. The first synthetic
pathway starts from indole-5-carbonitrile, which first reacts with
4-chlorobutyrochloride. A reduction of the keto-function results in the precursor
3-(4-chlorobutyl)-1H-indole-5-carbonitrile. In the second synthetic pathway,
5-nitrobenzofuran-2-carboxylic-acid-ethyl-ester is first reduced in a catalytic hydro-
genation to amine intermediate. The piperidine structure is built up by reaction with
bis(2-chloroethyl)amine. The condensation of the precursors of both synthesis
pathways results in vilazodone (Sorbera et al. 2001; Heinrich et al. 2004).
Pharmacology
Vilazodone is absorbed after oral administration and has a plasma half-life of 20–24 h.
When vilazodone is taken with food, the drug’s absolute bioavailability is 72%. It loses
50% of its bioavailability if this drug is taken without food (Cruz 2012; Schwartz et al.
2011; Wang et al. 2013). Steady state is achieved after approximately 4 days. It is a
selective serotonin reuptake inhibitor as well as a selective partial agonist for 5-HT1A-
receptors. Vilazodone blocks the serotonin reuptake pump (serotonin transporter),
desensitizes serotonin receptors (5-HT1A), and therefore presumably increases seroto-
nergic neurotransmission (Schwartz et al. 2011).
Vilazodone is metabolized extensively by the hepatic CYP3A4 enzyme. Its dose
should be reduced to 20 mg/day with concomitant use of any potent CYP3A4
inhibitors (Schwartz et al. 2011).
(continued)
Table 1 (continued)
1314
Study (RCT,
meta-analysis,
NIS; name/
acronym, authors, Study arms, comparators, Results (scores, response/ Dropout Relevant and severe side
year n, patients, country/ies and placebo remissions rates) rates effects
vilazodone for the treatment
of MDD in adults
Jain et al. (2014) Post hoc analyses were Vilazodone 40 mg/day or Montgomery-Åsberg
conducted on pooled data placebo in adult patients Depression Rating Scale
from two phase III, with major depressive (MADRS)
multicenter, 8-week, double- disorder
blind, randomized,
controlled trials (RCTs)
Croft et al. (2014) Eight-week, randomized (1: Vilazodone 40 mg/day Montgomery-Asberg Diarrhea and nausea; most
1), double-blind, placebo- with placebo in Depression Rating Scale incidences were mild in
controlled, parallel-group, outpatients with (MADRS) severity
fixed-dose study (placebo ¼ DSM-IV-TR-diagnosed
252, vilazodone ¼ 253) MDD
Mathews et al. Ten-week, multicenter, Vilazodone 20 or Montgomery-Åsberg Most common adverse
(2015) double-blind, placebo- 40 mg/day, citalopram Depression Rating Scale events (5% of vilazodone
controlled and active- 40 mg/day, or placebo (MADRS), Clinical Global patients, twice the rate of
controlled, fixed-dose trial, Impressions-Severity and placebo) were diarrhea,
1133 patients, (placebo ¼ sustained response nausea, vomiting
281; vilazodone (vilazodone 40 mg/day
20 mg/day ¼ 288; only), and insomnia
vilazodone 40 mg/day ¼
284; and citalopram ¼ 280)
Wang et al. (2013) Two pivotal 8-week, Dose titration up to Early onset of action, fewer
randomized, double- 2 weeks to reach a target sexual side effects, the
dose of 40 mg/day absence of known cardiac
F. Faltraco and J. Thome
blinded, placebo-controlled toxicity, and minimal effect
studies on weight gain
Zareifopoulos and Three RCTs (comprising Vilazodone (20-40 mg, HAMA reduction at week
Dylja (2017) 1453 patients), meta- mean dose ¼ 31.42 mg) 8, CGI-S reduction at week
analysis of all relevant 8, and CGI-I score at week 8
randomized controlled trials,
clinical trials on the use of
vilazodone in the treatment
of generalized anxiety
Vilazodone and Depressions
for 7 days, followed by 20 mg once daily for additional 7 days, and then an increase
to 40 mg once daily (Cruz 2012; Schwartz et al. 2011).
Clinical Studies
Two phase III studies enabled approval of vilazodone for treatment of major
depressive disorder and reported that vilazodone (40 mg/day) showed significant
superior efficacy over placebo (Rickels et al. 2009; Khan et al. 2011).
One pooled analysis aimed to assess the efficacy of vilazodone across a range of
symptoms and severities of depression. The study reported about significantly
improvement in depressive symptoms for the vilazodone-treated group compared
to placebo (Khan et al. 2014). A second study investigated the timing of depressive
symptom improvement and reported about vilazodone’s superior efficacy over
placebo in response rates (Jain et al. 2014). Two phase IV clinical trials confirmed
the efficacy and safety (Croft et al. 2014; Mathews et al. 2015).
Wang et al. reported that vilazodone exerts an antidepressive response after
1 week of treatment. After 8 weeks, it resulted in a 13% higher response compared
to placebo (Wang et al. 2013). Development for generalized anxiety disorder has
been stopped in 2017, because of the small benefit and high number of side effects
(Zareifopoulos and Dylja 2017) (Table 1).
Major side effects of vilazodone are diarrhea, nausea, headache, and dry mouth (may
affect up to 1 of 10 persons treated). Vilazodone carries an increased lethality
warning in young adult patients (may affect up to 1 of 1000 persons treated).
Adverse effects in relation to sexual dysfunction are reported, but the effects were
minimal with 1–2% over placebo rates (Howland 2011; Schwartz et al. 2011).
Cross-References
References
Croft HA, Pomara N, Gommoll C, Chen D, Nunez R, Mathews M. Efficacy and safety of
vilazodone in major depressive disorder: a randomized, double-blind, placebo-controlled trial.
J Clin Psychiatry. 2014;75(11):e1291–8.
Cruz MP. Vilazodone HCl (Viibryd): a serotonin partial agonist and reuptake inhibitor for the
treatment of major depressive disorder. P T. 2012;37(1):28–31.
Heinrich T, Böttcher H, Gericke R, Bartoszyk GD, Anzali S, Seyfried CA, Greiner HE, van
Amsterdam C. Synthesis and structure–activity relationship in a class of indolebutylpiperazines
as dual 5-HT1A receptor agonists and serotonin reuptake inhibitors. J Med Chem. 2004;47:
4684–92. https://doi.org/10.1021/jm040793q.
Howland RH. Vilazodone: another novel atypical antidepressant drug. J Psychosoc Nurs Ment
Health Serv. 2011;49(3):19–22.
Hu F, Su W. An investigation of the synthesis of vilazodone. J Chem Res. 2019; https://doi.org/10.
1177/1747519819893293.
Jain R, Chen D, Edwards J, Mathews M. Early and sustained improvement with vilazodone in adult
patients with major depressive disorder: post hoc analyses of two phase III trials. Curr Med Res
Opin. 2014;30(2):263–70.
Khan A, Cutler AJ, Kajdasz DK, Gallipoli S, Athanasiou M, Robinson DS, et al. A randomized,
double-blind, placebo-controlled, 8-week study of vilazodone, a serotonergic agent for the
treatment of major depressive disorder. J Clin Psychiatry. 2011;72(4):441–7.
Khan A, Sambunaris A, Edwards J, Ruth A, Robinson DS. Vilazodone in the treatment of major
depressive disorder: efficacy across symptoms and severity of depression. Int Clin
Psychopharmacol. 2014;29(2):86–92.
Mathews M, Gommoll C, Chen D, Nunez R, Khan A. Efficacy and safety of vilazodone 20 and
40 mg in major depressive disorder: a randomized, double-blind, placebo-controlled trial. Int
Clin Psychopharmacol. 2015;30(2):67–74.
Rickels K, Athanasiou M, Robinson DS, Gibertini M, Whalen H, Reed CR. Evidence for efficacy
and tolerability of vilazodone in the treatment of major depressive disorder: a randomized,
double-blind, placebo-controlled trial. J Clin Psychiatry. 2009;70(3):326–33.
Schwartz TL, Siddiqui UA, Stahl SM. Vilazodone: a brief pharmacological and clinical review of
the novel serotonin partial agonist and reuptake inhibitor. Ther Adv Psychopharmacol.
2011;1(3):81–7.
Sorbera LA, Rabasseda X, Silvestre J, Castaner J. Vilazodone hydrochloride. Antidepressant,
5-HT1A partial agonist, 5-HT reuptake inhibitor. Drugs Future. 2001;36:247–52.
Traynor K. Vilazodone approved for major depression. Am J Health Syst Pharm. 2011;68(5):366.
Wang SM, Han C, Lee SJ, Patkar AA, Masand PS, Pae CU. A review of current evidence for
vilazodone in major depressive disorder. Int J Psychiatry Clin Pract. 2013;17(3):160–9.
Zareifopoulos N, Dylja I. Efficacy and tolerability of vilazodone for the acute treatment of
generalized anxiety disorder: a meta-analysis. Asian J Psychiatr. 2017;26:115–22.
Serotonin-Norepinephrine Reuptake
Inhibitors (SNRIs): Venlafaxine,
Desvenlafaxine, Duloxetine, Milnacipran,
Levomilnacipran
Contents
Chemistry, Developmental History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1320
Pharmacology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1320
Pharmacokinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1321
Mechanism of Action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1322
Indications (of Marketed Products) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1322
Clinical Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1324
Desvenlafaxine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1335
Duloxetine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1335
Milnacipran . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1335
Levomilnacipran . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1336
Side Effects, Adverse Reactions, Toxicology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1336
Desvenlafaxine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1337
Duloxetine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1337
Milnacipran . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1337
Levomilnacipran . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1337
Pregnancy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1338
Discontinuation Symptoms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1338
Combination Therapy: Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1338
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1339
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1339
G. Laux (*)
Institute of Psychological Medicine (IPM), Soyen, Germany
MVZ Waldkraiburg, Center of Neuropsychiatry, Waldkraiburg, Germany
Department of Psychiatry and Psychotherapy, Ludwig-Maximilians-University (LMU), Munich,
Germany
e-mail: ipm@ipm-laux.de
A. Klimke
Vitos Waldkrankenhaus Köppern, Friedrichsdorf, Germany
Abstract
Approved members of the selective serotonin (5-HT) and norepinephrine
(NE) reuptake inhibitor (SNRI) class so long are venlafaxine, desvenlafaxine,
duloxetine, milnacipran, and levomilnacipran. They differ regarding binding
affinity to 5-HT/NE-transporter binding and uptake inhibition, pharmacokinetic
properties (half-life, metabolism), and clinical profile (e.g., analgesic properties).
Numerous randomized placebo-controlled clinical studies (RCTs) and
non-interventional studies have shown efficacy and effectiveness. Some studies
reported higher efficacy compared to serotonin selective antidepressants (SSRIs)
proposing these drugs for severe and so-called therapy-resistant depressions.
Typical side effects are nausea, trouble sleeping, and sexual dysfunction.
Pharmacology
All SNRIs are more potent 5-HT than NE reuptake inhibitors (Cusack et al. 1994;
Harvey et al. 2000; Richelson 2002). Venlafaxine and its active metabolite
O-desmethylvenlafaxine (Desvenlafaxine) inhibit the neural uptake of 5-HT, NE
Serotonin-Norepinephrine Reuptake Inhibitors (SNRIs): Venlafaxine,. . . 1321
Milnacipran Levomilnacipran
and with lower potency dopamine and has very low affinity for muscarinic, alpha
adrenergic, or histamine receptors (Bymaster et al. 2001).
Duloxetine and milnacipran inhibit 5-HT and NE uptake, duloxetine has weak
affinity for the dopamine transporter, both have weak or no significant affinity for
muscarinic, α-adrenergic, histaminergic, opioid, glutamate, and GABA receptors.
Analgesic properties are associated to sodium ion channel blockade.
Milnacipran exerts no significant actions on H1, α1, D1, D2, and mACh receptors
(Puozzo et al. 2002).
Pharmacokinetics
Venlafaxine has linear kinetics, is metabolized in the liver via the CYP2D6 isoen-
zyme to the active metabolite O-desmethylvenlafaxine (desvenlafaxine). When most
normal metabolizers take venlafaxine, approximately 70% of the dose is metabo-
lized into desvenlafaxine (Ereshefsky and Dugan 2000).
Duloxetine is a moderate CYP2D6 inhibitor (Knadler et al. 2011). Smoking is
associated with a decrease in duloxetine concentration (Fric et al. 2008).
1322 G. Laux and A. Klimke
Mechanism of Action
SNRIs have a broad spectrum of indications. They are used and in many countries
released for the treatment of major depressive disorder, anxiety disorders, such as
panic disorder, social anxiety disorder, obsessive-compulsive disorder (OCD), bulimia
Serotonin-Norepinephrine Reuptake Inhibitors (SNRIs): Venlafaxine,. . . 1323
1
1
1
9 1,6
30*-3
nervosa, and premenstrual dysphoric disorder (PMDD) (Stahl 2017; Shelton 2018;
Ghaemi 2019).
Venlafaxine is approved to treat major depressive disorder (MDD), general-
ized anxiety disorder (GAD), panic disorder, and social phobia. Off-label,
venlafaxine can be used for attention deficit disorder, fibromyalgia, diabetic
neuropathy, complex pain syndromes, hot flashes, migraine prevention, post-
traumatic stress disorder, obsessive-compulsive disorder, and premenstrual dys-
phoric disorder.
Starting dose in MDD is 75 mg/day, approved maximum doses are up to
375 mg/day in absence of response to lower doses due to wide inter-individual
differences in blood levels. Dose should be reduced in case of hepatic or renal
impairment.
Desvenlafaxine is approved for MDD in USA. Doses of 10 to 400 mg/day were
studied in clinical trials. There is no evidence that doses greater than 50 mg per day
provide additional benefit.
Duloxetine is used to treat major depressive disorder, generalized anxiety disor-
der, fibromyalgia, neuropathic pain, and stress urinary incontinence. Duloxetine was
approved for the pain associated with diabetic peripheral neuropathy. 2010 FDA
approved duloxetine to treat chronic musculoskeletal pain, including discomfort
from osteoarthritis and chronic lower back pain.
Starting dose for MDD is 60 mg/day, recommended maximal doses are 120 mg/day.
Milnacipran is prescribed for major depressive disorder, fibromyalgia, and neu-
ropathic pain/chronic pain.
Usual dosage is 50 mg b.i.d., begin at 25 mg b.i.d. increase up to 100 mg twice
daily, diminished in case of renal dysfunction.
Levomilnacipran initial dose is 20 mg orally once a day for 2 days, then increase
to 40 mg orally once a day. Maintenance dose: 40 to 120 mg orally once a day,
maximum dose: 120 mg/day.
1324 G. Laux and A. Klimke
Clinical Studies
Multiple RCTs proved SNRIs superior to placebo required for approvement (e.g.,
Khan et al. 1991; Entsuah et al. 1996 for venlafaxine), they are as effective as
tricyclic antidepressants but better tolerated (Shelton 2018; Ghaemi 2019). Table 2
gives a selection of RCTs between SNRIs and TCAs and SSRIs as well as some
major non-interventional/surveillance/observational studies (NIS). In conclusion,
efficacy and effectiveness of SNRIs is established.
A review concluded that second-generation antidepressants appear equally effec-
tive, although they may differ in efficacy (response rates) and acceptance (dropout
rates). In a network meta-analysis comparing 21 major antidepressants beside
amitriptyline, mirtazapine, and paroxetine the SNRIs venlafaxine, duloxetine, and
milnacipran are on the top 6 ranking regarding efficacy (Cipriani et al. 2018).
Treatment guidelines issued by the National Institute of Health and Clinical
Excellence (NICE), the American Psychiatric Association (APA), and the German
Association of Psychiatry, Psychotherapy and Psychosomatics (DGPPN) reflect the
viewpoint of equal efficacy.
However, some studies and reviews reported higher efficacy of venlafaxine
compared with SSRIs (Clerc et al. 1994; Smith et al. 2002), it was at least as
effective or more effective as the TCAs (Einarson et al. 1999; Stahl et al. 2002). A
re-evaluation of studies reported no better efficacy or tolerability conpared with
SSRIs (Weinmann et al. 2008).
According the Cochrane Collaboration Depression, Anxiety and Neurosis Review
Group Controlled Trials Register all RCTs comparing fluoxetine with any other AD
for patients with unipolar major depressive disorder 171 studies were included in the
analysis (24,868 participants). Fluoxetine was as effective as the TCAs, less effective
than venlafaxine which was more effective than milnacipran (Magni et al. 2013).
Efficacy of venlafaxine for the treatment of Generalized Anxiety Disorder (GAD)
has been established in various RCTs (e.g., Gelenberg et al. 2000; Rickels et al.
2000; Allgulander et al. 2001). Efficacy in the treatment of panic disorder has been
compared with paroxetine and placebo in 664 non-depressed adult outpatients; 75%
of patients were responders, nearly 45% achieved remission under 75 or150 mg/day
venlafaxine or paroxetine 40 mg/day. Venlafaxine and paroxetine showed greater
improvement than placebo. Panic-free rates after 12 weeks were 54% to 61%
compared with 35% for placebo, 75% of patients given active treatments were
responders, nearly 45% achieved remission. Placebo response rate was 55%, remis-
sion near 25% (Pollack et al. 1996).
N ¼ 440 adult outpatients with DSM-IV generalized social anxiety disorder were
randomly assigned for 6 months or longer to receive venlafaxine mean daily dose
202 mg, paroxetine (46 mg/day), or placebo for 12 weeks. Response rates were 58.6%
for venlafaxine, 62.5% for paroxetine, and 36.1% for placebo (Liebowitz et al. 2005).
A systematic review and network meta-analysis including 67 RCTs found superior
response compared to placebo for paroxetine mainly, escitalopram and sertraline.
Venlafaxine performed worse due to adverse events dropouts (Williams et al.
2020). Effects regarding neuropathic pain have been reported (Gallagher et al. 2015).
Table 2 Survey review SNRIs studies in MDD
Dropout rates,
N patients Study arms, comparators, Results Response-(R)/Remission-Rate Dis-continuation Relevant side
Study Duration Placebo (RR) rates effects
Venlafaxine (V)
Schweizer et al. N ¼ 60 V 75 mg/day, 225 mg/day, V > Pl all doses V nervousness,
1991 6 weeks 375 mg/day R: V 68%, Pl 31% sweating, nausea
RCT vs. Pl
Cunningham N ¼ 225 V vs. Trazodone (T) vs. Pl V, T > Pl 33% V nausea, T
et al. 1994 6 weeks V better in cognitive factors and dizziness,
N ¼ 96 retardation, T in sleep disturbance somnolence
Responders
1 year
continuation
Schweizer et al. N ¼ 224 V Ø182 mg/day R: V 90%, I 79% V 16%,
1994 6 weeks Imipramine (I) 176 mg/day Pl 53% I 25%
RCT Pl
Nierenberg et al. N ¼ 84 R: 30–40% according to HAMD,
1994 treatment- MADRS
Open study resistant out- Maintained for 3 months
and inpts.
12 weeks
Entsuah et al. N ¼ 304 N ¼ 185 V, N ¼ 62 Imipra- Relapse rates V 20%, I 31%, T 29%, Pl Completers V
Serotonin-Norepinephrine Reuptake Inhibitors (SNRIs): Venlafaxine,. . .
Table 2 (continued)
Dropout rates,
N patients Study arms, comparators, Results Response-(R)/Remission-Rate Dis-continuation Relevant side
Study Duration Placebo (RR) rates effects
Costa e Silva N ¼ 382 V 75–150 mg/ V ¼ F, higher V doses (>) F V Nausea,
et al. 1998 8 weeks day vs. Fluoxetine dizziness, F
RCT (F) 20–40 mg/day Nausea, insomnia
Samuelian and N ¼ 102 V or Clomipramine R: V 59–62%, V 13%, C 20% C anticholinergic
Hackett 1998 43 days (C) 150 mg/day C 43–54% TEAE,
RCT tachycardia,
#blood pressure
Einarson et al. N ¼ 4033 V vs. SSRIs citalopram, V > SSRIs, TCAs
1999 HAMD>15, fluoxetine, fluvoxamine, R: V 73.7%, SSRIs 61.1%, TCAs
Meta-analysis MADRS>18 paroxetine, sertraline vs. TCAs 57.9%
RCTs 44 trials amitriptyline, imipramine,
desipramine, nortriptyline
Rudolph and N ¼ 232 N ¼ 100 V 75–225 mg/ RR: V 37%, F 22%, Pl 18% V 6%, F 9%
Feiger 1999 8 weeks d vs. N ¼ 103 Fluoxetine
RCT (F) 20–60 mg/day
N ¼ 98 Pl
Poirier and N ¼ 122 V 200–300 mg/day, Paroxetine R: V 51.9%, P 32,7% V¼P
Boyer 1999 Treatment- (P) 30–40 mg/day RR: V 42.3%, P 20.0%
RCT resistant
depressions
Ballus et al. N ¼ 84 N ¼ 41 V 75–150 mg/ R: V 55%, P 29% V 39%, V Nausea (28%),
2000 HAMD > 17, day vs. N ¼ 43 Paroxetine RR: V 59%, P 31% P 26% P Headache (40%)
RCT Dysthymia (P) 20–40 mg/day
12 weeks
G. Laux and A. Klimke
Tzanakaki et al. N ¼ 109 N ¼ 55 V 75–225 mg/ V>F 22% Nausea V 5.5%, F
2000 Inpts with day vs. N ¼ 54 Fluoxetine R: V V 70%, F 66% Due to side 14.8%
RCT melancholia (F) 20–60 mg/day RR: V 41%, F 36% effects V 5%, F
HAMD >25 9%
6 weeks
Mehtonen et al. N ¼ 147 N ¼ 75 V 75–150 mg/ V>S V 21% V Nausea,
2000 8 weeks day, vs. N ¼ 72 Sertraline R: V 83%, S 68% S 17% sweating
RCT (S) 50–100 mg/day RR: V 67%, S 36% S Nausea, diarrhea
Thase et al. 2001 N ¼ 2045 N ¼ 851 V vs. N ¼ 748 SSRIs V > SSRIs
Pool analysis (Fluoxetine, paroxetine, RR: V 45%, SSRIs 35%,
8 RCTs fluvoxamine Pl 25%
vs. N ¼ 446 Pl
Entsuah and N ¼ 301 N ¼ 100 V 75–225 mg/ Response benefit: V 66%, F 53%, Pl
Gorman 2002 day vs. N ¼ 103 Fluoxetine 52%
Retrospective (F) 20–60 mg/day vs. N ¼ 98 Remission benefit: V 34%, F 19%, Pl
Global benefit- Pl 19%
risk (GBR)
Stahl et al. 2002 8 weeks V vs. SSRIs vs. Pl V > SSRIs > Pl R: V 67–71%, SSRIs
8 RCTs Pool 59–64%
analysis Pl 41–50%
Sauer et al. 2003 N ¼ 160 N ¼ 60 V 75–150 mg/day V ¼A A due to dry V >A
RCT 6 weeks N ¼ 57Amitriptyline Performance d2 test V > A mouth
(A) 75–150 mg/day
Serotonin-Norepinephrine Reuptake Inhibitors (SNRIs): Venlafaxine,. . .
Montgomery N ¼ 235 N ¼ 109 V 100–200 mg/ Relapse V 22%, Pl 55% V 21%, Pl 48%
et al. 2004 1 year day vs. N ¼ 116 Pl due to lack of
efficacy
Bielski et al. N ¼ 197 N ¼ 98 V 225 mg/day V¼E E>V
2004 8 weeks N ¼ 97 Escitalopram R: V 48.0%
RCT (E) 20 mg/day E 58.8%
RR: V 36.7%
E 41.2%
1327
(continued)
1328
Table 2 (continued)
Dropout rates,
N patients Study arms, comparators, Results Response-(R)/Remission-Rate Dis-continuation Relevant side
Study Duration Placebo (RR) rates effects
Thase et al. 2006 N ¼ 348 V 150–225 mg/ RR: V 33%, B 46% V more sexual
RCT 12 weeks day vs. Bupropion dysfunction
(B) 300–450 mg/day
Keller et al. N ¼ 258 V Ø 224.7 mg/day vs. Pl Recurrence V 28%, Pl 63%
2007/Kocsis 1 and 2 years 1 year V 23%, Pl 42%
et al. 2007/ 2 years V 28%
Kornstein et al. Pl 47%
2008
PREVENT-
Study, RCT
Nemeroff et al. N ¼ 8744 N ¼ 4191 V V > Fluoxetine Due to adverse
2008 Ø 151 mg/day RR difference 5.9% favoring V effects V 11%,
34 RCTs Meta- N ¼ 3621 SSRIs Drug-placebo differences V 13%, SSRIs 9%
analysis N ¼ 932 Pl SSRIs 6%
Schueler et al. N ¼ 18.180 N ¼ 12.816 V, N ¼ 4.528 V and D > Pl V, D > Pl V, D > Pl
2011 Duloxetine (D), different R and RR: SSRIs > V
54 RCTs Meta- SSRIs V>D V ¼ TCAs
analysis V ¼ TCAs D < SSRIs, V
Thase et al. 2011 N ¼ 259 N ¼ 160 V, Non-recurrence V 71.9%
PREVENT Recurrence N ¼ 99 Fluoxetine (F) F 55.8%
Study 2 years
RCT
Thase et al. 2016 N ¼ 1087 V 75–225 mg/day vs. Pl V > Pl at week 2, no sign. interaction V 9.4%,
Meta-analysis with baseline HAMD-scores Pl 3.6%
5 RCTs
G. Laux and A. Klimke
Fagiolini et al. N ¼ 324 N ¼ 158 V vs. N ¼ 166 V HAMD-total score – 16.4, V nausea
2020 8 weeks Trazodone (T) T – 15.4 T dizziness,
RCT somnolence
Desvenlafaxine (Dv)
Thase et al. 2009 N ¼ 2913 N ¼ 1805 Dv R: Dv 53% Pl 41% Dv 4–18%
Meta-analysis 8 weeks N ¼ 1108 Pl RR: Dv 32%, Pl 23% (increase with
9 RCTs dose)
Pl 3%
Meta-analysis
RCT
Laoutidis and 17 trials Dv 50,100,200, 400 mg/day Risk ratio for response 1.24, for Risk ratio 1.16 Risk ratio 1.98
Kioulos 2015 Pl Fluoxetine, Venlafaxine remission 1.29 vs. Pl, 0.90 or 0.82 in
Review, Meta- head-to-head trials. Not as efficacious
analysis RCTs as other antidepressants
Duloxetine (D)
Goldstein et al. N ¼ 160 D 40 mg/day vs. D 80 mg/ D 40 and 80 mg/day > Pl Depression Insomnia D
2004 8 weeks day vs. Paroxetine (P) 20 mg/ scores and pain severity 19.8%, P 8%
RCT day vs. Pl P ¼ Pl
RR D 80 mg/day 57%, P 34%
Nelson et al. N ¼ 90 N ¼ 47 D 60 mg/day RR: D 44.1% D 21.0%
2005 >55 years N ¼ 43 Pl Pl 16.1% Pl 6.7%
RCT 9 weeks Pain reduction D > Pl
Eckert and 22 studies D N ¼ 1280 vs. Fluoxetine D¼F D¼F
Serotonin-Norepinephrine Reuptake Inhibitors (SNRIs): Venlafaxine,. . .
Table 2 (continued)
Dropout rates,
N patients Study arms, comparators, Results Response-(R)/Remission-Rate Dis-continuation Relevant side
Study Duration Placebo (RR) rates effects
Perahia et al. N ¼ 392 N ¼ 93 D 80 mg/day D 80 and 120 mg/day > Pl D¼P
2006 8 months N ¼ 103 D120 mg/day D¼P
RCT N ¼ 97 P 20 mg/day
N ¼ 99 Pl
Perahia et al. N ¼ 288 D60–120 mg/day vs. Pl Relapse rate D ¼ Pl due to D ¼ Pl
2009 52 weeks D 14.4%, adverse events,
RCT Relapse Pl 33.1% vital signs, or
prevention weight
43 centers
(France,
Germany,
Italy, Russia,
Sweden,
USA)
Oakes et al. N ¼ 681 N ¼ 456 D 60 mg/day D > Pl at 4,8,16 and 20 weeks (not at D 15.3% Dry mouth,
2013 >65 years N ¼ 225 Pl 12 weeks) in depression and pain scales Pl 5.8% constipation,
Robinson et al. 12/24 weeks nausea, diarrhea,
2014 USA, France, fatigue
RCT Mexico, 2x > Pl
Puerto Rico
Cipriani et al. N ¼ 5735 pts D vs. Paroxetine, Escitalopram E and V > D D higher dropout
2012 (E), Fluoxetine, Venlafaxine, as E and V
Review 16 RCTs desvenlafaxine
G. Laux and A. Klimke
Milnacipran (M)
Macher et al. N ¼ 58 M 2 50 mg/day vs. Pl M > Pl M 10.3%, Pl M¼P
1989 Severe 55.2%
RCT depressions
(MADRS
>25)
5 weeks
Ansseau et al. N ¼ 45 M 50 mg/day vs. 100 mg/day M 100 ¼ A > M 50 More
1989 5 weeks vs. Amitriptyline (A) 150 mg/d anticholinergic
side effects,
weight gain, #
blood pressure
Kasper et al. N ¼ 842 M 2 50 mg/day vs. TCAs R: M 64%, TCA 67% M > TCAs,
1996 4–12 weeks Imipramine and Clomipramine RR: M 39%, TCAs 42% dysuria M
Meta-analysis 2 75 mg/day (2.1%) > TCAs
7 studies
Lopez-Ibor et al. M 250 mg/day vs. Fluoxetine M > SSRIs M fewer
1996 20 mg/days, Fluvoxamine R: M 64%, SSRIs 50% gastrointestinal
Meta-analysis 2100 mg/day RR: M 39%, SSRIs 28% side effects, M
RCTs more dysuria
Puech et al. 1997 N ¼ 1032 M 2 50 mg/ M ¼ I, > SSRIs M>I
Meta-analysis day vs. SSRIs, vs. Imipramine R: M 64%, SSRIS 50% M more dysuria, I
(I) more
Serotonin-Norepinephrine Reuptake Inhibitors (SNRIs): Venlafaxine,. . .
anticholinergic
side effects, SSRIs
nausea
Leinonen et al. N ¼ 107 N ¼ 52 M 200 mg/ C>M Due to TEAE M M insomnia
1997 26 weeks day vs. N ¼ 55 Clomipramine R; M 58%, C 72% 21%, C 38%; C dry mouth
RCT (C) 150 mg/day RR: M 45%, C 63% due to lack of
efficacy M 19%,
C 7%
1331
(continued)
1332
Table 2 (continued)
Dropout rates,
N patients Study arms, comparators, Results Response-(R)/Remission-Rate Dis-continuation Relevant side
Study Duration Placebo (RR) rates effects
Steen and Den Therapy- M 100–200 mg/ M¼C
Bour 1997 resistant day, vs. Clomipramine Poor antide-pressant activity
RCT depressions (C) 150 mg/day R: 30%
6 months
10 hospitals
NL
Tignol et al. N ¼ 219 M 2 50 mg/day M¼I 72 pts. I anticholinergic
1998 Elderly pts. Imipramine (I) 250 mg/day side effects
8 weeks
Guelfi et al. N ¼ 289 M 100 or 200 mg/ M¼F M more " heart
1998 inpatients day vs. Fluoxetine (F) 20 mg/ rate, F weight loss
RCT 12 weeks day
Rouillon et al. N ¼ 204 N ¼ 104 M 2 50 mg/ Recurrence M 16.3%, Pl 23.6% M 18%,
2000 1 year day vs. N ¼ 110 Pl Pl 25%
RCT
Clerc 2001 N ¼ 113 M 2 50 mg/day vs. Fluvox- R: M 78.9%, F 60.7% F tremor,
RCT 6 weeks amine (F) 2 100 mg/day drowsiness
Van Amerongen N ¼ 109 N ¼ 53 M 100 mg/ M¼I I>M M>I
et al. 2002 6 weeks day vs. N ¼ 56 Imipramine
RCT (I) 150 mg/day
Sechter et al. N ¼ 300 M 100 mg/day vs. Paroxetine M¼P P more
2004 6 weeks (P) 20 mg/day R: M 62.8%, P 64.9% discontinu-ation
RCT symptoms
G. Laux and A. Klimke
Papakostas et al. N ¼ 17.036 SNRIs vs. SSRIs SNRIs (>) SSRIs
2007 R: SNRIs 63.6%, SSRIs 59.3%
Meta-analysis
93 trials
Nakagawa et al. N ¼ 2.277 M vs. SSRIs, other SNRIs, M ¼ other AD M > TCAs due M < sedation,
2009 TCAs to side effects dizziness,
Cochrane meta- constipation, dry
analysis mouth
16 RCTs
Olie et al. 2010 N ¼ 195 M and Venlafaxine R: 70% M, 77% V 31% 72% M, 74% V
RCT 24 weeks (V) 100–200 mg/day RR: 52.2% M, 62.1% V (mild)Nausea,
dizziness,
hyperhidrosis,
dysuria (♂ M),
sexual dysfunction
(V)
Levomilnacipran (Lm)
Asnis et al. 2013 N ¼ 724 N ¼ 179 Pl, N ¼ 181 Lm40 R: Pl 29.1%, Lm 40 36.4%, Lm Pl 21.6%, Lm40 TEAEs Lm > Pl:
RCT outpatients mg/day, N ¼ 181 Lm 80 mg/ 80 37.3%, Lm 120 41.5% 27%, Lm Headache, nausea,
8 weeks day, N ¼ 83 Lm 120 mg/day RR: 80 32.4%, constipation,
19.4% vs. 21.6%, vs. 20.9% vs. 20.5% Lm120 35% hyperhidrosis,
vomiting, urinary
hesitation
Serotonin-Norepinephrine Reuptake Inhibitors (SNRIs): Venlafaxine,. . .
Montgomery N ¼ 553 N ¼ 276 Lm 75 or 100 mg/day, R: Lm 59.1%; Pl 42.2% Lm 9.4%, Pl Lm > Pl: Hyper-
et al. 2013 Outpatients N ¼ 277 Pl RR: 46.4% vs. 26.0% 6.5% due to hidrosis,
RCT 10 weeks MADRS adverse events constipation,
diarrhea,
tachycardia,
hypertension
(continued)
1333
1334
Table 2 (continued)
Dropout rates,
N patients Study arms, comparators, Results Response-(R)/Remission-Rate Dis-continuation Relevant side
Study Duration Placebo (RR) rates effects
Sambunaris N ¼ 434 Lm 40–120 mg/day Lm > Pl Pl 61.8%
et al. 2014 10 weeks vs. Pl Lm 81.6%
RCT Nausea, dizziness,
constipation,
tachycardia,
hypertension,
urinary hesitation,
hyperhidrosis,
insomnia,
ejaculation
disorder
Note: > significantly more effective than or better as or more frequent as, ¼ not significantly different, HAMD Hamilton Depression Scale, MADRS
Montgomery Asberg Depression Rating Scale, Pl ¼ Placebo, TEAE ¼ Treatment Emergent Adverse Events
G. Laux and A. Klimke
Serotonin-Norepinephrine Reuptake Inhibitors (SNRIs): Venlafaxine,. . . 1335
Desvenlafaxine
An integrated analysis compared 50 and 100 mg/day with placebo in adult out-
patients from 4317 pts. Statistically significant improvements versus placebo were
observed for all efficacy endpoints (Carrasco et al. 2016). An indirect comparison
meta-analysis showed no differences between Dv and venlafaxine regarding effi-
cacy, nausea, and dropout rates favored Dv (Coleman and Xavier 2012).
Duloxetine
A 2012 Cochrane Review did not find greater efficacy of duloxetine compared to
SSRIs and newer antidepressants. Additionally, the review found evidence that
duloxetine has increased side effects and reduced tolerability compared to other
antidepressants. It thus did not recommend duloxetine as a firstline treatment for
major depressive disorder. Duloxetine appears less tolerable than some other anti-
depressants (Cipriani et al. 2012).
Duloxetine is more effective than placebo in the treatment of generalized anxiety
disorder (GAD) (Carter and McCormack 2009). In a relapse prevention study (N ¼ 887;
mean age ¼ 43.3 years) over 26 weeks placebo-treated patients relapsed more fre-
quently (41.8%) than duloxetine-treated patients (13.7%) (Davidson et al. 2008).
Duloxetine is used in pain and stress urinary incontinence treatment.
One review of duloxetine found that it reduced pain and fatigue, and improved
physical and mental performance in fibromyalgia compared to placebo (Acuna
2008). Another systematic review noted that tricyclic antidepressants (imipramine
and amitriptyline), traditional anti-convulsants and opioids have better efficacy in
painful diabetic neuropathy than duloxetine (Wong et al. 2007). The comparative
data collected by reviewers in BMC Neurology indicated that amitriptyline, other
tricyclic antidepressants, and venlafaxine may be more effective for painful diabetic
neuropathy and fibromyalgia (Sultan et al. 2008). A newer review based on the
Cochrane Neuromuscular Group Specialized Register analyzed benefits and harms
of duloxetine for treating painful neuropathy and different types of chronic pain.
Eighteen trials with 6407 participants were included. Doses of 60 mg and 120 mg
daily were efficacious for treating pain in diabetic peripheral neuropathy, the effect in
fibromyalgia was attributed to greater improvement in mental symptoms than in
somatic physical pain. Minor side effects were common, direct comparisons of
duloxetine with other antidepressants and with other drugs, such as pregabalin
were recommended (Lunn et al. 2014).
Milnacipran
discontinuation for side effects or lack of efficacy (Papakostas et al. 2007). A meta-
analysis of a total of 16 randomized controlled trials with more than 2200 patients
concluded that there were no statistically significant differences in efficacy, accept-
ability, and tolerability when comparing milnacipran with other antidepressant
agents. However, compared with TCAs, significantly fewer patients taking
milnacipran dropped out due to adverse events (Ansseau et al. 1989; Nakagawa
et al. 2008, 2009).
A systematic review in 2015 showed moderate relief for a minority of people with
fibromyalgia. Milnacipran was associated with increased adverse events when
discontinuing use of the drug (Cording et al. 2015).
Levomilnacipran
A systematic review of all available clinical studies reported superiority over placebo
with a NNT for response vs. placebo of 9 and for remission 14. Improvement in
social functioning was also demonstrated (Citrome 2013). A review of seven RCTs
found Lm more effective than placebo in reducing depression as well as improving
functional impairment. In the placebo-controlled long-term study Lm was not
significantly superior to placebo (Asnis and Henderson 2015).
Common side effects of SNRIs are similar to SSRIs and mainly include gastroin-
testinal side-effects like nausea, loss of appetite, vomiting, diarrhea, and CNS
effects like trouble sleeping, shakiness, feeling tired, excessive yawning, severe
tinnitus, and sexual dysfunction (Clayton et al. 2002). Rates vary depending if
spontaneous reporting or direct questioning is applied.
Induction of suicidality by antidepressants is a point of frequent scientific and
medical discussion. Meta-analyses of randomized clinical trials (RCTs) indicate a
small increased risk for suicidal events in adolescents and young adults, but a
protective effect in older adults. In contrast, pharmacoepidemiologic studies show
a protective effect across the life span (Brent 2016).
Most commonly reported adverse effects of Venlafaxine are nausea, dizziness,
sweating, nervousness, and insomnia. It can cause abnormal bleeding, altered
platelet function, and in higher doses because of norepinephrine reuptake inhibition
tachycardia, tremor, and elevated blood pressure (Rudolph and Derivan 1996).
Serotonin syndromes have been reported rarely.
At lower doses the adverse effect profile of venlafaxine is similar to SSRIs, at
higher doses venlafaxine can produce adverse effects because of norepinephrine
reuptake inhibition: diaphoresis, tachycardia, tremors, anxiety, and mild but
sustained hypertension (Nelson 1997; Preskorn 1995). The elevated blood pressure
is a dose-dependent effect, doses of 300 mg per day or higher have a rate of 13%
(Feighner 1995; Thase 1998).
Serotonin-Norepinephrine Reuptake Inhibitors (SNRIs): Venlafaxine,. . . 1337
Desvenlafaxine
50 and 100 mg/day was generally safe and well tolerated (Carrasco et al. 2016).
Duloxetine
Milnacipran
The most frequently occurring adverse reactions (5% and greater than placebo)
were nausea, headache, constipation, dizziness, insomnia, hot flush, and hyperten-
sion. Effects on weight, QTc, or sexual functions are missing. A study with N ¼ 1871
pts. reported 7.9% dry mouth, 4.3% sweating, 2.1% dysuria (Lopez-Ibor et al. 1996).
In summary, side effects are comparable to SSRIS with fewer nausea but more
dysuria (Puech et al. 1997).
Levomilnacipran
The most commonly encountered AEs (incidence >5%) were nausea, hyperhidrosis,
tachycardia, and mild increase of blood pressure (Citrome 2013). Lm was weight
neutral, did not cause QTc prolongation, pulse and blood pressure were significantly
elevated, nausea and hyperhidrosis occurred in 17% vs. 6% under Pl and 9% vs. 2%
under Pl, respectively (Asnis and Henderson 2015).
Regarding safety of overdose SNRIs have low hazard indices compared to TCAs
and MAOIs (White et al. 2008). Acute overdosage is often manifested by emesis,
lethargy, ataxia, disturbances in heart rhythm, tremor, amnesia, confusion, coma, or
convulsions. The fatal toxicity index (deaths due to acute overdose) was reported to
be higher with venlafaxine than with SSRIs (Buckley and McManus 2002).
1338 G. Laux and A. Klimke
Pregnancy
Discontinuation Symptoms
Symptoms occur within a few days from drug discontinuation, are mild and last a
few weeks, variations are possible, however (Fava et al. 2018).
Typical symptoms can be summarized with the acronym FINISH: Flu-like,
Insomnia, Nausea, Imbalance, Sensory disturbance, Hyperarousal like nightmares,
vivid dreams, dizziness, feelings of electricity in the body (Gabriel and Sharma
2017). A review of 14 studies with more than 4000 patients reported discontinuation
symptoms in 27–86% (56% on average) of patients treated with antidepressants.
Nearly half of the patients reported severe symptoms (Davies and Read 2019). Data
regarding frequency, intensity, and duration vary however. Frequently intake is
terminated by patients’ own decision without symptoms.
People stopping venlafaxine commonly experience discontinuation symptoms
such as dysphoria, headaches, nausea, irritability, emotional lability, sensation of
electric shocks, and sleep disturbance. Venlafaxine has a higher rate of moderate to
severe discontinuation symptoms relative to other antidepressants.
SNRIs should not be taken with serotonergic agents like SSRIs, tramadol, fentanyl,
dextro-methorphane, pethidine, St John’s wort, tryptophan, or 5-HTP as the resulting
drug interaction could lead to serotonin syndrome. Concomitant use of serotonergic
Serotonin-Norepinephrine Reuptake Inhibitors (SNRIs): Venlafaxine,. . . 1339
drugs like triptans, tramadol, buspiron must be monitored, dose must be reduced by
half when strong inhibitors of CYP2D6 are coadministered.
SSRIs and SNRIs: Concomitant use of aspirin, nonsteroidal anti-inflammatory
drugs (NSAIDs), and anticoagulants can result in gastrointestinal and other bleeding
events.
Duloxetine: CYP1A2 inhibitors (Fluvoxamine, ciprofloxazine, enoxacine);
smoking.
Milnacipran: Clonidine, sympathomimetics (Spencer and Wilde 1998).
Levomilnacipran shows minimal drug–drug interactions.
Cross-References
References
Acuna C. Duloxetine for the treatment of fibromyalgia. Drugs Today. 2008;44:725–34.
Allgulander C, Hackett D, Salinas E. Venlafaxine extended release (ER) in the treatment of
generalized anxiety disorder: twenty-four-week placebo-controlled dose-ranging study. Br J
Psychiatry. 2001;179:15–22.
Ansseau M, von Frenckell R, Mertens C, de Wilde J, et al. Controlled comparison of two doses of
milnacipran (F 2207) and amitriptyline in major depressive inpatients. Psychopharmacology.
1989;98:163–8.
Ansseau M, Papart P, Troisfontaines B, Bartholome F, et al. Controlled comparison of milnacipran
and fluoxetine in major depression. Psychopharmacology. 1994;114:131–7.
Asnis GM, Bose A, Gommoll CP, Chen C, Greenberg WM. Efficacy and safety of levomilnacipran
sustained release 40 mg, 80 mg, or 120 mg in major depressive disorder: a phase 3, randomized,
double-blind, placebo-controlled study. J Clin Psychiatry. 2013;74:242–8.
Asnis GM, Henderson MA. Levomilnacipran for the treatment of major depressive disorder: a
review. Neuropsychiatr Dis Treat. 2015;11:125–35.
Ballus C, Quiros G, De Flores T, et al. The efficacy and tolerability of venlafaxine and paroxetine in
outpatients with depressive disorders or dysthymia. Int Clin Psychopharmacol. 2000;15:43–8.
Bérard A, Sheehy O, Zhao JP, Vinet É, et al. SSRI and SNRI use during pregnancy and the risk of
persistent pulmonary hypertension of the newborn. Br J Clin Pharmacol. 2017;83:1126–33.
Bielski RJ, Ventura D, Chang CC. A double-blind comparison of escitalopram and venlafaxine
extended release in the treatment of major depressive disorder. J Clin Psychiatry. 2004;65:
1190–6.
Brent DA. Antidepressants and suicidality. Psychiatr Clin North Am. 2016;39:503–12.
Buckley NA, McManus PR. Fatal toxicity of serotonergic and other antidepressant drugs: analysis
of United Kingdom mortality data. Br Med J. 2002;325:1332–3.
1340 G. Laux and A. Klimke
Bymaster FP, Dreshfield-Ahmad LJ, Threlkeld PG, Shaw JL. Comparative affinity of duloxetine
and venlafaxine for serotonin and norepinephrine transporters in vitro and in vivo, human
serotonin receptor subtypes, and other neuronal receptors. Neuropsychopharmacology.
2001;25:871–80.
Carrasco JL, Kornstein SG, McIntyre RS, et al. An integrated analysis of the efficacy and safety of
desvenlafaxine in the treatment of major depressive disorder. Int Clin Psychopharmacol.
2016;31:134–46.
Carter NJ, McCormack PL. Duloxetine: a review of its use in the treatment of generalized anxiety
disorder. CNS Drugs. 2009;23:523–41.
Cipriani A, Furukawa TA, Salanti G, Chaimani A, et al. Comparative efficacy and acceptability of
21 antidepressant drugs for the acute treatment of adults with major depressive disorder: a
systematic review and network meta-analysis. Lancet. 2018;391:1357–66.
Cipriani A, Koesters M, Furukawa TA, Nosè M, et al. Duloxetine versus other antidepressive agents
for depression. Cochrane Database Syst Rev. 2012;10:CD006533.
Citrome L. Levomilnacipran for major depressive disorder: a systematic review of the efficacy and
safety profile for this newly approved antidepressant—what is the number needed to treat, number
needed to harm and likelihood to be helped or harmed? Int J Clin Pract. 2013;67:1089–104.
Clayton AH, Pradko JF, Croft HA, et al. Prevalence of sexual dysfunction among newer antide-
pressants. J Clin Psychiatry. 2002;63:357–66.
Clerc G. Antidepressant efficacy and tolerability of milnacipran, a dual serotonin and noradrenaline
reuptake inhibitor: a comparison with fluvoxamine. Int Clin Psychopharmacol. 2001;16:145–51.
Clerc GE, Ruimy P, Verdeau-Palles J. A double-blind comparison of venlafaxine and fluoxetine in
patients hospitalized for major depression and melancholia. The Venlafaxine French Inpatient
Study Group. Int Clin Psychopharmacol. 1994;9:139–43.
Coleman KA, Xavier VY. An indirect comparison of the efficacy and safety of desvenlafaxine and
venlafaxine using placebo as the common comparator. CNS Spectr. 2012;17:131–41.
Cording M, Derry S, Phillips T, Moore RA, Wiffen PJ. Milnacipran for pain in fibromyalgia in
adults. Cochrane Database Syst Rev. 2015;10:CD008244.
Costa e Silva J. Randomized, double-blind comparison of venlafaxine and fluoxetine in outpatients
with major depression. J Clin Psychiatry. 1998;59:352–7.
Cunningham LA, Borison RL, Carman JS, Chouinard G, et al. A comparison of venlafaxine,
trazodone, and placebo in major depression. J Clin Psychopharmacol. 1994;14:99–106.
Cusack B, Nelson A, Richelson E. Binding of antidepressants to human brain receptors: focus on
newer generation compounds. Psychopharmacology. 1994;114:559–65.
Davidson JRT, Wittchen HU, Llorca PM, et al. Duloxetine treatment for relapse prevention in adults
with generalized anxiety disorder: a double-blind placebo-controlled trial. Eur Neuropsycho-
pharmacol. 2008;18:673–81.
Davies J, Read J. A systematic review into the incidence, severity and duration of antidepressant
withdrawal effects: are guidelines evidence-based? Addict behaviors. 2019;97:111–21.
Delini-Stula A. Milnacipran: an antidepressant with dual selectivity of action on noradrenaline and
serotonin uptake. Hum Psychopharmacol. 2000;15:255–60.
De Vries C, Gadzhanova S, Sykes MJ, Ward M, Roughead E. A systematic review and meta-
analysis considering the risk for congenital heart defects of antidepressant classes and individual
antidepressants. Drug Saf. 2021;44:291–312.
Dueñas H, Brnabic AJ, Lee A, Montejo AL. Treatment-emergent sexual dysfunction with SSRIs
and duloxetine: effectiveness and functional outcomes over a 6-month observational period. Int
J Psychiatry Clin Pract. 2011;15:242–54.
Eckert L, Lançon C. Duloxetine compared with fluoxetine and venlafaxine: use of meta-regression
analysis for indirect comparisons. BMC Psychiatry. 2006;6:30.
Einarson TR, Arikian SR, Casciano J, Doyle JJ. Comparison of extended-release venlafaxine,
selective serotonin reuptake inhibitors, and tricyclic antidepressants in the treatment of depres-
sion: a meta-analysis of randomized controlled trials. Clin Ther. 1999;21:296–308.
Serotonin-Norepinephrine Reuptake Inhibitors (SNRIs): Venlafaxine,. . . 1341
Knadler MP, Lobo E, Chappell J, Bergstrom R. Duloxetine: clinical pharmacokinetics and drug
interactions. Clin Pharmacokinetics. 2011;50:281–94.
Kocsis JH, Thase ME, Trivedi MH, Shelton RC, et al. Prevention of recurrent episodes of
depression with venlafaxine ER in a 1-year maintenance phase from the PREVENT study.
J Clin Psychiatry. 2007;68:1014–23.
Kornstein SG, Kocsis JH, Ahmed S et al. Assessing the efficacy of 2 years of maintenance treatment
with venlafaxine extended release in patients with recurrent major depression: a secondary
analysis of data from the prevent study. Int Clin Psychopharmacol 2008;23:357–63.
Laoutidis ZG, Kioulos KT. Desvenlafaxine for the acute treatment of depression: a systematic
review and meta-analysis. Pharmacopsychiatry. 2015;48:187–99.
Lecrubier Y. Milnacipran: the clinical properties of a selective serotonin and noradrenaline reuptake
inhibitor (SNRI). Hum Psychopharmacol. 1997;12:127–34.
Leinonen E, Leopola U, Mehtonen H, Rimon O-P. Long-term efficacy and safety of milnacipran
compared to clomipramine in patients with major depression. Acta Psychiatr Scand. 1997;96:
497–504.
Liebowitz MR, Gelenberg AJ, Munjack D. Venlafaxine extended release vs placebo and paroxetine
in social anxiety disorder. Arch Gen Psychiatry. 2005;62:190–8.
Lopez-Ibor J, Guelfi JD, Pletan Y, Tournoux A, Prost JF. Milnacipran and selective serotonin
reuptake inhibitors in major depression. Int Clin Psychopharmacol. 1996;11(Suppl 4):41–6.
Lunn MP, Hughes RA, Wiffen PJ. Duloxetine for treating painful neuropathy, chronic pain or
fibromyalgia. Cochrane Database Syst Rev. 2014;1:CD007115.
Macher JP, Sichel JP, Serre C, von Frenckell R, Huck JC, Demares JP. Double-blind placebo-
controlled study of milnacipran in hospitalized patients with major depressive disorders.
Neuropsychobiology. 1989;22:77–82.
Magni LR, Purgato M, Gastaldon C, Papola D, et al. Fluoxetine versus other types of pharmaco-
therapy for depression. Cochrane Database Syst Rev. 2013;17:CD004185.
Mehtonen OP, Sogaard J, Roponen P, Behnke K. Randomized, double-blind comparison of
venlafaxine and sertraline in outpatients with major depressive disorder. Venlafaxine
631 Study Group. J Clin Psychiatry. 2000;61:95–100.
Montgomery SA, Entsuah R, Hackett D, Kunz NR, Rudolph RL. Venlafaxine versus placebo in the
preventive treatment of recurrent major depression. J Clin Psychiatry. 2004;65:328–36.
Montgomery SA, Mansuy L, Ruth A, Bose A, et al. Efficacy and safety of levomilnacipran
sustained release in moderate to severe major depressive disorder: a randomized, double-
blind, placebo-controlled, proof-of-concept study. J Clin Psychiatry. 2013;74:363–9.
Nakagawa A, Watanabe N, Omori IM, Barbui C, et al. Efficacy and tolerability of milnacipran in the
treatment of major depression in comparison with other antidepressants: a systematic review and
meta-analysis. CNS Drugs. 2008;22:587–602.
Nakagawa A, Watanabe N, Omori IM, Barbui C, et al. Milnacipran versus other antidepressive
agents for depression. Cochrane Database Syst Rev. 2009;8:CD006529.
Nelson JC. Safety and tolerability of the new antidepressants. J Clin Psychiatry. 1997;58(6):26–31.
Nelson JC, Lu Pritchett Y, Martynov O, Yu JY, Mallinckrodt CH, Detke MJ. The safety and
tolerability of duloxetine compared with paroxetine and placebo: a pooled analysis of 4 clinical
trials. Prim Care Companion J Clin Psychiatry. 2006;8:212–9.
Nelson JC, Wohlreich MM, Mallinckrodt CH, Detke MJ, et al. Duloxetine for the treatment of
major depressive disorder in older patients. Am J Geriatr Psychiatry. 2005;13:227–35.
Nemeroff CB, Entsuah R, Benattia I, et al. Comprehensive analysis of remission (COMPARE) with
venlafaxine versus SSRIs. Biol Psychiatry. 2008;63:424–34.
Nierenberg AA, Feighner JP, Rudolph R, Cole J, Sullivan J. Venlafaxine for treatment-resistant
depression. J Clin Psychopharmacol. 1994;14:419–23.
Oakes TM, Katona C, Liu P, Robinson M, et al. Safety and tolerability of duloxetine in elderly
patients with major depressive disorder: a pooled analysis of two placebo-controlled studies. Int
Clin Psychopharmacol. 2013;28:1–11.
Serotonin-Norepinephrine Reuptake Inhibitors (SNRIs): Venlafaxine,. . . 1343
Olie JP, Gourion D, Montagne A, Rostin M, Poirier MF. Milnacipran and venlafaxine at flexible
doses (up to 200 mg/day) in the outpatient treatment of adults with moderate-to-severe major
depressive disorder: a 24-week randomized, double-blind exploratory study. Neuropsychiatr Dis
Treat. 2010;6:71–9.
Papakostas GI, Fava M. A meta-analysis of clinical trials comparing milnacipran, a serotonin—
norepinephrine reuptake inhibitor, with a selective serotonin reuptake inhibitor for the treatment
of major depressive disorder. Eur Neuropsychopharmacol. 2007;17:32–6.
Papakostas GI, Thase ME, Fava M, Nelson JC, Shelton RC. Are antidepressant drugs that combine
serotonergic and noradrenergic mechanisms of action more effective than the selective serotonin
reuptake inhibitors in treating major depressive disorder? A meta-analysis of studies of newer
agents. Biol Psychiatry. 2007;62:1217–27.
Perahia DGS, Wang F, Mallinckrodt CH, Walker DJ, Detke MJ. Duloxetine in the treatment of
major depressive disorder: a placebo- and paroxetine-controlled trial. Eur Psychiatry. 2006a;21:
367–78.
Perahia DGS, Pritchett YL, Desaiah D, Raskin J. Efficacy of duloxetine in painful symptoms: an
analgesic or antidepressant effect? Int Clin Psychopharmacol 2006b;21:311–7.
Perahia DGS, Maina G, Thase ME, et al. Duloxetine in the prevention of depressive recurrences: a
randomized, double-blind, placebo-controlled trial. J Clin Psychiatry. 2009;70:706–16.
Poirier MF, Boyer F. Venlafaxine and paroxetine in treatment-resistant depression: double-blind,
randomized comparison. Br J Psychiatry. 1999;175:12–6.
Pollack MH, Worthington JJ, Otto MW, et al. Venlafaxine for panic disorder: results from a double-
blind, placebo-controlled study. Psychopharmacol Bull. 1996;32:667–70.
Preskorn SH. Comparison of the tolerability of bupropion, fluoxetine, imipramine, nefazodone,
paroxetine, sertraline, and venlafaxine. J Clin Psychiatry. 1995;56:12–21.
Puech A, Montgomery SA, Prost JF, Solles A, Briley M. Milnacipran, a new serotonin and
norepinephrine reuptake inhibitor: an overview of its antidepressant activity and clinical toler-
ability. Int Clin Psychopharmacol. 1997;12:99–108.
Puozzo C, Panconi E, Deprez D. Pharmacology and pharmacokinetics of milnacipran. Int Clin
Psychopharmacology. 2002;17(Suppl 1):S25–35.
Richelson E. The clinical relevance of antidepressant interaction with neurotransmitter transporters
and receptors. Psychopharmacol Bull. 2002;36:133–50.
Rickels K, Pollack MH, Sheehan DV, Haskins JT. Efficacy of extended-release venlafaxine in
nondepressed outpatients with generalized anxiety disorder. Am J Psychiatry. 2000;157:
968–74.
Robinson M, Oakes TM, Raskin J, et al. Acute and long-term treatment of late-life major depressive
disorder: duloxetine versus placebo. Am J Geriatr Psychiatry. 2014;22:34–45.
Rouillon F, Warner B, Pezous N, Bisserbe JC. Milnacipran efficacy in the prevention of recurrent
depression: a 12-month placebo-controlled study. Int Clin Psychopharmacol. 2000;15:133–40.
Rudolph RL, Derivan AT. The safety and tolerability of venlafaxine hydrochloride: analysis of the
clinical trials database. J Clin Psychopharmacol. 1996;16(suppl 2):54S–9S.
Rudolph RL, Feiger AD. A double-blind, randomized, placebo-controlled trial of once-daily
venlafaxine extended release (XR) and fluoxetine for the treatment of depression. J Affect
Disord. 1999;56:171–81.
Sambunaris A, Bose A, Gommoll CP, Chen C, et al. A phase III, double-blind, placebo-controlled,
flexible-dose study of levomilnacipran extended-release in patients with major depressive
disorder. J Clin Psychopharmacol. 2014;34:47–56.
Samuelian JC, Hackett D. A randomized, double-blind, parallel-group comparison of venlafaxine
and clomipramine in outpatients with major depression. J Psychopharmacol. 1998;12:273–8.
Sansone RA, Sansone LA. Serotonin norepinephrine reuptake inhibitors: a pharmacological com-
parison. Innov Clin Neurosci. 2014;11:37–42.
Sauer H, Huppertz-Helmhold S, Dierkes W. Efficacy and safety of venlafaxine ER vs. amitriptyline
ER in patients with major depression of moderate severity. Pharmacopsychiatry. 2003;36:169–75.
1344 G. Laux and A. Klimke
Schueler YB, Koesters M, Wieseler B, Grouven U, et al. A systematic review of duloxetine and
venlafaxine in major depression, including unpublished data. Acta Psychiatr Scand. 2011;123:
247–65.
Schweizer E, Feighner J, Mandos LA, Rickels K. Comparison of venlafaxine and imipramine in the
acute treatment of major depression in outpatients. J Clin Psychiatry. 1994;55:104–8.
Schweizer E, Weise C, Clary C, Fox I, Rickels K. Placebo-controlled trial of venlafaxine for the
treatment of major depression. J Clin Psychopharmacol. 1991;11:233–6.
Sechter D, Vandel P, Weiller E, Pezous N, Cabanac F, Tournoux A. A comparative study of
milnacipran and paroxetine in outpatients with major depression. J Affect Disord. 2004;83:
233–6.
Shelton RC. Serotonin norepinephrine reuptake inhibitors. In: Macaluso M, Preskorn SH, editors.
Antidepressants. Handbook of experimental pharmacology, vol. 250. Springer Nature; 2018.
Smith D, Dempster C, Glanville J, Freemantle N, Anderson I. Efficacy and tolerability of
venlafaxine compared with selective serotonin reuptake inhibitors and other antidepressants: a
meta-analysis. Br J Psychiatry. 2002;180:396–404.
Spencer CM, Wilde MI. Milnacipran. A review of its use in depression. Drugs. 1998;56:405–27.
Stahl S. Essential psychopharmacology. The prescriber’s guide. Antidepressants. 6th
ed. Cambridge: Cambridge University Press; 2017.
Stahl SM, Entsuah R, Rudolph RL. Comparative efficacy between venlafaxine and SSRIs: a pooled
analysis of patients with depression. Biol Psychiatry. 2002;52:1166–74.
Steen A, Den Bour JA. A double-blind, six-month comparative study of milnacipran and clomip-
ramine in major depressive disorder. Int Clin Pharmacol. 1997;12:269–81.
Sultan A, Gaskell H, Derry S, Moore RA. Duloxetine for painful diabetic neuropathy and
fibromyalgia pain: systematic review of randomised trials. BMC Neurol. 2008;8:29.
Tatsumi M, Groshan K, Blakely RD, Richelson E. Pharmacological profile of antidepressants and
related compounds at human monoamine transporters. Eur J Pharmacol. 1997;340:249–58.
Thase ME. Efficacy and tolerability of once-daily venlafaxine extended release (XR) in outpatients
with major depression. The Venlafaxine XR 209 Study Group. J Clin Psychiatry. 1997;58:
393–8.
Thase ME. Effects of venlafaxine on blood pressure: a meta-analysis of original data from 3744
depressed patients. J Clin Psychiatry. 1998;59:502–8.
Thase ME, Entsuah AR, Rudolph RL. Remission rates during treatment with venlafaxine or
selective serotonin reuptake inhibitors. Br J Psychiatry. 2001;178:234–41.
Thase ME, Clayton AH, Haight BR, Thompson AH, et al. A double-blind comparison between
bupropion XL and venlafaxine XR: sexual functioning, antidepressant efficacy, and tolerability.
J Clin Psychopharmacol. 2006;26:482–8.
Thase ME, Kornstein SG, Germain JM, Jiang Q, et al. An integrated analysis of the efficacy of
desvenlafaxine compared with placebo in patients with major depressive disorder. CNS Spectr.
2009;14:144–54.
Thase ME, Gelenberg A, Kornstein SG, Kocsis JH, et al. Comparing venlafaxine extended release
and fluoxetine for preventing the recurrence of major depression: results from the PREVENT
study. J Psychiatr Res. 2011;45:412–20.
Thase ME, Asami Y, Wajsbrot D, Dorries K, et al. A meta-analysis of the efficacy of venlafaxine
extended release 75–225 mg/day for the treatment of major depressive disorder. Curr Med Res
Opin. 2016;33:317–26.
Tignol J, Pujol-Domenech J, Chartres JP, Leger JM, et al. Double-blind study of the efficacy and
safety of milnacipran and imipramine in elderly patients with major depressive episode. Acta
Psychiatr Scand. 1998;97:157–65.
Tzanakaki M, Guazelli M, Nimatoudis I, et al. Increased remission rates with venlafaxine compared
with fluoxetine in hospitalized patients with major depression and melancholia. Int Clin
Psychopharmacol. 2000;15:29–34.
Serotonin-Norepinephrine Reuptake Inhibitors (SNRIs): Venlafaxine,. . . 1345
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1348
Clinical Applications of MAOIs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1349
Pharmacokinetics/Metabolism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1351
Adverse Effects of MAOIs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1364
Drug-Drug Interactions with MAOIs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1366
Use of Drugs in Combination with MAOIs to Treat Depression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1367
Pharmacological and Neurochemical Actions of MAOIs in Addition to Inhibition
of MAO . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1368
Target Matching of MAOIs in Major Depressive Disorder (MDD) . . . . . . . . . . . . . . . . . . . . . . . 1369
Revitalization of Interest in MAOIs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1370
Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1370
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1371
Abstract
Monoamine oxidase inhibitors (MAOIs) first came into use as antidepressants in the
1950s and 1960s but were soon replaced as first-line drugs by tricyclic antidepressants
and then later by drugs like selective serotonin reuptake inhibitors (SSRIs). However,
there is a strong feeling among many prominent health professionals that the MAOIs
are effective antidepressants (and anxiolytics in some cases) that are underutilized
because of concerns about potential adverse effects, including food-drug and drug-
drug interactions, even though these are often avoidable and/or manageable. In this
J. H. Meyer · A. Santhirakumar
Centre for Addiction and Mental Health (CAMH) and Department of Psychiatry, University of
Toronto, Toronto, ON, Canada
e-mail: jeffrey.meyer@utoronto.ca; apitharani.santhirakumar@mail.utoronto.ca;
apitharani.santhirakumar@camh.ca
D. Matveychuk · A. Holt · G. B. Baker (*)
Neurochemical Research Unit, Department of Psychiatry, University of Alberta, Edmonton, AB,
Canada
e-mail: dmitriym@ualberta.ca; aholt@ualberta.ca; glen.baker@ualberta.ca
Introduction
Although MAOIs are no longer first-line drugs, there is general agreement that they
should definitely be considered for use in atypical depression (characterized by
pathological sensitivity to interpersonal loss or rejection, prominent anergia and
increased appetite, weight or sleep but lacking catatonic or melancholic features)
(Quitkin et al. 1993; Henkel et al. 2006; Ricken et al. 2017; Ulrich et al. 2017) and in
treatment-refractory depression (McGrath et al. 1987; Birkenhager et al. 2004;
Fiedorowicz and Swartz 2004; Amsterdam and Shults 2005: Kim et al. 2019).
However, there are numerous reports indicating that MAOIs are also effective in
the treatment of melancholic depression (endogenous depression) that is not
treatment-refractory (McGrath et al. 1984; Davidson et al. 1988; Gillman et al.
2020). In a systematic review comparing tranylcypromine and tricyclic
1350 J. H. Meyer et al.
Table 1 is a summary table that lists several important references that provide
information on many of the clinical studies that have been done on the use of MAOIs
in depression. Tables 2, 3, 4 and 5 list specific clinical studies for the most commonly
used MAOIs, i.e. phenelzine, tranylcypromine, moclobemide and EMSAM.
There are also reports in the literature on the effectiveness of MAOIs in treating
panic disorder, social anxiety disorder and post-traumatic stress disorder (Sheehan
et al. 1980; Stein et al. 2002; Versiani et al. 1996; Riederer et al. 2004; Zhang and
Davidson 2007; Baldwin et al. 2014; Cipriani et al. 2018; Williams et al. 2020;
Chamberlain et al. 2020).
Pharmacokinetics/Metabolism
Table 1 Selected examples of important papers discussing the use of MAOIs in depression
White and Simpson 1981: Review of the literature on MAOI TCA combinations
Davidson et al. 1988: RCT on efficacy of isocarboxazid
Quitkin et al. 1993: Review of several double-blind studies comparing efficacy of phenelzine and
imipramine in atypical depression
Thase et al. 1995: Comprehensive review and meta-analysis of effectiveness of phenelzine,
tranylcypromine and isocarboxazid compared to TCAs
De Lima and Hotopf 2003: Systematic review of RCTs comparing two or more active drug
treatments for dysthymia (now called persistent depressive disorder). Comparisons of MAOIs and
TCAs were included
Bonnet 2003. Overview of clinical trials with moclobemide and its interactions with CYP
enzymes
Fiedorowicz and Swartz 2004: Review of the use of MAOIs in atypical depression, bipolar
depression and TRD
Riederer et al. 2004: Review of clinical applications of MAOIs which includes detailed tables on
RCTs with tranylcypromine and moclobemide in treatment of depressive disorders
Amsterdam and Shults 2005: Retrospective study on efficacy and safety of MAOIs in early and
advanced TRD
Henkel et al. 2006: Meta-analysis of evidence regarding treatment of atypical depression. MAOIs
(phenelzine and moclobemide) were compared to TCAs and SSRIs
Mallinger et al. 2009. Retrospective study on use of MAOIs and the SSRI paroxetine in bipolar
depression
Shulman et al. 2013: Review of the historical background and use of the MAOIs (phenelzine,
isocarboxazid, tranylcypromine, moclobemide and selegiline) in treatment of depression
Thomas et al. 2015: Literature review and retrospective study on combining MAOIs (phenelzine,
tranylcypromine, isocarboxazid, selegiline) with other antidepressants or stimulants for
management of TRD
Entzeroth and Ratty 2017: Review of the following aspects of MAOIs; historical background,
pharmacology, potential adverse effects and potential use in a number of disorders, with a focus on
moclobemide in the section on depressive disorders
Ricken et al. 2017: Detailed review and meta-analysis containing extensive text and tables on the
studies of the use of tranylcypromine to treat subtypes of depression
Chockalingam et al. 2018: Literature review on MAOIs covering historical background,
comparison of effectiveness of MAOIs and TCAs in depressive disorders and the
underprescribing of MAOIs and TCAs
Amsterdam and Kim 2019: Retrospective study on effectiveness of MAOI-TCA combination
therapy for TRD
Gillman 2019: An overview of MAOI and TCA interactions which includes a detailed list of
references to studies on the use and safety of MAOI-TCA combinations for treating depressive
disorders
Kim et al. 2019: Retrospective review comparing effectiveness of MAOIs versus TCAs as
monotherapy for early stage and advanced TRD
Ulrich et al. 2020: Meta-analysis comparing tranylcypromine and TCAs in the treatment of
depression
Abbreviations: MAOI monoamine oxidase inhibitor, RCT randomized controlled trial, SSRI selec-
tive serotonin reuptake inhibitor, TRD treatment resistant depression
Table 2 Acute phase trials of phenelzine (PHZ) in depression
Methods
Diagnostic system, duration Randomized (n)/
Author diagnostic method (weeks) RX cells (dosages) completers (n) Comments
Agnew et al. DSM-I, CLIN Random, DB PHZ (45 mg/d) 4/4 Inpatients; mixed depressive diagnoses; PHZ ¼
(1961) (3 weeks) ISO 6/6 IMI > PBO
IMI (75 mg/d) 6/6
PBO 5/5
Rees and “Diagnostic Random, DB PHZ (90 mg/d) 20/20 Inpatients; PHZ > PBO
Davies (1961) Criteria of Royal (3 weeks) PBO 21/20
Bethlehem Hospital”
Leitch and CLIN Random, DB PHZ (45 mg/d) 24/22 Inpatients; endogenous depression; PHZ ¼ IMI
Seager (1963) (4 weeks) IMI (150 mg/d) 26/25
Martin (1963) No systematic Random, DB PHZ (45–60 mg/d) NR/47 In- (n ¼ 79) and outpatients (n ¼ 16); endogenous
criteria reported, (4 weeks) IMI (150–200 NR/49 depression;
CLIN mg/d) IMI PHZ
Glick (1964) Depression Random, DB PHZ (x ¼ 55 mg/d) NR/4 Outpatients;
Monoamine Oxidase Inhibitors in Depressive Disorders
(continued)
Table 2 (continued)
1354
Methods
Diagnostic system, duration Randomized (n)/
Author diagnostic method (weeks) RX cells (dosages) completers (n) Comments
Schildkraut DSM-I, CLIN Random PHZ (45–60 6/6 Inpatients; PHZ ¼ IMI > PBO
et al. (1964) (3 weeks) mg/d) 6/6
IMI (100–200 5/5
mg/d)
PBO
Brit. Med. Res. No systematic Random PHZ (60 mg/d) 65/50 Inpatients; predominantly endogenous;
Coun. (1965) criteria reported, (Rx only), IMI (200 mg/d) 65/58 PHZ ¼ PBO < IMI < ECT
CLIN DB PBO 65/51
(4 weeks) ECT (4–8 65/58
treatments)
Kay et al. Newcastle Random, DB PHZ (45 mg/d) 31/27 Outpatients; non-endogenous depression;
(1973) Rating Scale, (4 weeks) AMI (150 mg/d) 31/18 PHZ ¼ IMI (intention to threat); PHZ AMI
CLIN completers only
Robinson et al. Structured Random, DB PHZ (x ¼ 58.5 44/33 Outpatients with
(1973) Interview (6 weeks) mg/d) 43/27 nonendogenous depression;
Depression, DSM-II PBO PHZ > PBO
Raskin et al. DSM-II, CLIN Random, DB PHZ (x ¼ 45.5 110/78 Inpatients; mixed
(1974) (4 weeks) mg/d) 111/81 diagnoses; 9 site
PBO collaborative study; PHZ ¼ PBO
Ravaris et al. Structured Random, DB PHZ (total) 41/30 Outpatients; predominantly nonendogenous;
(1976) Diagnostic (6 weeks) 60 mg/d 20/14 PHZ (60 mg) > PBO ¼ PHZ (30 mg)
Interview, 30 mg/d 21/16
CLIN AMI (150 mg/d) 21/19
Davidson et al. Criteria of Random, DB PHZ (90 mg/d) 4/4 Inpatients;
(1977) Feighner et al., CLIN (3 weeks) IMI (150 mg/d) 6/6 PHZ ¼ IMI
Ravaris et al. SDI, RDC Random, DB PHZ (60 mg/d) 68/55 Outpatients;
(1980) (6 weeks) AMI (150 mg/d) 61/49 predominantly nonendogenous; PHZ ¼ AMI
J. H. Meyer et al.
Hamilton Criteria of Random (Rx PHZ (45–90 mg/d) NR/65 Mixed in- and
(1982) Feighner et al., CLIN only), open IMI (150–225 NR/65 outpatients (majority of Rx cases are outpatients);
label mg/d) NR/146 melancholia; newly referred depressed;
ECT (6–12 PHZ IMI < ECT
treatments)
Rowan et al. RDC, Random, DB PHZ (x ¼ 75 mg/d) 58/42 Outpatients; non-endogenous;
(1982) Newcastle Rating (6 weeks) AMI (x ¼ 62/44 PHZ ¼ AMI > PBO
Scale, 188 mg/d) 56/45
CLIN PBO
Kayser et al. RDC, SDI Random, DB PHZ (60 mg/d) NR/23 Outpatients; PHZ ¼ AMI; In “hysteroid
(1985) (6 weeks) AMI (150 mg/d) NR/24 dysphoria,” PHZ (9/9) > AMI (3/5)
Georgotas et al. RDC, CLIN Random, DB PHZ (x ¼ 54 mg/d) 30/20 Outpatients; age
(1986) (6 weeks) NTP (x ¼ 79 mg/d) 30/23 55–75; PHZ ¼ NTP > PBO
PBO 30/19
Kayser et al. DSM-III, SDI Random, DB PHZ (60 mg/d) NR/12 Outpatients; DSM-III melancholia;
(1988) (6 weeks) AMI (150 mg/d) NR/12 PHZ ¼ AMI
Liebowitz et al. RDC, CLIN Random, DB PHZ (x ¼ 73 mg/d) 56/34 Outpatients;
(1988) (6 weeks) IMI (x ¼ 52/38 atypical major and minor depression;
Monoamine Oxidase Inhibitors in Depressive Disorders
Table 2 (continued)
Methods
Diagnostic system, duration Randomized (n)/
Author diagnostic method (weeks) RX cells (dosages) completers (n) Comments
Quitkin et al.b RDC, CLIN Random, DB PHZ (x ¼ 73 mg/d) 33/30 Outpatients;
(1990) (6 weeks) IMI (x ¼ 37/34 atypical major and minor depression;
270 mg/d) 34/26 PHZ > IMI PBO
PBO
Quitkin et al.b RDC, CLIN Random, DB PHZ (x ¼ 69 mg/d) 43/35 Outpatients;
(1991) (6 weeks) IMI (x ¼ 37/29 atypical major and minor depression patients who
276 mg/d) failed 6 weeks of placebo treatment;
PHZ > IMI
McGrath et al. RDC, CLIN Random, DB PHZ (x ¼ 56/45 Outpatients; probable or definite atypical major
(1993) (6 weeks) 75 mg/dc) 33/22 and minor depression patients who failed
IMI (x ¼ treatment with the other compound;
274 mg/dc) PHZ > IMI
Table adapted from Thase et al. (1995); copyright clearance from Springer Nature was obtained
Abbreviations: DSM Diagnostic and Statistical Manual of Mental Disorders, RDC Research Diagnostic Criteria, SDI Structured Depression Interview, DB
double blind, CLIN clinical assessment, PHZ phenelzine, ISO isocarboxazid, IMI imipramine, PBO placebo, TCP tranylcypromine, ECT electroconvulsive
therapy, AMI amitriptyline, NTP nortriptyline, NR not reported
a
Marked improvement only as employed in this study (the moderate improvement rating does not appear to correspond to the contemporary use of CGI score
of 2)
b
Patients completing >4 weeks are considered completers
c
Dosages reported only for nonresponders
J. H. Meyer et al.
Table 3 Acute phase trials of tranylcypromine (TCP) in depression
Diagnostic system, Methods duration Randomized (n)/
Author diagnostic method (weeks) RX cells (dosages) completers (n) Comments
Bartholomew No systematic Random, DB TCP (x ¼ 43 mg/d) 51/42 Outpatients;
(1962) diagnostic criteria, (6 weeks) PBO 51/49 TCP > PBO
CLIN
Gottfries No systematic Random, DB TCP (30 mg/d) NR/25 Inpatients;
(1963) method, CLIN (15 days) PBO NR/25 TCP ¼ PBO
Glick (1964) Depression Random, DB TCP (37 mg/d) NR/4 Outpatients;
Rating Scale and (14 weeks) PHZ NR/6 sample too small to ascertain significance
Global Rating, PBO NR/6
CLIN
Richmond and No systematic Random, DB TCP (40 mg/d) NR/20 Outpatients;
Roberts (1964) criteria, CLIN (3 weeks) ISO 38/20 TCP > AMI + IMI
AMI (150 mg/d) NR/20
IMI (225 mg/d) NR/20
Monoamine Oxidase Inhibitors in Depressive Disorders
Spear et al. No systematic Random, DB TCP (30 mg/d) 37/34 Inpatients and
(1964) criteria, CLIN (3 weeks) IMI (150 mg/d) 41/36 outpatients (proportion not specified);
TCP ¼ IMI
Himmelhoch RDC, SADS Random, DB TCP (40 mg/d) 28/22 Outpatients;
et al. (1982) (6 weeksa) PBO 31/17 predominantly bipolar; anergic
depression;
TCP > PBO
Razani et al. DSM-III, CLIN Random, DB TCP (40 mg/d) 25/21 Inpatients (57%)
(1983) (4 weeks) AMI (293 mg/d) 28/20 and outpatients (43%);
TCP ¼ AMI
White et al. RDC, CLIN Random, DB TCP (x ¼ 44 mg/d) 63/37 Outpatients;
(1984) (4 weeks) NTP (109 mg/d) 61/40 TCP ¼ NTP PBO;
PBO 59/45 TCP > PBO
1357
(continued)
Table 3 (continued)
1358
sensorimotor performance);
psychomotor performance: MOC <
MAP, in nonresponders
Dierick et al. No specified diagnostic Random, DB MOC (300–600 mg/d) 32/29 Patient type not specified; MOC =
(1990) criteria (4 weeks) CLO (75–150 mg/d) 31/22 CLO; difficulty interpreting results
due to small numbers and
concomitant medication
Laux et al. No specified diagnostic Random, DB MOC (x̄ = 327 mg/d) 20a/15 Inpatients; predominantly
(1990) criteria (4 weeks) MAP (x̄ = 169 mg/d) 20a/20b endogenous depression; MOC =
MAP
(continued)
1359
1360
Table 4 (continued)
Methods
Diagnostic system, diagnostic duration Randomized (n)/
Author method (weeks) RX cells (dosages) completers (n) Comments
Lecrubier and DSM-III, CLIN Random, DB MOC (300–600 mg/d) 164/146c Outpatients; multicenter study; 50%
Guelfi (1990) (6 weeks) IMI (100–200 mg/d) 164/145c of patients endogenously depressed;
PBO 162/120c MOC = IMI > PBO
Larsen et al. Newcastle II Scale, HAM-D, Random, DB MOC (300–600 mg/d) 59/41 Outpatients; atypical, endogenous,
(1991) QSAD, DSM-III, CLIN (6 weeks) CLO (150–200 mg/d) 57/39 and reactive depression; MOC < ISO
ISO (30–40 mg/d) 51/39 < CLO
Bakish et al. DSM-III-R, HAM-D, CLIN Random, DB MOC (200–600 mg/d) 58/40 Outpatients; multi center study; MOC
(1992) (6 weeks) AMI (50–150 mg/d) 59/39 = AMI > PBO
PBO 56/28
Evans et al. DSM-III, HAM-D, CLIN Random, DB MOC (x̄ = 325 mg/d) 24/19 In- (n = 4) and outpatients (n = 44);
(1992) (4 weeks) AMI (x̄ = 123 mg/d) 24/17 multicenter study; MOC = AMI
Guelfi et al. Newcastle Rating Scale, ICD- Random, DB MOC (300–600 mg/d) 62/47 Inpatients; endogenous depression;
(1992) 9, MADRS, DSM-III, CLIN (6 weeks) CLO (100–200 mg/d) 67/56 MOC = CLO
DUAG (1993) DUAG diagnostic depression Random, DB MOC (400 mg/d) 57/51d Inpatients; MOC < CLO
scale, HAM-D, CLIN (6 weeks) CLO (150 mg/d) 58/56d
Philipp et al. Research diagnostic criteria, Random, DB MOC (400–500 mg/d) 118/94e Outpatients; MOC = DOX; MOC
(1993) HAM-D, CLIN (6 weeks) DOX (100–125 mg/d) 119/89e better tolerated
Rimon et al. DSM-III, HAM-D, CLIN Random, DB MOC (150–525 mg/d) 62/55 Inpatients and outpatients (proportion
(1993) (4 weeks) IMI (50–175 mg/d) 65/58 not specified); MOC = IMI
Williams et al. DSM-III, HAM-D, CLIN Random, DB MOC (300–600 mg/d) 62/49 In- (21% MOC, 33% FLU) and
(1993) (6 weeks) FLU (20–40 mg/d) 60/43 outpatient (79% MOC, 67% FLU);
MOC = FLU
J. H. Meyer et al.
UK DSM-III, HAM-D, CLIN Random, DB MOC (450 mg/d) NR/56 Patient type not specified; multi-
Moclobemide (6 weeks) IMI (150 mg/d) NR/50 center study; MOC = IMI = PBO
Study Group PBO NR/54
(1994)
Gachoud et al. DSM-III, HAM-D, CLIN Random, DB, MOC (300–400 mg/d) 66/54 Outpatients; MOC = MAP
(1994) (4 weeks) MAP (75–100 mg/d) 64/55
Lonnqvist et al. DSM-III-R, HAM-D, CLIN Random, DB MOC (300–450 mg/d) 102/84 Inpatients and outpatients (84%
(1994) (6 weeks) FLU (20–40 mg/d) 107/85 MOC, 79% FLU); MOC = FLU
Kragh-Sorensen DSM-III-R, HAM-D, CLIN Random, DB MOC (400 mg/d) 48/39a,f Patient type not specified; MOC =
et al. (1995) (6 weeks) CLO (150 mg/d) 48/30a,f CLO
Lecrubier et al. DSM-III, HAM-D, CLIN Random, DB MOC (400–600 mg/d) 98/68a Outpatients; nonmelancholic,
(1995) (12 weeks) CLO (100–150 mg/d) 93/62a nonpsychotic major depression MOC
= CLO
Nair et al. (1995) DSM-III-R, CLIN Random, DB MOC (400 mg/d) 36/15a Inpatients and outpatients; ages > 60;
(7 weeks) NOR (75 mg/d) 38/20a 3 inter-national sites; MOC PBO;
PBO 35/20a NOR > PBO
Reynaert et al. DSM-III-R, HAM-D, CLIN Random, DB MOC (300–600 mg/d) 51/38 In- (n = 65) and outpatients (n = 34);
(1995) (6 weeks) FLU (20–40 mg/d) 50/42 MOC = FLU
Monoamine Oxidase Inhibitors in Depressive Disorders
Lapierre et al. DSM-III-R, HAM-D, CLIN Random, DB MOC (x̄ = 440 123 66/53 Outpatients; MOC = FLU
(1997) (6 weeks) mg/d) 62/54
FLU (x̄ = 35 8 mg/
d)
Jouvent et al. MADRS, DSM-III-R, Random, DB MOC (450 mg/d) 65/62a Inpatients; MOC CLO
(1998) Abrams-Taylor Scale, ERD, (4 weeks) CLO (150 mg/d) 59/54a
CLIN
Sogaard et al. Random, DB MOC (300–450 mg/d) 97/84 Outpatients; atypical depression;
(1999) (12 weeks) SER (50–100 mg/d) 100/83 MOC SER
(continued)
1361
1362
Table 4 (continued)
Methods
Diagnostic system, diagnostic duration Randomized (n)/
Author method (weeks) RX cells (dosages) completers (n) Comments
Atypical Depression
Diagnostic Scale, HAM-D,
DSM-III-R, CLIN
Table adapted from Riederer et al. (2004)
Abbreviations: DSM = Diagnostic and Statistical Manual of Mental Disorders; HAM-D = Hamilton Depression Rating Scale; QSAD = Quantitative Scale for
Atypical Depression; ICD = International Classification of Diseases; MADRS = Montgomery-Asberg Depression Rating Scale; RDC = Research Diagnostic
Criteria; DUAG = Danish University Antidepressant Group; ERD = échelle de ralentissement dépressif; DB = double blind; CLIN = clinical assessment; MOC
= moclobemide; AMI = amitriptyline; PBO = placebo; CLO = clomipramine; ISO = isocarboxazid; DOX = doxepin; IMI = imipramine; FLU = fluoxetine; MAP
= maprotiline; NOR = nortriptyline; SER = sertraline
a
Data estimated from published figure
b
Patient dropout numbers not specified
c
Completers calculated with the percentages of dropouts mentioned in article
d
Number of randomizers and completers during active treatment
e
Additional 14 patients from both groups (proportion not specified) were withdrawn from study for protocol violation (169 valid completers)
f
Patients with HAM-D scores 16
J. H. Meyer et al.
Monoamine Oxidase Inhibitors in Depressive Disorders 1363
Table 5 Acute phase trials of selegiline transdermal system (STS; EMSAM) in depressiona
Diagnostic
system, Methods
diagnostic duration RX cells Randomized (n)/
Author method (weeks) (dosages) completers (n) Comments
Bodkin and DSM-IV, Random, STS 89/79 Outpatients;
Amsterdam HAM-D, DB (20 mg/d) 88/73 six-site
(2002) CLIN (6 weeks) PBO collaboration
study; tyramine-
restricted diet
during trial and
2 weeks
following end of
treatment;
STS > PBO
Amsterdam DSM-IV, Random, STS 149/108b Outpatients;
(2003) HAM-D, DB, (6 mg/d) 152/111b 16 investigative
CLIN (8 weeks) PBO sites; no dietary
restrictions;
STS > PBO
Feiger et al. DSM-IV, Random, STS 132/100 Patient type not
(2006) Mini DB, (6–12 133/106 specified; no
International (8 weeks) mg/d) dietary
Neuropsy- PBO restrictions;
chiatric STS > PBO
Interview,
HAM-D,
CGI-S,
CLIN
Table adapted from Jessen et al. (2008).
Abbreviations: STS selegiline transdermal system, DSM Diagnostic and Statistical Manual of
Mental Disorders, HAM-D Hamilton Depression Rating Scale, CGI-S Clinical Global Impressions
Scale, DB double blind, CLIN clinical assessment, PBO placebo
a
Only includes studies that targeted patients ages 18 and above
b
41 patients from each group discontinued; completers (n) calculated by subtracting 41 from
randomized (n)
is 0.3–2 h, and the elimination half-life after oral administration is around 2 h (Nair
et al. 1993). Chronic dosing of total daily doses of 300 mg to 1200 mg is associated
with MAO-A occupancies of 75% to 85% (Chiuccariello et al. 2016). Moclobemide
is metabolized extensively in the liver (Jauch et al. 1990), and at least, part of that
metabolism appears to be catalyzed by CYP2C19 (Waldmeier et al. 1994). In
addition, moclobemide is an inhibitor of CYPs 2C19, 2D6 and 1A2 (Gram et al.
1995; Bonnet 2003). Two major metabolites appearing in the plasma are a pharma-
cologically inactive lactam formed by oxidation of the morpholine moiety and the
N-oxide which retains weak MAO-A-inhibiting activity (Jauch et al. 1990). There
has been a great deal of research on the metabolism of selegiline, with extensive
metabolism to the () isomers of methamphetamine, amphetamine,
N-propargylamphetamine and selegiline-N-oxide reported (Heinonen et al. 1989),
1364 J. H. Meyer et al.
and much of this metabolism is catalyzed by several CYP enzymes (Salonen et al.
2003; Meyer 2017). Because of their long-lasting inhibition of MAO, the pharma-
codynamic half-life of the irreversible MAOIs is very much longer than their
pharmacokinetic half-life.
One of the most widely described adverse effects of irreversible MAOIs is the
“cheese effect,” which is actually a food-drug interaction. An increase in blood
pressure, the result of irreversible inhibitors of MAO-A (e.g., phenelzine,
tranylcypromine, isocarboxazid) preventing the oxidation of sympathomimetic
amines, primarily tyramine, in foodstuffs ingested while the patient is taking the
MAOI, may result in symptoms ranging from a headache to a full-blown hyperten-
sive crisis. Tyramine, a sympathomimetic amine that is present at relatively high
concentrations in some foods, is normally metabolized by MAO-A in the gut.
However, when such foods are ingested by patients taking irreversible,
non-selective MAOIs, the tyramine is not metabolized and enters sympathetic
varicosities supplying the vasculature. Displacement and release of noradrenaline,
which may itself be present in vesicles at elevated concentrations due to inhibition of
MAO-A, results in intense vasoconstriction. This effect is called the cheese effect
because in early studies on this phenomenon, one of the proposed food culprits was
aged cheese. The awareness of this effect means that patients receiving these drugs
receive a list of foods that should be avoided while taking these medications. This
adverse effect has been widely described in medical textbooks and has been respon-
sible in large part for the underutilization of MAOIs as antidepressants. However, as
indicated in numerous papers (e.g., Grady and Stahl 2012; Meyer 2017; Gillman
2018), many of the foods on the list do not contain as much tyramine as in past years
because of improved food processing. Although clinicians and patients should be
aware of this food-drug interaction, it is generally felt to be manageable with correct
dietary instructions (Gillman 2018), and the list of culprit foods has now been
updated (see Grady and Stahl 2012; Meyer 2017; and Chockalingam et al. 2018
for current information).
Nevertheless, concern about this adverse effect did lead to the development of
reversible inhibitors of MAO-A (RIMAs) (Nair et al. 1993; Waldmeier et al. 1994;
Fulton and Benfield 1996; Riederer et al. 2004), one of which, moclobemide, is used
in many countries, although not approved in the USA. Unlike phenelzine,
tranylcypromine and isocarboxazid, moclobemide is not an irreversible inhibitor of
MAO-A, but rather competes reversibly with substrates (including tyramine) for
access to the active site. Moclobemide retains antidepressant properties through
reducing the inactivation of noradrenaline and serotonin, but tyramine may still
compete for access to the enzyme, thereby minimizing risk of an increase in levels of
the dietary amine. Moclobemide also has the advantage that in situations where it is
decided to stop the MAOI, it is not necessary to wait for about 2 weeks while new
enzyme is synthesized before starting another drug; the drug change can be made
Monoamine Oxidase Inhibitors in Depressive Disorders 1365
after a much shorter time in the case of moclobemide. Brofaromine, a RIMA that
also has modest effects as a serotonin reuptake inhibitor, has also been reported to be
an effective antidepressant (Laux et al. 1990; Chouinard et al. 1993; Volz et al. 1996;
Lotufo-Neto et al. 1999; Entzeroth and Ratty 2017), but was not marketed for what
were apparently commercial reasons.
In another effort to deal with the cheese effect, the drug ()-deprenyl (selegiline)
a selective, irreversible MAO-B inhibitor, was tested as an antidepressant. Deprenyl
was actually synthesized in the 1960s as a potential psychic energizer, and it was
subsequently discovered that the () enantiomer (selegiline) was a strong selective
MAO-B inhibitor at low doses (Knoll and Magyar 1972). Early clinical studies with
oral selegiline, alone or in combination with the amino acids L-5-hydroxytryptophan
or L-phenylalanine, reported antidepressant activity (Mendlewicz and Youdim 1978,
1983; Birkmayer et al. 1984). Mendlewicz and Youdim (1983) reported, in a double-
blind placebo-controlled study, that oral selegiline at 15 mg/day for 40 days pro-
duced an improvement in Hamilton scores beginning at 8 days and a marked
improvement by 40 days of administration. Mann et al. (1989), also in a double-
blind placebo-controlled study, compared the effects of doses of selegiline of 10 mg/
day and 30 mg/day (the latter dose inhibits MAO-A as well as MAO-B), each dose
administered for 3 weeks; they found an antidepressant effect with the higher dose,
but not with the lower dose. In addition, plasma levels of selegiline fluctuate
considerably after oral dosing (Azzaro et al. 2007); so, in order to take advantage
of the potential antidepressant effect of selegiline and achieve consistent blood
plasma levels, it has been administered topically, as a patch (Bodkin and Amsterdam
2002: Tabi et al. 2020), minimizing inhibition in the gut while causing an antide-
pressant effect due to inhibition of MAO-A (and -B) in brain, a consequence of the
high brain-blood concentration ratio achieved. Fowler et al. (2015), in a [11C]
clorgyline PET study of patients treated with the selegiline patch for 28 days,
demonstrated a modest occupancy of brain MAO-A of 33 per cent. The transdermal
patch is marketed under the proprietary name, EMSAM. Oral selegiline and
rasagiline, another selective inhibitor of MAO-B, are used in the treatment of
Parkinson’s disease (Riederer et al. 2004; Riederer and Muller 2017;Tabi et al.
2020).
Although the cheese effect and hypertension have received a great deal of
attention, MAOIs administered under conditions in which there is not concomitant
consumption of potentially interacting foods or sympathomimetic drugs are more
likely to produce hypotension and associated dizziness (Fiedorowicz and Swartz
2004; Meyer 2017; Ricken et al. 2017; Kiani 2020). Although not well documented,
tranylcypromine may produce transient paradoxical hypertension in some patients in
an hour or two after administration (Gillman 2018). Weight gain and sexual dys-
function may occur with phenelzine and isocarboxazid, but these adverse effects
appear much less frequently with tranylcypromine and moclobemide (Ricken et al.
2017; Meyer 2017; Kiani 2020). Insomnia and dry mouth occur more often with
tranylcypromine than with the hydrazine MAOIs (Kiani 2020). Pyridoxine defi-
ciency and rare cases of hepatotoxicity have been reported with phenelzine
(Fiedorowicz and Swartz 2004). In a systematic review and meta-analysis
1366 J. H. Meyer et al.
The main drug-drug interactions of concern with MAOIs are those involved in
increased blood pressure caused by noradrenergic sympathomimetic drugs and
those that may cause serotonin syndrome (due to markedly increased serotonin
availability as a result of combining MAOIs with serotonin reuptake inhibitors).
With regard to noradrenergic drugs, product monographs indicate that the following
should be not be used in patients prescribed an MAOI: decongestants (phenyleph-
rine, pseudoephedrine), stimulants (amphetamine, methylphenidate, modafinil,
armodafinil), antidepressants that inhibit noradrenaline reuptake and some other
drugs including phentermine, local anaesthetics containing vasoconstrictors,
tramadol and tapentadol. However, in exceptional circumstances, some of these
drugs have been used cautiously in combination with MAOIs in treatment-resistant
depression (Grady and Stahl 2012; Israel 2015; Thomas et al. 2015), for example,
with some anaesthetics in situations in which surgeries are required and an MAOI
has been an essential antidepressant or in combination with certain antidepressants
when virtually all other antidepressant treatments have been unsuccessful. Patients
who are known abusers of cocaine, methamphetamine or other stimulants should not
be treated with MAOIs. With regard to opioids, there are no known direct interac-
tions between MAOIs and opioid mechanisms, but opioids with off-target effects
that influence the reuptake or release of serotonin such as meperidine, methadone,
tramadol and dextromethorphan or tapentadol which inhibits noradrenaline reuptake
are contraindicated. Other drugs that may cause serotonin syndrome in combination
with MAOIs include SSRIs, SNRIs, the tricyclics clomipramine and imipramine,
St. John’s wort, sumatriptan, chlorpheniramine and brompheniramine (Grady and
Stahl 2012). There is a wide range of symptoms in serotonin syndrome, the most
serious of these being hyperthermia, brain damage and even death (Grady and Stahl
Monoamine Oxidase Inhibitors in Depressive Disorders 1367
2012; Volpi-Abadie et al. 2013). The Hunter criteria used for diagnosis of serotonin
syndrome include clonus, agitation, diaphoresis, tremor and hyperreflexia (Dunkley
et al. 2003). For detailed discussions of serotonin syndrome/serotonin toxicity,
readers are referred to Gillman (2011b) and Buckley et al. (2014). Moclobemide
inhibits CYP2C19, CYP1A2 and CYP2D6 and may decrease catabolism of sub-
strates for those enzymes, although there have been relatively few reports of
clinically relevant drug interactions involving moclobemide (Bonnet 2003). Cimet-
idine has been reported to double plasma levels of moclobemide when the two drugs
are coadministered (Nair et al. 1993).
If discontinuation of a serotonin reuptake inhibitor is required in order to start a
patient on an MAOI, it may be necessary to wait for up to 2 weeks, depending on the
half-life of the serotonin reuptake inhibitor (but in the case of fluoxetine a 5-week
wait is required). If discontinuing an irreversible MAOI to switch to another drug, a
delay of 2 weeks is required to allow new MAO to be synthesized (Grady and Stahl
2012). In the case of the reversible MAOI moclobemide, the product monograph
typically recommends a washout period of 4–5 days. When switching from an
irreversible MAOI to another MAOI, a 2-week delay is also recommended, but it
is also acknowledged that the gap without antidepressant treatment may be associ-
ated with a worsening of symptoms in patients with severe depression, and there
have been reports of successful switches with much shorter delays (Szuba et al.
1997; Polnak et al. 2017), although such patients should be monitored closely.
A detailed table on switching based on the Maudsley Prescribing Guidelines is
available (Luft 2013). With regard to surgery on patients taking MAOIs, it has
generally been recommended that the use of MAOIs be discontinued prior to
surgery. However, such discontinuation may risk compromising the patient’s psy-
chiatric status, and there are reports indicating that discontinuation of the MAOI is
not always necessary (Van Haelst et al. 2012; Krings-Ernst et al. 2013).
The product monographs for MAOIs consistently report that the addition of an
MAOI to a failed tricyclic antidepressant trial is dangerous because of the risk of a
possible hypertensive or serotonergic crisis. There is a dissenting view in the
literature reporting that the tricyclic (with the exception of clomipramine, which
must be avoided) plus MAOI combination is effective and well-tolerated (White and
Simpson 1981; Gillman 2007, 2019; Amsterdam and Kim 2019; Ulrich et al. 2020
and references therein). It has been proposed that the combination of tricyclic
antidepressants that are strong noradrenaline reuptake inhibitors with MAOIs
might even make the MAOIs safer since the tricyclic attenuates the pressor response
to tyramine (Gillman 2007, 2011a). When MAOIs and tricyclics have been proposed
as combinations, it has been indicated that the two drugs should be started together
or the tricyclic should be given before the MAOI (White and Simpson 1981; Ulrich
et al. 2020). There are cases in the literature where very serious adverse effects
occurred when the tricyclic is started after the MAOI, but there are fewer reports of
1368 J. H. Meyer et al.
serious adverse events when the MAOI is given after the TCA is started (White and
Simpson 1981). Gillman (2019) has published a comprehensive list of references to
studies on MAOI-tricyclic antidepressant combinations.
Other drugs that have been used in combination with MAOIs include trazodone,
bupropion, lithium, stimulants such as pemoline, amphetamine and methylpheni-
date, and atypical (second-generation) antipsychotics (Tariot et al. 1986; Fawcett
et al. 1991; Kok et al. 2007; Israel 2015; Thomas et al. 2015; Meyer et al. 2017;
Meyer 2018; Kiani 2020). Lithium augmentation has been used effectively with
tranylcypromine (Tariot et al. 1986) and phenelzine (Kok et al. 2007). However, it
should be noted that all of these, with the exception of lithium addition, are generally
contraindicated in combination with MAOIs in product monographs. It is reported
that the addition of low-dose trazodone to MAOIs appears to improve tolerability to
the MAOIs by counteracting insomnia (Nierenberg and Keck 1989). However,
trazodone has some serotonin reuptake-inhibiting ability and is listed in the June
2020 position statement of the Royal College of Psychiatrists in the United Kingdom
as one of the drugs that MAOIs should not be prescribed alongside because of the
risk of inducing serotonin syndrome (Chamberlain et al. 2020). Tranylcypromine
and bupropion in combination have been reported to improve symptoms in
treatment-refractory depression (Pierre and Gitlin 2000); however, Thomas et al.
(2015) have indicated that patients on this combination should be monitored closely
for hypertension and, as mentioned earlier, this is contraindicated in most product
monographs. Stimulants have been reported to be effective in combination with
MAOIs for treatment-refractory depression (Fawcett et al. 1991; Feinberg 2004).
The stimulants may help in normalizing blood pressure in those patients experienc-
ing hypotension when taking an MAOI, but patients taking this drug combination
must be monitored carefully (Thomas et al. 2015), and again, this is contraindicated
in most product monographs. Meyer et al. (2017) published a case report, accom-
panied by a literature review, in which a combination of phenelzine, aripiprazole and
quetiapine was effective in a treatment-resistant patient. The authors concluded that
the atypical antipsychotics are effective for augmenting MAOIs in patients who have
responded inadequately. They also stated that there may be a risk for orthostasis in
middle-aged and older patients with MAOI-atypical antipsychotic combinations but
that with the exception of ziprasidone, the combination does not carry the problem of
potential serotonin syndrome (Meyer et al. 2017).
There has been concern expressed for several years about the pricing and avail-
ability of MAOIs and the lack of adequate education of health care professionals
regarding the appropriate use of these drugs (Fiedorowicz and Swartz 2004;
Gillman 2011a; Grady and Stahl 2012; Shulman et al. 2013). Members of the
International MAOI Expert Group recently published a statement about MAOIs
enumerating these issues and making recommendations for improvements for the
future (Gillman et al. 2020).
Summary
The MAOIs represent a class of antidepressant drugs that are underutilized in the
treatment of depressive disorders, primarily because of concerns about their potential
adverse effects and the paucity of education about these drugs in academic curricula.
In this review, we discussed the five MAOIs currently available, namely, phenelzine,
tranylcypromine, isocarboxazid (irreversible, nonselective), moclobemide (revers-
ible, MAO-A selective) and selegiline (irreversible, MAO-B selective at low doses
but non-selective at higher doses; used as a patch in depression). Although it is
generally agreed that the MAOIs are useful in the treatment of atypical depression
and treatment-refractory depression, there are also numerous reports on their efficacy
in melancholic depression, dysthymic disorder, bipolar depression and various
anxiety disorders with or without accompanying depression. There are also numer-
ous reports in the literature suggesting that the MAOIs are multifaceted drugs with
neuroprotective effects, although these effects may in some cases be due, at least in
part, to actions other than inhibition of MAO. Concern over the “cheese effect” has
deterred many physicians from prescribing MAOIs, leading to underuse, even
though this effect can be avoided. As a result of underutilization, these drugs are
currently in short supply worldwide (Gillman et al. 2020).
Acknowledgments The authors are grateful to CIHR (GBB, AH, DM) and the Offices of the
Provost and the Vice President (Research) (GBB), the Faculty of Medicine & Dentistry and the
Department of Psychiatry, University of Alberta (GBB, DM, AH) for funding. The technical and
secretarial support of Ms. Tricia Kent is gratefully acknowledged.
Conflicts of Interest Dr. Meyer has patents to detect MAO in brain and blood in mood and
commonly associated disorders and is developing natural health products to counter temporary
elevation in MAO-A in the early postpartum. Dr. Baker is an advisor for NeuraWell Therapeutics
about MAOIs, but that company has had no input into the preparation of this manuscript. The other
authors have no conflicts to declare.
Monoamine Oxidase Inhibitors in Depressive Disorders 1371
References
Agnew PC, Baran ID, Klapman HJ, et al. A clinical evaluation of four antidepressant drugs (Nardil,
Tofranil, Marplan, and Deprol). Am J Psychiatry. 1961;118:160–2. https://doi.org/10.1176/ajp.
118.2.160.
Amsterdam JD. A double-blind, placebo-controlled trial of the safety and efficacy of selegiline
transdermal system without dietary restrictions in patients with major depressive disorder. J Clin
Psychiatry. 2003;64(2):208–14. https://doi.org/10.4088/jcp.v64n0216.
Amsterdam JD, Kim TT. Relative effectiveness of monoamine oxidase inhibitor and tricyclic
antidepressant combination therapy for treatment-resistant depression. J Clin Psychopharmacol.
2019;39(6):649–52.
Amsterdam JD, Shults J. MAOI efficacy and safety in advanced stage treatment-resistant depres-
sion – a retrospective study. J Affect Disord. 2005;89:183–8.
Azzaro AJ, Ziemniak J, Kemper E, et al. Pharmacokinetics and absolute bioavailability of selegiline
following treatment of healthy subjects with the selegiline transdermal system (6 mg/24h): a
comparison with oral selegiline capsules. J Clin Pharmacol. 2007;47:1256–67.
Bach AW, Lan NC, Johnson DL, et al. cDNA cloning of human liver monoamine oxidase a and B:
molecular basis of differences in enzymatic properties. Proc Natl Acad Sci USA. 1988;85:4934–8.
Baker GB, Dewhurst WG. Biochemical theories of affective disorders. In: Dewhurst WG, Baker
GB, editors. Pharmacotherapy of affective disorders. Theory and practice. Beckenham: Croom
Helm Ltd.; 1985. p. 1–59.
Baker GB, Wong JT-F, Yeung JM, et al. Effects of the antidepressant phenelzine on brain levels of
γ-aminobutyric acid (GABA). J Affec Disorders. 1991;21:207–11.
Baker GB, Urichuk LJ, McKenna KF, et al. Metabolism of monoamine oxidase inhibitors. Cell Mol
Neurobiol. 1999;19:411–26.
Baker GB, Matveychuk D, MacKenzie EM, et al. Attenuation of the effects of oxidative stress by
the MAO-inhibiting antidepressant and carbonyl scavenger phenelzine. Chem Biol Interact.
2019;304:139–47.
Bakish D, Bradwejn J, Nair N, McClure J, Remick R, Bulger L. A comparison of moclobemide,
amitriptyline and placebo in depression: a Canadian multicentre study. Psychopharmacology.
1992;106(Suppl):S98–S101. https://doi.org/10.1007/BF02246248.
Baldwin DS, Anderson IM, Nutt DJ, et al. Evidence-based pharmacological treatment of anxiety
disorders, post-traumatic stress disorder and obsessive compulsive disorder: a revision of the
2005 guidelines from the British Association for Psychopharmacology. J Psychopharmacol.
2014;28(5):403–9.
Bartholomew AA. An evaluation of tranylcypromine (“Parnate”) in the treatment of depression.
Med J Aust. 1962;49(1):655–62. https://doi.org/10.5694/j.1326-5377.1962.tb27021.x.
Baumhackl U, Biziere K, Fischbach R, et al. Efficacy and tolerability of moclobemide compared
with imipramine in depressive disorder (DSM-III): an Austrian double-blind, multicentre study.
Br J Psychiatry Suppl. 1989;6:78–83.
Benson CA, Wong G, Tenorio G, et al. The MAO inhibitor phenelzine can improve functional
outcomes in mice with established clinical signs in experimental autoimmune encephalomyelitis
(EAE). Behav Brain Res. 2013;252:302–11.
Binda C, Newton-Vinson P, Hubalek F, et al. Structure of human monoamine oxidase B, a drug
target for the treatment of neurological disorders. Nature Struct Biol. 2002;9(1):22–6.
Birkenhager TK, van den Broek WW, Mulder PG, et al. Efficacy and tolerability of tranylcypromine
versus phenelzine: a double-blind study in antidepressant-refractory depressed patients. J Clin
Psychiatry. 2004;65(11):1505–10.
Birkmayer W, Riederer P, Linauer W, et al. L-Deprenyl plus L-phenylalanine in the treatment of
depression. J Neural Transm. 1984;59:81–7.
Bodkin JA, Amsterdam J. Transdermal selegiline in major depression: a double-blind, placebo
controlled parallel-group study in outpatients. Am J Psychiatry. 2002;159(11):1869–75.
Bonnet U. Moclobemide: therapeutic use and clinical studies. CNS Drug Rev. 2003;9(1):97–140.
1372 J. H. Meyer et al.
British Medical Research Council. Clinical trial of the treatment of depressive illness. Report to the
Medical Research Council by its clinical Psychiatry committee. Br Med J. 1965;1(5439):881–6.
https://doi.org/10.1136/bmj.1.5439.881.
Buckley N, Isbister GK, Dawson AH. Serotonin syndrome. Br Med J. 2014;348:g1626.
Chamberlain SR, Metastasio A, Stokes PRA, et al. The use of monoamine oxidase inhibitors in
psychiatric practice. Royal College of Psychiatrists. 2020;PS03/20:1–20.
Chen Z, Park J, Butler B, et al. Mitigation of sensory and motor deficits by acrolein scavenger
phenelzine in a rat model of spinal cord contusive injury. J Neurochem. 2016;138(2):328–38.
Chiuccariello L, House S, Miler L, et al. Elevated monoamine oxidase A binding during major
depressive episodes is associated with greater severity and reversed neurovegetative symptoms.
Neuropsychopharmacology. 2014;39:973–80.
Chiuccariello L, Cooke RG, Miler L, et al. Monoamine oxidase-A occupancy by moclobemide and
phenelzine: implications for the development of monoamine oxidase inhibitors. Int J Neuropsy-
chopharmacol. 2016:1–9.
Chockalingam R, Gott BM, Conway CR. Tricyclic antidepressants and monoamine oxidase
inhibitors: are they too old for a new look? In: Macaluso M, Preskorn S, editors. Handbook
of experimental pharmacology: antidepressants, vol. 250. Heidelburg: Springer; 2018. p. 37–47.
Chouinard G, Saxena B, Nair NP, et al. Brofaromine in depression; a Canadian multicenter placebo
trial and a review of standard drug comparative studies. Clin Neuropharmacol. 1993;16(Suppl
2):551–4.
Cipriani A, Williams T, Nikolakopoulou A, et al. Comparative efficacy and acceptability of
pharmacological treatments for post-traumatic stress disorder in adults: a network meta-analysis.
Psychol Med. 2018;48(12):1975–84.
Classen W, Laux G. Psychometric alterations in treatment with the MAO-A-inhibitor moclobemide.
J Neural Transm. 1990;Suppl 32:185–8.
Clineschmidt BV, Horita A. The monoamine oxidase catalyzed degradation of phenelzine-1-14C,
an irreversible inhibitor of monoamine oxidase – I. Studies in vitro. Biochem Pharmacol.
1969;18(5):1011–20.
Danish University Antidepressant Group (DUAG). Moclobemide: a reversible MAO-A-inhibitor
showing weaker antidepressant effect than clomipramine in a controlled multicenter study.
J Affect Disord. 1993;28(2):105–16. https://doi.org/10.1016/0165-0327(93)90039-m.
Davidson JR, McLeod MN, Kurland AA, et al. Antidepressant drug therapy in psychotic depres-
sion. Br J Psychiatry. 1977;131:493–6. https://doi.org/10.1192/bjp.131.5.493.
Davidson JRT, Giller EL, Zisook S, et al. An efficacy study of isocarboxazid and placebo in
depression, and its relationship to depressive nosology. Arch Gen Psychiatry. 1988;45:120–7.
De Lima MS, Hotopf M. A comparison of active drugs for the treatment of dysthymia. Cochrane
Database Syst Rev. 2003;3:CD004047.
Dierick M, Cattiez P, Franck G, et al. Moclobemide versus clomipramine in the treatment of
depression: a double-blind multicentre study in Belgium. Acta Psychiatr Scand Suppl.
1990;360:50–1. https://doi.org/10.1111/j.1600-0447.1990.tb05328.x.
Dunkley EJC, Isbister GK, Sibbritt D, et al. The hunter serotonin toxicity criteria: simple and
accurate diagnostic decision rules for serotonin toxicity. QJ Med. 2003;96(9):635–42.
Entzeroth M, Ratty AK. Monoamine oxidase inhibitors – revisiting a therapeutic principle. Open
J Depress. 2017;6(2):31–68.
Evans L, George T, O’Sullivan B, Mitchell P, Johnson G, Adena M. An Australian multicentre
study of moclobemide versus amitriptyline in the treatment of depression. Aust N Z J Psychi-
atry. 1992;26(3):454–8. https://doi.org/10.3109/00048679209072070.
Fawcett J, Kravitz HM, Zajecka JM, et al. CNS stimulant potentiation of monoamine oxidase
inhibitors in treatment-refractory depression. J Clin Psychopharmacol. 1991;11(20):127–32.
Feiger AD, Rickels K, Rynn MA, et al. Selegiline transdermal system for the treatment of major
depressive disorder: an 8-week, double-blind, placebo-controlled, flexible-dose titration trial.
J Clin Psychiatry. 2006;67(9):1354–61. https://doi.org/10.4088/jcp.v67n0905.
Monoamine Oxidase Inhibitors in Depressive Disorders 1373
Feinberg SS. Combining stimulants with monoamine oxidase inhibitors. A review of uses and one
possible additional indication. J Clin Psychiatry. 2004;65:1520–4.
Fiedorowicz JG, Swartz KL. The role of monoamine oxidase inhibitors in current psychiatric
practice. J Psychiatr Pract. 2004;10(4):239–48.
Fowler JS, Logan J, Volkow ND, et al. Evidence that formulations of the selective MAO-B
inhibitor, selegiline, which bypass first-pass metabolism, also inhibit MAO-A in the human
brain. Neuropsychopharmacology. 2015;40:650–7.
Fulton B, Benfield P. Moclobemide. An update of its pharmacological properties and therapeutic
use. Drugs. 1996;52(3):450–74.
Gachoud JP, Dick P, Kohler M. Comparison of the efficacy and tolerability of moclobemide and
maprotiline in depressed patients treated by general practitioners. Clin Neuropharmacol.
1994;17(Suppl 1):S29–37. https://doi.org/10.1097/00002826-199417001-00005.
Georgotas A, McCue RE, Hapworth W, et al. Comparative efficacy and safety of MAOIs versus
TCAs in treating depression in the elderly. Biol Psychiatry. 1986;21(12):1155–66. https://doi.
org/10.1016/0006-3223(86)90222-2.
Gillman PK. Tricyclic antidepressant pharmacology and therapeutic drug interactions updated. Br J
Pharmacol. 2007;151:737–48.
Gillman PK. Advances pertaining to the pharmacology and interactions of irreversible nonselective
monoamine oxidase inhibitors. J Clin Psychopharmacol. 2011a;31(1):66–74.
Gillman PK. CNS toxicity involving methylene blue: the exemplar for understanding and predicting
drug interactions that precipitate serotonin toxicity. J Psychopharmacol. 2011b;25(3):429–36.
Gillman PK. A reassessment of the safety profile of monoamine oxidase inhibitors: elucidating tired
old tyramine myths. J Neural Transm. 2018;125:1707–17.
Gillman PK. Overview: MAOI and TCA interactions. PsychoTropical Commen. 2019 (updated
Sept. 28, 2020). https://psychotropical.com/overview_maoi_and_tca_interactions/
Gillman PK, Feinberg SS, Fachtmann LJ. Revitalizing monoamine oxidase inhibitors: a call for
action. CNS Spectr. 2020;25:452–4.
Glick BS. Double-blind study of tranylcypromine and phenelzine in depression. Dis Nerv Syst.
1964;25:617–9.
Gottfries CG. Clinical trial with the monoamine oxidase inhibitor tranylcypromine on a psychiatric
clientele. Acta Psychiatr Scand. 1963;39(3):463–72. https://doi.org/10.1111/j.1600-0447.1963.
tb07476.x.
Grady MM, Stahl SM. Practical guide for prescribing MAOIs: debunking myths and removing
barriers. CNS Spectr. 2012;17:2–10.
Gram LF, Guentert TW, Grange S, et al. Moclobemide, a substrate of CYP2C19 and an inhibitor of
CYP2C19, CYP2D6 and CYP1A2: a panel study. Clin Pharmacol Therap. 1995;57(6):670–7.
Greenblatt M, Grosser GH, Wechsler H. Differential response of hospitalized depressed patients to
somatic therapy. Am J Psychiatry. 1964;120:935–43. https://doi.org/10.1176/ajp.120.10.935.
Guelfi JD, Payan C, Fermanian J, et al. Moclobemide versus clomipramine in endogenous depres-
sion. A double-blind randomised clinical trial. Br J Psychiatry. 1992;160:519–24. https://doi.
org/10.1192/bjp.160.4.519.
Hamilton M. The effect of treatment on the melancholias (depressions). Br J Psychiatry. 1982;140:
223–30. https://doi.org/10.1192/bjp.140.3.223.
Hampson DR, Baker GB, Coutts RT. A comparison of the neurochemical and pharmacological
properties of the stereoisomers of tranylcypromine. Cell Mol Biol. 1986;32:333–41.
Harris S, Johnson S, Duncan JW, et al. Evidence revealing deregulation of the KLF11-Mao A
pathway in association with chronic stress and depressive disorders. Neuropsychophar-
macology. 2014;40:1373–82.
Heinonen EH, Myllyla V, Sotaniemi K, et al. Pharmacokinetics and metabolism of selegiline. Acta
Neurol Scand Suppl. 1989;126:93–9.
Henkel V, Mergi R, Allgaier AK, et al. Treatment of depression with atypical features: a meta-
analytic approach. Psychiatry Res. 2006;141(1):89–101.
1374 J. H. Meyer et al.
Himmelhoch JM, Fuchs CZ, Symons BJ. A double-blind study of tranylcypromine treatment of
major anergic depression. J Nerv Ment Dis. 1982;170(10):628–34.
Himmelhoch JM, Thase ME, Mallinger AG, Houck P. Tranylcypromine versus imipramine in
anergic bipolar depression. Am J Psychiatry. 1991;148(7):910–6. https://doi.org/10.1176/ajp.
148.7.910.
Imlah NW, Fahy PT, Harrington JA. A comparison of two antidepressant drugs. Psychophar-
macologia. 1964;6(6):472–4. https://doi.org/10.1007/BF00429573.
Ishizaki J, Mimura M. Dysthymia and apathy: diagnosis and treatment. Depress Res Treat.
2011;2011:893905.
Israel JA. Combining stimulants and monoamine oxidase inhibitors: a reexamination of the
literature and a report of a new treatment combination. Primary Care Companion CNS Disord.
2015;17(6) https://doi.org/10.4088/PCC.15br01836.
Jauch R, Griesser E, Oesterhelt G, et al. Biotransformation of moclobemide in humans. Acta
Psychiatr Scand Suppl. 1990;360:87–90.
Jessen L, Kovalick LJ, Azzaro AJ. The selegiline transdermal system (Emsam): a therapeutic option
for the treatment of major depressive disorder. Pharm Therap. 2008;33(4):212–46.
Johnson S, Stockmeier CA, Meyer JH, et al. The reduction of R1, a novel repressor protein for
monoamine oxidase A, in major depressive disorder. Neuropsychopharmacol. 2011;36:2139–48.
Johnston JP. Some observations upon a new inhibitor of monoamine oxidase in brain tissue.
Biochem Pharmacol. 1968;17(7):1285–97.
Jouvent R, Le Houezec J, Payan C, et al. Dimensional assessment of onset of action of antidepres-
sants: a comparative study of moclobemide vs. clomipramine in depressed patients with blunted
affect and psychomotor retardation. Psychiatry Res. 1998;79(3):267–75. https://doi.org/10.
1016/s0165-1781(98)00046-8.
Kallem RR, Jillela B, Ravula AR, et al. Highly sensitive LC-MS/MS-ESI method for determination
of phenelzine in human plasma and its application to a human pharmacokinetic study. J
Chromatogr B. 2016;1022:126–32.
Kay DW, Garside RF, Fahy TJ. A double-blind trial of phenelzine and amitriptyline in depressed
out-patients. A possible differential effect of the drugs on symptoms. Br J Psychiatry. 1973;123
(572):63–7. https://doi.org/10.1192/bjp.123.1.63.
Kayser A, Robinson DS, Nies A, Howard D. Response to phenelzine among depressed patients
with features of hysteroid dysphoria. Am J Psychiatry. 1985;142(4):486–8. https://doi.org/10.
1176/ajp.142.4.486.
Kayser A, Robinson DS, Yingling K, Howard DB, Corcella J, Laux D. The influence of panic
attacks on response to phenelzine and amitriptyline in depressed outpatients. J Clin
Psychopharmacol. 1988;8(4):246–53.
Kennedy SH, Holt A, Baker GB. Monoamine oxidase inhibitors. In: Sadock BJ, Sadock VA, Ruiz P,
editors. Kaplan & Sadock’s comprehensive textbook of psychiatry, 9th ed. Philadelphia: Wolters
Kluwer; 2009. pp. 3154–64.
Kiani C. Tranylcypromine: its pharmacology, safety and efficacy. Am J Psychiatry Residents’
J. 2020;15(4):3–5.
Kim T, Xu C, Amsterdam JD. Relative effectiveness of tricyclic antidepressant versus monoamine
oxidase inhibitor monotherapy for treatment-resistant depression. J Affect Disord. 2019;250:
199–203.
Knoll J, Magyar K. Some puzzling pharmacological effects of monoamine oxidase inhibitors. Adv
Biochem Psychopharmacol. 1972;5:393–408.
Kok RM, Vink D, Heeren DJ, et al. Lithium augmentation compared with phenelzine in treatment-
resistant depression in the elderly: an open, randomized, controlled trial. J Clin Psychiatry.
2007;68(8):1177–85.
Kragh-Sorensen P, Muller B, Andersen JV, et al. Moclobemide versus clomipramine in depressed
patients in general practice. A randomized, double-blind, parallel, multicenter study. J Clin
Psychopharmacol. 1995;15(4 Suppl 2):24S–30S. https://doi.org/10.1097/00004714-
199508001-00005.
Monoamine Oxidase Inhibitors in Depressive Disorders 1375
Krings-Ernst I, Ulrich A, Adli M. Antidepressant treatment with MAO-inhibitors during general and
regional anesthesia: a review and case report of spinal anesthesia for lower extremity surgery
without discontinuation of tranylcypromine. Int J Clin Pharmacol Therapeut. 2013;51(10):763–70.
Lapierre YD, Joffe R, McKenna K, et al. Moclobemide versus fluoxetine in the treatment of major
depressive disorder in adults. J Psychiatry Neurosci. 1997;22(2):118–26.
Larsen JK, Gjerris A, Holm P, et al. Moclobemide in depression: a randomized, multicentre trial
against isocarboxazid and clomipramine emphasizing atypical depression. Acta Psychiatr
Scand. 1991;84(6):564–70. https://doi.org/10.1111/j.1600-0447.1991.tb03196.x.
Larsen JK, Krogh-Neilsen L, Brosen K. The monoamine oxidase inhibitor isocarboxazid is a
relevant treatment option in treatment-resistant depression – experience-based strategies in
Danish psychiatry. Health Care Curr Rev. 2016;4:2.
Laux G, Beckmann H, Classen W, Becker T. Moclobemide and maprotiline in the treatment of
inpatients with major depressive disorder. J Neural Transm Suppl. 1989;28:45–52.
Laux G, Classen W, Sofic E, et al. Clinical, biochemical and psychometric findings with the new
MAO-A-inhibitors moclebemide and brofaromine in patients with major depressive disorder.
J Neural Transm Suppl. 1990;32:189–95.
Lecrubier Y, Guelfi JD. Efficacy of reversible inhibitors of monoamine oxidase-a in various forms
of depression. Acta Psychiatr Scand Suppl. 1990;360:18–23. https://doi.org/10.1111/j.1600-
0447.1990.tb05319.x.
Lecrubier Y, Pedarriosse AM, Payan C, et al. Moclobemide versus clomipramine in non-
melancholic, nonpsychotic major depression. A Study group. Acta Psychiatr Scand. 1995;92
(4):260–5. https://doi.org/10.1111/j.1600-0447.1995.tb09580.x.
Leitch A, Seager CP. A trial of four anti-depressant drugs. Psychopharmacologia. 1963;4:72–7.
https://doi.org/10.1007/BF00429366.
Liebowitz MR, Quitkin FM, Stewart JW, et al. Antidepressant specificity in atypical depression.
Arch Gen Psychiatry. 1988;45(2):129–37. https://doi.org/10.1001/archpsyc.1988.
01800260037004.
Lonnqvist J, Sintonen H, Syvalahti E, et al. Antidepressant efficacy and quality of life in depression:
a double-blind study with moclobemide and fluoxetine. Acta Psychiatr Scand. 1994;89(6):363–
9. https://doi.org/10.1111/j.1600-0447.1994.tb01530.x.
Lopez-Munoz F, Alamo C. Monoaminergic neurotransmission: the history of the discovery of
antidepressants from 1950s until today. Curr Pharm Des. 2009;15:1563–86.
Lotufo-Neto F, Trivedi M, Thase ME. Meta-analysis of the reversible inhibitors of monoamine
oxidase type A moclobemide and brofaromine for the treatment of depression. Neuropsycho-
pharmacology. 1999;20(3):226–47.
Luft B. Antidepressant switching strategies. Graylands Hospital Drug Bulletin. 2013;20(1):ISSN
1323-1251.
Lyles GA, Holt A, Marshall CM. Further studies on the metabolism of methylamine by
semicarbazide-senstive amine oxidase activities in human plasma, umbilical artery and rat
aorta. J Pharm Pharmacol. 1990;42(5):332–8.
MacKenzie EM. Neurochemical and neuroprotective aspects of phenelzine and it active metabolite
β-phentlethylidenehydrazine. PhD Thesis, University of Alberta, 2009.
Mallinger AG, Frank E, Thase ME, et al. Revisiting the effectiveness of standard antidepressants in
bipolar disorder: are monoamine oxidase inhibitors superior? Psychopharmacol Bull. 2009;42:
64–74.
Mann JJ, Aarons SF, Wilner PJ, et al. A controlled study of the antidepressant efficacy and side
effects of (-)-deprenyl. A selective monoamine oxidase inhibitor. Arch Gen Psychiatry. 1989;46:
45–50.
Martin ME. A Comparative trial of imipramine and phenelzine in the treatment of depression. Br J
Psychiatry. 1963;109(459):279–85. https://doi.org/10.1192/bjp.109.459.279.
Matveychuk D, MacKenzie EM, Kumpula D, et al. Overview of the neuroprotective effects of the
MAO-inhibiting antidepressant phenelzine. Cell Mol Neurobiol. 2021; https://doi.org/10.1007/
s10571-021-01078-3.
1376 J. H. Meyer et al.
McGrath PJ, Quitkin FM, Harrison W, et al. Treatment of melancholia with tranylcypromine. Am J
Psychiatry. 1984;141:288–9.
McGrath PJ, Stewart W, Harrison W, et al. Treatment of tricyclic refractory depression with a
monoamine oxidase inhibitor antidepressant. Psychopharmacol Bull. 1987;23:169–72.
McGrath PJ, Stewart JW, Nunes EV, et al. A double-blind crossover trial of imipramine and
phenelzine for outpatients with treatment-refractory depression. Am J Psychiatry. 1993;150
(1):118–23. https://doi.org/10.1176/ajp.150.1.118.
Meister R, von Wolff A, Mohr H, et al. Comparative safety of pharmacologic treatments for
persistent depressive disorder: a systematic review and network meta-analysis. PLoS One.
2016;11(5):e0153380.
Mendlewicz J, Youdim MBH. Anti-depressant potentiation of 5-hydroxytryptophan by L-deprenyl,
an MAO “type B” inhibitor. J Neural Transm. 1978;43:279–86.
Mendlewicz J, Youdim MBH. L-Deprenil, a selective monoamine oxidase type B inhibitor, in the
treatment of depression: a double blind evaluation. Br J Psychiatry. 1983;142:508–11.
Meyer JM. A concise guide to monoamine oxidase inhibitors. Curr Psychiatr Ther. 2017;16(12):
15–23.
Meyer JM. A concise guide to monoamine oxidase inhibitors: how to avoid drug interactions. Curr
Psychiatr Ther. 2018;17(1):22–33.
Meyer JM, Cummings MA, Proctor G. Augmentation of phenelzine with aripiprazole and
quetiapine in a treatment-resistant patient with psychotic unipolar depression: case report and
literature review. CNS Spectr. 2017;22:391–6.
Meyer JH, Ginovart N, Boovariwala A, et al. Elevated monoamine oxidase A levels in the brain: an
explanation for the monoamine imbalance of major depression. Arch Gen Psychiatry. 2006;63
(11):1209–16.
Meyer JH, Wilson AA, Sagrati S, et al. Brain monoamine oxidase A binding in major depressive
disorder. Arch Gen Psychiatry. 2009;66(12):1304–12.
Moll E, Neumann N, Schmid-Burgk W, et al. Safety and efficacy during long-term treatment with
moclobemide. Clin Neuropharmacol. 1994;17(Suppl 1):S74–87.
Moriguchi S, Wilson AA, Miler L, et al. Monoamine oxidase B total distribution volume in the
prefrontal cortex of major depressive disorder. JAMA Psychiat. 2019;76(6):634–41.
Nair NP, Ahmed SK, Kin NM. Biochemistry and pharmacology of reversible inhibitors of MAO-A
agents: focus on moclobemide. J Psychiatry Neurosci. 1993;18(5):214–25.
Nair NP, Amin M, Holm P, et al. Moclobemide and nortriptyline in elderly depressed patients. A
randomized, multicentre trial against placebo. J Affect Disord. 1995;33(1):1–9. https://doi.org/
10.1016/0165-0327(94)00047-d.
Nibuya M, Morinobu S, Duman RS. Regulation of BDNF and trkB mRNA in rat brain by chronic
ECS and antidepressant treatments. J Neurosci. 1995;15:7539–47.
Nierenberg AA, Keck PE. Management of monoamine oxidase inhibitor-associated insomnia with
trazodone. J Clin Psychopharmacol. 1989;9:42–5.
Nolen WA, van de Putte JJ, Dijken WA, et al. L-5HTP in depression resistant to re-uptake
inhibitors. An open comparative study with tranylcypromine. Br J Psychiatry. 1985;147:
16–22. https://doi.org/10.1192/bjp.147.1.16.
Nolen WA, van de Putte JJ, Dijken WA, et al. Treatment strategy in depression. II. MAO inhibitors
in depression resistant to cyclic antidepressants: two controlled crossover studies with
tranylcypromine versus L-5-hydroxytryptophan and nomifensine. Acta Psychiatr Scand.
1988;78(6):676–83. https://doi.org/10.1111/j.1600-0447.1988.tb06403.x.
O’Carroll AM, Fowler CJ, Phillips JP, et al. The deamination of dopamine by human brain
monoamine oxidase. Specificity for the two enzyme forms in seven brain regions. Naunyn
Schmiedeberg’s Arch Pharmacol. 1983;322(3):198–202.
Paslawski T, Knaus E, Iqbal N, et al. β-Phenylethylidenehydrazine, a novel inhibitor of GABA
transaminase. Drug Devel Res. 2001;54:35–9.
Philipp M, Kohnen R, Benkert O. A comparison study of moclobemide and doxepin in major
depression with special reference to effects on sexual dysfunction. Int Clin Psychopharmacol.
1993;7(3–4):149–53. https://doi.org/10.1097/00004850-199300730-00005.
Monoamine Oxidase Inhibitors in Depressive Disorders 1377
Robinson D, Cooper TB, Jindal SP, et al. Metabolism and pharmacokinetics of phenelzine: lack of
evidence for acetylation pathway in humans. J Clin Psychopharmacol. 1985;5:337–57.
Rowan PR, Paykel ES, Parker RR. Phenelzine and amitriptyline: effects on symptoms of neurotic
depression. Br J Psychiatry. 1982;140:475–83. https://doi.org/10.1192/bjp.140.5.475.
Salonen JS, Nyman L, Boobis AR, et al. Comparative studies on the cytochrome P450 -associated
metabolism and interaction potential of selegiline between human liver-derived in vitro systems.
Drug Metab Dispos. 2003;31(9):1093–102.
Schildkraut JJ, Klerman GL, Hammond R, Friend DG. Excretion of 3-methoxy-4-hydroxymandelic
acid (VMA) in depressed patients treated with antidepressant drugs. J Psychiatr Res. 1964;2:
257–66. https://doi.org/10.1016/0022-3956(64)90012-3.
Schlessinger A, Geier E, Fan H, et al. Structure-based discovery of prescription drugs that interact
with the norepinephrine transporter, NET. Proc Natl Acad Sci USA. 2011;108:15810–5.
Sheehan DV, Ballenger J, Jacobsen G. Treatment of endogenous anxiety with phobic, hysterical,
and hypochondriacal symptoms. Arch Gen Psychiatry. 1980;37(1):51–9.
Shulman KI, Hermann N, Walker SE. Current place of monoamine oxidase inhibitors in the
treatment of depression. CNS Drugs. 2013;27:789–97.
Singh IN, Gilmer LK, Miller DM, et al. Phenelzine mitochondrial functional preservation and
neuroprotection after traumatic brain injury related to scavenging of the lipid-peroxidation
derived aldehyde 4-hydroxy-2-nonenal. J Cereb Blood Flow Metab. 2013;33(4):593–9.
Sogaard J, Lane R, Latimer P, et al. A 12-week study comparing moclobemide and sertraline in the
treatment of outpatients with atypical depression. J Psychopharmacol. 1999;13(4):406–14.
https://doi.org/10.1177/026988119901300412.
Song M-S, Matveychuk D, MacKenzie EM, et al. An update on amine oxidase inhibitors:
multifaceted drugs. Progr Neuropsychopharmacol Biol Psychiatry. 2013;44:118–24.
Spear FG, Hall P, Stirland JD. A comparison of subjective responses to imipramine and
tranylcypromine. Br J Psychiatry. 1964;110:53–5. https://doi.org/10.1192/bjp.110.464.53.
Stein DJ, Cameron A, Amrein R, et al. Moclobemide is effective and well tolerated in the long-term
pharmacotherapy of social anxiety disorder with or without comorbid anxiety disorder. Int Clin
Psychopharmacol. 2002;17(4):161–70.
Szoko E, Tabi T, Riederer P, et al. Pharmacological aspects of the neuroprotective effects of
irreversible MAO-B inhibitors, selegiline and rasagaline, in Parkinson’s disease. J Neural
Transm. 2018;125:1735–49.
Szuba MP, Hornig-Rohan M, Amsterdam JD. Rapid conversion from one monoamine oxidase
inhibitor to another. Expert Opin Drug Metab Toxicol. 1997;58(7):307–10.
Tabi T, Vecsel L, Youdim MB, et al. Selegiline: a molecule with innovative potential. J Neural
Transm. 2020;127:831–42.
Tariot PN, Murphy DL, Sunderland T, et al. Rapid antidepressant effect of addition of lithium to
tranylcypromine. J Clin Psychopharmacol. 1986;6(3):165–7.
Tatton W, Chalmers-Redman R, Tatton N. Neuroprotection by deprenyl and other propargylamines:
glyceraldehyde-3-phosphate dehydrogenase rather than monoamine oxidase-B. J Neural
Transm. 2003;110:509–15.
Thase ME, Mallinger AG, McKnight D, Himmelhoch JM. Treatment of imipramine-resistant
recurrent depression, IV: a double-blind crossover study of tranylcypromine for anergic bipolar
depression. Am J Psychiatry. 1992;149(2):195–8. https://doi.org/10.1176/ajp.149.2.195.
Thase ME, Trivedi MH, Rush AJ. MAOIs in the contemporary treatment of depression. Neuropsy-
chopharmacology. 1995;12(3):185–219. https://doi.org/10.1016/0893-133X(94)00058-8.
Thomas SJ, Shin M, Mcinnis MG, et al. Combination therapy with monoamine oxidase inhibitors
and other antidepressants or stimulants: strategies for the management of treatment-resistant
depression. Pharmacotherapy. 2015;35(4):433–49.
Tipton KF, Spires IP. Oxidation of 2-phenylethylhydrazine by monoamine oxidase. Biochem
Pharmacol. 1972;21(2):268–70.
UK Moclobemide Study Group. A multicentre comparative trial of moclobemide, imipramine and
placebo in major depressive disorder. Int Clin Psychopharmacol. 1994;9(2):109–13. https://doi.
org/10.1097/00004850-199400920-00007.
Monoamine Oxidase Inhibitors in Depressive Disorders 1379
Ulrich S, Ricken R, Adli M. Tranylcypromine in mind (Part I): review of pharmacology. Eur
Neuropsychopharmacol. 2017;27:697–713.
Ulrich S, Ricken R, Buspanavich P, et al. Efficacy and adverse effects of tranylcypromine and tricyclic
antidepressants in the treatment of depression. J Clin Psychopharmacol. 2020;40:63–74.
Van Haelst IMM, van Klei WA, Doodeman HJ, et al. Antidepressive treatment with monoamine
oxidase inhibitors and the occurrence of intraoperative hemodynamic events: a retrospective
observational cohort study. J Clin Psychiatry. 2012;73(8):1103–9.
Vazquez GH, Tondo L, Undurraga J, et al. Overview of antidepressant treatment of bipolar
depression. Int J Neuropsychopharmacol. 2013;16:1673–85.
Versiani M, Oggero U, Alterwain P, et al. A double-blind comparative trial of moclobemide
v. imipramine and placebo in major depressive episodes. Br J Psychiatry. 1989;155(Suppl
6):72–7.
Versiani M, Nardi AE, Mundim FD, et al. The long-term treatment of social phobia with
moclobemide. Int Clin Psychopharmacol. 1996;11(Suppl 3):83–8.
Volpi-Abadie J, Kaye AM, Kaye AD. Serotonin syndrome. Ochsner J. 2013;13(4):533–40.
Volz H-P, Gleiter CH, Waldmeier PC, et al. Brofaromine – a review of its pharmacological
properties and therapeutic use. J Neural Transm. 1996;103:217–45.
Waldmeier P, Amrein R, Schmid-Burgk W. Pharmacology and pharmacokinetics of brofaromine
and moclobemide in animals and humans. In: Kennedy SH, editor. Clinical advances in
monoamine oxidase inhibitor therapies. Washington, DC: American Psychiatric Press; 1994.
p. 33–59.
White K, Simpson G. Combined MAOI-tricyclic antidepressant treatment: a reevaluation. J Clin
Psychopharmacol. 1981;1:264–82.
White K, Razani J, Cadow B, et al. Tranylcypromine vs nortriptyline vs placebo in depressed
outpatients: a controlled trial. Psychopharmacology. 1984;82(3):258–62. https://doi.org/10.
1007/BF00427786.
Williams R, Edwards RA, Newburn GM, Mullen R, Menkes DB, Segkar C. A double-blind
comparison of moclobemide and fluoxetine in the treatment of depressive disorders. Int Clin
Psychopharmacol. 1993;7(3–4):155–8. https://doi.org/10.1097/00004850-199300730-00006.
Williams T, McCaul M, Schwarzer G, et al. Pharmacological treatments for social anxiety disorder
in adults: a systematic review and network meta-analysis. Acta Neuropsychiatr. 2020;32(4):
169–76.
Wood PL, Khan MA, Moskal JR, et al. Aldehyde load in ischemia-reperfusion brain injury:
neuroprotection by neutralization of reactive aldehydes with phenelzine. Brain Res.
2006;1122(1):184–90.
Youdim MBH, Collins GGS, Sandler M, et al. Human brain monoamine oxidase, multiple forms
and selective inhibitors. Nature. 1972;236:225–8.
Youdim MB, Edmondson D, Tipton KF. The therapeutic potential of monoamine oxidase inhibitors.
Nat Rev Neurosci. 2006;7:295–309.
Zhang W, Davidson JR. Post-traumatic stress disorder: an evaluation of existing pharmacotherapies
and new strategies. Expert Opin Pharmacother. 2007;8(12):1861–187.
Agomelatine and Depressions
Gerd Laux
Contents
Chemistry, Developmental History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1382
Pharmacology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1382
Pharmacokinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1382
Mechanism of Action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1382
Indications (of Marketed Products) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1383
Clinical Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1383
Depressive Disorders . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1383
Sleep Disorders, Insomnia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1383
Other Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1383
Long-Term Studies, Relapse Prevention . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1388
Side Effects, Adverse Reactions, and Toxicology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1388
Combination Therapy – Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1389
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1389
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1389
Abstract
Agomelatine is an antidepressant drug acting as melatonergic MT1- and MT2-
receptor agonist and selective serotonergic 5-HT2C-antagonist. The efficacy in
treatment of major depressive disorder was established in randomized, double-
blind, placebo controlled studies and a maintenance study in adult in- and out-
patients. Positive effects on sleep parameters have been reported. Side effects are
rare; weight gain, sexual dysfunction, or discontinuation syndromes are missing;
G. Laux (*)
Institute of Psychological Medicine (IPM), Soyen, Germany
MVZ Waldkraiburg, Center of Neuropsychiatry, Waldkraiburg, Germany
Department of Psychiatry and Psychotherapy, Ludwig-Maximilians-University (LMU), Munich,
Germany
e-mail: ipm@ipm-laux.de
The chemical structure of agomelatine is very similar to that of melatonin (Fig. 1).
Agomelatine was discovered and developed by Servier; it received approval from
the European Medicines Agency (EMA) in 2009 and 2010 in Australia.
Pharmacology
Pharmacokinetics
Mechanism of Action
The psychotropic effects of agomelatine are due to the synergy between its melatonergic
and 5-HT (hydroxytryptaminergic) effects. Agomelatine dose-dependently inhibits the
firing of cells in the SCN (the endogeneous master clock), synchronizes the circadian
rhythm to 24 h by phase advancing circadian rhythms, and resynchronizes circadian
rhythms (Lemoine et al. 2007; Kasper and Hamon 2009; Kasper et al. 2010; Dridi et al.
2013). Agomelatine affects transcriptional-translational molecular circadian clock,
neurogenesis, synaptic remodeling, and glutamate signaling.
Agomelatine and Depressions 1383
Clinical Studies
Reviews of clinical studies are given by Kennedy and Rizvi (2010), Howland 2011,
Guaiana et al. (2013), and Norman and Olver 2019.
Depressive Disorders
Agomelatine is also studied for its effects on sleep regulation. Studies report various
improvements in general quality of sleep metrics, as well as benefits in circadian
rhythm disorders (Dridi et al. 2013).
Outpatients with DSM-IV-TR-defined MDD received either agomelatine 25 to
50 mg (n ¼ 154) or sertraline 50 to 100 mg (n ¼ 159) during a 6-week, randomized,
double-blind treatment. Sleep latency and sleep efficiency improved significantly
more with agomelatine than with sertraline as did anxiety symptoms. Sleep-wake
cycle at week 1 reflected early improvement in sleep and daytime functioning
(Kasper et al. 2010).
In depressed patients, agomelatine increased slow wave sleep without modifica-
tion of rapid eye movement (REM) sleep amount or REM latency. From the first
week of treatment, onset of sleep and the quality of sleep were significantly
improved without daytime clumsiness as assessed by patients (Quera Salva et al.
2007).
Other Studies
It has been found more effective than placebo in the treatment of generalized anxiety
disorder (De Berardis et al. 2013). A 12-week international RCT of agomelatine
(N ¼ 139) vs. escitalopram (N ¼ 142) vs placebo (N ¼ 131) showed response rates
of 64%, 66.2%, and 36.6%, respectively. Remission rates were 36.7%, 31.7%, and
19.9%, respectively. Agomelatine was well tolerated with no more adverse events
1384
Table 1 Survey controlled clinical studies with agomelatine (A) in Major Depressive Disorder (MDD)
N, patients, Study arms, comparators,
Study duration placebo (Pl) Results Drop-out rates Relevant side effects
Lemoine N ¼ 332 Venlafaxine (V) 75–150 mg/day A ¼V A 4.2% Nausea, dizziness, tremor
et al. V >A
(2007)
RCT 6 weeks R: A 76%, V 71% V 13.2%
A > V sleep
parameters
Quera- N ¼ 138 Escitalopram (E) A¼E Nausea E > A
Salva et al.
(2011)
RCT 24 weeks 10–20 mg/day R:A 77%, E 74%
RR:A 48%, E 42%
A > in nighttime
sleep and daytime
condition
Laux et al. N ¼ 9283 R: 78.7% 5.3% headache, nausea
(2017)
Pooled 12 weeks RR: 34.5% 0.5% transaminase elevations
analysis of 52 weeks R: 75.9%
4 NIS RR: 47.5%
G. Laux
Laux N ¼ 3317 R 65.8% (>50% 25.8% (13.5% patient’s 1.7% headache, nausea 1.4%,
(2012) decrease in request, 9.1% insufficient dizziness 0.9%, 0.6% increase
svMADRS) efficacy, 5.2% transaminase levels (0.2%
NIS 12 weeks RR 54.8% intolerability) >3ULN)
(svMADRS < 12)
Kasper N ¼ 154 A Sertraline (S) 50–100 mg/day R: A 70%, 61.5% S A 2.6% Fatigue 5.9%, A > S, diarrhea
et al. and hyperhidrosis S > A
(2010)
RCT N ¼ 159 S RR: 32.7% A, S 11.3%
Agomelatine and Depressions
28.8% S
6 weeks A > S sleep wake
cycle, sleep
efficiency
Martinotti N ¼ 30 A Venlafaxine (V) 75–150 mg/day A greater efficacy on A 3.2% A 3.2%, V 39.2%
et al. anhedonia
(2012)
RCT N ¼ 30 V V 17.8%
8 weeks
Olie and N ¼ 238 Pl A > Pl A 14.4%, Pl 15.0%, due to Headache 5%Fatigue A > Pl,
Kasper side effects A 3.4%, Pl Headache Pl > A
(2007) 5.8%
RCT 6 weeks Effect size 0.41
R: A 54.3%, Pl 4% nausea, dizziness, n.s.
35.3%
Hale et al. N ¼ 515 N ¼ 252 A A>F A 11.9% A¼F
(2010)
RCT 8 weeks N ¼ 263 Fluoxetine (F) R: A 71.7% F 18.6% Liver enzyme elevations A
Severe MDD F 63.8% 4, F 1
(HAMD > 25)
(continued)
1385
1386
Table 1 (continued)
N, patients, Study arms, comparators,
Study duration placebo (Pl) Results Drop-out rates Relevant side effects
Heun et al. N ¼ 222 Pl A > Pl
(2013) 8 weeks N ¼ 151 A R: A 59.5%
Age > 65 years N ¼ 71 Pl Pl 38.6%
Guaiana N ¼ 4495 SSRIs, Venlafaxine R and RR A lower than venlafaxine A fewer side effects
et al.
(2013)
Review A ¼ SSRIs,
13 studies Venlafaxine
Singh et al. N ¼ 3661 N ¼ 1963 Pl A > Pl Effect size
(2012) 0.26
Meta- N ¼ 1698 SSRIs A vs. Comparators
analysis Effect size 0.11
Corruble N ¼ 324 N ¼ 164 A R: A 82.6% A 11.0%, E 18.1%
et al.
(2013)
RCT 24 weeks N ¼ 160 Escitalopram (E) E 81.3%
RR: A 69.5%
E 63.1%
Sleep A > E
G. Laux
Kasper N ¼ 1997 N ¼ 1001 A A > SSRIs/SNRI
et al.
(2013)
Pool N ¼ 996 Sertraline, fluoxetine, R: A 71%
analysis venlafaxine, escitalopram, SSRIs/SNRI 66%
paroxetine
Taylor et al. N ¼ 7460 N ¼ 3951 Pl A > Pl Effect size A 18% NA
(2014) 0.24
Meta- N ¼ 4559 Sertraline (S), A > S, A ¼ F, E, V Pl 19%
Agomelatine and Depressions
Agomelatine has a low incidence of nausea and dizziness and seems not to affect
sexual function, hematologic parameters, QTc interval, and weight. In studies,
agomelatine was well tolerated with no more adverse events than placebo (Stein
et al. 2014).
Data of the drug surveillance program in German-speaking countries of 184234
psychiatric inpatients revealed incidence rates of drug-induced liver injury for
agomelatine 0.33%, mianserine 0.36%, clomipramine 0.23%, and escitalopram
0.01% (Friedrich et al. 2016).
An observational cohort study 6-month evolution of 8453 depressed patients
(female: 67.7%; mean age: 49.1 years) reported cutaneous events in 1.7% of the
patients and increased hepatic transaminases values in 0.9% of the patients
(Gorwood et al. 2021). Hepatotoxic reactions like acute liver injury are an identified
Agomelatine and Depressions 1389
risk in the European risk management plan for agomelatine. Formal required are
liver function testing at baseline and 3, 6, 12, and 24 weeks. This monitoring is
mandated because of a low but important incidence of raised liver enzyme activity
and the risk of nonfatal “toxic hepatitis.” Summarizing agomelatine increased ALAT
and ASAT (liver enzymes) in 1.1% of patients (vs. 0.7% in placebo groups) (Laux
et al. 2017). On the other hand, a case-control study using data sources in Denmark,
Germany, Spain, and Sweden evaluated 3,238,495 new antidepressant and 74,440
agomelatine users. Agomelatine was not associated with an increased risk of ALI
hospitalization (Pladevall-Vila et al. 2019).
Agomelatine is contraindicated in patients with liver impairment, no
co-medication with strong CYP1A2 inhibitors like fluvoxamine or ciprofloxacine.
Clinical worsening, (hypo-) mania, and suicide risk must be monitored by all
antidepressants.
Cross-References
References
Cipriani A, Furukawa TA, Salanti G, Chaimani A, et al. Comparative efficacy and acceptability of
21 antidepressant drugs for the acute treatment of adults with major depressive disorder: a
systematic review and network meta-analysis. Lancet. 2018;391:1357–66.
Corruble E, de Bodinat C, Belaidi C, Goodwin G. Efficacy of agomelatine and escitalopram on
depression, subjective sleep and emotional experiences in patients with major depressive
1390 G. Laux
Millan MJ, Gobert A, Lejeune F, et al. The novel melatonin agonist agomelatine (S20098) is an
antagonist at 5-hydroxytryptamine2C receptors, blockade of which enhances the activity of
frontocortical dopaminergic and adrenergic pathways. J Pharmacol Exp Ther. 2003;306:954–64.
Montgomery SA, Kasper S. Severe depression and antidepressants: focus on a pooled analysis of
placebo-controlled studies on agomelatine. Int Clin Psychopharmacol. 2007;22:283–91.
Montgomery SA, Kennedy SH, Burrows GD, Lejoyeux, et al. Absence of discontinuation symp-
toms with agomelatine and occurrence of discontinuation symptoms with paroxetine: a ran-
domized, double-blind, placebo-controlled discontinuation study. Int Clin Psychopharmacol.
2004;19:271–80.
Norman TR, Olver JS. Agomelatine for depression: expanding the horizons? Expert Opin
Pharmacother. 2019;20:647–56.
Nussbaumer-Streit B, Greenblatt A, Kaminski-Hartenthaler A, et al. Melatonin and agomelatine for
preventing seasonal affective disorder. Cochrane Database Syst Rev. 2019;6:CD011271.
Olie JP, Kasper S. Efficacy of agomelatine, a MT1/MT2 receptor agonist with 5-HT2C antagonistic
properties, in major depressive disorder. Int J Neuropsychopharmacol. 2007;10:661–73.
Papakostas GI, Nielsen RZ, Dragheim M, Tonnoir B. Efficacy and tolerability of vortioxetine
versus agomelatine, categorized by previous treatment, in patients with major depressive
disorder switched after an inadequate response. J Psychiatr Res. 2018;101:72–9.
Pjrek E, Winkler D, Konstantinidis A, Willeit M, et al. Agomelatine in the treatment of seasonal
affective disorder. Psychopharmacology (Berl). 2007;190:575–9.
Pladevall-Vila M, Pottegård A, Schink T, Reutfors J, et al. Risk of acute liver injury in agomelatine
and other antidepressant users in four European countries: a cohort and nested case–control
study using automated health data sources. CNS Drugs. 2019;33:383–95.
Quera Salva MA, Vanier B, Laredo J, et al. Major depressive disorder, sleep EEG and agomelatine:
an open-label study. Int J Neuropsychopharmacol. 2007;10:691–6.
Quera-Salva MA, Hajak G, Philip P, Montplaisir J, et al. Comparison of agomelatine and
escitalopram on nighttime sleep and daytime condition and efficacy in major depressive disorder
patients. Int Clin Psychopharmacol. 2011;26:252–62.
Racagni G, Riva MA, Molteni R, Musazzi L, et al. Mode of action of agomelatine: synergy between
melatonergic and 5-HT(2C) receptors. World J Biol Psychiatry. 2011;12:574–87.
Singh SP, Singh V, Kar N. Efficacy of agomelatine in major depressive disorder: meta-analysis and
appraisal. Int J Neuropsychopharmacol. 2012;15:417–28.
Stein DJ, Ahokas A, Marquez MS, Höschl C, et al. Agomelatine in generalized anxiety disorder: an
active comparator and placebo-controlled study. J Clin Psychiatry. 2014;75:362–8.
Taylor D, Sparshatt A, Varma S, et al. Antidepressant efficacy of agomelatine: meta-analysis of
published and unpublished studies. BMJ. 2014;348:g1888.
Hypericum and Depression
Hans-Peter Volz
Contents
Chemistry, Developmental History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1394
Pharmacology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1394
Clinical Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1395
Side Effects/Adverse Effects and Toxicology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1396
Combination Therapy: Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1398
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1399
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1399
Abstract
St. John’s wort (also known as hypericum extract, HE) is a long used phythophar-
maceutical compound, showing, besides others, antidepressive properties. The
antidepressive activity of HE seems to be closely linked to only a few of the more
than 150 constituents, mainly hypericin and hyperforin. In several studies an
inhibitory effect on the mitochondrial monoamineoxidase (MAO) A and B could
be shown; also interactions with the GABAA- and GABAB-receptor were dem-
onstrated. Hyperforin also acts as a reuptake inhibitor of serotonin, dopamine,
noradrenaline, GABA, and L-glutamate. In animal studies, the antidepressive
effect of HE could be demonstrated. In several clinical trials, HE was more
effective compared to placebo and as effective as synthetical antidepressants in
mild to moderate depression. These single trials have been included in several
meta-analyses confirming the overall efficacy. The subjective tolerability is
excellent with placebo-level frequencies of adverse events, results being present
in single trials and also in several meta-analyses. Since HE might induce certain
CYP 450 isoenzymes and p-glycoprotein, pharmacokinetic interactions are
possible.
Pharmacology
Fig. 1 Chemical structure of four major constituents of HE. (According to Zirak et al. 2018)
2004; De Marchis et al. 2006; Grundmann et al. 2010; Crupi et al. 2011; Schmidt
und Butterweck 2015). In a cell-modell (transfected rat C6 glioblastoma cells) it
could be shown that hyperforin and hyperoside, both compounds of HE, lead to a
postsynaptic β-down-regulation (Jakobs et al. 2013).
Pharmacokinetic parameters are dependent on the extract used in such pharma-
cokinetic trials. Just to give one example, the results of Schulz et al. (2005) using
STW-3612 mg dry extract in 18 male patients in a single and multiple design is given
(means + standard deviation): After single administration, the AUC(01) [h∙ng/ml]
was 75.96 23.52 for hypericin, 93.03 29.40 for pseudohypericin,
1009.0 203.4 for hyperforin, and 318.70 130.82 for quercetin. Cmax [ng/ml]
was 3.14 1.57, 8.50 4.35, 83.5 27.8 resp. 47.7 22.5, tmax [h] was 8.1 1.8,
3.0 1.4, 3.0 1.4 resp. 1.17 0.52, and t1/2 [h] was 23.76 5.46, 25.39 10.18,
19.64 6.35 resp. 4.16 2.97.
Clinical Studies
Most of the randomized clinical trials (RCT) were performed in mild to moderate
depressed patients, with only some exceptions (Vorbach et al. 1997; Szegedi et al.
2005) investigating also severely depressed patients. In the meantime several meta-
analyses have been published, the most important will be summarized.
Linde et al. (2008) included a total of 29 trials (5.489 patients). Eighteen comparisons
versus placebo and 17 comparisons versus synthetic antidepressant compounds were
performed. Regarding the placebo comparisons, the combined response rate ratio was
for nine larger trials 1.28 (95% confidence interval [95% CI]: 1.10–1.49) and for nine
smaller trials 1.87 (95% CI: 1.22–2.87). This separation into larger and smaller trials
1396 H.-P. Volz
was done since the efficacy results of the placebo-HE comparisons were statistically
heterogeneous. The results of the trials comparing HE and standard antidepressants
were statistically homogeneous. Regarding the comparison with tri- or tetracyclic
antidepressants, the response rate ratios were 1.02 (95% CI: 0.90–1.15; 5 trials),
regarding the comparison with SSRIs, the response rate ratios were 1.00 (95% CI:
0.90–1.11; 12 trials). The authors state that trails from German-speaking countries
showed more favorable results for HE than trials from non-German-speaking countries.
Recently Ng et al. (2017) published another meta-analysis comprising 27 RCTs
comparing HE with SSRIs with a total of 3.808 patients. They found a comparable
response (pooled response rate ratio 0.983 [95% CI: 0.924–1.042, p < 0.001]) and
remission (pooled remission rate ratio 1.013 [95% CI: 0.892–1.134, p < 0.001]) rate
compared to standard SSRIs. The pooled standardized mean differences (SMD) from
baseline Hamilton Depression Rating Scale-scores (pooled SMD -0.068 [95% CI:
0.127–0.021, p < 0.001]) also support the significant clinical efficacy of HE.
Further reviews and meta-analyses were published by Apaydin et al. (2016),
Forsdike and Pirotta (2017), Zirak et al. (2018), Sarris et al. (2013) and Cui and
Zheng (2016) (see also Table 2).
Regarding long-term efficacy, one trial evaluated in a double-blind manner
moderately to severe depressed patients after the acute treatment phase of 6 weeks
(results reported by Szegedi et al. 2005) for another 16 weeks comparing the HE WS
5570 to paroxetine (Anghelescu et al. 2006) and found no differences between the
groups; however, the design was not suited to detect equal efficacy. In a placebo-
controlled long-term trial (Kasper et al. 2008), a state-of-the art design to show
efficacy (and tolerability) in the continuation and prophylactic (maintenance) phase
of long-term treatment was used. After an initial open treatment phase with WS 5570
for 6 weeks, responders were randomized to WS 5570 (3X300 mg/day) or placebo
and treated in a double-blind manner for an additional 26 weeks (continuation
phase), stable patients on WS 5570 were then again randomized to WS 5570 resp.
placebo, stable patients who have been already in the placebo group continued with
placebo for another 52 weeks (prophylaxis/maintenance phase). WS 5570 was very
efficacious (time to relapse, relapse rates) in the continuation phase, but in the
prophylaxis phase this superiority was only present initially in severely ill patients.
Regarding efficacy, it is noteworthy that – since the constituents of HE differ
considerably between the individual manufacturers – the efficacy cannot be extrap-
olated from one extract to another. As Kasper et al. (2010a) showed, WS 5572, LI
160, WS 5570, and ZE 117 are significantly more efficacious than placebo and at
least as efficacious with better tolerability than synthetic antidepressants, which is
also the case for STW3/-VI (Uebelhack et al. 2004; Gastpar et al. 2005, 2006).
Table 2 Reviews and meta-analyses mentioned in the text with the most important details
regarding methods and results
Author Type Method Result
Linde Meta-analysis A total of 29 studies with 5,489 HE showed superior efficacy
et al. patients have been included compared to placebo and equal
(2008) efficacy compared to
synthetical antidepressants.
Tolerability of HE was
superior to synthetical
antidepressants
Ng et al. Meta-analysis A total of 35 studies with 6,993 HE showed superior efficacy
(2017) patients have been included compared to placebo and equal
efficacy comped to SSRI.
Tolerability of HE was better
compared to SSRI
Forsdike Scoping 13 studies including 5,183 With the exception of
and review about physicians/care providers and Germany St. John’s wort was
Pirotta attitudes and a total population of 1,903,649 seldomly used due to little
(2017) experiences knowledge about efficacy data
of the single extracts
Zirak Narrative Available data of in vitro, HE may exert potent
et al. review in vivo, and clinical evidence antidepressant effects. It seems
(2018) to be also safe
Sarris Narrative Outlines the current evidence Evidence supports the use of
(2013) review of the efficacy of HE in HE for mild to moderate
common psychiatric disorders, depression and somatization
not only in depression, but also disorder, but not in other
in other disorders. Mechanisms psychiatric disorders.
of action, including emerging Differences in the quality and
pharmacogenetic data, safety, safety of HE need to be
and clinical considerations are addressed in recommending
also detailed extracts to the patients.
Potential interactions should
be kept in mind
Cui and Meta-analysis 27 studies with a total of 3,126 Equal efficacy of HE compared
Zheng patients have been included to SSRI and superior
(2016) tolerability
Abbr.: HE hypericum extract, SSRI serotonin reuptake inhibitor
trials due to adverse effects less frequently than those given older antidepressants
(odds ratio [OR]: 0.24 [95% CI: 0.13–0.46]) or SSRIs (OR: 0.53 [95% CI: 0.34–
0.83]). Similar results were found by Ng et al. (2017) who showed significantly
lower discontinuation/dropout rates compared to standard SSRIs (pooled OR 0.587
[95% CI: 0.478–0.697], p < 0.001).
The differential side effect profile in comparison to SSRIs was analyzed by
Kasper et al. (2010b), based on four RCTs (with a total of 1.661 outpatients
included). The HE WS 5570 (dose between 600 and 1.800 mg/day) was compared
to 20–40 mg/day paroxetine resp. to placebo. Across the four trials the percentage of
patients with an adverse event (AE) under HE was comparable to placebo (risk ratio
1398 H.-P. Volz
Table 3 Symptom clusters of adverse events (%) of patients affected and risk ratio with 95%
confidence interval (95& CI)
Placebo
WS 5570 Placebo Paroxetine vs. WS Paroxetine
(n ¼ 1.264) (n ¼ 271) (n ¼ 126) 5570 vs. WS 5570
Symptom
clusters N (%) N (%) N (%) RR (95% CI)
Sedationa 20 (1.6) 1 (0.4) 26 (20.6) 0.23 (0.04– 13.04 (7.51–
1.35) 22.44)
Anticholinergic 11 (0.9) 4 (1.5) 42 (33.3) 1.70 (0.57– 38.30 (20.42–
reactionsb 4.99) 71.72)
Typical SSRI 63 (5.0) 20/7.4) 52 (41.3) 1.48 (0.91– 8.26 (5.99–
symptomsc 2.38) 11.31)
Sexual 0 (0.0) 0 (0.0) 3 (2.4) – (7.84–)
dysfunctiond
Heart rhythme 2 (0.2) 0 (0.0) 2 (1.6) 0.00 (0.00– 10.03 (1.77–
8.92) 56.32)
Hypotension/ 11 (0.9) 3 (1.1) 28 (22.2) 1.27 (0.38– 25.54 (13.15–
vertigof 4.19) 49.37)
According to Kasper et al. (2010b)
a
Fatigue, somnolence, insomnia, flat affect
b
Constipation, constipation aggravated, dry mouth
c
Headache, migraine, headache aggravated, diarrhea, nausea, vomiting, insomnia, initial insomnia,
agitation, restlessness, irritability, nervousness
d
Libido decreased, loss of libido, ejaculation disorders
e
Bradycardia, palpitation
f
Hypotension, hypotension postural, dizziness, dizziness (ex vertigo), vertigo
[RR]: 1.1 [95% CI: 0.9–1.3]) and significantly lower than for paroxetine (RR: 2.4
[95% CI: 2.1–2.8]). For further details, see Table 3.
Regarding single AEs of HE, rare cases of gastro-intestinal disorders, allergic
skin reactions, tiredness, and restlessness are reported (Russo et al. 2014). Photo-
sensitivity seems only to occur after very high doses of HE and/or intense sun light
(Schulz et al. 2006).
Cross-References
References
Anghelescu IG, Kohnen R, Szegedi A, Klement S, Kieser M. Comparison of Hypericum extract
WS 5570 and paroxetine in ongoing treatment after recovery from an episode of moderate to
severe depression: results from a randomized multicenter study. Pharmacopsychiatry.
2006;39:213–9.
Apaydin EA, Maher AR, Shanman R, Booth MS, Miles JN, Sorbero ME, Hempel S. A systematic
review of St. John’s wort for major depressive disorder. Syst Rev. 2016;5:148.
Barnes J, Anderson LA, Phillipson JD. St John’s wort (Hypericum perforatum L.): a review of its
chemistry, pharmacology and clinical properties. J Pharm Pharmacol. 2001;53:583–600.
Berry-Bibee EN, Kim MJ, Tepper NK, Riley HE, Curtis KM. Co-administration of St. John’s wort
and hormonal contraceptives: a systematic review. Contraception. 2016;94:668–77.
Crupi R, Mazzon E, Marino A, La Spada G, Bramanti P, Battaglia F, Cuzzocrea S, Spina
E. Hypericum perforatum treatment: effect on behaviour and neurogenesis in a chronic stress
model in mice. BMC Complement Altern Med. 2011;11:7.
Crupi R, Abusamra YA, Spina E, Calapai G. Preclinical data supporting/refuting the use of
Hypericum perforatum in the treatment of depression. CNS Neurol Disord Drug Targets.
2013;12:474–86.
Cui YH, Zheng Y. A meta-analysis on the efficacy and safety of St John’s wort extract in depression
therapy in comparison with selective serotonin reuptake inhibitors in adults. Neuropsychiatr Dis
Treat. 2016;12:1715–23.
De Marchis GM, Bürgi S, Kientsch U, Honegger UE. Vitamin E reduces antidepressant-related
beta-adrenoceptor down-regulation in cultured cells. Comparable effects on St. John’s wort and
tricyclic antidepressant treatment. Planta Med. 2006;72:1436–7.
Forsdike K, Pirotta M. St John’s wort for depression: scoping review about perceptions and use by
general practitioners in clinical practice. J Pharm Pharmacol. 2017;71:117–28.
Gastpar M, Singer A, Zeller K. Efficacy and tolerability of hypericum extract STW3 in long-term
treatment with a once-daily dosage in comparison with sertraline. Pharmacopsychiatry.
2005;38:78–86.
Gastpar M, Singer A, Zeller K. Comparative efficacy and safety of a once-daily dosage of
hypericum extract STW3-VI and citalopram in patients with moderate depression: a double-
blind, randomized, multicentre, placebo-controlled study. Pharmacopsychiatry. 2006;39:66–75.
1400 H.-P. Volz
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1402
Pharmacology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1403
Constituents of Ginseng and Phytochemical Classification of Ginsenoside . . . . . . . . . . . . . . 1403
Pharmacological Activities in the Brain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1404
Antidepressant Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1406
Pharmacokinetics of Ginseng and Ginsenosides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1410
Side Effects/Adverse Reactions and Toxicology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1410
Drug-Drug Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1411
Clinical Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1411
Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1414
Financial and Competitive Interests Disclosure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1415
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1415
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1415
Abstract
Panax ginseng Meyer is a well-known traditional Chinese medicine and has been
proposed to heal any disease. Ginseng and the major ingredient saponins
(ginsenosides) have multiple pharmacological activities, such as antioxidant, anti-
aging, anti-inflammatory, and immunomodulatory activity and are used as anti-
depressive, anti-stress, anti-fatigue, and anxiolytic agents in depressive state. The
antidepressant-like functions have been proved in animal models of depression
prepared by physical and psychological stress. Antidepressant function of ginseng
is mainly due to modulation of the hypothalamic–pituitary–adrenal axis and
immune system and anti-inflammation rather than modulation of monoamine
transmitter systems, brain-derived neurotrophic factor expression, and hippocampal
neurogenesis in contrast to other phytochemicals. The clinical trials of ginseng have
Introduction
Pharmacology
Fig. 1 Chemical structure of major ginsenosides and the metabolism pathway. (a) Structure of PPD
and PPT. (b) Related derivatives. (c) Metabolic pathway of ginsenosides by intestinal microflora.
Rg3 and Rg1 have potent antidepressant activity and Rb1 anxiolytic activity. Glc, β-D-
glucopyranose; Ara(p), α-L-arabinose (pyranose); Ara(f), α-L-arabiose (fucose); Rha, α-L-
rhambopyranose; Xyl, β-D-xylopyranose
1404 M. Naoi et al.
(PPDs), protopanaxatriols (PPTs), and oleanolic acid (ginsenoside Ro). PPDs have
the sugar moieties attached to the 3-position of tripeptide, and include ginsenoside
Rb1, Rb2, Rb3, Rc, Rd., Rg3, Rg5, Rh2, and C-K, and notoginsenoside R1, whereas
PPTs have sugar moieties at the 6-position and include ginsenoside Re, Rg1, Rg2,
and Rf. Among 150 different types of ginsenosides, PPDs constitute more than 90%
of the total ginsenosides in Panax ginseng. Major bioactive ginsenosides include
Rb1, Rb2, Rb3, and notoginsenoside Fc and contain three to five sugars, whereas
minor ones are Rg3, Rh2, C-K, and 20-protopanaxaiol (20-PPD). The bioactivity
depends on the glycosylation patterns; the type, position, and number of sugar
moieties attached to the glycosidic bond at C-3 and C-6. Each ginsenoside has
different action mechanisms and tissue-specific effects, and exerts multiple func-
tions. Ginsenoside contents vary according to Panax species, the age and parts of
ginseng, season of harvest, and preservation and extraction methods.
Other constituents of ginseng also show pharmacological activities. Acidic poly-
saccharides, pectin combined with galacturonic acid, glucuronic acid, and mannuronic
acid, have antioxidant, anti-fatigue, immune-modulating, and antiproliferative effects
and show neuroprotection by activation of ERK/MAPK signaling pathway in cultured
neuronal cells. Polyacetylenes are unsaturated triple-bonded polyalcohols, including
panaxynol (heptadeca-1,9-diene-4,6-dine-3-ol), panaxydol (heptadeca-1-ene-9,10-
epoxy-4,6-dinye-3-ol), panaxytriol, panacacol, and panaxylene epoxide, and have
antioxidant and anticancer effects. Gintonin, a ginseng glycolipoprotein complex,
contains lysophosphatidic acids, linoleic acid, and lysophosphatidyl-inositol. Gintonin
binds to the lysophosphatidic acid (LPA) receptors, the G-protein coupled receptors
(GPCRs), GPR40 and GPR55; activates phosphatidylinostol-3-kinase (PI3K)/AKT
signaling pathway; exerts neuroprotection; and ameliorates cognition impairment in
Alzheimer’s, Parkinson’s, and Huntington’s diseases.
Ginseng and ginsenosides stimulate or inhibit the brain functions and protect
neurons in neurodegeneration and ischemia. They have anti-inflammatory and
immunomodulatory functions, promotion of neurotrophic factors (NTFs) expression
and neurogenesis, phytoestrogen-like activities, and modulation of neurotransmitter
systems [serotonin (5-hydroxytryptamine, 5-HT), noradrenalin (NA), dopamine
(DA), GABA, glutamate, acetylcholine] in animal models. They also affect cardio-
vascular, antineoplastic, and aphrodisiac effects, which will not be discussed here. In
general, PPD derivatives have neuroprotective function, whereas PPT derivatives
antidepressant activities. Rb3 compromises over 15% of Panax notoginseng and has
antidepressant effects in animal models. Rb1 and its metabolite C-K are associated
with the most pharmacological activities in cardiovascular, endocrine, and immune
system and neurodegenerative diseases. Rb1 is an active anxiolytic compound. Rb2
has anticancer, anti-adipocytic, antidiabetic, and antioxidant functions in animal
models. Rg3 is enriched in Korean red ginseng and has neuromodulatory, cognition
enhancing, anti-inflammatory, antioxidant, antiangiogenic, and anticancer activities.
Ginseng and Ginsenosides in Depression 1405
The 20(S)-epimer of Rg3 more easily dissolves in water than the 20(R)-, and has
proapoptotic and anticancer activity. Rg1 is the primary ingredient, induces NTFs,
promotes neurogenesis, and has sex hormone-like activities to exert potent antide-
pressant function. Rh3, Rc, Rg1, Rg2, and Rh1 have antiaging, memory promoting,
and neuroprotective activities in animal models, but clinical trials have presented
only insufficient evidences for the effectiveness.
Pharmacological function of ginseng is characterized by the intracellular signal
transduction mediated by non-selective receptors or channels without any intracellular
signal transduction pathway (Nah 2014). Ginsenosides inhibit cation influx, stimulate
anion influx, and decrease the excitability of neuronal cells (Fig. 2). Ginsenosides
modulate voltage-gated ion channel. Rg3 and Rh2 affect Na+ channel, and ginseng,
Rb1, and Rg1 L-type voltage-gated Ca2+ channel (VGCC) and show neuroprotection.
Fig. 2 Ion channels, receptors, and intracellular signal transduction of ginsenosides for antide-
pressant function. Ginsenosides regulate voltage-gated ion channel, such as voltage-gated calcium
channel (VGCC) and voltage-gated sodium (Na+) channel (VGNC), and also ligand-gated ion
channel, such as GABAA and glycine receptor, NMDA, and nicotinic acetylcholine receptor.
Gintonin binds to G-protein-coupled LPA receptors and activates p38/MAPK pathway. Signaling
pathways activate transcription and enhance BDNF expression, or suppress the expression of
cytokines to exert anti-inflammation. ERα and ERβ, estrogen receptors; RXRα-PPARγ, retinoid
X receptor α-peroxisome-proliferating receptor γ; PGC-1α, proliferator-activated receptor γ
coactivator-1a
1406 M. Naoi et al.
Rh2, Rg3, C-K, PPT, and Rd. inhibit cation-gated N-methyl-D-aspartate (NMDA)
receptors and neuronal Na+ channels and stimulate anion-gated GABAA receptors and
glycine receptors to show anxiolytic effects. Rg3 regulates also ligand-gated ion
channels, 5-HT, nicotine acetylcholine, and NMDA receptors; affects nuclear retinoid
X receptor α-peroxisome-proliferating receptor γ (RXRα-PPARγ); activates PPAR
signaling; and has anti-inflammation activity. Ginsenosides bind also to estrogen
receptor (ER), glucocorticoid receptor (GR), and insulin-like growth factor-1 (IGF-1)
receptors and activate tyrosine kinase, PI3K/Akt and ERK1/2, p38 kinases. PPD
derivatives (Rh2, C-K) have higher affinity to GR than the PPT (Re, Rg1), whereas
Rh1 and Rg3 function as ER ligands. Gintonin binds to G protein-coupled LPA
receptors and activates MAPK, PI3K, PKC, and Rho kinases through cytosolic
calcium as a second messenger. Ginsenosides activate these receptors, induce brain-
derived neurotrophic factor (BDNF), promote hippocampal neurogenesis, and have
anti-inflammation and anti-fatigue activity. On the other hand, bioactive polyphenols,
such as flavonoids, bind to tyrosine kinase B (TrkB), the BDNF receptor, and other
receptors coupled with downstream mediators to have trophic effects, including the
formation, stabilization, and potentiation of neuronal architecture.
Antidepressant Functions
by interfering with glycogen synthase kinase (GSK)-3β and induced BDNF expres-
sion in the prefrontal cortex and hippocampus of rodent depression models. G115
and ginsenosides (Rg1, Rg2, Rb1, Rb3, Rg3, Rg5) also increase BDNF and exert
antidepressant effects by activation of the BDNF/TrkB-CREB pathway. Rg1
increased BDNF expression and reversed decrease in hippocampal neurogenesis
and dendritic spine density in depression mouse models (Jiang et al. 2012).
Sesquiterpenoids from Panax ginseng root showed antidepressant effects by activa-
tion of BDNF/TrkB/NF-κB signal pathway, in addition to modulation of DA,
GABA, and glutamine systems.
In general, ginseng is safe and the root of Panax ginseng appears nontoxic in human
when ginseng consumption is limited under 1–2 g ginseng containing 4–5%
ginsenosides per day. However, ginseng is stimulant and may cause nervousness
or insomnia, if taken at high doses. “Ginseng-abuse syndrome” was described in 14
of 133 long-term ginseng users (3 g/day Panax ginseng root material), presenting
side effects of hypertension, restlessness, and skin rash. In a 15 g/day ginseng user,
confusion, depression, or depersonalization was observed (Siegel 1979). There were
five cases (two females, three males) of ginseng-associated mania, and the reported
Ginseng and Ginsenosides in Depression 1411
daily doses of ginseng were quite higher than recommended doses. Two cases of
new onset of manic psychosis without any prior psychiatric history were reported
after daily ginseng use and manic symptoms remitted within days by discontinuation
of ginseng (Norelli and Xu 2015). These results suggest that ginseng ingredients
might be psychotropic and affect mood function especially in BD. Other adverse
effects are related to the estrogen-like activity, such as vaginal bleeding, and
ginsenosides may effect particularly pregnant or breastfeeding women.
Drug-Drug Interactions
Drug transporters, such as protein organic anion transporter (OAT) and multiple drug
resistance (MDR), and drug metabolizing enzymes, cytochrome P450s (CYPs) and
P-glycoprotein (P-gp), play a vital role in modification of pharmacokinetics and dug-
drug interaction. Ginseng and ginsenosides are weak inhibitors of CYPs and trans-
porters, and severe harmful drug-ginseng interaction has never been reported in
clinical studies. During the use of ginseng, serotonin syndrome was detected in
patients treated with SSRI or SNRI (escitalopram, fluoxetine, paroxetine) and
ventricular arrhythmias in patients with haloperidol (Woron and Siwek 2018).
Adverse effects of ginseng and ginsenosides on chemotherapeutic, antiplatelet,
anticoagulant, and antidiabetic drugs have been never presented. For treatment of
fatigue in adjuvant chemotherapy, ginseng improved fatigue-related syndrome with-
out causing critical adverse events, but did not change survival rates and tumor
markers. Panax ginseng was reported to induce CYP3A activity in liver and
gastrointestinal tract, suggesting the interaction with CYP3A or P-gp substrates. In
rats, the co-administration of ginsenosides (Rg1, Re, Rd., Rf, Rb1, Rc, Rb2) and
warfarin the most common anticoagulant increased 7-hydroxywarfarin, a metabolite
by CYP2C9 and CYP3A4, in liver and attenuated anticoagulation. But, the clinical
data on the interaction of ginsenosides with warfarin were not consistent.
Deglycosylated ginsenosides increase bioavailability and elimination half-lives
and their interaction with drugs should be further investigated.
Clinical Studies
Clinical trials of ginseng and ginsenosides have been scarcely reported in depression
(Table 2). Siberian ginseng (Eleutherococcus senticosus), whose active constituents
are different from Panax ginseng, showed antidepressant activities in patients with
BD (Weng et al. 2007). Ginseng adjunctive administration to lithium improved 17-
Item-Hamilton Depression Rating Scale (HAMD-17) score, but the difference was
not statistically significant from lithium group. Korean red ginseng significantly
improved depressive state in women with residual symptoms of MDD (Jeong et al.
2015). Estrogen-like activity of Korean red ginseng is associated with improvement
of severe climacteric syndromes, including fatigue, insomnia, and depression in
postmenopausal subjects (Tode et al. 1999), but later trials with fermented red
Table 2 Representative clinical trials of ginseng and ginsenosides in patients with depressive diseases and chronic fatigue
1412
Adverse
Study Subjects. Sample size Aim and design Results Dropout rate effects
PCT 12 PM women with climacteric KRG (6 g/d 30 days) Improved CMI, STAI-A score
Tode et al. syndrome
(1999) (Control; 8 without syndrome)
Japan
DBRPCT 30 PM women Ginsosana No significant effects on HAD 11/30
Hartley et (Placebo, 27) (320 mg/d 12 weeks)
al. (2004) UK
DBRPCT BD patients Adjunct effects to lithium, Improved HAMD-17 Nausea,
Weng et al. 37 (Li + SG; f/m ¼ 21/16), 39 compared with fluoxetine rash
(2007) (Li + fluorexetine; f/m ¼ 24/15) Siberian ginseng
China (750 mg/d 6 weeks)
DBRPCT Healthy young adults G115 Improved working memory
Reay et al. 30 volunteers (200, 400 mg 8 days) Increased calmness
(2010) (f/m ¼ 15/15)
UK
Direct 35 MDD women with residual Effects as adjunct of SSRI, Improved DRSS, MADRS, CGI- 4 /35 GI
evaluation syndrome SNRI S scores syndromes
Jeong et al. South Korea KRG (3 g/d 4 weeks) Headache,
(2015) insomnia
DBRPCT Post-stress KRG No improved mood 4/32
Baek et al. 32 (f/m ¼ 15/17) (2 g/d 6 weeks) (4/31)
(2019) Control; 31 (f/m ¼ 16/15)
South Korea
DBRPT Idiopathic chronic fatigue KG extract Improved VAS score In 2 g/d Ginseng 1 g/d *AR
Kim et al. Ginseng 1 g 30 (1, 2 g/d 4 weeks) group group, 1/30
(2013) Ginseng 2 g 30 Improved NRS score in 1, 2 g/d Ginseng 2 g/d
Placebo 30 groups group, 1*/30
South Korea Placebo, 0/30
M. Naoi et al.
DBRPT Cancer-related fatigue Wisconsin ginsengb Improved MFSI-SF score Ginseng, 38/ 16 AR
Barton et Ginseng 171 (2000 mg/d 8 weeks) 171 13 AR
al. (2013) Placebo 170 Placebo 42/170
USA
DBRPT Volunteers 180 KRG Improved fatigue self-assessment Placebo, 4/60 1 AR
Zhang et Placebo 60 (f/m ¼ 30/21) (1.8 g, 3.6 g/d 4 weeks) score 1.8 g/d, 0/60 1 common
al. (2019) KRG 1.8 g 60 (f/m ¼ 43/19) 3.6 g/d, 2*/60 cold
KRG 3.6 g (f/m ¼ 43/17)
China
DBRPCT 15, Patients with chronic fatigues Fatigue No effects on VAS, BDI scales 1/25
Sung et al. (15 Placebo) KRG (3 g/d 6 weeks) Improved only in aged (Placebo, 2/25)
(2020) South Korea (>50 years) with moderate
fatigue
a
Ginseng and Ginsenosides in Depression
ginseng containing high C-K could not approve the beneficial effects. A Panax
ginseng and Ginkgo biloba combination (gincosan) therapy improved physiological
and cognition function in humans, but the synergy could not be confirmed. Effects of
Korean red ginseng were investigated in individuals exposed to high stress, but
beneficial effects could not be confirmed (Baek et al. 2019). Wisconsin ginseng
(Panax quinquefolius), Korean red ginseng, and ginsenosides improved chronic
fatigue-related symptoms, including quality of life and mood, but did not increase
physical activities (Sung et al. 2020). Meta-analyses of reported clinical results of
ginsenosides have not presented conclusive evidences for the effects in depression
and mental and physical fatigue.
Discussion
Ginseng the most popular TCM is still enigmatic medicine and the pharmacological
characteristics are hard to clarify from aspect of modern medicine. Clinical results on
antidepressant potency of ginseng and ginsenosides in MDD and BD should be
tempered because of small sample size and brief intervention duration, and the
benefit has been not fully approved. Ginseng and ginsenosides could improve
stress-induced depressive state in postmenopausal women and individual with
chronic fatigues, cancer, and frailty, which might be caused by the decline of “Qi”
according to the term of the TCM. Ginseng contains various multifunctional active
components, and they may afford ginseng the synergistic pharmacological effects.
The effects on “Qi” have been commonly evaluated by the holistic TCM criteria
based on symptomatic observation of patients, and not by reductionism- and evi-
dence-based modern medicine. Clinical studies of ginseng and ginsenosides should
be further continued by use of the consistent and quality-controlled samples and
more objective quantitative determination of the effects in more systematically
designed trials. The structure-activity relationship of ginsenosides should be further
investigated in order to develop new antidepressant agents.
The very poor bioavailability of ginsenosides in the brain may reduce the
antidepressant activity of ginsenosides. Co-administration of Ginkgo biloba extract
increased the permeability of ginsenoside (Rg1, Re, Rd., and Rb1) across the BBB
and improved the uptake in the brain. Ginkgo biloba extract activated A1 adenosine
receptor signal pathway, phosphorylated the BBB tight junction-relate proteins
(ezrin/radixin/moesin, myosin light chain), and increased the BBB permeability.
Modification of ginsenoside properties has been tried to increase bioavailability.
Highly hydrophilic properties of Rb1 inhibited direct absorbance in human body and
the cleavage of the four glycosides of 3-O-Glc-Glc- and 20-O-Glic-Glc- substitutes
by enzymatic and chemical reactions increased the bioavailability and activity. Rb1
was modified by hydrogenation, acetylation, and epoxidation, which enhanced the
potency to promote cell proliferation in ARPR-19 cells. Chemical sulfation of Rh2
increased water solubility and anti-inflammatory activities. The 3-hydroxy group of
PPD was modified with fatty acids and diacids, and the esters increased anticancer
activity. At present, it has not been reported whether such modification can promote
Ginseng and Ginsenosides in Depression 1415
Cross-References
References
Arring NM, Millstine D, Marks LA, Nal LM. Ginseng as treatment for fatigue: a systematic review.
J Altern Complement Med. 2018;24(7):624–33.
Baek JH, Heo JY, Fava M, Mischoulaon D, Choi KW, Na EJ, Cho H, Jeon HJ. Effect of Korean red
ginseng in individuals exposed to high stress levels: a 6-week, double-blind, randomized,
placebo-controlled trial. J Ginseng Res. 2019;43(3):402–7.
Barton DL, Liu H, Dakhi SR, et al. Wisconsin ginseng (Panax quinquefolius) to improve cancer-
related fatigue: a randomized, double-blind trial, N07C2. J Natl Cancer Inst. 2013;105(16):
1230–8.
Choi K. Botanical characteristics, pharmacological effects and medical components of Korean
Panax ginseng C A Meyer. Acta Pharmacol Sin. 2008;29(9):1109–18.
Hartley DE, Elsabagh S, File SE. Ginsosan (a combination of Ginkgo biloba and Panax ginseng):
the effects on mood and cognition of 6 and 12 weeks’ treatment in post-menopausal women.
Nutr Neurosci. 2004;7(5–6):325–33.
Jeong HG, Ko YH, Oh SY, Han C, Kim T, Joe SH. Effect of Korean red ginseng as an adjunct
treatment for women with residual symptoms of major depression. Asia Pac Psychiatry. 2015;
7(3):330–6.
Jiang B, Xiong Z, Yang J, Wang W, Wang Y, Hu ZL, Wang F, Chen JG. Anti-depressant-like effects
of ginsenoside Rg1 are due to activation of BDNA signalling pathway in the hippocampus. Br J
Pharmacol. 2012;166(6):1872–87.
Jin Y, Cui R, Zhao L, Fan J, Li B. Mechanisms of Panax ginseng action as an antidepressant. Cell
Prolif. 2019;52(6):e12606.
1416 M. Naoi et al.
Kim HG, Cho JH, Yoo AR, Lee JS, Han JM, Lee NH, Ahn YC, Son CG. Antifatigue effects of
Panax ginseng C.A. Mayer: a randomised, double-blind, placebo-controlled trial. PLoS One.
2013;8(4):e61271.
Lee SM, BAE B, Park H, Ahn N, Cho B, Cho Y, Kwak Y. Characterization of Korean red ginseng
(Panax ginseng Meyer): history, preparation method, and chemical composition. J Ginseng Res.
2015;39(4):384–91.
Nah SY. Ginseng ginsenoside pharmacology in the nervous system: involvement in the regulation
of ion channels and receptors. Front Physiol. 2014;5:98.
Norelli LJ, Xu C. Manic psychosis associated with ginseng: a report of two cases and discussion of
the literature. J Diet Suppl. 2015;12(2):119–25.
Reay JL, Scholey AB, Kennedy DO. Panax finseng (G115) improves aspect of working memory
performance and subjective rating of calmness in healthy young adult. Hum Psychopharmacol
Clin Exp. 2010;2596:462–71.
Siegel RK. Ginseng abuse syndrome. Problems with the panacea. JAMA. 1979;241(15):1614–5.
Sung WS, Kang HR, Jung CY, Park SS, Lee SH, Kim EJ. Efficacy of Korean red ginseng (Panax
ginseng) for middle-aged and moderate level of chronic fatigue patients: a randomized, double-
blind, placebo-controlled trial. Complement Ther Med. 2020;48:102246.
Tode T, Kikuchi Y, Hirata J, Kita T, Nakata H, Nagata I. Effect of Korean red ginseng on
psychological functions in patients with severe climacteric syndromes. Int J Glynecol Obstet.
1999;67(3):169–74.
Weng S, Tang J, Wang G, Wang X, Wang H. Comparison of the addition of Siberian ginseng
(Acanthopanax senticosus) versus fluoxetine to lithium for the treatment of bipolar disorder in
adolescents: a randomized, double-blind trial. Curr Ther Res Clin Exp. 2007;68(4):280–90.
Woron J, Siwek M. Unwanted effects of psychotropic drug interactions with medicinal products and
diet supplements containing plant extracts. Psychiatr Pol. 2018;52(6):983–96.
Yu SE, Mwesige B, Yi Y, Yoo BC. Ginsenosides: the need to move forward from bench to clinical
trials. J Ginseng Res. 2019;43(3):361–7.
Zhang H, Li Z, Zhou Z, Yang H, Zhong Z, Lou C. Antidepressant-like effects of ginsenosides: a
comparison of ginsenoside Rb3 and its four deglycosylated derivatives, Rg3, Rh2, compound K
and 20(S)-protopanaxadiol in mice model of despair. Pharmacol Biochem Behav. 2016;140:17–26.
Zhang L, Chen X, Cheng Y, et al. Safety and antifatigue effect of Korean red ginseng: a randomized,
double-blind, and placebo-controlled clinical trial. J Ginseng RE. 2019;43(4):676–83.
Zhao YN, Shao X, Ouyang LF, Chen L, Gu L. Quantitative detection of ginsenosides in brain
tissues after oral administration of high purity ginseng total saponins by using polyclonal
antibody against ginsenosides. Chin J Nat Med. 2018;16(3):175–83.
Vortioxetine and Depressions
Gerd Laux
Contents
Chemistry, Developmental History (Fig. 1) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1417
Pharmacology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1418
Pharmacokinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1418
Mechanism of Action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1418
Indications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1419
Clinical Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1419
Side Effects, Adverse Reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1421
Combination Therapy: Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1421
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1422
Abstract
Vortioxetine is a multimodal antidepressant acting on serotonin (5-HT) receptors
in several ways: as an antagonist on 5-HT3, 5-HT7, and 5-HT1D receptors, as a
partial agonist on 5-HT1B receptors, and as an agonist on 5-HT1A receptors;
furthermore, it inhibits the 5-HT transporter. The efficacy in treatment of major
depressive disorder was established in randomized, double-blind, placebo-con-
trolled studies and a maintenance study in adult in- and outpatients. In preclinical
animal studies as well as clinical studies, vortioxetine showed positive effects on
learning and memory. Dose-related nausea was the most common adverse
reaction.
G. Laux (*)
Institute of Psychological Medicine (IPM), Soyen, Germany
MVZ Waldkraiburg, Center of Neuropsychiatry, Waldkraiburg, Germany
Department of Psychiatry and Psychotherapy, Ludwig-Maximilians-University (LMU), Munich,
Germany
e-mail: ipm@ipm-laux.de
Pharmacology
Pharmacokinetics
Mechanism of Action
Indications
Clinical Studies
Study arms,
comparators, Dropout
Study N, Patients placebo Results rates Relevant side effects
Alvarez et al. 429 pts. Pl, venlafaxine Response rate V 68%, venla 72%, Pl 4%, vort Nausea, hyperhidrosis,
2012 EU, Asia, Canada XR 225 mg/d Pl 45% 10%, venla dry mouth
RCT 14%
Baldwin et al. 766 pts. Pl, duloxetine V > pl Pl 8%, vort Nausea, dizziness, dry
2012 EU, Asia, Canada 60 mg/d V ¼ duloxetine 9%, dulox mouth
RCT 12%
Katona et al. 2012 452 elderly pts. (mean age 70 years) Pl, duloxetine V response rate 53%, pl 35%, Pl 3%, Nausea (22%), dulox
RCT 60 mg/d dulox 63% dulox 10%, nausea, constipation,
V > Pl in cognition tests of speed vort 6% dry mouth, somnolence
of processing, verbal learning, and
memory
Mahableshwarkar 614 pts Pl, duloxetine V > Pl in MADRS and cognitive Nausea, diarrhea
et al. 2015a, b US multicenter 60 mg/d tests
RCT
McIntyre et al. Pl, duloxetine V improved cognition
2016 independent of depressive
Meta-analysis 3 symptoms
RCTs
Citrome 2016 8 studies for duloxetine, 3 for escitalopram, 5 V ¼ comparison drugs Overall tolerability
Indirect for levomilnacipran, 1 for sertraline, 4 for NNTs for response vs placebo (NNHs
comparison venlafaxine, 2 for vilazodone, 11 for 6–10 V > comparators
vortioxetine
Vieta et al. 2018 N ¼ 101 Escitalopram V favored for improvements of Nausea
RCT Nonresponders 10–20 mg/d cognition, functioning, and mood
symptoms
Papakostas et al. N ¼ 493 V 10–20 mg/d V > Agomelatine V 21% Rare side effects,
2018 SSRI/SNRI nonresponders Agomelatine Ago 26% Agomel > V
EU multicenter 25–50 mg/d
G. Laux
Vortioxetine and Depressions 1421
Fig. 3 shows the results of a pivotal study comparing vortioxetine with venlafaxine regarding
MADRS difference scores (Alvarez et al. 2012)
depressive symptoms (Al-Sukhri et al. 2015, McIntyre et al. 2016). Data of one
study are summarized in Fig. 4 (Herrera-Guzman et al. 2009).
In systematic reviews’ perspective, Cochrane Database concluded the place of
vortioxetine in the treatment of acute depression to be unclear; in comparison to
SNRIs no advantage has been found; major limitation is the lack of comparisons
with the first-line treatment SSRIs (Koesters et al. 2017).
The most common adverse reactions (incidence >5% and at least twice the rate of
placebo) are nausea, constipation, and vomiting. Table 2 shows the side effect (SE)
profile compared to venlafaxine and placebo (Alvarez et al. 2012).
Clinical worsening, (hypo-) mania, and suicide risk must be monitored by all
antidepressants.
Sexual dysfunction is common with SSRIs, vortioxetine showed benefits com-
pared to escitalopram, and rates were not significantly different from placebo in
short-term clinical trials (Jacobsen et al. 2015a, b, 2016).
References
Alvarez E, Perez V, Dragheim M, Loft H, Artigas F. A double-blind, randomized, placebo-
controlled, active reference study of Lu AA21004 in patients with major depressive disorder.
Int J Neuropsychopharmacol. 2012;15:589–600.
Vortioxetine and Depressions 1423
Al-Sukhri M, Maruschak NA, McIntyre RS. Vortioxetine: a review of efficacy, safety and tolera-
bility with a focus on cognitive symptoms in major depressive disorder. Expert Opin Drug Saf
2015;14:1291–304.
Areberg J, Sogaard B, Hojer AM. The clinical pharmacokinetics of LuAA21004 and its major
metabolite in healthy young volunteers. Basic Clin Pharmacol Toxicol. 2012;111:198–205.
Baldwin DS, Loft H, Dragheim M. A randomised, double-blind, placebo controlled, duloxetine-
referenced, fixed-dose study of three dosages of Lu AA21004 in acute treatment of major
depressive disorder (MDD). Eur Neuropsychopharmacol. 2012;22:482–91.
Baldwin DS, Florea I, Jacobsen PL, Zhong W, Nomikos GG. A meta-analysis of the efficacy of
vortioxetine in patients with major depressive disorder (MDD) and high levels of anxiety
symptoms. J Affect Disord. 2016;206:140–50.
Berhan A, Barker A. Vortioxetine in the treatment of adult patients with major depressive disorder: a
meta-analysis of randomized double-blind controlled trials. BMC Psychiatry. 2014;27:276.
Boulenger JP, Loft H, Florea I. A randomized clinical study of Lu AA21004 in the prevention of
relapse in patients with major depressive disorder. J Psychopharmacol. 2012;26:1408–16.
Citrome L. Vortioxetine for major depressive disorder: an indirect comparison with duloxetine,
escitalopram, levomilnacipran, sertraline, venlafaxine, and vilazodone, using number needed to
treat, number needed to harm, and likelihood to be helped or harmed. J Affect Disord. 2016;196:
225–33.
Connolly KR, Thase ME. Vortioxetine: a new treatment for major depressive disorder. Expert Opin
Pharmacother. 2016;17:421–31.
Herrera-Guzman I, Gudayol-Ferre E, Herrera-Guzman D, et al. Effects of selective serotonin
reuptake and dual serotonergic-noradrenergic reuptake treatments on memory and mental
processing speed in patients with major depressive disorder. J Psychiatr Res. 2009;43:855–63.
Jacobsen PL, Mahableshwarkar AR, Chen Y, Chrones L, Clayton AH. Effect of Vortioxetine vs.
Escitalopram on sexual functioning in adults with well-treated major depressive disorder
experiencing SSRI-induced sexual dysfunction. J Sex Med. 2015a;12:2036–48.
Jacobsen PL, Harper L, Chrones L, Chan S, Mahableshwarkar AR. Safety and tolerability of
vortioxetine (15 and 20 mg) in patients with major depressive disorder: results of an open-
label, flexible-dose, 52-week extension study. Int Clin Psychopharmacol. 2015b;30:255–64.
Jacobsen PL, Mahableshwarkar AR, Palo WA, Chen Y, Dragheim M, Clayton AH. Treatment-
emergent sexual dysfunction in randomized trials of vortioxetine for major depressive disorder
or generalized anxiety disorder: a pooled analysis. CNS Spectr. 2016;21:367–78.
Katona C, Hansen T, Olsen CK. A randomized, double-blind, placebo-controlled, duloxetine-
referenced, fixed-dose study comparing the efficacy and safety of Lu AA21004 in elderly
patients with major depressive disorder. Int Clin Psychopharmacol. 2012;27:215–23.
Koesters M, Ostuzzi G, Guaiana G, Breilmann J, Barbui C. Vortioxetine for depression in adults.
Cochrane Database Syst Rev. 2017;5(7):CD011520. https://doi.org/10.1002/14651858.
CD011520.pub2.
Mahableshwarkar AR, Zajecka J, Jacobson W, Chen Y, Keefe RS. A randomized, placebo-controlled,
active-reference, double-blind, flexible-dose study of the efficacy of Vortioxetine on cognitive
function in major depressive disorder. Neuropsychopharmacology. 2015a;40:2025–37.
Mahableshwarkar AR, Jacobsen PL, Chen Y, Serenko M, Trivedi MH. A randomized, double-blind,
duloxetine-referenced study comparing efficacy and tolerability of 2 fixed doses of vortioxetine
in the acute treatment of adults with MDD. Psychopharmacology. 2015b;232:2061–70.
Meeker AS, Herink MC, Haxby DG, Hartung DM. The safety and efficacy of vortioxetine for acute
treatment of major depressive disorder: a systematic review and meta-analysis. Syst Rev.
2015;4:21.
McIntyre RS, Harrison J, Loft H, Jacobson W, Olsen CK. The effects of vortioxetine on cognitive
function in patients with major depressive disorder(MDD): a meta-analysis of three randomized
controlled trials. Int J Neuropsychopharmacol 2016;19:pyw055.
Mørk A, Pehrson A, Brennum LT, Nielsen SM, Zhong H, Lassen AB, Miller S, Westrich L, Boyle
NJ, Sánchez C, Fischer CW, Liebenberg N, Wegener G, Bundgaard C, Hogg S, Bang-Andersen
B, Stensbøl TB. Pharmacological effects of Lu AA21004: a novel multimodal compound for the
treatment of major depressive disorder. J Pharmacol Exp Ther. 2012;340:666–75.
1424 G. Laux
Nomikos GG, Tomori D, Zhong W, Affinito J, Palo W. Efficacy, safety, and tolerability of
vortioxetine for the treatment of major depressive disorder in patients aged 55 years or older.
CNS Spectr. 2017;22:348–62.
Papakostas GI, Nielsen R, Dragheim M, Tonnoir B. Efficacy and tolerability of vortioxetine versus
agomelatine, categorized by previous treatment, in patients with major depressive disorder
switched after an inadequate response. J Psychiatr Res 2018;101:72–79.
Thase ME, Mahableshwarkar AR, Dragheim M, Loft H, Vieta E. A meta-analysis of randomized,
placebo-controlled trials of vortioxetine for the treatment of major depressive disorder in adults.
Eur Neuropsychopharmacol. 2016;26:979–93.
Thase ME, Danchenko N, Brignone M, Florea I, Diamand F, Jacobsen PL, Vieta E. Comparative
evaluation of vortioxetine as a switch therapy in patients with major depressive disorder. Eur
Neuropsychopharmacol. 2017;27:773–81.
Vieta E, Sluth LB, Olsen CK. The effects of vortioxetine on cognitive dysfunction in patients with
inadequate response to current antidepressants in major depressive disorder: a short-term,
randomized, double-blind, exploratory study versus escitalopram. J Affect Disord. 2018;227:
803–9.
Wagner G, Schultes MT, Titscher V, Teufer B, Klerings I, Gartlehner G. Efficacy and safety of
levomilnacipran, vilazodone and vortioxetine compared with other second-generation antide-
pressants for major depressive disorder in adults: a systematic review and network meta-
analysis. J Affect Disord. 2018;228:1–12.
Part IV
Mood Stabilizers
Andreas Menke
Mood Stabilizer: Definition
Andreas Menke
Contents
Description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1427
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1428
Abstract
The term mood stabilizer is commonly used in the context of the treatment of
bipolar disorder. However, there is no consensus on the exact definition among
researchers or clinicians. A widely accepted agreement recommends that a mood
stabilizer should have at least two of the four properties: antimanic, antidepres-
sant, preventing relapse of depression, and preventing relapse of mania, without
increasing the risk for a switch in the opposite episode. While lithium still remains
the gold standard, also anti-convulsants and atypical antipsychotics are used for
maintenance treatment of bipolar disorder.
Description
The term mood stabilizer is commonly used in the context of the treatment of bipolar
disorder. However, there is no consensus on the exact definition among researchers
or clinicians (Ketter 2018). Roughly, there are mood-stabilizing agents treating from
above, ameliorating manic symptoms, or treating from below, thus improving
depressive symptoms. A possible agreement that is widely accepted recommends
that a drug should have at least two of the four properties: antimanic, antidepressant,
preventing relapse of depression, and preventing relapse of mania, without
A. Menke (*)
Department of Psychiatry, Psychosomatics and Psychotherapy, University Hospital of Würzburg,
Würzburg, Germany
e-mail: menke_a@ukw.de
increasing the risk for a switch in the opposite episode (Bauer and Mitchner 2004).
Lithium has been shown to be clearly one of the most effective drugs for preventing
both manic and depressive episodes and thus remains the gold standard (Kessing
et al. 2018; Miura et al. 2014). Randomized placebo-controlled clinical trials in
bipolar disorder have further proved the efficacy in the maintenance treatment of
anti-convulsants such as valproate and lamotrigine (Miura et al. 2014), as well as
atypical antipsychotics such as quetiapine, olanzapine, and aripiprazole (Lindstrom
et al. 2017). In addition, there is limited evidence that mood stabilizers are also
effective in schizoaffective disorder (Vieta 2010), a disorder on a spectrum between
bipolar disorder and schizophrenia, and borderline personality disorder (Belli et al.
2012). There is an emerging approach to categorize drugs using the Neuroscience-
based Nomenclature for Psychotropic Medications (NbN, www.ecnp.eu/nomencla
ture), that is based on the mechanisms of the drugs. However, due to insufficient
understanding of mechanisms of the mood-stabilizing agents, it is challenging to
apply this taxonomy on this heterogeneous class (Ketter 2018). In the following
chapters, we will address the classification and indications of mood stabilizers, the
pharmacology and biochemistry, and the course of the therapy including adverse
events, and we will review the mood-stabilizing agents such as lithium, valproate,
lamotrigine, carbamazepine, risperidone, olanzapine, quetiapine, and asenapine.
References
Bauer MS, Mitchner L. What is a “mood stabilizer”? An evidence-based response. Am J Psychiatry.
2004;161:3–18.
Belli H, Ural C, Akbudak M. Borderline personality disorder: bipolarity, mood stabilizers and
atypical antipsychotics in treatment. J Clin Med Res. 2012;4:301–8.
Kessing LV, Bauer M, Nolen WA, Severus E, Goodwin GM, Geddes J. Effectiveness of mainte-
nance therapy of lithium vs other mood stabilizers in monotherapy and in combinations:
a systematic review of evidence from observational studies. Bipolar Disord. 2018. PMID:
29441712. https://doi.org/10.1111/bdi.12623
Ketter TA. Definition of the term “mood stabilizer”. Bipolar Disord. 2018;20:74–5.
Lindstrom L, Lindstrom E, Nilsson M, Hoistad M. Maintenance therapy with second generation
antipsychotics for bipolar disorder – a systematic review and meta-analysis. J Affect Disord.
2017;213:138–50.
Miura T, Noma H, Furukawa TA, Mitsuyasu H, Tanaka S, Stockton S, Salanti G, Motomura K,
Shimano-Katsuki S, Leucht S, Cipriani A, Geddes JR, Kanba S. Comparative efficacy and
tolerability of pharmacological treatments in the maintenance treatment of bipolar disorder: a
systematic review and network meta-analysis. Lancet Psychiatry. 2014;1:351–9.
Vieta E. Developing an individualized treatment plan for patients with schizoaffective disorder:
from pharmacotherapy to psychoeducation. J Clin Psychiatry. 2010;71(Suppl 2):14–9.
Mood Stabilizers: Classification, Indications
and Differential Indications
Andreas Menke
Contents
Classification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1430
Lithium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1430
Anti-convulsants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1430
Antipsychotics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1431
Indications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1431
Differential Indications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1432
Unipolar Depression/Major Depressive Disorder (MDD) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1432
Schizoaffective Disorder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1432
Borderline Personality Disorder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1432
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1433
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1433
Abstract
Pharmacological long-term maintenance treatment of bipolar disorder usually con-
sists of mood stabilizing agents. While there is no consensus on the exact definition
of mood stabilizer, it is widely accepted that a mood stabilizing agent should have at
least two of the four properties: anti-manic, anti-depressant, preventing relapse of
depression, and preventing relapse of mania. Thus, mood stabilizing agents belong
to a heterogeneous class of drugs acting through many different mechanisms.
Lithium still remains the gold standard in the maintenance treatment of bipolar
disorder. In addition, anti-convulsants such as valproate and lamotrigine and atypical
A. Menke (*)
Department of Psychiatry, Psychosomatics and Psychotherapy, University Hospital of Würzburg,
Würzburg, Germany
e-mail: Menke_A@ukw.de
Classification
Due to the recurrent and chronic nature of bipolar disorder, a long-term maintenance
treatment is essential. This long-term management usually consists of pharmacolog-
ical treatments, psychological therapies, and lifestyle approaches (Vieta et al. 2018).
The pharmacological approach typically comprises the mood stabilizers, a hetero-
geneous class of drugs acting through many different mechanisms (Baldessarini
et al. 2018).
Lithium
Lithium still remains the first-line maintenance treatment for bipolar disorder, for
example, a network meta-analysis observed a risk ratio of 0.62 (95% credible
interval 0.53–0.72) for the risk of recurrence or relapse compared with placebo
(Miura et al. 2014). Despite the many decades of research and clinical use, the
mechanisms of lithium have been only partially understood. There is some evidence
that lithium restores electrolyte balance, but newer research shows that lithium
affects multiple steps in cellular signaling. Key nodes of these regulatory networks
are CREB (cAMP response element-binding protein), GSK3 (glycogen synthase
kinase 3), and Na+-K+ ATPase (Alda 2015). In addition, lithium exerts also chrono-
biological effects (Alda 2015).
Anti-convulsants
Only some anti-convulsants have mood stabilizing effects, while others that exert
similar pharmacodynamics effects lack the mood stabilizing properties (Baldessarini
et al. 2018). Currently, valproate, lamotrigine, and carbamazepine are clinically used
for long-term maintenance treatment of bipolar disorder (Baldessarini et al. 2018).
Of note, valproate is not recommended for the treatment of women of childbearing
age, because this agent has teratogenic effects and increases, for example, the risks of
spina bifida and low IQ in off-spring (Yatham et al. 2018). While the exact mech-
anisms of action are still not fully understood, there are various hypotheses:
inhibiting voltage-sensitive sodium channels, boosting the actions of GABA, reduc-
ing excitatory glutamate neurotransmission, and the regulation of downstream signal
transduction cascades (Landmark 2007).
Mood Stabilizers: Classification, Indications and Differential Indications 1431
Antipsychotics
Indications
Generally mood stabilizing agents are indicated for the maintenance treatment of
bipolar disorder, i.e., preventing a relapse of depression and/or mania. However, the
mood stabilizer should be selected with respect of the patient’s predominant polarity,
that is, defined as the condition the patient experiences at least twice as often as the
other conditions (Vieta et al. 2018). The polarity index is a metric tool that classifies
the maintenance treatment as predominant antidepressive or as predominant anti-
manic (Carvalho et al. 2014). For example, patients who experience more depressive
episodes may benefit from lamotrigine, while patients who experience more manic
episodes may improve with atypical antipsychotics such as risperidone (see Fig. 1)
(Baldessarini et al. 2012; Grande et al. 2012). Of note, Lithium and quetiapine have a
polarity index around 1, thus these agents are equally suitable for the prevention of
depressive and manic episodes (Vieta et al. 2018).
Fig. 1 Polarity index: The polarity index of drugs used for the maintenance treatment of patients
with bipolar disorders is the ratio of the number of patients needed to treat for prevention of
depression to the number of patients needed to treat for prevention of mania on the basis of results of
randomized placebo-controlled trials 173. This index classifies therapies as those with an antimanic
prophylactic effect and those with an antidepressant prophylactic effect. A polarity index of
1 reflects an equal efficacy in preventing manic and depressive episodes. (Reprinted by permission
from Springer Nature: Nature Reviews Disease Primers (Bipolar disorders, Eduard Vieta, Michael
Berk, Thomas G. Schulze, André F. Carvalho, Trisha Suppes et al.), (2018))
1432 A. Menke
Differential Indications
Schizoaffective Disorder
Cross-References
References
Alda M. Lithium in the treatment of bipolar disorder: pharmacology and pharmacogenetics. Mol
Psychiatry. 2015;20:661–70.
Angst J, Gamma A, Sellaro R, Lavori PW, Zhang H. Recurrence of bipolar disorders and major
depression. A life-long perspective. Eur Arch Psychiatry Clin Neurosci. 2003;253:236–40.
Baldessarini RJ, Undurraga J, Vazquez GH, Tondo L, Salvatore P, Ha K, Khalsa HM, Lepri B, Ha
TH, Chang JS, Tohen M, Vieta E. Predominant recurrence polarity among 928 adult interna-
tional bipolar I disorder patients. Acta Psychiatr Scand. 2012;125:293–302.
Baldessarini RJ, Tondo L, Vazquez GH. Pharmacological treatment of adult bipolar disorder. Mol
Psychiatry. 2018;24:198.
Belli H, Ural C, Akbudak M. Borderline personality disorder: bipolarity, mood stabilizers and
atypical antipsychotics in treatment. J Clin Med Res. 2012;4:301–8.
Bond DJ, Lam RW, Yatham LN. Divalproex sodium versus placebo in the treatment of acute bipolar
depression: a systematic review and meta-analysis. J Affect Disord. 2010;124:228–34.
Canadian Agency for Drugs and Technologies in Health (CADTH). Aripiprazole for borderline
personality disorder: a review of the clinical effectiveness. In: Aripiprazole for borderline
personality disorder: a review of the clinical effectiveness. Ottawa: Canadian Agency for Drugs
and Technologies in Health; 2017.
Carvalho AF, Quevedo J, McIntyre RS, Soeiro-de-Souza MG, Fountoulakis KN, Berk M,
Hyphantis TN, Vieta E. Treatment implications of predominant polarity and the polarity
index: a comprehensive review. Int J Neuropsychopharmacol. 2014;18(2):1–11.
Crawford MJ, Sanatinia R, Barrett B, Cunningham G, Dale O, Ganguli P, Lawrence-Smith G,
Leeson V, Lemonsky F, Lykomitrou G, Montgomery AA, Morriss R, Munjiza J, Paton C,
Skorodzien I, Singh V, Tan W, Tyrer P, Reilly JG, LABILE Study Team. The clinical effective-
ness and cost-effectiveness of lamotrigine in borderline personality disorder: a randomized
placebo-controlled trial. Am J Psychiatry. 2018;175:756–64.
1434 A. Menke
Crossley NA, Bauer M. Acceleration and augmentation of antidepressants with lithium for depres-
sive disorders: two meta-analyses of randomized, placebo-controlled trials. J Clin Psychiatry.
2007;68:935–40.
Fawcett JA. Lithium combinations in acute and maintenance treatment of unipolar and bipolar
depression. J Clin Psychiatry. 2003;64(Suppl 5):32–7.
Geddes JR, Carney SM, Davies C, Furukawa TA, Kupfer DJ, Frank E, Goodwin GM. Relapse
prevention with antidepressant drug treatment in depressive disorders: a systematic review.
Lancet. 2003;361:653–61.
Geddes JR, Calabrese JR, Goodwin GM. Lamotrigine for treatment of bipolar depression: inde-
pendent meta-analysis and meta-regression of individual patient data from five randomised
trials. Br J Psychiatry. 2009;194:4–9.
Grande I, Balanza-Martinez V, Jimenez-Arriero M, Iglesias Lorenzo FG, Franch Valverde JI, de
Arce R, Zaragoza S, Cobaleda S, Vieta E, SIN-DEPRES Group. Clinical factors leading to
lamotrigine prescription in bipolar outpatients: subanalysis of the SIN-DEPRES study. J Affect
Disord. 2012;143:102–8.
Gunderson JG, Choi-Kain LW. Medication management for patients with borderline personality
disorder. Am J Psychiatry. 2018;175:709–11.
Guzzetta F, Tondo L, Centorrino F, Baldessarini RJ. Lithium treatment reduces suicide risk in
recurrent major depressive disorder. J Clin Psychiatry. 2007;68:380–3.
Hansen R, Gaynes B, Thieda P, Gartlehner G, Deveaugh-Geiss A, Krebs E, Lohr K. Meta-analysis
of major depressive disorder relapse and recurrence with second-generation antidepressants.
Psychiatr Serv. 2008;59:1121–30.
Landmark CJ. Targets for antiepileptic drugs in the synapse. Med Sci Monit. 2007;13:RA1–7.
Mazza M, Marano G, Traversi G, Carocci V, Romano B, Janiri L. Cariprazine in bipolar depression
and mania: state of the art. CNS Neurol Disord Drug Targets. 2018;17:723.
Miura T, Noma H, Furukawa TA, Mitsuyasu H, Tanaka S, Stockton S, Salanti G, Motomura K,
Shimano-Katsuki S, Leucht S, Cipriani A, Geddes JR, Kanba S. Comparative efficacy and
tolerability of pharmacological treatments in the maintenance treatment of bipolar disorder: a
systematic review and network meta-analysis. Lancet Psychiatry. 2014;1:351–9.
Morsel AM, Morrens M, Sabbe B. An overview of pharmacotherapy for bipolar I disorder. Expert
Opin Pharmacother. 2018;19:203–22.
Nelson JC, Papakostas GI. Atypical antipsychotic augmentation in major depressive disorder: a
meta-analysis of placebo-controlled randomized trials. Am J Psychiatry. 2009;166:980–91.
Nierenberg AA, Husain MM, Trivedi MH, Fava M, Warden D, Wisniewski SR, Miyahara S, Rush
AJ. Residual symptoms after remission of major depressive disorder with citalopram and risk of
relapse: a STARD report. Psychol Med. 2010;40:41–50.
Paolini E, Mezzetti FA, Pierri F, Moretti P. Pharmacological treatment of borderline personality
disorder: a retrospective observational study at inpatient unit in Italy. Int J Psychiatry Clin Pract.
2017;21:75–9.
Solomon DA, Keller MB, Leon AC, Mueller TI, Lavori PW, Shea MT, Coryell W, Warshaw M,
Turvey C, Maser JD, Endicott J. Multiple recurrences of major depressive disorder. Am J
Psychiatry. 2000;157:229–33.
Ventriglio A, Vincenti A, Centorrino F, Talamo A, Fitzmaurice G, Baldessarini RJ. Use of mood
stabilizers for hospitalized psychotic and bipolar disorder patients. Int Clin Psychopharmacol.
2011;26:88–95.
Vieta E. Developing an individualized treatment plan for patients with schizoaffective disorder:
from pharmacotherapy to psychoeducation. J Clin Psychiatry. 2010;71(Suppl 2):14–9.
Vieta E, Berk M, Schulze TG, Carvalho AF, Suppes T, Calabrese JR, Gao K, Miskowiak KW,
Grande I. Bipolar disorders. Nat Rev Dis Primers. 2018;4:18008. https://doi.org/10.1038/
nrdp.2018.8.
Vincenzi B, Greene CM, Ulloa M, Parnarouskis L, Jackson JW, Henderson DC. Lithium or
valproate adjunctive therapy to second-generation antipsychotics and metabolic variables in
patients with schizophrenia or schizoaffective disorder. J Psychiatr Pract. 2016;22:175–82.
Mood Stabilizers: Classification, Indications and Differential Indications 1435
Weisler RH, Calabrese JR, Thase ME, Arvekvist R, Stening G, Paulsson B, Suppes T. Efficacy of
quetiapine monotherapy for the treatment of depressive episodes in bipolar I disorder: a post hoc
analysis of combined results from 2 double-blind, randomized, placebo-controlled studies. J
Clin Psychiatry. 2008;69:769–82.
Yatham LN, Kennedy SH, Parikh SV, Schaffer A, Bond DJ, Frey BN, Sharma V, Goldstein BI,
Rej S, Beaulieu S, Alda M, MacQueen G, Milev RV, Ravindran A, O’Donovan C, McIntosh D,
Lam RW, Vazquez G, Kapczinski F, McIntyre RS, Kozicky J, Kanba S, Lafer B, Suppes T,
Calabrese JR, Vieta E, Malhi G, Post RM, Berk M. Canadian Network for Mood and Anxiety
Treatments (CANMAT) and International Society for Bipolar Disorders (ISBD) 2018 guidelines
for the management of patients with bipolar disorder. Bipolar Disord. 2018;20:97–170.
Yildiz A, Vieta E, Leucht S, Baldessarini RJ. Efficacy of antimanic treatments: meta-analysis of
randomized, controlled trials. Neuropsychopharmacology. 2011;36:375–89.
Mood Stabilizers: Pharmacology
and Biochemistry
Leif Hommers
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1438
Lithium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1438
Anti-convulsants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1440
Antipsychotics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1442
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1445
Abstract
Drugs successfully applied in the treatment of bipolar disorder have been termed
mood stabilizers and include different pharmacological classes: antipsychotic
agents, anti-convulsants, and lithium.
Contrary to the development of antipsychotic and antidepressive drugs, no
drug has been developed on biological models or hypotheses of bipolar disorder.
The most commonly applied drug lithium has been introduced based on findings
by serendipity, and no unique pharmacological mechanism of mood stabilization
was identified so far. However, several mood stabilizers share a modulation of
inositol- and GSK-3β-mediated pathways leading to neuroprotection and
increased neuronal plasticity.
Future hypothesis-driven drug development is presently limited since
(I) bipolar disorder results from a large spectrum of molecular mechanisms also
associated and shared with other psychiatric morbidities and (II) animal models
allow to address only neurobiological aspects well conserved across species.
L. Hommers (*)
Zentrum für Psychische Gesundheit, Universitätsklinikum Würzburg, Würzburg, Germany
e-mail: L.Hommers@ihp-labor.de
Introduction
Drugs successfully applied in the treatment of bipolar disorder have been termed
mood stabilizers and include different pharmacological classes: lithium, antipsy-
chotic agents, and anti-convulsants. Lithium has been in medical use to treat
different conditions since the early nineteenth century. However, its effect on bipolar
disorder was only recognized in the late 1940s, and it represents the first mood
stabilizer established and still commonly used today (Cade 1949). Later on, it was
noted that anti-convulsants show not only activity against epileptic seizures. Mood-
stabilizing effect of carbamazepine has been recognized since the 1970s, of valproic
acid since the 1980s, and of lamotrigine since the 1990s. While the effect of
antipsychotic drugs in treating acute mania has been recognized for a long time,
research was focusing on mood stabilizing of antipsychotic agents only recently, and
the typical drugs applied are asenapine, olanzapine, quetiapine, and risperidone class
(López-Muñoz et al. 2018). This chapter reviews the molecular pharmacology of
drugs proposed to be mood stabilizers.
Lithium
Lithium (Li+) is the smallest alkali metal sharing some chemical characteristics with
sodium and potassium, but without any identified physiological role. Li+ is com-
monly applied as lithium carbonate, lithium citrate, or lithium sulfate. It becomes
almost completely absorbed from the gastrointestinal tract with peak plasma con-
centrations at about 4 h after intake of commonly applied slow-release lithium
carbonate. There is no relevant protein binding, and lithium is evenly distributed
in extracellular compartments. However, concentrations in organs may vary, and Li+
concentrations in liquor were reported to reach only about 50% of serum concen-
trations with a slow equilibration between liquor and serum due to a slow passage of
the blood-brain barrier, as it was also noted for intracellular Li+ concentrations in
erythrocytes. Lithium is eliminated by renal excretion at about 95% of a single dose.
While it is freely filtered in the nephron, the majority of filtered Li+ becomes
reabsorbed resulting in a Li+ clearance of only 20% of that of creatinine. The
elimination half-time is about 24 h, and excretion upon discontinuation follows
two phases with 6–12 h of rapid excretion and 10–14 days of slow excretion. Li+
has a narrow therapeutic window. The concentration-response relationship points
toward serum levels to be maintained between 0.5 and 1.2 mmol/l in order to balance
therapeutic and unwanted side effects (Hiemke et al. 2018). Depending on single or
twice daily dosing schemes, peak and trough levels are expected to show a greater or
smaller variation. There is no consensus whether single or twice daily dosing leads to
better tolerability with equivalent efficacy. Moreover, there is no consensus, whether
therapeutic drug monitoring of Li+ serum levels should be generally conducted 12 h
after the last dose or at trough conditions, in case of single dosing 24 h after the
last dose.
Mood Stabilizers: Pharmacology and Biochemistry 1439
Anti-convulsants
Anti-convulsants have been developed for the treatment of epileptic seizures. They
represent a diverse class of drugs with different molecular pharmacological mech-
anisms. The anti-seizure activity of these drugs is primarily related to mechanisms
enhancing inhibitory neuronal activity (e.g., GABA-mediated signaling) as well as
mechanisms inhibiting excitatory neuronal activity (e.g., Na+ and Ca2+ currents).
However, these mechanisms may be involved in, but are not considered to be
primarily responsible for, the mood-stabilizing effects of valproic acid, carbamaze-
pine, and lamotrigine. It has been hypothesized that mechanisms induced resembling
molecular actions of lithium are responsible for mood-stabilizing effects and most
data has been accumulated for valproic acid.
Valproic acid. Valproic acid is a simple branched-chain carboxylic acid. It is
absorbed completely after oral administration within 1–4 h. Extended release for-
mulations lead to a delay for up to several hours. 78% to 94% of valproic acid are
bound to plasma proteins. Valproic acid undergoes complex hepatic metabolism.
The major pathways are glucuronidation by UGT enzymes and β-oxidation in the
mitochondria. Valproic acid may also be hydroxylated by CYP2C9 and CYP2C19
and to a minor extent by CYP2A6 and CYP2B6. Valproic acid weakly inhibits
CYP2A6, CYP2C9, CYP2C19, CYP3A4/5, and UGT enzymes; however, it is a
strong inhibitor of UGT1A4, important for the inhibition of lamotrigine metabolism
by valproic acid. Complex drug-drug pharmacokinetic interactions have frequently
been observed for valproic putatively due to inhibition of drug oxidation, suggesting
to apply therapeutic drug monitoring when valproic acid is administered. The
therapeutic range when applied as an anti-seizure or mood stabilizer is small and
ranges from 50 μg/l to 100 μg/l (Hiemke et al. 2018).
The antiepileptic properties of valproic acid were discovered by serendipity when
it was used as a vehicle for compounds being screened. A number of molecular
pharmacological mechanisms have been identified. Valproic acid potentiates the
inhibitory effects of GABA. It inhibits GABA transaminase, succinic semialdehyde
dehydrogenase, and alpha-ketoglutarate dehydrogenase, resulting in increased
GABA levels during initiation of treatment. It suppresses glutamatergic neurotrans-
mission and NMDA receptor activity, and this mechanism may also be important for
the efficacy when treating mania. However, the underlying processes are presently
not fully understood. In rodent models, administration of valproic acid leads to an
increase in norepinephrine, dopamine, and serotonin brain levels, and the link
between GABAergic and serotonergic neurotransmission via presynaptic serotonin-
ergic 5-HT1 receptors may also be of high relevance for its mood-stabilizing
properties. Moreover, valproic acid inhibits high-frequency firing of neurons in
models of epileptic seizures. However, direct inhibition of voltage-gated Na+ chan-
nels, as well as T-type Ca2+ channels, appears to be only relevant at higher concen-
trations of valproic acid in vivo, suggesting that impairment of neuronal ATP
production due to inhibitory effects on cerebral glucose metabolism resulting from
the inhibition of alpha-ketoglutarate dehydrogenase may equally be relevant
(Johannessen and Johannessen 2003). Valproic acid inhibits HDACs at therapeutic
Mood Stabilizers: Pharmacology and Biochemistry 1441
shown to increase growth cone area, presumably unrelated to GSK-3β and HDAC
(Williams et al. 2002). Due to its structural relation, carbamazepine may inhibit the
serotonin transporter weakly at therapeutic concentrations (Sarker et al. 2010).
Recently, carbamazepine was related to increasing global methylation levels, resem-
bling lithium and valproic acid (Pisanu et al. 2018).
Lamotrigine. Lamotrigine is a phenyltriazine derivative and becomes completely
absorbed within 1–4 h from the gastrointestinal tract upon oral intake. It is metab-
olized primarily by glucuronidation, predominantly by UGT1A4 with a plasma half-
life of about 24 h. The half-time may largely vary on the UGT1A4 genotype. About
50% are bound to plasma proteins (Hiemke et al. 2018).
Lamotrigine was initially developed as an antifolate agent, but its established
pharmacological mechanism is use- and voltage-dependent inhibition of Na+ chan-
nels comparable to carbamazepine. Since lamotrigine appears to be effective in a
broader range of epileptic syndromes, it has been speculated that other mechanisms
than those shared with carbamazepine may be involved in its mood-stabilizing
effects, e.g., lamotrigine was reported to weakly inhibit voltage-gated Ca2+ channels
and to show antiglutamatergic activity, as well as a weak inhibition of the serotonin
transporter, but direct interference with monoaminergic, NMDA, or GABA recep-
tors is not known. Neuroprotective effects have been observed in translational
models, but no unique mechanism was identified so far (Ketter et al. 2003). There
is conflicting data concerning the activation of the Akt/GSK-3β (Rogawski and
Löscher 2004a; Aubry et al. 2009).
Antipsychotics
binding of the physiological ligand and increase the activity of the receptor. In the
presence of high levels of physiological ligand, a partial agonist may therefore lead
to physiologically inhibitory effects on a signaling pathway (Kenakin 2002).
Dopaminergic D2 and serotoninergic 5-HT2A receptors are considered to be the
most important molecular targets of antipsychotic drugs mediating antipsychotic
effects. Drug-specific inhibition of histaminergic, adrenergic, and muscarinergic
receptors is commonly related to several side effects of antipsychotic drugs, but
depending on the class and exact agent. Among dopaminergic and serotoninergic
receptors, several lines of evidence suggest that agonistic effects at D1, 5-HT1A,
5-HT2C, and 5-HT4 as well as antagonistic effects at D3, D4, 5-HT3, 5-HT6, and
5-HT7 may additionally contribute to antipsychotic, anxiolytic, and antidepressive
effects of antipsychotic drugs. It has also been suggested that not only the affinity
toward a given receptor but also the rates of association to and dissociation from the
receptor of each drug modulate their pharmacological profile. This hypothesis was
used to explain the prevalence of hyperprolactinemia and parkinsonism for different
antipsychotic agents upon comparable dopamine D2 receptor binding in vivo when
reaching therapeutic steady-state serum concentrations (Kapur and Seeman 2001).
Additionally, allosteric activation of NMDA as well as AMPA receptors, inhibition
of glycine transporters, activation of metabotropic glutamate receptors, and (partial)
activation of cholinergic as well as muscarinergic receptors among many other
putative mechanisms may contribute to the pharmacological effects of antipsychotic
drugs. Agonism at metabotropic glutamate receptors has specifically been related to
the mood-stabilizing effects of some antipsychotic drugs (Miyamoto et al. 2012). In
the synaptic cleft, G protein-coupled receptors may not only act as independent
monomers, as in several in vitro experiments, but also form oligomeric homo- and
heteromers (Ferré et al. 2014). This may then result in various modulatory effects on
receptor-mediated intracellular signaling upon binding of ligand. Allosteric modu-
lation and biased agonism at heteromers represent important mechanisms: while a
homomer may only activate canonically one class of G proteins, any ligand may now
have distinctive effects on the activation of different receptors and intracellular
signaling cascades, depending on the composition of the heteromer. This may
range from allosteric activation of the secondary receptor to antagonistic effects on
one canonical pathway (e.g., inhibition of G protein activation) but (partial) agonistic
effects on a second pathway (e.g., β-arrestin signaling) (Kenakin 2009). These
phenomena have been reported in molecular studies of antipsychotic agents (Masri
et al. 2008), and it has been suggested that antipsychotic drugs counteract psychotic
symptoms related to neuronal dysbalance of Gi- and Gq-dependent signaling by
binding to a heteromeric 5-HT2A:mGluR2-receptor complex, thereby re-balancing
Gi- and Gq-dependent signaling (Fribourg et al. 2011). Downstream, signaling may
then be modulated and converge on the level of Akt/GSK-3β (Beaulieu and
Gainetdinov 2011). However, the importance toward treatment response and side
effects has not been ultimately established so far.
Asenapine. Asenapine is a tetracyclic dibenzooxepinpyrrole derivative with some
structural resemblance toward the antidepressants mianserine and mirtazapine. It is
rapidly absorbed when applied sublingually and undergoes extensive first-pass
1444 L. Hommers
metabolism to about 65% of a single dose. Upon ingestion, the bioavailability is only
minimal. It is bound to plasma proteins to a large extent. It becomes directly
glucuronidated by UGT1A4 and metabolized by CYP1A2 as well as to a smaller
extent by CYP2D6 and CYP3A4. The half-life is about 24 h. Asenapine is a strong
antagonist at 5-HT2A/C, 5-HT7, and D2 and D3 receptors but also at H1 and α1/2
receptors. It is a partial agonist at 5-HT1A receptors. No significant interaction with
muscarinergic receptors has been identified (Hiemke et al. 2018; Aringhieri et al.
2018).
Olanzapine. Olanzapine is a thiobenzodiazepine derivative structurally resem-
bling clozapine. It is well absorbed within 4–6 h after ingestion, and about 40% of
a single dose undergo first-pass metabolism. It is bound to plasma protein to a
large extent. It becomes metabolized by CYP1A2 and to a smaller extent by
CYP2A6, CYP2C8, and CYP2D6 or becomes directly glucuronidated. The half-
life is between 33 and 52 h and increases with age. Olanzapine is an inverse
agonist predominantly at 5-HT2A/C and H1 receptors. It is an antagonist at D2, D3,
5-HT7, M1, M2, and M4 receptors as well as adrenergic α1/2 receptors. The
affinity for the 5-HT2A receptor is considerably higher than for the D2 receptor.
Inhibitory effects at BDNF receptors have been reported (Hiemke et al. 2018;
Aringhieri et al. 2018).
Quetiapine. Quetiapine is a dibenzothiazepine structurally resembling clozapine.
It is well absorbed within 2 h after ingestion. It is bound to plasma protein to a large
extent. It becomes metabolized by CYP3A4 and to a smaller extent by CYP2D6. The
half-life is about 8 h. Quetiapine is a strong antagonist at H1 receptors. To a lesser
extent, it is an antagonist at α1/2, 5-HT2A/C, 5-HT7, and D2/3 receptors as well and a
partial agonist at 5-HT1A receptors. Minor antagonistic effects have been reported at
muscarinergic as well as BDNF receptors. Its metabolite norquetiapine additionally
inhibits the norepinephrine transporter and shows antagonistic properties at
muscarinergic receptors (Hiemke et al. 2018; Aringhieri et al. 2018).
Risperidone. Risperidone is a benzisoxazole derivative. It is rapidly absorbed
within 1 h after ingestion, and about 90% are bound to plasma proteins. It becomes
rapidly metabolized by CYP2D6 and to a smaller extent by CYP3A4 within less
than 12 h into its active metabolite 9-hydroxyrisperidone. 9-Hydroxyrisperidone
becomes predominantly excreted renally unchanged, with the remainder being
metabolized by at least four different pathways, leading to a terminal half-life of
about 24 h. Both are substrates of MDRD1. 9-Hydroxyrisperidone is also available
as paliperidone. Risperidone is an antagonist at D2 and 5-HT7 receptors as well as
an inverse agonist at 5-HT2, α2C, and D3 receptors. Like olanzapine, its affinity for
the 5-HT2A receptor is larger than for the D2 receptor. Only minimal antagonism at
H1 as well as BDNF receptors has been identified, and no significant interaction
with muscarinergic receptors has been noted so far. While it must be noted that due
to its longer half-life 9-hydroxyrisperidone is primarily mediating the pharmaco-
logical effects of risperidone, only minor differences concerning the receptor
binding profile of both agents have been identified (Hiemke et al. 2018; Aringhieri
et al. 2018).
Mood Stabilizers: Pharmacology and Biochemistry 1445
References
Alda M. Lithium in the treatment of bipolar disorder: pharmacology and pharmacogenetics. Mol
Psychiatry. 2015;20(6):661–70.
Aringhieri S, et al. Molecular targets of atypical antipsychotics: from mechanism of action to
clinical differences. Pharmacol Ther. 2018;192:20. Ahead of print
Aubry J-M, et al. Early effects of mood stabilizers on the Akt/GSK-3beta signaling pathway and on
cell survival and proliferation. Psychopharmacology. 2009;205(3):419–29.
Beaulieu J-M, Gainetdinov RR. The physiology, signaling, and pharmacology of dopamine recep-
tors. Pharmacol Rev. 2011;63(1):182–217.
Beaulieu J-M, et al. A beta-arrestin 2 signaling complex mediates lithium action on behavior. Cell.
2008;132(1):125–36.
Berridge MJ, Downes CP, Hanley MR. Neural and developmental actions of lithium: a unifying
hypothesis. Cell. 1989;59(3):411–9.
Cade JF. Lithium salts in the treatment of psychotic excitement. Med J Aust. 1949;2(10):349–52.
Chiu C-T, et al. Therapeutic potential of mood stabilizers lithium and valproic acid: beyond bipolar
disorder. Pharmacol Rev. 2013;65(1):105–42.
Coppen A, Shaw DM. The distribution of electrolytes and water in patients after taking lithium
carbonate. Lancet. 1967;2(7520):805–6.
Davis KL, et al. Dopamine in schizophrenia: a review and reconceptualization. Am J Psychiatry.
1991;148(11):1474–86.
Ferré S, et al. G protein-coupled receptor oligomerization revisited: functional and pharmacological
perspectives. Pharmacol Rev. 2014;66(2):413–34.
Fribourg M, et al. Decoding the signaling of a GPCR heteromeric complex reveals a unifying
mechanism of action of antipsychotic drugs. Cell. 2011;147(5):1011–23.
Ghasemi M, Dehpour AR. The NMDA receptor/nitric oxide pathway: a target for the therapeutic
and toxic effects of lithium. Trends Pharmacol Sci. 2011;32(7):420–34.
Gurvich N, Klein PS. Lithium and valproic acid: parallels and contrasts in diverse signaling
contexts. Pharmacol Ther. 2002;96(1):45–66.
Hiemke C, et al. Consensus guidelines for therapeutic drug monitoring in neuropsychophar-
macology: update 2017. Pharmacopsychiatry. 2018;51(1–02):9–62.
Johannessen CU, Johannessen SI. Valproate: past, present, and future. CNS Drug Rev. 2003;9
(2):199–216.
Jope RS. Lithium and GSK-3: one inhibitor, two inhibitory actions, multiple outcomes. Trends
Pharmacol Sci. 2003;24(9):441–3.
Kapur S, Seeman P. Does fast dissociation from the dopamine d(2) receptor explain the action of
atypical antipsychotics? A new hypothesis. Am J Psychiatry. 2001;158(3):360–9.
Kenakin T. Efficacy at G-protein-coupled receptors. Nat Rev Drug Discov. 2002;1(2):103–10.
Kenakin TP. Cellular assays as portals to seven-transmembrane receptor-based drug discovery. Nat
Rev Drug Discov. 2009;8(8):617–26.
Ketter TA, Manji HK, Post RM. Potential mechanisms of action of lamotrigine in the treatment of
bipolar disorders. J Clin Psychopharmacol. 2003;23(5):484–95.
López-Muñoz F, et al. A history of the pharmacological treatment of bipolar disorder. Int J Mol Sci.
2018;19(7):2143.
Manji HK, Lenox RH. Signaling: cellular insights into the pathophysiology of bipolar disorder. Biol
Psychiatry. 2000;48(6):518–30.
Manji HK, Zarate CA. Molecular and cellular mechanisms underlying mood stabilization in bipolar
disorder: implications for the development of improved therapeutics. Mol Psychiatry. 2002;7
Suppl 1(S1):S1–7.
Masri B, et al. Antagonism of dopamine D2 receptor/beta-arrestin 2 interaction is a common
property of clinically effective antipsychotics. Proc Natl Acad Sci U S A. 2008;105
(36):13656–61.
1446 L. Hommers
Contents
Preliminary Pathophysiological Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1448
Important Generic Side Effects of Mood Stabilizers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1453
Specific Side Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1456
Contraindications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1459
Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1460
Future Perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1464
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1465
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1465
Abstract
The key mood stabilizers used in clinical practice are a heterogeneous group of
medications with multiple disparities. Included among this group are lithium,
certain antiepileptic drugs such as carbamazepine, lamotrigine, and valproate, as
well as certain atypical antipsychotics such as asenapine, olanzapine, quetiapine,
and risperidone. The side effect profile is varied, but these medications can have
an impact on the central nervous system as well as the cardiac, gastrointestinal,
endocrine, renal, and the hematopoietic systems. Sexual and allergic side effects
are also of significant importance. Among the mood stabilizers, valproate treat-
ment seemingly carries the highest risk for birth defects. Other specific side
effects of clinical importance include lithium toxicity which can lead to ataxia,
rigor, cerebral seizures, and shock; decreased bone marrow function under
carbamazepine treatment; and Stevens-Johnson syndrome and Lyell’s syndrome
H. Himmerich (*)
Department of Psychological Medicine, King’s College London, London, UK
e-mail: hubertus.himmerich@kcl.ac.uk
A. Hamilton
Bethlem Royal Hospital, South London and Maudsley NHS Foundation Trust, London, UK
e-mail: Amy.Hamilton@slam.nhs.uk
Table 1 (continued)
Implications
Mood for clinical
stabilizer Side effects Contraindications Interactions management
Lamotrigine Headache Severe hepatic Metabolized by Low starting
Somnolence dysfunction UGTs dose, slow dose
Aggressiveness Severe renal No combination increase
Nausea, vomiting dysfunction with sertraline Patient to be
Stevens-Johnson No combination informed about
syndrome with ethinyl the possibility
Lyell’s syndrome estradiol of life-
Quincke’s edema Caution: threatening
combination with skin reactions.
sedating Stevens-
medications Johnson
Combination with syndrome and
valproate: Lyell’s
lamotrigine levels syndrome:
to be measured lamotrigine to
regularly be stopped
immediately
Caution:
history of
allergic skin
reactions
Valproate Drowsiness Severe hepatic Metabolized by Pregnancy to
Nausea, vomiting dysfunction UGTs, CYP2A6, be excluded
Tremor Severe renal CYP2B6, before start of
Paresthesia dysfunction CYP2C9, treatment
Birth defects: spina Coagulation CYP2C19 Caution:
bifida, valproate disorder Combination with history of
syndrome Pancreatitis CYP2A6-, allergic skin
Acute CYP2B6-, reactions
intermittent CYP2C9-, or
porphyria CYP2C19-
Pregnancy inducing
substances:
valproate levels to
be measured
regularly
Combinations with
anticoagulants or
TCAs to be
avoided
Caution:
combination with
sedating
medications
Antipsychotics
Asenapine Somnolence Long QT Metabolized via Attention to be
Anxiety syndrome CYP1A2, UGTs paid to weight
EPS Severe hepatic Caution when gain
(continued)
Mood Stabilizers: Side Effects, Contraindications, and Interactions 1451
Table 1 (continued)
Implications
Mood for clinical
stabilizer Side effects Contraindications Interactions management
Weight gain dysfunction combined with
Prolonged QTc Severe renal CYP1A2 inhibitors
Malignant dysfunction Smokers may need
neuroleptic Hyperthyroidism higher doses
syndromea History of Caution:
malignant combination with
neuroleptic sedating
syndrome medications
People > 65 y.o. Caution:
combination with
QTc-prolonging
medications
Olanzapine Drowsiness Narrow-angle Metabolized by Attention to be
EPS glaucoma CYP1A2, paid to weight
Orthostatic Prostatic CYP2D6 gain,
hypotension hyperplasia Caution when dyslipidemia,
Weight gain Diabetes combined with and diabetes
Disturbances of Long QT CYP1A2 inhibitors
glucose and lipid syndrome Smokers may need
metabolism Severe hepatic higher doses
Prolonged QTc dysfunction Caution:
Sexual dysfunction Severe renal combination with
Blood count changes dysfunction sedating
Malignant History of medications
neuroleptic malignant Caution:
syndrome neuroleptic combination with
Postinjection syndrome QTc-prolonging
syndrome medications
(IM administration)
Quetiapine EPS Long QT Metabolized by Attention to be
Prolonged QTc syndrome CYP3A4 paid to weight
Weight gain Severe hepatic Pharmacologically gain
Disturbances of dysfunction active metabolite: Attention to be
glucose and lipid Severe renal norquetiapine paid to thyroid
metabolism dysfunction Not to be hormones
Decrease in the History of combined with
thyroid hormones malignant CYP3A4 inhibitors
Cardiomyopathya neuroleptic Combination with
Epilepsy syndrome CYP3A4 inducers:
Malignant loss of the
neuroleptic therapeutic effect
syndrome Caution:
combination with
sedating
medications
Caution:
combination with
QTc-prolonging
medications
(continued)
1452 H. Himmerich and A. Hamilton
Table 1 (continued)
Implications
Mood for clinical
stabilizer Side effects Contraindications Interactions management
Risperidone Headache Long QT Metabolized via Attention to be
Somnolence syndrome CYP2D6, paid to
EPS Parkinson’s CYP3A4 prolactin levels
Prolonged QTc disease Active metabolite:
Prolactin elevation Prolactin- paliperidone
Respiratory tract dependent tumors CYP3A4 inducers:
infections Severe hepatic loss of therapeutic
Gastrointestinal dysfunction effect
disturbances Severe renal CYP2D6
Malignant dysfunction inhibitors: reduced
neuroleptic Epilepsy tolerability
syndrome History of Caution:
malignant combination with
neuroleptic sedating
syndrome medications
Caution:
combination with
QTc-prolonging
medications
Abbreviations: TCA tricyclic antidepressant, EPS extrapyramidal motor symptoms, UGT
glucuronosyltransferase, CYP cytochrome P450 isoenzyme
a
Only sporadic cases reported
Side effects of the central nervous system: Central nervous system (CNS) effects are
among the most frequent side effects associated with mood stabilizers. Lithium, for
example, can elicit tremor, cognitive problems, tiredness, shakiness of the hands, and
increased thirst. Intoxication with lithium can lead to nausea, somnolence, psycho-
motor retardation, dysarthria, ataxia, rigor, cerebral seizures, and shock. Potential
CNS side effects of antiepileptic drugs include dizziness, somnolence, aggressive-
ness, confusion, ataxia, and double or blurred vision (Jinhua et al. 2019). The most
important adverse effects of antipsychotics like olanzapine, quetiapine, and risper-
idone on the brain are extrapyramidal motor symptoms (EPS) like parkinsonism and
an increase in appetite and weight, most probably elicited by their antihistaminergic
effects (Divac et al. 2014). Atypical antipsychotics such as olanzapine can also have
unfavorable effects on glucose and lipid metabolism (Himmerich et al. 2015).
However, mood stabilizers like carbamazepine and lithium which do not have
antihistaminergic properties are also capable of inducing weight gain (Himmerich
1454 H. Himmerich and A. Hamilton
et al. 2005). Mood stabilizers often cause disturbances of sleep and sleep-wake
regulation such as tiredness, somnolence, nightmares, somnambulism, or sleepless-
ness (Rumble et al. 2015).
Cardiac side effects: As mood stabilizers modulate ion currents at cell membranes
and thus the formation and development of electrical impulses and conduction, they
bear a risk for eliciting disturbances in the conduction system of the heart. Such side
effects could present in the form of bradycardic arrhythmias, sinus node dysfunc-
tions, delays in conduction, atrioventricular blocks, or bundle branch blocks. The
electrocardiogram (ECG) may show an increase in QTc time trigger or torsade de
pointes. This highlights the importance of regular ECG monitoring. When com-
mencing mood stabilizers or considering a dose increase, extra vigilance and caution
should be taken regarding the risk associated with concomitant use of additional
medications that prolong QTc interval. The presence of other risk factors for torsade
de pointes must also be considered (Wu et al. 2015; Mehta and Vannozzi 2017;
Zaccara et al. 2017).
Gastrointestinal side effects: As with any psychiatric and somatic medication,
mood stabilizers can lead to a variety of gastrointestinal symptoms including
anticholinergic effects such as dry mouth and constipation but also diarrhea, nausea,
feelings of fullness and loss of appetite and weight, and other gastrointestinal
complaints. The liver or pancreas may be affected as reflected by liver enzyme
elevation, pancreatic enzyme elevation, dyspepsia, or pain (Jinhua et al. 2019;
Severance et al. 2015).
Hormonal and renal side effects: Lithium treatment can alter the function of the
thyroid gland and the kidneys, for example, it reduces the urine-concentrating ability
and leads to hypothyroidism, hyperparathyroidism, and weight gain. Very high
lithium doses and serum concentrations are particularly associated with impaired
renal function, hypothyroidism, and hypercalcemia. Overall, the findings on long-
term renal effects of lithium treatment remain inconsistent (Dineen et al. 2017).
The antidopaminergic effects of antipsychotic drugs in the tuberoinfundibular sys-
tem mediate their neuroendocrinological side effects such as prolactin increase
(Bostwick et al. 2009).
Allergic side effects: Mood stabilizers can have various effects on the skin; this is
particularly true with antiepileptic drugs like carbamazepine and lamotrigine. These
effects can range from general skin lesions to Stevens-Johnson syndrome or toxic
epidermal necrolysis. There is data suggesting that the major histocompatibility
complex (MHC) class IA gene, also known as human leukocyte antigen (HLA)-A
gene, conveys a risk for allergic side effects of the skin in certain populations
(Błaszczyk et al. 2015).
Hematopoietic disorders: Disorders of the hematopoietic system including eosin-
ophilia, neutropenia, leucopenia, thrombocytopenia, and agranulocytosis can arise
during treatment with various different mood stabilizers such as lithium, antiepilep-
tics like carbamazepine and valproate, or antipsychotics such as olanzapine (Curtis
2014).
Sexual dysfunction: Sexual dysfunction is common in patients with bipolar
disorder. However, the risk of this is further increased in treatment with some
Mood Stabilizers: Side Effects, Contraindications, and Interactions 1455
mood stabilizers. Lithium treatment carries some risk of sexual dysfunction (Elnazer
et al. 2015); however, carbamazepine treatment is associated with a high-risk for the
development of various sexual side effects. This is because carbamazepine has been
reported to alter the protein binding of testosterone in the serum by reducing free
testosterone (Calabrò and Bramanti 2013). Libido and orgasmic disorders were
specifically reported in women with bipolar disorder during valproate treatment
(Verrotti et al. 2016). During treatment with antipsychotics, sexual dysfunction
may be a result of their antidopaminergic effect which leads to an increase in
prolactin release (Montejo et al. 2015).
Birth defects: Birth defects are particularly prevalent in treatment with valproate;
exposure during pregnancy is associated with about three times as many major
abnormalities as usual, mainly spina bifida. More rarely, a “valproate syndrome” is
seen, which has specific characteristics including facial features that tend to evolve
with age, a triangle-shaped forehead, tall forehead with bifrontal narrowing,
epicanthic folds, medial deficiency of eyebrows, flat nasal bridge, broad nasal root,
anteverted nares, shallow philtrum, long upper lip and thin vermillion borders, thick
lower lip, and small downturned mouth. Children of mothers taking valproate during
pregnancy are at risk of lower intelligence quotients and have a higher probability of
autism. Women who intend to become pregnant should therefore switch to a
different medication or decrease their dose of valproate (Wiedemann et al. 2017).
Women who take lithium during the first trimester are also more likely to have a baby
with a birth defect, compared to pregnant women who have a mental disorder but do
not take lithium (Patorno et al. 2017).
Suicide risk: Several years ago, the “Food and Drug Administration” (FDA) had
issued a warning that anti-convulsant treatment was linked to a higher risk of suicidal
ideation and suicidal behavior. However, a meta-analysis of 5,130,795 patients
showed that there was in fact a decrease in the risk of suicide among people
with epilepsy or bipolar disorder treated with anti-convulsants (Arana et al. 2010).
However, even though the risk was relativized by meta-analyses, the warnings were
included in anti-convulsant patient information leaflets as stipulated by the FDA.
In an observational longitudinal study with 199 patients, the use of carbamazepine,
lamotrigine, and valproate showed no increased suicide rate in patients with bipolar
disorder (Leon et al. 2012). However, irrespective of the contradictory data, when
using anti-convulsants and other mood stabilizers, clinicians must carefully pay due
attention to the risk of suicidal ideation, suicide attempts, and suicide. Such caution
arises from the fact that bipolar disorder is associated with an increased risk of
suicide. It is estimated that 25% to 50% of patients with bipolar disorder will attempt
suicide at least once over their lifetime and that between 8% and 19% will complete
suicide. However, it appears that the introduction of lithium treatment was the most
important breakthrough in the prevention of suicide in bipolar disorder (Latalova
et al. 2014).
Risk populations: Certain populations fall into a higher risk category where mood
stabilizers are concerned. It is important to be aware of the specific patient groups
where these risks are elevated. One such group is those with cardiovascular system
defects and cardiac disease such as orthostatic dysregulation, heart failure, cardiac
1456 H. Himmerich and A. Hamilton
arrythmias, or QTc prolongation. People with kidney or liver problems are also at
risk, because mood stabilizers are metabolized and eliminated through these organs.
Almost all mood stabilizers have the potential to affect the hematopoietic system;
therefore, having abnormalities in the blood count should be recognized as a severe
risk factor for further complications where a mood stabilizer is being considered.
Considering all of this, it is paramount that we carry out a full physical examination;
obtain the necessary blood tests, including electrolytes, liver function tests, and
pancreatic enzymes and full blood count; and perform an ECG before commencing
mood stabilizers (Benkert and Hippius 2019; Mehta and Vannozzi 2017; Zaccara
et al. 2017).
As these problems frequently occur in elderly patients, this age group must be
deemed a specific risk population. In this regard, it is worth mentioning that
risperidone is the only currently approved atypical antipsychotic for patients with
dementia and pronounced psychotic or behavioral disorders.
Lithium: Many patients take lithium on a long-term basis without any unwanted
effects. Side effects often occur at the beginning of lithium treatment but disappear
spontaneously. Initial side effects should, therefore, not necessarily lead to discon-
tinuation of treatment. Thus, it is of importance to inform the patient about these
potentially transient side effects and provide some reassurance that they may even-
tually settle down but should be monitored closely.
Lithium can have a number of psychiatric and neurological side effects such as
tremor, cognitive problems, tiredness, and muscle weakness; renal side effects such
as polyuria and polydipsia, renal impairment and dysfunction, and glomerulonephri-
tis; endocrine adverse effects such as goiter, hypothyroidism, hyperparathyroidism,
and weight gain; and cardiovascular side effects such as repolarization disturbances
and arrythmias. It may also lead to unwanted effects regarding fluid and electrolyte
balance such as facial or ankle edema, the hematopoietic system like leukocytosis,
the skin such as alopecia, folliculitis, pruritus, or exacerbations of psoriasis; and the
gut like diarrhea, nausea, feelings of fullness, and loss of appetite. The most common
side effects include increased urination, shakiness of the hands, and increased thirst.
If used during pregnancy, lithium can cause birth defects. It appears to be safe to use
while breastfeeding. For further information on specific side effects of lithium, see
chapter ▶ “Mood Stabilizers: Lithium.”
Special note: As lithium is eliminated renally, renal function should be tested
before initiating lithium therapy. Symptoms of lithium toxicity can occur at concen-
trations above 1.2 mmol/l, in some cases even at lower lithium plasma concentra-
tions. Signs of lithium toxicity include nausea, vomiting, diarrhea, hand tremor,
fatigue, psychomotor slowdown, vigilance reduction, dizziness, dysarthria, ataxia,
rigor, hyperreflexia, fasciculations, cerebral seizures, and shock. In extreme
cases, lithium toxicity can lead to loss of consciousness, coma, and cardiac arrest
(Paulzen et al. 2019).
Mood Stabilizers: Side Effects, Contraindications, and Interactions 1457
Valproate: Common side effects of the central nervous system include drowsiness,
nausea, vomiting, tremor, and paresthesia. Occasionally, headache, irritability, hyper-
activity, confusion, stupor, spasticity, or ataxia may occur. Another frequent side effect
is hyperammonemia. Hypersalivation, weight gain or weight loss, diarrhea, and
hepatic problems including liver failure, pancreatitis, and pancreas damage are uncom-
mon but possible and serious adverse events. Rare events under valproate treatment
comprise peripheral edema, bleeding, and impaired bone marrow function including
aplasia, agranulocytosis, macrocytic anemia or macrocytosis, lupus erythematosus,
and erythema multiforme. Further possible but rare adverse drug effects include
disorders of the sexual organs such as dysmenorrhea or polycystic ovaries, temporary
hair loss, as well as Stevens-Johnson syndrome, Lyell’s syndrome, and reversible
hypothermia in isolated cases. In some studies, diminished bone density has been
detected after long-term administration of valproate. For further information on
specific side effects of valproate, see chapter ▶ “Mood Stabilizers: Valproate.”
Special note: When valproate is taken together with anticoagulants or anti-
aggregants, this may increase the risk of bleeding. Therefore, in these cases, it is
recommended to regularly monitor blood clotting. In high-risk patients who are
immobilized over long periods without sun exposure or low calcium intake, Vitamin
D substitution should be considered. Due to the risk of birth defects, pregnancy
should be excluded before starting treatment with valproate (Paulzen et al. 2019).
Asenapine: Very frequent side effects of asenapine are anxiety and somnolence.
Frequent side effects include appetite and weight gain, akathisia, dystonia, muscle
rigidity, dyskinesia, parkinsonism, sedation, fatigue, dizziness, taste disorders, oral
hypesthesia, and liver enzyme elevation. Occasionally, hyperglycemia, syncope,
dysarthria, seizures, prolonged QTc time, bundle branch block, hypotension, ortho-
stasis, swollen tongue, sexual dysfunction, amenorrhea, EPS, sinus bradycardia and
tachycardia, dysphagia, glossodynia, and oral paresthesia are seen. In rare cases,
neutropenia, accommodation disorders, malignant neuroleptic syndrome, rhabdo-
myolysis, pulmonary embolism, gynecomastia, galactorrhea, and sleepwalking have
been described (Müller and Benkert 2019). For further information on specific side
effects of asenapine, see chapter ▶ “Mood Stabilizers: Asenapine.”
Olanzapine: The risk of weight gain and metabolic syndrome is very high with
olanzapine treatment. However, beneficial psychotherapeutic, behavioral, and phar-
macological measures for the prevention and reduction of weight gain associated
with olanzapine have meanwhile been tested successfully.
Very frequent side effects of olanzapine include drowsiness, increased appetite
and weight gain, hyperprolactinemia, and orthostatic hypotension. Further adverse
effects frequently encountered are fatigue, asthenia, dizziness, akathisia, parkinson-
ism, dyskinesia, eosinophilia, leukopenia, neutropenia, increased glucose, triglycer-
ide, and cholesterol levels, glucosuria, transient elevation of liver enzymes, transient
anticholinergic effects such as constipation and dry mouth, rash, arthralgia, edema,
erectile dysfunction, decreased libido, and fever. Rare side effects are bradycardia,
QTc time prolongation, malignant neuroleptic syndrome, hepatitis including hepa-
tocellular or cholestatic liver damage, development or worsening of diabetes plus
accompanying ketoacidosis or coma, thrombocytopenia, hypothermia, pancreatitis,
Mood Stabilizers: Side Effects, Contraindications, and Interactions 1459
Contraindications
Interactions
taken when combining those mood stabilizers with other medications that prolong
QTc time. If arterial hypotension occurs during treatment with the antipsychotics
olanzapine, quetiapine, or risperidone, it should not be treated with epinephrine but
norepinephrine, because of the risk of a further drop in blood pressure due to
a reverse epinephrine response. The effect of dopamine agonists such as bromocrip-
tine or L-dopa will be weakened under treatment with antidopaminergic medication
like the antipsychotic mood stabilizers.
Lithium: Lithium has interactions with other psychopharmacological agents but
also with drugs used for the treatment of physical diseases.
It has been discussed in the literature that monoamine oxidase inhibitors (MAOIs)
may potentially increase lithium plasma levels. Selective serotonin and norepineph-
rine reuptake inhibitors (SSNRIs), duloxetine and venlafaxine, have been reported to
possibly increase lithium-induced neurotoxicity. A severe and serious side effect
of such a combination is central serotonin syndrome. Tricyclic antidepressants
(TCA) can increase tremor. However, lithium has been reported to improve the
antidepressant efficacy of SSNRIs, duloxetine, venlafaxine, and TCAs, when used to
augment their antidepressant effect. According to the scientific literature, there is no
indication of a specific interaction between lithium and mirtazapine.
Lithium has been shown to increase side effects of antipsychotics including EPS
and sometimes neurotoxicity leading to electroencephalographic (EEG) abnormali-
ties, delirium, and convulsions. It also possibly increases the risk of malignant
neuroleptic syndrome. In very rare cases, irreversible movement disorders with
persistent EEG changes have been described. Lithium also increases the risk for
severe side effects including neurotoxicity of carbamazepine and phenytoin.
Angiotensin-converting enzyme (ACE) inhibitors and angiotensin II receptor
blockers (ARBs) amplify the side effects of lithium while having the potential to
increase its plasma concentration and thus increasing the risk of nephrotoxicity;
therefore, weekly lithium plasma concentration measurements are recommended in
the first 8 weeks of such combinations. Further drugs associated with an increase
of lithium plasma concentrations and thus lithium-induced side effects or toxicity
include antibiotics like ampicillin, tetracyclines, aminoglycosides, and metronida-
zole as well as certain nonsteroidal anti-inflammatory drugs (NSAIDs) such as
cyclooxygenase (COX)-2 inhibitors, calcium channel blockers, ketamine, and
methyl-dopamine. This is why it is recommended to measure lithium plasma levels
frequently. Lithium may attenuate the hypotensive effect of clonidine. If lithium is
delivered together with digoxin, the risk of arrhythmias is reinforced, but the
antimanic effect of lithium may be weakened. Diuretics such as thiazide diuretics
like hydrochlorothiazide, potassium-sparing diuretics like amiloride and triamterene,
as well as loop diuretics like furosemide decrease lithium renal clearance, posing the
risk of lithium toxicity. Acetazolamide, sodium bicarbonate, and methylxanthines
such as theophylline and caffeine lead to an increased lithium excretion with
decreased lithium plasma concentrations. Potassium iodide increases lithium’s
thyrostatic action. When used together with muscle relaxants such as pancuronium
and suxamethonium, lithium can lead to a prolonged neuromuscular blockade.
Lithium can mitigate the hypertensive effect of sympathomimetics. The use
1462 H. Himmerich and A. Hamilton
CYP2C19 by increasing the drug levels of TCAs in the plasma. Concomitant use of
valproate and anticoagulants or antiaggregants like acetylsalicylic acid leads to an
increased bleeding tendency, which is why regular checks of blood coagulation
markers are recommended. Valproate inhibits the glucuronidation and thus the
degradation of lamotrigine. Lamotrigine, in turn, induces glucuronosyltransferases
and thus accelerates the metabolism of valproate. When combined with olanzapine,
it can cause both a drop and an increase in the plasma concentration of olanzapine
since valproate inhibits glucuronosyltransferases (UGTs) in the short term but
induces UGTs and CYP1A2 in the long term. When combined with enzyme-
inducing substances such as carbamazepine, imipenem, mefloquine, meropenem,
panipenem, phenytoin, primidone, or rifampicin, plasma levels of valproate should
be controlled regularly, as the levels may fall.
Asenapine: Asenapine is metabolized via CYP1A2. Therefore, when
co-administered with CYP1A2 inhibitors like fluvoxamine and ciprofloxacin, the
dose should be reduced. Smokers may need to have their dose increased, because of
induction of CYP1A2.
Olanzapine: When combined with valproate, treatment with olanzapine is asso-
ciated with an increased risk of thrombo-, leuko-, or neutrocytopenia. This combi-
nation with valproate can cause both a drop and an increase of the plasma
concentration of olanzapine, since valproate inhibits UGTs short term but induces
and UGTs and CYP1A2 long term. The combination should therefore only be used
with repeated measurements of plasma concentrations of olanzapine and valproate.
Extra care should be taken when combining olanzapine with inhibitors of CYP1A2,
e.g., fluvoxamine; this will lead to slowed down degradation and thus increase in
olanzapine plasma level. Smokers have on average 30% lower plasma concentra-
tions than non-smokers, and an increase in plasma concentration can be observed
during the week after termination of smoking.
Quetiapine: Beware of combinations with other medications that prolong the QTc
time or can lead to hypokalemia such as amiodarone, quinidine, chlorpromazine,
gatifloxacin, pentamidine, procainamide, methadone, moxifloxacin, thioridazine, or
ziprasidone. Quetiapine should not be combined with CYP3A4 inhibitors such as
clarithromycin and erythromycin, because this may lead to a five- to eightfold
increase in plasma concentrations of quetiapine. CYP3A4 inducers like carbamaz-
epine, phenytoin, or St. John’s wort lower the plasma level and lead to a possible loss
of the effect of quetiapine. Lamotrigine lowers quetiapine plasma levels most
probably by induction of glucuronidation. Therefore, in this combination, measure-
ment of quetiapine plasma levels and, if necessary, dose adjustments are necessary, if
lamotrigine is discontinued.
Risperidone: As risperidone is metabolized by CYP3A4 and CYP2D6, combi-
nation with carbamazepine or St. John’s wort preparations should happen alongside
close monitoring and control of plasma concentrations, because both substances
induce CYP3A4. When combined with CYP2D6 inhibitors like paroxetine or
terbinafine, an increase in risperidone plasma concentration and reduction in risper-
idone tolerability can be expected. Risperidone may enhance the effects
of antihypertensives.
1464 H. Himmerich and A. Hamilton
Future Perspectives
Pharmacogenetic testing: Until recently, psychiatrists had little choice but to pre-
scribe medications without knowing in advance how their patients might genetically
respond. Pharmacogenetic testing, while not yet widely implemented in clinical
psychiatry, offers promising and exciting prospects for the future. It is assumed
that this technology can be implemented soon to predict therapeutic effects, potential
side effects, and drug interactions during therapy with psychopharmacological
agents such as mood stabilizers in an individual patient. One of the biggest barriers
to implementing this is the current lack of translation of evidence-based recommen-
dations and standardization of genetic testing panels. However, minimum gene and
allele sets for pharmacogenetic testing in psychiatry have been proposed. These sets
include cytochrome P450 enzymes as well as MHC genes like HLA-A (Bousman
et al. 2019).
Psychopharmacology of the gut-brain axis: Our knowledge about how the human
gut microbiome is involved in the development and regulation of physiological
systems is increasing, and evidence has accumulated to suggest a role for the gut
microbiome in drug response to psychopharmacological agents and the side effects
of psychiatric drug treatment (Kanji et al. 2018). However, the complex relationships
between the enteric nervous system, the gut microbiota, and the central nervous
system during psychiatric conditions and the mechanisms how the relations between
these systems are influenced by psychopharmacological agents are currently far from
being understood.
Inflammation and mood stabilizers: There is growing evidence that bipolar and
unipolar affective disorders are linked to changes within the immune system (Mao
et al. 2018). Additionally, mood stabilizers have been shown to modulate cytokine
signaling, for example, the interleukin (IL)-1β, IL-2, and IL-22 and tumor necrosis
factor (TNF)-α signaling pathways (Himmerich et al. 2013). Therefore, it seems
promising to abandon the relatively narrow pharmacological perspective that
predominantly looks at how neurotransmitter signaling is involved in the pathophys-
iology of psychiatric disorders and their therapy but to broaden the view to include
the signaling pathways of immune cells within the brain and the body such as
astrocytes, microglia, dendritic cells, macrophages, and T helper cells. This will
potentially help to explain side effects such as bone marrow toxicity with carbamaz-
epine; Stevens-Johnson syndrome and Lyell’s syndrome associated with
lamotrigine; liver failure or pancreatitis caused by valproate; swollen tongue from
asenapine; rhabdomyolysis caused by olanzapine; jaundice, angioedema, and
Mood Stabilizers: Side Effects, Contraindications, and Interactions 1465
Cross-References
References
Arana A, Wentworth CE, Ayuso-Mateos JL, Arellano FM. Suicide-related events in patients treated
with antiepileptic drugs. N Engl J Med. 2010;363:542–51.
Benkert O, Hippius H. Kompendium der Psychiatrischen Pharmakotherapie. 12th ed. Heidelberg/
Berlin: Springer; 2019.
Błaszczyk B, Lasoń W, Czuczwar SJ. Antiepileptic drugs and adverse skin reactions: an update.
Pharmacol Rep. 2015;67:426–34.
Bostwick JR, Guthrie SK, Ellingrod VL. Antipsychotic-induced hyperprolactinemia.
Pharmacotherapy. 2009;29:64–73.
Bousman C, Maruf AA, Müller DJ. Towards the integration of pharmacogenetics in psychiatry:
a minimum, evidence-based genetic testing panel. Curr Opin Psychiatry. 2019;32:7–15.
Calabrò RS, Bramanti P. Carbamazepine-related sexual disorders: beyond hormonal changes! J Sex
Med. 2013;10:1440.
Can A, Schulze TG, Gould TD. Molecular actions and clinical pharmacogenetics of lithium therapy.
Pharmacol Biochem Behav. 2014;123:3–16.
Choi H, Morrell MJ. Review of lamotrigine and its clinical applications in epilepsy. Expert Opin
Pharmacother. 2003;4:243–51.
Chopko TC, Lindsley CW. Classics in chemical neuroscience: risperidone. ACS Chem Nerosci.
2018;9:1520–9.
Citrome L. Asenapine review, part I: chemistry, receptor affinity profile, pharmacokinetics and
metabolism. Expert Opin Drug Metab Toxicol. 2014;10:893–903.
Curtis BR. Drug-induced immune neutropenia/agranulocytosis. Immunohematology.
2014;30:95–101.
Dailey JW, Reith ME, Steidley KR, Milbrandt JC, Jobe PC. Carbamazepine-induced release of
serotonin from rat hippocampus in vitro. Epilepsia. 1998;39:1054–63.
Dineen R, Bogdanet D, Thompson D, Thompson CJ, Behan LA, McKay AP, Boran G, Wall C,
Gibney J, O’Keane V, Sherlock M. Endocrinopathies and renal outcomes in lithium therapy:
impact of lithium toxicity. QJM. 2017;110:821–7.
1466 H. Himmerich and A. Hamilton
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1471
Anti-convulsants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1471
Lithium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1472
Antipsychotics (AP) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1473
Course and Duration of Treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1475
Acute Treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1475
Mood Stabilizers for Manic Symptoms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1476
Mood Stabilizers for Depressive Symptoms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1477
Mood Stabilizers for Mixed Symptoms/Mixed Features . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1478
Mood Stabilizers for Agitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1479
Maintenance Treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1480
Withdrawal Syndromes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1482
Anti-convulsants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1483
Lithium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1483
Antipsychotics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1484
Resistance to Therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1484
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1486
G. Schoretsanitis (*)
Psychiatry Research, The Zucker Hillside Hospital, New York, USA
e-mail: george.schor@gmail.com
M. Paulzen
Department of Psychiatry, Psychotherapy and Psychosomatics and JARA – Translational Brain
Medicine, RWTH Aachen University, Aachen, Germany
Alexianer Hospital Aachen, Aachen, Germany
e-mail: m.paulzen@alexianer.de
Abstract
Despite different available pharmacological frameworks, mood stabilizers are
widely acknowledged as a group of agents prescribed in the treatment of bipolar
disorders. In this chapter we use an indication-based approach to describe acute and
maintenance treatment with mood stabilizers encompassing lithium, anti-convul-
sants, and antipsychotics in adult patients. Primary and secondary clinical evidence
as well established guidelines are reviewed focusing on the included agents.
Further, we elaborate on clinical manifestations related to the discontinuation of
mood stabilizers, although these drugs do not typically possess the potential for
withdrawal syndromes. Lastly, therapeutic options for treatment-refractory forms of
bipolar disorders including pharmacological and non-pharmacological interven-
tions are discussed. Summing up, the armamentarium of mood stabilizers has not
been essentially revised during the past decades; older agents still comprise a core
part of routine pharmacotherapy in bipolar disorders. However, prescription of
mood stabilizers remains a tricky business, and clinicians need to develop a
comprehensive understanding of profiles of action and adverse drug reactions in
order to enhance efficacy and safety.
Abbreviations
5-HT Serotonin
ADR Adverse drug-induced reaction
AP Antipsychotic treatment
BP-I Bipolar I disorder
CANMAT Canadian Network for Mood and Anxiety Treatments
CYP Cytochrome P450
ECT Electroconvulsive therapy
EMA European Medicines Agency
ER Extended-release
FDA US Food and Drug Administration
FGA First-generation antipsychotics
GABA Gamma-aminobutyric acid
GSK-3 Glycogen synthase kinase 3
HDAC Histone deacetylase
IR Immediate-release
ISBD International Society for Bipolar Disorders
LAI Long-acting injectable
P-gp P-Glycoprotein
PKC Protein kinase C
RAI Rapid-acting injectable
SGA Second-generation antipsychotics
TMS Transcranial magnetic stimulation
UGT Uridine 50 -diphospho-glucuronosyltransferase
Mood Stabilizers: Course and Duration of Therapy, Withdrawal Syndromes, and. . . 1471
Introduction
There is no unanimously accepted definition for mood stabilizers, and the clinical
utility of this category or classification may be disputed (Malhi et al. 2018). In fact,
there is not a distinct pattern of common properties for the agents we may refer to as
mood stabilizers other than their application in the treatment of bipolar disorders.
This approach reproduces a clear indication-based nomenclature (Zohar et al. 2015).
In fact, the group of mood stabilizers encompasses medications of strikingly hetero-
geneous pharmacologic profiles, originally introduced for utterly different indica-
tions. By trying to narrow the definition of mood stabilizers, groups or classes of
drugs, such as antidepressants, that have been included in guidelines over the years
(Nierenberg et al. 2006) are no longer suggested for the treatment of bipolar
disorders without a concomitantly applied mood stabilizer. Another corollary of
this issue is that mood stabilizers are not exclusively prescribed for bipolar disorders
but may be used in terms of off-label therapeutic regimens in other psychiatric
diseases as well (Fond et al. 2018; Crawford et al. 2018; Rajaratnam et al. 2017).
The scope of this chapter is to review evidence for compounds that are used in the
treatment of bipolar disorders in adults foregoing a rigorous definition of mood
stabilizers. We focus on medications holding approval from the US Food and Drug
Administration, FDA, or the European Medicines Agency, EMA, for the treatment
of bipolar disorders. Agents with approval for bipolar disorders in children or
adolescents will not be discussed in this chapter.
An established classification of mood stabilizers follows the pharmacological class of
these drugs: (1) a subset of anti-convulsants, (2) lithium, and (3) antipsychotics (AP).
Following an alternative perspective, we may classify these agents based on the polarity
index (Popovic et al. 2012); this newly introduced concept is based on antimanic and/or
antidepressant preventive properties of mood stabilizers prescribed in maintenance
treatment of bipolar disorders. According to this classification, at one pole of the polarity
index, the anti-convulsant lamotrigine is the unique agent with relapse-preventing
properties in bipolar disorders with predominantly depressive episodes, whereas on
the other side, risperidone long-acting injectable (LAI) and aripiprazole display pro-
nounced antimanic properties. Another conceptualization framework is integrating
knowledge about distinct properties of different drugs targeting acute symptoms or
addressing prophylactic properties (Malhi et al. 2018). This elaborate four-domain
model (antimanic vs. antidepressant and acute vs. maintenance) aims to assist clinicians
during the decision-making process by considering additional treatment aspects.
Anti-convulsants
HDAC. Nevertheless, the precise mechanisms of action remain only partially under-
stood, and current knowledge predominantly derives from in vitro models. Compa-
rably undiscovered is the pharmacokinetics of valproate, which is mainly
metabolized via UGT1A3, UGT1A6, and UGT2B7, but also via CYP2A6,
CYP2B6, CYP2C9, CYP2C19, and by β-oxidation. Furthermore, valproate slightly
inhibits cytochrome P450 (CYP) 2C9 as well as other isoenzymes with clinical
effects remaining unclear. Valproate is particularly efficacious for acute mania
(Bowden et al. 1994), for which it holds an approval from FDA in both available
forms, valproic acid and divalproex (a stable coordination compound comprised of
sodium valproate and valproic acid). The latter is the enteric-coated counterpart of
the first and may be better tolerable (Schwartz et al. 2000).
Lamotrigine is a second-generation anti-convulsant presumably acting on sodium
channels. Its main metabolic pathway is a glucuronidation catalyzed by uridine
diphosphate-glucuronosyltransferase 1A4 (UGT1A4) leading to its major metabolite,
lamotrigine-2-N-glucuronide (Paulzen et al. 2018). Lamotrigine shows antidepressant
properties in the treatment of major depressive disorder; however, its approval is only
in maintenance treatment for the prophylaxis of depressive episodes in bipolar disorder
when depressive episodes are predominant. Since lamotrigine was reported to be of
low teratogenic risk, there has been a shift in prescription patterns of anti-convulsants
in pregnant women with lamotrigine now being the most commonly prescribed anti-
convulsant or mood-stabilizing drug (Petersen et al. 2016).
The third anti-convulsant approved for patients with bipolar disorders is carba-
mazepine. Its primarily CYP3A4/5- and CYP2C8-based metabolic pathway (Hata et
al. 2008) leads to carbamazepine 10,11-epoxide, which also possesses anti-convul-
sant activity. Like other mood stabilizers, carbamazepine acts at sodium channels in
terms of its anti-convulsant action, whereas its mood-stabilizing properties may be
based on inositol-signaling mechanisms. Its extended-release (ER) form holds
approval for the treatment of acute mania. The prescription of carbamazepine usually
introduces a challenge for clinicians due to its high potential for pharmacokinetic
interactions, with carbamazepine acting as a strong inductor mainly on CYP3A4 and
P-glycoprotein (P-gp) (Schoretsanitis et al. 2016).
In the course of the past decades, some other anti-convulsant agents have also
been prescribed as mood stabilizers: levetiracetam, topiramate, tiagabine,
gabapentin, zonisamide, and oxcarbazepine have received considerable attention.
Notwithstanding, the evidence displaying a considerable effect for these drugs is still
lacking, and therefore, these therapeutic regimens remain off-label.
Lithium
Introduced in psychiatry almost 70 years ago (Cade 1949), lithium is probably the
mood stabilizer with the longest clinical application and still the reference drug in the
treatment of bipolar disorders. The mechanisms of action for this prototypical mood
stabilizer have been investigated better than any other agent from this group. Lithium
seems to share common pathways with valproate (Harwood and Agam 2003); modes
Mood Stabilizers: Course and Duration of Therapy, Withdrawal Syndromes, and. . . 1473
Antipsychotics (AP)
research in the past decades. After almost 70 years, it is still prescribed, though in
smaller numbers than in the past. It possesses affinity for a large group of receptors, but
its primary effect is considered the dopaminergic antagonism. Its main metabolic
pathway comprises a 7-hydroxylation catalyzed by CYP2D6 and partially by
CYP1A2 (Yoshii et al. 2000). Another FGA, probably the most widely prescribed,
haloperidol, is approved by the EMA, but not the FDA for acute mania. Haloperidol is a
high-affinity D2-receptor antagonist and has various metabolic pathways involving
CYP2D6 and CYP3A4 as well as UGT. The most usual adverse effect related to
haloperidol treatment is extrapyramidal symptoms. Across the agents described in this
section, loxapine is the only drug that is delivered in an inhalable application form and
holds approval for agitation in schizophrenia or bipolar disorder. Although being
originally classified as an FGA, it has been debated that it mainly acts as a SGA with
moderate affinity for D1- and D2-dopamine as well as 5-HT2-receptors (Glazer 1999).
The application device of the drug leads to rapid absorption via inhalation of the agent,
which is then hepatically metabolized through CYP1A2, CYP3A4, and CYP2D6.
Slightly deviating from the principle of dopaminergic antagonism for antipsy-
chotics, lurasidone has a high affinity for D2-receptors, but does not bind to D4-
receptors, which has been suggested as contributing to procognitive effects (Murai et
al. 2014). The hepatic metabolism of lurasidone involves CYP3A4 with metabolites
lacking clinical relevance. Due to its reasonable safety profile, lurasidone may
qualify as a reliable option in treatment of older patients (Vasudev et al. 2018).
Also noteworthy is that its indications are depressive episodes.
One of the first SGAs entering the market was olanzapine, which is approved for
bipolar disorders as monotherapy or augmentation to lithium or valproate. It has
been also approved in the treatment of bipolar disorders in combination with
fluoxetine, although this product has been available only limited to several countries.
Mainly metabolized by CYP1A2, olanzapine acts on the dopaminergic system, but
also displays a considerable affinity for histaminic receptors. This affinity may
explain the propensity of olanzapine to cause metabolic abnormalities, which in
turn is the major complication of olanzapine-based maintenance treatment
(Schoretsanitis et al. 2018b). The rapid-acting injectable form of olanzapine holds
approval for acute agitation in bipolar disorders.
Very similar to olanzapine is quetiapine in terms of its chemical structure.
Quetiapine is marketed in two pharmacokinetically distinct forms, as an immediate-
and as an extended-release formulation, and is metabolized by CYP3A4 leading to
multiple metabolites, which barely contribute to psychotropic effects (Mauri et al.
2018). Its action might rather relate to serotonergic antagonism, whereas
quetiapine’s affinity for D2-receptors is low. Quetiapine holds indication for acute
treatment of manic episodes as monotherapy or as an adjunctive treatment to lithium
or valproate, acute treatment of depressive episodes, and maintenance treatment (as
monotherapy or combined with lithium or valproate).
Risperidone is a benzisoxazole derivative acting on both serotonergic and dopa-
minergic neurotransmitter systems. Its metabolic pathways include CYP2D6 and
CYP3A4, and the active metabolite, 9-hydroxyrisperidone, displays approximately
70% of the pharmacological activity of the parent compound. For risperidone, the
Mood Stabilizers: Course and Duration of Therapy, Withdrawal Syndromes, and. . . 1475
LAI formulation has been also approved as monotherapy or adjunctive therapy for
maintenance treatment of bipolar I disorder (BP-I).
Lastly, ziprasidone is a SGA with approval for acute manic or mixed episodes of
bipolar disorders. It antagonizes D2- and 5-HT2A-receptors and is highly metabo-
lized following multiple oxidations. Patients under ziprasidone-based treatment may
be at risk for QTc prolongation, and therefore, monitoring is suggested.
Apart from the FDA-approved antipsychotics for the treatment of bipolar disor-
ders, there is pharmacoepidemiological data for other antipsychotics as well
(LÄhteenvuo et al. 2018). Although off-label, pharmacological interventions may
be very effective; as these agents do not hold approval for bipolar disorders,
clinicians need to proceed with caution.
The indications for mood stabilizers include acute and maintenance treatment of
bipolar disorders. It is plausible that these two types of treatment options introduce
quintessentially different challenges for clinicians, as they differ in terms of goal,
course, and duration. The propensity of bipolar episode to changes regarding the
polarity and the severity of the symptoms introduces additional challenges for
clinicians. A typical pitfall in the management of bipolar episodes is the prescription
of agents, which may effectively treat depressive symptoms at the expense of
inducing manic symptoms (Salvadore et al. 2010). In fact, the heterogeneity of the
action profiles for the various mood stabilizers furthermore sets obstacles for pre-
scribers. For instance, some of these medications are starkly efficacious in amelio-
rating acute mania, but do not possess antidepressant properties, whereas other
agents reduce the risk of reoccurrence of mood symptoms, but are ineffective in
the treatment of acute episodes. Therefore, clinicians need to be highly aware of the
risk for rapid changes in psychopathology, while a more comprehensive understand-
ing of treatment options will facilitate improvement of treatment outcomes.
Acute Treatment
Clinicians would claim that the most characteristic acute bipolar presentation referred
for treatment is manic states, but symptom clusters may also include hypomanic,
depressive, or mixed states (Table 1).
The following agents hold FDA approval for acute treatment: aripiprazole (oral
and rapid-acting injection), asenapine (as monotherapy or adjunctive to lithium or
valproate), carbamazepine, cariprazine, chlorpromazine, lithium, olanzapine (oral as
monotherapy, adjunctive to fluoxetine, lithium or valproate and rapid-acting injec-
tion), quetiapine (as monotherapy or adjunctive to lithium or valproate), risperidone
(oral as monotherapy or adjunctive to lithium or valproate and long-acting),
valproate, and ziprasidone. Haloperidol is approved for the indication of mania by
the EMA, but not the FDA.
1476 G. Schoretsanitis and M. Paulzen
Table 1 Medications with FDA indication for acute treatment of bipolar disorders
Drugs Mania Mixed Depression
Aripiprazole √ √ x
Asenapine √ √ x
Asenapine + lithium/valproate √ √ x
Carbamazepine ER √ √ x
Cariprazine √ √ x
Chlorpromazine √ x x
Divalproex sodium √ √ x
Haloperidol √a x x
Lithium (ER and IR) √ x x
Lurasidone x x √
Lurasidone + lithium/valproate x x √
Olanzapine √ √ x
Olanzapine + fluoxetine x x √
Olanzapine + lithium/valproate √ √ x
Quetiapine IR √ x √
Quetiapine IR + lithium/valproate √ x x
Quetiapine ER √ √ √
Quetiapine ER + lithium/valproate √ √ x
Risperidone √ √ x
Risperidone + lithium/valproate √ √ x
Valproic acid √ √ x
Ziprasidone √ √ x
FDA US Food and Drug Administration, EMA European Medicines Agency, ER extended-release,
IR immediate-release
a
Approved by the EMA, but not the FDA
Lithium remains the most intensively studied agent in the treatment of bipolar
disorders over the last years. It is the only mood stabilizer with antisuicidal effects
and is widely embraced as the mainstay of treatment for patients with suicidal
ideation. Therapeutic drug monitoring of lithium concentrations is an inherent part
of a treatment with lithium. Measuring blood concentrations enables clinicians to
minimize toxicity and enhance treatment efficacy, even for drugs with narrow
therapeutic windows. For patients with acute manic episodes, lithium concentrations
ranging between 0.6 and 1.0 mmol/l are suggested (Malhi et al. 2016). Early partial
response to lithium may predict remission of manic symptoms (Machado-Vieira et
al. 2013). For patients not responding to lithium, anti-convulsants, i.e., valproate and
carbamazepine, may offer efficacious treatment alternatives (Grunze et al. 2009). In
particular, patients with psychotic features or rapid cycling may profit from
lamotrigine or valproate (Lieberman and Tasman 2006). These patients may also
respond to carbamazepine when not responding to lithium. As valproate and
Mood Stabilizers: Course and Duration of Therapy, Withdrawal Syndromes, and. . . 1477
Clinicians have been barely aware of the clinical and neurobiological differences
between unipolar and bipolar depression, and the treatment of depressive presentations
either as unipolar depressive episodes or as bipolar depression has been less extensively
investigated compared to acute mania. The FDA has provided approval only for three
agents, i.e., quetiapine, lurasidone, and the combination of olanzapine and fluoxetine
for acute depressive episodes. The antidepressant effect of quetiapine is dose-indepen-
dent, at least over 300 mg daily (Yatham et al. 2018). For lurasidone, clinicians may
consider its reasonable safety profile and, therefore, a valuable role in the treatment of
acute depression in older patients (Sajatovic et al. 2016). Although evidence for
olanzapine as monotherapy in acute bipolar depression is equivocal, the combination
1478 G. Schoretsanitis and M. Paulzen
of olanzapine and fluoxetine holds FDA approval. The prescription of this therapeutic
regimen may be limited through tolerability issues (Silva et al. 2013), as fluoxetine does
not seem to reverse the metabolic effects of olanzapine (Bustillo et al. 2003).
Apart from the approved agents, there is robust evidence for some more com-
pounds in terms of off-label therapies. As previously mentioned, lithium is a first-
line medication for patients with suicidal ideations. When prescribed for depressive
patients, lithium concentrations in the patient’s blood are suggested to be in a range
between 0.4 and 0.8 mmol/l (Malhi et al. 2016). Valproate may be also considered
when first-line agents fail to be effective (Yatham et al. 2018). Furthermore, a recent
meta-analysis implies a potential benefit for a subset of antidepressants as an add-on
therapy in the short term, whereas long-term prescription contributes to the risk of
treatment emergent affective switches into mania or hypomania (Mcgirr et al. 2016).
The criterion of early response, i.e., response to an agent within 2 weeks, applies in
the treatment of acute depression as well.
Apart from the unipolar or pure forms of mood symptom clusters such as manic or
depressive episodes, clinicians may also encounter mixed states related to bipolar
disorders. According to the nomenclature of the Diagnostic and Statistical Manual of
Mental Disorders, 5th revision, DSM-5, the former description of “mixed symptoms”
has been changed into “with mixed features.” Presentations of mixed features are
frequently trickier to recognize and even more complicated to efficiently treat. Mixed
features may be frequently misclassified as manic or depressive presentations and often
receive less recognition as a distinct clinical pattern. However, reservations need to be
made, as the clinical practice and wisdom from the other two bipolar presentations
cannot be carried to mixed states. Nevertheless, it is no surprise that the armamentarium
of approved mood stabilizers with effects on mixed states looks very similar to the
group of agents approved for manic symptoms; exceptions regard chlorpromazine,
lithium, and the immediate-release form of quetiapine (as monotherapy or adjunctive),
which only hold approval for manic symptoms, but not for mixed states.
The reasons underlying these exceptions may relate to marketing issues. How-
ever, for lithium, it is of particular interest that it is, indeed, considered less
efficacious for patients with mixed features (Swann et al. 1997). Notwithstanding,
as patients with mixed features may display an enhanced propensity to switch from
depressive to manic symptoms and back (Fornaro et al. 2018), lithium may be
considered, though in combination with other mood stabilizers. The WFSBP Guide-
lines primarily suggest olanzapine as monotherapy or in combination for the treat-
ment of manic syndromes with mixed features, with aripiprazole monotherapy as a
second-line option (Grunze et al. 2018). The role of SGAs for mixed features is also
highlighted by the CANMAT guidelines (Yatham et al. 2018). These guidelines do
not only include approved agents but also antipsychotics that are prescribed off-label
such as paliperidone, which is suggested as first-line monotherapy for manic symp-
toms in acute episodes with mixed features.
Mood Stabilizers: Course and Duration of Therapy, Withdrawal Syndromes, and. . . 1479
Across the licensed agents, loxapine offers a less invasive alternative to injectable
medications, because it is applied as an inhalation powder. The therapeutic outcome
may be dose-dependent, and 10 mg loxapine may counter agitation symptoms more
successfully than 5 mg (Dundar et al. 2016). A recent study directly compared up to
two doses of 9.1 mg inhaled loxapine with a single intramuscular aripiprazole
injection for agitation in a mixed sample of patients with schizophrenia or bipolar
disorders (San et al. 2018). The comparison favored loxapine for all assessments
during 24 h after drug administration. Lastly, rapid-acting intramuscular injectable
olanzapine is also approved for agitation in bipolar disorders. Comparative analyses
highlighted the superiority of olanzapine in reducing agitation symptoms 1 h after
application (Dundar et al. 2016).
Apart from the approved agents, there is also data that supports the application of
some additional psychotropic drugs. For instance, intramuscular lorazepam has been
reported as equally efficacious as olanzapine (Meehan et al. 2001) and is also
suggested as first-line treatment for the management of agitation by the CANMAT
(Yatham et al. 2018). The same guidelines include across second-line options
asenapine, haloperidol (as monotherapy or in combination), risperidone, and
ziprasidone, for which previous primary data has provided support (Baldacara et
al. 2011). Data analyzing different agents than asenapine may be less easy to
interpret as they mainly derive from samples with mixed diagnoses. Nevertheless,
asenapine and risperidone offer oral treatment options, which may be preferred over
intramuscular routes for milder forms of agitation (Garriga et al. 2016).
Maintenance Treatment
al. 2018). Nevertheless, lithium may be less efficacious in patients with a history of
substance abuse (Maj 2003), probably because these patients are more likely to
relapse into depression (Kemp et al. 2009). Combinations of agents may be consid-
ered for patients that did not benefit from monotherapy (Table 3).
When choosing across the various psychotropic agents, clinicians also need to
take into account the patient’s risk for depressive or manic episodes. For instance,
lamotrigine and quetiapine may be more efficacious in preventing depressive epi-
sodes on the one hand (Popovic et al. 2012; Grunze et al. 2013). On the other hand,
risperidone LAI, aripiprazole, lithium, and ziprasidone are more successful in
preventing manic episodes (Yatham et al. 2018; Li et al. 2017). Data for olanzapine
is rather equivocal (Pan et al. 2014; Popovic et al. 2012). Apart from the number (or
severity) of new episodes, time to recurrence of mood episodes presents an addi-
tional valuable measure for clinicians when assessing efficacy of mood stabilizers in
maintenance treatment. For instance, time to recurrence in olanzapine-medicated
patients was longer than in risperidone LAI-medicated patients, while it was longer
for quetiapine compared to lithium (Grunze et al. 2013). Although more comparative
trials have been conducted, differences between mood stabilizers have been fre-
quently nonsignificant. Perhaps, this is due to pooling of depressive and manic
episodes when investigating time to recurrence for mood stabilizers.
In the literature, there is also some support for a few off-label treatments. Various
international guidelines encompass valproate across first-line pharmacotherapies
(Yatham et al. 2018; Goodwin et al. 2016). The marketer company has pursued
approval of valproate in the European Union, but has been rejected due to lack of
placebo-controlled studies (EMA 2011). Moreover, there may be some place for
other LAI agents apart from risperidone and aripiprazole; in a large naturalistic trial
long-acting injections were associated with lower risk of psychiatric hospitalizations
(LÄhteenvuo et al. 2018). The effects of perphenazine and haloperidol LAI essen-
tially contributed to these findings. The evidence for LAI agents indirectly highlights
the role of compliance in recurrence of mood episodes. Further, there is a single trial
reporting prophylactic properties for manic episodes for oral paliperidone
(Berwaerts et al. 2012), while a naturalistic small cohort study reported encouraging
1482 G. Schoretsanitis and M. Paulzen
results for patients stabilized and continued on the middle-term with lurasidone
(Schaffer et al. 2016). Last, the BAP guidelines describe carbamazepine across
long-term treatments particularly for patients not responding to lithium (Goodwin
et al. 2016).
The prescription of mood stabilizers for long periods confers lifelong vulnera-
bility for adverse mood stabilizer-induced drug reactions. There is a variety of
ADRs depending on the individual safety profile of the prescribed agent, although
combinations of mood stabilizers are in general more burdensome (Galling et al.
2015). Chronic treatment with lithium poses a potential risk of acute or chronic
nephrotoxicity, so that regular monitor schedules are necessary, although data
about potentially toxic effects of lithium on renal function remain inconsistent.
For long-term lithium treatment, recommended plasma levels range between 0.6
and 0.8 mmol/l (Malhi et al. 2016). Many of the mood stabilizers and in particular
antipsychotics are linked to metabolic abnormalities in a dose- and treatment
duration-independent way (Bak et al. 2014). For patients at high risk for a
metabolic syndrome, agents such as aripiprazole and ziprasidone may be treatment
of choice in using antipsychotics, whereas olanzapine and quetiapine should only
be considered, when all other options are excluded. As women may be at higher
risk for the hyperprolactinemia sequelae, they may be good candidates for partial
dopamine agonists, when AP treatment is considered; aripiprazole (and in Europe
off-label cariprazine) may be less prone in inducing hyperprolactinemia compared
to other AP. For the majority of the mood stabilizers, data for teratogenic risk is
barely available. However, there is an increasing consensus regarding the terato-
genic potential of valproate, so that it should be avoided in women of reproductive
age or pregnant patients (Patel et al. 2018). Valproate may be also not the first-line
treatment for patients with hepatic deficits. In this subgroup of patients, long-acting
agents may be indicated as they display a limited first-pass metabolism
(Schoretsanitis et al. 2018a). Another particularly vulnerable patient subgroup
includes elderly patients. There is a dearth of evidence for the safety and efficacy
issues in these individuals. However, post hoc data support off-label prescription
of lurasidone in elderly patients, while open-label data support asenapine and
aripiprazole as well (Vasudev et al. 2018).
Withdrawal Syndromes
Anti-convulsants
Lithium
Lithium is the prototypic substance for which bipolar rebound phenomena have been
described and investigated during the past decades. Research has addressed short-
and long-term effects of the discontinuation of lithium monotherapy. Researchers
have also acknowledged recurrence due to the natural course of bipolar disorders,
which poses difficulties in disentangling the inherent relapse risk of bipolar disorders
from lithium discontinuation effects.
Lithium discontinuation may at first support the remission of lithium-induced
adverse reactions (Christodoulou and Lykouras 1982), but has also been linked to
early recurrence of bipolar episodes with an enhanced risk within the first weeks
after discontinuation (Verdoux and Bourgeois 1993). In fact, the frequency of mood
episodes in patients with discontinued treatment was comparable to the frequency in
unmedicated patients. Abrupt versus gradual discontinuation of lithium may
enhance the risk of early recurrence (Suppes et al. 1993). There have also been
reports of psychotic states following lithium withdrawal (Klein et al. 1981). Patients
relapsing during discontinuation are less responsive to lithium in case of a
reintroduction of lithium (Cakir et al. 2017). Actually, this is an old notion postu-
lating that patients who discontinued lithium more than once develop a lithium-
refractoriness. However, it is uneasy to clarify if this effect is tolerance effect
1484 G. Schoretsanitis and M. Paulzen
Antipsychotics
Resistance to Therapy
Resistance to therapy with mood stabilizers reflects a very serious problem with a
high burden given its strikingly high prevalence (Geddes and Miklowitz 2013).
When mood stabilizers do not work, clinicians may extrapolate their clinical
Mood Stabilizers: Course and Duration of Therapy, Withdrawal Syndromes, and. . . 1485
experience from refractory major depression; this strategy may be of success, but
also hides considerable traps. Before activating their knowledge for treatment-
resistant major depression, prescribers need to exhaust evidence-based options for
bipolar disorders (Goodwin et al. 2016). In fact, adherence has to be assessed,
because it is a common problem in these patients (LECLERC et al. 2013). In
treatment-refractory bipolar depression, clinicians may consider lamotrigine,
which might have positive effects on depressive symptoms (Nierenberg et al.
2006). In principle, combinations of agents are more likely to be effective than
monotherapies (Galling et al. 2015). Therefore, for patients not responding to a
mood stabilizer treatment, augmenting with a second mood stabilizer offers a
reasonable second step. Nevertheless, the decision-making algorithm is restricted
to an offering of additionally reliable options for patients not responding to a
combination of mood stabilizers.
There are two treatment options that have consistently received attention by the
majority of established guidelines, whereas there may be some space for novel
therapies as well. First, modern guidelines place electroconvulsive therapy (ECT)
across options when patients do not respond to first- or second-line therapies
(Fountoulakis et al. 2017; Goodwin et al. 2016; Yatham et al. 2018). However, a
randomized clinical trial for ECT in treatment-refractory was not available till
recently; Schoeyen and colleagues reported larger reduction in depressive symptoms
and higher response rates for patients receiving ECT for 6 weeks compared to
patients allocated to algorithm-based pharmacological treatment (Schoeyen et al.
2015). Older data had previously implied that continuation as well as maintenance
ECT may accelerate treatment response (Sikdar et al. 1994). Second, based on the
superiority of clozapine over every other pharmacotherapy in schizophrenia,
research has expanded on its application in affective disorders with encouraging
results. Nevertheless, a controlled trial for treatment-refractory bipolar disorder is yet
to be conducted. Guidelines consider clozapine as monotherapy or adjunctive for
patients not responding to mood stabilizers (Yatham et al. 2018; Goodwin et al.
2016; Grunze et al. 2009). When prescribed in patients with bipolar disorders,
clozapine may need to be titrated slower than in other patients (Calabrese et al.
1996). Clinical improvement is likely to be observed within 2 months from cloza-
pine initiation (Green et al. 2000), and the presence of psychotic features did not
predict treatment outcome (Suppes et al. 1999).
Literature provides some evidence for additional options in treatment-refractory
bipolar disorders. These encompass pharmacological and non-pharmacological inter-
ventions. Data for agents such as pramipexole, memantine, and pregabalin show
efficacy in small samples, but need further support by studies with larger groups of
patients (Goodwin et al. 2016; Hui Poon et al. 2015). Antidepressants also introduce a
plausible approach, specifically for treatment-resistant bipolar depressive episodes.
Furthermore, the enormous excitement for ketamine has pushed researchers and
clinicians to use it for treatment-resistant bipolar disorder. Meta-analytical data report
drastic short-term reduction of depressive symptoms and suicidal ideation after a
single-dose of intravenous ketamine (Parsaik et al. 2015); however, as in unipolar
depression, these effects are limited through their short duration. Lastly, there may be a
1486 G. Schoretsanitis and M. Paulzen
role for thyroid replacement therapies for these patients. A recent trial investigating
patients with rapid cycling bipolar disorder reinvigorated the interest for the link
between bipolar disorders and thyroid economy (Walshaw et al. 2018).
Research has also addressed the effects of non-pharmacological treatments such
as stimulation-modulation therapies. Repetitive transcranial magnetic stimulation
(TMS) for depressive episodes has received considerable interest, although data
often derive from mixed samples (unipolar and bipolar depression). Trials have
failed to highlight its role as add-on to pharmacotherapy (Hu et al. 2016; Fitzgerald
et al. 2016). However, a particular modality of TMS, known as deep TMS, applying
a different type of coil was recently shown to be efficacious for treatment-resistant
bipolar depression with treatment response being twofold higher than in placebo
(Tavares et al. 2017). Data for vagal nerve stimulation has not been encouraging.
Future research activity needs to address many questions regarding standardization
of these processes, before they enter clinical routine.
References
Arora M, Praharaj SK. Olanzapine discontinuation emergent recurrence in bipolar disorder. Indian J
Psychol Med. 2014;36:170–3.
Bak M, Fransen A, Janssen J, Van Os J, Drukker M. Almost all antipsychotics result in weight gain:
a meta-analysis. PLoS One. 2014;9:e94112.
Baldacara L, Sanches M, Cordeiro DC, Jackoswski AP. Rapid tranquilization for agitated patients in
emergency psychiatric rooms: a randomized trial of olanzapine, ziprasidone, haloperidol plus
promethazine, haloperidol plus midazolam and haloperidol alone. Rev Bras Psiquiatr.
2011;33:30–9.
Berwaerts J, Melkote R, Nuamah I, Lim P. A randomized, placebo- and active-controlled study of
paliperidone extended-release as maintenance treatment in patients with bipolar I disorder after
an acute manic or mixed episode. J Affect Disord. 2012;138:247–58.
Bowden CL, Brugger AM, Swann AC, Calabrese JR, Janicak PG, Petty F, Dilsaver SC, Davis JM,
Rush AJ, Small JG, Al E. Efficacy of divalproex vs lithium and placebo in the treatment of
mania. The Depakote Mania Study Group. JAMA. 1994;271:918–24.
Buoli M, Esposito CM, Godio M, Caldiroli A, Serati M, Altamura AC. Have antipsychotics a
different speed of action in the acute treatment of mania? A single-blind comparative study.
J Psychopharmacol. 2017;31:1537–43.
Bustillo JR, Lauriello J, Parker K, Hammond R, Rowland L, Bogenschutz M, Keith S. Treatment of
weight gain with fluoxetine in olanzapine-treated schizophrenic outpatients. Neuropsychophar-
macology. 2003;28:527–9.
Cade JF. Lithium salts in the treatment of psychotic excitement. Med J Aust. 1949;2:349–52.
Cakir S, Yazici O, Post RM. Decreased responsiveness following lithium discontinuation in bipolar
disorder: a naturalistic observation study. Psychiatry Res. 2017;247:305–9.
Calabrese JR, Kimmel SE, Woyshville MJ, Rapport DJ, Faust CJ, Thompson PA, Meltzer HY.
Clozapine for treatment-refractory mania. Am J Psychiatry. 1996;153:759–64.
Calabrese JR, Sanchez R, Jin N, Amatniek J, Cox K, Johnson B, Perry P, Hertel P, Such P, Salzman
PM, Mcquade RD, Nyilas M, Carson WH. Efficacy and safety of Aripiprazole once-monthly in
the maintenance treatment of bipolar I disorder: a double-blind, placebo-controlled, 52-week
randomized withdrawal study. J Clin Psychiatry. 2017;78:324–31.
Chateauvieux S, Morceau F, Dicato M, Diederich M. Molecular and therapeutic potential and
toxicity of valproic acid. J Biomed Biotechnol. 2010;2010. https://doi.org/10.1155/2010/
479364.
Mood Stabilizers: Course and Duration of Therapy, Withdrawal Syndromes, and. . . 1487
Garriga M, Pacchiarotti I, Kasper S, Zeller SL, Allen MH, Vazquez G, Baldacara L, San L,
McAllister-Williams RH, Fountoulakis KN, Courtet P, Naber D, Chan EW, Fagiolini A, Moller
HJ, Grunze H, Llorca PM, Jaffe RL, Yatham LN, Hidalgo-Mazzei D, Passamar M, Messer T,
Bernardo M, Vieta E. Assessment and management of agitation in psychiatry: expert consensus.
World J Biol Psychiatry. 2016;17:86–128.
Geddes JR, Miklowitz DJ. Treatment of bipolar disorder. Lancet. 2013;381:1672–82.
Glazer WM. Does loxapine have “atypical” properties? Clinical evidence. J Clin Psychiatry.
1999;60(Suppl 10):42–6.
Goikolea JM, Colom F, Capapey J, Torres I, Valenti M, Grande I, Undurraga J, Vieta E. Faster onset of
antimanic action with haloperidol compared to second-generation antipsychotics. A meta-analysis
of randomized clinical trials in acute mania. Eur Neuropsychopharmacol. 2013;23:305–16.
Gonzalez D, Bienroth M, Curtis V, Debenham M, Jones S, Pitsi D, George M. Consensus statement
on the use of intramuscular aripiprazole for the rapid control of agitation in bipolar mania and
schizophrenia. Curr Med Res Opin. 2013;29:241–50.
Goodwin GM, Haddad PM, Ferrier IN, Aronson JK, Barnes T, Cipriani A, Coghill DR, Fazel S,
Geddes JR, Grunze H, Holmes EA, Howes O, Hudson S, Hunt N, Jones I, Macmillan IC,
McAllister-Williams H, Miklowitz DR, Morriss R, Munafo M, Paton C, Saharkian BJ, Saunders
K, Sinclair J, Taylor D, Vieta E, Young AH. Evidence-based guidelines for treating bipolar
disorder: revised third edition recommendations from the British Association for Psychophar-
macology. J Psychopharmacol. 2016;30:495–553.
Green AI, Tohen M, Patel JK, Banov M, Durand C, Berman I, Chang H, Zarate C Jr, Posener J, Lee
H, Dawson R, Richards C, Cole JO, Schatzberg AF. Clozapine in the treatment of refractory
psychotic mania. Am J Psychiatry. 2000;157:982–6.
GrÜnder G, Carlsson A, Wong DF. Mechanism of new antipsychotic medications: occupancy is not
just antagonism. Arch Gen Psychiatry. 2003;60:974–7.
Grunze H, Vieta E, Goodwin GM, Bowden C, Licht RW, Azorin JM, Yatham L, Mosolov S, Moller
HJ, Kasper S, Members of the WFSBP Task Force on Bipolar Affective Disorders Working on
this topic. The World Federation of Societies of Biological Psychiatry (WFSBP) Guidelines for
the Biological Treatment of Bipolar Disorders: acute and long-term treatment of mixed states in
bipolar disorder. World J Biol Psychiatry. 2018;19:2–58.
Grunze H, Vieta E, Goodwin GM, Bowden C, Licht RW, Moller HJ, Kasper S. The World
Federation of Societies of Biological Psychiatry (WFSBP) guidelines for the biological treat-
ment of bipolar disorders: update 2009 on the treatment of acute mania. World J Biol Psychiatry.
2009;10:85–116.
Grunze H, Vieta E, Goodwin GM, Bowden C, Licht RW, Moller HJ, Kasper S, Disorders
WTFOTGFB. The World Federation of Societies of Biological Psychiatry (WFSBP) guidelines
for the biological treatment of bipolar disorders: update 2012 on the long-term treatment of
bipolar disorder. World J Biol Psychiatry. 2013;14:154–219.
Harwood AJ, Agam G. Search for a common mechanism of mood stabilizers. Biochem Pharmacol.
2003;66:179–89.
Hata M, Tanaka Y, Kyoda N, Osakabe T, Yuki H, Ishii I, Kitada M, Neya S, Hoshino T. An
epoxidation mechanism of carbamazepine by CYP3A4. Bioorg Med Chem. 2008;16:5134–48.
Hayes J, Prah P, Nazareth I, King M, Walters K, Petersen I, Osborn D. Prescribing trends in bipolar
disorder: cohort study in the United Kingdom THIN primary care database 1995-2009. PLoS
One. 2011;6:e28725.
Hendrick V, Altshuler LL, Szuba MP. Is there a role for neuroleptics in bipolar depression? J Clin
Psychiatry. 1994;55:533–5.
Hiemke C, Bergemann N, Clement HW, Conca A, Deckert J, Domschke K, Eckermann G, Egberts
K, Gerlach M, Greiner C, GrÜnder G, Haen E, Havemann-Reinecke U, Hefner G, Helmer R,
Janssen G, Jaquenoud E, Laux G, Messer T, Mossner R, Muller MJ, Paulzen M, Pfuhlmann B,
Riederer P, Saria A, Schoppek B, Schoretsanitis G, Schwarz M, Gracia MS, Stegmann B,
Steimer W, Stingl JC, Uhr M, Ulrich S, Unterecker S, Waschgler R, Zernig G, Zurek G,
Baumann P. Consensus guidelines for therapeutic drug monitoring in neuropsychophar-
macology: update 2017. Pharmacopsychiatry. 2018;51:9–62.
Mood Stabilizers: Course and Duration of Therapy, Withdrawal Syndromes, and. . . 1489
Hu SH, Lai JB, Xu DR, Qi HL, Peterson BS, Bao AM, Hu CC, Huang ML, Chen JK, Wei N, Hu JB,
Li SL, Zhou WH, Xu WJ, Xu Y. Efficacy of repetitive transcranial magnetic stimulation with
quetiapine in treating bipolar II depression: a randomized, double-blinded, control study. Sci
Rep. 2016;6:30537.
Hui Poon S, Sim K, Baldessarini RJ. Pharmacological approaches for treatment-resistant bipolar
disorder. Curr Neuropharmacol. 2015;13:592–604.
Kemp DE, Gao K, Ganocy SJ, Elhaj O, Bilali SR, Conroy C, Findling RL, Calabrese JR. A 6-
month, double-blind, maintenance trial of lithium monotherapy versus the combination of
lithium and divalproex for rapid-cycling bipolar disorder and co-occurring substance abuse or
dependence. J Clin Psychiatry. 2009;70:113–21.
Kessing LV, Vradi E, Andersen PK. Nationwide and population-based prescription patterns in
bipolar disorder. Bipolar Disord. 2016;18:174–82.
Klein HE, Broucek B, Greil W. Lithium withdrawal triggers psychotic states. Br J Psychiatry.
1981;139:255–6.
LÄhteenvuo M, Tanskanen A, Taipale H, Hoti F, Vattulainen P, Vieta E, Tiihonen J. Real-world
effectiveness of pharmacologic treatments for the prevention of rehospitalization in a Finnish
nationwide cohort of patients with bipolar disorder. JAMA Psychiatry. 2018;75:347–55.
LECLERC E, Mansur RB, Brietzke E. Determinants of adherence to treatment in bipolar disorder: a
comprehensive review. J Affect Disord. 2013;149:247–52.
Li DJ, Tseng PT, Stubbs B, Chu CS, Chang HY, Vieta E, Fornaro M, Carvalho AF, Solmi M,
Veronese N, Chen TY, Chen YW, Lin PY, Chow PC. Efficacy, safety and tolerability of
aripiprazole in bipolar disorder: an updated systematic review and meta-analysis of randomized
controlled trials. Prog Neuro-Psychopharmacol Biol Psychiatry. 2017;79:289–301.
Lieberman JA, Tasman A. Handbook of psychiatric drugs. West Sussex: Wiley; 2006.
Machado-Vieira R, Luckenbaugh DA, Soeiro-De-Souza MG, Marca G, Henter ID, Busnello JV,
Gattaz WF, Zarate CA Jr. Early improvement with lithium in classic mania and its association
with later response. J Affect Disord. 2013;144:160–4.
Maj M. The effect of lithium in bipolar disorder: a review of recent research evidence. Bipolar
Disord. 2003;5:180–8.
Malhi GS, Gershon S, Outhred T. Lithiumeter: version 2.0. Bipolar Disord. 2016;18:631–41.
Malhi GS, Porter R, Irwin L, Hamilton A, Morris G, Bassett D, Baune BT, Boyce P, Hopwood MJ,
Mulder R, Parker G, Mannie Z, Outhred T, Das P, Singh AB. Defining a mood stabiliser: novel
framework for research and clinical practice. BJPsych Open. 2018;4:278–81.
Mauri MC, Paletta S, Di Pace C, Reggiori A, Cirnigliaro G, Valli I, Altamura AC. Clinical
pharmacokinetics of atypical antipsychotics: an update. Clin Pharmacokinet. 2018;58(9):1217–8.
Mcgirr A, Vohringer PA, Ghaemi SN, Lam RW, Yatham LN. Safety and efficacy of adjunctive
second-generation antidepressant therapy with a mood stabiliser or an atypical antipsychotic in
acute bipolar depression: a systematic review and meta-analysis of randomised placebo-con-
trolled trials. Lancet Psychiatry. 2016;3:1138–46.
Meehan K, Zhang F, David S, Tohen M, Janicak P, Small J, Koch M, Rizk R, Walker D, Tran P,
Breier A. A double-blind, randomized comparison of the efficacy and safety of intramuscular
injections of olanzapine, lorazepam, or placebo in treating acutely agitated patients diagnosed
with bipolar mania. J Clin Psychopharmacol. 2001;21:389–97.
Murai T, Nakako T, Ikeda K, Ikejiri M, Ishiyama T, Taiji M. Lack of dopamine D4 receptor affinity
contributes to the procognitive effect of lurasidone. Behav Brain Res. 2014;261:26–30.
Newport DJ, Stowe ZN, Viguera AC, Calamaras MR, Juric S, Knight B, Pennell PB, Baldessarini
RJ. Lamotrigine in bipolar disorder: efficacy during pregnancy. Bipolar Disord. 2008;10:432–6.
Nierenberg AA, Ostacher MJ, Calabrese JR, Ketter TA, Marangell LB, Miklowitz DJ, Miyahara S,
Bauer MS, Thase ME, Wisniewski SR, Sachs GS. Treatment-resistant bipolar depression: a
STEP-BD equipoise randomized effectiveness trial of antidepressant augmentation with
lamotrigine, inositol, or risperidone. Am J Psychiatry. 2006;163:210–6.
Ogawa Y, Tajika A, Takeshima N, Hayasaka Y, Furukawa TA. Mood stabilizers and antipsychotics
for acute mania: a systematic review and meta-analysis of combination/augmentation therapy
versus monotherapy. CNS Drugs. 2014;28:989–1003.
1490 G. Schoretsanitis and M. Paulzen
Pan PY, Lee MS, Lo MC, Yang EL, Yeh CB. Olanzapine is superior to lamotrigine in the prevention
of bipolar depression: a naturalistic observational study. BMC Psychiatry. 2014;14:145.
Parsaik AK, Singh B, Khosh-Chashm D, Mascarenhas SS. Efficacy of ketamine in bipolar
depression: systematic review and meta-analysis. J Psychiatr Pract. 2015;21:427–35.
Patel N, Viguera AC, Baldessarini RJ. Mood-stabilizing anticonvulsants, spina bifida, and folate
supplementation: commentary. J Clin Psychopharmacol. 2018;38:7–10.
Paulzen M, Stingl J, Augustin M, Saßmannshausen H, Franz C, GrÜnder G, Schoretsanitis G.
Comprehensive measurements of intrauterine and postnatal exposure to lamotrigine. Clin
Pharmacokinet. 2018;58(4):535–43.
Petersen I, McCrea RL, Sammon CJ, Osborn DP, Evans SJ, Cowen PJ, Freemantle N, Nazareth I.
Risks and benefits of psychotropic medication in pregnancy: cohort studies based on UK
electronic primary care health records. Health Technol Assess. 2016;20:1–176.
Popovic D, Reinares M, Goikolea JM, Bonnin CM, Gonzalez-Pinto A, Vieta E. Polarity index of
pharmacological agents used for maintenance treatment of bipolar disorder. Eur Neuropsycho-
pharmacol. 2012;22:339–46.
Rajaratnam K, Xiang YT, Tripathi A, Chiu HF, Si TM, Chee KY, Avasthi A, Grover S, Chong MY,
Kuga H, Kanba S, He YL, Lee MS, Yang SY, Udomratn P, Kallivayalil RA, Tanra AJ, Maramis
MM, Shen WW, Sartorius N, Kua EH, Tan CH, Mahendran R, Shinfuku N, Sum MY,
Baldessarini RJ, Sim K. Clinical use of mood stabilizers with antidepressants in Asia: report
from the research on Asian psychotropic prescription patterns for antidepressants (REAP-AD)
projects in 2004 and 2013. J Clin Psychopharmacol. 2017;37:255–9.
Sajatovic M, Forester BP, Tsai J, Kroger H, Pikalov A, Cucchiaro J, Loebel A. Efficacy of
Lurasidone in adults aged 55 years and older with bipolar depression: post hoc analysis of 2
double-blind, placebo-controlled studies. J Clin Psychiatry. 2016;77:e1324–31.
Saksa JR, Baker CB, Woods SW. Mood-stabilizer-maintained, remitted bipolar patients: taper and
discontinuation of adjunctive antipsychotic medication. Gen Hosp Psychiatry. 2004;26:233–6.
Salvadore G, Quiroz JA, Machado-Vieira R, Henter ID, Manji HK, Zarate CA Jr. The neurobiology
of the switch process in bipolar disorder: a review. J Clin Psychiatry. 2010;71:1488–501.
San L, Estrada G, Oudovenko N, Montanes F, Dobrovolskaya N, Bukhanovskaya O, Popov M,
Vieta E. Placid study: a randomized trial comparing the efficacy and safety of inhaled loxapine
versus intramuscular aripiprazole in acutely agitated patients with schizophrenia or bipolar
disorder. Eur Neuropsychopharmacol. 2018;28:710–8.
Sansone RA, Sawyer RJ. Aripiprazole withdrawal: a case report. Innov Clin Neurosci.
2013;10:10–2.
Schaffer CB, Schaffer LC, Nordahl TE, Stark NM, Gohring CE. An open trial of Lurasidone as an
acute and maintenance adjunctive treatment for outpatients with treatment-resistant bipolar
disorder. J Clin Psychopharmacol. 2016;36:88–9.
Schoeyen HK, Kessler U, Andreassen OA, Auestad BH, Bergsholm P, Malt UF, Morken G,
Oedegaard KJ, Vaaler A. Treatment-resistant bipolar depression: a randomized controlled trial
of electroconvulsive therapy versus algorithm-based pharmacological treatment. Am J Psychi-
atry. 2015;172:41–51.
Schoretsanitis G, De Leon J, Haen E, Stegmann B, Hiemke C, Grunder G, Paulzen M. Pharmaco-
kinetics of risperidone in different application forms – comparing long-acting injectable and oral
formulations. Eur Neuropsychopharmacol. 2018a;28:130–7.
Schoretsanitis G, Haen E, Grunder G, Stegmann B, Schruers KR, Hiemke C, Lammertz SE, Paulzen
M. Pharmacokinetic drug-drug interactions of mood stabilizers and Risperidone in patients
under combined treatment. J Clin Psychopharmacol. 2016;36:554–61.
Schoretsanitis G, Paulzen M, Unterecker S, Schwarz M, Conca A, Zernig G, Grunder G, Haen E,
Baumann P, Bergemann N, Clement HW, Domschke K, Eckermann G, Egberts K, Gerlach M,
Greiner C, Havemann-Reinecke U, Hefner G, Helmer R, Janssen G, Jaquenoud-Sirot E, Laux
G, Messer T, Mossner R, Muller MJ, Pfuhlmann B, Riederer P, Saria A, Schoppek B, Silva
Gracia M, Stegmann B, Steimer W, Stingl JC, Uhr M, Ulrich S, Waschgler R, Zurek G, Hiemke
C. Tdm in psychiatry and neurology: a comprehensive summary of the consensus guidelines for
therapeutic drug monitoring in neuropsychopharmacology, update 2017; a tool for clinicians.
World J Biol Psychiatry. 2018b;19:162–74.
Mood Stabilizers: Course and Duration of Therapy, Withdrawal Syndromes, and. . . 1491
Schwartz TL, Massa JL, Gupta S, Al-Samarrai S, Devitt P, Masand PS. Divalproex sodium versus
valproic acid in hospital treatment of psychotic disorders. Prim Care Companion J Clin
Psychiatry. 2000;2:45–8.
Sharma PS, Kongasseri S, Praharaj SK. Outcome of mood stabilizer discontinuation in bipolar
disorder after 5 years of euthymia. J Clin Psychopharmacol. 2014;34:504–7.
Sikdar S, Kulhara P, Avasthi A, Singh H. Combined chlorpromazine and electroconvulsive therapy
in mania. Br J Psychiatry. 1994;164:806–10.
Silva MT, Zimmermann IR, Galvao TF, Pereira MG. Olanzapine plus fluoxetine for bipolar
disorder: a systematic review and meta-analysis. J Affect Disord. 2013;146:310–8.
Suppes T, Baldessarini RJ, Faedda GL, Tondo L, Tohen M. Discontinuation of maintenance
treatment in bipolar disorder: risks and implications. Harv Rev Psychiatry. 1993;1:131–44.
Suppes T, Webb A, Paul B, Carmody T, Kraemer H, Rush AJ. Clinical outcome in a randomized 1-
year trial of clozapine versus treatment as usual for patients with treatment-resistant illness and a
history of mania. Am J Psychiatry. 1999;156:1164–9.
Swann AC, Bowden CL, Morris D, Calabrese JR, Petty F, Small J, Dilsaver SC, Davis JM.
Depression during mania. Treatment response to lithium or divalproex. Arch Gen Psychiatry.
1997;54:37–42.
Tavares DF, Myczkowski ML, Alberto RL, Valiengo L, Rios RM, Gordon P, De Sampaio-Junior B,
Klein I, Mansur CG, Marcolin MA, Lafer B, Moreno RA, Gattaz W, Daskalakis ZJ, Brunoni
AR. Treatment of bipolar depression with deep TMS: results from a double-blind, randomized,
parallel group, sham-controlled clinical trial. Neuropsychopharmacology. 2017;42:2593–601.
Vasudev A, Chaudhari S, Sethi R, Fu R, Sandieson RM, Forester BP. A review of the pharmaco-
logical and clinical profile of newer atypical antipsychotics as treatments for bipolar disorder:
considerations for use in older patients. Drugs Aging. 2018;35(10):887–95.
Verdolini N, Hidalgo-Mazzei D, Murru A, Pacchiarotti I, Samalin L, Young AH, Vieta E, Carvalho
AF. Mixed states in bipolar and major depressive disorders: systematic review and quality
appraisal of guidelines. Acta Psychiatr Scand. 2018;138(3):196–222.
Verdoux H, Bourgeois M. Short-term sequelae of lithium discontinuation. Encéphale.
1993;19:645–50.
Viguera AC, Whitfield T, Baldessarini RJ, Newport DJ, Stowe Z, Reminick A, Zurick A, Cohen LS.
Risk of recurrence in women with bipolar disorder during pregnancy: prospective study of mood
stabilizer discontinuation. Am J Psychiatry. 2007;164:1817–24; quiz 1923.
Walshaw PD, Gyulai L, Bauer M, Bauer MS, Calimlim B, Sugar CA, Whybrow PC. Adjunctive
thyroid hormone treatment in rapid cycling bipolar disorder: a double-blind placebo-controlled
trial of levothyroxine (L-T4) and triiodothyronine (T3). Bipolar Disord. 2018;20(7):594–603.
Williams RS, Cheng L, Mudge AW, Harwood AJ. A common mechanism of action for three mood-
stabilizing drugs. Nature. 2002;417:292–5.
Yatham LN, Kennedy SH, Parikh SV, Schaffer A, Bond DJ, Frey BN, Sharma V, Goldstein BI, Rej
S, Beaulieu S, Alda M, Macqueen G, Milev RV, Ravindran A, O'donovan C, McIntosh D, Lam
RW, Vazquez G, Kapczinski F, McIntyre RS, Kozicky J, Kanba S, Lafer B, Suppes T, Calabrese
JR, Vieta E, Malhi G, Post RM, Berk M. Canadian Network for Mood and Anxiety Treatments
(CANMAT) and International Society for Bipolar Disorders (ISBD) 2018 guidelines for the
management of patients with bipolar disorder. Bipolar Disord. 2018;20:97–170.
Yoshii K, Kobayashi K, Tsumuji M, Tani M, Shimada N, Chiba K. Identification of human
cytochrome P450 isoforms involved in the 7-hydroxylation of chlorpromazine by human liver
microsomes. Life Sci. 2000;67:175–84.
Zarate CA Jr, Tohen M. Double-blind comparison of the continued use of antipsychotic treatment
versus its discontinuation in remitted manic patients. Am J Psychiatry. 2004;161:169–71.
Zhu B, Tunis SL, Zhao Z, Baker RW, Lage MJ, Shi L, Tohen M. Service utilization and costs of
olanzapine versus divalproex treatment for acute mania: results from a randomized, 47-week
clinical trial. Curr Med Res Opin. 2005;21:555–64.
Zohar J, Stahl S, Moller HJ, Blier P, Kupfer D, Yamawaki S, Uchida H, Spedding M, Goodwin GM,
Nutt D. A review of the current nomenclature for psychotropic agents and an introduction to the
neuroscience-based nomenclature. Eur Neuropsychopharmacol. 2015;25:2318–25.
Mood Stabilizers: Lithium
Janusz K. Rybakowski
Contents
Developmental History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1494
Chemical Formula: Graphics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1496
Physicochemical Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1496
Pharmacology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1496
Pharmacokinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1496
Mechanisms of Action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1497
Clinical Indications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1499
Mood Stabilization in Bipolar Disorder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1499
Comparison of Lithium with Other Mood-Stabilizing Drugs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1501
Excellent Lithium Responders . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1502
Genetic Studies of Lithium Response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1503
Lithium in the Treatment of Acute Episodes of Mood Disorders . . . . . . . . . . . . . . . . . . . . . . . . . 1504
Anti-suicidal Action in Mood Disorders . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1504
Antiviral and Immunomodulatory Action in Mood Disorders . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1505
Neuroprotective Action in Mood Disorders . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1506
Lithium in Dementia and Neurodegenerative Disorders . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1507
Side Effects/Adverse Reactions and Toxicology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1509
The Long-Term Effects of Lithium on Kidney, Thyroid, and Parathyroid Function . . . . . . 1510
Effect of Lithium on Neurocognitive Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1511
Lithium in Pregnancy and the Postpartum Period . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1513
Lithium Intoxication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1513
Lithium Interaction with Other Drugs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1514
Combination Therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1515
Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1516
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1516
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1516
J. K. Rybakowski (*)
Department of Adult Psychiatry, Poznan University of Medical Sciences, Poznan, Poland
e-mail: janusz.rybakowski@gmail.com
Abstract
The developmental history of lithium and its physicochemical properties, phar-
macokinetics, and mechanisms of action are presented. Among the latter, the
effect on the phosphatidylinositol (PI) system and the inhibition of the glycogen
synthase kinase 3beta (GSK-3β) activity are the most important. Lithium makes a
prototype of the mood-stabilizing drug, is regarded as the drug of the first choice
for the prophylaxis of bipolar disorders, and, as a monotherapy, surpasses all
other mood stabilizers. The drug can also be used in the treatment of acute
episodes of mood disorders, especially for the augmentation of antidepressants
in treatment-resistant depression. Lithium possesses significant anti-suicidal
properties, the strongest among all mood stabilizers. The drug exerts antiviral,
especially against herpes infection, and immunomodulatory influence. The evi-
dence has also been accumulated for the neurotrophic and neuroprotective effects
of lithium. Neuroimaging studies in bipolar subjects showed an increase in the
volume of some brain structures during lithium administration. Lithium therapy is
associated with numerous side effects, most of them can be successfully man-
aged. Given lithium superiority for the prophylaxis of mood disorders, together
with its anti-suicidal, immunomodulatory, and neuroprotective effects, the drug is
currently greatly underprescribed, and its therapeutic potential in patients with
bipolar disorder is not sufficiently utilized.
Recently, epidemiological data suggest an association between lithium intake
and dementia risk reduction. In experimental models of neurodegenerative dis-
orders, lithium exhibits significant therapeutic activity. Promising results have
also been obtained in some clinical studies.
Developmental History
Lithium history goes back to as early as the beginning of the world since lithium
appeared as one of the first elements after the Big Bang. This was mentioned by
recently deceased, the prominent English theoretical physicist and mathematician,
Stephen Hawking (1942–2018), in his book A Brief History of Time published
30 years ago (Hawking 1988). In the second century A.D., the outstanding Roman
physician, Soranus of Ephesus (98–138 A.D.), recommended that people with
nervous disorders should drink alkaline mineral waters with, as it transpired later,
a significant content of lithium ions (Gerdtz 1994).
The existence of lithium as a chemical element has been known for 200 years as it
was discovered in 1817 thanks to the research of a Swedish chemist, Johann August
Arfwedson (1792–1841), working in the Stockholm laboratory of famous chemistry
and pharmacy professor, Jons Jacob Berzelius (1779–1848). Arfwedson obtained
lithium carbonate from the mineral petalite that occurs on the Swedish island Utő
(Arfwedson 1818). In the middle of the nineteenth century, chemists found that the
salt lithium urate was among the most soluble urates. This prompted to introduce
lithium salts to treat gout and other rheumatic diseases in which an excess of
Mood Stabilizers: Lithium 1495
urate deposits was suspected. Such treatment was initiated by the English physician,
Alfred Baring Garrod (1819–1907) (Garrod 1859).
The first to use lithium for the treatment of acute mania was American physician,
William Alexander Hammond, in New York, and it was the bromide salt of lithium
(Hammond 1871). However, the real credit for introducing lithium for the treatment
of mood disorders should be given to the Danish doctor and scholar, Carl Lange
(1834–1900), also known as the co-author of the first neurobiological concept of the
emotions (the James-Lange theory). In 1886, Carl Lange published in Denmark his
treatise on the periodic depressive states “Om Periodiske Depressionstilstande og
deres Patogenese” (Lange 1886). This was produced 9 years later in the German
language as “Periodische Depressionzustände und ihre Pathogenesis auf dem Boden
der harnsauren Diathese” (On periodical depressions and their pathogenesis in the
context of uric acid abnormality) (Lange 1895). The important issue in this mono-
graph was a pathogenic concept of depression as a condition connected with an
excess of uric acid, something like a “gout of the brain.” This prompted Lange to
administer lithium as lithium carbonate to eliminate a possible uric acid excess by
forming the lithium urate in depressive patients. To this aim, he also used the dose
recommended by Garrod, i.e., around 25 mmol lithium (equalling about 1000 mg of
lithium carbonate) per day. According to Felber (1987), he might treat with lithium
as much as 2000 patients. While Carl Lange was giving lithium to ambulatory
patients in his private practice, this treatment was taken up by his brother, Fritz
Lange, the superintendent of Middelfart psychiatric hospital who administered it to
psychiatric inpatients.
Half a century after Carl Lange’s endeavors of lithium treatment, the Australian
psychiatrist John F. Cade (1912–1980) working in Melbourne researched on bio-
logical causes of mental disorders, using guinea pigs as experimental animals. Cade
found that the urine of patients in a manic state showed particular toxicity for guinea
pigs and reached the conclusion that such patients had probably an excess of urates.
He observed that after administering lithium urate, a significant decrease of toxic
symptoms and calming down of the animals occurred. However, a similar effect was
also obtained after giving lithium carbonate, indicating the primary effect of lithium
ions. Cade carried out a clinical self-experiment by taking himself lithium carbonate.
When it turned out that this did not lead to poisoning, he decided to give lithium
carbonate to ten patients with acute and chronic manic states. The results were
spectacular. Even though it was not possible to exclude spontaneous remission
among some patients with acute symptoms of mania, the significant improvements
observed in patients among whom manic symptoms had lasted for many months
were remarkable. Cade even hypothesized: “The effect on patients with pure psy-
chotic excitement – that is, true manic attacks – is so specific that it inevitably leads
to speculation as to the possible etiological significance of a deficiency in the body of
lithium ions in the genesis of this disorder” (Cade 1949).
Cade’s publication in the Medical Journal of Australia in 1949, in which he
described the antimanic effect of lithium, may be acknowledged as the evidence of
the introduction of lithium to modern psychiatric therapy. This publication could
also mark the beginning of modern clinical psychopharmacology because it
1496 J. K. Rybakowski
Physicochemical Properties
Lithium (from Greek word “lithos,” meaning “stone”) is a chemical element with
symbol Li and atomic number 3, the lightest metal, and the first member of the
family of alkali metals. Its atomic structure is simple: the nucleus comprising three or
four neutrons and three protons, and there are three electrons on two valence orbitals.
Naturally occurring lithium (3Li) is composed of two stable isotopes, lithium-6 and
lithium-7, where lithium-7 makes about 92.5% of the atoms (Fig. 1).
Lithium resembles sodium and potassium as the elements of the same group 1A
of the periodic table and is interacting with these ions in the mechanisms of
membrane transport. Moreover, lithium exhibits a similarity to magnesium, belong-
ing to the group IIA, to which it bears a diagonal relationship in the periodic table. It
has been supposed that many lithium biological effects may involve an interaction or
competition between lithium and magnesium ions.
Pharmacology
Pharmacokinetics
+ Protons (3)
N – Nucleons (7)
+N
– Neutrons (4)
N+ + e N
N – – Electrons (3)
e
Nuclide (Nucleus)
release preparations result in the peak about 12 h after ingestion. Lithium is distrib-
uted quickly throughout the body. The balance between blood (extracellular) and
tissue (intracellular) concentration is maintained by several transport mechanisms.
The concentrations of lithium in cells (e.g., erythrocytes) and in the brain are lower
than the plasma concentration. However, there is a significant correlation between
these parameters. Lithium is eliminated in 97% via the kidneys, and a negligible
amount is excreted in sweat and feces. The elimination half-life of lithium ranges
from 10 to 42 h (average 20 h) which means that lithium can be administered in one
daily dose (Alda 2006).
The lithium ion is filtered freely through the glomerular membrane and is in 80%
reabsorbed in the proximal tubule, proportionally to the reabsorption of sodium and,
generally, not reabsorbed in the distal nephron. Lithium clearance is usually
8–40 ml/min, on the average 20–30% of the glomerular filtration rate (GFR), and
can be calculated as the daily dosage of mmol lithium, divided by 1.44 and next by
the serum lithium concentration in mmol/l. Lithium concentration in serum after a
given dose is depending on lithium clearance. The changes in lithium clearance can
be due to the changes in the GFR, the variations of antinatriuretic and natriuretic
mechanisms, and the activation of lithium reabsorption in the distal nephron.
Lithium clearance is decreased in older age what makes a necessity to adjust the
dose of lithium, accordingly, in such patients. On the other hand, lithium clearance is
increased in pregnancy. Therefore, the dose of lithium can be increased during
pregnancy to achieve a desirable level and decreased after delivery when the
clearance declines. Most pharmacokinetic interactions of lithium are connected
with an effect on renal lithium clearance (Schou and Kampf 2006).
For prophylaxis of affective recurrences, lithium concentrations of 0.6–0.8 mmol/
l are recommended, and higher concentration can be attained for the treatment of
mania. However, the concentrations above 1.2 should be avoided as they may lead to
lithium intoxication. The measurement of lithium concentration in serum should be
performed 10–12 h after the last dose.
Mechanisms of Action
The mechanisms of lithium in mood disorders made the topic of several reviews
(Pasquali et al. 2010; Quiroz et al. 2010; Alda 2015; Malhi and Outhred 2016). The
authors concluded that among a variety of biological systems, the most important
mechanisms of lithium action are connected with intracellular signaling, especially
with the effect of lithium on the phosphatidylinositol (PI) system and with the
inhibition by lithium of the glycogen synthase kinase-3beta (GSK-3β). Also, a
neurotrophic and neuroprotective effect of lithium has been highlighted as an
important element of its therapeutic action in mood disorders but also in making
lithium a therapeutic candidate to be used in neurodegenerative disorders.
Lithium is acting on several stages of the phosphatidylinositol (PI) system, and
this activity has been thought of as important for its therapeutic action in mood
disorders. The inositol depletion hypothesis of bipolar mood disorder formulated
1498 J. K. Rybakowski
nearly 30 years ago by Berridge et al. (1989) proposes that lithium attenuates PI
signaling. Lithium inhibits inositol monophosphatase-1 which ameliorates inositol
depletion-related mitochondrial dysfunction. Changes in IP signaling measured as a
spread of growth cones were postulated to make a common effect of the first
generation mood stabilizers (lithium, valproates, and carbamazepine) (Williams et
al. 2004). Another part of intracellular signaling which has long been implicated in
the mechanism of lithium action is the system of adenylyl cyclases which converts
ATP into cyclic adenosine monophosphate (cAMP). Mann et al. (2009) suggest that
the inhibition of adenylyl cyclase isoforms by lithium and carbamazepine may be
related to their antidepressant effect. The important element of this system is also the
cAMP response element-binding protein (CREB), the regulator of gene expression.
The third component of the lithium mechanism connected with intracellular signal-
ing is inhibition by lithium of protein kinase C (PKC). New findings show that
lithium acts to inhibit PKC through a myristoylated alanine-rich C kinase substrate
(MARCKS) pathway. This mechanism of lithium is shared with another mood-
stabilizing drug, valproate, and was a basis to introduce PKC inhibitor, tamoxifen,
to the treatment of mania.
The evidence has been accumulating using various experimental models showing
that lithium inhibits glycogen synthase kinase-3beta (GSK-3β) activity. GSK-3β is a
serine/threonine kinase that regulates gene transcription, synaptic plasticity, apopto-
sis, cellular structure and resilience, and the circadian cycle, all of which are
implicated in the pathophysiology of mood disorders. Therefore, the GSK-3β inhi-
bition by lithium can make the main mechanism of therapeutic action in mood
disorders (Jope 2011). GSK-3β is also a key enzyme in the metabolism of amyloid
precursor protein and in the phosphorylation of the tau protein, playing the main
pathogenetic role in Alzheimer’s disease (AD). Due to the inhibition of GSK-3β,
lithium arrests the development of neurofibrillary tangles in mutant tau transgenic
mice with advanced neurofibrillary pathology (Leroy et al. 2010), and, in the
Drosophila fly, adult-onset model of the AD ameliorates amyloid-beta pathology
(Sofola et al. 2010).
Lithium exerts a stimulatory action on the brain-derived neurotrophic factor
(BDNF), the most important member of the neurotrophin family, necessary for the
survival and function of neurons. BDNF modulates the activity of such neurotrans-
mitters like glutamate, gamma-aminobutyric acid, dopamine and serotonin.
Transcription of the BDNF gene is activated by CREB. In experimental studies, it
was demonstrated that lithium activates CREB and increases BDNF expression. In
clinical studies, lithium treatment increases the blood level of BDNF.
Lithium treatment results in an increase of B-cell leukemia 2 (Bcl-2) in the brain
of experimental animals. Bcl-2 is an important protein for cellular resilience and
plasticity, exerting mostly anti-apoptotic effects. The expression of Bcl-2 is activated
by CREB. Lithium also increases the expression of Bcl-2-associated athanogene
(bag-1) which is known to attenuate glucocorticoid receptor nuclear translocation,
thus potentiating the anti-apoptotic effect.
Lithium is a monovalent cation, similar to sodium and potassium, and the first
theories of the mechanisms of lithium action were connected with electrolytes and
Mood Stabilizers: Lithium 1499
membrane transport. The early studies pointed to increased residual sodium during
acute episodes (depression or mania) and its normalization in the course of lithium
treatment. In the 1970s, the lithium-sodium countertransport system was discovered
as the main mechanism of lithium transport out of the cells. This system was
supposed to be deficient in bipolar disorder, reflected by an increased erythrocyte/
plasma lithium ratio in such patients. It was also found that lithium can enhance the
activity of the sodium-potassium adenosine triphosphatase (ATPase) which can be
decreased in mood disorders.
The interest in the pathogenetic role of altered neurotransmission in mood
disorder resulted in identifying meaningful effects of lithium on several neurotrans-
mitters. In experimental studies, it was demonstrated that lithium increases serotonin
transmission through multiple mechanisms and inhibits increased dopaminergic
activity. It was also found that lithium can reduce the glutamate-induced
excitotoxicity and also reverses and repair oxidative stress connected with
excitotoxicity.
Disturbances of circadian rhythm belong to the high-order modulatory biological
systems implicated in mood disorders. Many experimental and clinical studies
demonstrated that lithium could ameliorate and rectify circadian rhythm due to its
effect on the GSK-3β and PI system and by modulating the expression of certain
clock genes. In our molecular genetic study, we found an association between the
prophylactic effect of lithium and polymorphisms in the biological clock genes
(Rybakowski et al. 2014).
Recently, Pisanu et al. (2018) suggest that the epigenome might be also a target of
lithium and other mood stabilizers, as they can favorably interfere with the epige-
netic mechanisms such as DNA methylation and histone deacetylase activity.
Clinical Indications
Depressive Illness (Duke and Hochman 1992). However, the most influential about
treating the illness with lithium was the autobiography An Unquiet Mind, written by
a professional, Kay Redfield Jamison, the co-author (together with Frederick
Goodwin) of the fundamental textbook Manic-Depressive Illness (Jamison 1996).
Another backlash on lithium therapy emerged in the second part of 1990
containing allegations that there has been inadequate methodological evidence that
lithium has a beneficial effect and that it is ineffective in the long-term treatment of
bipolar disorder and may be associated with various form of harm, being “old but
flourishing blunder” (Moncrieff 1997). However, three meta-analyses of the twenty-
first century have amply confirmed the prophylactic effectiveness of lithium in
bipolar disorder. The first analysis performed by Geddes et al. (2004) including 5
randomized controlled trials with 770 patients showed that lithium was significantly
more effective than a placebo in preventing all affective relapses, being slightly
better against manic than against depressive recurrences. In the second study, Nivoli
et al. (2010) analyzed long-term controlled trials lasting at least half a year, with
1,561 patients, of whom 534 were receiving lithium. They noticed that earlier
research had suggested the nearly equal effectiveness of lithium against both
mania and depression; however, nowadays, lithium prophylaxis is regarded as
more effective against manic than against depressive relapses. The most recent
meta-analysis has been performed by Severus et al. (2014). Including seven trials
(1,580 patients) comparing lithium with placebo, the authors concluded that lithium
was significantly superior to placebo in preventing any mood episodes and manic
episodes. In some analyses, lithium was also better than a placebo in preventing
depressive episodes.
Nowadays, lithium is regarded as the drug of the first choice for long-term
prevention of mood recurrences in bipolar disorders. In a proportion of patients, it
is administered in combination with other mood stabilizers.
Lithium can be also of use for long-term prevention of depressive recurrences in
unipolar depression. A recent meta-analysis showed that lithium monotherapy is
significantly more effective for prevention of depressive recurrences than placebo
and is better, albeit nonsignificantly that antidepressant monotherapy and combina-
tion of lithium with an antidepressant is superior to an antidepressant alone in this
respect (Undurraga et al. 2019).
The studies comparing lithium with other first-generation mood stabilizers (carba-
mazepine and valproates) were the MAP (Multicenter study of long-term treatment
of Affective and schizoaffective Psychoses) and the BALANCE (Bipolar Affective
disorder Lithium/ANti-Convulsant Evaluation). In the MAP study, lasting 2.5 years,
lithium was better than carbamazepine in patients with bipolar I disorder, and both
drugs had similar efficacy in bipolar II illness. In subjects experiencing mood-
incongruent delusions and psychiatric comorbidity, a better effect of carbamazepine
was observed (Kleindienst and Greil 2000). In the BALANCE study, bipolar patients
1502 J. K. Rybakowski
The term “excellent lithium responders” (ER) was introduced by the Canadian
psychiatrist, Paul Grof (1999), for patients which on monotherapy with lithium
experienced a dramatic change in their life as their mood episodes were fully
prevented. We followed-up for 10 years 60 patients who started lithium prophylaxis
in the 1970s and 49 patients beginning this procedure in the 1980s. Those without
mood episodes during this period (ER) made 35% of the first group and 27% of the
second one, roughly one-third of bipolar subjects treated longitudinally with lithium
(Rybakowski et al. 2001). Grof (2010) suggests that lithium responders can be
characterized by distinct mood episodes, with full remissions between them, the
absence of other psychiatric morbidity, and frequent history of bipolar illness in their
families, reminding the aspects of the illness, defined by Emil Kraepelin (1899) as
“manisch-depressives Irresein.”
Mood Stabilizers: Lithium 1503
The quality of prophylactic lithium response makes a good topic for molecular
genetic studies. A review of genetic influences on the efficacy of lithium prophylaxis
was made by the author of this chapter (Rybakowski 2013). The genes studied for
their association with lithium response have been those connected with neurotrans-
mitters (serotonin, dopamine, and glutamate); second messengers, phosphatidy-
linositol (PI), cyclic adenosine monophosphate (cAMP) and protein kinase C
(PKC) pathways); and glycogen synthase kinase 3-β (GSK-3β), substances involved
in neuroprotection (brain-derived neurotrophic factor (BDNF)), in circadian
rhythms, and a number of other miscellaneous genes.
In 2009, an initiative of the National Institutes of Mental Health and the Interna-
tional Group for the Study of Lithium-Treated Patients (IGSLI) resulted in the
formation of the International Consortium on Lithium Genetics (ConLiGen), aiming
for the first genome-wide association study (GWAS) of lithium response (Schulze et
al. 2010). In 2013, the first report of the key phenotypic measures of the “Retro-
spective Criteria of Long-Term Treatment Response in Research Subjects with
Bipolar Disorder” scale, known as the Alda’s scale, used by ConLiGen was pre-
sented (Manchia et al. 2013). The first GWAS of lithium response was published in
2016, in which 2563 patients were included coming from 22 participating sites of the
ConLiGen. A single locus of four-linked single nucleotide polymorphisms (SNPs)
on chromosome 21 met genome-wide significance criteria for association with
1504 J. K. Rybakowski
lithium response. This region contains two genes for long, noncoding RNAs
(lncRNAs) which are important regulators of gene expression in the central nervous
system (Hou et al. 2016).
A recent study of the ConLiGen suggests some specificity of lithium for bipolar
disorder as patients with a higher polygenic score for schizophrenia had a worse
response to prophylactic lithium administration (International Consortium on
Lithium Genetics 2018).
The most important study of lithium in mania after Cade’s report (Cade 1949) was
that of Danish psychiatrist Mogens Schou (1918–2005). He and his colleagues
studied the effectiveness of lithium among 38 patients in a manic state. In the
study, for the first time, a placebo was used. A spectacular improvement was noted
in 12 patients, improvement in 15, and a lack of effect in 3 of them (Schou et al.
1954). Since then lithium has been widely and successfully used in the treatment of
mania both as monotherapy and combined with other drugs. A meta-analysis shows
a significant therapeutic effect of lithium in moderate and severe manic episodes
(Storosum et al. 2007). However, recent comparative analyses suggest superiority of
antipsychotic drugs over lithium in severe manic episodes (Cipriani et al. 2011).
Since the 1970s, the reports showing a possibility of the therapeutic effect of
lithium in a depressive episode have been published and in some American guidelines
of 1980–1990 lithium were recommended to treat a depressive episode in the course of
bipolar disorder. In 1981, Canadian psychiatrists showed that adding lithium to
antidepressant drugs among people whose therapy does not bring satisfying results
brings a substantial, and sometimes very quick, improvement of mood (De Montigny
et al. 1981). Since then evidence has been systematically gathered that lithium causes
potentiation of the effectiveness of antidepressant drugs. Rybakowski and Matkowski
(1992) showed that the augmenting effect of lithium is better in depression in the
course of bipolar disorder than in unipolar disorder. A review on this subject, in
particular, research comparing the effect of lithium with a placebo by the German
psychiatrists, demonstrates that lithium is an effective remedy augmenting the effect of
antidepressant drugs in treatment-resistant depression both bipolar and unipolar and a
successful effect may be expected in at least 50% of patients (Crossley and Bauer
2007; Bauer et al. 2014). In most current therapeutic standards, the strategy of
potentiation of antidepressant drugs by lithium is recommended in the first place.
The evidence that long-term lithium treatment can decrease mortality, primarily by
preventing suicide, has been accumulated since the early 1990s. Coppen et al. (1991)
observing 103 patients attending lithium clinic for 11 years concluded that lithium
Mood Stabilizers: Lithium 1505
reverses the excess mortality associated with mood disorders, mainly by preventing
suicide. The researchers assembled in the IGSLI group analyzing 827 patients with
bipolar and schizoaffective disorder given lithium treatment for more than 6 months
observed that the mortality of these patients did not differ significantly from that of
the general population (Müller-Oerlinghausen et al. 1992). Fifteen years later, a
meta-analysis from Harvard University group, including 45 papers containing data
on suicides committed while taking lithium (on average for 1.5 years) and 34 papers
registered suicides of persons not receiving lithium, showed that the risk of com-
mitting suicide was five times lower among patients taking lithium than those
subjected to other forms of treatment (Baldessarini et al. 2006). A recent meta-
analysis performed by Cipriani et al. (2013) with 6,674 patients concluded that
lithium was significantly better than a placebo in reducing the number of suicides
and deaths from any cause both in bipolar disorder and recurrent depression and
superior to other mood stabilizers or antidepressants.
Currently, the anti-suicidal effect of lithium in mood disorders is well
documented, and, in many comparative papers, it has been shown that it is signif-
icantly greater than other mood-stabilizing drugs. For example, a recent meta-
analysis did not show an effect of valproates on suicide risk in bipolar disorder
(Chen et al. 2019)
While being on lithium prevents suicide, its discontinuation significantly
increases this risk. The anti-suicidal effect is not significantly correlated with the
quality of prevention of mood recurrences by lithium, which points to the specificity
of such an effect exerted by lithium (Lewitzka et al. 2015)
In recent years, the anti-suicidal effect of lithium has been also demonstrated with
trace levels of this drug. In studies performed in Japan, Austria, and the USA, a
negative correlation between suicides and lithium concentrations in drinking water
was demonstrated (Ohgami et al. 2009; Kapusta et al. 2011; Blüml et al. 2013).
These results may reactivate Cade’s speculation he made when administered lithium
to manic patients, suggesting that lithium can be a trace element for mood disorders.
Researchers from the University of Birmingham showed that lithium in 5–30 mmol/l
concentration inhibits replication of the herpes simplex virus (Skinner et al. 1980).
The mechanism of action here probably involves blockage of synthesis of the virus’s
DNA by lithium or competition with magnesium ions catalyzing enzymatic reac-
tions of the virus. At that time also descriptions of cases of labial herpes remissions
while using lithium appeared. Labial herpes is caused by an infection of herpes
simplex virus type 1 (HSV-1) and occurs in approximately 1/3 of the population, and
its course is characterized by frequent recurrences.
Retrospective research of labial herpes occurrence in patients receiving lithium
for prophylactic purposes was carried out within a collaborative study of the
Department of Adult Psychiatry, Poznan University of Medical Sciences, and the
Department of Psychiatry of the University of Pennsylvania. In the Polish population
1506 J. K. Rybakowski
of 28 persons with recurrent labial herpes, during lithium therapy, the full cessation
of recurrence of herpes occurred in 13 patients, among 7 the frequency of recur-
rences decreased, among 6 it remained at the same level, and among 2 increased. The
general decrease in recurrence frequency was 64%. A better effect was observed in
patients in whom lithium concentration in the serum was higher than 0.65 mmol/l
and erythrocyte concentration exceeded 0.35 mmol/l. The American population
comprised two groups, 52 persons in each, matched with sex, age, and the duration
of systematic pharmacological treatment. In the first group, including mainly
patients with bipolar disorder treated with lithium, the frequency of labial herpes
recurrences in comparison with the 5-year period before the treatment decreased by
73%. In the second group, including patients with recurrent depression receiving
antidepressant drugs, no significant difference was observed (Rybakowski and
Amsterdam 1991).
Lithium can profoundly influence hematological and immunological system. An
increase of leukocytes by lithium treatment was observed nearly 70 years ago
(Radomski et al. 1950) and subsequently confirmed in many studies. Such a property
of lithium made some therapeutic uses also beyond psychiatry (e.g., in oncology). In
recent two decades, it was also found that lithium can mitigate the immune-endo-
crine component of the pathogenesis of the bipolar disorder, such as the acute-phase
reaction, production of pro-inflammatory cytokines, and excessive activation of the
hypothalamic-pituitary-adrenal axis (Rybakowski 2000). The anti-inflammatory
effect of lithium was the subject of a review 4 years ago (Nassar and Azab 2014).
Recently, we examined the impact of long-term lithium administration on very small
embryonic-like stem cells (VSELs) and the mRNA expression of pluripotency and
glial markers, in peripheral blood, showing that lithium can alleviate excessive
regenerative and inflammatory processes in bipolar disorder (Ferensztajn-
Rochowiak et al. 2018).
was reported regardless of mood state and diagnostic subtype. In the IGSLI study,
bipolar patients receiving lithium had larger hippocampal volumes compared to non-
lithium patients and similar to healthy controls (Hajek et al. 2014). Two studies
compared the effects of lithium with those of anti-convulsants and antipsychotics,
possessing mood-stabilizing properties. Lyoo et al. (2010) found that lithium caused
an increase in gray matter volume, peaking at weeks 10–12, and maintained through
16 weeks of treatment and that this increase was associated with positive clinical
response. By contrast, valproate-treated patients did not show grey matter volume
changes over time. Germana et al. (2010) found that the volume of gray matter in the
subgenual anterior cingulate gyrus on the right and the postcentral gyrus, the
hippocampus/amygdala complex, and the insula on the left was greater in patients
on lithium treatment compared to all other treatments. Thus, there is considerable
evidence for lithium causing an increase in cerebral grey matter volume both in
healthy subjects and in patients with bipolar disorders which may be associated with
its neuroprotective effect at a clinical level. Such a replicated and consistent evidence
for this has not been demonstrated for any other mood-stabilizing drugs.
The results of population studies reviewed by Donix and Bauer (2016) suggest an
association between lithium treatment and dementia risk reduction or reduced
dementia severity. In two papers from the Danish University of Copenhagen in
which the nationwide register of lithium prescriptions was used, it was found that in
patients who continued to take lithium, the rate of dementia decreased to the same
level as the rate for the general population. However, in persons treated with anti-
convulsant drugs, the risk of dementia increased with the duration of treatment
(Kessing et al. 2008). Continued treatment with lithium was also associated with a
reduced rate of dementia in patients with bipolar disorder, in contrast to
continued treatment with anti-convulsants, antidepressants, or antipsychotics
(Kessing et al. 2010).
Three studies concerning lithium treatment of AD or mild cognitive impairment
(MCI) were meta-analyzed by Matsunaga et al. (2015), suggesting some benefit
from lithium treatment. Of interest, one of these studies (Nunes et al. 2013) evaluated
the effect of a microdose of 300 μg lithium, administered to AD patients once daily
for 15 months. The treated group showed no decreased performance in the mini-
mental state examination test, in opposition to the lower scores observed for the
control group during this time. A possible effect of trace doses of lithium has been
recently suggested in a Danish study of Kessing et al. (2017) showing a negative
association between the incidence of dementia and lithium concentration in drinking
water. Concurring with this was a recent American study demonstrating that changes
in Alzheimer’s dementia mortality were negatively correlated with trace lithium in
drinking water (Fajardo et al. 2018). Taking into account the association between
suicidality and lithium in drinking water mentioned earlier, again, speculation on a
possible role of lithium as a trace element for mental health may be reasonable.
1508 J. K. Rybakowski
Table 1 Major meta-analyses and reviews on lithium therapeutic efficacy in the last decade
Subject of
study Authors Design Main results
Prevention of Nivoli et al. Six randomized controlled High effectiveness of
episodes in (2010) trials lasting at least half a year: lithium against mania and
bipolar 1,561 bipolar I and II patients, slightly less against
disorder randomizing 534 to lithium depression
Prophylaxis against mania:
lithium = valproate <
olanzapine
Prophylaxis against
depression:
lithium<lamotrigine
Prevention of Severus et Lithium versus placebo: seven Lithium more effective than
episodes in al. (2014) trials (1,580 participants) placebo in preventing:
bipolar overall mood episodes
disorder (RR 0.66)
manic episodes (RR 0.52)
depressive episodes
(RR 0.78)
Lithium versus anti- Preventing manic episodes:
convulsants: seven trials (1,305 lithium > anti-convulsants
participants) No difference in preventing
overall and depressive
episodes
Augmentation Bauer et al. 10 placebo controlled trials Mean response rate
of (2014) 30 open trials Lithium 41.2%; placebo
antidepressants 14.4%
Good efficacy of lithium
augmentation of TCA and
SSRI
No comparator superior to
lithium augmentation
Prevention of Undurraga Long-term lithium treatment: Lithium monotherapy >
episodes in et al. 21 trials (846 participants) placebo (7 trials)
unipolar (2019) Lithium monotherapy =
disorder antidepressant
Augmentation monotherapy (5 trials)
of Lithium + antidepressant >
antidepressants antidepressant
monotherapy (9 trials)
Augmentation of Lithium > placebo (OR
antidepressants 12 trials (541 2.34)
participants)
Treatment of Matsunaga Three clinical trials (232 Lithium > placebo in
Alzheimer’s et al. participants) decreasing cognitive
disease (2015) decline
Standardized mean
difference 0.41
Mood Stabilizers: Lithium 1509
The issues of lithium side effects and toxicity as well as their prevalence and
management strategies were recently reviewed by Gitlin (2016). In many cases,
adverse reactions to lithium present a significant burden and may play a role in non-
adherence and also in lithium discontinuation. These effects may appear already at
the initial stage of lithium therapy but also after several years of such treatment.
However, in most cases, they can be manageable. The adverse effects of lithium
perceived as posing a significant challenge for its long-term administration include
mostly kidney and thyroid effects.
Gastrointestinal side effects, such as nausea and diarrhea, occur in about 10–20% of
lithium-treated patients early in the treatment and tend to decrease during long-term
therapy. They may be connected with serum lithium concentration and with a type of
preparation (regular or slow-release). Both vomiting and diarrhea appear during lithium
intoxication. Weight gain is pretty common during lithium treatment although its
prevalence differs in various studies, pertaining, on the average, to 25% of patients.
This abnormality is sometimes bothersome, and, in some patients, it can result in lithium
discontinuation. Therapeutic options for lithium-induced weight gain include dietary
control and physical activity, adjunctive treatment with topiramate, and, in severe and
resistant cases, switching to another mood stabilizer (Rybakowski and Suwalska 2006).
Polyuria and polydipsia may occur within several weeks of lithium treatment. The
main reason is a decrease in renal concentrating capacity caused by lithium. This
side-effect can alleviate after reducing lithium dose. Therefore the optimal manage-
ment when these symptoms appear is keeping lithium levels as low as possible. If the
effect of lithium is favorable, severe polyuria can be treated with amiloride. Polyuria
normally disappears after lithium discontinuation. Some decrease of lithium-induced
renal concentrating capacity may persist for many years during long-term lithium
administration (Rybakowski et al. 2012).
The tremor, occurring in about 20% of lithium-treated patients, is a fine tremor of
the upper limbs occurring as postural or action tremor. It looks like an exaggeration
of physiologic tremor with a similar frequency range between 8 and 13 Hz, differing
from resting Parkinsonian tremor, which has a frequency of 4–6 Hz. Predisposing
factors include high lithium level; older age; and concomitant treatment with the
antidepressant, anti-convulsant, or antipsychotic drugs. Lithium-induced tremor is
usually mild and often disappears after dose reduction. However, if a reduction of
lithium dosage is not possible and if the tremor is considered clinically relevant (i.e.,
1510 J. K. Rybakowski
The most serious concern connected with long-term lithium use is the possibility of
lithium-induced interstitial nephropathy. This complication can develop after
10–20 years of treatment and leads to increased creatinine concentration and a
decreased glomerular filtration rate (GFR) (Rybakowski et al. 2012). In a recent
IGSLI study, Tondo et al. (2017) analyzed data from 312 patients with bipolar
disorder, coming from 12 participating centers, receiving lithium carbonate for
8–48 (mean 18) years. The lowering of GFR amounted to 0.71% with each year of
age and 0.92% with each year of lithium administration. Nearly 1/3 of subjects had
the value of GFR <60 mL/min/1.73 m2 more than once, more frequently after
15 years of lithium administration and after 55 years of age, and 18.1% of patients
had that value more than twice. However, no case of an end-stage renal failure was
detected.
In a small fraction of patients receiving long-term lithium administration, pro-
gressive renal damage may occur. Such a situation frequently fosters discontinuation
of lithium and replacing with other mood stabilizer. However, a decision about
Mood Stabilizers: Lithium 1511
stopping lithium should be taken with caution, especially in good responders, since
other mood stabilizers may not be equally efficacious. This may result in a high risk
of relapse of the illness and a further treatment-resistant course (Rybakowski et al.
2012). In patients with lithium-induced nephropathy, kidney function should be
closely and frequently monitored, and some guidelines for managing such patients
were recently published (Severus and Bauer 2013).
The influence of lithium on the thyroid gland is one of the key side effects in long-
term therapy with this drug. Lithium is accumulated in the thyroid gland at three- to
fourfold higher concentrations as compared to its plasma levels. Its administration
results in a reduced production with release inhibition of thyroid hormones, altering
the immune processes of this gland. Fifty years ago, Schou et al. (1968) were the first
to report the incidence of goiter in lithium-treated patients. According to the current
knowledge, the most common thyroid side effects associated with long-term lithium
treatment are goiter and hypothyroidism. Hyperthyroidism is a rare complication.
Lithium may also interfere with thyroid immunity (Kraszewska et al. 2014).
In bipolar disorder, the thyroid function should be examined in the context of the
role of the thyroid gland, the hypothalamic-pituitary-thyroid axis, and the thyroid
autoimmunity in the pathophysiology of this illness. Recently we studied 98 bipolar
subjects receiving lithium for mean 19 years and 39 subjects, matched for gender and
duration of the illness, never receiving lithium. The concentration of the thyroid-
stimulating hormone (TSH) and the volume of the thyroid gland were significantly
higher in patients receiving lithium. However, the frequency of hypothyroidism in
the course of the illness was similar in both groups (24% vs. 18%), higher in women
than in men (32% vs. 7% and 22% vs. 8%, respectively). In lithium-treated subjects,
hypothyroidism usually developed in the first years of lithium therapy, and all
hypothyroid patients were successfully treated with levothyroxine. No difference
in thyroid autoantibodies was found between the two groups (Kraszewska et al
2019a, b). In another study, we showed that patients who have been taking lithium
for 10–20 years had similar indexes of thyroid function as the patients in which the
drug was given for 20 years or more (Kraszewska et al. 2015).
Lithium increases renal calcium reabsorption and stimulates the release of the
parathyroid hormone. Long-term lithium administration causes a 10% increase of
serum calcium, and the cases of hyperparathyroidism have also been described. In
asymptomatic patients, monitoring of calcium level is recommended. In case of
serious hyperparathyroidism, the condition should be treated similarly as primary
cases.
performed by Wingo et al. (2009) including 276 lithium-treated and 263 similar
subjects lithium-free showed that lithium treatment of mood disorder patients was
associated with a small, but significant, impairment in immediate verbal learning and
memory and longer duration of treatment – with impairment of psychomotor
performance. On the other hand, the results of animal research point to lithium
exerting a favorable effect on cognitive function as demonstrated in various models.
Lithium treatment protects irradiated hippocampal neurons from apoptosis and
improves cognitive performance in irradiated mice. Such a pro-cognitive effect of
lithium was associated with the inhibition of GSK-3beta and an increase in Bcl-2
protein expression (Yazlovitskaya et al. 2006). In another study, using three different
positive reinforcement spatial cognitive tasks in rats, it was demonstrated that
lithium magnifies learning in all three tasks which may be associated with enhancing
hippocampal synaptic plasticity (Nocjar et al. 2007).
Clinical studies assessing the effect of lithium on cognitive functions were
recently reviewed (Rybakowski 2016). It seems that lithium treatment may
not negatively affect previously impaired cognitive functions in bipolar patients.
Spanish researchers (Lopez-Jaramillo et al. 2010) found a lower performance on
episodic verbal and visual-verbal memory tests of bipolar patients, compared to the
healthy subjects but no differences between lithium-treated and bipolar patients with
no medication intake and concluded that lithium therapy had no deleterious effect on
cognition. We have attempted to correlate cognitive functions in lithium-treated
patients with a quality of lithium prophylactic effect. We found that non-responders
to lithium had significantly worse performance on many domains of the Wisconsin
Card Sorting Test, compared to excellent and partial responders (Rybakowski et al.
2009). In another study, using neuropsychological tests from a CANTAB battery
which measured spatial working memory and sustained attention, we demonstrated
that bipolar patients who are excellent lithium responders have cognitive functions
comparable to those of matched control subjects, thereby probably constituting a
specific subgroup of bipolar patients in which long-term lithium administration can
produce complete normality in this respect (Rybakowski and Suwalska 2010).
Out of possible mechanisms mitigating the negative effect of lithium on cognitive
functions, the most important could be related to the prevention by lithium of
affective recurrences. Studies performed on bipolar patients have shown a correla-
tion between existing neuropsychological deficits, a greater number of affective
episodes, and a more severe course of the illness. The neurobiological components
mitigating a negative effect of lithium on cognitive functions could mostly be due to
its neurotrophic and neuroprotective effects. The most important neurobiological
processes in this respect are the stimulation of the BDNF system and inhibition of
GSK-3β. An antiviral effect of lithium, especially against herpes viruses, may also be
taken into account. Dickerson et al. (2004) demonstrated that infection with herpes
simplex virus type 1 was an independent predictor of decreased cognitive function-
ing (mostly immediate verbal memory) in bipolar patients, while a study of
Rybakowski and Amsterdam (2001) showed that lithium treatment exerted a signif-
icant anti-herpes activity. Finally, using lithium in the lowest possible dose could
also be a protective factor against possible cognitive problems.
Mood Stabilizers: Lithium 1513
In women with bipolar disorder, there is an increased risk for a recurrence of illness
during pregnancy. Therefore it is agreed upon that women with the previous
favorable effect of lithium should continue using lithium during pregnancy. Gener-
ally, such a procedure is thought to be relatively safe; however, in each case, the cost-
benefit ratio should be considered between the effectiveness of lithium at reducing
relapse and increased risk of potential complications with lithium use. A recent meta-
analysis included data from pregnant women and their children from 6 international
cohorts based in the community (Denmark, Sweden, and Canada) and clinics (the
Netherlands, UK, and the USA), identifying 727 lithium-exposed out of 22,124
eligible pregnancies. Lithium exposure was not associated with any of the predefined
pregnancy complications or delivery outcomes. Such an exposure during the first
trimester was associated with a 1.7-fold increased risk of major malformations, but
for major cardiac malformations, the difference was not significant. A 1.6-fold
increased risk for neonatal readmission within 28 days of birth was also seen in
the lithium-exposed group (Munk-Olsen et al. 2018). Women using lithium should
remember that during pregnancy lithium clearance increases. Therefore, starting
with the second trimester, there is a necessity to increase the lithium dose. Also,
there is a need to reduce the dose of lithium or temporarily stop its use for 1 or 2 days
before expected delivery.
Lithium is excreted in breast milk. In infants, the plasma levels of lithium may
reach 30–50% of the mother. Therefore, there is a need for the mother to reconsider
breastfeeding or to reduce lithium dose and monitor closely the infant for any signs
of toxicity.
Lithium Intoxication
The review of lithium interaction with other drugs was recently made by Finley
(2016). Pharmacokinetic interactions with lithium are mainly connected with the
effect of other drugs on renal lithium clearance. A decrease of lithium clearance
can be due to stimulation on proximal tubular sodium reabsorption with concom-
itant reabsorption of lithium. This makes a mechanism of decreasing lithium
clearance by thiazides, amiloride, and spironolactone. Furosemide exerts only
minimal effect in this respect. Drugs that are lowering blood pressure such as
angiotensin-converting enzyme (ACE) inhibitors, beta-blocking agents, or calcium
entry blocker, verapamil, can decrease lithium clearance by reducing renal perfu-
sion pressure. However, other calcium entry blockers such as nifedipine or
isradipine can cause an increase in lithium clearance by producing afferent arteri-
olar vasodilatation. The same mechanism pertains to xanthines, such as aminoph-
ylline, theophylline, and caffeine. Patients who abruptly stop excessive drinking of
coffee or tea may be at risk of decreasing lithium clearance which may even result
in intoxication. On the other hand, a decrease of lithium clearance using nonste-
roidal anti-inflammatory drugs (NSAIDs) is due to inducing by these drugs the
reabsorption of lithium in the distal neuron. Aspirin does not seem to have such an
effect. Accelerating lithium elimination can be also obtained through decreasing
lithium reabsorption in the proximal tubule by osmotic diuresis (e.g., mannitol),
carbonic anhydrase inhibitor, acetazolamide, and sodium bicarbonate (Schou and
Kampf 2006).
Pharmacodynamic drug interactions of lithium are mostly connected with
neural transmission. Since lithium can attenuate dopaminergic transmission and
enhance serotonergic one, this may be significant for the interaction of lithium with
antipsychotic and antidepressant drugs. Concomitant use of lithium with typical
antipsychotic drugs, with strong antidopaminergic activity, such as haloperidol,
can result in an increase of neurotoxicity, e.g., exaggeration of extrapyramidal
symptoms, disturbances of consciousness, and a higher risk of the neuroleptic
malignant syndrome. Such effects are dose-dependent. Conversely, lithium may
attenuate the effect of dopaminergic substances such as cocaine or amphetamine.
The use of lithium with serotonergic antidepressant drugs (tricyclic drug clomip-
ramine and all selective serotonin reuptake inhibitors – SSRI) may increase the risk
for serotonin syndrome. Using lithium with any tricyclic antidepressant may cause
an intensification of the muscular tremor. Concomitant use of lithium and anti-
convulsants, especially in higher concentrations may increase the neurotoxicity. It
can also be noticed that carbamazepine used with lithium can decrease lithium-
induced polyuria and lithium can prevent carbamazepine-induced leukopenia.
Lithium can cause a prolongation of action of a neuromuscular blocker,
suxamethonium, by interfering with the cholinergic system. Therefore, lithium
withdrawal has been recommended before the course of electroconvulsive therapy,
where suxamethonium is used.
The significant main interactions of lithium with other drugs are shown in Table 2.
Mood Stabilizers: Lithium 1515
Table 2 Pharmacokinetic and pharmacodynamic interactions between lithium and other drugs
Pharmacokinetic interactions (the effect via renal lithium clearance)
Drugs decreasing lithium clearance (the increase of lithium concentration and toxicity)
Diuretics: thiazides, amiloride (furosemide – minimal effect)
Angiotensin-converting enzyme (ACA) inhibitors: captopril, enalapril
Non-steroidal anti-inflammatory drugs (NSAID): indomethacin, diclofenac, piroxicam, ibuprofen
(aspirin – minimal effect)
Drug increasing lithium clearance (accelerating lithium elimination)
Xanthines: aminophylline, caffeine, theophylline
Osmotic diuretics: glucose, mannitol
Carbonic anhydrase inhibitors: acetazolamide
Sodium bicarbonate
Pharmacodynamic interactions (the effect via neural transmission)
Typical antipsychotic drugs, e.g., haloperidol: increase of neurotoxicity – an exaggeration of
extrapyramidal symptoms, disturbances of consciousness, higher risk of the neuroleptic malignant
syndrome
Psychostimulant drugs (cocaine, amphetamine) – attenuation of euphoriant action
Serotonergic antidepressant drugs, e.g., clomipramine, selective serotonin reuptake inhibitors
(SSRI): higher risk of the serotonin syndrome
Tricyclic antidepressant drugs: intensification of muscular tremor
Anti-convulsant drugs: increase of neurotoxicity (carbamazepine may decrease lithium-induced
polyuria)
Neuromuscular blockers (suxamethonium) – prolonged duration of action
Combination Therapy
Conclusions
Cross-References
References
Adityanjee, Munshi KR, Thampy A. The syndrome of irreversible lithium-effectuated neurotoxic-
ity. Clin Neuropharmacol. 2005;28:38–48.
Alda M. Pharmacokinetics of lithium. In: Bauer M, Grof P, Müller-Oerlinghausen B, editors.
Lithium in neuropsychiatry. The comprehensive guide. Boca Raton: Taylor & Francis; 2006.
p. 321–8.
Alda M. Lithium in the treatment of bipolar disorder: pharmacology and pharmacogenetics. Mol
Psychiatry. 2015;20:661–70.
Arfwedson A. Untersuchungen einiger bei der Eisen Grube von Utö vorkommenden Fossilien und
von einem darin gefundenen neuen feuerfesten Alkali. Schweiggers Journal für Chemie und
Physik. 1818;22:93–120.
Baastrup PC. The use of lithium in manic-depressive psychoses. Compr Psychiatry.
1964;5:396–408.
Baastrup PC, Schou M. Lithium as a prophylactic agent. Its effect against recurrent depression and
manic-depressive psychosis. Arch Gen Psychiatry. 1967;16:162–72.
Mood Stabilizers: Lithium 1517
Garrod AB. Gout and rheumatic gout. London: Walton and Maberly; 1859.
Geddes JR, Burgess S, Kawton K, Jamison K, Goodwin GM. Long-term lithium therapy for bipolar
disorder: systematic review and meta-analysis of randomized controlled trials. Am J Psychiatry.
2004;161:217–22.
Gerdtz J. Mental illness and the Roman physician: the legacy of Soranus of Ephesus. Hosp
Community Psychiatry. 1994;45:485–7.
Geddes JR, Goodwin GM, Rendell J Azorin J-M, Cipriani A, Ostacher MJ, et al. Lithium plus
valproate combination therapy versus monotherapy for relapse prevention in bipolar I disorder
(BALANCE): a randomized open-label trial. Lancet. 2010;375:385–95.
Germana C, Kempton MJ, Sarnicola A, Christodoulou T, Haldane M, Hadjulis M, et al. The effects
of lithium and anticonvulsants on brain structure in bipolar disorder. Acta Psychiatr Scand.
2010;122:481–7.
Gitlin M. Lithium side effects and toxicity: prevalence and management strategies. Int J Bipolar
Disord. 2016;4:27.
Goodwin GM, Bowden CL, Calabrese JR, Grunze H, Kasper S, White R, et al. A pooled analysis of
2 placebo-controlled 18-month trials of lamotrigine and lithium maintenance in bipolar I
disorder. J Clin Psychiatry. 2004;65:432–41.
Grof P. Excellent lithium responders: people whose lives have been changed by lithium prophy-
laxis. In: Birch NJ, Gallicchio VS, Becker RW, editors. Lithium: 50 years of psychopharma-
cology, new perspectives in biomedical and clinical research. Cheshire: Weidner Publishing
Group; 1999. p 36–51.
Grof P. Sixty years of lithium responders. Neuropsychobiology. 2010;62:27–35.
Hajek T, Weiner MW. Neuroprotective effects of lithium in human brain? Food for thought. Curr
Alzeimer Res. 2016;13:862–72.
Hajek T, Bauer M, Simhandl C, Rybakowski J, O’Donovan C, Pfennig A, et al. Neuroprotective
effect of lithium on hippocampal volumes in bipolar disorder independent of long-term treat-
ment response. Psychol Med. 2014;44:507–17.
Hammond WA. Treatise on diseases of the nervous system. New York: Appleton; 1871.
Hartigan GP. The use of lithium salts in affective disorders. Br J Psychiatry. 1963;109:810–4.
Hawking S. A brief history of time. From the Big Bang to back holes. New York: Bantam Books;
1988.
Hou L, Heilbronner U, Degenhardt F, Adli M, Akiyama K, Akula N, et al. Genetic variants
associated with response to lithium treatment in bipolar disorder: a genome-wide association
study. Lancet. 2016;387:1085–93.
International Consortium on Lithium Genetics (ConLi+Gen), Amare AT, Schubert KO, Hou L,
Clark SR, Papiol S, et al. Association of polygenic score for schizophrenia and HLA antigen and
inflammation genes with response to lithium in bipolar affective disorder: a genome-wide
association study. JAMA Psychiat. 2018;75:65–74.
Jamison KR. An unquiet mind. A memoir of moods and madness. New York: Alfred A. Knopf;
1996.
Jope RS. Glycogen synthase kinase-3 in the etiology and treatment of mood disorders. Front Mol
Neurosci. 2011;4:16.
Kapusta ND, Mossaheb N, Etzersdorfer E, Hlavin G, Thau K, Willeit M, et al. Lithium in drinking
water and suicide mortality. Br J Psychiatry. 2011;198:346–50.
Katagiri H, Takita Y, Tohen M, Higuchi T, Kanba S, Takahashi M. Safety and efficacy of olanzapine
monotherapy and olanzapine with a mood stabilizer in 18-week treatment of manic/mixed
episodes for Japanese patients with bipolar I disorder. Curr Med Res Opin. 2012;28:701–13.
Kessing LV, Sondergard L, Forman JL, Andersen PK. Lithium treatment and the risk of dementia.
Arch Gen Psychiatry. 2008;65:1331–5.
Kessing LV, Forman JL, Andersen PK. Does lithium protect against dementia ? Bipolar Disord.
2010;12:97–4.
Kessing LV, Gerds TA, Knudsen NN, Jørgensen LF, Kristiansen SM, Voutchkova D, et al.
Association of lithium in drinking water with the incidence of dementia. JAMA Psychiat.
2017;74:1005–10.
Mood Stabilizers: Lithium 1519
Kessing LV, Bauer M, Nolen WA, Severus E, Goodwin GM, Geddes J. Effectiveness of mainte-
nance therapy of lithium vs other mood stabilizers in monotherapy and in combinations: a
systematic review of evidence from observational studies. Bipolar Disord. 2018; https://doi.org/
10.1111/bdi.12623.. [Epub ahead of print]
Ketter TA, Miller S, Dell’Osso B, Wang PW. Treatment of bipolar disorder: review of evidence
regarding quetiapine and lithium. J Affect Disord. 2016;191:256–73.
Kraszewska A, Ziemnicka K, Jończyk-Potoczna K, Sowiński J, Rybakowski JK. Thyroid structure
and function in long-term lithium-treated and lithium-naïve bipolar patients. Hum
Psychopharmacol. 2019a;34:e2708.
Kraszewska A, Ziemnicka K, Sowiński J, Ferensztajn-Rochowiak E, Rybakowski JK. No connec-
tion between long-term lithium treatment and antithyroid antibodies. Pharmacopsychiatry.
2019b;52:232–6.
Kleindienst N, Greil W. Differential efficacy of lithium and carbamazepine in the prophylaxis of
bipolar disorder: results of the MAP study. Neuropsychobiology. 2000;42(suppl 1):2–10.
Kraepelin E. Psychiatrie. Ein Lehrbuch für Studierende und Ärzte. 6th ed. Leipzig: Barth; 1899.
Kraszewska A, Abramowicz M, Chłopocka-Woźniak M, Sowiński J, Rybakowski J. The effect of
lithium on thyroid function in patients with bipolar disorder. Psychiatr Pol. 2014;48:417–28.
Kraszewska A, Chlopocka-Wozniak M, Abramowicz M, Sowinski J, Rybakowski JK. A cross-
sectional study of thyroid function in 66 patients with bipolar disorder receiving lithium for
10–44 years. Bipolar Disord. 2015;17:375–80.
Lange C. Om Periodiske Depressionstilstande og deres Patogenese. Lund: Copenhagen; 1886.
Lange C. Periodische Depressionzustände und ihre Pathogenesis auf dem Boden der harnsäuren
Diathese. Hamburg/Leipzig: Verlag von Leopold Voss; 1895.
Leroy K, Ando K, Heraud C, Yilmaz Z, Authelet M, Boeynaems JM, et al. Lithium treatment arrests
the development of neurofibrillary tangles in mutant tau transgenic mice with advanced neuro-
fibrillary pathology. J Alzheimers Dis. 2010;19:705–19.
Lewitzka U, Severus E, Bauer R, Ritter P, Müller-Oerlinghausen B, Bauer M. The suicide
prevention effect of lithium: more than 20 years of evidence-a narrative review. Int J Bipolar
Disord. 2015;3:32.
Lopez-Jaramillo C, Lopera-Vasquez J, Ospina Duque J, et al. Lithium treatment effects on the
neuropsychological functioning of patients with bipolar I disorder. J Clin Psychiatry.
2010;71:1055–60.
Luria SE. A slot machine, a broken test tube: an autobiography. New York: Harper and Row; 1984.
Lyoo K, Dager SR, Kim JE, Yoon SJ, Friedman SD, Dunner DL, et al. Lithium-induced grey matter
volume increase as a neural correlate of treatment response in bipolar disorder. A longitudinal
brain imaging study. Neuropsychopharmacology. 2010;35:1743–50.
Machado-Vieira R. Lithium, stress, and resilience in bipolar disorder: deciphering this key homeo-
static synaptic plasticity regulator. J Affect Disord. 2018;233:92–9.
Malhi GS, Outhred T. Therapeutic mechanisms of lithium in bipolar disorder: recent advances and
current understanding. CNS Drugs. 2016;30:931–49.
Manchia M, Adli M, Akula N, Ardau R, Aubry JM, Backlund L, et al. Assessment of response to
lithium maintenance treatment in bipolar disorder: a Consortium on Lithium Genetics
(ConLiGen) report. PLoS One. 2013;8:e65636.
Mann L, Heldman E, Bersudsky Y, Vatner SF, Ishikawa Y, Almog O, et al. Inhibition of specific
adenylyl cyclase isoforms by lithium and carbamazepine, but not valproate, may be related to
their antidepressant effect. Bipolar Disord. 2009;11:885–96.
Marcus R, Khan A, Rollin L, Morris B, Timko K, Carson W. Efficacy of aripiprazole adjunctive to
lithium or valproate in the long-term treatment of patients with bipolar I disorder with an
inadequate response to lithium or valproate monotherapy: a multicenter, double-blind, random-
ized study. Bipolar Disord. 2011;13:133–44.
Matsunaga S, Kishi T, Annas P, Basun H, Hampel H, Iwata N. Lithium as a treatment for
Alzheimer’s disease: a systematic review and meta-analysis. J Alzheimers Dis. 2015;48:403–10.
Moncrieff J. Lithium: evidence rconsidered. Br J Psychiatry. 1997;171:113–9.
1520 J. K. Rybakowski
Moore GJ, Bebchuk JM, Wilds IB, Chen G, Manji HK. Lithium-induced increase in human brain
grey matter. Lancet. 2000;356:1241–2.
Müller-Oerlinghausen B, Ahrens B, Grof E, Grof P, Lenz G, Schou M, et al. The effect of long-term
lithium treatment on the mortality of patients with manic-depressive and schizoaffective illness.
Acta Psychiatr Scand. 1992;86:218–22.
Munk-Olsen T, Liu X, Viktorin A, Brown HK, Di Florio A, D’Onofrio BM, et al. Maternal and
infant outcomes associated with lithium use in pregnancy: an international collaborative meta-
analysis of six cohort studies. Lancet Psychiatry. 2018;5:644–52.
Nassar A, Azab AN. Effects of lithium on inflammation. ACS Chem Neurosci. 2014;5:451–8.
Nivoli AMA, Murru A, Vieta E. Lithium: still a cornerstone in the long-term treatment in bipolar
disorder? Neuropsychobiology. 2010;62:27–35.
Nocjar C, Hammonds MD, Shim SS. Chronic lithium treatment magnifies learning in rats.
Neuroscience. 2007;150:774–88.
Nunes MA, Viel TA, Buck HS. Microdose lithium treatment stabilized cognitive impairment in
patients with Alzheimer’s disease. Curr Alzheimer Res. 2013;10:104–7.
Ohgami H, Terao T, Shiotsuki I, Ishii N, Iwata N. Lithium levels in drinking water and risk of
suicide. Br J Psychiatry. 2009;195:464–5.
Pasquali L, Busceti CL, Fulceri F, Paparelli A, Fornai F. Intracellular pathways underlying the
effects of lithium. Behav Pharmacol. 2010;21:473–92.
Permoda-Osip A, Abramowicz M, Kraszewska A, Suwalska A, Chłopocka-Woźniak M,
Rybakowski JK. Kidney, thyroid and other organ functions after 40 years or more of lithium
therapy: a case series of five patients. Ther Adv Psychopharmacol. 2016;6:277–82.
Pisanu C, Papadima EM, Del Zompo M, Squassina A. Understanding the molecular mechanisms
underlying mood stabilizer treatments in bipolar disorder: potential involvement of epigenetics.
Neurosci Lett. 2018;669:24–31.
Quiroz JA, Machado-Vieira R, Zarate CA, Manji HK. Novel insights into lithium’s mechanism of
action: neurotrophic and neuroprotective effects. Neuropsychobiology. 2010;62:50–60.
Radomski J, Fuyat HN, Nelson AA, Smith PK. The toxic effects, excretion and distribution of
lithium chloride. J Pharmacol Exp Ther. 1950;100:429–40.
Remlinger-Molenda A, Wojciak P, Michalak M, Karczewski J, Rybakowski JK. Selected cytokine
profiles during remission in bipolar patients. Neuropsychobiology. 2012;66:193–8.
Rybakowski JK. Antiviral and immunomodulatory effect of lithium. Pharmacopsychiatry.
2000;33:159–64.
Rybakowski JK. Two generations of mood stabilizers. Int J Neuropsychopharmacol.
2007;10:709–11.
Rybakowski JK. Genetic influences on response to mood stabilizers in bipolar disorder: current
status of knowledge. CNS Drugs. 2013;27:165–73.
Rybakowski JK. Neurological, cognitive and neuroprotective effects of treatment used in bipolar
disorder. In: Yildiz A, Ruiz P, Nemeroff CB, editors. The bipolar book. History, neurobiology
and treatment. New York: Oxford University Press; 2015. p. 403–15.
Rybakowski JK. Effect of lithium on neurocognitive functioning. Curr Alzheimer Res.
2016;13:887–93.
Rybakowski JK. Meaningful aspects of the term ‘mood stabilizer’. Bipolar Disord. 2018;20:391–2.
Rybakowski JK, Amsterdam JD. Lithium prophylaxis and recurrent labial herpes infections.
Lithium. 1991;2:43–7.
Rybakowski J, Matkowski K. Adding lithium to antidepressant therapy: factors related to thera-
peutic potentiation. Eur Neuropsychopharmacol. 1992;2:161–5.
Rybakowski JK, Chłopocka-Woźniak M, Suwalska A. The prophylactic effect of long-term lithium
administration in bipolar patients entering treatment in the 1970s and 1980s. Bipolar Disord
2001;3:63–67.
Rybakowski JK, Suwalska A. Gastrointestinal, metabolic and body weight changes during treat-
ment with lithium. In: Bauer M, Grof P, Müller-Oerlinghausen B, editors. Lithium in neuro-
psychiatry. The comprehensive guide. Boca Raton: Taylor & Francis; 2006. p. 283–94.
Mood Stabilizers: Lithium 1521
Johannes M. Hennings
Contents
Chemistry, Developmental History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1524
Pharmacology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1525
Pharmacokinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1525
Mechanism of Action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1525
Chemically Related Drugs: Oxcarbazepine and Eslicarbazepine . . . . . . . . . . . . . . . . . . . . . . . . . . 1526
Indications (of Marketed Products) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1526
Clinical Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1527
Side Effects/Adverse Reactions and Toxicology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1527
Dose-Related Effects Related to Drug Mechanism of Action, and Toxic Effects . . . . . . . . . 1532
Not Necessarily Dose-Related Effects, Cutaneous Adverse Reactions . . . . . . . . . . . . . . . . . . . . 1532
Teratogenic Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1533
Combination Therapy – Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1533
Summary for Clinical Use in Psychiatric Indications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1534
Cross-References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1534
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1534
Abstract
Carbamazepine is an antiepileptic drug that is chemically related to tricyclic
antidepressants. In psychiatric indications, it has been first used as an antimanic
treatment in bipolar disorder in the 1970s. Carbamazepine is now approved as
second-line treatment in bipolar disorder with best evidence in the treatment of
acute manic episodes. Carbamazepine is also effective in the management of
alcohol and benzodiazepine withdrawal syndromes. It was somehow useful in
various other psychiatric indications, including aggressive and impulsive behav-
ior, suicidality, and as an add-on in refractory schizophrenia. However, these
approaches are poorly supported by data. The clinical use of carbamazepine, on
J. M. Hennings (*)
kbo Isar-Amper-Klinikum Munich, Haar/Munich, Bavaria, Germany
e-mail: Johannes.hennings@kbo.de
the other hand, is limited by its potential to interact with various other drugs,
some important side effects, especially in risk populations, and the presence of
somehow better alternatives in most indications. The carbamazepine-related
drugs oxcarbazepine and eslicarbazepine may have a more favorable side effect
and drug-interaction profile compared to its mother compound, but data are still
limited for the use in psychiatric indications.
Pharmacology
Pharmacokinetics
Carbamazepine itself is poorly soluble in water and its absorption is slow and
erratical after oral administration (McNamara 2001). Peak concentrations occur
after 4–8 h after oral ingestion. Carbamazepine distributes rapidly into all tissues,
but most of the drug will remain to plasma protein due to its high protein binding
property (about 75%). It is almost entirely metabolized in the liver by epoxidation
and hydroxylation. The elimination half-life is approximately 35 h, but it has been
reported to be lower in children (about 10 h) and higher in elderly patients
(30–50 h). The active metabolite carbamazepine-10,11-epoxide which may reach
50% of the parent compound plasma concentration, emerge from hepatic oxidation
via CYP3A4 and to a lesser extent CYP3A5 and CYP2C8 (Thorn et al. 2011).
The 10,11-epoxide is further metabolized and excreted partly as inactive glucuro-
nides in the urine. Carbamazepine upregulates transcription of CYP3A4 and other
enzymes, thus inducing its own metabolism. Subsequently increased clearance
may require dose adjustments after 2 or 3 week of treatment (Fricke-Galindo
et al. 2018).
Therapeutic drug concentrations in plasma are indicated between 4 and 12 μg/ml,
corresponding to 600–1200 mg/day. Nevertheless, there is no simple relationship
between the dose of carbamazepine and plasma concentrations, and dosing was not
specifically optimized for bipolar disorder (Baldessarini et al. 2019; Fountoulakis
et al. 2017).
Mechanism of Action
The prodrug oxcarbazepine and its active 10-hydroxy derivate eslicarbazepine are
structurally related to carbamazepine and presumably act also via binding to the
open channel conformation of the VSSC. Eslicarbazepine acetate is a once daily
antiepileptic drug that has been approved in 2009 by the European Medicines
Agency and in 2013 by the FDA and Health Canada as an adjunctive AED. After
oral administration, eslicarbazepine acetate undergoes extensive first pass hydrolysis
to its major active metabolite eslicarbazepine that represents about 95% of circulat-
ing active moieties (Soares-da-Silva et al. 2015). Compared to carbamazepine and
oxcarbazepine, eslicarbazepine has a higher affinity to the inactivated state of
the VSSC, resulting in a greater inhibitory effect on rapidly firing neurons.
Eslicarbazepine is further much more selective for the epileptogenic Cav3.2 calcium
channels (Soares-da-Silva et al. 2015).
While similar clinical effects as an antiepileptic agent has been described
compared to carbamazepine, the mood stabilizing properties of oxcarbazepine
seem to be less pronounced. Nevertheless, compared to carbamazepine,
oxcarbazepine seems to be less sedating, be less bone marrow toxic, and to
have fewer CYP3A4 interactions (Stahl 2013). Compared to carbamazepine,
eslicarbazepine is not metabolized to carbamazepine-10,11-epoxide, which is
accused to be related to enzymatic CYP450 induction and side effects
(Fricke-Galindo et al. 2018; McNamara 2001). Eslicarbazepine has shown an
improved tolerability, a better drug availability (once-daily administration) and
fewer drug-interactions compared to carbamazepine in the treatment focal
seizures (Fricke-Galindo et al. 2018) and, potentially, also in refractory bipolar
disorder. However, clinical efficacy in bipolar disorder, including acute
mania, has not yet been proven in a recent phase II clinical study (Grunze et al.
2018).
Clinical Studies
In the 1970s, first clinical studies demonstrated mood stabilizing and antimanic proper-
ties of carbamazepine (López-Muñoz et al. 2018). Despite its wide clinical use in
various psychiatric indications, the evidence based on randomized placebo-controlled
trials (RCT) is limited, even for the treatment of bipolar disorder (Baldessarini et al.
2019). Acute antimanic effects have been shown in two randomized placebo-controlled
trials with 427 subjects in total (Baldessarini et al. 2019). Carbamazepine was not
significantly different to placebo in one long-term trial (up to 12 months), and carba-
mazepine was inferior compared to lithium in two long-term trials. The evidence for the
use in bipolar depression is limited to one RCT (Baldessarini et al. 2019).
According to recent guidelines of the World Federation of Societies of Biological
Psychiatry (Grunze et al. 2018), carbamazepine and oxcarbazepine remain second-
line treatments in bipolar disorder. There is a grade 2 (fair research-based evidence)
recommendation of the International College Neuropsychopharmacology (CINP)
for carbamazepine for the acute treatment, while grades for oxcarbazepine and
eslicarbazepine are 3 (some evidence from comparative studies without placebo
arm or from post-hoc analyses) and 5 (negative data), respectively (Fountoulakis
et al. 2017). In acute mixed episodes, there is a grade 3 recommendation for
depressive and manic symptomatology, mirroring that the evidence is limited due
to only few available studies and syndromal overlap of this entity with other
conditions such as borderline personality disorder (BPD). Maintenance treatment
for manic or depressive episode prevention with carbamazepine has a level 4
recommendation (inconclusive data or poor quality of RCTs). There is no recom-
mendation of neither carbamazepine, oxcarbazepine, or eslicarbazepine in the NICE
guidelines (National Collaborating Centre for Mental Health (UK) 2018).
In contrast to some former reports of its potential use in refractory cases of
schizophrenia and schizoaffective disorder, carbamazepine cannot be recommended
as a routine use in the treatment of schizophrenia, even not if augmented to an
existing antipsychotic medication according to a recent meta-analysis (Leucht et al.
2014). Some smaller studies further suggests effectivity for behavioral control in
BPD as well as agitation, aggression, and impulsive behavior with or without
dementia (Belli et al. 2012; Gallagher and Herrmann 2014; Jones et al. 2011).
Carbamazepine and oxcarbazepine has shown mood-stabilizing properties in atten-
tion-deficit hyperactivity disorder (ADHD) (Davids et al. 2006; Silva et al. 1996).
Recommendations for carbamazepine’s various psychiatric indications are summa-
rized in Table 1, illustrating its wide use in off-label indications with limited
evidence from controlled clinical studies (Perucca and Gilliam 2012).
Table 1 (continued)
Approved Dose
Psychiatric indication rangea Summary of
indication (FDA) (mg/die) recommendation Reference(s)
N/A Possible mood-stabilizing Davids et al.
effect in adult ADHD for (2006), Silva
oxcarbazepine et al. (1996),
(300–1500 mg/die). Some Gorman et al.
preliminary evidence for (2015)
carbamazepine in child
and adolescent ADHD,
but recommendation
against use in child and
adolescent conduct
disorder due to ADRs and
limited behavioral effects.
Schizophrenia No 600–1200 Off-label use for Leucht et al.
refractory schizophrenia (2014)
and schizoaffective
disorder. According to a
recent meta-analysis, no
evidence for superiority
over placebo, even not in
combination with an
antipsychotic.
Suicidality No N/A Some effect in mood Chen et al.
disorders that cannot (2019), Tondo
clearly separated from and Baldessarini
primary anti-aggressive (2016)
effects, but less
pronounced than reported
for lithium. Negative
results in a recent meta-
analysis.
Substance use No 600–1200 Treatment of alcohol Welsh et al.
withdrawal symptoms, (2018), Fond et
but inconclusive evidence al. (2019), Soyka
in withdrawal seizure et al. (2017)
prevention compared to
benzodiazepines.
Reducing withdrawal in
benzodiazepine taper.
a
Note that dose recommendations derive from studies using occasionally various dosages, and
dosages may further be beyond this range in individual cases (e.g., in case of drug interactions).
Therefore, the dosage may be primarily adjusted according to the plasma concentration range
recommended for the treatment of epilepsia or bipolar disorder, i.e., 4 (6) to 12 μg/ml (Benkert
and Hippius 2019; Fountoulakis et al. 2017). Lower dosages (100–300 mg/die) have been used
throughout in child, adolescent and geriatric indications. N/A: not available (e.g., doses not
evaluated in studies). CINP: International College Neuropsychopharmacology, AED: antiepileptic
drug, BPD: Borderline personality disorder, ADHD: attention deficit hyperactivity disorder.
Table 2 Adverse drug reactions, clinical presentation, and recommendations
1530