Accepted Manuscript
Dendritic spines: Revisiting the physiological role
Saravana Babu Chidambaram, A.G. Rathipriya, Srinivasa Rao
Bolla, Abid Bhat, Bipul Ray, Arehslly Marappa Mahalakshmi,
Thamilarasan Manivasagam, Arokiasamy Justin Thenmozhi,
Musthafa Mohamed Essa, Gilles J. Guillemin, Ramesh Chandra,
Meena Kishore Sakharkar
PII:
DOI:
Reference:
S0278-5846(18)30607-9
https://doi.org/10.1016/j.pnpbp.2019.01.005
PNP 9574
To appear in:
Progress in Neuropsychopharmacology & Biological Psychiatry
Received date:
Revised date:
Accepted date:
22 July 2018
4 January 2019
12 January 2019
Please cite this article as: Saravana Babu Chidambaram, A.G. Rathipriya, Srinivasa Rao
Bolla, Abid Bhat, Bipul Ray, Arehslly Marappa Mahalakshmi, Thamilarasan
Manivasagam, Arokiasamy Justin Thenmozhi, Musthafa Mohamed Essa, Gilles J.
Guillemin, Ramesh Chandra, Meena Kishore Sakharkar , Dendritic spines: Revisiting the
physiological role. Pnp (2018), https://doi.org/10.1016/j.pnpbp.2019.01.005
This is a PDF file of an unedited manuscript that has been accepted for publication. As
a service to our customers we are providing this early version of the manuscript. The
manuscript will undergo copyediting, typesetting, and review of the resulting proof before
it is published in its final form. Please note that during the production process errors may
be discovered which could affect the content, and all legal disclaimers that apply to the
journal pertain.
ACCEPTED MANUSCRIPT
Dendritic Spines: Revisiting the Physiological Role
Saravana Babu Chidambaram1,1,* csaravanababu@gmail.com, Rathipriya AG2,1 , Srinivasa
Rao Bolla3 , Abid Bhat1 , Bipul Ray1 ,
Manivasagam4 ,
Arehslly Marappa Mahalakshmi1 , Thamilarasan
Arokiasamy Justin Thenmozhi4 ,
6
Musthafa Mohamed
7
Guillemin , Ramesh Chandra , Meena Kishore Sakharkar
8,**
Essa5 ,
Gilles J.
meena.sakharkar@usask.ca
1
D
MA
NU
SC
RI
PT
Dept of Pharmacology, JSS College of Pharmacy, JSS Academy of Higher Education &
Research (JSSAHER), Mysuru, Karnataka 570015, India
2
Food and Brain Research Foundation, Chennai, Tamil Nadu, India
3
Department of Anatomy, College of Medicine, Imam Abdulrahman Bin Faisal University,
Damam, Kingdom of Saudi Arabia
4
Department of Biochemistry and Biotechnology, Faculty of Science, Annamalai University,
Annamalainagar, Tamilnadu, India
5
Department of Food Science and Nutrition, CAMS, Sultan Qaboos, University, Muscat,
Oman
6
Neuropharmacology group, Faculty of Medicine and Health Sciences, Deb Bailey MND
Research Laboratory, Macquarie University, Sydney, NSW 2109, Australia
7
Department of Chemistry, University of Delhi, Delhi, India - 110007 and Ambedkar Centre
for BioMedical Research, Delhi University, Delhi, India - 110007
8
College of Pharmacy and Nutrition, University of Saskatchewan, 107, Wiggins Road,
Saskatoon, SK S7N 5C9, Canada
*
AC
ABSTRACT
CE
P
TE
Correspondence to: Saravana Babu Chidambaram, MPharm, PhD, FASc(AW), Associate
Professor, Department of Pharmacology, JSS College of Pharmacy, JSS Academy of Higher
Education & Research (JSS AHER), Mysuru, Karnataka 570015, India.
**
Correspondence to: Meena Kishore Sakharkar, PhD, Associate Professor, College of
Pharmacy and Nutrition, University of Saskatchewan, 107, Wiggins Road, Saskatoon, SK,
Canada S7N 5C9.
Dendritic spines are small, thin, specialized protrusions from neuronal dendrites, primarily
localized in the excitatory synapses. Sophisticated imaging techniques revealed that dendritic
spines are complex structures consisting of a dense network of cytoskeletal, transmembrane
and scaffolding molecules, and numerous surface receptors. Molecular signaling pathways,
mainly Rho and Ras family small GTPases pathways that converge on actin cytoskeleton,
regulate the spine morphology and dynamics bi-directionally during synaptic activity. During
synaptic plasticity the number and shapes of dendritic spines undergo radical reorganizations.
1
Equal contribution.
ACCEPTED MANUSCRIPT
Long-term potentiation (LTP) induction promote spine head enlargement and the formation
and stabilization of new spines. Long-term depression (LTD) results in their shrinkage and
retraction. Reports indicate increased spine density in the pyramidal neurons of autism and
Fragile X syndrome patients and reduced density in the temporal gyrus loci of schizophrenic
patients. Post-mortem reports of Alzheimer’s brains showed reduced spine number in the
hippocampus and cortex. This review highlights the spine morphogenesis process, the
activity-dependent structural plasticity and mechanisms by which synaptic activity sculpts the
RI
PT
dendritic spines, the structural and functional changes in spines during learning and memory
using LTP and LTD processes. It also discusses on spine status in neurodegenerative diseases
and the impact of nootropics and neuroprotective agents on the functional restoration of
SC
dendritic spines.
Keywords: dendritic spines, synaptic plasticity, spine dynamics, dendritic spine pathology,
AC
CE
P
TE
D
MA
NU
neurological diseases
ACCEPTED MANUSCRIPT
1. Introduction
In 1888, Ramón Y Cajal, a neuroanatomist discovered the existence of thin protrusions
emerging from the surface of certain neurons which he termed "Dendritic spines". Gray's
(1959) electron microscopy study defined that dendritic spines are postsynaptic structures
which play subtle functions in synaptic communications. Since dendritic spines comprise the
most important parts of excitatory synapses (McKinney, 2010; Nimchinsky et al., 2002) their
RI
PT
morphology and density is reported to play a crucial functional role in synaptic plasticity, and
consequently, in the learning and memory processes. Synaptic development, maintenance and
plasticity in both physiological and pathological conditions are frequently associated with
abnormalities in the morphology and numbers of dendritic spines (Fiala et al., 2002).
SC
However, the exact molecular mechanisms involved in synaptic plasticity formation and role
of dendritic spines remain a matter of debate. The mechanisms of synapse formation and the
NU
effects of therapeutic interventions in neural circuitry in neurological disorders have not been
elaborated. In this review, the authors illustrate the major molecular mechanisms of dendritic
MA
spines in terms of the long-term potentiation (LTP) and long-term depression (LTD)
processes; the fundamental variations in dendritic spines in acute and chronic neurological /
TE
regulation of dendritic spines.
D
neurodegenerative disorders; and the effects of neuroprotective and nootropic drugs on
2. Dendritic spines morphogenesis
In the early postnatal stage, the dendrites are devoid of spines and synapses. During
CE
P
synaptogenesis, finger-like actin-rich protrusions called 'filopodia' emerge from dendrites and
form synapses with adjacent axons (Portera and Yuste, 2001; Zhang and Benson, 2000; Ziv
and Smith, 1996). Normally, filopodia measure 3-40 µm in length (García-López et al., 2010)
AC
and are highly mobile structures that extend or retract transiently to form synaptic contacts
with adjacent dendritic shafts or to a new small protospine. Upon synaptic contact, these
filaments
morphologically
transform into
spines with functional synapses termed
as
protospines (Dailey and Smith, 1996; Portera-Cailliau et al., 2003). Later, the protospines
undergo morphological changes and turn either into mushroom or thin dendritic spines
(Majewska et al., 2006). The formation and maturation of spines fluctuates constantly
depending upon the synaptic strength and activities (Hotulainen and Hoogenraad, 2010).
However, not all the filopodia transform into spines (Ziv and Smith, 1996). Only a small
percentage of filopodia (0.2%) stimulated during synaptic activity mature into spines
(Majewska et al. 2006). Classically, dendritic spines consist of bulbous heads attached to the
ACCEPTED MANUSCRIPT
dendritic shafts by narrow necks (Nimchinsky et al., 2002; Rao et al., 1998). The size, shape
and number of dendritic spines vary heterogeneously even within a single dendritic segment
of a given neuron. The number of dendritic spines varies from 1 to 10 spines/um length of a
dendrite depending on the type of neuron and the developmental stage (Calabrese et al.,
2006). Dendritic spines in a matured brain are typically <3 µm in length with a spherical head
0.5-1.5 µm in diameter which is connected to the parent dendrite by a thin neck (<0.5 µm in
diameter) (Phillips et al., 2015).
RI
PT
Generally, dendritic spines are classified into 4 types on the basis of their morphology (Harris
et al., 1992; Peters and Kaiserman-Abramof, 1970) as mentioned below (Fig. 1).
mushroom spines with a large head and a narrow neck (more stable).
•
thin spines with a smaller head and a narrow neck (less stable).
•
stubby spines without an obvious constriction between the head and the attachment to
SC
•
branched or cup-shaped spines with two heads attached to a single narrow neck.
3. Dendritic spines - ultrastructure
MA
•
NU
the shaft.
3.1. Main structural components of dendritic spines
Major components of dendritic spines include cytoskeleton structures (actin), cytoskeleton
D
proteins, cell membrane receptors (NMDA, AMPA and metabotropic receptors), small
TE
GTPase and associated proteins, post-synaptic density (PSD) region, micro RNA, mRNA
3.1.1. Actins
CE
P
binding proteins, transcription factors, extra-cellular matrix and adhesion molecules (Fig. 2).
Actins are highly dynamic structures that act as the main cytoskeleton components of
AC
dendritic spines (Bosch and Hayashi, 2012; Koleske, 2013; Penzes and Rafalovich, 2012).
Dendritic spines contain high levels of actin, approximately 6 times higher than dendritic
shafts (Honkura et al., 2008). Actin exists either in a polymerized filamentous form
(filamentous actin / F-actin) or as a soluble depolymerized monomeric form (globular actin /
G-actin). In dendritic spines, F-actin forms the major proportion, whilst G-actin contributes to
only about 12% of the total actin content (Honkura et al., 2008; Star et al., 2002). F-actin
structure varies from branched filament networks to thick bundles of cross-linked filaments,
which provide support and stability to dendrites (Blanchoin et al., 2014). Cross-talks between
F- and G-actins allows morphological changes of dendritic spines during synaptic activities.
These results suggest that mainly actin polymerization and/or depolymerization determines
ACCEPTED MANUSCRIPT
the motility, growth and shape of the dendritic spines during maturation process (Blanchoin
et al., 2014; Pollard and Cooper, 2009). In fact, most of the morphological changes in spines
are mediated by actin polymerization (Segal and Andersen, 2000).
3.1.2. Myosin
Myosin II, a hexameric polypeptide consists of 2 heavy chains and 2 pairs of regulatory light
chains (Ryu et al., 2006a). Myosin IIB, the major non-muscle myosin II isoform found in
RI
PT
brain binds and contracts actin filaments. It is essential for normal spine morphology and
dynamics (Kawamoto and Adelstein, 1991; Ryu et al., 2006b; Vicente-Manzanares and
Horwitz, 2010a). Recent mass spectrometry data show abundant presence of myosin heavy
SC
chain IIB in the post synaptic density (PSD) region of neurons in rats (Jordan et al., 2004;
Peng et al., 2004). Loss-of-function experiments using both pharmacologic (blebbistatin, a
NU
small molecule inhibitor of myosin II) and genetic (myosin IIB-RNAi, hippocampal neurons
transfected with a pSuper plasmid expressing small hairpin RNAs against rat myosin IIB)
approaches led to depletion of mushroom-type spines and their replacement with irregular
MA
filopodia-like processes confirming the essential necessity of myosin IIB for spine
morphology and normal motility (Ryu et al., 2006b). Short-term inhibition of myosin IIB
activity induces immature filopodia-like spines and results in a corresponding disruption of
D
LTP and memory acquisition (Rex et al., 2010; Ryu et al., 2006b).
TE
The immunocytochemical localization of myosin IIB at synapses indicates its crucial role in
shaping the dendritic spine head. Myosin IIB immunoreactivity is found to be more intense in
soma
and
proximal dendrites
CE
P
the
in
immature rat hippocampal neurons vs.
the
immunoreactivity was more concentrated in the neck to the head of dendritic spines in the
more mature neurons. Myosin IIB activity is regulated by phosphorylation on residues Thr18
AC
and/or Ser19 in its regulatory light chain; simultaneous phosphorylation on both residues
promotes maximal myosin ATPase activity and the formation of large actin bundles (Ikebe
and Hartshorne, 1985; Vicente-Manzanares et al., 2009; Vicente-Manzanares and Horwitz,
2010b). Myosin IIB mediates and regulates the emergence of filopodia from the dendritic
shaft, spine elongation, and maturation into mushroom spines indicating its role as main
coordinator during the different stages of spine development (Yuste and Bonhoeffer, 2004).
3.1.3. GTPases
The Rho family of small GTPases (particularly RhoA, Rnd1, Rac1 and Cdc42) are the main
regulators of actin and have a profound influence on spine morphogenesis (Ishikawa et al.,
ACCEPTED MANUSCRIPT
2003; Newey et al., 2005). The Ras family of GTPases and their downstream MAP kinase
signaling pathways also regulate spine morphogenesis. A major postsynaptic inhibitor of Ras
signaling is SynGAP (synaptic Ras GAP), which is enriched in PSD. Even though both Rap
GTPases and Ras exhibit a close relationship in synaptic plasticity, they play opposite roles.
Ras promotes long term potentiation (LTP), while Rap1 and Rap2 mediate long term
depression (LTD) and depotentiation, respectively (Zhu et al., 2005). In contrast to RasGAP
and SynGAP, the postsynaptic RapGAP SPAR (spine associated RapGAP) enhances spine
RI
PT
growth (Pak et al., 2001; Pak and Sheng, 2003). Ras and Rap act antagonistically in PSD
matrices to regulate the spine morphology and synapse strengths.
SC
3.1.4. PSD region
In the adult mouse neocortex, the majority (almost 96%) of dendritic spines are characterized
NU
by large complex electron-dense structures defined as PSD that mark synaptic contacts
(Arellano et al., 2007). Within the PSD matrix, synaptic vesicles comprising the excitatory
isoxazolepropionic
-
glutamate
acid
and
(AMPA),
its
receptors
-
N-methyl- D-aspartate
MA
neurotransmitter
α-amino-3-hydroxy-5-methyl-4(NMDA)
and
metabotropic
glutamate receptors are established (Calabrese et al., 2006; Kennedy, 2000; Nimchinsky et
al., 2002). AMPA receptors are found at the edges of PSD while NMDA receptors are
D
located centrally within the PSD (Kharazia and Weinberg, 1997). Morphological alterations
TE
of dendritic spines mainly depend on NMDA receptor’s activation and contribute to the
formation of new synapses during LTP induction and the regression of existing synapses
CE
P
during LTD induction (Matsuzaki et al., 2004; Park et al., 2006; Zhou et al., 2004).
The PSD surfaces range from small discs to large irregular shapes that can be perforated by
electron lucent regions (Bourne and Harris, 2008). A recent protein analysis of PSD revealed
AC
numerous properties such as actin binding and cytoskeletal proteins (actin, N-catenin,
Arp2/3, calponin, cortactin, cofilin, drebrin, myosin IIB, myosin VI, neurabin I, neurabin II,
spinophilin, profilin I/II, synaptopodin and SPIN90); small GTP-ases (ARF6, Cdc42, Rac1,
Rap1, Rap2, Rem2, RhoA, Rif and Rnd) and associated proteins (Rho-GEF: ARHGEF6/PIX,
ARHGEF7/PIX, Dock180, GEFT, kalirin-7, Lfc/GEF-H1, Tiam1, Vav and Rho-GAP:
alpha1chimerin, oligophrenin1, p190RhoGAP, p250GAP, RhoGAP2, SrGAP2); cell surface
receptors (beta2-nAChR, GABBAR, GluA2, GluN1, GluN2B, Npn-2, NgR1, PGC-1);
adhesion
molecules
(alpha3-integrin,
alpha5-integrin,
arcadlin,
DSCAM,
N-cadherin,
neuroligins (1,2,3,4), syndecan-2, telencephalin, APP, vezatin); receptor tyrosine kinases
(EphB1/2/3, EphA4/ephrin-A3, ErbB2/B4, p75NTR, TrkB); other kinases (PAK, PAK1,
ACCEPTED MANUSCRIPT
PAK3, CaMKII, CaMKII-alpha, CDKL5, DCLK1, DGKf, LIMK-1, MARK4); post synaptic
scaffold proteins (G-protein coupled receptor kinase-interacting protein 1, PSD-95, Homer1a,
Homer1b, intersectin-s, PSD95, SAP102, TANC1/2, PAR-3, PAR-6 and Shank1,2,3) and
adaptor proteins (afadin, IRSp53, numb); and micro RNA, mRNA binding proteins, and other
transcription factors that are involved in various signaling cascades in dendritic spines during
synaptic plasticity (Husi et al., 2000; Maiti et al., 2015; Walikonis et al., 2000). The scaffold
protein - PSD-95, found closely in the postsynaptic membrane zone interacts directly with
RI
PT
several NMDA receptor subunits (Fujita and Kurachi, 2000; Kim and Sheng, 2004),
membrane proteins and many cytoplasmic signaling molecules and thus it (PSD-95)
3.1.5. Other cellular organelles in dendritic spines
SC
facilitates signal conduction in PSD (von Bohlen und Halbach, 2009) (Fig. 3).
NU
Myosin II, a hexameric polypeptide, consists of 2 heavy chains and 2 pairs of regulatory light
chains (Ryu et al., 2006a). Myosin IIB is the major non-muscle myosin II isoform found in
the brain. It binds and contracts actin filaments and is essential for normal spine morphology
MA
and dynamics (Kawamoto and Adelstein, 1991; Ryu et al., 2006a; Vicente-Manzanares et al.,
2009). Recent mass spectrometry data shows an abundant presence of myosin - heavy chain
IIB in the PSD region of neurons in rats (Jordan et al., 2004; Peng et al., 2004). Some forms
D
of smooth endoplasmic reticulum are found in the spines of hippocampal CA1 and Purkinje
TE
neurons (Spacek and Harris, 1997), while in few larger spines, endoplasmic reticulum
transforms itself into a specialized organelle called 'spine apparatus' (Gray and Guillery,
CE
P
1963). Endocytic machineries located within specialized subdomains have also been noted in
spines (Rácz et al., 2004; Song and Huganir, 2002). Stimulation of the synapses translocate
mitochondria into spines; further the number and functions of mitochondria determine the
AC
spine morphogenesis in dendrites (Li et al., 2004). Additionally, polyribosomes and protein
translational machineries are rooted at the base of spines (Ostroff et al., 2002). Such
specialized molecular assemblies determine the shape and function of the spine and facilitate
the postsynaptic neurons to biochemically respond to glutamate or other transmembrane
signals (Hering and Sheng, 2001; Smart and Halpain, 2000; Zhang and Benson, 2001).
3.2. Molecular mechanisms involved in dendritic spines morphogenesis
New outgrowth of dendritic spines or changes in existing spine morphology during LTP
mostly results from the modulation of actin filaments (Fukazawa et al., 2003; Kim and
Lisman, 1999; Lin et al., 2005; Okamoto et al., 2004; Ouyang et al., 2005). Actin
ACCEPTED MANUSCRIPT
polymerization can be regulated directly by actin-binding proteins (mainly cofilin and
profilin) which in turn are controlled by many upstream signaling pathways (particularly
PAK kinase, LIM kinase, calcineurin phosphatase and slingshot phosphatase) involved in
synaptic activity (Chen et al., 2007; Halpain et al., 1998; Meng et al., 2002). Several studies
reveal that Arp2/3 complex stimulates the nucleation of new actin filaments and the
formation of branches (Matsuzaki et al., 2004; Nägerl et al., 2004; Okamoto et al., 2004).
LTP causes an early transient phase of actin depolymerization which allows for structural
RI
PT
plasticity and motility leading to new spine formation or changes in spine morphology,
followed by subsequent polymerization of F-actin. The above-mentioned changes in-turn
result in the long-term stabilization of these dendritic changes and consolidation of LTP
SC
(Ouyang et al., 2005). Actin-stabilizing proteins such as phosphorylated-cofilin (Gohla and
Bokoch, 2002), profilin (Ackermann and Matus, 2003) and gelsolin (Star et al., 2002) are
NU
translocated to the spine heads by LTP-mediated NMDA receptor activation (Okamoto et al.,
2004).
As the major function of spines is to compartmentalize the chemical changes within the
MA
individual synapses, the length and diameter of the neck of the spine influence the degree and
dynamics of postsynaptic intracellular Ca2+ elevation, which is mediated by NMDA receptor
activation. Small spines show greater increases in Ca2+ owing to reduced leaks through the
D
thinner spine neck (Matsuzaki et al., 2004; Noguchi et al., 2005). The enlargement of the
TE
spine head largely depends upon the strength of synaptic stimuli. This is presumably related
to higher levels of AMPA-type glutamate receptors in the larger spines (Kasai et al., 2003a;
CE
P
Matsuzaki et al., 2001).
With the maturation of synapses, the stimulation of NMDA receptors promotes actin
mobilization to nascent synaptic sites leading to the emergence of new motile filopodia. The
AC
acquisition of AMPA receptors (Shi et al., 1999) and recruitment of adhesion molecules to
synaptic junctions (Bozdagi et al., 2000) suppresses the motility of actin and stabilizes the
morphology of spine (Matus, 2000). In dissociated hippocampal cells, the blockage of the
neuronal activity by tetrodotoxin decreases the number of spines or leads to the emergence of
filopodia (Kang et al., 2009; McKinney et al., 1999; Verpelli et al., 2010). Furthermore,
blocking the AMPA receptor by NBQX results in spine loss (Kang et al., 2009; McKinney et
al., 1999). Blocking NMDA receptors by AP5 (selective NMDAr antagonist) leads to the
appearance of filopodia-like processes without reducing the density of total dendritic spines,
indicating different roles for the two types of receptors in the maintenance and maturation of
spines (McKinney et al., 1999). Loss-of-function experiments via both pharmacologic
ACCEPTED MANUSCRIPT
(blebbistatin, a small molecule inhibitor of myosin II) and genetic (myosin IIB-RNAi,
hippocampal neurons transfected with a pSuper plasmid expressing small hairpin RNAs
against rat myosin IIB) approaches led to depletion of mushroom-type spines and their
replacement with irregular filopodia-like processes confirming the necessity of myosin IIB
for spine morphology and normal motility (Ryu et al., 2006a).
4. Advanced methods for studying dendritic spines morphology and dynamics
RI
PT
Dendritic spines, the fundamental units of synaptic activity undergo various structural
aberrations in a variety of neurological diseases. An accurate and unbiased structural
characterization of dendritic spines are vital for the understanding of their involvement in the
SC
development of neural circuitry, synaptic activity and synaptic failure in neurological
diseases (Fiala et al., 2002; Maiti et al., 2015). The most commonly used methods are Golgior fluorescent dye staining in post-mortem brain tissues.
Fortunately,
recent
NU
Cox
technological advances in microscopy, molecular biology, computational biology, immunelabeling techniques, and protein and genetic engineering have now made imaging dendritic
MA
spine dynamics in the living brain not a distinct reality (Mancuso et al., 2013).
4.1. Golgi staining and fluorescent-labeling methods
D
In classical staining techniques, thick sections of post-mortem brain tissues are immersed in a
TE
solution containing potassium chromate and dichromate and chloride salts of heavy metals
(silver or mercury) for several weeks. After procedures such as the dehydration of the tissue
CE
P
and clearing, these Golgi-impregnated tissues are mounted on microscope slides and viewed
under bright field microscopy to study the structure of individual dendritic spines and
dendritic arborizations (Gipson and Olive, 2017). Main advantages of staining (Golgi, Golgi-
AC
Cox and rapid or modified) methods are less costs, and clear images of the entire neuron. The
procedure is also time efficient and the samples are resistant to fading or photobleaching over
time. Major disadvantages of these staining techniques include limited capacity to determine
the neurochemical phenotypes of impregnated neurons, inconsistent specificity, overlapping
and out-of-focus dendritic segments. Staining methods cannot be applied to visualize the
cultured living neurons, and this may cause underestimation of spine numbers due to the twodimensional nature of the obtained images (Gipson and Olive, 2017). These disadvantages
have been overcome by the development of fluorescent Golgi stains (Koyama and Tohyama,
2013), confocal laser scanning microscopy (Tredici et al., 1993), and immunofluorescence
procedures that allow the identification of labeled neurons (Pinto et al., 2012; Spiga et al.,
ACCEPTED MANUSCRIPT
2011). Following their labeling with fluorescent dyes, images are taken on a confocal laser
scanning microscope and serial images obtained are processed for three-dimensional
reconstruction, and then, the density and morphology of dendritic spines is assessed. Other
fluorescent labeling methods include the use of various commercially available tracer dyes,
fluorochrome-labeled antibodies, and genetically encoded fluorescent proteins such as green
fluorescent protein (GFP) or yellow fluorescent protein (YFP) (Malinow et al., 2010;
Staffend and Meisel, 2011). Among these methods, viral-transfected fluorescent protein
immuno-labeling techniques, and transgenic animal engineering have been
RI
PT
engineering,
helpful in elucidating the detailed structure and dynamics of dendritic spines seen in different
SC
brain diseases (Sala and Segal, 2014) (Fig. 4).
4.2. Advanced microscopic tools
NU
Using optical microscope and electron microscope (especially single sections), researchers
classically defined the clear distinction between stubby, thin, and mushroom shaped spines
(Peters and Kaiserman-Abramof, 1970). Many advanced microscopic techniques such as
MA
confocal laser scanning microscopy and two-photon microscopy/multi-photon laser scanning
microscopy overcome the diffraction barrier of light to provide spatial resolution from the
microscale to the nanoscale (Maiti et al., 2015). Digital recording of spine images, including
D
semi-automated software-guided tracing systems (example NeuroZoom, Neurolucida), were
TE
later developed to characterize the fine 3-D structure of dendritic spines (Glaser and Glaser,
CE
P
1990; Sala and Segal, 2014).
4.2.1. Two-photon laser scanning microscopy (Fig. 5):
Widely used as the long wavelength-excitation light helps to penetrate the intact
AC
nervous tissue to a depth of several hundred microns (Denk et al., 1990).
Neurons of interest are typically labelled with fluorescent proteins such as GFP or
YFP either in transgenic mouse lines or after labelling through viral transduction in in
vitro models, and in in vivo studies up to depths of approximately 800 μm can be
studied (Mittmann et al., 2011).
Within a single spine, two-photon microscopy activates proteins tagged with
photoactivatable GFP enabling the study of the dynamics of dendritic spines during
activation-dependent plasticity (like LTP) (Gray et al., 2006; Steiner et al., 2008).
ACCEPTED MANUSCRIPT
Direct structure-function relationships in dendritic spines during synaptic activation
are studied using two-photon uncaging of a caged glutamate compound (Branco et al.,
2010; Branco and Häusser, 2011; Govindarajan et al., 2011). The disadvantage of this
methodology is that the quality of the picture through a thinned skull, without the
incorporation of other methods like microendoscopy, declines at distances of more
RI
PT
than 50 μm from the brain surface (Isshiki and Okabe, 2014).
4.2.2. Super-resolution dendritic spine imaging techniques:
Recent developments in imaging technologies enable researchers to produce very vibrant
images of the dendritic spines of living neurons and show extraordinary details in spine
SC
structure and dynamics. These include
Saturated stimulated emission depletion microscopy which enables us to view the
NU
morphology of each spine in detail. The thickness of the spine neck or subtle changes in
the head shape and size (from thin to mushroom shaped spines) at a resolution less than
70 nm (Hell, 2007) is shown. It provides images of dendritic spines at a nanoscale
MA
resolution (Nägerl et al., 2008) and also in the living brain (Berning et al., 2012;
Heilemann, 2010).
Photoactivation localization microscopy (Betzig et al., 2006) and stochastic optical
D
reconstruction microscopy (Rust et al., 2006) rely on the sequential, statistical excitation
TE
of a fraction of fluorophores spaced at distances larger than the diffraction limit
(Giannone et al., 2010; Huang et al., 2010). Both approaches give a spatial resolution of
CE
P
around 20 nm and are commonly suitable in fixed cell cultures.
Digital holographic quantitative phase microscopy (Marquet et al., 2013) and fiber-optic
endo-microscopes (Barretto and Schnitzer, 2012; Gu et al., 2014; Murari et al., 2011) are
AC
also used to study the distributions and dynamics of receptors and signaling molecules in
dendritic spines.
4.3. Viral-mediated neuronal tracing
In recent decades, viral strains are genetically engineered by recombinant technology and
used as tools to map and trace the motor, visual, and auditory neural circuits in in vivo set-up.
Advancements in viral vector technology led to the development of new tracing tools using
viruses which help the researchers to distinctly label the neural pathways embedded in
complex
brain
circuits
underlying
particular
brain
functions.
Viral-mediated
tracing
ACCEPTED MANUSCRIPT
technology uses natural viruses that infect, persist, and migrate within neurons. These viral
tracers can also spread across the synaptic connections either in retrograde or anterograde
directions (Brian et al., 2017). These tracers provide an excellent imaging toolbox to
precisely map the input and output connectivity of specific neuronal population or a
particular cell type (Callaway and Luo, 2015; Junyent et al., 2015; Oyibo et al., 2014;
Defalco et al., 2001; Gradinaru et al., 2010; Schwarz et al., 2015; Ugolini, 2010). Viruses
which replicate and spread across the presynaptic connections are known as retrograde trans-
RI
PT
synaptic viral tracers (Oztas, 2003), and those viruses that replicate and spread across the
downstream post-synaptic connections are known as analogous anterograde trans-synaptic
tracers (Oztas, 2003, Wickersham et al., 2007) (Fig. 6)
SC
Most common retrograde trans-synaptic viral tracers include Rabies Virus (RABV) and the
Bartha strain of pseudorabies virus (Dietzschold et al., 2008). These retrograde trans-synaptic
NU
viral tracers replicate and spread in a synapse-specific manner and are used traditionally for
mapping presynaptic inputs to transduced neurons (Wickersham et al., 2007; Wall et al.,
MA
2010; Defalco et al., 2001; Ekstrand et al., 2008, Callaway, 2008, Luo et al., 2008, Ugolini,
2008 and 2011). Remarkably, RABV is the most promising retrograde tracer tool to map the
synaptically connected neuronal populations (Kelly and Strick, 2000; Wickersham et al.,
D
2007a; Ugolini, 2011). However, the native RABV spreads across multiple serial synapses
TE
non-specifically, providing ambiguous results on the interpretation of the synaptic steps
crossed at any given time. Thus, it was difficult to map the input and output connections
CE
P
precisely between rabies labeled neurons. To prevent the polysynaptic expansion of RABV,
glycoprotein (GP) gene deleted-RABV, a variant of RABV has been developed recently,
using trans-complementation approach, which enables the scientists to label only a neuronal
AC
connection via a single synaptic step. Monosynaptic neural circuit tracing using GP-deleted
RABV (RVdG) has been a generally used tool to study the anatomy and function of neural
circuitry (Callaway and Luo, 2015; Wickersham et al., 2007b). Using trans-complementation
strategy, a recombinant RABV with its GP gene replaced with the coding sequence for
enhanced GFP (EGFP) was developed (RVdG-EGFP). Even though RVdG-EGFP cannot
spread beyond initially infected neurons, it can replicate within the viral core, and produce
marked levels of EGFP to brightly label the fine dendritic and axonal structures (Wickersham
et al., 2007). By pseudotyping GP-deleted RABV with EnvA (envelope protein of avian
sarcoma and leukosis virus), it became much easier to target RABV infection to specific
transduced cell types or even single neurons, and to trans-synaptically map only their direct
ACCEPTED MANUSCRIPT
presynaptic inputs in the mammalian central nervous system (CNS) (Wickersham et al.,
2007b). This new methodology enabled the tracing of monosynaptic connections of defined
neurons (Wickersham et al., 2007b; Callaway, 2008; Marshel et al., 2010). By combining this
methodology with the expression of a variety of genes, optical or electrophysiology methods,
and behavioral assays, the researches have access to perform integrative studies about the
neural circuits and their physiological functions (Callaway, 2008; Arenkiel and Ehlers, 2009;
Despite
the
anterograde
successful development
trans-synaptic
tools
for
of retrograde
tagging
RI
PT
Osakada et al., 2011; Wickersham and Feinberg, 2012).
trans-synaptic
viruses,
analogous
postsynaptical neural pathways labeling
downstream of targeted neurons remains an active area of investigation. Analogous
SC
anterograde trans-synaptic viral tracers use viruses like Herpes Simplex Virus (HSV, Norgren
et al., 1998, McGovern et al., 2012), Vesicular stomatitis virus (Beier et al., 2011), Sindbis
NU
virus (Beier et al., 2001) and Adeno-associated virus (AAV, Chamberlin et al., 1998). Even
though HSV and vesicular stomatitis virus migrate through neurons only where it can
MA
replicate, they spread across several synaptic connections labeling a large neuron population
(Ugolini et al., 1989). But, the neurotoxicity of these viruses and their uncontrollable spread
across multiple serial synapses limits their applications in precisely mapping the neural
D
circuits (Lo and Anderson, 2011; Beier et al., 2011). In brain regions, AAV tracers spread
TE
across multiple synaptic connections and transduce wide variety of neuronal populations.
Though, AAV vectors expressing various fluorescence reporter genes have been widely used
CE
P
for anterograde circuit mapping, mostly it provides mixed results (Salegio et al., 2013;
Hutson et al., 2015; Harris et al., 2012; Oh et al., 2014). Recently, studies using AAV1 as
tracing tool reveal anterograde trans-synaptic tagging down the axon (Castle et al., 2013;
AC
Castle et al., 2014). Combination of AAV with a conditional expression strategy (such as
two-step viral injection approach) enabled the researchers to map the output synapses of
distinct subpopulations of superior colliculus neurons that receive different corticocollicular
inputs and demonstrate their unique functional roles in driving different types of defense
behavior (flight and freezing). A recent study (Brian et al., 2017) shows that AAV1 and
AAV9 viruses are effective anterograde trans-synaptic tracers for mapping input-defined
functional neural pathways, and pseudotyping AAV1 with GFP -tagged CAG failed to show
any GFP + cell bodies in regions postsynaptic to the primary visual cortex. AAV provides an
adaptable platform that can be combined and customized easily with up-to-date tools,
promoters and recombinase-dependent expression control systems (Cai et al., 2013).
ACCEPTED MANUSCRIPT
Application of two-step viral injection method provides a direct means for examining the
potential role of downstream structures in mediating a function or behavior observed from
activation of a given upstream structure (Brian et al., 2017). Altogether, these results suggest
that AAV-based anterograde trans-synaptic tagging approach is the most successful model to
screen the functional circuits across multiple synaptic steps due to its ability to express high
levels of protein with lower levels of toxicity and restricted spread to first-order downstream
RI
PT
neurons.
A perfect comprehensive model for neuronal tracing and network mapping requires anteroand retrograde labeling, synaptic restriction, and controlled monosynaptic and polysynaptic
spreading across the targeted neuronal species. These viral-mediated tracing methods are less
SC
neurotoxic and provide accurate mapping of neuronal connections when compared with other
neuroscience methods such as electrophysiological recordings or conventional fluorescent
NU
tracers. Trans-synaptic spread of viral tracers can be controlled genetically and thus neural
circuitry of specific neuronal population can be studied in detail (Nassi et al., 2015). In
MA
contrast to classical polysynaptic viral tracers, improved neuronal tracing can be achieved by
optogenetically controlled transgenes, pseudotyping technology and other genetic targeting
approaches (Ginger et al., 2013), which may reveal not only the anatomy of neuronal
D
connections, but also provide for visualizing neuron connections that are functionally
4.4. Optogenetics
TE
relevant, i.e. activity-dependent neuronal tracing.
CE
P
The term ‘‘optogenetics’’ was coined first by Deisseroth et al. (2006) to describe genetically
targeted photoreceptor expression in neurons for their selective activation or inhibition with
light. Optogenetic experiments provide a platform to study the physiology of specific
AC
neuronal types and projections in normal and disease-related models. Optogenetics comprise
a combination of optical and genetic techniques to map the synaptic activation of specific
neurons within intact or living neural circuits (Gobbo et al., 2016). The precision of
optogenetics has provided major experimental leverage (Lin et al., 2011) and has provided
insights into neural circuit function and dysfunction.
The basis of optogenetic approach is to label the fast light-sensitive membrane channels at
synapses by delivering RNA and protein regulatory sequences into specific neurons using
viral vectors. These RNA and protein-targeting genes induce the expression of reporter
proteins, including optogenetic probes, at activated synapses. This results in selective
ACCEPTED MANUSCRIPT
mapping of previously active synapses and the reactivation of neurons particularly only at
these synaptic sites (Mahalaxmi, 2017).
Main components of optogenetic toolbox include
(i) Opsin, the functional unit of optogenetics which are defined as small membrane bound
proteins that integrate with minute organic retinal particles and functions as a light receptor.
Commonly,
opsins are of microbial origin or proteins like bacteriorhodopsins, the
RI
PT
halorhodopsins and the channel-rhodopsins (Deisseroth, 2015). Opsin variants provide both
unique advantages and individual limitations in controlling cellular activity or neuronal
signaling (Zhang et al., 2007, 2008 and 2011, Nagel et al., 2002 and 2003, Berndt et al., 2009
SC
and 2011, Gradinaru et al., 2010, Chow et al., 2010, Gunaydin et al., 2010, Mattis et al.,
2011).
NU
(ii) Light-delivery method such as fiber optics. The numerical aperture, diameter and mode of
the fiber have an impact on the spread of light. Optical fibers are most commonly used in
MA
light delivery method to deep neural somas or axon terminals. (Yizhar et al., 2011, Aravanis
et al., 2007, Atasoy et al., 2008, Kuhlman and Huang, 2008, Tsai et al., 2009).
(iii) Targeting strategies to specific cell somas or axons widely include viral vector to express
D
the specific opsin in target cells (Gradinaru et al., 2010, Atasoy et al., 2008, Kuhlman and
TE
Huang, 2008, Tsai et al., 2009). Viral vectors widely used in optogenetic methods include
lentivirus and AAV to induce the expression of opsin into the injection site. Herpes simplex
CE
P
virus or rabies virus are generally avoided due to their high toxicity (Mahalaxmi Mohan,
2017). Many kinds of viral transduction can restrict cell-body expression of opsin to the
AC
injection site and can be used in wild-type animals or Cre recombinase lines.
Genetically encoded optical sensor proteins permit the researchers to monitor several cellular
parameters like ion and metabolite levels, enzyme activities, and membrane voltage. Spatiotemporal precision in neuronal control has been attained using optogenetic actuators and
sensors (Packer et al., 2012). The two types of genetically encoded photosensitive proteins
used in optogenetic toolbox (Dugue et al., 2012; Miesenbock, 2009) include
actuators or reporters (which control the neuronal activity in a light-dependent
manner) such as channel-rhodopsin, halorhodopsin and archaerhodopsin.
ACCEPTED MANUSCRIPT
sensors or indicators (which display the neuronal activity) for calcium (Aequorin,
Cameleon, GCaMP), Chloride (Clomeleon) or membrane voltage (mermaid) (Packer
et al., 2012).
Genetically defined, cell-type-specific expression of optogenetic probes has been improved to
a level where neuronal circuits can be more meticulously untangled and the activity of
dispersed cell populations controlled or monitored in situ (Rost et al., 2017). Optogenetic
RI
PT
methods facilitate the selective activation or inhibition of individual spines that affects the
presynaptic terminal in the intact, alive animal, thus proving to be an essential tool to record
the behavioral changes in both healthy and diseased models (Packer et al., 2012). Most
optogenetic manipulations are localized with limited discrimination throughout the neurons
SC
(Rost et al., 2017). Recent developments have begun now to target and/or control specific
subcellular compartments using optogenetic actuators. The promising benefits of subcellular
NU
compartment-specific optogenetic manipulations include re-localization of organelles van
Wyk et al., 2015) and neurotransmitter receptors (Sinnen et al., 2017), control of spine size
MA
(Hayashi-Takagi et al., 2015), and light-activated filling of synaptic vesicles (Rost et al.,
2015). This approach is widely used tool to identify the mechanisms underlying pathogenetic
process in neoplastic and degenerative diseases. Both activity-dependent double-strand DNA
D
breakage in neurons (Suberbielle et al., 2013) and acceleration of amyloid-β (Aβ) plaques as
TE
the disease progresses (Yamamoto et al., 2012) in mouse model of Alzheimer’s disease (AD),
and altered striatal inhibition in mouse model of Huntington’s disease (Cepeda et al., 2013)
identified
Stimulation
of
using defined
CE
P
are
neurons
using
neuronal populations expressing microbial opsin genes.
optogenetic
methods
shows
regenerative
results
like
optogenetic activation of cones using halorhodopsin restored blindness due to neuronal loss
AC
in retinitis pigmentosa (Busskamp et al., 2010) and improved neuropsychiatric sleep
disorders (Adamantidis et al., 2007). Optogenetic stimulation of spiral ganglion in deaf mice
restored auditory activity (Hernandez et al., 2014).
As dendritic spines are the major anatomical structures of neural plasticity, optogenetics
enables the researchers to directly visualize the changes in the spines and synapses during
potentiation in an activity dependent manner. Optogenetics when used in combination with
protein targeting and RNA-targeting sequences has been shown to enrich the expression of
opsins at the synapses, especially in the dendritic spines . To facilitate the targeting of opsins
specifically at the postsynaptic sites in dendritic spines, combination of RNA targeting and
regulatory sequences with a short amino acid tag are used. These target a specific protein in
ACCEPTED MANUSCRIPT
the spine (Gobbo et al., 2016). Similar to the endogenous synaptic proteins, these optogenetic
reporter proteins also anchor and relocalize within the spines following local translation in an
activity-dependent manner. These synaptic membrane proteins are synthesized primarily in
the spine apparatus (Steward and Reeves, 1988; Spacek and Harris, 1997). These findings are
consistent with observation found with the Arc-regulatory sequence used in synapse
potentiation (Steward and Worley, 2001; Minatohara et al., 2015). These reports suggest that
the regulation of translation conferred by Arc UTRs is maintained when connected to a
RI
PT
different protein coding sequence like a membrane-tagged form of Cherry fluorescent
protein. Interestingly, ArcSYP-Ch (fast-spiking ChETA-Cherry22-MS2 engineered with the
entire Arc 5'- and 3'UTRs tagged with SYP) and Arc-Ch (ChETA-Cherry-MS2 engineered
SC
with Arc 5’ and 3’UTR) are specifically found in bigger spines and are expressed locally only
at stimulated spines (Gobbo et al., 2016). Using Arc RNA regulatory sequences, a Channel
NU
rhodopsin variant is expressed at synapses undergoing potentiation, thus providing a novel
tool to map and reactivate these sites selectively and specifically. Hayashi-Takagi et al.
(2015) developed a novel synaptic optoprobe, AS-PaRac1 (activated synapse targeting
MA
photoactivatable Rac1), that labels recently potentiated spines specifically in the motor cortex
during motor learning task. In vivo imaging revealed that light induces selective shrinkage of
AS-PaRac1-containing spines, which severely disrupts the motor learning process. Hence,
D
optogenetics is a versatile and novel tool to tag spines and synapses and to enhance the local
TE
expression of an opsin of the Channel rhodospin family (Gobbo et al., 2016).
CE
P
Major limitation in optogenetics include potential for toxicity (of viral vectors) at very high
expression levels or long-term expression (of opsins) in neurons. Improvement in modes of
light delivery is immediate requirement to promote the accuracy and efficiency of
AC
optogenetic tool. Induction of heat by light emitted from optogenetic fiber may be
detrimental to cell health, hence this is another limitation in optogenetic toolbox. To verify
the
exact
functioning
immunohistochemistry,
of
optogenetic
tools,
confirmation
using
electrophysiology,
behavioral assessments or other assays are crucial for data
interpretation (Tye and Deisseroth, 2011).
5. Molecular functions of dendritic spines
In addition to transmitting the synaptic signals from synapse to the neuronal cell body recent
detailed analyses of the structure, density and function of dendritic spines have reported its
ACCEPTED MANUSCRIPT
cardinal role on the complex regulation of different molecular signaling mechanisms upon
neuronal activation and synaptic plasticity. Apart from acting as a single organic unit the
dendrites are typically compartmentalized into many individual processing units called
“dendritic spines”.
Dendritic spines usually receive synaptic contact from a single presynaptic terminal,
including glutamatergic synapses. In 2002, Nimchinsky et al. made a striking discovery that
dendritic spines have the remarkable capacity of modulating the original synaptic signals.
RI
PT
Molecular understanding of dendritic spines has revealed that they serve as biochemical
compartments while the narrow neck provides an isolated compartment where the
biochemical signals are localized without disturbing the neighboring synapses along the
SC
parent dendrite. This compartmentalization confines the membrane trafficking to a localized
region. Philips et al. (2015), refined the earlier findings and indicated that the spine head
NU
serves as a localized biochemical compartment, while the spine neck is thought to function as
a diffusional barrier for intracellular organelles, signaling ions and signaling molecules.
Using electron microscopy, a good correlation of the spine head size with PSD size and
MA
receptor expression levels on synaptic stimulation are confirmed.
Existing evidences suggest that each dendritic spine functions as a partially autonomous
compartment with its own membrane-trafficking events which regulates the components in-
D
and out-side of the spine membrane. Typically, delicate molecular organization of the pre-
TE
and post-synaptic apparatus requires positional information to manifest the correct placement
of pre- and post-synaptic elements. Postsynaptic potentials could be electrically altered via
CE
P
structural variability in passive membrane properties and by active regulation via voltagegated ion channels within spines (Tsay and Yuste, 2004). The electrical properties and
biochemical compartmentalization of spines enables them to modulate the expression of
AC
receptors and activation of intracellular signaling pathways and second messengers, such as
calcium (Koch and Zador, 1993; Majewska et al., 2000; Matsuzaki et al., 2001; Noguchi et
al., 2005; Yuste, 2010).
Once the synapses have been established, the next important steps are to ensure the synaptic
specificity and to strengthen or weaken synaptic connections (Zhang and Benson, 2000)
depending upon the synaptic stimulation. Hence, in both synaptogenesis and synaptic
plasticity basic processes involved in learning and memory depend on the physiology of
dendritic spines. One common hypothesis regarding the dendritic spines is that either the
formation of new spines or a morphological change in existing spine occur during the
ACCEPTED MANUSCRIPT
learning process. Furthermore, when these changes are strengthened by new or repeated
synaptic connections, dendritic spines serve as the anatomical locus of memory storage.
Recently, many studies revealed that dendritic spines are the major site of synaptic plasticity
(Bosch and Hayashi, 2012; Bourne and Harris, 2008; Penzes and Rafalovich, 2012; Urbanska
et al., 2012). During both structural and synaptic plasticity, actin filaments and myosin in
dendritic spines undergo significant changes to aid the alterations depending on the synaptic
stimuli and strength. Apart from the modifications noted in dendritic spines morphology,
RI
PT
variations in its density are also shown clearly using various cellular models of synaptic
plasticity. During brain development an increase in the number of dendritic spines followed
by pruning appears to be dependent upon many factors (Urbanska et al., 2012). Furthermore,
SC
a recent study confirms that once the synaptic connectivity is established between the
neurons, dendritic spines turnover continues (Koleske, 2013). During the process of building
NU
new neural networks, both dendritic branches and new dendritic spines continue to develop,
while maturation and/or retraction of novel dendritic spines depends specifically upon the
stimuli in synaptogenesis. Some studies have report that synaptic plasticity mainly involves
MA
the dendritic spines in adolescent brain vs. the dendritic tree in the adult brain.
6. Density and dynamics of dendritic spines
D
In mammalian brain, most of the excitatory neurons consist of dendritic spines (Harris and
TE
Kater, 1994; Hering and Sheng, 2001). Generally, the number of dendritic spines depends
upon the age, cell type, position along the dendrite and also on the method of measurement
CE
P
applied. Most commonly, the pyramidal neurons of neocortex, medium spiny neurons of the
striatum and the Purkinje cells of the cerebellum consist of dendritic spines (Hering and
Sheng, 2001). Dendritic spine density shows a heterogenous distribution throughout the
AC
dendritic branch, and the spine number also varies among different cortical regions and
layers. Commonly, more than 100,000 spines can be found in the dendritic arbor of a single
neuron (Yuste, 2010). The numbers of dendritic spines range from 0.2 to 3.5 spines/µm of
dendrite in the human postmortem cortex (Benavides-Piccione et al., 2013). In adult
hippocampal CA1 pyramidal and granule cells, dendritic spines density ranges 2-4 spines/µm
of dendrite (Harris et al., 1992; Sorra et al., 1998) vs. 10-15 spines/µm in Purkinje cells
(Harris and Stevens, 1989; Harvey and Napper, 1988), indicating the independent regulation
of synaptic system among diverse parts of the dendritic tree. In both humans and macaque
monkeys, the neurons of frontal and orbitofrontal cortices show higher dendritic spine density
when compared with the neurons of the primary visual and somatosensory cortices (Elston,
ACCEPTED MANUSCRIPT
2003; Jacobs et al., 2001). Among 8 cortical areas studied in human brain, layer III basal
dendritic spine density is found to be higher in the prefrontal and orbitofrontal cortical areas,
and about 40% higher in the frontal polar cortex (area 10) than in the primary somatosensory
cortex (Jacobs et al., 2001). Further research analysis reveals that basal dendrites of
pyramidal neurons in layer-III of the cortex (Brodmann areas 10, 11, and 12) have 3 times
more dendritic spines when compared to the neurons of the primary visual cortex (area 17),
and 2 times more density than the parietal visual cortex (area 7a) of the macaque monkey
RI
PT
(Elston, 2003). Basal dendrites show a higher dendritic spine density in the neurons from the
layer-III cortex as compared to the prefrontal and orbitofrontal cortex of the adult human
brain (Elston, 2003; Nimchinsky et al., 2002; Oga et al., 2013). These results postulate that
SC
higher order cortical regions involved in higher grade of processing require more synaptic
connections and hence more dendritic spines.
NU
Turnover of dendritic spine number is a normal physiological process during brain
development. Although the majority of the dendritic spines are stable for longer periods of
time, a significant proportion of the dendritic spines transiently appear and disappear during
MA
synaptic activity driven by experience-dependent remodeling of specific neuronal circuits,
and these transient changes of the number of spines are called dendritic spine turnover (Knott
and Holtmaat, 2008). Advancements in imaging techniques like chronic two-photon imaging
D
have emerged as a powerful tool to study the spine motility in different cortical areas during
TE
various developmental stages (Grutzendler et al., 2002; Trachtenberg et al., 2002). In a
neuronal development model, researchers show that a postsynaptic neuron receiving a local
CE
P
synaptic input projects in to thin filopodia and then matures into a dendritic spine upon
continuous synaptic inputs. Those filopodia that do not receive inputs may be pruned (Bhatt
et al., 2009). Using many animal models, researchers have shown that
The density of dendritic spines is highly dynamic at early postnatal ages (Lendvai et al.,
AC
2000) spine turnover decreases with age and spine density reaches its peak level after
significant pruning during adolescent life (Zuo et al., 2005).
The total number of spines decrease to a relatively stable level over time due to similar
rates of spine formation and elimination in adult life (Hofer et al., 2009; Xu et al., 2009;
Zhang and Benson, 2000; Zuo et al., 2005).
Spine turnover and morphological changes in dendritic spines continue to occur in the
adult brain after the induction of experience-dependent plasticity.
ACCEPTED MANUSCRIPT
Spine turnover varies greatly during postnatal development (Xu et al., 2009), affecting
approximately 10-15% of spines per 24 h in very young mice. By using a transcranial twophoton imaging technique, Zuo et al. (2005) revealed that dendritic spines in layer 5
pyramidal neurons become progressively more stable over months with less pruning in barrel,
primary motor and frontal cortices of adult mice. In young adult mice (1-2 months old), the
percentage of stable spines ranges from approximately 55% in the somatosensory cortex
(Holtmaat et al., 2005; Trachtenberg et al., 2002) to 75% in the visual cortex (Grutzendler et
RI
PT
al., 2002; Majewska and Sur, 2003). In mature adult mice (4–5 months old), more than 70%
of the dendritic spines in the somatosensory cortex show increased stability and more than
90% of the dendritic spines in the visual cortex are persistent for more than 30 days
SC
(Holtmaat et al., 2005; Trachtenberg et al., 2002; Zuo et al., 2005). More than 70% of the
dendritic spines in the adult barrel cortex remain stable for more than 18 months. Another
NU
study declared that the majority of adult spines are stable over weeks, while 21.4% of the
total spines in adolescent mice have transient and persistent loss in few days (Holtmaat et al.,
2005). Experiments using repetitive confocal imaging and multiphoton microscopy over days
MA
to months in living mice have confirmed the long-held belief that spines and their synapses
can form and retract throughout adulthood (Grutzendler et al., 2002; Trachtenberg et al.,
2002). It is also thereby confirmed that the mature brain retains the capacity to form new
D
synapses and remodel its neural circuitry throughout life.
TE
The process and frequency of spine dynamics appears to vary in different brain regions and
also exhibits cell-type specificity (Holtmaat et al., 2005). In rat hippocampal slices, spine
CE
P
density reaches its maximum level at the third postnatal week, and then it remains relatively
static during the following weeks. However, in the mouse sensory cortex, spine numbers
show a steady decline from adolescence till the late adulthood. Dendritic spines in the mouse
AC
somatosensory cortex show shorter half-life (hours to days) in adulthood (Trachtenberg et al.,
2002); whereas in the mouse primary visual cortex, the number of dendritic spines is reported
to be dynamic during adolescent life and then, it becomes remarkably stable in adulthood
(Grutzendler et al., 2002). Studies have shown that the spine turnover is affected by sensory
activities, i.e. the synaptic networks (Holtmaat et al., 2006). Mice raised in total darkness
have shown fewer dendritic spines on the apical dendrites of pyramidal cells which strongly
suggests on the importance of afferent inputs (Valverde, 1967). Deafferentation of
hippocampal granule cells, following lesions of the entorhinal cortex, results in a decrease in
spine number - an effect that is reversed with reinnervation (Parnavelas et al., 1974).
ACCEPTED MANUSCRIPT
Studies of spine dynamics in higher order cortical brains have shown that during song
learning process, the increase in the rate of spine turnover is associated with an enhanced
capacity for song imitation in the forebrain nucleus HVC of zebra finches (Barretto and
Schnitzer, 2012). In another in vivo study, it is shown that fear conditioning increases the rate
of spine elimination, whereas fear extinction increases the rate of dendritic spines formation
in the layer 5 pyramidal neurons of the frontal cortex in living mice (Lai et al., 2012). Spine
dynamics is inhibited by the activation of glutamatergic receptors in hippocampal cultures,
RI
PT
while the addition of antagonists to NMDA receptors has no effect (Fischer et al., 2000),
indicating that the motility of dendritic spines is not linked with the expression of AMPA
and NMDA receptors or with the influx of calcium ions which would stabilize the actin
SC
filaments. Matsuzaki et al. (2004), reported that changes in the size and shape of dendritic
spines correlates with the individual synaptic activity during adulthood. It is reasonable to
NU
speculate that the turnover of spines and synapses varies with different frequencies in
different brain circuits. Altogether, these results show that the overall turnover of dendritic
spines accounts for the synaptic plasticity underlying the development of novel neural
MA
circuitry in accordance with the new learning process. Turnover of dendritic spines also
declines in the absence of significant inputs, thus balance between formation and pruning of
dendritic spines is important in stabilizing the crucial synaptic connections in the neural
D
circuitry. Researchers have proven that spine dynamics also varies between the normal and
TE
diseased brains. Thus, detailed investigation on the turnover of dendritic spines in synaptic
regulation and molecular mechanisms behind their structural changes is crucial to understand
CE
P
the physiology of brain function and dysfunction. These findings may pave the way for the
identification of therapeutic targets for various neurological disorders.
AC
7. Plasticity of dendritic spines
During brain development from the embryo till adulthood, plasticity of dendritic spines is the
most striking crucial phenomenon in the formation of new neuronal circuits. Many
neuroscientists have proven the morphological diversity of spine dynamics like spine
motility, changes in their shape and size occur depending upon the strength of the synaptic
stimuli using advanced imaging technologies (Calabrese et al., 2006). These imaging results
suggest that dendritic spines are highly dynamic structures that characterize the fundamental
morphological changes of synaptic plasticity. The plasticity of dendritic spines can be
categorized broadly into short-term and long-term plasticity.
ACCEPTED MANUSCRIPT
Short term plasticity (phase of formation and development of new spines) mainly
includes the emergence of nascent spines which occurs within 30 minutes of LTP
induction in dissociated neuronal slice cultures (Segal, 2005). A glutamate uncaging
study shows that new spines protrude from the dendritic shaft of layer 2–3 pyramidal
neurons in mouse brain slices and within 2 minutes they behave as mature spines (Sala
and Segal, 2014). On the contrary, another in vivo study using time-lapse imaging
reported that nascent spines in the neocortex require a minimum period of 4 days to
RI
PT
become mature (Knott et al., 2006). These nascent spines are often very long and have
frequent filopodia-like shapes. Generally, nascent spines may take few hours to a single
day to become structurally and functionally mature.
Long-term plasticity (phase of maturation and stabilization of newly formed spines)
SC
mainly includes an increase in dendritic spine density and spine head size along with a
NU
decrease in their overall mean length, the number of dendritic filopodia, and spine
motility (Dunaevsky et al., 1999; Nimchinsky et al., 2001). The initial expansion of
spines mainly depends on the activation of NMDA and glutamate receptors in the spine
MA
head, while a long-lasting expansion depends upon the activation of kinases (Yang et al.,
2008). Approximately half of the newly-formed spines in the neurons of mouse barrel
D
cortex become stable (Trachtenberg et al., 2002) and mature after their emergence (Fu et
al., 2012). Time taken for the maturation of nascent spines ranges from several hours to
TE
almost 1 or more days depending on strength of the stimulus and functional requirements
(Holtmaat and Svoboda, 2009).
CE
P
An in vivo study reported that following extensive motor learning, alterations in spine
turnover for long term result in a rearrangement of synaptic connectivity in the mouse
neocortex (Holtmaat et al., 2005) and these changes may last for many days after the training
AC
(Dayan and Cohen, 2011; Sala and Segal, 2014). The maturation process continues until the
dendritic spines make synaptic contact and form synapses that conduct the transmitting
signals. However, the generation or elimination of spines depends upon the nature of the
stimuli (weak or strong), functional requirements (learning new tasks or fear conditioning)
and local environment (stress or enrichment). Both the morphology and density of dendritic
spines vary in response to various other factors like the environmental enrichment,
pharmacological therapy, hormonal changes, and learning skills (Fiala et al., 2002; Yuste and
Bonhoeffer, 2001). In both physiological and pathological conditions, these activity-
ACCEPTED MANUSCRIPT
dependent structural changes of dendritic spines are noted in the primary sensory, motor and
frontal cortical areas (Rochefort and Konnerth, 2012).
7.1. Dendritic spines morphological changes in response to LTP and LTD induction
In the neural circuitry, multiple mechanisms like the formation of new synapses, the
destabilization of existing synapse, and turnover of dendritic spines underlie the experiencedependent plasticity processes (Bourne and Harris, 2012; Caroni et al., 2012; Hill and Zito,
RI
PT
2013; Holtmaat and Svoboda, 2009; Kasai et al., 2010). Such a synaptic plasticity is believed
to be the key fundamental process for learning and memory (Kandel and Schwartz, 1982). In
the experimental setup, synaptic plasticity is demonstrated in terms of LTP and LTD
SC
induction (using stimuli of different frequency) by electrophysiological means which could
be otherwise defined as long-lasting enhancement or the reduction of synaptic transmission,
NU
respectively (Bliss and Lomo, 1973; Bliss and Collingridge, 1993; Malenka and Bear, 2004;
Malenka and Nicoll, 1999) (Fig. 7).
Excitatory synapses contain AMPA and NMDA receptors localized on dendritic spines with
MA
basal synaptic transmission largely mediated via AMPA. Intense synaptic stimulation opens
NMDA receptors leading to long-lasting changes in postsynaptic AMPA receptor number
and LTP (Bliss and Lomo, 1973), thereby resulting in the growth of dendritic spines. During
D
the downstream signaling cascade of LTP-induced NMDA receptor activation, the binding of
TE
CaMKII to NMDA receptor GluN2B subunit is important for the induction and maintenance
of LTP (Barria and Malinow, 2005; Lisman et al., 2002; Sanhueza et al., 2011; Zhou et al.,
CE
P
2007) and new spine outgrowth (Hamilton et al., 2012; Hill and Zito, 2013).
In dissociated hippocampal neurons, the activation of NMDA receptors via chemicallyinduced LTP (Engert and Bonhoeffer, 1999; Maletic-Savatic et al., 1999) increases the
AC
connectivity of specific neurons in 4 different ways:
Rapid enlargement of spine heads (Kopec et al., 2006; Lang et al., 2004; Park et al.,
2006) that precedes the increase in AMPA receptor abundance (Kopec et al., 2006).
Increase in neck width and decrease in length of spines (Fifková and Anderson,
1981).
Rapid formation of new spines (Lin et al., 2004; Nägerl et al., 2004; Park et al., 2006).
Stabilization of newly-formed spines (Harvey and Svoboda, 2007; Matsuzaki et al.,
2004; Tanaka et al., 2008).
ACCEPTED MANUSCRIPT
After the induction of LTP, enlargement of spine head size is mainly due to the appearance of
more perforated and complex PSD (Popov et al., 2004) followed by the accumulation of
more number of multiple presynaptic buttons, more AMPA and NMDA receptors (Kopec et
al., 2006) and F-actin levels (Kramár et al., 2006; Lin et al., 2005) along with a significant
increase in the number of recycling endosomes, coated vesicles (Harris et al., 1992) and
amorphous vesicular clumps (Park et al., 2006), polyribosomes (Ostroff et al., 2002), and
mitochondria (Li et al., 2004). These functional changes in spine structure account for
development of new learning and memory processes.
RI
PT
enhanced synaptic transmission and remodeling of synaptic connections underlying the
In vivo studies conducted to monitor the individual spines over weeks to months found that
SC
smaller spines are more labile (Trachtenberg et al., 2002). On the other hand, the larger
spines are more functionally active, but show less plasticity and sometimes remain
NU
morphologically resilient for the animal’s entire adult life (Grutzendler et al., 2002). Many
studies have shown that morphological differences contribute to a spine’s response to LTPinducing stimuli via diffusional properties, whereas other groups have found neck diffusion
MA
to be greatly influenced by pre- and post-synaptic activity and depolarization (Bloodgood and
Sabatini, 2005; Grunditz et al., 2008). Although the relationship between spine morphology
and activity-induced plasticity is still under debate, generally, the larger, more stable,
D
mushroom and stubby spines are designated as “memory spines” (Trachtenberg et al., 2002)
TE
and thin dynamic spines are defined as “learning spines” due to their enhanced ability to
2002).
CE
P
undergo structural changes (Holtmaat et al., 2005; Kasai et al., 2003b; Trachtenberg et al.,
Hebbian LTP and LTD are well-studied forms of synaptic plasticity that form the cellular
basis of hippocampal-dependent learning and memory. In synaptic studies, Hebbian plasticity
AC
shows direct correlation with the volumetric changes in the corresponding dendritic spines
(Govindarajan et al., 2011; Harvey and Svoboda, 2007; Matsuzaki et al., 2004).
The induction of LTP either by tetanic stimulation or by theta-burst stimulation induces a
marked increase in spine densities (Muller et al., 2000) and formation of new, mature and
probably active synapses (Toni et al., 1999). But, LTD induced by low-frequency stimulation
causes a distinct reduction in both spine densities (Bastrikova et al., 2008; Monfils and
Teskey, 2004) and spine size within the CA1 area of the hippocampus, which in turn
contributes to the activity-dependent elimination of synaptic connections (Zhou et al., 2004).
Thus, dendritic spines can undergo bidirectional morphological changes in response to
neuronal activity (Nägerl et al., 2004). Two-photon imaging shows that LTP induction using
ACCEPTED MANUSCRIPT
high-frequency stimuli induce both an increase and a decrease in spine densities (Engert and
Bonhoeffer, 1999; Nägerl et al., 2004), indicating a strong correlation between the synaptic
stimulation and alterations in dendritic spines. Studies have shown that LTP increases the
turnover of protrusions and formation of new spines that preferentially grow in close
proximity to
activated
synapses and
become functional (Bastrikova
et al., 2008).
Controversial report has also been reported, as induction of LTP at one spine, reduces the
threshold stimuli for neighboring spines, sometimes causing retraction of these spines
RI
PT
(Harvey and Svoboda, 2007). Thus, LTP-induced activity can induce long-lasting changes in
the hippocampal pyramidal network.
LTP consists of 2 distinct phases:
Early LTP phase is described as the “short-lasting phase” (about 1 hour) which includes
SC
the activation or rapid insertion of post synaptic (AMPA-glutamate receptors) into
NU
existing synapses and post-translational modification of synaptic proteins. This phase is
also known as LTP induction phase (Malenka, 2003; Malenka and Nicoll, 1999; Malinow
and Malenka, 2002; Malinow and Tsien, 1990; Stevens and Wang, 1994).
Late LTP phase is described as the “long-lasting phase” of LTP which includes
MA
translational and transcriptional changes involved in protein synthesis; and structural
D
changes (C H Bailey and Kandel, 1993; Yuste and Bonhoeffer, 2001), including
synaptogenesis (Bourne and Harris, 2012). This phase is also known as the LTP
TE
maintenance phase (Bolshakov et al., 1997; Bolshakov and Siegelbaum, 1995; Reymann
and Frey, 2007; Toni et al., 1999; TROMMALD, 1990; Voronin et al., 1995).
CE
P
When LTP is inhibited by post-synaptic inhibition via exocytosis, protein synthesis or protein
kinase A (important signaling factors involved in LTP induction phase), an increase in
dendritic spine volume can still occur; however, the process become unstable and they
AC
collapse immediately (Yang et al., 2008). Finally, with regard to the spine expansion, LTP
induction phase requires different signaling pathways, while maintenance phase may require
similar molecular pathways. Common signaling pathways involved in the induction and
maintenance phase of LTP associated with the increase in dendritic spines are still unclear.
On the other hand, low frequency stimulation of the synapse activates NMDA receptors to
produce LTD (Lynch et al., 1977) which causes the removal of postsynaptic AMPA receptors
and loss of dendritic spines (Lee et al., 2002; Lüscher et al., 1999; Man et al., 2000; Nägerl et
al., 2004; Snyder et al., 2001; Xiao et al., 2001; Zhou et al., 2004).
The main changes observed after LTD induction are
ACCEPTED MANUSCRIPT
steady reduction in the size of dendritic spines (Nägerl et al., 2004; Okamoto et al.,
2004)
accelerated shrinkage of spine head (Zhou et al., 2004).
few spines and/or filopodia retracted (Zhou et al., 2004).
In glutamate uncaging model, live imaging of dendritic spines following LTD induction (Oh
et al., 2013) shows spine shrinkage mainly on the stimulated spine, but not neighboring
spines. During LTD induction, size reduction is noted in both large and small spines, but the
RI
PT
mechanisms by which they occur are different. The retraction of small spines depends on
NMDA receptor, while that of large spines requires both NMDA and metabotropic glutamate
receptors. Both LTD and spine shrinkage require the activation of NMDA receptors and
SC
calcineurin (Zhou et al., 2004), while the electrophysiological changes are mediated through
PP1/2A signaling; and spine remodeling is mediated by the activation of cofilin (a family of
binding
proteins),
indicating
the
involvement of various downstream signaling
NU
actin
pathways in dendritic spine shrinkage and LTD. These results provide clear evidence for the
stable alteration of dendritic spines in neural circuits during experience-dependent plasticity
MA
(Hill and Zito, 2013). Several studies have shown that NMDA receptor hypofunction leads to
reduction in spine numbers (Brigman et al., 2010; Ramsey et al., 2011; Roberts et al., 2009;
TE
D
Ultanir et al., 2007).
7.1.1. Homeostatic plasticity
CE
P
Homeostatic synaptic plasticity (HSP) is generally defined as a negative feedback mechanism
through which the density of dendritic spines and population of synapses are maintained
within optimum range globally in brain and help stabilize its neuronal activity (Turrigiano et
al., 1998). In contrast to Hebbian plasticity which exerts input specificity (Barrionuevo and
AC
Brown, 1983) and fundamentally promotes the number of dendritic spines and increases the
synaptic strength, HSP is expressed over a wide range of synapses and induces a non-specific
decrease in global activity that proportionally scales all synapses of a given neuron (up or
down). Altogether, HSP maintains the relative strengths of all synaptic inputs into the neuron,
and makes the neuronal network function optimally (Turrigiano and Nelson 2000 and 2004,
Turrigiano, 2011). Both in vivo and in vitro studies have shown the presence of HSP,
supporting its crucial role in stabilizing the neural network activity and optimizing its
functions (Desai et al., 2002). During HSP, modulations in AMPA receptor trafficking to and
from the post-synaptic density, and coherent presynaptic changes that result in altered
ACCEPTED MANUSCRIPT
neurotransmitter release leads to appropriate modulations in synaptic strength (Murthy et al.,
2001; O’Brien et al., 1998). In the dendritic spines of mouse hippocampal neurons, induction
of HSP via activity blockade shows a generalized increase in the size of dendritic spines and
it prolongs LTP induction in dendritic spines with a concomitant increase in growth rate and
a reduction in plasticity threshold. Computational studies have postulated that HSP can
influence previous Hebbian events at a synapse (Rabinowitch and Segev, 2008). HSP
enhances the magnitude of synaptic potentiation by promoting the growth of unstimulated
RI
PT
neighboring small spines and also by enhancing the structural plasticity at clustered inputs
following Hebbian plasticity (activation of one input) within the dendrite, at single synapses.
HSP enhances the synaptic activity in nearby group of synapses in an activity-dependent
SC
manner within a distance of 5 µm of the stimulated synapse, over which plasticity related
proteins can spread across the co-active synapses (Harvey et al., 2008). These results suggest
NU
that induction of Hebbian plasticity on a background of HSP results in compromised input
specificity, as neighboring small spines grow and thus, modulate the neuron’s response to
activity by promoting the sensitivity of spines to smaller stimulus. In addition, several studies
MA
indicate that HSP also induces local homeostatic changes within dendrites (Branco et al.,
2008; Ju et al., 2004; Liu, 2004; Sutton et al., 2006). Many electrophysiological studies show
that synaptic threshold modulation after HSP is found in small spines than large spines, and
D
in a reversible manner, suggest that synaptic scaling mechanisms are separable from
TE
plasticity mechanisms. Thus, HSP drives the local expression of Hebbian plasticity and
reduces input specificity. Few reports (Toyoizumi et al., 2014, Keck T, et al. 2017) show that
CE
P
stability of neuronal responses depends on the operation of slow HSP and fast Hebbian
plasticity at different sites. Together, these results highlight the important role of HSP in
modulating synaptic efficacy and promote learning without compromising previous memory
AC
(Stryker, 2017).
7.2. Spine remodeling during process of learning and memory
Alterations in the structure and/or number of synapses in neural circuitry are important
substrates of memory formation following learning. Rapid remodeling of spines is the main
focus in learning-related synaptic plasticity in the matured brain. Generalized hypothesis
regarding the role of dendritic spines in learning and memory have shown a periodic change
from the previous notion that density of dendritic spines has a correlation with learning to the
idea that morphological alterations within dendritic spines show a stronger and direct
ACCEPTED MANUSCRIPT
association with new learning and memory formation than spine turnover (van der Zee,
2015). A two-photon microscopy study shows that motor learning tasks rapidly induce the
formation of new spines in the pyramidal neurons of motor cortex in mice. Most of the newly
formed spines persist for weeks and months following training. Interestingly, the motor task
performance positively correlates to the amount of new spines formed (Xu et al., 2009).
Similarly, mice trained for novel forelimb skills show the development of new dendritic
spines in clusters in layer V of pyramidal neurons in the motor cortex, during the acquisition
RI
PT
phase of learning. The clustered new spines possess a bigger head size and persist even after
the training had stopped when compared to the non-clustered new spines (Fu et al., 2012).
Using enhanced YFP-expressing mice, a spine dynamic study on the primary motor cortex
SC
following a motor learning (accelerated rotarod running) task or in the barrel cortex following
sensory experiences (environmental enrichment) reports experience-induced de novo spine
NU
formation within two days of training (Yang et al., 2009). Some of these newly formed spines
persist for months by subsequent exposure to novel experiences. The degree of dendritic
spine formation shows linear correlation with the degree of learning acquisition (Xu et al.,
MA
2009) along with the prolonged maintenance of the acquired skill (Yang et al., 2009). The
results of the above mentioned studies suggest that repetitive tasks induce formation of new
de novo dendritic spines during learning, thereby providing a morphological basis for spatial
D
coding of motor memory storage by creating new synaptic connections in neural circuitry of
TE
the mammalian brain (Stepanyants and Chklovskii, 2005). Moreover, rapid spinogenesis
during the initial learning is followed by significant pruning of dendritic spines, making the
CE
P
total spine number return to the basal levels after prolonged training (Yu and Zuo, 2011).
Pruning occurs selectively for dendritic spines that have existed before training, while new
learning-induced dendritic spines are stabilized during subsequent training and endure long
AC
enough even after the training has ended (Xu et al., 2009).
Few researchers postulate that dendritic spines with small diameters and/or lengths are
unstable, easily eliminated, and involved in memory acquisition, while dendritic spines with
bigger head diameters are more stable and involved in consolidation processes and long-term
memory formation (Kasai et al., 2010, 2003b). Recent research analysis reveals that the
formation of new memories in a conditioning paradigm shows a significant increase in
dendritic spine density in CA1 pyramidal neurons in both adult male (Jedlicka et al., 2008;
Leuner et al., 2003) and female (Beltrán-Campos et al., 2011) rats. Rats exposed to a spatially
challenging environment, like Morris water maze, show an increased spine density on CA1
pyramidal neurons, and also display faster and precise learning ability (Moser et al., 1994). In
ACCEPTED MANUSCRIPT
novel object recognition and object placement tasks, trained rats astoundingly discriminate
newer objects, and show higher spine density in the CA1 region of hippocampus and layer
II/III of medial prefrontal cortex (Eilam-Stock et al., 2012). Altogether, these results suggest
that an increase in spine density enhances the number of synapses per neuron resulting in a
strong neural connectivity, which can be attributed to the mechanisms underlying memory
consolidation and retrieval process.
An electron microscopy study (O’Malley et al., 1998) confirms an increase in the number of
RI
PT
dendritic spines in the rat dentate gyrus after the training of passive avoidance learning task.
Another confocal microscopy study (Moser et al., 1994; Moser and Moser, 1998) reports
similar findings like an increase in the total dendritic spine number per unit length in the
SC
basal dendrites of CA1 pyramidal cells, following the spatial training of rats in a complex
environment. In the dentate middle molecular layer of rats trained in the passive avoidance
NU
task, a significant and transient increase in dendritic spine density is observed after 6 hours of
training session, but dendritic spine density is reported to return to basal levels after 72 hours
of training (O’Malley et al., 1998). Similarly, a significant time-dependent increase in the
MA
number of spines is noted in the dentate gyrus of the adult rat, following water maze training
(O’Malley et al., 2000). Transient changes in spine density are suggested to increase the
synapse turnover rate, which concurrently induce changes in the connectivity pattern of the
D
neuronal network in the long-term memory consolidation process.
TE
During synaptic stimulation, thin spines that are mostly responsive to the stimuli (either low
or high in strength) are labeled as ‘learning spines’, while mushroom spines that are stable for
CE
P
longer periods are labeled as ‘memory spines’ (Sala et al., 2008). Many researchers show that
thick spines remain stable for a month, while thin spines are transient (Geinisman, 2000;
Holtmaat et al., 2005). In response to LTP and environmental enrichment, branched and
AC
mushroom spines show greater changes vs. modest alterations seen in thin and stubby spines
(Desmond and Levy, 1986; Geinisman et al., 1989; Johansson and Belichenko, 2002;
Trommald et al., 1996). These results indicate that thick, mushroom and branched spines may
be responsible for the development and maintenance of long-term memory.
In both pathological studies and live time-lapse imaging of dendrites in animal models, the
induction of LTP causes a rapid increase in dendritic spine number or expansion of spine size
(Desmond and Levy, 1988; Engert and Bonhoeffer, 1999; Lang et al., 2004; Maletic-Savatic
et al., 1999; Matsuzaki et al., 2004; Trommald et al., 1996). Additionally, abnormalities in
dendritic spines and synapses have been frequently found in different syndromic and nonsyndromic forms of mental retardation in people, suggesting again on the role of spines in
ACCEPTED MANUSCRIPT
cognitive function and learning (Armstrong et al., 1995; Hinton et al., 1991; Huttenlocher,
1974; Kaufmann and Moser, 2000; Marin-Padilla, 1976; Purpura, 1974; Takashima et al.,
1981).
7.3. Spine remodeling under conditions of excitotoxicity
Recently, many researchers have reported that dendritic spines undergo changes during
excitotoxic conditions induced by either high intracellular glutamate or calcium levels. In
RI
PT
cultured hippocampal neurons, application of glutamate shows a dual effect (elongation to
shrinkage) on the same group of dendritic spines by activating different cascades of
postsynaptic second messengers, depending on the magnitude and/or duration of the
2+
i[Ca ]
SC
intracellular calcium2+ i[Ca2+] influx produced by it. Mostly, glutamate-induced
increase in dendritic spines occurs via activation of NMDA receptors (Korkotian and Segal,
NU
1999). During excitotoxicity, a rapid and marked loss of dendritic spines followed by loss of
actin filaments is reported (Segal et al., 2000). Prolonged application of glutamate to the
hippocampal neurons (over 2 hours) induces an NMDA receptor-dependent decrease in
MA
dendritic spine length by about 30% (Segal, 1995), while short applications induce a slow
onset increase in spine length (Korkotian and Segal, 1999).
In cultured pyramidal neurons, flash photolysis of caged glutamate shows that larger spines
D
are affected more by glutamate application than small spines, suggesting that large spines are
TE
the basic substrates of functional synaptic connections, while small thin ones are only on their
way to mature and become functional (Kasai et al., 2003b). Cultured hippocampal neurons
increase
in
CE
P
exposed to DL-2-amino-7-phosphonovalerate, a NMDA receptor antagonist, show an
AMPA
receptor-mediated
excitatory
postsynaptic
currents
and
higher
synchronization among neuronal cells. An enhancement in synchronicity is followed by the
AC
formation of novel dendritic spines which will become functional when they are innervated
by active presynaptic terminals and by pruning of existing spines (Goldin et al., 2001).
Hippocampus of neonatal rats exposed to monosodium glutamate shows both neuronal loss
and morphological changes in the surviving pyramidal CA1 cells. The number of dentate
granule cells and CA1 pyramidal neurons in glutamate-treated rats are 11.5% less than those
counted in sodium chloride-treated control rats. Moreover, after a glutamate treatment, the
dentate granule cells show more dendrites and branched spines, along with a greater number
of thin and mushroom spines. These results strongly suggest that the neuronal loss and
cytoarchitectural modifications in the surviving neurons can lead to functional alterations in
ACCEPTED MANUSCRIPT
the hippocampal integrative activity due to an early excitotoxicity (González-Burgos et al.,
2009).
7.4. Dendritic spines remodeling under conditions of enriched environment
The environment plays a major role in synaptic plasticity, and density of dendritic spines is
believed to be a sensitive measure of environmental influence, more than dendritic branching
(Bryan and Riesen, 1989). Animals exposed to an enriched environment (reared in larger
RI
PT
cages with toys, tunnels, and obstacles, and in groups to allow for ample opportunities for
problem-solving and complex social interactions) show greater number of dendritic spines in
hippocampus (Kozorovitskiy et al., 2005) and improved performance of several learning
SC
tasks (Bruel-Jungerman et al., 2005). In addition to aerobic exercise and healthy diet,
enriched environments can have a promising effect on the recovery of damaged dendritic
NU
spines and/or dendritic spine loss in diseased brains. Experimental rats reared in enriched
environments show increased formation of new neurons along with increased synaptic and
neuronal plasticity. This is reflected in stimulation of new dendritic branching and synapse
MA
formation, along with increase in the genetic expression of neurotrophic factors (BDNF,
NGF, GDNF, and other growth factors). Altogether, these changes in neural circuitry lead to
enhanced
learning
and
memory.
These
results suggest that exposure
to
enriched
D
environments can be one of the potential ways to optimize the synaptic plasticity, and can
TE
delay progression of several neurological disorders (Fratiglioni et al., 2004; Mora et al., 2012;
Pham et al., 2002). In a behavioral study, rats housed in enriched environments show
CE
P
increased expression of brain-derived neurotrophic factor gene in comparison to rats housed
in isolation (Falkenberg et al., 1992). Using 3-D confocal laser scanning microscopy, intact
rats exposed to an enriched environment reveal a marked increase in the number of dendritic
AC
spines in the layers II/III and V/VI pyramidal neurons of the somatosensory cortex vs. rats
housed in standard environment. On the contrary, spontaneously hypertensive rats with an
infarct brain (induced by occlusion of middle cerebral artery distal to the striatal branches)
housed in an enriched environment for 3 weeks show a significantly high number of dendritic
spines in layers II/III pyramidal neurons (in the contralateral hemisphere) when compared to
the rats in infarct group housed under standard environment (Johansson and Belichenko,
2002). Major morphological changes observed in dendritic spines after an exposure to
environmental enrichment include an increase in dendritic branching of cortical and
hippocampal pyramidal neurons, spine density, number of synapses, polyribosomes, the
number of perforated PSD and the mean PSD length (Altschuler, 1979; Diamond et al., 1975;
ACCEPTED MANUSCRIPT
Globus et al., 1973; Greenough, 1975; Greenough et al., 1985; Kozorovitskiy et al., 2005,
2005; Leggio et al., 2005; Rampon et al., 2000; Schapiro and Vukovich, 1970; Volkmar and
Greenough, 1972). A recent primate study confirms that environmental enrichment enhances
dendritic branching and spine density in the pyramidal neurons of the CA1 region of the
hippocampus and prefrontal cortex (Kozorovitskiy et al., 2005). Many studies have proven
that environmental enrichment prevents the age-related loss of dendritic spines in rats (Saito
et al., 1994) and fasten the recovery on both morphological and behavioral basis in mouse
RI
PT
models of fragile X syndrome (FXS) (Restivo et al., 2005). Additionally, environmental
enrichment reduces the Aβ-plaque deposition in an animal model of AD by increasing the Aβ
degrading endopeptidase activity (Lazarov et al., 2005). Animals reared in rich environment
SC
after trimming the whiskers show higher functional plasticity in the barrel cortex (Rema et
al., 2006). Altogether, these studies suggest that enriched environment plays an important
higher
cognitive skills,
learning and
NU
role in enhancing the optimal synaptic connections of neural circuits, which in turn promotes
memory in healthy individuals.
Also,
enriched
environment positively regulates the neural circuity to combat the spine or neuronal loss in
MA
diseased brains and delays the progression of neurodegenerative diseases.
7.5. Spine changes associated with aging
dendritic
spines and synapses during aging. In some individuals, aberrant
TE
circuitry,
D
Even though aging is a normal physiological process, several changes are observed in neural
alterations in the dendritic spine number, morphology and synapses results in impaired
CE
P
cognition, working memory and behavior. Although neurons, dendrites and dendritic spines
undergo significant attrition during aging, changes in dendritic spine pathologies vary in
different brain regions (Dickstein et al., 2013). In cerebellar Purkinje cells, about a 17%
AC
decrease in dendritic spine density is reported in 26-month vs. 6-month old rats (Rogers et al.,
1984). In rat amygdala, an age-related increase of dendrite length and unchanged spine
densities are discovered (Braidy et al., 2011; Marcuzzo et al., 2007; Rubinow et al., 2009). In
1980, Cupp and Uemura et al., revealed marked reduction in the dendritic branch order,
number of branches, total dendritic length, and spine density in the apical and basal dendritic
spines of layers III and IV of pyramidal cells in the prefrontal cortex of rhesus monkeys
(aged 27–28 years). Aged primates (27–32 years of age) showed a loss of apical dendritic
tufts of pyramidal cells in layer I of area 46 when compared with young monkeys (6–9 years
of age) (Peters et al., 1998). An age-related, significant and consistent decline in the total
number of dendritic spines and spine density are observed on both the apical and basal
ACCEPTED MANUSCRIPT
dendrites of layer 3 pyramidal neurons of prefrontal area 46 in old vs. young macaque
monkeys. Only the second branch order of the apical dendrites shows specific age-related
reduction in dendritic length and segment numbers (Duan et al., 2003). A recent study
confirms that the pyramidal neurons of aged Rhesus monkeys show 33% of dendritic spine
loss and almost 50% of thin spines loss in the prefrontal cortex region along with impaired
performance on the Delayed Non-matching to Sample task (Dumitriu et al., 2010).
Neurologically intact aged subjects (above 50 years) show decrease in total dendritic length
RI
PT
(9-11%) and decrease in the numbers of spines (approximately 50%) in layer III of pyramidal
cells in prefrontal and occipital areas when compared with young aged subjects (below or
equal to 50 years) (Jacobs et al., 1997). In the substantia nigra of human, severe spine loss is
SC
observed in aged (70–91 years) compared to middle-aged subjects (Cruz-Sánchez et al.,
1995). These data strongly support aging-related loss of dendritic spine density and turnover
MA
cognitive functions like learning and memory.
NU
results in reduced synaptic connectivity, which in turn causes significant decline in higher
8. Dendritic spines status in acute and chronic neurological disorders
Both post-mortem studies of patients and transgenic animal models have shown a strong link
D
between the aberrant dendritic spine structure and number, altered synaptic plasticity in
TE
critical regions of the brain in neuropsychiatric, and neurodegenerative disorders (Table. 1).
Dysmorphogenesis of dendritic spines can lead to defective or excessive synapse function
CE
P
and connectivity, which in turn disrupts the neural circuitry. Many researchers have
successfully proven the direct correlation of dendritic spine anomalies and/or dendritic spine
loss with the impaired synaptic connectivity and plasticity that results in altered mental status
AC
and impaired cognitive, motor and learning skills (Kasai et al., 2010; Penzes et al., 2011). In
addition, the post-mortem reports also confirm the presence of malformation or loss of
dendritic spines in patients with epilepsy, stroke, schizophrenia, dementia, major depression
and chronic substance abuse (Fiala et al., 2002; Glantz and Lewis, 2001; Nimchinsky et al.,
2002; Swann et al., 2000).
8.1. Alzheimer’s disease
AD is an age-related neurodegenerative disease and the common leading cause of dementia
and death in the elderly (Braidy et al., 2011; Jack et al., 2011). Main symptomatic features of
AD are memory deficits, dementia, and a decline of cognitive and intellectual performances
ACCEPTED MANUSCRIPT
(Kelley and Petersen, 2007). Cardinal neuropathology of AD includes accumulation of
Aβ protein, which is called senile plaque extracellularly and phosphorylated tau, which is
called neurofibrillary tangles intracellularly (Holtzman et al., 2011; Selkoe and Schenk,
2003). Several studies suggest that synaptic loss and dendritic spine dysfunctions can be
significant preceding and contributing factors to neuronal loss and degenerative pathology in
AD (Masliah et al., 1991; Scheff et al., 2006; Terry et al., 1991). Early postmortem brain
samples of many AD patients persistently show marked synapse and dendritic spine loss
RI
PT
(DeKosky and Scheff, 1990; Knobloch and Mansuy, 2008; Walsh and Selkoe, 2004) and
other dendritic changes (DeKosky and Scheff, 1990; Selkoe, 2002) in the hippocampus and
throughout the cortex, when compared with age-matched control brains. The loss of dendritic
SC
spines and synapse in the hippocampus and throughout the cortex, the principal areas that are
commonly affected by AD-related pathology (DeKosky and Scheff, 1990; Walsh and Selkoe,
NU
2004), give a morphological evidence for the lower mental status (DeKosky and Scheff,
1990; Selkoe, 1989).
A recent study reports approximately 45% dendritic spine loss in both the neocortex and
MA
hippocampus of AD patients vs. cognitively normal controls (Serrano-Pozo et al., 2011).
Another in vivo study reports about a 50% dendritic spine loss near the amyloid plaques
region in the living animal model of AD using multiphoton microscopy (Spires et al., 2005).
D
Primary hippocampal neurons challenged with Aβ oligomers show dendritic spine loss along
TE
with an increase in premature dendritic spines and varicosities formation (Attar et al., 2012;
Maiti et al., 2011, 2010). Tg2576 mice (a widely used AD mouse model that expresses
CE
P
mutant human amyloid precursor protein (APP)) show lower spine density in both CA1 and
dentate gyrus (Jacobsen et al., 2006; Lanz et al., 2003). Notable cognitive decline is found in
these mutant Tg2576 mice when the dendritic spines are depleted. An over-expressed human
AC
APP animal model of AD shows significant spine loss and neurite dystrophy around amyloid
plaques, which alters the neuronal circuits and results in cognitive decline (Garcia-Alloza et
al., 2006; Spires et al., 2005; Spires-Jones et al., 2007; Tsai et al., 2004).
In AD, excessive accumulation of soluble Aβ leads to synaptic depression and removal of
AMPA receptors, which results in synapse and spine loss (Selkoe, 2008; Yu and Lu, 2012).
Molecular mechanisms involved in Aβ toxicity in dendritic spines and synapses are believed
to include signaling cascades from Aβ to tau and PSD-95, involving tau kinases like PAR1/MARK and its activating kinase LKB1 (Yu and Lu, 2012). Toxicity studies on Aβ suggest
that Aβ can also affect LTP and LTD by modulating glutamate receptor-dependent signaling
pathways (Hsieh et al., 2006; Li et al., 2009; Shankar et al., 2007), triggering aberrant
ACCEPTED MANUSCRIPT
patterns of the neural network (Palop et al., 2007), inducing mitochondrial dysfunction (Lin
and Beal, 2006) and lysosomal failure (Nixon and Cataldo, 2006). Cultured neuronal cells
from Tg2576 mutant APP transgenic mice (mimicking Aβ-induced synaptic dysfunction)
show synaptic changes like fewer smaller postsynaptic compartments and fewer enlarged
active presynaptic compartments. A reduction in PSD-95 is observed to be the prime change
in the synaptic components of mutant mice (Kim and Sheng, 2004). In the hippocampus of
AD patients and in animal models of AD, the activation of p21-activated kinase (PAK, actin
RI
PT
regulator and downstream effector molecule of Rac) is markedly reduced and delocalized
(Penzes et al., 2003; Zhao et al., 2006). Brain tissues from AD patients and transgenic animal
models of AD show severe decrease in drebrin, a postsynaptic protein involved in actin
SC
stability in dendritic spines. Dendritic spine and synaptic loss in AD by abnormal signaling
pathways results in aberrant neural circuitry, which underlies basic causative factor for all the
NU
cognitive deficits and dementia.
8.2. Parkinson’s disease
MA
Parkinson's disease (PD) is characterized by a dramatic and selective loss of dopaminergic
neurons in the substantia nigra pars compacta along with a gradual and significant decrease in
dopamine levels (Alexander, 2004; Dauer and Przedborski, 2003; Sathiya et al., 2013). Major
D
symptomatic features of PD are tremor, bradykinesia, rigidity, impaired gait and abnormal
TE
posture. Main neuropathology of PD is the aggregation of alpha-synuclein in the midbrain
dopaminergic neurons (Dauer and Przedborski, 2003). Using Golgi-Braitenberg method,
CE
P
neuropathological changes such as a decrease in dendritic length, dendritic spine loss and the
formation of different dendritic varicosities are found in the substantia nigra pars compacta
neurons of PD patients (Patt et al., 1991). Striatal medium spiny neurons show a greater loss
AC
of dendritic spines in PD patients and in animal models of PD (Villalba and Smith, 2013). In
animal models of PD, rats with unilateral degeneration of the nigrostriatal dopaminergic
system induced by 6-hydroxydopamine show distinct striatal spine loss (about 20%) in the
caudate-putamen complex (Ingham et al., 1989). Further investigations reveal that the spine
loss is accompanied by a reduction of the total number of asymmetric glutamatergic synapses
in the striatum (Ingham et al., 1998). Several post-mortem striatal tissue samples from PD
patients show spine loss (Stephens et al., 2005; Zaja-Milatovic et al., 2005). Monkeys with 1methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP)-induced PD also show marked spine
loss along with the inhibition of spinogenesis in the medium spiny neurons of the striatum
(Stephens et al., 2005; Villalba et al., 2009; Villalba and Smith, 2010). The degree of spine
ACCEPTED MANUSCRIPT
loss is in accordance with the pattern of dopamine denervation (Smith and Villalba, 2008;
Villalba et al., 2009; Zaja-Milatovic et al., 2005). A recent research study shows that MPTPlesioned monkeys have about a 50% striatal dendritic spine loss and a significant reduction in
striatal dopamine levels, while about a 25% dendritic spine loss is noted in the nucleus
accumbens (Villalba et al., 2009). The 3D ultra-structural analysis reveals an increase in the
spine head size along with an increase in length, complexity and the extent of perforations of
PSD of corticostriatal and thalamostriatal glutamatergic synapses in the striatum of
RI
PT
Parkinsonian monkeys (Villalba and Smith, 2011a) and a massive growth of the spine
apparatus (Villalba and Smith, 2011b, 2010). The PSD length and number of perforated
synapses are significantly higher in dendritic spines of caudate nucleus neurons in 3 PD
SC
patients vs. 3 matched controls (Anglade et al., 1996). The most common pathway for
dendritic spine loss in PD is the activation of cortical glutamatergic system that results in
NU
glutamate excitotoxicity in striatal neurons (Calabresi et al., 1996; Mallet et al., 2006;
Wichmann and Delong, 2007). Both spine loss and inhibition of development of novel spines
MA
in the striatal regions accounts for the progression of PD.
8.3. Huntington’s disease
Huntington’s disease (HD) is an autosomal dominant,
progressive neurodegenerative
D
disorder, which usually causes movement, cognitive and psychiatric disorders with a broad
TE
variety of signs and symptoms. Common symptomatic features are involuntary movements of
the face and body; abnormalities in gait, posture and balance; obsessive-compulsive behavior,
CE
P
and dementia (Bertram and Tanzi, 2005). The main neuropathological feature is the abnormal
accumulation of the huntingtin protein intracellularly due to the repetition of glutamine
(Evans-Galea et al., 2013). Brain samples obtained from both human HD patients and various
AC
transgenic and knock-in animal models of HD show a progressive loss of dendritic spines in
the cortical and the striatal neurons (Ferrante et al., 1991; Guidetti et al., 2001). Post-mortem
brain tissues from patients with HD show severe dendritic degeneration. Patients with mild
HD show an abnormal bending of distal dendrites along with a marked increase in dendritic
branching, spine density and spine size in the striatal area (Ferrante et al., 1991; Graveland et
al., 1985). However, during the advanced stages of HD, degenerative changes like dendritic
arbor truncation, focal swellings on dendrites and decrease in spine density are noted. Brain
samples from HD patients show a remarkable increase in the number of primary dendritic
segments along with an increase in total dendritic length, total surface area, dendritic
branching and the number of dendritic branching points in layer-3 and -5 of pyramidal
ACCEPTED MANUSCRIPT
neurons when compared to the control participants. The total number and density of dendritic
swellings are markedly increased in layers-3 and -5 of the pyramidal neurons of HD brains.
Although the degeneration of prefrontal, cortical, and pyramidal neurons are noted at a
smaller scale in HD patients, the plasticity of the remaining pyramidal neurons occurs
temporarily as a compensatory response (Sotrel et al., 1993). Transgenic R6/1 and R6/2 mice
expressing a larger CAG trinucleotide expansion (closely related animal model of human
HD) show progressive cognitive, psychiatric and motor symptoms (Mangiarini et al., 1996;
RI
PT
Miller and Bezprozvanny, 2010; Nithianantharajah et al., 2008; Spires et al., 2004). In
YAC128 mouse model of HD, reduced striatal dendritic spine density and other spinal
morphological anomalies along with distinct decline in cognitive and motor skills are noted at
SC
16 weeks of age (Slow et al., 2003; Xie et al., 2010). These results demonstrate a significant
positive correlation between the loss of dendritic spines and dendritic anomalies and the
NU
progression of motor and cognitive deficits in HD.
MA
8.4. Intellectual disability or mental retardation
Intellectual disability is a generalized neurodevelopmental disorder and its prevalence ranges
from children below 3 years of age to adults. The main characteristic features include
D
impairment in intelligence quotient and adaptive functioning along with a wide spectrum of
TE
emotional and behavior disorders. Classical sign intellectual disability is a nonverbal
intelligence quotient below 70. Several functional studies show that numerous genes coding
CE
P
for synaptic proteins and involved in synapse and dendritic spine formation are mutated in
intellectual disability patients (Verpelli and Sala, 2012). A common neuropathological
feature is
decreased dendritic spine number with normal morphology in intellectual disability
AC
patients (Fiala et al., 2002). Several genes linked to intellectual disability and autism code for
proteins implicated in the Rac1/Cdc42/PAK and Rho pathways, like the upstream regulators
of Rac/Cdc42 such as PIX/ARHGEF6 and srGAP3/MEGAP (Slit Robo Rho GTPaseactivating protein 3), or downstream effectors such as PAK3 and LIMK and the RhoGAP
protein oligophrenin-1. Dysregulated protein synthesis or deficits in the regulation of
dendritic cytoskeleton, synaptic anomaly, alterations in dendritic spine density, and/or
distortions in spine shape are reported in several forms of mental retardation (Fiala et al.,
2002; Maul et al., 2008; Newey et al., 2005; Vanderklish and Edelman, 2005). Maul et al.,
(2008) show a direct correlation of abnormal hippocampal dendritic spine morphology and
length with impaired spatial memory in the angiotensin type II receptor knockout mice. Both
ACCEPTED MANUSCRIPT
these pathological features of dendritic spines are also detected in some cases of mental
retardation. Other forms of mental retardation e.g. Tuberous sclerosis type I, fetal alcohol
syndrome, non-syndromic mental retardation or Trisomy 21 are also shown to have
abnormalities
in
dendritic
spines
(Armstrong,
2005;
Kaufmann and
Moser,
2000;
Nimchinsky et al., 2002). Hence, genetic mutation-mediated decrease in spine density and
morphological abnormalities results in aberrant neural circuitry throughout the brain which
RI
PT
impairs the intelligence and higher cognitive functions of the individuals significantly.
8.5. Autism spectrum disorder
Autism spectrum disorders (ASD, which includes autism, Asperger’s syndrome) are defined
SC
as severe neurodevelopmental disorders characterized by low intelligence quotient, impaired
social communication and adaptation, disordered language, and repetitive and stereotypic
NU
behavior (Matsuzaki et al., 2004). Autism symptoms generally appear in early childhood
(Courchesne et al., 2007) and are associated variably with mental retardation, seizures, sleep
disorders, intellectual disability and gastrointestinal problems. In comparison to age-matched
MA
control subjects, the mean dendritic spine densities are higher in the cortical pyramidal
neurons of ASD subjects (Hutsler and Zhang, 2010). In the analyses of golgi-impregnated
ASD brain slices restricted to the apical dendrites of pyramidal cells, greater spine densities
D
are found commonly within layer II of the frontal, temporal and parietal lobes and within
TE
layer V of the temporal lobe in ASD subjects. High spine densities are associated with lower
levels of cognitive functioning and decreased brain weights in ASD subjects (Penzes et al.,
CE
P
2011). Brain samples of ASD patients show a higher dendritic spine density with reduced
dendritic spine elimination in layer V pyramidal neurons in the temporal lobe (Tang et al.,
2014). In a
genetic research study on ASD patients, major mutations in cell adhesion
AC
molecules like neurexins (Nrxn1) and neuroligins (Nlgn1, Nlgn3, and Nlgn4) genes are found
to have a stronger correlation with ASD (Dalva et al., 2007; Dean and Dresbach, 2006;
Ruddy et al., 2015; Südhof, 2008). Abnormal expression of Shank-3, Homer or cortactin is
found in ASD patients (Südhof, 2008). In both the mouse models of intellectual disability and
autism,
when
functional
and
behavioral
phenotypes
are
successfully
restored
by
pharmacological and genetic managements, a parallel recovery in spine alterations is
observed,
indicating
the
cardinal role of dendritic spines in the pathogenesis of
neurodevelopment diseases (Landi et al., 2011; Pop et al., 2014; Restivo et al., 2005; Tropea
et al., 2009).
ACCEPTED MANUSCRIPT
8.6. Fragile X syndrome
Fragile X syndrome (FXS) is the most common congenital intellectual disorder. FXS is more
common in males than in females. Characteristic features include mental retardation (like
reduced intellectual ability), impaired visuospatial processing, and developmental delay
(Walter et al., 2009). The concurrent prevalence of ASD ranges 15-60% in patients with
FXS, while 2–5% of autistic children have FXS (Hagerman et al., 2009). The absence of
FMR protein, caused by the excessive expansion of CGG sequence (more than 200 repeats)
RI
PT
in the fmr1 gene, results in the development of FXS (Willemsen et al., 2011). FMR protein is
a RNA-binding protein that regulates the transport and translation of a subset of neuronal
mRNAs (PSD95, elongation factor 1a, SAPAP3/4, AMPA receptor subunit GluR1/2, NMDA
SC
receptors, CaMKII, and other cytoskeletal proteins) to synapses (Bassell and Warren, 2008)
and the activation of metabotropic glutamate receptors (Antar et al., 2006, 2004; Bagni and
NU
Greenough, 2005; Bassell and Warren, 2008; De Rubeis and Bagni, 2010; Restivo et al.,
2005; Wang, 2015).
Patients with FXS show an increased number of long dendritic spine and a reduced number
MA
of shorter dendritic spines in both the temporal and visual cortical areas when compared to
the control subjects. In both the cortical areas of FXS patients, the distal segments of apical
and basilar dendrites show a greater dendritic spine density than the controls. FXS patients
D
also show significantly more dendritic spines with immature morphology and fewer dendritic
TE
spines with a more mature morphology (Irwin et al., 2001). Golgi-impregnated cerebral
cortex of Fragile X knockout mice (transgenic mice that lack Fmr1 gene expression) show
CE
P
long, thin and tortuous dendritic spines and increased spine density on the apical dendrites of
layer V of the pyramidal cells when compared to the dendritic spines noted in the wild-type
mice. Such morphological and density variations result in an impaired developmental
AC
organizational of synapse stabilization and turnover processes (Comery et al., 1997). In an in
vivo study (Cruz-Martín et al., 2010), during the first two postnatal weeks, Fragile X
knockout mice exhibit an abnormally high spine turnover and over-abundance of immature
protrusion subtypes (filopodia) in layer 2/3 neurons of the somatosensory cortex. On the
contrary, the wild-type mice show a rapid decrease in dendritic spine dynamics; immature
small filopodia and protospines are replaced by mushroom spines. Generally, FMR protein
controls the balance of the mGluR5 and AMPA trafficking, but the overexpression of
metabotropic glutamate receptors causes synaptic damage and spine abnormalities leading to
the behavioral deficits seen in FXS (Cook et al., 2014). The absence of synaptic proteins
results in aberrant spine development, like an increased number of long, thin, and immature
ACCEPTED MANUSCRIPT
filopodia-like spines in cortical neurons of FXS patients (Sutton and Schuman, 2005; Weiler
et al., 2004). Put together, these results show that the development of immature dendritic
spines and impaired spine turnover results in the cortical areas of FXS patients with the
development of intellectual disability.
8.7. Rett syndrome
Rett syndrome is a rare neurological disorder commonly found in childhood, especially in
clinical features
are
hypotonia,
RI
PT
girls. Its characteristic features include wide range of motor to intellectual deficits. Main
difficulties in breathing,
speaking,
chewing,
feeding,
crawling, and walking, teeth grinding, impaired posture, delay in milestone achievements,
SC
wringing movements, seizures, walking on the toes, and a gradual decline in cognitive
development and intellectual abilities (Collins et al., 2004; Samaco et al., 2012). Rett
NU
syndrome is caused by sporadic X-linked mutation in the methyl-CpG-binding protein-2
(MECP2) gene, which encodes methyl cytosine binding protein-2, a protein that regulates
gene transcription by binding to methylated DNA, which is an essential protein in normal
MA
brain development (Gonzales and LaSalle, 2010). The MECP2 protein plays a vital role in
the regulation of synaptic development, synaptic plasticity and dendritic spine structure
(Chapleau et al., 2009; Jiang et al., 2013). Another gene found to be mutated in Rett
D
syndrome is CDKL5. CDKL5 is found to be involved in the positive regulation of dendritic
TE
spine morphology and synapse activity by phosphorylating NGL-1 and alleviating its
correlation with PSD95 in synapse (Ricciardi et al., 2012). Brain samples of Rett syndrome
CE
P
patients show a decrease in neuronal size and an increase in neuronal cell density (Bauman et
al., 1995) along with a marked decrease in dendritic arborization (Armstrong et al., 1995) and
spine dysgenesis (P. V. Belichenko et al., 1994; Chapleau et al., 2009; Kaufmann et al.,
AC
1995). Brain slices of Rett syndrome patients show a lower spine density and a decreased
number of mushroom-type spines in the cortex and hippocampus (Armstrong et al., 1995; P.
V. Belichenko et al., 1994; Chapleau et al., 2009). In animal models, MECP2-deficient mice
show impaired dendritic complexity (Fukuda et al., 2005; Nguyen et al., 2012), lower
dendritic spine density (Castro et al., 2014; Chapleau et al., 2012; Landi et al., 2011; Tropea
et al., 2009) and a lack of mushroom spines in cortical and hippocampal neurons (Baj et al.,
2014; Chao et al., 2007). These results suggest that both spine dysgenesis and decreased
dendritic arborization in many cortical areas is the key mechanism for the development of
motor and cognitive deficits in Rett syndrome. Therefore, selective restoration of spine
ACCEPTED MANUSCRIPT
density and morphological abnormalities of dendritic spines using therapeutics may delay the
disease progression in Rett’s patients.
8.8. Down’s syndrome
Down’s syndrome is a highly prevalent neurodevelopment genetic disorder. It is caused by
the presence of a third copy of human chromosome 21 which regulates the transcription of
almost 300 genes (Haas et al., 2013) and thus, it is also known as trisomy 21. The
RI
PT
pathological hallmark features in Down Syndrome patients are abnormal physical growth and
spine dysgenesis resulting in a cognitive deficits including learning and memory (Haas et al.,
2013; Takashima et al., 1981; W. Moser, 1995). Other symptoms include poor immune
SC
function (Hammer and McPhee, 2014), stunted growth, flattened nose and head, decreased
muscle tone, protruding tongue, strabismus, slanted eyes, short neck and abnormal teeth.
NU
Down’s Syndrome patients show a higher risk of leukemia, thyroid diseases, congenital heart
defects, epilepsy, and mental disorders (Hickey et al., 2012). The dendritic spine number is
generally normal in Down’s syndrome patients, but decreases quickly around 20 years of age
MA
(Takashima et al., 1994).
Using the mouse models of Downs Syndrome (Ts1Rhr and Tc mice), it was shown that,
markedly fewer thin spines are found in Ts1Rhr mice (aged 3 weeks) and significantly fewer
D
mushroom spines with a concomitant increase in the number of stubby spines are reported in
TE
Tc1 mice (aged 3 months). Mouse models of Down’s syndrome show significantly bigger
spine heads in the dendritic spines of the motor cortex along with a decreased dendritic spine
CE
P
number in the hippocampal dentate gyrus (Belichenko et al., 2009, 2007). Rapid Golgi
method reveal a marked reduction in dendritic spine number in the apical (middle, distal and
oblique segments) and basilar (thick and thin segments) dendritic arbors of CA1 and CA2-3
AC
segments of the hippocampal neurons in Down's syndrome patients with no evidence of AD.
A significant decrease of dendritic spines in every segment is noted in Down's syndrome
patients associated with AD when compared to age-matched controls and Down's syndrome
patients without AD. In Down's syndrome (either associated or not to AD), thin basilar
dendrites are severely affected; while in AD patients, CA1 pyramids are more severely
affected than the pyramidal neurons of CA2-3 areas (Ferrer and Gullotta, 1990). The
developmental morphology of the visual cortical neurons in newborns and older infants with
Down's syndrome shows shorter basilar dendrites and decreased numbers of spines with
altered morphology and defective cortical layering (Takashima et al., 1981). Altogether, these
results suggest that dendritic spine abnormalities are an important cause for cognitive decline
ACCEPTED MANUSCRIPT
in Down syndrome patients in comparison to the neuronal structural defects (Haas et al.,
2013).
8.9. Schizophrenia
Schizophrenia is a complex neuropsychiatric disorder affecting the cognitive and sociopsychological skills, with an incidence rate of 0.5–1% of the population (Awad and
Voruganti, 2008; Wu et al., 2005). Mostly, schizophrenia appears in late adolescence or early
RI
PT
adulthood (Lewis and Lieberman, 2000) and is more commonly seen in males (Aleman et al.,
2003; Grossman et al., 2008; McGrath et al., 2004). The main symptoms are social isolation,
agitation,
compulsive
hallucinations,
behavior,
paranoia,
disorganized
behavior, excitability, hostility,
disturbing thoughts, amnesia, memory loss, mental
SC
delusions,
aggression,
confusion, slowness in activity, altered emotional expression or lack of emotional response,
with cognition impairment,
executive
NU
circumstantial speech, incoherent speech, and repetitive movements (Yoon et al., 2013) along
dysfunction,
hallucination, , and problems with
working memory (Elvevåg and Goldberg, 2000; Lesh et al., 2011). Main pathology of
MA
schizophrenia is due to the interference inconnectivity and deficit in association among
various brain regions. Of the 100 genetic variants involved in schizophrenia, DISC1, ERBB4,
and NRG1 are found to show constant association with schizophrenia in relation of synapse
D
and spine function (Lipska et al., 2006; Mei and Xiong, 2008; Schumacher et al., 2009;
TE
Shamir et al., 2012; St Clair et al., 1990; Walsh et al., 2008). Most important
neuropathological feature of the schizophrenic brain is the loss of grey matter, like the
CE
P
shrinkage or loss of neurons in the dorsolateral prefrontal cortex, superior temporal gyrus,
subicular complex, and CA3 area of hippocampus (Glausier and Lewis, 2013).
In layer 3 of the prefrontal and temporal cortical areas, spine density is approximately 60%
AC
lower in schizophrenia subjects (Garey et al., 1998). Post-mortem brain samples of
schizophrenic subjects show approximately 20% lower spine density on the basilar dendrites
of pyramidal neurons located in layer deep 3 of the dorsolateral prefrontal cortex area 46 and
in layer 3 of the primary visual cortical area 17 compared with the normal healthy controls
and psychiatric subjects (Glantz and Lewis, 2001; Kolluri et al., 2005). In deep layer 3 of the
primary and association auditory cortices areas 41 and 42, the dendritic spine density is
significantly decreased in 15 schizophrenic subjects vs. normal comparison subjects (Sweet
et al., 2009). In schizophrenic brains, both the pyramidal neurons (in frontal and temporal
cortices) and medium spiny neurons (of the striatum) show reduced dendritic spine number.
Postmortem brain slices of schizophrenic patients show smaller pyramidal somas and reduced
ACCEPTED MANUSCRIPT
neuropil.
These
results
altogether
suggest
that
the
neurobehavioral
alterations
in
schizophrenic patients are mainly due to deficits in dendritic spines, which in turn results in
altered neuronal circuitry (Glausier and Lewis, 2013).
8.10. Depression
In today’s urbanized life style, the increased levels of struggles and competition affect the
mental health leading to anxiety, stress and depression. Depression is a severe psychiatric
RI
PT
disorder affecting approximately 20% of the American population (Kendler et al., 2006; Patel
et al., 2001) and is more common in females than in males (Bangasser and Valentino, 2014;
Kessler et al., 2005; Levinstein and Samuels, 2014; Thompson et al., 2015). The most
SC
common symptoms are anxiety, apathy, fatigue, mood swings, grieving, excess of sleepiness
or insomnia, excessive hunger, loss of appetite, irritability, agitation, lack of concentration,
NU
slowness in activity, and suicidal thoughts. Depression is closely associated with alterations
in dendritic spine morphology and spine density.
Short-term or moderate stressful conditions make individuals more adaptive and augment the
MA
neuronal function in most cases, but severe or chronic stressors are harmful and eventually
disrupt the ability to handle normal stress resulting in fear, anxiety and depression in later
stages of life (de Kloet et al., 2015; Lupien and McEwen, 1997; McEwen, 2005; Radley and
D
Morrison, 2005). Compared to healthy individuals, depressive patients show a smaller
TE
hippocampus due to a marked reduction in dendritic arborization and dendritic spine density
in hippocampal neurons (Bremner et al., 2000; Caetano et al., 2004; Drevets et al., 2008;
CE
P
Lorenzetti et al., 2009; Sawyer et al., 2012; Videbech and Ravnkilde, 2004). Patients with
major depressive disorders also show reductions in dendritic spines of hippocampal regions
(Law et al., 2004).
AC
Under chronic stress, the neurons of the brain regions like amygdala, hippocampus, and
prefrontal cortex show remodeling of the synaptic connectivity, neuronal degeneration,
impairment of neuroplasticity and dendritic spine loss (Christoffel et al., 2011; Radley and
Morrison, 2005). Meanwhile, the neurons of the amygdala and nucleus accumbens exhibit an
increase in spine density. These alterations induced by chronic stress are often accompanied
by depression-like behaviors like cognitive disturbances, poor concentration and negative
thoughts (Gualtieri and Morgan, 2008). Chronic stress induces the release of stress hormones
(glucorcorticoid and epinephrine) that can activate many kinases and phosphatases resulting
in the activation of the cortactin pathway. Phosphorylated (activated) cortactin promotes the
polymerization and branching of actin in dendritic spines resulting in spine dysfunction and
ACCEPTED MANUSCRIPT
an altered signal transmission across the synapse (Leuner and Shors, 2013). Under chronic
stress, increased secretin of corticosterone is found to be the major cause for the induction of
depression-like behaviors, synaptic dysfunction and dendritic spine loss in the hippocampus
of male mice and rats (Wang et al., 2013).
Treatment with antidepressants significantly increase the spine density in hippocampal CA1
and BDNF mRNA in a rat brain (Altar, 1999). Since BDNF and its specific receptor trkB are
found to play an important role in the regulation of hippocampal spine densities and spine
RI
PT
shape (Chapleau et al., 2008), significant reductions in density of dendritic spines are noted in
trkB-deficient mice (von Bohlen und Halbach et al., 2008, 2006). Dysregulation of fibroblast
growth factor 2 seen in individuals with major depression is found to correlate with the
SC
altered mean length of hippocampal dendritic spines (Zechel et al., 2009). Based on these
data, it is clear that aberrant alterations in spine density and morphology of hippocampal
NU
neurons are important mechanisms underlying the development of depressive neural
circuitry. Hence, therapeutic targets aimed at restoring the spine loss and dysfunction in
cortical regions like amygdala, hippocampus and prefrontal cortex may help the depressive
MA
patients significantly.
8.11. Epilepsy
D
Epilepsy is a chronic neurological disease characterized by recurrent epileptic seizures
TE
(Wong and Guo, 2013). The incidence rate of epilepsy is approximately 1% of the world
population and it is associated with high risk of morbidity and mortality (Dodrill, 2002; Elger
CE
P
et al., 2004). Epilepsy can be defined as a multifactorial disease as several neurobiological,
cognitive, psychological, environmental and social factors contribute to this condition
(Todorova et al., 2006). Several factors like tumors, strokes, head trauma, previous infections
AC
of the central nervous system, genetic abnormalities, exposure to environmental toxins
(pesticides, herbicides, insecticides, xenobiotic etc.), neurodegenerative diseases and a
variety of metabolic disorders can be possible causes for epileptic seizures (Todorova et al.,
2006). Epileptic seizures can be classified as focal and generalized seizures. Symptoms
commonly found
in focal seizures are muscle contractions followed by relaxation,
contractions on just one side of your body, abnormal head or eye movements, numbness,
tingling, abdominal pain, tachycardia, repetitive movements, sweating, nausea, pupillary
dilation, vision changes, hallucinations, mood changes and blackouts. The symptoms found
in generalized tonic-clonic seizures are fatigue, temporary confusion, depression, anxiety,
ACCEPTED MANUSCRIPT
fear, uncontrollable jerking movements of the arms and legs, loss of consciousness, tingling,
dizziness, muscle stiffening, urinary incontinence, atonia, and tongue biting.
Histological sections of rat brain with acute seizures or chronic epilepsy induced by
convulsant drugs (pilocarpine, monosodium glutamate or kainate) or electrical kindling show
a frequent loss of dendritic spines and varicose swelling of dendrites (Ampuero et al., 2007;
González-Burgos et al., 2009; Isokawa, 1998; Jiang et al., 1998; Nishizuka et al., 1991;
Willmore et al., 1980). On the contrary, an increase in dendrites or dendritic spines is also
RI
PT
reported rarely (Bundman et al., 1994; Spigelman et al., 1998; Suzuki et al., 1997). In vitro
seizure models including epileptiform bursting in brain slice-cultures have provided the
evidence for marked spine loss and other dendritic changes (Drakew et al., 1996; Müller et
SC
al., 1993; Nishimura et al., 2008; Zhao et al., 2006). In vivo time-lapse studies of acute
seizure models show marked emergence of dendritic injury following seizures, which is
NU
characterized by the temporary beading of dendrites that resolved within a couple hours after
a seizure and then followed by constant loss of dendritic spines (Mizrahi et al., 2004; Rensing
et al., 2005; Zeng et al., 2007). In rats with pilocarpine-induced acute status epilepticus, the
MA
dendrites of dentate granule cells show a generalized spine loss immediately after the
seizures. However, these spine losses are transient, recovery and dynamic changes in the
shape and density of dendritic spines are noted during the chronic phase of seizures.
D
Pathological studies of cortical and hippocampal samples of patients with intractable epilepsy
TE
show a number of dendritic abnormalities, which mainly included a significant loss of
dendritic spines (Swann et al., 2000; Wong, 2005). Hippocampal pyramidal neurons and
CE
P
dentate granule cells of patients with temporal lobe epilepsy show dendritic spine loss either
alone or concomitantly with varicose swelling of dendritic branches (Pavel V. Belichenko et
al., 1994; Isokawa and Levesque, 1991; Scheibel et al., 1974; Schewe et al., 1998; von
AC
Campe et al., 1997). Similarly, spine loss and dendritic swellings are noted in the pyramidal
neurons of the neocortex, including sites distant from the primary epileptogenic focus
(Multani et al., 1994). In humans with temporal lobe epilepsy, the proximal dendrites of
dentate granule cells reveal greater spine density (Isokawa, 2000). Other less commonly
found dendritic abnormalities are changes in dendritic length, shape and branching patterns,
along with a focal increase in dendritic spines in patients with neocortical and hippocampal
epilepsy (Pavel V. Belichenko et al., 1994; Isokawa, 1997; Multani et al., 1994; Schewe et
al., 1998; von Campe et al., 1997). Hippocampal neurons of temporal-lobe epileptic patients
show dendritic swelling, varicosities formation, bigger spine heads and electron-dense spines
on deteriorating dendrites (Fiala et al., 2002). Morphological abnormalities in dendrites
ACCEPTED MANUSCRIPT
(spine loss and dendritic beading) are often found in human pathological studies and animal
models of epilepsy, which contribute mainly for the neurological dysfunction along with
other neurological/cognitive deficits (Wong, 2008). Epileptic seizure induced spine loss is
mediated mostly by the cofilin pathway, which is involved in the positive regulation of actin
polymerization and spine dynamics (Kang et al., 2011). The activation of PAK/LIMK or
calcineurin pathway by epileptic seizures-induced glutamate excitotoxicity are the other
common pathways involved in disrupting of spine cytoskeleton in turn to spine collapse
RI
PT
(Wong and Guo, 2013). These studies reveal the distinct correction of spine beading, spine
loss and dendritic swelling with abnormal neural circuitry which causes outburst of seizure
SC
activity.
8.12. Stroke
NU
As the brain consumes about 20% of the total energy generated in the body, adequate blood
flow is crucial to meet the high demand of oxygenation for the generation of energy. As
neurons hold a higher metabolic rate, the demand of oxygen supply is generally high, and
MA
neurons are vulnerable to hypoxic stress (Baburamani et al., 2012). The clinical condition of
interrupted or reduced blood flow to the brain results in stroke. The impairment of cerebral
circulation due to a blocked artery (like thrombus formation) is defined as ischemic stroke
D
(incident rate approximately 87%) or due to leaking/rupture of blood vessel (like aneurysms)
TE
is defined as a hemorrhagic stroke (incident rate about 13%; Naranjo et al., 2013). Ischemia
is generally defined as a reduced blood flow to a particular region of brain which causes the
CE
P
deprivation of adequate oxygenation, leading to tissue hypoxia (reduced oxygen) or anoxia
(absence of oxygen). Common symptoms include fatigue, vertigo, gait disturbances, loss of
balance and coordination, blurred or double vision, temporary vision loss, difficulty in
AC
speaking and swallowing, slurred speech, speech loss, numbness or weakness in limbs and
facial muscles, headache and muscle paralysis.
After 20 minutes of hypoxia, increases in dendritic varicosities and spine loss are observed
(Hasbani et al., 2001). Early symptoms of neuronal injury due to ischemia include synaptic
failure or the failure of neurotransmission due to alterations in dendritic spine shape like
varicosite formation or constriction (Hori and Carpenter, 1994). The main sign of
ischemia/hypoxia is energy failure due to the impairment of mitochondrial oxidative
phosphorylation (Solaini et al., 2010). Additionally, excessive Ca2+ influx through glutamate
receptors causes calcium excitotoxicity which impairs actin dynamics by activating actin
depolymerization via the cofilin pathway (Kang et al., 2011) and causing mitochondrial
ACCEPTED MANUSCRIPT
damage leading to high levels of reactive oxygen species (Solaini et al., 2010). Therefore,
neuronal death and dendritic spine loss are the main consequences of ischemia/hypoxia
(Akulinin et al., 1997; Kolb and Gibb, 1993).
The experimental models of cerebral ischemia used in various in vitro and in vivo studies
report spine loss within minutes to hours following the insult (Babu and Ramanathan, 2011;
Corbett et al., 2006; Hasbani et al., 2001; Jourdain et al., 2003). Focal ischemia induced by
photo-thrombosis causes spine loss in layer 1 apical dendritic tufts of the ischemic YFP-
RI
PT
labelled layer 5 neurons (Enright et al., 2007; Kalaivani et al., 2014; Zhang et al., 2005).
Golgi-Cox labelling technique traces the rapid reduction of neurons in sensorimotor cortex
layer 2/3 and 5 of mice with photo-thrombotically induced stroke. Substantial loss in spine
SC
density (approximately 38%) is recorded during the first 24 h following insult and the
reduction continues during the delayed phase (approximately 25% at 6 hours) in the
NU
penumbra area of cortex. These effects are restricted mainly to the peri-infarct regions (less
than 200 m from the core area / infarct border (Brown et al., 2008).
In mouse models of transient global ischemia, rapid distortion of dendrites along with fewer
MA
number of loss of spines are noted following ischemia in the somatosensory cortices.
Majority of the disrupted synaptic structures are rapidly reinstated following reperfusion,
suggesting that the restored dendrites can survive the initial ischemia challenge. Spines on the
D
restored dendrites undergo rapid and sustained structural reorganization following the
TE
transient ischemia. Such results suggest that the increase in spine/synapse number may
represent an adaptive response of post-ischemic neurons to the compromised cortical (Babu
CE
P
and Ramanathan, 2011; Zhu et al., 2017) functions.
8.13. Prion disease
AC
Prion diseases belong to a rare variety of fatal neurodegenerative disorders of animals and
humans called transmissible spongiform encephalopathies. They may occur in sporadic
(Jakob-Creutzfeldt
disease),
genetic
(genetic
Jakob-Creutzfeldt
disease,
Gerstmann-
Straussler-Scheinker syndrome, and fatal familial insomnia), and acquired (kuru, variant
Jakob-Creutzfeldt disease, and iatrogenic Jakob-Creutzfeldt disease) forms (Geschwind,
2015). They are caused by the conversion of the normal (cellular) prion protein (PrP) with a
primarily α-helical structure into an abnormal disease-associated form of the protein called
the prion with a primarily β-pleated sheet structure (Prusiner, 1998) inside the brain. As prion
proteins are not broken down by proteases, they can accumulate extracellularly within the
CNS to form plaques, called as prion plaques, which disrupt neuronal morphology (Brundin
ACCEPTED MANUSCRIPT
et al., 2010; Prusiner, 1998). Eventually, progressive neuron destruction causes brain tissue to
become filled with holes in a sponge-like or spongiform pattern (Brundin et al., 2010).
Cognitive deficits (memory loss, aphasia, dysphasia, mental confusion, decreased alertness,
apraxia) are the most commonly noted symptom, followed by cerebellar (difficulty in
walking, altered gait and posture, limb ataxia), constitutional (dizziness, vertigo, lethargy,
fatigue, sleep disturbance, headache, urinary incontinence, weight loss, palpitations), and
behavioral
(agitation,
irritability,
depression,
anger/aggression,
apathy,
mania,
social
RI
PT
isolation, panic attack) symptoms. In both animal models of prion diseases (Belichenko et al.,
2000; Sisková et al., 2009) and in the brain samples of patients with prion disease, called
Creutzfeldt-Jakob disease (Landis et al., 1981), reductions in dendritic spine number and
SC
emergence of varicosities are reported. Mice infected with scrapie prions show an accelerated
reduction in synaptic density along with degeneration of the presynaptic compartment and
NU
dendritic spine loss in neurons of hippocampus and cerebral cortex which are followed by
nerve cell loss (Brown et al., 2001). In vivo two-photon imaging in YFP-H mice inoculated
with Rocky Mountain Laboratory prion strain over several months show a steady decline in
MA
density of dendritic spines in layer-5 cortical neurons. Another study on prions disease shows
that cellular prion protein is a mediator of Aβ-induced synaptic dysfunction and spine loss as
both treatment with anti-PrP antibodies or PrP knockout reverses oligomeric Aβ-induced
8.14. Sleep disorders
TE
D
impairment of synaptic plasticity and spatial memory deficits in mice (Laurén et al., 2009).
CE
P
Sleep is a fundamental neurophysiological mechanism which plays a crucial role in synaptic
development, maintenance and plasticity, as extensive synaptogenesis is known to occur in
early post-natal life when sleeping hours are longest. Apart from playing a major role in
AC
memory consolidation (Grønli et al., 2014), sleep helps in the elimination of toxins from
brain and alleviates the abnormal dendritic spines in various neurodegenerative and
neuropsychiatric diseases (Picchioni et al., 2014). Sleep disorders have proven to worsen
several neuropsychiatric and neurodegenerative disorders (Dorris et al., 2008; Halbower et
al., 2006; O’Brien and Gozal, 2004). Many researchers postulate a strong link between spine
dysfunction and degeneration in sleep-deprived animal models and in patients with sleep
disorders. Sleep loss or sleep deprivation has been shown to have detrimental effects on
dendritic spines signaling, actin polymerization, and dynamic changes in dendritic spines.
Experimentally, sleep loss can decrease synaptic signaling molecules (PSD95, Shank,
Homer, Kalirin, cortactin), synaptic cell adhesion molecules (neurexins and neuroligins),
ACCEPTED MANUSCRIPT
neurotropic factors (BDNF and TrkB) and basal glutamate levels (Davis et al., 2004; Yang et
al.,
2014).
In mice,
sleep
deprivation for 5
hours increases the expression of
phosphodiesterase-4 which in turn reduces PKA activation and expression of several
immediate early genes (Vecsey et al., 2009). All of these molecular changes interfere with
signaling pathways involved in F-actin stabilization or perturb synaptic protein expressions
inside DS resulting in spine instability. Using transcranial two-photon microscopy and
EEG/EMG recordings, the turnover of dendritic protrusions (dendritic spines and filopodia)
RI
PT
is found to be higher over 2 hours in both the awake and sleep states in the developing mouse
somatosensory cortex. The formation of dendritic spines is similar between the wakeful and
sleep states, but the elimination of dendritic spines is lower in wakefulness state than sleep
SC
state. Similar results are reported on dynamics of dendritic spines and filopodia over 12-hour
light/dark cycle (Yang and Gan, 2012). Sleep after motor learning task enhances the
NU
formation of dendritic spines in layer V pyramidal neurons in mouse primary motor cortex
(Yang et al., 2014). All these data suggest the crucial role of sleep for healthy dendritic spines
MA
and direct link between sleep disorders and dendritic spine abnormalities.
9. Effects of therapeutic regimens and new chemical leads on dendritic spines
9.1. Anti-psychotics
D
Several research studies prove that dopamine plays a major stimulatory role in dendritic spine
TE
growth in primary dissociated hippocampal neurons (Critchlow et al., 2006). It was also
observed that haloperidol, a dopaminergic D2 receptor blocker, reduces spinophilin (Ouimet
CE
P
et al., 2004). Treatment with chlorpromazine and risperidone (D2 dopamine receptor
blockers) or with olanzapine (an atypical neurolepticum and D4-receptor antagonist)
increases the expression of microtubule-associated protein 2 in the dentate gyrus of rats (Law
AC
et al., 2004). In rats, the depletion of dopamine levels causes a marked reduction in spine
length and spine density in the nucleus accumbens (Meredith et al., 1995). D-amphetamine
and methylphenidate are common psychostimulants, widely used to treat narcolepsy and
attention-deficit/hyperactivity disorder (W. Alexander Morton. and Gwendolyn G. Stockton,
2000, David J Heal et al., 2013). Interestingly, treatment with drugs that elevate dopamine
levels such as methylphenidate, cocaine and D-amphetamine increases the pre-synaptic
proteins GAP-43 and synaptophysin (Stroemer et al., 1998) that stimulate the dendritic
growth (Levitt et al., 1997) and increase the dendritic spine density (Lee et al., 2006; Zehle et
al., 2007).
ACCEPTED MANUSCRIPT
9.2. CNS stimulants
In rats, repeated intraperitoneal treatment with D-amphetamine sulphate (3 mg/kg) or cocaine
HCl (15 mg/kg) for 5 consecutive days for 4 weeks significantly increases the dendritic
arborization and the density of dendritic spines on medium spiny neurons in the shell of the
nucleus accumbens and pyramidal cells in the medial prefrontal cortex. The brain samples of
cocaine-treated rats show increased branching of dendrites and spine density on the basilar
dendrites of pyramidal cells (Robinson and Kolb, 1999). In an earlier context drug-associated
RI
PT
study on learning, David and colleagues reported that rats treated with amphetamine (1
ml/kg, i.p.) and immediately paired with a specific chamber spend significantly more time in
the amphetamine-paired (AMPH CPP) than the unpaired chamber during the conditioned
SC
place preference test (Rademacher et al., 2010). Another important finding was that neurons
from the basolateral nucleus of the amygdala of amphetamine-paired group show an increase
NU
in the total number of asymmetric (Gray Type I, excitatory) synapses contacting spines and
dendrites (Figge et al., 2013; Rademacher et al., 2010) and symmetric (Gray Type II,
inhibitory or modulatory) synapses contacting the dendrites of neurons in the basolateral
MA
nucleus of the amygdala (Figge et al., 2013) than the saline-treated controls. In a recent
stereological analysis (Rademacher et al., 2015), the above mentioned findings were further
augmented by proving an increase in asymmetric (excitatory) and symmetric (inhibitory or
D
modulatory) inputs contacting the gamma-aminobutyric acid (GABA)-ergic interneurons.
TE
Specifically, increases in excitatory synapses onto dendrites of GABAergic interneurons and
inhibitory or modulatory synapses onto dendrites were observed in AMPH CPP group, but
CE
P
not in the delayed pairing group. The results suggest that the changes were mainly due to the
formation of context-drug associations rather than repeated exposures to amphetamine.
AC
9.3. MK-801, a NMDA antagonist
In mouse striatal neurons, acute exposure to MK-801 (0.2 mg/kg, i.p.) markedly reduces
spine density after 24 hours; but has no distinct effects on the spine density of cortical
pyramidal neurons. Administration of MK-801 for 7 days substantially reduces spine density
in the striatum of wild type mice (Ramsey et al., 2011). Similarly, subchronic administration
of MK-801 (0.2 mg/kg/day for 7 days) causes a transient decrease in dendritic spine density
in medium spiny neurons in striatum, but induces a marked reduction in spine density in the
cortex (Ruddy et al., 2015).
9.4. Ketamine, a NMDA antagonist
ACCEPTED MANUSCRIPT
Acute treatment with a low dose of ketamine (10 mg/kg, i.p.) increases spine density in rat
brain after 24 hrs post-injection (Li et al., 2010). A single low dose of ketamine (10 mg/kg,
i.p.) injection produces an increase in spine density in layer V of the pyramidal neurons of the
cortex, while mice treated with a higher dose of ketamine (20 mg/kg) do not show such
effects. However, there were no changes in the striatal spine density following the
administration of ketamine (Ruddy et al., 2015).
Similarly, administration of ketamine (20 mg/kg once daily for 7 days) markedly decreases
RI
PT
spine density in layer V of the pyramidal neurons in the cortex, but not in the striatum.
Ketamine (5 and 10 mg/kg, i.p.) activates the mammalian target of rapamycin (mTOR)
signaling proteins (eukaryotic initiation factor 4E binding protein 1, p70S6 kinase and
SC
mTOR), both transiently and dose-dependently, in the synapto-neurosomes of prefrontal
cortex, which also returns to basal levels 2 h after administration. Ketamine also producesa
NU
rapid and transient enhancement in the phosphorylated and activated forms of extracellular
signal-regulated kinases (ERK, including ERK1 and ERK2) and protein kinase B (Li et al.,
2010). Pre-treatment with NBQX (10 mg/kg, i.p.) blocked ketamine (10 mg/kg, i.p.) induced
MA
activation of phospho-mTOR, phosphor-4E binding protein 1, and phosphor-p70S6K, as well
as the other upstream signaling of kinases like phospho-ERK and phospho-Akt (Li et al.,
D
2010).
TE
9.5. Agomelatine, a melatonergic antidepressant
Agomelatine is known to be a safe antidepressant with relatively no significant withdrawal
CE
P
symptoms (Hickie and Rogers, 2011, de Bodinat C et al., 2010). Exposure to agomelatine
completely normalizes the stress-affected cell survival by completely reversing the reduced cFos and double cortin expressions, but only partially reverses the synapsin-1 expression in rat
AC
hippocampal dentate gyrus (Dagyte et al., 2010; Dagytė et al., 2011). Chronic oral
administration of agomelatine (40 mg/kg for 20 weeks) enhances the spatial memory in rats
in the Morris water maze task. It causes significant volumetric and numerical enhancements
in granular and pyramidal neurons in the dentate gyrus and CA1-3 subregions; and a marked
increase in the number of the mushroom and stubby types of dendritic spines, without
affecting the number of thin-shaped spines. The above mentioned reports suggest that
agomelatine produces nootropic effects through spine structural changes (Demir Özkay et al.,
2015).
9.6. 9-Methyl-beta-carboline, a heterocyclic amine of beta-carboline family
ACCEPTED MANUSCRIPT
A neuro-morphological study reports that 9-methyl-beta-carboline administration stimulates
the growth of granule cells in the dentate gyrus of rats. Rats treated with 9-methyl-betacarboline (0.2 umol/100 g, i.p., for 10 days) show improvements in spatial learning in the
radial arm maze along with increased hippocampal dopamine levels and an increase in the
spine length, number of dendritic spines and the dendritic complexity in several concentric
rings of granule cells in dentate gyrus (Gruss et al., 2012). Also, other in vitro studies
(Hamann et al., 2008; Polanski et al., 2010) show that treatment with 9-methyl-beta-carboline
RI
PT
(10-150 uM for 48 hours) increases the number of dopaminergic neurons in a dose-dependent
manner. Newly formed dopaminergic neurons show an increase in the number and
SC
ramification of neurites.
9.7. Vinpocetine, a synthetic derivative of the Vinca Alkaloid extracted from the periwinkle
Vinpocetine,
a
phosphodiesterase
NU
plant
inhibitor,
known
for
great
potential in cognitive
enhancement (DeNoble VJ, 1987, Deshmukh R et al., 2009, Filgueiras CC et al., 2010,
MA
Ishihara K et al., 1989) and effective anti-inflammatory action (Alexandre E. Medina, 2010).
Two-photon laser scanning microscopy revealed that perfusion with vinpocetine significantly
increases the dendritic spine motility in the neocortical layer 2/3 pyramidal neurons of rats
D
(Lendvai et al., 2003). Rapid increases in the length of existing spines often appears as
TE
growths from the heads of mushroom spines. The formation of new spines is more frequent
than the complete retraction of spines during vinpocetine administration. Veratridine, a
CE
P
steroid-derived neurotoxin alkaloid, enhances sodium influx, but fails to induce any change
in spine motility. The results indicate that rapid changes in spine shape and size occur when
calcium and sodium influx is altered by vinpocetine. Spine motility induced by vinpocetine
AC
might be associated with microtubule alterations.
10. Modification of life style for maintenance of healthy dendritic spines
Many researchers have shown that life style plays a main role in neuronal vulnerability,
synaptic plasticity and cognitive development. Factors such as reduction of chronic stress,
healthy food and nutritional diets along with physical (aerobic and cardiac) and cognitive
exercises enhances the molecular mechanisms of dendritic signaling and plasticity (Mora,
2013). Chronic stress negatively regulates dendritic spine morphology, dynamicity and
function along with synaptic plasticity and secretion of stress hormones (corticosterone). This
ACCEPTED MANUSCRIPT
affects neuronal energetics leading to spine or neuronal loss mainly during advanced age
(Mora, 2013; Mora et al., 2012). Avoiding chronic stress is essential for the maintenance of
significantly better neural and synaptic functioning.
A diet rich in antioxidants can nurture our brain and provides better protection from oxidative
damage during advanced age (Colman et al., 2009; Johnson et al., 2007). Docosahexaenoic
acid from fish oil has been shown to improve cognitive processes by increasing neurogenesis
and the secretion of neurotropic factors (brain-derived growth factor, nerve growth factor,
RI
PT
glial derived neurotropic factor and ciliary neurotropic factor) in humans (Gómez-Pinilla,
2008) and positively regulates the genes important for maintaining synaptic function and
plasticity in rodents (Wu et al., 2007). A caloric restriction diet is found to reduce dendritic
SC
spine loss along with concomitant increases in neurogenesis which in turn enhances the
cognitive functions of an individual (Johnson et al., 2007). In transgenic mouse model of AD,
NU
a long-term and regular intake of pomegranate, dates and figs is shown to reduce the
neuroinflammatory effects (Essa et al., 2015). A study suggests that dietary phytochemicals,
particularly flavonoids, have neuroprotective effects by protecting neurons from stress-
MA
induced injury, by suppressing neuroinflammation and by promoting LTP and synaptic
plasticity. These beneficial properties are mediated by their ability to interact with signaling
pathways (lipid kinase) and a number of synaptic proteins are known to play a crucial role in
D
LTP, acquisition, consolidation and storage of memory in humans (Spencer, 2008). Regular
TE
aerobic exercise is proven to enhance neurogenesis, synaptic signaling and dendritic spine
number in elderly people. Daily exercise protects neuronal and synaptic loss, and delays the
CE
P
onset or progression of age-related neurodegenerative diseases (Cotman et al., 2007; Gitler,
2011; Hillman et al., 2008).
AC
11. Summary
Dendritic spines are believed to be an important anatomical focus of neural connectivity and
plasticity for many decades. Ultra-structural components of dendritic spines include actins,
myosins, endosomes, GTPase, PSD95, spine apparatus, NMDA and AMPA receptors. Many
investigations provide evidence on the prominent role of dendritic spines in the formation of
new neural circuitry as well as synaptic plasticity. Numerous in vivo and in vitro studies have
proven that dendritic spines undergo change in size, shape and number throughout life. The
density of spines varies with age, peaking at adolescence and then the number slowly
declines near adulthood. As the brain matures, the turnover and the density of dendritic
ACCEPTED MANUSCRIPT
spines gets stabilized, while the rate of elimination slows down. Even in the adult brain,
morphological and volumetric changes in dendritic spines mark the stimuli-induced synaptic
plasticity and novel synaptogenesis. The development of advanced imaging techniques like
two-photon laser scanning microscopy combined with glutamate caging and super-resolution
spine imaging,
crystal-clear neuroanatomical details of dendritic spines unfolded the
mysteries behind the synaptic connectivity and plasticity. With recent technologies such as
viral-mediated neuronal tracing and optogenetics, in vivo imaging studies have taken a great
RI
PT
leap forward. These viral-mediated genetic and fluorescent techniques enable the researchers
to target and decipher the specific neuronal population or neural networks involved in a
specific behavior and motor activities. Novel synaptogenesis and alterations in synaptic
SC
connections can be visualised directly using these latest technologies.
Depending upon the synaptic stimulation, dendritic spines undergo speciific morphological
NU
and volumetric changes. Generally, LTP induction causes the enlargement of spine head and
increases spine density, while LTD induction causes a decrease in spine size and density.
MA
During LTP induction, emergence and stabilization of the novel spines, and maturation of
thin spines to mushroom or stubby spines results in strengthening of the neural network,
while LTD induces significant pruning of the existing spines and prevents the maturation of
D
filopodia to mature spines. Multiple evidences reveal that the morphology of spines and their
TE
synaptic complexes depletes under the conditions of excitotoxicity.
Many imaging studies have proven that both, changes in the morphology of individual
CE
P
dendritic spines and spine density are related to neuronal plasticity, learning and memory.
During the learning process, both dendritic spine density and maturation of spines enhance
significantly. During memory formation, the neural circuitry gets stablilized as the synaptic
AC
connectiivity gets strengthened along with a prominent increase in the density of mushroom
spines. However, till date the molecular mechanisms underlying the correlation between the
morphological and volumetric changes in dendritic spines with learning and memory
processes are not clearly understood.
Several investigations on spines have shown that the signaling pathways that control spine
morphogenesis have to be tightly regulated at multiple levels. Dysregulation of these
signaling pathways, either by hyperactivation or hypoactivation leads to damaging effects on
cognitive function, learning and memory.
ACCEPTED MANUSCRIPT
The dynamic nature of spines mediated by various signaling cascades that characterize the
synaptic plasticity of the brain either by strengthening or pruning the synaptic connections
depends upon the induction (LTP or LTP), environment (stress or enrichment), and trauma
(injury or disorders). The two common types of synaptic plasticity studied extensively are
Hebbian plasticity and HSP. Hebbian plasticity is input specific and significantly increases
the dendritic spine density and synaptic strength. HSP operates in a wide range of synapses
spine density at an optimal level.
RI
PT
and stabilizes the neural network activities and optimizes their functions by maintaining the
Experimental studies have shown strong and clear links between reduction in the spine
density
and
progressive
decline
in
cognitive
and
motor skills.
However,
detailed
SC
investigations on the morphological, physiological, biochemical and molecular levels of
dendritic spines are required. Therapeutics targeted at
spine recovery promote the motor
symtoms.
Understanding
the
cellular
NU
skills and quality of life and also delay the progression as well as the onset of degenerative
mechanisms
and
signaling
pathways
of spine
physiological and
diseased
MA
morphogenesis and their role in structural plasticity of neural networks under both normal
conditions are cardinal for understanding spine function,
morphology and their role in cogitive function and behaviour. Specific neuronal stimulation
D
processes using latest imaging techniques will help decode the molecular processess
TE
underlying basic homeostasis to higher-order cognitive functions. Detailed investigations on
stabilizing the abnormalities of dendritic spines will lay the foundation for powerful
AC
References
CE
P
therapeutic tools for various neurodegenerative and psychiatric diseases.
Ackermann, M., Matus, A., 2003. Activity-induced targeting of profilin and stabilization
of dendritic spine morphology. Nat. Neurosci. 6, 1194–1200.
Adamantidis, A.R., Zhang, F., Aravanis, A.M., Deisseroth, K., de Lecea, L., 2007.
Neural substrates of awakening probed with optogenetic control of hypocretin neurons.
Nature. 450, 420-24.
Aravanis, A.M., Wang, L.P., Zhang, F., Meltzer, L.A., Mogri, M.Z., Schneider, M.B.,
Deisseroth, K., 2007. An optical neural interface: in vivo control of rodent motor cortex
with integrated fiberoptic and optogenetic technology. J. Neural Eng. 4, S143–S156.
ACCEPTED MANUSCRIPT
Arenkiel, B.R., Ehlers, M.D., 2009. Molecular genetics and imaging technologies for
circuitbased neuroanatomy. Nature 461, 900–90
Atasoy, D., Aponte, Y., Su, H.H., Sternson, S.M., 2008. A FLEX switch targets
Channelrhodopsin-2 to multiple cell types for imaging and long-range circuit mapping.
J. Neurosci. 28, 7025–7030.
Akulinin, V.A., Stepanov, S.S., Semchenko, V.V., Belichenko, P.V., 1997. Dendritic
RI
PT
changes of the pyramidal neurons in layer V of sensory-motor cortex of the rat brain
during the postresuscitation period. Resuscitation 35, 157–164.
Aleman, A., Kahn, R.S., Selten, J.-P., 2003. Sex differences in the risk of schizophrenia:
evidence from meta-analysis. Arch. Gen. Psychiatry 60, 565–571.
Alexander,
G.E.,
2004.
Biology
of
Parkinson’s
SC
disease:
pathogenesis
and
pathophysiology of a multisystem neurodegenerative disorder. Dialogues Clin Neurosci
NU
6, 259–280.
Altar, C.A., 1999. Neurotrophins and depression. Trends Pharmacol. Sci. 20, 59–61.
Altschuler, R.A., 1979. Morphometry of the effect of increased experience and training
MA
on synaptic density in area CA3 of the rat hippocampus. J. Histochem. Cytochem. 27,
1548–1550.
Ampuero, E., Dagnino-Subiabre, A., Sandoval, R., Zepeda-Carreño, R., Sandoval, S.,
D
TE
Viedma, A., Aboitiz, F., Orrego, F., Wyneken, U., 2007. Status epilepticus induces
region-specific changes in dendritic spines, dendritic length and TrkB protein content of
CE
P
rat brain cortex. Brain Res 1150, 225–238.
Anglade, P., Mouatt-Prigent, A., Agid, Y., Hirsch, E.C., 1996. Synaptic Plasticity in the
Caudate Nucleus of Patients with Parkinson’s Disease. Neurodegeneration 5, 121–128.
Antar, L.N., Afroz, R., Dictenberg, J.B., Carroll, R.C., Bassell, G.J., 2004. Metabotropic
AC
glutamate receptor activation regulates fragile x mental retardation protein and FMR1
mRNA localization differentially in dendrites and at synapses. J. Neurosci. 24, 2648–
2655.
Antar, L.N., Li, C., Zhang, H., Carroll, R.C., Bassell, G.J., 2006. Local functions for
FMRP in axon growth cone motility and activity-dependent regulation of filopodia and
spine synapses. Mol. Cell. Neurosci. 32, 37–48.
Arellano, J.I., Benavides-Piccione, R., DeFelipe, J., Yuste, R., 2007. Ultrastructure of
Dendritic Spines: Correlation Between Synaptic and Spine Morphologies. Front
Neurosci 1, 131–143.
ACCEPTED MANUSCRIPT
Armstrong, D., Dunn, J.K., Antalffy, B., Trivedi, R., 1995. Selective dendritic alterations
in the cortex of Rett syndrome. J. Neuropathol. Exp. Neurol. 54, 195–201.
Armstrong, D.D., 2005. Neuropathology of Rett syndrome. J. Child Neurol. 20, 747–
753.
Attar, A., Ripoli, C., Riccardi, E., Maiti, P., Puma, L., D, D., Liu, T., Hayes, J., Jones,
M.R., Lichti-Kaiser, K., Yang, F., Gale, G.D., Tseng, C., Tan, M., Xie, C.-W.,
Straudinger, J.L., Klärner, F.-G., Schrader, T., Frautschy, S.A., Grassi, C., Bitan, G.,
RI
PT
2012. Protection of primary neurons and mouse brain from Alzheimer’s pathology by
molecular tweezers. Brain 135, 3735–3748.
Awad, A.G., Voruganti, L.N.P., 2008. The burden of schizophrenia on caregivers: a
Bailey, C.H., Kandel, E.R., 1993. Structural Changes Accompanying Memory Storage.
Babu,
C.S.,
Ramanathan,
NU
Annual Review of Physiology 55, 397–426.
SC
review. Pharmacoeconomics 26, 149–162.
M., 2011. Post-ischemic administration of nimodipine
following focal cerebral ischemic-reperfusion injury in rats alleviated excitotoxicity,
MA
neurobehavioural alterations and partially the bioenergetics. International Journal of
Developmental Neuroscience 29, 93–105.
Baburamani, A.A., Ek, C.J., Walker, D.W., Castillo-Melendez, M., 2012. Vulnerability
D
of the developing brain to hypoxic-ischemic damage: contribution of the cerebral
TE
vasculature to injury and repair? Front Physiol 3, 424.
Bagni, C., Greenough, W.T., 2005. From mRNP trafficking to spine dysmorphogenesis:
CE
P
the roots of fragile X syndrome. Nat. Rev. Neurosci. 6, 376–387.
Baj, G., Patrizio, A., Montalbano, A., Sciancalepore, M., Tongiorgi, E., 2014.
Developmental and maintenance defects in Rett syndrome neurons identified by a new
AC
mouse staging system in vitro. Front Cell Neurosci 8, 18.
Bangasser, D.A., Valentino, R.J., 2014. Sex differences in stress-related psychiatric
disorders: neurobiological perspectives. Front Neuroendocrinol 35, 303–319.
Barretto, R.P.J., Schnitzer, M.J., 2012. In Vivo Optical Microendoscopy for Imaging
Cells Lying Deep within Live Tissue. Cold Spring Harb Protoc 2012, 071464.
Barria, A., Malinow, R., 2005. NMDA receptor subunit composition controls synaptic
plasticity by regulating binding to CaMKII. Neuron 48, 289–301.
Bassell, G.J., Warren, S.T., 2008. Fragile X syndrome: loss of local mRNA regulation
alters synaptic development and function. Neuron 60, 201–214.
ACCEPTED MANUSCRIPT
Bastrikova, N., Gardner, G.A., Reece, J.M., Jeromin, A., Dudek, S.M., 2008. Synapse
elimination accompanies functional plasticity in hippocampal neurons. PNAS 105, 3123–
3127.
Bauman, M.L., Kemper, T.L., Arin, D.M., 1995. Pervasive neuroanatomic abnormalities
of the brain in three cases of Rett’s syndrome. Neurology 45, 1581–1586.
Belichenko, P.V., Brown, D., Jeffrey, M., Fraser, J.R., 2000. Dendritic and synaptic
alterations
of
hippocampal
pyramidal
neurones
in
scrapie‐ infected
mice.
RI
PT
Neuropathology and applied neurobiology 26, 143–149.
Belichenko, P.V., Kleschevnikov, A.M., Masliah, E., Wu, C., Takimoto-Kimura, R.,
Salehi, A., Mobley, W.C., 2009. Excitatory-inhibitory relationship in the fascia dentata
SC
in the Ts65Dn mouse model of Down syndrome. J. Comp. Neurol. 512, 453–466.
Belichenko, P.V., Kleschevnikov, A.M., Salehi, A., Epstein, C.J., Mobley, W.C., 2007.
NU
Synaptic and cognitive abnormalities in mouse models of Down syndrome: exploring
genotype-phenotype relationships. J. Comp. Neurol. 504, 329–345.
Belichenko, P. V., Oldfors, A., Hagberg, B., Dahlström, A., 1994. Rett syndrome: 3-D
MA
confocal microscopy of cortical pyramidal dendrites and afferents. Neuroreport 5, 1509–
1513.
Belichenko, Pavel V., Sourander, P., Malmgren, K., Nordborg, C., von Essen, C.,
morphology in epileptogenic cortex from TRPE patients, revealed by
TE
Dendritic
D
Rydenhag, B., Lindström, S., Hedström, A., Uvebrant, P., Dahlström, A., 1994.
intracellular Lucifer Yellow microinjection and confocal laser scanning microscopy.
CE
P
Epilepsy Research 18, 233–247.
Beltrán-Campos, V., Prado-Alcalá, R.A., León-Jacinto, U., Aguilar-Vázquez, A.,
Quirarte, G.L., Ramírez-Amaya, V., Díaz-Cintra, S., 2011. Increase of mushroom spine
AC
density in CA1 apical dendrites produced by water maze training is prevented by
ovariectomy. Brain Research 1369, 119–130.
Beier, K.T., Saunders, A., 2011. Anterograde or retrograde transsynaptic labeling of
CNS neurons with vesicular stomatitis virus vectors. Proc. Natl. Acad. Sci. USA.
108,15414–15419.
Berndt, A., Schoenenberger, P., Mattis, J., Tye, K.M., Deisseroth, K., Hegemann, P.,
Oertner, T.G., 2011. High-efficiency channelrhodopsins for fast neuronal stimulation at
low light levels. Proc. Natl Acad. Sci. USA 108, 7595–7600.
Berndt, A., Yizhar, O., Gunaydin, L.A., Hegemann, P., Deisseroth, K., 2009. Bi-stable
neural state switches. Nature Neurosci. 12, 229–234.
ACCEPTED MANUSCRIPT
Busskamp, V., Duebel, J., Balya, D., Fradot, M., Viney, T.J., Siegert, S., Groner, A.C.,
Cabuy, E., Forster, V., Seeliger, M., Biel, M., Humphries, P., Paques, M., Mohand-Said,
S., Trono, D., Deisseroth, K., Sahel, J.A., Picaud, S., Roska, B., 2010. Genetic
reactivation of cone photoreceptors restores visual responses in retinitis pigmentosa.
Science. 329,413-7.
Benavides-Piccione, R., Fernaud-Espinosa, I., Robles, V., Yuste, R., DeFelipe, J., 2013.
RI
PT
Age-Based Comparison of Human Dendritic Spine Structure Using Complete ThreeDimensional Reconstructions. Cereb Cortex 23, 1798–1810.
Berning, S., Willig, K.I., Steffens, H., Dibaj, P., Hell, S.W., 2012. Nanoscopy in a
Living Mouse Brain. Science 335, 551–551.
Bertram, L., Tanzi, R.E., 2005. The genetic epidemiology of neurodegenerative disease.
SC
J Clin Invest 115, 1449–1457.
Betzig, E., Patterson, G.H., Sougrat, R., Lindwasser, O.W., Olenych, S., Bonifacino,
NU
J.S., Davidson, M.W., Lippincott-Schwartz, J., Hess, H.F., 2006. Imaging Intracellular
MA
Fluorescent Proteins at Nanometer Resolution. Science 313, 1642–1645.
Bhatt, D.H., Zhang, S., Gan, W.-B., 2009. Dendritic spine dynamics. Annu. Rev. Physiol.
71, 261–282.
Blanchoin, L., Boujemaa-Paterski, R., Sykes, C., Plastino, J., 2014. Actin Dynamics,
D
TE
Architecture, and Mechanics in Cell Motility. Physiological Reviews 94, 235–263.
Bliss, T.V., Lomo, T., 1973. Long-lasting potentiation of synaptic transmission in the
CE
P
dentate area of the anaesthetized rabbit following stimulation of the perforant path. J.
Physiol. (Lond.) 232, 331–356.
Bliss, T.V.P., Collingridge, G.L., 1993. A synaptic model of memory: long-term
potentiation in the hippocampus. Nature 361, 31–39.
Bloodgood, B.L., Sabatini, B.L., 2005. Neuronal activity regulates diffusion across the
AC
neck of dendritic spines. Science 310, 866–869.
Bolshakov, V.Y., Golan, H., Kandel, E.R., Siegelbaum, S.A., 1997. Recruitment of New
Sites of Synaptic Transmission During the cAMP-Dependent Late Phase of LTP at
CA3–CA1 Synapses in the Hippocampus. Neuron 19, 635–651.
Bolshakov, V.Y., Siegelbaum, S.A., 1995. Regulation of hippocampal transmitter
release during development and long-term potentiation. Science 269, 1730–1734.
Bosch, M., Hayashi, Y., 2012. Structural plasticity of dendritic spines. Curr. Opin.
Neurobiol. 22, 383–388.
ACCEPTED MANUSCRIPT
Bourne, J.N., Harris, K.M., 2012. Nanoscale analysis of structural synaptic plasticity.
Current Opinion in Neurobiology, Synaptic structure and function 22, 372–382.
Bourne, J.N., Harris, K.M., 2008. Balancing structure and function at hippocampal
dendritic spines. Annu. Rev. Neurosci. 31, 47–67.
Bozdagi, O., Shan, W., Tanaka, H., Benson, D.L., Huntley, G.W., 2000. Increasing
Numbers of Synaptic Puncta during Late-Phase LTP. Neuron 28, 245–259.
Braidy, N., Guillemin, G.J., Mansour, H., Chan-Ling, T., Poljak, A., Grant, R., 2011.
RI
PT
Age Related Changes in NAD+ Metabolism Oxidative Stress and Sirt1 Activity in
Wistar Rats. PLOS ONE 6, e19194.
Branco, T., Clark, B.A., Häusser, M., 2010. Dendritic Discrimination of Temporal Input
Branco, T., Häusser, M., 2011. Synaptic Integration Gradients in Single Cortical
NU
Pyramidal Cell Dendrites. Neuron 69, 885–892.
SC
Sequences in Cortical Neurons. Science 329, 1671–1675.
Branco, T., Staras, K., Darcy, K.J., Goda, Y., 2008. Local dendritic activity sets release
probability at hippocampal synapses. Neuron 59, 475–485
Bremner, J.D., Narayan, M., Anderson, E.R., Staib, L.H., Miller, H.L., Charney, D.S.,
MA
2000. Hippocampal volume reduction in major depression. Am J Psychiatry 157, 115–
118.
Brigman, J.L., Wright, T., Talani, G., Prasad-Mulcare, S., Jinde, S., Seabold, G.K.,
D
TE
Mathur, P., Davis, M.I., Bock, R., Gustin, R.M., Colbran, R.J., Alvarez, V.A.,
Nakazawa, K., Delpire, E., Lovinger, D.M., Holmes, A., 2010. Loss of GluN2B-
CE
P
containing NMDA receptors in CA1 hippocampus and cortex impairs long-term
depression, reduces dendritic spine density, and disrupts learning. J. Neurosci. 30, 4590–
4600.
Brown, C.E., Wong, C., Murphy, T.H., 2008. Rapid Morphologic Plasticity of Peri-
AC
Infarct Dendritic Spines After Focal Ischemic Stroke. Stroke 39, 1286–1291.
Brown, D., Belichenko, P., Sales, J., Jeffrey, M., Fraser, J.R., 2001. Early loss of
dendritic spines in murine scrapie revealed by confocal analysis. NeuroReport 12, 179.
Bruel-Jungerman, E., Laroche, S., Rampon, C., 2005. New neurons in the dentate gyrus
are involved in the expression of enhanced long-term memory following environmental
enrichment. Eur. J. Neurosci. 21, 513–521.
Brundin, P., Melki, R., Kopito, R., 2010. Prion-like transmission of protein aggregates in
neurodegenerative diseases. Nat. Rev. Mol. Cell Biol. 11, 301–307.
ACCEPTED MANUSCRIPT
Bryan, G.K., Riesen, A.H., 1989. Deprived somatosensory-motor experience in
stumptailed monkey neocortex: Dendritic spine density and dendritic branching of layer
IIIB pyramidal cells. J. Comp. Neurol. 286, 208–217.
Bundman, M.C., Pico, R.M., Gall, C.M., 1994. Ultrastructural plasticity of the dentate
gyrus granule cells following recurrent limbic seizures: I. Increase in somatic spines.
Hippocampus 4, 601–610.
Cai, D., Cohen, K.B., Luo, T., Lichtman, J.W., Sanes, J.R., 2013. Improved tools for the
RI
PT
Brainbow toolbox. Nat. Methods 10, 540–547.
Callaway, E.M., 2008. Transneuronal circuit tracing with neurotropic viruses. Curr.
Opin. Neurobiol. 18, 617–623
Callaway, X.E.M., Luo, L., 2015. Monosynaptic Circuit Tracing with Glycoprotein-
SC
Deleted Rabies Viruses. J. Neurosci. 35,8979–8985.
Castle, M.J, Gershenson, Z.T., Giles, A.R., Holzbaur, E.L.F., Wolfe, J.H., 2014. Adeno-
NU
associated virus serotypes 1, 8, and 9 share conserved mechanisms for anterograde and
retrograde axonal transport. Hum. Gene Ther. 25,705–720.
Castle, M.J., Perlson, E, Holzbaur, E.L., Wolfe, J.H., 2013. Long-distance Axonal
MA
Transport of AAV9 Is Driven by Dynein and Kinesin-2 and Is Trafficked in a Highly
Caetano, S.C., Hatch, J.P., Brambilla, P., Sassi, R.B., Nicoletti, M., Mallinger, A.G.,
TE
D
Motile Rab7-positive Compartment. Mol. Ther. 22,1–13
Frank, E., Kupfer, D.J., Keshavan, M.S., Soares, J.C., 2004. Anatomical MRI study of
CE
P
hippocampus and amygdala in patients with current and remitted major depression.
Psychiatry Res 132, 141–147.
Calabrese, B., Wilson, M.S., Halpain, S., 2006. Development and Regulation of
Dendritic Spine Synapses. Physiology 21, 38–47.
Calabresi, P., Pisani, A., Mercuri, N.B., Bernardi, G., 1996. The corticostriatal
AC
projection: from synaptic plasticity to dysfunctions of the basal ganglia. Trends
Neurosci. 19, 19–24.
Caroni, P., Donato, F., Muller, D., 2012. Structural plasticity upon learning: regulation
and functions. Nat Rev Neurosci 13, 478–490.
Castro, J., Garcia, R.I., Kwok, S., Banerjee, A., Petravicz, J., Woodson, J., Mellios, N.,
Tropea, D., Sur, M., 2014. Functional recovery with recombinant human IGF1 treatment
in a mouse model of Rett Syndrome. Proc. Natl. Acad. Sci. U.S.A. 111, 9941–9946.
ACCEPTED MANUSCRIPT
Cepeda, C., Galvan, L., Holley, S.M., Rao, S.P., André, V.M., Botelho, E.P., Chen, J.Y.,
Watson, J.B., Deisseroth, K., Levine, M.S., 2013. Multiple sources of striatal inhibition
are differentially affected in Huntington's disease mouse models. J Neurosci. 33, 7393406.
Chamberlin, N.L., Du, B., de Lacalle, S., Saper, C.B., 1998. Recombinant adenoassociated virus vector: use for transgene expression and anterograde tract tracing in the
CNS. Brain Res. 793, 169–75.
Chow, B.Y., Han, X., Dobry, A.S., Qian, X., Chuong, A.S., Li, M., Henninger, M.A.,
RI
PT
Belfort, G.M., Lin, Y., Monahan, P.E., Boyden, E.S., 2010. High-performance
genetically targetable optical neural silencing by light-driven proton pumps. Nature 463,
SC
98–102.
Chao, H.-T., Zoghbi, H.Y., Rosenmund, C., 2007. MeCP2 controls excitatory synaptic
NU
strength by regulating glutamatergic synapse number. Neuron 56, 58–65.
Chapleau, C.A., Boggio, E.M., Calfa, G., Percy, A.K., Giustetto, M., Pozzo-Miller, L.,
MA
2012. Hippocampal CA1 Pyramidal Neurons of Mecp2 Mutant Mice Show a Dendritic
Spine Phenotype Only in the Presymptomatic Stage. Neural Plasticity.
Chapleau, C.A., Calfa, G.D., Lane, M.C., Albertson, A.J., Larimore, J.L., Kudo, S.,
D
Armstrong, D.L., Percy, A.K., Pozzo-Miller, L., 2009. Dendritic spine pathologies in
TE
hippocampal pyramidal neurons from Rett syndrome brain and after expression of Rettassociated MECP2 mutations. Neurobiol. Dis. 35, 219–233.
Chapleau, C.A., Carlo, M.E., Larimore, J.L., Pozzo-Miller, L., 2008. The actions of
CE
P
BDNF on dendritic spine density and morphology in organotypic slice cultures depend
on the presence of serum in culture media. Journal of Neuroscience Methods 169, 182–
190.
Chen, L.Y., Rex, C.S., Casale, M.S., Gall, C.M., Lynch, G., 2007. Changes in synaptic
AC
morphology accompany actin signaling during LTP. J. Neurosci. 27, 5363–5372.
Christoffel, D.J., Golden, S.A., Russo, S.J., 2011. Structural and synaptic plasticity in
stress-related disorders. Rev Neurosci 22, 535–549.
Collins, A.L., Levenson, J.M., Vilaythong, A.P., Richman, R., Armstrong, D.L.,
Noebels, J.L., David Sweatt, J., Zoghbi, H.Y., 2004. Mild overexpression of MeCP2
causes a progressive neurological disorder in mice. Hum. Mol. Genet. 13, 2679–2689.
Colman, R.J., Anderson, R.M., Johnson, S.C., Kastman, E.K., Kosmatka, K.J., Beasley,
T.M., Allison, D.B., Cruzen, C., Simmons, H.A., Kemnitz, J.W., Weindruch, R., 2009.
ACCEPTED MANUSCRIPT
Caloric Restriction Delays Disease Onset and Mortality in Rhesus Monkeys. Science
325, 201–204.
Comery, T.A., Harris, J.B., Willems, P.J., Oostra, B.A., Irwin, S.A., Weiler, I.J.,
Greenough, W.T., 1997. Abnormal dendritic spines in fragile X knockout mice:
Maturation and pruning deficits. PNAS 94, 5401–5404.
Cook, D., Nuro, E., Murai, K.K., 2014. Increasing our understanding of human cognition
through the study of Fragile X Syndrome. Dev Neurobiol 74, 147–177.
Corbett, D., Giles, T., Evans, S., McLean, J., Biernaskie, J., 2006. Dynamic changes in
RI
PT
CA1 dendritic spines associated with ischemic tolerance. Experimental Neurology 202,
133–138.
Cotman, C.W., Berchtold, N.C., Christie, L.-A., 2007. Exercise builds brain health: key
SC
roles of growth factor cascades and inflammation. Trends Neurosci. 30, 464–472.
Courchesne, E., Pierce, K., Schumann, C.M., Redcay, E., Buckwalter, J.A., Kennedy,
NU
D.P., Morgan, J., 2007. Mapping early brain development in autism. Neuron 56, 399–
413.
Critchlow, H.M., Maycox, P.R., Skepper, J.N., Krylova, O., 2006. Clozapine and
MA
haloperidol differentially regulate dendritic spine formation and synaptogenesis in rat
hippocampal neurons. Mol Cell Neurosci 32, 356–365.
Cruz-Martín, A., Crespo, M., Portera-Cailliau, C., 2010. Delayed Stabilization of
D
TE
Dendritic Spines in Fragile X Mice. J. Neurosci. 30, 7793–7803.
Cruz-Sánchez, F.F., Cardozo, A., Tolosa, E., 1995. Neuronal Changes in the Substantia
CE
P
Nigra with Aging: A Golgi Study. J Neuropathol Exp Neurol 54, 74–81.
Dagytė, G., Crescente, I., Postema, F., Seguin, L., Gabriel, C., Mocaër, E., Boer, J.A.D.,
Koolhaas, J.M., 2011. Agomelatine reverses the decrease in hippocampal cell survival
AC
induced by chronic mild stress. Behavioural Brain Research 218, 121–128.
Dagyte, G., Trentani, A., Postema, F., Luiten, P.G., Den Boer, J.A., Gabriel, C., Mocaër,
E., Meerlo, P., Van der Zee, E.A., 2010. The novel antidepressant agomelatine
normalizes hippocampal neuronal activity and promotes neurogenesis in chronically
stressed rats. CNS Neurosci Ther 16, 195–207.
Dailey, M.E., Smith, S.J., 1996. The Dynamics of Dendritic Structure in Developing
Hippocampal Slices. J. Neurosci. 16, 2983–2994.
Dalva, M.B., McClelland, A.C., Kayser, M.S., 2007. Cell adhesion molecules: signalling
functions at the synapse. Nat. Rev. Neurosci. 8, 206–220.
ACCEPTED MANUSCRIPT
Dauer, W., Przedborski, S., 2003. Parkinson’s disease: mechanisms and models. Neuron
39, 889–909.
Davis, K.F., Parker, K.P., Montgomery, G.L., 2004. Sleep in infants and young children:
Part two: common sleep problems. Journal of Pediatric Health Care 18, 130–137.
Dayan, E., Cohen, L.G., 2011. Neuroplasticity subserving motor skill learning. Neuron
72, 443–454.
de Kloet, A.D., Liu, M., Rodríguez, V., Krause, E.G., Sumners, C., 2015. Role of
RI
PT
neurons and glia in the CNS actions of the renin-angiotensin system in cardiovascular
control. Am. J. Physiol. Regul. Integr. Comp. Physiol. 309, R444-458.
de Bodinat, C., Guardiola-Lemaitre, B., Mocaër, E., Renard, P., Muñoz, C., Millan, M.J.,
development. Nat Rev Drug Discov. 9, 628-42.
De Rubeis, S., Bagni, C., 2010. Fragile X mental retardation protein control of neuronal
NU
SC
2010. Agomelatine, the first melatonergic antidepressant: discovery, characterization and
mRNA metabolism: Insights into mRNA stability. Mol. Cell. Neurosci. 43, 43–50.
Dean, C., Dresbach, T., 2006. Neuroligins and neurexins: linking cell adhesion, synapse
MA
formation and cognitive function. Trends Neurosci. 29, 21–29.
DeKosky, S.T., Scheff, S.W., 1990. Synapse loss in frontal cortex biopsies in
Demir Özkay, Ü., Söztutar, E., Can, Ö.D., Üçel, U.İ., Öztürk, Y., Ulupinar, E., 2015.
Effects
TE
D
Alzheimer’s disease: correlation with cognitive severity. Ann. Neurol. 27, 457–464.
of long-term agomelatine
treatment
on the cognitive performance
and
CE
P
hippocampal plasticity of adult rats. Behav Pharmacol 26, 469–480.
Denk, W., Strickler, J.H., Webb, W.W., 1990. Two-photon laser scanning fluorescence
microscopy. Science 248, 73–76.
Desmond, N.L., Levy, W.B., 1988. Synaptic interface surface area increases with long-
AC
term potentiation in the hippocampal dentate gyrus. Brain research 453, 308–314.
Desmond, N.L., Levy, W.B., 1986. Changes in the postsynaptic density with long-term
potentiation in the dentate gyrus. J. Comp. Neurol. 253, 476–482.
DeNoble, V.J., 1987. Vinpocetine enhances retrieval of a step-through passive avoidance
response in rats. Pharmacol Biochem Behav. 26,183–186.
Deshmukh, R., Sharma, V., Mehan, S., Sharma, N., Bedi, K.L., 2009. Amelioration of
intracerebroventricular streptozotocin induced cognitive dysfunction and oxidative stress
by vinpocetine - a PDE1 inhibitor. Eur J Pharmacol. 620,49–56.
ACCEPTED MANUSCRIPT
DeFalco, J., Tomishima, M., Liu, H., Zhao, C., Cai, X., Marth, J.D., Enquist, L.,
Friedman, J.M., 2001. Virus-Assisted Mapping of Neural Inputs to a Feeding Center in
the Hypothalamus. Science 291, 2608-13
Deisseroth, K., 2015. Optogenetics: 10 years of microbial opsins in neuroscience. Nat
Neurosci. 18, 1213–1225.
Deisseroth, K., Feng, G., Majewska, A.K., Miesenbock, G., Ting, A., Schnitzer, M.J.,
2006.
Next-Generation Optical Technologies for Illuminating Genetically Targeted
RI
PT
Brain Circuits. J. Neurosci. 26, 10380-6.
Desai, N.S., Cudmore, R.H., Nelson, S.B., Turrigiano, G.G., 2002. Critical periods for
experience-dependent synaptic scaling in visual cortex. Nat. Neurosci. 5, 783–789.
Dietzschold, B., Li, J., Faber, M., Schnell, M., 2008. Concepts in the pathogenesis of
SC
rabies. Future Virol. 3, 481–490.
Diamond, M.C., Lindner, B., Johnson, R., Bennett, E.L., Rosenzweig, M.R., 1975.
NU
Differences in occipital cortical synapses from environmentally enriched, impoverished,
MA
and standard colony rats. J. Neurosci. Res. 1, 109–119.
Dickstein, D.L., Weaver, C.M., Luebke, J.I., Hof, P.R., 2013. Dendritic spine changes
associated with normal aging. Neuroscience 251, 21–32.
Dodrill, C.B., 2002. Progressive cognitive decline in adolescents and adults with
D
TE
epilepsy. Prog. Brain Res. 135, 399–407.
Dorris, L., Scott, N., Zuberi, S., Gibson, N., Espie, C., 2008. Sleep problems in children
CE
P
with neurological disorders. Developmental neurorehabilitation 11, 95–114.
Drakew, A., Mu¨ller, M., Ga¨hwiler, B.H., Thompson, S.M., Frotscher, M., 1996. Spine
loss in experimental epilepsy: Quantitative light and electron microscopic analysis of
intracellularly stained CA3 pyramidal cells in hippocampal slice cultures. Neuroscience
AC
70, 31–45.
Drevets, W.C., Price, J.L., Furey, M.L., 2008. Brain structural and functional
abnormalities in mood disorders: implications for neurocircuitry models of depression.
Brain Struct Funct 213, 93–118.
Duan, H., Wearne, S.L., Rocher, A.B., Macedo, A., Morrison, J.H., Hof, P.R., 2003.
Age-related Dendritic and Spine Changes in Corticocortically Projecting Neurons in
Macaque Monkeys. Cereb Cortex 13, 950–961.
Dumitriu, D., Hao, J., Hara, Y., Kaufmann, J., Janssen, W.G.M., Lou, W., Rapp, P.R.,
Morrison, J.H., 2010. Selective changes in thin spine density and morphology in monkey
ACCEPTED MANUSCRIPT
prefrontal cortex correlate with aging-related cognitive impairment. J. Neurosci. 30,
7507–7515.
Dunaevsky, A., Tashiro, A., Majewska, A., Mason, C., Yuste, R., 1999. Developmental
regulation of spine motility in the mammalian central nervous system. Proc Natl Acad
Sci U S A 96, 13438–13443.
Edward, M., Callaway., 2008. Transneuronal circuit tracing with neurotropic viruses.
Curr Opin Neurobiol. 18, 617–623.
Ekstrand,
M.I.,
Enquist,
L.W., Pomeranz, L.E., 2008. The alpha-herpesviruses:
RI
PT
molecular pathfinders in nervous system circuits. Trends Mol. Med. 14, 134–140
Eilam-Stock, T., Serrano, P., Frankfurt, M., Luine, V., 2012. Bisphenol-A Impairs
SC
Memory and Reduces Dendritic Spine Density in Adult Male Rats. Behav Neurosci 126,
175–185.
Elger, C.E., Helmstaedter, C., Kurthen, M., 2004. Chronic epilepsy and cognition.
NU
Lancet Neurol 3, 663–672.
Elston, G.N., 2003. Cortex, cognition and the cell: new insights into the pyramidal
MA
neuron and prefrontal function. Cereb. Cortex 13, 1124–1138.
Elvevåg, B., Goldberg, T.E., 2000. Cognitive impairment in schizophrenia is the core of
Engert, F., Bonhoeffer, T., 1999. Dendritic spine changes associated with hippocampal
TE
D
the disorder. Crit Rev Neurobiol 14, 1–21.
long-term synaptic plasticity. Nature 399, 66–70.
Enright, L.E., Zhang, S., Murphy, T.H., 2007. Fine Mapping of the Spatial Relationship
CE
P
between Acute Ischemia and Dendritic Structure Indicates Selective Vulnerability of
Layer V Neuron Dendritic Tufts within Single Neurons in Vivo. J Cereb Blood Flow
Metab 27, 1185–1200.
Essa, M.M., Subash, S., Akbar, M., Al-Adawi, S., Guillemin, G.J., 2015. Long-Term
AC
Dietary Supplementation of Pomegranates, Figs and Dates Alleviate Neuroinflammation
in a Transgenic Mouse Model of Alzheimer’s Disease. PLoS One 10.
Evans-Galea, M.V., Hannan, A.J., Carrodus, N., Delatycki, M.B., Saffery, R., 2013.
Epigenetic modifications in trinucleotide repeat diseases. Trends in Molecular Medicine
19, 655–663.
Falkenberg, T., Mohammed, A.K., Henriksson, B., Persson, H., Winblad, B., Lindefors,
N., 1992. Increased expression of brain-derived neurotrophic factor mRNA in rat
ACCEPTED MANUSCRIPT
hippocampus is associated with improved spatial memory and enriched environment.
Neurosci. Lett. 138, 153–156.
Ferrante, R.J., Kowall, N.W., Richardson, E.P., 1991. Proliferative and degenerative
changes in striatal spiny neurons in Huntington’s disease: a combined study using the
section-Golgi method and calbindin D28k immunocytochemistry. J. Neurosci. 11, 3877–
3887.
Ferrer, I., Gullotta, F., 1990. Down’s syndrome and Alzheimer’s disease: dendritic spine
RI
PT
counts in the hippocampus. Acta Neuropathol. 79, 680–685.
Fiala, J.C., Spacek, J., Harris, K.M., 2002. Dendritic spine pathology: cause or
consequence of neurological disorders? Brain Res. Brain Res. Rev. 39, 29–54.
Fifková, E., Anderson, C.L., 1981. Stimulation-induced changes in dimensions of stalks
SC
of dendritic spines in the dentate molecular layer. Experimental Neurology 74, 621–627.
NU
https://doi.org/10.1016/0014-4886(81)90197-7
Filgueiras, C.C., Krahe, T.E., Medina, A.E., 2010. Phosphodiesterase type 1 inhibition
improves learning in rats exposed to alcohol during the third trimester equivalent of
MA
human gestation. Neurosci Lett. 473,202–207.
Figge, D.A., Rahman, I., Dougherty, P.J., Rademacher, D.J., 2013. Retrieval of
D
contextual memories increases activity-regulated cytoskeleton-associated protein in the
TE
amygdala and hippocampus. Brain Struct Funct 218, 1177–1196.
Fischer, M., Kaech, S., Wagner, U., Brinkhaus, H., Matus, A., 2000. Glutamate
CE
P
receptors regulate actin-based plasticity in dendritic spines. Nat. Neurosci. 3, 887–894.
Fratiglioni, L., Paillard-Borg, S., Winblad, B., 2004. An active and socially integrated
lifestyle in late life might protect against dementia. Lancet Neurol 3, 343–353.
Fu, M., Yu, X., Lu, J., Zuo, Y., 2012. Repetitive motor learning induces coordinated
of
AC
formation
clustered
dendritic
spines
in
vivo.
Nature
483,
92–95.
https://doi.org/10.1038/nature10844
Fujita, A., Kurachi, Y., 2000. SAP Family Proteins. Biochemical and Biophysical
Research Communications 269, 1–6.
Fox, K., Stryker, M., 2017. Integrating Hebbian and homeostatic plasticity: introduction.
Philos Trans R Soc Lond B Biol Sci. 372, 20160413.
Furuta, T., Tomioka, R., Taki, K., Nakamura, K., Tamamaki, N., Kaneko, T., 2001. In
Vivo Transduction of Central Neurons Using Recombinant Sindbis Virus: Golgi-like
ACCEPTED MANUSCRIPT
Labeling of Dendrites and Axons with Membrane-targeted Fluorescent Proteins. The
Journal of Histochemistry and Cytochemistry. 49, 1497–1507.
Fukazawa, Y., Saitoh, Y., Ozawa, F., Ohta, Y., Mizuno, K., Inokuchi, K., 2003.
Hippocampal LTP Is Accompanied by Enhanced F-Actin Content within the Dendritic
Spine that Is Essential for Late LTP Maintenance In Vivo. Neuron 38, 447–460.
Fukuda, T., Itoh, M., Ichikawa, T., Washiyama, K., Goto, Y., 2005. Delayed maturation
RI
PT
of neuronal architecture and synaptogenesis in cerebral cortex of Mecp2-deficient mice.
J. Neuropathol. Exp. Neurol. 64, 537–544.
Garcia-Alloza, M., Dodwell, S.A., Meyer-Luehmann, M., Hyman, B.T., Bacskai, B.J.,
2006. Plaque-Derived Oxidative Stress Mediates Distorted Neurite Trajectories in the
SC
Alzheimer Mouse Model. J Neuropathol Exp Neurol 65, 1082–1089.
García-López, P., García-Marín V., Freire, M., 2010. Dendritic Spines and Development:
NU
Towards a Unifying Model of Spinogenesis—A Present Day Review of Cajal's
Histological Slides and Drawings. Neural Plast. 769
Garey, L.J., Ong, W.Y., Patel, T.S., Kanani, M., Davis, A., Mortimer, A.M., Barnes,
MA
T.R., Hirsch, S.R., 1998. Reduced dendritic spine density on cerebral cortical pyramidal
neurons in schizophrenia. J. Neurol. Neurosurg. Psychiatr. 65, 446–453.
Geinisman, Y., 2000. Structural synaptic modifications associated with hippocampal
D
TE
LTP and behavioral learning. Cereb. Cortex 10, 952–962.
Geinisman, Y., Morrell, F., de Toledo-Morrell, L., 1989. Perforated synapses on double-
CE
P
headed dendritic spines: a possible structural substrate of synaptic plasticity. Brain
Research 480, 326–329.
Geschwind, M.D., 2015. Prion diseases. Continuum (Minneapolis, Minn.) 21, 1612.
Giannone, G., Hosy, E., Levet, F., Constals, A., Schulze, K., Sobolevsky, A.I., Rosconi,
AC
M.P., Gouaux, E., Tampé, R., Choquet, D., Cognet, L., 2010. Dynamic Superresolution
Imaging of Endogenous Proteins on Living Cells at Ultra-High Density. Biophysical
Journal 99, 1303–1310.
Gipson, C.D., Olive, M.F., 2017. Structural and functional plasticity of dendritic spines root or result of behavior? Genes Brain Behav. 16, 101–117.
Ginger, M., Haberl, M., Conzelmann, K.K., Schwarz, M.K., Frick, A., 2013. Revealing
the secrets of neuronal circuits with recombinant rabies virus technology. Front Neural
Circuits. 7, 2.
ACCEPTED MANUSCRIPT
Gobbo, F., Marchetti, L., Alia, C., Luin, S., Cattaneo, A., 2016. Activity-dependent
expression of reporter proteins at dendritic spines for synaptic activity mapping and
optogenetic stimulation. BioXRiv.
Gradinaru, V., Zhang, F., Ramakrishnan, C., Mattis, J., Prakash, R., Diester, I., Goshen,
I., Thompson, K.R., Deisseroth, K., 2010. Molecular and cellular approaches for
diversifying and extending optogenetics. Cell 141, 154–165.
Gradinaru, V., Zhang, F., Ramakrishnan, C., Mattis, J., Prakash, R., Diester, I., Goshen,
RI
PT
I., Thompson, K.R., Deisseroth, K., 2010. Molecular and cellular approaches for
diversifying and extending optogenetics. Cell. 141,154–165.
Gunaydin, L.A., Yizhar, O., Berndt, A., Sohal, V.S., Deisseroth, K., Hegemann, P.,
SC
2010. Ultrafast optogenetic control. Nature Neurosci. 13, 387–392.
Gitler, A.D., 2011. Neuroscience. Another reason to exercise. Science 334, 606–607.
Glantz, L.A., Lewis, D.A., 2001. Dendritic spine density in schizophrenia and
NU
depression. Arch. Gen. Psychiatry 58, 203.
Glaser, J.R., Glaser, E.M., 1990. Neuron imaging with neurolucida — A PC-based
MA
system for image combining microscopy. Computerized Medical Imaging and Graphics,
307–317.
Glausier,
J.R.,
Lewis,
D.A.,
2013.
Dendritic spine pathology in schizophrenia.
TE
D
Progress in Imaging in the Neurosciences Using Microcomputers and Workstations 14,
Neuroscience 251, 90–107.
Globus, A., Rosenzweig, M.R., Bennett, E.L., Diamond, M.C., 1973. Effects of
CE
P
differential experience on dendritic spine counts in rat cerebral cortex. Journal of
comparative and physiological psychology 82, 175.
Gohla, A., Bokoch, G.M., 2002. 14-3-3 Regulates Actin Dynamics by Stabilizing
AC
Phosphorylated Cofilin. Current Biology 12, 1704–1710.
Goldin, M., Segal, M., Avignone, E., 2001. Functional Plasticity Triggers Formation and
Pruning of Dendritic Spines in Cultured Hippocampal Networks. J. Neurosci. 21, 186–
193.
Gómez-Pinilla, F., 2008. Brain foods: the effects of nutrients on brain function. Nat Rev
Neurosci 9, 568–578.
Gobbo, F., Marchetti, L., Alia, C., Luin, S., Cattaneo, A., 2016. Activity-dependent
expression of reporter proteins at dendritic spines for synaptic activity mapping and
optogenetic stimulation. BioXRiv.
ACCEPTED MANUSCRIPT
Gonzales, M.L., LaSalle, J.M., 2010. The role of MeCP2 in brain development and
neurodevelopmental disorders. Curr Psychiatry Rep 12, 127–134.
González-Burgos, I., Velázquez-Zamora, D.A., Beas-Zárate, C., n.d. Damage and
plasticity in adult rat hippocampal trisynaptic circuit neurons after neonatal exposure to
glutamate excitotoxicity. International Journal of Developmental Neuroscience 27, 741–
745.
Govindarajan, A., Israely, I., Huang, S.-Y., Tonegawa, S., 2011. The Dendritic Branch Is
RI
PT
the Preferred Integrative Unit for Protein Synthesis-Dependent LTP. Neuron 69, 132–
146.
Graveland, G.A., Williams, R.S., DiFiglia, M., 1985. Evidence for degenerative and
SC
regenerative changes in neostriatal spiny neurons in Huntington’s disease. Science 227,
770–773.
Gray, E.G., Guillery, R.W., 1963. A note on the dendritic spine apparatus. J Anat 97,
NU
389-392.3.
Gray, N.W., Weimer, R.M., Bureau, I., Svoboda, K., 2006. Rapid Redistribution of
MA
Synaptic PSD-95 in the Neocortex In Vivo. PLOS Biology 4, e370.
Greenough, W.T., 1975. Experiential modification of the developing brain. Am. Sci. 63,
37–46.
Greenough, W.T., Hwang, H.M., Gorman, C., 1985. Evidence for active synapse
D
TE
formation or altered postsynaptic metabolism in visual cortex of rats reared in complex
environments. Proc. Natl. Acad. Sci. U.S.A. 82, 4549–4552.
Grønli, J., Soulé, J., Bramham, C.R., 2014. Sleep and protein synthesis-dependent
CE
P
synaptic plasticity: impacts of sleep loss and stress. Front Behav Neurosci 7.
Grossman, L.S., Harrow, M., Rosen, C., Faull, R., Strauss, G.P., 2008. Sex differences in
AC
schizophrenia and other psychotic disorders: a 20-year longitudinal study of psychosis
and recovery. Compr Psychiatry 49, 523–529.
Grunditz, Å., Holbro, N., Tian, L., Zuo, Y., Oertner, T.G., 2008. Spine Neck Plasticity
Controls Postsynaptic Calcium Signals through Electrical Compartmentalization. J.
Neurosci. 28, 13457–13466.
Gruss, M., Appenroth, D., Flubacher, A., Enzensperger, C., Bock, J., Fleck, C., Gille, G.,
Braun, K., 2012. 9-Methyl-β-carboline-induced cognitive enhancement is associated
with elevated hippocampal dopamine levels and dendritic and synaptic proliferation.
Journal of Neurochemistry 121, 924–931.
ACCEPTED MANUSCRIPT
Grutzendler, J., Kasthuri, N., Gan, W.-B., 2002. Long-term dendritic spine stability in
the adult cortex. Nature 420, 812–816.
Gu, M., Bao, H., Kang, H., 2014. Fibre-optical microendoscopy. Journal of Microscopy
254, 13–18.
Gualtieri, C.T., Morgan, D.W., 2008. The frequency of cognitive impairment in patients
with anxiety, depression, and bipolar disorder: an unaccounted source of variance in
clinical trials. J Clin Psychiatry 69, 1122–1130.
Guidetti, P., Charles, V., Chen, E.-Y., Reddy, P.H., Kordower, J.H., Whetsell, W.O.,
RI
PT
Schwarcz, R., Tagle, D.A., 2001. Early Degenerative Changes in Transgenic Mice
Expressing Mutant Huntingtin Involve Dendritic Abnormalities but No Impairment of
SC
Mitochondrial Energy Production. Experimental Neurology 169, 340–350.
Haas, M.A., Bell, D., Slender, A., Lana-Elola, E., Watson-Scales, S., Fisher, E.M.C.,
NU
Tybulewicz, V.L.J., Guillemot, F., 2013. Alterations to Dendritic Spine Morphology, but
Not Dendrite Patterning, of Cortical Projection Neurons in Tc1 and Ts1Rhr Mouse
Models of Down Syndrome. PLOS ONE 8, e78561.
Hagerman, R.J., Berry-Kravis, E., Kaufmann, W.E., Ono, M.Y., Tartaglia, N.,
MA
Lachiewicz, A., Kronk, R., Delahunty, C., Hessl, D., Visootsak, J., Picker, J., Gane, L.,
Tranfaglia, M., 2009. Advances in the treatment of fragile X syndrome. Pediatrics 123,
Halbower, A.C., Degaonkar, M., Barker, P.B., Earley, C.J., Marcus, C.L., Smith, P.L.,
TE
D
378–390.
Prahme, M.C., Mahone, E.M., 2006. Childhood obstructive sleep apnea associates with
CE
P
neuropsychological deficits and neuronal brain injury. PLoS Med. 3, e301.
Halpain, S., Hipolito, A., Saffer, L., 1998. Regulation of F-actin stability in dendritic
spines by glutamate receptors and calcineurin. J. Neurosci. 18, 9835–9844.
Hamann, J., Wernicke, C., Lehmann, J., Reichmann, H., Rommelspacher, H., Gille, G.,
2008.
AC
9-Methyl-beta-carboline
up-regulates
the
appearance
of
differentiated
dopaminergic neurones in primary mesencephalic culture. Neurochem. Int. 52, 688–700.
Hamilton, A.M., Oh, W.C., Vega-Ramirez, H., Stein, I.S., Hell, J.W., Patrick, G.N.,
Zito, K., 2012. Activity-Dependent Growth of New Dendritic Spines Is Regulated by the
Proteasome. Neuron 74, 1023–1030.
Hammer, G.D., McPhee, S.J., 2014. Pathophysiology of Disease: An Introduction to
Clinical Medicine 7/E. McGraw Hill Professional.
Harris, K.M., Jensen, F.E., Tsao, B., 1992. Three-dimensional structure of dendritic
spines and synapses in rat hippocampus (CA1) at postnatal day 15 and adult ages:
ACCEPTED MANUSCRIPT
implications for the maturation of synaptic physiology and long-term potentiation. J.
Neurosci. 12, 2685–2705.
Harris, K.M., Kater, S.B., 1994. Dendritic spines: cellular specializations imparting both
stability and flexibility to synaptic function. Annu. Rev. Neurosci. 17, 341–371.
Harris, K.M., Stevens, J.K., 1989. Dendritic spines of CA 1 pyramidal cells in the rat
hippocampus:
serial
electron
microscopy
with
reference
to
their
biophysical
characteristics. J. Neurosci. 9, 2982–2997.
Harvey, C.D., Svoboda, K., 2007. Locally dynamic synaptic learning rules in pyramidal
RI
PT
neuron dendrites. Nature 450, 1195–1200.
Harvey, R.J., Napper, R.M., 1988. Quantitative study of granule and Purkinje cells in the
SC
cerebellar cortex of the rat. J. Comp. Neurol. 274, 151–157.
Hasbani, M.J., Viquez, N.M., Goldberg, M.P., 2001. NMDA receptors mediate hypoxic
NU
spine loss in cultured neurons. Neuroreport 12, 2731–2735.
Harris, J.A., Wook Oh, S., Zeng. H., 2012. Adeno-associated viral vectors for
anterograde axonal tracing with fluorescent proteins in nontransgenic and Cre driver
MA
mice. Curr. Protoc. Neurosci.; 1:1–18.
Harvey, C.D., Yasuda, R., Zhong, H., Svoboda, K., 2008. The spread of Ras activity
triggered by activation of a single dendritic spine. Science 321, 136–140.
Hayashi-Takagi, A., Yagishita, S., Nakamura, M., Shirai, F., Wu, Y.I., Loshbaugh, A.L.,
D
TE
Kuhlman, B., Hahn, K.M., Kasai, H., 2015. Labelling and optical erasure of synaptic
memory traces in the motor cortex. Nature 525, 333–338
Hernandez, V.H., Gehrt, A., Zhizi, J., Gerhard, H., Jeschke, M., Strenzke, N., Tobias,
CE
P
M., 2014. Optogenetic stimulation of the auditory pathway. J Clin Invest. 124, 1114-29.
Hutson, T.H., Kathe, C., Moon, L.D.F., 2015. Trans-neuronal transduction of spinal
AC
neurons following cortical injection and anterograde axonal transport of a bicistronic
AAV1 vector. Gene Ther. 23, 231– 236.
Heilemann, M., 2010. Fluorescence microscopy beyond the diffraction limit. Journal of
Biotechnology, BioImaging - Contributions from Biology, Physics and Informatics 149,
243–251.
Heal, D.J., Smith, S.L., Gosden, J., Nutt, D.J., 2013. Amphetamine, past and present – a
pharmacological and clinical perspective. J Psychopharmacol 27, 479–496.
Hell, S.W., 2007. Far-Field Optical Nanoscopy. Science 316, 1153–1158.
ACCEPTED MANUSCRIPT
Hering, H., Sheng, M., 2001. Dendritic spines: structure, dynamics and regulation. Nat.
Rev. Neurosci. 2, 880–888.
Hickey, F., Hickey, E., Summar, K.L., 2012. Medical update for children with Down
syndrome for the pediatrician and family practitioner. Advances in pediatrics 59, 137–
157.
Hickie, I.B., Rogers, N.L., 2011. Novel melatonin-based therapies: potential advances in
RI
PT
the treatment of major depression. Lancet. 378, 621-31.
Hill, T.C., Zito, K., 2013. LTP-Induced Long-Term Stabilization of Individual Nascent
Dendritic Spines. J. Neurosci. 33, 678–686.
Hillman, C.H., Erickson, K.I., Kramer, A.F., 2008. Be smart, exercise your heart:
SC
exercise effects on brain and cognition. Nat. Rev. Neurosci. 9, 58–65.
Hinton, V.J., Brown, W.T., Wisniewski, K., Rudelli, R.D., 1991. Analysis of neocortex
NU
in three males with the fragile X syndrome. Am. J. Med. Genet. 41, 289–294.
Hofer, S.B., Mrsic-Flogel, T.D., Bonhoeffer, T., Hübener, M., 2009. Experience leaves a
MA
lasting structural trace in cortical circuits. Nature 457, 313–317.
Holtmaat, A., Svoboda, K., 2009. Experience-dependent structural synaptic plasticity in
the mammalian brain. Nat. Rev. Neurosci. 10, 647–658.
Holtmaat, A., Wilbrecht, L., Knott, G.W., Welker, E., Svoboda, K., 2006. Experience-
D
TE
dependent and cell-type-specific spine growth in the neocortex. Nature 441, 979–983.
Holtmaat, A.J.G.D., Trachtenberg, J.T., Wilbrecht, L., Shepherd, G.M., Zhang, X.,
CE
P
Knott, G.W., Svoboda, K., 2005. Transient and persistent dendritic spines in the
neocortex in vivo. Neuron 45, 279–291.
Holtzman, D.M., Morris, J.C., Goate, A.M., 2011. Alzheimer’s disease: the challenge of
the second century. Sci Transl Med 3, 77sr1.
Honkura, N., Matsuzaki, M., Noguchi, J., Ellis-Davies, G.C.R., Kasai, H., 2008. The
AC
subspine organization of actin fibers regulates the structure and plasticity of dendritic
spines. Neuron 57, 719–729.
Hori, N., Carpenter, D.O., 1994. Functional and Morphological Changes Induced by
Transient in Vivo Ischemia. Experimental Neurology 129, 279–289.
Hotulainen, P., Hoogenraad, C.C., 2010. Actin in dendritic spines: connecting dynamics
to function. J. Cell Biol. 189, 619–629.
ACCEPTED MANUSCRIPT
Hsieh, H., Boehm, J., Sato, C., Iwatsubo, T., Tomita, T., Sisodia, S., Malinow, R., 2006.
AMPAR removal underlies Abeta-induced synaptic depression and dendritic spine loss.
Neuron 52, 831–843.
Huang, B., Babcock, H., Zhuang, X., 2010. Breaking the diffraction barrier: superresolution imaging of cells. Cell 143, 1047–1058.
Husi, H., Ward, M.A., Choudhary, J.S., Blackstock, W.P., Grant, S.G., 2000. Proteomic
analysis of NMDA receptor-adhesion protein signaling complexes. Nat. Neurosci. 3,
RI
PT
661–669.
Hutsler, J.J., Zhang, H., 2010. Increased dendritic spine densities on cortical projection
neurons in autism spectrum disorders. Brain Res. 1309, 83–94.
Huttenlocher, P.R., 1974. Dendritic development in neocortex of children with mental
SC
defect and infantile spasms. Neurology 24, 203–203.
Ishihara, K., Katsuki, H., Sugimura, M., Satoh, M., 1989. Idebenone and vinpocetine
augment
long-term
potentiation
in
hippocampal
slices
in
the
guinea
pig.
MA
Neuropharmacology. 28,569–573.
NU
Ikebe, M., Hartshorne, D.J., 1985. Phosphorylation of smooth muscle myosin at two
distinct sites by myosin light chain kinase. J. Biol. Chem. 260, 10027–10031.
Ingham, C.A., Hood, S.H., Arbuthnott, G.W., 1989. Spine density on neostriatal
D
TE
neurones changes with 6-hydroxydopamine lesions and with age. Brain Research 503,
334–338.
Ingham, C.A., Hood, S.H., Taggart, P., Arbuthnott, G.W., 1998. Plasticity of Synapses
CE
P
in the Rat Neostriatum after Unilateral Lesion of the Nigrostriatal Dopaminergic
Pathway. J. Neurosci. 18, 4732–4743.
Irwin, S.A., Patel, B., Idupulapati, M., Harris, J.B., Crisostomo, R.A., Larsen, B.P.,
AC
Kooy, F., Willems, P.J., Cras, P., Kozlowski, P.B., Swain, R.A., Weiler, I.J., Greenough,
W.T., 2001. Abnormal dendritic spine characteristics in the temporal and visual cortices
of patients with fragile-X syndrome: a quantitative examination. Am. J. Med. Genet. 98,
161–167.
Ishikawa, Y., Katoh, H., Negishi, M., 2003. A role of Rnd1 GTPase in dendritic spine
formation in hippocampal neurons. J. Neurosci. 23, 11065–11072.
Isokawa, M., 2000. Remodeling dendritic spines of dentate granule cells in temporal
lobe epilepsy patients and the rat pilocarpine model. Epilepsia 41(suppl. 6).
ACCEPTED MANUSCRIPT
Isokawa, M., 1998. Remodeling dendritic spines in the rat pilocarpine model of temporal
lobe epilepsy. Neurosci Lett 258, 73–76.
Isokawa, M., 1997. Preservation of dendrites with the presence of reorganized mossy
fiber collaterals in hippocampal dentate granule cells in patients with temporal lobe
epilepsy. Brain Res 744, 339–343.
Isokawa, M., Levesque, M.F., 1991. Increased NMDA responses and dendritic
degeneration in human epileptic hippocampal neurons in slices. Neuroscience Letters
RI
PT
132, 212–216.
Isshiki, M., Okabe, S., 2014. Evaluation of cranial window types for in vivo two-photon
imaging of brain microstructures. Microscopy (Oxf) 63, 53–63.
Jack, C.R., Albert, M.S., Knopman, D.S., McKhann, G.M., Sperling, R.A., Carrillo,
SC
M.C., Thies, B., Phelps, C.H., 2011. Introduction to the recommendations from the
Institute
on
Aging-Alzheimer’s
Association
workgroups
on
diagnostic
NU
National
guidelines for Alzheimer’s disease. Alzheimers Dement 7, 257–262.
Jacobs, B., Driscoll, L., Schall, M., 1997. Life-span dendritic and spine changes in areas
MA
10 and 18 of human cortex: a quantitative Golgi study. Journal of Comparative
Neurology 386, 661–80.
Jacobs, B., Schall, M., Prather, M., Kapler, E., Driscoll, L., Baca, S., Jacobs, J., Ford, K.,
D
Wainwright, M., Treml, M., 2001. Regional dendritic and spine variation in human
TE
cerebral cortex: a quantitative golgi study. Cereb. Cortex 11, 558–571.
Jacobsen, J.S., Wu, C.-C., Redwine, J.M., Comery, T.A., Arias, R., Bowlby, M.,
CE
P
Martone, R., Morrison, J.H., Pangalos, M.N., Reinhart, P.H., Bloom, F.E., 2006. Earlyonset behavioral and synaptic deficits in a mouse model of Alzheimer’s disease. PNAS
103, 5161–5166. https://doi.org/10.1073/pnas.0600948103
Jedlicka, P., Vlachos, A., Schwarzacher, S.W., Deller, T., 2008. A role for the spine
AC
apparatus in LTP and spatial learning. Behavioural Brain Research, Biobehavioural
Plasticity 192, 12–19.
Jiang, M., Ash, R.T., Baker, S.A., Suter, B., Ferguson, A., Park, J., Rudy, J., Torsky,
S.P., Chao, H.-T., Zoghbi, H.Y., Smirnakis, S.M., 2013. Dendritic Arborization and
Spine Dynamics Are Abnormal in the Mouse Model of MECP2 Duplication Syndrome.
J Neurosci 33, 19518–19533. https://doi.org/10.1523/JNEUROSCI.1745-13.2013
Jiang, Y.H., Armstrong, D., Albrecht, U., Atkins, C.M., Noebels, J.L., Eichele, G.,
Sweatt, J.D., Beaudet, A.L., 1998. Mutation of the Angelman ubiquitin ligase in mice
ACCEPTED MANUSCRIPT
causes increased cytoplasmic p53 and deficits of contextual learning and long-term
potentiation. Neuron 21, 799–811.
Johansson, B.B., Belichenko, P.V., 2002. Neuronal plasticity and dendritic spines: effect
of environmental enrichment on intact and postischemic rat brain. J. Cereb. Blood Flow
Metab. 22, 89–96.
Johnson, J.B., Summer, W., Cutler, R.G., Martin, B., Hyun, D.-H., Dixit, V.D., Pearson,
M., Nassar, M., Telljohann, R., Tellejohan, R., Maudsley, S., Carlson, O., John, S.,
RI
PT
Laub, D.R., Mattson, M.P., 2007. Alternate day calorie restriction improves clinical
findings and reduces markers of oxidative stress and inflammation in overweight adults
with moderate asthma. Free Radic. Biol. Med. 42, 665–674.
Jordan, B.A., Fernholz, B.D., Boussac, M., Xu, C., Grigorean, G., Ziff, E.B., Neubert,
SC
T.A., 2004. Identification and Verification of Novel Rodent Postsynaptic Density
NU
Proteins. Mol Cell Proteomics 3, 857–871.
Jourdain, P., Fukunaga, K., Muller, D., 2003. Calcium/Calmodulin-Dependent Protein
Kinase II Contributes to Activity-Dependent Filopodia Growth and Spine Formation. J.
MA
Neurosci. 23, 10645–10649.
Ju, W., Morishita, W., Tsui, J., Gaietta, G., Deerinck, T.J., Adams, S.R., Garner, C.C.,
Tsien, R.Y., Ellisman, M.H., Malenka, R.C., 2004.
Activity-dependent regulation of
Junyent, F., Kremer, E.J., 2015. CAV-2 - Why a canine virus is a neurobiologist’s best
TE
D
dendritic synthesis and trafficking of AMPA receptors. Nat. Neurosci. 7, 244–253
CE
P
friend. Curr. Opin. Pharmacol.24, 86–93.
Kalaivani, P., Ganesh, M., Sathiya, S., Ranju, V., Gayathiri, V., Babu, C.S., 2014.
Alteration in bioenergetic regulators, SirT1 and Parp1 expression precedes oxidative
stress
in
rats
subjected
to
transient
cerebral focal ischemia: molecular
and
2766.
AC
histopathologic evidences. Journal of Stroke and Cerebrovascular Diseases 23, 2753–
Kandel, E.R., Schwartz, J.H., 1982. Molecular biology of learning: modulation of
transmitter release. Science 218, 433–443.
Kang, D.E., Roh, S.E., Woo, J.A., Liu, T., Bu, J.H., Jung, A.-R., Lim, Y., 2011. The
Interface
between
Cytoskeletal
Aberrations
and
Mitochondrial
Alzheimer’s Disease and Related Disorders. Exp Neurobiol 20, 67–80.
Dysfunction
in
ACCEPTED MANUSCRIPT
Kang, M.-G., Guo, Y., Huganir, R.L., 2009. AMPA receptor and GEF-H1/Lfc complex
regulates dendritic spine development through RhoA signaling cascade. PNAS 106,
3549–3554.
Kasai, H., Hayama, T., Ishikawa, M., Watanabe, S., Yagishita, S., Noguchi, J., 2010.
Learning rules and persistence of dendritic spines. Eur. J. Neurosci. 32, 241–249.
Kasai, H., Matsuzaki, M., Noguchi, J., Yasumatsu, N., Nakahara, H., 2003a. Structurestability-function relationships of dendritic spines. Trends Neurosci. 26, 360–368.
Kasai, H., Matsuzaki, M., Noguchi, J., Yasumatsu, N., Nakahara, H., 2003b. Structure-
RI
PT
stability-function relationships of dendritic spines. Trends Neurosci. 26, 360–368.
Kaufmann, W.E., Moser, H.W., 2000. Dendritic anomalies in disorders associated with
SC
mental retardation. Cereb. Cortex 10, 981–991.
Kaufmann, W.E., Naidu, S., Budden, S., 1995. Abnormal expression of microtubule-
NU
associated protein 2 (MAP-2) in neocortex in Rett syndrome. Neuropediatrics 26, 109–
113.
Kawamoto, S., Adelstein, R.S., 1991. Chicken nonmuscle myosin heavy chains:
MA
differential expression of two mRNAs and evidence for two different polypeptides. The
Journal of Cell Biology 112, 915–924.
Kelley, B.J., Petersen, R.C., 2007. Alzheimer’s disease and mild cognitive impairment.
Neurol Clin 25, 577–609, v.
Kendler, K.S., Gatz, M., Gardner, C.O., Pedersen, N.L., 2006. A Swedish national twin
TE
D
study of lifetime major depression. Am J Psychiatry 163, 109–114.
Kennedy, M.B., 2000. Signal-processing machines at the postsynaptic density. Science
CE
P
290, 750–754.
Kessler, R.C., Chiu, W.T., Demler, O., Walters, E.E., 2005. Prevalence, Severity, and
AC
Comorbidity of Twelve-month DSM-IV Disorders in the National Comorbidity Survey
Replication (NCS-R). Arch Gen Psychiatry 62, 617–627.
Keck, T., Toyoizumi, T., Chen, L., Doiron, B., Feldman, D.E., Fox, K., Gerstner, W.,
Haydon, P.G., Hübener, M., Lee, H.K., Lisman, J.E., Rose, T., Sengpiel, F., Stellwagen,
D., Stryker, M.P., Turrigiano, G.G., van Rossum, M.C., 2017. Integrating Hebbian and
homeostatic plasticity: the current state of the field and future research directions. Phil.
Trans. R. Soc. B 372, 20160158
Kelly, R.M., Strick, P.L., 2000. Rabies as a transneuronal tracer of circuits in the central
nervous system. J. Neurosci. Methods 103, 63–71.
ACCEPTED MANUSCRIPT
Kharazia, V.N., Weinberg, R.J., 1997. Tangential synaptic distribution of NMDA and
AMPA receptors in rat neocortex. Neuroscience Letters 238, 41–44.
Kim, C.H., Lisman, J.E., 1999. A role of actin filament in synaptic transmission and
long-term potentiation. J. Neurosci. 19, 4314–4324.
Kim, E., Sheng, M., 2004. PDZ domain proteins of synapses. Nat. Rev. Neurosci. 5,
771–781.
Knobloch, M., Mansuy, I.M., 2008. Dendritic spine loss and synaptic alterations in
Alzheimer’s disease. Mol. Neurobiol. 37, 73–82.
Knott, G., Holtmaat, A., 2008. Dendritic spine plasticity--current understanding from in
vivo studies. Brain Res Rev 58, 282–289.
Knott, G.W., Holtmaat, A., Wilbrecht, L., Welker, E., Svoboda, K., 2006. Spine growth
SC
RI
PT
precedes synapse formation in the adult neocortex in vivo. Nat. Neurosci. 9, 1117–1124.
Koch, C., Zador, A., 1993. The function of dendritic spines: devices subserving
NU
biochemical rather than electrical compartmentalization. J. Neurosci. 13, 413–422.
Kolb, B., Gibb, R., 1993. Possible anatomical basis of recovery of function after
MA
neonatal frontal lesions in rats. Behav. Neurosci. 107, 799–811.
Koleske, A.J., 2013. Molecular mechanisms of dendrite stability. Nat. Rev. Neurosci. 14,
536–550.
Kolluri, N., Sun, Z., Sampson, A.R., Lewis, D.A., 2005. Lamina-Specific Reductions in
D
162, 1200–1202.
Kopec, C.D., Li, B., Wei, W., Boehm, J., Malinow, R., 2006. Glutamate receptor
CE
P
TE
Dendritic Spine Density in the Prefrontal Cortex of Subjects With Schizophrenia. AJP
exocytosis and spine enlargement during chemically induced long-term potentiation. J.
Neurosci. 26, 2000–2009.
Korkotian, E., Segal, M., 1999. Release of calcium from stores alters the morphology of
AC
dendritic spines in cultured hippocampal neurons. Proc Natl Acad Sci U S A 96, 12068–
12072.
Koyama, Y., Tohyama, M., 2013. A novel, Golgi-Cox-based fluorescent staining
method for visualizing full-length processes in primary rat neurons. Neurochem. Int. 63,
35–41.
Kozorovitskiy, Y., Gross, C.G., Kopil, C., Battaglia, L., McBreen, M., Stranahan, A.M.,
Gould, E., 2005. Experience induces structural and biochemical changes in the adult
primate brain. Proc. Natl. Acad. Sci. U.S.A. 102, 17478–17482.
ACCEPTED MANUSCRIPT
Kramár, E.A., Lin, B., Rex, C.S., Gall, C.M., Lynch, G., 2006. Integrin-driven actin
polymerization consolidates long-term potentiation. PNAS 103, 5579–5584.
Kuhlman, S.J., Huang, Z.J., 2005. High-resolution labeling and functional manipulation
of specific neuron types in mouse brain by Cre-activated viral gene expression. PLoS
ONE 3, e2005.
Lai, C.S.W., Franke, T.F., Gan, W.-B., 2012. Opposite effects of fear conditioning and
RI
PT
extinction on dendritic spine remodelling. Nature 483, 87.
Landi, S., Putignano, E., Boggio, E.M., Giustetto, M., Pizzorusso, T., Ratto, G.M., 2011.
The short-time structural plasticity of dendritic spines is altered in a model of Rett
syndrome. Sci Rep 1, 45.
Landis, D.M., Williams, R.S., Masters, C.L., 1981. Golgi and electronmicroscopic
SC
studies of spongifom encephalopathy. Neurology 31, 538–538.
Lang, C., Barco, A., Zablow, L., Kandel, E.R., Siegelbaum, S.A., Zakharenko, S.S.,
NU
2004. Transient expansion of synaptically connected dendritic spines upon induction of
MA
hippocampal long-term potentiation. Proc. Natl. Acad. Sci. U.S.A. 101, 16665–16670.
Lanz, T.A., Carter, D.B., Merchant, K.M., 2003. Dendritic spine loss in the hippocampus
of young PDAPP and Tg2576 mice and its prevention by the ApoE2 genotype.
Laurén, J., Gimbel, D.A., Nygaard, H.B., Gilbert, J.W., Strittmatter, S.M., 2009. Cellular
TE
D
Neurobiology of Disease 13, 246–253.
prion protein mediates impairment of synaptic plasticity by amyloid-beta oligomers.
CE
P
Nature 457, 1128–1132.
Law, A.J., Weickert, C.S., Hyde, T.M., Kleinman, J.E., Harrison, P.J., 2004. Reduced
spinophilin but not microtubule-associated protein 2 expression in the hippocampal
formation in schizophrenia and mood disorders: molecular evidence for a pathology of
AC
dendritic spines. Am J Psychiatry 161, 1848–1855.
Lazarov, O., Robinson, J., Tang, Y.-P., Hairston, I.S., Korade-Mirnics, Z., Lee, V.M.-Y.,
Hersh,
L.B., Sapolsky, R.M., Mirnics, K., Sisodia, S.S., 2005. Environmental
enrichment reduces Abeta levels and amyloid deposition in transgenic mice. Cell 120,
701–713.
Lee, K.-W., Kim, Y., Kim, A.M., Helmin, K., Nairn, A.C., Greengard, P., 2006.
Cocaine-induced dendritic spine formation in D1 and D2 dopamine receptor-containing
medium spiny neurons in nucleus accumbens. Proc. Natl. Acad. Sci. U.S.A. 103, 3399–
3404.
ACCEPTED MANUSCRIPT
Lee, S.H., Liu, L., Wang, Y.T., Sheng, M., 2002. Clathrin Adaptor AP2 and NSF
Interact with Overlapping Sites of GluR2 and Play Distinct Roles in AMPA Receptor
Trafficking and Hippocampal LTD. Neuron 36, 661–674.
Leggio, M.G., Mandolesi, L., Federico, F., Spirito, F., Ricci, B., Gelfo, F., Petrosini, L.,
2005. Environmental enrichment promotes improved spatial abilities and enhanced
dendritic growth in the rat. Behav. Brain Res. 163, 78–90.
Lendvai, B., Stern, E.A., Chen, B., Svoboda, K., 2000. Experience-dependent plasticity
RI
PT
of dendritic spines in the developing rat barrel cortex in vivo. Nature 404, 876–881.
Lendvai, B., Zelles, T., Rozsa, B., Vizi, E.S., 2003. A vinca alkaloid enhances
morphological dynamics of dendritic spines of neocortical layer 2/3 pyramidal cells.
SC
Brain Research Bulletin 59, 257–260.
Lesh, T.A., Niendam, T.A., Minzenberg, M.J., Carter, C.S., 2011. Cognitive control
NU
deficits in schizophrenia: mechanisms and meaning. Neuropsychopharmacology 36,
316–338.
Leuner, B., Falduto, J., Shors, T.J., 2003. Associative memory formation increases the
MA
observation of dendritic spines in the hippocampus. J. Neurosci. 23, 659–665.
Leuner, B., Shors, T.J., 2013. Stress, anxiety, and dendritic spines: what are the
connections? Neuroscience 251, 108–119.
Levinstein, M.R., Samuels, B.A., 2014. Mechanisms underlying the antidepressant
D
TE
response and treatment resistance. Front Behav Neurosci 8.
Levitt, P., Harvey, J.A., Friedman, E., Simansky, K., Murphy, E.H., 1997. New evidence
CE
P
for neurotransmitter influences on brain development. Trends Neurosci. 20, 269–274.
Lewis, D.A., Lieberman, J.A., 2000. Catching up on schizophrenia: natural history and
neurobiology. Neuron 28, 325–334.
Li, N., Lee, B., Liu, R.-J., Banasr, M., Dwyer, J.M., Iwata, M., Li, X.-Y., Aghajanian,
AC
G., Duman, R.S., 2010. mTOR-Dependent Synapse Formation Underlies the Rapid
Antidepressant Effects of NMDA Antagonists. Science 329, 959–964.
Li, S., Hong, S., Shepardson, N.E., Walsh, D.M., Shankar, G.M., Selkoe, D., 2009.
Soluble oligomers of amyloid Beta protein facilitate hippocampal long-term depression
by disrupting neuronal glutamate uptake. Neuron 62, 788–801.
Li, Z., Okamoto, K.-I., Hayashi, Y., Sheng, M., 2004. The importance of dendritic
mitochondria in the morphogenesis and plasticity of spines and synapses. Cell 119, 873–
887. https://doi.org/10.1016/j.cell.2004.11.003
ACCEPTED MANUSCRIPT
Liu, G., 2004. Local structural balance and functional interaction of excitatory and
inhibitory synapses in hippocampal dendrites. Nat. Neurosci. 7, 373–379
Lin, B., Kramár, E.A., Bi, X., Brucher, F.A., Gall, C.M., Lynch, G., 2005. Theta
stimulation polymerizes actin in dendritic spines of hippocampus. J. Neurosci. 25, 2062–
2069.
Lin, H., Huganir, R., Liao, D., 2004. Temporal dynamics of NMDA receptor-induced
RI
PT
changes in spine morphology and AMPA receptor recruitment to spines. Biochemical
and Biophysical Research Communications 316, 501–511.
Lin, S.C., Deisseroth, K., Henderson, J.M., 2011. Optogenetics: background and
concepts for neurosurgery. Neurosurgery. 69, 1-3.
Lin, M.T., Beal, M.F., 2006. Alzheimer’s APP mangles mitochondria. Nat Med 12,
SC
1241–1243.
Lipska, B.K., Peters, T., Hyde, T.M., Halim, N., Horowitz, C., Mitkus, S., Weickert,
NU
C.S., Matsumoto, M., Sawa, A., Straub, R.E., Vakkalanka, R., Herman, M.M.,
MA
Weinberger, D.R., Kleinman, J.E., 2006. Expression of DISC1 binding partners is
reduced in schizophrenia and associated with DISC1 SNPs. Hum. Mol. Genet. 15, 1245–
1258.
Lisman, J., Schulman, H., Cline, H., 2002. The molecular basis of CaMKII function in
D
TE
synaptic and behavioural memory. Nat. Rev. Neurosci. 3, 175–190.
Lorenzetti, V., Allen, N.B., Fornito, A., Yücel, M., 2009. Structural brain abnormalities
117, 1–17.
CE
P
in major depressive disorder: a selective review of recent MRI studies. J Affect Disord
Lupien, S.J., McEwen, B.S., 1997. The acute effects of corticosteroids on cognition:
AC
integration of animal and human model studies. Brain Res. Brain Res. Rev. 24, 1–27.
Lüscher, C., Xia, H., Beattie, E.C., Carroll, R.C., von Zastrow, M., Malenka, R.C.,
Nicoll, R.A., 1999. Role of AMPA receptor cycling in synaptic transmission and
plasticity. Neuron 24, 649–658.
Luo, L., Callaway, E.M., Svoboda, K., 2008. Genetic dissection of neuralcircuits.
Neuron 57, 634–660.
Lynch,
G.S., Dunwiddie, T., Gribkoff, V., 1977. Heterosynaptic depression: a
postsynaptic correlate of long-term potentiation. Nature 266, 737–739.
ACCEPTED MANUSCRIPT
Maiti, P., Lomakin, A., Benedek, G.B., Bitan, G., 2010. Despite its role in assembly,
methionine 35 is not necessary for amyloid β-protein toxicity. J Neurochem 113, 1252–
1262.
Maiti, P., Manna, J., Ilavazhagan, G., Rossignol, J., Dunbar, G.L., 2015. Molecular
regulation of dendritic spine dynamics and their potential impact on synaptic plasticity
and neurological diseases. Neurosci Biobehav Rev 59, 208–237.
Maiti, P., Piacentini, R., Ripoli, C., Grassi, C., Bitan, G., 2011. Surprising toxicity and
RI
PT
assembly behaviour of amyloid β-protein oxidized to sulfone. Biochemical Journal 433,
323–332.
Majewska, A., Brown, E., Ross, J., Yuste, R., 2000. Mechanisms of calcium decay
SC
kinetics in hippocampal spines: role of spine calcium pumps and calcium diffusion
through the spine neck in biochemical compartmentalization. J. Neurosci. 20, 1722–
NU
1734.
Majewska, A., Sur, M., 2003. Motility of dendritic spines in visual cortex in vivo:
Sci. U.S.A. 100, 16024–16029.
MA
changes during the critical period and effects of visual deprivation. Proc. Natl. Acad.
Majewska, A.K., Newton, J.R., Sur, M., 2006. Remodeling of Synaptic Structure in
Sensory Cortical Areas In Vivo. J. Neurosci. 26, 3021–3029.
Malenka, R.C., 2003. Synaptic Plasticity and AMPA Receptor Trafficking. Annals of the
D
Malenka, R.C., Bear, M.F., 2004. LTP and LTD: an embarrassment of riches. Neuron
44, 5–21.
CE
P
TE
New York Academy of Sciences 1003, 1–11.
Malenka, R.C., Nicoll,
and R.A., 1999. Long-Term Potentiation--A Decade of
Progress? Science 285, 1870–1874.
Maletic-Savatic, M., Malinow, R., Svoboda, K., 1999. Rapid dendritic morphogenesis in
AC
CA1 hippocampal dendrites induced by synaptic activity. Science 283, 1923–1927.
Malinow, R., Hayashi, Y., Maletic-Savatic, M., Zaman, S.H., Poncer, J.-C., Shi, S.-H.,
Esteban, J.A., Osten, P., Seidenman, K., 2010. Introduction of Green Fluorescent Protein
(GFP) into Hippocampal Neurons through Viral Infection. Cold Spring Harb Protoc
2010, 5406.
Malinow, R., Malenka, R.C., 2002. AMPA Receptor Trafficking and Synaptic Plasticity.
Annual Review of Neuroscience 25, 103–126.
Malinow, R., Tsien, R.W., 1990. Presynaptic enhancement shown by whole-cell
recordings of long-term potentiation in hippocampal slices. Nature 346, 177–180.
ACCEPTED MANUSCRIPT
Mallet, N., Ballion, B., Le Moine, C., Gonon, F., 2006. Cortical inputs and GABA
interneurons imbalance projection neurons in the striatum of parkinsonian rats. J.
Neurosci. 26, 3875–3884.
Man, H.Y., Lin, J.W., Ju, W.H., Ahmadian, G., Liu, L., Becker, L.E., Sheng, M., Wang,
Y.T., 2000. Regulation of AMPA receptor-mediated synaptic transmission by clathrindependent receptor internalization. Neuron 25, 649–662.
Mancuso, J.J., Chen, Y., Li, X., Xue, Z., Wong, S.T.C., 2013. Methods of dendritic spine
RI
PT
detection: From Golgi to high-resolution optical imaging. Neuroscience, Dendritic Spine
Plasticity in Brain Disorders 251, 129–140.
Mancuso, J.J., Kim, J., Lee, S., Tsuda, S., Chow, N.B.H., Augustine, G.J., 2010.
SC
Optogenetic probing of functional brain circuitry. Expt Physiol 96, 26-33.
Mahalaxmi, M., 2017. Optogenetics- A brief review. IJPSR. 8, 440-445
Mancuso, J.J., Kim, J., Lee, S., Tsuda, S., Chow, N.B.H., Augustine, G.J., 2010.
NU
Optogenetic probing of functional brain circuitry. Expt Physiol 96, 26-33.
Marshel, J.H., Mori, T., Nielsen, K.J., Callaway, E.M., 2010. Targeting single neuronal
MA
networks for gene expression and cell labeling in vivo. Neuron 67, 562–574.
Mattis, J., Tye, K.M., Ferenczi, E.A., Ramakrishnan, C., O'Shea, D.J., Prakash, R.,
Gunaydin, L.A., Hyun, M., Fenno, L.E., Gradinaru, V., Yizhar, O., Deisseroth, K., 2011.
D
Principles for applying optogenetic tools derived from direct comparative analysis of
TE
microbial opsins. Nature Methods 9, 159–172.
McGovern, A.E., Davis-Poynter, N., Rakoczy, J., Phipps, S., Simmons, D.G., Mazzone,
CE
P
S.B., 2012. Anterograde neuronal circuit tracing using a genetically modified herpes
simplex virus expressing EGFP. J Neurosci Methods. 209, 158–67
Mangiarini, L., Sathasivam, K., Seller, M., Cozens, B., Harper, A., Hetherington, C.,
AC
Lawton, M., Trottier, Y., Lehrach, H., Davies, S.W., Bates, G.P., 1996. Exon 1 of the
HD gene with an expanded CAG repeat is sufficient to cause a progressive neurological
phenotype in transgenic mice. Cell 87, 493–506.
Marcuzzo, S., Dall’oglio, A., Ribeiro, M.F.M., Achaval, M., Rasia-Filho, A.A., 2007.
Dendritic spines in the posterodorsal medial amygdala after restraint stress and ageing in
rats. Neurosci. Lett. 424, 16–21.
Marin-Padilla, M., 1976. Pyramidal cell abnormalities in the motor cortex of a child with
Down’s syndrome. A Golgi study. J. Comp. Neurol. 167, 63–81.
ACCEPTED MANUSCRIPT
Marquet, P., Depeursinge, C., Magistretti, P.J., 2013. Exploring Neural Cell Dynamics
with Digital Holographic Microscopy. Annual Review of Biomedical Engineering 15,
407–431.
Masliah, E., Mallory, M., Hansen, L., Alford, M., Albright, T., DeTeresa, R., Terry, R.,
Baudier, J., Saitoh, T., 1991. Patterns of aberrant sprouting in alzheimer’s disease.
Neuron 6, 729–739.
Matsuzaki, M., Ellis-Davies, G.C., Nemoto, T., Miyashita, Y., Iino, M., Kasai, H., 2001.
pyramidal neurons. Nat. Neurosci. 4, 1086–1092.
RI
PT
Dendritic spine geometry is critical for AMPA receptor expression in hippocampal CA1
Matsuzaki, M., Honkura, N., Ellis-Davies, G.C.R., Kasai, H., 2004. Structural basis of
SC
long-term potentiation in single dendritic spines. Nature 429, 761–766.
Matus, A., 2000. Actin-Based Plasticity in Dendritic Spines. Science 290, 754–758.
Medina, A.E., 2010.
NU
Vinpocetine as a potent antiinflammatory agent. Proc Natl Acad
Sci U S A. 107, 9921–9922.
Maul, B., von Bohlen und Halbach, O., Becker, A., Sterner-Kock, A., Voigt, J.-P.,
MA
Siems, W.-E., Grecksch, G., Walther, T., 2008. Impaired spatial memory and altered
dendritic spine morphology in angiotensin II type 2 receptor-deficient mice. J. Mol.
McEwen, B.S., 2005. Glucocorticoids, depression, and mood disorders: structural
TE
D
Med. 86, 563–571.
remodeling in the brain. Metab. Clin. Exp. 54, 20–23.
McGrath, J.A., Nestadt, G., Liang, K.-Y., Lasseter, V., Wolyniec, P.S., Fallin, M.D.,
CE
P
Thornquist, M.H., Luke, J.R., Pulver, A.E., 2004. Five latent factors underlying
schizophrenia: analysis and relationship to illnesses in relatives. Schizophrenia bulletin
30, 855–873.
McKinney, R.A., 2010. Excitatory amino acid involvement in dendritic spine formation,
AC
maintenance and remodelling. J. Physiol. (Lond.) 588, 107–116.
McKinney, R.A., Capogna, M., Dürr, R., Gähwiler, B.H., Thompson, S.M., 1999.
Miniature synaptic events maintain dendritic spines via AMPA receptor activation. Nat.
Neurosci. 2, 44–49.
Mei, L., Xiong, W.-C., 2008. Neuregulin 1 in neural development, synaptic plasticity
and schizophrenia. Nat. Rev. Neurosci. 9, 437–452.
Meng, W., Swenson, L.L., Fitzgibbon, M.J., Hayakawa, K., Ter Haar, E., Behrens, A.E.,
Fulghum, J.R., Lippke, J.A., 2002. Structure of mitogen-activated protein kinase-
ACCEPTED MANUSCRIPT
activated protein (MAPKAP) kinase 2 suggests a bifunctional switch that couples kinase
activation with nuclear export. J. Biol. Chem. 277, 37401–37405.
Meredith, G.E., Ypma, P., Zahm, D.S., 1995. Effects of dopamine depletion on the
morphology of medium spiny neurons in the shell and core of the rat nucleus
accumbens. J. Neurosci. 15, 3808–3820.
Miller, B.R., Bezprozvanny, I., 2010. Corticostriatal circuit dysfunction in Huntington’s
disease: intersection of glutamate, dopamine and calcium. Future Neurol 5, 735–756.
Minatohara, K., Akiyoshi, M., Okuno, H., 2015. Role of immediate-early genes in
RI
PT
synaptic plasticity and neuronal ensembles underlying the memory trace. Frontiers in
Molecular Neuroscience 8.
Mittmann, W., Wallace, D.J., Czubayko, U., Herb, J.T., Schaefer, A.T., Looger, L.L.,
SC
Denk, W., Kerr, J.N.D., 2011. Two-photon calcium imaging of evoked activity from L5
NU
somatosensory neurons in vivo. Nat. Neurosci. 14, 1089–1093.
Mizrahi, A., Crowley, J.C., Shtoyerman, E., Katz, L.C., 2004. High-Resolution In Vivo
Imaging of Hippocampal Dendrites and Spines. J. Neurosci. 24, 3147–3151.
Monfils, M.-H., Teskey, G.C., 2004. Induction of long-term depression is associated
MA
with decreased dendritic length and spine density in layers III and V of sensorimotor
neocortex. Synapse 53, 114–121.
Mora, F., 2013. Successful brain aging: plasticity, environmental enrichment, and
D
TE
lifestyle. Dialogues Clin Neurosci 15, 45–52.
Morton, W.A., Stockton, G.G., 2000. Methylphenidate abuse and psychiatric side
CE
P
effects. J Clin Psychiatry. 2, 159–164.
Mora, F., Segovia, G., Del Arco, A., de Blas, M., Garrido, P., 2012. Stress,
neurotransmitters, corticosterone and body-brain integration. Brain Res. 1476, 71–85.
Moser, M.-B., Moser, E.I., 1998. Functional differentiation in the hippocampus.
AC
Hippocampus 8, 608–619.
Moser, M.B., Trommald, M., Andersen, P., 1994. An increase in dendritic spine density
on hippocampal CA1 pyramidal cells following spatial learning in adult rats suggests the
formation of new synapses. PNAS 91, 12673–12675.
Muller, D., Toni, N., Buchs, P.A., 2000. Spine changes associated with long-term
potentiation. Hippocampus 10, 596–604.
ACCEPTED MANUSCRIPT
Müller, M., Gähwiler, B.H., Rietschin, L., Thompson, S.M., 1993. Reversible loss of
dendritic spines and altered excitability after chronic epilepsy in hippocampal slice
cultures. PNAS 90, 257–261.
Multani, P., Myers, R.H., Blume, H.W., Schomer, D.L., Sotrel, A., 1994. Neocortical
Dendritic Pathology in Human Partial Epilepsy: A Quantitative Golgi Study. Epilepsia
35, 728–736.
Murari, K., Zhang, Y., Li, S., Chen, Y., Li, M.-J., Li, X., 2011. Compensation-free, all-
RI
PT
fiber-optic, two-photon endomicroscopy at 1.55 μm. Opt. Lett., OL 36, 1299–1301.
Murthy, V.N., Schikorski, T., Stevens, C.F., Zhu, Y., 2001. Inactivity produces increases
in neurotransmitter release and synapse size. Neuron 32, 673–682.
Nägerl, U.V., Eberhorn, N., Cambridge, S.B., Bonhoeffer, T., 2004. Bidirectional
SC
activity-dependent morphological plasticity in hippocampal neurons. Neuron 44, 759–
NU
767.
Nägerl, U.V., Willig, K.I., Hein, B., Hell, S.W., Bonhoeffer, T., 2008. Live-cell imaging
MA
of dendritic spines by STED microscopy. PNAS 105, 18982–18987.
Nagel, G., Szellas, T., Huhn, W., Kateriya, S., Adeishvili, N., Berthold, P., Ollig, D.,
Hegemann, P., Bamberg, E., 2003. Channelrhodopsin-2, a directly light gated cationNagel, G., Ollig, D., Fuhrmann, M., Kateriya, S., Musti, A.M., Bamberg, E., Hegemann,
TE
D
selective membrane channel. Proc. Natl Acad. Sci. USA 100, 13940 –13945.
P., 2002. Channelrhodopsin-1: a light-gated proton channel in green algae. Science 296,
CE
P
2395–2398.
Nassi, J.J., Cepko, C.L., Born, R.T., Beier, K.T., 2015. Neuroanatomy goes viral! Front
Neuroanat. 9, 80.
Norgren, R.B. Jr., Lehman, M.N., 1998. Herpes simplex virus as a transneuronal tracer.
AC
Neurosci Biobehav Rev. 22, 695–708.
Newey, S.E., Velamoor, V., Govek, E.-E., Van Aelst, L., 2005. Rho GTPases, dendritic
structure, and mental retardation. J. Neurobiol. 64, 58–74.
Nguyen, M.V.C., Du, F., Felice, C.A., Shan, X., Nigam, A., Mandel, G., Robinson, J.K.,
Ballas, N., 2012. MeCP2 is critical for maintaining mature neuronal networks and global
brain anatomy during late stages of postnatal brain development and in the mature adult
brain. J. Neurosci. 32, 10021–10034.
Nimchinsky, E.A., Oberlander, A.M., Svoboda, K., 2001. Abnormal development of
dendritic spines in FMR1 knock-out mice. J. Neurosci. 21, 5139–5146.
ACCEPTED MANUSCRIPT
Nimchinsky, E.A., Sabatini, B.L., Svoboda, K., 2002. Structure and function of dendritic
spines. Annu. Rev. Physiol. 64, 313–353.
Nishimura,
M.,
Owens,
J.,
Swann,
J.W.,
2008.
Effects of Chronic Network
Hyperexcitability on the Growth of Hippocampal Dendrites. Neurobiol Dis 29, 267–277.
Nishizuka, M., Okada, R., Seki, K., Arai, Y., Iizuka, R., 1991. Loss of dendritic
synapses in the medial amygdala associated with kindling. Brain Res 552, 351–355.
Nithianantharajah, J., Barkus, C., Murphy, M., Hannan, A.J., 2008. Gene-environment
RI
PT
interactions modulating cognitive function and molecular correlates of synaptic plasticity
in Huntington’s disease transgenic mice. Neurobiol. Dis. 29, 490–504.
Nixon,
R.A.,
Cataldo,
A.M.,
2006.
Lysosomal
system
pathways: genes
to
SC
neurodegeneration in Alzheimer’s disease. J. Alzheimers Dis. 9, 277–289.
Noguchi, J., Matsuzaki, M., Ellis-Davies, G.C.R., Kasai, H., 2005. Spine-neck geometry
NU
determines NMDA receptor-dependent Ca2+ signaling in dendrites. Neuron 46, 609–
622.
O’Brien, L.M., Gozal, D., 2004. Neurocognitive dysfunction and sleep in children: from
MA
human to rodent. Pediatr. Clin. North Am. 51, 187–202.
O’Brien, R.J., Kamboj, S., Ehlers, M.D., Rosen, K.R., Fischbach, G.D., Huganir, R.L.,
1998. Activity-dependent modulation of synaptic AMPA receptor accumulation. Neuron
TE
D
21, 1067-78.
Oga, T., Aoi, H., Sasaki, T., Fujita, I., Ichinohe, N., 2013. Postnatal development of
CE
P
layer III pyramidal cells in the primary visual, inferior temporal, and prefrontal cortices
of the marmoset. Front Neural Circuits 7.
Oh, W.C., Hill, T.C., Zito, K., 2013. Synapse-specific and size-dependent mechanisms
of spine structural plasticity accompanying synaptic weakening. Proc. Natl. Acad. Sci.
AC
U.S.A. 110, E305-312.
Oh, S.W., Harris, J.A., Ng, L., Winslow, B., Cain, N., Mihalas, S., Wang, Q., Lau, C.,
Kuan, L., Henry, A.M., 2014. A mesoscale connectome of the mouse brain. Nature.
508,207–214.
Osakada, F., Mori, T., Cetin, A.H., Marshel, J.H., Virgen, B., Callaway, E.M., 2011.
New rabies virus variants for monitoring and manipulating activity and gene expression
in defined neural circuits. Neuron 71, 617–631.
ACCEPTED MANUSCRIPT
Oyibo, H.K., Znamenskiy, P., Oviedo, H.V., Enquist, L.W., Zador, A.M., 2014. Longterm Cre-mediated
retrograde
tagging
of neurons
using
a
novel recombinant
pseudorabies virus. Front. Neuroanat.8,86.
Oztas, E., 2003. Neuronal Tracing. Neuroanatomy. 2, 2–5.
Okamoto, K.-I., Nagai, T., Miyawaki, A., Hayashi, Y., 2004. Rapid and persistent
modulation
of
actin
dynamics
regulates
postsynaptic
reorganization
underlying
RI
PT
bidirectional plasticity. Nat. Neurosci. 7, 1104–1112.
O’Malley, A., O’Connell, C., Murphy, K.J., Regan, C.M., 2000. Transient spine density
increases in the mid-molecular layer of hippocampal dentate gyrus accompany
consolidation of a spatial learning task in the rodent. Neuroscience 99, 229–232.
O’Malley, A., O’Connell, C., Regan, C.M., 1998. Ultrastructural analysis reveals
SC
avoidance conditioning to induce a transient increase in hippocampal dentate spine
NU
density in the 6 hour post-training period of consolidation. Neuroscience 87, 607–613.
Ostroff, L.E., Fiala, J.C., Allwardt, B., Harris, K.M., 2002. Polyribosomes redistribute
MA
from dendritic shafts into spines with enlarged synapses during LTP in developing rat
hippocampal slices. Neuron 35, 535–545.
Ouimet, C.C., Katona, I., Allen, P., Freund, T.F., Greengard, P., 2004. Cellular and
D
subcellular distribution of spinophilin, a PP1 regulatory protein that bundles F-actin in
TE
dendritic spines. J. Comp. Neurol. 479, 374–388.
Ouyang, Y., Wong, M., Capani, F., Rensing, N., Lee, C.-S., Liu, Q., Neusch, C.,
CE
P
Martone, M.E., Wu, J.Y., Yamada, K., Ellisman, M.H., Choi, D.W., 2005. Transient
decrease in F-actin may be necessary for translocation of proteins into dendritic spines.
Eur. J. Neurosci. 22, 2995–3005.
Packer, A.M., Peterka, D.S., Hirtz, J.J., Prakash, R., Deisseroth, K., Yuste, R., 2012.
optogenetics
AC
Two-photon
of
dendritic
spines
and
neural
circuits.
Nature
Methods 9, 1202–1205
Pak, D.T., Yang, S., Rudolph-Correia, S., Kim, E., Sheng, M., 2001. Regulation of
dendritic spine morphology by SPAR, a PSD-95-associated RapGAP. Neuron 31, 289–
303.
Pak, D.T.S., Sheng, M., 2003. Targeted protein degradation and synapse remodeling by
an inducible protein kinase. Science 302, 1368–1373.
ACCEPTED MANUSCRIPT
Packer, A.M., Peterka, D.S., Hirtz, J.J., Prakash, R., Deisseroth, K., Yuste, R., 2012.
Two-photon
optogenetics
of
dendritic
spines
and
neural
circuits.
Nature
Methods 9, 1202–1205
Palop, J.J., Chin, J., Roberson, E.D., Wang, J., Thwin, M.T., Bien-Ly, N., Yoo, J., Ho,
K.O., Yu, G.-Q., Kreitzer, A., Finkbeiner, S., Noebels, J.L., Mucke, L., 2007. Aberrant
excitatory neuronal activity and compensatory remodeling of inhibitory hippocampal
RI
PT
circuits in mouse models of Alzheimer’s disease. Neuron 55, 697–711.
Park, M., Salgado, J.M., Ostroff, L., Helton, T.D., Robinson, C.G., Harris, K.M., Ehlers,
M.D., 2006. Plasticity-induced growth of dendritic spines by exocytic trafficking from
recycling endosomes. Neuron 52, 817–830.
Parnavelas, J.G., Lynch, G., Brecha, N., Cotman, C.W., Globus, A., 1974. Spine Loss
SC
and Regrowth in Hippocampus following Deafferentation. Nature 248, 71.
Patel, V., Abas, M., Broadhead, J., Todd, C., Reeler, A., 2001. Depression in developing
NU
countries: lessons from Zimbabwe. Bmj 322, 482–484.
Patt, S., Gertz, H.-J., Gerhard, L., Cervos-Navarro, J., 1991. Pathological changes in
MA
dendrites of substantia nigra neurons in Parkinson’s disease: a Golgi study. Histol
Histopathol 6, 373–380.
Peng, J., Kim, M.J., Cheng, D., Duong, D.M., Gygi, S.P., Sheng, M., 2004.
D
TE
Semiquantitative proteomic analysis of rat forebrain postsynaptic density fractions by
mass spectrometry. J. Biol. Chem. 279, 21003–21011.
Penzes, P., Beeser, A., Chernoff, J., Schiller, M.R., Eipper, B.A., Mains, R.E., Huganir,
CE
P
R.L., 2003. Rapid induction of dendritic spine morphogenesis by trans-synaptic ephrinBEphB receptor activation of the Rho-GEF kalirin. Neuron 37, 263–274.
Penzes, P., Cahill, M.E., Jones, K.A., VanLeeuwen, J.-E., Woolfrey, K.M., 2011.
AC
Dendritic spine pathology in neuropsychiatric disorders. Nat Neurosci 14, 285–293.
Penzes, P., Rafalovich, I., 2012. Regulation of the actin cytoskeleton in dendritic spines.
Adv. Exp. Med. Biol. 970, 81–95.
Peters, A., Kaiserman-Abramof, I.R., 1970. The small pyramidal neuron of the rat
cerebral cortex. The perikaryon, dendrites and spines. Am. J. Anat. 127, 321–355.
Peters, A., Sethares, C., Moss, M.B., 1998. The effects of aging on layer 1 in area 46 of
prefrontal cortex in the rhesus monkey. Cereb. Cortex 8, 671–684.
Pham, T.M., Winblad, B., Granholm, A.-C., Mohammed, A.H., 2002. Environmental
influences on brain neurotrophins in rats. Pharmacol. Biochem. Behav. 73, 167–175.
ACCEPTED MANUSCRIPT
Phillips, M., Raju, N., Rubinsky, L., Rubinsky, B., 2015. Modulating electrolytic tissue
ablation with reversible electroporation pulses. Technology 03, 45–53.
Picchioni, D., Reith, R.M., Nadel, J.L., Smith, C.B., 2014. Sleep, Plasticity and the
Pathophysiology of Neurodevelopmental Disorders: The Potential Roles of Protein
Synthesis and Other Cellular Processes. Brain Sci 4, 150–201.
Pinto, L., Mateus-Pinheiro, A., Morais, M., Bessa, J.M., Sousa, N., 2012. Immuno-Golgi
as
a
Tool
for
Analyzing
Neuronal
3D-Dendritic
Characterized Neurons. PLOS ONE 7, e33114.
of
9-methyl-β-carboline:
dopaminergic
neurons
coupled
stimulation,
with
protection
anti-inflammatory
Neurochemistry 113, 1659–1675.
and
regeneration
effects.
Journal
of
of
NU
Pollard, T.D., Cooper, J.A., 2009. Actin, a central player in cell shape and movement.
Science 326, 1208–1212.
Phenotypically
Polanski, W., Enzensperger, C., Reichmann, H., Gille, G., 2010. The exceptional
properties
in
SC
Structure
RI
PT
Pop, A.S., Levenga, J., Esch, C.E.F. de, Buijsen, R.A.M., Nieuwenhuizen, I.M., Li, T.,
phenotype
in
KO
mice
MA
Isaacs, A., Gasparini, F., Oostra, B.A., Willemsen, R., 2014. Rescue of dendritic spine
with
the
mGluR5
antagonist
AFQ056/Mavoglurant.
Psychopharmacology 231, 1227–1235.
Gabbott,
D
Popov, V.I., Davies, H.A., Rogachevsky, V.V., Patrushev, I.V., Errington, M.L.,
P.L.A., Bliss, T.V.P., Stewart, M.G., 2004. Remodelling of synaptic
morphology
but
TE
unchanged
synaptic
density
during
late
phase
long-term
CE
P
potentiation(ltp): A serial section electron micrograph study in the dentate gyrus in the
anaesthetised rat. Neuroscience 128, 251–262.
Portera, C.C., Yuste, R., 2001. [On the function of dendritic filopodia]. Rev Neurol 33,
AC
1158–1166.
Portera-Cailliau, C., Pan, D.T., Yuste, R., 2003. Activity-regulated dynamic behavior of
early dendritic protrusions: evidence for different types of dendritic filopodia. J.
Neurosci. 23, 7129–7142.
Prusiner, S.B., 1998. The prion diseases. Brain Pathol. 8, 499–513.
Purpura, D.P., 1974. Dendritic spine “dysgenesis” and mental retardation. Science 186,
1126–1128.
Rabinowitch, I., Segev, I., 2008. Two opposing plasticity mechanisms pulling a single
synapse. Trends Neurosci. 31,377-83.
ACCEPTED MANUSCRIPT
Rácz, B., Blanpied, T.A., Ehlers, M.D., Weinberg, R.J., 2004. Lateral organization of
endocytic machinery in dendritic spines. Nat Neurosci 7, 917–918.
Rademacher, D.J., Mendoza-Elias, N., Meredith, G.E., 2015. Effects of Context-Drug
Learning on Synaptic Connectivity in the Basolateral Nucleus of the Amygdala in Rats.
Eur J Neurosci 41, 205–215.
Rademacher, D.J., Rosenkranz, J.A., Morshedi, M.M., Sullivan, E.M., Meredith, G.E.,
2010. Amphetamine-Associated Contextual Learning is Accompanied by Structural and
RI
PT
Functional Plasticity in the Basolateral Amygdala. J Neurosci 30, 4676–4686.
Radley, J.J., Morrison, J.H., 2005. Repeated stress and structural plasticity in the brain.
Ageing Res. Rev. 4, 271–287.
Rampon, C., Jiang, C.H., Dong, H., Tang, Y.P., Lockhart, D.J., Schultz, P.G., Tsien,
SC
J.Z., Hu, Y., 2000. Effects of environmental enrichment on gene expression in the brain.
NU
Proc. Natl. Acad. Sci. U.S.A. 97, 12880–12884.
Ramsey, A.J., Milenkovic, M., Oliveira, A.F., Escobedo-Lozoya, Y., Seshadri, S.,
Salahpour, A., Sawa, A., Yasuda, R., Caron, M.G., 2011. Impaired NMDA receptor
MA
transmission alters striatal synapses and DISC1 protein in an age-dependent manner.
Proc. Natl. Acad. Sci. U.S.A. 108, 5795–5800.
Rao, A., Kim, E., Sheng, M., Craig, A.M., 1998. Heterogeneity in the Molecular
D
Composition of Excitatory Postsynaptic Sites during Development of Hippocampal
Rema, V., Armstrong-James, M., Jenkinson, N., Ebner, F.F., 2006. Short exposure to an
enriched
environment accelerates plasticity in the barrel cortex of adult rats.
CE
P
TE
Neurons in Culture. J. Neurosci. 18, 1217–1229.
Neuroscience 140, 659–672.
Rensing, N., Ouyang, Y., Yang, X.-F., Yamada, K.A., Rothman, S.M., Wong, M., 2005.
AC
In vivo imaging of dendritic spines during electrographic seizures. Ann. Neurol. 58,
888–898. https://doi.org/10.1002/ana.20658
Restivo, L., Ferrari, F., Passino, E., Sgobio, C., Bock, J., Oostra, B.A., Bagni, C.,
Ammassari-Teule,
M.,
2005.
Enriched
environment
promotes
behavioral
and
morphological recovery in a mouse model for the fragile X syndrome. Proc. Natl. Acad.
Sci. U.S.A. 102, 11557–11562.
Rex, C.S., Gavin, C.F., Rubio, M.D., Kramar, E.A., Chen, L.Y., Jia, Y., Huganir, R.L.,
Muzyczka, N., Gall, C.M., Miller, C.A., Lynch, G., Rumbaugh, G., 2010. Myosin IIb
regulates actin dynamics during synaptic plasticity and memory formation. Neuron 67,
603–617.
ACCEPTED MANUSCRIPT
Reymann, K.G., Frey, J.U., 2007. The late maintenance of hippocampal LTP:
Requirements,
phases,
‘synaptic
tagging’,
‘late-associativity’
and
implications.
Neuropharmacology, LTP: Forty Unforgettable Years. A Festschrift in Honour of
Professor Tim Bliss FRS 52, 24–40.
Ricciardi, S., Ungaro, F., Hambrock, M., Rademacher, N., Stefanelli, G., Brambilla, D.,
Sessa, A., Magagnotti, C., Bachi, A., Giarda, E., Verpelli, C., Kilstrup-Nielsen, C., Sala,
C., Kalscheuer, V.M., Broccoli, V., 2012. CDKL5 ensures excitatory synapse stability
RI
PT
by reinforcing NGL-1-PSD95 interaction in the postsynaptic compartment and is
impaired in patient iPSC-derived neurons. Nat. Cell Biol. 14, 911–923.
Roberts, A.C., Díez-García, J., Rodriguiz, R.M., López, I.P., Luján, R., Martínez-
SC
Turrillas, R., Picó, E., Henson, M.A., Bernardo, D.R., Jarrett, T.M., Clendeninn, D.J.,
López-Mascaraque, L., Feng, G., Lo, D.C., Wesseling, J.F., Wetsel, W.C., Philpot, B.D.,
NU
Pérez-Otaño, I., 2009. Downregulation of NR3A-containing NMDARs is required for
synapse maturation and memory consolidation. Neuron 63, 342–356.
Robinson, T.E., Kolb, B., 1999. Alterations in the morphology of dendrites and dendritic
MA
spines in the nucleus accumbens and prefrontal cortex following repeated treatment with
amphetamine or cocaine. Eur. J. Neurosci. 11, 1598–1604.
Rochefort, N.L., Konnerth, A., 2012. Dendritic spines: from structure to in vivo
Rost, B.R., Schneider-Warme, F., Schmitz, D., Hegemann, P., 2017. Optogenetic Tools
TE
D
function. EMBO Rep 13, 699–708.
for Subcellular Applications in Neuroscience. Neuron 96, 572-603
Rost, B.R., Schneider, F., Grauel, M.K., Wozny, C., Bentz, C., Blessing, A.,
CE
P
Rosenmund, T., Jentsch, T.J., Schmitz, D., Hegemann, P., Rosenmund, C., 2015.
1852.
AC
Optogenetic acidification of synaptic vesicles and lysosomes. Nat. Neurosci. 18, 1845–
Rogers, J., Zornetzer, S.F., Bloom, F.E., Mervis, R.E., 1984. Senescent microstructural
changes in rat cerebellum. Brain Research 292, 23–32.
Rubinow, M.J., Drogos, L.L., Juraska, J.M., 2009. Age-related dendritic hypertrophy
and sexual dimorphism in rat basolateral amygdala. Neurobiology of aging 30, 137–146.
Ruddy, R.M., Chen, Y., Milenkovic, M., Ramsey, A.J., 2015. Differential effects of
NMDA receptor antagonism on spine density. Synapse 69, 52–56.
Rust, M.J., Bates, M., Zhuang, X., 2006. Sub-diffraction-limit imaging by stochastic
optical reconstruction microscopy (STORM). Nature Methods 3, 793.
ACCEPTED MANUSCRIPT
Ryu, J., Liu, L., Wong, T.P., Wu, D.C., Burette, A., Weinberg, R., Wang, Y.T., Sheng,
M., 2006a. A Critical Role for Myosin IIB in Dendritic Spine Morphology and Synaptic
Function. Neuron 49, 175–182.
Ryu, J., Liu, L., Wong, T.P., Wu, D.C., Burette, A., Weinberg, R., Wang, Y.T., Sheng,
M., 2006b. A critical role for myosin IIb in dendritic spine morphology and synaptic
function. Neuron 49, 175–182.
Saito, S., Kobayashi, S., Ohashi, Y., Igarashi, M., Komiya, Y., Ando, S., 1994.
RI
PT
Decreased synaptic density in aged brains and its prevention by rearing under enriched
environment as revealed by synaptophysin contents. J. Neurosci. Res. 39, 57–62.
Sala, C., Cambianica, I., Rossi, F., 2008. Molecular mechanisms of dendritic spine
Sala, C., Segal, M., 2014. Dendritic Spines: The Locus of Structural and Functional
NU
Plasticity. Physiological Reviews 94, 141–188.
SC
development and maintenance. Acta Neurobiol Exp (Wars) 68, 289–304.
Salegio, E.A., Samaranch, L., Kells, A.P., Mittermeyer, G., Sebastian, W.S., Zhou, S.,
Beyer, J., Forsayeth, J., Bankiewicz, K.S., 2013. Axonal transport of adeno-associated
MA
viral vectors is serotype-dependent. Gene Ther. 20,348–352.
Schwarz, L.A., Miyamichi, K., Gao, X.J., Beier, K.T., Weissbourd, B., DeLoach, K.E.,
Ren, J., Ibanes, S., Malenka, R.C., Kremer, E.J., 2015. Viral-genetic tracing of the
Sinnen, B.L, Bowen, A.B., Forte, J.S., Hiester, B.G., Crosby, K.C., Gibson, E.S.,
TE
D
input–output organization of a central noradrenaline circuit. Nature.524,88–92.
Dell’Acqua., M.L., Kennedy, M.J. 2017. Optogenetic control of synaptic composition
CE
P
and function. Neuron 93, 646–660
Suberbielle, E., Sanchez, P.E., Kravitz, A.V., Wang, X., Ho, K., Eilertson, K., Devidze,
N., Kreitzer, A.C., Mucke, L., 2013. Physiologic brain activity causes DNA double-
AC
strand breaks in neurons, with exacerbation by amyloid-β. Nat Neurosci. 16, 613-621
Samaco, R.C., Mandel-Brehm, C., McGraw, C.M., Shaw, C.A., McGill, B.E., Zoghbi,
H.Y., 2012. Crh and Oprm1 mediate anxiety-related behavior and social approach in a
mouse model of MECP2 duplication syndrome. Nat Genet 44, 206–211.
Sanhueza, M., Fernandez-Villalobos, G., Stein, I.S., Kasumova, G., Zhang, P., Bayer,
K.U., Otmakhov, N., Hell, J.W., Lisman, J., 2011. Role of the CaMKII/NMDA receptor
complex in the maintenance of synaptic strength. J. Neurosci. 31, 9170–9178.
Sathiya, S., Ranju, V., Kalaivani, P., Priya, R.J., Sumathy, H., Sunil, A.G., Babu, C.S.,
2013. Telmisartan attenuates MPTP induced dopaminergic degeneration and motor
ACCEPTED MANUSCRIPT
dysfunction through regulation of α-synuclein and neurotrophic factors (BDNF and
GDNF) expression in C57BL/6J mice. Neuropharmacology 73, 98–110.
Sawyer, K., Corsentino, E., Sachs-Ericsson, N., Steffens, D.C., 2012. Depression,
Hippocampal Volume Changes, and Cognitive Decline in a Clinical Sample of Older
Depressed Outpatients and Non-depressed Controls. Aging Ment Health 16, 753–762.
Schapiro, S., Vukovich, K.R., 1970. Early Experience Effects upon Cortical Dendrites:
A Proposed Model for Development. Science 167, 292–294.
Scheff, S.W., Price, D.A., Schmitt, F.A., Mufson, E.J., 2006. Hippocampal synaptic loss
RI
PT
in early Alzheimer’s disease and mild cognitive impairment. Neurobiol. Aging 27, 1372–
1384.
Scheibel, M.E., Crandall, P.H., Scheibel, A.B., 1974. The Hippocampal-Dentate
SC
Complex in Temporal Lobe Epilepsy. Epilepsia 15, 55–80.
Schewe, J.C., Zuschratter, W., Riederer, B., Schramm, J., Elger, C.E., Wiestler, O.D.,
NU
Blumcke, I., 1998. Cellular Pathology Of Hippocampal Neurons In Ammon’s Horn
Sclerosis. Epilepsia 39.
Schumacher, J., Laje, G., Jamra, R.A., Becker, T., Mühleisen, T.W., Vasilescu, C.,
MA
Mattheisen, M., Herms, S., Hoffmann, P., Hillmer, A.M., Georgi, A., Herold, C.,
Schulze, T.G., Propping, P., Rietschel, M., McMahon, F.J., Nöthen, M.M., Cichon, S.,
sample
and
TE
central European
D
2009. The DISC locus and schizophrenia: evidence from an association study in a
from a meta-analysis across different European
populations. Hum Mol Genet 18, 2719–2727.
Segal, M., 2005. Dendritic spines and long-term plasticity. Nat. Rev. Neurosci. 6, 277–
CE
P
284. Segal, M., 1995. Imaging of calcium variations in living dendritic spines of cultured
rat hippocampal neurons. J. Physiol. (Lond.) 486 ( Pt 2), 283–295.
Segal, M., Andersen, P., 2000. Dendritic spines shaped by synaptic activity. Current
AC
Opinion in Neurobiology 10, 582–586.
Segal, M., Clarke, D.J., Maddox, P., Salmon, E.D., Bloom, K., Reed, S.I., 2000.
Coordinated Spindle Assembly and Orientation Requires Clb5p-Dependent Kinase in
Budding Yeast. J Cell Biol 148, 441–452.
Selkoe, D.J., 2008. Soluble oligomers of the amyloid β-protein impair synaptic plasticity
and behavior. Behavioural Brain Research, Biobehavioural Plasticity 192, 106–113.
Selkoe, D.J., 2002. Alzheimer’s disease is a synaptic failure. Science 298, 789–791.
Selkoe, D.J., 1989. Amyloid β protein precursor and the pathogenesis of Alzheimer’s
disease. Cell 58, 611–612.
ACCEPTED MANUSCRIPT
Selkoe, D.J., Schenk, D., 2003. Alzheimer’s disease: molecular understanding predicts
amyloid-based therapeutics. Annual review of pharmacology and toxicology 43, 545–
584.
Serrano-Pozo, A., Frosch, M.P., Masliah, E., Hyman, B.T., 2011. Neuropathological
Alterations in Alzheimer Disease. Cold Spring Harb Perspect Med 1.
Shamir, A., Kwon, O.-B., Karavanov, I., Vullhorst, D., Leiva-Salcedo, E., Janssen, M.J.,
Buonanno, A., 2012. The importance of the nrg-1/erbb4 pathway for synaptic plasticity
RI
PT
and behaviors associated with psychiatric disorders. J Neurosci 32, 2988–2997.
Shankar, G.M., Bloodgood, B.L., Townsend, M., Walsh, D.M., Selkoe, D.J., Sabatini,
B.L., 2007. Natural oligomers of the Alzheimer amyloid-beta protein induce reversible
SC
synapse loss by modulating an NMDA-type glutamate receptor-dependent signaling
pathway. J. Neurosci. 27, 2866–2875.
Shi, S.H., Hayashi, Y., Petralia, R.S., Zaman, S.H., Wenthold, R.J., Svoboda, K.,
NU
Malinow, R., 1999. Rapid spine delivery and redistribution of AMPA receptors after
synaptic NMDA receptor activation. Science 284, 1811–1816.
Sisková, Z., Page, A., O’Connor, V., Perry, V.H., 2009. Degenerating synaptic boutons
MA
in prion disease: microglia activation without synaptic stripping. Am. J. Pathol. 175,
1610–1621.
Slow, E.J., van Raamsdonk, J., Rogers, D., Coleman, S.H., Graham, R.K., Deng, Y., Oh,
D
TE
R., Bissada, N., Hossain, S.M., Yang, Y.-Z., Li, X.-J., Simpson, E.M., Gutekunst, C.-A.,
Leavitt, B.R., Hayden, M.R., 2003. Selective striatal neuronal loss in a YAC128 mouse
Smart, F.M., Halpain, S., 2000. Regulation of dendritic spine stability. Hippocampus 10,
542–554.
Smith, Y., Villalba, R., 2008. Striatal and extrastriatal dopamine in the basal ganglia: An
AC
CE
P
model of Huntington disease. Hum. Mol. Genet. 12, 1555–1567.
overview of its anatomical organization in normal and Parkinsonian brains. Mov. Disord.
23, S534–S547.
Snyder, E.M., Philpot, B.D., Huber, K.M., Dong, X., Fallon, J.R., Bear, M.F., 2001.
Internalization of ionotropic glutamate receptors in response to mGluR activation. Nat.
Neurosci. 4, 1079–1085.
Solaini, G., Baracca, A., Lenaz, G., Sgarbi, G., 2010. Hypoxia and mitochondrial
oxidative metabolism. Biochim. Biophys. Acta 1797, 1171–1177.
Song, I., Huganir, R.L., 2002. Regulation of AMPA receptors during synaptic plasticity.
Trends Neurosci. 25, 578–588.
ACCEPTED MANUSCRIPT
Sorra, K.E., Fiala, J.C., Harris, K.M., 1998. Critical assessment of the involvement of
perforations, spinules, and spine branching in hippocampal synapse formation. J. Comp.
Neurol. 398, 225–240.
Sotrel, A., Williams, R.S., Kaufmann, W.E., Myers, R.H., 1993. Evidence for neuronal
degeneration and dendritic plasticity in cortical pyramidal neurons of Huntington’s
disease A quantitative Golgi study. Neurology 43, 2088–2088.
Spacek, J., Harris, K.M., 1997. Three-dimensional organization of smooth endoplasmic
RI
PT
reticulum in hippocampal CA1 dendrites and dendritic spines of the immature and
mature rat. J. Neurosci. 17, 190–203.
Spencer, J.P., 2008. Food for thought: the role of dietary flavonoids in enhancing human
SC
memory, learning and neuro-cognitive performance: Symposium on ‘Diet and mental
health.’ Proceedings of the Nutrition Society 67, 238–252.
Spiga, S., Acquas, E., Puddu, M.C., Mulas, G., Lintas, A., Diana, M., 2011.
NU
Simultaneous Golgi-Cox and immunofluorescence using confocal microscopy. Brain
Struct Funct 216, 171–182.
Spigelman, I., Yan, X.-X., Obenaus, A., Lee, E.Y.-S., Wasterlain, C.G., Ribak, C.E.,
MA
1998. Dentate granule cells form novel basal dendrites in a rat model of temporal lobe
epilepsy. Neuroscience 86, 109–120.
Spires, T.L., Grote, H.E., Varshney, N.K., Cordery, P.M., Dellen, A. van, Blakemore,
D
TE
C., Hannan, A.J., 2004. Environmental Enrichment Rescues Protein Deficits in a Mouse
Model of Huntington’s Disease, Indicating a Possible Disease Mechanism. J. Neurosci.
CE
P
24, 2270–2276.
Spires, T.L., Meyer-Luehmann, M., Stern, E.A., McLean, P.J., Skoch, J., Nguyen, P.T.,
Bacskai, B.J., Hyman, B.T., 2005. Dendritic spine abnormalities in amyloid precursor
AC
protein transgenic mice demonstrated by gene transfer and intravital multiphoton
microscopy. J. Neurosci. 25, 7278–7287.
Spires-Jones, T.L., Meyer-Luehmann, M., Osetek, J.D., Jones, P.B., Stern, E.A.,
Bacskai, B.J., Hyman, B.T., 2007. Impaired spine stability underlies plaque-related spine
loss in an Alzheimer’s disease mouse model. Am. J. Pathol. 171, 1304–1311.
St Clair, D., Blackwood, D., Muir, W., Carothers, A., Walker, M., Spowart, G., Gosden,
C., Evans, H.J., 1990. Association within a family of a balanced autosomal translocation
with major mental illness. Lancet 336, 13–16.
Staffend, N.A., Meisel, R.L., 2011. DiOlistic labeling in fixed brain slices: phenotype,
morphology and dendritic spines. Curr Protoc Neurosci CHAPTER, Unit2.13.
ACCEPTED MANUSCRIPT
Star, E.N., Kwiatkowski, D.J., Murthy, V.N., 2002. Rapid turnover of actin in dendritic
spines and its regulation by activity. Nat. Neurosci. 5, 239–246.
Steiner, P., Higley, M.J., Xu, W., Czervionke, B.L., Malenka, R.C., Sabatini, B.L., 2008.
Destabilization of the Postsynaptic Density by PSD-95 Serine 73 Phosphorylation
Inhibits Spine Growth and Synaptic Plasticity. Neuron 60, 788–802.
Stepanyants,
A., Chklovskii, D.B., 2005. Neurogeometry and potential synaptic
connectivity. Trends Neurosci. 28, 387–394.
Stephens, B., Mueller, A.J., Shering, A.F., Hood, S.H., Taggart, P., Arbuthnott, G.W.,
RI
PT
Bell, J.E., Kilford, L., Kingsbury, A.E., Daniel, S.E., Ingham, C.A., 2005. Evidence of a
breakdown of corticostriatal connections in Parkinson’s disease. Neuroscience 132, 741–
Stevens, C.F., Wang, Y., 1994. Changes in reliability of synaptic function as a
Steward, O., Worley, P. F., 2001. Selective targeting of newly synthesized Arc mRNA to
active synapses requires NMDA receptor activation. Neuron 30, 227–240.
Steward, O., Reeves, T. M., 1988. Protein-synthetic machinery beneath postsynaptic sites
on CNS neurons: association between polyribosomes and other organelles at the synaptic
site. The Journal of Neuroscience 8, 176–184.
MA
NU
mechanism for plasticity. Nature 371, 704–707.
SC
754.
Stroemer, R.P., Kent, T.A., Hulsebosch, C.E., 1998. Enhanced neocortical neural
D
sprouting, synaptogenesis, and behavioral recovery with D-amphetamine therapy after
TE
neocortical infarction in rats. Stroke 29, 2381–2393.
Südhof, T.C., 2008. Neuroligins and neurexins link synaptic function to cognitive
CE
P
disease. Nature 455, 903–911.
Sutton, M.A., Schuman, E.M., 2005. Local translational control in dendrites and its role
in long-term synaptic plasticity. J. Neurobiol. 64, 116–131.
Sutton, M.A., Ito, H.T., Cressy, P., Kempf, C., Woo, J.C., Schuman, E.M., 2006.
AC
Miniature neurotransmission stabilizes synaptic function via tonic suppression of local
dendritic protein synthesis. Cell 125, 785–799.
Suzuki, A., Ueno, N., Hemmati-Brivanlou, A., 1997. Xenopus msx1 mediates epidermal
induction and neural inhibition by BMP4. Development 124, 3037–3044.
Suberbielle, E., Sanchez, P.E., Kravitz, A.V., Wang, X., Ho, K., Eilertson, K., Devidze,
N., Kreitzer, A.C., Mucke, L., 2013. Physiologic brain activity causes DNA doublestrand breaks in neurons, with exacerbation by amyloid-β. Nat Neurosci. 16, 613-621
ACCEPTED MANUSCRIPT
Swann, J.W., Al-Noori, S., Jiang, M., Lee, C.L., 2000. Spine loss and other dendritic
abnormalities in epilepsy. Hippocampus 10, 617–625.
Sweet, R.A., Henteleff, R.A., Zhang, W., Sampson, A.R., Lewis, D.A., 2009. Reduced
Dendritic
Spine Density In Auditory Cortex Of Subjects With Schizophrenia.
Neuropsychopharmacology 34, 374–389.
Takashima, S., Becker, L.E., Armstrong, D.L., Chan, F., 1981. Abnormal neuronal
development in the visual cortex of the human fetus and infant with down’s syndrome. A
RI
PT
quantitative and qualitative golgi study. Brain Research 225, 1–21.
Takashima, S., Iida, K., Mito, T., Arima, M., 1994. Dendritic and histochemical
development and ageing in patients with Down’s syndrome. J Intellect Disabil Res 38 (
SC
Pt 3), 265–273.
Tanaka, J.-I., Horiike, Y., Matsuzaki, M., Miyazaki, T., Ellis-Davies, G.C.R., Kasai, H.,
NU
2008. Protein synthesis and neurotrophin-dependent structural plasticity of single
dendritic spines. Science 319, 1683–1687.
Tang, G., Gudsnuk, K., Kuo, S.-H., Cotrina, M.L., Rosoklija, G., Sosunov, A., Sonders,
MA
M.S., Kanter, E., Castagna, C., Yamamoto, A., Yue, Z., Arancio, O., Peterson, B.S.,
Champagne, F., Dwork, A.J., Goldman, J., Sulzer, D., 2014. Loss of mTOR-dependent
macroautophagy causes autistic-like synaptic pruning deficits. Neuron 83, 1131–1143.
Terry, R.D., Masliah, E., Salmon, D.P., Butters, N., DeTeresa, R., Hill, R., Hansen, L.A.,
D
TE
Katzman, R., 1991. Physical basis of cognitive alterations in Alzheimer’s disease:
synapse loss is the major correlate of cognitive impairment. Ann. Neurol. 30, 572–580.
Thompson, S.M., Kallarackal, A.J., Kvarta, M.D., Van Dyke, A.M., LeGates, T.A., Cai,
CE
P
X., 2015. An excitatory synapse hypothesis of depression. Trends Neurosci. 38, 279–
294.
Todorova, M.T., Mantis, J.G., Le, M., Kim, C.Y., Seyfried, T.N., 2006. Genetic and
AC
environmental interactions determine seizure susceptibility in epileptic EL mice. Genes,
Brain and Behavior 5, 518–527.
Toni, N., Buchs, P.-A., Nikonenko, I., Bron, C.R., Muller, D., 1999. LTP promotes
formation of multiple spine synapses between a single axon terminal and a dendrite.
Nature 402, 421–425.
Toyoizumi, T., Kaneko, M., Stryker, M.P., Miller, K.D., 2014. Modeling the dynamic
interaction of Hebbian and homeostatic plasticity. Neuron. 84, 497-510.
ACCEPTED MANUSCRIPT
Trachtenberg, J.T., Chen, B.E., Knott, G.W., Feng, G., Sanes, J.R., Welker, E., Svoboda,
K., 2002. Long-term in vivo imaging of experience-dependent synaptic plasticity in
adult cortex. Nature 420, 788–794.
Tredici, G., Di Francesco, A., Miani, A., Pizzini, G., 1993. Real Complete ThreeDimensional Reconstruction of Golgi-Impregnated Neurons by Means of a Confocal
Laser Scanning Microscope. NeuroImage 1, 87–93.
Trommald, m., 1990. Dendritic spine changes in rat dentate granule cells associated with
RI
PT
long-term potentiation. Neurotoxicity of excitatory amino acids.
Trommald, M., Hulleberg, G., Andersen, P., 1996. Long-term potentiation is associated
with new excitatory spine synapses on rat dentate granule cells. Learn. Mem. 3, 218–
SC
228.
Tropea, D., Giacometti, E., Wilson, N.R., Beard, C., McCurry, C., Fu, D.D., Flannery,
NU
R., Jaenisch, R., Sur, M., 2009. Partial reversal of Rett Syndrome-like symptoms in
MeCP2 mutant mice. Proc. Natl. Acad. Sci. U.S.A. 106, 2029–2034.
Tsai, H.C., Zhang, F., Adamantidis, A., Stuber, G.D., Bonci, A., de Lecea, L.,
MA
Deisseroth, K., 2009. Phasic firing in dopaminergic neurons is sufficient for behavioral
conditioning. Science 324, 1080–1084.
Tye, K.M., Deisseroth, K., 2012. Optogenetic investigation of neural circuits underlying
Turrigiano, G.G., Nelson, S.B., 2004. Homeostatic plasticity in the developing nervous
TE
D
brain disease in animal models. Nat Rev. Neurosci. 13, 251-66
system. Nat. Rev. Neurosci. 5, 97–107
Turrigiano, G.G., Nelson, S.B., 2000. Hebb and homeostasis in neuronal plasticity. Curr.
CE
P
Opin. Neurobiol. 10, 358–364
Turrigiano, G., 2011. Too Many Cooks? Intrinsic and synaptic homeostatic mechanisms
AC
in cortical circuit refinement. Annu. Rev. Neurosci. 34, 89–103.
Turrigiano, G.G., Leslie, K.R., Desai, N.S., Rutherford, L.C., Nelson, S.B., 1998.
Activity dependent scaling of quantal amplitude in neocortical neurons. Nature 391,
892–896.
Tsai, J., Grutzendler, J., Duff, K., Gan, W.-B., 2004. Fibrillar amyloid deposition leads
to local synaptic abnormalities and breakage of neuronal branches. Nat. Neurosci. 7,
1181–1183.
Tsay, D., Yuste, R., 2004. On the electrical function of dendritic spines. Trends
Neurosci. 27, 77–83.
ACCEPTED MANUSCRIPT
Ugolini, G., 2010. Advances in viral transneuronal tracing. J. Neurosci. Methods. 194,2–
20.
Ugolini, G., 2011. Rabies virus as a transneuronal tracer of neuronal connections. Adv.
Virus Res. 79, 165–202
Ugolini, G., Kuypers, H.G., Strick, P.L., 1989. Transneuronal transfer of herpes virus
from peripheral nerves to cortex and brainstem. Science 243, 89-91
Ultanir, S.K., Kim, J.-E., Hall, B.J., Deerinck, T., Ellisman, M., Ghosh, A., 2007.
RI
PT
Regulation of spine morphology and spine density by NMDA receptor signaling in vivo.
Proc Natl Acad Sci U S A 104, 19553–19558.
Urbanska, M., Swiech, L., Jaworski, J., 2012. Developmental plasticity of the dendritic
van Wyk, M., Pielecka-Fortuna, J., Lo¨ wel, S., and Kleinlogel, S. (2015). Restoring the
ON
Switch
in
Blind
Retinas: Opto-mGluR6,
NU
SC
compartment: focus on the cytoskeleton. Adv. Exp. Med. Biol. 970, 265–284.
a Next-Generation,
Cell-Tailored
Optogenetic Tool. PLoS Biol. 13, e1002143
Valverde, F., 1967. Apical dendritic spines of the visual cortex and light deprivation in
MA
the mouse. Experimental brain research 3, 337–52.
van der Zee, E.A., 2015. Synapses, spines and kinases in mammalian learning and
D
memory, and the impact of aging. Neuroscience & Biobehavioral Reviews, Brain,
TE
Memory and Development: The Imprint of Gabriel Horn 50, 77–85.
Vanderklish, P.W., Edelman, G.M., 2005. Differential translation and fragile X
CE
P
syndrome. Genes Brain Behav. 4, 360–384.
Vecsey, C.G., Baillie, G.S., Jaganath, D., Havekes, R., Daniels, A., Wimmer, M.,
Huang, T., Brown, K.M., Li, X.-Y., Descalzi, G., Kim, S.S., Chen, T., Shang, Y.-Z.,
AC
Zhuo, M., Houslay, M.D., Abel, T., 2009. Sleep deprivation impairs cAMP signalling in
the hippocampus. Nature 461, 1122–1125.
Verpelli, C., Piccoli, G., Zibetti, C., Zanchi, A., Gardoni, F., Huang, K., Brambilla, D.,
Di Luca, M., Battaglioli, E., Sala, C., 2010. Synaptic activity controls dendritic spine
morphology by modulating eEF2-dependent BDNF synthesis. J. Neurosci. 30, 5830–
5842.
Verpelli, C., Sala, C., 2012. Molecular and synaptic defects in intellectual disability
syndromes. Curr. Opin. Neurobiol. 22, 530–536.
ACCEPTED MANUSCRIPT
Vicente-Manzanares, M., Horwitz, A.R., 2010a. Myosin light chain mono- and diphosphorylation
differentially
regulate
adhesion
and
polarity
in
migrating
cells.
Biochem. Biophys. Res. Commun. 402, 537–542.
Vicente-Manzanares, M., Horwitz, A.R., 2010b. Myosin light chain mono- and diphosphorylation
differentially
regulate
adhesion
and
polarity
in
migrating
cells.
Biochemical and Biophysical Research Communications 402, 537–542.
Vicente-Manzanares, M., Ma, X., Adelstein, R.S., Horwitz, A.R., 2009. Non-muscle
RI
PT
myosin II takes centre stage in cell adhesion and migration. Nat. Rev. Mol. Cell Biol. 10,
778–790.
Videbech, P., Ravnkilde, B., 2004. Hippocampal volume and depression: a meta-
SC
analysis of MRI studies. American Journal of Psychiatry 161, 1957–1966.
Villalba, R.M., Lee, H., Smith, Y., 2009. Dopaminergic denervation and spine loss in the
NU
striatum of MPTP-treated monkeys. Experimental Neurology 215, 220–227.
Villalba, R.M., Smith, Y., 2013. Differential striatal spine pathology in Parkinson’s
disease and cocaine addiction: a key role of dopamine? Neuroscience 251, 2–20.
Villalba, R.M., Smith, Y., 2011a. Neuroglial Plasticity at Striatal Glutamatergic
MA
Synapses in Parkinson’s Disease. Front Syst Neurosci 5.
Villalba, R.M., Smith, Y., 2011b. Differential Structural Plasticity of Corticostriatal and
D
Thalamostriatal Axo-Spinous Synapses in MPTP-Treated Parkinsonian Monkeys. J
Villalba, R.M., Smith, Y., 2010. Striatal Spine Plasticity in Parkinson’s Disease. Front
CE
P
Neuroanat 4.
TE
Comp Neurol 519, 989–1005.
Volkmar, F.R., Greenough, W.T., 1972. Rearing complexity affects branching of
dendrites in the visual cortex of the rat. Science 176, 1445–1447.
von Bohlen und Halbach, O., 2009. Structure and function of dendritic spines within the
AC
hippocampus. Annals of Anatomy - Anatomischer Anzeiger 191, 518–531.
von Bohlen und Halbach, O., Krause, S., Medina, D., Sciarretta, C., Minichiello, L.,
Unsicker, K., 2006. Regional- and age-dependent reduction in trkB receptor expression
in the hippocampus is associated with altered spine morphologies. Biol. Psychiatry 59,
793–800.
von Bohlen und Halbach, O., Minichiello, L., Unsicker, K., 2008. TrkB but not trkC
receptors are necessary for postnatal maintenance of hippocampal spines. Neurobiology
of Aging 29, 1247–1255.
ACCEPTED MANUSCRIPT
von Campe, G., Spencer, D.D., de Lanerolle, N.C., 1997. Morphology of dentate granule
cells in the human epileptogenic hippocampus. Hippocampus 7, 472–488.
Voronin, L., Byzov, A., Kleschevnikov, A., Kozhemyakin, M., Kuhnt, U., Volgushev,
M., 1995. Neurophysiological analysis of long-term potentiation in mammalian brain.
Behavioural Brain Research, Neurobiology of Memory 66, 45–52.
W. Moser, H., 1995. A role for gene therapy in mental retardation. Ment Retard Dev
Reviews - Mental Retard dev disabil res 1, 4–6.
RI
PT
Disabil Res Rev 1:4-6. Mental Retardation and Developmental Disabilities Research
Walikonis, R.S., Jensen, O.N., Mann, M., Provance, D.W., Mercer, J.A., Kennedy,
M.B., 2000. Identification of proteins in the postsynaptic density fraction by mass
Wiegert, J.S.,
SC
spectrometry. J. Neurosci. 20, 4069–4080.
Oertner, T.G.,
2013. Long-term depression triggers the selective
NU
elimination of weakly integrated synapses. PNAS E4510–E4519
Walsh, D.M., Selkoe, D.J., 2004. Deciphering the Molecular Basis of Memory Failure in
Alzheimer’s Disease. Neuron 44, 181–193.
Walsh, T., McClellan, J.M., McCarthy, S.E., Addington, A.M., Pierce, S.B., Cooper,
MA
G.M., Nord, A.S., Kusenda, M., Malhotra, D., Bhandari, A., Stray, S.M., Rippey, C.F.,
Roccanova, P., Makarov, V., Lakshmi, B., Findling, R.L., Sikich, L., Stromberg, T.,
D
Merriman, B., Gogtay, N., Butler, P., Eckstrand, K., Noory, L., Gochman, P., Long, R.,
TE
Chen, Z., Davis, S., Baker, C., Eichler, E.E., Meltzer, P.S., Nelson, S.F., Singleton, A.B.,
Lee, M.K., Rapoport, J.L., King, M.-C., Sebat, J., 2008. Rare structural variants disrupt
543.
CE
P
multiple genes in neurodevelopmental pathways in schizophrenia. Science 320, 539–
Walter, E., Mazaika, P., Reiss, A., 2009. Insights into brain development from
AC
neurogenetic syndromes: evidence from fragile x syndrome, williams syndrome, turner
syndrome and velocardiofacial syndrome. Neuroscience 164, 257–271.
Wang, G., Cheng, Y., Gong, M., Liang, B., Zhang, M., Chen, Y., Zhang, C., Yuan, X.,
Xu, J., 2013. Systematic correlation between spine plasticity and the anxiety/depressionlike phenotype induced by corticosterone in mice. NeuroReport 24, 682.
Wang, H., 2015. Fragile X mental retardation protein: from autism to neurodegenerative
disease. Front Cell Neurosci 9.
Weiler, I.J., Spangler, C.C., Klintsova, A.Y., Grossman, A.W., Kim, S.H., BertainaAnglade, V., Khaliq, H., de Vries, F.E., Lambers, F.A., Hatia, F., 2004. Fragile X mental
retardation protein is necessary for neurotransmitter-activated protein translation at
ACCEPTED MANUSCRIPT
synapses. Proceedings of the National Academy of Sciences of the United States of
America 101, 17504–17509.
Wall, N.R., Wickersham, I.R., Cetin, A., De La Parra, M., Callaway, E.M., 2010.
Monosynaptic
circuit
tracing
in
vivo
through
Credependent
targeting
and
complementation of modified rabies virus. Proc. Natl. Acad. Sci. U.S.A. 107, 21848–
21853.
Wickersham, I.R., Feinberg, E.H., 2012. New technologies for imaging synaptic
RI
PT
partners. Curr. Opin. Neurobiol. 22, 1–7
Wickersham, I.R., Finke, S., Conzelmann, K.K., Callaway, E. M., 2007a. Retrograde
neuronal tracing with a deletion-mutant rabies virus. Nat. Methods 4, 47–49.
Wickersham, I.R., Lyon, D.C., Barnard, R.J., Mori, T., Finke, S., Conzelmann, K.K.,
SC
Young, J.A., Callaway, E.M., 2007b. Monosynaptic restriction of transsynaptic tracing
NU
from single, genetically targeted neurons. Neuron 53, 639–647.
Wichmann, T., Delong, M.R., 2007. Anatomy and physiology of the basal ganglia:
MA
relevance to Parkinson’s disease and related disorders. Handb Clin Neurol 83, 1–18.
Willemsen, R., Levenga, J., Oostra, B.A., 2011. CGG repeat in the FMR1 gene: size
matters. Clin. Genet. 80, 214–225.
Willmore, J.L., Ballinger, W.E., Boggs, W., Sypert, G.W., Rubin, J.J., 1980. Dendritic
in
Rat
Isocortex
within
an
Iron-induced
Chronic Epileptic Focus.
TE
Alterations
D
Neurosurgery 7, 142–146.
Wong, M., 2008. Stabilizing dendritic stucture as a novel therapeutic approach for
CE
P
epilepsy. Expert Rev Neurother 8, 907–915.
Wong, M., 2005. Modulation of dendritic spines in epilepsy: cellular mechanisms and
functional implications. Epilepsy & Behavior 7, 569–577.
Wong, M., Guo, D., 2013. Dendritic spine pathology in epilepsy: cause or consequence?
AC
Neuroscience 251, 141–150.
Wu, A., Ying, Z., Gomez-Pinilla, F., 2007. Omega-3 fatty acids supplementation
restores mechanisms that maintain brain homeostasis in traumatic brain injury. J.
Neurotrauma 24, 1587–1595.
Wu, E.Q., Birnbaum, H.G., Shi, L., Ball, D.E., Kessler, R.C., Moulis, M., Aggarwal, J.,
2005. The economic burden of schizophrenia in the United States in 2002. J Clin
Psychiatry 66, 1122–1129.
ACCEPTED MANUSCRIPT
Xiao, G., Harhaj, E.W., Sun, S.C., 2001. NF-kappaB-inducing kinase regulates the
processing of NF-kappaB2 p100. Mol. Cell 7, 401–409.
Xie, Y., Hayden, M.R., Xu, B., 2010. BDNF Overexpression in the Forebrain Rescues
Huntington’s Disease Phenotypes in YAC128 Mice. J Neurosci 30, 14708–14718.
Xu, P., Duong, D.M., Seyfried, N.T., Cheng, D., Xie, Y., Robert, J., Rush, J.,
Hochstrasser, M., Finley, D., Peng, J., 2009. Quantitative proteomics reveals the
function of unconventional ubiquitin chains in proteasomal degradation. Cell 137, 133–
RI
PT
145.
Yamamoto, K., Tanei, Z.I., Hashimoto, T., Wakabayashi, T., Okuno, H., Naka, Y.,
Yizhar, O., Fenno, L.E., Fukayama, M., Bito, H., Cirrito, J.R., Holtzman, D.M.,
SC
Deisseroth, K., Iwatsubo, T., 2015. Chronic optogenetic activation augments aβ
pathology in a mouse model of Alzheimer disease. Cell Rep. 11, 859-865.
Yang, G., Gan, W.-B., 2012. Sleep contributes to dendritic spine formation and
NU
elimination in the developing mouse somatosensory cortex. Dev Neurobiol 72, 1391–
1398.
Yang, G., Lai, C.S.W., Cichon, J., Ma, L., Li, W., Gan, W.-B., 2014. Sleep promotes
MA
branch-specific formation of dendritic spines after learning. Science 344, 1173–1178.
Yang, G., Pan, F., Gan, W.-B., 2009. Stably maintained dendritic spines are associated
Yang, Y., Wang, X., Frerking, M., Zhou, Q., 2008. Spine Expansion and Stabilization
TE
D
with lifelong memories. Nature 462, 920–924.
Associated with Long-term Potentiation. J Neurosci 28, 5740–5751.
Yizhar O, Fenno, L.E., Davidson, T.J., Mogri, M., Deisseroth, K., 2011. Optogenetics in
CE
P
neural systems. Neuron 71, 9–34.
Yoon, J.H., Minzenberg, M.J., Raouf, S., D’Esposito, M., Carter, C.S., 2013. Impaired
AC
prefrontal-basal ganglia functional connectivity and substantia nigra hyperactivity in
schizophrenia. Biol. Psychiatry 74, 122–129.
Yu, W., Lu, B., 2012. Synapses and Dendritic Spines as Pathogenic Targets in
Alzheimer’s Disease. Neural Plasticity.
Yu, X., Zuo, Y., 2011. Spine plasticity in the motor cortex. Curr. Opin. Neurobiol. 21,
169–174.
Yuste, R., 2010. Dendritic Spines. MIT Press.
Yuste, R., Bonhoeffer, T., 2004. Genesis of dendritic spines: insights from ultrastructural
and imaging studies. Nat. Rev. Neurosci. 5, 24–34.
ACCEPTED MANUSCRIPT
Yuste, R., Bonhoeffer, T., 2001. Morphological changes in dendritic spines associated
with long-term synaptic plasticity. Annu. Rev. Neurosci. 24, 1071–1089.
Zaja-Milatovic, S., Milatovic, D., Schantz, A.M., Zhang, J., Montine, K.S., Samii, A.,
Deutch, A.Y., Montine, T.J., 2005. Dendritic degeneration in neostriatal medium spiny
neurons in Parkinson disease. Neurology 64, 545–547.
Zechel, S., Unsicker, K., von Bohlen und Halbach, O., 2009. Fibroblast growth factor-2
deficiency
affects
hippocampal
spine
morphology,
but
not
hippocampal
RI
PT
catecholaminergic or cholinergic innervation. Dev. Dyn. 238, 343–350.
Zehle, S., Bock, J., Jezierski, G., Gruss, M., Braun, K., 2007. Methylphenidate treatment
recovers stress-induced elevated dendritic spine densities in the rodent dorsal anterior
SC
cingulate cortex. Dev Neurobiol 67, 1891–1900.
Zeng, L.-H., Xu, L., Rensing, N.R., Sinatra, P.M., Rothman, S.M., Wong, M., 2007.
NU
Kainate Seizures Cause Acute Dendritic Injury and Actin Depolymerization In Vivo. J
Neurosci 27, 11604–11613.
Zhang, S., Boyd, J., Delaney, K., Murphy, T.H., 2005. Rapid Reversible Changes in
MA
Dendritic Spine Structure In Vivo Gated by the Degree of Ischemia. J. Neurosci. 25,
5333–5338.
Zhang, W., Benson, D.L., 2001. Stages of synapse development defined by dependence
Zhang, W., Benson, D.L., 2000. Development and molecular organization of dendritic
TE
D
on F-actin. J. Neurosci. 21, 5169–5181.
spines and their synapses. Hippocampus 10, 512–526.
Zhang, F., Prigge, M., Beyrière, F., Tsunoda, S.P., Mattis, J., Yizhar, O., Hegemann,
CE
P
P.,Deisseroth, K., 2008. Red-shifted optogenetic excitation: a tool for fast neural control
derived from Volvox carteri. Nature Neurosci. 11, 631–633.
Zhang, F., Wang, L.P., Brauner, M., Liewald, J.F., Kay, K., Watzke, N., Wood, P.G.,
AC
Bamberg, E., Nagel, G., Gottschalk, A., Deisseroth, K., 2007. Multimodal fast optical
interrogation of neural circuitry. Nature 446, 633–639.
Zhang, F., Vierock, J., Yizhar, O., Fenno, L.E., Tsunoda, S., Kianianmomeni, A., Prigge,
M., Berndt, A., Cushman, J., Polle, J., Magnuson, J., Hegemann, P., Deisseroth, K.,
2011. The microbial opsin family of optogenetic tools. Cell 147, 1446–1457.
Zingg. B., Chou, X.L., Zhang, Z.G., Mesik, L., Liang, F., Tao, H.W., Zhang, L.I., 2017.
AAV-mediated Anterograde Transsynaptic Tagging: Mapping Input-Defined Functional
Neural Pathways for Defense Behavior. Neuron. 93, 33–47.
ACCEPTED MANUSCRIPT
Zhao, L., Ma, Q.-L., Calon, F., Harris-White, M.E., Yang, F., Lim, G.P., Morihara, T.,
Ubeda, O.J., Ambegaokar, S., Hansen, J.E., Weisbart, R.H., Teter, B., Frautschy, S.A.,
Cole, G.M., 2006. Role of p21-activated kinase pathway defects in the cognitive deficits
of Alzheimer disease. Nat. Neurosci. 9, 234–242.
Zhou, Q., Homma, K.J., Poo, M., 2004. Shrinkage of dendritic spines associated with
long-term depression of hippocampal synapses. Neuron 44, 749–757.
Zhou, Y., Takahashi, E., Li, W., Halt, A., Wiltgen, B., Ehninger, D., Li, G.-D., Hell,
RI
PT
J.W., Kennedy, M.B., Silva, A.J., 2007. Interactions between the NR2B receptor and
CaMKII modulate synaptic plasticity and spatial learning. J. Neurosci. 27, 13843–
13853.
Zhu, L., Wang, L., Ju, F., Ran, Y., Wang, C., Zhang, S., 2017. Transient global cerebral
SC
ischemia induces rapid and sustained reorganization of synaptic structures. J Cereb
NU
Blood Flow Metab 37, 2756–2767.
Zhu, Y., Pak, D., Qin, Y., McCormack, S.G., Kim, M.J., Baumgart, J.P., Velamoor, V.,
Auberson, Y.P., Osten, P., van Aelst, L., Sheng, M., Zhu, J.J., 2005. Rap2-JNK
MA
Removes Synaptic AMPA Receptors during Depotentiation. Neuron 46, 905–916.
Ziv, N.E., Smith, S.J., 1996. Evidence for a role of dendritic filopodia in synaptogenesis
and spine formation. Neuron 17, 91–102.
Zuo, Y., Yang, G., Kwon, E., Gan, W.-B., 2005. Long-term sensory deprivation prevents
D
CE
P
TE
dendritic spine loss in primary somatosensory cortex. Nature 436, 261–265.
AC
Figure 1. Types of spines based on shapes. Filopodium – thin elongated protrusions; Thin
spines – small head connected to the dendritic shaft by a narrow neck; Cup spines – two
small heads connected to the dendritic shaft by a short neck; Mushroom spines - large head
connected to the dendritic shaft by a narrow neck; Stubby spines – Short bulbous structure
with no distinction of head and neck
Figure 2. Major components of dendritic spines. Synaptic activation (via AMPA and NMDA
receptors by glutamate) alters structural molecule like PSD-95 which in turn stimulates
RhoGTPases, or protein kinases signals. RhoA, Rnd1, Rac1 and Cdc42 are main regulators of
dendritic spine morphology and synapse strengths.
Figure 3: Transmission electron microscopy of a dendritic spine (magnification 50 × 103
times). Red arrow denotes presynaptic density which contains vesicles; Green arrow denotes
postsynaptic density and star denotes spine apparatus (Maiti et al., 2015, reproduced with
permission from Hiroyuki Kamiguchi, Editor Neuroscience Research journal)
Figure 4: Single neuronal cell with axon, dendritic tree and spines visualized using
fluorescent Golgi method. (a) RGM colour image (low magnification) and (b, c) gray scale
ACCEPTED MANUSCRIPT
images (high magnification). Arrows indicate (b) axon or (c) spines (reproduced with
permission from Yassine Amrani, Editor OA Anatomy)
Figure 5: Two photon microscopy imaging illustrating changes in a dendritic spine. (A) A
charge coupled device image with the black box showing the areas of the photon where two
photon images were taken (Scale bars: 500 μm). (B) A low-magnification image of a
dendritic spine in the motor cortex of mouse (Scale bar 20 μm). (C) Repetitive pictures of
dendrite under higher-magnification showing newly formed dendritic spines with arrows
RI
PT
denoting eliminated spines and stars showing filopodia (Scale bar 2 μm) (Xinzhu et al., 2014,
reproduced with permission from Benjamin Werth, Editor JOVE)
CE
P
TE
D
MA
NU
SC
Figure 6. Viral-mediated neuronal tracing (Example: RABV). (A) Wild-type RABV genome
consists of five proteins including GP. (B) deletion of GP prevents trans-synaptic spreading
and restricts the spread in a single-step vector. The gene of interest usually tagged with a
fluorescent protein like GFP or YFP, and inserted in the space of deleted (GP). These viral
transfected neurons can be visualized using fluorescent screening. This technique also
permits the expression of various genetically encoded tools to visualize neuronal circuits,
e.g., biosensors, synapse markers, activators/repressors of neuronal activities. (C) GP-deleted
RABV is pseudotyped either with its native GP (that provides normal infection capabilities)
or an engineered surface protein (like EnvA that provides cell-specific infection).(D) GPdeleted RABV tagged with its native GP is carried by axon terminals and used to trace
neurons retrogradely. (E) EnvA pseudotyped GP-deleted RABV only infects the neurons
expressing cognate (TVA) receptors. Trans-complementation strategy allows EnvA
pseudotyped GP-deleted RABV to cross one synaptic step and infect the presynaptic neurons
specifically, as it cannot infect other presynaptic cells due to the lack of envelope GP. This
limiting strategy makes this tracing tool as a valuable one to decipher a mono-synaptic event
in the selective neuronal population.
AC
Figure 7. Structural changes in a dendritic spine (a) repetitive LTP induction (via activation
of AMPA and NMDA receptors) causes enlargement of spine head. (b) repetitive LTD
induction causes spine head shrinkage (due to actin depolarization)
ACCEPTED MANUSCRIPT
Table.1. Abnormalities in dendritic spines in various neurological disorders
Neurological /
S
neurodegenerative
No:
Dendritic spine pathology
References
condition
Marked spine loss, increase in
Alzheimer’s
varicosities formation
RI
PT
premature dendritic spines and
1. disease
14, 272, 247,
245, 203
Decrease in dendritic length,
2. Parkinson’s disease
dendritic spine loss and varicosities
SC
formation
342, 459
In mild cases, significant increase in
NU
dendritic branching, spine density
and spine size are seen.
MA
Huntington’s
3. disease
In advanced cases, decrease in
115, 144
dendritic arborization, focal
D
swellings on dendrites and decrease
Intellectual
TE
in spine density are reported.
CE
P
4. disability or mental
retardation
AC
Autism spectrum
5. disorder
6.
Both changes in dendritic spine
117, 318,
452, 296
density and/or spine shape
Higher dendritic spine density
188, 345
Increase in dendritic spine density
Fragile X
and many dendritic spines show
syndrome
immature morphology like long,
193, 82
thin, torturous nature
7. Rett syndrome
Size of neurons gets reduced along
12, 29, 70,
with increase in neuronal cell
125, 319
ACCEPTED MANUSCRIPT
density; lower dendritic spine
density especially mushroom-type
spines; and impaired dendritic
arborization
Dendritic spines show larger spine
heads along with distinct reduction
in DS number and altered spine
RI
PT
8. Down’s syndrome
morphology
Marked reduction in spine number
SC
9. Schizophrenia
NU
Spine dysfunction, impaired signal
10.Depression
27, 28
134, 235
256, 471
transmission across synapse and
MA
dendritic spine loss
Frequent loss of dendritic spines,
D
varicose swelling of dendrites and
dendritic abnormalities such as
314, 39, 65,
changes in dendritic length, shape
29, 195
TE
11.Epilepsy
432, 472,
CE
P
and branching patterns
AC
12.Stroke
13.Prion disease
Spine loss and reduction in spine
number
Significant decrease in spine density
and dendritic spine loss
112, 500
56, 246
Harmful effects on signaling
14.Sleep disorders
molecules, polymerization of Factin and dynamics of dendritic
spines
89, 486
ACCEPTED MANUSCRIPT
Highlights
Detailed review of dendritic spine structure, morphology and morphogenesis, volumetric
alterations of dendritic spines in various neurological diseases.
Spine density is generally believed to get stabilized during adult life, but recent research
analyses reveal that learning, practising and training new skills indeed causes remarkable
morphological and volumetric changes in spine turnover, which in turns alters the neural
RI
PT
circuitry.
Alterations in the dendritic spine turnover and shape is the central dogmatic factor in
synaptic plasticity.
Advanced development in higher resolution imaging techniques describes the molecular
SC
structural role of the dendritic spines in new synapse formation and alteration in the neural
circuits.
The manuscript gives significant details on the experimental proof of dendritic spine
NU
alterations using LTP and LTD stimulations in diverse experimental setup.
The status of spines in various neurological and neurodegenerative diseases in discussed
MA
elaborately which includes, Rett syndrome, Down's syndrome, Prions disease, Fragile X
syndrome, schizophrenia, epilepsy etc.,
D
Remarkable effects of various pharmacological agents on dendritic spines and synapses
CE
P
TE
status are discussed.
AC
Figure 1
Figure 2
Figure 3
Figure 4
Figure 5
Figure 6
Figure 7