Biomedical Chemistry - Current Trends and Developments
Biomedical Chemistry - Current Trends and Developments
Biomedical Chemistry - Current Trends and Developments
This work is licensed under the Creative Commons Attribution-NonCommercial-NoDerivs 3.0 license,
which means that the text may be used for non-commercial purposes, provided credit is given to
the author. For details go to http://creativecommons.org/licenses/by-nc-nd/3.0/.
ISBN 978-3-11-046875-5
e- ISBN 978-3-11-046887-8
www.degruyteropen.com
3.1 Amino Acids and Peptides in Medicine: Old or New Drugs? 178
3.1.1 Introduction 178
3.1.1.1 Amino acids: biological and chemical concepts 178
3.1.2 Amino Acids and Drug Development 182
3.1.2.1 Rationale for Drug Design 182
3.1.2.2 Amino Acid Prodrug in Drug Delivery 185
3.1.2.3 L-type Amino Acid Transporter 192
3.1.2.4. Variability of Amino Acid Application to Exclusive Drugs 194
3.1.3 Peptides for Biomedicine 197
3.1.3.1 Antimicrobial Peptides (AMPs) 198
3.1.3.1.1 AMPs: Mechanism of Action and Peptide Families 199
3.1.3.1.2 α-Helical Peptides without Cys Residues 201
3.1.3.1.3 Peptides Containing Disulfide Bridges 202
3.1.3.1.4 Peptides Rich in Pro, Gly, His, Arg and Trp Residues 202
3.1.3.3 Peptides: Scaffolding Materials in Tissue Engineering 206
3.1.3.3.1 Peptide-based Biopolymers 206
3.1.3.3.2 Strategies to Create Scaffolds as Instructive Extracellular
Microenvironments for Tissue Engineering - Peptide-conjugated
Polymers to Mimic Natural ECM 207
3.1.3.3.3 Peptide-based Biomaterials Responsive to Environmental Cues 208
3.1.3.3.4 Self-assembling Peptides as Biomaterials 209
3.1.3.4 Therapeutic Peptides and Market 211
3.1.3.4.1 Chemical Strategies to Improve Peptide Biological Activity 211
3.1.3.4.2 Market 213
3.1.4 Conclusions 215
References 216
Index 343
Preface
Biomedical Chemistry focuses on creating molecules that help advance the
understanding and treatment of diseases. Basic chemical ideas and determination
of disease etiology are approached by developing techniques to ensure optimum
interaction between drugs and human cells.
This book provides readers with an understanding of how fundamental chemical
concepts are used to combat some diseases. The authors explain the interdisciplinary
nexus of chemistry with biology, physics, pharmacy and medicine. The results of
chemical research can be applied to understand chemical processes in cells and in
the body, and new methods for drug transportation.
Biomedical Chemistry: Current Trends and Developments is an excellent resource
for students and researchers in health-related fields with frontier topics in medicinal
and pharmaceutical chemistry, organic chemistry and biochemistry.
Nuno Vale
University of Porto, Portugal
Section 1: Chemical Principles in Drug Design and
Discovery
Diana C. G. A. Pinto, João M. P. Pereira, Artur M. S. Silva*1
1.1 Functional Groups of Biomolecules and their
Reactions
Abstract: This chapter starts with a general introduction on some concepts needed to
understand the reactivity of organic functional groups. Elementary reaction mecha-
nisms are then presented according to their functionality with relevant biological
examples. These reactions explain the vast majority of transformations involving bio-
molecules. Finally, two examples on the application of the presented concepts are
given, namely the metabolism of fatty acids and reactivity of penicillin. Both of these
examples call for various types of reactions showing the diversity and simplicity of
biological transformations when analysed step by step.
Table 1.1.1: Common functional groups present in biomolecules. In parentheses are the names of the
families of compounds, where the group has the highest priority in the compound.
C C
C C C
C
O N
OH
R-oxyalkyl (ethers)
hydroxyl (alcohols) amino (amines)
C S
C C C S C
SH S
sulfhydryl (thiols/mercaptanes) sulfide (thioethers/sulfides) disulfide (disulfides)
O O
N
C
C C
C C
H C C
formyl (aldehydes)
dialkylcarbonyl (ketones)
imino (imines)
O O O
C C C C
C OH C O C N
C C C P
C P O
C S C O
O O
O
sulfanyl O
carbonyl (thioesters) alkyl phosphate (phosphates) acyl phosphate (mixed
anhydrides)
4 Functional Groups of Biomolecules and their Reactions
Table 1.1.2: Possible orbital combinations in the carbon atom, their spatial configuration and geometry
with inter-bond angles.
1 x 2s + 3 x 2p 4 x sp3
tetrahedral
1 x 2s + 2 x 2p 3 x sp2
trigonal planar
1 x 2s + 2 x 2p 2 x sp
linear
All of the hybrid orbitals may be used to form σ (sigma) molecular orbitals by fusion
with s or with other hybrid orbitals from other atoms (according to Molecular Orbital
Theory). If there are remaining 2p orbitals in the carbon atoms (in the case of sp2
and sp hybrid orbitals) they are used to form π (pi) molecular orbitals by lateral
combination with other adjacent 2p orbitals (Fig. 1.1.1). A simple bond is formed with
a single σ-bond; the double bond is formed with a σ-bond and a π-bond; the triple
bond is formed with a σ-bond and two π-bonds.
Figure 1.1.1: A simplified representation of the spatial distribution of two adjacent p orbitals (a) and
their combination into a π bond (b) between two sp2 carbon atoms. The π-bonds in a triple bond
would be along orthogonal planes to each other.
Acids and Bases Versus Electrophiles and Nucleophiles 5
O O O δ−
or
C C C δ+
Scheme 1.1.1: Carbonyl group resonance structures and equivalent notation, explicitly showing the
bond polarity with partial charges.
Acids and bases are very important in biological transformations, as most require
some form of acid or basic catalysis to occur. The simplest acid-base theory is the
Brønsted-Lowry theory, which states that acids are molecules that donate protons
(hydrogen ions, H+) and bases/alkalis are molecules that accept protons. For example,
a carboxylic acid can donate a proton to a base, such as an amine, in a reversible
proton-transfer reaction (Scheme 1.1.2).
6 Functional Groups of Biomolecules and their Reactions
O H O H
+ N
+ N H
1
R OH R H R O R1 H
acid base conjugate base conjugate acid
Scheme 1.1.2: Example of a proton transfer reaction. This specific reaction explains why it is difficult
for condensations to happen directly between an amine and a carboxylic acid, as the non-ionic
forms of these molecules are more reactive (Chapter 1.1.4.5).
Acids can differ in their ability to donate protons, being classified as strong or weak
according to the extent of deprotonation. A strong acid will have a stable conjugate
base (or weak conjugate base), resulting in ready donation of a proton. Table 1.1.3 lists
the acidity of some typical functional groups (water and ammonium acidity are also
given for comparison). The acidity is measured by the acidity constant, Ka or by its pKa
(Scheme 1.1.3), where a stronger acid has a smaller pKa and a weaker acid has a larger
pKa. The same approach can be applied to bases and their strength.
HA + H2 O A- + H3 O+
The problem with the Brönsted/Lowry definition is that it only covers the compounds
that donate or accept protons. The more general and widely used model is the Lewis
definition. A Lewis acid is a molecule that accepts a pair of electrons and a Lewis base
is a molecule that donates a pair of electrons. To accept electrons, a Lewis acid must
have a vacant low-energy orbital. As a consequence, many species, including H+ itself,
metal cations such as Mg2+ and Zn2+, and neutral species such as boron trifluoride
(BF3) and carbon dioxide (CO2) are Lewis acids.
Lewis acids and bases are involved in many biological reactions, such as the
transformation of carbon dioxide into hydrogen carbonate (Scheme 1.1.4). Lewis
bases use unshared electrons to form new bonds with other atoms and are usually
referred as nucleophiles (“nucleus-loving”, versus Lewis acids as electrophiles,
“electron-loving”). The terms “electrophile” and “nucleophile” are commonly used
Acids and Bases Versus Electrophiles and Nucleophiles 7
Table 1.1.3: Relative acidity strengths of some functional groups and common reactants.
CH3COH
ammonium 9.26
NH4+
alkylthiol CH3SH 10.3
CH3CCH2COCH3
water H 2O 15.74
ketone O 19.3
CH3CCH3
thioester O 21
CH3CSCH 3
ester O 25
CH3COCH3
δ− O
O carbonic
δ+ anhydrase
HO + C C
HO O
O δ−
Lewis base Lewis acid
Scheme 1.1.4: Example of a biochemical reaction involving a Lewis acid and a Lewis base. Reversible
conversion of carbon dioxide to hydrogen carbonate can occur rapidly in the active site of a carbonic
anhydrase, the equilibrium being regulated by factors such as pH.
8 Functional Groups of Biomolecules and their Reactions
O O
H O
HO OH
cis trans
HO C
C C
C OH
H H
O H
maleic acid fumaric acid
m.p. 135oC m.p. 287oC
Figure 1.1.2: Maleic and fumaric acid structures with their respective melting points, highlighting the
importance of cis-trans isomerism for physical properties.
Stereoisomerism and Chirality 9
A compound is defined as chiral if it does not possess any planes of symmetry. This
implies that two chiral molecules, constructed in such a way that one is the reflection
of the other upon a plane, are not superimposable. In this case, these two molecules
are classified as enantiomers. One can refer to carbons as asymmetrical or chiral
whenever all of its substituents are unique, and are also referred to as stereogenic
centers (stereocenters), as they can cause stereoisomerism in the molecule.
A molecule can have more than one stereocenter, such as 1,2-dimethylcyclopropane
with two asymmetric carbons. Molecules that differ in the configuration of one or
more (but not all) stereocenters, whilst being stereoisomers but not enantiomers,
are called diastereomers. Having that in mind, it follows that any of the trans-1,2-
dimethylcyclopropane enantiomers is a diastereomer of cis-1,2-dimethylcyclopropane,
since it is equivalent to switching only one of the methyl groups. The cis-stereoisomer
does not have enantiomers because it has a plane of symmetry. Structures such as cis-
1,2-dimethylcyclopropane are called a meso structures.
plane of
reflection
H H
H 3C
CH 3
CH2 H2C
H3C CH 3
H H
will rotate light by the same amount but anticlockwise. Racemic mixtures (racemates)
are mixtures with same amount of both enantiomers of a compound and have a null
optical rotation.
Biological systems are highly stereospecific ‒ in general, only one stereoisomer
is reactive towards an enzyme or a certain receptor. As a result, racemate resolution
is of extreme importance for drug synthesis, particularly when one enantiomers has a
negative effect. This problem is often solved by two different approaches:
1. Enzymatic resolution: the racemate is transformed in such a way that the
resulting product may be then cleaved by an enzyme (e.g. lyase). Because of the
stereospecificity of the enzyme, only one of the compounds is cleaved, enabling
separation of cleaved and uncleaved product.
2. Diastereomeric resolution: the racemate is made to react with a specific,
enantiomerically pure reagent (e.g. L-tartaric acid, to form an ester). The obtained
products are diastereomers, which unlike enantiomers possess different
properties and are therefore separable. After separation, the reverse reaction is
performed to return the original enantiomers.
Reactions that occur in living organisms follow the same rules of those occurring
in the laboratory. The solvent, temperature and almost certainly the catalyst can be
different, but the fundamental reaction mechanisms are the same. So conveniently,
common organic reaction mechanisms can be used to understand the equivalent
biological transformations.
The following happens in one concerted step: a nucleophile (N) attacks the carbon
atom as a more electronegative group (X) leaves, with inversion of stereochemistry
through an unstable transition state. Note that N can be a neutral protic nucleophile
that deprotonates after the substitution. This type of mechanism is typical of primary
alkyl halides substitutions. One reaction often performed in a laboratory is an
O-methylation using methyl iodide. The oxygen atom (of an alcohol, for example) acts
as nucleophile substituting the iodide which is a very good leaving group.
example, geranyl diphosphate cleaves at the C-O bond and the corresponding allylic
carbocation, well stabilized by resonance, is attacked by water which deprotonates
to produce geraniol (Scheme 1.1.7).
-O O-
O
R P R 1
P
- O O-
O
a b
Scheme 1.1.9: a) Phosphorane intermediate ‒ analogue of the activated complex ‒ in red are the
axial positions (collinear) and in blue are the equatorial positions (coplanar); b) metaphosphate
(planar, stabilised by resonance) ‒ analogue of a carbocation.
Electrophilic reagents can react with compounds that are electron rich in certain
exposed regions, from which alkenes are a typical example. In these systems, the π
bond results from overlapping of p orbitals and provides regions of increased electron
density above and below the plane of the molecule. π electrons are more loosely bound
than those of a σ-bond so they can interact more easily with a positively charged
electrophilic species, forming a new σ-bond and a carbocation (Scheme 1.1.10). This
in turn rapidly reacts with a nucleophile to form another σ-bond. In this case, the
nucleophile is either an anion or a neutral moiety with free pairs of electron that will
become neutral again eliminating a group or an atom.
14 Functional Groups of Biomolecules and their Reactions
E Nu
C C E C C E C C Nu
In simple terms, aromaticity may be described as a chemical property that arises from
delocalization of electrons in a ring. These electrons may be provided by conjugated
unsaturation, lone pair electrons or even orbitals; the system becomes aromatic when
agreeing with the Huckel rule, which are state that the number of the total conjugated
electrons must be 4n+2 electrons. It is a particularly strong form of resonance
stabilization making aromatic compounds more stable than expected otherwise. The
orbital alignment required for aromatic stabilization turns the aromatic (aryl) moieties
planar (Figs. 1.1.4 and 1.1.5). Because of the intermediate character of its bonds (Fig.
1.1.4) the reactivity of aromatic compounds is not identical to the reactivity of other
unsaturated compounds.
Figure 1.1.5: Cytosine is an example of a biomolecule exhibiting aromaticity in its imidic acid form.
Furan is an example of a heteroaromatic compound where a lone pair of electrons of the heteroatom
is delocalized into the ring to allow aromaticity. The Fig. explicitly shows all the resonance forms in
the rings of the compounds.
The reactivity of the aromatic ring is enhanced if electron donating (hydroxyl, amino
and alkoxyl) groups are attached to the ring, increasing the electron density and
making the attack to the electrophile easier.
Common Mechanisms in Biological Chemistry 17
Scheme 1.1.14: Reaction mechanism for the synthesis of DMAT from tryptophan and dimethylallyl
phosphate.
18 Functional Groups of Biomolecules and their Reactions
Scheme 1.1.15: Dehydrogenation of ethane to form ethene through two different mechanisms:
proton-hydride transfer and a radical mechanism, both leading to the elimination of what is equi-
valent to a hydrogen molecule (the arrows without origin represent electron transfers to/from other
molecules).
neighbouring hydrogen atom (elimination). Steric hindrance both at the base and at
the carbocation, as well as stronger bases promote elimination.
H B H
C C C C C C + BH+
X + X-
B + BH+
C C C C
+ X-
X
E2 mechanisms occur through a concerted transfer of a set of electron pairs from the
base to the more electronegative group (X), the latter leaving as its anion and a vicinal
proton is transferred to the base (Scheme 1.1.17). This mechanism is preferred if no
stable ionic intermediate can be formed. The double bond is formed in a step involving
the two reagents, thus the elimination mechanism is bimolecular. It is noteworthy
that the E2 mechanism competes with the SN2 mechanism, as both reactions involve a
base and an electronegative leaving group. More steric hindrance and stronger bases
favour an E2 mechanism over SN2.
H BH+ X-
B
C C C C C C
X X
E1cB mechanism is the symmetric version of the E1 mechanism. First, the proton
is removed from the main molecule to form its conjugate base, a carbanion, which
promotes the elimination of the electrophilic group. This mechanism is preferred
whenever a carbanion intermediate is stabilized, in most of the cases by resonance
(Scheme 1.1.18). This mechanism is particularly relevant in biological transformation,
as it is by far the most frequent elimination mechanism because of the high occurrence
of carbonyl compounds which form relatively stable carbanions.
This type of reaction happens between a nucleophile and a carbonyl group where a
pair of electrons from the nucleophile is transferred to the carbonyl carbon (Scheme
1.1.19).
δ−
O O
Nu
+
Cδ C
R R
Nu R R
1 2 1 2
If the nucleophile is neutral and gives a pair of electrons then it will acquire a positive
charge, which in turn is compensated by a deprotonation of the introduced group
(as in the formation of imines; Scheme 1.1.20c). If the nucleophile is anionic, no
positive charge is generated. In both cases, a negative charge is generated at the most
electronegative atom, the carbonyl oxygen, which is usually neutralized through
protonation (Scheme 1.1.20). Scheme 1.1.20 illustrates some of typical nucleophilic
addition reactions of different nucleophiles to aldehydes and ketones.
All the reactions depicted in Scheme 1.1.20 are called “direct” or [1,2]-additions.
A special case of nucleophilic addition reactions is the Michael or [1,4]-conjugate
addition. This reaction is quite frequent in biochemical pathways and consists of
a nucleophile addition to the β–position of an α,β-unsaturated carbonyl system
(Scheme 1.1.21).
Common Mechanisms in Biological Chemistry 21
a b c d
- Nu = RNH2
Nu = H Nu = ROH Nu = H2N-NH2
O O O O
C C R C R C NH 2
H O N N
H H2 H2
OH OH OH OH
C C R C R C NH 2
H O N N
H H
OR
C R C R C NH2
O N N
acetal imine hydrazone
Scheme 1.1.20: Nucleophilic addition reactions: a) with a hydride ion as nucleophile; b) with an
alcohol as nucleophile; c) with an amine as nucleophile; and d) with a hydrazine as nucleophile.
The reaction steps are the same as in any other nucleophilic carbonyl addition reaction
followed by dehydration:
1. nucleophilic attack of the nucleophilic amino moiety to the electrophilic carbo-
nyl carbon;
2. deprotonation of the positive nitrogen atom and protonation of the negative
oxygen atom;
3. elimination of water by protonation of the hydroxyl group and deprotonation at
the nitrogen atom.
The first step of this mechanism is shared with nucleophilic addition reactions
and consists in the nucleophilic addition to the carbonyl group. The second step is
a regeneration of the carbonyl group with elimination of the anion. It is important
to note that the whole reaction is reversible unless some product stabilization is
provided, such as if the leaving group is a stable anion, which is unlikely to be able
to attack the carbonyl again. Acyl phosphates are considered activated analogues of
carboxylic acids, as the leaving phosphate is a very stable anion and it is unlikely
24 Functional Groups of Biomolecules and their Reactions
that a better leaving group is attached to a carbonyl group. This shifts the equilibrium
towards product formation. On the other hand, a simple carboxylic acid would have
the hydroxyl as leaving group which is a strong nucleophile capable of adding onto
the carbonyl again. This is the reason why esters, thioesters and acyl phosphates
play a major role in promoting substitution reactions within biological systems. In
the laboratory environment it is more common the use of carboxylic acid anhydrides
and acyl chlorides or bromides. The resulting anions (carboxylates and halogens,
respectively) are very weak bases promoting the completeness of the reaction. Because
of that, these compounds are very sensitive to hydrolysis thus rendered useless in
biological systems.
Scheme 1.1.25: Simplified mechanism for the synthesis of aspirin from salicylic acid and acetic anhy-
dride. This reaction is normally performed under acid catalysis with sulfuric or phosphoric acids.
Scheme 1.1.26: Simplified mechanism of the transesterification reaction of aspirin with the serine
residue.
Scheme 1.1.27: General mechanism for an aldol condensation reaction. The aldol adduct suffers a
dehydration for which the mechanism is not specified since the precise step sequence may vary
according to catalyst used.
Firstly, the enolate is generated from an ester by α-deprotonation with a base. The
enolate performs an acyl substitution followed by elimination of the alcoxyl group,
giving a β-keto ester (Scheme 1.1.28).
Oxidation and reduction reactions, or simply redox reactions, are a complex class of
reactions with diverse mechanisms.
The procedure for determining the oxidation state of a carbon atom is similar to
other compounds (Fig. 1.1.6):
Common Mechanisms in Biological Chemistry 29
1. For each bond to less electronegative atoms, such as a hydrogen atoms, count
as -1
2. For each bond to more electronegative atoms, such as oxygen atoms, count as +
3. Bonds between carbon atoms do not affect the oxidation state, unlike other
elements.
H H H CH3 H OH H OPO3 2- O
C C C C C
H H H H H H H3C H H H
-4 -3 -2 -1 0
H SH Cl CH 3 O O
C C C C
H 3C NH2 H3C OH H3C OH H 2N OH
+1 +2 +3 +4
Figure 1.1.6: Compounds with the oxidation state of the highlighted carbon indicated below each
structure. In order: methane, ethane, methanol, ethyl phosphate, formaldehyde, (R)-1-aminoethane-
1-thiol, 1,1-dichloroethan-1-ol, acetic acid and carbamic acid.
Scheme 1.1.31: General mechanism for the oxidation of an alcohol to a carbonyl compound with a
metal or its complex (M). Note that the M leaves in a lower oxidation state.
Figure 1.1.7: Structure of NAD+ and NADH. The structures of NADP+ and NADPH are identical with
exception of the highlighted hydroxyl group (in magenta), which is replaced by a phosphate group.
Scheme 1.1.32: Mechanism of an alcohol oxidation by NAD+ or NADP+ ‒ the base may be provided by
residues of an enzyme.
Common Mechanisms in Biological Chemistry 31
Scheme 1.1.33: Mechanism of an alcohol reduction of a carbonyl group by NAD or NADP, in this case
of an acetylated acyl carrier protein (ACP), an important part in the fatty acid synthesis.
Scheme 1.1.35: Mechanism of disulphide linkage formation by a glutathione dimer, releasing two
reduced glutathione molecules.
32 Functional Groups of Biomolecules and their Reactions
The catabolism of fatty acids (saturated or unsaturated) starts with chemical activation
by esterification with coenzyme A (Scheme 1.1.37), with the following steps occurring
in an acyl-CoA synthase:
1. The carboxylate acts as a nucleophile to attack the double P‒O bond in ATP;
2. The diphosphate group leaves, resulting in an activated acid in the form of a
mixed anhydride (acyl AMP). (PPi is a good leaving group as it is stable and not
nucleophilic)
3. Another acyl substitution is performed on the mixed anhydride by the thiol
moiety in CoA (a moderate nucleophile). Again, this is facilitated because AMP is
a good leaving group.
With the thioester (activated fatty acid) formed, β-oxidation may now occur through a
sequence of dehydrogenation, hydration and dehydrogenation reactions:
1. The Cα-Cβ single bond is oxidised to a double bond, with flavin adenine nucleotide
(FAD) as the oxidant. This step occurs in a family of acyl-CoA dehydrogenases
where the products are FADH2 (reduced FAD) and α,β-unsaturated acyl-CoA. The
mechanistic details of this step are not yet fully resolved, though it is known that
one of the α protons may be first attacked by 376-Glu and a β-hydride is abstracted
from the fatty acid by FAD.
acyl CoA a ,b-unsaturated acyl CoA
(thioester) FAD FADH2
O O
H2 H
C C
C SCoA C SCoA
H2 H
The pathway is repeated until all of the fatty acid is oxidised. If, during the cleavage
process, a cis-oriented unsaturation is encountered, the stereochemistry of the double
bond is switched to trans by an isomerase and the reaction sequence continues as
for saturated fatty acids, since the enoyl-CoA hydratase is stereospecific for a trans-
configuration.
The Organic Mechanisms of Biological Transformations 35
Historically, penicillin antibiotics are one of the most important classes of antibiotics
and are worth studying since many different reactions in its synthesis and mode of
action are involved. The essential feature in penicillin class antibiotic’s structure is
shown in Fig. 1.1.8. Penicillins are part of the bigger class of β-lactam antibiotics,
named because the β-lactam ring is the active site of the compound. A lactam is a
cyclic amide, and the β means there are 2 carbons in between the O and N atoms in
their structure.
Figure 1.1.8: On top, the penicillin core structure is depicted. The β-lactam ring is highlighted in
magenta, and in yellow the thiazolidine ring. R is a variable group and the structures on the bottom
are those of the named antibiotics in this group. There are many other known possible groups that
improve certain properties (absorption, hydrolysis, resistance, etc.).
2. Another acyl substitution reaction involves a second peptide chain that displaces
the link to the hydroxyl of a serine residue resulting in the two peptide chains
being linked.
Scheme 1.1.38: Mechanism of the acyl substitution reaction occurring with penicillin and the serine
residue from transpeptidase.
The Organic Mechanisms of Biological Transformations 37
Scheme 1.1.40: Mechanism for the synthesis of isopenicillin-N by the action of isopenicillin-N
synthase.
Substitutions of the side chain can also be performed by penicillin acylase (PA) to
give other types of penicillin. PA catalyses the hydrolysis and synthesis of the amide
bond to some side chain. The use of this enzyme in vitro enables the formation of
semi-synthetic penicillin.
The Organic Mechanisms of Biological Transformations 39
The most known intervention of coenzyme NAD+ is its role in the ethanol metabolism,
where it acts as a common hydride acceptor (Scheme 1.1.42).
Flavin adenine dinucleotide (FAD) (Fig. 1.1.10) is another coenzyme used in oxidation
reactions and is normally associated tightly to a protein, usually designated as a
flavoprotein. For example, in chorismate synthase, FAD is linked through the flavin
ring C(2)=O to one protein histidine residue (His106) (Scheme 1.1.44). FAD is considered
a more versatile coenzyme than NAD+, because it can participate in oxidation
reactions by several different mechanisms. For example, oxidation of dihydrolipoate
involves a nucleophilic attack on C-4a whereas oxidation of an amino acid involves a
nucleophilic attack on the N-5 position (Scheme 1.1.43).
References
Allen, K. N., & Dunaway-Mariano, D. (2004). Phosphoryl group transfer: evolution of a catalytic
scaffold. Trends in Biochemical Sciences, 29(9), 495–503.
Cleland, W., & Hengge, A. (1995). Mechanisms of phosphoryl and acyl transfer. FASEB J, 9(15),
1585–1594.
Fushinobu, S., Nishimasu, H., Hattori, D., Song, H.-J., & Wakagi, T. (2011). Structural basis for the
bifunctionality of fructose-1,6-bisphosphate aldolase/phosphatase. Nature, 478(7370), 538–
541.
Gebler, J. C., Woodside, A. B., & Poulter, C. D. (1992). Dimethylallyltryptophan synthase. An enzyme-
catalyzed electrophilic aromatic substitution. Journal of the American Chemical Society, 114(19),
7354–7360.
Imai, S.-i. & Guarente, L. (2014). NAD+ and sirtuins in aging and disease. Trends in Cell Biology,
24(8), 464-470.
IUPAC, Commission on Nomenclature of Organic Chemistry. (1993). A Guide to IUPAC Nomenclature
of Organic Compounds (Recommendations 1993). Blackwell Scientific publications.
Bibliography 43
Kirchberg, K., Kim, T.-Y., Haase, S. & Alexiev, U. (2010). Functional interaction structures of the
phochromic retinal protein rhodopsin. Photochemical & Photobiological Sciences, 9, 226-233.
Kitzing, K., Auweter, S., Amrhein, N. & Macheroux, P. (2004). Mechanism of Chorismate Synthase.
The Journal of Biological Chemistry, 279(5), 9451-9461.
Lietzan, A. D., Lin, Y. & St. Maurice, M. (2014). The role of biotin and oxamate in the
carboxyltransferase reaction of pyruvate carboxylase. Archives of Biochemistry and Biophysics,
562, 70-79.
McMurry, J. E. & Bagley, T. P. (2005). The Organic Chemistry of Biological Pathways. Robeerts and
Company Publishers.
Yuan, H. & Marmorstein, R. (2012). Structural basis for sirtuin activity and inhibition. Journal of
Biological Chemistry, 287(14), 42428-42435.
Francisca Lopes, Maria M. M. Santos and Rui Moreira*2
1.2 Designing Covalent Inhibitors: A Medicinal
Chemistry Challenge
1.2.1 Introduction
The design of enzyme inhibitors is one of the most attractive research topics in
Medicinal Chemistry. In particular, covalent inhibitors provide the opportunity of
combining concepts of chemical reactivity and mechanisms of organic reactions with
the structural features required for optimal molecular recognition in order to obtain
the appropriate reactivity and selectivity profile towards the desired enzyme target.
Typically, these inhibitors present an electrophilic functionality capable of reacting
irreversibly with catalytic amino acid residues containing a nucleophilic group (e.g.
serine, threonine and cysteine) (Powers, 2002; Santos, 2007). In spite of its tremendous
potential (Robertson, 2007; Potashman, 2009), designing selective covalent inhibitors
remains a challenging task as the electrophilic groups present in many inhibitor
structures can also react with other macromolecules leading to deleterious (off-
target) events, or can be scavenged by ubiquitous low-molecular-weight nucleophiles
such as glutathione leading to sub-optimal drug concentration at the site of action
(Johansson, 2012). However, there are several examples of covalent inhibitors that are
widely used drugs, including acetyl salicylic acid (the active ingredient of Aspirin),
orlistat (anti-obesity drug) and ampicillin (antibiotic) (Fig. 1.2.1). Many of these drugs
were not originally designed as irreversible inhibitors and their exact mechanism of
action was often discovered afterwards. A recent example is the case of clopidogrel
(antiplatelet agent), which was found to require activation by cytochrome P450 in the
liver to generate an active metabolite containing a free thiol capable of reacting with
a cysteine residue of adenosine 5´-diphosphate (ADP) receptor to form a covalent
disulfide adduct (Fig. 1.2.1). Overall, nearly 30% of the enzymes that are inhibited by
marketed drugs are irreversibly inhibited via covalent modification, which highlights
the therapeutic potential of covalent inhibitors (Robertson, 2005; Robertson, 2007;
Johansson, 2012).
Francisca Lopes, Maria M. M. Santos and Rui Moreira: Research Institute for Medicines (iMed.ULis-
boa), Faculty of Pharmacy, Universidade de Lisboa, Av. Prof. Gama Pinto, 1649-003 Lisboa, Portugal,
*Email: rmoreira@ff.ul.pt
Figure 1.2.1: Drugs acting as covalent inhibitors including their protein target(s) with active-site nuc-
leophiles. Arrows indicate the reaction of the enzyme’s nucleophilic residue with the electrophilic
site at the drug or its metabolite resulting in covalent modification of the target.
covalent irreversible inhibitors. These approaches will be dealt in more detail in the
following sections.
N
H 2N CO2 H NH2
Figure 1.2.2: Mechanism-based covalent inhibitors. For detailed description on their mechanism of
action see Silverman, 1992.
p90 ribosomal protein S6 kinase RSK2, led to the discovery of the first reversible
covalent kinase inhibitor (Fig. 1.2.3B) (Serafimova, 2012). Additional scaffolds were
discovered de novo using a fragment-based ligand approach, where a library of low
molecular-weight cyanoacrylamides was screened against several kinase assays
(Fig. 1.2.3B) (Miller, 2013).
Figure 1.2.3: A) Thiol reactivity of electron-deficient olefins; B) Reversible covalent inhibitors that
selectively target the non-catalytic cysteine-436 present in the C-terminal domain of the p90 riboso-
mal protein S6 kinase RSK2.
48 Designing Covalent Inhibitors: A Medicinal Chemistry Challenge
Figure 1.2.4: β-Elimination half-lives for the adducts from electron-deficient acrylonitriles and
β-mercaptoethanol (BME) – adapted from Krishnan, 2014.
Figure 1.2.5: Selected examples of Michael acceptors that act as cysteine protease inhibitors.
50 Designing Covalent Inhibitors: A Medicinal Chemistry Challenge
These compounds inhibit cysteine proteases by forming covalent bonds with the
active site thiol of cysteine proteases. In general, they are stable, and need the catalytic
machinery of the cysteine proteases for activation. Of all the compounds developed,
two Michael acceptor cysteine protease inhibitors, K-777 and Ruprintrivir (Fig. 1.2.6),
have entered clinical trials.
Figure 1.2.6: Cruzain inhibitor K777 and human rhinovirus 3C protease inhibitor Rupintrivir.
Cruzain, a cysteine protease that belongs to clan CA, is a key protease required for
the survival of Trypanosoma cruzi, the ethiologic agent of Chagas’ disease (McGrath,
1995; Wilkinson, 2009). For that reason, a promising group of drug leads for Chagas’
disease is cysteine protease inhibitors targeting cruzain. In 1998, McKerrow’s
group reported a dipeptide vinyl sulfone, K-777 (Fig. 1.2.6), that rescued mice from
a lethal Trypanosoma cruzi infection by inhibition of cruzain (Engel, 1998). Proof
of mechanism came from the crystal structure of inhibitor K777 bound to cruzain,
where the catalytic cysteine residue (Cys25) is covalently linked to the β-carbon of the
vinylsulfone (Fig. 1.2.7). K777 was shown to be safe and efficacious in animal models
of acute and chronic Chagas disease (Doyle, 2007; Barr, 2005). In particular, this vinyl
sulfone protected Beagle dogs from cardiac damage during infection by T. cruzi (Barr,
2005). The results of preclinical trials showed that K777 was non-mutagenic, well-
tolerated, and demonstrated efficacy in models of acute and chronic Chagas’ disease
in both mice and dogs (Kerr, 2009). On the basis of these results the inhibitor entered
clinical trials, but in 2013 the clinical assays were stopped due to tolerability findings
at low dose in primates and dogs (http://www.dndi.org/diseases-projects/portfolio/
k777.html).
Case Study 2: From Covalent Inhibitors to Hybrid Drugs 51
Figure 1.2.7: The crystal structure of inhibitor K777 bound to cruzain (pdb 2OZ2). The insert shows
the interaction of K777 with the primary recognition site (S2) and secondary recognition sites (S1 and
S3) of cruzain.
The human rhinovirus 3C protease (3CP) belongs to clan PA(C). Because the rhinoviral
protease (3CP) is specific to rhinovirus, inhibitors of this protease should have few
side effects in humans. Several compounds containing a Michael acceptor moiety
were described as potent Human rhinovirus (HRV) 3C protease inhibitors (Santos,
2007). Specifically, Rupintrivir (also known as Ruprintrivir or AG7088) was developed
by Agouron Pharmaceuticals to treat HRV infections (Fig. 1.2.6). Rupintrivir is an α,β-
unsaturated carbonyl that irreversibly inhibits the HRV 3C protease and shows potent,
broad-spectrum anti-HRV activity in vitro (Hayden, 2003). The phase I and phase II
results showed good safety and pharmacokinetic profiles. Although it successfully
completed the initial Phase II trials (Hayden, 2003), due to lack of efficacy in natural
infection studies, further development of Rupintrivir was stopped in PhaseII/III trials
(http://www.drugdevelopment-technology.com/projects/ag7088/).
Falcipain-2 (FP-2) and falcipain-3 (FP-3) are cysteine proteases, located in the digestive
vacuole of the malaria parasite and involved in the digestion of host hemoglobin
essential for the survival of Plasmodium falciparum, the causing agent of the most
lethal form of malaria. Rosenthal and co-workers demonstrated that inhibition
of these enzymes allowed the cure of murine models of the disease and thus they
constitute a validated target for its treatment. Generally, the inhibitors of falcipains
have a backbone recognition element coupled with an electrophilic warhead such
as aldehydes, fluoro methyl ketones, vinyl esters, vinyl sulfones (Fig. 1.2.6), and
chalcones that can react with the catalytic cysteine residue (Shenai, 2003). However,
the intrinsic reactivity of many of these warheads has prompted the development
of hybrid compounds in order to modulate reactivity, improve selectivity and to
avoid the emergence of resistant strains. Several hybrid compounds were prepared
combining endoperoxides with different inhibitors of hemoglobin digestion. The goal
was to bring together two drugs that can act in the digestive vacuole by two different
mechanisms of action and taking advantage of the iron (II) coming from the digestion
of the host hemoglobin (Meunier, 2008).
vinyl sulfone (Fig. 1.2.9). Importantly, the authors were able to detect the intact parent
vinyl sulfone in infected erythrocytes 48h post-treatment with the hybrid compounds,
which strongly suggests that the vinyl sulfone inhibitor remained within infected
erythrocytes long after the fast-acting endoperoxide moiety had exerted its effects.
These results indicate that the intrinsic activity of the vinyl sulfone partner can be
masked, suggesting that a tetraoxane-based delivery system offers the potential to
attenuate the off-target effects of known drugs (e.g. irreversible enzyme inhibitors
such as Michael acceptors).
Electrophilic warhead
H O
O O H
O N SO 2Ph
N
H
O O O O O
O Ph
O
1,2,4,5-tetraoxane-vinyl sulfone hybrid
Electrophilic warhead
Electrophilic warhead
O O Br
H O O N
N S H
O N Ph N
H N N
O O O N
O
Ph
the pyrimidine nitrile tetraoxane hybrids were shown to deliver the corresponding
peptidomimetic pyrimidine nitrile inside the parasite upon activation by Fe(II).
Figure 1.2.9: Tetraoxane-vinylsulfone hybrid compounds and their activation by Fe(II) from host
hemoglobin in infected erythrocytes (Oliveira, 2013). Insert: delivery of the Michael acceptor (vinyl
sulfone) via iron binding to oxygen-2 leads to the swelling of the parasitic digestive vacuole.
The same group extended the concept to a new type of hybrids combining a 1,2,4-
trioxolane moiety with peptidic cysteine protease inhibitors. They demonstrated that
in the presence of Fe(II) the 1,2,4-trioxolane scaffold decomposes into a potentially
toxic carbon radical that alkylates the heme in vitro, and a Michael acceptor that acts
as a cysteine protease inhibitor (Fig. 1.2.11). The most potent compound of the series
presented an IC50 value in the low nanomolar region (35 nM) against the 3D7 strain of
P. falciparum. Remarkably, although the same compound showed inhibitory activity
of FP-2 in the submicromolar range (0.5 µM), the corresponding aldehyde released
upon Fe(II) activation (Fig. 1.2.11) was shown to inhibit FP-2 with an IC50 value of
16 nM, thus confirming 1,2,4-trioxolane scaffold as a site-specific delivery system for
electrophilic aldehydes and ketones (Gibbons, 2010).
56 Designing Covalent Inhibitors: A Medicinal Chemistry Challenge
Figure 1.2.11: Fragmentation of 1,2,4-trioxolane to secondary carbon centred radical species and
protease inhibitor, active against falcipain-2 from P. falciparum. Adapted from Gibbons, 2010.
1.2.5 Conclusions
Designing selective covalent inhibitors remains one of the most attractive areas of
Medicinal Chemistry, where concepts of Organic Chemistry and Biochemistry are
used to discover compounds that react preferentially with the active site of the target
enzyme. The chemical basis to design irreversible and reversible covalent inhibitors
is now well established, providing the medicinal chemist with a variety of solutions to
modulate intrinsic reactivity and reduce the likelihood of toxicity issues. The chemical
toolbox is further expanded with the concept of molecular hybridization to modulate
the reactivity of electrophilic warheads or even to mask these chemical functionalities.
This chemical toolbox is already delivering interesting drug candidates into the
pipelines of pharmaceutical companies, small biotech and academic laboratories.
References
Barr, S. C., Warner, K. L., Kornreic, B. G., Piscitelli, J., Wolfe, A., Benet, L., & McKerrow, J. H.
(2005). A cysteine protease inhibitor protects dogs from cardiac damage during infection by
Trypanosoma cruzi. Antimicrobial Agents and Chemotherapy, 49, (12), 5160-5161.
Capela, R., Oliveira, R., Gonçalves, L. M., Domingos, A., Gut, J., Rosenthal, P. J., Lopes, F., & Moreira,
R. (2009). Artemisinin-dipeptidyl Vinyl Sulfone Hybrid Molecules: Design, Synthesis and
Preliminary SAR for Antiplasmodial Activity and Falcipain-2 Inhibition. Bioorganic & Medicinal
Chemistry Letters, 19, 3229-3232.
Coterón, J. M., Catterick, D., Castro, J., Chaparro, M. J., Díaz, B., Fernández, E., Ferrer, S., Gamo, F.
J., Gordo, M., Gut, J., de Las Heras, L., Legac, J., Marco, M., Miguel, J., Muñoz, V., Porras, E.,
de La Rosa, J. C., Ruiz, J. R., Sandoval, E., Ventosa, P., Rosenthal, P. J., & Fiandor, J. M. (2010).
Falcipain Inhibitors: Optimization Studies of the 2-Pyrimidinecarbonitrile Lead Series. Journal
of Medicinal Chemistry, 53, 6129–6152.
Doyle, P. S., Zhou, Y. M., Engel, J. C., & McKerrow, J. H. (2007). A cysteine protease inhibitor cures
Chagas’ disease in an immunodeficient-mouse model of infection. Antimicrobial Agents and
Chemotherapy, 51(11), 3932-3939.
Dragovich, P. S., Prins, T. J., Zhou, R., Johnson, T. O., Hua, Y., Luu, H. T., Sakata, S. K., Brown, E. L.,
Maldonado, F. C., Tuntland, T., Lee, C. A., Fuhrman, S. A., Zalman, L. S., Patick, A. K., Matthews,
D. A., Wu, E. Y., Guo, M., Borer, B. C., Nayyar, N. K., Moran, T., Chen, L. J., Rejto, P. A., Rose, P.
W., Guzman, M. C., Dovalsantos, E. Z., Lee, S., McGee, K., Mohajeri, M., Liese, A., Tao, J. H.,
Kosa, M. B., Liu, B., Batugo, M. R., Gleeson, J. P. R., Wu, Z. P., Liu, J., Meador, J. W., & Ferre, R.
A. (2003). Structure-based design, synthesis, and biological evaluation of irreversible human
rhinovirus 3C protease inhibitors. 8. Pharmacological optimization of orally bioavailable
2-pyridone-containing peptidomimetics. Journal of Medicinal Chemistry, 46(21), 4572-4585.
Ehmke, V., Heindl, C., Rottmann, M., Freymond, C., Schweizer, W. B., Brun, R., Stich, A.,
Schirmeister, T., & Diederich, F. (2011). Potent and Selective Inhibition of Cysteine Proteases
from Plasmodium Falciparum and Trypanosoma Brucei. ChemMedChem, 6, 273–278.
Ehmke, V., Quinsaat, J. E. Q., Rivera-Fuentes, P., Heindl, C., Freymond, C., Rottmann, M., Brun, R.,
Schirmeister, T., & Diederich, F. (2012). Tuning and Predicting Biological Affinity: Aryl Nitriles as
Cysteine Protease Inhibitors. Organic & Biomolecular Chemistry, 10, 5764–5768.
Ekici, O. D., Li, Z. Z., Campbell, A. J., James, K. E., Asgian, J. L., Mikolajczyk, J., Salvesen, G. S.,
Ganesan, R., Jelakovic, S., Gruetter, M. G., & Powers, J. C. (2006). Design, synthesis, and
evaluation of aza-peptide Michael acceptors as selective and potent inhibitors of caspases-2,
-3, -6, -7, -8, -9, and -10. Journal of Medicinal Chemistry, 49(19), 5728-5749.
Engel, J. C., Doyle, P. S., Hsieh, I., & McKerrow, J. H. (1998). Cysteine protease inhibitors cure an
experimental Trypanosoma cruzi infection. Journal of Experimental Medicine, 188(4), 725-734.
Gibbons, P., Verissimo, E., Araujo, N. C., Barton, V., Nixon, G. L., Amewu, R. K., Chadwick, J., Stocks,
P. A., Biagini, G. A., Srivastava, A., Rosenthal, P. J., Gut, J., Guedes, R. C., Moreira, R., Sharma,
R., Berry, N., Cristiano, L. S., Shone, A. E., Ward, S. A., & O’Neill, P. M. (2010). Endoperoxide
Carbonyl Falcipain 2/3 Inhibitor Hybrids: Toward Combination Chemotherapy of Malaria
Through a Single Chemical Entity. Journal of Medicinal Chemistry, 53, 8202-8206.
Glória, P. M. C., Coutinho, I., Gonçalves, L. M., Baptista, C., Soares, J., Newton, A. S., Moreira,
R., Saraiva, L., & Santos, M. M. M. (2011). Aspartic vinyl sulfones: inhibitors of a caspase-3-
dependent pathway”, European Journal of Medicinal Chemistry, 46, 2141-2146.
Graczyk, P. P. (1999). Caspase inhibitors - A chemist’s perspective. Restorative Neurology and
Neuroscience, 14(1), 1-23.
Hanzlik, R. P., & Thompson, S. A. (1984). Vinylogous Amino-acid Esters - A New Class of Inactivators
for Thiol Proteases. Journal of Medicinal Chemistry, 27(6), 711-712.
58 Designing Covalent Inhibitors: A Medicinal Chemistry Challenge
Hayden, F. G., Turner, R. B., Gwaltney, J. M., Chi-Burris, K., Gersten, M., Hsyu, P., Patick, A. K., Smith,
G. J., & Zalman, L. S. (2003). Phase II, randomized, double-blind, placebo-controlled studies of
ruprintrivir nasal spray 2-percent suspension for prevention and treatment of experimentally
induced rhinovirus colds in healthy volunteers. Antimicrobial Agents and Chemotherapy, 47(12),
3907-3916.
Johansson, M. H. (2012). Reversible Michael Additions: Covalent Inhibitors and Prodrugs. Mini-
Reviews in Medicinal Chemistry, 12(13), 1330-1344.
Kathman, S. G., Xu, Z., & Statsyuk, A. V. (2014). A Fragment-Based Method to Discover Irreversible
Covalent Inhibitors of Cysteine Proteases. Journal of Medicinal Chemistry, 57, 4969-4974.
Kerr, I. D., Lee, J. H., Farady, C. J., Marion, R., Rickert, M., Sajid, M., Pandey, K. C., Caffrey, C. R.,
Legac, J., Hansell, E., McKerrow, J. H., Craik, C. S., Rosenthal, P. J., & Brinen, L. S. (2009). Vinyl
Sulfones as Antiparasitic Agents and a Structural Basis for Drug Design. Journal of Biological
Chemistry, 284(38), 25697-25703.
Krishnan, S., Miller, R. M., Tian, B., Mullins, R. D., Jacobson, M. P., & Taunton J. (2014). Design of
Reversible, Cysteine-Targeted Michael Acceptors Guided by Kinetic and Computational Analysis.
Journal of the American Chemical Society, 136, 12624−12630.
Kumar, S. P., Glória, P. M. C., Gonçalves, L. M., Gut, J., Rosenthal, P. J., Moreira, R., & Santos, M. M.
M. (2012). Squaric acid: a valuable scaffold for developing antimalarials?, MedChemCommun,
3, 489-493.
Liu, S., & Hanzlik, R. P. (1992). Structure-activity-relationships for Inhibition of Papain by Peptide
Michael Acceptors. Journal of Medicinal Chemistry, 35(6), 1067-1075.
Lucas, S. D., Costa, E., Guedes, R. C., & Moreira, R. (2013) Targeting COPD: Advances on low-
molecular-weight inhibitors of human neutrophil elastase. Medicinal Research Reviews, 33,
E73-E101
McGrath, M. E., Eakin, A. E., Engel, J. C., McKerrow, J. H., Craik, C. S., & Fletterick, R. J. (1995). The
Crystal-structure of Cruzain - A Therapeutic Target for Chagas-disease. Journal of Molecular
Biology, 247(2), 251-259.
Meunier, B. (2008). Hybrid molecules with a dual mode of action: Dream or reality? Accounts of
Chemical Research, 41, 69-77.
Miller, R. M., Paavilainen, V. O., Krishnan, S., Serafimova, I. M., & Taunton, J. (2013). Electrophilic
fragment-based design of reversible covalent kinase inhibitors. Journal of the American
Chemical Society, 135, 5298-5301.
Oliveira, R., Newton, A. S., Guedes, R. C., Miranda, D., Amewu, R. K., Srivastava, A., Gut, J.,
Rosenthal, P. J., O’Neill, P. M., Ward, S. A., Lopes, F., & Moreira, R. (2013). An Endoperoxide-
based Hybrid Approach to Deliver Falcipain Inhibitors Inside Malaria Parasites. ChemMedChem,
8, 1528-1536.
Oliveira, R., Guedes, R. C., Meireles, P., Albuquerque, I. S., Gonçalves, L. M., Pires, E., Bronze, M. R.,
Gut, J., Rosenthal, P. J., Prudêncio, M., Moreira, R., O’Neill, P., & Lopes, F. (2014). Tetraoxane-
pyrimidine nitrile as dual stage antimalarials. Journal of Medicinal Chemistry, 57, 4916-4923.
Olson, J. E., Lee, G. K., Semenov, A., & Rosenthal, P. J. (1999). Antimalarial effects in mice of orally
administered peptidyl cysteine protease inhibitors. Bioorganic & Medicinal Chemistry, 7(4),
633-638.
O’Neill, P. M., Stocks, P. A., Pugh, M. D., Araujo, N. C., Korshin, E. E., Bickley, J. F., Ward, S. A.,
Bray, P. G., Pasini, E., Davies, J., Verissimo, E., & Bachi, M. D. (2004). Design and Synthesis
of Endoperoxide Antimalarial Prodrug Models. Angewandte Chemie International Edition, 43,
4193-4197.
Potashman, M. H., & Duggan, D. E. (2009). Covalent modifiers: an orthogonal approach to drug
design. Journal of Medicinal Chemistry, 52, 1231–1246.
Powers, J. C., Asgian, J. L., Ekici, O. D., & James K. E. (2002). Irreversible inhibitors of serine,
cysteine, and threonine proteases. Chemical Reviews, 102, 4639–4750.
References 59
Pritchard, R. B., Lough, C. E., Currie, D. J., & Holmes, H. L. (1968). Equilibrium Reactions of
N-Butanethiol with Some Conjugated Heteroenoid Compounds. Canadian Journal of Chemistry,
46, 775–781.
Robertson, J.G. (2005). Mechanistic basis of enzyme-targeted drugs. Biochemistry, 44, 5561–5571.
Robertson, J.G. (2007). Enzymes as a special class of therapeutic target: clinical drugs and modes of
action. Current Opinion in Structural Biology, 17, 674–679.
Roush, W. R., Cheng, J. M., Knapp-Reed, B., Alvarez-Hernandez, A., McKerrow, J. H., Hansell, E., &
Engel, J. C. (2001). Potent second generation vinyl sulfonamide inhibitors of the trypanosomal
cysteine protease cruzain. Bioorg. ACS Medicinal Chemistry Letters, 11(20), 2759-2762.
Santos, M. M. M., & Moreira, R. (2007). Michael acceptors as cysteine protease inhibitors. Mini-
Reviews in Medicinal Chemistry, 7(10), 1040-1050.
Serafimova, I. M., Pufall, M. A., Krishnan, S., Duda, K., Cohen, M. S., Maglathlin, R. L., McFarland, J.
M., Miller, R. M., Frödin, M., Taunton, J. (2012). Reversible targeting of noncatalytic cysteines
with chemically tuned electrophiles. Nature Chemical Biology, 8, 471-476.
Shenai, B. R., Lee, B. J., Alvarez-Fernandes, A., Chong, P. Y., Emal, C. D., Neitz, R. J., Roush, W. R.,
& Rosenthal, P. J. (2003). Structure-activity Relationships for Inhibition of Cysteine Protease
Activity and Development of Plasmodium falciparum by Peptidyl Vinyl Sulfones. Antimicrobial
Agents and Chemotherapy, 47, 154-160.
Singh, J., Petter, R.C., Baillie, T.A., & Whitty, A. (2011). The resurgence of covalent drugs. Nature
Reviews Drug Discovery, 10, 307–317.
Tan, J., George, S., Kusov, Y., Perbandt, M., Anemueller, S., Mesters, J. R., Norder, H., Coutard,
B., Lacroix, C., Leyssen, P., Neyts, J., & Hilgenfeld, R. (2013). 3C Protease of Enterovirus 68:
Structure-Based Design of Michael Acceptor Inhibitors and Their Broad-Spectrum Antiviral
Effects against Picornaviruses. Journal of Virology, 87(8), 4339-4351.
Thompson, S. A., Andrews, P. R., & Hanzlik, R. P. (1986). Carboxyl-modified Amino-acids and
Peptides as Protease Inhibitors. Journal of Medicinal Chemistry, 29(1), 104-111.
Wilkinson, S. R., & Kelly, J. M. (2009). Trypanocidal drugs: mechanisms, resistance and new targets.
Expert Reviews in Molecular Medicine, 11, e31.
Wilson, A. J., Kerns, J. K., Callahan, J. F., & Moody, C. J. (2013). Keap Calm, and Carry on Covalently.
Journal of Medicinal Chemistry, 56(19), 7463-7476.
Section 2: Chemical Basis of Drug Action and
Diseases
Ana Fortuna, Gilberto Alves, Amílcar Falcão3
2.1 Pharmacokinetics and Bioanalysis to Improve
Drug Development
Abbreviations
2.1.1 Introduction
The first decade of the 21st century was a challenging period for the pharmaceutical
industries focused on drug research. Several changes in technology and market
dynamics have occurred over the last years and they inexorably revolutionized
the drug discovery and development (DDD) process. The tremendous progress
in biomedical knowledge and the completion of the human genome sequencing
project have accelerated the discovery of the molecular basis of human diseases
and have allowed the identification of important new biological therapeutic targets.
In addition, the synergistic interaction between rational drug design, recombinant
biotechnology and combinatorial chemistry prompted a fast increment of the number
of new chemical entities (NCEs) with molecular characteristics propitious to interact
with specific biological targets (Ohlmeyer, 2010).
Despite this exponential increase in the number of NCEs, pharmaceutical
industry is suffering from a lack of productivity considering the number of new drugs
that have successfully reached the market (Keserü, 2009). According to recent news, it
64 Pharmacokinetics and Bioanalysis to Improve Drug Development
is estimated that only one in ten drug candidates that enter the clinical development
phase will reach the market and the success rate for drugs to be approved considering
all the therapeutic areas is only 11% (Tamimi, 2009).
The current climate of downward pressure on healthcare costs combined with
the need of expensive technologies and greater information about drugs under
development prompted pharmaceutical companies to develop paradigm-shifting
strategies and tactics in order to cut drug development costs, fill the depleting product
pipelines and reverse the trend of dropping productivity. A well-succeeded front-
loading approach consisted on selecting, in the early drug discovery stage, the drug
candidates that are not only pharmacologically active but also those having desirable
pharmacokinetic properties. It is particularly disheartening when the in vitro potency
is achieved, but the lead compound fails in vivo because, for instance, it is so short-
lived that its clinical benefits would be minimal. In fact, since the pharmaceutical
industries embraced the early pharmacokinetic screening of drug candidates, the
attrition rates of DDD were substantially improved and, therefore, it is currently
straightforward to make a decision on the pharmacokinetic desirability of each drug
candidate in the beginning of DDD processes and not only later in the non-clinical
and clinical development stages (Saxena, 2008; Gallo, 2010). This demand expanded
the investigation on pharmacokinetics into the initial phases of drug discovery and
several in silico, in vitro and in vivo models have been developed and improved to
optimize the pharmacokinetic behavior of drug candidates at different stages of DDD.
Moreover, pharmacokinetic principles are also applied during the post approval
period of a drug, particularly for therapeutic drug monitoring (which has recently
been defined as therapeutic drug management, TDM), evaluation of line-extension
medicinal products and generic formulations, particularly through bioavailability/
bioequivalence (BA/BE) studies.
Taking into account these new attempts, a close involvement between
pharmacokinetics and bioanalysis (defined as the determination of drug
concentrations on biological specimens) became required in order to obtain data
with a high quality level and allow an adequate interpretation and ultimately rational
decision-making. On the other hand, the increasing number of NCEs that must be
investigated and the application of pharmacokinetic screening at initial phases of
DDD increases the number of samples that must be analyzed and, consequently,
new bioanalytical techniques with higher sensitivity, selectivity and speed have
emerged. Indeed, decades ago, high performance liquid chromatography (HPLC)
coupled with ultraviolet (UV) or fluorescence detection and gas chromatography
(GC) with mass spectrometry (MS) ensured the acquisition of plasma concentration
data to define drug exposure in animal tests and clinical studies. In current practices,
liquid chromatography (LC) coupled to MS or tandem mass spectrometry (MS/MS) are
becoming essential due to their higher throughput capacity to identify and quantify
the parent drugs and metabolites in biological matrices, enabling fast and effective
decision-making. Furthermore, new strategies for sample preparation have also been
Pharmacokinetic on Drug Discovery and Development Process 65
The pharmacokinetic principles applied during the process of DDD have undergone a
revolution in the last decades, as it was understood that the efficacy of any successful
drug depends not only on its pharmacological potency but also on its absorption,
distribution, metabolism and excretion (ADME) characteristics.
Over the time, it became clear that drug candidates with low and/or high variable
bioavailability and/or short half-life in animal studies were frequently prevented
to progress to clinical trials. Furthermore, poor pharmacokinetic properties of new
drug candidates in phase I clinical trials accounted for their clinical failure, while the
incidence of severe (toxic) adverse side effects resulting from drug metabolism and
drug-drug interactions (DDI) observed during phase II and III clinical trials may lead to
the termination of potential new drugs or the withdrawal of marketed drugs. Although
ADME studies are undeniably important throughout the stages of (non)-clinical drug
development, some of the aforementioned failures could have been avoided if the
pharmacokinetic deficits of the drug candidate were identified earlier in the discovery
phase. The recognition that understanding the ADME processes of test compounds
would potentially prevent their unsuccessful progression into clinical development
phases expanded the ADME/pharmacokinetics interest from its traditional role as
non-clinical safety support towards the early stage of drug discovery. Today, ADME/
66 Pharmacokinetics and Bioanalysis to Improve Drug Development
Table 2.1.1: The role of pharmacokinetics in different stages of drug discovery and development
process (Panchagnula, 2000; Chien, 2005; Bhogal, 2008; Zhang, 2012).
Continued
Table 2.1.1: The role of pharmacokinetics in different stages of drug discovery and
development process (Panchagnula, 2000; Chien, 2005; Bhogal, 2008; Zhang, 2012).
Continued
Table 2.1.1: The role of pharmacokinetics in different stages of drug discovery and
development process (Panchagnula, 2000; Chien, 2005; Bhogal, 2008; Zhang, 2012).
Determination of eventual
influence of other drugs on the
pharmacokinetic profile;
ADME, absorption, distribution, metabolism and excretion; NCE, new chemical entity
Table 2.1.2: Major in vitro model systems used to investigate the absorption, distribution and
metabolism of new chemical entities in initial phases of drug discovery.
Table 2.1.2: Major in vitro model systems used to investigate the absorption, distribution and
metabolism of new chemical entities in initial phases of drug discovery.
Transfected MDCK cell line MDCK are previously transfected with Evaluate the efflux transport
P-glycoprotein gene to over-express the via P-glycoprotein.
transporter. Identification of
P-glycoprotein substrates and
5 days of incubation is usually required inhibitors in drug discovery.
to initiate the studies. Evaluation of drug-drug
Bi-directional studies and compound interaction potential.
quantification by HPLC with simple
sample preparation for bioanalysis.
Table 2.1.2: Major in vitro model systems used to investigate the absorption, distribution and
metabolism of new chemical entities in initial phases of drug discovery.
Table 2.1.2: Major in vitro model systems used to investigate the absorption, distribution and
metabolism of new chemical entities in initial phases of drug discovery.
Table 2.1.2: Major in vitro model systems used to investigate the absorption, distribution and
metabolism of new chemical entities in initial phases of drug discovery.
BCRP, breast cancer resistance protein; Caco-2, human colon adenocarcinoma cell line; CYP, cytochrome
P450; DDI, drug-drug interaction; HPLC, high performance liquid chromatography; HTS, high throughput
screening assay; MDCK, Madin-Darby canine kidney cell line; PAMPA, parallel artificial membrane
permeability assay; Papp, apparent permeability; UGT, UDP glucuronosyltransferase.
Pharmacokinetic on Drug Discovery and Development Process 75
–– Relevance: the result of the screen should have good concordance with the cor-
responding in vivo properties of the drug, requiring validation of the screen with
standard compounds of known animal or human performance;
–– Effectiveness: the cut-off criterion should eliminate a substantial fraction of com-
pounds without eliminating every compound, otherwise the screen would merely
add a useless extra step that delays the discovery process;
–– Speed: the experimental procedure must be fast enough to keep pace with the
input rate of new compounds from chemistry;
–– Robustness: the experimental procedure must be applicable to a wide variety of
chemical structures;
–– Accuracy and reproducibility: any screening procedure has an inevitable charac-
teristic error rate, because it is usually necessary to sacrifice some accuracy or
precision to achieve the desired speed. Thus, when a large number of compounds
is carried through a particular screen model, some of the compounds will be clas-
sified incorrectly. False positives are tolerable, but false negatives may be lost
forever if the failure eliminates them from further testing.
It is also essential to regularly validate the in vitro models against in vivo data to ensure
that decisions based on in vitro results are sufficiently predictive (Burton, 2010).
Once exhibiting these characteristics, in vitro assays contribute to understanding
the underlying mechanisms of drug absorption and disposition, which are invaluable
to establishing structure-activity relationships to guide subsequent chemical synthesis
of new drugs, avoiding potential obstacles in drug development. It is also important
to highlight that in vitro data are in current practice recognized as good predictors of
in vivo end-points. Even the Food and Drug Administration (FDA) accepts the
importance of in vitro data, as it can inclusively avoid certain clinical studies (FDA,
2000, 2003, 2006). For instance, since a negative result in CYP induction for in vitro
human-based studies using primary hepatocyte cultures always seems to lead to a
negative induction in human clinical studies, FDA typically waives clinical DDI studies
if the drug candidate is tested as negative in a human in vitro CYP induction study.
However, the relative simplicity of in vitro systems for studying ADME/
pharmacokinetics in comparison to real in vivo drug behavior means that absolute
correlations between in vitro and in vivo findings are often difficult to establish
routinely, particularly for humans. For this reason, in vitro screens are frequently
viewed as a means to rank compounds for further study rather than for outright
rejection, so the in vivo pharmacokinetic studies performed in animals during lead
characterization remain crucial after in vitro evaluation.
Without requiring full pharmacokinetic characterization, the in vivo studies
performed at this stage ensure that lead compounds have appropriate pharmacokinetic
properties to be evaluated in non-clinical pharmacology and safety studies (Table 2.1.1).
In fact, when combined with in vitro results, they quickly assess the bioavailability of a
compound, giving a good indication of the suitability for advancement to pre-clinical
76 Pharmacokinetics and Bioanalysis to Improve Drug Development
and clinical trials. The current trend is to use rodents (mice and rats) as first animal
species to test drug exposure because they are less expensive and require smaller
amount of test compound. These in vivo pharmacokinetic studies in rodents include
single oral and intravenous pharmacokinetic studies and provide information such as
drug oral bioavailability, distribution, clearance and duration of exposure for a drug
and its metabolites. Thus, they can help to identify some limitations of a new chemical
series regarding their ADME characteristics, such as low absorption or high clearance,
leading to undesirable pharmacokinetic profiles. Subsequently, in vitro models such as
Caco-2 can be used to optimize absorption of compounds from the same chemical series.
The microsomal stability assay can also be applied to select more stable compounds.
Although a continuous demand for in vivo pharmacokinetic studies remains
at discovery stages, most of these studies are still conducted in a traditional, low-
throughput manner in many pharmaceutical companies. Furthermore, compound
quantities as well as practical and ethical issues preclude in vivo studies performed
at a sufficient rate to investigate the high number of new drug candidates (Balani,
2005). Conventional pharmacokinetic studies include the intravenous and oral
administration of each compound to groups of 6−8 animals with blood samples taken
at 6−12 different points of time (Alves, 2008), requiring a high number of animals,
time and intensive bioanalysis. Consequently, in order to make timely and meaningful
contributions in early discovery programs, in vivo pharmacokinetic analysis has been
efficiently modified and accomplished through recent strategies such as pooling of
plasma samples and fewer sampling time points per study, reducing the number of
samples analyzed per each compound (Table 2.1.3). Interestingly, similar pooling
strategies have also been applied in order to determine brain concentrations and
estimate the concentration ratios in relation to plasma levels (Amore, 2010).
The most promising compounds found in discovery stages are then evaluated in at
least two relevant animal species to obtain pharmacodynamic, pharmacokinetic, and
toxicological information. The animal species are chosen according to the previous in
vitro analysis.
The extensive pharmacokinetic studies performed at this stage include single
and multiple dose administration with several ascending doses, metabolic profiling
and evaluation of drug potential to induce/inhibit drug-metabolizing enzymes.
Major objectives of these studies are depicted in Table 2.1.1. Compartmental and non-
compartmental methods are used to determine multiple pharmacokinetic parameters,
including the maximum concentration (Cmax), time to reach the Cmax (tmax), plasma
drug concentration-time curve (AUC), apparent volume of distribution at steady state
(Vss), clearance (CL), elimination half-life (t1/2) and absolute or relative bioavailability,
providing characterization of the pharmacokinetic profiles for a test compound.
Table 2.1.3: Comparison between the strategies employed to increase the throughput of in vivo pharmacokinetic studies during drug discovery stages (Ward,
2001; Cheung, 2005; Li, 2013).
AUC, area under the curve; CL, systemic clearance; Cmax, maximum concentration; DDI, drug-drug interaction; F, systemic bioavailability; LLOQ, lower limit of
quantification; NCEs, new chemical entities; t1/2, half-life time; Vss, apparent volume of distribution at steady state
Pharmacokinetics and Bioanalysis to Improve Drug Development
Pharmacokinetic on Drug Discovery and Development Process 79
In parallel, toxicity tests are always performed in this stage and include genotoxicity,
safety pharmacology in all biological systems, reproductive toxicology in male and
female animals and long-term carcinogenicity. In this context, toxicokinetics, which is
defined as the generation of pharmacokinetic data under the conditions of the toxicity
studies themselves, is required as an integral component of non-clinical toxicity
studies (Guideline ICH S3A). The primary objective of toxicokinetics is to describe the
systemic exposure achieved in animals and its relationship to the dose level and the
time course of the toxicity study. These objectives are mainly achieved by measuring
the plasma (or whole blood or serum) concentrations of the parent compound and/
or metabolite(s) through time and determining the plasma (or whole blood or serum)
AUC and Cmax. Subsequently, toxicokinetics relates the exposure achieved in toxicity
studies with the toxicological findings to assess the relevance of these results on
human clinical safety. Toxicokinetics is thus recognized as an integral part of non-
clinical testing programs that enhances the value of the toxicological data generated,
not only to understand the toxicity results but also to predict the risk and safety in
humans. Together with bioavailability evaluation, metabolite profile determination
and plasma drug levels associated with toxicity, toxicokinetics determines the
choice of the no observed adverse effect level, which is essential to define the human
equivalent dose and the maximum recommended starting dose in first-in-human
clinical trials (FDA, 2005).
At non-clinical development, disposition studies are also performed using
radiolabeled compounds administered to animals to provide useful mass balance
data, first-pass metabolism from vascular and portal vein cannulation animal models,
biliary excretion conducted on bile-duct cannulated animals and tissue distribution,
where the organs can be removed and drug quantified in order to evaluate whether drug
accumulation may be involved. In all these studies, identification and quantification
of drug (radiolabeled or otherwise) and main metabolites is required, implying the
application of bioanalytical techniques adequately developed and validated for that
purpose.
Clinical development initiates with the first administration of the drug to humans
and all the stages are strictly based on specific guidelines from the International
Conference on Harmonisation (ICH), European Medicines Agency (EMA) and FDA (ICH,
1997; EMA, 1998 2008; FDA, 2008a; Milton, 2009). The main goal of phase I clinical
trials is to investigate whether the compound is safe and well tolerated in humans
(Table 2.1.1). Pharmacokinetic properties must also be evaluated and the development
of the drug can be stopped if its elimination half-life is found to be too short, too
long or if it has poor bioavailability. Phase I clinical trials usually start with single
sub-therapeutic dose studies which are escalated gradually followed by multiple
dose studies and if the drug is not well tolerated, it is dropped from development.
These closely-monitored trials help define the safe dosing range and whether the
compound should move on to further development phases (Duff, 2007). Thus, from
the pharmacokinetic point of view, absorption and elimination phases of the plasma
80 Pharmacokinetics and Bioanalysis to Improve Drug Development
drug toxicity and reduce therapeutic costs. In practice, it involves the measurement
of drug concentrations in plasma or serum samples, and the interpretation of these
results in an attempt to adjust the dose to that particular patient. Not all drugs are
candidates for TDM, as it must meet the following criteria: (1) a narrow therapeutic
range; (2) significant inter-individual variability in systemic exposure at a given dose;
(3) a clear relationship between blood exposure and clinical effect; and (4) a validated
bioanalytical method to measure drug concentration in plasma or serum.
As described in the previous section, in an attempt to diminish the attrition rate of DDD
process, the pharmaceutical industry has put more emphasis on pharmacokinetic
evaluation in the beginning of NCEs discovery. Consequently, bioanalysis emerged as
an essential support tool to characterize ADME/pharmacokinetics of NCEs through all
stages of DDD and not only in the development ones. Around the world, bioanalysis
shares the main goal of quantifying drugs (and/or metabolites) in biological matrices,
ensuring that high-quality data for valid scientific decision making is obtained.
For that, several initiatives are taken in day-to-day laboratory practices, such as
various kinds of validation, inclusion of independent quality control (QC) samples
or calibration curves mimicking real samples as close as possible. Considering the
distinct objectives and techniques employed in each stage of DDD to characterize
ADME/pharmacokinetics, it is evident that bioanalysis must be adapted to each
particular case, requiring different sample preparation and analytical methodologies
which do not need the same quality validation results.
Table 2.1.4 suggests the general characteristics of bioanalytical methods and
the validation requirements for each DDD stage. In opposition to the regulated
bioanalysis, which is focused on the rigorous development and validation of highly
specific bioanalytical methods for a selected drug candidate in a particular biological
fluid, the most striking feature of discovery bioanalysis is its breadth and diversity
of in vitro and in vivo studies. Consequently, a high number of sample matrices are
generated and a high number of different NCEs are simultaneously measured. As a
result, the throughput, capacity and turnaround requirements are generally much
higher in discovery phases in order to make timely decisions and process the elevated
number of samples. At this stage, early screening tests conducted in vitro and in vivo
can be performed following either good laboratory practices or not, with no real need
to input significant resources in an attempt of developing a rugged and fully validated
method. Instead, only some parameters, such as accuracy, precision and selectivity
must be contained under certain well-accepted boundaries to guarantee that the
results can be used for critical decision making. Thus, at drug discovery stages, the
ideal bioanalytical assay should be simple, straightforward, rapid, largely applicable,
efficient and allow significant throughput in order to be a successful screening tool
Table 2.1.4: Bioanalysis and method validation requirements suggested during drug discovery and development.
82
Drug Discovery • Very fast • Absorption screening (n-octanol/water, PAMPA, • Early study of detection and chromatographic
• High-throughput Caco-2 cells ) conditions
• Straightforward • Metabolic stability screening (microsomes, • Stability assays of stock solution
• Moderate quality data hepatocytes) • Former sample extraction procedure
• In vitro cytochrome P450 inhibitory screening • Selectivity (no sample endogenous interferences,
(microsomes, human recombinant enzymes) no interference of co-administered leads in cassette
• Plasma protein binding (ultrafiltration, dosing and/or cassette analysis)
equilibrium dialysis) • Calibration curve (low number of standards, n = 3-5)
• In vivo pharmacokinetic screening (cassette • Preliminary quantification range
dosing, pooling samples) • Reasonable precision and accuracy
Preclinical Drug • Fast • Pharmacokinetic characterization in rodents, •Sample extraction procedure optimization
Development • Moderate-throughput dogs, monkeys (single and multiple dose studies, •Selectivity (no sample endogenous interferences, no
• High quality data absolute bioavailability) interference of parent drug/metabolites)
• Route-dependent pharmacokinetic disposition •Internal standard selection
(intravenous/oral) •Defined range of standard calibration curve
• Toxicokinetics characterization of parent drug •LLOQ and ULOQ establishment
and its metabolites in toxicology species (dose/ •Intra and interday precision and accuracy assessment
exposure levels, estimating the first dose to (3-5 validation runs)
Pharmacokinetics and Bioanalysis to Improve Drug Development
Table 2.1.4: Bioanalysis and method validation requirements suggested during drug discovery and development.
Continued
Clinical Drug • As fast as possible • Pharmacokinetics in healthy humans, target • Finalization of extraction, chromatographic
Development • Low-throughput patient population and special populations and detection conditions (based on anticipated
• Very high quality data (paediatric, geriatric, hepatic and renal concentrations following the lowest dose in humans)
impairment) • Full validation is required: selectivity, linearity, LLOQ,
• Food effect and relative bioavailability (solution/ LOD, accuracy, precision, recovery, parent drug/
suspension vs solid dosage form) metabolites stability experiments
• Drug-drug interactions with agents commonly co- • Verification of robustness/ruggedness of the assay
prescribed or with narrow safety window
Caco-2, human colon adenocarcinoma cell line; LLOQ, lower limit of quantification; LOD, limit of detection; PAMPA, parallel artificial membrane permeability
assay; ULOQ, upper limit of quantification.
Bioanalysis and Validation Requirements on DDD
83
84 Pharmacokinetics and Bioanalysis to Improve Drug Development
capable of quantifying several NCEs subjected to the in vitro and in vivo tests performed
in initial screening experiments.
As the drug candidate enters to the development stages, matrix complexity
increases as well as the rigor needed to accurately quantify the NCEs in biological
samples and then estimate the pharmacokinetic parameters. The throughput of
the method may decrease, but the quality of data must be higher and validation
should be formalized and mandated as per the required norms. The most widely
accepted guidelines for validation of bioanalytical methods, particularly in
the pharmaceutical and medical sciences, are the Technical Requirements for
Registration of Pharmaceuticals for Human Use guideline Q2(R1) issued by the
ICH (2005), the Guidance for Industry, Bioanalytical Method Validation by the FDA
(2001) and the Guideline on Bioanalytical Method Validation recently issued by the
European Medicines Agency (EMA) in 2011. Typical validation parameters and the
requirements of each guideline are presented in Table 2.1.5. Selectivity is referred to as
the ability of a bioanalytical method to unequivocally differentiate and quantify the
analytes of interest in the presence of other sample components. Endogenous matrix
components, metabolites, and decomposition products or even co-prescribed drugs
include the most common interferences. Thus, it is necessary to establish that the
chromatographic signal produced is only due to the analyte and not as a result of the
co-elution of other compounds. A bioanalytical method is considered to be selective if
the lack of response in the blank biological matrix at the retention time of the analytes
is demonstrated at the lower limit of quantification (LLOQ). The determination of
method sensitivity, linearity, precision and accuracy is required by the three guidelines
and usually performed at three concentration levels in several replicates (typically
three to five). Linearity should be demonstrated within the defined calibration range
with the quality of the calibration curve as an aspect of the greatest importance in
bioanalytical method validation. In fact, as Almeida et al. (2002) stated, the quality of
bioanalytical data is highly dependent on the quality of the calibration curve used to
extrapolate the analyte concentrations in unknown samples. The calibration curve is
the relationship between instrumental response and the concentrations of the analyte
and it must be generated for each analyte, using a sufficient number of calibration
standards (Table 2.1.5). Precision, which expresses the closeness degree among
individual measures of an analyte when the same procedure is applied repeatedly to
multiple aliquots of a homogenous matrix, is frequently performed on the same day
(intra-day precision or repeatability) and over different days (intermediate precision
or inter-day precision), both expressed by the relative standard deviation (RSD),
which is the absolute value of coefficient of variation (CV). Accuracy, which describes
the closeness between the mean experimental data obtained by the method and the
corresponding nominal concentration, is also determined intra- and inter-daily and
is expressed by the relative error (RE), as a percentage, which is the ratio between
experimental and nominal concentrations, or by bias, which is the deviation from the
nominal value as a percentage. Matrix effects, carry-over and dilution integrity have
Table 2.1.5: The ICH, FDA and EMA validation parameters and respective limits instituted.
Accuracy 3 concentration levels 3 concentration levels Accuracy ± 15% 4 concentration levelsb Accuracy ± 15% c
(Bias %) 3 replicates for each one 5 replicates for each one 5 replicates for each one
Precision 3 concentration levels 3 concentration levels Precision ≤ 15% 4 concentration levelsb Precision ≤ 15%
(RSD or CV %) 3 replicates for each one 5 replicates for each one 5 replicates for each one
With-in run and between-run
Recovery (%) NR 3 replicates Precise and consistent NR
Carry over effect NR NR Blank sample analysis following the Interference should be
high concentration standard ≤ 20% of LLOQ and
5% for internal
standard
Bioanalysis and Validation Requirements on DDD
85
Table 2.1.5: The ICH, FDA and EMA validation parameters and respective limits instituted.
Continued
86
Robustness √ NR NR
Stability NR Stability of stock and working solution of Stability of stock and working solution of standards, Stability
standards, Stability of the analyte in matrix of the analyte in matrix includes:
includes: long-term storage, short-term stability at room temperature
long-term storage, short-term temperature or sample processing temperature, post-preparative stability
stability, post-preparative stability, freeze–thaw at room temperature or under the storage conditions to be
cycle. used during the study (dry extract or in the injection phase),
freeze–thaw cycle.
Mean concentration at each level within ± 15% of nominal
concentration.
a
Limits defined for all calibration standards with exception of LLOQ (in LLOQ ± 20%);
Pharmacokinetics and Bioanalysis to Improve Drug Development
b
One must correspond to the LLOQ
c
At LLOQ ± 20%
√, Parameter required but its limits are not referenced; LLOQ, Lower limit of quantification; LOD, Limit of detection; NR, Not required; RSD, relative standard
deviation (corresponds to the absolute value of coefficient of variation); ULOQ, Upper limit of quantification.
Bioanalysis and Validation Requirements on DDD 87
been recently integrated in the guideline issued by EMA as well as stability evaluation.
From a practical point of view, these parameters can be estimated from the analysis of
QC samples, which represent the future real samples.
During pre-clinical development, bioanalytical data may integrate the regulatory
submission process and influence clinical trials design, and, therefore, the method
may differ from the original assay of the discovery phase. Frequently, an internal
standard (IS) is added to samples in order to compensate errors that can occur during
sample preparation and chromatographic analysis, ensuring the reliability of analyte
quantification. Validation experiments should be performed maintaining constant
IS concentration, where the chromatographic response is given by analyte/IS peak
area or height ratio. A stable isotope-labeled derivative of the parent compound is
increasingly being used as IS. However, when not possible, a compound structurally
similar to the analyte and sharing identical behaviors during the entire bioanalytical
protocol is used.
It is noteworthy that various animal species and matrices are frequently
investigated within pre-clinical studies, requiring significant effort to fully validate
all the assays if they are independently developed to each species/matrix. In these
situations, a partial validation or cross validation may be sufficient to manage time
and resources more efficiently at this stage. Using cross validation, the analyst will rely
on the ability of the full validation protocol established in the matrix of one species
to predict the concentrations of NCE in a similar matrix of other species. However, the
scope of partial validation must be justified and certain parameters must be assessed
separately in the sample type that is being subjected to cross validation (FDA, 2001;
EMA, 2011), including stability and selectivity. In essence, although a significant
shortcut of the procedures is undertaken, it cannot compromise either the data quality
or the data integrity.
As the molecule advances into clinical trials, the assays developed to analyze
human samples need to be more sensitive, rugged and robust to some small variations
in matrices. Particularly in phase I clinical trials, high sensitivity is demanded to
ensure that the lowest effective doses can be identified. In fact, if the samples are from
first-time-into-humans studies, resolution is more important than throughput since
the number of samples is low but the fate of the compound is unknown, requiring
complete resolution of the analyte peak from all eventual matrices interferences.
Additionally, stability must be evaluated again because matrices are distinct from those
analyzed in pre-clinical stages. Similarly, it is also required to evaluate the method
selectivity, also considering the metabolites formed in vivo and drugs probably co-
prescribed to the patients in order to avoid their interference with the drug candidate
under investigation. Accordingly, it is desirable to develop flexible and robust assays
that enable minor alterations in chromatographic conditions to circumvent eventual
interfering peaks if necessary.
In current practice, MS detection is employed in the analysis of the majority
of pharmacokinetic samples from clinical studies, and the issues of selectivity and
88 Pharmacokinetics and Bioanalysis to Improve Drug Development
specificity should not pose a major concern for the majority of NCEs at this stage.
However, other elements associated with LC/MS assays have to be assessed, including
the matrix effect, ion suppressing effect and, in the event of selective ion monitoring
(SIM), the generation of a common m/z ion for quantification of multiple peaks
(Srinivas 2008). It is also important to emphasize that the number of samples increases
significantly in the later phases II/III and IV clinical trials and high-throughput,
although desirable, may not be the best choice as compared to improving analytical
efficiency. At these stages, advantages can be found in automating the bioanalytical
process using fast, sensitive and reproducible chromatography with detection and
data-capturing capabilities that lead with high amount of data. In the next section,
some strategies within this scope will be outlined.
In the world of DDD, speed is essential to ensure a competitive position and success
in the highly-competitive life sciences marketplace. Therefore, it is important to
employ innovative thinking in all areas of DDD and identify laboratory technologies/
methodologies that reduce the cycle time from milestone to milestone in conformation
with all the requirements laid out for a regulated environment. To address this
challenge, besides the advances observed in the HTS in vitro and in vivo methodologies
broached in section 1.2, high-throughput bioanalysis has been continuously designed
and refined to improve the speed, quality and efficiency necessary to support the
increasing number of samples from ADME/pharmacokinetic studies. Table 2.1.6
represents the flexibility in today’s number and types of hyphenated bioanalytical
techniques, providing several recent bioanalytical techniques employed to investigate
the pharmacokinetics of NCEs in all stages of DDD programs. Study objectives, sample
preparation and validation parameters evaluated are also summarized therein.
As expected, samples obtained from in vitro and in vivo are predominantly
analyzed by LC (Table 2.1.6). Indeed, chromatography is critical to the success of
any bioanalytical assay as it allows the separation of the analytes from endogenous
substances and metabolites. Resolution of the peaks correspondent to analytes
and metabolites is essential to prevent an overestimation of the concentration of
parent compound and subsequently incorrect determination of pharmacokinetic
parameters and drug exposure. Even using highly selective MS or MS/MS detection, a
prior chromatographic separation (although simpler than those required with other
detection methods) is essential because the co-elution of endogenous or exogenous
compounds may cause ion suppression or ion enhancement, both detrimental to
the development of a sensitive method. Reversed-phase LC is the chromatographic
technique most widely employed in bioanalysis of drugs and their metabolites during
DDD due to its application to small molecules, which are separated by their different
Table 2.1.6: Examples of pharmacokinetic investigations performed during DDD, the bioanalytical conditions employed and respective parameters that were
validated.
In vitro metabolic Estimate the half-life time of Samples filtration before HPLC HPLC-PDA NR Gupta,
stability in the prodrugs in Caco-2 cell analysis. C18 reversed-phase column (5 μm, 4.6 × 2013
intestine homogenates 250 mm)
Verify the conversion to the parent Mobile phase: water and ACN contained
drug (GOCarb). 0.1% TFA
Flow rate: 1 mL min-1
In vitro intestinal Evaluate the epithelial transport Samples filtered in 96-well filter LC-MS NR
permeability across Caco-2 monolayers; plates before analysis. C18 column (5 μm, 2.1 × 50 mm) with guard
Identify the most permeable column.
prodrug candidate. Mobile phase: water and ACN, containing
0.1% formic acid
Flow rate: 0.2 mL min-1
Positive ESI mode.
In vivo mice Oral (10 mg kg-1) and intravenous NR LC-MS/MS NR
pharmacokinetic administration (1 mg kg-1) to Swiss C18 (5 µm, 150 × 2.1 mm) column
studies Webster mice. Mobile phase: 0.1% formic acid/ACN
Determine Cmax, tmax, AUC and F of Flow rate: 0.2 mL min-1.
the prodrug and active metabolite Data acquired by MRM in ESI positive
(GOCarb). mode.
Bioanalysis & Pharmacokinetics, a Synergistic Partnership on DDD
89
Continued
Table 2.1.6: Examples of pharmacokinetic investigations performed during DDD, the bioanalytical conditions employed and respective parameters that
90
were validated.
BMS 690514
pH 5.0
(B) 100% ACN
Flow rate of 1.0 mL min-1
Continued
Table 2.1.6: Examples of pharmacokinetic investigations performed during DDD, the bioanalytical conditions employed and respective parameters that
were validated.
In vitro Evaluate the binding plasma Extraction with Cyclone HTLC LC-ESI/MS/MS Validation Zhang,
ultrafiltration proteins. column (50 × 0.5 mm, 60 µm) C18 column (30 × 2.1 mm, 3 µm) according to 2006
Perform a HTS ultrafiltration and the mobile phase composed Mobile phase: 80% ACN and 20% 2mM international
method employing the 96-well of ACN and 2 mM ammonium ammonium formate (pH 3.0). guidelines.
Bioanalysis & Pharmacokinetics, a Synergistic Partnership on DDD
were validated.
In vivo rat Determine the absolute Blood samples were placed HPLC-MS/MS Linearity;
pharmacokinetic bioavailability in rats after oral into heparinized tubes and C18 column (50 × 2.1 mm, 1.7 μm). Recovery,
dose response administration (5, 15, 30 mg kg-1) centrifuged to obtain plasma. Quantification by MRM Accuracy;
and compare it to that of the oral IS: lamivudine Analysis of Cytarabine and prodrug: Precision.
aqueous solution of cytarabine (8 For Cytarabine and prodrug Mobile phase: Water containing 2%
mg kg-1). extraction methanol and methanol containing 0.1%
Determination of the Plasma pretreatment by SPE formic acid.
concentrations of the prodrug, cation-exchange SPE cartridges. ESI source set in positive ionization.
Pharmacokinetics and Bioanalysis to Improve Drug Development
Continued
Table 2.1.6: Examples of pharmacokinetic investigations performed during DDD, the bioanalytical conditions employed and respective parameters that
were validated.
In vitro CYP Evaluate the effect of CYP PP with ice-cold ACN and direct UHPLC-QTOF-MS Linearity; Spaggiari,
metabolism inhibitors on each isoenzyme analysis of the supernatant. C18 column (100 × 2.1 mm; 2.5 μm) Inter and intra- 2014
studies in human activity using a specific substrate Mobile phase: day precision;
liver microsomes probe cocktail in human liver (A) water with 0.1% formic acid; Matrix effect.
microsomes. (B) ACN with 0.1% formic acid
Flow rate: 400 µL min-1
Q-TOF-MS system was operated in wide-
pass quadrupole mode
In vitro direct Evaluate the inhibition potential of PP with ice-cold ACN; LC-/MS/MS Linearity; Lee, 2013
and metabolism- test compounds on 8 human CYP Supernatant was put in 96-well C18 column (150 × 4.6mm, 5 µm) protected Accuracy and
dependent CYP enzymes. plates for analysis. by a C18 guard column (2.0 × 4.0 mm) precision.
inhibition studies IS: Carbamazepine Mobile phase:
(A ) 0.1% v/v formic acid in water;
(B) 0.1% v/v formic acid in acN.
Flow rate: 0.7 mL min-1
API MS/MS operated in the positive and
negative ion mode
Quantitation was performed by MRM.
Bioanalysis & Pharmacokinetics, a Synergistic Partnership on DDD
93
Continued
Table 2.1.6: Examples of pharmacokinetic investigations performed during DDD, the bioanalytical conditions employed and respective parameters that
94
were validated.
In vitro cocktail CYP Evaluate the influence of NCEs PP with ACN. HPLC-MS/MS NR Otten,
inhibition studies in several CYP isoforms, using Plates centrifugation and direct RP-Amide column (2.1 × 50 mm, 5.0 µm) 2012
specific substrates and human injection of the supernatant Mobile phase:
liver microsomes. previously diluted. (A) Aqueous 0.1% acetic acid and 25:75
Quantify the main metabolite IS: labetalol. (v/v) methanol:ACN
of each substrate probe in the (B) 0.1% acetic acid.
presence and absence of the NCE. Flow rate: 0.6 mL min-1
API Triple Quadropole LC-MS used in the
positive ESI mode.
SRM to identify each probe substrate,
their metabolites and IS.
Clinical Simultaneous quantification of Blood centrifugation to obtain HPLC-MS/MS Full validation Oh, 2012
pharmacokinetic five CYP substrate probes and their plasma. C18 column (100 × 2.1 mm, 3.5 µm) according to
study main metabolites in plasma. Dual liquid-liquid extraction Mobile phase: FDA guideline.
Administration of the CYP with ethyl acetate. (A) 0.1% formic acid in water
substrate cocktail to healthy Organic layers evaporation (B) 0.1% formic acid in acetonitrile
volunteers and quantify the parent followed of reconstitution with Flow rate: 0.2 mL min-1
Pharmacokinetics and Bioanalysis to Improve Drug Development
drugs and main metabolites. methanol. Linear ion trap mass spectrometer triple
IS: Propranolol quadrupole mass spectrometer equipped
with a turbo ion spray source.
ESI positive ion mode
Quantification by MRM.
Continued
Table 2.1.6: Examples of pharmacokinetic investigations performed during DDD, the bioanalytical conditions employed and respective parameters that
were validated.
In vitro Caco-2 Determination of Papp through Sample direct injection HPLC-UV NR Yan, 2011
Caco-2 monolayer; No IS C18 column (5 μm, 200 × 4.6 μm)
Identify the most permeable Mobile phase: mixture (25:75) of
prodrug candidate and very if: methanol 0.05 M and KH2PO4 buffer
- it is a substrate of the solution
oligopeptide transporter (PepT1); Flow rate:1.0 mL min-1,
- influence the uptake of Wavelength: 250 nm.
glycylsarcosine.
In vivo rat Determine the absolute Blood samples were placed UHPLC-MS/MS NR
pharmacokinetic bioavailability of the previously into heparinized tubes and Hydrophilic interaction column (50 ×
dose response selected prodrug candidate in rats centrifuged to obtain plasma. 2.1 mm, 1.7 μm)
after oral administration (5, 15, Plasma pretreatment by SPE Mobile Phase: water containing 0.1%
30 mg kg-1). using HLB cartridges (1 cc, formic acid and methanol (85:15, v/v)
Similar studies performed co- 30 mg) ESI source set in positive ionization.
administering the prodrug with IS: lamivudine Quantification by MRM.
glycylsarcosine.
Bioanalysis & Pharmacokinetics, a Synergistic Partnership on DDD
95
Continued
Table 2.1.6: Examples of pharmacokinetic investigations performed during DDD, the bioanalytical conditions employed and respective parameters that
96
were validated.
In vitro mice Determine the Papp through Sample centrifugation prior LiChroCART Purospher Star (C18, 3 µm, 55 Partial Fortuna,
intestine CD-1 mice intestine membrane analysis. × 4 mm; Merck KGaA) column. Validation: 2012
permeability mounted in Ussing chamber; Mobile Phase: water–methanol– Linearity;
Identify substrates of acetonitrile (64:30:6, v/v/v) Intra- and
P-glycoprotein. UV: 235 nm inter-day
precision and
accuracy.
In vivo In vivo pharmacokinetic Plasma: Chiral column LiChroCART 250-4 Full validated Fortuna,
pharmacokinetic characterization of the Blood samples were collected ChiraDex. according to 2013
studies test compounds and their to heparinized tubes and UV: 235 nm International
corresponding metabolites after centrifuged to obtain plasma. Mobile phase: water and methanol guidelines.
oral administration to mice. For CBZ, OXC, ESL, S-Lic, R-Lic and CBZ-E
Calculate Cmax, tmax, AUC, t1/2 and Brain: and main metabolites:
MRT in plasma and brain. Brain tissue was homogenized in Flow rate: 0.9 mL min-1
a 0.1 M sodium phosphate buffer IS: Chloramphenicol
pH 5 (4 mL g-1) and centrifuged. For BIA 2-059, BIA2-024, BIA 2-265 and
Plasma (300 µL) and brain their main metabolites:
Pharmacokinetics and Bioanalysis to Improve Drug Development
Continued
Table 2.1.6: Examples of pharmacokinetic investigations performed during DDD, the bioanalytical conditions employed and respective parameters that
were validated.
Clinical Studies Investigate the effect of gender, Blood samples were collected LC-MS Linearity; Falcão,
food and hepatic failure in the into tubes containing lithium LichroCART 250-4 ChiraDex analytical Precision and 2007;
biodisposition of eslicarbazepine heparin and centrifuged to column (β-cyclodextrin, 5 µm), a accuracy (with Almeida,
after oral administration of obtain plasma. LichroCART 4-4 ChiraDex guard column 3 QCs) 2008
eslicarbazepine acetate. (β-cyclodextrin, 5 µm).
Plasma pretreatment with Mobile Phase:
Schleicher and C18/100 mg 96- (A) 0.2 mM sodium acetate
well SPE plate. (B) 0.2 mM sodium acetate, MeOH.
Final extract was reconstituted in ESI source set in positive ion mode.
water:methanol.
IS: 10,11-dihydrocarbamazepine
Bioanalysis & Pharmacokinetics, a Synergistic Partnership on DDD
97
Continued
Table 2.1.6: Examples of pharmacokinetic investigations performed during DDD, the bioanalytical conditions employed and respective parameters that
98
were validated.
Continued
Table 2.1.6: Examples of pharmacokinetic investigations performed during DDD, the bioanalytical conditions employed and respective parameters that
were validated.
Randomized, Investigate the pharmacokinetics Blood samples were collected LC/MS/MS Full validation Cundy,
double-blind, and tolerability after immediate- into tubes containing K2EDTA C18 HPLC column (50 × 4.6 mm) according to 2004
placebo-controlled release capsules of gabapentin and centrifuged to obtain Mobile phase: international
design. enacarbil oral administration at plasma. (A) water/ACN/formic acid (95:5:1); guidelines.
5 ascending single doses and Plasma samples were quenched (B) ACN/water/formic acid (95:5:1).
compared with commercialized with methanol immediately Flow rate: 2.0 mL min-1
gabapentine capsules. before analysis. Detector was an API MS/MS and detection
Urine samples had no performed with MRM.
preparation before analysis
were validated.
Continued
Table 2.1.6: Examples of pharmacokinetic investigations performed during DDD, the bioanalytical conditions employed and respective parameters that
were validated.
Bioequivalence Quantify glimepiride in Addition of o-phosphoric acid LC-MS/MS Full validation Kundlik,
study plasma samples from ongoing (2.5%) to plasma samples which C18 column (100 × 3 mm, 3.0 µm) according to 2012
development of an immediate- are loaded onto HLB 96 well Mobile phase was ACN and 2 mM international
release formulation. cartridges. ammonium format, pH 3.5 with formic guidelines
acid
Flow rate: 0.5 mL min-1
Triple quadrupole mass spectrometer
equipped with an ESI ion source in
positive ionization SRM mode.
HM781-36B, a new anticancer drug
In vitro metabolism Profile and identify the Centrifugation and supernatant LC-MS/MS NR Kim, 2013
in liver metabolites in microsomes and evaporation to dryness. C18 column (150 × 2.1 mm, 3.5 µm)
recombinant CYPs. Analysis of the residue after Mobile phase: ammonium acetate
reconstitution. (10 mM, pH 3.6)/ACN
Flow rate: 0.2 mL min-1
Use of MS, MS2 and MS3.
Bioanalysis & Pharmacokinetics, a Synergistic Partnership on DDD
101
Continued
Table 2.1.6: Examples of pharmacokinetic investigations performed during DDD, the bioanalytical conditions employed and respective parameters that
102
were validated.
Continued
Table 2.1.6: Examples of pharmacokinetic investigations performed during DDD, the bioanalytical conditions employed and respective parameters that
were validated.
In vivo Determination of Cmax, tmax, AUC, Blood was collected into EDTA LC–MS/MS-MRM Linearity
pharmacokinetic t1/2,F and Cl after oral (10 mg tubes and centrifuged to obtain NR
evaluation kg-1) or intravenous (3 mg kg-1) plasma.
administrations PP with ACN containing the IS.
In vivo Calculate Cmax, tmax, AUC, MRT and Blood samples collection UHPLC-MS/MS Full validation Wang,
pharmacokinetic Cl after Psoralea corylifolia extract into heparinized tubes and C18 (2.1 × 50 mm,1.7 µm) according to 2014
study or benzofuran glycoside fraction centrifuged to obtain plasma. Mobile phase: FDA guidelines
oral administration to Wistar rats. PP with methanol; water addition (A) methanol
and injection of the supernatant. (B) 0.1% formic acid aqueous
IS: sophoricoside Flow rate: 0.2 mL min-1
Triple quadrupole mass spectrometry
Data acquisition by MRM in ESI positive
mode.
Bioanalysis & Pharmacokinetics, a Synergistic Partnership on DDD
103
Continued
Table 2.1.6: Examples of pharmacokinetic investigations performed during DDD, the bioanalytical conditions employed and respective parameters that
104
were validated.
In vitro hepatic Stability evaluation in PP with methanol and direct HPLC-MS/MS NR Chhonker,
and intestinal rat intestinal and hepatic analysis of the supernatant. C18 column (5 μm, 4.6 × 150 mm) column, 2014
metabolism microsomes; preceded with a guard column.
Prediction of the first pass effect Mobile Phase: 10 mM ammonium acetate
on drug bioavailability; (pH 4): methanol
In vivo Calculate Cmax, tmax, AUC, Vd, Blood samples collected Flow rate: 0.6 mL min-1 Full validation,
pharmacokinetic t1/2, Cl and F after oral (50 mg into micro-centrifuge tubes ESI source set in positive ionization. according
evaluation kg-1) and intravenous (5 mg kg-1) containing heparin and Quantification by MRM to US FDA
administrations. centrifuged to obtain plasma. guideline.
Plasma pretreatment by SPE (1
mL, C18) cartridges.
Evaporation of eluents and
analysis of the dry residue
reconstituted in methanol.
IS: Phenacetin
15 derivatives of Sulfonamide with antitumor activity
In vitro metabolic All test compounds were Centrifugation and direct LC-MS/MS NR Belka,
Pharmacokinetics and Bioanalysis to Improve Drug Development
stability and incubated for 2 h with human liver injection of the resulting C18 column (3.0 × 100 mm, 2.7 µm) 2014
elucidation of microssomes in order to: supernatant. Mobile Phase:
metabolites - quantify the remaining parent (A) 0.1% formic acid in water
structures compound; (B) 0.1% formic acid in ACN
- estimate their metabolic Flow rate: 0.25 mL min-1
stability; MS detection with ESI positive mode.
- separate and identify the MS/MS detection with Q-TOF analyzer
metabolites generated. to collect the fragmentation spectra of
biotransformation products.
Continued
Table 2.1.6: Examples of pharmacokinetic investigations performed during DDD, the bioanalytical conditions employed and respective parameters that
were validated.
In vitro cell Investigate the Papp through MDCK Direct injection of apical and UV/Vis Spectrometry at 330 nm NR Ninomiya,
permeability cell monomlayers. basolateral solutions. 2011
evaluation Quantification of tricin
concentrations initially in apical
solution and at the end in the
basolateral solution.
In vivo systemic Determination of Cmax, tmax, AUC, Blood centrifugation to obtain HPLC-UV Linearity; Ninomiya,
pharmacokinetic t1/2,Vd and Cl of tricin after a single plasma. C18 column (250 × 4.6 mm, 5 μm) Inter-and intra- 2011;
studies oral dose of prodrugs (100, 300 or Liquid-liquid extraction with Mobile phase: ammonium acetate buffer day precision Cai 2003
1000 mg kg-1) to Crl:CD rats. acetone; water addition to the (0.1 M, pH 5.1) and methanol plus EDTA. (with 3QCs);
organic layer, centrifugation and Flow-rate: 1 mL min-1. Stability.
inject ion of the supernatant. Tricin concentration in plasma was
determined by an IS method.
ACN, acetonitrile; APCI, atmospheric pressure chemical ionization; API, atmospheric pressure ionization; Ara-U, 1-ß-D-arabinofuranosyluracil; AUC, area under
the curve; BBB , blood-brain barrier; BCRP, breast cancer resistance protein; CE, collision energy; Cl, systemic clearance; DAD, diode array detection; DP,
declustering potential; F, absolute bioavailability; FDA, food and drug administration; GOCarb, guanidine oseltamivir carboxylate; HPLC, high performance
liquid chromatography; IS, internal standard; MRM, multiple reaction monitoring; K2EDTA, dipotassium salt of ethylenediaminetetraacetic acid; MRT, mean
residence time; NR, not reported; PDA, photodiode array; PP, protein precipitation; QC, quality control samples; SPE, solid phase extraction; TFA, Trifluoroacetic
acid; t1/2, half-life time; UHPLC, ultra-high-performance liquid chromatography; UV, ultraviolet; Vd, volume of distribution;
Bioanalysis & Pharmacokinetics, a Synergistic Partnership on DDD
105
106 Pharmacokinetics and Bioanalysis to Improve Drug Development
affinity to the hydrophobic stationary phase in relation to the mobile phase. However,
compounds with low octanol-water partition coefficients and the majority of common
metabolites may hamper the development of a reliable reversed-phase LC method,
as they present a very short or non-existent retention time. Thus, the demand for
analytical laboratories to increase sample throughput provided the impetus for HPLC
column manufacturers to introduce new stationary phases and column geometries
in order to increase speed and sensitivity. In this context, monolithic columns were
created and have been employed with fast gradients and high flow rates for the direct
analysis of several pharmaceutical compounds in biological samples (Zhou, 2005;
Huang, 2006), but these columns have not been very frequently employed during
DDD. Indeed, ultra-high pressure liquid chromatography (UHPLC) seems to be the
most important achievement in LC evolution over the last decade and its application
is increasing in the pharmaceutical field (Table 2.1.6). This technology retains the
practicality and principles of traditional HPLC system but it is capable of providing
liquid flow at pressures higher than 10,000 psi and columns packed with small
particles (< 2 µm) that can withstand these pressures (Guillarme, 2013; Rodriguez-
Aller, 2013; Fekete, 2014; Nishi, 2014). It is well-known that the theoretical plate height
is proportional to the particle size, so using particle sizes down to 1.7 µm decreases
theoretical plate height and consequently produces a significant gain in efficiency that
does not diminish at increased flow rates or linear velocities (Gilpin, 2008). UHPLC is
becoming increasingly popular (Table 2.1.6) because it enables high speed analysis,
superior resolution, increased sensitivity and reduced overall cost per analysis,
ensuring that higher quality information is gained and laboratory productivity is
optimized using sub-2 µm particle technology. However, certain practical concerns
may need improvement, including sample introduction, reproducibility and the
possible formation of temperature gradients within UHPLC columns at such high
pressures. Moreover, extremely narrow sample plugs are required to minimize sample
volume effect on peak broadening.
Currently, LC/MS and LC/MS/MS take the bulk of the work during pharmacokinetic
evaluation in DDD programs (Table 2.1.6), particularly due to the inherent specificity
and sensitivity of these techniques, which dramatically increase throughput for
the quantitative determination of drugs and metabolites in biological matrices. In
addition, these methods reduce the need of exhaustive chromatographic separations
prior to detection, reducing analysis time. Even though both LC/MS and LC/MS/MS
are very expensive and not available to all research laboratories, bioanalysts have
been devoted to improve MS and MS/MS hardware and software, not only to identify
and quantify metabolites in a variety of in vitro and in vivo assays during all stages
of DDD, but also to allow users to benefit from lower LLOQs and ease of use. In fact,
reduced LLOQ values are essential in order to guarantee strong characterization during
the elimination phase of a compound. Besides their high sensitivity and specificity,
MS and MS/MS detections are based on a combination of the unique parent and
fragment masses of each compound, eliminating the need for baseline separation
Bioanalysis & Pharmacokinetics, a Synergistic Partnership on DDD 107
and achieving fast analyses. Prior to MS detection, analyte ionization is required and
can be achieved by electrospray ionization (ESI), atmospheric pressure ionization
(API) or atmospheric pressure chemical ionization (APCI) modes. MS detection of
ionized analytes has been carried out in a linear scan mode in which a range of
m/z values is constantly monitored. In a more selective and sensitive mode called
SIM, one specific m/z value (or more) is selected for the monitoring. Single-reaction
monitoring (SRM) or multiple-reaction monitoring (MRM) modes can also be used:
in SRM mode, a specific product ion of a specific parent ion is detected, producing
very simple plots (usually containing only a single peak). MRM delivers a unique
fragment ion that can be monitored and quantified in a very complicated matrix,
making the MRM plot ideal for sensitive and specific quantifications. Commonly,
several kinds of tandem mass spectrometers including the triple stage quadrupole,
the three-dimensional ion-trap, linear ion-traps/quadrupole coupled with time-of-
flight (Q-TOF) and hybrid triple quadrupole linear ion trap mass spectrometers have
been used in DDD.
Nevertheless, classical detection modes such as UV and fluorescence, although
less employed in current practice than MS or MS/MS detections, continue to be
applied in the early in vitro bioavailability or metabolic stability screening and
remain valuable in cases where sensitivity is not paramount or where LC/MS is not
economically viable.
a specific incubation time, samples are taken from both sides of the monolayers
for bioanalytical determination of the drug concentration. Most of these in vitro
techniques are calibrated within the laboratory using a set of commercial drugs with
known human absorption fraction, which are plotted against the experimentally
obtained apparent permeability (Fortuna, 2012). In our laboratory, we have developed
an Ussing chamber technique employing mouse jejunum segments in order to predict
the human intestinal absorption fraction and the involvement of P-glycoprotein in
drug absorption. The technique was initially validated with reference compounds
and then applied to compare the absorption of nine derivatives of carbamazepine, a
classical antiepileptic drug (Fortuna, 2012). An HPLC-UV technique was successfully
employed in order to quantify each compound in our samples. However, a major
drawback is one frequently ascribed to the use of HPLC-UV in these analyses, which
is that for poorly permeable drugs, the LLOQ achieved is often not compatible with
the amount of drug present in the samples. At this point, some companies are opting
to develop LC-MS techniques or automated LC-MS/MS approaches to circumvent this
limitation (Table 2.1.6).
Nevertheless, the major application of LC-MS and LC-MS/MS during in vitro
ADME analysis regards drug metabolism. In fact, due to the publication of the FDA
guidelines on metabolites in safety testing (MIST) and DDI, metabolite profiling and
screening for metabolites in plasma samples have been recently included as part of
new drug discovery stages, implying the development of bioanalytical techniques
that quickly identify and quantify the parent drug and its metabolites. Information
generated from these studies could therefore enable the scientist to predict earlier
the major metabolite exposure in animals and humans, allowing a better strategy to
be developed. For instance, for compounds with extensive metabolism, identification
of metabolites may help medicinal chemists to modify the NCE to block the sites of
metabolism. On the other hand, active metabolites may have better pharmacokinetic
profiles than the dosed NCE and may become a new lead compound.
Moreover, the DDI guidance of EMA recommends the identification of the enzymes
involved in the formation and elimination pathway of metabolites that contribute
to more than 50% of the in vivo pharmacological effect (EMA, 2012). Since CYP
enzymes are the most frequently involved in drugs metabolism, the identification of
the specific isoforms involved in the metabolism of new compounds at early in vitro
stages become crucial, specifically whether the drug candidate is metabolized by
single or multiple enzyme isoforms and whether highly polymorphic enzymes, such
as CYP2D6 and CYP2C19, contribute to their metabolic clearance. In addition, these
in vitro techniques (mainly employing microsomes and hepatocyte cell lines) can
also be useful in identifying drug candidates that are inhibitors or inducers of CYP
isoforms, and can therefore predict their DDI potential. For these investigations, CYP
probe substrates must be used and their respective metabolites must be quantified
in the presence and absence of the drug candidate. Thus, it is not surprising that
significant efforts have been devoted to develop rapid chromatographic techniques
Bioanalysis & Pharmacokinetics, a Synergistic Partnership on DDD 109
DDD (Table 2.1.2). Devices based on multi-well formats have been recently developed
(Fung 2003; Zhang 2006; Plumb 2008; van Liempd, 2011; Zamek-Gliszczynski, 2011)
and employed to investigate NCEs (Table 2.1.6). Briefly, they consist of 48 or 96-well
Teflon base plates with a semi-permeable membrane that allows the passage only
of the unbound drug. When equilibrium between both membrane sides is reached
in the equilibrium dialysis technique, the total drug concentration is determined in
the plasma compartment while the free drug concentration is determined in buffer
compartment in order to calculate the percentage of drug bound to plasma proteins.
On the other hand, ultrafiltration is currently emerging as a faster and simpler
alternative to equilibrium dialysis once the centrifugation applied (approximately
2000 × g) accelerates the passage of the unbound drug through the membrane (Fung
2003; Zhang 2006). Particularly because of the low free drug levels that can be achieved
in the buffer compartment or in the filtrate, both in vitro techniques require the
application of LC-MS/MS methods to accurately quantify the drugs (Table 2.1.6). Both
MRM (Zamek-Gliszczynski, 2011) and SRM (van Liempd, 2011) detection modes seem to
be successful. Sample preparation prior to chromatographic analysis is very simple, and
relies on protein precipitation with acetonitrile (van Liempd, 2011, Zamek-Gliszczynski,
2011) or a mixture of acetonitrile/0.1% formic acid aqueous solution (90/10, v/v) (Plumb,
2008) followed by centrifugation. It is important to highlight that the bioanalytic
methodologies employed to quantify the drugs were at least partially validated in order
to confirm their acceptability for each specific compound (Table 2.1.6).
2.1.5 Conclusions
References
Alves, G., Fortuna, A., Falcão, A. (2008) High Performance Liquid Chromatography and its Impact in
the Development of Chiral Drugs: a Review. Trends in Chromatography, 4, 1-10.
Almeida, L., Potgieter, J.H., Maia, J., Potgieter, M.A., Mota, F., Soares-da-Silva, P., (2008)
Pharmacokinetics of eslicarbazepine acetate in patients with moderate hepatic impairment.
European Journal of Clinical Pharmacology, 64(3), 267-273.
Amore, B.M., Gibbs, J.P., Emery, M.G., (2010) Application of In Vivo Animal Models to Characterize
the Pharmacokinetic and Pharmacodynamic Properties of Drug Candidates in Discovery
Settings. Combinatorial Chemistry & High Throughput Screening, 13(2), 207-218.
Balani, S.K., Miwa, G.T., Gan, L.S, Wu, J.T., Lee, F.W., (2005) Strategy of Utilizing In Vitro and In Vivo
ADME Tools for Lead Optimization and Drug Candidate Selection. Current Topics in Medicinal
Chemistry, 5(11), 1033-1038.
Belka, M., Hewelt-Belka, W., Sławiński, J., Bączek, T., (2014) Mass Spectrometry Based Identification
of Geometric Isomers during Metabolic Stability Study of a New Cytotoxic Sulfonamide
Derivatives Supported by Quantitative Structure-Retention Relationships. PLoS One, 9(6),
e98096.
Bhogal, N., Balls, M. (2008) Translation of New Technologies: From Basic Research to Drug Discovery
and Development. Current Drug Discovery Technologies, 5(3), 250-262.
Boer, de T., Meulman, E., Meijering, H., Wieling, J., Dogterom, P., Lass, H., (2012) Development
and validation of automated SPE-HPLC-MS/MS methods for the quantification of asenapine,
a new antipsychotic agent, and its two major metabolites in human urine. Biomedical
Chromatography, 26(12), 1461-1463.
Burton, P.S., Goodwin, J.T. (2010) Solubility and Permeability Measurement and Applications in Drug
Discovery. Combinatorial Chemistry & High Throughput Screening, 13(2), 101-111.
Carlson, T.J., Fisher, M.B., (2008) Recent Advances in High Throughput Screening for ADME
Properties. Combinatorial Chemistry & High Throughput Screening, 11, 258-264.
Cai, H., Verschoyle, R.D., Steward, W.P., Gescher, A.J., (2003) Determination of the flavone tricin in
human plasma by high-performance liquid chromatography. Biomedical Chromatography, 17(7),
435-439.
Chen, D., Lal, R., Zomorodi, K., Atluri, H., Ho, J., Luo, W., Tovera, J., Bonzo, D., Cundy, K., (2012)
Evaluation of Gabapentin Enacarbil on Cardiac Repolarization: A Randomized, Double-Blind,
Placebo- and Active-Controlled, Crossover Thorough QT/QTc Study in Healthy Adults. Clinical
Therapeutics, 34(2), 351-362.
Cheung, B.W., Cartier, L.L., Russlie, H.Q., Sawchuk, R.J., (2005) The application of sample pooling
methods for determining AUC, AUMC and mean residence times in pharmacokinetic studies.
Fundamental & Clinical Pharmacology, 19(3), 347-354.
Chien, J.Y., Friedrich, S., Heathman, M.A., de Alwis, D.P., Sinha, V., (2005) Pharmacokinetics/
Pharmacodynamics and the Stages of Drug Development: Role of Modeling and Simulation.
AAPS Journal, 7(3), E544-559.
Chhonker, Y.S., Chandasana, H., Kumar, D., Mishra, S.K., Srivastava, S., Balaramnavar, V.M.,
Gaikwad, A.N., Kanojiya, S., Saxena, A.K., Bhatta, R.S., (2014) Pharmacokinetic and metabolism
studies of rohitukine in rats by high performance liquid-chromatography with tandem mass
spectrometry. Fitoterapia, 97, 34-42.
Cundy, K.C., Branch, R., Chernov-Rogan, T., Dias, T., Estrada, T., Hold, K., Koller, K., Liu, X., Mann,
A., Panuwat, M., Raillard, S.P., Upadhyay, S., Wu, Q.Q., Xiang, J.N., Yan, H., Zerangue, N.,
Zhou, C.X., Barrett, R.W., Gallop, M.A., (2004) XP13512 [(+/-)-1-([(alpha-Isobutanoyloxyethoxy)
carbonyl] aminomethyl)-1-cyclohexane Acetic Acid], A Novel Gabapentin Prodrug: I.
Design, Synthesis, Enzymatic Conversion to Gabapentin, and Transport by Intestinal Solute
Transporters. Journal of Pharmacology and experimental therapeutics, 311(1), 315-323.
114 Pharmacokinetics and Bioanalysis to Improve Drug Development
Di, L., Kerns, E.H., Ma, X.J., Huang, Y., Carter, G.T., (2008) Applications of High Throughput
Microsomal Stability Assay in Drug Discovery. Combinatorial Chemistry & High Throughput
Screening, 11(6), 469-476.
Duff, G. (2007) Guidelines for phase 1 clinical trials. In: The Association of the British Pharmaceutical
Industry, London, Association of the British Pharmaceutical Industry (pp. 1-49).
Eerkes, A., Shou, W.Z., Naidong, W., (2003) Liquid/liquid extraction using 96-well plate format in
conjunction with hydrophilic interaction liquid chromatography-tandem mass spectrometry
method for the analysis of fluconazole in human plasma. Journal of Pharmaceutical and
Biomedical Analysis, 31(5), 917-928.
EMA, Note for guidance on toxicokinetics: a guidance for assessing systemic exposure in toxicology
studies (CPMP/ICH/384/95).
EMA (1998) Note for Guidance of General considerations for clinical trials (CPMP/ICH/291/95).
EMA (2004a) Guideline on pharmacokinetics of medicinal products in patients with impaired renal
function. European Agency for the Evaluation of Medical Products.
EMA (2004b) Guideline on the role of pharmacokinetics in the development of medicinal products in
the paediatric population. European Agency for the Evaluation of Medical Products.
EMA (2005). Guideline on the evaluation of the pharmacokinetics of medicinal products in patients
with impaired hepatic function. European Agency for the Evaluation of Medical Products.
EMA (2008) Note for Guidance on Non-clinical safety studies for the conduct of human clinical trials
and marketing authorization for pharmaceuticals (CPMP/ICH/286/95).
EMA (2011) Guideline on Bioanalytical Method Validation. European Agency for the Evaluation of
Medical Products.
EMA (2012) Guideline on Investigation of Drug Interactions. European Agency for the Evaluation of
Medical Products.
Falcão, A., et al. (2007) Effect of Gender on the Pharmacokinetics of Eslicarbazepine Acetate (BIA
2-093), a New Voltage-gated Sodium Channel Blocker. Biopharmaceutics and Drug Disposition,
28(5), 249-256.
FDA (2000) Guidance for Industry, Waiver of In Vivo Bioavailability and Bioequivalence Studies
for Immediate-Release Solid Oral Dosage Forms Based on a Biopharmaceutics Classification
System. U.S. Department of Health and Human Services.
FDA (2001) Guidance for Industry, Bioanalytical Method Validation. U.S. Department of Health and
Human Services.
FDA (2003) Guidance for Industry, Bioavailability and Bioequivalence Studies for Orally Administered
Drug Products — General Considerations. U.S. Department of Health and Human Services.
FDA (2006) Guidance for Industry, Drug Interaction Studies - Study Design, Data Analysis, and
Implications for Dosing and Labeling. U.S. Department of Health and Human Services.
FDA (2008a) Guidance for Industry, CGMP for Phase I investigational drugs. U.S. Department of
Health and Human Services.
FDA (2008b) Guidance for Industry: Safety Testing of Drug Metabolites. U.S. Department of Health
and Human Services.
FDA (2012) Guidance for Industry: Drug Interaction Studies – Study Design, Data Analysis,
Implications for Dosing, and Labeling Recommendations. U.S. Department of Health and
Human Services.
Fekete, S., Kohler, I., Rudaz, S., Guillarme, D., (2014) Importance of instrumentation for fast liquid
chromatography in pharmaceutical analysis. Journal of Pharmaceutical and Biomedical
Analysis, 87, 105-119.
Frederick, C.B., Obach, R.S. (2010) Metabolites in Safety Testing: “MIST” for the Clinical
Pharmacologist. Clinical Pharmacology & Therapeutics, 87(3), 345-350.
Fortuna, A., Alves, G., Falcão, A., Soares-da-Silva, P., (2012) Evaluation of the permeability and
P-glycoprotein efflux of carbamazepine and several derivatives across mouse small intestine by
the Using chamber technique. Epilepsia, 53(3), 529-538.
References 115
Fortuna, A., Alves, G., Soares-da-Silva, P., Falcão, A., (2012) Pharmacokinetics, brain distribution and
plasma protein binding of carbamazepine and nine derivatives: New set of data for predictive in
silico ADME models. Epilepsy Research, 107(1-2): 37-50.
Fretland, A., Monshouwer, M. (2010) Enzyme Induction: Translating Multiple Approaches, Assays,
Endpoints, and Opinions into a Valuable Induction Screening Strategy. Combinatorial
Chemistry & High Throughput Screening, 13(2), 135-144.
Fung, E.N., Chen, Y.H., Lau, Y.Y., (2003) Semi-automatic high-throughput determination of plasma
protein binding using a 96-well plate filtrate assembly and fast liquid chromatography–tandem
mass spectrometry. Journal of Chromatography B, 795(2), 187-194.
Gallo, J.M. (2010) Pharmacokinetic/Pharmacodynamic-Driven Drug Development. Mount Sinai
Journal of Medicine, 77(4), 381-388.
Gong, J., Gan, J., Caceres-Cortes, J., Christopher, L.J., Arora, V., Masson, E., Williams, D., Pursley, J.,
Allentoff, A., Lago, M., Tran, S.B., Iyer, R.A., (2011) Metabolism and Disposition of [14C]Brivanib
Alaninate after Oral Administration to Rats, Monkeys, and Humans. Drug Metabolism and
Disposition, 39(5), 891-903.
Guillarme, D., Veuthey, J.L. (2013) State-of-the art of (UHP)LC–MS(–MS) techniques and their
practical application. Journal of Chromatography A, 1291, 1.
Guilpin, R.K., Zhou, W. (2008) Ultrahigh-Pressure Liquid Chromatography: Fundamental Aspects of
Compression and Decompression Heating. Journal of Chromatographic Science, 46(3), 248-253.
Gupta, D., Varghese Gupta, S., Dahan, A., Tsume, Y., Hilfinger, J., Lee, K.D., Amidon, G.L., (2013)
Increasing Oral Absorption of Polar Neuraminidase Inhibitors: A Prodrug Transporter Approach
Applied to Oseltamivir Analogue. Molecular Pharmaceutics, 10(2), 512-522.
ICH (1997), General considerations for clinical trials E8.
ICH (2005) International Conference on Harmonisation, Validation of Analytical Procedures: Text and
Methodology Q2 (R1). Conference on Harmonization.
Keserü, G.M., Makara, G.M. (2009) The influence of lead discovery strategies on the properties of
drug candidates. Natural Reviews and Drug Discovery, 8(3), 203-212.
Kim, E., Kim, H., Suh, K., Kwon, S., Lee, G., Park, N.H., Hong, J., (2013) Metabolite identification
of a new tyrosine kinase inhibitor, HM781-36B, and a pharmacokinetic study by liquid
chromatography/tandem mass spectrometry. Rapid Communications in Mass Spectrometry,
27(11), 1183-1195.
Kozakai, K., Yamada, Y., Oshikata, M., Kawase, T., Suzuki, E., Haramaki, Y., Taniguchi, H., (2012)
Reliable high-throughput method for inhibition assay of 8 cytochrome P450 isoforms using
cocktail of probe substrates and stable isotope-labeled internal standards. Drug Metabolism
and Pharmacokinetics, 27(5), 520-529.
Kuentz, M., Nick, S., Parrott, N., Röthlisberger, D., (2006) A strategy for preclinical formulation
development using GastroPlusTM as pharmacokinetic simulation tool and a statistical screening
design applied to a dog study. European Journal of Pharmaceutical Sciences, 27(1), 91-99.
Kundlik, M.L., Zaware, B.H., Kuchekar, S.R., (2012) Rapid and Specific Approach for Direct
Measurement of Glimepiride in Human Plasma by LC–ESI-MS–MS Employing Automated 96
Well Format: Application to a Bioequivalence Study. Journal of Chromatographic Science, 50(1),
64-70.
Lal, R., Sukbuntherng, J., Luo, W., Chen, D., Vu, A., Tovera, J., Cundy, K.C., (2009) Pharmacokinetics
and Tolerability of Single Escalating Doses of Gabapentin Enacarbil: A Randomized-Sequence,
Double-Blind, Placebo-Controlled Crossover Study in Healthy Volunteers. Clinical Therapeutics,
31(8), 1776-1786.
Lal, R., Sukbuntherng, J., Luo, W., Huff, F.J., Zou, J., Cundy, K.C., (2010) The effect of food with
varying fat content on the clinical pharmacokinetics of gabapentin after oral administration of
gabapentin enacarbil. International Journal of Clinical Pharmacology and Therapeutics, 48(2),
120-128.
116 Pharmacokinetics and Bioanalysis to Improve Drug Development
Lal, R., Sukbuntherng, J., Luo, W., Tovera, J., Lassauzet, M.L., Cundy, K.C., (2013) Population
Pharmacokinetics and Pharmacodynamics of Gabapentin After Administration of Gabapentin
Enacarbil. Journal of Clinical Pharmacology, 53(1), 29-40.
Lavé, T., Parrott, N., Grimm, H.P., Fleury, A., Reddy, M., (2007) Challenges and opportunities with
modelling and simulation in drug discovery and drug development. Xenobiotica, 37(10-11),
1239-1310.
Lee, K.S., Kim, S.K., (2011) Direct and metabolism-dependent cytochrome P450 inhibition assays for
evaluating drug–drug interactions. Journal of Applied Toxicology, 33(2), 100-108.
Li, C., Liu, B., Chang, J., Groessl, T., Zimmerman, M., He, Y.Q., Isbell, J., Tuntland, T., (2013) A modern
in vivo pharmacokinetic paradigm: combining snapshot, rapid and full PK approaches to
optimize and expedite early drug discovery. Drug Discovery Today. 18(1-2), 71-78.
Lindegardh, N., Tarning, J., Toi, P.V., Hien, T.T., Farrar, J., Singhasivanon, P., White, N.J., Ashton, M.,
Day, N.P., (2009) Quantification of artemisinin in human plasma using liquid chromatography
coupled to tandem mass spectrometry. Journal of Pharmaceutical and Biomedical Analysis,
49(3), 768-773.
Lv, H., Zhang, X., Sharma, J., Reddy, M.V., Reddy, E.P., Gallo, J.M., (2013) Integrated Pharmacokinetic-
Driven Approach to Screen Candidate Anticancer Drugs for Brain Tumor Chemotherapy. AAPS
Journal, 15(1), 250-257.
Mallet, C.R., Lu, Z., Fisk, R., Mazzeo, J.R., Neue, U.D., (2003) Performance of an ultra-low elution-
volume 96-well plate: drug discovery and development applications. Rapid Communications in
Mass Spectrometry, 17(2), 163-170.
Marathe, P., Tang, Y., Sleczka, B., Rodrigues, D., Gavai, A., Wong, T., Christopher, L., Zhang, H.,
(2010) Preclinical Pharmacokinetics and In Vitro Metabolism of BMS-690514, a Potent Inhibitor
of EGFR and VEGFR2. Journal of Pharmaceutical Sciences, 99(8), 3579-3593.
Mayr, L.M., Bojanic, D., (2009) Novel trends in high-throughput screening. Current Opinion in
Pharmacology, 9(5), 580-588.
Milton, M.N., Horvath, C.J. (2009) The EMEA Guideline on First-in-Human Clinical Trials and Its
Impact on Pharmaceutical Development. Toxicologic Pathology, 37(3), 363-371.
Murphy, A.T., Witcher, D.R., Luan, P., Wroblewski, V.J., (2007) Quantitation of hepcidin from human
and mouse serum using liquid chromatography tandem mass spectrometry. Blood, 110(3),
1048-1054.
Ninomiya, M., Tanaka, K., Tsuchida, Y., Muto, Y., Koketsu, M., Watanabe, K., (2011) Increased
Bioavailability of Tricin-Amino Acid Derivatives via a Prodrug Approach. Journal of Medicinal
Chemistry, 54(4), 1529-1536.
Nishi, H., Nagamatsu, K. (2014) New trends in the LC separation analysis of pharmaceuticals – High
performance separation by ultra-high-performance liquid chromatography (UHPLC) with core-
shell particle C18 columns. Analytical Sciences, 30, 205-211.
Ohlmeyer, M., Zhou, M.-M. (2010) Integration of Small-Molecule Discovery in Academic Biomedical
Research. Mount Sinai Journal of Medicine, 77(4), 350-357.
Oh, K.S., Park, S.J., Shinde, D.D., Shin, J.G., Kim, D.H., (2012) High-sensitivity liquid
chromatography–tandem mass spectrometry for the simultaneous determination of five
drugs and their cytochrome P450-specific probe metabolites in human plasma. Journal of
Chromatography B, 895-896, 56-64.
Otten, J.N., Hingorani, G.P., Hartley, D.P., Kragerud, S.D., Franklin, R.B., (2012) An In Vitro, High
Throughput, Seven CYP Cocktail Inhibition Assay for the Evaluation of New Chemical Entities
Using LC-MS/MS. Drug Metabolism Letters, 5(1), 17-24.
Panchagnula, R., Thomas, N.S. (2000) Biopharmaceutics and pharmacokinetics in drug research.
International Journal of Pharmaceutics, 201(2), 131-150.
Parrrott, N., Lavè, T. (2002) Prediction of intestinal absorption: comparative assessment of
GASTROPLUSTM and IDEATM. European Journal of Pharmaceutical Sciences, 17(1-2), 51-61.
References 117
Pereira, D.A., Williams, J.A. (2007) Origin and evolution of high throughput screening. British Journal
of Pharmacology, 152(1), 53-61.
Pillai, V.C., Strom, S.C., Caritis, S.N., Venkataramanan, R., (2013) A sensitive and specific CYP
cocktail assay for the simultaneous assessment of human cytochrome P450 activities in
primary cultures of human hepatocytes using LC–MS/MS. Journal of Pharmaceutical and
Biomedical Analysis, 74, 126-132.
Plumb, R.S., Potts, W.B. 3rd, Rainville, P.D., Alden, P.G., Shave, D.H., Baynham, G., Mazzeo, J.R.,
(2008) Addressing the analytical throughput challenges in ADME screening using rapid
ultra-performance liquid chromatography/tandem mass spectrometry methodologies. Rapid
Communications in Mass Spectrometry, 22(14), 2139-2152.
Rhodes, S.P., Otten, J.N., Hingorani, G.P., Hartley, D.P., Franklin, R.B., (2011) Simultaneous
assessment of cytochrome P450 activity in cultured human hepatocytes for compound-
mediated induction of CYP3A4, CYP2B6, and CYP1A2. Journal of Pharmacological and
Toxicological Methods, 63(3), 223-226.
Rodriguez-Aller, M., Gurny, R., Veuthey, J.L., Guillarme, D., (2013) Coupling ultra-high pressure liquid
chromatography with mass spectrometry: Constraints and possible applications. Journal of
Chromatography A, 1292, 2-18.
Saxena, V., Panicucci, R., Joshi, Y., Garad, S., (2008) Developability Assessment in Pharmaceutical
Industry: An Integrated Group Approach for Selecting Developable Candidates. Journal of
Pharmaceutical Sciences, 98(6), 1962-1979.
Spaggiari, D., Geiser, L., Daali, Y., Rudaz, S., (2014) A cocktail approach for assessing the in
vitro activity of human cytochrome P450s: An overview of current methodologies. Journal of
Pharmaceutical and Biomedical Analysis, 101, 221-237
Srinivas N.R. (2008) Changing need for bioanalysis during drug development. Biomedical
Chromatography, 22(3), 235-243.
Sun, Y., Sun, J., Shi, S., Jing, Y., Yin, S., Chen, Y., Li, G., Xu, Y., He, Z., (2009) Synthesis, Transport and
Pharmacokinetics of 5′-Amino Acid Ester Prodrugs of 1-ß-D-Arabinofuranosylcytosine. Molecular
Pharmaceutics, 6(1), 315-325.
Tamimi, N.A.M., Ellis, P. (2009) Drug Development: From Concept to Marketing! Nephron Clinical
Practice, 113(3), c125-c131.
van Liempd, S., Morrison, D., Sysmans, L., Nelis, P., Mortishire-Smith, R., (2011) Development and
Validation of a Higher-Throughput Equilibrium Dialysis Assay for Plasma Protein Binding.
Journal of Laboratory Automation, 16(1), 56-67.
Volpe, D.A., (2010) Application of Method Suitability for Drug Permeability Classification. AAPS
Journal, 12(4), 670-678.
Volpe, D.A., (2011) Drug-permeability and transporter assays in Caco-2 and MDCK cell lines. Future
Medicinal Chemistry, 3(16), 2063-2077.
Wan, H., Holmen, A.G. (2009) High Throughput Screening of Physicochemical Properties and In
Vitro ADME Profiling in Drug Discovery. Combinatorial Chemistry & High Throughput Screening,
12(3), 315-329.
Wang, Y.F., Liu, Y.N., Xiong, W., Yan, D.M., Zhu, Y., Gao, X.M., Xu, Y.T., Qi, A.D., (2014) A UPLC–MS/
MS method for in vivo and in vitro pharmacokinetic studies of psoralenoside, isopsoralenoside,
psoralen and isopsoralen from Psoralea corylifolia extract. Journal of Ethnopharmacology,
151(1), 609-617.
Ward, K.W., Proksch, J.W., Levy, M.A., Smith, B.R., (2001) Development of an in vivo preclinical
screen model to estimate absortion and bioavailability of xenobiotics. Drug Metabolism and
Disposition, 29(1), 82-88.
Xie, R., Wen, J., Wei, H., Fan, G., Zhang, D., (2010) High-throughput determination of faropenem in
human plasma and urine by on-line solid-phase extraction coupled to high-performance liquid
chromatography with UV detection and its application to the pharmacokinetic study. Journal of
Pharmaceutical and Biomedical Analysis, 52(1), 114-121.
118 Pharmacokinetics and Bioanalysis to Improve Drug Development
Xue, Y.J., Pursley, J., Arnold, M.E., (2004) A simple 96-well liquid–liquid extraction with a mixture of
acetonitrile and methyl t-butyl ether for the determination of a drug in human plasma by high-
performance liquid chromatography with tandem mass spectrometry. Journal of Pharmaceutical
and Biomedical Analysis, 34(2), 369-378.
Yang, A.Y., Sun, L., Musson, D.G., Zhao, J.J., (2005) Application of a novel ultra-low elution volume
96-well solid-phase extraction method to the LC/MS/MS determination of simvastatin and
simvastatin acid in human plasma. Journal of Pharmaceutical and Biomedical Analysis, 38(3),
521-527.
Yan, Z., Sun, J., Chang, Y., Liu, Y., Fu, Q., Xu, Y., Sun, Y., Pu, X., Zhang, Y., Jing, Y., Yin, S., Zhu, M.,
Wang, Y., He, Z., (2011) Bifunctional Peptidomimetic Prodrugs of Didanosine for Improved
Intestinal Permeability and Enhanced Acidic Stability: Synthesis, Transepithelial Transport,
Chemical Stability and Pharmacokinetics. Molecular Pharmaceutics, 8(2): 319-329.
Zamek-Gliszczynski, M.J., Ruterbories, K.J., Ajamie, R.T., Wickremsinhe, E.R., Pothuri, L., Rao, M.V.,
Basavanakatti, V.N., Pinjari, J., Ramanathan, V.K., Chaudhary, A.K., (2011) Validation of 96-well
Equilibrium Dialysis with Non-radiolabeled Drug for Definitive Measurement of Protein Binding
and Application to Clinical Development of Highly-Bound Drugs. Journal of Pharmaceutical
Sciences, 100(6), 2498-2507.
Zhang, D., Luo, G., Ding, X., Lu, C. (2012) Preclinical experimental models of drug metabolism and
disposition in drug discovery and development. Acta Pharmaceutica Sinica B, 2(6), 549-561.
Zhang, J., Musson, G. (2006) Investigation of high-throughput ultrafiltration for the determination
of an unbound compound in human plasma using liquid chromatography and tandem mass
spectrometry with electrospray ionization. Journal of Chromatography B, 843(1), 47-56.
Zhu, M., Zhang, D., Zhang, H., Shyu, W.C., (2009) Integrated Strategies for Assessment of Metabolite
Exposure in Humans During Drug Development: Analytical Challenges and Clinical Development
Considerations. Biopharmaceutics and Drug Disposition, 30(4), 163-184.
Ma-Li Wong, Martin Lewis and Julio Licinio*4
2.2 Translational Research in Endocrinology and
Neuroimmunology Applied to Depression
Ma-Li Wong, Martin Lewis, Julio Licinio: Mind and Brain Theme, South Australian Health and Medical
Research Institute, PO Box 11060 Adelaide SA 5001, South Australia, Australia; Department of
Psychiatry, Flinders University School of Medicine, Sturt Road, Bedford Park 5042, South Australia,
Australia, *Email: julio.licinio@anu.edu.au
Four related ligands (CRH, urocortin I, II and III), two receptors (CRHR1 and 2),
and a binding protein (CRHBP) that acts as an endogenous antagonist have been
recognized in this system (Aubry, 2013; Heinrichs, 1997). For the purpose of this
chapter, we will focus on the CRH and the CRHR1, as both are widely distributed in
the brain, and CRH has high affinity to CRHR1 and poor affinity to CRHR2.
CRH is the key hypothalamic factor in the HPA axis, it stimulates the pituitary
gland to release adrenocorticotropin hormone (ACTH) and indirectly regulates
glucocorticoids secretion and its production in the adrenal cortex. Besides its
critical role in the HPA axis, CRH plays multiple additional roles in the stress
response; it is implicated in coordinating the behavioral, neuroendocrine,
autonomic and neurovegetative aspects of the stress response; it is relevant in
functions such as the activation the LC-NE system, the sympathetic nervous
system, catecholamine synthesis in the adrenal cortex, fear-related behavior,
and the inhibition of exploratory behavior, food consumption, and growth and
reproduction functions (Chrousos & Gold, 1992; Gold, 1988a, 1988b). CRH also
modulates anxiety behavior, as corroborated by findings in studies of mice lacking
CRHR1, which display decreased anxiety (Smith, 1998) and of CRHR1 antagonists,
such as Antalarmin, which is a non-peptide selective CRHR1 antagonist.
The Stress Response 121
2.2.2.2 The Locus Ceruleus Norepinephrine (LC-NE) System and Other Central Nervous
System (CNS) Structures that Modulate the Stress System
other (Fuster, 2000, 2001; Smith & Jonides, 1999), and activation of the prefrontal
cortex and consequent restraint of the stress and the sympathetic nervous systems
promote flexibility in cognition and affect (Smith & Jonides, 1999).
Stress affects many central and peripheral systems in the body, and the immune
system is critical for the promotion of survival; therefore, it has key roles during the
stress response and immune functions can be enhanced or suppressed by stressors.
Endocrine and cytokine mediators modulate the immune function during short-term
acute stress, which can modulate immune responsiveness encompassing humoral
and cellular aspects of both innate and adaptive immune responses.
The humoral changes related to stress response to psychological stressors, such
as the Trier Social Stress Test, include significant increases in the concentration of
cytokines, including interleukin-6 (IL-6) and IL-1ß (Altemus, 2001; Aschbacher,
2012; Pace, 2006; Prather, 2013; Puterman, 2014; Steptoe, 2007). Cytokines are small
proteins released by the immune system, typically under inflammatory situations.
These increased circulating cytokine levels may enhance the immune system
during acute stress, contribute to survival, and are also related to emotional states.
Peripheral IL-6 levels during stress reaction are related to the experience of anger;
it is likely that an angry individual will engage in an aggressive confrontation and
sustain injuries and an enhanced immune response will help promote wound healing
(Puterman, 2014).
It is now understood that peripheral cytokines modulate brain functions during
physiological conditions, where they can regulate neuronal processes, including
stress, inflammatory challenges, sickness behaviour, feeding, sleep, learning and
memory (Vitkovic, 2000a; 2000b), and that the communication between the brain
and the immune system is bidirectional. Our lab has contributed to the understanding
of many of the central aspects of cytokine response. We described the expression of
cytokines and immunomediators in the brain during baseline and after inflammatory
challenges (Licinio, 1991, 1992; Wong, 1997; 1996; Wong & Licinio, 1994); during
systemic inflammation, there is a high expression of central IL-1ß. Contrary to the
potent systemic counter regulatory anti-inflammatory response, in the brain the
expression of counter regulatory cytokines, such as IL-1 receptor antagonist and IL-
10, is much lower, which supports that differential regulation of anti-inflammatory
cytokines in the CNS and periphery (Wong, 1997).
Acute stress also orchestrates a massive redistribution of immune cells in the body,
which is consistent with the critical role the immune system has on survival; therefore,
functions such as wound healing and immune surveillance (Dhabhar, 2012) would be
enhanced during acute stress. Stress responsive hormones such as NE, corticosterone
and epinephrine (EPI) influence many different subsets of immune cells. NE and EPI
The Effect of Chronic Stress and MDD in Dysregulating the Core Stress System 123
are released early in the stress response; while, NE increases leucocyte numbers,
mobilizing immune cells, including neutrophils, monocytes, and lymphocytes to
enter the blood; EPI mobilizes neutrophils and monocytes into the blood, but directs
lymphocytes to leave the circulation to specific tissues, such as skin. Corticosterone is
then released and mobilizes immune cells to leave the blood towards tissues.
Acute restraint stress increases the numbers of T cells, such as memory and
effector helper cells, in sentinel lymphnodes (Dhabhar & Viswanathan, 2005). Stress-
induced increments in T cell memory may stimulate the increase of infiltrating
lymphocytes and macrophage detected on antigen re-exposure several months later
and this process is driven by increased levels of CD4+ T helper (Th) cells type 1 (Th1)
cytokines IL-2, interferon gamma (IFN-γ) and tumor necrosis factor alpha (TNF-α).
In chronic or long-term stress the physiological stress response continues long after
the stressor has ended, resulting in a prolonged exposure to stress hormones and
stress-related reactions, or is activated repeatedly as a result of continued exposure to
the stressor. Chronic activation of the HPA axis and sympathetic nervous system are
common in MDD, melancholic subtype; circulating levels of cortisol are elevated and
it is commonly accepted that the central CRH level is increased (or inappropriately
non-suppressed for the level of hypercortisolism). Elevations of the cerebrospinal
fluid (CSF) 24-hour levels of NE and plasma cortisol indicate that there is a persistent
stress-system dysfunction in MDD, melancholic subtype, which is not conditional
on the conscious awareness of stress (Wong, 2000). Circadian measures of CSF NE
and CSF CRH and peripheral cortisol suggest mutually reinforcing bidirectional
relationships between brain NE and CRH pathways that promotes increased central
noradrenergic, adrenomedullary, and adrenocortical secretion, which promote a state
of hyperarousal and anxiety found in MDD (Gold, 2005; Wong, 2000).
In MDD, atypical subtype, patients display fatigue, lethargy, hypersomnia,
and hyperphagia, which are related to suppression of the stress system mediators
(Chrousos & Gold, 1992; Gold, 1988a, 1988b). Patients with atypical features have
been reported to be eucortisolemic and have low CSF CRH levels (Geracioti, 1997;
Gold, 1995). However, detecting decrement in HPA axis is challenging. In response
to synthetic CRH, these patients show delays in peripheral ACTH response and very
attenuated cortisol responses (Schulte, 1984).
We have reported the involvement of the CRHR1 gene in antidepressant response;
anxious patients homozygous for the GAG haplotype of CRHR1 gene had a better
antidepressant response (Licinio, 2004); as pre-clinical studies suggested that
antidepressants suppress CRH gene expression (Brady, 1992; 1991; Reul, 1993) and
clinical studies reported that antidepressants reduce CSF CRH concentrations (De
124 Translational Research in Endocrinology and Neuroimmunology Applied to Depression
Bellis, 1993; Heuser, 1998; Veith, 1993), downregulation of CRH activity could be a
common pathway for the therapeutic actions of antidepressants.
Chronic stress can also result in dysregulation with activation or suppression of
the immune system; clinical and pre-clinical research support a role of immune system
dysregulation in MDD with activation of the cellular and humoral immune responses
(Kronfol & Remick, 2000; Licinio & Wong, 1999). Increased circulating lymphocytes
and monocytes have been described (Herbert & Cohen, 1993; Rothermundt, 2001;
Seidel, 1996). Increased circulating IL-6 levels have been described in MDD and these
levels were significantly correlated with self-reported mood ratings (Alesci, 2005).
Pro-inflammatory cytokines, such as IL-1ß and IL-6, stimulate the HPA axis (Plata-
Salaman, 1991). Stress-induced depressive-like behaviours also increase peripheral
IL-6 levels in rodents (Himmerich, 2013), and recently brain IL-6 has been implicated
in facilitating cognitive flexibility in the orbitofrontal cortex (Donegan, 2014). A
depressive syndrome can be associated with cytokine administration, typically
during IFN-α treatment of hepatitis C or malignant melanoma (Bonaccorso, 2000;
Dieperink, 2004; Hauser, 2002; Kraus, 2003; Schaefer, 2004), the psychopathological
presentation can include changes in mood or cognition, psychomotor retardation,
fatigability, sleep disturbances and decreased appetite (Capuron & Miller, 2004;
Loftis, 2006; Raison, 2006).
Dysregulation of the T-cell arm of the immune system has corroborated the role
of immunomediators in depression (Kronfol & Remick, 2000; Pavon, 2006). Evidence
supporting both the predominance of T-cells Th1 or Th2 have increased and the
balance of these two subtypes of cytokine-producing T cells Th1/Th2 regulates the
immune response to pathogens. These two main lineages of cells differentiate from
naïve CD4+ T lymphocytes and have distinct cytokine profiles. Th1 patterns are related
to increased inflammation and autoimmune conditions, and the production of IFN-γ,
TNF-α or IL-1; conversely, Th2 patterns are related to allergic responses and asthma,
and the production of IL-6, IL-10 or IL-13 (Kimball, 2005; Pavon, 2006; Scott, 2007).
Data from our lab supports that there is an imbalance of Th1/Th2 activities with
a predominance of Th1 in MDD; we found increased chemokine CXCL10/IP10 levels,
which decreased with antidepressant treatment, and decreased IL-13 levels; although
no distinct Th1 or Th2 cytokine patterns was found in our MDD patients (Wong, 2008).
Nevertheless, different types of MDD may exhibit divergent immune profiles, and
decreased Th1 activation has been described in melancholic depressed patients, which
is accompanied by a marked HPA axis activation. Non-melancholic MDD patients show
increased inflammation features with elevated monocyte count and levels of IL-1ß and
α2-macroglobulin (Kaestner, 2005; Rothermundt, 2001). Post-partum depression also
shows a predominance of Th1 activation (Maes, 2002; Ostensen, 2005), and symptoms
of anxiety and depression in the early post-partum period are related to increased
levels of pro-inflammatory cytokine IL-8 (Maes, 2002).
In the brain, stress-induced increased glucocorticoids enhances immune function,
promotes microglia proliferation and activation, and a loss of astrocytes number and
References 125
Optimal performance of the CNS requires that the core stress system be maintained
within a particular range, and emotional responses involve the integration of several
areas of the brain. Superb collaboration of several CNS and peripheral systems in a
sequentially coordinated manner are required for the optimal response of the stress
system, which will ultimately determine the ability of an organism to survive under a
threat. Dysregulation of the stress system is implicated in MDD; melancholia involves
the hyperactivity of the stress system and atypical features are related to the down-
regulation of the stress response; likewise, dysregulation of the immune response
implicates the activation or suppression of humoral and cellular immune components.
The core stress and immune systems interact with other CNS pathways that have also
been reported to be dysregulated in MDD, including the serotonergic, glutamatergic,
and kynurenine pathways (Muller & Schwarz, 2007).
References
Alesci, S., Martinez, P. E., Kelkar, S., Ilias, I., Ronsaville, D. S., Listwak, S. J., Gold, P. W. (2005). Major
depression is associated with significant diurnal elevations in plasma interleukin-6 levels,
a shift of its circadian rhythm, and loss of physiological complexity in its secretion: clinical
implications. Journal of Clinical Endocrinology & Metabolism, 90(5), 2522-2530. doi: 10.1210/
jc.2004-1667
Altemus, M., Rao, B., Dhabhar, F. S., Ding, W., & Granstein, R. D. (2001). Stress-induced changes in
skin barrier function in healthy women. Journal of Investigative Dermatology, 117(2), 309-317.
doi: 10.1046/j.1523-1747.2001.01373.x
Arnsten, A. F. (2000). Through the looking glass: differential noradenergic modulation of prefrontal
cortical function. Neural Plasticity, 7(1-2), 133-146. doi: 10.1155/NP.2000.133
126 Translational Research in Endocrinology and Neuroimmunology Applied to Depression
Aschbacher, K., Epel, E., Wolkowitz, O. M., Prather, A. A., Puterman, E., & Dhabhar, F. S. (2012).
Maintenance of a positive outlook during acute stress protects against pro-inflammatory
reactivity and future depressive symptoms. Brain, Behavior, and Immunity, 26(2), 346-352. doi:
10.1016/j.bbi.2011.10.010
Aston-Jones, G., & Cohen, J. D. (2005). An integrative theory of locus coeruleus-norepinephrine
function: adaptive gain and optimal performance. Annual Review of Neuroscience, 28, 403-450.
doi: 10.1146/annurev.neuro.28.061604.135709
Aston-Jones, G., Rajkowski, J., Kubiak, P., Valentino, R. J., & Shipley, M. T. (1996). Role of the locus
coeruleus in emotional activation. Progress in Brain Research, 107, 379-402.
Aubry, J. M. (2013). CRF system and mood disorders. Journal of Chemical Neuroanatomy, 54, 20-24.
doi: 10.1016/j.jchemneu.2013.09.003
Basterzi, A. D., Aydemir, C., Kisa, C., Aksaray, S., Tuzer, V., Yazici, K., & Goka, E. (2005). IL-6 levels
decrease with SSRI treatment in patients with major depression. Human Psychopharmacology,
20(7), 473-476. doi: 10.1002/hup.717
Berridge, C. W., & Waterhouse, B. D. (2003). The locus coeruleus-noradrenergic system: modulation of
behavioral state and state-dependent cognitive processes. Brain Research Reviews, 42(1), 33-84.
Bonaccorso, S., Meltzer, H., & Maes, M. (2000). Psychological and behavioural effects of interferons.
Current Opinion in Psychiatry, 13(6), 673-677.
Brady, L. S., Lynn, A. B., Whitfield, H. J., Jr., Kim, H., & Herkenham, M. (1992). Intrahippocampal
colchicine alters hypothalamic corticotropin-releasing hormone and hippocampal steroid
receptor mRNA in rat brain. Neuroendocrinology, 55(2), 121-133.
Brady, L. S., Whitfield, H. J., Jr., Fox, R. J., Gold, P. W., & Herkenham, M. (1991). Long-term
antidepressant administration alters corticotropin-releasing hormone, tyrosine hydroxylase,
and mineralocorticoid receptor gene expression in rat brain. Therapeutic implications. Journal
of Clinical Investigation, 87(3), 831-837. doi: 10.1172/JCI115086
Cahill, L., & McGaugh, J. L. (1998). Mechanisms of emotional arousal and lasting declarative
memory. Trends in Neuroscience, 21(7), 294-299.
Capuron, L., & Miller, A. H. (2004). Cytokines and psychopathology: lessons from interferon-alpha.
Biol Psychiatry, 56(11), 819-824. doi: 10.1016/j.biopsych.2004.02.009
Chrousos, G. P., & Gold, P. W. (1992). The concepts of stress and stress system disorders. Overview
of physical and behavioral homeostasis. JAMA, 267(9), 1244-1252.
Clark, D. A., & Beck, A. T. (2010). Cognitive theory and therapy of anxiety and depression:
convergence with neurobiological findings. Trends in Cognitive Sciences, 14(9), 418-424. doi:
10.1016/j.tics.2010.06.007
Cullinan, W. E., Herman, J. P., Battaglia, D. F., Akil, H., & Watson, S. J. (1995). Pattern and time course
of immediate early gene expression in rat brain following acute stress. Neuroscience, 64(2),
477-505.
Curtis, A. L., Leiser, S. C., Snyder, K., & Valentino, R. J. (2012). Predator stress engages
corticotropin-releasing factor and opioid systems to alter the operating mode of locus
coeruleus norepinephrine neurons. Neuropharmacology, 62(4), 1737-1745. doi: 10.1016/j.
neuropharm.2011.11.020
Czeh, B., Simon, M., Schmelting, B., Hiemke, C., & Fuchs, E. (2006). Astroglial plasticity in the
hippocampus is affected by chronic psychosocial stress and concomitant fluoxetine treatment.
Neuropsychopharmacology, 31(8), 1616-1626. doi: 10.1038/sj.npp.1300982
De Bellis, M. D., Gold, P. W., Geracioti, T. D., Jr., Listwak, S. J., & Kling, M. A. (1993). Association of
fluoxetine treatment with reductions in CSF concentrations of corticotropin-releasing hormone
and arginine vasopressin in patients with major depression. American Journal of Psychiatry,
150(4), 656-657.
Deak, T., Nguyen, K. T., Ehrlich, A. L., Watkins, L. R., Spencer, R. L., Maier, S. F., . . . Gold, P. W.
(1999). The impact of the nonpeptide corticotropin-releasing hormone antagonist antalarmin
References 127
on behavioral and endocrine responses to stress. Endocrinology, 140(1), 79-86. doi: 10.1210/
endo.140.1.6415
Dhabhar, F. S., Malarkey, W. B., Neri, E., & McEwen, B. S. (2012). Stress-induced redistribution
of immune cells--from barracks to boulevards to battlefields: a tale of three hormones--
Curt Richter Award winner. Psychoneuroendocrinology, 37(9), 1345-1368. doi: 10.1016/j.
psyneuen.2012.05.008
Dhabhar, F. S., & Viswanathan, K. (2005). Short-term stress experienced at time of immunization
induces a long-lasting increase in immunologic memory. American Journal of Physiology
- Regulatory, Integrative and Comparative Physiology, 289(3), R738-744. doi: 10.1152/
ajpregu.00145.2005
Dieperink, E., Ho, S. B., Tetrick, L., Thuras, P., Dua, K., & Willenbring, M. L. (2004). Suicidal ideation
during interferon-alpha2b and ribavirin treatment of patients with chronic hepatitis C. General
Hospital Psychiatry, 26(3), 237-240. doi: 10.1016/j.genhosppsych.2004.01.003
Donegan, J. J., Girotti, M., Weinberg, M. S., & Morilak, D. A. (2014). A novel role for brain
interleukin-6: facilitation of cognitive flexibility in rat orbitofrontal cortex. Journal of
Neuroscience, 34(3), 953-962. doi: 10.1523/JNEUROSCI.3968-13.2014
Frommberger, U. H., Bauer, J., Haselbauer, P., Fraulin, A., Riemann, D., & Berger, M. (1997).
Interleukin-6-(IL-6) plasma levels in depression and schizophrenia: comparison between the
acute state and after remission. European Archives of Psychiatry and Clinical Neuroscience,
247(4), 228-233.
Fuster, J. M. (2000). Memory networks in the prefrontal cortex. Progress in Brain Research, 122,
309-316.
Fuster, J. M. (2001). The prefrontal cortex--an update: time is of the essence. Neuron, 30(2), 319-333.
Geracioti, T. D., Jr., Loosen, P. T., & Orth, D. N. (1997). Low cerebrospinal fluid corticotropin-releasing
hormone concentrations in eucortisolemic depression. Biological Psychiatry, 42(3), 165-174.
doi: 10.1016/S0006-3223(96)00312-5
Gold, P. W., Goodwin, F. K., & Chrousos, G. P. (1988a). Clinical and biochemical manifestations of
depression. Relation to the neurobiology of stress (1). New England Journal of Medicine, 319(6),
348-353. doi: 10.1056/NEJM198808113190606
Gold, P. W., Goodwin, F. K., & Chrousos, G. P. (1988b). Clinical and biochemical manifestations of
depression. Relation to the neurobiology of stress (2). New England Journal of Medicine, 319(7),
413-420. doi: 10.1056/NEJM198808183190706
Gold, P. W., Licinio, J., Wong, M. L., & Chrousos, G. P. (1995). Corticotropin releasing hormone in the
pathophysiology of melancholic and atypical depression and in the mechanism of action of
antidepressant drugs. Annals of the New York Academy of Sciences, 771, 716-729.
Gold, P. W., Wong, M. L., Goldstein, D. S., Gold, H. K., Ronsaville, D. S., Esler, M., . . . Kling, M.
A. (2005). Cardiac implications of increased arterial entry and reversible 24-h central and
peripheral norepinephrine levels in melancholia. Proceedings of the National Academy of
Sciences USA, 102(23), 8303-8308. doi: 10.1073/pnas.0503069102
Habib, K. E., Weld, K. P., Rice, K. C., Pushkas, J., Champoux, M., Listwak, S., Gold, P. W. (2000).
Oral administration of a corticotropin-releasing hormone receptor antagonist significantly
attenuates behavioral, neuroendocrine, and autonomic responses to stress in primates.
Proceedings of the National Academy of Sciences USA, 97(11), 6079-6084.
Hauser, P., Khosla, J., Aurora, H., Laurin, J., Kling, M. A., Hill, J., Howell, C. D. (2002). A prospective
study of the incidence and open-label treatment of interferon-induced major depressive
disorder in patients with hepatitis C. Molecular Psychiatry, 7(9), 942-947. doi: 10.1038/
sj.mp.4001119
Heinrichs, S. C., Lapsansky, J., Lovenberg, T. W., De Souza, E. B., & Chalmers, D. T. (1997).
Corticotropin-releasing factor CRF1, but not CRF2, receptors mediate anxiogenic-like behavior.
Regulatory Peptides, 71(1), 15-21.
128 Translational Research in Endocrinology and Neuroimmunology Applied to Depression
Herbert, T. B., & Cohen, S. (1993). Depression and immunity: a meta-analytic review. Psychological
Bulletin, 113(3), 472-486.
Heuser, I., Bissette, G., Dettling, M., Schweiger, U., Gotthardt, U., Schmider, J., Holsboer, F. (1998).
Cerebrospinal fluid concentrations of corticotropin-releasing hormone, vasopressin, and
somatostatin in depressed patients and healthy controls: response to amitriptyline treatment.
Depression and Anxiety, 8(2), 71-79.
Himmerich, H., Fischer, J., Bauer, K., Kirkby, K. C., Sack, U., & Krugel, U. (2013). Stress-induced
cytokine changes in rats. European Cytokine Network, 24(2), 97-103. doi: 10.1684/
ecn.2013.0338
Kaestner, F., Hettich, M., Peters, M., Sibrowski, W., Hetzel, G., Ponath, G., Rothermundt, M. (2005).
Different activation patterns of proinflammatory cytokines in melancholic and non-melancholic
major depression are associated with HPA axis activity. Journal of Affective Disorders, 87(2-3),
305-311. doi: 10.1016/j.jad.2005.03.012
Kessler, R. C., Chiu, W. T., Demler, O., Merikangas, K. R., & Walters, E. E. (2005). Prevalence, severity,
and comorbidity of 12-month DSM-IV disorders in the National Comorbidity Survey Replication.
Archives of General Psychiatry, 62(6), 617-627. doi: 10.1001/archpsyc.62.6.617
Kessler, R. C., McGonagle, K. A., Zhao, S., Nelson, C. B., Hughes, M., Eshleman, S., Kendler, K. S.
(1994). Lifetime and 12-month prevalence of DSM-III-R psychiatric disorders in the United States.
Results from the National Comorbidity Survey. Archives of General Psychiatry, 51(1), 8-19.
Kimball, A. B., Jacobson, C., Weiss, S., Vreeland, M. G., & Wu, Y. (2005). The psychosocial burden of
psoriasis. American Journal of Clinical Dermatology, 6(6), 383-392.
Kraus, M. R., Schafer, A., Faller, H., Csef, H., & Scheurlen, M. (2003). Psychiatric symptoms in
patients with chronic hepatitis C receiving interferon alfa-2b therapy. Journal of Clinical
Psychiatry, 64(6), 708-714.
Kronfol, Z., & Remick, D. G. (2000). Cytokines and the brain: implications for clinical psychiatry.
American Journal of Psychiatry, 157(5), 683-694.
Kubera, M., Lin, A. H., Kenis, G., Bosmans, E., van Bockstaele, D., & Maes, M. (2001). Anti-
Inflammatory effects of antidepressants through suppression of the interferon-gamma/
interleukin-10 production ratio. Journal of Clinical Psychopharmacology, 21(2), 199-206.
LeDoux, J. E. (1995). Emotion: clues from the brain. Annual Review of Psychology, 46, 209-235. doi:
10.1146/annurev.ps.46.020195.001233
Leonard, B. E. (2001). The immune system, depression and the action of antidepressants. Progress
in Neuropsychopharmacology & Biological Psychiatry, 25(4), 767-780.
Licinio, J., O’Kirwan, F., Irizarry, K., Merriman, B., Thakur, S., Jepson, R., Wong, M. L. (2004).
Association of a corticotropin-releasing hormone receptor 1 haplotype and antidepressant
treatment response in Mexican-Americans. Molecular Psychiatry, 9(12), 1075-1082. doi:
10.1038/sj.mp.4001587
Licinio, J., & Wong, M. L. (1999). The role of inflammatory mediators in the biology of major
depression: central nervous system cytokines modulate the biological substrate of depressive
symptoms, regulate stress-responsive systems, and contribute to neurotoxicity and
neuroprotection. Molecular Psychiatry, 4(4), 317-327.
Licinio, J., & Wong, M. L. (2011). Advances in depression research: 2011. Molecular Psychiatry, 16(7),
686-687. doi: 10.1038/mp.2011.74
Licinio, J., Wong, M. L., & Gold, P. W. (1991). Localization of interleukin-1 receptor antagonist mRNA in
rat brain. Endocrinology, 129(1), 562-564. doi: 10.1210/endo-129-1-562
Licinio, J., Wong, M. L., & Gold, P. W. (1992). Neutrophil-activating peptide-1/interleukin-8 mRNA is
localized in rat hypothalamus and hippocampus. Neuroreport, 3(9), 753-756.
Loftis, J. M., Hauser, P., Macey, T. A., & Lowe, J. D. (2006). Can rodents be used to model interferon-
alpha-induced depressive symptoms? Progress in Neuropsychopharmacology & Biological
Psychiatry, 30(7), 1364-1365; author reply 1366. doi: 10.1016/j.pnpbp.2006.04.004
References 129
Lohoff, F. W. (2010). Overview of the genetics of major depressive disorder. Current Psychiatry
Reports, 12(6), 539-546. doi: 10.1007/s11920-010-0150-6
Lopez, A. D., & Murray, C. C. (1998). The global burden of disease, 1990-2020. Nature Medicine,
4(11), 1241-1243. doi: 10.1038/3218
Maes, M., Song, C., Lin, A. H., Bonaccorso, S., Kenis, G., De Jongh, R., . . . Scharpe, S. (1999).
Negative immunoregulatory effects of antidepressants: inhibition of interferon-gamma and
stimulation of interleukin-10 secretion. Neuropsychopharmacology, 20(4), 370-379. doi:
10.1016/S0893-133X(98)00088-8
Maes, M., Verkerk, R., Bonaccorso, S., Ombelet, W., Bosmans, E., & Scharpe, S. (2002). Depressive
and anxiety symptoms in the early puerperium are related to increased degradation of
tryptophan into kynurenine, a phenomenon which is related to immune activation. Life
Sciences, 71(16), 1837-1848.
McEwen, B. S. (1998). Protective and damaging effects of stress mediators. New England Journal of
Medicine, 338(3), 171-179. doi: 10.1056/NEJM199801153380307
Muller, N., & Schwarz, M. J. (2007). The immune-mediated alteration of serotonin and glutamate:
towards an integrated view of depression. Molecular Psychiatry, 12(11), 988-1000. doi: 10.1038/
sj.mp.4002006
Myint, A. M., Leonard, B. E., Steinbusch, H. W., & Kim, Y. K. (2005). Th1, Th2, and Th3 cytokine
alterations in major depression. Journal of Affective Disorders, 88(2), 167-173. doi: 10.1016/j.
jad.2005.07.008
Narita, K., Murata, T., Takahashi, T., Kosaka, H., Omata, N., & Wada, Y. (2006). Plasma
levels of adiponectin and tumor necrosis factor-alpha in patients with remitted major
depression receiving long-term maintenance antidepressant therapy. Progress in
Neuropsychopharmacology & Biological Psychiatry, 30(6), 1159-1162. doi: 10.1016/j.
pnpbp.2006.03.030
Obuchowicz, E., Kowalski, J., Labuzek, K., Krysiak, R., Pendzich, J., & Herman, Z. S. (2006).
Amitriptyline and nortriptyline inhibit interleukin-1 release by rat mixed glial and microglial
cell cultures. International Journal of Neuropsychopharmacology, 9(1), 27-35. doi: 10.1017/
S146114570500547X
Ostensen, M., Forger, F., Nelson, J. L., Schuhmacher, A., Hebisch, G., & Villiger, P. M. (2005).
Pregnancy in patients with rheumatic disease: anti-inflammatory cytokines increase in
pregnancy and decrease post partum. Annals of the Rheumatic Diseases, 64(6), 839-844. doi:
10.1136/ard.2004.029538
Pace, T. W., Mletzko, T. C., Alagbe, O., Musselman, D. L., Nemeroff, C. B., Miller, A. H., & Heim, C.
M. (2006). Increased stress-induced inflammatory responses in male patients with major
depression and increased early life stress. American Journal of Psychiatry, 163(9), 1630-1633.
doi: 10.1176/appi.ajp.163.9.1630
Pavon, L., Sandoval-Lopez, G., Eugenia Hernandez, M., Loria, F., Estrada, I., Perez, M., . . . Heinze, G.
(2006). Th2 cytokine response in Major Depressive Disorder patients before treatment. Journal
of Neuroimmunology, 172(1-2), 156-165. doi: 10.1016/j.jneuroim.2005.08.014
Plata-Salaman, C. R. (1991). Immunoregulators in the nervous system. Neuroscience & Biobehavioral
Reviews, 15(2), 185-215.
Prather, A. A., Puterman, E., Epel, E. S., & Dhabhar, F. S. (2013). Poor sleep quality potentiates
stress-induced cytokine reactivity in postmenopausal women with high visceral abdominal
adiposity. Brain, Behavior, and Immunity, doi: 10.1016/j.bbi.2013.09.010
Puterman, E., Epel, E. S., O’Donovan, A., Prather, A. A., Aschbacher, K., & Dhabhar, F. S. (2014).
Anger is associated with increased IL-6 stress reactivity in women, but only among those low
in social support. International Journal of Behavioral Medicine, 21(6), 936-945. doi: 10.1007/
s12529-013-9368-0
130 Translational Research in Endocrinology and Neuroimmunology Applied to Depression
Raison, C. L., Capuron, L., & Miller, A. H. (2006). Cytokines sing the blues: inflammation and the
pathogenesis of depression. Trends in Immunology, 27(1), 24-31. doi: 10.1016/j.it.2005.11.006
Reul, J. M., Stec, I., Soder, M., & Holsboer, F. (1993). Chronic treatment of rats with the
antidepressant amitriptyline attenuates the activity of the hypothalamic-pituitary-
adrenocortical system. Endocrinology, 133(1), 312-320. doi: 10.1210/endo.133.1.8391426
Rothermundt, M., Arolt, V., Fenker, J., Gutbrodt, H., Peters, M., & Kirchner, H. (2001). Different
immune patterns in melancholic and non-melancholic major depression. European Archives of
Psychiatry and Clinical Neuroscience, 251(2), 90-97.
Sapolsky, R. M. (1985). A mechanism for glucocorticoid toxicity in the hippocampus: increased
neuronal vulnerability to metabolic insults. Journal of Neuroscience, 5(5), 1228-1232.
Schaefer, M., Horn, M., Schmidt, F., Schmid-Wendtner, M. H., Volkenandt, M., Ackenheil, M.,
Schwarz, M. J. (2004). Correlation between sICAM-1 and depressive symptoms during adjuvant
treatment of melanoma with interferon-alpha. Brain, Behavior, and Immunity, 18(6), 555-562.
doi: 10.1016/j.bbi.2004.02.002
Schulte, H. M., Chrousos, G. P., Avgerinos, P., Oldfield, E. H., Gold, P. W., Cutler, G. B., Jr., &
Loriaux, D. L. (1984). The corticotropin-releasing hormone stimulation test: a possible aid
in the evaluation of patients with adrenal insufficiency. Journal of Clinical Endocrinology &
Metabolism, 58(6), 1064-1067. doi: 10.1210/jcem-58-6-1064
Scott, K. M., Von Korff, M., Ormel, J., Zhang, M. Y., Bruffaerts, R., Alonso, J., Haro, J. M. (2007).
Mental disorders among adults with asthma: results from the World Mental Health Survey.
General Hospital Psychiatry, 29(2), 123-133. doi: 10.1016/j.genhosppsych.2006.12.006
Seidel, A., Arolt, V., Hunstiger, M., Rink, L., Behnisch, A., & Kirchner, H. (1996). Major depressive
disorder is associated with elevated monocyte counts. Acta Psychiatrica Scandinavica, 94(3),
198-204.
Sluzewska, A., Rybakowski, J. K., Laciak, M., Mackiewicz, A., Sobieska, M., & Wiktorowicz, K. (1995).
Interleukin-6 serum levels in depressed patients before and after treatment with fluoxetine.
Annals of the New York Academy of Sciences, 762, 474-476.
Smith, E. E., & Jonides, J. (1999). Storage and executive processes in the frontal lobes. Science,
283(5408), 1657-1661.
Smith, G. W., Aubry, J. M., Dellu, F., Contarino, A., Bilezikjian, L. M., Gold, L. H., Lee, K. F. (1998).
Corticotropin releasing factor receptor 1-deficient mice display decreased anxiety, impaired
stress response, and aberrant neuroendocrine development. Neuron, 20(6), 1093-1102.
Snyder, K., Wang, W. W., Han, R., McFadden, K., & Valentino, R. J. (2012). Corticotropin-releasing
factor in the norepinephrine nucleus, locus coeruleus, facilitates behavioral flexibility.
Neuropsychopharmacology, 37(2), 520-530. doi: 10.1038/npp.2011.218
Steptoe, A., Hamer, M., & Chida, Y. (2007). The effects of acute psychological stress on circulating
inflammatory factors in humans: a review and meta-analysis. Brain, Behavior, and Immunity,
21(7), 901-912. doi: 10.1016/j.bbi.2007.03.011
Sullivan, P. F., Daly, M. J., & O’Donovan, M. (2012). Genetic architectures of psychiatric disorders: the
emerging picture and its implications. Nature Reviews Genetics, 13(8), 537-551. doi: 10.1038/
nrg3240
Szuster-Ciesielska, A., Tustanowska-Stachura, A., Slotwinska, M., Marmurowska-Michalowska, H., &
Kandefer-Szerszen, M. (2003). In vitro immunoregulatory effects of antidepressants in healthy
volunteers. Polish Journal of Pharmacology, 55(3), 353-362.
Tempier, R., Meadows, G. N., Vasiliadis, H. M., Mosier, K. E., Lesage, A., Stiller, A., Lepnurm,
M. (2009). Mental disorders and mental health care in Canada and Australia: comparative
epidemiological findings. Social Psychiatry and Psychiatric Epidemiology, 44(1), 63-72. doi:
10.1007/s00127-008-0409-y
References 131
Veith, R. C., Lewis, N., Langohr, J. I., Murburg, M. M., Ashleigh, E. A., Castillo, S. (1993). Effect of
desipramine on cerebrospinal fluid concentrations of corticotropin-releasing factor in human
subjects. Psychiatry Research, 46(1), 1-8.
Vitkovic, L., Bockaert, J., & Jacque, C. (2000). “Inflammatory” cytokines: neuromodulators in normal
brain? Journal of Neurochemistry, 74(2), 457-471.
Vitkovic, L., Konsman, J. P., Bockaert, J., Dantzer, R., Homburger, V., & Jacque, C. (2000). Cytokine
signals propagate through the brain. Molecular Psychiatry, 5(6), 604-615.
Vos, T., Haby, M. M., Barendregt, J. J., Kruijshaar, M., Corry, J., & Andrews, G. (2004). The burden
of major depression avoidable by longer-term treatment strategies. Archives of General
Psychiatry, 61(11), 1097-1103. doi: 10.1001/archpsyc.61.11.1097
Wong, M. L., Bongiorno, P. B., Rettori, V., McCann, S. M., & Licinio, J. (1997). Interleukin (IL) 1beta,
IL-1 receptor antagonist, IL-10, and IL-13 gene expression in the central nervous system and
anterior pituitary during systemic inflammation: pathophysiological implications. Proceedings
of the National Academy of Sciences, 94(1), 227-232.
Wong, M. L., Dong, C., Maestre-Mesa, J., & Licinio, J. (2008). Polymorphisms in inflammation-related
genes are associated with susceptibility to major depression and antidepressant response.
Molecular Psychiatry, 13(8), 800-812. doi: 10.1038/mp.2008.59
Wong, M. L., Kling, M. A., Munson, P. J., Listwak, S., Licinio, J., Prolo, P., Gold, P. W. (2000).
Pronounced and sustained central hypernoradrenergic function in major depression with
melancholic features: relation to hypercortisolism and corticotropin-releasing hormone.
Proceedings of the National Academy of Sciences, 97(1), 325-330.
Wong, M. L., & Licinio, J. (1994). Localization of interleukin 1 type I receptor mRNA in rat brain.
Neuroimmunomodulation, 1(2), 110-115.
Wong, M. L., & Licinio, J. (2001). Research and treatment approaches to depression. Nature Reviews
Neuroscience, 2(5), 343-351. doi: 10.1038/35072566
Wong, M. L., & Licinio, J. (2004). From monoamines to genomic targets: a paradigm shift for drug
discovery in depression. Nature Reviews Drug Discovery, 3(2), 136-151. doi: 10.1038/nrd1303
Wong, M. L., Rettori, V., al-Shekhlee, A., Bongiorno, P. B., Canteros, G., McCann, S. M., . . . Licinio,
J. (1996). Inducible nitric oxide synthase gene expression in the brain during systemic
inflammation. Nature Medicine, 2(5), 581-584.
Woolley, C. S., Gould, E., & McEwen, B. S. (1990). Exposure to excess glucocorticoids alters dendritic
morphology of adult hippocampal pyramidal neurons. Brain Research, 531(1-2), 225-231.
Cidália Pereira, Rosário Monteiro, Maria João Martins*5
2.3 Understanding the Metabolic Syndrome Using
a Biomedical Chemistry Profile
2.3.1 Introduction
For quite some time, it has been identified that high blood pressure, dyslipidemia
[increased triglycerides and reduced high-density lipoprotein (HDL) cholesterol
levels], impaired glucose homeostasis and abdominal obesity take place concurrently
more than by random, supporting the existence of the metabolic syndrome (MetSyn).
Additionally, hyperuricemia, a prothrombotic state, oxidative stress, chronic low-
grade inflammation, increased levels of apolipoprotein-B and small dense low-
density lipoprotein (LDL) cholesterol (contributing to atherogenic dyslipidemia),
non-alcoholic fatty liver disease and/or non-alcoholic steatohepatitis, obstructive
sleep apnea and/or polycystic ovarian disease (Fulop, 2006; Alberti, 2009; Roberts,
2009; Ma, 2012; Matsuda, 2013; Mule, 2014; Carson, 2015) are quite often present on
the MetSyn, although not yet included in its current/actual definition. Taking this into
consideration, it is not surprising that the MetSyn associates with an increased risk
of type 2 diabetes mellitus (T2DM) and atherosclerotic cardiovascular disease (Fulop,
2006; Qiao, 2007; Carson, 2015).
Epidemiological and experimental evidence has demonstrated beneficial effects of
dietary magnesium, calcium, potassium and bicarbonate on the MetSyn or some of its
individual components (Luft, 1990; Van Leer, 1995; Schorr, 1996; Whelton, 1997; Franch,
2004; Rylander, 2004; Schoppen, 2004; Franzoni, 2005; Karppanen, 2005; Schoppen,
2005; He, 2006; Feldeisen, 2007; Schoppen, 2007; Champagne, 2008; Rylander, 2008;
Volpe, 2008; Perez-Granados, 2010; Adeva, 2011; Rice, 2011; Lee, 2013).
Natural mineral-rich waters are good sources of highly absorbable and bioavailable
minerals (such as calcium, magnesium and potassium) and bicarbonate (Heaney,
1989; Bohmer, 2000; Kessler, 2000; Sabatier, 2002; Bacciottini, 2004; Kiss, 2004;
Heaney, 2006; Karagulle, 2006; Petraccia, 2006; Karagulle, 2007; Sabatier, 2011). So,
the particular composition of natural mineral-rich waters would be responsible for
their favorable effects. In fact, the World Health Organization has recognized that
Cidália Pereira: Department of Biochemistry, Faculty of Medicine, University of Porto, Alameda Prof.
Hernâni Monteiro, 4200-319, Porto, Portugal; Escola Superior de Saúde, Instituto Politécnico de Lei-
ria, Campus 2 – Morro de Lena - Alto do Vieiro, Apartado 4137, 2411-901, Leiria, Portugal.
Rosário Monteiro, Maria João Martins: Department of Biochemistry, Faculty of Medicine, Universi-
ty of Porto, Alameda Prof. Hernâni Monteiro, 4200-319, Porto, Portugal ; Instituto de Investigação
e Inovação em Saúde, Universidade do Porto, Rua Alfredo Allen 208, 4200-135 Porto, Portugal.
*Email: mmartins@med.up.pt.
Chronic deficiency of magnesium (in animal models with low magnesium intake) is
associated with hypertension and increased heart rate (and somewhat higher plasma
corticosterone levels) as well as dyslipidemia, insulin resistance and oxidative stress
(Caddell, 1991; Balon 1994; Laurant, 1999; Busserolles, 2003; Takaya, 2012).
Clinical and experimental studies point to magnesium intake/status being
inversely associated with the risk of hypertension, T2DM and coronary heart disease.
Additionally, magnesium intake may decrease triglycerides and increased HDL-
cholesterol circulating levels (Balon, 1994; Touyz, 2003; Takaya, 2004; Barbagallo,
2007; Belin, 2007; Abete, 2011; Heer, 2015).
Individuals with MetSyn (or with some of its individual components) frequently
show reduced magnesium status and reduced magnesium intake as compared with
non-MetSyn (or healthy) subjects (Barbagallo, 2007; Belin, 2007; Evangelopoulos,
2008; Abete, 2011; Heer, 2015). Interestingly, hypomagnesaemia has been associated
with metabolic abnormalities characteristic of MetSyn in the absence of obesity and,
Magnesium and MetSyn/MetSyn Features – Associated Mechanisms 135
Serum calcium levels are increased in individuals with MetSyn and positively
correlated with serum triglycerides and blood pressure values (Park, 2012; Cho,
2011). Calcium and/or dairy products ingestion has been associated with decreased
adipose tissue mass, lower blood pressure, better circulating lipid profile (increased
HDL-cholesterol and decreased total-cholesterol and LDL-cholesterol levels as well as
decreased total-cholesterol to HDL-cholesterol ratio and increased HDL-cholesterol to
LDL-cholesterol ratio) and prevention/reduced incidence of insulin resistance, T2DM
and MetSyn (Teegarden, 2006; van Meijl, 2008; Tremblay, 2009; Abete, 2011; Rice,
2011). The inconsistency between serum calcium levels and dietary calcium intake
values with MetSyn still needs to be clarified (Park, 2012; Cho, 2011).
The increase of fecal excretion of fat and inhibition of intestinal bile salts
absorption (with the consequent increase of cholesterol conversion into bile acids
in the liver) as well as the stimulation of lipolysis and inhibition of lipogenesis
improve plasma lipid profile and decrease adipose tissue mass as a consequence of
calcium intake (Abete, 2011; Rice, 2011). Regulation of parathyroid hormone and 1,25-
dihydroxycholecalciferol levels by dietary calcium mediates its effects on fat mass
and insulin signaling/action (Teegarden, 2006). There is also evidence that adequate
138 Understanding the Metabolic Syndrome Using a Biomedical chemistry profile
secretion (Norbiato, 1984; Lee, 2013). Serum potassium levels are closely regulated
by homeostatic mechanisms and depend on dietary potassium intake and potassium
excretion (and its regulators) as well as on partitioning between intracellular and
extracellular spaces (modulated by insulin, catecholamines and thyroid hormone)
(Chatterjee, 2011). Renal potassium excretion is primarily controlled by sodium
delivery to the distal nephron and urine flow, vasopressin levels, acid-base status
(also hormone regulated, as above) and the renin-angiotensin-aldosterone system
(Chatterjee, 2011). As a consequence of the strict control of serum potassium levels,
dietary and serum potassium levels are not inevitably associated (Chatterjee, 2011).
with cortisol excess, the latter being prevented by bicarbonate administration (Zhang,
2009; Adeva, 2011; Pereira, 2011; Pereira, 2012b; Pereira, 2012c).
In this context, once more, natural mineral-rich water consumption would be
extremely beneficial.
2.3.8 Conclusion
Acknowledgements: This work has been supported by FCT (Fundação para a Ciência
e Tecnologia, PEst-OE/SAU/UI0038/2014) and Instituto de Investigação e Inovação em
Saúde, University of Porto, Porto, Portugal. We thank Unicer Bebidas, S.A. (Leça do Balio,
Matosinhos, Portugal) the supply of the Portuguese natural mineral-rich water tested.
References 141
References
Abete, I., Goyenechea, E., Zulet, M. A., & Martinez, J. A. (2011). Obesity and metabolic syndrome:
potential benefit from specific nutritional components. Nutrition, Metabolism & Cardiovascular
Diseases, 21(S2), B1-15.
Adeva, M. M., & Souto, G. (2011). Diet-induced metabolic acidosis. Clin Nutr 30(4): 416-421.
Al-Solaiman, Y., Jesri, A., Mountford, W. K., Lackland, D. T., Zhao, Y., & Egan, B. M. (2010). DASH
lowers blood pressure in obese hypertensives beyond potassium, magnesium and fibre. Journal
of Human Hypertension, 24(4), 237-246.
Alberti, K. G., Eckel, R. H., Grundy, S. M., Zimmet, P. Z., Cleeman, J. I., Donato, K. A., Fruchart, J. C.,
James, W. P., Loria, C. M., & Smith, S. C. Jr. (2009). Harmonizing the metabolic syndrome: a joint
interim statement of the International Diabetes Federation Task Force on Epidemiology and
Prevention; National Heart, Lung, and Blood Institute; American Heart Association; World Heart
Federation; International Atherosclerosis Society; and International Association for the Study of
Obesity. Circulation, 120(16), 1640-1645.
Bacciottini, L., Tanini, A., Falchetti, A., Masi, L., Franceschelli, F., Pampaloni, B., Giorgi, G., & Brandi,
M. L. (2004). Calcium bioavailability from a calcium-rich mineral water, with some observations
on method. Journal of Clinical Gastroenterology, 38(9), 761-766.
Balon, T. W., Jasman, A., Scott, S., Meehan, W. P., Rude, R. K., & Nadler, J. L. (1994). Dietary magnesium
prevents fructose-induced insulin insensitivity in rats. Hypertension, 23(6 Pt 2), 1036-1039.
Barbagallo, M., & Dominguez, L. J. (2007). Magnesium metabolism in type 2 diabetes mellitus, metabolic
syndrome and insulin resistance. Archives of Biochemistry and Biophysics, 458(1), 40-47.
Bastos, P., Araújo, J. R., Azevedo, I., Martins, M. J., & Ribeiro, L. (2014). Effect of a natural mineral-
rich water on catechol-O-methyltransferase function. Magnesium research, 27(3),131-141.
Belin, R. J., & He, K. (2007). Magnesium physiology and pathogenic mechanisms that contribute to
the development of the metabolic syndrome. Magnes Res 20(2): 107-129.
Benedetti, S., Benvenuti, F., Nappi, G., Fortunati, N. A., Marino, L., Aureli, T., De Luca S., Pagliarani,
S., & Canestrari, F. (2009). Antioxidative effects of sulfurous mineral water: protection against
lipid and protein oxidation. European Journal of Clinical Nutrition, 63(1), 106-112.
Blaustein, M. P., Zhang, J., Chen, L., & Hamilton, B. P. (2006). How does salt retention raise blood
pressure? American Journal of Regulatory, Integrative, and Comparative Physiology, 290(3),
R514-523.
Bohmer, H., Muller, H., & Resch, K. L. (2000). Calcium supplementation with calcium-rich mineral
waters: a systematic review and meta-analysis of its bioavailability. Osteoporosis International,
11(11), 938-943.
Botvineva, L. A., Nikitin, E. N., Mel’nikova, L. N., & Akaeva, E. A. (2010). The use of drinking mineral
waters and the fiber-enriched diet for the treatment of patients with type 2 diabetes mellitus.
Voprosy kurortologii, fizioterapii, i lechebnoi fizicheskoi kultury, Mar-Apr(2), 13-16 (In Russian).
Brandolini, M., Gueguen, L., Boirie, Y., Rousset, P., Bertiere, M. C., & Beaufrere, B. (2005). Higher
calcium urinary loss induced by a calcium sulphate-rich mineral water intake than by milk in
young women. British Journal of Nutrition, 93(2), 225-231.
Busserolles, J., Gueux, E., Rock, E., Mazur, A., & Rayssiguier, Y. (2003). High fructose feeding of
magnesium deficient rats is associated with increased plasma triglyceride concentration and
increased oxidative stress. Magnesium research, 16(1), 7-12.
Bussière, L., Mazur, A., Gueux, E., Nowacki, W., & Rayssiguier, Y. (1995). Triglyceride-rich lipoproteins
from magnesium-deficient rats are more susceptible to oxidation by cells and promote
proliferation of cultured vascular smooth muscle cells. Magnesium research, 8(2), 151-157.
Cacho, J., Sevillano, J,. de Castro, J., Herrera, E., & Ramos, M.P. (2008). Validation of simple indexes
to assess insulin sensitivity during pregnancy in Wistar and Sprague-Dawley rats. American
Journal of Physiology Endocrinology and Metabolism, 295(5), E1269-E1276.
142 Understanding the Metabolic Syndrome Using a Biomedical chemistry profile
Fulop, T., Tessier, D., & Carpentier, A. (2006). The metabolic syndrome. Pathologie Biologie (Paris)
54(7), 375-386.
Geleijnse, J. M., Grobbee, D. E., & Kok, F. J. (2005). Impact of dietary and lifestyle factors on the
prevalence of hypertension in Western populations. Journal of Human Hypertension, 19 (S3),
S1-4.
Guerrera, M. P., Volpe, S. L., & Mao, J. J. (2009). Therapeutic uses of magnesium. American Family
Physician, 80(2): 157-162.
Guerrero-Romero, F., & Rodriguez-Moran, M. (2013). Serum magnesium in the metabolically-obese
normal-weight and healthy-obese subjects. European Journal of Internal Medicin, 24(7), 639-
643.
Gueux, E., Azais-Braesco, V., Bussière, L., Grolier, P., Mazur, A., & Rayssiguier, Y. (1995). Effect
of magnesium deficiency on triacylglycerol-rich lipoprotein and tissue susceptibility to
peroxidation in relation to vitamin E content. British Journal of Nutrition 74(6), 849-856.
He, K., Song, Y., Belin, R. J., & Chen, Y. (2006). Magnesium intake and the metabolic syndrome:
epidemiologic evidence to date. Journal of the cardiometabolic syndrome, 1(5), 351-355.
Heaney, R. P. (2006). Absorbability and utility of calcium in mineral waters. American Journal of
Clinical Nutrition, 84(2), 371-374.
Heaney, R. P., Smith, K. T., Recker, R. R., & Hinders, S. M. (1989). Meal effects on calcium absorption.
American Journal of Clinical Nutrition, 49(2), 372-376.
Heer, M., & Egert, S. (2015). Nutrients other than carbohydrates: their effects on glucose
homeostasis in humans. Diabetes/Metabolism Research and Reviews, 31(1), 14-35.
Huerta, M. G., Roemmich, J. N., Kington, M. L., Bovbjerg, V. E., Weltman, A. L., Holmes, V. F., Patrie,
J. T., Rogol, A. D., & Nadler, J. L. (2005). Magnesium deficiency is associated with insulin
resistance in obese children. Diabetes Care, 28(5), 1175-1181.
Ikari, A., Atomi, K., Kinjo, K., Sasaki, Y., & Sugatani, J. (2010). Magnesium deprivation inhibits a
MEK-ERK cascade and cell proliferation in renal epithelial Madin-Darby canine kidney cells. Life
Sciences, 86(19-20), 766-773.
Karagulle, O., Kleczka, T., Vidal, C., Candir, F., Gundermann, G., Kulpmann, W. R., Gehrke, A., &
Gutenbrunner, C. (2006). Magnesium absorption from mineral waters of different magnesium
content in healthy subjects. Forsch Komplementmed, 13(1), 9-14.
Karagulle, O., Smorag, U., Candir, F., Gundermann, G., Jonas, U., Becker, A. J., Gehrke, A., &
Gutenbrunner, C. (2007). Clinical study on the effect of mineral waters containing bicarbonate
on the risk of urinary stone formation in patients with multiple episodes of CaOx-urolithiasis.
World Journal of Urology, 25(3), 315-323.
Karppanen, H., Karppanen, P., & Mervaala, E. (2005). Why and how to implement sodium,
potassium, calcium, and magnesium changes in food items and diets? Journal of Human
Hypertension, 19(Suppl 3), S10-19.
Kessler, T., & Hesse, A. (2000). Cross-over study of the influence of bicarbonate-rich mineral water
on urinary composition in comparison with sodium potassium citrate in healthy male subjects.
British Journal of Nutrition, 84(6), 865-871.
Kiss, S. A., Forster, T., & Dongo, A. (2004). Absorption and effect of the magnesium content of a mineral
water in the human body. Journal of the American College of Nutrition, 23(6), 758s-762s.
Kitada, M., Kume, S., Kanasaki, K., Takeda-Watanabe, A., & Koya, D. (2013a). Sirtuins as possible
drug targets in type 2 diabetes. Current Cancer Drug Targets, 14(6), 622-636.
Kitada, M., & Koya, D. (2013b). SIRT1 in Type 2 Diabetes: Mechanisms and Therapeutic Potential.
Diabetes and Metabolism Journal, 37(5), 315-325.
Kunes, J., Zicha J., & Jelinek, J. (2004). The role of chloride in deoxycorticosterone hypertension:
selective sodium loading by diet or drinking fluid. Physiological Research, 53(2), 149-154.
Laurant, P., Dalle, M., Berthelot, A., & Rayssiguier, Y. (1999). Time-course of the change in blood
pressure level in magnesium-deficient Wistar rats. British Journla of Nutrition, 82(3), 243-251.
144 Understanding the Metabolic Syndrome Using a Biomedical chemistry profile
Lecube, A., Baena-Fustegueras, J. A., Fort, J. M., Pelegri, D., Hernandez, C., & Simo, R. (2012). Diabetes
is the main factor accounting for hypomagnesemia in obese subjects. PLoS One, 7(1), e30599.
Lee, H., Lee, J., Hwang, S. S., Kim, S., Chin, H. J., Han, J. S., & Heo, N. J. (2013). Potassium intake and
the prevalence of metabolic syndrome: the Korean National Health and Nutrition Examination
Survey 2008-2010. PLoS One, 8(1), e55106.
Li, Y., Wong K., Giles, A., Jiang, J., Lee, J. W., Adams, A. C., Kharitonenkov, A., Yang, Q., Gao, B.,
Guarente, L., & Zang, M. (2014). Hepatic SIRT1 attenuates hepatic steatosis and controls energy
balance in mice by inducing fibroblast growth factor 21. Gastroenterology, 146(2), 539-549.e537.
Luft, F. C., Zemel, M. B., Sowers, J. A., Fineberg, N. S., & Weinberger, M. H. (1990). Sodium
bicarbonate and sodium chloride: effects on blood pressure and electrolyte homeostasis in
normal and hypertensive man. Journal of Hypertension, 8(7), 663-670.
Ma, K., Jin, X., Liang, X., Zhao, Q., & Zhang X. (2012). Inflammatory mediators involved in the progression
of the metabolic syndrome. Diabetes/Metabolism Research and Reviews, 28(5), 388-394.
MacFadyen, R. J., Barr, C. S., & Struthers, A. D. (1997). Aldosterone blockade reduces vascular
collagen turnover, improves heart rate variability and reduces early morning rise in heart rate in
heart failure patients. Cardiovascular Research, 35(1), 30-34.
Maraver, Fi, Vitoria, I., Ferreira-Pêgo, C., Armijo, F., & Salas-Salvadó, J. (2015). Magnesium in tap and
bottled mineral water in spain and its contribution to nutritional recommendations. Nutr Hosp,
31(n05), 2297-2312.
Markiewicz-Gorka, I., Zawadzki, M., Januszewska, L., Hombek-Urban, K., & Pawlas, K. (2011).
Influence of selenium and/or magnesium on alleviation alcohol induced oxidative stress in rats,
normalization function of liver and changes in serum lipid parameters. Human & Experimental
Toxicology, 30(11), 1811-1827.
Matsuda, M., & Shimomura, I. (2013). Increased oxidative stress in obesity: implications for
metabolic syndrome, diabetes, hypertension, dyslipidemia, atherosclerosis, and cancer.
Obesity Research & Clinical Practice, 7(5), e330-341.
McNair, P., Christensen, M. S., Christiansen, C., Madsbad, S., & Transbol, I. (1982). Renal
hypomagnesaemia in human diabetes mellitus: its relation to glucose homeostasis. European
Journal of Clinical Investigation, 12(1), 81-85.
Mule, G., Calcaterra, I., Nardi, E., Cerasola, G., & Cottone, S. (2014). Metabolic syndrome in
hypertensive patients: An unholy alliance. World Journal of Cardiology, 6(9), 890-907.
Nobre, J. L., Lisboa, P. C., Santos-Silva, A. P., Lima, N. S., Manhaes, A. C., Nogueira-Neto,
J. F., Cabanelas, A., Pazos-Moura, C. C., Moura, E. G., & de Oliveira, E. (2011). Calcium
supplementation reverts central adiposity, leptin, and insulin resistance in adult offspring
programed by neonatal nicotine exposure. Journal of Endocrinology, 210(3), 349-359.
Norbiato, G., Bevilacqua, M., Meroni, R., Raggi, U., Dagani, R., Scorza, D., Frigeni, G., & Vago, T. (1984).
Effects of potassium supplementation on insulin binding and insulin action in human obesity:
protein-modified fast and refeeding. European Journal of Clinical Investigation, 14(6), 414-419.
Odermatt, A. (2011). The Western-style diet: a major risk factor for impaired kidney function and
chronic kidney disease. American Journal of Physiology Renal Physiology, 301(5), F919-931.
Olatunji, L. A., & Soladoye, A. O. (2007). Increased magnesium intake prevents hyperlipidemia and
insulin resistance and reduces lipid peroxidation in fructose-fed rats. Pathophysiology, 14(1), 11-15.
Oron-Herman, M., Kamari, Y., Grossman, E., Yeger, G., Peleg, E., Shabtay, Z., Shamiss, A., & Sharabi,
Y. (2008). Metabolic syndrome: comparison of the two commonly used animal models.
American Journal of Hypertension, 21(9), 1018-1022.
Paredes, S., & Ribeiro, L. (2014). Cortisol: the villain in Metabolic Syndrome? Revista da Associação
Médica Brasileira, 60(1), 84-92.
Park, S. H., Kim, S. K., & Bae, Y. J. (2012). Relationship between serum calcium and magnesium
concentrations and metabolic syndrome diagnostic components in middle-aged Korean men.
Biological Trace Element Research, 146(1), 35-41.
References 145
Pereira, C., Monteiro, R., Constância, M., & Martins, M. J. (2012c). 11β-Hydroxysteroid
Dehydrogenases in the Regulation of Tissue Glucocorticoid Availability. Steroids - From
Physiology to Clinical Medicine. S. Ostojic, InTech: 83-106.
Pereira, C., Passos, E., Neves, D., Ascensão, A., Magalhães, J., Monteiro, R., & Martins, M. J. (2013).
Modulation of stress pathways by a hypersaline sodium-rich carbonated natural mineral water
seems to improve the hepatic insulin signalling in an animal model of metabolic syndrome.
5th International Symposium Nutrition, Oxygen Biology and Medicine - Société Française de
Recherche sur les Radicaux Libres.
Pereira, C. D., Azevedo, I., Monteiro, R., & Martins, M. J. (2012b). 11beta-Hydroxysteroid dehydrogenase
type 1: relevance of its modulation in the pathophysiology of obesity, the metabolic syndrome and
type 2 diabetes mellitus. Diabetes, Obesity and Metabolism 14(10): 869-881.
Pereira, C. D., Collado, M. C., Passos, E., Azevedo, I., Monteiro, R., & Martins, M. J. (2014c). Comment
to: Luo (2013) International Journal of Cardiology, 168(4), 4454-6. International Journal of
Cardiology, 172(2), 512-514.
Pereira, C. D., Martins, M. J., Azevedo, I., & Monteiro, R. (2011). 11β-Hydroxysteroid Dehydrogenase
Type 1 and the Metabolic Syndrome. Steroids - Clinical Aspect. P. H. Abduljabbar, InTech: 25-50.
Pereira, C. D., Severo, M., Araújo, J. R., Guimarães, J. T., Pestana, D., Santos, A., Ferreira, R.,
Ascensão, A., Magalhães, J., Azevedo, I., Monteiro, R., & Martins, M. J. (2014a). Relevance of a
hypersaline sodium-rich naturally sparkling mineral water to the protection against metabolic
syndrome induction in fructose-fed sprague-dawley rats: a biochemical, metabolic, and redox
approach. International Journal of Endocrinology, 2014 (Article ID 384583).
Pereira, C. D., Severo, M., Neves, D., Ascensao, A., Magalhaes, J., Guimaraes, J. T., Monteiro,
R., & Martins, M. J. (2015). Natural mineral-rich water ingestion improves hepatic and fat
glucocorticoid-signaling and increases sirtuin 1 in an animal model of metabolic syndrome,
Hormone Molecular Biology and Clinical Investigation, 21(2), 149-157.
Pereira, C. D., Severo, M., Rafael, L., Martins, M. J., & Neves, D. (2014b). Effects of natural mineral-
rich water consumption on the expression of Sirtuin 1, vascular endothelial growth factor,
angiopoietins, and receptors in the erectile tissue of rats with fructose-induced Metabolic
Syndrome. Asian Journal of Andrology, 16(4), 631-638.
Pereira, C. D., Silva, M. B., Monteiro, R., & Martins, M. J. (2012a). Modulation of 11ß-HSD1 by chronic
ingestion of a hypersaline sodium-rich carbonated Portuguese natural mineral water in an animal
model of the Metabolic Syndrome. Revista Portuguesa de Cirurgia Suplemento novembro: 33.
Perez-Granados, A. M., Navas-Carretero, S., Schoppen, S., & Vaquero, M. P. (2010). Reduction in
cardiovascular risk by sodium-bicarbonated mineral water in moderately hypercholesterolemic
young adults. Journal of Nutritional Biochemistry, 21(10), 948-953.
Petraccia, L., Liberati, G., Masciullo, S. G., Grassi, M., & Fraioli, A. (2006). Water, mineral waters and
health. Clinical Nutrition, 25(3), 377-385.
Poikolainen, K., & Alho, H. (2008). Magnesium treatment in alcoholics: a randomized clinical trial.
Subst Abuse Treat Prev Policy 3: 1.
Polizio, A.H., Gonzales, S., Muñoz, M.C., Peña, C., & Tomaro, M.L. (2006). Behaviour of the anti-
oxidant defence system and heme oxygenase-1 protein expression in fructose-hypertensive
rats. Clinical and Experimental Pharmacology and Physiology, 33(8), 734-739.
Polushina, N. D. (1998). [The effect of mineral water on serotonin and insulin production (an experimental
study)]. Voprosy kurortologii, fizioterapii, i lechebnoi fizicheskoi kultury, Jul-Aug (4), 9-10.
Polushina, N. D., Vladimirov, V. I., & Frolkov, V. K. (2002). [Effect of drinking mineral water on
hormonal regulation of glucose homeostasis after mastectomy for breast cancer]. Voprosy
kurortologii, fizioterapii, i lechebnoi fizicheskoi kultury, Nov-Dec (6), 19-21.
Qiao, Q., Gao, W., Zhang, L., Nyamdorj, R., & Tuomilehto, J. (2007). Metabolic syndrome and
cardiovascular disease. Annals of Clinical Biochemistry, 44(Pt 3), 232-263.
146 Understanding the Metabolic Syndrome Using a Biomedical chemistry profile
Rayssiguier, Y., Gueux, E., Nowacki, W., Rock, E., & Mazur, A. (2006). High fructose consumption
combined with low dietary magnesium intake may increase the incidence of the metabolic
syndrome by inducing inflammation. Magnesium research, 19(4), 237-243.
Rayssiguier, Y., Libako, P., Nowacki, W., & Rock, E. (2010). Magnesium deficiency and metabolic syndrome:
stress and inflammation may reflect calcium activation. Magnesium research, 23(2), 73-80.
Rice, B. H., Cifelli, C. J., Pikosky, M. A., & Miller, G. D. (2011). Dairy components and risk factors for
cardiometabolic syndrome: recent evidence and opportunities for future research. Advances in
Nutrition, 2(5), 396-407.
Roberts, C. K., & Sindhu, K. K. (2009). Oxidative stress and metabolic syndrome. Life Sciences,
84(21-22), 705-712.
Rylander, R. (2008). Drinking water constituents and disease. Journal of Nutrition, 138(2), 423s-425s.
Rylander, R., & Arnaud, M. J. (2004). Mineral water intake reduces blood pressure among subjects
with low urinary magnesium and calcium levels. BMC Public Health, 4, 56.
Sabatier, M., Arnaud, M. J., Kastenmayer, P., Rytz, A., & Barclay, D. V. (2002). Meal effect on
magnesium bioavailability from mineral water in healthy women. The American Journal of
Clinical Nutrition, 75(1), 65-71.
Sabatier, M., Grandvuillemin, A., Kastenmayer, P., Aeschliman, J. M., Bouisset, F., Arnaud, M. J.,
Dumoulin, G., & A. Berthelot (2011). Influence of the consumption pattern of magnesium from
magnesium-rich mineral water on magnesium bioavailability. British Journal of Nutrition,
106(3), 331-334.
Santos, A., Martins, M. J., Guimaraes, J. T., Severo, M., & Azevedo, I. (2010). Sodium-rich carbonated
natural mineral water ingestion and blood pressure. Revista Portuguesa de Cardiologia, 29(2),
159-172.
Schoppen, S., Perez-Granados, A. M., Carbajal, A., Oubina, P., Sanchez-Muniz, F. J., Gomez-Gerique,
J. A., & Vaquero, M. P. (2004). A sodium-rich carbonated mineral water reduces cardiovascular
risk in postmenopausal women. Journal of Nutrition, 134(5), 1058-1063.
Schoppen, S., Perez-Granados, A. M., Carbajal, A., Sarria, B., Sanchez-Muniz, F. J., Gomez-Gerique,
J. A., & Pilar Vaquero, M. (2005). Sodium bicarbonated mineral water decreases postprandial
lipaemia in postmenopausal women compared to a low mineral water. British Journal of
Nutrition, 94(4), 582-587.
Schoppen, S., Sanchez-Muniz, F. J., Perez-Granados, M., Gomez-Gerique, J. A., Sarria, B., Navas-
Carretero, S., & Pilar Vaquero, M. (2007). Does bicarbonated mineral water rich in sodium
change insulin sensitivity of postmenopausal women? Nutricion Hospitalaria, 22(5), 538-544.
Schorr, U., Distler, A., & Sharma, A. M. (1996). Effect of sodium chloride- and sodium bicarbonate-
rich mineral water on blood pressure and metabolic parameters in elderly normotensive
individuals: a randomized double-blind crossover trial. Journal of Hypertension, 14(1), 131-135.
Simunic, S., & Pintac, L. (1990). [Use of the Varazdinske Toplice mineral water in the treatment of
diabetes mellitus]. Reumatizm, 37(1-6), 71-74.
Sun, K., Su, T., Li, M., Xu, B., Xu, M., Lu, J., Liu, J., Bi, Y. & Ning, G. (2014). Serum potassium level is
associated with metabolic syndrome: A population-based study. Clin Nutr 33(3): 521-527.
Takaya, J., Higashino, H., & Kobayashi, Y. (2004). Intracellular magnesium and insulin resistance.
Magnesium Research, 17(2), 126-136.
Takaya, J., Iharada, A., Okihana, H., & Kaneko, K. (2011a). Magnesium deficiency in pregnant
rats alters methylation of specific cytosines in the hepatic hydroxysteroid dehydrogenase-2
promoter of the offspring. Epigenetics, 6(5), 573-578.
Takaya, J., Iharada, A., Okihana, H., & Kaneko, K. (2011b). Upregulation of hepatic 11beta-
hydroxysteroid dehydrogenase-1 expression in calcium-deficient rats. Annals of Nutrition and
Metabolism, 59(2-4), 73-78.
Takaya, J., Iharada, A., Okihana, H., & Kaneko, K. (2012). Down-regulation of hepatic
phosphoenolpyruvate carboxykinase expression in magnesium-deficient rats. Magnesium
Research, 25(3), 131-139.
References 147
Teegarden, D. (2006). Dietary Calcium and the Metabolic Syndrome. Calcium in Human Health.
Nutrition and Health. C. Weaver and R. P. Heaney, Humana Press: 401-409.
Touyz, R. M. (2003). Role of magnesium in the pathogenesis of hypertension. Molecular Aspects of
Medicine, 24(1-3), 107-136.
Tremblay, A., & Gilbert, J. A. (2009). Milk products, insulin resistance syndrome and type 2 diabetes.
Journal of the American College of Nutrition, 28(S1), 91s-102s.
Tsatsoulis, A., Mantzaris, M. D., Bellou, S., & Andrikoula, M. (2013). Insulin resistance: an adaptive
mechanism becomes maladaptive in the current environment - an evolutionary perspective.
Metabolism, 62(5), 622-633.
Van Leer, E. M., Seidell, J. C., & Kromhout, D. (1995). Dietary calcium, potassium, magnesium and
blood pressure in the Netherlands. International Journal of Epidemiology, 24(6), 1117-1123.
van Meijl, L. E., Vrolix, R., & Mensink, R. P. (2008). Dairy product consumption and the metabolic
syndrome. Nutrition Research Reviews, 21(2), 148-157.
van Raalte, D. H., & Diamant, M. (2014). Steroid diabetes: from mechanism to treatment?
Netherlands Journal of Medicine, 72(2), 62-72.
Vaskonen, T. (2003). Dietary minerals and modification of cardiovascular risk factors. Journal of
Nutritional Biochemistry, 14(9), 492-506.
Venu, L., Kishore, Y. D., & Raghunath, M. (2005). Maternal and perinatal magnesium restriction
predisposes rat pups to insulin resistance and glucose intolerance. Journal of Nutrition, 135(6),
1353-1358.
Venu, L., Padmavathi, I. J., Kishore, Y. D., Bhanu, N. V., Rao, K. R., Sainath, P. B., Ganeshan, M., &
Raghunath, M. (2008). Long-term effects of maternal magnesium restriction on adiposity and
insulin resistance in rat pups. Obesity (Silver Spring), 16(6), 1270-1276.
Vitoria, I., Maraver, F., Ferreira-Pêgo, C., Armijo, F., Moreno Aznar, L., & Salas-Salvadó, J. (2014). The
calcium concentration of public drinking waters and bottled mineral waters in Spain and its
contribution to satisfying nutritional needs. Nutrición Hospitalaria, 30(1), 188-199.
Volpe, S. L. (2008). Magnesium, the metabolic syndrome, insulin resistance, and type 2 diabetes
mellitus. Critical Reviews in Food Science and Nutrition, 48(3), 293-300.
Wells, I. C. (2008). Evidence that the etiology of the syndrome containing type 2 diabetes
mellitus results from abnormal magnesium metabolism. Canadian Journal of Physiology and
Pharmacology, 86(1-2), 16-24.
Whelton, P. K., He, J., Cutler, J. A., Brancati, F. L., Appel, L. J., Follmann, D., & Klag, M. J. (1997).
Effects of oral potassium on blood pressure. Meta-analysis of randomized controlled clinical
trials. JAMA, 277(20), 1624-1632.
Xu, C., Bai, B., Fan, P., Cai, Y., Huang, B., Law, I. K., Liu, L., Xu, A., Tung, C., Li, X., Siu, F. M., Che, C.
M., Vanhoutte, P. M., & Wang, Y. (2013). Selective overexpression of human SIRT1 in adipose
tissue enhances energy homeostasis and prevents the deterioration of insulin sensitivity with
ageing in mice. American Journal of Translational Research, 5(4), 412-426.
Zemel, M. B., & Sun, X. (2008). Dietary calcium and dairy products modulate oxidative and
inflammatory stress in mice and humans. Journal of Nutrition, 138(6), 1047-1052.
Zhang, L., Curhan, G. C., & Forman, J. P. (2009). Diet-dependent net acid load and risk of incident
hypertension in United States women. Hypertension, 54(4), 751-755.
Ziomber, A., Machnik, A., Dahlmann, A., Dietsch, P., Beck, F. X., Wagner, H., Hilgers, K. F., Luft, F. C.,
Eckardt, K. U. & Titze, J. (2008). Sodium-, potassium-, chloride-, and bicarbonate-related effects
on blood pressure and electrolyte homeostasis in deoxycorticosterone acetate-treated rats.
American Journal of Physiology Renal Physiology, 295(6), F1752-1763.
Andrea Lobo & Teresa Summavielle*6
2.4 Brain Neurochemistry and Cognitive
Performance: Neurotransmitter Systems
Dopamine appears early in the embryonic CNS, and plays a key role in neural
development. Interfering with the normal interaction between dopamine and
dopamine receptors, during developmental periods is long known to alter the standard
embryonic development, disrupting neuronal proliferation, cell migration, neuronal
maturation, pruning and proper connectivity.
Dopaminergic neurons are concentrated in three major cell groups: the substantia
nigra (SN) the ventral tegmental area (VTA), and a cell group of dopamine-containing
cells in the caudal extension of the SN (Fallon, 1978; 1985a; Deutch, 1988). The SN and
VTA regions are strongly implicated in cognition, affective and behavioral disorders
(Roeper, 2013), which is partially explained by a relevant overlap with the brain
reward and limbic systems (Koob, 1988; 1992). The main efferent connections from the
SN-VTA are well described and can be summarized in three groups: the mesostriatal
or nigrostriatal projections, the mesolimbic projections and the mesocortical
projections (sometimes referred together as mesocorticolimbic). The mesostriatal
pathway originates largely from dopaminergic neurons in the ventral sheet of the
SNc projecting into the dorsal striatum (Fallon, 1985b; 1988; Voorn, 1988). In the
dorsal striatum, dopaminergic terminals synapse with GABAergic and cholinergic
neurons, resulting in either excitatory or inhibitory effects depending on the pattern
of dopaminergic receptors displayed at the synapse. Dopamine in the dorsal striatum
plays an important role in controlling patterns of motor activity, reward-based
learning and sequencing, which constitute a complete motor act and represent a
cognitive function (Calabresi, 2007; Leisman, 2014). The striatum receives major
glutamatergic inputs from the cortex and can control dopaminergic neurons in the
SN-VTA through striatal feedback projections (Voorn, 1988; Robbins, 1992; Calabresi,
2000). The mesolimbic pathway arises predominantly from the dorsal sheet of the
SNc, SNl and dorsal VTA (Fallon, 1995). These projections innervate, among other
areas, the amygdala, and the nucleus accumbens (Fallon, 1985b; 1988), which plays
a major role in the limbic-motor integration (Nicola, 2000), underlying complex
adaptive behaviors involved in brain stimulation reward and self-administration of
psychoactive drugs (Woodward, 1999; Schramm, 2002). In the nucleus accumbens,
DA seems to modulate limbic hippocampal and amygdalar inputs (Mogenson, 1988).
Dopamine in this brain region was also shown to be involved in stimulus conditioned
responsiveness (Bassareo, 2002) and locomotor activity. The mesocortical pathways
of dopaminergic projections arise mainly from the VTA and medial SN, densely
innervating the prefrontal, cingulated, suprarhinal and entorhinal cortices, and the
dorsal and ventral hippocampus and the visual cortex. Dopamine was described
to act in the prefrontal cortex (PFC), by modulating signals received from relevant
thalamic afferents (Thierry, 1988). Under stressful situations, dopamine metabolism
is increased in the PFC and, therefore, dopamine seems to be also relevant in plastic
responses to cope with the stressors (Sullivan, 1998; 2004). Dopamine is also involved
in the control of memory-guided behavior in the prefrontal cortex (Goldman-Rakic,
1992) and in cognitive tasks (Weinberger, 1988; Puig, 2014). It is noteworthy that these
pathways are under a complex net of direct stimulation and feedback regulation
within the dopaminergic system.
In the CNS, serotonin (5-hydroxytriptamine, 5-HT) is a key player in mood control.
Alterations in 5-HT function have been implicated in several clinical states, including
affective disorders, obsessive-compulsive disorders, schizophrenia, anxiety states,
150 Brain Neurochemistry and Cognitive Performance: Neurotransmitter Systems
phobic disorders, eating disorders, migraine, and sleep disorders (Stockmeier, 2003;
Albert, 2014). There is also evidence that 5-HT is involved in learning and memory
processes (Meneses, 1999; Faulkner, 2014).
The majority of serotonergic neurons are located in the midline raphe nuclei
(Tork, 1990). Serotonergic neurons in the raphe nuclei have extensive projections to
virtually all areas of the brain and spinal cord (Tork, 1990), but the brain regions
usually implicated in cognition and mood receive 5-HT projections mainly from the
dorsal raphe nucleus (DRN) and median raphe nucleus (Halliday, 1995). The majority
of the ascending 5-HT fibers originate at MRN or DRN, while a caudal cluster gives rise
to almost all the descending 5-HT fibers (Steinbusch, 1981; Halliday, 1995). The early
development of the 5-HT cell bodies and their extensive net of connections through all
the brain, are indicative of the important role that serotonin plays in the development
and functioning of the CNS. Diencephalic structures, the basal forebrain, septal
regions, hippocampal formation and the cerebellum all receive ascending projections
from the DRN and MRN (Halliday, 1995). These projections are essentially non-
overlapping and distribute to separate sites. Fibers from the DRN primarily project
to the SNc, VTA, amygdala, striatum, lateral preoptic area, substantia innominata,
nucleus accumbens and several regions of the cortex, while MRN fibers distribute
mainly to midline and para-midline structures (Vertes, 1997; 1999). Direct synaptic
connections between raphe serotonergic terminals and DA neurons were first
demonstrated in the VTA (Herve, 1987), since then, a close relationship between the
serotonergic and the dopaminergic systems has been well established.
Synthesis of dopamine in DA-neurons starts with the metabolization of L-tyrosine
into 3,4-dihydroxiphenylalanine (L-DOPA) by TH (Nagatsu, 1964; 1998), followed
by almost immediate conversion of L-DOPA into dopamine by L-aromatic amino
acid decarboxilase (L-AADC) (Deutch, 1999). Dopamine is stored into vesicles in
the presynaptic terminal of dopaminergic neurons, which not only prevents DA
from degrading, but also delays its diffusion to the extracellular space. Dopamine
is released into the synapse by calcium dependent exocytotic mechanisms. Once in
the synaptic cleft, DA can interact with presynaptic autoreceptors (D2-like receptors
that include D2, D3 and D4), regulating DA synthesis, release and neuronal firing-
rate; or with postsynaptic receptors (D1-like, that include D1 and D5), modulating the
response of the postsynaptic neuron. The D1 receptor is the most widespread receptor,
the most expressed and exclusively postsynaptic (Vallone, 2000). The D2 receptor
is found mainly expressed by GABAergic neurons (Civelli, 1991). Importantly, these
receptors are known to form heteromers, which can have relevant functions in the
dopaminergic brain (Hasbi, 2011; Perreault, 2014). The dopamine transporter protein
(DAT) seems to be targeted by a complex net of regulation mechanisms, which depend
primarily on the extracellular levels of DA. This transporter is a major target for
psychostimulant drugs, such as cocaine and amphetamines, which can easily block it
or use it to enter the dopaminergic terminal, leading to increased extracellular levels
of DA and increased oxidative stress (Reith, 1997; Chen, 2000). DAT increased levels
Monoaminergic Neurotransmission and Cognition 151
are potentially dangerous to the neuron (due also to its ability to transport toxins). A
positive correlation between the levels of DAT expression and neuroinjury has been
demonstrated (Miller, 1999). Storing of DA depends on a vesicular transporter protein,
the vesicular monoamine transporter (VMAT). If left free on the cytoplasm, DA is
rapidly inactivated by the mitochondrial enzyme MAO-A, simultaneously producing
H2O2.
Serotonin is synthesized from the amino acid tryptophan that is hydroxylated into
5-hydroxytryptophan (5-HTP) by the rate-limiting enzyme tryptophan hydroxylase
(TPH). The serotonin precursor, 5-HTP, is decarboxylated by the L-AADC originating
the neurotransmitter 5-HT. Synthesis of serotonin depends mainly on the availability of
tryptophan, which is present in high levels in the plasma and crosses the blood-brain
barrier by active transport. Dietary levels of tryptophan may affect substantially the
levels of serotonin in the brain. In the presynaptic terminal, 5-HT is stored in vesicles
by a process identical to the one described for dopamine. Interaction of extracellular
5-HT with 5-HT autoreceptors regulates synthesis and release of 5-HT. Inactivation
of released 5-HT is mainly attained by 5-HT reuptake through the membrane carrier-
protein 5-HT transporter (SERT). Once inside the neuron, 5-HT can be reacumulated
into vesicles, or inactivated by the mitochondrial enzyme MAO-B, that metabolizes
5-HT into 5-hydroxyindoleacetic acid (5-HIAA) simultaneously producing H2O2.
Inhibition of 5-HT neuron firing activity is mediated through the somatodendritic
5-HT1A autoreceptor, increasing the membrane hyperpolarization (Pineyro, 1999).
Serotonin receptors represent the most complex family of neurotransmitter receptors.
With the exception of the 5-HT3 receptor, which is a ligand-gated ion channel, all
5-HT receptors belong to the G-protein-coupled receptor superfamily (Hoyer, 2002).
Due to its role in antidepressant drug action, 5-HT receptors have been the target of
intense research (Pineyro, 1999). These receptors have been divided, according to
their primary function, in three main groups. The first group comprises the 5-HT1-
like receptors (5-HT1A, 5-HT1B and 5-HT1D), which couple preferentially to Gi/o proteins
(pertussis toxin-sensitive) to inhibit cAMP formation (Hoyer, 2002). In the raphe nuclei,
5-HT1A receptors are somatodendritic and act like autoreceptors inhibiting cell firing,
while postsynaptic 5-HT1A receptors, are present in a number of limbic structures,
particularly in the hippocampus (Hoyer, 2002). The 5-HT1A receptor was associated
with modulation of anxiety-related behaviors (Heisler, 1998) and also with locomotor
activity (Carey, 2001). The 5-HT1B receptors are expressed mainly in the basal ganglia,
striatum, nucleus accumbens and frontal cortex, and are thought to act as terminal
autoreceptors; in addition, they can also serve as terminal heteroreceptors, controlling
the release of other neurotransmitters (Hoyer, 2002). The 5-HT2-like receptor class
comprises the 5-HT2A, 5-HT2B and 5-HT2C (formerly 5-HT1C) receptors, which couple
preferentially to G proteins that employ protein kinase C, (Blaukat, 2000) to increase
the hydrolysis of inositol phosphates and increase cytosolic calcium (Hoyer, 2002). In
the CNS, the 5-HT2A receptors are mainly located in the cortex, claustrum and basal
ganglia, and are particularly important to the behavioral effects of drugs of abuse (Filip,
152 Brain Neurochemistry and Cognitive Performance: Neurotransmitter Systems
2001; Munzar, 2002; Porras, 2002). The 5-HT2B and 5-HT2C receptors are restricted to a
few brain regions and were reported to be implicated in mediating hypo/hyperphagia,
grooming frequency and anxiety behavior (Hoyer, 2002). The 5-HT3 receptors are
intrinsic ligand-gated channels, which trigger rapid depolarization due to transient
inward current. The 5-HT7 receptor, in accordance with its distribution in the limbic
system and thalamocortical regions, plays a relevant role in the pathophysiology of
affective and mood disorders (Vanhoenacker, 2000; Hoyer, 2002).
Accumulating evidence shows that optimal brain function relies on a complex
equilibrium of neurotransmitters and respective receptors in well-organized
circuitries. For most of these players, an inverted “U”-shaped curve is the best fit
to describe the relation between either sub-optimal or supra-optimal levels and a
specific behavioral performance (Cools, 2011; Yadid, 1998). For DA this seems to be
particularly relevant. Balanced DA transmission is essential for correct modulation
of the cognitive process. Deregulation of the DA function at the frontocortical level
will impact on decision-making, judgment capacity, planning, working-memory,
attention, cognitive flexibility, and impulsivity (Brozoski, 1979; Cai, 1997; Fischer,
2010). As such, the dopaminergic system is often targeted by pharmacologic
approaches meant to improve the cognitive function in several diseases (but also in
healthy individuals). A common example is the prescription of methylphenidate, a
DAT blocker, to attention-deficit hyperactivity disorder (ADHD) patients (Agay, 2010;
Gamo, 2010). Likewise, most typical and atypical antipsychotic drugs display a
significant affinity to dopaminergic (and serotonergic) receptors. Deficient decision-
making, judgment capacity and impulse control are likely to result in increased risk-
tanking and facilitate drug experimentation, binge intoxication and development of
addictive behaviors (either substance or non-substance dependent).
Addiction to drugs refers to the loss of control over drug intake or the compulsivity
to seek and consume drugs despite adverse effects. Addiction only occurs through
repeated exposure and implies adaptations of several pathways at the molecular
and cellular levels (Nestler, 2001). Behavioral adaptations in addiction result from
the dysfunctional interaction of three key neural systems: an hyperactive/impulsive
amygdala-striatum loop that promotes automatic and repetitive behaviors; an
hypoactive/reflective PFC, impairing decision-making; and the insula-dependent
decision-making processes related to uncertain risk/reward, which potentiates
impulsivity and weakens the cognitive function (Evans, 2008; Noel, 2013). In addicted
persons, the uncontrollable urge to obtain drugs and relapse is associated with a
pathological balance between excitatory and inhibitory neurotransmission circuits,
leading to persistent behavioral abnormalities (Winder, 2002).
Changes in DA balance are also implicated in the physiopathology of other
psychiatric conditions, such as schizophrenia, and anxiety and mood disorders
(Willner, 2005; Chaudhury, 2013; Tye, 2013). In schizophrenia, manifestation of
the positive symptoms (hallucinations, delusions and psychosis) is a consequence
of an hiperactive mesolimbic DA system, while hipofrontality is linked to reduced
Monoaminergic Neurotransmission and Cognition 153
Glutamate is a ubiquitous anionic amino acid that exists in all cell types, but
in the brain it acts as a signaling molecule, being stored and released from the
glutamatergic neurons subpopulation. Glutamate is considered the main excitatory
neurotransmitter in the CNS, used by around half of the neurons in the brain
(Fonnum, 1984; Liguz-Lecznar, 2007). Glutamate is derived from glutamine through
enzymatic conversion involving phosphate-activated glutaminase (PAG) (Albrecht,
2007), and is stored in small synaptic vesicles at nerve terminals by the action of
vesicular glutamate transporters 1 and 2 (VGLUT1 and 2). Following membrane
depolarization and Ca2+ entry into cells, synaptic vesicles fuse with the plasma
membrane and release the glutamate by exocytosis into the synaptic cleft. After
being released, glutamate is transported back to the neuron or into the glial cells by
the action of excitatory amino acid transporters (EAATs). In astrocytes, glutamate is
converted to glutamine by glutamine synthase, and glutamine is transferred back
to neurons, probably through the sequential action of amino acid system N and A
transporters (Albrecht, 2007; Liguz-Lecznar, 2007; Lee, 2010). Once released from the
presynaptic nerve terminal, glutamate binds to specific receptors in the postsynaptic
membrane to conduct excitatory transmission. Pre-synaptic glutamate receptors act
in the modulation of glutamate release. The effects of glutamate are mediated by
activation of ionotropic or metabotropic receptors, which differ in their molecular,
biochemical, physiological and pharmacological properties (Kew, 2005; Kim,
2001). The ionotropic glutamate receptors have been classified into three distinct
subgroups, α.amino-3-hydroxy-5-methyl-4-isoxazolepropionic (AMPA), N-methyl-D-
aspartate (NMDA) and Kainate (KA) receptors (Dingledine, 1999; Mayer, 2004). AMPA
and kainate receptors are responsible for most of the fast excitatory transmission
in the vertebrate CNS. They are voltage-independent ion channels and permeable
to Na+ and K+, leading to a net depolarizing influx of cations upon activation by
glutamate (Swanson, 1997). AMPA receptors are composed of four possible subunits,
GluA1-4, which associate in different stoichiometries to form receptors with distinct
properties (Greger, 2007). NMDA receptors are ligand-gated ion channels that exhibit
strong voltage dependence owing to the blocking of the receptor channels at negative
membrane potentials by extracellular magnesium. As a result, these receptors
contribute little to the postsynaptic response during low-frequency synaptic activity.
However, when the cell is depolarized, Mg2+ dissociates from its binding site within
the NMDAR channel, allowing Ca2+ and Na+ to enter the dendritic spine (Cull-Candy,
2001). Functional NMDA receptors are heterotrimeric complexes containing both
GluN1 and GluN2 subunits (Prybylowski, 2004). Metabotropic glutamate receptors
are coupled to G proteins (which in turn stimulate second messenger signaling
pathways), and, as such, they mediate slower synaptic responses, occurring over
seconds and minutes, rather than milliseconds as occurs for ionotropic glutamate
receptors. There are three groups of mGluR, distinguished based on sequence
Glutamate Neurotransmission and Cognition 155
homology, signal transduction mechanisms and agonist selectivity (Pin, 2002; Kim,
2008; Niswender, 2010).
Glutamatergic neurons play crucial roles in physiological mechanisms, such
as synaptic plasticity mechanism like long-term potentiation (LTP) and long-
term depression (LTD) underlying processes of learning and memory. However,
disturbances in glutamatergic signaling contribute to the pathogenesis of several
neurological disorders, such as ischemic stroke, schizophrenia, epilepsy and
neurodegenerative disorders (Dong, 2009; Szydlowska, 2010). Neuronal excitotoxicity
refers to the injury of neurons resulting from a prolonged exposure to glutamate and
consequent overactivation of both ionotropic and metabotropic glutamate receptors
(Choi, 1988; Kroemer, 2009). The associated sustained influx of ions into neurons,
and particularly, calcium overload through NMDA receptors and calcium release
from intracellular stores triggered by mGluRs is highly neurotoxic, leading to the
activation of enzymes that will degrade membranes, proteins, and nucleic acids,
which ultimately lead to cell death (Friedman, 2006; Dong, 2009).
Stroke is a complex and devastating disease, constituting the second leading
cause of death worldwide and the leading cause of acquired disability in adults
(Brouns, 2009; Lloyd-Jones, 2009; Moskowitz, 2010). Stroke can be subdivided into
ischemic and hemorrhagic (Doyle, 2008). Ischemic strokes are more frequent and
are caused by a thrombosis or an embolic occlusion of a cerebral blood vessel, more
frequently the middle and anterior cerebral arteries, or a general hypo-perfusion,
all of which result in a constraint of blood flow to the brain, reducing the delivery
of substrates, particularly oxygen and glucose, and ATP production (Dirnagl, 1999;
Doyle, 2008; Lloyd-Jones, 2009). Because of its high metabolic activity, together with
large concentrations of glutamate (Choi, 1992), the brain is particularly vulnerable to
ischemic insults. The ischemic core is the irreversibly damaged tissue characterized
by less than 20% of normal blood flow levels, reduced ATP levels and irreversible
energetic failure (Lo, 2008b). Cells in the core are killed rapidly by proteolysis, lypolysis,
bioenergetic failure and collapse of ion homeostasis (Doyle, 2008; Brouns, 2009).
In the peripheral areas of stroke, between the normal brain and the damaged core,
lies the ischemic penumbra or peri-infarct zone (Astrup, 1981). In this region, blood
flow deficits are less severe, hardly sufficient to support basal ATP levels and normal
ionic gradients (Moskowitz, 2010). Therefore, in the penumbra region the tissue is
functionally impaired but potentially salvageable, and can be rescued by enhancing
blood flow or interfering with the ischemic cascade (Lo, 2008b; 2008a; Brouns, 2009).
With time, the infarct core expands into the ischemic penumbra, so, accurate detection
of this tissue at risk can help to identify patients who might benefit from treatment
with recombinant tissue plasminogen activator (rtPA), the only approved drug to
treat acute stroke patients, by promoting clot lysis and reperfusion, therefore restore
blood flow at an early time point (NINDS, 1995; Paciaroni, 2009). The decreased ATP
production leads to the dysfunction of energy-dependent ion transport pumps, and to
the consequent depolarization of neurons and glia (Katsura, 1994; Martin, 1994). The
156 Brain Neurochemistry and Cognitive Performance: Neurotransmitter Systems
Na+/K+ ATPase at the plasma membrane of neurons maintains high K+ and low Na+
intracellular concentrations, that are essential for the propagation of action potentials.
After ischemia, the inhibition of ATP synthesis by mitochondria leads to a rapid ATP
consumption, which causes a neuronal membrane depolarization with the release
of K+ and Na+ entry into cells (Caplan, 2009). Energy failure also impedes the plasma
membrane Ca2+ ATPase to maintain the very low calcium concentrations usually
found within the cells (Doyle, 2008). This, together with the activation of voltage-
dependent calcium channels, leads to the release of neurotransmitters, especially
glutamate, which plays a critical role in the ischemic damage (Nicholls, 1990; Brouns,
2009). A large concentration gradient of glutamate is maintained across the plasma
membrane by sodium-dependent glutamate transporters located at the plasma
membrane, which maintain a high cytosolic glutamate concentration (approximately
10 mM) when compared to the synaptic concentration (in the micromolar range) (Hsu,
1998). Membrane depolarization during ischemia also induces a reversal of glutamate
uptake carriers and enables glutamate to exit the cells along its concentration gradient,
further increasing its accumulation in the extracellular space (Nicholls, 1990; Rossi,
2000). The extracellular accumulation of glutamate leads to excitotoxic stimulation
of synaptic and extra-synaptic ionotropic and metabotropic glutamate receptors,
which will result in neuronal dysfunction and death (Choi, 1988; Sims, 1995; Kroemer,
2009). In particular, the stimulation of AMPA and NMDA receptors causes an influx
of Na+, and NMDA receptors are also characterized by a high Ca2+ permeability and
conductance properties (Dong, 2009). Activation of group I mGluR receptors, which
includes mGluR1 and mGluR5, increases the intracellular inositol trisphospate (IP3),
thereby activating protein kinase C and releasing Ca2+ from neuronal stores (Pin, 2002;
Friedman, 2006). Therefore, overactivation of NMDA receptors and mGluR contribute
to a [Ca2+]i overload. AMPA receptors are not normally calcium permeable due to the
GluA2 subunit, but the downregulation of this subunit after ischemia increases the
calcium permeability of these receptors, and thus allows AMPA receptors to contribute
to delayed cell death (Liu, 2006; Peng, 2006).
Glutamatergic signaling and excitotoxic mechanisms play a role in chronic
neurodegenerative disorders. Alzheimer’s Disease is characterized by 3
neuropathological hallmarks: deposits of amyloid beta (Aβ) peptides that form
senile plaques, neurofibrillary tangles of hyperphosphorylated tau and neuronal
loss (Parameshwaran, 2008). Several studies have demonstrated that excitatory
synaptic transmission and plasticity are impaired in the hippocampus of AD, and
particularly the glutamatergic system is altered leading to synaptic dysfunction and
neurodegeneration (Parameshwaran, 2008). Several studies demonstrated that levels
of VGLUTs are decreased, particularly VGLUT1, in AD patient’s cortices, and that Aβ
peptides accumulate preferentially in glutamatergic neurons, expressing VGLUT1/2
(Sokolow, 2012). Also, EAAT levels are reduced, indicating that glutamate reuptake
from the synaptic cleft is compromised (Scott, 2011). Furthermore, glutamate synthase
levels are reduced, increasing glutamate levels in astrocytes, which can be released
Glutamate Neurotransmission and Cognition 157
to the synaptic cleft (Robinson, 2001). These effects may increase glutamate levels at
the synapse, triggering excitotoxic mechanisms that will contribute to cell death in
AD (Dong, 2009). It has been hypothesized that changes in AMPA receptors (AMPARs)
number and function play an important role in AD pathogenesis. Various studies
reported a downregulation of AMPARs by Aβ peptides at initial stages of the disease,
probably related to a synaptic failure underlying initial cognitive impairment, prior
to neuronal loss (Selkoe, 2002). This reduction of AMPARs levels contributes to loss
of synaptic function at initial stages of the disease, and is related to caspase cleavage
or endocytosis of the receptors triggered by Aβ peptides (Chan, 1999; Chang, 2006).
Moreover, Aβ affects proteins related to AMPARs insertion and stabilization at the
plasma membrane, such as post-synaptic density protein 95 (PSD-95) in a NMDA
receptors (NMDARs) dependent mechanism (Roselli, 2005). Excitotoxic neuronal
death triggered by excessive glutamate at the synapse and sustained Ca2+ influx
through NMDARs is believed to be one of the major causes of neurodegeneration in AD
(Harkany, 2000). Abnormal NMDARs upregulation contribute to elevated production
of Aβ, resulting in increased glutamate levels and activation of NMDARs as disease
progresses, indicating that NMDARs play a key role in Aβ induced neurotoxicity
(Miguel-Hidalgo, 2002). It is also believed that a functional downregulation of NMDARs
at initial stages of the disease is related to a compromised glutamatergic function. Aβ
peptides reduce surface NMDARs levels at the synapse through endocytosis, which
depresses the glutamatergic function (Snyder, 2005). Moreover, Aβ facilitates LTD in a
NMDA-dependent mechanism, indicating that in AD not only receptor levels, but also
their function, is impaired and may be linked to reduced glutamatergic transmission
associated with cognitive deficits (Kim, 2001).
Parkinson’s disease (PD) main pathological hallmarks are the loss of
dopaminergic neurons in the substantia nigra pars compacta (SNc) projecting to the
corpus striatum, and the presence of cytoplasmic inclusions of protein aggregates,
composed mainly of α-synuclein, and other proteins, called Lewi bodies (Betarbet,
2002). Several studies demonstrated that dopaminergic neurodegeneration in SNc
is mainly associated with mitochondrial dysfunction and oxidative stress. However,
it is widely accepted that glutamate excitotoxicity also plays an important role
in PD pathology (Blandini, 2010). The dopaminergic denervation of the striatum
associated with SNc neurodegeneration triggers changes in the basal ganglia
circuitry which are associated with the motor neurological symptoms of the disease,
namely tremors and rigidity (Blandini, 2000). Within the basal ganglia circuitry,
glutamate neurotransmission occurs in the projections from cortical areas to the
striatum and to the subthalamic nucleus (STN), and from the STN to the substantia
nigra, therefore playing an important role in the mechanisms related to PD motor
symptoms (Blandini, 2000). Also, glutamate excitotoxic effects may be relevant to
the neurodegenerative processes occurring in PD. Dopaminergic neurons from SNc
are particularly vulnerable to excitotoxicity triggered by a bioenergetic failure related
to an impairment of mitochondrial function (Erecinska, 1990). Also, the presence
158 Brain Neurochemistry and Cognitive Performance: Neurotransmitter Systems
high density of DA neurons that project to the striatum (Trevitt, 2002). DA regulates
GABAergic signaling in the SN and to the thalamus (Aceves, 1995; Timmerman,
1997). METH increases glutamate in the striatum through a polysynaptic pathway,
characterized by an increase in striatonigral GABA transmission, which will decrease
nigrothalamic GABAergic signaling, disinhibiting thalamocortical glutamatergic
transmission, ultimately causing an increase in glutamatergic release in striatum via
the corticostriatal pathway. The increase of glutamate in the striatum will contribute
to the degeneration of DA terminals, leading to a long-term depletion of DA in this
brain region (Mark, 2004). It was also shown that regulation of VGLUT1 expression
and function via this polysynaptic pathway facilitates vesicular accumulation and
glutamate release in the striatum after METH administration, contributing to a
sustained increase in glutamatergic transmission in the corticostriatal pathway (Mark,
2007). The increase in glutamate release might lead to an overstimulation of NMDARs
and consequent oxidative stress contributing to neuronal damage (Gunasekar, 1995).
Recent evidence also shows that glutamatergic inputs, especially those contacting
GABAergic PV-positive interneurons, may be developmentally regulated by repeated
exposure to cannabinoids, which can result in disruption of the glutamatergic
facilitation of PV-positive interneuron function and underlie the cognitive impairments
seen in young adults that chronically abuse these drugs. Overstimulation of the CB1
receptor over the adolescent developmental period seems to alter frontal circuits
leading to a schizophrenic-like disorder (Caballero, 2012).
The amino acid γ-aminobutyric (GABA) is the main inhibitory neurotransmitter in the
CNS. This neurotransmitter is synthesized by the enzyme glutamic acid descarboxylase
(GAD) which catalyzes the decarboxylation of glutamate. Although the expression
of GABA in the nervous system was first described in 1950 by Eugene Roberts and
Jorge Awapara, it was only accepted as a neurotransmitter more than 10 years later
(Roberts, 1950; Del Castillo, 1964). The difficulty in the identification of GABA as a
neurotransmitter came from its enormous abundance in the vertebrate brain (which
is about 1000 fold higher than monoamine transmitters), its simple structure and
its role in the Krebs cycle, suggesting that it was likely more involved in metabolism
than in intercellular signaling (Schuske, 2004). Cellular release of GABA may be
mediated by several different mechanisms (Saransaari, 1992): (1) GABA can be released
from neurons by exocytosis through synaptic vesicles, which is the most common
mechanism of GABA release under physiological conditions; (2) it may simply leak
through plasma membranes; (3) the plasma membrane GABA transporters-GAT may
be reversed (due to changes in the electrochemical gradients); and (4) finally, ion
channels in the membranes may also mediate GABA release despite the size of this
molecule. The release of GABA in extra-synaptic sites activates non-synaptic GABAA
GABAergic Neurotransmission and Cognition 161
Research in cerebral ischemia and excitotoxic neuronal damage has been mainly
focused on the excitatory mediators and much less attention was given to the changes
in GABAergic activity (Schwartz-Bloom, 2001). The release of GABA in the ischemic
brain and the consequent activation of GABAA receptors may be neuroprotective
through reduction of membrane depolarization. Although Cl- entry through GABAA
receptors in association with overactivation of glutamate receptors may further
increase the influx of water and cell swelling, strategies to increase GABAergic
neurotransmission, targeting both sides of the synapse, have been quite efficient in
models of ischemia (Schwartz-Bloom, 2001). The impairment of GABAergic synaptic
transmission in brain ischemia is partly due to a down-regulation of synaptic GABAA
receptors, which may contribute to the ongoing neuronal excitability and possibly to
neuronal death (Schwartz-Bloom, 2001). Furthermore, exposure of hippocampal slices
to oxygen- and glucose-deprivation was shown to induce an early release of GABA
by exocytosis, followed by a delayed phase of neurotransmitter release mediated by
reversal of the plasma membrane transporter (Allen, 2004).
The balance between excitation and inhibition is an important mechanism
in epilepsy, with inhibitory GABAergic regulation considered the main disorder
in epilepsy (Schindler, 2008). Evidence shows that during the onset of the status
epilepticus endocytosis of GABAA receptors takes place in the hippocampus,
decreasing post-synaptic inhibitory currents (Naylor, 2005a; 2005b). Glutamate and
GABA transporters also affect the excitatory/inhibitory balance. Glutamate transporter
activity could trigger GAT reversal activity in astrocytes, by elevating intracellular Na+
levels. This glutamate/GABA exchange may act as a negative feedback to decrease
excitation during seizures (Gadea, 2001; Heja, 2012; Wu, 2014). Neuronal network
remodeling is also common in epilepsy, including loss of GABAergic interneurons and
axon growth. The formation of new excitatory circuits may increase seizure intensity
(Wu, 2014). Excitatory/inhibitory balance determination shows that inhibition
superimposes on pyramidal neurons during initiation of seizure events, shifts
towards excitation during seizures propagation, and eventually inhibition dominates
again during seizures termination (Murase, 2014). The mechanism underlying such
homeostatic changes involves increased glutamate signaling and alterations in
GABAergic transmission, related to suppressed presynaptic inhibition and GABAergic
vesicular depletion (Trevelyan, 2013).
Several evidences show that GABA plays an important role in the cognitive features
of schizophrenia (Gonzalez-Burgos, 2011). Most domains of GABAergic signaling are
altered in schizophrenic patients, such as in the expression of GAD 65 and 67, and GAT1
(Pierri, 1999; Hashimoto, 2008). Alterations in GABAA receptors were also described
in schizophrenic patients (Fatemi, 2013), depending on the targeted brain region,
GABAergic transmission can be either increased or decreased (Deidda, 2014). The
imbalance of excitatory/inhibitory stimulus has also been linked to schizophrenia-
related GABAergic dysfunction (Lewis, 2005). Hypofunction of NMDARs and increased
D2 activity leads to hypostimulation of GABAergic neurons (Lewis, 2005). GABAergic
GABAergic Neurotransmission and Cognition 163
effects that scale from social disinhibition to impaired motor control and decision-
making, followed by mild to severe ataxia and hippocampal dysfunction, and
eventually ending in sedation, coma, cardiac arrest and possible death (Tabakoff, 1996;
2013; Gilpin, 2008; Koob, 2014). Chronic exposure to alcohol leads to compensatory
adaptations in the reward circuitry, and to the development of alcohol related
behaviors, such as tolerance, meaning that over time one will need a higher alcohol
dose to obtain the same reward effect. When alcohol consumption is discontinued,
several withdrawal effects occur, such as tremors, seizures, insomnia, and confusion.
These effects may result from the hyperactive adaptations suffered under prolonged
use that are no longer balanced by the alcohol inhibitory effects. The consequent
increase in glutamatergic activity may lead to toxicity and cell death (Becker, 2014).
Under chronic exposure to psychostimulants the GABAergic component of the reward
system is also deeply affected, reduced GABAergic inhibition of the dopaminergic
mesocorticolimbic pathway contributes to addiction by disrupting the frontal cortical
circuits that regulate motivation, drive, and self-control, increasing the motivational
salience of drug-associated stimuli (Volkow, 2002)
2.4.4 Gliotransmitters
References
Abi-Dargham, A., O. Mawlawi, et al. (2002). Prefrontal dopamine D1 receptors and working memory
in schizophrenia. The Journal of Neuroscience, 22(9), 3708-3719.
Aceves, J., B. Floran, et al. (1995). D-1 receptor mediated modulation of the release of gamma-
aminobutyric acid by endogenous dopamine in the basal ganglia of the rat. Progress in Neuro-
psychopharmacology & Biological Psychiatry, 19(5), 727-739.
Agay, N., E. Yechiam, et al. (2010). Non-specific effects of methylphenidate (Ritalin) on cognitive
ability and decision-making of ADHD and healthy adults. Psychopharmacology, 210(4): 511-519.
Albert, P. R., F. Vahid-Ansari, et al. (2014). Serotonin-prefrontal cortical circuitry in anxiety and
depression phenotypes: pivotal role of pre- and post-synaptic 5-HT1A receptor expression.
Frontiers in Behavioral Neuroscience, 8, 199.
Albrecht, J., U. Sonnewald, et al. (2007). Glutamine in the central nervous system: function and
dysfunction. Frontiers in Bioscience, 12, 332-343.
Allen, N. J., D. J. Rossi, et al. (2004). Sequential release of GABA by exocytosis and reversed uptake
leads to neuronal swelling in simulated ischemia of hippocampal slices. The Journal of
Neuroscience, 24(15), 3837-3849.
Ampe, B., A. Massie, et al. (2007). “NMDA-mediated release of glutamate and GABA in the
subthalamic nucleus is mediated by dopamine: an in vivo microdialysis study in rats.” Journal
of neurochemistry 103(3): 1063-1074.
Araque, A., G. Carmignoto, et al. (2014). Gliotransmitters travel in time and space. Neuron 81(4):
728-739.
Astrup, J., B. K. Siesjo, et al. (1981). Thresholds in cerebral ischemia - the ischemic penumbra.
Stroke, 12(6), 723-725.
Baker, D. A., K. McFarland, et al. (2003). Neuroadaptations in cystine-glutamate exchange underlie
cocaine relapse. Nature Neuroscience, 6(7), 743-749.
Balsters, J. H., R. G. O’Connell, et al. (2011). Donepezil impairs memory in healthy older subjects:
behavioural, EEG and simultaneous EEG/fMRI biomarkers. PLoS One, 6(9), e24126.
Bassareo, V., M. A. De Luca, et al. (2002). Differential expression of motivational stimulus properties
by dopamine in nucleus accumbens shell versus core and prefrontal cortex. The Journal of
Neuroscience, 22(11), 4709-4719.
Bauer, D., D. Gupta, et al. (2008). Abnormal expression of glutamate transporter and transporter
interacting molecules in prefrontal cortex in elderly patients with schizophrenia. Schizophrenia
Research, 104(1-3), 108-120.
Becker, H. C. and P. J. Mulholland (2014). Neurochemical mechanisms of alcohol withdrawal.
Handbook of Clinical Neurology, 125, 133-156.
Ben-Ari, Y. (2002). Excitatory actions of gaba during development: the nature of the nurture. Nature
reviews. Neuroscience, 3(9), 728-739.
Betarbet, R., T. B. Sherer, et al. (2002). Mechanistic approaches to Parkinson’s disease
pathogenesis. Brain Pathology, 12(4) 499-510.
Bettler, B. and J. Y. Tiao (2006). Molecular diversity, trafficking and subcellular localization of GABAB
receptors. Pharmacology & Therapeutics, 110(3), 533-543.
Blandini, F. (2010). An update on the potential role of excitotoxicity in the pathogenesis of
Parkinson’s disease. Functional Neurology, 25(2), 65-71.
Blandini, F., G. Nappi, et al. (2000). Functional changes of the basal ganglia circuitry in Parkinson’s
disease. Progress in Neurobiology, 62(1), 63-88.
Blaukat, A., A. Barac, et al. (2000). G protein-coupled receptor-mediated mitogen-activated protein
kinase activation through cooperation of Gαq and Gαi signals. Molecular and Cellular Biology,
20(18), 6837-6848.
References 167
Bohnen, N. I., D. I. Kaufer, et al. (2007). Cortical cholinergic denervation is associated with
depressive symptoms in Parkinson’s disease and parkinsonian dementia. Journal of Neurology
and Neurosurgery Psychiatry, 78(6), 641-643.
Brouns, R. and P. P. De Deyn (2009). The complexity of neurobiological processes in acute ischemic
stroke. Clinical Neurology Neurosurgery, 111(6), 483-495.
Brozoski, T. J., R. M. Brown, et al. (1979). Cognitive deficit caused by regional depletion of dopamine
in prefrontal cortex of rhesus monkey. Science, 205(4409), 929-932.
Caballero, A. and K. Y. Tseng (2012). Association of Cannabis Use during Adolescence, Prefrontal CB1
Receptor Signaling, and Schizophrenia. Frontiers in Pharmacology, 3, 101.
Cai, J. X. and A. F. Arnsten (1997). Dose-dependent effects of the dopamine D1 receptor agonists
A77636 or SKF81297 on spatial working memory in aged monkeys. Journal of Pharmacology
and Experimental Therapeutics, 283(1), 183-189.
Calabresi, P., D. Centonze, et al. (2000). Synaptic transmission in the striatum: from plasticity to
neurodegeneration. Progress in Neurobiology, 61(3), 231-265.
Calabresi, P., B. Picconi, et al. (2007). Dopamine-mediated regulation of corticostriatal synaptic
plasticity. Trends in Neurosciences, 30(5), 211-219.
Campbell, V., N. Berrow, et al. (1993). GABAB receptor modulation of Ca2+ currents in rat sensory
neurones by the G protein G(0): antisense oligonucleotide studies. The Journal of Physiology,
470, 1-11.
Caplan, L. R. (2009). Caplan’s Stroke. A Clinical Approach, Butterworth Heinemann.
Cardinal, R. N., J. A. Parkinson, et al. (2002). Emotion and motivation: the role of the amygdala,
ventral striatum, and prefrontal cortex. Neuroscience and Biobehavioral reviews, 26(3), 321-
352.
Carey, R. J., G. DePalma, et al. (2001). Cocaine and serotonin: a role for the 5-HT1A receptor site in the
mediation of cocaine stimulant effects. Behavioral Brain Research, 126(1-2), 127-133.
Carlson, K. M., J. M. Andresen, et al. (2009). Emerging pathogenic pathways in the spinocerebellar
ataxias. Current Opinion in Genetics & Development, 19(3), 247-253.
Chan, S. L., W. S. Griffin, et al. (1999). Evidence for caspase-mediated cleavage of AMPA receptor
subunits in neuronal apoptosis and Alzheimer’s disease. Journal of Neuroscience Research,
57(3), 315-323.
Chang, E. H., M. J. Savage, et al. (2006). AMPA receptor downscaling at the onset of Alzheimer’s
disease pathology in double knockin mice. Proceedings of the National Academy of Sciences of
the United States of America, 103(9), 3410-3415.
Chaudhury, D., J. J. Walsh, et al. (2013). Rapid regulation of depression-related behaviours by control
of midbrain dopamine neurons. Nature, 493(7433), 532-536.
Chen, J. W., D. E. Naylor, et al. (2007). Advances in the pathophysiology of status epilepticus. Acta
neurologica Scandinavica. Supplementum, 186, 7-15.
Chen, N. H. and M. E. A. Reith (2000). Structure and function of the dopamine transporter. European
Journal of Pharmacology, 405(1-3), 329-339.
Choi, D. W. (1988). Calcium-mediated neurotoxicity: relationship to specific channel types and role in
ischemic damage. Trends in Neuroscience, 11(10), 465-469.
Choi, D. W. (1992). Excitotoxic cell death. J Neurobiol, 23(9), 1261-1276.
Civelli, O., J. Bunzow, et al. (1991). Molecular biology of the dopamine D2 receptor. NIDA Res Monogr,
111, 45-53.
Clark, H. B., E. N. Burright, et al. (1997). Purkinje cell expression of a mutant allele of SCA1 in
transgenic mice leads to disparate effects on motor behaviors, followed by a progressive
cerebellar dysfunction and histological alterations. The Journal of Neuroscience, 17(19), 7385-
7395.
Cohen Kadosh, R., S. Soskic, et al. (2010). Modulating neuronal activity produces specific and long-
lasting changes in numerical competence. Current Biology, 20(22), 2016-2020.
168 Brain Neurochemistry and Cognitive Performance: Neurotransmitter Systems
Colibazzi, T., B. E. Wexler, et al. (2013). Anatomical abnormalities in gray and white matter of the
cortical surface in persons with schizophrenia. PLoS One, 8(2), e55783.
Cools, R. and M. D’Esposito (2011). Inverted-U-shaped dopamine actions on human working memory
and cognitive control. Biology Psychiatry, 69(12), e113-125.
Cousins, D. A., K. Butts, et al. (2009). The role of dopamine in bipolar disorder. Bipolar Disorders,
11(8), 787-806.
Coyle, J. T., A. Basu, et al. (2012). Glutamatergic synaptic dysregulation in schizophrenia: therapeutic
implications. Handbook of Experimental Pharmacology, 213, 267-295.
Cull-Candy, S., S. Brickley, et al. (2001). NMDA receptor subunits: diversity, development and
disease. Current Opinion Neurobiolgy, 11(3), 327-335.
Deidda, G., I. F. Bozarth, et al. (2014). Modulation of GABAergic transmission in development and
neurodevelopmental disorders: investigating physiology and pathology to gain therapeutic
perspectives. Frontiers in Cellular Neuroscience, 8, 119.
Del Castillo, J., W. C. De Mello, et al. (1964). Inhibitory action of gamma-aminobutyric acid (GABA) on
Ascaris muscle. Experientia, 20(3), 141-143.
Deutch, A. Y., M. Goldstein, et al. (1988). S Telencephalic projections of the A8 dopamine cell group.
Annuals of the New York Academy of Science, 537, 27-50.
Deutch, A. Y. and R. H. Roth (1999). Neurotransmitters. Fundamental Neuroscience. M. J. Zigmond, F.
E. Bloom, S. C. Landis, J. L. Roberts and L. R. Squire. San Diego, Academic Press: 193-234.
Deutsch, S. I., R. B. Rosse, et al. (2001). A revised excitotoxic hypothesis of schizophrenia:
therapeutic implications. Clinical Neuropharmacology, 24(1), 43-49.
Dingledine, R., K. Borges, et al. (1999). The glutamate receptor ion channels. Pharmacol Rev, 51(1),
7-61.
Dirnagl, U., C. Iadecola, et al. (1999). Pathobiology of ischaemic stroke: an integrated view. Trends in
Neuroscience, 22(9), 391-397.
Dong, X. X., Y. Wang, et al. (2009). Molecular mechanisms of excitotoxicity and their relevance to
pathogenesis of neurodegenerative diseases. Acta Pharmacologica Sin, 30(4), 379-387.
Dornhege, G., B. Blankertz, et al. (2006). Combined optimization of spatial and temporal filters for
improving brain-computer interfacing. IEEE Trans Biomedical Eng, 53(11), 2274-2281.
Doyle, K. P., R. P. Simon, et al. (2008). Mechanisms of ischemic brain damage. Neuropharmacology,
55(3), 310-318.
Erecinska, M. and F. Dagani (1990). Relationships between the neuronal sodium/potassium pump
and energy metabolism. Effects of K+, Na+, and adenosine triphosphate in isolated brain
synaptosomes. The Journal of General Physiology, 95(4), 591-616.
Evans, J. S. (2008). Dual-processing accounts of reasoning, judgment, and social cognition. Annual
Review of Psychology, 59, 255-278.
Falkenberg, L. E., R. Westerhausen, et al. (2014). Impact of glutamate levels on neuronal response
and cognitive abilities in schizophrenia. NeuroImage. Clinical, 4, 576-584.
Fallon, J. H. (1988). Topographic organization of ascending dopaminergic projections. Annuals of the
New York Academy of Science, 537, 1-9.
Fallon, J. H. and S. E. Loughlin (1985a). Substantia nigra. In G. Paxinos and J. Watson (Eds.), The rat
nervous system: A handbook for neuroscientists (pp. 353-374). Sydney: Academic Press.
Fallon, J. H. and S. E. Loughlin (1985b). Substantia nigra. The Rat Nervous System. G. Paxinos. New
York, Academic Press: 353-374.
Fallon, J. H. and S. E. Loughlin (1995). Substantia nigra. In G. Paxinos (Ed.), The Rat Nervous System
(pp. 215-237). Sydney: Academic Press.
Fallon, J. H. and R. Y. Moore (1978). S Catecholamine innervation of the basal forebrain. IV.
Topography of the dopamine projection to the basal forebrain and neostriatum. Journal of
Comparative Neurology, 180(3), 545-580.
References 169
Farrant, M. and Z. Nusser (2005). Variations on an inhibitory theme: phasic and tonic activation of
GABA(A) receptors. Nature reviews. Neuroscience, 6(3), 215-229.
Fatemi, S. H., T. D. Folsom, et al. (2013). mRNA and protein expression for novel GABAA receptors
theta and rho2 are altered in schizophrenia and mood disorders; relevance to FMRP-mGluR5
signaling pathway. Translational Psychiatry, 3, e271.
Faulkner, P. and J. F. Deakin (2014). The role of serotonin in reward, punishment and behavioural
inhibition in humans: Insights from studies with acute tryptophan depletion. Neuroscience and
Biobehavior Reviews, 3, 365-78.
Filip, M., E. Nowak, et al. (2001). On the role of serotonin 2A/2C receptors in the sensitization to
cocaine. The Journal of Physiology & Pharmacology, 52(3), 471-481.
Fischer, H., L. Nyberg, et al. (2010). Simulating neurocognitive aging: effects of a dopaminergic
antagonist on brain activity during working memory. Biology & Psychiatry 67(6): 575-580.
Fonnum, F. (1984). Glutamate: a neurotransmitter in mammalian brain. Journal of Neurochemistry,
42(1), 1-11.
Franke, A. G., C. Bonertz, et al. (2011). Non-medical use of prescription stimulants and illicit use of
stimulants for cognitive enhancement in pupils and students in Germany. Pharmacopsychiatry,
44(2), 60-66.
Friedman, L. K. (2006). Calcium: a role for neuroprotection and sustained adaptation. Molecular
Intervention, 6(6), 315-329.
Gadea, A. and A. M. Lopez-Colome (2001). Glial transporters for glutamate, glycine, and GABA: II.
GABA transporters. Journal of Neuroscience Research, 63(6), 461-468.
Gamo, N. J., M. Wang, et al. (2010). Methylphenidate and atomoxetine enhance prefrontal function
through alpha2-adrenergic and dopamine D1 receptors. Journal of the American Academy of
Child and Adolescent Psychiatry, 49(10), 1011-1023.
Gatchel, J. R. and H. Y. Zoghbi (2005). Diseases of unstable repeat expansion: mechanisms and
common principles. Nature reviews. Genetics, 6(10), 743-755.
Gerfen, C. R. (1989). The neostriatal mosaic: striatal patch-matrix organization is related to cortical
lamination. Science, 246(4928), 385-388.
Ghasemzadeh, M. B., L. K. Permenter, et al. (2003). Homer1 proteins and AMPA receptors modulate
cocaine-induced behavioural plasticity. The European Journal of Neuroscience, 18(6), 1645-1651.
Gilpin, N. W. and G. F. Koob (2008). Neurobiology of alcohol dependence: focus on motivational
mechanisms. Alcohol Research in Health, 31(3), 185-195.
Glannon, W. (2009). Stimulating brains, altering minds. Journal of Medical Ethics 35(5): 289-292.
Goldman-Rakic, P. S. (1992). Dopamine-mediated mechanisms of the prefrontal cortex. The
Neurosciences, 4, 149-159.
Gonzalez-Burgos, G., K. N. Fish, et al. (2011). GABA neuron alterations, cortical circuit dysfunction
and cognitive deficits in schizophrenia. Neural Plasticity, 2011, 723184.
Greger, I. H., E. B. Ziff, et al. (2007). Molecular determinants of AMPA receptor subunit assembly.
Trends in Neuroscience, 30(8), 407-416.
Grouselle, D., R. Winsky-Sommerer, et al. (1998). Loss of somatostatin-like immunoreactivity in the
frontal cortex of Alzheimer patients carrying the apolipoprotein epsilon 4 allele. Neuroscience
Letters, 255(1), 21-24.
Gunasekar, P. G., A. G. Kanthasamy, et al. (1995). NMDA receptor activation produces concurrent
generation of nitric oxide and reactive oxygen species: implication for cell death. Journal of
Neurochemistry, 65(5), 2016-2021.
Halliday, G., A. Harding, et al. (1995). Serotonin and tachykinin systems. In G. Paxinos (Ed.), The rat
nervous system (pp. 1205-1256). New York: Academic Press.
Harkany, T., I. Abraham, et al. (2000). beta-amyloid neurotoxicity is mediated by a glutamate-
triggered excitotoxic cascade in rat nucleus basalis. The European Journal of Neuroscience,
12(8), 2735-2745.
170 Brain Neurochemistry and Cognitive Performance: Neurotransmitter Systems
Hasbi, A., B. F. O’Dowd, et al. (2011). Dopamine D1-D2 receptor heteromer signaling pathway in the
brain: emerging physiological relevance. Molecular Brain, 4, 26.
Hashimoto, T., H. H. Bazmi, et al. (2008). Conserved regional patterns of GABA-related transcript
expression in the neocortex of subjects with schizophrenia. The American Journal of Psychiatry,
165(4), 479-489.
Heisler, L. K., H. M. Chu, et al. (1998). Elevated anxiety and antidepressant-like responses in
serotonin 5-HT1A receptor mutant mice. Procedures of the National Academy of Scieces 95(25):
15049-15054.
Heja, L., G. Nyitrai, et al. (2012). Astrocytes convert network excitation to tonic inhibition of neurons.
BMC Biology, 10, 26.
Herve, D., V. M. Pickel, et al. (1987). Serotonin axon terminals in the ventral tegmental area of the
rat: fine structure and synaptic input to dopaminergic neurons. Brain Research, 435(1-2), 71-83.
Hoyer, D., J. P. Hannon, et al. (2002). Molecular, pharmacological and functional diversity of 5-HT
receptors. Pharmacology Biochemistry & Behavior, 71(4), 533-554.
Hsu, C. Y. (1998). Ischemic Stroke: From Basic Mechanisms to New Drug Development, Karger.
Ikeda, Y., K. A. Dick, et al. (2006). Spectrin mutations cause spinocerebellar ataxia type 5. Nature
Genetics, 38(2), 184-190.
Iuculano, T. and R. Cohen Kadosh (2013). The mental cost of cognitive enhancement. The Journal of
neuroscience: the official journal of the Society for Neuroscience, 33(10), 4482-4486.
Iuculano, T. and R. Cohen Kadosh (2014). Preliminary evidence for performance enhancement
following parietal lobe stimulation in Developmental Dyscalculia. Frontier Human Neuroscience,
8, 38.
Kalivas, P. W. and P. Duffy (1997). “Dopamine regulation of extracellular glutamate in the nucleus
accumbens.” Brain research 761(1): 173-177.
Kamenetz, F., T. Tomita, et al. (2003). APP processing and synaptic function. Neuron, 37(6), 925-937.
Katsura, K., T. Kristian, et al. (1994). Energy metabolism, ion homeostasis, and cell damage in the
brain. Biochemistry Soic Trans, 22(4), 991-996.
Kew, J. N. and J. A. Kemp (2005). Ionotropic and metabotropic glutamate receptor structure and
pharmacology. Psychopharmacology (Berl), 179(1), 4-29.
Kim, C. H., J. Lee, et al. (2008). Metabotropic glutamate receptors: phosphorylation and receptor
signaling. Journal of Neurosci Research, 86(1), 1-10.
Kim, J. H., R. Anwyl, et al. (2001). Use-dependent effects of amyloidogenic fragments of (beta)-
amyloid precursor protein on synaptic plasticity in rat hippocampus in vivo. The Journal of
Neuroscience, 21(4), 1327-1333.
Kish, S. J., J. Tong, et al. (2008). Preferential loss of serotonin markers in caudate versus putamen in
Parkinson’s disease. Brain, 131(Pt 1), 120-131.
Koob, G. F. (1992). Drugs of abuse: anatomy, pharmacology and function of reward pathways. Trends
in Pharmacological Sciences, 13(5), 177-184.
Koob, G. F. (2014). Neurocircuitry of alcohol addiction: synthesis from animal models. Handbook of
Clinical Neurology, 125, 33-54.
Koob, G. F. and N. R. Swerdlow (1988). S The functional output of the mesolimbic dopamine system.
Annuals of the New York Academy of Science, 537, 216-227.
Kroemer, G., L. Galluzzi, et al. (2009). Classification of cell death: recommendations of the
Nomenclature Committee on Cell Death 2009. Cell Death & Differentiation, 16(1), 3-11.
Lawrie, S. M. and S. S. Abukmeil (1998). Brain abnormality in schizophrenia. A systematic and
quantitative review of volumetric magnetic resonance imaging studies. The British Journal of
Psychiatry, 172, 110-120.
Lee, A. and D. V. Pow (2010). Astrocytes: Glutamate transport and alternate splicing of transporters.
International Journal of Biochemical & Cellullar Biology, 42(12), 1901-1906.
References 171
Lee, Y. S. and A. J. Silva (2009). The molecular and cellular biology of enhanced cognition. Nature
Review. Neuroscience, 10(2), 126-140.
Leisman, G., O. Braun-Benjamin, et al. (2014). Cognitive-motor interactions of the basal ganglia in
development. Frontiers in Systems & Neuroscience, 8, 16.
Lewis, D. A., T. Hashimoto, et al. (2005). Cortical inhibitory neurons and schizophrenia. Nature
reviews. Neuroscience, 6(4), 312-324.
Liguz-Lecznar, M. and J. Skangiel-Kramska (2007). Vesicular glutamate transporters (VGLUTs): the
three musketeers of glutamatergic system. Acta Neurobiologica Experimental, 67(3), 207-218.
Liu, B., M. Liao, et al. (2006). Ischemic insults direct glutamate receptor subunit 2-lacking AMPA
receptors to synaptic sites. The Journal of Neurosciece, 26(20), 5309-5319.
Lloyd-Jones, D., R. Adams, et al. (2009). Heart disease and stroke statistics - 2009 update: a report
from the American Heart Association Statistics Committee and Stroke Statistics Subcommittee.
Circulation, 119(3), 480-486.
Lo, E. H. (2008a). Experimental models, neurovascular mechanisms and translational issues in
stroke research. Br J Pharmacol, 153 Suppl 1, S396-405.
Lo, E. H. (2008b). A new penumbra: transitioning from injury into repair after stroke. Nature
Medicine, 14(5), 497-500.
Luscher, C., L. Y. Jan, et al. (1997). G protein-coupled inwardly rectifying K+ channels (GIRKs) mediate
postsynaptic but not presynaptic transmitter actions in hippocampal neurons. Neuron, 19(3),
687-695.
Marczynski, T. J. (1998). GABAergic deafferentation hypothesis of brain aging and Alzheimer’s
disease revisited. Brain Research Bulletin, 45(4), 341-379.
Mark, K. A., M. S. Quinton, et al. (2007). Dynamic changes in vesicular glutamate transporter 1
function and expression related to methamphetamine-induced glutamate release. The Journal
of neuroscience, 27(25), 6823-6831.
Mark, K. A., J. J. Soghomonian, et al. (2004). High-dose methamphetamine acutely activates the
striatonigral pathway to increase striatal glutamate and mediate long-term dopamine toxicity.
The Journal of neuroscience, 24(50), 11449-11456.
Martin, R. L., H. G. Lloyd, et al. (1994). The early events of oxygen and glucose deprivation: setting
the scene for neuronal death? Trends in Neuroscience, 17(6), 251-257.
Mayer, M. (2004). Structure and function of glutamate receptors. Annuals of the New York Academy
of Science, 1038, 125-130.
McFarland, K., C. C. Lapish, et al. (2003). Prefrontal glutamate release into the core of the nucleus
accumbens mediates cocaine-induced reinstatement of drug-seeking behavior. The Journal of
euroscience, 23(8), 3531-3537.
McIntire, S. L., R. J. Reimer, et al. (1997). Identification and characterization of the vesicular GABA
transporter. Nature, 389(6653), 870-876.
Meldrum, B. (2007). Status epilepticus: the past and the future. Epilepsia, 48 Suppl 8, 33-34.
Meneses, A. (1999). 5-HT system and cognition. Neuroscience Biobehavior Review 23(8): 1111-1125.
Metzger, T. and E. Hildt (2011). Cognitive Enhancement. In J. Illes and B. J. Sahakian (Eds.), Oxford
Handbook of Neuroethics (pp. 245-264). Oxford: Oxford University Pres.
Meyer-Lindenberg, A. (2011). Neuroimaging and the question of neurodegeneration in
schizophrenia. Progress in Neurobiology, 95(4), 514-516.
Miguel-Hidalgo, J. J., X. A. Alvarez, et al. (2002). Neuroprotection by memantine against
neurodegeneration induced by beta-amyloid(1-40). Brain Research, 958(1), 210-221.
Miller, E. K. (2000). The prefrontal cortex and cognitive control. Nature reviews. Neuroscience, 1(1),
59-65.
Miller, G. W., R. R. Gainetdinov, et al. (1999). Dopamine transporters and neuronal injury. Trends in
Pharmacological Sciences, 20(10), 424-429.
Minkeviciene, R., S. Rheims, et al. (2009). Amyloid beta-induced neuronal hyperexcitability triggers
progressive epilepsy. The Journal of Neuroscience, 29(11), 3453-3462.
172 Brain Neurochemistry and Cognitive Performance: Neurotransmitter Systems
Paciaroni, M., V. Caso, et al. (2009). The concept of ischemic penumbra in acute stroke and
therapeutic opportunities. European Neurology, 61(6), 321-330.
Palop, J. J. and L. Mucke (2010). Amyloid-beta-induced neuronal dysfunction in Alzheimer’s disease:
from synapses toward neural networks. Nature Neuroscience, 13(7), 812-818.
Parameshwaran, K., M. Dhanasekaran, et al. (2008). Amyloid beta peptides and glutamatergic
synaptic dysregulation. Experimental Neurology, 210(1), 7-13.
Peng, P. L., X. Zhong, et al. (2006). ADAR2-dependent RNA editing of AMPA receptor subunit GluR2
determines vulnerability of neurons in forebrain ischemia. Neuron, 49(5), 719-733.
Perreault, M. L., B. F. O’Dowd, et al. (2014). Dopamine D(1)-D(2) receptor heteromer regulates
signaling cascades involved in addiction: potential relevance to adolescent drug susceptibility.
Developmental Neuroscience, 36(3-4), 287-296.
Pierri, J. N., A. S. Chaudry, et al. (1999). Alterations in chandelier neuron axon terminals in the prefrontal
cortex of schizophrenic subjects. The American Journal of Psychiatry, 156(11), 1709-1719.
Pin, J. P. and F. Acher (2002). The metabotropic glutamate receptors: structure, activation mechanism
and pharmacology. Current Drug Targets CNS Neurologic Disorders, 1(3), 297-317.
Pinard, A., R. Seddik, et al. (2010). GABAB receptors: physiological functions and mechanisms of
diversity. Advances in pharmacology, 58, 231-255.
Pineyro, G. and P. Blier (1999). Autoregulation of serotonin neurons: role in antidepressant drug
action. Pharmacology Reviews, 51(3), 533-591.
Plitman, E., S. Nakajima, et al. (2014). Glutamate-mediated excitotoxicity in schizophrenia: A review.
European Neuropsychopharmacology, 24(10), 1591-1605.
Porras, G., V. Di Matteo, et al. (2002). 5-HT2A and 5-HT2C/2B receptor subtypes modulate dopamine
release induced in vivo by amphetamine and morphine in both the rat nucleus accumbens and
striatum. Neuropsychopharmacology, 26(3), 311-324.
Prybylowski, K. and R. J. Wenthold (2004). N-Methyl-D-aspartate receptors: subunit assembly and
trafficking to the synapse. Journal of Biology & Chemistry, 279(11), 9673-9676.
Puig, M. V., J. Rose, et al. (2014). Dopamine modulation of learning and memory in the prefrontal
cortex: insights from studies in primates, rodents, and birds. Frontiers in Neural Circuits, 8, 93.
Ragan, C. I., I. Bard, et al. (2013). What should we do about student use of cognitive enhancers? An
analysis of current evidence. Neuropharmacology, 64, 588-595.
Rao, J. S., M. Kellom, et al. (2012). Dysregulated glutamate and dopamine transporters in
postmortem frontal cortex from bipolar and schizophrenic patients. Journal of Affective
Disorders, 136(1-2), 63-71.
Reinikainen, K. J., L. Paljarvi, et al. (1988). A post-mortem study of noradrenergic, serotonergic and
GABAergic neurons in Alzheimer’s disease. Journal of the Neurological Sciences, 84(1), 101-116.
Reith, M. E., C. Xu, et al. (1997). Pharmacology and regulation of the neuronal dopamine transporter.
European Journal of Pharmacology, 324(1), 1-10.
Repantis, D., P. Schlattmann, et al. (2010). Modafinil and methylphenidate for neuroenhancement in
healthy individuals: A systematic review. Pharmacological Research, 62(3), 187-206.
Riedel, O., J. Klotsche, et al. (2010). Frequency of dementia, depression, and other neuropsychiatric
symptoms in 1,449 outpatients with Parkinson’s disease. Journal Neurology, 257(7), 1073-1082.
Rissman, R. A. and W. C. Mobley (2011). Implications for treatment: GABAA receptors in aging, Down
syndrome and Alzheimer’s disease. Journal of Neurochemistry, 117(4), 613-622.
Robbins, T. W. and B. J. Everrit (1992). Functions of dopamine in the dorsal and ventral striatum. The
Neurosciences, 4, 119-127.
Roberts, E. and S. Frankel (1950). gamma-Aminobutyric acid in brain: its formation from glutamic
acid. The Journal of Biological Chemistry, 187(1), 55-63.
Robinson, S. R. (2001). Changes in the cellular distribution of glutamine synthetase in Alzheimer’s
disease. Journal of Neuroscience Research, 66(5), 972-980.
174 Brain Neurochemistry and Cognitive Performance: Neurotransmitter Systems
Roeper, J. (2013). Dissecting the diversity of midbrain dopamine neurons. Trends Neurosci 36(6):
336-342.
Roselli, F., M. Tirard, et al. (2005). Soluble beta-amyloid1-40 induces NMDA-dependent degradation
of postsynaptic density-95 at glutamatergic synapses. The Journal of Neuroscience, 25(48),
11061-11070.
Ross, C. A. and M. A. Poirier (2004). Protein aggregation and neurodegenerative disease. Nature
Medicine, 10 Suppl, S10-17.
Rossato, J. I., L. R. Bevilaqua, et al. (2009). Dopamine controls persistence of long-term memory
storage. Science, 325(5943), 1017-1020.
Rossi, D. J., T. Oshima, et al. (2000). Glutamate release in severe brain ischaemia is mainly by
reversed uptake. Nature, 403(6767), 316-321.
Sagne, C., S. El Mestikawy, et al. (1997). Cloning of a functional vesicular GABA and glycine
transporter by screening of genome databases. FEBS letters, 417(2), 177-183.
Salter, M. W. and S. Beggs (2014). Sublime microglia: expanding roles for the guardians of the CNS.
Cell, 158(1), 15-24.
Samii, A., J. G. Nutt, et al. (2004). Parkinson’s disease. Lancet, 363(9423), 1783-1793.
Sanchez, R. M., S. Koh, et al. (2001). Decreased glutamate receptor 2 expression and enhanced
epileptogenesis in immature rat hippocampus after perinatal hypoxia-induced seizures. The
Journal of Neuroscience, 21(20), 8154-8163.
Saransaari, P. and S. S. Oja (1992). Release of GABA and taurine from brain slices. Progress in
Neurobiology, 38(5), 455-482.
Schafer, D. P., E. K. Lehrman, et al. (2013). The “quad-partite” synapse: microglia-synapse
interactions in the developing and mature CNS. Glia, 61(1), 24-36.
Schindler, K. A., S. Bialonski, et al. (2008). Evolving functional network properties and
synchronizability during human epileptic seizures. Chaos, 18(3), 033119.
Schramm, N. L., R. E. Egli, et al. (2002). LTP in the mouse nucleus accumbens is developmentally
regulated. Synapse, 45(4), 213-219.
Schuske, K., A. A. Beg, et al. (2004). The GABA nervous system in C. elegans. Trends in
Neurosciences, 27(7), 407-414.
Schwartz-Bloom, R. D. and R. Sah (2001). Gamma-Aminobutyric acid(A) neurotransmission and
cerebral ischemia. Journal of Neurochemistry, 77(2), 353-371.
Scott, H. A., F. M. Gebhardt, et al. (2011). Glutamate transporter variants reduce glutamate uptake in
Alzheimer’s disease. Neurobiology of Aging, 32(3), 553 e551-511.
Selkoe, D. J. (2002). Alzheimer’s disease is a synaptic failure. Science, 298(5594), 789-791.
Sellings, L. H. and P. B. Clarke (2003). Segregation of amphetamine reward and locomotor
stimulation between nucleus accumbens medial shell and core. The Journal of Neuroscience,
23(15), 6295-6303.
Serra, H. G., C. E. Byam, et al. (2004). Gene profiling links SCA1 pathophysiology to glutamate
signaling in Purkinje cells of transgenic mice. Human Molecular Genetics, 13(20), 2535-2543.
Seyedabadi, M., G. Fakhfouri, et al. (2014). The role of serotonin in memory: interactions with
neurotransmitters and downstream signaling. Experimental Brain Research, 232(3), 723-738.
Shimizu, K., K. Matsubara, et al. (2003). Paraquat leads to dopaminergic neural vulnerability in
organotypic midbrain culture. Neuroscience Research, 46(4), 523-532.
Shimo, Y. and T. Wichmann (2009). Neuronal activity in the subthalamic nucleus modulates the release
of dopamine in the monkey striatum. The European Journal of Neuroscience, 29(1), 104-113.
Sims, N. R. and E. Zaidan (1995). Biochemical changes associated with selective neuronal death
following short-term cerebral ischaemia. International Journal of Biochemistry & Cellular
Biology, 27(6), 531-550.
Singer, A. C. and L. M. Frank (2009). Rewarded outcomes enhance reactivation of experience in the
hippocampus. Neuron, 64(6), 910-921.
References 175
Snyder, E. M., Y. Nong, et al. (2005). Regulation of NMDA receptor trafficking by amyloid-beta. Nature
Neuroscience, 8(8), 1051-1058.
Sokolow, S., S. H. Luu, et al. (2012). Preferential accumulation of amyloid-beta in presynaptic
glutamatergic terminals (VGluT1 and VGluT2) in Alzheimer’s disease cortex. Neurobiology of
Disease, 45(1), 381-387.
Steinbusch, H. W. (1981). Distribution of serotonin-immunoreactivity in the central nervous system of
the rat-cell bodies and terminals. Neuroscience, 6(4), 557-618.
Stockmeier, C. A. (2003). Involvement of serotonin in depression: evidence from postmortem and
imaging studies of serotonin receptors and the serotonin transporter. Journal of Psychiatric
Research, 37(5), 357-373.
Sullivan, E. M. and P. O’Donnell (2012). Inhibitory interneurons, oxidative stress, and schizophrenia.
Schizophrenia bulletin, 38(3), 373-376.
Sullivan, R. M. (2004). Hemispheric asymmetry in stress processing in rat prefrontal cortex and the
role of mesocortical dopamine. Stress, 7(2), 131-143.
Sullivan, R. M. and A. Gratton (1998). Relationships between stress-induced increases in medial
prefrontal cortical dopamine and plasma corticosterone levels in rats: role of cerebral laterality.
Neuroscience, 83(1), 81-91.
Sun, W., E. McConnell, et al. (2013). Glutamate-dependent neuroglial calcium signaling differs
between young and adult brain. Science, 339(6116), 197-200.
Surmeier, D. J., J. Ding, et al. (2007). “D1 and D2 dopamine-receptor modulation of striatal glutamatergic
signaling in striatal medium spiny neurons.” Trends in neurosciences 30(5): 228-235.
Swanson, G. T., S. K. Kamboj, et al. (1997). Single-channel properties of recombinant AMPA
receptors depend on RNA editing, splice variation, and subunit composition. The Journal of
Neuroscience, 17(1), 58-69.
Swanson-Park, J. L., C. M. Coussens, et al. (1999). A double dissociation within the hippocampus
of dopamine D1/D5 receptor and beta-adrenergic receptor contributions to the persistence of
long-term potentiation. Neuroscience, 92(2), 485-497.
Szydlowska, K. and M. Tymianski (2010). Calcium, ischemia and excitotoxicity. Cell Calcium, 47(2),
122-129.
Tabakoff, B. and P. L. Hoffman (1996). Alcohol addiction: an enigma among us. Neuron, 16(5), 909-
912.
Tabakoff, B. and P. L. Hoffman (2013). The neurobiology of alcohol consumption and alcoholism: an
integrative history. Pharmacology, biochemistry, and behavior, 113, 20-37.
Theberge, J., K. E. Williamson, et al. (2007). Longitudinal grey-matter and glutamatergic losses in
first-episode schizophrenia. The British journal of psychiatry, 191, 325-334.
Thierry, A. M., J. Mantz, et al. (1988). Influence of the mesocortical / prefrontal dopamine neurons on
their target cells. Annuals of the New York Academy of Science, 537, 101-111.
Timmerman, W. and B. H. Westerink (1997). Electrical stimulation of the substantia nigra reticulata:
detection of neuronal extracellular GABA in the ventromedial thalamus and its regulatory
mechanism using microdialysis in awake rats. Synapse, 26(1), 62-71.
Tork, I. (1990). Anatomy of the serotonergic system. Ann N Y Acad Sci, 600, 9-34.
Trevelyan, A. J. and C. A. Schevon (2013). How inhibition influences seizure propagation.
Neuropharmacology, 69, 45-54.
Trevitt, T., B. Carlson, et al. (2002). Interactions between dopamine D1 receptors and gamma-
aminobutyric acid mechanisms in substantia nigra pars reticulata of the rat: neurochemical and
behavioral studies. Psychopharmacology, 159(3), 229-237.
Tye, K. M., J. J. Mirzabekov, et al. (2013). Dopamine neurons modulate neural encoding and
expression of depression-related behaviour. Nature, 493(7433), 537-541.
Vallone, D., R. Picetti, et al. (2000). Structure and function of dopamine receptors. Neuroscience
Biobehavioral Reviews, 24(1), 125-132.
176 Brain Neurochemistry and Cognitive Performance: Neurotransmitter Systems
Vanhoenacker, P., G. Haegeman, et al. (2000). 5-HT7 receptors: current knowledge and future
prospects. Trends in Pharmacological Scieces, 21(2), 70-77.
Vertes, R. P. and A. M. Crane (1997). Distribution, quantification, and morphological characteristics
of serotonin-immunoreactive cells of the supralemniscal nucleus (B9) and pontomesencephalic
reticular formation in the rat. Journal of Comparative Neurology, 378(3), 411-424.
Vertes, R. P., W. J. Fortin, et al. (1999). Projections of the median raphe nucleus in the rat. Journal of
Comparative Neurology, 407(4), 555-582.
Volkow, N. D., J. S. Fowler, et al. (2002). Role of dopamine, the frontal cortex and memory circuits in
drug addiction: insight from imaging studies. Neurobiology of Learning & Memory, 78(3), 610-
624.
Voorn, P. (1988). The dopaminergic innervation of the striatum in th rat. Department of Anatomy and
Embryology. Amsterdam, Vrije Universiteit, 175 p.
Walker, R. H., C. Moore, et al. (2012). “Effects of subthalamic nucleus lesions and stimulation upon
corticostriatal afferents in the 6-hydroxydopamine-lesioned rat.” PloS one 7(3): e32919.
Wang, D. D. and A. R. Kriegstein (2011). Blocking early GABA depolarization with bumetanide results
in permanent alterations in cortical circuits and sensorimotor gating deficits. Cerebral cortex,
21(3), 574-587.
Wasterlain, C. G., J. M. Bronstein, et al. (1992). Translocation and autophosphorylation of brain
calmodulin kinase II in status epilepticus. Epilepsy Research. Supplement, 9, 231-238.
Wasterlain, C. G., H. Liu, et al. (2009). Molecular basis of self-sustaining seizures and
pharmacoresistance during status epilepticus: The receptor trafficking hypothesis revisited.
Epilepsia, 50 Suppl 12, 16-18.
Weinberger, D. R., K. F. Berman, et al. (1988). Mesocortical dopaminergic function and human
cognition. Annual of the New york Academy of Sciences, 537, 330-338.
Willner, P., A. S. Hale, et al. (2005). Dopaminergic mechanism of antidepressant action in depressed
patients. Journal of Affective Disorders, 86(1), 37-45.
Winder, D. G., R. E. Egli, et al. (2002). Synaptic plasticity in drug reward circuitry. Current Molecular
Medicine, 2(7), 667-676.
Woodward, D. J., J. Y. Chang, et al. (1999). Mesolimbic neuronal activity across behavioral states. Ann
N Y Acad Sci, 877, 91-112.
Wu, Y., D. Liu, et al. (2014). Neuronal networks and energy bursts in epilepsy. Neuroscience.
Yadid, G., Y. Fraenkel, et al. (1998). Strychnine, glycine, and adrenomedullary secretion. Advances in
Pharmacology, 42, 604-606.
Yao, W. D., R. R. Gainetdinov, et al. (2004). Identification of PSD-95 as a regulator of dopamine-
mediated synaptic and behavioral plasticity. Neuron, 41(4), 625-638.
Ziemann, U., W. Paulus, et al. (2008). Consensus: Motor cortex plasticity protocols. Brain
Stimulation, 1(3), 164-182.
Zoghbi, H. Y. and H. T. Orr (2000). Glutamine repeats and neurodegeneration. Annual Review of
Neuroscience, 23, 217-247.
Section 3: Strategies to Develop New and Better Drugs
Nuno Vale*7
3.1 Amino Acids and Peptides in Medicine: Old
or New Drugs?
3.1.1 Introduction
Amino acids (AA) are molecules containing an amine group, a carboxylic acid group
and a side-chain that varies between different amino acids. AA are an important
class of cell signalling molecules, involved in the regulation of gene expression and
the protein phosphorylation cascade, as well precursors of hormone synthesis and
low-molecular nitrogenous substances (Wu, 2009). The 20 natural AA are commonly
found in proteins and are also referred to as alpha amino acids (Fig. 3.1.1). The α-amino
acids differ in the nature of the side-chain (R group) attached to their α-carbon, which
can vary in size from just one hydrogen atom in glycine to a large heterocyclic group
in tryptophan. α-Amino acids are typically divided into three categories, based on
the properties of their side-chain. The first category contains α-amino acids with
relatively nonpolar R groups (glycine, alanine, valine, leucine, isoleucine, proline,
phenylalanine, tryptophan, and methionine), the second contains α-amino acids
with uncharged but polar R groups (serine, threonine, cysteine, tyrosine, asparagine,
glutamine), and the third contains α-amino acids with charged R groups (aspartic
acid, glutamic acid, lysine, arginine and histidine) (Vig, 2013).
Proteins are key players in many vital processes in living organisms. They
transport substances, catalyze chemical reactions, pump ions or recognize signalling
molecules. The complexity and variety of proteins is tremendous: in the human
body alone, there are more than 100,000 different proteins at work. But almost all of
them are made up of just twenty different AA. Only a few highly specialized proteins
additionally contain selenocysteine (Sec, Fig. 3.1.2), the very rare 21st amino acid
discovered in 1986. Sec is identical to cysteine (Cys), except that it contains selenium
instead of sulphur. Selenocysteine is recognized as the 21st amino acid in ribosome-
mediated protein synthesis and its specific incorporation is directed by the UGA codon
(Stadtman, 1996).
Nuno Vale: Department of Chemistry and Biochemistry, Faculty of Sciences of University of Porto, Rua
do Campo Alegre, 687, 4169-007 Porto, Portugal; Research Unit on Applied Molecular Biosciences/
UCIBIO-REQUIMTE, Faculty of Sciences of University of Porto, Rua do Campo Alegre, 687, 4169-007
Porto, Portugal, *Email: nuno.vale@fc.up.pt
OH NH
S
HN
H COOH
Glycine Alanine Valine Leucine Isoleucine Methionine Proline Phenylalanine Tyrosine Tryptophan
(Gly) (Ala) (Val) (Leu) (Ile) (Met) (Pro) (Phe) (Tyr) (Trp)
Non-polar
Charged (basic)
Non Charged
Polar
N NH2 Charged (acidic)
NH2 HN
O NH2
NH N HO O O
O
OH OH SH
NH 2
OH
Lysine Argynine Hystidine Aspartic Glutamic Asparagine Glutamine Serine Threonine Cysteine
(Lys) (Arg) (His) (Asp) (Glu) (Asn) (Gln) (Ser) (Thr) (Cys)
(Z)
O N O
H
N
(R)
HSe (S) OH (R) (S) OH
NH 2 O NH2
L-Selenocysteine L-Pyrrolysine
(Sec) (Pyl)
Selenium is in the same family below sulfur in the periodic table and the two
elements share many similar chemical properties. Sec has a distinct functional
advantage because the selenol group is more fully ionized than the thiol group of Cys
at physiological pH (biosynthesis of Sec and de novo synthesis of Cys can be observed
in Fig. 3.1.3). When Sec is replaced with Cys, the catalytic activity of a selenoenzyme
is drastically reduced (Stadtman, 2000).
A big surprise was the discovery of a 22nd amino acid in methane-producing archaea
of the family Methanosarcinaceae in 2002, pyrrolysine (Pyl, Fig. 3.1.2). It is genetically
encoded in a similar manner as that of selenocysteine and the other twenty amino
acids. The archaea use this unusual amino acid in proteins required for energy
180 Amino Acids and Peptides in Medicine: Old or New Drugs?
conversion. Pyrrolysine is located in the catalytic center of the proteins carrying it and
is essential for their function. The energy generation process of the archaebacteria
would not work without pyrrolysine (Srinivasan, 2002). Biosynthesis of this amino
acid is based on the radical S-adenosyl-L-methionine (SAM) protein PylB, which
mediates a lysine mutase reaction and produces 3-methylornithine, which is then
ligated to a second molecule of lysine by PylC before oxidation by PylD to produce
pyrrolysine (Gaston, 2011). In addition, Sec and Pyl are often referred as the natural
amino acids.
Among more than 300 AA in nature, only 20 of them (α-AA) serve as building blocks
of proteins. However, represented in Fig. 3.1.4 are non-protein α-AA (e.g. ornithine,
citrulline, and homo-cysteine) and non-α AA (e.g. taurine and β-alanine), which
also play important roles in cell metabolism (Curis, 2007; Perta-Kajan, 2007; Manna,
2009).
Figure 3.1.3: Biosynthesis of Sec and de novo synthesis of Cys. The complete synthesis of Sec on its
tRNA in eukaryotes and archaea uses selenite and ATP are substrates for SPS2, yielding selenophos-
phate. This interacts with SecS and the intermediate, likely dehydroalanine, is used to generate Sec-
tRNA[Ser]Sec (shown on the right in the upper pathway). In the lower pathway, the de novo synthesis of
Cys is shown in mammals, where sulfide and ATP are substrates for SPS2 and yield thiophosphate,
which interacts with SecS and yields Cys-tRNA[Ser]Se from the SecS intermediate (once again, likely
dehydroalanine) (adapted from Turanov, 2011).
Introduction 181
O O O
H 2N (S) OH H2 N N (S) OH
H
NH2 NH2
L-Ornithine L-Citrulline
O O O
HS HO S
(S) OH NH 2 H 2N OH
O
NH2 Taurine β-Alanine
L-Homocysteine
Figure 3.1.4: Structures of non-protein α-AA (ornithine, citrulline, homo-cysteine) and non-α-AA
(taurine, β-alanine).
the regions of the brain, brain activity and animal species studied. It also presents
different functions, which have been studied for their potential in neurology as a
trophic factor in brain development, regulating calcium transport, in the integrity of
the eardrum, as an osmoregulator, neurotransmitter and neuromodulator as well as
for its neuroprotective action (Wu & Prentice, 2010).
All α-amino acids except glycine have a chiral α-carbon and can exist in two optical
isomers, L-form and D-form. The L-form of amino acids occurs naturally and prodrugs
utilizing these amino acids are generally activated by naturally occurring enzymes.
Both L- and D-amino acid prodrugs tend to have very similar physicochemical
properties but the latter is generally more stable to hydrolysis by naturally occurring
enzymes (Vig, 2003). This property of amino acids are often utilized by medicinal
chemists to develop stable amino acid prodrugs. In addition to the natural amino
acids and their D-forms, extensive arrays of synthetic amino acids and di-/tri-peptides
are commercially available for medicinal chemists as promoieties (Ma, 2003).
Table 3.1.1: Advantages of using amino acids in drug development (adapted from Vig, 2013).
Advantage
• Large structural diversity
• Commercial availability
• Side chain functional groups e.g. amine, carboxylic acid, alcohol, and thiol
• Commonly used linkage chemistry between amino acid and parent including ester,
amide, carbonate and carbamate bonds
• Amino acids are building blocks for proteins and are generally regarded as safe
• Can potentially target carrier-mediated transporters for improved delivery across cell
membranes
• Many amino acid prodrugs have been commercially developed and help patients
Introducing an amino acid, either a natural or its derivative, to a parent drug usually
increases the water solubility by several orders of magnitude through an ionized
carboxylate anion or ammonium cation. Consequently, there are several amino acid
ester prodrugs that are investigated as water soluble derivatives for oral administration
(Chan, 1998; Mulholland, 2001; Gingrich, 2003; Marathe, 2009). Amino acid esters
(Bundgaard, 1984) amino acid amides (Pochopin, 1995; Bradshaw, 2002; Mittal, 2007;
Rasheed, 2011) or other amino acid examples like amidoximes (Kotthaus, 2011) or
ureas (Hemenway, 2010) have been investigated for parenteral use, albeit to a lesser
extent. In proline, the side-chain links to the α-amino group, meaning it is the only
amino acid containing a secondary amine at this position. The pKa of the α-amino
group is typically in the range of 8.8 to 10.6 and the acidic α-carboxyl group has a
pKa typically between 1.7 and 2.6. The dissociation constant of both the α-amino and
α-carboxyl groups are affected by each other and the side chain. The α-amino group
in its charged state is strongly electron withdrawing and makes the α-carboxyl group
more acidic, thereby lowering its pKa as compared with the typical carboxylic acid
pKa of approximately 4.5. At physiological pH, the predominant form of α-amino
acids contains both a negative carboxylate and a positive ammonium group. This
molecular state is known as a zwitterion and has minimum solubility at the isoelectric
point (mid-point between 2 pKa). However, in most α-amino acid prodrugs either the
carboxylic acid or amine group is linked to a parent drug, abolishing the zwitterionic
form. Exceptions are α-amino acids that have an additional functional group in the
side chain amenable to prodrug formation (e.g. -SH in Cys, -OH in Tyr, -NH2 in Lys,
-COOH in Asp and Glu) which consequently can maintain a double charge. Water
solubility of α-amino acids themselves is largely a function of the polar or nonpolar
nature of the side chain. An increase in hydrocarbon content of the side chain (R
group) from Gly to Val or Leu decreases water solubility. Consequently, amino acids
with shorter hydrocarbon side chains resulted in higher aqueous solubility versus
amino acids with longer hydrocarbon side chains for a series of α-amino acid amide
prodrugs of dapsone (Fig. 3.1.5, Pochopin, 1995).
Moreover, α-amino acids with uncharged but polar R groups are generally more
soluble than the corresponding α-amino acids having nonpolar R groups. This trend is
in line with the observations from several studies, where phenylalanine as a promoiety
resulted in the least soluble prodrug in a series of amino acid prodrugs (Pochopin,
1995; Rautio, 1999). Among the α-amino acids with ionizable R groups, Lys renders
relatively high water solubility over the wide pH range due to the ionized ε-amine
regardless of prodrug type (e.g. ester or amide) (Pochopin, 1995; Nam, 2003).
Amino Acids and Drug Development 185
H
H2 N S N R
Figure 3.1.5: Structures of dapsone (R=H) and prodrugs with amino acids.
Amino acid prodrugs are expected to pose synthetic challenges similar to other
prodrugs. Amino acid prodrug linkage chemistry is well-established, therefore many
of the synthetic challenges will depend on the number and type of protecting groups
required to mask the active groups on the amino acid and parent drug. Furthermore,
all amino acids except glycine introduce a stereocenter to the prodrug. Like other
prodrugs, amino acid prodrugs increase the molecular weight of the parent, requiring
mindfulness of the increase in dose due to the “non-active” contribution of the
prodrug relative to the bioavailability enhancement with prodrug (Vig, 2013).
Many drugs suffer from an extensive first-pass metabolism leading to drug inactivation
and/or production of toxic metabolites, which makes them attractive targets for prodrug
design. The classical prodrug approach, which involves enzyme-sensitive covalent
linkage between the parent drug and a carrier moiety, is a well-established strategy to
overcome bioavailability/toxicity issues. However, the development of prodrugs that
can regenerate the parent drug through non-enzymatic pathways has emerged as an
alternative approach in which prodrug activation is not influenced by inter- and intra-
individual variability that affects enzymatic activity. As discussed above, amino acids
are excellent promoieties to increase the aqueous solubility of the parent drug. To
increase the dissolution rate further, amino acid prodrugs can be converted to their
salt forms. In the following sections, we will present some examples of commercial
as well as investigational amino acid prodrugs that have been designed to improve
1) oral bioavailability, 2) sustained drug delivery, 3) intravenous drug delivery, 4) to
target drugs to their site of action and 5) improve enzymatic stability.
Prodrug strategies have arguably been successful for a number of clinically-
used therapeutic agents. However, prodrug research encounters various challenges
and requires additional work in preclinical and clinical settings, much of which can
be attributed to understanding the bioconversion mechanisms of prodrugs. Many
enzymes involved in prodrug activation are subject to interindividual variabilities in
186 Amino Acids and Peptides in Medicine: Old or New Drugs?
their activities. The main factors contributing to this variability are intrinsic, especially
polymorphisms in the genes encoding the enzymes, but can also be extrinsic (i.e.
interactions caused by other drugs and xenobiotics). Both intrinsic and extrinsic
factors may cause insufficient or excessive conversion of the prodrugs into their active
forms. Moreover, interspecies differences in enzyme activation represent another
hurdle to the prediction of human disposition of certain prodrugs (Hutunen, 2011).
The enzymatic hydrolysis of amino acid esters and amides by various hydrolases
(e.g. esterases and/or peptidases) is far more effective than chemical hydrolysis.
The half-lives of prodrugs are usually several orders of magnitude shorter in blood
or tissue homogenates than in aqueous solutions. Often the activating enzymes are
unidentified. However, it has been suggested that the biphenyl hydrolase-like protein
human valacyclovirase (VACVase) is at least partly responsible for hydrolysis of the
prodrugs valacyclovir and valganciclovir (Fig. 3.1.6) and might be involved in the
activation of other amino acid prodrugs as well (Kim, 2003; 2004; Lai, 2004). As the
specificity of this enzyme resides mainly in the amino acid acyl promoiety (and to a
lesser extent in the alcohol moiety of a parent drug) and that the α-amino group in
the substrate is important for activity, the enzyme can be better defined as an α-amino
acid ester prodrug-activating enzyme (Burnette, 1995; Lai, 2004). VACVase prefers
small, hydrophobic (valine, proline) or aromatic side chains (phenylalanine) over
the charged amino acids (lysine, aspartic acid). It has shown clear stereoselectivity
for L-valine over D-valine prodrugs irrespective of the parent drug while exhibiting
comparable hydrolytic activity toward D-phenylalanine and L-phenylalanine esters,
as indicated in a study of various nucleoside analogue prodrugs (Kim, 2004). The
success of valacyclovir and valganciclovir has proved that an L-valine prodrug
approach is a very efficient option to improve the oral bioavailability by utilizing
amino acid transporters. Therefore, there are several other L-valyl ester prodrugs,
including nucleoside analogues levovirin valinate, valopicitabine, and valtorcitabine
currently undergoing clinical evaluation.
L-Valyl ester prodrugs are probably the first commercial examples of amino
acid prodrugs that utilize intestinal transporters to increase the permeability and
consequently poor oral bioavailability of their parent drugs. Valacyclovir (Valtrex®)
was the pioneer of L-valyl ester prodrugs, which achieved 3-5-times higher oral
bioavailability (> 60%) (Soul-Lawton, 1995) than that of its parent drug, acyclovir
(10−20%) (de Miranda and Blum, 1983). Valacyclovir is absorbed by the PepT1 and
ATB0,+ transporters and is then rapidly hydrolyzed to acyclovir, predominantly by
biphenyl hydrolase-like protein (valacyclovirase) (Kim, 2003; Hatanaka, 2004).
Finally, acyclovir triphosphate acts as an antiherpetic agent by inhibiting viral DNA
replication. Soon, after the discovery of valacyclovir, valganciclovir (Valcyte®) was
designed. Valganciclovir is also absorbed by the PepT1 and ATB0,+ transporters then
bioactivated by valacyclovirase (Kim, 2003; Umapathy, 2004). The oral bioavailability
of ganciclovir is approximately 60%, which is almost 10-times higher than that from
oral ganciclovir (6−8%) (Jung & Dorr, 1999).
Amino Acids and Drug Development 187
O O
N H 2N N H2N
HN HN OH
N O N O
H 2N N H2N N
O O O
O
Valacyclovir Valganciclovir
valacyclovirase
O O
N N
HN HN OH
OH N OH
N H 2N N
H 2N N
O O
Acyclovir Ganciclovir
O O
N N
HN HN OH
O O O
N N O P O P O P OH
H 2N N H 2N N
OH OH OH
O O
O O O Ganciclovir triphosphate
Acyclovir triphosphate O P O P O P OH
OH OH OH
Figure 3.1.6: Bioconversion of valacyclovir and valganciclovir to their parent drugs by valacyclovi-
rase and to their corresponding triphosphates, firstly by thymidine kinase and secondly by cellular
kinases.
The successful results obtained for valacyclovir and valganciclovir, amino acid ester
prodrugs of acyclovir and ganciclovir, respectively, has prompted the investigations
of amino acids as promoieties for other agents. This success has been attributed to
their enhanced intestinal transport via oligopeptide transporters. The work of Han
188 Amino Acids and Peptides in Medicine: Old or New Drugs?
prove that amino acid ester prodrugs significantly improve the cellular uptake of the
parent drugs via peptide transport mechanism, though there is no peptide bond in
their structures (Han, 1998). The fact that some cancer epithelial cells are rich in this
type of transporter suggests their use for the delivery of peptidomimetic anticancer
agents (Gonzalez, 1998; Nakanishi, 2000; Vig, 2003). Anticancer drugs are primarily
cytotoxic agents and exert their antitumor activity by interfering with some aspects
of DNA replication, repair, translation or cell division. However, these agents are not
“magic bullets”, as they do not destroy tumor cells while sparing the normal cells
(Sinhababu, 1996).
The prodrug strategy is once again used with anticancer agents as a way to
alleviate their toxicity. Different prodrug strategies have been employed to not only
to improve solubility, transport and pharmacokinetic properties, but also to enable
selective activation in target tissues. As a means to reduce the toxic effects of these
agents, prodrugs designed for selective activation in target tissues are by far the most
efficient and attractive option. However, an enzyme or transporter that is exclusively
or preferentially expressed in these tissues is a prerequisite for selective targeting
(Landowski, 2006; Sinhababu, 1996).
There are numerous studies of amino acid derivatives of the clinically-effective
anticancer agents floxuridine (Vig, 2003) and gemcitabine (Song, 2005) concerning the
activation of the prodrug. Unlike the desired rapid activation required for valacyclovir,
extensive intestinal activation of floxuridine and gemcitabine prodrugs would lead
to severe intestinal toxicity. In this regard, Song and Landowski demonstrated that
amino acid ester prodrugs provided resistance to deamination of gemcitabine (Song,
2005) and to floxuridine cleavage (Landowski, 2005a), respectively.
The studies developed with the amino acid esters of fluxoridine were consistent
with previous findings: 5’-L-valyl floxuridine was the most efficiently transported
floxuridine prodrug, exhibiting the highest PEPT1-mediated transport and permeability
across Caco-2 monolayers. The length and stereochemistry of the amino acid moiety
side chain influences the transport efficiency of floxuridine prodrugs. The slightly
more branched isoleucyl side chain reduced the transport of these prodrugs to half,
but the branching at γ carbon (as in leucine) side chain decreases this transport
even further. The permeability of 5’-monoester prodrugs of floxuridine across Caco-2
monolayers was significantly higher than that of the parent drug and also reflected
a profound promoiety dependency. Thus, the permeability of the 5’-L-valyl prodrug
was roughly 2- and 5-fold higher than the permeability of 5’-L-isoleucyl and 5’-leucyl
floxuridine prodrugs, respectively (Landowski, 2005a).
The metabolic conversion of floxuridine to 5-fluorouacil (5-FU) following systemic
delivery has been shown to be detrimental to the therapeutic efficacy of floxuridine.
The mechanism of action of these two drugs are well understood, where the toxicity
of 5-FU is predominantly caused by 5-FU incorporation into RNA. Unlike 5-FU,
floxuridine is specifically incorporated into DNA, which leads to the minimization of
adverse effects. It is important to highlight that floxuridine has shown to inhibit cell
Amino Acids and Drug Development 189
proliferation 10- to 100-fold more than 5-FU. However, floxuridine is rapidly converted
to 5-FU in many tissues (including the liver) by the enzyme thymidine phosphorylase
(Tsume, 2008). As a consequence, higher doses of floxuridine are required to maintain
clinical efficacy, which leads to greater toxicity. Therefore, protection of floxuridine
prodrugs against this enzyme is essential to enhance therapeutic efficacy at low doses
and obviate toxicity.
All amino acid ester prodrugs examined by Landowski and co-workers were stable
to glycosidic bond cleavage by thymidine phosphorylase. It is therefore possible to
conclude that modification of one or both of the free hydroxyl groups on the sugar
moiety provide protection from glycosidic bond cleavages. The rate of conversion of the
prodrugs to the parent drug after transport would determine floxuridine disposition
and therapeutic action (Landowski, 2005a). As previously reported to structure,
stereochemistry and the site of esterification of the amino acid promoiety affect the
rates of activation of floxuridine prodrugs. Therefore, the wide range of variations
for prodrug structure suggests that the hydrolysis rate can be tailored to produce a
prodrug with the desired half-life (Landowski, 2005a; Vig, 2003).
The roughly 5- to 12-fold higher activity in Caco-2 cell homogenates compared
with pH 7.4 buffer suggests the predominance of enzymatic bioconversion of the
prodrugs. The results obtained for the leucyl ester prodrugs indicate that they would
not be suitable candidates. In comparison, isoleucyl ester prodrugs of floxuridine
are enzymatically more stable than the valyl ester floxuridine prodrugs and the
reference drug valacyclovir (Landowsky, 2005b). The combined results of the in vitro
studies suggest that isoleucyl monoesters of floxuridine may be potentially promising
candidates for improving oral bioavailability of floxuridine in vivo. The prodrugs
given orally could improve the intestinal uptake of floxuridine as well as shield it from
unwanted degradation (Landowsky, 2005b). Various dipeptides and peptidomimetics
have also been tested to characterize the hPEPT1 transporter and improve its affinity,
and mono amino acid ester prodrugs have been evaluated as hPEPT1 substrates
(Landowsky, 2005a; 2005b; Song, 2005). Based on those reports, six amino acids were
chosen to be N-terminal amino acids of the dipeptide and paired with three others to
test the hypothesis that molecular sizes may structurally affect its ester bond stability
(Tsume, 2008). In this study, dipeptide prodrugs appeared to be less stable in pH
7.4 buffers than the corresponding mono amino acid ester prodrugs. Since no mono
amino acid ester prodrug degradation products were detected, it is quite likely that
the dipeptide monoester prodrugs degrade through parallel pathways, as previously
suggested for some anti-viral dipeptide prodrugs.
Although gemcitabine is clinically effective in the treatment of advanced or
metastatic pancreatic cancer, it also exhibits various side effects which are attributed
to its inability to distinguish between normal or target cells. It is known that
gemcitabine exerts its antiproliferative activity via multiple mechanisms of action. It
is initially phosphorylated intracellularly by deoxycytidine kinase and subsequently
by nucleotide kinases to its active metabolites, gemcitabine diphosphate and
190 Amino Acids and Peptides in Medicine: Old or New Drugs?
NH2 NH2
HN HN
O N O N
HO O O
O O
F F
R NH 2
OH F OH F
Figure 3.1.7: Amino acid-gemcitabine prodrug (R = L-Val, L-Ile, L-Phe, D-Val, D-Phe).
The chemical stability and rapid enzymatic bioconversion characteristic of the 5’-
L-valyl-gemcitabine derivative suggests it potential in enhancing oral absorption
of gemcitabine. On the other hand, the 5’-L-isoleucyl-gemcitabine showed a slow
bioconversion in Caco-2 cells and in human plasma, as well as an unusual resistance
to cytidine deaminase deactivation. This way a longer systemic circulation half-life is
possible which may facilitate the targeting of cells over expressing hPEPT-1 transporter
(Song, 2005).
Spontaneous cyclization of oligopeptides and prodrugs with amino acids in
solution does not easily occur without the intervention of specific enzymes, with the
exception of dipeptides that easily cyclize to piperazine-2,5-diones or diketopiperazines
(DKP, 1). The DKP scaffold is widely found in compounds of biological interest
and could serve as a drug template with appropriately arrayed pharmacophores
Amino Acids and Drug Development 191
(Gomes, 2007). The major role of DKP in prodrug design falls in the domain of approach
(ii): by linking adequate dipeptide carriers to a drug, a prodrug can be created which
undergoes a strictly chemical cyclization-elimination process via intramolecular
aminolysis of the dipeptide moiety to a DKP, with simultaneous departure of the free
parent drug (Fig. 3.1.8, Gomes, 2003).
Peptide derivatives of some drugs have been prepared and evaluated as prodrug
candidates where prodrug activation processes involved DKP formation. One example
are the peptide conjugates of the cytotoxic agent vinblastine, designed as potential
prodrugs targeted at prostate cancer cells. In vitro and in vivo evaluation showed
the best derivative was a conjugate bearing an octapeptide segment attached by an
ester linkage to position 4 of vinblastine. This conjugate was found to undergo fast
(t1/2 = 12 min) and specific cleavage of the Gln-Ser peptide bond by prostate enzymes,
and additional data from metabolism studies supported that the final spontaneous
vinblastine release was driven by DKP formation from a dipeptidyl intermediate
(Brady, 2002).
O R2
H2 N Parent Drug
N
H
R1 O
R1
NH
HN
R2
Parent Drug
The formation of DKPs was reported as a major degradation pathway for simple alkyl
esters (Larsen, 2004) and dipeptide p-nitroanilides (Goolcharran, 1998). Dipeptides
have been thus proposed as drug carriers to deliver the parent drug through enzyme-
independent processes, namely via DKP formation. Dipeptides are also readily
accessible carriers that can be easily modified to optimize the rate of release of the
parent drug (Shan, 1997). Some authors have therefore considered that dipeptides
might play a crucial role as carriers for hydroxyl-containing drugs. This hypothesis
was developed in a systematic study of the reactivity of paracetamol dipeptide
esters, as this phenolic drug represents an excellent leaving group in the course of
intramolecular dipeptide ester aminolysis (Santos, 2005) and is known to originate
hepatotoxic metabolites (Bertoloni, 2006). Dipeptide esters of paracetamol were
found to be quantitatively hydrolysed to the parent drug and corresponding DKPs at
physiological pH and temperature. Additional evidence that the rate of paracetamol
release depended on the structure of the dipeptide carrier also supported an
intramolecular pathway. Paracetamol esterification with dipeptides also led to
significant decrease or even elimination of the drug’s hepatotoxic effects on mice,
reinforcing the great potential of dipeptide carriers in prodrug design (Santos, 2005).
From what has been described, it is clear that a major drawback of dipeptide-based
prodrugs is their susceptibility to non-specific peptidases. However, this problem can
be easily solved by incorporating at least one non-natural amino acid. For instance,
enzymatically stable dipeptides (e.g. containing α-aminoisobutyric acid or Sar) have
been successfully employed in prodrugs of cytarabine (Wipf, 1996) and cyclosporine
A (CsA) (Hamel, 2004). Dipeptide esters of CsA showed high thermodynamic stability,
differential conversion rates under physiological conditions and strongly increased
water solubility, offering a novel route for the design of CsA prodrugs (Hamel, 2004).
Unnatural amino acids were also used to construct prodrugs of 5-fluorodeoxyuridine
(FdU) to improve prodrug stability against protease/esterase action. Interestingly, these
compounds were designed to be resistant to certain enzymes but to be susceptible to
others. In fact, the antibacterial prodrugs of FdU developed by Wei and Pei are first
activated in vivo by peptide deformylase, after which spontaneous intramolecular
aminolysis of the resulting dipeptide carrier forms a DKP, with simultaneous release
of FdU (Wei, 2000).
O
O-Tyr (L) (2)
R
Peptide transporters appear to be attractive targets in prodrug design for their high
capacity, broad substrate specificity, strong expression in the intestinal epithelium and
low occurrence of functional polymorphisms. Conjugating a specific amino acid with
a drug can make it a substrate for PepT1 to enhance the absorption of the parent drug.
Two excellent examples of marketed prodrugs that exploit carrier-mediated transport
are valacyclovir (Valtrex; GlaxoSmithKline) and valganciclovir (Valcyte; Roche) (Cao,
2012). They are L-Val esters (i.e. valine as the promoiety) of acyclovir and ganciclovir,
which both have limited and variable oral bioavailability owing to their high polarity.
These amino acid prodrugs increased the intestinal permeation of their parent drugs
by 3−10-fold, and their membrane transport was mediated predominantly through
the di- and tripeptide transporter (hPepT1) expressed in intestinal epithelial cells
(Han, 1998; Sugawara, 2000).
Primaquine (PQ, 4, Fig. 3.1.9) is the only generally available anti-malarial that
prevents relapse in vivax and ovale malaria, and is the only potent gametocytocide in
falciparum malaria. Primaquine becomes increasingly important as malaria-endemic
countries move towards elimination, and although it is widely recommended, it is
commonly not given to malaria patients because of haemolytic toxicity in subjects who
are glucose-6-phosphate dehydrogenase (G6PD) deficient (gene frequency typically
3−30% in malaria endemic areas; > 180 different genetic variants) (Ashley, 2014). PQ
is often associated with serious adverse effects because of its toxic metabolites (Vale,
2009). Blocking the terminal primary amine in PQ may represent a huge improvement
in terms of PQ bioavailability, as it can significantly reduce the extent of PQ conversion
into carboxyPQ, the principal metabolite. N-Acylation of anti-malarials with amino
acids and oligopeptides has been used in several works aimed at improving drug
transport into malaria-infected erythrocytes and over the last two decades, these
modifications started to be also seen as a means to avoid premature PQ inactivation
by oxidative deamination to carboxyPQ (Kirk, 2001). An earlier work describes the
synthesis of N-cysteinyl-PQ that was subsequently coupled to carrier proteins, and
both the anti-malarial activity in vivo and the acute lethal toxicity of the protein-drug
conjugates were evaluated. The causal prophylactic activity of the lactosaminated
serum albumin conjugate was two times higher than that of the free drug, whereas
Amino Acids and Drug Development 195
its acute lethal toxicity was at least 6.5-fold lower than that of PQ [mean lethal dose
(LD50) > 85 mg of PQ base kg-1]. This yielded a therapeutic index for the conjugate at
least 12 times higher than that of the free parent drug (Hofsteenge, 1986).
Oligopeptide derivatives of PQ such as PQ-Lys-Leu-d-Val, PQ-Lys-Leu-Ala and
PQ-Lys-Leu-l-Val have been prepared and tested for radical curative anti-malarial
activity against P. cynomolgi in rhesus monkeys and blood-schizontocidal activity
against P. berghei in mice. The d-Val-containing derivative was found to be less
toxic and more active than both its l-Val counterpart and PQ, whereas the activity
of the Ala-bearing compound was comparable to that of the l-Val derivative (Philip,
1988). PQ N-acylation with amino acids or oligopeptides yields structures that are
amenable to proteolytic degradation. Despite this bioreversibility being desirable in
a prodrug, fast conversion into PQ will not overcome the problems associated with
PQ-based therapies. Thus, additional protection of the peptide moiety in peptidyl-
PQ compounds became a new goal so structures could be obtained with resistance
to early proteolytic cleavage and bioreversibility at target tissues or cells to release
the active molecule (Bundgaard and Moss, 1989). Two groups of tetrapeptides with
general formula PQ-Y-Ala-Leu-X-NH2 were also synthesized. In the first group, Leu,
Tyr, Lys and Asp were used in the Y positions and Ala at X. In the second group, Ala,
Tyr, Lys and Asp were used in X positions, while Y was Leu. The peptide derivatives
were then coupled to polyacryl starch microparticles (via the N-terminal amino acid
α-amino group), forming lysosomotropic drug carriers (Borissova, 1995).
Dipeptide derivatives of PQ (PQ-Arg-Phe, PQ-Arg-Lys and PQ-Ala-Phe) were also
synthesized and tested against Chagas disease. PQ-Arg-Lys was seen to be active
against T. cruzi development inside host cells, probably by interfering with the initial
steps of trypomastigote-amastigote transformation. This derivative was more active
than the other two, so it seems that specific cleavage has an important role in PQ release
(Chung, 1997). Portela have also evaluated the dipeptide derivatives of PQ as novel
transmission-blocking anti-malarials. In contrast with PQ, none of the compounds
were able to block oocyst production in mosquitoes’ midguts at 3.75 mg kg-1 but all
were found to completely inhibit the formation of oocysts at 15 mg kg-1, where
N-acetylPQ is not active at this dose. As a whole, this work has demonstrated that
acylation of the PQ’s primary amino group effectively prevents the early conversion of
PQ into carboxyPQ while confirming that the presence of a terminal amino group, as in
the dipeptide derivatives of PQ, is essential for gametocytocidal activity (Portela, 1999).
Gomes worked on the further transformation of the amino acid moiety linked to the PQ
primary amine with the introduction of an imidazolidin-4-one ring at the amino acid’s
α-amino group (5, Fig. 3.1.9). Condensation of the N-aminoacyl derivatives of PQ with
carbonyl compounds (ketones or aldehydes) was suggested by these authors as a means
to protect the amino acid moiety against premature proteolytic degradation (Gomes,
2004). Imidazolidin-4-ones were highly stable in human plasma, with a weak or null
degree of PQ release after 3 days of incubation. Moreover, compound 5 hydrolyzed 50-
100 times slower in aqueous buffer at physiological pH and T than the corresponding
196 Amino Acids and Peptides in Medicine: Old or New Drugs?
R4
O
N N O
NH
HN HN N
NH 2 N
H R3
R2
R1
4 6
O
O
N O
N O
HN
N HN
R1
N
R1
R2
NH
R2
N
5 R3 NH 2
R3
7
O
R4
Figure 3.1.9: Structures do PQ and its imidazolidin-4-ones derived from L-amino acids and carbonyl
compounds.
Peptides for Biomedicine 197
To fulfil their therapeutic function, most drug-like compounds must pass through the
cellular membrane to enter a cell. Small molecules that obey Lipinski’s rule of five
(no more than 5 hydrogen bond donors, not more than 10 hydrogen bond acceptors,
a molecular mass less than 500 Da and an octanol-water partition coefficient not
greater than 5) have a reasonable chance of penetrating the cell membrane. However,
an increasing number of therapeutics, such as biologics, cannot be considered
‘small molecules’ and these therapeutic are often unable to enter the cell unassisted
(Liskamp, 2014). One of the vehicles used to facilitate the passage of these compounds
into cells are peptides.
Peptides and proteins play a central role in numerous biological and physiological
processes in living organisms: they are involved as hormones and neurotransmitters
in intercellular communication, act as antibodies in the immune system to protect
organisms against foreign invaders, and are also involved in the transport of various
substances through biological membranes. Peptides have significant advantages over
other small molecules in terms of specificity/affinity for targets and toxicity profiles,
and over antibodies in terms of tissue penetration owing to their smaller size. A large
number of peptide-based drugs are now being marketed and the number of candidates
entering clinical evaluation in recent years is steadily increasing. Therefore, the
synthesis of such structures has been a major focus of organic chemistry for over a
century in order to improve the prospects for synthetic therapeutic peptides. In this
section, the relevance of peptides in biomedicine will be discussed.
198 Amino Acids and Peptides in Medicine: Old or New Drugs?
Parasitic diseases are most common in tropical regions with poverty and undeveloped
health care systems. Many parasitic diseases have been out of focus on national and
international agendas and increasing drug resistance combined a lack of vaccines in
several countries make this very alarming. To combat this reality, it is imperative to
understand the relationship between insect and host associated to parasitic disease.
Insects are widely distributed and developed in various ecological niches and serve
as vectors of many parasitic diseases in humans and animals, suggesting outstanding
strategies for defending against pathogens such as microorganisms by overcoming and
adapting to different environmental conditions (Lowenberger, 2001). After infection,
the parasites are presented to the host immune system and induce the production of
defence compounds such as proteins and peptides. Several of these humoral response
peptides exert antibacterial, antifungal, or antiviral properties (Bulet, 1999) and are
known as small cationic peptides, host defense peptides or antimicrobial peptides
(AMPs) (Bell, 2011).
AMPs mostly contain 15−45 amino acid residues and are generally cationic at
physiological pH, often with an amphipathic character and encoded by separated
genes. They form the first line of host defence against pathogenic infections and are
a key component of the ancient innate immune system. AMPs have been identified in
various species ranging from bacteria and frogs to mammals, including humans. In
insects, AMPs are synthesized in the fat body, blood cells (hemocytes) or epithelia and
are released into the hemolymph, the insect blood. In vertebrates, AMPs are present
in amphibian skin secretions (Simmaco, 1999) and epithelia (Ganz & Weiss, 1997;
Bals, 1998). In mammals, they were also observed in lymphocytes (Agerberth, 2000)
and leukocytes (Sorensen, 1997). Because of their broad activity against microbes and
expression triggered by various infections, AMPs are currently intensely examined as
potential antiparasitic compounds (Vale, 2014).
In 2004, the antimicrobial peptide database (APD, http://aps.unmc.edu/AP/
main.php) reported a significant number of AMPs that have been discovered at both
the gene and protein levels (Wang Z & Wang G, 2003). The APD was later updated
and expanded to allow users to search peptide families (bacteriocins, cyclotides, or
defensins), peptide sources (fish, frogs or chicken), post-translationally modified
peptides (amidation, oxidation, lipidation, glycosylation or D- amino acids), and
peptide binding targets (membranes, proteins, DNA/RNA, LPS or sugars) (Wang,
2009).
The most prominent models are the carpet mechanism, the barrel-stave model, and the
toroidal pore model. Overall, the mechanism of interaction of AMPs with membranes
results either in the translocation of peptides into the cell, the formation of pores
embedded in the membrane to enable a flux of ions and molecules, or in membrane
permeabilization/disruption (Fig. 3.1.10).
A
D
The potency and range of the antimicrobial activity of AMPs vary depending on
various parameters, such as the amino acid composition of the structure and its
stability under different environmental conditions, as well as on parameters of the
target cell such as membrane composition and structural features (Fig. 3.1.11). As the
200 Amino Acids and Peptides in Medicine: Old or New Drugs?
In current practice, it is known that AMPs not only target cell membranes but
may have intracellular targets. The AMP buforin II kills E. coli without lysis of the
cell membrane. Buforin II penetrates the cell membrane, accumulates inside the
bacterial cell and binds to the DNA and RNA, which leads to cell death by inhibiting
cellular functions (Park, 1998). Pleurocidin inhibits DNA and protein synthesis in
E. coli without damaging the cytoplasmic membrane (Patrzykat, 2002). Other AMPs
inhibit enzymatic activities inside the target cell such as phyrrhocoricin, apidaecin,
and drosocin, which interact with the E. coli heat shock protein DnaK and inhibit
protein folding (Kragol, 2001). Some AMPs are capable of inducing apoptotic cell
death, including the breakdown of mitochondrial membrane potential and activation
of caspase 3-like activity, such as in Leishmania (Kulkarni, 2006). AMPs may have
different target sites and act in more than one mechanism to kill the same species, and
different AMPs may act synergistically (Fieck, 2010).
According Boman, AMPs can be classified into three groups: a) linear, often
α-helical peptides free of cysteine residues; b) peptides containing disulfide bridges,
giving peptides a β-sheet structure; and c) peptides with an overrepresentation in
certain amino acids, such as proline, arginine, tryptophan, or histidine (Boman,
2003).
3.1.3.1.4 Peptides Rich in Pro, Gly, His, Arg and Trp Residues
This group includes AMPs such as apidaecins, short-chain Pro-rich peptides that
may adopt a polyPro helical type II structure, possibly forming the structural basis
to bind specific targets and confer antibacterial activity (Li, 2006). The mechanism
of membrane action does not include the formation of pores but is energy-driven,
resulting in a transporter-mediated model (Castle, 1999).
As for the Pro-rich antimicrobial peptide class, the Gly rich peptides have variable
sizes and do not show clear sequence consensus, apart from the high proportion of
Gly in their primary sequence. These peptides are in general longer than AMP from
other classes and between 25 to 50% of their amino acids are glycine. They have
disordered structure in water but tend to self-order when in contact with artificial
membranes (Bruston, 2007). Attacins are a group of six glycine-rich AMPs and can
be grouped into four basic (A-D) and two acidic (E-F) peptides, probably derived
from two attacin genes (Yi, 2014). Attacin inhibits the synthesis of outer membrane
proteins of E. coli by blocking the transcription of the respective genes (Carlsson,
1991). This is presumably achieved by an indirect mechanism, where attacin binds
to lipopolysaccharides (which serve as a receptor for attacin) but does not enter the
bacterial cell (Carlsson, 1998).
The archetypical Trp rich peptide is indolicin, which unlike the amphipathic alpha
helical structure of the cecropin class of peptides, has a linear structure (no disulfide
Peptides for Biomedicine 203
bridges) and no particular secondary structure in water. Some authors suggest that
structural changes and strong membrane affinity are key to the antimicrobial activity
of indolicin (Ladokhin & White, 2001). Trp-rich AMP sequences contain more than
25% of Trp. This peptide has the ability to permeate bacterial membranes and,
depending on its three dimensional shape, inhibits DNA synthesis by binding to its
strand (Hsu, 2005).
His-rich amphipathic cationic peptides are peptides with 25% of their amino acids
represented by His. They show a global cationic amphipathic helical structure. They
trigger microorganism membrane disruption when the peptide adopts an alignment
parallel to the membrane surface. However, pore formation is not essential for their
high antimicrobial activity (Mason, 2009). Clavinin (van Kan, 2002) and daptomycin
(Jeu & Fung, 2004) are two studied members of this antimicrobial peptide class.
X Y
X Y
H H2
a) S S b) C N c) C S
O O
H H H
d) N C N e) C O f) C N N
H
Figure 3.1.12: The structure of a drug-linker-peptide (CPP) conjugate developed in our research
group. X and Y represent the common functional groups used to connect either the drug or the
peptide to the linker. X may be similar to or different than Y. Here, the nature of the X bond and the
drug peptide conjugation chemistry has been classified according to the nature of the X bond: a)
disulfide bond, b) amide, c) thioether, d) carbamate ester, e) carboxylic acid ester or f) hydrazone
bond.
Sequences of common CPPs are provided in Table 3.1.2. The common virtue of all CPPs
is the ability to efficiently pass through cell membranes while carrying a wide variety
of cargos inside cells (Fig. 3.1.13) (Capolovici, 2014). Interestingly, CPP sequences are
known to vary considerably as seen by examining Table 3.1.2. There are, however,
several similarities between the structural natures of these short peptides. Almost
every CPP sequence involves positively charged amino acids. In fact, a chain of
arginines form one of the most widely used CPPs (Myrberg, 2008). The membranolytic
properties of a given CPP can also be governed by its secondary structure, specifically
its helicity. It has been shown that peptides with an α-helical region can enter cells
more efficiently.
Peptides for Biomedicine 205
(Derossi, 1994)
MAP KLALKLALKALKAALKLAa Amphipathic model peptide
(Oehlke, 1998)
Transportan/ GWTLNS/ Galanin-Lys-mastoparan
TP10 AGYLLGKINLKALAALAKKILa
(Futaki, 2001)
MPG GALFLGFLGAAGSTMGAb Hydrophobic domain from the fusion
sequence of HIV gp41 and NLS of SV40
(Morris, 1997)
T-antigen
Pep1 KETWWETWWTEWSQPKKKRKVb NLS from Simian Virus 40 large T antigen
and reverse transcriptase of HIV-1
(Chaloin, 1998)
pVEC LLIILRRRIRKQAHAHSKa VE-cadherin
(Säälik, 2004)
YTA2 YTAIAWVKAFIRKLRKa MMP cleavage site as seeding sequence
(Lindgreen, 2006)
YTA4 IAWVKAFIRKLRKGPLGa MMP cleavage site as seeding sequence
(Lindgreen, 2006)
M918 MVTVLFRRLRIRRACGPPRVRVa Tumor suppressor protein p14ARF
(El-Andaloussi, 2007)
CADY GLWRALWRLLRSLWRLLWRAb Derived from PPTG1 peptide, W and
charged amino acids
(Crombez, 2007)
a
C-terminal amide. b C-terminal cysteamide.
206 Amino Acids and Peptides in Medicine: Old or New Drugs?
Initially, CPPs were composed of only natural amino acids but recently have included
non-primary and even unnatural amino acids (Farrera-Sinfreu, 2005). Furthermore,
when lysine residues are replaced with ornithine residues, the peptide becomes more
resistant to cellular degradation (Ezzat, 2011; Angeles-Boza, 2010). Cargo delivery
efficiency can also be improved by changing the structure of the peptides, such as
modification into dendrimers (Angeles-Boza, 2010; Bode, 2014), cyclic peptides
(Mandal, 2011; Northfield, 2014), and the often-used strategy of modifying the side
chains of CPPs, such as in TP10 derivatives (Andaloussi, 2011; Oskolkov, 2011).
Recently, Vale’s research group developed new CPP-Gemcitabine conjugates for drug
delivery in cancer therapy (Vale, 2016).
Further investigations into the structure and mechanisms of uptake of CPP-cargo
complexes will be required to evaluate CPPs as potential delivery tools for biomolecules
in vitro and in vivo models. However, considering evidence from numerous studies,
CPPs have the potential to become a universal tool to carry therapeutic molecules
across cellular membranes without a risk of toxicity or inflammatory reactions.
A B
synthetic polymers to improve potency and specificity (Vlieghe, 2010). In this context,
peptides can be used as targeting moieties, as carriers to provide transport across cellular
membranes, and to modify the bioactivity of the original compound/material. In the field
of biomaterials, peptides have also been extensively used as cell-instructive motifs with
different roles, namely to promote cell-adhesion to otherwise non-adhesive polymers,
as well as proliferation and differentiation (Collier & Segma, 2011; Wojtowicz, 2010).
Finally, peptides by themselves are paving the way as new biomaterials. This is the case
for a new class of materials named self-assembling peptides (SAPs), which self-assemble
into nanofiber-like hydrogen bond networks under physiological concentrations of salt
solutions and have found numerous applications as three-dimensional matrices in the
biomedical field (Collier & Segma, 2011; Matson & Stupp, 2012).
processing and sterilization steps required for biomaterials synthesis, many of which
cause protein denaturation. The use of RGD, as compared with native ECM proteins,
also minimizes the risk of immune reactivity or pathogen transfer. Another benefit
is that the synthesis of RGD peptides is relatively simple and inexpensive, which
facilitates translation into the clinic (Bellis, 2011).
β-Hairpin peptides β-Hairpin secondary structures (two β-strands linked by a kink) can
(Loo, 2013) be rationally designed to self-assemble into ibrillary macromolecular
scaffolds
α-helical coiled coil α-helices result from the hydrogen bonding of a peptide backbone amide
(Stephanopoulos, 2013) hydrogen to a backbone carbonyl four residues away. This motif results
in a right-handed helical twist configuration. The formation of α-helices is
determined by the amino acid sequence and is dependent on hydrophilic
and hydrophobic amino acid residues organizing into a hydrophobic and
a hydrophilic face. The gelation can be controlled using heat to provide
a versatile scaffold for cell culture, especially when gel dissolution at
elevated temperatures is desired. The α-helical peptide gels proved
highly cytocompatible, and PC12 cells cultured in them were able to
proliferate and differentiate
The obstacles that remain for the development of peptides as therapeutics are
significant. As in all drug development, in vivo efficacy might differ from promising
in vitro models. The reason can be the result of interactions with many in vivo
components in the circulation, gastrointestinal tract, dermis, relative clearance and
affinity to target receptors (Lin & Lowman, 2003).
The main limitations generally attributed to therapeutic peptides are (i) oral
bioavailability (injection is normally required), (ii) short half-life because of rapid
degradation by proteolytic enzymes of the digestive system and blood plasma, (iii)
rapid removal from the circulation by the liver or kidneys, and (iv) high synthetic and
production costs (Bray, 2003; Picherean & Allary, 2005).
However, therapeutic peptides are becoming increasingly attractive for the
discovery and development of new generations of drugs. With the recent addition
of new methodologies, peptides can be engineered to have long plasma half-
lives and low immunogenicity (Salo, 2006), along with alternative routes of
administration. Production of synthetic therapeutic peptides has become possible for
the pharmaceutical industry with recent developments of SPPS, initially developed
by Merrifield. SPPS is crucial in the early steps of preclinical research and in the
production of peptide-based active pharmaceutical ingredients (APIs) (Bruckdorfer,
2004). SPPS is especially suited for medium-sized peptides (up to 80 amino acids
residues), which comprise the majority of therapeutically-relevant peptides.
Table 3.1.4: Lead peptide chemical optimization (adapted from Vlieghe, 2010).
Lead Attribute
3 Simplification and/or optimization of the structure after alanine scanning (Ala-scan) and/
or D-scanning (D-scan) to eliminate potential sites of cleavage (notably by endopeptidases)
and to determine important functional groups involved in the interaction with the target of
interest
4 Isosteric, or not, amide bond replacement between two amino acids: NH-amide alkylation,
the carbonyl function of the peptide bond can be replaced by CH2 (reduced bond: -CH2-NH-),
C(S) (endothio-peptide, -C(S)-NH-) or PO2H (phosphonamide, -P(O)OH-NH-). NH-amide bond
can be exchanged by O (depsipeptide, -CO-O-), S (thioester, -CO-S-) or CH2 (ketomethylene,
-CO-CH2-). The peptide bond can also be modified: retro-inverso bond (-NH-CO-), methylene-
oxy bond (-CH2-), thiomethylene bond (-CH2-S-), carba bond (-CH2-CH2-), hydroxyethylene
bond (-CHOH-CH2-) and so on, to increase plasma stability of the peptide sequence (notably
towards endopeptidases)
5 Cyclization of the peptide sequence (between side chains or ends of the peptide sequence:
head to tail, N-backbone to N-backbone, end to N-backbone, end to side chain, side chain to
N-backbone, side chain to side chain) through disulfides (disulfide-bond cyclization scan),
lanthionine, dicarba, hydrazine or lactam bridges, to decrease the conformational flexibility
of linear peptides and reduce hydrogen bonding, enhance membrane permeability, and
most importantly to increase their stability against proteolysis by endo- and exopeptidases
6 Substitution of a natural amino acid residue by an unnatural amino acid (D-configuration),
an N-methyl-α-amino acid, a non-proteogenic constrained amino acid or a β-amino acid, to
increase plasma stability (e.g. resistance to endopeptidases) of the peptide and/or affinity
(activity) for its target
7 Deletion of one or more consecutive amino acid(s) and combinatorial deletion with two or
more positions omitted independently throughout the sequence
Alternative routes for the administration of peptide-based drugs have been improved
in recent years and novel peptide delivery technologies have emerged, including
controlled-release parental drug, mucosal route, oral route and transdermal route
(Pettit & Gombotz, 1998). In spite of some limitations, synthetic therapeutic peptides
present numerous advantages compared with their homologous compounds (proteins
Peptides for Biomedicine 213
and antibodies) and with small organic molecules, thanks to progress in chemical
synthesis and routes of administration (Vlieghe, 2010). In Table 3.1.5, the principal
advantages of peptides over proteins/antibodies or over small organic molecules that
make up traditional medicines are presented.
Table 3.1.5: Advantages of peptides over other drug candidates as proteins/antibodies or organic
molecules (Loffet, 2002; Lien & Lowman, 2003; Witt & Davis, 2006; Vlieghe, 2010; McGregor, 2008;
Hummel, 2006).
3.1.3.4.2 Market
The USA and EU are the major markets for drugs of all kinds, so first approvals for
peptide therapeutics have occurred primarily in one of these two regions. All 19
peptides approved in the USA during the period 2001 to 2012 were first approved in
either the USA or EU. The notable expansion of peptide therapeutic development in
the late 1990s and 2000s led to an unprecedented number of marketing approvals
in 2012, and has provided a robust pipeline that should deliver numerous approvals
during the remainder of the 2010 (Kaspar & Reichert, 2013).
In the US, annual sales of peptide drugs exceed 13 billion, representing 1.5% of
drug sales globally. Protein drugs such as therapeutic antibodies also represent a
larger share with the combined biopharmaceutical market valued at over 70 billion. In
Europe, Germany and the UK account for 63% of the peptide therapeutic market with
France, Italy, Scandinavia and Spain making up the rest of the major users (Hervé,
2008).
214 Amino Acids and Peptides in Medicine: Old or New Drugs?
Figure 3.1.14: Therapeutic areas for peptides in clinical studies (on 15 February 2013; CNS-central
nervous system; CV-cardiovascular; adapted from (Kaspar & Reichert, 2013).
The commercial value of peptide therapeutics that have been marketed for years is
well-established, but could substantially increase as recently approved products and
those on the horizon gain market share. Peptides will probably be used intensively
in the near future for various applications in the treatment of CNS diseases, notably
in the design of peptide regulators of protein activity, if they are able across the BBB
by absorptive-mediated transport (Hervé, 2008; Pardridge, 2003). CMT- or RMT-based
peptides (de Boer & Gaillard, 2007a; 2007b) developed for CNS drug targeting will also
certainly contribute to the development of peptide-based prodrugs with facilitated
Conclusions 215
access to the CNS. Also, in current trials conducted to treat cancer, gene disorders, and
autoimmune diseases, CPP-based therapies are being used. Gene-targeted therapies
have a huge potential to fight rare genomic diseases and will be associated to the
usage of CPP to deliver therapeutic agents into cells (Figueiredo, 2014).
3.1.4 Conclusions
as they are later removed by washing and filtration of the peptidyl-resin contained in
the vessel, which is usually attached to a vacuum filtration system.
The progression of peptide compounds into clinical therapy goes along with
the increasing need for new therapeutic strategies due to the limitations of current
biomaterials for engineering complex tissues. The chemical synthesis of peptides
enables the conjugation of other small molecules or incorporation of non-natural
amino acids by design. Conjugation of small molecules to a peptide opens up the
possibility for greater chemical diversity, analogous to small-molecule medicinal
chemistry approaches for developing high-affinity, high-specificity molecular
recognition increasing the range of potential applications of these molecules in the
growing area of tissue engineering.
Although the interaction of CPPs with cellular membranes often determines their
uptake efficiency, CPPs are first and foremost delivery vectors. This means that they
must efficiently deliver various cargo molecules to their designated location, whether
it is in the cytoplasm or nucleus. CPPs were discovered 20 years ago based on the
potency of several proteins to enter cells. In recent years CPPs have been associated
with biodistribution and efficacy of oral biodrug delivery. Nevertheless, the recent
progress of CPPs as new carriers for intracellular cargo delivery were focused in
tumor cells, where only small quantity of drugs were used as cargo. We propose that
new methodologies will be used to design and develop CPP-drug complexes, in that
CPPs transport their cargo inside cells using endocytosis. Each modification will be
carefully designed to avoid problems that could lead to low synthetic yields, poor
solubility, aggregation or toxicity. The in vivo delivery of drugs (as cargo) requires a
longer drug circulation time, which in turn requires a more stable complex between
the CPP and its drug, as it introduces different hydrophilic groups to an already cell-
permeable peptide. When using CPPs as a delivery system, one must consider that
drug delivery must often be highly specific.
Chemical strategies can improve the in vivo plasma residence time and, after that,
introduce new peptides as drugs in the market. This way, the commercial value of
peptides therapeutic will be well-established.
References
Agerberth, B.; Charo, J.; Werr, J.; Olsson, B.; Idali, F.; Lindbom, L.; Kiessling, R.; Jornvall, H.; Wigzell,
H.; Gudmundsson, G. H. (2000). The human antimicrobial and chemotactic peptides ll-37 and
alpha-defensins are expressed by specific lymphocyte and monocyte populations. Blood, 96,
3086-3093.
Aggeli, A.; Bell, M.; Carrick, L. M.; Fishwick, C. W.; Harding, R.; Mawer, P. J.; Radford, S. E.; Strong, A.
E.; Boden, N. (2003). pH as a trigger of peptide β-sheet self-assembly and reversible switching
between nematic and isotropic phases. Journal of the American Chemical Society, 125, 9619-
9628.
Aizawa, Y.; Owen, S. C.; Shoichet, M. S. (2012). Polymers used to influence cell fate in 3D geometry:
new trends. Progress in Polymer Science, 37, 645-658.
Andaloussi, S. E.; Lehto,T.; Mäger, I.; Rosenthal-Aizman, K.; Oprea, I. I.; Simonson, O. E.; Sork, H.;
Ezzat, K.; Copolovici, D. M.; Kurrikoff, K.; Viola, J. R.; Zaghloul, E. M.; Sillard, R.; Johansson,
H. J.; Hassane, F. S.; Guterstam, P.; Suhorutsenko, J.; Moreno, P. M.; Oskolkov, N.; Hälldin, J.;
Tedebark, U.; Metspalu, A.; Lebleu, B.; Lehtiö, J.; Smith, C. I.; Langel, Ü. (2011). Design of a
peptide-based vector, PepFect6, for efficient delivery of siRNA in cell culture and systemically in
vivo. Nucleic Acids Research, 39, 3972-3987.
Angeles-Boza, A. M.; Erazo-Oliveras, A.; Lee, Y. J.; Pellois, J. P. (2010). Beneration of endosomolytic
reagents by branching of cell-penetrating peptides: tools for the delivery of bioactive
compounds to live cells in cis or trans. Bioconjugate Chemistry, 21, 2164-2167.
Araújo, M. J.; Bom, J.; Capela, R.; Casimiro, C.; Chambel, P.; Gomes, P.; Iley, J.; Lopes, F.; Morais, J.;
Moreira, R.; Oliveira, E.; Rosário, V.; Vale, N. (2005). Imidazolidin-4-one derivatives of primaquine
as novel transmission-blocking antimalarials. Journal of Medicinal Chemistry, 48, 888-892.
Ashley, E. A.; Recht, J.; White, N. J. (2014). Primaquine: the risks and the benefits. Malaria Journal,
13, 418-424.
Bals, R.; Wang, X.; Zasloff, M.; Wilson, J. M. (1998). The peptide antibiotic ll-37/hcap-18 is expressed
in epithelia of the human lung where it has broad antimicrobial activity at the airway surface.
Proceedings of the National Academy of Sciences of USA, 95, 9541-9546.
Beevers, A. J.; Dixon, A. M. (2010). Helical membrane peptides to modulate cell function. Chemical
Society Reviews, 39, 2146-2157.
Bell, A. (2011). Antimalarial peptides: the long and the short of it. Current Pharmaceutical Design, 17,
2719-2731.
Bellis, S. L. (2011). Advantages of RGD peptides for directing cell association with biomaterials.
Biomaterials, 32, 4205-4210.
Bertoloni, A.; Ferrari, A.; Ottani, A.; Guerzoni, S.; Tacchi, R.; Leone, S. (2006). Paracetamol: new
vistas of an old drug. CNS Drug Reviews, 12, 250-275.
Birolo, L.; Sandmeier, E.; Christen, P.; John, R. A. (1995). The roles of Tyr70 and Tyr225 in aspartate
aminotransferase assessed by analysing the effects of mutations on the multiple reactions of
the subtrate analogue serine o-sulphate. European Journal of Biochemistry, 232, 859-864.
Boer, A. G.; Gaillard, P. J. (2007a). Drug targeting to the brain. Annual Review of Pharmacology and
Toxicology, 47, 323-355.
Boer, A. G.; Gaillard, P. J. (2007b). Strategies to improve drug delivery across the blood-brain barrier.
Clinical Pharmacokinetics, 46, 553-576.
Bode, S. A.; Wallbrecher, R.; Brock, R.; van Hest, J. C. M.; Löwik, D. W. P. M. (2014). Activation of
cell-penetrating peptides by disulfide bridge formation of truncated precursors. Chemical
Communications, 50, 415-417.
Boman H. G. (2003). Antibacterial peptides: basic facts and emerging concepts. Journal of Internal
Medicine, 254, 197-215.
218 Amino Acids and Peptides in Medicine: Old or New Drugs?
Borissova, R.; Lammek, B.; Stjärnkvist, P.; Sjöholm, I. (1995). Biodegradable microspheres. 16.
Synthesis of primaquine–peptide spacers for lysosomal release from starch microparticles.
Journal of Pharmaceutical Sciences, 84, 249-255.
Bradshaw, T. D.; Bibby, M. C.; Double, J. A.; Fichtner, I.; Cooper, P. A.; Alley, M. C.; Donohue, S.;
Stinson, S. F.; Tomaszewjski, J. E.; Sausville, E. A.; Stevens, M. F. (2002). Preclinical evaluation
of amino acid prodrugs of novel antitumor 2-(4-amino-3-methylphenyl)benzothiazoles.
Molecular Cancer and Therapy, 1, 239-246.
Brady, S. F.; Pawluczyk, J. M.; Lumma, P. K.; Feng, D. M.; Wai, J. M.; Jones, R.; DeFeo-Jones, D.; Wong,
B. K.; Miller-Stein, C.; Lin, J. H.; Oliff, A.; Freidinger, R. M.; Garsky, V. M. (2002). Design and
synsthesis of a pro-drug of vinblastine targeted at treatment of prostate cancer with enhanced
efficacy and reduced systemic toxicity. Journal of Medicinal Chemistry, 45, 4706-4715.
Bray, B. L. (2003). Large-scale manufacture of peptide therapeutic by chemical synthesis. Nature
Reviews Drug Discover, 2, 587-293.
Bulet, P.; Hetru, C.; Dimarcq, J. L.; Hoffmann, D. (1999). Antimicrobial peptides in insects; structure
and function. Developmental and Comparative Immunology, 23, 329-344.
Bundgaard, H.; Larsen, C.; Thorbek, P. (1984). Prodrugs as drug delivery systems XXVI. Preparation
and enzymatic hydrolysis of various water-soluble amino acid esters of metronidazole.
International Journal of Pharmaceutics, 18, 67-77.
Bundgaard, H.; Moss, J. (1989). Prodrugs of peptides IV: Bioreversible derivatization of the
pyroglutamyl group by N-acylation and N-aminomethylation to effect protection against
pyroglutamyl aminopeptidase. Journal of Pharmaceutical Sciences, 78, 122-126.
Burnette, T. C.; Harrington, J. A.; Reardon, J. E.; Merrill, B. M.; Miranda, P. (1995). Purification and
characterization of a rat liver enzyme that hydrolyzes valaciclovir, the L-valyl ester prodrug of
acyclovir. Journal of Biological Chemistry, 270, 15827-15831.
Bruckdorfer, T.; Marder, O.; Albericio, F. (2004). From production of peptides in milligram
amounts for research to multi-tons quantities for drugs of the future. Current Pharmaceutical
Biotechnology, 5, 29-43.
Bruston, F.; Lacombe, C.; Zimmermann, K.; Piesse, C.; Nicolas, P.; El Amri, C. (2007). Structural
malleability of plasticins: preorganized conformations in solution and relevance for
antimicrobial activity. Biopolymers, 86, 42-56.
Cao, F.; Jinghao, J.; Yin, Z.; Gao, Y.; Sha, L.; Lai, Y.; Ping, Q.; Zhang, Y. (2012). Ethylene glycol-
linked amino acid diester prodrugs of oleanolic acid for PepT1-mediated transport: tynthesis,
intestinal permeability and pharmacokinetics. Molecular Pharmaceutics, 9, 2127-2135.
Carlsson, A.; Engstrom, P.; Palva, E. T.; Bennich, H. (1991). Attacin, an antibacterial protein from
Hyalophora cecropia, inhibits synthesis of outer membrane proteins in Escherichia coli by
interfering with omp gene transcription. Infection and Immunity, 59, 3040-3045.
Carlsson, A.; Nystrom, T.; de Cock, H.; Bennich, H. (1998). Attacin - an insect immune protein-
binds lps and triggers the specific inhibition of bacterial outer-membrane protein synthesis.
Microbiology, 144, 2179-2188.
Castle, M.; Nazarian, A.; Yi, S. S.; Tempst, P. (1999). Lethal effects of apidaecin on Escherichia coli
involve sequential molecular interactions with diverse targets. Journal of Biological Chemistry,
274, 32555-32564.
Cha, C.; Liechty, W. B.; Khademhosseini, A.; Peppas, N. A. (2012). Designing biomaterials to direct
stem cell fate. ACS Nano, 6, 9353-9358.
Chaloin, L.; Vidal, P.; Lory, P.; Mery, J.; Lautredou, N.; Divita, G.; Heitz, F. (1998). Design of carrier
peptide-oligonucleotide conjugates with rapid membrane translocation and nuclear
localization properties. Biochemical and Biophysical Communications, 243, 601-608.
Chan, G.; Mooney, D. J. (2008). New materials for tissue engineering: towards greater control over
the biological response. Trends in Biotechnology, 26, 382-392.
References 219
Chan, O. H.; Schmid, H. L.; Stilgenbauer, L. A.; Howson, W.; Horwell, D. C.; Stewart, B. H. (1998).
Evaluation of a targeted prodrug strategy of enhance oral absorption of poorly water-soluble
compounds. Pharmaceutical Research, 15, 1012-1018.
Chambel, P.; Capela, R.; Lopes, F.; Iley, J.; Morais, J.; Gouveia, L.; Gomes, J. R. G.; Gomes, P.; Moreira,
R. (2006). Reactivity of imidazolidin-4-one derivatives of primaquine: implications for prodrug
design. Tetrahedron, 62, 9883-9891.
Chau, Y.; Tan, F. E.; Langer, R. (2004). Synthesis and Characterization of
Dextran−Peptide−Methotrexate Conjugates for Tumor Targeting via Mediation by Matrix
Metalloproteinase II and Matrix Metalloproteinase IX. Bioconjugate Chemistry, 15, 931-941.
Christensen, B.; Flink, J.; Merrifield, R. B.; Mauzerall, D. (1988). Channel-forming properties
of cecropins and related model compounds incorporated into planar lipid membranes.
Proceedings of the National Academy of Sciences of USA, 85, 5072-5076.
Chung, M. C.; Gonçalves, M. F.; Colli, W.; Ferreira, E. I.; Miranda, M. T. M. (1997). Synthesis and in
vitro evaluation of potential antichagasic dipeptide prodrugs of primaquine. Journal of
Pharmaceutical Sciences, 86, 1127-1131.
Collier, J. H.; Segura, T. (2011). Evolving the use of peptides as components of biomaterials.
Biomaterials, 32, 4198-4204.
Copolovici, D. M.; Langel, K.; Eriste, E.; Langel, Ü. (2014). Cell-penetrating peptides: design,
synthesis and applications. ACS Nano, 8, 1972-1994.
Cornford, E. M.; Young, D.; Paxton, J. W.; Finlay, G. J.; Wilson, W. R.; Pardridge, W. M. (1992).
Melphalan penetration of the blood-brain barrier via the neutral amino acid transporter in
tumor-bearing brain. Cancer Research, 52, 138-143.
Cui, H.; Webber, M. J.; Stupp, S. I. (2010). Self‐assembly of peptide amphiphiles: From molecules to
nanostructures to biomaterials. Peptide Science, 94, 1-18.
Cundy, K. C.; Branch, R.; Chernov-Rogan, T.; Dias, T.; Estrada, T.; Hold, K.; Koller, K.; Liu, X.; Mann,
A.; Panuwat, M.; Raillard, S. P.; Upadhyay, S.; Wu, Q. Q.; Xiang, J. N.; Yan, H.; Zerangue, N.;
Zhou, C. X.; Barrett, R. W.; Gallop, M. A. (2004). XP13512 [(+/-)-1-([(alphaisobutanoyloxyethox
y) carbonyl] aminomethyl)-1-cyclohexane acetic acid], a novel gabapentin prodrug: I. Design,
synthesis, enzymatic conversion to gabapentin, and transport by intestinal solute transporters.
Journal of Pharmacology and Experimental Therapeutics, 311, 315-323.
Cuthbertson, B. J.; Bullesbach E. E.; Fievet, J.; Bachere, E.; Gross, P. S. (2004). A new class
(penaeidin class 4) of antimicrobial peptides from the Atlantic white shrimp (Litopenaeus
setiferus) exhibits target specificity and an independent proline-richdomain function.
Biochemical Journal, 381, 79-86.
Curis, E.; Crenn, P.; Cynober, L. (2007). Citrulline and the gut. Current Opinion in Clinical Nutrition
and Metabolic Care 10, 620-626.
Crombez, L.; Aldrian-Herrada, G.; Konate, K.; Nguyen, Q. N.; McMaster, G. K.; Brasseur, R.; Heitz,
F.; Divita, G. (2009). A new potent secondary amphipathic cell-penetrating peptide for siRNA
delivery into mammalian cells. Molecular Therapy, 17, 95-103.
Derossi, D.; Joliot, A. H.; Chassaing, G.; Prochiantz, A. (1994). The third helix of the Antennapedia
homeodomain translocates through biological membranes. Journal of Biological Chemistry,
269, 10444-10450.
Destoumieux, D.; Munoz, M.; Bulet, P.; Bachere, E. (2000). Penaeidins, a family of antimicrobial
peptides from penaeid shrimp (Crunstacea, Decapoda). Cellular and Molecular Life Sciences,
57, 1260-1271.
Duro-Castano, A.; Conejos-Sánchez, I.; Vicent, M. J. (2014). Peptide-Based Polymer Therapeutics.
Polymers, 6, 515-551.
Ekengren, S.; Hultmark, D. (1999). Drosophila cecropin as an antifungal agent. Insect Biochemistry
and Molecular Biology, 29, 965-972.
220 Amino Acids and Peptides in Medicine: Old or New Drugs?
El-Andaloussi, S.; Johansson, H. J.; Holm, T.; Langel, U. (2007). A novel cell-penetrating peptide,
M918, for efficient delivery of proteins and peptide nucleic acids. Molecular Therapy, 15, 1820-
1826.
Elliott, G.; O’Hare, P. (1997). Intercellular trafficking and protein delivery by a herpesvirus structural
protein. Cell 88, 223-233.
Ezzat, K.; Andaloussi, S. E. L.; Zaghloul, E. M.; Lehto, T.; Lindber, S.; Moreno, P. M. D.; Viola, J. R.;
Magdy, T.; Abdo, R.; Guterstam, P.; Sillard, R.; Hammond, S. M.; Wood, M. J. A.; Arzumanov,
A. A.; Gait, M. J.; Smith, Ci. I. E.; Hällbrink, M.; Langel, Ü. (2011). pepFecr 14, a novel cell-
penetrating peptide for oligonucleotide delivery in solution and as solid formulation. Nucleic
Acids Research, 39, 5284-5298.
Farrera-Sinfreu, J.; Giralt, E.; Castel, S.; Albericio, F.; Royo, M. (2005). Cell-penetrating cis-γ-amino-L-
proline-derived peptides. Journal of the American Chemical Society, 127, 9459-9468.
Fieck, A.; Hurwitz, I.; Kang, A. S.; Durvasula, R. (2010). Trypanosoma cruzi: Synergistic cytotoxicity
of multiple amphipathic anti-microbial peptides to T. cruzi and potential bacterial hosts.
Experimental Parasitology, 125, 342-347.
Figueiredo, I. R.; Freire, J. M.; Flores, L.; Veiga, A. S. (2014). Castanho, M. A. R. B. Cell-penetrating
peptides: a tool for effective delivery in gene-targeted therapies. IUBMB Life, 66, 182-194.
Fleischer, S.; Dvir, T. (2013). Tissue engineering on the nanoscale: lessons from the heart. Current
Opinion in Biotechnology, 24, 664-671.
Fonseca, K. B.; Maia, F. R.; Cruz, F. A.; Andrade, D.; Juliano, M. A.; Granja, P. L.; Barrias, C. C. (2013).
Enzymatic, physicochemical and biological properties of MMP-sensitive alginate hydrogels.
Soft Matter, 9, 3283-3292.
Frankel, A. D.; Pabo, C. O. (1988). Cellular uptake of the tat protein from Human Immunodeficiency
virus. Cell, 55, 1189-1193.
Futaki, S.; Suzuki, T.; Ohashi, W.; Yagami, T.; Tanaka, S.; Ueda, K.; Sugiura, Y. (2001). Arginine-rich
peptides: An Abundant source of membrane-permeable peptides having potential as carriers
for intracellular protein delivery. Journal of Biology Chemistry, 276, 5836-5840.
Ganz, T.; Weiss, J. (1997). Antimicrobial peptides of phagocytes and epithelia, Seminars in
Hematology, 34, 343-354.
Gaston, M.A.; Zhang, L.; Green-Church, K. B.; Krzycki, J. A. (2011). The complete biosynthesis of the
genetically encoded amino acids pyrrolysine from lysine. Nature, 471, 647-650.
Gingrich, D. E.; Reddy, D. R.; Iqbal, M. A.; Singh, J.; Aimone, L. D.; Angeles, T. S.; Albom, M.; Yang,
S.; Ator, M. A.; Meyer, S. L.; Robinson, C.; Ruggeri, B. A.; Dionne, C. A.; Vaught, J. L.; Mallamo,J.
P.; Hudkins,R. L. (2003). A new class of potent vascular endothelial growth factor receptor
tyrosine kinase inhibitors: structure-activity relationships for a series of 9-alkoxymethyl-12-
(3-hydroxypropyl)indeno[2,1-a]pyrrolo[3,4-c] carbazole-5-ones and the identification of CEP-
5214 and its dimethylglycine ester prodrug clinical candidate CEP-7055. Journal of Medicinal
Chemistry, 46, 5375-5388.
Gomes, P.; Gomes, J. R. B.; Rodrigues, M.; Moreira, R. (2003). Amino acids as selective sulfonamide
acylating agents. Tetrahedron, 59, 7473-7480.
Gomes, P.; Araújo, M. J.; Rodrigues, M.; Vale, N.; Azevedo, Z.; Iley, J.; Chambel, P.; Morais, J.; Moreira,
R. (2004). Synthesis of imidazolidin-4-one and 1H-imidazo[2,1-a]isoindole-2,5(3H,9bH)-dione
derivatives of primaquine: scope and limitations. Tetrahedron, 60, 5551-5562.
Gomes, P.; Vale, N.; Moreira, R. (2007). Cyclization-activated prodrugs. Molecules, 12, 2484-2506.
Gonzalez, D. E.; Covitz, K-M. Y.; Sadée, W.; Mrsny, R. J.; (1998). An oligopeptide transporter is
expressed at high levels in the pancreatic carcinoma cell lines AsPc-1 and Capan-2. Cancer
Research, 58, 519-525.
Goolcharran, C.; Borchardt, R. T. (1998). Kinetics of diketopiperazine formation using model
peptides. Journal of Pharmaceutical Sciences, 87, 283-288.
References 221
Gupta, D.; Gupta, S. V.; Lee, K. D.; Amidon, G. D. (2009). Chemical and enzymatic stability of
amino acid prodrugs containing methoxy, ethoxy and propylene glycol linkers, Molecular
Pharmaceutics, 6, 1604-1611.
Green, M.; Loewenstein, P. M. (1988). Autonomous functional domains of chemically synthesized
human immunodeficiency virus tat trans-activator protein. Cell, 55, 1179-1188.
Grillo, M. A.; Colombatto, S. (2004). Arginine revisited: minireview article. Amino Acids, 26, 345-351.
Gynther, M.; Laine, K.; Ropponen, J.; Leppanen, J.; Mannila, A.; Nevalainen, T.; Savolainen, J.;
Jarvinen, T.; Rautio, J. (2008). Large neutral amino acid transporter enables brain drug delivery
via prodrugs. Journal of Medicinal Chemistry, 51, 932-936.
Haas, H. L.; Sergeeva, O. A.; Selbach, O. (2008). Histamine in the nervous system. Physiology
Review, 88, 1183-1241.
Han, H.-K.; Amidom, G. L. (2010). Targeted prodrug design to optimize drug delivery. AAPS
PharmaSci, 2, 48-58.
Hamel, A. R.; Hubler, F.; Carrupt, A.; Wenger, R. M.; Mutter, M. (2004). Cyclosporin A prodrugs:
design, synthesis and biophysical properties. Journal of Peptide Research, 63, 147-154.
Han, H.; de Vrueh, R. L.; Rhie, J. K.; Covitz, K. M.; Smith, P. L.; Lee, C. P.; Oh, D. M.; Sadee, W.;
Amidon, G. L. (1998). 5’-Amino acid esters of antiviral nucleosides, acyclovir, and AZT are
absorbed by the intestinal PEPT1 peptide transporter. Pharmacetical Research, 15, 1154-1159.
Hatanaka, T.; Haramura, M.; Fei, Y. J.; Miyauchi, S.; Bridges, C. C.; Ganapathy, P. S.; Smith, S. B.;
Ganapathy, V.; Ganapathy, M. E. (2004). Transport of amino acid-based prodrugs by the Na+-
and Cl(−) -coupled amino acid transporter ATB(0,+) and expression of the transporter in tissues
amenable for drug delivery. Journal of Pharmacology and Experimental Therapeutics, 308,
1138-1147.
Hauser, C. A.; Zhang, S. (2010). Designer self-assembling peptide nanofiber biological materials.
Chemical Society Reviews, 39, 2780-2790.
Hemenway, J. N.; Jarho, P.; Henri, J. T.; Nair, S. K.; VanderVelde, D.; Georg, G. I.; Stella, V. J. (2010).
Preparation and physicochemical characterization of a novel water-soluble prodrug of
carbamazepine. Journal of Pharmaceutical Sciences, 99, 1810-1825.
Heitz, F.; Morris, M. C.; Divita, G. (2009). Twenty years of cell-penetrating peptides: from molecular
mechanisms to therapeutics. British Journal of Pharmacology, 157, 195-206.
Hervé, F.; Ghinea, N.; Scherrmann, J. M. (2008). CNS delivery via adsorptive transcytosis. AAPS
Journal, 10, 455-472.
Hofsteenge, J.; Capuano, A.; Altszuler, R.; Moore, S. (1986). Carrier-linked primaquine in the
chemotherapy of malaria. Journal of Medicinal Chemistry, 29, 1765-1769.
Hsiao, L. L.; Howard, R. J.; Aikawa, M.; Taraschi, T. F. (1991). Modification of host cell membrane
lipid composition by the intra-erythrocytic human malaria parasite Plasmodium falciparum.
Biochemical Journal, 274, 121-132.
Hsu, C. H.; Chen, C.; Jou, M. L.; Lee, A. Y.; Lin, Y. C.; Yu, Y. P.; Huang, W. T.; Wu S. H. (2005). Structural
and DNA-binding studies on the bovine antimicrobial peptide, indolicidin: evidence for multiple
conformations involved in binding to membranes and DNA. Nucleic Acids Research, 33, 4053-
4064.
Huang, P.; Chubb, S.; Hertel, L. W.; Grindey, G. B.; Plunkett, W. (1991). Action of 2’,2’-
difluorodeoxycytidine on DNA synthesis. Cancer Research, 51, 6110-6117.
Hummel, G.; Reineke, U.; Reimer, U. (2006). Translating peptides into small molecules. Molecular
BioSystems, 2, 499-508.
Hutunen, K. M.; Raunio, H.; Rautio, J. (2011). Prodrugs-from serendipity to rational design.
Pharmacological Reviews, 63, 750-771.
Ischakov, R.; Adler-Abramovich, L.; Buzhansky, L.; Shekhter, T.; Gazit, E. (2013). Peptide-based
hydrogel nanoparticles as effective drug delivery agents. Bioorganic and Medicinal Chemistry,
21, 3517-3522.
222 Amino Acids and Peptides in Medicine: Old or New Drugs?
Jeu, L.; Fung, H. B. (2004). Daptomycin: a cyclic lipopeptide antimicrobial agent. Clinical
Therapeutics, 1728-1757.
Johnstone, B.; Alini, M.; Cucchiarini, M.; Dodge, G. R.; Eglin, D.; Guilak, F.; Madry, H.; Mata, A.;
Mauck, R. L.; Semino, C. E. (2013). Tissue engineering for articular cartilage repair-the state of
the art,. European Cells and Materials, 25, 248-267.
Jung, D.; Dorr, A. (1999). Single-dose pharmacokinetics of valganciclovir in HIV- and CMV-
seropositive subjects. Journal of Clinical Pharmacology 39, 800-804.
Kageyama, T.; Nakamura, M.; Matsuo, A.; Yamasaki, Y.; Takakura, Y.; Hashida, M.; Kanai, Y.; Naito,
M.; Tsuruo, T.; Minato, N.; Shimohama, S. (2000). The 4F2hc/LAT1 complex transports L-DOPA
across the blood-brain barrier. Brain Research, 879, 115-121.
Kan, E. J.; Demel, R. A.; Breukink, E.; van der Bent., A.; de Kruijff, B. (2002). Clananin permeabilizes target
membranes via two distinctly different pH-dependent mechanisms. Biochemistry, 18, 7529-7539.
Kanai, Y.; Segawa, H.; Miyamoto, K., Uchino, H.; Takeda, E.; Endou, H. (1998). Expression cloning
and characterization of a transporter for large neutral amino 367 acids activated by the heavy
chain of 4F2 antigen (CD98). Journal of Biological Chemistry, 273, 23629-23632.
Kaspar, A. A.; Reichert, J. M. (2013). Future directions for peptide therapeutics developments. Drug
Discovery Today, 18, 807-817.
Killian, D. M.; Hermeling, S.; Chikhale, P. J. (2007). Targeting the cerebrovascular large neutral amino
acid transporter (LAT1) isoform using a novel disulfide-based brain drug delivery system. Drug
Delivery, 14, 25-31.
Kim, I.; Chu, X. Y.; Kim, S.; Provoda, C. J.; Lee, K. D.; Amidon, G. L. (2003). Identification of a human
valacyclovirase: biphenyl hydrolase-like protein as valacyclovir hydrolase. Journal of Biology
Chemistry, 278, 25348-25356.
Kim, I.; Song, X.; Vig, B. S.; Mittal, S.; Shin, H. C.; Lorenzi,P. J.; Amidon, G. L. (2004). A novel
nucleoside prodrug-activating enzyme: substrate specificity of biphenyl hydrolase-like protein.
Molecular Pharmaceutics, 1, 117-127.
Kirk, K. (2001). Membrane transport in the malaria-infected erythrocyte. Physiological Reviews, 81,
495-537.
Kotthaus, J.; Hungeling, H.; Reeh, C.; Kotthaus, J.; Schade, D.; Wein, S.; Wolffram, S.; Clement, B.
(2011). Synthesis and biological evaluation of L-valine-amidoximeesters as double prodrugs of
amidines. Bioorganic and Medicinal Chemistry, 19, 1907-1914.
Kragol, G.; Lovas, S.; Varadi, G.; Condie, B. A.; Hoffmann, R.; Otvos, L. Jr. (2001). The antibacterial
peptide pyrrhocoricin inhibits the atpase actions of DnaK and prevents chaperone-assisted
protein folding. Biochemistry, 40, 3016-3026.
Kulkarni, M. M.; McMaster, W. R.; Kamysz, E.; Kamysz, W.; Engman, D. M.; McGwire, B. S. (2006).
The major surface-metalloprotease of the parasitic protozoan, Leishmania, protects against
antimicrobial peptide-induced apoptotic killing. Molecular Microbiology, 62, 1484-1497.
Ladner, R. C.; Sato, A. K.; Gorzelany, J.; de Sousa, M. (2004). Phage display-derived peptides as
therapeutic alternatives to antibodies. Drug Discovery Today, 15, 525-529.
Ladokhin, A. S.; White S. H. (2001). Protein chemistry at membrane interfaces: non-additivity of
electrostatic and hydrophobic interactions. Journal of Molecular Biology, 309, 543-552.
Lai, L.; Xu, Z.; Zhou, J.; Lee, K. D.; Amidon, G. L. (2008). Molecular basis of prodrug activation by
human valacyclovirase, an alpha-amino acid ester hydrolase. Journal of Biological Chemistry,
283, 9318-9327.
Lal, R.; Sukbuntherng, J.; Tai, E. H.; Upadhyay, S.; Yao, F.; Warren, M. S.; Luo, W.; Bu, L.; Nguyen,
S.; Zamora, J.; Peng, G.; Dias, T.; Bao, Y.; Ludwikow, M.; Phan, T.; Scheuerman, R. A.; Yan,
H.; Gao, M.; Wu, Q. Q.; Annamalai, T.; Raillard, S. P; Koller, K.; Gallop, M. A.; Cundy, K. C.
(2009). Arbaclofen placarbil, a novel R-baclofen prodrug: improved absorption, distribution,
metabolism, and elimination properties compared with R-baclofen. Journal of Pharmacology
and Experimental Therapy, 330, 911-921.
References 223
Landowski, C. P.; Vig, B. S.; Song, X.; Amidon, G. L. (2005a). Targeted delivery to PEPT1 -
overexpresssing cells: acidic, basic, and secondary floxuridine amino acid ester prodrugs.
Molecular Cancer Therapeutics, 4, 659-667.
Landowski, C. P.; Sang, X.; Lorenzi, P. L.; Hilfinger, J. M.; Amidon, G. L. (2005b). Floxuridine amino
acid ester prodrugs: enhancing Cacol-2 permeability and resistance to glycosidic bond
metabolism. Pharmaceutical Research, 22, 1510-1518.
Larsen, S. W.; Ankersen, M.; Larsen, C. (2004). Kinetics of degradation and oil solubility of ester prodrug
of a model dipeptide (Gly-Phe). European Journal of Pharmaceutical Sciences, 22, 399-408.
Lehrer, R.; Barton, A.; Daher, K. A.; Harwig S. S. L.; Ganz T.; Selsted M. E. (1989). Interaction of
human defensins with Escherichia coli. Journal of Clinical Investigation, 84, 553-561.
Li, W. F.; Ma, G. X.; Zhou, X. X. (2006). Apidaecin-type peptides: biodiversity, structure-function
relationships and mode of action. Peptides, 27, 2350-2359.
Lin, C.-C.; Anseth, K. (2009). PEG Hydrogels for the Controlled Release of Biomolecules in
Regenerative Medicine. Pharmaceutical Research, 26, 631-643.
Lin, S., Lowman, H. B. (2003). Therapeutic peptides. Trends in Biotechnology, 21, 556-562.
Lien, S.; Lowman, H. B. (2003). Therapeutic Peptides. Trends in Biotechnology, 21, 556-562.
Lindgren, M.; Rosenthal-Aizman, K.; Saar, K.; Eiríksdóttir, E.; Jiang, Y.; Sassian, M.; Ostlund, P.;
Hällbrink, M.; Langel, Ü. (2006). Overcoming methotrexate resistance in breast cancer tumour
cells by the use of a new cell-penetrating peptide. Biochemical Pharmacology, 71, 416-425.
Lipinski, C. A.; Lombardo, F.; Dominy, B. W.; Feeney, P. J. (1997). Experimental and computational
approaches to estimate solubility and permeability in drug discovery and development
settings. Advanced Drug Delivery Reviews, 23, 3-25.
Lipinski, C. A. (2000). Drug-like properties and the causes of poor solubility and poor permeability.
Journal Pharmacological and Toxicological Methods, 44, 235-249.
Lipinski, C. A. (2004). Lead- and drug-like compounds: the rule-of-five revolution. Drug Discovery
Today Technologies, 1, 337-341.
Liskamp, R. M. J. (2014). Bicycling into cells. Nature Chemistry, 6, 855-857.
Loffet, A. (2002). Peptides as drugs: is there a market? Journal of Peptide Sciences, 8, 1-7.
Loo, Y.; Zhang, S.; Hauser, C. A. E. (2012). From short peptides to nanofibers to macromolecular
assemblies in biomedicine. Biotechnology Advances, 30, 593-603.
Lowenberger, C. (2001). Innate immune response of Aedes aegypti. Insect Biochemistry and
Molecular Biology, 31, 219-229.
Ludtke, S. J.; He, K.; Heller, W. T.; Harroun, T. A.; Yang, L.; Huang H. W. (1996). Membrane pores
induced by magainin. Biochemistry, 35, 13723-13728.
Ma, J. S. (2003). Unnatural amino acids in drug discovery. Chimica Oggi, 21, 65-68.
Manna, P.; Sinha, M.; Sil, P. C. (2009). Taurine plays a beneficial role against cadmium-induced
oxidative renal dysfunction. Amino Acids, 36, 417-428.
Mandal, D.; Shirazi, A. N.; Parang, K. (2011). Cell-penetrating homochiral cyclic peptides as nuclear-
targeting molecular transporters. Angewandte Chemie International Edition, 50, 9633-9637.
Marathe, P. H.; Kamath, A. V.; Zhang, Y.; D’Arienzo, C.; Bhide, R.; Fargnoli, J. (2009). Preclinical
pharmacokinetics and in vitro metabolism of brivanib (BMS-540215), a potent VEGFR2 inhibitor
and its alanine ester prodrug brivanib alaninate. Cancer Chemotherapy and Pharmacology, 65,
55-66.
Mart, R. J.; Osborne, R. D.; Stevens, M. M.; Ulijn, R. V. (2006). Peptide-based stimuli-responsive
biomaterials. Soft Matter, 2, 822-835.
Mason, A. J.; Moussaoui, W.; Abdelrahman, T.; Boukhari, A.; Bertani, P.; Marquette
A.; Shooshtarizaheh, P.; Moulay, G.; Boehm, N.; Guerold, B.; Sawers, R. J.; Kichler A.; Metz-
Boutigue, M. H.; Candolfi, E.; Prévost, G.; Bechinger, B. (2009). Structural determinants of
antimicrobial and antiplasmodial activity and selectivity in histidine-rich amphipathic cationic
peptides. Journal of Biological Chemistry, 284, 119-133.
224 Amino Acids and Peptides in Medicine: Old or New Drugs?
Matson, J. B.; Zha, R. H.; Stupp, S.I. (2011). Peptide self-assembly for crafting functional biological
materials. Current Opinion in Solid State and Materials Science, 15, 225-235.
Matson, J. B.; Stupp, S. I. (2012). Self-assembling peptide scaffolds for regenerative medicine.
Chemical Communications, 48, 26-33.
McGregor, D. P. (2008). Discovering and improving novel peptide therapeutics. Current Opinion in
Pharmacology, 8, 616-619.
Melo, M. N.; Ferre, R.; Castanho, M. A. (2009). Antimicrobial peptides: linking partition, activity and
high membrane-bound concentrations, Nature Review Microbiology, 7, 245-250.
Milletti, F. (2012). Cell-penetrating peptides: classes, origin, and current landscape. Drug Discovery
Today, 17, 850-860.
Miranda, P.; Blum, M. R. (1983). Pharmacokinetics of acyclovir after intravenous and oral
administration, J. Antimicrobial Agents Chemotherapy, 12(Suppl. B), 29-37.
Mittal, S.; Song, X.; Vig, B. S.; Amidon, G. L. (2007). Proline prodrug of melphalan targeted to
prolidase, a prodrug activating enzyme overexpressed in melanoma. Pharmaceutical Research,
24, 1290-1298.
Mizuma, T.; Ohta, K.; Hayashi, M.; Awazu, S. (1992). Intestinal active absorption of sugar-conjugated
compounds by glucose transport system: implication of improvement of poorly absorbable
drugs. Biochemical Pharmacology, 43, 2037-2039.
Morris, M. C.; Vidal, P.; Chaloin, L.; Heitz, F.; Divita, G. A (1997). New peptide vector for efficient
delivery of oligonucleotides into mammalian cells. Nucleic Acids Research 25, 2730-2736..
Mulholland, P. J.; Ferry, D. R.; Anderson, D.; Hussain, S. A.; Young, A. M.; Cook, J. E.; Hodgkin, E.;
Seymour, L. W.; Kerr, D. J. (2001). Pre-clinical and clinical study of QC12, a water-soluble, pro-
drug of quercetin, Annals of Oncology, 12, 245-248.
Myrberg, H.; Zhang, L.; Mäe, M.; Langel, Ü. (2008). Design of a tumor-homing cell-penetrating
peptide. Bioconjugate Chemistry, 19, 70-75.
Nam, N. H.; Kim, Y.; You, Y. J.; Hong, D. H.; Kim, H. M.; Ahn, B. Z. (2003). Water soluble prodrugs of
the antitumor agent 3-[(3-amino-4-methoxy)phenyl]-2-(3,4,5-trimethoxyphenyl) cyclopent-2-
ene-1- one. Bioorganic and Medicinal Chemistry, 11, 1021-1029.
Nakanishi, T.; Tamai, I.; Takaki, A.; Tsuji, A. (2000). Cancer cell targeted drug delivery utilizing
oligopeptide transport activity. International Journal of Cancer, 88, 274-280.
Northfield, S. E:; Wang, C. K.; Schroeder, C. I.; Durek, T.; Kan, M. W.; Swedberg, J. E.; Craik, D. J.
(2014). Dissulfide-rich macrocyclic peptides as templates in drug design. European Journal
Medicinal Chemistry, 77, 248-257.
Oehlke, J.; Scheller, A.; Wiesner, B.; Krause, E.; Beyermann, M.; Klauschenz, E.; Melzig, M.; Bienert,
M. (1998). Cellular uptake of an alpha-helical amphipathic model peptide with the potential to
deliver polar compounds into the cell interior non-endocytically. Biochimica et Biophysica Acta,
Biomembranes, 1414, 127-139.
O’Leary, L. E.; Fallas, J. A.; Bakota, E. L.; Kang, M. K.; Hartgerink, J. D. (2011). Multi-hierarchical self-
assembly of a collagen mimetic peptide from triple helix to nanofibre and hydrogel. Nature
Chemistry, 3, 821-828.
Orbach, R.; Adler-Abramovich, L.; Zigerson, S.; Mironi-Harpaz, I.; Seliktar, D.; Gazit, E. (2009). Self-
assembled fmoc-peptides as a platform for the formation of nanostructures and hydrogels.
Biomacromolecules, 10, 2646-2651.
Oskolkov, N.; Arukuusk, P.; Copolovici, D. M.; Lindberg, S.; Margus, H.; Padari, K.; Pooga, M.;
Langel, Ü. (2011). Nickects, phosphorylated derivatives of transportan 10 for cellular delivery of
oligonucleotides. International Journal of Peptide Research and Therapeutics, 17, 147-157.
Page-McCaw, A.; Ewald, A. J.; Werb, Z. (2007). Matrix metalloproteinases and the regulation of tissue
remodelling. Nature Reviews. Molecular Cell Biology, 8, 221-233.
Pardridge, W. M. (2003). Blood-brain barrier drug targeting: the future of brain drug development.
Molecular Interventions, 3, 90-105.
References 225
Park, C. B.; Kim, H. S.; Kim, S. C. (1998). Mechanism of action of the antimicrobial peptide buforin
ii: Buforin ii kills microorganisms by penetrating the cell membrane and inhibiting cellular
functions. Biochemical and Biophysical Research Communications, 244, 253-257.
Patrzykat, A.; Friedrich, C. L.; Zhang, L.; Mendoza, V.; Hancock, R. E. (2002). Sublethal
concentrations of pleurocidin-derived antimicrobial peptides inhibit macromolecular synthesis
in Escherichia coli. Antimicrobial Agents Chemotherapy, 46, 605-614.
Pettit, D. K.; Gombotz, W. R. (1998). The development of site-specific drug-delivery systems for
protein and peptide biopharmaceuticals. Trends in Biotechnology, 16, 343-349.
Patterson, J.; Hubbell, J. A. (2011). SPARC-derived protease substrates to enhance the plasmin
sensitivity of molecularly engineered PEG hydrogels. Biomaterials, 32, 1301-1310.
Perta-Kajan, J.; Twardowski, T.; Jakubowski, H. (2007). Mechanisms of homocysteine toxicity in
humans. Amino Acids, 32, 561-572.
Philip, A.; Kepler, J. A.; Johnson, B. H.; Carrol, F. I. (1988). Peptide derivatives of primaquine as
potential antimalarial agents. Journal of Medicinal Chemistry, 31, 870-874.
Pichereau, C.; Allary, C. (2005). Therapeutic peptides under the spotlight. European
Biopharmaceutical Review Winter Issue, 88-91.
Pochan, D. J.; Schneider, J. P.; Kretsinger, J.; Ozbas, B.; Rajagopal, K.; Haines, L. (2003). Thermally
reversible hydrogels via intramolecular folding and consequent self-assembly of a de novo
designed peptide. Journal of the American Chemical Society, 125, 11802-11803.
Pochopin, N. L.; Charman, W. N.; Stella, V. J. (1995). Amino acid derivatives of dapsone as water-
soluble prodrugs. International Journal of Pharmaceutics, 121, 157-167.
Pooga, M.; Hällbrink, M.; Zorko, M.; Langel, Ü. (1998). Cell penetration by transportan. The FASEB
Journal, 12, 67-77.
Portela, M. J.; Moreira, R.; Valente, E.; Constantino, L.; Iley, J.; Pinto, J.; Rosa, R.; Cravo, P.; Rosário, V.
E. (1999). Dipeptide Derivatives of Primaquine as Transmission-Blocking Antimalarials: Effect
of Aliphatic Side-Chain Acylation on the Gametocytocidal Activity and on the Formation of
Carboxyprimaquine in Rat Liver Homogenates. Pharmaceutical Research, 16, 949-955.
Pouny, Y.; Rapaport, D.; Mor, A.; Nicolas, P.; Shai, Y. (1992). Interaction of antimicrobial dermaseptin
and its fluorescently labeled analogues with phospholipid membranes. Biochemistry, 31,
12416-12423.
Powers, J. -P. S. and Hancock, R. E. W. (2003). The relationship between peptide structure and
antibacterial activity. Peptides, 24, 1681-1691.
Pretzel, J.; Mohring, F.; Rahlfs, S.; Becker, K. (2013). Antiparasitic peptides. Advances in Biochemical
Engineering Biotechnology, 135, 157-192.
Raghuraman, H.; Chattopadhyay, A. (2007). Melittin: a membrane-active peptide with diverse
functions. Bioscience Reports, 27, 189-223.
Rasheed, A.; Kumar, C. K.; Mishra, A. (2011). Synthesis, hydrolysis studies and phamacodynamic
profiles of amide prodrugs of dexibuprofen with amino acids. Journal of Enzyme Inhibition and
Medicinal Chemistry, 26, 688-695.
Rautio, J.; Nevalainen, T.; Taipale, H.; Vepsalainen, J.; Gynther, J.; Pedersen, T.; Jarvinen, T. (1999).
Synthesis and in vitro evaluation of aminoacyloxyalkyl esters of 2-(6-ethoxy-2-naphthyl)
propionic acid as novel naproxen prodrugs for dermal drug delivery. Pharmaceutical Research,
16, 1172-1178.
Rowley, J. A.; Madlambayan, G.; Mooney, D. J. (1999). Alginate hydrogels as synthetic extracellular
matrix materials. Biomaterials, 20, 45-53.
Ruoslahti, E.; Pierschbacher, M. (1987). New perspectives in cell adhesion: RGD and integrins.
Science 238, 491-497.
Säälik, P.; Elmquist, A.; Hansen, M.; Padari, K.; Saar, K.; Viht, K.; Langel, Ü.; Pooga, M. (2004).
Protein cargo delivery properties of cell-penetrating peptides. A comparative study.
Bioconjugate Chemistry, 15(2004), 1246-1253.
226 Amino Acids and Peptides in Medicine: Old or New Drugs?
Turanov, A. A.; Xu, X. M.; Carlson, B. A. Yoo, M. H.; Gladyshev, V. N.; Hatfield, D. L. (2011).
Biosynthesis of selenocysteine, the 21st amino acid in the genetic code, and a novel pathway for
cysteine biosynthesis. Advances in Nutrition, 2, 122-128.
Turk, B. E.; Huang, L. L.; Piro, E. T.; Cantley, L. C. (2001). Determination of protease cleavage site
motifs using mixture-based oriented peptide libraries. Nature Biotechnology 19, 661-667.
Umapathy, N. S.; Ganapathy, V.; Ganapathy, M. E. (2004). Transport of amino acid esters and
the amino-acid-based prodrug valganciclovir by the amino acid transporter ATB(0,+).
Pharmaceutical Research, 21, 1303-1310.
Vale, N.; Moreira, R.; Gomes, P. (2005). Quimioterapia da Malária: um século de desenvolvimento de
antimaláricos. Química, 99, 57-69.
Vale, N.; Collins, M. S.; Gut, J.; Ferraz, R.; Rosenthal, P. J.; Cushion, M. T.; Moreira, R.; Gomes,
P. (2008a). Anti-Pneumocystis carinii and antiplasmodial activities of primaquine-derived
imidazolidin-4-ones. Bioorganic and Medicinal Chemistry Letter, 18, 485-488.
Vale, N.; Matos, J.; Gut, J.; Nogueira, F.; Rosário, V.; Rosenthal, P. J.; Moreira, R.; Gomes, P. (2008b).
Imidazolidin-4-one peptidomimetic derivatives of primaquine: synthesis and antimalarial
activity. Bioorganic and Medicinal Chemistry Letter, 18, 4150-4153.
Vale, N.; Matos, J.; Moreira, R.; Gomes, P. (2008c). Amino acids as selective acylating agents:
regioselective N1-acylation of imidazolidin-4-one derivatives of the antimalarial drug
primaquine. Tetrahedron, 64, 11144-11149.
Vale, N.; Nogueira, F.; do Rosário, V. E.; Gomes, P.; Moreira, R. (2009a). Primaquine dipeptide
derivatives bearing an imidazolidin-4-one moiety at the N-terminus as potential antimalarial
prodrugs. European Journal of Medicinal Chemistry, 44, 2506-2516.
Vale, N.; Prudêncio, M.; Marques, C. A.; Collins, M.; Gut, J.; Nogueira, F.; Matos, J.; Rosenthal, P. J.;
Cushion, M. T.; do Rosário, V. E.; Mota M. M.; Moreira, R.; Gomes, P. (2009b). Imidazoquines as
antimalarial and antipneumocystis agents. Journal of Medicinal Chemistry, 52, 7800-7807.
Vale, N.; Aguiar, L.; Gomes, P. Antimicrobial peptides: a new class of antimalarial drugs? (2014).
Frontiers in Pharmacology, 5, 275.
Vale, N.; Ferreira, A.; Fernandes, I.; Mateus, N.; Gomes, P. (2016). New peptide-gemcitabine
conjugates for Cancer Therapy. (Submitted).
Veber, D. F.; Johnson, S. R.; Cheng, H. Y.; Smith, B. R.; Ward, K. W.; Kopple, K. D. (2002). Molecular
properties that influence the oral bioavailability of drug candidates. Journal of Medicinal
Chemistry, 45, 2615-2623.
Vig, B. S.; Lorenzi, P. J.; Mittal, S.; Landowski, C. P.; Shin, H. C.; Mosberg, H. I.; Hilfinger, J. M.;
Amidon, G. L. (2003). Amino acid ester prodrugs of floxuridine: synthesis and effects of
structure, stereochemistry, and site of esterification on the rate of hydrolysis. Pharmaceutical
Research, 20, 1381-1388.
Vig, B. S.; Lorenzi, P. L.; Mittal, S.; Landowski, C. P.; Shin, H-C.; Mosberg, H. I.; Hilfinger, J. M.;
Amidon, G. L. (2003). Amino acid ester prodrugs of Floxuridine: synthesis and effects of
structure, stereochemistry and site of esterification on the rate of hydrolysis. Pharmaceutical
Research, 20, 1381-1388.
Vig, B. S.; Huttunen, K. M.; Laine, K.; Rautio, J. (2013). Amino acids as promoieties in prodrug design
and development. Advanced Drug Delivery Reviews, 65, 1370-1385.
Vlieghe, P.; Lisowski, V.; Martinez, J.; Khrestchatisky, M. (2010). Synthetic therapeutic peptides:
science and market. Drug Discovery Today, 15, 40-56.
Walker, I.; Nicholls, D.; Irwin, W. J.; Freeman, S. (1994). Drug delivery via active transport at the
blood-brain barrier: affinity of a prodrug of phosphonoformate for the large amino acid
transporter. International Journal of Pharmaceutics, 104, 157-167.
Wang, Y.; Welty, D. F. (1996). The simultaneous estimation of the influx and efflux blood-brain barrier
permeabilities of gabapentin using a microdialysis-pharmacokinetic approach. Pharmaceutical
Research, 13, 398-403.
228 Amino Acids and Peptides in Medicine: Old or New Drugs?
Wang, Z.; Wang, G. (2004). APD: The antimicrobial peptide database. Nucleic Acids Research, 32,
D520-D592.
Wang, G.; Li, X.; Wang, Z. (2009). APD2: the updated antimicrobial peptide database and its
application in peptide design. Nucleic Acids Research, 37, D933-D937.
Wei, Y.; Pei, D. (2000). Activation of antibacterial prodrugs by peptide deformylase. Bioorganic and
Medicinal Chemistry Letter, 10, 1073-1076.
Wimley, W. C.; Selsted, M. E.; White, S. H. (1994). Interactions between human defensins and lipid
bilayers: evidence for formation of multimeric pores. Protein Science, 3, 136-1373.
Wipf, P.; Li, W. L.; Adeyeye, C. M.; Rusnak, J. M.; Lazo, J. S. (1996). Synthesis of chemoreversible
prodrugs of ara-C with variable time-release profiles. Biological evaluation of their apoptotic
activity. Bioorganic and Medicinal Chemistry, 4, 1585-1596.
Witt, K. A.; Gillespie, T. J.; Huber, J. D.; Egleton, R. D.; Davis, T. P. (2001). Peptide drug modifications
to enhance bioavailability and blood-brain barrier permeability. Peptides, 22, 2239-2243.
Witt, K. A.; Davis, T. P. (2006). CNS drug delivery: Opioid peptides and the blood-brain barrier. AAPS
Journal, 8, E76-E88.
Wojtowicz, A. M.; Shekaran, A.; Oest, M. E.; Dupont, K. M.; Templeman, K. L.; Hutmacher, D. W.;
Guldberg, R. E.; García, A. J. (2010). Coating of biomaterial scaffolds with the collagen-mimetic
peptide GFOGER for bone defect repair. Biomaterials, 31, 2574-2582.
Woodley, J. F. (1994). Enzymatic barrier for GI peptide and protein delivery. Critical Review in
Therapeutic Drug Carrier Systems, 11, 61-95.
Wu, G. (2009). Amino acids: metabolism, functions, and nutrition. Amino Acids, 37, 1-17.
Wu, J. Y.; Prentice, H. (2010). Role of taurine in the central nervous system. Journal of Biomedical
Sciences, 17, S1.
Yang, L.; Harroun, T. A.; Weiss, T. M.; Ding L.; Huang, H. W. (2001). Barrel-stave model or toroidal
model? A case study on melittin pores. Biophysical Journal, 81, 1475-1485.
Yeaman, M. R.; Yount, N. Y. (2003). Mechanisms of antimicrobial peptide action and resistance.
Pharmacology Reviews, 55, 27-55.
Yi, H. Y.; Chowdhury, M.; Huang, Y. D.; Yu, X. Q. (2014). Insect antimicrobial peptides and their
applications. Applied Microbiology Biotechnology, 98, 5807-5822.
Zasloff, M. (1987). Magainins, a class of antimicrobial peptides from Xenopus skin: isolation,
characterization of two active forms, and partial cDNA sequence of a precursor. Proceedings of
the National Academy of Sciences of USA, 84, 5449-5453.
Zhou, M.; Smith, A. M.; Das, A. K.; Hodson, N. W.; Collins, R. F.; Ulijn, R: V.; Gough, J. E. (2009). Self-
assembled peptide-based hydrogels as scaffolds for anchorage-dependent cells. Biomaterials,
30, 2523-2530.
James J. Chambers*8
3.2 Targeting Calcium-mediated Excitotoxicity in the
CNS
3.2.1 Introduction
At the heart of our ability to recognize, remember, and perform basic motor functions,
is the highly coordinated placement of neuronal receptors at the interface between
adjacent cells that control the flux of metal ions across the membrane (Kerchner &
Nicoll, 2008; Ribrault, 2011; Wang, 2012). Glutamate is the neurotransmitter that
mediates the majority of fast, excitatory neurotransmission in the central nervous
system (Cotman & Monaghan, 1986). Vesicles of glutamate are released from the
presynaptic cell when action potentials depolarize the membrane of synaptic boutons
(terminals) and vesicles fuse with the membrane (Debelleroche & Bradford, 1977).
Glutamate then diffuses across the synapse and activates postsynaptic glutamate
receptors, initiating postsynaptic current flow via ion entry and depolarizing that
cell, thus completing chemical synaptic communication (Moore & Buchanan, 1993).
Despite decades of inquiry, we still do not fully understand the elegant electrical and
chemical neuronal signals that are generated in a highly complex tangle of minuscule
structures.
Communication between cells in the nervous system requires neurons to quickly
alter their membrane potential, transiently entering a depolarized, excited state to
pass action potentials along to neighboring neurons, and then rapid repolarization
to be prepared for new incoming information. Over-excitation of neurons by various
means maintains a depolarized state for longer than usual and can contribute to
neurological disease. Some of these diseases, which are linked to neuronal over-
activity, can also cause cell damage and death via a process called excitotoxicity.
Excitotoxicity is thought to play a role in neuronal cell and network damage following
stroke and traumatic brain injury and, quite possibly, in the progressive damage
associated with a number of congenital and sporadic neurodegenerative disorders
such as Alzheimer’s, stroke, Huntington’s, Parkinson’s, Amyotrophic Lateral Sclerosis
(ALS), and Multiple Sclerosis (MS).
Receptors located at the neuronal membrane help regulate these fluctuations
in membrane potential by controlling the opening and closing of transmembrane
ion channels, which variously allow cations and anions to flow into and out of the
neuron down their electrical and chemical concentration gradient. During typical
communication at an excitatory synapse, the presynaptic cell releases glutamate, the
primary excitatory neurotransmitter, after an influx of calcium that enters through
voltage-gated calcium channels. Glutamate crosses the synaptic cleft by diffusion and
binds to receptors on the postsynaptic cell. Some glutamate receptors are ionotropic
ligand-gated channels that are found on the postsynaptic neuron. When an agonist
such as glutamate binds, the associated ion channel opens, allowing monovalent
cations, primarily sodium, to flow into the neuron. As the positive charge flows into
the neuron, the membrane potential of that postsynaptic cell depolarizes, and if it is
sufficiently depolarized, other receptors located nearby on the membrane will open
as well, eventually allowing calcium to enter the postsynaptic cell.
Glutamate is the most prevalent excitatory neurotransmitter in the central
nervous system (CNS) and glutamate receptors (GluRs) play a key role in both normal
excitatory neurotransmission, and in modulating both normal and constructive
plasticity as well as excitotoxic biochemistry, the main topic of this chapter (Cotman
& Monaghan, 1986; Debelleroche & Bradford, 1977; Moore & Buchanan, 1993). There
are two major types of GluRs that differ in the way they influence neuronal response
to excitatory neurotransmission. The first are the metabotropic glutamate receptors
(mGluRs) which, upon binding to glutamate, set off an intracellular biochemical
cascade via second messenger signaling. While mGluRs are certainly targets of interest
for negating the excitotoxic effects of over-stimulation, they are not the focus of this
work, however, a well-written and expansive review of mGluRs and pharmacological
agents was published recently (Williams & Dexter, 2014). Of more importance here,
ionotropic glutamate receptors (iGluRs) mediate fast responses to glutamate release
and also, as their name implies, allow ions to enter or leave the cell via an intrinsic ion
channel. iGluRs are typically heteromultimeric, integral membrane proteins that are
composed of four subunits that form a transmembrane ion channel. This channel is
allosterically connected to the agonist-binding site and upon binding to glutamate or
an exogenous agonist, the channel portion opens and allows cations to pass into the
cell which directly depolarize the membrane potential.
The postsynaptic iGluR composition has been found to mediate the ionic makeup
of the current that results from glutamate binding. The iGluRs are subdivided into
NMDA receptors and non-NMDA receptors with the latter being further divided into
AMPA receptors and kainate receptors. The AMPA receptors, named after the synthetic
agonist α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid, are located on the
acceptor, or postsynaptic, dendrite where they bind neurotransmitters, causing a
conformational change in the receptor protein structure that opens intrinsic sodium
ion channels. AMPA receptors are responsible for the vast majority of excitatory
Glutamate and Glutamate Receptors 231
neurotransmission and they open quickly in response to glutamate and close fast as
well. The influx of sodium ions alters the electrical potential of the postsynaptic cell,
producing a local signal that, after summation and integration at the neuron and then
circuit level, ultimately results in cognition or action. Another family of iGluRs is
the kainate receptors. These appear to be more related to AMPA receptors, but they
are still the subject of ongoing research and their roles in either normal signaling
or excitotoxicity has not been fully explored. If the cellular polarization is above a
certain threshold, neighboring NMDA receptors, named after the synthetic agonist
N-methyl-D-aspartate, fully or partially eject an ion channel-bound magnesium
ion allowing calcium to flow into the cell. The entry of calcium then sets off myriad
intracellular events, not least of which is activation of calcium-activated kinases and,
eventually, synaptic potentiation (Lisman, 2012). A change in the communication
strength between adjacent neurons arises from trafficking of receptors to or from the
synaptic zone and it is presently thought that this molecular trafficking plays a crucial
role in synaptic plasticity, which, in turn, is believed to underlie memory formation
and loss.
AMPA receptors are assembled from subunits GluA1 through GluA4 and are typically
comprised of heterotetramers, quite often containing GluA2 subunits. These
receptors are known to play a critical role in long-term potentiation (LTP), one of the
mechanisms thought to underlie memory formation as well as maladaptive and even
neurodegenerative plasticity (Kang, 2009). Most AMPA receptors are not permeable to
calcium ions, due to the presence of the GluA2 subunit. In the GluA2 subunit RNA, a
key modification is made to translate a codon to arginine instead of a glutamine in the
ion channel pore. The incorporation of arginine in the ion conduction pore results in a
channel that does not allow divalent cations, such as calcium, to pass.
A subset of AMPA receptors that are currently of considerable interest in the
neurobiology field and in particular the field of neurodegeneration are those that lack
the GluA2 subunit. GluA2-containing channels are quite often calcium-impermeable
due to an mRNA codon editing event (Araki, 2010; Isaac, 2007). The precise role of
GluA2-lacking receptors in synaptic plasticity or, for that matter, normal synaptic
signaling, is still the subject of ongoing debate in the field (Henley, 2011). While it
is thought that most functional AMPA receptors contain an edited GluA2 subunit,
there is now considerable evidence for a subpopulation of GluA2-lacking receptors
that result in calcium-permeable AMPA receptors (Keller, 1992; McDermott, 2003;
Szabo, 2012; Wright & Vissel, 2012). The enzyme Adenosine deaminase acting on
RNA 2 (ADAR2) specifically mediates RNA editing of the glutamine/arginine (Q/R) site
of GluA2 subunit of the AMPA receptors and it has been found that motor neurons
expressing Q/R site-unedited GluA2 undergo slow death in conditional ADAR2
232 Targeting Calcium-mediated Excitotoxicity in the CNS
knockout mice (Hideyama, 2012). This line of research also found that unedited
GluA2 mRNA was expressed in a large proportion of motor neurons from ALS patients
while motor neurons from normal and disease control subjects expressed only edited
GluA2 mRNA. ADAR2 was significantly down-regulated in all the motor neurons of
ALS patients, more extensively in those expressing Q/R site-unedited GluA2 mRNA
than those expressing only Q/R site-edited GluA2 mRNA. These findings suggest that
antagonism of calcium-permeable AMPA receptors could slow disease progression, a
topic we will discuss later in this manuscript.
Kainate receptors are also heterotetrameric ion channels assembled from subunits
GluA5-7 and KA1-2 (also known as GluK1-2). The properties of kainate channels are
similar to AMPARs in that they allow ion flux if glutamate activates the channel.
Further, kainate receptors are mostly impermeable to calcium. One main difference
between AMPA and kainate receptors is their subcellular location. While most
AMPA receptors are located at the postsynaptic membrane, kainate receptors have
been found on the presynaptic side as well as the postsynaptic side. It remains
unclear if presynaptic kainate receptor stimulation by glutamate results in more or
less vesicular release of glutamate into the synapse, though in the case of neuronal
necrosis and ischemia, it very likely does. These are discussed later in this chapter.
The function of postsynaptic kainate receptors is similar to that of AMPA receptors
in that depolarization via glutamate-evoked kainate channel current can unplug the
magnesium ion blocking NMDA receptors, thus leading to calcium entry into the
postsynaptic spine.
NMDA receptors are heterotetramers and are assembled from three major families
of subunits. The NR1 subunit is ubiquitously expressed, there are four NR2 subunits
(A-D), and two NR3 members (A and B). Most NMDA receptors that have been studied
to date contain both NR1 and NR2 subunits and they share the common attributes of
being activated by glutamate but normally being blocked by a magnesium ion that is
removed upon strong stimulation of the postsynaptic spine. Besides glutamate, it was
found that glycine was a required co-agonist to elicit NMDA channel opening.
NMDA receptors have slow gating kinetics, but once open, are highly permeable to
both calcium and sodium. Because of the large ion permeability to calcium, this influx
generates intracellular calcium waves or transients (also the term “calcium spikes” is
popular in the literature) that set off myriad calcium-mediated biochemical cascades.
The NMDA receptors are bound to postsynaptic cytoskeletal scaffolding proteins that
The Role of Calcium in Normal Neuronal Biochemistry 233
are collectively known as the postsynaptic density. This scaffolding also supports
kinases and other proteins that have downstream signaling roles, likely to receive the
calcium transient as efficiently as possible. Under normal circumstances, glutamate is
quickly removed from the synaptic cleft by excitatory amino acid transporters (EAAT)
located on nearby glial cells. This ensures that only a limited amount of calcium will
be able to enter the postsynaptic neuron.
NMDA receptors are opened when the neuronal membrane potential is depolarized
to a sufficient level to expel the bound magnesium ion. Opening of these ionotropic
receptors leads to a rise in postsynaptic calcium concentration, which has been linked
to long-term potentiation (LTP) via protein kinase activation. It is thought that strong
depolarization fully expels the magnesium plug, whereas milder depolarization
partially displaces the magnesium, resulting in less access to calcium through NMDA
receptors. This latter process is thought to underlie long-term depression (LTD), in
which lower concentrations of postsynaptic calcium activates protein phosphatases
and initiates partial dismantling of the synaptic connection.
The activated protein kinases that arise from the high calcium concentration
phosphorylate a variety of postsynaptic targets including some AMPA receptors.
These post-translational modifications have been shown to enhance single channel
conduction through these channels, thereby potentiating the connection between
these two neurons. Further, these kinases can phosphorylate the protein machinery
that determines the number of postsynaptic AMPA receptors, thus opening more
“docking” sites for new AMPA receptors from nearby locations. In the end, the result
of controlled calcium entry into a synaptic spine is more AMPA receptors that are
more permeable to cations. Of course, protein phosphatases can act to reverse these
potentiation modifications to return the synapses to the pre-LTP levels.
Besides direct activation of postsynaptic kinases, calcium entry can activate a
second messenger cascade to turn on gene transcription of proteins such as CaMKII
and PKAII leading to relatively high local concentrations of these kinases in the
synaptic spine. Both of these kinases have been found to play a role in LTP processes,
such as spine volume, addition of new AMPA receptors, and direct phosphorylation of
ion channels. It is this local effect of LTP that allows for selective potentiation of single
synapses and, because of this, each synaptic spine could be considered isolated from
other postsynaptic zones as all of the changes in kinase activity are highly restricted
to that local spine. This addressable and highly spatial system is likely related to long-
term memory storage.
234 Targeting Calcium-mediated Excitotoxicity in the CNS
3.2.4 Excitotoxicity
magnesium ion that normally blocks NMDA receptors occurs. This, of course, then
results in uncontrolled calcium entry through the NMDA receptors into a previously
unaffected neuron, thus propagating the cellular insult. In addition, it has been
shown that the NMDA receptors found on oligodendrocytes, the cells that provide
myelin sheaths the axons in the CNS, are activated as well, thereby exacerbating
excitotoxicity. Once in the cell, the increased calcium activates kinases, but also has an
effect on mitochondria. In response to high calcium, the mitochondrial permeability
transition pore is opened and it is presently thought that pore opening may allow the
release of reactive oxygen species, further contributing to apoptosis and slowing or
halting the local production of adenosine triphosphate (ATP). Lack of ATP quickly
eliminates the electrochemical gradient across the membrane of some ions, which
in turn shuts down glutamate transporters, the proteins that remove glutamate from
extracellular space, to terminate glutamatergic signaling. The result of this excitotoxic
cascade is the accumulation of glutamate in the extracellular milieu which can spread
to neighboring neurons.
is the high degree of selectivity that the disease shows for motor neurons over all other
cells in the body, even though the mutant form of SOD1 protein is in many other cell
types in the patient. A possible mechanism for this selectivity is based on excitatory
neurotransmitter-mediated neuropathology, in particular glutamate. It is presently
believed that motor neurons are especially susceptible to excitotoxicity because
they are known to receive strong glutamatergic input. Additionally, spinal motor
neurons have recently been found to express calcium-permeable AMPA receptors on
their surface. Multiple lines of neurobiological research have resulted in the current
understanding that calcium-permeable AMPA receptors play crucial roles in synaptic
signaling and plasticity in the CNS. An overabundance of these receptors, coupled
with glutamatergic excitation, could overwhelm the calcium buffering capacity of a
cell, resulting in metabolic/mitochondrial breakdown followed by cell death.
Presently, riluzole (Rilutek) is the only United States Food and Drug
Administration approved therapeutic to treat ALS. The specific biological target of
riluzole is controversial, but it is thought that the drug acts by reducing excitatory
neurotransmission, resulting in the influx of less calcium into motor neurons slowing
the progression of the disease. Unfortunately, riluzole is not a cure and only offers
some of those suffering with ALS around 3 additional months of life.
3.2.6.5 Stroke
During a stroke, the blood flow to part of the brain is disrupted either due to a blockage
in a blood vessel feeding the brain (ischemic stroke) or due to a rupture in one of these
vessels (hemorrhagic stroke). Neurons in the area of the brain that are supplied by
the affected vessel are often deprived of the required supply of oxygen and glucose
and are thus unable to continue operating the energy-demanding ion transporters
that establish the negative membrane potential. With a loss of these pumps, the
membrane begins to depolarize and when it has sufficiently depolarized to the point
that magnesium plugs have been ejected from the NMDA receptors, calcium ions enter
the neuron at the slightest exposure to glutamate. As already discussed, this calcium
entry then initiates excitotoxic molecular mechanisms (Lipton, 2006).
After a stroke, excess calcium is also able to enter neurons by a non-glutamate
receptor pathway. The Transient Receptor Potential (TRP) channels allow calcium
Diseases that are Potentially Exacerbated by Calcium-mediated Excitotoxicity 239
The progression of Parkinson’s disease involves the loss of the dopaminergic neurons
in the substantia nigra. These neurons help control muscle movement and as they
degenerate, patients lose control of voluntary and spontaneous muscle movement
presenting symptoms that include resting tremors, slowed movement, difficulty with
movement, and balance problems. In addition, a number of cognitive symptoms
typically develop as the disease progresses, including memory loss, depression, and
eventually dementia. The mechanism behind the initial neuronal degeneration is
only known in a small percentage of cases, termed familial Parkinson’s, which has
been linked to an inherited genetic mutation. However, like AD, most other cases
are sporadic Parkinson’s and there is no known cause. However, recent research
has implicated glutamate excitotoxicity as a factor as the reduction in dopaminergic
transmission leads to a loss of regulation of striatal neurons, resulting in an in increase
in glutamatergic activity (Koutsilieri & Riederer, 2007; Meissner, 2011; Mony, 2009).
240 Targeting Calcium-mediated Excitotoxicity in the CNS
Intracranial injury, more commonly know as traumatic brain injury (TBI), occurs when
a head injury results in damage to neuronal structure and integrity. There are two
phases of TBI; the acute initial damage from the insult and the secondary injury that
are caused by excitotoxicity. Much like the progressive neurodegenerative diseases
already discussed and acute stroke, glutamate dysregulation leads to calcium influx
into neurons close to the region of initial insult, which results in necrosis of neurons.
TBI is complicated as there may or may not be blood flow alterations, hypoxia,
swelling, and intracranial pressure. Presently, no pharmacological intervention is
available to treat this secondary, excitotoxic cascade that results from TBI. Like other
excitotoxic diseases, NMDA receptor antagonists showed promise in animal studies
but failed to show efficacy in human clinical trials.
Homeostatic plasticity is the broad term used to describe all of the molecular plastic
changes that a neuron uses to govern and adjust its own intrinsic excitability. One
common underlying method that neurons use to make these adjustments is synaptic
scaling (Thalhammer & Cingolani, 2014). That is, the neuron will adjust the properties
of ion channels to meet a certain set point. This is a process that was first observed
in neurons adjusting their excitatory response to glutamate release after chronic
manipulations of their activity. In vitro experiments with neurons have demonstrated
that chronic blockage of either NMDA receptors or AMPA receptors results in a
process that looks like LTP (Lee & Chung, 2014). There have been reports of enhanced
trafficking of iGluRs to the surface of the cell and delivery to synaptic locations after
treatments. Thus, we draw a parallel to these findings and suggest that full blockage
of iGluRs could lead to homeostatic plasticity induction and this would be counter-
productive to reducing excitotoxicity, since it would result in an increased expression
of just the channels that clinicians are trying to modulate. We hypothesize that a
partial antagonist will be effective in reducing spine loss and neuron death due to
excitotoxic calcium influx. In fact, one of the few pharmacological agents that actually
demonstrates some efficacy on slowing excitotoxic effects is memantine, a partial and
weak antagonist of NMDA receptors (Gardoni & Di Luca, 2006).
With an increasingly large aging population, the numbers of people afflicted with the
diseases of excitotoxicity will continue to grow. Treatment of the effects of calcium-
mediated excitotoxicity should be at the forefront of our collective efforts to treat
these diseases. In too few cases, there are obvious genetic targets that can potentially
be addressed with the new genome editing technology that is currently maturing.
However, for the vast majority of sporadic cases of disease, the root cause is either
undetected genomic mutations or environmental or developmental insults. For these
categories, genome editing has no known target and thus, must be tackled at the site
of molecular damage. With calcium-mediated excitotoxicity implicated in so many
neurological diseases, the prize for the scientist who solves this problem will likely be
substantial and society will be indebted to them.
References
Araki, Y., Lin, D. T., Huganir, R. L. (2010). Plasma membrane insertion of the AMPA receptor GluA2
subunit is regulated by NSF binding and Q/R editing of the ion pore. Proceedings of the
National Academy of Sciences, 107, 11080-11085.
Ballard, C., Gauthier, S., Corbett, A., (2011). Alzheimer’s disease. Lancet, 377, 1019-1031.
Beique, J. C., Na, Y., Kuhl, D., (2011). Arc-dependent synapse-specific homeostatic plasticity.
Proceedings of the National Academy of Sciences, 108, 816-821.
Benatar, M., & Wuu, J. (2012). Presymptomatic studies in ALS: rationale, challenges, and approach.
Neurology, 79, 1732-1739.
Carriedo, S. G., Yin, H. Z., & Weiss, J. H. (1996). Motor neurons are selectively vulnerable to AMPA/
kainate receptor-mediated injury in vitro. Journal of Neuroscience, 16, 4069-4079.
Corona, J. C., & Tapia, R. (2007). Ca2+-permeable AMPA receptors and intracellular Ca2+ determine
motoneuron vulnerability in rat spinal cord in vivo. Neuropharmacology, 52, 1219-1228.
Corona, J. C., Tovar-y-Romo, L. B., & Tapia, R. (2007). Glutamate excitotoxicity and therapeutic
targets for amyotrophic lateral sclerosis. Expert Opinion on Therapeutic Targets, 11, 1415-1428.
Cotman, C. W., & Monaghan, D. T. (1986). Anatomical organization of excitatory amino acid receptors
and their properties. Advances in Experimental Medicine and Biology, 203, 237-252.
Debelleroche, J. S., & Bradford, H. F. (1977). Site of Origin of Transmitter Amino-Acids Released by
Depolarization of Nerve-Terminals Invitro. Journal of neurochemistry, 29, 335-343.
References 243
Gardoni, F., & Di Luca, M. (2006). New targets for pharmacological intervention in the glutamatergic
synapse. European Journal of Pharmacology, 545, 2-10.
Grosskreutz, J., Van Den Bosch, L., & Keller, B. U. (2010). Calcium dysregulation in amyotrophic
lateral sclerosis. Cell Calcium, 47, 165-174.
Hardingham, G. E., & Bading, H. (2010). Synaptic versus extrasynaptic NMDA receptor signalling:
implications for neurodegenerative disorders. Nature Reviews Neuroscience, 11, 682-696.
Henley, J. M., Barker, E. A., & Glebov, O. O. (2011). Routes, destinations and delays: recent advances
in AMPA receptor trafficking. Trends in Neurosciences, 34, 258-268.
Hideyama, T., Yamashita, T., Aizawa, H., (2012). Profound downregulation of the RNA editing enzyme
ADAR2 in ALS spinal motor neurons. Neurobiology of Disease, 45, 1121-1128.
Hu, N. W., Ondrejcak, T., & Rowan, M. J. (2012). Glutamate receptors in preclinical research on
Alzheimer’s disease: update on recent advances. Pharmacology Biochemistry and Behavior,
100, 855-862.
Isaac, J. T., Ashby, M. C., & McBain, C. J. (2007). The role of the GluR2 subunit in AMPA receptor
function and synaptic plasticity. Neuron, 54, 859-871.
Kang, J. E., Lim, M. M., Bateman, R. J., (2009). Amyloid-beta dynamics are regulated by orexin and
the sleep-wake cycle. Science, 326, 1005-1007.
Keller, B. U., Hollmann, M., Heinemann, S., (1992). Calcium influx through subunits GluR1/GluR3
of kainate/AMPA receptor channels is regulated by cAMP dependent protein kinase. EMBO
Journal, 11, 891-896.
Kerchner, G. A., & Nicoll, R. A. (2008). Silent synapses and the emergence of a postsynaptic
mechanism for LTP. Nature Reviews Neuroscience, 9, 813-825.
Koutsilieri, E., & Riederer, P. (2007). Excitotoxicity and new antiglutamatergic strategies in
Parkinson’s disease and Alzheimer’s disease. Parkinsonism & Related Disorders, 13(Suppl 3),
S329-331.
Lau, A., & Tymianski, M. (2010). Glutamate receptors, neurotoxicity and neurodegeneration. Pflugers
Archiv, 460, 525-542.
Lee, K. Y., & Chung, H. J. (2014). NMDA receptors and L-type voltage-gated Ca(2)(+) channels mediate
the expression of bidirectional homeostatic intrinsic plasticity in cultured hippocampal
neurons. Neuroscience, 277, 610-623.
Li, S., Hong, S., Shepardson, N. E., (2009). Soluble oligomers of amyloid Beta protein facilitate
hippocampal long-term depression by disrupting neuronal glutamate uptake. Neuron, 62, 788-801.
Lipton, S. A. (2006). Paradigm shift in neuroprotection by NMDA receptor blockade: memantine and
beyond. Nature Reviews Drug Discovery, 5, 160-170.
Lisman, J., Yasuda, R., & Raghavachari, S. (2012). Mechanisms of CaMKII action in long-term
potentiation. Nature Reviews Neuroscience, 13, 169-182.
McDermott, C. M., LaHoste, G. J., Chen, C., (2003). Sleep deprivation causes behavioral, synaptic,
and membrane excitability alterations in hippocampal neurons. Journal of Neurosciences, 23,
9687-9695.
Meissner, W. G., Frasier, M., Gasser, T., (2011). Priorities in Parkinson’s disease research. Nature
Reviews Drug Discovery, 10, 377-393.
Meyer, D. K., Lindemeyer, A. K., Wilmes, T., (2012). GluA and GluN receptors regulate the
surface density of GluN receptor subunits in cultured neocortical interneurons. Journal of
Neurochemistry, 121, 597-606.
Milnerwood, A. J., Gladding, C. M., Pouladi, M. A., (2010). Early increase in extrasynaptic NMDA
receptor signaling and expression contributes to phenotype onset in Huntington’s disease
mice. Neuron, 65, 178-190.
Mony, L., Kew, J. N., Gunthorpe, M. J., (2009). Allosteric modulators of NR2B-containing NMDA
receptors: molecular mechanisms and therapeutic potential. British Journal of Pharmacology,
157, 1301-1317.
244 Targeting Calcium-mediated Excitotoxicity in the CNS
Moore, L. E., & Buchanan, J. T. (1993). The effects of neurotransmitters on the integrative properties
of spinal neurons in the lamprey. Journal of Experimental Biology, 175, 89-114.
Okamoto, S. I., Pouladi, M. A., Talantova, M., (2009). Balance between synaptic versus extrasynaptic
NMDA receptor activity influences inclusions and neurotoxicity of mutant huntingtin. Nature
Medicine, 15, 1407-U1408.
Renton, A. E., Chio, A., & Traynor, B. J. (2014). State of play in amyotrophic lateral sclerosis genetics.
Nature Neuroscience, 17, 17-23.
Ribrault, C., Sekimoto, K., & Triller, A. (2011). From the stochasticity of molecular processes to the
variability of synaptic transmission. Nature Reviews Neroscience, 12, 375-387.
Suhs, K. W., Fairless, R., Williams, S. K., (2014). N-methyl-D-aspartate receptor blockade is
neuroprotective in experimental autoimmune optic neuritis. Journal of Neuropathology &
Experimental Neurology, 73, 507-518.
Szabo, A., Somogyi, J., Cauli, B., (2012). Calcium-permeable AMPA receptors provide a common
mechanism for LTP in glutamatergic synapses of distinct hippocampal interneuron types.
Journal of Neuroscience, 32, 6511-6516.
Szydlowska, K., & Tymianski, M. (2010). Calcium, ischemia and excitotoxicity. Cell Calcium, 47, 122-
129.
Thalhammer, A., & Cingolani, L. A. (2014). Cell adhesion and homeostatic synaptic plasticity.
Neuropharmacology, 78, 23-30.
Turner, M. R., Hardiman, O., Benatar, M., (2013). Controversies and priorities in amyotrophic lateral
sclerosis. Lancet Neurology, 12, 310-322.
Van Damme, P., Van Den Bosch, L., Van Houtte, E., (2002). GluR2-dependent properties of AMPA
receptors determine the selective vulnerability of motor neurons to excitotoxicity. Journal of
Neurophysiology, 88, 1279-1287.
Van Den Bosch, L., & Robberecht, W. (2000). Different receptors mediate motor neuron death
induced by short and long exposures to excitotoxicity. Brain Research Bulletine, 53, 383-388.
Van den Bosch, L., Van Damme, P., Vleminckx, V., (2002). An alpha-mercaptoacrylic acid derivative
(PD150606) inhibits selective motor neuron death via inhibition of kainate-induced Ca2+ influx
and not via calpain inhibition. Neuropharmacology, 42, 706-713.
Wang, G., Gilbert, J., & Man, H. Y. (2012). AMPA receptor trafficking in homeostatic synaptic
plasticity: functional molecules and signaling cascades. Neural Plasticity, 2012, 825364.
Wen, J. A., & Barth, A. L. (2012). Synaptic lability after experience-dependent plasticity is not
mediated by calcium-permeable AMPARs. Frontiers in Molecular Neuroscience, 5, 15.
Williams, C. J., & Dexter, D. T. (2014). Neuroprotective and symptomatic effects of targeting group III
mGlu receptors in neurodegenerative disease. Journal of Neurochemistry, 129, 4-20.
Wright, A., & Vissel, B. (2012). The essential role of AMPA receptor GluR2 subunit RNA editing in the
normal and diseased brain. Frontiers in Molecular Neuroscience, 5, 34.
Rebecca Fransson, Christian Sköld and Anja Sandström*9
3.3 Strategies for Conversion of Peptides to Peptido-
mimetic Drugs
Figure 3.3.1: Drug distribution in the human body is determined by its ADME properties.
Rebecca Fransson, Christian Sköld and Anja Sandström: Division of Organic Pharmaceutical
Chemistry, Department of Medicinal Chemistry, BMC, Box 574, Uppsala University, SE-751 23
Uppsala, Sweden, *Email: anja.sandstrom@orgfarm.uu.se
Furthermore, peptides have a large degree of conformational flexibility, and can fold
into complex tertiary structures crucial for their molecular recognition and their
ability to produce a biological response. In order to fully benefit from the potential
of biologically active peptides in chemical biology and in drug discovery, different
approaches to overcome their limitations are needed.
There are several strategies available that can be utilized for the development
of drugs that modulates the physiological events that are triggered by peptides. One
common approach is high-throughput screening (HTS) of libraries of small molecules.
This approach only considers the macromolecular target and can be utilized if the target
is known and a suitable biochemical assay is available. Peptide-based therapeutics
is another, currently growing, area that instead focuses on the biologically active
peptide and aims to circumvent the limitation of peptides by introducing various
modifications (e.g. PEG and phosphoesters linkages) or by employing different
ways of administration (e.g. parenteral, mucosal, oral and transdermal routes). As
a complement to HTS and peptide modifications, rational design based on step-wise
transformation of peptides into low-molecular-weight and bioavailable drug-like
molecules that mimic the action of peptides, i.e. peptidomimetics, is a viable way to
overcome the problems associated with peptides (Hruby, 2002; Olson, 1993; Ripka,
1998; Vagner, 2008). Peptidomimetics are molecules with significantly reduced
peptide character that mimic the bioactive conformation of peptides, and thus retain
the ability to interact with the biological target and cause the same biological effect
(Grauer, 2009). As these compounds are non-peptides which often possess desired
improved pharmacokinetic properties, such as better absorption, metabolic stability
and/or bioavailability. There are several successful examples where a rational design
approach has resulted in approved drugs, e.g. the development of protease inhibitors
such as angiotensin-converting enzyme (ACE) inhibitors, HIV protease inhibitors,
and hepatitis C virus (HCV) inhibitors.
The activity at the target and the exposure (e.g. concentration and duration)
determine the efficacy of a drug. In the body, several barriers to drug exposure can
be found, for example cell membranes, metabolic enzymes, efflux transporters,
and binding proteins. How a compound performs at a specific barrier is connected
to its drug properties. In the GI tract, compounds can cross the cellular membrane
barrier by three major mechanisms (van de Waterbeemd, 2001). The two most
common are transcellular absorption, i.e. passive transfer by diffusion across the
lipid membranes, and paracellular absorption, which proceeds through aqueous
pores at the tight junctions between the cells. The third mechanism is active uptake
by transport proteins that usually transport nutrients across the membrane (Kerns,
2008; van de Waterbeemd, 2001). The most important mechanism for drug absorption
is passive diffusion, and about 95% of all commercial drugs are absorbed by this route
(Kerns, 2008). Metabolizing enzymes in the GI tract, e.g. the cytochrome P450 (CYP)
enzymes and proteolytic enzymes as well as efflux transporters, e.g. P-glycoprotein
(PgP), are expressed, which limit the oral absorption of compounds (Fig. 3.3.3) (van
de Waterbeemd, 2001). In addition to avoiding the efflux of PgP and metabolism by
gut wall enzymes, good permeability is important to maximize the oral absorption
of a compound. In general, therapeutic peptides suffer from short half-lives due to
rapid degradation by proteolytic enzymes of the GI tract and blood plasma, rapid
clearance from the circulation by liver and kidneys, as well as limited permeability
across physiological barriers because of their hydrophilic structure (Vlieghe, 2010).
Especially CYP3A4 and PgP have been shown to have a significant impact on the
248 Strategies for Conversion of Peptides to Peptidomimetic Drugs
During rational peptide lead optimization various in vitro property assays (e.g.
solubility, permeability, chemical stability, metabolism, protein binding and transport)
are used to assess the drug-like properties of the generated peptide analogues (Kerns,
2008). Extensive in vitro profiling regarding the PK data of lead compounds in the
early stages of development can thus provide the medicinal chemist with information
on the structure-property relationship important for the further development of orally
active compounds.
A B
i. N- to C-terminal Internal N-methylation
i
α-Carbon methylation C-terminal methylation
ii
R O
H
R O R N
H H2 N OH
N OH
H2 N N O R
H iv ii. Side chain to
O R O
side chain N-terminal methylation Side chain methylation
iii
iii. Backbone to backbone iv. Backbone to side chain
C D
H
O S
S
N N H 2N
H H 2N O
HO
O OH
H O
N OH β-turn mimetic β-turn mimetic
O R OH
H R
Retro-inversion Hydroxyethylene (E)-alkene N
O S O
H 2N N
N N
H H OH
O N3
Reduced Reduced Thioamide OH
O
β-turn mimetic γ-turn mimetic
Figure 3.3.4: A) Cyclization strategies that can be performed in a peptide sequence to introduce
global constraints (Rizo & Gierasch, 1992). B) Positions in a peptide that can be methylated.
Methylation of both the backbone atoms and the side chains can introduce local constraints (Rizo,
1992). C) Isosteric replacement of the peptide bond can introduce local constraints. It should be
noted that such replacements are not always real constraints, but alter the overall conformational
behavior of the peptide backbone to varying degrees. In some cases the flexibility is increased (Rizo,
1992; Sewald, 2002). D) Secondary structure mimetics can induce a desired conformation when
introduced into the peptide backbone (Rosenström, 2006; Schmidt, 1998; Sewald, 2002).
with the environment (Rose, 1985). Therefore, to limit the conformational flexibility,
conformational constraints can be introduced into the peptide to provide information
about the bioactive conformation. In many cases the constrained part of the peptide is
not completely rigid (e.g., using side chain cyclizations as constraints) and therefore
several differently constrained analogues of the peptide may be required in order to
derive a putative receptor-bound conformation. When peptides bind to their receptors
they become an integral part of the protein structure and thus can be predicted to
adopt secondary structure motifs as part of the bioactive conformation (Hruby, 2002).
Therefore, conformational constraints or organic scaffolds that induce or mimic
secondary structures can give valuable information when searching for the bioactive
conformation and may provide a valuable starting point for the development of
peptidomimetic drugs.
ψi+ 1
Ri+1 O φ Ri+2
i+2
ψi+2 R i+1 ψi+1
φi+1 φi+1
N O O
HN H HN
H R i+2
N R i+3 N N
O H
Ri
O H
NH O Ri O
A B
HO
OH
H
N
O
O H HN
N
HN N N
O O
C D
Figure 3.3.5: A β-turn (A) and a γ-turn (B) and examples of a β-turn mimetic (C) (Rosenström, 2006)
and a γ-turn mimetic (D) (Rosenström, 2005). The backbone torsion angles phi (φ) and psi (ψ) are
indicated for the relevant residues in the peptide turns.
Table 3.3.1 Idealized backbone torsion angles for the most common β-turn types (Venkatachalam,
1968; Wilmot, 1988) and the backbone torsion angles for the classic and inverse γ-turn (Matthews,
1972; Nemethy, 1972).
The γ-turn is not as common as the β-turn but has been found to be adopted by
peptides as short as tripeptides (Motta, 2005). A γ-turn spans over three amino acid
residues with a hydrogen bond between COi and NHi+2, resulting in the shape of a
seven-membered ring. There are two types of γ-turns: the classic γ-turn with the i+1
residue side chain in an axial geometry and the inverse γ-turn with the side chain in
an equatorial geometry (Table 3.3.1) (Matthews, 1972; Nemethy, 1972). Although the
classic γ-turn was proposed first, it has been shown to be very rare and the inverse
γ-turn is the significantly more common of the two (Milnerwhite, 1988). The classic
γ-turn conformation causes reversal of the backbone direction and is mainly found
in β-hairpin structures, as a tight turn, forming an antiparallel β-sheet structure.
The inverse γ-turn can be found as backbone kinks and rarely causes reversal of the
backbone direction (Milnerwhite, 1988).
The information gained from the investigation of the bioactive conformation
can provide indications regarding the presence of a turn structure in the bioactive
conformation. Attempts to replace this part of the peptide with a turn mimetic would
then be a rational step towards a peptide mimicking compound. However, how does
one know if a particular organic scaffold indeed does mimic a turn, and in that case,
which specific class of turns? One way of characterizing a potential turn mimicking
moiety is to compare the geometry with experimental turn structures found in proteins
(Claerhout, 2012; Rosenström, 2006; Whitby, 2011). The great abundance of protein
3D structures determined by X-ray crystallography in the Protein Data Bank (Berman,
2000) provides a good source of experimental protein secondary structures that can
be used for this purpose. Using this method, both β- and γ-turn mimicking scaffolds
have been characterized, which are exemplified below.
scaffold C has been used to replace amino acids in the center of Ang II to yield high
affinity Ang II type 1 (AT1) and Ang II type 2 (AT2) receptor ligands (Table 3.3.2, entries
2 and 3).
Figure 3.3.6: Low energy conformation of β-turn mimetic scaffold C (dark grey carbons) superimpo-
sed to a type II β-turn (light grey carbons; PDB ID 1H2C, chain A, sequence Ile142-Pro143-Asp144-His145).
For clarity, only Cβ is shown for the side chains and non-polar hydrogen atoms are omitted.
Table 3.3.2: AT1- and AT2 receptor affinity of Ang II and Ang II mimicking pseudopeptides.
HN
H2 N NH
Continued
Table 3.3.2: AT1- and AT2 receptor affinity of Ang II and Ang II mimicking pseudopeptides.
H 2N NH
a
Rat liver membranes
b
Pig uterus myometrium
OH
O
HN Cα 3.8 Å 3.8-3.9 Å O
H i+1 92 o H HN
N N N
Cα i+2 44o 43-44 o N
O Cα i O H O
NH 5.5 Å
O NH
O
Figure 3.3.7: Comparison of Cα distances and angles in an ideal γ-turn model peptide (left) and
corresponding atoms in the γ-turn mimicking scaffold D (right).
256 Strategies for Conversion of Peptides to Peptidomimetic Drugs
Although the distances and angles correspond well between an idealized γ-turn and
scaffold D, such an analysis does not include the incoming and outgoing directions of
the peptide backbone or the direction of the side chain of residue i+1. As in the case
of the β-turn mimetic classification, one way of characterizing the γ-turn mimetic is to
compare relevant parts to a determined 3D structure of a γ-turn. In Fig. 3.3.8, an energy-
minimized conformation of scaffold D can be seen superimposed on a γ-turn found in
a protein 3D structure, which shows the good match between the structures.
Figure 3.3.8: Low energy conformation of scaffold D superimposed on an inverse γ-turn (light grey
carbons, PDB ID 1NNF, chain A, sequence Thr82-Ala83-Gln84). For clarity, only Cβ is shown for the side
chains and non-polar hydrogen atoms are omitted.
3.3.2.2.1 Strategy
A SAR study of the two peptide leads, SP1–7 and EM-2, was initiated, with the intention
of identifying the pharmacophoric groups. This design strategy included Ala scans,
truncations and C- and N-terminal modifications of the two target peptides (Fig.
3.3.9). Thus, a series of peptide analogs, in which each amino acid residue of the two
target peptides was replaced sequentially with an alanine, was synthesized. In the
truncation studies, one amino acid at a time was removed from the N-terminal. Both
C-terminal carboxylic acids and carboxamides were included.
NH NH 2
N-terminal acylation
NH 2
O
N
H2 N O OH
O N H
N N N
O H H
O O
N
O H O
N-terminal truncations Substitution with a primary amide
O
H 2N
Alanine substitution of
each amino acid residue
HO
N O
H2 N H
O N NH 2
N
O H
O
N-terminal truncations
Figure 3.3.9: Illustration of the modifications of A) SP1–7 and B) EM-2, used in the SAR study.
Table 3.3.3: Ki values of SP1–7 and EM-2 analogs for inhibition of [3H]-SP1–7 binding to rat spinal cord
membrane.
H 2N NH
NH NH 2 HO
NH 2
O
N N O
H2 N O OH H2 N H
O N H O N NH 2
N N N N
O H O H
O H O O
SP1-7 N EM-2
O H O
O
H 2N
Alanine-substituted SP1–7
8 H-Arg-Pro-Lys-Pro-Gln-Ala-Phe-OH 365 ± 3
Alanine-substituted EM-2
13 H-Tyr-Pro-Phe-Ala-NH2 1460 ± 15
Table 3.3.3: Ki values of SP1–7 and EM-2 analogs for inhibition of [3H]-SP1–7 binding to rat spinal
Continued
cord membrane.
27 H-Phe-NH2 5028 ± 31
The results were remarkably consistent for SP1–7 and EM-2 peptides (Table 3.3.3).
Substitution with alanine in the N-terminal part of SP1–7 was well tolerated without
affecting the binding affinity significantly (cf. 1 with 4, 5 and 6), although replacement
of the basic amino acid arginine rendered a 10-fold lower affinity (cf. 1 and 3). Likewise,
the three N-terminal amino acid residues in EM-2 (tyrosine, proline and the internal
phenylalanine) could be substituted with an alanine and still retain binding affinity
(cf. 2 with 10, 11 and 12). However, removal of the two primary amide functions of
the side chains of the glutamine in the C-terminal of SP1–7 resulted in a considerable
decrease in affinity (7 and 8). The C-terminal phenylalanine was absolutely crucial for
strong affinity in both SP1–7 and EM-2. Replacement of the C-terminal phenylalanine in
SP1–7 gave analog 9, which was devoid of affinity. Making the same substitution in EM-2
resulted in compound 13 with a Ki value of 1460 nM. The finding that the C-terminal
phenylalanine plays such an important role in binding affinity is in line with the
peptide scan discussed above, where the C-terminal fragment SP1–6 possessed very
weak binding affinity to the SP1–7 binding site (Botros, 2006; Igwe, 1990). Moreover, the
potency of SP1–7 showed a five-fold increase upon amidation of the terminal carboxyl
group (15). Similarly, the binding affinity of EM-2 was reduced by a factor of four upon
A Case study of Rational Peptide Lead Optimization: Development of Small and Constrained... 261
removal of the amide function (16). Since SP1–7 is a proteolytic product of SP resulting
in a C-terminal carboxylic acid, it was surprising that the amidated analog 15 led to
improved binding affinity.
As deduced from the two Ala scans, the N-terminal parts of SP1–7 and EM-2 does
not seem to be engaged in molecular recognition or binding to the target protein, a fact
that prompted synthesis of the truncated analogs 17–24 and 25–27. For SP1–7, removal
of the N-terminal arginine rendered the hexapeptide 17 with a 20-fold reduction
in affinity compared to SP1–7 and a little more than 2-times lower affinity than the
alanine derivative 3. The affinity could, however, be recovered by amidation of 17,
resulting in peptide 18, which was 10 times more potent. Further truncation down to
the tripeptide level was possible without loss of affinity, and C-terminal amidation of
all the truncated SP1–7 analogs improved the affinity 5–10-fold. Hence, the tripeptide
H-Gln-Gln-Phe-NH2 (24) exhibited a Ki of 1.9 nM.
For EM-2, simultaneous removal of the two N-terminal amino acids tyrosine and
proline improved the binding affinity six times, resulting in the notable discovery of
the dipeptide H-Phe-Phe-NH2 (26) with a Ki value similar to that of SP1–7 itself. It should
be emphasized that the Tyr-Pro sequence is the critical fragment for binding to the
µ-receptor, whereas the two C-terminal phenylalanines are not, facts that highlight
the double nature of EM-2 (Fichna, 2007; Kruszynski, 2005; Okada, 2003). In the
Ala scan of EM-2, substitution of the internal phenylalanine was accepted (12), but
truncation down to a single phenylalanine resulted in 27 with no binding affinity (Ki
of 5028 nM).
The binding features of the new dipeptide lead compound H-Phe-Phe-NH2 were
further explored via the synthesis of peptides 28–32 (Table 3.3.4). As observed for SP1–7
and EM-2, the C-terminal function should be a primary amide. The corresponding
carboxylic acid was devoid of activity (cf. 26 with 28). All four stereoisomers of H-Phe-
Phe-NH2 (26, 30, 31 and 32) were synthesized and evaluated. The natural l-Phe-l-Phe
isomer (26) was found to be preferred, followed by the d,d compound (31), although
with a 40-fold lower affinity. As mentioned above, incorporation of d-amino acids into
neuropeptides can change their biological function from an agonist to an antagonist,
e.g. d-SP1–7, which makes it interesting to evaluate the analogs 30, 31 and 32 in animal
studies concerning their functional activities.
Due to the smaller size of 26 in comparison to the heptapeptide SP1–7, lower
selectivity can be expected. Moreover, 26 resembles ligands for the NK3 receptor
(Boden, 1995; 1994). Hence, the possible binding affinity of 26 to the human neurokinin
receptors NK1 and NK3 was studied. Binding was evaluated in agonist radioligand
binding assays relying on the displacement of [Sar9, Met(O2)11]-SP from NK-1 receptors,
and [MePhe7]-NKB from NK-3 receptors (Anthes, 2002; Heuillet, 1993), 26 was tested at
a concentration of 10 µM, but showed no affinity for any of the receptors.
262 Strategies for Conversion of Peptides to Peptidomimetic Drugs
Table 3.3.4: Ki values of Phe-Phe analogs for inhibition of [3H]-SP1–7 binding to rat spinal cord
membrane.
O
H
N (S)
H2 N (S) NH 2
O
O
H
N (S)
H2 N (S) OH
O
29 18.5 ± 1.7
O O
H
N (S)
N (S) NH 2
H
O
Phe-Phe Analogs
30 540 ± 20
O
H
N (R)
H2 N (S) NH 2
O
31 64 ± 2
O
H
N (R)
H2 N (R) NH 2
O
32 175 ± 13
O
H
N (S)
H2 N (R) NH 2
O
a
S configuration = l configuration and R configuration = d configuration.
A Case study of Rational Peptide Lead Optimization: Development of Small and Constrained... 263
H-Phe-Phe-NH2
SP1-7
Figure 3.3.10: The antinociceptive effect of SP1–7 (1), SP1–7-NH2 (15) and H-Phe-Phe-NH2 (26) in
non-diabetic (left) and diabetic (right) mice. The antinociceptive effect was evaluated by the AUC
calculated from the time-response curve of tail-flick latency. Each column represents the mean with
S.E.M. (n = 6).
264 Strategies for Conversion of Peptides to Peptidomimetic Drugs
3.3.2.3.1 Strategy
The potent dipeptide lead H-Phe-Phe-NH2 (26), discussed above, was chosen for
further optimization studies with the overall aim of developing metabolically stable
and selective SP1–7 analogs. The introduction of local constraints can enhance stability,
selectivity and bioavailability. The intestinal permeability is an important factor in the
development of orally bioavailable drugs. In the intestine, the di/tri-peptide transporter
PepT1 enables the absorption of small peptides from the digestion of dietary proteins.
This transport system has also been shown to transport a variety of peptidomimetic
drugs, such as β-lactam antibiotics and ACE inhibitors and might be exploited in
order to increase the absorption of our small compounds (Brandsch, 2009; Brodin,
2002; Rubio-Aliaga, 2002). A known problem with peptides is their susceptibility to
efflux. For peptides targeting functions in the CNS, uptake in the brain, i.e. crossing
the BBB, is a crucial factor. As a defense mechanism preventing harmful substances
from entering the brain, the BBB is equipped with efflux transporters (Witt, 2001). PgP
is one of the most important, and can actively transport substances out of the brain
(Giacomini, 2010). Such transporters can be an obstacle to entering the CNS.
A series of H-Phe-Phe-NH2 analogs (33–43, Tables 3.3.5 and 3.3.6) incorporating
different types of constraints were designed, synthesized and evaluated regarding
their binding affinity, stability, uptake and permeability (Fig. 3.3.11). N-methyl
and α-methyl amino acids were incorporated, substituting one residue at a time.
Furthermore, β-methylation of the phenylalanine side chain was used to reduce the
conformational flexibility, which can be advantageous upon binding. This approach
has been successful in other projects in obtaining neuropeptide analogs resistant to
metabolism while still retaining their biological activity (Veber, 1985). Both N- and
C-terminal rigidifications were accomplished by the introduction of a 3-phenylproline
derivative (Sewald, 2002).
Table 3.3.5: Binding affinity (Ki values) and metabolic stability (Clint and t1/2) of the methylated H-Phe-
Phe-NH2 analogs
34 70 ± 3 64 ± 11 22 ± 4
O
H
N (S)
H 2N (S) NH 2
O
35 9.4 ± 0.1 92 ± 0 15 ± 1
O
N (S)
H 2N (S) NH 2
O
36 26 ± 1 38 ± 15 40 ± 16
O
H
N (S)
H 2N (S) NH 2
O
a
Ki value determined on the same occasion as for 33–46. b Previously determined Ki value c The meta-
bolic stability data are expressed as mean ± SD.
d
Clint = in vitro intrinsic clearance. e t1/2 = in vitro half-life.
266 Strategies for Conversion of Peptides to Peptidomimetic Drugs
Table 3.3.6: Binding affinity (Ki values) and metabolic stability (Clint and t1/2) of the rigidified and
C-terminal-modified H-Phe-Phe-NH2 analogs
38a 34 ± 3 28 ± 3 50 ± 6
(S) O
(S)
H
N (S)
N NH 2
H
O
39a 51 ± 2 16 ± 9 103 ± 6
(R)
O
(R) H
N (S)
N NH 2
H
O
O NH 2
O
42a 18 ± 1 39 ± 4 36 ± 3
O
H 2N (S)
HN (S)
(S)
O NH 2
43 a
68 ± 1 16 ± 7 98 ± 4
O
H2N (S)
HN (R)
(R)
O NH 2
a
The stereochemistry of each diastereomer pair was estimated from the pharmacophore model. b
IC50 value. Analog 40 was tested at a 2:1 ratio with analog 41. It was tested once, at six different con-
centrations, in triplicate. c The metabolic stability data are expressed as mean ± SD. d Clint = in vitro
intrinsic clearance. e t1/2 = in vitro half-life.
268 Strategies for Conversion of Peptides to Peptidomimetic Drugs
Table 3.3.7: Active uptake and permeability data for the methylated H-Phe-Phe-NH2 analogs.
37 0.3 ± 0.1 0.4 ± 0.1 0.7 0.3 ± 0.1 0.9 ± 0.2 0.3 3
O
H
N (S)
H 2N (S) N
H
O
a
The results are expressed as mean ± SD. b Papp = apparent permeability coefficient. c a–b = apical to
basolateral. d b–a = basolateral to apical.
Conclusion 269
A Papp value below 0.2 × 10-6 cm s-1 indicates low permeability, a Papp value ranging from
0.2 × 10-6 cm s-1 to 1.6 × 10-6 cm s-1 indicates moderate permeability and a Papp value above
1.6 × 10-6 cm s-1 indicates high permeability, in this particular setting (Bergström, 2003).
The permeability in the a–b direction increased substantially for all the methylated
analogs (ranging from 0.3 × 10-6 cm s-1 to 20 × 10-6 cm s-1) compared to H-Phe-Phe-NH2
(26, 0.02 × 10-6 cm s-1). All the compounds, except 37, were classified as having high
permeability, with the highest permeability observed for the α-methylated analogs
34 and 36, 18 × 10-6 cm s-1 and 20 × 10-6 cm s-1, respectively. However, the methylated
compounds were not actively transported, as can be seen from the PepT1/K1 ratio.
Furthermore, the peptides also displayed efflux (the ba/ab ratio ranging from 3 to 14),
where compounds 35 and 36 showed the highest tendency towards efflux and were
equal to 26.
Compared to the methylated analogs, the rigidified and the C-terminal-
phenylalanine-modified analogs possessed much lower permeability (ranging from
0.02 × 10-6 cm s-1 to 4.4 × 10-6 cm s-1). Satisfyingly, the high affinity analog 40 turned out
to have high permeability (3.6 × 10-6 cm s-1). In this series, the efflux was much higher;
8−95 times, in the b–a direction than in the a–b direction. The stereochemistry of
the compounds also seemed to influence the predisposition for efflux. Thus, the
compounds 38, 40 and 42 all showed lower efflux ratios than their diastereomeric
counterparts 39, 41 and 43. The replacement of the phenylalanine in the N- or
C-terminal by the 3-phenylproline moiety seemed advantageous in improving the
active uptake of these compounds, since a slight increase in the PepT1/K1 ratio was
observed (cf. 26 vs. 38, 39, 40 and 41).
3.3.3 Conclusion
The case study presented herein is an illustrative example of rational design of drug-like
molecules mimicking the actions of a bioactive peptide. It highlights the importance
working in an iterative manner, where structural modifications of compounds are
evaluated both from a pharmacodynamic and pharmacokinetic perspective.
The optimization process, starting with the bioactive heptapeptide SP1-7 and the
tetrapeptide EM-2, resulted in the remarkable discovery of the dipeptide H-Phe-Phe-
NH2, which was equipotent to endogenous SP1-7 and had a higher binding affinity
than EM-2. However, in the ADME assessment, the dipeptide showed poor metabolic
stability and permeability, and suffered from high efflux rates. By introducing local
constraints in the C-terminal of H-Phe-Phe-NH2, via the introduction of the cis
3-phenyl-pyrrolidine moiety, the pharmacokinetic properties could be substantially
improved and the binding affinity retained.
It should be emphasized that although numerous strategies and attempts for
transforming a bioactive peptide into small peptidomimetics have been reported, it
is not straightforward, it is case dependent, and there is no guarantee of success.
270 Strategies for Conversion of Peptides to Peptidomimetic Drugs
The decreasing number of approved drugs during recent years has put enormous
pressure on the pharmaceutical industry, resulting in a revival of interest in peptides
as potential drug candidates (Vlieghe, 2010). By using synthetic strategies to limit
metabolism and exploring alternative routes of administration, a number of peptidic
drugs have been brought to the market (Vlieghe, 2010), showing that it is not necessary
to remove the peptide character completely, and that small pseudopeptides can be
useful as drugs.
Table 3.3.8: Active uptake and permeability data for the rigidified and C-terminal-modified H-Phe-
Phe-NH2 analogs.
O NH 2
O
O NH 2
O
HN (S)
(S)
O NH 2
O
H2N (S)
HN (R)
(R)
O NH 2
a
The results are expressed as mean ± SD. b Papp = apparent permeability coefficient. c a–b = apical to
basolateral. d b–a = basolateral to apical.
References 271
References
Anthes, J. C., Chapman, R. W., Richard, C., et al. (2002). SCH 206272: a potent, orally active tachykinin
NK1, NK2, and NK3 receptor antagonist. European Journal of Pharmacology, 450(2), 191-202.
Bergström, C. A. S., Strafford, M., Lazorova, L., et al. (2003). Absorption classification of oral drugs
based on molecular surface properties. Journal of Medicinal Chemistry, 46(4), 558-570.
Berman, H. M., Westbrook, J., Feng, Z., et al. (2000). The Protein Data Bank. Nucleic Acids Research,
28(1), 235-242.
Boden, P., Eden, J. M., Hodgson, J., et al. (1994). The rational development of small molecule tachykinin
NK3 receptor selective antagonists - the utilization of a dipeptide chemical library in drug design.
Bioorganic & Medicinal Chemistry Letters, 4(14), 1679-1684.
Boden, P., Eden, J. M., Hodgson, et al. (1995). The development of a novel series of non-peptide
tachykinin NK3 receptor selective antagonists. Bioorganic & Medicinal Chemistry Letters, 5(16),
1773-1778.
Botros, M., Hallberg, M., Johansson, T., et al. (2006). Endomorphin-1 and endomorphin-2 differentially
interact with specific binding sites for substance P (SP) aminoterminal SP1-7 in the rat spinal
cord. Peptides, 27(4), 753-759.
Brandsch, M. (2009). Transport of drugs by proton-coupled peptide transporters: pearls and pitfalls.
Expert Opinion on Drug Metabolism & Toxicology, 5(8), 887-905.
Brodin, B., Nielsen, C. U., Steffansen, et al. (2002). Transport of peptidomimetic drugs by the intestinal
di/tri-peptide transporter, PepT1. Pharmacology & Toxicology, 90(6), 285-296.
Carlsson, A., Ohsawa, M., Hallberg, et al. (2010). Substance P1-7 induces antihyperalgesia in diabetic
mice through a mechanism involving the naloxone-sensitive sigma receptors. European Journal
of Pharmacology, 626(2-3), 250-255.
Chang, M. M., Leeman, S. E., & Niall, H. D. (1971). Amino-Acid Sequence of Substance P. Nature-New
Biology, 232(29), 86-87.
Claerhout, S., Sharma, S., Sköld, C. et al. (2012). Synthesis of functionalized furopyrazines as
restricted dipeptidomimetics. Tetrahedron, 68(14), 3019-3029.
Culman, J., & Unger, T. (1995). Central Tachykinins - Mediators of Defense Reaction and Stress
Reactions. Canadian Journal of Physiology and Pharmacology, 73(7), 885-891.
De Araujo, J. E., Huston, J. P., & Brandao, M. L. (2001). Opposite effects of substance P fragments
C (anxiogenic) and N (anxiolytic) injected into dorsal periaqueductal gray. European Journal of
Pharmacology, 432(1), 43-51.
Ebner, K., Muigg, P., Singewald, G., et al. (2008). Substance P in stress and anxiety: NK-1 receptor
antagonism interacts with key brain areas of the stress circuitry. Annals of the New York
Academy of Sciences, 1144(Neural Signaling Opportunities for Novel Diagnostic Approaches and
Therapies), 61-73.
Erspamer, V. (1981). The tachykinin peptide family. Trends in NeuroSciences, 4(11), 267-269.
Fichna, J., Janecka, A., Costentin, et al. (2007). The endomorphin system and its evolving
neurophysiological role. Pharmacological Reviews, 59(1), 88-123.
Fransson, R., Botros, M., Nyberg, F., et al. (2008). Small peptides mimicking substance P (1-7) and
encompassing a C-terminal amide functionality. Neuropeptides, 42(1), 31-37.
Fransson, R., Botros, M., Sköld, C., et al. (2010a). Discovery of dipeptides with high affinity to the
specific binding site for substance P1-7. Journal of Medicinal Chemistry, 53(6), 2383-2389.
Fransson, R., Carlsson, A., Sköld, et al. (2010b). Design and Synthesis of Small Substance P (1-7)
Mimetics. Paper presented at the INRC 2010, Malmö, Sweden.
Fransson, R., Sköld, C., Kratz, J. M., et al. (2013). Constrained H-Phe-Phe-NH2 Analogues with High
Affinity to the Substance P 1-7 Binding Site and with Improved Metabolic Stability and Cell
Permeability. Journal of Medicinal Chemisrty, 56(12), 4953-4965.
272 Strategies for Conversion of Peptides to Peptidomimetic Drugs
Giacomini, K. M., Huang, S. M., Tweedie, D. J., et al. (2010). Membrane transporters in drug
development. Nature Reviews: Drug Discovery, 9(3), 215-236.
Grauer, A., & Konig, B. (2009). Peptidomimetics - A Versatile Route to Biologically Active Compounds.
European Journal of Organic Chemistry, (30), 5099-5111.
Hallberg, M., & Nyberg, F. (2003). Neuropeptide conversion to bioactive fragments--an important
pathway in neuromodulation. Current Protein & Peptide Science, 4(1), 31-44.
Harrison, S., & Geppetti, P. (2001). Substance P. International Journal of Biochemistry and Cell Biology,
33(6), 555-576.
Hasenohrl, R. U., Gerhardt, P., & Huston, J. P. (1991). Naloxone blocks conditioned place preference
induced by substance P and [pGlu6]-SP(6-11). Regulatory peptides, 35(3), 177-187.
Heuillet, E., Menager, J., Fardin, V., et al. (1993). Characterization of a Human Nk1 Tachykinin Receptor
in the Astrocytoma Cell-Line U-373 Mg. Journal of Neurochemistry, 60(3), 868-876.
Houston, J. B. (1994). Utility of in-Vitro Drug-Metabolism Data in Predicting in-Vivo Metabolic-
Clearance. Biochemical Pharmacology, 47(9), 1469-1479.
Hruby, V. J. (1982). Conformational Restrictions of Biologically-Active Peptides Via Amino-Acid Side-
Chain Groups. Life Sciences, 31(3), 189-199. Hruby, V. J. (2002). Designing peptide receptor
agonists and antagonists. Nature Reviews: Drug Discovery, 1(11), 847-858.
Hubatsch, I., Ragnarsson, E. G. E., & Artursson, P. (2007). Determination of drug permeability and
prediction of drug absorption in Caco-2 monolayers. Nature Protocols, 2(9), 2111-2119.
Humphrey, M. J., & Ringrose, P. S. (1986). Peptides and Related Drugs - a Review of Their Absorption,
Metabolism, and Excretion. Drug Metabolism Reviews, 17(3-4), 283-310.
Huston, J. P., Hasenoehrl, R. U., Boix, F., et al. (1993). Sequence-specific effects of neurokinin substance P on
memory, reinforcement, and brain dopamine activity. Psychopharmacology, 112(2-3), 147-162.
Hutchinson, E. G., & Thornton, J. M. (1994). A Revised Set of Potentials for Beta-Turn Formation in
Proteins. Protein Science, 3(12), 2207-2216.
Igwe, O. J., Kim, D. C., Seybold, V. S., et al. (1990). Specific binding of substance P aminoterminal
heptapeptide [SP(1-7)] to mouse brain and spinal cord membranes. Journal of Neuroscience,
10(11), 3653-3663.
Kerns, E. H., & Di, L. (2008). Drug-like Properties: Concepts, Structure Design and Methods. London,
UK: Elsevier Inc.
Kim, H. O., & Kahn, M. (2000). A merger of rational drug design and combinatorial chemistry:
Development and application of peptide secondary structure mimetics. Combinatorial Chemistry
and High Throughput Screening, 3(3), 167-183.
Kramer, M. S., Cutler, N., Feighner, J., et al. (1998). Distinct mechanism for antidepressant activity by
blockade of central substance P receptors. Science, 281(5383), 1640-1645.
Kreeger, J. S., & Larson, A. A. (1993). Substance P-(1-7), a substance P metabolite, inhibits withdrawal
jumping in morphine-dependent mice. European Journal of Pharmacology, 238(1), 111-115.
Kruszynski, R., Fichna, J., do-Rego, J.-C., et al. (2005). Synthesis and biological activity of N-methylated
analogs of endomorphin-2. Bioorganic &Medicinal chemistry, 13(24), 6713-6717.
Lewis, P. N., Momany, F. A., & Scheraga, H. A. (1973). Chain Reversals in Proteins. Biochimica Et
Biophysica Acta, 303(2), 211-229.
Matthews, B. W. (1972). Gamma-Turn - Evidence for a New Folded Conformation in Proteins.
Macromolecules, 5(6), 818-819.
Milnerwhite, E. J., Ross, B. M., Ismail, R., et al. (1988). One Type of Gamma-Turn, Rather Than the
Other Gives Rise to Chain-Reversal in Proteins. Journal of Molecular Biology, 204(3), 777-782.
Motta, A., Reches, M., Pappalardo, L., et al. (2005). The preferred conformation of the tripeptide
Ala-Phe-Ala in water is an inverse gamma-turn: Implications for protein folding and drug design.
Biochemistry, 44(43), 14170-14178.
Nemethy, G., & Printz, M. P. (1972). Gamma-Turn, a Possible Folded Conformation of Polypeptide Chain
- Comparison with Beta-Turn. Macromolecules, 5(6), 755-
References 273
Nyberg, F., Hallberg, A., Hallberg, M., et al. (2010). Application: WO Patent No. 2009-
IB530202010004535.
Obach, R. S. (1999). Prediction of human clearance of twenty-nine drugs from hepatic microsomal
intrinsic clearance data: An examination of in vitro half-life approach and nonspecific binding to
microsomes. Drug Metabolism and Disposition, 27(11), 1350-1359.
Ohsawa, M., Carlsson, A., Asato, M., K et al. (2011). The effect of substance P1-7 amide on nociceptive
threshold in diabetic mice. Peptides, 32(1), 93-98.
Ohsawa, M., Carlsson, A., Koizumi, T., et al. (2010). Changes of Opioidergic Systems in Diabetic Painful
Neuropathy. Paper presented at the INRC 2010, Malmö, Sweden.
Okada, Y., Fujita, Y., Motoyama, T., et al. (2003). Structural studies of [2’,6’-dimethyl-L-tyrosine1]
endomorphin-2 analogues: enhanced activity and cis orientation of the Dmt-Pro amide bond.
Bioorganic & medicinal chemistry, 11(9), 1983-1994.
Olson, G. L., Bolin, D. R., Bonner, M. P., et al. (1993). Concepts and progress in the development of
peptide mimetics. Journal of Medicinal Chemistry, 36(21), 3039-3049.
Otsuka, M., & Yoshioka, K. (1993). Neurotransmitter functions of mammalian tachykinins. Physiological
Reviews, 73(2), 229-308.
Piercey, M. F., Dobry, P. J. K., Einspahr, F. J., et al. (1982). Use of Substance-P Fragments to Differentiate
Substance-P Receptors of Different Tissues. Regulatory Peptides, 3(5-6), 337-349.
Richardson, J. S. (1981). The anatomy and taxonomy of protein structure. Advances in Protein
Chemistry, 34, 167-339.
Ripka, A. S., & Rich, D. H. (1998). Peptidomimetic design. Current Opinion in Chemical Biology, 2(4),
441-452.
Rizo, J., & Gierasch, L. M. (1992). Constrained peptides: models of bioactive peptides and protein
substructures. Annual Review of Biochemistry, 61, 387-418.
Rose, G. D., Gierasch, L. M., & Smith, J. A. (1985). Turns in Peptides and Proteins. Advances in Protein
Chemistry, 37, 1-109.. (2006). Design, synthesis, and incorporation of a beta-turn mimetic in
angiotensin II forming novel pseudopeptides with affinity for AT(1) and AT(2) receptors. Journal of
Medicinal Chemistry, 49(20), 6133-6137.
Rosenström, U., Sköld, C., Lindeberg, G., et al. (2006). Design, synthesis, and incorporation of a
beta-turn mimetic in angiotensin II forming novel pseudopeptides with affinity for AT(1) and AT(2)
receptors. Journal of Medicinal Chemistry, 49(20), 6133-6137.
Rubio-Aliaga, I., & Daniel, H. (2002). Mammalian peptide transporters as targets for drug delivery.
Trends in Pharmacological Sciences, 23(9), 434-440.
Sakurada, C., Watanabe, C., & Sakurada, T. (2004). Occurrence of substance P(1-7) in the metabolism
of substance P and its antinociceptive activity at the mouse spinal cord level. Methods and
Findings in Experimental and Clinical Pharmacology, 26(3), 171-176.
Schmidt, B., & Kuhn, C. (1998). Racemic, yet diastereomerically pure azido acids as both gamma-turn
and inverse gamma-turn mimetics for solid-phase peptide synthesis. Synlett, (11), 1240-1242.
Sewald, N., & Jakubke. H.-D. (2002). Peptides: Chemistry and Biology: Wiley-VCH Verlag GmbH:
Weinheim.
Skilling, S. R., Smullin, D. H., & Larson, A. A. (1990). Differential-Effects of C-Terminal and N-Terminal
Substance-P Metabolites on the Release of Amino-Acid Neurotransmitters from the Spinal-Cord -
Potential Role in Nociception. Journal of Neuroscience, 10(4), 1309-1318.
Stewart, B. H., Chan, O. H., Lu, R. H., et al. (1995). Comparison of Intestinal Permeabilities Determined
in Multiple in-Vitro and in-Situ Models - Relationship to Absorption in Humans. Pharmaceutical
Research, 12(5), 693-699.
Wacher, V. J., Silverman, J. A., Zhang, Y. C., et al. (1998). Role of P-glycoprotein and cytochrome
P450 3A in limiting oral absorption of peptides and peptidomimetics. Journal of Pharmaceutical
Sciences, 87(11), 1322-1330.
274 Strategies for Conversion of Peptides to Peptidomimetic Drugs
Vagner, J., Qu, H. C., & Hruby, V. J. (2008). Peptidomimetics, a synthetic tool of drug discovery. Current
Opinion in Chemical Biology, 12(3), 292-296. Van de Waterbeemd, H., Camenisch, G., Folkers, G.,
et al. (1998). Estimation of blood-brain barrier crossing of drugs using molecular size and shape,
and H-bonding descriptors. Journal of Drug Targeting, 6(2), 151-165.
van de Waterbeemd, H., Smith, D. A., Beaumont, K., et al. (2001). Property-based design: optimization
of drug absorption and pharmacokinetics. Journal of Medicinal Chemistry, 44(9), 1313-1333.
Wang, B., Nimkar, K., Wang, W., et al. (1999). Synthesis and evaluation of the physicochemical
properties of esterase-sensitive cyclic prodrugs of opioid peptides using coumarinic acid and
phenylpropionic acid linkers. Journal of Peptide Research, 53(4), 370-382.
Veber, D. F., & Freidinger, R. M. (1985). The Design of Metabolically-Stable Peptide Analogs. Trends in
Neurosciences, 8(9), 392-396.
Venkatachalam, C. M. (1968). Stereochemical Criteria for Polypeptides and Proteins .V. Conformation
of a System of 3 Linked Peptide Units. Biopolymers, 6(10), 1425-1436.
Whitby, L. R., Ando, Y., Setola, V., et al. (2011). Design, Synthesis, and Validation of a beta-Turn Mimetic
Library Targeting Protein-Protein and Peptide-Receptor Interactions. Journal of the American
Chemical Society, 133(26), 10184-10194.
Wiktelius, D., Khalil, Z., & Nyberg, F. (2006). Modulation of peripheral inflammation by the substance
P N-terminal metabolite substance P1-7. Peptides, 27(6), 1490-1497.
Wiley, R. A., & Rich, D. H. (1993). Peptidomimetics Derived from Natural-Products. Medicinal Research
Reviews, 13(3), 327-384.
Wilmot, C. M., & Thornton, J. M. (1988). Analysis and Prediction of the Different Types of Beta-Turn
in Proteins. Journal of Molecular Biology, 203(1), 221-232. Witt, K. A., Gillespie, T. J., Huber, J.
D., et al. (2001). Peptide drug modifications to enhance bioavailability and blood-brain barrier
permeability. Peptides, 22(12), 2329-2343.
Vlieghe, P., Lisowski, V., Martinez, J., et al. (2010). Synthetic therapeutic peptides: science and market.
Drug Discovery Today, 15(1/2), 40-56.
Von Euler, U. S., & Gaddum, J. H. (1931). An unidentified depressor substance in certain tissue extracts.
Journal of Physiology, 72(1), 74-87.
Zhou, Q., Carlsson, A., Botros, M., et al. (2009). The C-terminal amidated analogue of the Substance
P (SP) fragment SP1-7 attenuates the expression of naloxone-precipitated withdrawal in morphine
dependent rats. Peptides, 30, 2418-2422.
Zhou, Q., Frändberg, P.-A., Kindlundh, A. M. S., et al. (2003). Substance P(1-7) affects the expression
of dopamine D2 receptor mRNA in male rat brain during morphine withdrawal. Peptides, 24(1),
147-153.
Zubrzycka, M., & Janecka, A. (2000). Substance P: transmitter of nociception (Minireview). Endocrine
Regulations, 34(4), 195-201.
Brandon S. Pybus*10
3.4 Synthetic Vs. Natural Bioactive Compounds
Against Tropical Disease
3.4.1 Introduction
“The angel of disease and death, ascending from his oozy bed, along the marshy
margins of the bottom grounds […] floats in his aerial chariot, and in seasons favorable
to his progress, spreads mortal desolation as he flies”, so it was written in an Ohio
newspaper article from 1820 (Findley, 1968). From this quote one can see how parasitic
disease plagued the early frontiersmen of the 19th century. However, the unhappy
coexistence of humans and parasitic disease goes back much further into the past.
In fact, the earliest recorded descriptions of a disease with symptoms consistent
with those of malaria date back to 2700 BC from imperial China. During this period
traditional Chinese medical practitioners discovered the use of sweet wormwood
(Artemesia annua) to treat this mysterious fever. Meanwhile, on the other side of the
planet Quechuas from South America used the bark of the cinchona tree in the treatment
of similar fevers. Jesuit monks brought this bark from the new world and introduced
Brandon S. Pybus: Division of Experimental Therapeutics, Military Malaria Research Program, Walter
Reed Army Institute of Research, 503 Robert Grant Ave, Silver Spring, MD 20910, USA,
*Email: brandon.s.pybus.mil@mail.mil
its powerful medicinal properties to Europe during the 17th century, however, the active
component, quinine, was not isolated until 1820 (Faurant, 2011). Quinine is known
for several unpleasant and potentially severe adverse reactions including cinchonism,
hearing impairment, increased risk of hemolysis in G6PD deficient patients, and is
linked to a severe syndrome in some patients, dubbed blackwater fever. Modern efforts
suggest the latter may be the result of redox active metabolites (Marcsisin, 2013). As we
will explore, many of the ongoing efforts in anti-malarial therapy revolve around not
only circumventing resistance, but in mitigating many of the safety risks associated
with historically used drugs. These reasons and the outbreak of war in malaria endemic
regions of the world led to a surge of interest in the development of safe and effective
drugs to treat malaria that we will explore further.
“Doctor […] this shall be a long war if for every division I have facing the enemy I
must count on a second division in the hospital with malaria and a third division
convalescing from this debilitating disease!” Thus opined Gen. Douglas MacArthur
concerning the ravages of malaria during the second world war (Coates, 1963). This
constant struggle with non-combat related injury from disease was a tremendous
burden on medical logistical chains, morale, and overall readiness on fronts across
Africa, Asia, and even in parts of Europe. This reality was a major shaping force for a
massive effort to prevent and treat malaria.
During WWI, Germany had no access to quinine. As a result, a large-scale effort
ensued in the intervening years between WWI and WWII in which the Germans
synthesized and screened thousands of compounds for anti-malarial activity.
Among these were several 8-aminoquinolines, atabrine (quinacrine), and a drug
they discarded for toxicity called Resochin (chloroquine) (Coates, 1963; Vale, 2009).
After access to quinine from Indonesia was cut to Allied forces by the entry of Japan
into WWII, US and British chemists began a similar push resulting in the adoption of
atabrine as the drug of choice for malaria treatment and prophylaxis during the later
years of the war. Post-war, powerful new anti-malarial drugs began to emerge as the
culmination of follow-on efforts by US and British chemists to exploit the original
efforts by Germany. Many of these drugs are still in use today, with robust analog
efforts being employed around them to overcome developing resistance.
For natural products, quinine is a success story that has endured for centuries.
Quinine is an aryl amino alchohol derived from the bark of the cinchona tree (Fig.
3.4.1). Despite its discovery over 400 years ago, it remains an enormously important
drug in the treatment of malaria in the developing world. Quinine is rapidly absorbed,
both orally and parenterally, broadly distributed throughout the body, and has proven
highly effective for the treatment of uncomplicated or severe malaria (Achan, 2011).
In the age of resistance, quinine has made a resurgence in usage. Although quinine
resistance has been reported, it is generally low-grade, and is largely found in Asia
and South America (Noedl, 2006; Parola, 2001).
Perhaps a larger issue still in play for quinine is tolerance. Quinine has several
notable and potential severe adverse events associated with usage. Most common side-
effects include tinnitus and hearing impairment, but more severe effects can include
vertigo, vomiting, abdominal pain, or hypotension (Achan, 2011). The most severe
and potentially least understood is a syndrome called blackwater fever characterized
by hemolysis and hemoglobinuria (George, 2009). Recent studies suggest a potential
link between CYP mediated oxidative metabolites of quinine and this potentially fatal
reaction (Marcsisin, 2013). It should also be noted that while clinically different from
the hemolytic events observed with the 8-aminoquinolines (which we will discuss
in a later section), as with the 8-aminoquinolines a link may exist between glucose-
6-phosphate (G6PD) deficiency and hemolysis (Hue, 2009). This is significant as it
suggests common metabolic pathways for some quinoline drugs that should be avoided
or at least considered when developing new drugs in these classes. For quinine, the
formation of redox active quinones may lead to increased oxidative stress under
conditions of high parasitemia or in G6PD deficient individuals that ultimately results
in the destruction of red-cells through mechanisms which are not fully understood
(Fig. 3.4.2) (Marcsisin, 2013).
Arguably one of the most successful classes of drugs to arise from WWII efforts is
the 4-aminoquinolines. Fig. 3.4.3 illustrates two members of the class, chloroquine and
amodiaquine, and can serve as a generic paradigm for its structure. Chloroquine was
278 Synthetic Vs. Natural Bioactive Compounds Against Tropical Disease
originally discovered by Hans Andersag in 1934 while working for Bayer, however, the
drug was discarded for perceived toxicity. After its re-discovery by British and American
chemists, it quickly won favor as a safe and effective anti-malarial (CDC, August 18,
2014). In fact, chloroquine was used extensively in post-war eradication efforts. While
the mechanism of action for all 4-aminoquinolines is not exactly known, they are
known to be effective in treating only erythrocytic stages of infection. Poor compliance
or poor management of care with such drugs can rapidly lead to resistance. This was
exactly the case with chloroquine. By the early 1950’s, chloroquine resistance was
identified along the Thai-Cambodian border and Colombia. Within two decades it had
spread to every malaria endemic region of the world (Farooq & Mahajan, 2004).
While the mechanism of action for the 4-aminoquinolines is not completely
understood, the most widely cited hypothesis is that accumulation in the digestive
vacuole of the parasite interferes with hemoglobin digestion. Resistance is believed
to be conferred by the presence of a chloroquine specific P-glycoprotein pump (Foley
& Tilley, 1998). Several loci in the P. falciparum genome have been implicated in this
resistance (Farooq & Mahajan, 2004). This mechanism, or a variant thereof may well
be important in quinine, mefloquine, or other quinoline resistance mechanisms,
however, it should be noted that more lipophilic quinolines do not concentrate in the
food vacuole to the same extent as chloroquine, and therefore other mechanisms need
to be explored (including potential alternative targets of efficacy) when considering
these drugs.
A considerable effort is still ongoing to circumvent resistance through analog
campaigns or co-administration with other drugs (Hanboonkunupakarn, 2014; Le
Garlantezec, 2014; I. Opsenica, 2011; I. M. Opsenica, 2013; Thriemer, 2014). While
it is tempting to synthesize analogs of a compound with emerging or established
resistance, in the author’s opinion, such efforts should be entered into with full
knowledge that selection pressures in malaria endemic areas quickly erode the
efficacy of new drugs from old classes. It is advisable to pursue a combination
approach wherein fast acting short half-life drugs (e.g. artemesinins) are combined
with long half-life (longer exposure of parasite to drug) quinolines prone to emergence
of resistance. Further, future efforts in this class should pay careful attention to early
signs of cross resistance in chloroquine resistant strains. While small jumps in IC50
in a resistant strain may not raise alarms if those IC50s are still significantly less than
the anticipated maximum concentrations achieved in plasma post-exposure, they
could be indicative of shared mechanisms of resistance that might worsen over time
with large exposures in environments with poorly controlled administration or where
monotherapy is used.
Mefloquine, a quinoline methanol, was developed by the US Army in the 1970s,
and was considered very desirable for both treatment and prophylaxis due to its
long half-life (Croft, 2007). In recent years mefloquine usage has decreased due to
emerging resistance, and more significantly, poor tolerance (Farooq & Mahajan,
2004; Milner, 2010). Mefloquine has been linked to neurological side-effects including
Modern Efforts in Antimalarial Drug Development 279
vertigo, loss of balance, and polyneuropathy, which have recently been labeled as
potentially irreversible by the FDA (Nevin, 2014). Extensive efforts have shown
promise in improving the therapeutic index of mefloquine by opening or removing
the piperadine side-chain (Dow, 2006; 2011; Milner, 2011). With the notoriously poor
predictive power of models of CNS toxicity and the stigma attached to the quinoline
methanol class, these efforts have largely been abandoned. Further, recent trials with
enantiomerically pure mefloquine (currently marketed as a racemic mixture) have
shown little promise in improving tolerability (Nevin, 2014).
3.4.4.2 8-Aminoquinolines
Another powerful class of anti-malarial drugs to emerge from war efforts is the
8-aminoquinolines. Although still quinoline based, this class has several key features
which make it unique. One primary feature of this class is its ability to act as a causal
prophylactic, or to prevent the initial infection of parasites in the liver. Further, the
class has powerful anti-hypnozoite activity. Hypnozoites being the dormant liver stage
of P. vivax and P. ovale, this imparts a separate prophylactic use, namely presumptive
anti-relapse therapy or PART. Combined with this unique exoerythrocytic activity
is the gametocytocidal activity, whereby 8-aminoquinolines kill the sexual stages
of the parasite blocking further transmission of infection. This combination of
activities makes this class highly attractive for prophylaxis, treatment of relapsing
strains of malaria, and/or elimination efforts. The class has also shown in vitro anti-
leishmanicidal activity and has clinical utility in the treatment of pneumocystic
pneumonia, and trypanosomiasis (Vale, 2009). However, as with other classes that
we have reviewed here tolerability is a substantial problem for the 8-aminoquinolines.
From this class, currently only primaquine is clinically available for these uses.
Administration of 8-aminoquinolines is known to cause methemoglobinemia
and hemolysis in G6PD deficient individuals. Both of these effects are linked to
oxidative metabolites rather than the parent drug itself (Ganesan, 2009; Ganesan,
2012). Interestingly, these same metabolites were recently shown to play a crucial
role in activity (Bennett, 2013; Marcsisin, 2014; Pybus, 2013). For both primaquine
and tafenoquine (8-aminoquinoline, currently in clinical development), activity is
mediated by CYP 2D6 (Fig. 3.4.2). The mechanism of activity for the 8-aminoquinolines
is not fully understood, however, contemporary thought is that redox active metabolites
produce peroxides and superoxides (oxidative stress), which ultimately interferes
with electron transport in the parasite. Not surprisingly, what is also unknown at
present is whether the toxicity can be mitigated while preserving efficacy. Work in
this area is of paramount importance, as new 8-aminoquinolines with similar profiles
to primaquine are of little clinical utility.
The WHO currently calls for a 30 mg dose of primaquine as a chemoprophylaxis
in areas where P. vivax is the predominant risk, however significant fractions of the
280 Synthetic Vs. Natural Bioactive Compounds Against Tropical Disease
population (10% in Europeans and as much as 50% in some Asians) possess 2D6 alleles
with reduced function (Deye & Magill, 2014). It is therefore likely that primaquine
prophylaxis and even anti-relapse therapy will fail in many of these individuals.
All things considered, the importance of this class cannot be overemphasized and
it cannot be abandoned without more study. As stated previously, primaquine is
currently the only clinically available drug for the treatment of relapsing malaria.
As such, the challenges posed for future 8-aminoquinoline development include
re-routing metabolism through a non-CYP 2D6 pathway, circumventing hemolytic
toxicity, or both. The primaquine enamine bulaquine showed some promise in recent
animal studies at improving the therapeutic index (Lal, 2003; Mehrotra, 2007). This
appears to be due to differential pharmacokinetic exposure patterns as compared
to parent primaquine. The addition of the enamine group on the side-chain creates
an almost time-release effect as it is hydrolyzed to release primaquine. This leads
to an interesting hypothesis that perhaps while the same metabolites may well be
responsible for efficacy and toxicity, efficacy could be exposure driven while toxicity
is concentration dependent, providing an in-road to improvements in therapeutic
index. Although in the author’s opinion, circumstantial evidence overwhelmingly
points to an inextricable link between efficacy and toxicity in this class that may
make it difficult, if not impossible, to support the large-scale development of a new
candidate.
Although sweet wormwood was used for over 2000 years in China to treat malaria, the
active ingredient, artemisinin, was not identified until the 1970’s (Fig. 3.4.5) (Faurant,
2011). At present, several artemisinin compounds are clinically available for use
including artemisinin, artesunate, dihydroartemisinin (DHA), and artemether. Other
experimental drugs exist but are not clinically available. Due to poor tolerability
and hence compliance with quinine, artemisinin compounds have drastically risen
in popularity for the treatment of falciparum malaria (Shanks, 2006). Artemisinin
resistance was first reported in Southeast Asia in 2008, and evidence suggests that
it is spreading (Ashley, 2014; Dondorp, 2009; Thriemer, 2014). Widespread resistance
to the most powerful new weapon in the fight against malaria could de-rail many
ongoing efforts to eradicate the disease in the endemic world. As such, it is widely
acknowledged that combination therapy should be the standard of care. The WHO
currently recommends artemisinin combination therapy (ACT) consisting of DHA-
piperaquine, followed with a 0.75 mg kg-1 single dose of primaquine for the treatment
of uncomplicated malaria (WHO, 2010).
At present, research is still ongoing as to the exact killing mechanism for the
functional endoperoxide bridge contained in the artemisinins, however, it is thought
to be a result of membrane depolarization and subsequent interference with electron
Natural Products in the Treatment of Malaria 281
For the treatment of malaria and many other tropical diseases, limitations in funding
have led to discoveries in the area of repurposing of drugs commonly used for
other indications. Broad spectrum antibiotics have proven highly successful in the
treatment of malaria. Doxycycline, a synthetic tetracycline, was shown to have partial
prophylactic efficacy against malaria in the 1970’s (Andrews, 2014). In fact, despite
its own set of tolerance issues, doxycycline is now the prophylactic drug of choice
for the US Army due to more serious concerns with mefloquine (Kime, 2012). It is also
effective for use in treatment when partnered with a fast acting anti-malarial like
quinine or quinidine (Tan & Centers for Disease Control and Prevention, 2011). Other
antibiotics including clindamycin and sulfonamide antibiotics like sulphadiazine and
sulphadoxine have also proven effective in the treatment of malaria (Andrews, 2014).
Many of these target folate synthesis, which is a common pathway for other classical
anti-malarials. Antifolate drugs block the synthesis of tetrahydrofolate, which in
turn shuts down nucleic acid synthesis (Shanks, 2006). Among these is dapsone.
Dapsone, a sulfone, was first used to treat leprosy but was introduced by GSK as
Lapdap (dapsone/chloroproguanil) in the late 1990s. This drug was removed from
the market in 2008 due to hemolytic anemia similar to that of the 8-aminoquinolines.
Exploration of this strategy still continues with significant effort. Several examples can
be found in the literature with protease inhibitors, antibiotics, and quinolones found
to act against various targets in the parasite (Rosenthal, 2003). With pre-existing and
well established safety profiles, re-purposing of old drugs is an attractive strategy if
for no other reason than bypassing several regulatory hurdles and saving money in
development. Another potential benefit for the developing world is often the cost of
these drugs once marketed. Doxycycline, for example, is pennies per dose compared
to Malarone (atovaquone/proguanil) and inexpensive drugs are far more likely in
that regard to make a meaningful impact on the fight against parasitic disease world-
wide.
2014). Natural sources including plants, microbes, and animals provide libraries with
rich chemical diversity, however, isolation and purification of active components
presents a significant challenge in natural product development. The complexity of
isolation is highly dependent on structure, which can often be quite complex with
compounds isolated from natural sources. This of course leads to a second challenge
for the development of natural products, namely synthesis. Once a hit is identified
from a mixture, large quantities are often needed for further pre-clinical or clinical
testing. The likelihood of identifying a clinic-ready compound from a natural source
in today’s regulatory environment is small. The more likely scenario involves the
identification of a hit compound, which can be potentially modified to improve drug-
like characteristics or decrease potential liabilities from toxicity. As we will discuss
in the next section, the source of starting material for a drug development project is
largely irrelevant once a hit is identified, as all compounds should be gated through the
same testing scheme to ensure the end product matches the target product profile.
In any drug candidate’s journey down a development pipeline there are many
pitfalls. Sadly, activity against a biological target or even in vivo efficacy does not
make a successful drug. Careful consideration should be given early to liabilities
that might lead to late stage failure, which could be costly and drain resources vital
to the discovery of a potential candidate. In 1991, the leading cause of attrition for
drug candidates was poor pharmacokinetics and/or bioavailability (40%), yet by
2000 this had decreased to less than 10% (Kola & Landis, 2004). This is largely
attributable to the advent of in vitro screens for physiochemical properties likely to
create downstream liabilities. The placement of these screens early in the pipeline
forces compounds with poor solubility, bioavailability, and metabolic stability to fail
during far cheaper stages of development and spares the opportunity cost otherwise
missed in development of these compounds. Fig. 3.4.6 outlines a generic paradigm
for screening compounds. In this paradigm early hits from in vitro activity screens are
passed through an absorption, distribution, metabolism, and excretion (ADME) gate
before passing into more expensive animal models. The source of compounds could
be from commercially available large libraries, libraries of natural product extracts,
small scale synthesis, etc. The process remains unchanged. However, feedback
from all stages should be iteratively incorporated into the next round of screening.
Acceptable physiochemical properties for a drug ultimately depend on the target
product profile for intended use; however, it is reasonable to assume that compounds
with low solubility or poor permeability may not fit a profile for oral prophylaxis, for
example. It should also be noted that many of these properties go hand-in-hand, and
that it is often necessary to strike a balance. For example, lipophilicity, permeability,
and metabolic instability tend to correlate since CYPs tend to favor lipophilic drugs.
Considerations for Anti-parasitic Drug Development 283
From this example, if one were trying to develop a long half-life lipophilic drug to
concentrate in the food vacuole of a parasite, it might be necessary to trade off some
lipophilicity (and with it potentially some biological activity) to enhance half-life. The
subtleties of this interplay of course depend again on the intended use of the drug.
Feedback from all stages of screening should be used to inform later rounds. In such
a manner, successive rounds of screening should get closer and closer to the desired
chemical space with optimum physiochemical properties and biological activity for
intended use.
From the example of the 8-aminoquinolines, a very modern consideration for drug
development arises. Pro-drugs should be fully metabolically characterized, with all
pertinent pathways identified. Consideration should be given to the pharmacogenomic
make-up of the target population. Pro-drugs activated by CYPs 3A4, 2D6, or 2C19 may
lack efficacy in some populations due to high polymorphism. If at all possible, non-
CYP mediated conversion is desirable to avoid such liabilities in later development.
Likewise, toxicity can be affected in a similar manner if, for example, the formation of
a reactive metabolite is mediated by a highly polymorphic CYP.
Table 3.4.1: Approaches to antimalarial drug discovery and development. Adapted from Philip J.
Rosenthal’s review on Antimalarial drug discovery: old and new approaches (Rosenthal, 2003).
Approach Examples
Optimize therapy with existing drugs Amodiaquine/sulfadoxine/pyrimethamine
(combinations) Amodiaquine/artesunate
Artesunate/sulfadoxine/pyrimethamine
Artesunate/mefloquine
Artemether/lumefantrine
Chlorproguanil/dapsone
Chlorproguanil/dapsone/artesunate
Atovaquone/proguanil
Develop analogs of existing drugs New aminoquinolines
New endoperoxides
New folate antagonists
Natural Products New natural products
Repurposing of drugs from other Folate antagonists
therapeutic areas Antibiotics
Atovaquone
Iron chelators
Reversing drug resistance Verapamil, desipramine, trifluoperazine, chlorpheniramine
Discovery of compounds active Antibiotics, Quinolones, etc.
against novel targets
284 Synthetic Vs. Natural Bioactive Compounds Against Tropical Disease
Figure 3.4.2: Metabolism of quinine leads to the formation of redox active quinones. Adapted from
Marcsisin 2013 (Marcsisin, 2013).
Figure 3.4.3: Two important members of the 4-aminoquinoline class: A. Amodiaquine B. Chloroquine
Considerations for Anti-parasitic Drug Development 285
Figure 3.4.4: CYP 2D6 mediated metabolism of primaquine (Pybus, 2013). Several potentially oxi-
dative metabolites formed by CYP 2D6 may be key in primaquine activity. Carboxyprimaquine is the
primary metabolite found is plasma. Its production is likely MAO mediated and the pathway is not
shown.
286 Synthetic Vs. Natural Bioactive Compounds Against Tropical Disease
Figure 3.4.5: Artemisinin was identified as the active ingredient in sweet wormwood in the 1970s.
Several synthetic artemisinins and other endoperoxides now exist for the treatment of malaria.
Figure 3.4.6: Generic paradigm for drug discovery that uses ADME properties to down-select poten-
tial candidates prior to expensive and labor intensive animal modeling. Dashed lines indicate a
screen that is not necessarily a mandatory gate, but which can aid in the interpretation of potential
later stage failure.
References
Achan, J., Talisuna, A. O., Erhart, A., Yeka, A., Tibenderana, J. K., Baliraine, F. N., (2011). Quinine, an
old anti-malarial drug in a modern world: Role in the treatment of malaria. Malaria Journal, 10,
144-2875-10-144.
Andrews, K. T., Fisher, G., & Skinner-Adams, T. S. (2014). Drug repurposing and human parasitic
protozoan diseases. International Journal for Parasitology. Drugs and Drug Resistance, 4(2),
95-111.
Antoine, T., Fisher, N., Amewu, R., O’Neill, P. M., Ward, S. A., & Biagini, G. A. (2014). Rapid kill of
malaria parasites by artemisinin and semi-synthetic endoperoxides involves ROS-dependent
References 287
Kime, P. (2012, April 11). New concerns rising over antimalaria drug. Army Times,
Kola, I., & Landis, J. (2004). Can the pharmaceutical industry reduce attrition rates? Nature Reviews.
Drug Discovery, 3(8), 711-715.
Lal, J., Mehrotra, N., & Gupta, R. C. (2003). Analysis and pharmacokinetics of bulaquine and its
major metabolite primaquine in rabbits using an LC-UV method--a pilot study. Journal of
Pharmaceutical and Biomedical Analysis, 32(1), 141-150.
Lanteri, C. A., Chaorattanakawee, S., Lon, C., Saunders, D. L., Rutvisuttinunt, W., Yingyuen, K.,
(2014). Ex vivo activity of endoperoxide antimalarials, including artemisone and arterolane,
against multidrug-resistant plasmodium falciparum isolates from cambodia. Antimicrobial
Agents and Chemotherapy, 58(10), 5831-5840.
Le Garlantezec, P., Richard, C., Broto, H., & Rapp, C. (2014). Treatment of imported plasmodium
falciparum malaria: Role of the combination of dihydroartemisinin and piperaquine.
[Traitement du paludisme d’importation a Plasmodium falciparum : place de l’association
dihydroartemisinine-piperaquine] Medecine Et Sante Tropicales,
Marcsisin, S. R., Jin, X., Bettger, T., McCulley, N., Sousa, J. C., Shanks, G. D., (2013). CYP450
phenotyping and metabolite identification of quinine by accurate mass UPLC-MS analysis: A
possible metabolic link to blackwater fever. Malaria Journal, 12, 214-2875-12-214.
Marcsisin, S. R., Sousa, J. C., Reichard, G. A., Caridha, D., Zeng, Q., Roncal, N., (2014). Tafenoquine
and NPC-1161B require CYP 2D metabolism for anti-malarial activity: Implications for the
8-aminoquinoline class of anti-malarial compounds. Malaria Journal, 13, 2-2875-13-2.
Mehrotra, N., Lal, J., Puri, S. K., Madhusudanan, K. P., & Gupta, R. C. (2007). In vitro and in
vivo pharmacokinetic studies of bulaquine (analogue of primaquine), a novel antirelapse
antimalarial, in rat, rabbit and monkey--highlighting species similarities and differences.
Biopharmaceutics & Drug Disposition, 28(5), 209-227.
Milner, E., McCalmont, W., Bhonsle, J., Caridha, D., Cobar, J., Gardner, S., (2010). Anti-malarial
activity of a non-piperidine library of next-generation quinoline methanols. Malaria Journal, 9,
51-2875-9-51.
Milner, E., Sousa, J., Pybus, B., Melendez, V., Gardner, S., Grauer, K., (2011). Characterization of in
vivo metabolites of WR319691, a novel compound with activity against plasmodium falciparum.
European Journal of Drug Metabolism and Pharmacokinetics, 36(3), 151-158.
Mohammed, T., Erko, B., & Giday, M. (2014). Evaluation of antimalarial activity of leaves of
acokanthera schimperi and croton macrostachyus against plasmodium berghei in swiss albino
mice. BMC Complementary and Alternative Medicine, 14, 314-6882-14-314.
Nevin, R. L. (2014). Idiosyncratic quinoline central nervous system toxicity: Historical insights into
the chronic neurological sequelae of mefloquine. International Journal for Parasitology.Drugs
and Drug Resistance, 4(2), 118-125.
Noedl, H., Krudsood, S., Chalermratana, K., Silachamroon, U., Leowattana, W., Tangpukdee, N.,
(2006). Azithromycin combination therapy with artesunate or quinine for the treatment of
uncomplicated plasmodium falciparum malaria in adults: A randomized, phase 2 clinical trial in
thailand. Clinical Infectious Diseases : An Official Publication of the Infectious Diseases Society
of America, 43(10), 1264-1271.
Oliveira, R., Guedes, R. C., Meireles, P., Albuquerque, I. S., Goncalves, L. M., Pires, E., (2014).
Tetraoxane-pyrimidine nitrile hybrids as dual stage antimalarials. Journal of Medicinal
Chemistry, 57(11), 4916-4923.
Opsenica, I., Burnett, J. C., Gussio, R., Opsenica, D., Todorovic, N., Lanteri, C. A., (2011). A chemotype
that inhibits three unrelated pathogenic targets: The botulinum neurotoxin serotype A light
chain, P. falciparum malaria, and the ebola filovirus. Journal of Medicinal Chemistry, 54(5), 1157-
1169.
Opsenica, I. M., Tot, M., Gomba, L., Nuss, J. E., Sciotti, R. J., Bavari, S., (2013). 4-amino-7-
chloroquinolines: Probing ligand efficiency provides botulinum neurotoxin serotype A light
References 289
chain inhibitors with significant antiprotozoal activity. Journal of Medicinal Chemistry, 56(14),
5860-5871.
Parola, P., Ranque, S., Badiaga, S., Niang, M., Blin, O., Charbit, J. J., (2001). Controlled trial of
3-day quinine-clindamycin treatment versus 7-day quinine treatment for adult travelers with
uncomplicated falciparum malaria imported from the tropics. Antimicrobial Agents and
Chemotherapy, 45(3), 932-935.
Pybus, B. S., Marcsisin, S. R., Jin, X., Deye, G., Sousa, J. C., Li, Q., (2013). The metabolism of
primaquine to its active metabolite is dependent on CYP 2D6. Malaria Journal, 12(1), 212.
Rosenthal, P. J. (2003). Antimalarial drug discovery: Old and new approaches. The Journal of
Experimental Biology, 206(Pt 21), 3735-3744.
Shanks, G. D. (2006). Treatment of falciparum malaria in the age of drug resistance. Journal of
Postgraduate Medicine, 52(4), 277-280.
Singh, P., Anand, A., & Kumar, V. (2014). Recent developments in biological activities of chalcones:
A mini review. European Journal of Medicinal Chemistry, 85, 758-777.
Tan, K. R., Magill, A. J., Parise, M. E., Arguin, P. M., & Centers for Disease Control and Prevention.
(2011). Doxycycline for malaria chemoprophylaxis and treatment: Report from the CDC expert
meeting on malaria chemoprophylaxis. The American Journal of Tropical Medicine and Hygiene,
84(4), 517-531.
Thriemer, K., Van Hong, N., Rosanas-Urgell, A., Phuc, B. Q., Ha, D. M., Pockele, E., (2014). Delayed
parasite clearance after treatment with dihydroartemisinin-piperaquine in plasmodium
falciparum malaria patients in central vietnam. Antimicrobial Agents and Chemotherapy,
Traore, M. S., Diane, S., Diallo, M. S., Balde, E. S., Balde, M. A., Camara, A., (2014). In vitro
antiprotozoal and cytotoxic activity of ethnopharmacologically selected guinean plants. Planta
Medica, 80(15), 1340-1344.
Vale, N., Moreira, R., & Gomes, P. (2009). Primaquine revisited six decades after its discovery.
European Journal of Medicinal Chemistry, 44(3), 937-953.
WHO (Ed.). (2010). Guidelines for the treatment of malaria (2nd ed.). Geneva: World Health
Organization.
Mónica Ferreira, Patrick Almeida, Mohammad-Ali Shahbazi,
Alexandra Correia and Hélder A. Santos*11
3.5.1 Introduction
Cancer is one of the leading causes of death in the present days, killing millions of
people every year (Sutradhar, 2014). Cancer is a pathological problem derived from
genetic instability and multiple molecular alterations, caused by uncontrollable cell
division which invade the surrounding tissue and destroy it (Mody, 2011). More than
8.2 million people died from cancer in 2012, and around 60% of new cases occurred in
Africa and Asia, but 30% of cancers were prevented (WHO, 2014). In 2014, more than
1.6 million new cancer cases and more than half million cancer deaths are estimated to
occur in the United States (ACS, 2014). During the most recent 5 years of existing data
(2006−2010), delay-adjusted cancer incidence rates declined slightly in men (by 0.6%
per year) and were stable in women, while cancer death rates decreased by 1.8% per
year in men and by 1.4% per year in women. The combined cancer death rates (deaths
per 100,000 inhabitants) have been continuously declining for 2 decades, from 215.1
in 1991 to 171.8 in 2010. This 20% decline translates to the avoidance of more than 1.3
million cancer deaths (almost 1 million among men and more than 0.3 million among
women) during this time period (Siegel, 2014). In Europe, around one quarter of the
total population is affected by new cancer cases, with about 3.2 million new patients
every year (WHO, 2014). Breast, colorectal, lung, liver and stomach cancers are some
of the most common cases of cancers.
While there have been many advances in cancer treatment, such as chemotherapy
and radiotherapy for cancer, they are still far from being ideal. The problematic issue
is to achieve the desired concentration of therapeutic agents at the tumor site, thereby
destroying cancer cells, while minimizing the damage to normal cells. In order to
overcome some of the problems of conventional anticancer therapeutics, cancer
nanotechnology has been implemented, which indicates a major breakthrough in
cancer detection, diagnosis and treatment of cancer (Misra, 2010).
Nanotechnology is one of the fastest areas of growth and development in the 21st
century, where several material types (e.g., organic, inorganic, and polymeric based),
medicines and devices are used to manipulate matter with size in the range of 1−100
nm (Reddy, 2011) (Fig. 3.5.1). Some of the major challenges in nanotechnology are a
proper identification of neoplasic early biomarkers, as well as the understanding of
their evolution, screening and early detection, development of technology to provide
different selective therapeutic nanoplatforms and the ability to pass through biological
and physical barriers (Pinheiro AV1, 2011). Some nanosystems are developed to overcome
drug resistance, since they control the efflux of P-glycoprotein, the major mechanism
of drug resistance (Dong, 2010). Therefore, nanotechnology has brought new hope for
cancer detection due to the development of nanozised probes (Mody, 2011).
292 Current Trends and Developments for Nanotechnology in Cancer
Figure 3.5.1: A summary of nanoparticles that have been explored as carriers for drug delivery in
cancer therapy, together with illustrations of their biophysicochemical properties. Reprinted with
permission from (Sun, 2014).
study the biophysical interactions (Sumer, 2008). Drug development and imaging
nanomedicines are thus interconnected, due to the advantages of the application
and non-invasiveness in high yield imaging studies. All imaging modalities can be
used in drug studies to provide anatomical, pharmacokinetic and pharmacodynamic
information (Wang, 2010).
Table 3.5.1: Nanomedicines approved by one or more regulatory bodies. Reprinted from (Wang, 2013).
The current nanoplatforms used for the purpose of targeting drug delivery to cancer
tissues can also be addressed to selectively deliver imaging agents (Chen, 2014).
For example, the first evaluation of the effectiveness of a particular formulation to
a patient can be performed with imaging agents, in order to verify that the delivery
system will primarily target the cancerous tissues, even before any drug regimen test
is initiated (Toy, 2014). However, imaging is not only used for detection, it is also
important to determine the stage and the precise locations of cancer. Distribution of
multistage systems have shown potential to meet the challenges of drug targeting and
overcome biological barriers (Miele, 2012). Image-based nanomedicines have also a
crucial importance in cancer diagnostics due to their highly sensitive detection probes
(Pan, 2010). Imaging techniques, which are methods of producing images of the body,
are an important element of early detection of cancer diseases (Shukla-Dave, 2014).
Transformation procedures leading to malignancy can be detected by a matter of
routine screening, non-invasive means, such as standard proteomic analysis of blood
samples or in vivo imaging of molecular profiles (Bellisola, 2012). Some of the main
challenges for nanomedicines are the development for the detection and monitoring
of cancer markers in vivo, creation of technological platforms for early detection of
cancer biomarkers ex vivo, and engineering of nanoparticles to prevent biophysical
and biological barriers (Ferrari, 2005).
In this chapter we focus on the current trends and developments of nanotechnology
for cancer, highlighting some of the most advanced drug delivery nanosystems,
targeting strategies, nanotools used for cancer imaging and diagnostics, as well as
recent nanotechnological approaches used for cancer immunotherapy.
Cancer drug delivery is a demanding subject that requires deep and comprehensive
research. Thus, this area of research has been under intensive investigation during the
last few years. For example, problems associated with selective targeting, improved
penetration through biological barriers and a controlled release of therapeutic cargos
over time are details that influence the design of nanocarriers for drug delivery.
Therefore, recently, sophisticated nanosystems have been developed in order to
provide delivery of one or more therapeutic molecules in a controlled, selective and
“smart” way to the desired cancer cells/tissues.
Conventional systems for cancer therapy that are available in the market have limited
and unspecific access to tumor sites, and thus, it is well-known that patients experience
serious side effects when treated with anticancer drugs, frequently associated to low
Drug Delivery Nanosystems in Cancer Therapy 295
therapeutic efficacy (Siegel, 2011). Another factor that needs to be considered for
drug delivery is the physicochemical properties of the drugs themselves. Many of
the anticancer drugs used in the clinic at the present are rather hydrophobic, and
thus, their poor solubility may lead to local toxicity associated with the fact that they
may not be soluble enough to go through the aqueous environment surrounding the
tumor cells and cross the cell membrane to ultimately reach the intracellular targets
(Owen, 2012). On the other hand, successful utilization of hydrophilic drugs (e.g.,
macromolecules such as proteins or nucleic acids) has been stalled by a number of
obstacles, such as poor cell internalization, because of their inability to cross the lipid
bilayer of the cell membrane, as well as short half-life in the bloodstream, due to poor
stability against proteolytic and hydrolytic degradation (Fattal, 2009; Ishihara, 2010;
Sun, 2014).
Therefore, by applying nanotechnology, many scientists have focused on
investigating new ways to develop novel drug delivery systems with the aim to
maintain high therapeutic drug levels at the malignant cell sites and as low as
possible in healthy cells, hence overcoming and improving the poor physicochemical
properties of the drug (Grinberg, 2014; Sun, 2014). For this purpose, several strategies
to both target the tumor site and release the drug(s) in a controlled fashion have been
developed in order to selectively deliver therapeutic cargos over time. In this section,
we will briefly focus on the most recent controlled release strategies and give some
examples of the nanosystems used to achieve the abovementioned aims.
It is possible to fine-tune and control the release of payloads from the nanocarrier
through diverse mechanisms. For example, for porous materials, controlling the pore
size and the surface chemistry of the pore walls can lead to different diffusion release
profiles, either improving or sustaining the release of the loaded cargos. Porous hollow
Fe3O4 nanoparticles were able to provide the delivery of cisplatin via a slow diffusion-
controlled process, exhibiting different kinetic profiles of cisplatin depending of the
pore gap sizes (Cheng, 2009). Porous silicon (PSi) materials are another good example
of a biomaterial widely used for biomedical applications (Santos, 2014). The small
and tunable size of the pores, high surface area and large pore volume enables high-
drug-loading capacity. The confinement of large amounts of drug molecules, together
with their interaction with the nanocarriers, and depending on the physicochemical
properties of the PSi materials, leads to a tunable and controllable release of cargos
(Mäkilä, 2014; Salonen, 2008). Controlled release of drug molecules can be also
achieved through the uniform incorporation of the drug into matrix-based systems
(water-insoluble polymer carriers). These systems enable sustained drug release
because the diffusion of drug molecules from the inner matrix of the polymers to the
surface takes time and may depend on the physicochemical properties of the drug,
on the interactions between the drug and the polymeric matrix, and on the density
of the matrix itself (Sun, 2014). Ma have synthesized a PCL-Tween 80 copolymer
nanoparticulate matrix from ε-caprolactone and Tween 80 to maintain the release
of cargos through diffusion for as long as 28 days, with a better in vitro cytotoxicity
296 Current Trends and Developments for Nanotechnology in Cancer
towards C6 glioma cells compared to the commercial docetaxel (Taxotere) at the same
drug concentration (Ma, 2011). Sustained release can similarly be achieved by using
nanocarriers made of erodible or degradable materials. The release can be tailored by
tuning the erosion kinetics of nanoparticles through the selection of biodegradable
polymers, such as poly(lactic acid) (PLA), poly(lactic-co-glycolic acid) (PLGA),
polyethylene glycol (PEG), polycaprolactone (PCL) among others (Sun, 2014), and
through different encapsulation methods (Zilberman, 2008). Sun have successfully
constructed a biodegradable nanoparticle based on a triblock copolymer PEG-b-PCL-
b-poly (2-aminoethylethylene phosphate) (Fig. 3.5.2) with a proven fruitful delivery of
small interfering RNA (siRNA) and paclitaxel for synergistic tumor suppression (Sun,
2011).
Nowadays, the most advanced systems are not so simple and combine a series of
different functionalities that render tunable release of cargos, triggered release
properties, selectivity, and the possibility of tracking the nanocarriers by imaging
and even combined therapy of several active substances. Such systems are further
described in the sections below.
Nanoparticulate systems have been widely used for drug delivery when combined to
molecules that have the particularity to react to certain physiological or pathological
changes in the surrounding environment (Jhaveri, 2014). This is an approach that
Drug Delivery Nanosystems in Cancer Therapy 297
Figure 3.5.3: Stimulus-responsive delivery strategies for tumor targeting. Abbreviations: EPR, enhan-
ced permeability and retention effect; GSH, glutathione. Reprinted with permission from (Zhu, 2013).
298 Current Trends and Developments for Nanotechnology in Cancer
Figure 3.5.5: Schematic representation of the self-assembly, accumulation at the tumor site, uptake
by tumor cells, and triggered intracellular drug release of redox-responsive PEG-SS-PTX/PTX nano-
particles. The programmed drug release undergoes two phases: release of loaded PTX and conjuga-
ted PTX. Abbreviations: PTX, paclitaxel, GSH, glutathione, EPR, enhanced permeability and retention
effect. Reprinted with permission from (Chuan, 2014).
Drug Delivery Nanosystems in Cancer Therapy 301
Variations in the composition and expression of local enzymes are other potential
factors that may provide an opportunity to selectively deliver drugs efficiently in the
tumor sites via an enzyme-triggered mechanism. One of the most popular approaches
is based on the altered expression of matrix metalloproteinases (MMPs), since in the
MMP family, the up-regulation of MMP2 and MMP9 is generally recognized as involved
in the invasion, progression and metastasis of most human tumors (Mansour, 2003).
Zhu developed a multifunctional complex drug delivery system that has prolonged
blood circulation time, targets specifically to tumor cells, responds to the up-regulated
matrix metalloprotease 2 (MMP2) in the tumor microenvironment and has enhanced
intracellular delivery by coupling specific moieties to the surface of a liposomal
nanovector (Zhu, 2012). They were able to show a significant increase in the tumor
targetability, as well as tumor cell internalization of the nanocarrier, and thus, this
approach may be useful for selective drug delivery of cargos in the tumor sites. Other
up-regulated enzymes, namely cancer-associated enzymes, have been utilized for
the same purpose in liposomal formulations, yielding positive results of fast release
kinetics in the specific tumor site (Basel, 2011).
Cancer is a multifactorial disease and often the clinical treatment requires the
administration of more than one drug molecule in order to achieve successful
therapeutic outcomes. In addition, cancer cells frequently become multidrug resistant,
by developing defense mechanisms against the therapeutic approach, like expression
of different receptors/target proteins, self-repairing mechanisms or/and alterations in
their own metabolism, evolving towards a weaker response to the treatment (Holohan,
2013). The combination of different therapeutics can lead to reduced drug resistance
and synergize the therapeutic effect of different active substances, and thus, a dual
positive outcome on the improvement of therapeutic effectiveness and decrease of
deleterious effects (Hu, 2010). Despite the positive advantages of the association of
multidrugs in a delivery system, it is rather challenging to efficiently co-load different
drug molecules, often with distinct physicochemical properties, in the same ordinary
nanocarrier (Hu, 2010; Kratz, 2012; Zhang, 2011), and even more challenging to deliver
such cargos in vivo, where it is absolutely necessary to supply the proper drug ratio at
the tumor level (Mayer, 2006). Several nanoparticulate systems such as lipid-based,
inorganic and polymeric nanoparticles, among other types, have been used to co-
load two or more different payloads adopting different strategies (Hu, 2010). Some
include the physical adsorption of drugs to the nanovector, with the disadvantage
that only drugs with similar physicochemical properties can be loaded simultaneously
(Herranz-Blanco, 2014; Liu, 2012; 2014a). It is also possible to co-load molecules with
different physicochemical properties via a sequential adsorption procedure (Liu,
2013a). Others have followed the incorporation of therapeutics into the particle matrix
302 Current Trends and Developments for Nanotechnology in Cancer
during particle preparation (Doane, 2013), and covalently bonded the drug molecules
to the polymeric backbone after particle preparation (Aryal, 2011).
Despite the obstacles encountered in cancer therapy, extensive progress has been
made so far and nowadays there are several nanoformulations that combine two or
more different anticancer drugs (Liao, 2014; Tardi, 2009) or even two different classes
of therapeutics (Deng, 2013; Liu, 2014b; Saad, 2008). For example, a nanoparticulate
system was developed by Deng using a layer-by-layer deposition approach (Deng,
2013). Liposomes were chosen to be encapsulated by polycationic PLA through the
layer-by-layer deposition, and finally coated by a layer of hyaluronic acid that enhances
the in vivo stability of the nanosystem (Poon, 2011). This system co-delivered siRNA
molecules and doxorubicin loaded in the core of the liposomes, which were able to
knockdown a drug resistance pathway in tumor cells (Whitehead, 2009). This means
they can be applied in the treatment of triple-negative breast cancer form, which is
an estrogen, progesterone and human epidermal growth factor receptor 2 (HER2)-
negative cancer type with a more aggressive effect than other breast cancers (Kassam,
2009; Livasy, 2006). The successful co-delivery of both siRNA and doxorubicin was
achieved, as well as significantly enhanced DXR efficacy by 4-fold in vitro (Fig. 3.5.6).
The in vivo results showed an 8-fold decrease in tumor volume compared to the
control treatments, with no observed toxicity to the animals. An ultimate outer layer
of hyaluronic acid (HA) was deposited to increase the stability of the system (Deng,
2013).
Figure 3.5.6: The co-delivery of siRNA and doxorubicin using the modular doxorubicin-liposome/
PLA/siRNA/PLA/HA LbL nanoparticle platform. (A) Schematics of the siRNA-doxorubicin LbL liposo-
mes. (B) Release profile of the two therapeutics components: siRNA and doxorubicin from the LbL
liposomal nanoparticles in tissue culture medium over 72 h (n = 3). (C) Evaluation of the siRNA-
enhanced cytotoxicity of the combo therapy in MDA-MB-468 cells. The results represent mean ±
standard deviation (n = 3; P < 0.05). Reprinted with permission from (Deng, 2013).
Cancer Targeting 303
Another study that combines siRNA and the co-delivery of paclitaxel in a two-in-one
micelleplex system showed that it is possible to reduce by hundreds of times the dose
of paclitaxel given by associating it with siRNA, and equally achieve a synergistic
therapeutic effect (Sun, 2011). This and other systems are able to provide the co-
delivery of different drugs and therapeutic classes to the tumor sites in vivo, giving
hope to the decrease of harmful side effects of the anticancer drugs.
As described before, there are a variety of strategies that can be applied to
combine different molecules for treatment of tumor diseases. In addition, there is
the possibility that the nanocarrier itself is modified in such a way that it also has
some therapeutic activity. Interestingly, the following study combined doxorubicin
to the pro-apoptotic effect of the nanocarrier itself by means of incorporation of a
sphingolipid C6-ceramide into the liposomal formulation, targeting at the same time
nucleolin, a protein that is greatly expressed in cancer and endothelial cells of tumor
angiogenic blood vessels (Shi, 2007). In this work, the sinergistic combination of
DXR:C6-Cer in a 1:2 molar ratio was identified and tested, which increased the cell
toxicity potential, and thus, enabled at least a 4–6-fold dose reduction of doxorubicin,
independently of the cell line tested. Hence, this approach might be valid to overcome
drug resistance and, at the same time, decrease the deleterious effects of doxorubicin
(Fonseca, 2014).
overexpressed by the cancer cells, with the ultimate goal of defining strategies for
specifically targeting the tumor tissues. In this context, remarkable advances in
nanotechnology led to the emergence of targeting nanodelivery systems as innovative
as complex, which exploit both passive and active mechanisms for targeting an
extensive variety of malignant tissues (Fig. 3.5.7) (Davis, 2008; Farokhzad, 2009;
Ferrari, 2005; Nicolas, 2013). This section briefly describes the fundamentals behind
tumor targeting, scrutinizes different strategies that have been employed for targeting
cancer tissues, and highlights the most recent breakthroughs on designing and
developing targeting drug delivery nanoplatforms for application in cancer therapy.
Figure 3.5.7: Strategies adopted for drug targeting and localization of nanosystems to tumor cells and
tissues. (A) Passive targeting. Circulating nanoparticles passively extravasate in solid tumor tissue
via the enhanced permeability of blood vessels, i.e., through the disorganized and leaky vasculature
surrounding the solid tumor coupled with the absence of lymphatic drainage, and preferentially
accumulate in tumor cells (the EPR effect). (a) The drug is released into the extracellular matrix and
diffuses through the cells and tissue. (B) Active targeting. Once nanoparticles passively extravasate
and concentrate in the target tissue via the EPR effect, the presence of ligands grafted onto the
nanoparticle surface enable active targeting of xthe nanoparticles to receptors that are overexpressed
on tumor cells or tissue, resulting in enhanced uptake and internalization via receptor-mediated
endocytosis. (b) Tumor-specific ligands on the nanoparticles bind to cell surface receptors, triggering
internalization of the nanoparticles into the cell through endosomes on which, due to an internal
acidic pH, the drug is released from the nanoparticles and diffuses into the cytoplasm. (C) Active
targeting to endothelial cells. Nanoparticles can be targeted to bind to angiogenic endothelial cell
surface receptors with the aims of enhancing drug accumulation in the tumor endothelium, thereby
inhibiting growth of blood vessels supplying the tumor rather than inhibiting tumor cells per se (c);
and improving delivery of chemotherapeutic agents to tumor cells via the EPR effect with the potential
to act synergistically in targeting both the vascular tissue and tumor cells. Adapted and reproduced
with permission from (Lammers, 2012). Abbreviation: EPR, enhanced permeability and retention.
Cancer Targeting 305
Figure 3.5.8: Histological data of the nanoparticle uptake in tumors when different sized nanopartic-
les were used in vivo. Reproduced with permission from (Perrault, 2009).
Cancer Targeting 307
Figure 3.5.9: 2D micro-PET cross-sections images, created using a filtered back-projection, of mice
bearing a DLD-1 tumor (n = 3) at 36 h post-injection of 70 (A), 100 (B), and 160 nm (C) 64Cu-labeled
Dox-SMNPs. Each image was scaled to its own maximum. The 70 nm Dox-SMNPs show the highest
tumor-specific uptake among the three studies. Abbreviations: Dox-SMNPs, doxorubicin-loaded
supramolecular magnetic nanoparticles. Adapted and reproduced with permission from (Lee,
2013b).
Similarly to the size, the surface charge of nanoparticles exert an effect on their
behavior in the systemic circulation, including the circulation time, interaction
with elements of the RES and predisposition for interacting with the tumor cells,
affecting cellular association and internalization (Alexis, 2008; He, 2010). In this
regard, positively charged nanocarriers are recognized as owing more propensity for
interacting and, consequently, accumulating at the tumor tissues, as suggested by
studies performed in animal models (He, 2010; Hu-Lieskovan, 2005). An enhanced
interaction between positively charged nanoparticles and the endothelial cells of the
308 Current Trends and Developments for Nanotechnology in Cancer
Table 3.5.2: Examples of non-targeted nanosystems in clinical use for anticancer therapy. Adapted
with permission from (Sanna, 2014).
Figure 3.5.10: The physicochemical properties of the ligand and the nanoparticles affect their blood
circulation profiles, their biodistribution and their ability to be internalized by cancer cells. Reprodu-
ced with permission from (Bertrand, 2014).
The optimization of the targeted nanoparticles’ size is therefore crucial for their
extravasation from the tumor vasculature and accumulation at the tumor size.
Rationally, when a ligand functionalization is intended, the increase on the
hydrodynamic radius after conjugation of the targeting moiety must be taken into
consideration, particularly in the case of large ligands, such as antibodies, proteins or
aptamers, under penalty of negatively influencing the targeting efficiency by hindering
the accumulation of the nanocarrier in the tumor (Humrich, 2010). The size of targeted
nanoparticles was also revealed to impact their intracellular trafficking, with smaller
particles being increasingly found deposited in the cytoplasm and nucleus of cancer
cells, in a study conducted by Lee (Kapsenberg, 2003). In parallel, the influence of the
shape of actively targeted nanoparticles on their cell internalization by cancer cells
has also been recently investigated (Coosemans, 2014; Hanke, 2013; Subbotin, 2014).
The surface of the actively targeted nanoparticles also impact on the interaction
between those and the targeted cells’ surface, therefore contributing for the efficacy
of the targeting. Cationic nanoparticulate systems are acknowledged for more
extensively interacting with negatively-charged cell membranes, resulting in a non-
specific increased cellular association and internalization (Offringa, 2009). However,
an excess in the positive charge of the nanoparticles might increase toxicity and
stimulate immunologic response (Moghimi, 2005). When the nanoparticles’ surface
is functionalized with a targeting moiety, the surface charge is not only influenced
312 Current Trends and Developments for Nanotechnology in Cancer
by the intrinsic charge of the ligand and the nanoplatform, but also by the ligand
density.
Besides the surface charge, the hydrophobicity of the nanocarriers can also
interfere with the way they interact with targeted cells, since the adsorption of plasma
proteins onto the surface of non-sterically stabilized nanoparticles may decrease
the capacity of those to bind to the targeted receptor (Homma, 2005). Contrarily,
the nanoparticles’ surface functionalization with high molecular weight polymers,
such as PEG, may hinder the targeting moieties to efficiently bind to their antigen
(Lowin, 1995; Peters, 1991). The development of a PEG-shielded actively-targeted
nanoplatform with prolonged circulation time and capable of undergoing cleavage
of the steric protection in the tumor interstitium, consequently exposing the targeting
moiety to the cancer cells, might constitute an interesting approach demanding
further investigation (Arroyo, 2002).
Finally, the density of the targeting molecules functionalizing the nanoparticles’
surface significantly influences the avidity of targeting nanotherapeutics to the
complementary receptor. Theoretically, a higher density of the targeting moieties
generates a favorable binding enthalpy in the system, consequently inducing the
interaction between the ligand and the antigen (Callard, 1977; Liu, 2007), and also
stimulates the overexpression of the receptor at the cells’ surface, therefore resulting
in an enhanced internalization of the actively targeted nanoparticles by the tumor
cells. Nevertheless, various aspects, including the inappropriate orientation of the
targeting ligand, the steric hindrance of subjacent molecules and a competitive
behavior towards the receptor binding may revert this relation between the ligand
density and the efficiency of cell association (James, 1981; Valdez, 2000).
3.5.3.2.3.3 Aptamers
Nucleic acid-based aptamers are small, single-stranded RNA or DNA oligonucleotides,
where the conformational structure can be designed, rendering them the capability
of binding antigens with high affinity and specificity (Bellisola, 2012; Pan, 2010).
Candidates for targeting a specific receptor have been screened with oligonucleotide
libraries using high throughput screening methodologies. The Promising aptamers are
subsequently selected and amplified in detriment of the remaining ones. Recently, an
in vitro chemical technique denominated “systemic evolution of ligands by exponential
enrichment” (SELEX) has been explored for identification of the referred candidates
(Wang, 2013). This technique has been used by Langer and Farokhzad for developing
customized controlled-release nanoplatforms capable of specifically targeting the
prostate-specific membrane antigen (PMSA) (Ferris, 2010; Ma, 2013; Monjanel, 2014;
Nembrini, 2011; Tao, 2012). Aptamers present attractive intrinsic features to be used as
targeting ligands, such as smaller size and reduced immugenicity, in comparison with
mAb or other macromolecules, resulting in improved stability and biodistribution
profiles (Oki, 2012; Pawelec, 1999). Similarly to the other classes of targeting moieties,
the use of aptamers presents some limitations, which hinder their clinical success.
First of all, the susceptibility of nucleic acids to the action of nucleases, the stability
of the aptamer conjugation onto the nanoparticles’ surface becomes a major concern.
Second of all, the negative charge of the aptamers attributed to the phosphodiester
backbone is suspected to impact the systemic circulation kinetics of the actively-
targeted nanosystem. Finally, although presenting a relatively small molecular size,
the stiffness of the nucleic acid conformation may be translated into an increased
hydrodynamic radius of the targeting nanosystem after ligand conjugation, hence
influencing its accumulation at the targeted site, as previously discussed in this
section.
Cancer Targeting 315
Table 3.5.3: Tumor-targeted nanomedicines in clinical development. Adapted and reprinted with
permission from (Bertrand, 2014).
properties and antitumor activity. It was shown that mice treated with ovalbumin-
encapsulated CPTEG:CPH particles elicited very high CD8+ T cell responses on days 14
and 20, leading to a delay of the tumor progression and extended survival time. These
results suggest that the immunoadjuvant properties of the nanoparticles can enhance
antigen-specific immuno-cellular responses, and thus, they can be potentially applied
for the promotion of antitumor responses in cancer patients.
Figure 3.5.11: Confocal microscopy imaging of the PLGA nanoparticles in human DCs. (A) 4′,6-
Diamidino-2-phenylindole channel. (B) Fluorescein isothiocyanate channel. (C) 4′,6-Diamidino-2-
phenylindole and fluorescein isothiocyanate channels overlaid. (D) 4′,6-Diamidino-2-phenylindole,
fluorescein isothiocyanate, and reflection channels overlaid. Reprinted with permission from (Ma,
2012).
Since DCs have the intrinsic ability of phagocyting foreign material, passive DC
targeting for cancer immunotherapy is possible by directing the nanoparticles to sites
rich in DCs. For this purpose, the route of administration of the nanoparticles is very
important. For example, while subcutaneously administered small nanoparticles (<
100 nm) can be quickly transported into the lymphatic system to interact with the
DC localized in the lymph node (Cubas, 2009; Manolova, 2008; Reddy, 2007), larger
nanoparticles (> 500 nm) stay for a longer time in the injection site of the skin and
predominantly interact with the DCs or monocytes available in the skin (Manolova,
2008; Moon, 2012). To avoid the size dependency of passive DC targeting to the lymph
node, direct injection into the blood is suggested as an efficient way to reach both
blood and splenic DCs and induce strong immune responses by a variety of particles
(Perche, 2011; Tacken, 2011). However, the macrophage uptake is still one of the main
drawbacks of the cancer immunotherapy using this strategy.
322 Current Trends and Developments for Nanotechnology in Cancer
DCs also express many cell surface receptors, such as mannose (Chenevier, 2002;
Ramakrishna, 2004), CD11c (Cruz, 2014), DEC-205 (Cruz, 2014), Langerin (Flacher,
2010), DC-SIGN (Cruz, 2010; Varga, 2013), clec9A (Caminschi, 2008), and DCIR2
(Dudziak, 2007), which can facilitate actively targeted cellular endocytosis of the
nanoparticles via appropriate surface modification with the ligands that target these
receptors on the surface of the DCs (Foged, 2002). It has already been reported that the
conjugation of such targeting moieties on the surface of the nanoparticles increases
their uptake by the DCs and enhances the DC maturation, as evidenced by the high
secretion of cytokine and up-regulation of CD83 and CD86 surface maturation markers
(Kempf, 2003). The type of the target receptor can have a substantial effect on the
immunological response of nano-immunotherapeutics. A study by Dudziak showed
that the targeting antigens to DEC-205 results mainly in cross-presentation of antigens
to the CD8 T cells, whereas antigens targeted to DCIR2 are preferentially presented via
the MHC class II molecules to the CD4 T cells (Dudziak, 2007). In addition, Idoyaga
(Idoyaga, 2011) have shown that, while antigens can be targeted to the splenic CD8α+
DC subset by antigen conjugation to antibodies against Langerin and DEC205, DCIR2
antibodies can specifically target CD8α‒ DCs. In addition, it has been proved that
effective active DC targeting can be achieved by reducing the nonspecific interaction
of the nanoparticles with plasma constituents and other cells in the bloodstream,
by introducing a hydrophilic biomaterial consisting of PEG onto the surface of the
particles (Hillaireau, 2009).
Polysaccharides (Weber, 2010), PLA (Primard, 2010), PLGA (Binjawadagi, 2014),
polyanhydrides (Camacho, 2011), and polyphosphazene (Garlapati, 2012) are among
the most widely studied nanomaterials prepared for immunostimulative purposes.
Different techniques have been employed to prepare the nano-immunotherapeutics
using the abovementioned polymers, including spray drying, emulsification/
solvent evaporation, and coacervation (Kalkanidis, 2006; Oyewumi, 2010). All
of these methods can produce polymeric nanoparticles with a large surface area
that improve the interaction between immune cells and the immunostimulative
payload (Bachmann, 2010), leading to enhanced uptake of antigens by DCs and
improved immune responses (Silva, 2013). The main reason for the great interest in
immunostimulatory potential of the abovementioned nanoparticles is the superior
biocompatibility, versatile chemistry, high protection of the loaded biomolecules, high
loading efficiency, efficient endocytosis and high presentation of the immunogenic
molecules (Binjawadagi, 2014; Camacho, 2011; Garlapati, 2012; Jiang, 2005; Primard,
2010; Weber, 2010). For example, some of these nanocarriers have shown endosomal
escape properties, potentiating the proteosome-dependent processing of the
immunogenic payload and cross-presentation through MHC class I (Jia, 2013; Silva,
2013). Another benefit of these nanoparticles is the ability to control the release pattern
of the loaded immunostimulative compound and to act as an adjuvant and increase
the therapeutic effect of the formulation (Cohen, 1994; Nandakumar, 2012; Oyewumi,
2010). For example, Gómez (Gomez, 2006) have reported the ability of poly(methyl
Nanotechnology and Immunotherapy 323
Figure 3.5.12: Anti-OVA IgG1 and IgG2a titers in sera of female BALB/c mice after intradermal immu-
nisation with 10 μg of ovalbumin as follows: ovalbumin solution (OVA), ovalbumin adsorbed to Alum
(OVA-Alum), blank nanoparticles (NP), ovalbumin coated nanoparticles (NP I), ovalbumin encapsu-
lated nanoparticles (NP II) and 1,3-diaminopropane cross-linked ovalbumin encapsulated nano-
particles (NP III and NP IV). The antibody titer is defined as the reciprocal dilution giving an optical
density. Reprinted with permission from (Gomez, 2006).
intensive research studies are still needed in order to develop nano-formulations that
can be produced in large scale and applied in the clinic.
The early diagnosis of tumors through screening based on imaging might contribute
to the reduction in the mortality for certain types of cancers. Real-time imaging is a
most valuable technique that enables visualization of the tissue morphology and cell
function, thus allowing the identification of pathological situations that generally
cannot be aquired in in vitro analysis (Fass, 2008). Imaging techniques are being
combined with nanotechnology, rendering, in many cases, extraordinarily sensitive
and powerful diagnostic and imaging tools that, by other means, were not available
with small or microscale tools, thus enhancing the tissue contrast or allowing
the identification of specific biological changes (Toy, 2014). As it is necessary to
specifically target the tumor sites in order to achieve a more successful outcome out
of the treatment, it is also of utmost importance that the diagnostic and imaging tools,
such as those extensively used in the medical practice: Magnetic Resonance Imaging
(MRI), Computed Tomography (CT) and Positron Emission Topography (PET), are
reliable enough to facilitate the right medical decisions, which can be improved by
alliyng them to nanotechnology (Lee, 2013a; Perez-Medina, 2014).
Contrast agents are frequently employed to assist and complement the
visualization of abnormalities in a diagnostic image, interacting with the incident
radiation to yield noticeable alterations in the final aquired images. They increase
the power of the imaging techniques since there is an increased sensitivity, allowing
the diagnosis of many diseases that are not possible to detect using conventional
methods (Power, 2011). However, nanotechnology is impacting this area of diagnostic
imaging by opening new avenues of novel imaging techniques, and by allowing
the enhancement of the sensitivity, pharmacokinetics, pharmacodinamics and
biocompatibility features of several contrast agents, as well as providing a high signal-
to-noise ratio (Rosen JE, 2011). It is becoming inevitable to not associate imaging and
therapeutic properties in a single nanoplatform. Multifunctional nanocarriers are
becoming more and more fashionable and many studies have been focusing on the
production of multistage nanoparticulate systems.
The employment of nanoparticulate systems to image and diagnose cancer has
been extensively applied and new approaches have been developed (Choi, 2012).
Some advantages of using nanoparticulate-based probes for imaging purposes is
attained to their low cytotoxicity profiles and physicochemical properties (Santos,
2013a). In general, a surface area-to-volume ratio and effective surface functionality,
are particularities used as tools to tune the nanoparticle properties for tissue or
cell targeting in vivo through high affinity to certain cell biomarkers (Davis, 2008).
Biochemical changes in vivo, such as enhanced receptor density in a certain tumor
Cancer Imaging, Diagnostics and Multifuctional Nanosystems 325
type could be imaged and measured by probing with specific nanoparticulate systems
avid to bind to it (Jin, 2014; Satpathy, 2015). These probes shall attain certain features
that allow them to be detected and promote their accumulation at the site of interest,
be safe and biocompatible, as well as to have limited side effects and to avoid the
hostile environments encountered in vivo, such as avoidance of interactions with
plasma proteins and the recognition by the phagocytic cells and consequent removal
from the systemic circulation by the mononuclear phagocyte system (Santos, 2013a).
For example, superparamagnetic iron oxide nanoparticles (SPIONs), along with many
other clinical applications, are some of the nanoparticulate systems that are most
used as highly effective contrast agents for MRI diagnosis of solid tumors (Ittrich,
2013; Rosen, 2012). Compared to other traditional contrast agents, SPIONs exhibit
several advantageous properties, among them the greater magnetic signal strength
and longer lasting contrast enhancement, low cytotoxic effects, biodegradability and
improved delineation of tumor margins (Corot, 2006; Varallyay, 2002; Wang, 2001).
Such nanoparticulate systems have been used to target solid tumors, either in a passive
way, (Zolata, 2014), taking advantage of the leaky and damaged tumor vasculature
through the enhanced permeability and retention effect (Brannon-Peppas, 2004;
Rosen, 2012), by targeting actively the tumor sites by attaching a targeting ligand to
the surface of the SPIONs, or by taking into account the tumor pathological features,
such its lower pH microenvironment. The internalization of SPIONs enables a longer
and effective imaging time by improvement of the contrast-enhancing effect of the
particles through the accumulation of a high number of SPIONs in the tumor tissue,
therefore being a more promissing and sensitive molecular diagnostics approach than
passive targeting imaging. To make it possible to biofunctionalize the SPIONs, it is
necessary to provide them with a surface coating system in order to create a desirable
platform for conjugation of targeting ligands or to make the SPIONs sensitive to tumor
characteristic microenvironment (Fig. 3.5.13) (Jin, 2014; Xie, 2010).
Figure 3.5.13: Schematic representation illustration of the polymer micelle composite system of
surface-engineered SPION for imaging and therapy applications. Adapted from (Jin, 2014).
326 Current Trends and Developments for Nanotechnology in Cancer
Recently, several studies utilizing SPIONs for the active targeting of several types
of cancer were conducted and showed promising results (Kievit, 2012; Ling, 2014;
Melancon, 2011; Zhang, 2014a; Zhang, 2014c). “CLIO” nanoparticles, which are
SPIONs cross-linked with a dextran coating, have appeared as a powerful multistage
nanovector for targeted imaging and diagnostic applications, allowing a wide and
versatile conjugation of targeting moieties and compatible with diagnostic imaging
MRI, optical and PET techniques (Tassa, 2011). Imaging and diagnosis can be achieved
with this nanoscaled platform, as well as the delivery of therapeutic cargos to the tumor
sites, making multifunctional systems or “theranostic” (Choi, 2012). For example, in
a recent study, CLIO nanoparticles were developed with the purpose of producing a
multifunctional platform. Taking advantage of the eminent levels of MMPs in the tumor
microenvironment, in particular MMP-14, which has an important role in the tumor
invasion features and provides a direct cellular target for prodrug activation (Devy,
2009; Landry, 2005), it was found that functionalization of the CLIO nanoparticulate
system with the MMP-14 cleavable peptide-conjugate of azademethylcolchicine leads
to vascular collapse of the tumor (Reyes-Aldasoro, 2008). This theranostic nanovector
provided an opportunity to image selectively the tumor site and, at the same time,
track the drug payloads delivered to the tumour site. In addition, the conjugation of
a drug molecule to nanoparticles allowed for a higher amount of drug cargo to be
delivered, leading to a more effective therapeutic outcome and, again, decreasing the
associated toxicity of the anticancer drug to healthy tissues (Ansari, 2014).
PSi nanomaterials are other example of materials that have been proven to be
biocompatible and possess remarkable features for biomedical applications, including
imaging and drug delivery (Liu, 2013b; Santos, 2013b; Shahbazi, 2012). Besides the
vast medical applications (Park, 2009; Santos, 2014; Shahbazi, 2012), and taking into
account a special property of this material, its photoluminescence (Anglin, 2008;
Hua, 2005), PSi has been explored for clinical diagnosis and imaging applications
(Godin, 2011; Santos, 2013a), such as acting as a self-reporting drug delivery systems
using optical imaging in living animals (Cheng, 2008), multicolor near infrared
imaging in live mice (Erogbogbo, 2010), incorporation of gadolidium (Ananta, 2010)
and SPIONs inside the nanoporous structure for MRI applications (Kinsella, 2011), as
well as radiolabeling of PSi with radioisotopes for imaging purposes (Bimbo, 2010;
Sarparanta, 2012).
Liposomal formulations have also been used as theranostic nanoplatforms for
both cancer therapeutics and imaging, considering its safety and wide clinical use.
Recent studies included the encapsulation of Magnevist, a contrast agent for MRI
with in vivo contrast-enhancing effects into liposomes, which showed, together
with doxorubicin, an improved biocompatibility when coated with hyaluronic acid
ceramide polymer, aimed for targeted imaging and drug delivery (Park, 2014). SPIONs
have been successfully encapsulated inside immuno-liposomes targeting endothelial
cells in tumor vasculature and applied in MRI imaging (Zhang, 2014a). Multimodal
fluorescent labeled liposomes with thermo-responsive release properties of the
Conclusions and Future Prospects 327
contrast agents are some of the more advanced imaging platforms fabricated so far
(Kokuryo, 2014), and together with near infrared (NIR) fluorescent dyes, have been
similarly loaded inside theranostic liposomes with improved label efficiency (Xie,
2014).
The abovementioned nanoparticulate systems are revolutionary for the diagnostic
and imaging fields, but other materials such as gold nanomaterials, carbon nanotubes
and silica nanoparticles have also shown light absorbing properties, that when
scattered or emited, yield specific types of diagnostic/therapeutic signals, e.g. under
ultrasound, heat, Raman or fluorescence signals. These types of nanoparticles may
function as theranostic nanoparticles on their own and/or can be associated with
other molecules for increases in targeting and selectivity (Choi, 2012).
Nanotechnology has provided the opportunity to explore new clues for the diagnosis
and treatment of various diseases, in particular, cancer. With the advances of
nanomedicine, the direct use of nanoparticles in prevention, diagnostic and treatment
of cancer, has become a reality. In the past decades, the use of nanotechnology
for simultaneous drug delivery, imaging and diagnostics has grown exponentially.
There is a great evolution of the biological and physical resources to improve the use
of nanoparticles for finding the best treatment for cancer. The controlled size and
multifunctionality are the main reasons for the growing applications of nanomedicines
as anticancer agents. Nanoparticles can synergistically improve the immunostimulative
effect of antigens and other immunogenic molecules. However, using nanoparticles
as anticancer immune adjuvants and delivery of immunostimulative biomolecules
still suffers from several important drawbacks, including the problems associated
with the scaling-up of the system, expensive preparation processes, and limitations
for producing sterile products
Tremendous and fast advancements are being achieved to produce better and
improved nanosystems able to provide precise information for a better understanding
of the tumor diseases, as well as for facilitating medical decisions in the clinic,
adding therapeutic functions to the novel nanoplatforms that can, in the near future,
change the practical course of the treatment and diagnosis of cancer. The bench-
to-bedside translation of the nanomedicines and technologies have introduced a
new era in the design and development of innovative but simultaneously complex
targeting nanoparticles for delivery of therapeutic and diagnostic agents to tumors.
Accordingly, despite the great versatility and promising features observed for such
anticancer therapeutic systems, intensive research studies are still needed at the pre-
clinical level in order to develop nano-formulations that can be produced in large
scale and applied in the clinic.
328 Current Trends and Developments for Nanotechnology in Cancer
References
ACS. (2014). Cancer. American Cancer Society, Retrieved 27th October 2014, from http://www.
cancer.org/.
Alexis, F., Pridgen, E., Molnar, L. K., et al. (2008). Factors affecting the clearance and biodistribution
of polymeric nanoparticles. Molecular Pharmaceutics, 5(4), 505-515.
Allen, T. M. (2002). Ligand-targeted therapeutics in anticancer therapy. Nature Reviews Cancer, 2(10),
750-763.
Almeida, J. P., Figueroa, E. R., Drezek, R. A. (2014). Gold nanoparticle mediated cancer
immunotherapy. Nanomedicine, 10(3), 503-514.
Ananta, J. S., Godin, B., Sethi, R., et al. (2010). Geometrical confinement of gadolinium-based
contrast agents in nanoporous particles enhances t1 contrast. Nature Nanotechnol, 5(11), 815-
821.
Anglin, E. J., Cheng, L., Freeman, W. R., et al. (2008). Porous silicon in drug delivery devices and
materials. Advanced Drug Delivery Reviews, 60(11), 1266-1277.
Annabi, N., Tamayol, A., Uquillas, J. A., et al. (2014). 25th anniversary article: Rational design and
applications of hydrogels in regenerative medicine. Advanced Materials, 26(1), 85-123.
Ansari, C., Tikhomirov, G. A., Hong, S. H., et al. (2014). Development of novel tumor-targeted
theranostic nanoparticles activated by membrane-type matrix metalloproteinases for combined
cancer magnetic resonance imaging and therapy. Small, 10(3), 566-575, 417.
Arias, J. L. (2011). Drug targeting strategies in cancer treatment: An overview. Mini-Reviews in
Medicinal Chemistry, 11(1), 1-17.
Arroyo, A., Modriansky, M., Serinkan, F. B., et al. (2002). Nadph oxidase-dependent oxidation and
externalization of phosphatidylserine during apoptosis in me2so-differentiated hl-60 cells.
Role in phagocytic clearance. Journal of Biological Chemistry, 277(51), 49965-49975.
Aryal, S., Hu, C.-M. J., Zhang, L. (2011). Polymeric nanoparticles with precise ratiometric control over
drug loading for combination therapy. Molecular Pharmceutics, 8(4), 1401-1407.
Bachmann, M. F., Jennings, G. T. (2010). Vaccine delivery: A matter of size, geometry, kinetics and
molecular patterns. Nature Reviews Immunology, 10(11), 787-796.
Basel, M. T., Shrestha, T. B., Troyer, D. L., et al. (2011). Protease-sensitive, polymer-caged liposomes:
A method for making highly targeted liposomes using triggered release. ACS Nano, 5(3), 2162-
2175.
Baxevanis, C. N. (2008). Antibody-based cancer therapy. Expert Opinion on Drug Discovery, 3(4),
441-452.
Bellisola, G., Sorio, C. (2012). Infrared spectroscopy and microscopy in cancer research and
diagnosis. American Journal of Cancer Research, 2(1), 1-21.
Berretta, F., St-Pierre, J., Piccirillo, C. A., et al. (2011). Il-2 contributes to maintaining a balance
between cd4+foxp3+ regulatory t cells and effector cd4+ t cells required for immune control of
blood-stage malaria infection. Journal of Immunology, 186(8), 4862-4871.
Bertrand, N., Leroux, J. C. (2012). The journey of a drug-carrier in the body: An anatomo-physiological
perspective. Journal of Control Release, 161(2), 152-163.
References 329
Bertrand, N., Wu, J., Xu, X., et al. (2014). Cancer nanotechnology: The impact of passive and active
targeting in the era of modern cancer biology. Advanced Drug Delivery Reviews, 66, 2-25.
Bimbo, L. M., Denisova, O. V., Mäkilä, E., et al. (2013). Inhibition of influenza a virus infection in vitro
by saliphenylhalamide-loaded porous silicon nanoparticles. ACS Nano, 7(8), 6884-6893.
Bimbo, L. M., Sarparanta, M., Santos, H. A., et al. (2010). Biocompatibility of thermally
hydrocarbonized porous silicon nanoparticles and their biodistribution in rats. ACS Nano, 4(6),
3023-3032.
Binjawadagi, B., Dwivedi, V., Manickam, C., et al. (2014). Adjuvanted poly(lactic-co-glycolic)
acid nanoparticle-entrapped inactivated porcine reproductive and respiratory syndrome
virus vaccine elicits cross-protective immune response in pigs. International Journal of
Nanomedicine, 9, 679-694.
Bonsignori, M., Pollara, J., Moody, M. A., et al. (2012). Antibody-dependent cellular cytotoxicity-
mediating antibodies from an hiv-1 vaccine efficacy trial target multiple epitopes and
preferentially use the vh1 gene family. Journal of Virology, 86(21), 11521-11532.
Brannon-Peppas, L., Blanchette, J. O. (2004). Nanoparticle and targeted systems for cancer therapy.
Advanced Drug Delivery Reviews, 56(11), 1649-1659.
Cabral, H., Matsumoto, Y., Mizuno, K., et al. (2011). Accumulation of sub-100 nm polymeric micelles
in poorly permeable tumours depends on size. Nature Nanotechnology, 6(12), 815-823.
Callard, R. E., Basten, A., Waters, L. K. (1977). Immune function in aged mice. Ii. B-cell function. Cell
Immunology, 31(1), 26-36.
Camacho, A. I., Da Costa Martins, R., Tamayo, I., et al. (2011). Poly(methyl vinyl ether-co-maleic
anhydride) nanoparticles as innate immune system activators. Vaccine, 29(41), 7130-7135.
Caminschi, I., Proietto, A. I., Ahmet, F., et al. (2008). The dendritic cell subtype-restricted c-type
lectin clec9a is a target for vaccine enhancement. Blood, 112(8), 3264-3273.
Carmeliet, P., Jain, R. K. (2000). Angiogenesis in cancer and other diseases. Nature, 407(6801), 249-
257.
Chandramohan, V., Mitchell, D. A., Johnson, L. A., et al. (2013). Antibody, t-cell and dendritic cell
immunotherapy for malignant brain tumors. Future Oncology, 9(7), 977-990.
Chandrasekhar, P., Shahid-Mohammad, S., Debnath, S., et al. (2013). A review on latest trends in
oncho nanotechnology. International Journal of Pharmaceutical Development and Technology,
3(2), 106-109.
Chen, C. Y., Huang, D., Yao, S., et al. (2012). Il-2 simultaneously expands foxp3+ t regulatory and
t effector cells and confers resistance to severe tuberculosis (tb): Implicative treg-t effector
cooperation in immunity to tb. Journal of Immunology, 188(9), 4278-4288.
Chen, Z. Y., Wang, Y. X., Lin, Y., et al. (2014). Advance of molecular imaging technology and targeted
imaging agent in imaging and therapy. Biomed Research International, 2014, 819324.
Chenevier, P., Grandjean, C., Loing, E., et al. (2002). Grafting of synthetic mannose receptor-ligands
onto onion vectors for human dendritic cells targeting. Chemical Communications (Camb)(20),
2446-2447.
Cheng, K., Peng, S., Xu, C., et al. (2009). Porous hollow fe3o4 nanoparticles for targeted delivery and
controlled release of cisplatin. Journal of American Chemical Society, 131(30), 10637-10644.
Cheng, L., Anglin, E., Cunin, F., et al. (2008). Intravitreal properties of porous silicon photonic
crystals: A potential self-reporting intraocular drug-delivery vehicle. British Journal of
Ophthalmology, 92(5), 705-711.
Cho, K. J., Wang, X., Nie, S. M., et al. (2008). Therapeutic nanoparticles for drug delivery in cancer.
Clinical Cancer Research, 14(5), 1310-1316.
Choi, K. Y., Liu, G., Lee, S., (2012). Theranostic nanoplatforms for simultaneous cancer imaging and
therapy: Current approaches and future perspectives. Nanoscale, 4(2), 330-342.
Chuan, X., Song, Q., Lin, J., et al. (2014). Novel free-paclitaxel-loaded redox-responsive nanoparticles
based on a disulfide-linked poly(ethylene glycol)-drug conjugate for intracellular drug
330 Current Trends and Developments for Nanotechnology in Cancer
delivery: Synthesis, characterization, and antitumor activity in vitro and in vivo. Molecular
Pharmaceutics, 11(10), 3656-3670.
Cohen, S., Alonso, M. J., Langer, R. (1994). Novel approaches to controlled-release antigen delivery.
International Journal of Technology Assessment in Health Care, 10(1), 121-130.
Coosemans, A., Tuyaerts, S., Vanderstraeten, A., et al. (2014). Dendritic cell immunotherapy in
uterine cancer. Human Vaccines & Immunotherapeutics, 10(7), 1822-1827.
Corot, C., Robert, P., Idee, J. M., et al. (2006). Recent advances in iron oxide nanocrystal technology
for medical imaging. Advanced Drug Delivery Reviews, 58(14), 1471-1504.
Correia-Pinto, J. F., Csaba, N., Alonso, M. J. (2013). Vaccine delivery carriers: Insights and future
perspectives. International Journal of Pharmaceutics, 440(1), 27-38.
Coti, K. K., Belowich, M. E., Liong, M., et al. (2009). Mechanised nanoparticles for drug delivery.
Nanoscale, 1(1), 16-39.
Couvreur, P., Vauthier, C. (2006). Nanotechnology: Intelligent design to treat complex disease.
Pharmaceutical Research, 23(7), 1417-1450.
Cruz, L. J., Rosalia, R. A., Kleinovink, J. W., et al. (2014). Targeting nanoparticles to cd40, dec-205 or
cd11c molecules on dendritic cells for efficient cd8 t cell response: A comparative study. Journal
of Controlled Release, 192C, 209-218.
Cruz, L. J., Tacken, P. J., Fokkink, R., et al. (2010). Targeted plga nano- but not microparticles
specifically deliver antigen to human dendritic cells via dc-sign in vitro. Journal of Controlled
Release, 144(2), 118-126.
Cubas, R., Zhang, S., Kwon, S., et al. (2009). Virus-like particle (vlp) lymphatic trafficking and
immune response generation after immunization by different routes. Journal of Immunotherapy,
32(2), 118-128.
Danhier, F., Feron, O., Preat, V. (2010). To exploit the tumor microenvironment: Passive and active
tumor targeting of nanocarriers for anti-cancer drug delivery. Journal of Controlled Release,
148(2), 135-146.
Danquah, M. K., Zhang, X. A., Mahato, R. I. (2011). Extravasation of polymeric nanomedicines across
tumor vasculature. Advanced Drug Delivery Reviews, 63(8), 623-639.
Davis, I. D., Chen, W., Jackson, H., et al. (2004). Recombinant ny-eso-1 protein with iscomatrix
adjuvant induces broad integrated antibody and cd4(+) and cd8(+) t cell responses in humans.
Proceedings of the National Academy of Sciences, 101(29), 10697-10702.
Davis, M. E., Chen, Z., Shin, D. M. (2008). Nanoparticle therapeutics: An emerging treatment
modality for cancer. Nature Reviews Drug Discovery, 7(9), 771-782.
Deng, Z. J., Morton, S. W., Ben-Akiva, E., et al. (2013). Layer-by-layer nanoparticles for systemic
codelivery of an anticancer drug and sirna for potential triple-negative breast cancer treatment.
ACS Nano, 7(11), 9571-9584.
Devy, L., Huang, L., Naa, L., et al. (2009). Selective inhibition of matrix metalloproteinase-14 blocks
tumor growth, invasion, and angiogenesis. Cancer Research, 69(4), 1517-1526.
Diwan, M., Elamanchili, P., Cao, M., et al. (2004). Dose sparing of cpg oligodeoxynucleotide vaccine
adjuvants by nanoparticle delivery. Current Drug Delivery, 1(4), 405-412.
Doane, T., Burda, C. (2013). Nanoparticle mediated non-covalent drug delivery. Advanced Drug
Delivery Reviews, 65(5), 607-621.
Dobrovolskaia, M. A., Patri, A. K., Simak, J., et al. (2012). Nanoparticle size and surface charge
determine effects of pamam dendrimers on human platelets in vitro. Molecular Pharmceutics,
9(3), 382-393.
Dong, X., Mumper, R. J. (2010). Nanomedicinal strategies to treat multidrug-resistant tumors: Current
progress. Nanomedicine (Lond), 5(4), 597-615.
Dreher, M. R., Liu, W., Michelich, C. R., et al. (2006). Tumor vascular permeability, accumulation, and
penetration of macromolecular drug carriers. Journal of National Cancer Institute, 98(5), 335-
344.
References 331
Dudziak, D., Kamphorst, A. O., Heidkamp, G. F., et al. (2007). Differential antigen processing by
dendritic cell subsets in vivo. Science, 315(5808), 107-111.
Egilmez, N. K., Harden, J. L., Rowswell-Turner, R. B. (2012). Chemoimmunotherapy as long-term
maintenance therapy for cancer. Oncoimmunology, 1(4), 563-565.
Elzoghby, A. O., Samy, W. M., Elgindy, N. A. (2012). Albumin-based nanoparticles as potential
controlled release drug delivery systems. Journal of Controlled Release, 157(2), 168-182.
Erogbogbo, F., Yong, K.-T., Roy, I., et al. (2010). In vivo targeted cancer imaging, sentinel lymph node
mapping and multi-channel imaging with biocompatible silicon nanocrystals. ACS Nano, 5(1),
413-423.
Fang, J., Nakamura, H., Maeda, H. (2011). The epr effect: Unique features of tumor blood vessels for
drug delivery, factors involved, and limitations and augmentation of the effect. Advanced Drug
Delivery Reviews, 63(3), 136-151.
Farokhzad, O. C., Langer, R. (2009). Impact of nanotechnology on drug delivery. ACS Nano, 3(1),
16-20.
Fasol, U., Frost, A., Buchert, M., et al. (2012). Vascular and pharmacokinetic effects of endotag-1 in
patients with advanced cancer and liver metastasis. Annals of Oncology, 23(4), 1030-1036.
Fass, L. (2008). Imaging and cancer: A review. Molecular Oncology, 2(2), 115-152.
Fattal, E., Barratt, G. (2009). Nanotechnologies and controlled release systems for the delivery of
antisense oligonucleotides and small interfering rna. British Journal of Pharmacology, 157(2),
179-194.
Fernandez-Fernandez, A., Manchanda, R., McGoron, A. J. (2011). Theranostic applications of
nanomaterials in cancer: Drug delivery, image-guided therapy, and multifunctional platforms.
Applied Biochemistry and Biotechnology, 165(7-8), 1628-1651.
Ferrari, M. (2005). Cancer nanotechnology: Opportunities and challenges. Nature Reviews Cancer,
5(3), 161-171.
Ferris, R. L., Jaffee, E. M., Ferrone, S. (2010). Tumor antigen-targeted, monoclonal antibody-based
immunotherapy: Clinical response, cellular immunity, and immunoescape. Journal of Clinical
Oncology, 28(28), 4390-4399.
Fifis, T., Gamvrellis, A., Crimeen-Irwin, B., et al. (2004). Size-dependent immunogenicity:
Therapeutic and protective properties of nano-vaccines against tumors. Journal of Immunology,
173(5), 3148-3154.
Flacher, V., Tripp, C. H., Stoitzner, P., et al. (2010). Epidermal langerhans cells rapidly capture and
present antigens from c-type lectin-targeting antibodies deposited in the dermis. Journal of
Investigative Dermatology, 130(3), 755-762.
Foged, C., Brodin, B., Frokjaer, S., et al. (2005). Particle size and surface charge affect particle
uptake by human dendritic cells in an in vitro model. International Journal of Pharmacuetics,
298(2), 315-322.
Foged, C., Sundblad, A., Hovgaard, L. (2002). Targeting vaccines to dendritic cells. Pharmaceutics
Research, 19(3), 229-238.
Fonseca, N. A., Gomes-da-Silva, L. C., Moura, V., et al. (2014). Simultaneous active intracellular
delivery of doxorubicin and c6-ceramide shifts the additive/antagonistic drug interaction of
non-encapsulated combination. Journal of Controlled Release, 196(0), 122-131.
Foster, S., Duvall, C. L., Crownover, E. F., et al. (2010). Intracellular delivery of a protein antigen with
an endosomal-releasing polymer enhances cd8 t-cell production and prophylactic vaccine
efficacy. Bioconjugate Chemistry, 21(12), 2205-2212.
Gajewski, T. F., Woo, S. R., Zha, Y., et al. (2013). Cancer immunotherapy strategies based on
overcoming barriers within the tumor microenvironment. Current Opinion on Immunology,
25(2), 268-276.
Garlapati, S., Garg, R., Brownlie, R., et al. (2012). Enhanced immune responses and protection
by vaccination with respiratory syncytial virus fusion protein formulated with cpg
332 Current Trends and Developments for Nanotechnology in Cancer
Idoyaga, J., Lubkin, A., Fiorese, C., et al. (2011). Comparable t helper 1 (th1) and cd8 t-cell immunity
by targeting hiv gag p24 to cd8 dendritic cells within antibodies to langerin, dec205, and
clec9a. Proceedings of the National Academy of Sciences, 108(6), 2384-2389.
Ishihara, T., Mizushima, T. (2010). Techniques for efficient entrapment of pharmaceuticals in
biodegradable solid micro/nanoparticles. Expert Opinion on Drug Delivery, 7(5), 565-575.
Ittrich, H., Peldschus, K., Raabe, N., et al. (2013). Superparamagnetic iron oxide nanoparticles in
biomedicine: Applications and developments in diagnostics and therapy. Rofo, 185(12), 1149-
1166.
Jain, R. K. (1987). Transport of molecules across tumor vasculature. Cancer Metastasis Reviews, 6(4),
559-593.
Jain, R. K. (1998). The next frontier of molecular medicine: Delivery of therapeutics. Nature Medicine,
4(6), 655-657.
Jain, R. K., Stylianopoulos, T. (2010). Delivering nanomedicine to solid tumors. National Reviews on
Clinical Oncology, 7(11), 653-664.
James, S. L., Labine, M., Sher, A. (1981). Mechanisms of protective immunity against schistosoma
mansoni infection in mice vaccinated with irradiated cercariae. I. Analysis of antibody and
t-lymphocyte responses in mouse strains developing differing levels of immunity. Cell
Immunology, 65(1), 75-83.
Janib, S. M., Moses, A. S., MacKay, J. A. (2010). Imaging and drug delivery using theranostic
nanoparticles. Advanced Drug Delivery Reviews, 62(11), 1052-1063.
Jhaveri, A., Deshpande, P., Torchilin, V. (2014). Stimuli-sensitive nanopreparations for combination
cancer therapy. Journal of Controlled Release, 190, 352-370.
Jia, F., Liu, X., Li, L., et al. (2013). Multifunctional nanoparticles for targeted delivery of immune
activating and cancer therapeutic agents. Journal of Controlled Release, 172(3), 1020-1034.
Jiang, W., Gupta, R. K., Deshpande, M. C., et al. (2005). Biodegradable poly(lactic-co-glycolic acid)
microparticles for injectable delivery of vaccine antigens. Advanced Drug Delivery Reviews,
57(3), 391-410.
Jin, R., Lin, B., Li, D., et al. (2014). Superparamagnetic iron oxide nanoparticles for mr imaging and
therapy: Design considerations and clinical applications. Current Opinion in Pharmacology,
18C, 18-27.
Joshi, V. B., Geary, S. M., Carrillo-Conde, B. R., et al. (2013). Characterizing the antitumor response in
mice treated with antigen-loaded polyanhydride microparticles. Acta Biomaterialia, 9(3), 5583-
5589.
Jung, T., Kamm, W., Breitenbach, A., et al. (2001). Tetanus toxoid loaded nanoparticles from
sulfobutylated poly(vinyl alcohol)-graft-poly(lactide-co-glycolide): Evaluation of antibody
response after oral and nasal application in mice. Pharmaceutical Research, 18(3), 352-360.
Kalkanidis, M., Pietersz, G. A., Xiang, S. D., et al. (2006). Methods for nano-particle based vaccine
formulation and evaluation of their immunogenicity. Methods, 40(1), 20-29.
Kamaly, N., Xiao, Z. Y., Valencia, P. M., et al. (2012). Targeted polymeric therapeutic nanoparticles:
Design, development and clinical translation. Chemical Society Reviews, 41(7), 2971-3010.
Kapsenberg, M. L. (2003). Dendritic-cell control of pathogen-driven t-cell polarization. Nature
Reviews Immunology, 3(12), 984-993.
Kassam, F., Enright, K., Dent, R., et al. (2009). Survival outcomes for patients with metastatic triple-
negative breast cancer: Implications for clinical practice and trial design. Clinical Breast Cancer,
9(1), 29-33.
Kasturi, S. P., Skountzou, I., Albrecht, R. A., et al. (2011). Programming the magnitude and
persistence of antibody responses with innate immunity. Nature, 470(7335), 543-547.
Katare, Y. K., Panda, A. K., Lalwani, K., et al. (2003). Potentiation of immune response from polymer-
entrapped antigen: Toward development of single dose tetanus toxoid vaccine. Drug Delivery,
10(4), 231-238.
334 Current Trends and Developments for Nanotechnology in Cancer
Kazzaz, J., Neidleman, J., Singh, M., et al. (2000). Novel anionic microparticles are a potent adjuvant
for the induction of cytotoxic t lymphocytes against recombinant p55 gag from hiv-1. Journal of
Controlled Release, 67(2-3), 347-356.
Kempf, M., Mandal, B., Jilek, S., et al. (2003). Improved stimulation of human dendritic cells by
receptor engagement with surface-modified microparticles. Journal of Drug Targeting, 11(1),
11-18.
Kendall, M. (2006). Engineering of needle-free physical methods to target epidermal cells for DNA
vaccination. Vaccine, 24(21), 4651-4656.
Kersten, G., Hirschberg, H. (2004). Antigen delivery systems. Expert Review of Vaccines, 3(4), 453-
462.
Kievit, F. M., Stephen, Z. R., Veiseh, O., et al. (2012). Targeting of primary breast cancers and
metastases in a transgenic mouse model using rationally designed multifunctional spions. ACS
Nano, 6(3), 2591-2601.
King, J., Waxman, J., Stauss, H. (2008). Advances in tumour immunotherapy. QJM, 101(9), 675-683.
Kinsella, J. M., Ananda, S., Andrew, J. S., et al. (2011). Enhanced magnetic resonance contrast of fe3o4
nanoparticles trapped in a porous silicon nanoparticle host. Advanced Materials, 23(36), H248-253.
Kipp, J. E. (2004). The role of solid nanoparticle technology in the parenteral delivery of poorly water-
soluble drugs. International Jounal of Pharmaceutics, 284(1-2), 109-122.
Kipper, M. J., Shen, E., Determan, A., et al. (2002). Design of an injectable system based on
bioerodible polyanhydride microspheres for sustained drug delivery. Biomaterials, 23(22),
4405-4412.
Klippstein, R., Pozo, D. (2010). Nanotechnology-based manipulation of dendritic cells for enhanced
immunotherapy strategies. Nanomedicine, 6(4), 523-529.
Kokuryo, D., Nakashima, S., Ozaki, F., et al. (2015). Evaluation of thermo-triggered drug release in
intramuscular-transplanted tumors using thermosensitive polymer-modified liposomes and
MRI. Nanomedicine: Nanotechnology, Biology and Medicine, 11(1), 229–238.
Kratz, F., Warnecke, A. (2012). Finding the optimal balance: Challenges of improving conventional
cancer chemotherapy using suitable combinations with nano-sized drug delivery systems.
Journal of Controlled Release, 164(2), 221-235.
Krishnan, S. R., George, S. K. Nanotherapeutics in cancer prevention, diagnosis and treatment. In
Pharmacology and Therapeutics, (Ed. Thatha S. J.). Gowder, ISBN 978-953-51-1620-2, (2014),
DOI: 10.5772/58419.
Kuppusamy, P., Li, H., Ilangovan, G., et al. (2002). Noninvasive imaging of tumor redox status and its
modification by tissue glutathione levels. Cancer Research, 62(1), 307-312.
Lammers, T., Kiessling, F., Hennink, W. E., et al. (2012). Drug targeting to tumors: Principles, pitfalls
and (pre-) clinical progress. Journal of Controlled Release, 161(2), 175-187.
Landry, R., Jacobs, P. M., Davis, R., et al. (2005). Pharmacokinetic study of ferumoxytol: A new
iron replacement therapy in normal subjects and hemodialysis patients. American Journal of
Nephrology, 25(4), 400-410.
Le Garff-Tavernier, M., Herbi, L., de Romeuf, C., et al. (2014). Antibody-dependent cellular
cytotoxicity of the optimized anti-cd20 monoclonal antibody ublituximab on chronic
lymphocytic leukemia cells with the 17p deletion. Leukemia, 28(1), 230-233.
Lee, C. H., Cheng, S. H., Huang, I. P., et al. (2010). Intracellular ph-responsive mesoporous silica
nanoparticles for the controlled release of anticancer chemotherapeutics. Angewandte Chemie
International Edition, 49(44), 8214-8219.
Lee, G. Y., Qian, W. P., Wang, L., et al. (2013a). Theranostic nanoparticles with controlled release of
gemcitabine for targeted therapy and mri of pancreatic cancer. ACS Nano, 7(3), 2078-2089.
Lee, J. H., Chen, K. J., Noh, S. H., et al. (2013b). On-demand drug release system for in vivo cancer
treatment through self-assembled magnetic nanoparticles. Angewandte Chemie International
Edition, 52(16), 4384-4388.
References 335
Lee, S., Margolin, K. (2011). Cytokines in cancer immunotherapy. Cancers (Basel), 3(4), 3856-3893.
Lesterhuis, W. J., de Vries, I. J., Adema, G. J., et al. (2004). Dendritic cell-based vaccines in cancer
immunotherapy: An update on clinical and immunological results. Annals of Oncology, 15(S4),
iv145-151.
Liao, L., Liu, J., Dreaden, E. C., et al. (2014). A convergent synthetic platform for single-nanoparticle
combination cancer therapy: Ratiometric loading and controlled release of cisplatin,
doxorubicin, and camptothecin. Journal of the American Chemical Society, 136(16), 5896-5899.
Ling, D., Park, W., Park, S. J., et al. (2014). Multifunctional tumor ph-sensitive self-assembled
nanoparticles for bimodal imaging and treatment of resistant heterogeneous tumors. Journal of
the American Chemical Society, 136(15), 5647-5655.
Liu, C., Noorchashm, H., Sutter, J. A., et al. (2007). B lymphocyte-directed immunotherapy promotes
long-term islet allograft survival in nonhuman primates. Nature Medicine, 13(11), 1295-1298.
Liu, D., Bimbo, L. M., Mäkilä, E., et al. (2013a). Co-delivery of a hydrophobic small molecule and a
hydrophilic peptide by porous silicon nanoparticles. Journal of Controlled Release, 170(2), 268-
278.
Liu, D., Mäkilä, E., Zhang, H., et al. (2013b). Nanostructured porous silicon-solid lipid
nanocomposite: Towards enhanced cytocompatibility and stability, reduced cellular
association, and prolonged drug release. Advanced Functional Materials, 23(15), 1893-1902.
Liu, D. F., Zhang, H. B., Mäkilä, E., et al. (2015). Microfluidic assisted one-step fabrication of porous
silicon/acetalated dextran nanocomposites for precisely controlled combination chemotherapy.
Biomaterials, 39, 249-259.
Liu, Q., Zhang, J., Sun, W., et al. (2012). Delivering hydrophilic and hydrophobic chemotherapeutics
simultaneously by magnetic mesoporous silica nanoparticles to inhibit cancer cells.
International Journal of Nanomedicine, 7, 999-1013.
Liu, R., Liao, P., Liu, J., et al. (2011). Responsive polymer-coated mesoporous silica as a ph-sensitive
nanocarrier for controlled release. Langmuir, 27(6), 3095-3099.
Liu, Y., Fang, J., Joo, K. I., (2014). Codelivery of chemotherapeutics via crosslinked multilamellar
liposomal vesicles to overcome multidrug resistance in tumor. PLoS One, 9(10), e110611.
Livasy, C. A., Karaca, G., Nanda, R., et al. (2006). Phenotypic evaluation of the basal-like subtype of
invasive breast carcinoma. Modern Pathology, 19(2), 264-271.
Lohr, J. M., Haas, S. L., Bechstein, W. O., et al. (2012). Cationic liposomal paclitaxel plus gemcitabine
or gemcitabine alone in patients with advanced pancreatic cancer: A randomized controlled
phase II trial. Annals of Oncology, 23(5), 1214-1222.
Lowin, B., Peitsch, M. C., Tschopp, J. (1995). Perforin and granzymes: Crucial effector molecules
in cytolytic t lymphocyte and natural killer cell-mediated cytotoxicity. Current Topics in
Microbiology and Immunology, 198, 1-24.
Ma, M., Zhang, Y., Gong, H., et al. (2013). Silica-coated magnetite nanoparticles labeled by
nimotuzumab, a humanised monoclonal antibody to epidermal growth factor receptor:
Preparations, specific targeting and bioimaging. Journal of Nanoscience and Nanotechnology,
13(10), 6541-6545.
Ma, W., Chen, M., Kaushal, S., et al. (2012). Plga nanoparticle-mediated delivery of tumor antigenic
peptides elicits effective immune responses. International Journal of Nanomedicine, 7, 1475-
1487.
Ma, Y., Zheng, Y., Zeng, X., et al. (2011). Novel docetaxel-loaded nanoparticles based on pcl-tween 80
copolymer for cancer treatment. International Journal of Nanomedicine, 6, 2679-2688.
MacEwan, S. R., Callahan, D. J., Chilkoti, A. (2010). Stimulus-responsive macromolecules and
nanoparticles for cancer drug delivery. Nanomedicine (Lond), 5(5), 793-806.
Maeda, H. (2001). The enhanced permeability and retention (EPR) effect in tumor vasculature: The
key role of tumor-selective macromolecular drug targeting. Advances in Enzyme Regulation, 41,
189-207.
336 Current Trends and Developments for Nanotechnology in Cancer
Maeda, H., Wu, J., Sawa, T., et al. (2000). Tumor vascular permeability and the epr effect in
macromolecular therapeutics: A review. Journal of Controlled Release, 65(1-2), 271-284.
Mäkilä, E., Ferreira, M. P., Kivela, H., et al. (2014). Confinement effects on drugs in thermally
hydrocarbonized porous silicon. Langmuir, 30(8), 2196-2205.
Mallapragada, S. K., Narasimhan, B. (2008). Immunomodulatory biomaterials. International Journal
of Pharmaceutics, 364(2), 265-271.
Manolova, V., Flace, A., Bauer, M., et al. (2008). Nanoparticles target distinct dendritic cell
populations according to their size. European Journal of Immunology, 38(5), 1404-1413.
Mansour, A. M., Drevs, J., Esser, N., et al. (2003). A new approach for the treatment of malignant
melanoma: Enhanced antitumor efficacy of an albumin-binding doxorubicin prodrug that is
cleaved by matrix metalloproteinase 2. Cancer Research, 63(14), 4062-4066.
Markman, J. L., Rekechenetskiy, A., Holler, E., et al. (2013). Nanomedicine therapeutic approaches to
overcome cancer drug resistance. Advanced Drug Delivery Reviews, 65(13-14), 1866-1879.
Marrack, P., McKee, A. S., Munks, M. W. (2009). Towards an understanding of the adjuvant action of
aluminium. Nature Reviews Immunology, 9(4), 287-293.
Maruyama, K. (2011). Intracellular targeting delivery of liposomal drugs to solid tumors based on epr
effects. Advanced Drug Delivery Reviews, 63(3), 161-169.
Matsumura, Y., Maeda, H. (1986). A new concept for macromolecular therapeutics in cancer
chemotherapy: Mechanism of tumoritropic accumulation of proteins and the antitumor agent
smancs. Cancer Research, 46(12 Pt 1), 6387-6392.
Mayer, L. D., Harasym, T. O., Tardi, P. G., et al. (2006). Ratiometric dosing of anticancer drug
combinations: Controlling drug ratios after systemic administration regulates therapeutic
activity in tumor-bearing mice. Molecular Cancer Therapeutics, 5(7), 1854-1863.
Melancon, M. P., Lu, W., Zhong, M., et al. (2011). Targeted multifunctional gold-based nanoshells
for magnetic resonance-guided laser ablation of head and neck cancer. Biomaterials, 32(30),
7600-7608.
Meng, H., Xue, M., Xia, T., et al. (2010). Autonomous in vitro anticancer drug release from
mesoporous silica nanoparticles by ph-sensitive nanovalves. Journal of the American Chemical
Society, 132(36), 12690-12697.
Miele, E., Spinelli, G. P., Miele, E., et al. (2012). Nanoparticle-based delivery of small interfering rna:
Challenges for cancer therapy. International Journal of Nanomedicine, 7, 3637-3657.
Misra, R., Acharya, S., Sahoo, S. K. (2010). Cancer nanotechnology: Application of nanotechnology
in cancer therapy. Drug Discovery Today, 15(19-20), 842-850.
Mody, H. R. (2011). Cancer nanotechnology: Recent trends and developments. Internet Journal of
Medical Update - EJOURNAL, 6(1).
Moghimi, S. M., Hunter, A. C., Murray, J. C. (2005). Nanomedicine: Current status and future
prospects. FASEB J, 19(3), 311-330.
Monjanel, H., Deville, L., Ram-Wolff, C., et al. (2014). Brentuximab vedotin in heavily treated hodgkin
and anaplastic large-cell lymphoma, a single centre study on 45 patients. British Journal of
Haematology, 166(2), 306-308.
Monjazeb, A. M., Hsiao, H. H., Sckisel, G. D., et al. (2012). The role of antigen-specific and non-
specific immunotherapy in the treatment of cancer. Journal of Immunotoxicology, 9(3), 248-258.
Moon, J. J., Suh, H., Li, A. V., et al. (2012). Enhancing humoral responses to a malaria antigen
with nanoparticle vaccines that expand tfh cells and promote germinal center induction.
Proceedings of the National Academy of Sciences, 109(4), 1080-1085.
Nakayama, M., Akimoto, J., Okano, T. (2014). Polymeric micelles with stimuli-triggering systems for
advanced cancer drug targeting. Journal of Drug Targeting, 22(7), 584-599.
Nandakumar, K. S., Shakya, A. K. ( 2012). Polymers as immunological adjuvants: An update on recent
developments. Journal of Bioscience and Biotechnology, 1(3), 199-210.
References 337
Poon, Z., Lee, J. B., Morton, S. W., et al. (2011). Controlling in vivo stability and biodistribution in
electrostatically assembled nanoparticles for systemic delivery. Nano Letters, 11(5), 2096-2103.
Power, S., Slattery, M. M., Lee, M. J. (2011). Nanotechnology and its relationship to interventional
radiology. Part i: Imaging. CardioVascular and Interventional Radiology, 34(2), 221-226.
Prabhakar, U., Maeda, H., Jain, R. K., et al. (2013). Challenges and key considerations of the
enhanced permeability and retention effect for nanomedicine drug delivery in oncology. Cancer
Research, 73(8), 2412-2417.
Prang, N., Preithner, S., Brischwein, K., et al. (2005). Cellular and complement-dependent
cytotoxicity of ep-cam-specific monoclonal antibody mt201 against breast cancer cell lines.
British Journal of Cancer, 92(2), 342-349.
Primard, C., Rochereau, N., Luciani, E., et al. (2010). Traffic of poly(lactic acid) nanoparticulate
vaccine vehicle from intestinal mucus to sub-epithelial immune competent cells. Biomaterials,
31(23), 6060-6068.
Ramakrishna, V., Treml, J. F., Vitale, L., et al. (2004). Mannose receptor targeting of tumor antigen
pmel17 to human dendritic cells directs anti-melanoma t cell responses via multiple hla
molecules. Journal of Immunology, 172(5), 2845-2852.
Rappuoli, R., Aderem, A. (2011). A 2020 vision for vaccines against hiv, tuberculosis and malaria.
Nature, 473(7348), 463-469.
Reddy, A. K., Rathinaraj, B. S., Prathyusha, P., et al. (2011). Emerging trends of nanotechnology in
cancer therapy. International Journal of Pharmaceutical & Biological Archives, 2(1).
Reddy, S. T., van der Vlies, A. J., Simeoni, E., et al. (2007). Exploiting lymphatic transport and
complement activation in nanoparticle vaccines. Nature Biotechnology, 25(10), 1159-1164.
Reichert, S., Welker, P., Calderon, M., et al. (2011). Size-dependant cellular uptake of dendritic
polyglycerol. Small, 7(6), 820-829.
Reischl, D., Zimmer, A. (2009). Drug delivery of sirna therapeutics: Potentials and limits of
nanosystems. Nanomedicine, 5(1), 8-20.
Remant, B. K., Chandrashekaran, V., Cheng, B., et al. (2014). Redox potential ultrasensitive
nanoparticle for the targeted delivery of camptothecin to her2-positive cancer cells. Molecular
Pharmaceutics, 11(6), 1897-1905.
Reyes-Aldasoro, C. C., Wilson, I., Prise, V. E., et al. (2008). Estimation of apparent tumor vascular
permeability from multiphoton fluorescence microscopic images of p22 rat sarcomas in vivo.
Microcirculation, 15(1), 65-79.
Rosen, J. E., Chan, L., Shieh, D.-B., et al. (2012). Iron oxide nanoparticles for targeted cancer imaging
and diagnostics. Nanomedicine: Nanotechnology, Biology and Medicine, 8(3), 275-290.
Rosen JE, Y. S., Meerasa A., Verma M., Gu F.X., et al. (2011). Nanotechnology and diagnostic imaging:
New advances in contrast agent technology. Journal of Nanomedicine & Nanotechnology, 2(5),
115. DOI: 10.4172/2157-7439.1000115.
Rosenberg, S. A. (1988). The development of new immunotherapies for the treatment of cancer using
interleukin-2. A review. Annals of Surgery, 208(2), 121-135.
Roy, A., Singh, M. S., Upadhyay, P., et al. (2013). Nanoparticle mediated co-delivery of paclitaxel
and a tlr-4 agonist results in tumor regression and enhanced immune response in the tumor
microenvironment of a mouse model. International Journal of Pharmaceutics, 445(1-2), 171-180.
Saad, M., Garbuzenko, O. B., Minko, T. (2008). Co-delivery of sirna and an anticancer drug for
treatment of multidrug-resistant cancer. Nanomedicine, 3(6), 761-776.
Salonen, J., Kaukonen, A. M., Hirvonen, J., et al. (2008). Mesoporous silicon in drug delivery
applications. Journal of Pharmaceutical Sciences, 97(2), 632-653.
Sangha, R., Butts, C. (2007). L-blp25: A peptide vaccine strategy in non small cell lung cancer.
Clinical Cancer Research, 13(15 Pt 2), s4652-4654.
Sanna, V., Pala, N., Sechi, M. (2014). Targeted therapy using nanotechnology: Focus on cancer.
International Journal of Nanomedicine, 9, 467-483.
References 339
Santos, H. A., Bimbo, L. M., Herranz, B., et al. (2013a). Nanostructured porous silicon in preclinical
imaging: Moving from bench to bedside. Journal of Materials Research, 28(02), 152-164.
Santos, H. A., Mäkilä, E., Airaksinen, Anu J., et al. (2014). Porous silicon nanoparticles for
nanomedicine: Preparation and biomedical applications. Nanomedicine, 9(4), 535-554.
Santos, H. A., Peltonen, L., Limnell, T., et al. (2013). Mesoporous materials and nanocrystals for
enhancing the dissolution behavior of poorly water-soluble drugs. Current Pharmaceutical
Biotechnology, 14(10), 926-938.
Sarparanta, M. P., Bimbo, L. M., Mäkilä, E. M., et al. (2012). The mucoadhesive and gastroretentive
properties of hydrophobin-coated porous silicon nanoparticle oral drug delivery systems.
Biomaterials, 33(11), 3353-3362.
Satpathy, M., Zielinski, R., Lyakhov, I., et al. (2015). Optical imaging of ovarian cancer using her-2
affibody conjugated nanoparticles. Methods in Molecular Biology, 1219, 171-185.
Sawant, R. R., Torchilin, V. P. (2012). Challenges in development of targeted liposomal therapeutics.
The AAPS Journal, 14(2), 303-315.
Schmitt-Sody, M., Strieth, S., Krasnici, S., et al. (2003). Neovascular targeting therapy: Paclitaxel
encapsulated in cationic liposomes improves antitumoral efficacy. Clinical Cancer Research,
9(6), 2335-2341.
Scott, A. M., Wolchok, J. D., Old, L. J. (2012). Antibody therapy of cancer. Nature Reviews Cancer,
12(4), 278-287.
Shahbazi, M.-A., Herranz, B., Santos, H. A. (2012). Nanostructured porous si-based nanoparticles for
targeted drug delivery. Biomatter, 2(4), 296-312.
Sharma, S., Mukkur, T. K., Benson, H. A., et al. (2009). Pharmaceutical aspects of intranasal delivery
of vaccines using particulate systems. Journal of Pharmaceutical Sciences, 98(3), 812-843.
Shi, C., Guo, X., Qu, Q., et al. (2014). Actively targeted delivery of anticancer drug to tumor cells by
redox-responsive star-shaped micelles. Biomaterials, 35(30), 8711-8722.
Shi, H., Huang, Y., Zhou, H., et al. (2007). Nucleolin is a receptor that mediates antiangiogenic and
antitumor activity of endostatin. Blood, 110(8), 2899-2906.
Shi, J. J., Xiao, Z. Y., Kamaly, N., et al. (2011). Self-assembled targeted nanoparticles: Evolution of
technologies and bench to bedside translation. Accounts of Chemical Research, 44(10), 1123-
1134.
Shima, F., Uto, T., Akagi, T., et al. (2013). Size effect of amphiphilic poly(gamma-glutamic acid)
nanoparticles on cellular uptake and maturation of dendritic cells in vivo. Acta Biomaterialia,
9(11), 8894-8901.
Shukla-Dave, A., Hricak, H. (2014). Role of mri in prostate cancer detection. NMR in Biomedicine,
27(1), 16-24.
Siegel, R., Ma, J., Zou, Z., et al. (2014). Cancer statistics, 2014. CA: A Cancer Journal for Clinicians,
64(1), 9-29.
Siegel, R., Ward, E., Brawley, O., et al. (2011). Cancer statistics, 2011: The impact of eliminating
socioeconomic and racial disparities on premature cancer deaths. CA: A Cancer Journal for
Clinicians, 61(4), 212-236.
Signori, E., Iurescia, S., Massi, E., et al. (2010). DNA vaccination strategies for anti-tumour effective
gene therapy protocols. Cancer Immunology, Immunotherapy, 59(10), 1583-1591.
Silva, J. M., Videira, M., Gaspar, R., et al. (2013). Immune system targeting by biodegradable
nanoparticles for cancer vaccines. Journal of Controlled Release, 168(2), 179-199.
Singh, M., Chakrapani, A., O’Hagan, D. (2007). Nanoparticles and microparticles as vaccine-delivery
systems. Expert Review of Vaccines, 6(5), 797-808.
Sloat, B. R., Sandoval, M. A., Hau, A. M., et al. (2010). Strong antibody responses induced by protein
antigens conjugated onto the surface of lecithin-based nanoparticles. Journal of Controlled
Release, 141(1), 93-100.
340 Current Trends and Developments for Nanotechnology in Cancer
Sorkin, A., von Zastrow, M. (2002). Signal transduction and endocytosis: Close encounters of many
kinds. Nature Reviews Molecular Cell Biology, 3(8), 600-614.
Strebhardt, K., Ullrich, A. (2008). Paul ehrlich’s magic bullet concept: 100 years of progress. Nature
Reviews Cancer, 8(6), 473-480.
Strieth, S., Eichhorn, M. E., Werner, A., et al. (2008). Paclitaxel encapsulated in cationic liposomes
increases tumor microvessel leakiness and improves therapeutic efficacy in combination with
cisplatin. Clinical Cancer Research, 14(14), 4603-4611.
Subbotin, V. M. (2014). Dendritic cell-based cancer immunotherapy: The stagnant approach and a
theoretical solution. Drug Discovery Today, 19(7), 834-837.
Sultana, S., Khan, M. R., Kumar, M., et al. (2013). Nanoparticles-mediated drug delivery approaches
for cancer targeting: A review. Journal of Drug Targeting, 21(2), 107-125.
Sumer, B., Gao, J. (2008). Theranostic nanomedicine for cancer. Nanomedicine (Lond), 3(2), 137-140.
Sun, T.-M., Du, J.-Z., Yao, Y.-D., et al. (2011). Simultaneous delivery of sirna and paclitaxel via a “two-
in-one” micelleplex promotes synergistic tumor suppression. ACS Nano, 5(2), 1483-1494.
Sun, T., Zhang, Y. S., Pang, B., et al. (2014). Engineered nanoparticles for drug delivery in cancer
therapy. Angewandte Chemie International Edition, 53(46), 12320-12364.
Sutradhar, K. B., Amin, M. L. (2014). Nanotechnology in cancer drug delivery and selective targeting.
ISRN Nanotechnology, 2014(1), 1-12.
Tacken, P. J., Zeelenberg, I. S., Cruz, L. J., et al. (2011). Targeted delivery of tlr ligands to human and
mouse dendritic cells strongly enhances adjuvanticity. Blood, 118(26), 6836-6844.
Taniguchi, H., Imai, K. (2012). [An overview of antibody therapy against cancer]. Nihon Rinsho,
70(12), 2087-2092.
Tao, L., Zhang, K., Sun, Y., et al. (2012). Anti-epithelial cell adhesion molecule monoclonal antibody
conjugated fluorescent nanoparticle biosensor for sensitive detection of colon cancer cells.
Biosensors and Bioelectronics, 35(1), 186-192.
Tardi, P., Johnstone, S., Harasym, N., et al. (2009). In vivo maintenance of synergistic
cytarabine:Daunorubicin ratios greatly enhances therapeutic efficacy. Leukemia Research,
33(1), 129-139.
Tarn, D., Ashley, C. E., Xue, M., et al. (2013). Mesoporous silica nanoparticle nanocarriers:
Biofunctionality and biocompatibility. Accounts of Chemical Research, 46(3), 792-801.
Tassa, C., Shaw, S. Y., Weissleder, R. (2011). Dextran-coated iron oxide nanoparticles: A versatile
platform for targeted molecular imaging, molecular diagnostics, and therapy. Accounts of
Chemical Research, 44(10), 842-852.
Taurin, S., Nehoff, H., Greish, K. (2012). Anticancer nanomedicine and tumor vascular permeability;
where is the missing link? Journal of Controlled Release, 164(3), 265-275.
Tomic, S., Ethokic, J., Vasilijic, S., et al. (2014). Size-dependent effects of gold nanoparticles uptake
on maturation and antitumor functions of human dendritic cells in vitro. PLoS One, 9(5),
e96584.
Torchilin, V. (2011). Tumor delivery of macromolecular drugs based on the epr effect. Advanced Drug
Delivery Reviews, 63(3), 131-135.
Torchilin, V. P. (2006). Multifunctional nanocarriers. Advanced Drug Delivery Reviews, 58(14), 1532-
1555.
Toy, R., Bauer, L., Hoimes, C., et al. (2014). Targeted nanotechnology for cancer imaging. Advanced
Drug Delivery Reviews, 76C, 79-97.
Trombetta, E. S., Mellman, I. (2005). Cell biology of antigen processing in vitro and in vivo. Annual
Review of Immunology, 23, 975-1028.
Valdez, R. A., McGuire, T. C., Brown, W. C., et al. (2000). An in vivo model to investigate lymphocyte-
mediated immunity during acute hemoparasitic infections. Use of a monoclonal antibody to
selectively deplete cd4+ t lymphocytes from thymectomized calves. Annals of the New York
Academy of Sciences, 916, 233-236.
References 341
van den Berg, J. H., Oosterhuis, K., Hennink, W. E., et al. (2010). Shielding the cationic charge
of nanoparticle-formulated dermal DNA vaccines is essential for antigen expression and
immunogenicity. Journal of Controlled Release, 141(2), 234-240.
Varallyay, P., Nesbit, G., Muldoon, L. L., (2002). Comparison of two superparamagnetic viral-sized
iron oxide particles ferumoxides and ferumoxtran-10 with a gadolinium chelate in imaging
intracranial tumors. American Journal of Neuroradiology, 23(4), 510-519.
Varga, N., Sutkeviciute, I., Guzzi, C., et al. (2013). Selective targeting of dendritic cell-specific
intercellular adhesion molecule-3-grabbing nonintegrin (dc-sign) with mannose-based
glycomimetics: Synthesis and interaction studies of bis(benzylamide) derivatives of a
pseudomannobioside. Chemistry, 19(15), 4786-4797.
Vivek, R., Thangam, R., NipunBabu, V., et al. (2014). Multifunctional her2-antibody conjugated
polymeric nanocarrier-based drug delivery system for multi-drug-resistant breast cancer
therapy. ACS Applied Materials & Interfaces, 6(9), 6469-6480.
Wang, M., Thanou, M. (2010). Targeting nanoparticles to cancer. Pharmacology Research, 62(2),
90-99.
Wang, R., Billone, P. S., Mullett, W. M. (2013). Nanomedicine in action: An overview of cancer
nanomedicine on the market and in clinical trials. Journal of Nanomaterials, 2013, 629681.
Wang, Y. X., Hussain, S. M., Krestin, G. P. (2001). Superparamagnetic iron oxide contrast agents:
Physicochemical characteristics and applications in mr imaging. European Radiology, 11(11),
2319-2331.
Weber, C., Drogoz, A., David, L., et al. (2010). Polysaccharide-based vaccine delivery systems:
Macromolecular assembly, interactions with antigen presenting cells, and in vivo
immunomonitoring. Journal of Biomedical Materials Research Part A, 93(4), 1322-1334.
Weiner, L. M., Adams, G. P. (2000). New approaches to antibody therapy. Oncogene, 19(53), 6144-
6151.
Whitehead, K. A., Langer, R., Anderson, D. G. (2009). Knocking down barriers: Advances in sirna
delivery. Nature Reviews Drug Discov, 8(2), 129-138.
WHO. (2014). Cancer. World Health Organization Retrieved 27th October, 2014, from http://www.
who.int/topics/cancer/en/.
Wong, C., Stylianopoulos, T., Cui, J., et al. (2011). Multistage nanoparticle delivery system for deep
penetration into tumor tissue. Proceedings of the National Academy of Sciences, 108(6), 2426-
2431.
Wu, K. Y., Wu, M., Fu, M. L., et al. (2006). A novel chitosan cpg nanoparticle regulates cellular and
humoral immunity of mice. Biomedical and Environmental Sciences, 19(2), 87-95.
Xia, T., Kovochich, M., Liong, M., et al. (2009). Polyethyleneimine coating enhances the cellular
uptake of mesoporous silica nanoparticles and allows safe delivery of sirna and DNA
constructs. ACS Nano, 3(10), 3273-3286.
Xiang, S. D., Wilson, K., Day, S., et al. (2013). Methods of effective conjugation of antigens to
nanoparticles as non-inflammatory vaccine carriers. Methods, 60(3), 232-241.
Xie, J., Lee, S., Chen, X. (2010). Nanoparticle-based theranostic agents. Advanced Drug Delivery
Reviews, 62(11), 1064-1079.
Xie, R., Dong, L., Huang, R., et al. (2014). Targeted imaging and proteomic analysis of tumor-
associated glycans in living animals. Angewandte Chemie International Edition, 53(51), 14082-
14086.
Yanes, R. E., Tamanoi, F. (2012). Development of mesoporous silica nanomaterials as a vehicle for
anticancer drug delivery. Therapeutic Delivery, 3(3), 389-404.
Yang, K. N., Zhang, C. Q., Wang, W., et al. (2014). Ph-responsive mesoporous silica nanoparticles
employed in controlled drug delivery systems for cancer treatment. Cancer Biology & Medicine,
11(1), 34-43.
342 Current Trends and Developments for Nanotechnology in Cancer
Yang, S., Li, N., Chen, D., et al. (2013). Visible-light degradable polymer coated hollow mesoporous
silica nanoparticles for controlled drug release and cell imaging. Journal of Materials Chemistry
B, 1(36), 4628-4636.
Yoshida, T., Mei, H., Dorner, T., et al. (2010). Memory b and memory plasma cells. Immunology
Reviews, 237(1), 117-139.
Yotsumoto, S., Aramaki, Y., Kakiuchi, T., et al. (2004). Induction of antigen-dependent interleukin-12
production by negatively charged liposomes encapsulating antigens. Vaccine, 22(25-26), 3503-
3509.
Yu, B., Tai, H. C., Xue, W., et al. (2010). Receptor-targeted nanocarriers for therapeutic delivery to
cancer. Molecular Membrane Biology, 27(7), 286-298.
Zhang, H., Wang, G., Yang, H. (2011). Drug delivery systems for differential release in combination
therapy. Expert Opinion on Drug Delivery, 8(2), 171-190.
Zhang, L., Zhou, H., Belzile, O., et al. (2014). Phosphatidylserine-targeted bimodal liposomal
nanoparticles for in vivo imaging of breast cancer in mice. Journal of Controlled Release, 183(0),
114-123.
Zhang, Q., Neoh, K. G., Xu, L., et al. (2014b). Functionalized mesoporous silica nanoparticles with
mucoadhesive and sustained drug release properties for potential bladder cancer therapy.
Langmuir, 30(21), 6151-6161.
Zhang, S., Gong, M., Zhang, D., et al. (2014c). Thiol-peg-carboxyl-stabilized fe(2)o (3)/Au
nanoparticles targeted to cd105: Synthesis, characterization and application in mr imaging of
tumor angiogenesis. European Journal of Radiology, 83(7), 1190-1198.
Zhao, Y. L., Li, Z., Kabehie, S., et al. (2010). Ph-operated nanopistons on the surfaces of mesoporous
silica nanoparticles. Journal of the American Chemical Society, 132(37), 13016-13025.
Zhu, L., Kate, P., Torchilin, V. P. (2012). Matrix metalloprotease 2-responsive multifunctional
liposomal nanocarrier for enhanced tumor targeting. ACS Nano, 6(4), 3491-3498.
Zhu, L., Torchilin, V. P. (2013). Stimulus-responsive nanopreparations for tumor targeting. Integrative
Biology (Camb), 5(1), 96-107.
Zilberman, M., Grinberg, O. (2008). Hrp-loaded bioresorbable microspheres: Effect of copolymer
composition and molecular weight on microstructure and release profile. Journal of
Biomaterials Applications, 22(5), 391-407.
Zolata, H., Abbasi Davani, F., Afarideh, H. (2014). Synthesis, characterization and theranostic
evaluation of indium-111 labeled multifunctional superparamagnetic iron oxide nanoparticles.
Nuclear Medicine and Biology, DOI: 10.1016/j.nucmedbio.2014.09.007.
Index
E V, 5, 8, 10, 13, 17-20, 25, 29, 32, 43-44, 46, formal charge 13
48, 53, 56-59, 112-116, 118, 125-132, 140- fructose 27, 134, 137, 141, 146
148, 161, 166-169, 171-176, 180, 182, 184, fumaric acid 8
186, 192, 194, 200-204, 216-228, 239, 243- Functional Groups V, 2-8, 10, 12, 14, 16, 18, 20,
244, 246-249, 251, 257, 261, 263-265, 267, 22, 24, 26, 28, 30, 32, 34, 36, 38, 40, 42,
271-272, 278, 286-289, 291, 295, 304-305, 182, 204
310, 314, 327-332, 334-341 GABAergic VII, 149-150, 153, 159-164, 168, 171,
E-isomer 8 173, 238
electron density 5, 13, 17 gas chromatography (GC) 64
electron donating 17 gemcitabine 188-190, 226, 334-335
electronegativity 5 geometry 4, 32, 200, 217, 253, 255, 328
electron-poor 7 Gliotransmitters VII, 164, 166
electron rich 13 global burden of disease 119, 129
electrons 5-7, 13-17, 20 glucocorticoid 130, 136, 138-139
electron withdrawing 17, 184 glucocorticoids 120, 124, 131, 134, 136, 138
electrophile species 10 glutamate VII, VIII, 129, 154-164, 166, 168-174,
elderly 80, 133, 142, 146, 166 181, 229-243
Electrophiles V, 5-7, 59 guidelines for validation of bioanalytical
electrophilic V, VI, 13-14, 16-17, 20, 22-23, 42, methods 84
44-45, 48, 52-56, 58 11β-hydroxysteroid dehydrogenase 136, 138,
Electrophilic Addition Reactions V, 13, 22 145
electrophilic aromatic substitution V, 16-17, 42 heart rate 120, 134, 136, 143-144
Eliminations Reactions V, 18 hepatic function 114
empty orbitals 15 hepatitis C 124, 127-128, 246
Enantiomerism V, 9 heptapeptide 257, 261, 263, 269, 272
enantiomers 9-11 heteroaromatic 16
endoperoxide 52-54, 57-58, 280-281, 287 heterolytically 11
endoperoxides IX, 52, 280, 285-286 high performance liquid chromatography (HPLC)
endothelial dysfunction 133-134 64
enolate 26-27 high-throughput screening (HTS) 66, 246
Enzymatic resolution 10 Hippocrates 119
enzyme 10, 13-14, 24-25, 27, 30, 33, 35-39, homeostasis 126, 132-134, 137-140, 143-145,
44-46, 48, 52-53, 56, 69, 108, 115, 151, 160, 147, 155, 161, 170, 313
186, 188-189, 208, 218, 222, 224-225, 231, Huckel rule 15
234, 242-243, 246, 297, 335 Huntington’s disease VIII, 234, 238, 243
epinephrine (EPI) 122 Hybrid Drugs VI, 51, 53, 55
ester 10, 24, 26-27, 117, 184, 186-192, 204, 218, hybrid orbitals 4
220, 222-223, 226-227 hydride 18, 21, 30, 39
esterification 24, 33, 189-190, 192, 227 hydrogen carbonate 6-7, 139
ethers 5 hydrogen ions 5
European Medicine Agency 62, 79 hydroxyl 17, 23-24, 29-30, 33, 35-36, 182, 189
excitotoxic VIII, 156-158, 162-163, 167-169, 230, hypercortisolism 123, 131
234-235, 237-241 hypertension 133-134, 138, 140-144, 146-147
excretion 62, 65, 69, 79, 133, 137-139, 142, imines 20
245, 272, 282 immediate-early gene 121
FAD VI, 33, 40-41, 237 immune cells 122-123, 320, 322
fatty acids V, 2, 33, 35, 181 immune system VII, 122, 124, 128, 197-198,
fear-related behavior 120 236, 292, 316-317, 319, 329, 339
fight, flight, freeze or fawn response 120 induction 69, 75, 115, 117, 134, 145, 241, 319-
Food and Drug Administration 62, 75, 105, 236 320, 333, 336, 341
346 Index
inductive effect 11 lipid 133, 136-137, 141, 144, 200, 218, 221, 227,
inhibition VI, 36, 43, 45, 50, 52-53, 57-59, 69, 247, 295, 335
115-117, 120-121, 129, 135, 137-140, 151, liquid chromatography (LC) 64
156, 161-164, 169-170, 172, 175, 218, 225- liquid-liquid extraction (LLE) 112
226, 244, 259-260, 262, 321, 329-330 liver 44, 53, 66, 111, 132, 134, 137-138, 144,
in silico 64, 66, 115 189, 211, 218, 225, 247-248, 255, 265, 279,
intermediate 11, 13, 15-16, 19-20, 27, 53, 84, 291, 331
180, 191, 215 lone pair electrons 15
internal standard (IS) 87 lower limit of quantification (LLOQ) 84
International Conference on Harmonisation lymphocyte 216, 335, 337
(ICH) 79 macrophage 123, 321
Intramolecular Addition V, 14 magnesium VII, 132-137, 139-147, 154, 231-233,
in vitro VI, 38, 51-53, 55, 64, 66, 70-76, 80-81, 235, 238
84, 88, 106-113, 116-117, 125, 130, 189, 191, Major depressive disorder (MDD) 119
193, 196-197, 206, 211, 219, 223, 225, 241- marker 134
242, 246, 249, 265, 267, 273, 279, 282, market VIII, 63-64, 80, 211, 213-214, 216, 223,
288-289, 295, 302, 310, 314, 320, 324, 227, 270, 274, 281, 290, 294, 341
329-331, 336, 340 mass spectrometry (MS) 64
in vivo VI, 53, 64, 66, 75-78, 81, 84, 87-88, mefloquine 278-279, 281, 287-288
106-111, 113-114, 116-118, 170, 173, 189, melanoma 124, 130, 224, 336-337
191-194, 197, 206, 209, 211, 216-217, 242, membrane 63, 66, 74, 83, 107, 109-110, 135,
246, 263, 266, 282, 288, 294, 301-303, 151, 154, 156-157, 159-163, 192-194, 197,
306, 310, 313, 320, 324-326, 329-331, 334, 199-203, 215, 217-218, 221-224, 229-230,
337-342 232-233, 235, 238, 241-243, 247-248, 250,
isopenicillin-N synthase 37-38 259-260, 262, 272, 280, 286, 295, 305,
IUPAC 2, 42 314, 341
imaging X, 170, 175-176, 214, 290, 292-294, meso structures 9
296, 315, 321, 324-335, 337-342 metabolic acidosis 139, 141
imidazolidin-4-one 195-197, 217-218, 220, 227 metabolic stability 107, 111, 113, 226, 245-246,
immune surveillance 122 251, 265-267, 269, 271, 282
immunomediators 122, 124 metabolism V, 2, 14, 33, 39, 62, 65, 69-74, 79,
immunotherapy X, 290, 294, 316-321, 323, 108, 112-113, 115-118, 125, 130, 135-136,
328-332, 334-337, 339-340 138-139, 141-144, 146-147, 149, 160, 168,
inflammation 122, 124, 130-132, 136, 138-139, 170, 180, 185, 191, 215, 222-223, 228, 245,
146, 274 247-249, 264, 270-273, 280, 282, 284-285,
insulin 133-139, 141-147 288, 301, 332
interleukin-6 (IL-6) 122, 136, 319 metabolites 63-64, 66, 76, 79-81, 84, 87-88,
kainate receptors VIII, 154, 230-232 106, 108-116, 181, 185, 189, 192, 194, 207,
β-keto ester 27 209, 215, 257, 273, 276-277, 279-280, 285,
ketones 5, 20, 22, 25, 52, 55, 195 288
kynurenine pathways 125 metabolites in safety testing (MIST) 108
β-lactam 35-36, 264 mice 50, 58, 76, 110-111, 120, 130, 137, 144,
leaving group 11, 17-19, 23-24, 29, 33, 35, 192 147, 167, 170, 174, 192, 195-196, 232, 243,
leptin 134, 138, 144 263, 271-273, 287-288, 307, 319-320, 323,
leucocyte 123 326, 329, 333, 336, 341-342
Lewis acid 6-7 Michael acceptors VI, 45-46, 48-49, 53, 57-59
Lewis base 6-7 microglia 124, 164, 174
limbic system 120, 152 microsomal stability 66, 76, 114
linalyl diphosphate 14 microsomes 108, 111, 265, 273
linear 106-107, 135, 201-203, 250 molecular orbitals 4
Index 347
molecule 2, 5-6, 9, 13, 18, 20, 23, 26, 41, 54, 198-204, 206-209, 211-219, 221-228, 237,
87, 154, 160, 180-181, 195, 241, 271, 299, 246, 249-261, 263, 265, 267, 269-274, 314,
301, 315, 317, 326, 335, 340 320, 331, 335, 338
Monoaminergic VII, 148-149, 151, 153, 165 Peptidomimetic IX, 53-54, 118, 188, 197, 227,
monocyte 124, 130, 216 245-246, 248-254, 256, 258, 260, 262,
mood disorder 119 264, 266, 268, 270-274
mood ratings 124 P-glycoprotein 107-108, 115, 247, 273, 278, 291
multifunctional 290, 298, 301, 303, 319, 324, Pharmacokinetics VI, 62, 64-70, 72, 74-76, 78,
326, 331, 333-336, 340-342 80-82, 84, 86, 88-118, 209, 217-218, 221,
multiple-reaction monitoring (MRM) 107 223, 274, 282, 287-288, 324
NAD VI, 29-31, 34, 39-40, 42 phosphates 5, 11, 23-24, 151
neuroendocrine 119-121, 127, 130 plasma 64, 66, 69, 76, 79-81, 108-118, 123,
neurovegetative functions 121 125, 127, 129, 133-137, 139-142, 151, 154,
neutrophil 58, 225 156-157, 159-160, 162, 175, 190, 195, 197,
noradrenergic neurons 121 200, 211, 215-216, 226, 242, 247-248, 278,
nucleophiles V, 5-7, 20, 44-45 285, 312, 320, 322, 325, 336, 341
nucleophilic addition reactions 20-21, 23 plasmodium 52, 57, 59, 221, 275, 286-289
Nucleophilic Aromatic Substitutions V, 17 polarised 7
Nucleophilic Substitution Reactions V, 10-12 β–position 20
Nanomedicine 290, 292, 315, 327-330, 333- potassium VII, 132-134, 138-144, 146-147, 168
338, 340-341 precision 65, 75, 81, 84
nanoparticles X, 203, 221, 290-292, 294-296, preclinical 50, 112, 115-116, 118, 185, 211, 217,
298, 300-302, 304-308, 310-323, 325-342 223, 243, 290, 303, 313, 338
Nanotechnology IX, X, 290-292, 294-296, 298, pre-clinical development 87
300, 302-304, 306, 308, 310, 312, 314, primaquine 194, 217-221, 225, 227, 279-280,
316-324, 326-332, 334-338, 340, 342 285-289
nanovaccines 317-321, 336 prodrug VII, 54, 58, 114-116, 182, 184-186, 188-
natural mineral-rich waters VII, 132-133, 139 195, 215, 218-227, 326, 336
neurodegenerative diseases 168, 234, 240 protonation 20, 23, 38, 299
neuromodulation 165, 272 protons 5-6, 33
neuropeptide 246, 256, 264, 272 proline 178, 184, 186, 201, 224, 260-261, 266
neurotransmissor 181 pseudopeptides 246, 254-255, 270, 273
nicotine 138, 144 quality control 63, 81, 105
obese 135-137, 139, 141, 143-144 quinidine 281
obesity 132, 134, 138, 141-142, 144-145, 147 quinine IX, 275-278, 280-281, 284, 286, 288
oligopeptide 63, 107, 187, 190, 195, 207, 220, quinoline IX, 277-279, 287-288
224 racemates 10
optical rotation 9-10 Racemic mixtures 10
overlapping 13 radical 18, 37, 41, 55-56, 180, 195, 239
overweight 136, 139, 142 rats 76, 110, 113, 115, 128, 130, 134, 136-138,
β-Oxidation Pathway V, 14, 33 141-142, 144-147, 175, 257, 263, 274, 329
Oxidations V, 28 reactive oxygen species 137, 169, 234-235, 239
parasympathetic nervous system 121 redox reactions 28-29
Parkinson VIII, 153, 157, 166-167, 170, 172-174, Reductions V, 28, 126
229, 234, 239-240, 243 relative standard deviation (RSD) 84
particle size 106, 320, 331-332 renal function 114
passive targeting X, 292, 304-306, 308, 325 resistance 35, 51, 59, 62, 74, 105, 134-139, 141-
patient 81, 135, 156, 236, 294 144, 146-147, 188, 190, 195, 198, 222-223,
penicillin V, VI, 2, 35-38 228, 275-278, 280, 286-288, 291, 301-303,
peptide VIII, IX, 35-36, 48, 58, 188, 191-195, 305, 329, 332, 335-336
348 Index
resonance 5, 12-16, 20, 27, 36, 170, 324, 328, Th1 123-125, 129, 332
334 Th3 129
resonance structures 5, 27 Th1 cytokines 125
retinol 32 Th1 patterns 124
reversed-phase 88, 106 Th2 patterns 124-125
Rupintrivir VI, 50-51 thiazolidine 35-37
safety 51, 63, 65, 75, 79-80, 108, 114-115, 182, thioesters 5, 24, 34
276, 281, 317-318, 326 thiols V, 5, 31, 46
saturated fatty acids 35 toxicokinetics 79, 114
Schiff base 32 trans V, 8, 32, 34, 168, 170, 217
Semicarbazone V, 22 trans-configuration 8, 35
simple bond 4 transesterification 24-25
Single-reaction monitoring (SRM) 107 transforming growth factor beta (TGF-ß) 125
solid-phase extraction (SPE) 111 transpeptidase VI, 35-36
stable anion 14, 18, 23 Trier Social Stress Test 122
stereocenters 9 triple bond 4
stereogenic centers 9 targeting VIII, IX, X, 50-51, 58-59, 162-163, 165,
Stereoisomerism V, 8-9 188, 190, 193, 207, 209, 214-215, 217-218,
stereospecific 10, 27, 33, 35 222, 224, 229-230, 232, 234, 236, 238-
steric hindrance 19, 312 240, 242, 244, 256, 264, 274, 290, 292,
stress, chronic 132 294, 297, 303-315, 317, 320-322, 324-332,
stress system VI, VII, 119-121, 123, 125-126 334-342
stronger acid 6 tetraoxane 53-54
substrate 10-11, 36, 39, 42, 45, 109, 128, 186, therapeutic VIII, 44, 48, 58-59, 63-64, 80-81,
193-194, 215, 222 124, 126, 143, 153, 164, 168, 172, 185, 188-
sulfide 180 189, 195, 197, 206, 209, 211-216, 218, 222-
schizophrenia 127, 149, 152, 155, 158, 162-163, 223, 225, 227-228, 236, 242-243, 245, 247,
166-173, 175 274, 277, 279-280, 290-292, 294-295, 297,
sclerosis VIII, 229, 234-236, 242-244 301, 303, 305, 308, 310, 316-317, 319, 322-
Secondary Structure Mimetics IX, 250-251, 272 324, 326-327, 329, 331-333, 336, 339-341
self-assembling peptides VIII, 207, 209-210 thiol 33, 37, 44, 47-48, 50, 57, 133, 179, 182
serotonin 129, 145, 149-151, 153, 167, 169-170, thyroid 138
173-175, 181 tryptophan 17, 129, 151, 169, 178, 201
serum 79, 81, 116, 130, 133, 135, 137-139, 142- ultra-high pressure liquid chromatography
146, 194, 313 (UHPLC) 106
sodium VII, 114, 133, 138, 140, 142-144, 146, Unimolecular Elimination V, 18-19
168, 215, 230-232 Unimolecular Nucleophilic Substitution V, 11
Sprague-Dawley 134, 137, 142, 145 α,β-unsaturated carbonyl system 20
Stress-induced LC activation 121 unshared electrons 6
stroke VIII, 155, 166-168, 170-172, 229, 234, urocortin 120
238-240 Ussing chamber 107-108
structure-activity relationships (SARs) 249 vacant low-energy orbital 6
sulphadiazine 281 valacyclovir 186-189, 194, 222
sulphadoxine 281 valganciclovir 186-187, 194, 221, 226
sympathetic nervous system 120-121, 123, 136 Validation VI, 65, 75, 81-88, 112-115, 117-118,
tandem mass spectrometry (MS 64 142, 274
t cell 123, 320, 330, 338 vascular 79, 136-138, 141, 144-145, 220, 304-
α-terpineol V, 14 305, 326, 330-331, 335, 338, 340
tertiary carbons 11 vicinal carbons 11, 18
tetrahedral 13 vinylic systems 8
Index 349
Vision Pathway V, 32
vitamin A 32
volume of distribution 63, 76, 78, 105
weaker acid 6
Western diet 139-140, 142
Wistar 134, 136, 138, 142, 144
World Health Organization 22, 132, 142, 289,
341
wound healing 122, 207-208
Z 8, 57-58, 88, 107, 114, 116-118, 128-129, 141,
144, 168, 172, 198, 218, 220, 222, 224, 227,
242, 271, 274, 329-330, 333-334, 337, 339,
342
Z-isomer 8