Metabolism, Pharmacokinetics and Toxicity of Functional Groups
Metabolism, Pharmacokinetics and Toxicity of Functional Groups
Metabolism, Pharmacokinetics
and Toxicity of Functional Groups
Impact of Chemical Building Blocks on ADMET
Metabolism, Pharmacokinetics and Toxicity of Functional
Groups
Impact of Chemical Building Blocks on ADMET
RSC Drug Discovery Series
Editor-in-Chief
Professor David Thurston, London School of Pharmacy, UK
Series Editors:
Dr David Fox, Pfizer Global Research and Development, Sandwich, UK
Professor Salvatore Guccione, University of Catania, Italy
Professor Ana Martinez, Instituto de Quimica Medica-CSIC, Spain
Dr David Rotella, Wyeth Research, USA
Edited by
Dennis A. Smith
Sandwich Laboratories, Pfizer Global Research and Development, Kent, UK
RSC Drug Discovery Series No. 1
ISBN: 978-1-84973-016-7
ISSN: 2041-3203
A catalogue record for this book is available from the British Library
Apart from fair dealing for the purposes of research for non-commercial purposes or for
private study, criticism or review, as permitted under the Copyright, Designs and Patents
Act 1988 and the Copyright and Related Rights Regulations 2003, this publication may not
be reproduced, stored or transmitted, in any form or by any means, without the prior
permission in writing of The Royal Society of Chemistry or the copyright owner, or in the
case of reproduction in accordance with the terms of licences issued by the Copyright
Licensing Agency in the UK, or in accordance with the terms of the licences issued by the
appropriate Reproduction Rights Organization outside the UK. Enquiries concerning
reproduction outside the terms stated here should be sent to The Royal Society of
Chemistry at the address printed on this page.
The RSC is not responsible for individual opinions expressed in this work.
When it was suggested to me that the area of Drug Metabolism and Phar-
macokinetics was lacking in works available for the Medicinal Chemist I was
somewhat surprised. On reflection a study of many of the volumes revealed
that although the scholarship was outstanding many had not been written to
provide information to the medicinal chemist as they encountered new com-
pound series.
So often the change of project and perhaps even the chance of chemical lead
series takes the chemist into an area that he has not encountered before. The
changes in absorption, distribution, metabolism and excretion properties
(ADME) can be profound. For instance a chemist working on aminergic
GPCR receptors will be used to compounds cleared by metabolism, with large
volumes of distribution and often good access to the CNS. Results from screens
often do not need to be re-intepreted to allow for protein binding effects.
Perhaps a switch to a non-aminergic GPCR will cause the chemist now to
work with acidic molecules. Now high intrinsic clearance can be disguised by
high protein binding, metabolic clearance is augmented by significant drug
transporter effects and some of the metabolic steps are reversible (e.g. acyl
glucuronides).
This volume attempts to fill this void. Metabolism, Pharmacokinetics and
Toxicity of Functional Groups tries to do what it says on the cover. Our
definition of the key functional groups seemed right at the outset. As we have
assembled the work it is pleasing to see the holistic nature of the concept. Many
of the chapters build off concepts described in others but each can be viewed as
a seperate entity and one we hope chemists find rewarding as they move into
new chemical areas or perhaps revisit others.
The aim of all the authors was to impart around 300 years of collective
knowledge and wisdom about the impact of ADME on Drug Discovery and
v
vi Preface
hopefully aid all those active in this noble profession. We like to think that
someone, somewhere, has used a concept in this volume that has led to a new
medicine, even better if it is medicines in the plural. If this turns out to be true
our time will have been well spent. All the authors are associated with Pfizer
and we would like to thank the company for encouraging the production of
this work.
Contents
1.1 Introduction 1
1.2 Launched Drugs 6
1.2.1 Target Space of Launched Drugs 7
1.2.2 Chemical Space of Launched Drugs 9
1.2.3 Molecular Properties of Launched Drugs 10
1.2.4 Polypharmacology 13
1.3 Drugs Bound to their Targets 15
1.3.1 Comparison of Binding Sites of Drugs Bound
in their Biological Targets 25
1.3.2 Phosphodiesterase 5 (PDE5) Drugs 25
1.3.3 Cyclooxygenase (Cox) Drugs 26
1.3.4 Classes of Drugs with High Structural
Similarity 28
1.4 Privileged Substructures in Drugs 28
1.4.1 Examples of Privileged Substructures 30
1.4.2 Benzodiazepines 30
1.4.3 Arylsulfonamides and Drugs Derived from
them 31
1.4.4 Chemokine Receptor 5 (CCR5) 37
1.4.5 Diaryl Heterocycles such as Cyclooxygenase
Inhibitors (COX-2) 39
1.4.6 Aminoheterocycles as Kinases Inhibitors 41
1.4.7 HMG-CoA Reductase Inhibitors 44
1.5 Discussion of Privileged Substructures and Chemical
Space 44
vii
viii Contents
2.1 Introduction 61
2.1.1 Physicochemical Principles for ADME 62
2.1.2 Physicochemistry Summary 65
2.2 Delivery of Drugs and Bioavailability 65
2.2.1 Oral Delivery 66
2.2.2 Intranasal Delivery 73
2.2.3 Inhaled Delivery 73
2.2.4 Sublingual Delivery 73
2.2.5 Rectal Delivery 74
2.2.6 Transdermal Delivery 74
2.2.7 Subcutaneous and Intramuscular
Administration 74
2.3 Tissue Distribution of Drugs 75
2.3.1 Distribution to the Central Nervous System 80
2.4 Clearance, Extraction, Metabolism and Excretion 82
2.4.1 Clearance 82
2.4.2 Clearance by the Liver 84
2.4.3 Metabolism 84
2.4.4 Biliary Elimination 91
2.4.5 Clearance by the Kidney 92
2.4.6 Clearance Summary 93
2.5 Toxicology related to ADME 94
References 94
3.1 Introduction 99
3.2 Carboxylic Acid Containing Non-steroidal
Anti-inflammatory Drugs (NSAIDs) 103
3.2.1 Discovery of Aspirin 103
3.2.2 Mode of Inhibition of COX Activity by
NSAIDs 106
3.2.3 Molecular and Structural Basis for COX
Inhibition by NSAIDs 106
3.3 Carboxylic Acid Containing b-Lactam Antibiotics 107
3.3.1 Discovery of Penicillins 109
3.3.2 Mechanism of Action of b-Lactam Antibiotics 109
Contents ix
Chapter 4 Primary, Secondary and Tertiary Amines and their Isosteres 168
D. K. Walker, R. M. Jones, A. N. R. Nedderman and
P. A. Wright
9.3.4
Nucleoside Deaminase TSAI 435
9.3.5
Inosine 5-monophosphate Dehydrogenase
TSAI 440
9.3.6 Aspartate Carbamyl Transferase TSAI 441
9.3.7 Glycosidase Inhibitor TSAI 442
9.4 Conclusions 443
9.5 Abbreviations 444
References 445
1.1 Introduction
The major focus of the research-based pharmaceutical industry is the discovery
of safe, efficacious, new chemical entities (NCEs) for therapeutic targets. The
pharmaceutical industry can look back at a history of successful innovations,
indicated by the fact that there are currently just over 1400 unique drugs on the
market.
The success of the industry can be measured in, for example, the increase in life
expectancy in men and women over the last four decades. For instance, a child born
in the United States in 2005 can expect to live nearly 78 years (77.9 years). The
increase in life expectancy represents a continuation of a long-running trend. Life
expectancy has increased from 75.8 years in 1995 and from 69.6 years in 1955.
(www.cdc.gov/nchs/pressroom/07newsreleases/lifeexpectancy.htm). Although there
are multiple factors which potentially contribute to the increase in life expectancy,
like for example diet and life style, the development and availability of new drugs
appear to have made a substantial contribution.
Equally impressive, the impact of the industry can also be highlighted by
the increase in five-year-survival rates for cancer when diagnosed 1975–1977
1
2 Chapter 1
compared to when diagnosed in 2000 (www.phrma.org/files/PhRMA%-
202009%20Profile%20FINAL.pdf). Between 1975 and 1979, the five-year sur-
vival rate for cancer was just 50%; by 2000, survival had risen to 67%. Survival is
increasing dramatically for many forms of cancer. The rate of five-year survival
went up 21% for breast cancer, 42% for prostate cancer, 28% for colon and
rectum cancer, and 25% for lung and bronchial cancer.
Drug discovery is a complex multivariate process, but the basic requirements
for orally administered NCEs include novelty and patentability, intrinsic
potency, oral bioavailability, no toxicological effects in humans, and a sig-
nificant advantage over existing accepted therapies (if applicable). A schematic
representation of the drug discovery process in the United States is shown in
Figure 1.1.
Although it is possible to predict, with varying accuracy, what a NCE will do
when orally administered to humans, the full potential of a NCE is not known
until it has been tested in clinical trials. Therefore, any investment made will not
yield any return until the NCE is on the market, which could be in the region of
ten years after patenting, and for the majority of compounds there will be no
return at all to offset the enormous costs of drug discovery and development.
Therefore, drug discovery is a high risk business with massive, long-term up-
front investments aiming at discovering the few blockbusters that are on the
market at any one time. In addition, the pharmaceutical industry is one of the
most research-intensive industries; in the United States, an average of 16% of
sales is spent on R&D, second only to the aerospace industry (www.nsf.gov/
statistics). The global pharmaceutical market is worth $553.4 billion in the top
ten markets alone (Table 1.1).
The top ten marketed drugs and their revenue between June 2007 and June
2008 are shown in Table 1.2; they account for a total of $67.4 billion,1 which is
only 12.2% of total sales in the top ten markets.
Among the top ten drugs, Pfizer’s Lipitor is by far the biggest seller, $5.5
billion ahead of a cohort of three drugs, Plavix, Nexium and Serentide with
sales of $8.3, $7.7 and $7.5 billion, respectively. The top ten therapies are shown
Figure 1.1 Schematic representation of the drug discovery process with typical time
frames and attrition rates from drug discovery to FDA approval (adapted
from www.phrma.org/files/PhRMA%202009%20Profile%20FINAL. pdf).
Drugs and their Structural Motifs 3
Table 1.1 Top ten pharmaceutical markets worldwide and their revenue for
the period June 2007 to June 2008.
Country Sales June 2007–June 2008 ($ billions) Share of global sales (%)
Table 1.2 Top ten marketed drugs worldwide for the period June 2007 to June
2008.
Sales June
2007–June
Name Compound Marketer Indication 2008
in Table 1.3 and account for 36.5% of global sales. The annual sales figures
indicate that oncologics are by far the biggest revenue stream for the phar-
maceutical industry, followed by lipid regulators. Interestingly, Pfizer’s Lipitor
alone accounts for almost half of lipid regulator sales.
Historically, big pharmaceutical companies delivered. The secret of their
success was simple: pharmaceutical companies brought a huge number of
innovative products to the market that genuinely helped sick people, and so
4 Chapter 1
Table 1.3 Top ten drug therapies and their annual global sales for the period
June 2007 to June 2008.
Therapy Sales June 2007–June 2008 Share of global sales(%)
60
New molecular entities
50 Biologic license application
40
30
20
10
0
1996 1997 1998 1999 2000 2001 2002 2003 2004 2005 2006 2007 2008
Figure 1.2 New molecular entities (NMEs) and biologic license applications
approved by the US FDA’s Center for Drug Evaluation and Research
between 1996 and 2008.
were readily prescribed, which generated solid sales.2 Even during times of
economic hardship, drugs continued to be an essential purchase. During this
flourishing period from the mid-1980s to the beginning of this decade, major
drug companies routinely generated double-digit growth in sales year after
year.
However, the pharmaceutical industry’s investment in R&D has also risen
steeply over the last 20 years, with R&D spending of $47.9 billion in 2007
compared with $26 billion in 2000 and $8.4 billion in 1990, and an average cost
of $1.3 billion for bringing a new drug to market—an increase of 65% since
2000 (www.phrma.org/files/PhRMA%202009%20Profile%20FINAL.pdf).3
Despite this increased investment in research and development, the number of
new molecular entities (NMEs) has not increased in line with rising investment;
in fact it declined between 1996 to 2008 from 54 to 21 (Figure 1.2).4–15
Drugs and their Structural Motifs 5
Figure 1.2 could suggest that there has been a significant decline in inno-
vation rates in the pharmaceutical industry over the last decade. The reasons
for this decline have been reviewed extensively16 and several causes have been
indicated as contributors to the R&D decline.17 Among these are for example
submaximal optimisation of resources and the inability to control costs as well
as negative impact of mergers and acquisitions, which have been grouped
together as factors internal to R&D. Alongside these, external reasons for the
decline include evolving healthcare, regulatory burden, lack of regulatory
harmonisation as well as changes in tolerance for risk.17,18 Looking back over
recent decades, total approvals by the US Food and Drug Administration
(FDA) reached a record high of 381 entities in the decade between 1995 and
2004 compared with the two previous decades (241 in the decade 1985–1994
and 190 in the decade 1975–1984). Thus, it would appear that a myopic focus
on near-term performance has given rise to a perception that bears very little
relationship to the actual innovation rates of the pharmaceutical industry in the
last decade.19
However, the issue of high attrition rates in drug discovery and development
still remains, without which the innovation rates would be even higher and,
potentially, would keep better track with the enormous increases in R&D
investment. Only about 11% of compounds entering clinical development ever
reach the market, being withdrawn for reasons such as efficacy (25%), tox-
icology (24%), clinical safety (12%), drug metabolism and pharmacokinetics
(DMPK, 8%), formulation (1%) and portfolio-related and other reasons
(30%).20 Therefore, out of the 70% of failures caused by specific effects, the
majority of 61% can be attributed to lack of efficacy, toxicology and clinical
safety, whereas DMPK (physicochemical properties, or drug likeness, of the
drug candidate itself) accounts for only 8% of attrition.21 However, the actual
proportion may be higher since some reported attrition, which was attributed
to lack of efficacy, might be due at least in part to poor DMPK.20 A similar
proportion of 7% was discussed as having inappropriate absorption, dis-
tribution, metabolism and excretion (ADME) properties among NCEs between
1964 and 1985.22 In addition, apparently only about 30% of marketed pre-
scription drugs produce revenues that match or exceed average R&D costs.22
The apparent decrease in productivity in the entire pharmaceutical industry
has put enormous financial pressures on individual companies and their share
price—one of the measures of confidence of investors in future profitability.23
Although the underlying reasons for this decline in productivity are complex,
many factors have been suggested, such as for example increasing clinical
development costs, FDA approval standards and political pressures on drug
pricing.20
One of the key reasons for the decline in productivity is without doubt the
high rate of attrition at all stages of the drug discovery process from failures in
the early pre-clinical stages to the very expensive late stage failures in the clinic
or even post-launch. Although exact figures on attrition in drug discovery are
difficult to derive due to the sparseness of publicly available data, it is clear that
success rates of discovery projects over the last decade, perhaps in part due to
6 Chapter 1
the very high attrition rates, have not been able to match expectations in terms
of productivity targets.
Therefore, attempts to reduce attrition early in the drug discovery process
have been a major focus over the last decade. During that time, the application
of guidelines linked to the concept of drug-likeness (in particular absorption)
such as the ‘rule of five’24 (see Section 2.1.1 for details) has gained wide
acceptance25 as an approach to reducing attrition in drugs.26,27 However,
despite this acceptance, an analysis of recent trends revealed that the physical
properties of molecules that are currently being synthesised in leading drug
discovery companies differ significantly from those of recently discovered oral
drugs and compounds in clinical development.26 This was particularly notable
for lipophilicity, where the consequences of a significant increase include a
greater likelihood of lack of selectivity and attrition in drug development.26
Physicochemical properties of molecules are completely under the control of
medicinal chemists and can be easily calculated for very large numbers, in some
cases for hundreds of thousands of designed structures prior to synthesis.
Close monitoring of physical properties during a drug discovery programme
and compound series selection based on orthogonal attrition risks as indicated
by compound properties and chemical scaffold may provide the medicinal
chemist with opportunities to significantly reduce attrition rates, which are
currently estimated at 93–96%.28
In this chapter, we focus on the relationship of molecular properties and
functional groups of compounds on their interactions with biological targets,
which can potentially impact on their pharmacological profile and their
potential attrition risks.29
35
30
25
20
15
10
0
Rhodopsin-like
receptors
Ligand-gated ion
Penicillin-binding
Myeloperoxidas
Neurotransmitter
topoisomerase
Fibronectin type
Cytochrome
Others
Voltage-gated
symporter family
ion channels
Nuclear
Type II DNA
P450
GPCRs
channels
protein
e-like
III
Figure 1.3 The gene family share as a percentage of all FDA-approved drugs for the
top ten families. Beyond the ten most commonly drugged families, there
are a further 120 domain families or singletons for which only a few drugs
have been successfully launched. The data is based on 1357 dosed com-
ponents from 420 000 approved products, FDA, December 2005.
Drugs and their Structural Motifs 9
molecular properties. The usefulness of the concept of druggability from a
medicinal chemistry standpoint has been summarised by pointing out that the
rule of five (Ro5) and its extensions have generated awareness about the
importance of pharmacokinetic parameters for drug discovery and develop-
ment. In addition, the concept of druggability has led to the realisation that
there may be whole families of proteins for which it is either extremely chal-
lenging or impossible to design small molecules with acceptable oral
bioavailability.50
Another concept related to druggability is ligand efficiency,51,52 which gen-
erates a quantitative relationship between drugs and their biological targets,
and is defined as the binding energy per non-hydrogen atom in a particular
molecule. This concept can be very useful for lead selection by normalising
binding energy for molecular weight, but also for differentiation between gene
families that have a high or low probability of binding Ro5-compliant small
molecules based on an analysis of experimentally determined ligand-binding
energies for a particular target. These concepts of maximal affinity and ligand
efficiency have been developed further into a computational approach to pre-
dict druggability.53
In the past ten years or so, expectations in the pharmaceutical industry
have been raised as many companies have invested significantly in high-
throughput technologies that would make use of information derived from
the sequencing of the human genome.54 Therefore, it would seem that
companies are now well-placed to take advantage of the discovery of new
targets that have appeared in the post-genomic era. However, there appears to
be a reduced likelihood of delivering a preclinical drug development candi-
date against a new target, which could lead to a temptation to concentrate on
more established targets to reduce risk in current development portfolios.54
More recent in silico approaches such as high-throughput electronic biology
may help in identifying, for example, previously unknown complex relation-
ships between targets as well as compounds and targets in biological pathways
on a large scale in order to support many parallel work streams in a drug
discovery portfolio.55
700
600
500
400
300
200
100
0
P450
Transporter
GPCR aminergic
Ion channel
NHR
Other
PDE
Kinase
GPCR lipid
GPCR peptidic
Protease
Figure 1.4 Average molecular weight of oral drugs for gene families.
Although the data seems to indicate that overall molecular weight of dis-
covery and development compounds has increased over recent decades, it has
been suggested that the upward trend can be explained largely by variations in
the target portfolios of pharmaceutical companies. Most notably, a significant
decrease of biogenic amine GPCR drugs in the recent decades (43% to 28%)
and increases in protease and peptidic GPCR targeted drugs may explain much
of the overall molecular weight trend. Variation in properties over time for a
given family may result from varying pharmaceutical interest in its members
(e.g. serine proteases, metalloproteases, etc.).63
The central assumption in the applicability of standard rules for drug likeness
is that the target of interest requires molecular properties similar to those of the
average drug. Since bioavailability results from the interactions of drugs with
the same biological systems (i.e. membranes in the gastrointestinal tract as well
as metabolic enzymes like P450s and transporters), it is possible that well-
defined ranges of molecular properties can account for favourable interactions
with those systems. For certain proteomic families, application of standard
rules of drug likeness could bias the composition of corporate screening col-
lections away from the molecular properties needed for achieving high affinity
(e.g. for protein–protein interactions).67 The need to balance bioavailability
and affinity suggests that modified rules of drug likeness need to be adopted for
certain target classes.63
12 Chapter 1
The properties showing the clearest influence on the successful passage of a
candidate drug through the different stages of development are molecular
weight and lipophilicity. Statistical analysis shows that the mean molecular
weight of orally administered drugs in development decreases on passing
through each of the different clinical phases and gradually converges toward
the mean molecular weight of marketed oral drugs. It is also clear that the most
lipophilic compounds are being discontinued from development.66 This work
supports Lipinski’s findings that there are limiting factors to the molecular
weight and lipophilicity of a candidate drug that are reflected in the current
physicochemical property profiles of the marketed oral drug data set. This
study also suggests that these limiting values of physicochemical properties are
not historical artefacts but are under physiological control.66 In addition, an
analysis of the difference between drugs and their original lead compounds
shows that, for the majority of cases, only small structural changes are made in
the lead to drug process.68
Therefore, it can only be advantageous if a screening collection already
contains drug-like compounds with the right physical properties which carry a
lower risk of attrition during drug development. This has been developed into
the lead-like paradigm, which states that lead compounds need to be left-shifted
in terms of molecular weight and lipophilicity compared to drugs in order to
allow for the additional molecular weight and lipophilicity which has histori-
cally been added in the lead to drug process.69 Recent studies have shown that
molecular weight and log D are the most important factors in determining the
permeability of drug candidates.70 It has also been shown that the log D limits
are dependent on molecular weight, and rules have been defined for log D limits
required to achieve 450% chance of high permeability for a given MW band
(Table 1.5).
Although both molecular weight and log D have been linked to permeability,
log D may be the more important factor since it is an expression of multiple
molecular properties such as hydrogen bond donors and acceptors, lipophilicity
as well as dipole and polarisability, which are linked to physicochemical events
like permeability and binding.
o300 40.5
300–350 41.1
350–400 41.7
400–450 43.1
450–500 43.4
4500 44.5
Drugs and their Structural Motifs 13
Differentiation between drugs and non-drugs on the basis of molecular
properties has been reported in the literature, either based on statistical
approaches71 or analysis of structural features72 with an accuracy of between 71
and 92%, respectively. However, these approaches are mostly limited to
oral drugs and it has been shown that, for example, inhaled drugs reside in a
region of molecular property space which is very different from that of oral
drugs.73
In addition to predicting druggability for biological targets which can bind
small molecules, there has been considerable interest in targeting protein–
protein interactions with small molecules. Protein–protein interactions are
highly attractive for drug discovery because they are involved in a large number
of disease pathways where therapeutic intervention would bring widespread
benefit. Recent successes have challenged the widely held belief that these
targets are ‘undruggable’.74,75 Targeting interfaces between proteins has huge
therapeutic potential, but discovering small-molecule drugs that disrupt pro-
tein–protein interactions is an enormous challenge, which is being faced with
the support of for example bioinformatics approaches.76 This vast new field of
drug discovery has enormous potential but is outside the remit of this chapter;
for further reading we refer you to recent successes reported in the
literature.67,74,75
1.2.4 Polypharmacology
The understanding of pharmacological space is one of the fundamental aspects
of drug discovery, relating to off-target activity and in turn to compound
attrition. Pharmacological space has been mapped recently by Paolini et al.
through large-scale data integration of proprietary and published screening
data.39 They have assigned 2876 targets to protein sequences from 55 organ-
isms, with biologically active chemical tools for 1306 proteins. After removing
redundancies in the mammalian genes due to orthologs among species, 836
genes could be unambiguously identified in the human genome for which small-
molecule chemical tools with biological activity of IC50o10 mM have been
discovered. This number drops to 529 when a perhaps more realistic threshold
of 100 nM is applied (Table 1.6).
Of the pharmacological targets selected, 158 human proteins have been
identified as the primary modes-of-action for approved small-molecule drug
targets, with oral small-molecule drugs primarily targeting only 141 human
proteins.
A key question in global pharmacological space is how extensive is pro-
miscuity, which is defined as the specific binding of a chemical to more than one
target. Considering each pair of targets in turn, if two proteins both bind to the
same ligand, they can be considered as interacting in chemical space even if they
have no other interaction in physical space or similarity in sequence space. The
concept of ‘target-hopping’, where chemical matter for one target can be
considered as the basis for leads or tools for another target, has historically
14 Chapter 1
39
Table 1.6 Pharmacological target space.
Human targets with Human targets with
Gene family o10 mM binding affinity o100 nM binding affinity
been an extremely fruitful method of drug discovery.77,78 Of all the 276 122
active compounds found in the database used by Paolini et al.,39 65% have
recorded activity for one target, whereas 35% are reported to hit more than one
target.
The observed polypharmacology interaction network for human proteins
was mapped to navigate polypharmacology relationships between targets. The
entire protein interaction network consists of 700 proteins connected by 12 119
interactions for all compounds below the affinity threshold of IC50 10 mM, and
with a difference in affinity of up to three orders of magnitude between two
targets. Promiscuity can be considered from the perspective of both the com-
pound and the pharmacological target, to measure compound selectivity and
target overlap.79,80 Table 1.7 shows the top ten promiscuous targets taken from
ref. 39.
Although different definitions of promiscuity result in different rankings, the
same target classes (aminergic GPCRs, cytochrome P450s and protein kinases)
appear at the top regardless of the method used in the analysis (Table 1.6).
Aminergic GPCRs and protein kinases exhibit the greatest intra- as well as
inter-gene family promiscuity. The data set used for this work is a sparse
matrix, since activity data for each compound is mostly limited to only a few
targets. There are indications that molecular properties and the potentially
Drugs and their Structural Motifs 15
39
Table 1.7 Top ten most promiscuous targets.
Number Protein Gene family
O
O O
HN
O
H2N
HN N N
HN
O N
Chapter 1
N
Argatroban Thrombin Anti-coagulant 1dwc
HN
O
NH2 O S
NH
H 2N N
N O
O
Drugs and their Structural Motifs
OH
O N
N N
H
O
O
O
O
O
Chapter 1
Celecoxib (Celebrex) (structure of ana- Cyclooxygenase 2 Inflammation, pain, 6cox
logue SC-544) arthritis
F
N F
N F
H2N
S
O O
Cl O
OH
O
19
O
20
Table 1.8 (continued )
PDB
Drug name Structure Target Indication code
N
H
N OH
H2N N
O
H2N N N
O OH
OH
Penciclovir O Herpes simplex type-1 thymi- Herpes 1ki3
N dine kinase
HN
H2N N N
OH
Chapter 1
OH
Ritonavir HIV-1 protease HIV 1rl8
S N
N S
N
O O
O HN
HN OH
HN
O
Drugs and their Structural Motifs
N
O
NH
O H
N
HN N
O HO
N
OH
21
N
22
Table 1.8 (continued )
PDB
Drug name Structure Target Indication code
S
O
HO OH
N
H H
N H
N
H
O
O N
O
HN
HO
O
HN
O
Chapter 1
Atazanavir (Reyataz) HIV-1 protease HIV 2aqu
N
O OH O
H H
N N N O
O N N
H H
O O
F O
F
F N HO O
H
N
S
O O
HO
NH
O H
O
O O
H
23
24
Table 1.8 (continued )
PDB
Drug name Structure Target Indication code
N N N
O
Cl
F
F F
H2N N N
Chapter 1
Drugs and their Structural Motifs 25
1.3.1 Comparison of Binding Sites of Drugs Bound in their
Biological Targets
One of the concepts discussed above is druggability, which relates to the like-
lihood of success for discovery of a drug based on the molecular properties of a
small molecule ligand and the binding site properties of the receptor. The
druggability concept is also based on the assumption that there are particular
features of a binding site which enable it to bind small molecules with sufficient
potency in order to meet requirements for a drug.
Rules have been derived in order to quantify druggability87 and make pre-
dictions based on analysis of the receptor structure as to whether a target is
likely to be druggable.53 Some of the parameters linked to druggability are:53
the degree to which the binding site is buried inside the protein;
the curvature of the binding pocket;
the topology of multiple pockets and their relative positions in the binding
site;
lipophilicity;
polarity;
the ability to form hydrogen bonds.
Figure 1.5 Tamiflu bound to neuraminidase (PDB 1l7f, left), Sildenafil bound to
phosphodiesterase 5 (PDB 1udt, right) and Iressa bound to EGFR kinase
(PDB 2ity, bottom). Carbon atoms are yellow, light blue and purple,
respectively; oxygen atoms are red, nitrogen atoms are blue and sulfur
atoms are yellow. Surfaces are coloured by atom type: carbon ¼ green,
nitrogen ¼ blue, oxygen ¼ red, sulfur ¼ yellow.
able to bind to the same binding site, occupying significantly different areas in
the receptor (e.g. such as at the bottom and to the right of Figure 1.6).
Therefore, it appears that although sildenafil and vardenafil may contain
substructures which bind strongly to phosphodiesterase, they are two very
different interpretations of the phosphodiesterase pharmacophore which can
both show significant levels of activity in the receptor sufficient for the desired
pharmacological effect.
This ability to bind diverse substructures which are expressing a similar
pharmacophore does not seem to be specific to PDE5, but is observed perhaps
in an even more striking example of cyclooxygenase as highlighted below.
Figure 1.6 Overlay of experimental binding orientations for three marketed drugs for
erectile dysfunction: sildenafil (light grey), vardenafil (dark grey) and
tadalafil (white). Compound structures are shown in Table 1.8.
Figure 1.7 Overlay of ibuprofen (white), flurbiprofen (black), diclofenac (medium grey),
celecoxib analogue SC544 (light grey) and salicylic acid (dark grey). The acid
groups of diclofenac (top right) and ibuprofen, flurbiprofen and salicylic acid
(bottom) are 9.1Å apart. Comound structures are shown in Table 1.8.
28 Chapter 1
Although a CF3 group is considered to be a bioisostere for an acid group, in
this case the CF3 group overlaps with the methyl groups of ibuprofen and
flurbiprofen, rather than the acid groups, and is therefore a bioisostere for a
lipophilic rather than a polar group. A more detailed analysis of the binding of
Cox inhibitors (particularly Cox-2 inhibitors) is presented in Chapter 5.1.2.
1.4.2 Benzodiazepines
In the late 1980s, Evans and co-workers first defined the concept of privileged
structures.95 They worked on the development of novel non-peptidic chole-
cystokinin (CKK) receptor antagonists for the treatment of gastrointestinal
disorders (e.g. pancreatitis and gastroesophageal reflux) based on analogues of
the natural product asperlicin via structural modification of anxiolytic
Drugs and their Structural Motifs 31
110
benzodiazepine drugs such as diazepam. This work resulted in the devel-
opment of devazepide (MK-329) as the first non-peptidic benzodiazepine
antagonist, highly selective for cholecystokinin-1 (CCK-1) (IC50 0.8 nM).111–
113,93–95
Interestingly, this work involved key elements of structural planning
and molecular simplification from the natural product asperlicin. It was sug-
gested that both the benzodiazepine and tryptophan subunits of asperlicin are
key elements of the pharmacophore for molecular recognition be the CCK-1
receptors. Natural product guided development of CCK-1 antagonists is shown
in Figure 1.8.
Evans et al. recognised that this was not the first successful incorporation of
benzodiazepines into bioactive molecules. In fact the benzodiazepine motif
constitutes a broad class of neuroactive compounds acting as ligands to ion
channels and GPCRs. Notable examples of anxiolytic drugs of this class are
diazepam and lorazepam, which are ligands of central nervous system (CNS)
gabaergic receptors.114 In addition, there are the extensive numbers of CCK-1
diazepine containing ligands and numerous other applications as ligands to
other GPCRs such as k-opioid agonists like tifluadom115 for the treatment of
visceral pain, antithrombitic platelet activation factor (PAF) antagonists,116
analgesic and anti-inflammatory neurokinine (NK-1) receptor antagonists,117
and GPIIbIIIa receptor antagonists with antithrombitic profiles.118 In addition,
multiple classes of enzyme inhibitors have been developed that contain the
benzodiazepine unit; for example, HIV reverse transcriptase inhibitors such as
nevirapine,119 RAS-farnesyltransferase inhibitors for the treatment of cancer
(e.g. BMS-214662120). This diversity of bioactivity for benzodiazepines across a
broad range of classes of drug target led Evans to define this group as privileged.
A group of representative diazepine-containing drugs is shown in Figure 1.9.
Examining reasons for benzodiazepines to be privileged in this manner has
led to them being identified as b-turn peptidomimetics.121–123,92 The presence of
such structural motifs that are complementary to an array of primary and
secondary structural elements in proteins offers a potential explanation for the
promiscuous nature of the binding of many recurring scaffolds.
O
N
N
OH H H H
N
NH
N
O
O
Benzodiazepine
asperlicin
Tryptophan
Cl N
diazepam
O O
N
N
H
N HN
devazepide
H O
N N
OH O
Cl N N HN
Cl F
S
Iorazepam tifluadom
BZD/Gabba k-opioid agonist
H O
N H O
N
CF3
N
N N N
N
O
CF3
NK1 antagonist nevirapine
RT inhibitor
HN
N O
N
O
N
N
Ar O
NC N CO2H
S GPIIbIIIa antagonist
O S
O
BMS-214662
RAS-farnesyl
transferase inhibitor
N
HN HN
H2N S O H2N S N
O O
O O
HN
H2N S N
O
O
Figure 1.11 Structures of sulfacetamide (top left), sulfadiazine (top right) and
sulfadimidine.
HOOC Cl N
NH2 H2N NH
S S S
O O O O O O
H
F3C N
Cl NH
H2N NH
S S H2N
S COOH
O O O O
O O
N
HN
O
O
O
O
S COOH
S COOH O
O
NH2
NH2
Figure 1.12 Structures of diuretics: carzenide (top left), chlorothiazide (top right),
bendrofluazide (middle left), frusemide (middle right), bumetanide
(bottom left) and piretanide (bottom right).
N
N O
O O
O O
S N
S S
N N H
H H
NH2 NH2
O O
O O
O O
S N N
N H S H
N
H H
Cl
O
O
O
S N
N H
H
N
H
Cl O
O
O
OCH3
S N
N H
H
N N
H
O NH F
N OH
O
F F
N
F
N
N
O N
N O
F S O
N O
O
S
O NH2
S
O
O
F3C O H2N
O N Cl
N S
N
Me Cl
celecoxib refecoxib
COX-2 inhibitor COX-2 inhibitor N
IC50 40 nM IC50 20 nM CGS-2466
SO2NH2 SO2Me A3 antagonist, PDE4 inhibitor,
p38 inhibitor IC50 40 nM
AcS Cl
O
N
N Me
H
NH
N N N
N N
N N Cl
BuO Cl
rimonabant
SB-203580 glucagen antagonist
CB1 inverse agonist
p38 inhibitor
Ki 56 nM
IC50 600 nM
F OBu Cl
HN OH
N N
N OH
N
N H2N N N
H 2N N
O
Abacavir Acyclovir
OH
Kinase Hinge
HBA
HBD
Adenine Selectivity
Gate site
site Keeper
Activation
loop site
DFG-out site
Ribose
site
Asp
Phosphate
site
OH
R
and residue selection within this site across kinases. Furthermore, the ATP site
is well adapted to recognise the ATP purine ring head group via a series of
hydrogen bonding interactions.140 A schematic of ATP recognition by kinases
is shown in Figure 1.17.
Aranov et al. from Vertex discussed in detail the concept of kinase privileged
hinge-binding fragments and the notion of kinase targeted libraries whereby
libraries of known promiscuous kinase inhibitor scaffolds were made to spe-
cifically target SAR in kinase drug space.109 They investigated the idea of
kinase likeness in the context of kinase privileged fragments. Initially they
carried out an analogous method to that described by Murcko97,102 to define
the structures of kinase inhibitors in the context of their framework and
side chain atoms. The analysis was performed on 119 published kinase inhi-
bitors and revealed that the structural diversity at the level of rings and
linkers was relatively low. A combination of four rings and eight linkers was
found to describe 90% of the dataset.109 In particular, amino-substituted
Drugs and their Structural Motifs 43
N H N H
N N N N
Cl
O O F O O
Gefitinib Erlotinib
O O
N
N
HN
N H
N H N N
N
OH N
Cl N
KDR inhibitor TGFβ inhibitor
F F
OH OH O OH OH O
O OH OH
O O N
S
N N N
Cerivastatin R
OH OH O
NH
N OH
O
Atorvastatin
F
HO O
HO O
O
O O
O
O
H O
CH3 H
CH3
HO
Pravastatin CH3 Lovastatin
HO O HO O
O O
O O
O O
H H
CH3 CH3
MeO
OH
OH H
N N
NH2
O N O Cl
OMe
F F
OH OH O OH OH O
O OH OH
O O N
S
N N N
O O
OH OH
1.8 Abbreviations
ADMET absorption, distribution, metabolism, excretion, toxicity
ADP adenosine diphosphate
ATP adenosine triphosphate
DMPK drug metabolism and pharmacokinetics
CCK cholecystokinin
CMC Comprehensive Medicinal Chemistry
CNS central nervous system
COX cyclooxygenase inhibitor
EGFR epidermal growth factor receptor
Drugs and their Structural Motifs 53
FDA Food and Drug Administration
GABA gamma-aminobutyric acid
GPCR G-protein coupled receptor
GTP guanosine triphosphate
HERG human ether-a-go-go related gene
HIV human immunodeficiency virus
HMG-CoA 3-hydroxy-3-methyl-glutaryl-coenzyme A
HTS high throughput screening
MW molecular weight
NCE new chemical entity
NHR nuclear hormonal receptor
NME new molecular entity
NMR nuclear magnetic resonance
NSAID non-steroidal anti-inflammatory drug
PAF platelet activation factor
PDB Protein Data Bank
PDE phosphodiesterase
Ro5 rule of five
R&D research and development
SAR structure–activity relationship
TPSA topological polar surface area
References
1. S. J. Ainsworth, Chem. Eng. News, 2008, 86, 15–24.
2. N. N. Malik, Drug Discov. Today, 2008, 13, 909–912.
3. N. N. Malik, Expert Opin. Drug Discov., 2009, 4, 15–19.
4. B. Gaudilliere and P. Berna, Annu. Rep. Med. Chem., 2000, 35, 331–
356.
5. B. Gaudilliere, P. Bernardelli and P. Berna, Annu. Rep. Med. Chem., 2001,
36, 293–318.
6. P. Bernardelli, B. Gaudilliere and F. Vergne, Annu. Rep. Med. Chem.,
2002, 37, 257–277.
7. C. Boyer-Joubert, E. Lorthiois and F. Moreau, Annu. Rep. Med. Chem.,
2003, 38, 347–374.
8. S. Hegde and J. Carter, Annu. Rep. Med. Chem., 2004, 39, 337–368.
9. S. Hegde and M. Schmidt, Annu. Rep. Med. Chem., 2005, 40, 443–473.
10. S. Hegde and M. Schmidt, Annu. Rep. Med. Chem., 2006, 41, 439–477.
11. S. Hegde and M. Schmidt, Annu. Rep. Med. Chem., 2007, 42, 505–553.
12. S. Hegde and M. Schmidt, Annu. Rep. Med. Chem., 2008, 43, 455–497.
13. J. Owens, Nat. Rev. Drug Discov., 2007, 6, 99–101.
14. B. Hughes, Nat. Rev. Drug Discov., 2009, 8, 93–96.
15. B. Hughes, Nat. Rev. Drug Discov., 2008, 7, 107–109.
16. E. Schmidt and D. Smith, Drug Discov. Today, 2005, 10, 1031–1039.
17. R. R. Ruffolo, Expert Opin. Drug Discov., 2006, 1, 99–102.
54 Chapter 1
18. H.-G. Eichler, F. Pignatti, B. Flamion, H. Leufkens and A. Breckenridge,
Nat. Rev. Drug Discov., 2008, 7, 818–826.
19. E. F. Schmid and D. A. Smith, Annu. Rep. Med. Chem., 2005, 40,
431–441.
20. P. D. Leeson and A. M. Davis, J. Med. Chem., 2004, 47, 6338–6348.
21. H. Kubinyi, Nat. Rev. Drug Discov., 2003, 2, 665–668.
22. H. Grabowski, J. Vernon and J. A. DiMasi, Pharmacoeconomics, 2002,
20, 11–29.
23. J. A. Vernon, Expert Opin. Drug Discov., 2009, 4, 21–22.
24. C. A. Lipinski, F. Lombardo, D. B. W. and P. J. Feeney, Adv. Drug Del.
Rev., 1997, 23, 3–25.
25. C. A. Lipinski, Drug Discov. Today: Technologies, 2004, 1, 337–341.
26. P. D. Leeson and B. Springthorpe, Nat. Rev. Drug Discov., 2007, 6,
881–890.
27. P. D. Leeson, A. M. Davis and J. Steele, Drug Discov. Today: Technol-
ogies, 2004, 1, 189–195.
28. I. Kola and J. Landis, Nat. Rev. Drug Discov., 2004, 3, 711–715.
29. U. A. K. Betz, Drug Discov. Today, 2005, 10, 1057–1063.
30. I. Cavero and H. R. Kaplan, Expert Opin. Drug Discov., 2008, 3,
1145–1154.
31. W. Sneader, Drug Prototypes and their Exploitation, Wiley, London, 1996.
32. E. M. John, Annu. Rep. Med. Chem., 2008, 43, 525–544.
33. J. P. Overington, B. Al-Lazikani and A. L. Hopkins, Nat. Rev. Drug
Discov., 2006, 5, 993–996.
34. S. J. Haggarty, Curr. Opin. Chem. Biol, 2005, 9, 296–303.
35. IHGS Consortium, Nature, 2001, 409, 860–921.
36. IHGS Consortium, Nature, 2004, 431, 931–945.
37. A. L. Hopkins and C. R. Groom, Nat. Rev. Drug Discov., 2002, 1,
727–730.
38. A. P. Orth, S. Batalov, M. Perrone and S. K. Chanda, Expert Opin. Ther.
Targets, 2004, 8, 587–596.
39. G. V. Paolini, R. H. B. Shapland, W. P. van Hoorn, J. S. Mason and A. L.
Hopkins, Nat. Biotech., 2006, 24, 805–815.
40. P. R. Caron, M. D. Mullican, R. D. Mashal, K. P. Wilson, M. S. Su and
M. A. Murcko, Curr. Opin. Chem. Biol., 2001, 5, 464–470.
41. C. J. Harris and A. P. Stevens, Drug Discov. Today, 2006, 11, 880–888.
42. T. Klabunde, Brit. J. Pharmacol., 2007, 152, 5–7.
43. M. Casesa and J. Mestres, Drug Discov. Today, 2009, 14, 479–485.
44. P. Bamborough, D. Drewry, G. Harper, G. K. Smith and K. Schneider,
J. Med. Chem., 2008, 51, 7898–7914.
45. J. Drews, Nat. Biotech., 1996, 14, 1516–1518.
46. J. Drews and S. Ryser, Nat. Biotech., 1997, 15, 1318–1319.
47. A. L. Hopkins and C. R. Groom, Ernst Schering Res. Found. Workshop,
2003, 42, 11–17.
48. G. Vistoli, A. Pedretti and B. Testa, Drug Discov. Today, 2007, 13,
285–294.
Drugs and their Structural Motifs 55
49. Y. Sugiyama, Drug Discov. Today, 2005, 10, 1577–1579.
50. T. H. Keller, P. Arkadius and Y. Zheng, Curr. Opin. Chem. Biol., 2006,
10, 357–361.
51. I. D. Kuntz, K. Chen, K. A. Sharp and P. A. Kollman, Proc. Natl. Acad.
Sci. U.S.A., 1999, 96, 9997–10002.
52. A. L. Hopkins, C. R. Groom and A. Alex, Drug Discov. Today, 2004.
53. A. C. Cheng, R. G. Coleman, K. T. Smyth, Q. Cao, P. Soulard, D. R.
Caffrey, A. C. Salzberg and E. S. Huang, Nat. Biotechnol., 2007, 25,
71–75.
54. S. Carney, Drug Discov. Today, 2005, 10, 1012–1013.
55. W. T. Loging, L. Harland and B. Williams-Jones, Nat. Rev. Drug Discov.,
2007, 6, 220–230.
56. C. M. Dobson, Nature, 2004, 432, 824–828.
57. C. A. Lipinski and A. L. Hopkins, Nature, 2004, 432, 855–861.
58. J. Rosen, J. Gottfries, S. Muresan, A. Backlund and T. I. Oprea, J. Med.
Chem., 2009, 52, DOI: 10.1021/jm801514w.
59. M. Gualtieri, F. Banéres-Roquet, P. Villain-Guillot, M. Pugnière and
J.-P. Leonetti, Curr. Med. Chem., 2009, 16, 390–393.
60. Y. Yamanishi, M. Araki, A. Gutteridge, W. Honda and M. Kanehisa,
Bioinformatics, 2008, 24, i232–i240.
61. M. Vieth, M. G. Siegel, R. E. Higgs, I. A. Watson, D. H. Robertson,
K. A. Savin, G. L. Durst and P. A. Hipskind, J. Med. Chem., 2004, 47,
224–232.
62. M. Feher and J. M. Schmidt, J. Chem. Inf. Comput. Sci., 2003, 43,
218–227.
63. M. Vieth and J. J. Sutherland, J. Med. Chem., 2006, 49, 3451–
3453.
64. M. S. Lajiness, M. Vieth and J. Erickson, Curr. Opin. Drug Discov. Dev.,
2004, 7, 470–477.
65. C. A. Lipinski, F. Lombardo, B. W. Dominy and P. J. Feeney, Adv. Drug
Del. Rev., 2001, 46, 3–26.
66. M. C. Wenlock, R. P. Austin, P. Barton, A. M. Davis and P. D. Leeson,
J. Med. Chem., 2003, 46, 1250–1256.
67. A. V. Veselovsky and A. I. Archakov, Curr. Comput. Aided Drug Des.,
2007, 3, 51–58.
68. T. I. Oprea, A. M. Davis, S. J. Teague and P. D. Leeson, J. Chem. Inf.
Comput. Sci., 2001, 41, 1308–1315.
69. M. M. Hann, A. R. Leach and G. Harper, J. Chem. Inf. Comput. Sci.,
2001, 41, 856–864.
70. M. J. Waring, Bioorg. Med. Chem. Lett., 2009, 19, 2844–2851.
71. M. C. Hutter, J. Chem. Inf. Model., 2006, 47, 186–194.
72. M. Wagener and V. J. van Geerestein, J. Chem. Inf. Comput. Sci., 2000,
40, 280–292.
73. T. J. Ritchie, C. N. Luscombe and S. J. F. Macdonald, J. Chem. Inf.
Model., 2009, 49, 1025–1032.
74. J. A. Wells and C. L. McClendon, Nature, 2007, 450, 1001–1009.
56 Chapter 1
75. J. C. Fuller, N. J. Burgoyne and R. M. Jackson, Drug Discov. Today,
2008, 14, 155–161.
76. N. J. Burgoyne and R. M. Jackson, Bioinformatics, 2006, 22, 1335–
1342.
77. C. G. Wermuth, J. Med. Chem., 2004, 47, 1303–1314.
78. S. Van Gestel and V. Schuermans, Drug Dev. Rev., 1986, 8, 1–13.
79. M. Vieth, R. E. Higgs, D. H. Robertson, M. Shapiro, E. A. Gragg and
H. Hemmerle, Biochim. Biophys. Acta, 2004, 1697, 243–257.
80. M. Vieth, J. J. Sutherland, D. H. Robertson and R. M. Campbell, Drug
Discov. Today, 2005, 10, 839–846.
81. J. D. Hughes, J. Blagg, D. A. Price, S. Bailey, G. A. DeCrescenzo, R. V.
Devraj, E. Ellsworth, Y. M. Fobian, M. E. Gibbs, R. W. Gilles,
N. Greene, E. Huang, T. Krieger-Burke, J. Loesel, T. Wager, L. Whiteley
and Y. Zhang, Bioorg. Med. Chem. Lett., 2008, 18, 4872–4875.
82. A. F. Fliri, W. T. Loging, P. F. Thadeio and R. A. Volkmann, Proc. Natl.
Acad. Sci. U.S.A., 2005, 102, 261–266.
83. C. M. Krejsa, D. Horvath, S. L. Rogalski, J. E. Penzotti, B. Mao,
F. Barbosa and J. C. Migeon, Curr. Opin. Drug Disc. Dev., 2003, 6,
470–480.
84. A. F. Fliri, W. T. Loging, P. F. Thadeio and R. A. Volkmann, Nat. Chem.
Biol., 2005, 1, 389–397.
85. A. F. Fliri, W. T. Loging, P. F. Thadeio and R. A. Volkmann, J. Med.
Chem., 2005, 48, 6918–6925.
86. M. Congreve, C. W. Murray and T. L. Blundell, Drug Discov. Today,
2005, 10, 895–907.
87. T. A. Halgren, J. Chem. Inf. Model., 2009, 49, 377–389.
88. E. Hedon and J. Arrons, C. R. Hebd. Seances Acad. Sci., 1899, 129,
778.
89. J. Fischer and C. R. Ganellin, Analogue-based Drug Discovery, Wiley-
VCH, Weinheim, 2006.
90. G. M. Maggiora, J. Chem. Inf. Model., 2006, 46, 1535–1535.
91. E. J. Ariens, A. J. Beld, J. F. Rodrigues de Miranda and A. M. Simonis, in
The Receptors: A Comprehensive Treatise, ed. R. D. O’Brien, Plenum
Press, New York, 1979, vol. 1, pp. 33–91.
92. R. W. DeSimone, K. S. Currie, S. A. Mitchell, J. W. Darrow and D. A.
Pippin, Comb. Chem. High Throughput Screening, 2004, 7, 473–493.
93. M. G. Bock, R. M. DiPardo, K. E. Rittle, B. E. Evans, R. M. Freidinger,
D. F. Veber, R. S. L. Chang, T. B. Chen, M. E. Keegan and V. J. Lotti, J.
Med. Chem., 1986, 29, 1941–1945.
94. B. E. Evans, M. G. Bock, K. E. Rittle, R. M. Di Pardo, W. L. Whitter, D.
F. Veber, P. S. Anderson and R. M. Freidinger, Proc. Natl. Acad. U.S.A.,
1986, 83, 4918–4922.
95. B. E. Evans, K. E. Rittle, M. G. Bock, R. M. DiPardo, R. M. Freidinger,
W. L. Whitter, G. F. Lundell, D. F. Veber, P. S. Anderson, R. S. L.
Chang, V. J. Lotti, D. J. Cerino, T. B. Chen, P. J. Kling, K. A. Kinkel, J.
P. Speinger and J. Hirshfield, 1988, 31, 2235–2246.
Drugs and their Structural Motifs 57
96. A. A. Patchett and R. P. Nargund, Annu. Rep. Med. Chem., 2000, 35,
289–298.
97. G. W. Bemis and M. A. Murcko, J. Med. Chem., 1996, 39, 2887–2893.
98. G. Muller, Drug Discov. Today, 2003, 8, 681–691.
99. D. M. Schnur, Abstracts, 36th Middle Atlantic Regional Meeting of the
American Chemical Society, Princeton, NJ, June 8–11, 2003, 19.
100. D. M. Schnur, M. A. Hermsmeier and A. J. Tebben, J. Med. Chem., 2006,
49, 2000–2009.
101. T. Guo and D. W. Hobbs, Assay Drug Dev. Technol., 2003, 1,
579–592.
102. G. W. Bemis and M. A. Murcko, J. Med. Chem., 1999, 42, 5095–5099.
103. X. Q. Lewell, D. Judd, S. Watson and M. Hann, J. Chem. Inf. Comput.
Sci., 1998, 38, 511–522.
104. P. J. Hajduk, M. Bures, J. Praestgaard and S. W. Fesik, J. Med. Chem.,
2000, 43, 3443–3447.
105. M. G. Siegel and M. Vieth, Drug Discov. Today, 2006, 12, 71–79.
106. J. J. Sutherland, R. E. Higgs, I. Watson and M. Vieth, J. Med. Chem.,
2008, 51, 2689–2700.
107. E. Gianti and L. Sartori, J. Chem. Inf. Model., 2008, 48, 2129–2139.
108. C. D. Duarte, E. J. Barreiro and C. A. M. Fraga, Mini Rev. Med. Chem.,
2007, 7, 1108–1119.
109. A. M. Aronov, B. McClain, C. Stuver Moody and M. A. Murcko, J. Med.
Chem., 2008, 51, 1214–1222.
110. R. S. L. Chang, V. J. Lotti, R. L. Monaghan, J. Birnbaum, E. O. Stapley,
M. A. Goetz, G. Albers-Schonberg, A. A. Patchett and J. M. Liesch,
et al., Science, 1985, 230, 177–179.
111. M. G. Bock, R. M. DiPardo, B. E. Evans, K. E. Rittle, D. F. Veber, R. M.
Freidinger, R. S. L. Chang and V. J. Lotti, J. Med. Chem., 1988, 31,
176–181.
112. M. G. Bock, R. M. DiPardo, B. E. Evans, K. E. Rittle, W. L. Whitter,
V. M. Garsky, K. F. Gilbert, J. L. Leighton and K. L. Carson, et al.,
J. Med. Chem., 1993, 36, 4276–4292.
113. M. G. Bock, R. M. DiPardo, B. E. Evans, K. E. Rittle, W. L. Whitter,
D. F. Veber, R. M. Freidinger, R. S. L. Chang, T. B. Chen and V. J. Lotti,
J. Med. Chem., 1990, 33, 450–455.
114. L. H. Sternbach, J. Med. Chem., 1979, 22, 1–7.
115. D. Roemer, H. H. Buescher, R. C. Hill, R. Maurer, T. J. Petcher,
H. Zeugner, W. Benson, E. Finner, W. Milkowski and P. W. Thies,
Nature, 1982, 298, 759–760.
116. A. Walser, T. Flynn, C. Mason, H. Crowley, C. Maresca, B. Yaremko
and M. O’Donnell, J. Med. Chem., 1991, 34, 1209–1221.
117. D. R. Armour, N. M. Aston, K. M. L. Morriss, M. S. Congreve, A. B.
Hawcock, D. Marquess, J. E. Mordaunt, S. A. Richards and P. Ward,
Bioorg. Med. Chem. Lett., 1997, 7, 2037–2042.
118. R. S. McDowell, B. K. Blackburn, T. R. Gadek, L. R. McGee,
T. Rawson, M. E. Reynolds, K. D. Robarge, T. C. Somers, E. D.
58 Chapter 1
Thorsett, M. Tischler, R. R. Webb II and M. C. Venuti, J. Am. Chem.
Soc., 1994, 116, 5077–5083.
119. K. D. Hargrave, J. R. Proudfoot, K. G. Grozinger, E. Cullen, S. R.
Kapadia, U. R. Patel, V. U. Fuchs, S. C. Mauldin and J. Vitous, et al., J.
Med. Chem., 1991, 34, 2231–2241.
120. J. T. Hunt, C. Z. Ding, R. Batorsky, M. Bednarz, R. Bhide, Y. Cho, S.
Chong, S. Chao, J. Gullo-Brown, P. Guo, S. H. Kim, F. Y. F. Lee, K.
Leftheris, A. Miller, T. Mitt, M. Patel, B. A. Penhallow, C. Ricca, W. C.
Rose, R. Schmidt, W. A. Slusarchyk, G. Vite and V. Manne, J. Med.
Chem., 2000, 43, 3587–3595.
121. W. C. Ripka, G. V. De Lucca, A. C. Bach II, R. S. Pottorf and J. M.
Blaney, Tetrahedron, 1993, 49, 3609–3628.
122. W. C. Ripka, G. V. De Lucca, A. C. Bach II, R. S. Pottorf and J. M.
Blaney, Tetrahedron, 1993, 49, 3593–3608.
123. R. A. Fecik, K. E. Frank, E. J. Gentry, S. R. Menon, L. A. Mitscher and
H. Telikepalli, Med. Res. Rev., 1998, 18, 149–185.
124. IG Farbenindustrie, Ger. Pat., 607537, 1935.
125. W. Sneader, Drug Discovery: A History, Wiley, Chichester, 2005.
126. E. Fromm and J. Wittmann, Ber., 1909, 41, 2264–2273.
127. G. Wozel, Int. J. Dermatol., 1989, 28, 17–21.
128. T. Mann and D. Keilin, Nature, 1940, 146, 164–168.
129. P. W. Feit, J. Med. Chem., 1971, 14, 432–439.
130. G. Erhart, Naturwissenschaften, 1956, 43, 93.
131. F. J. Marshall and M. V. J. Sigal, J. Org. Chem., 1980, 23, 927–929.
132. E. van der Horst, Y. Okuno, A. Bender and A. P. Ijzerman, J. Chem. Inf.
Model., 2009, 49, 348–360.
133. D. A. Price, D. Armour, M. de Groot, D. Leishman, C. Napier, M.
Perros, B. L. Stammen and A. Wood, Bioorg. Med. Chem. Lett., 2006, 16,
4633–4637.
134. T. D. Penning, J. J. Talley, S. R. Bertenshaw, J. S. Carter, P. W. Collins,
S. Docter, M. J. Graneto, L. F. Lee, J. W. Malecha, J. M. Miyashiro, R.
S. Rogers, D. J. Rogier, S. S. Yu, G. D. Anderson, E. G. Burton, J. N.
Cogburn, S. A. Gregory, C. M. Koboldt, W. E. Perkins, K. Seibert, A. W.
Veenhuizen, Y. Y. Zhang and P. C. Isakson, J. Med. Chem., 1997, 40,
1347–1365.
135. P. Prasit, Z. Wang, C. Brideau, C. C. Chan, S. Charleson, W. Cromlish,
D. Ethier, J. F. Evans, A. W. Ford-Hutchinson, J. Y. Gauthier, R.
Gordon, J. Guay, M. Gresser, S. Kargman, B. Kennedy, Y. Leblanc, S.
Leger, J. Mancini, G. P. O’Neill, M. Ouellet, M. D. Percival, H. Perrier,
D. Riendeau, I. Rodger, P. Tagari, M. Therien, P. Vickers, E. Wong, L. J.
Xu, R. N. Young, R. Zamboni, S. Boyce, N. Rupniak, M. Forrest, D.
Visco and D. Patrick, Bioorg. Med. Chem. Lett., 1999, 9, 1773–1778.
136. A. Cuenda, J. Rouse, Y. N. Doza, R. Meier, P. Cohen, T. F. Gallagher, P.
R. Young and J. C. Lee, FEBS Lett., 1995, 364, 229–233.
137. K. H. Bleicher, H. -J. Böhm, K. Müller and A. I. Alanine, Nat. Rev. Drug
Discov., 2003, 2, 369–378.
Drugs and their Structural Motifs 59
138. S. J. F. Macdonald, M. D. Dowle, L. A. Harrison, P. Shah, M. R.
Johnson, G. G. Inglis, G. D. Clarke, R. A. Smith, D. Humphreys, C. R.
Molloy, A. Amour, M. Dixon, G. Murkitt, R. E. Godward, T. Padfield,
T. Skarzynski, O. M. Singh, K. A. Kumar, G. Fleetwood, S. T. Hodgson,
G. W. Hardy and H. Finch, Bioorg. Med. Chem. Lett., 2001, 11, 895–
898.
139. K. A. Jacobson, Drug Dev. Res., 2001, 52, 178–186.
140. M. W. Karaman, S. Herrgard, D. K. Treiber, P. Gallant, C. E. Atteridge,
B. T. Campbell, K. W. Chan, P. Ciceri, M. I. Davis, P. T. Edeen, R.
Faraoni, M. Floyd, J. P. Hunt, D. J. Lockhart, Z. V. Milanov, M. J.
Morrison, G. Pallares, H. K. Patel, S. Pritchard, L. M. Wodicka and P. P.
Zarrinkar, Nat. Biotech., 2008, 26, 127–132.
141. G. Yoshino, T. Kazumi, T. Kasama, I. Iwatani, M. Iwai, A. Inui,
M. Otsuki and S. Baba, Diabetes Res. Clin. Pract., 1986, 2, 179–181.
142. A. Endo, M. Kuroda and Y. Tsuyita, J. Antibiot., 1976, 29, 1346–
1348.
143. A. Endo, Y. Tsuyita, M. Kuroda and K. Tanzawa, Eur. J. Biochem., 1977,
77, 31–36.
144. N. Serizawa, K. Nakagawa and K. Hamano, J. Antibiot., 1983, 36,
604–607.
145. D. J. Triggle, Biochem. Pharmacol., 2009, 78, 217–223.
146. R. A. Wiley and D. H. Rich, Med. Res. Rev., 1993, 13, 327–384.
147. D. A. Smith and E. F. Schmid, Curr. Opin. Drug Discov. Dev., 2006, 9,
38–46.
148. M. T. Keating and M. C. Sanguinetti, Cell, 2001, 104, 569–580.
149. D. A. Price, D. Armour, M. de Groot, D. Leishman, C. Napier,
M. Perros, B. L. Stammen and A. Wood, Curr. Topics Med. Chem., 2008,
8, 1140–1151.
150. J. Scheiber, J. L. Jenkins, S. C. K. Sukuru, A. Bender, D. Mikhailov,
M. Milik, K. Azzaoui, S. Whitebread, J. Hamon, L. Urban, M. Glick and
J. W. Davies, J. Med. Chem., 2009, 52, 3103–3107.
151. J. Blagg, Annu. Rep. Med. Chem., 2006, 41, 353–368.
152. I. Walker and H. Newell, Nat. Rev. Drug Discov., 2008, 8, 15–16.
153. N. N. Taleb, The Black Swan: The Impact of the Highly Improbable,
Random House, New York, 2007.
154. A. Bakker, A. Caricasole, G. Gaviraghi, G. Pollio, G. Robertson, G. C.
Terstappen, M. Salerno and P. Tunici, ChemMedChem, 2009, 4, 1–12.
155. A. Bender, J. Scheiber, M. Glick, K. Azzaoui, J. Hamon, L. Urban,
S. Whitebread and J. L. Jenkins, ChemMedChem, 2007, 2, 861–873.
156. K. Azzaoui, J. Hamon, B. Faller, S. Whitebread, E. Jacoby, A. Bender,
J. L. Jenkins and L. Urban, ChemMedChem, 2007, 2, 874–880.
157. M. Waters, G. Boorman, P. Bushel, M. Cunningham, R. Irvin, A. Merrick,
K. Olden, R. Paules, J. Selkirk, S. Stasiewicz, B. Weis, B. van Houten,
N. Walker and R. Tennant, Environ. Health Perspect., 2003, 111,
811–824.
158. G. Brambilla and A. Martelli, Mutat. Res., 2008, 681, 209–229.
60 Chapter 1
159. E. David, T. Tramontin and R. Zemmel, Nat. Rev. Drug Discov., 2009, 8,
609–610.
160. A. L. Hopkins and A. Polinsky, Annu. Rep. Med. Chem., 2006, 41,
425–437.
161. M. Goodman, Nat. Rev. Drug Discov., 2008, 7, 795.
162. A. Hopkins, J. Lanfear, C. Lipinski and L. Beeley, Annu. Rep. Med.
Chem., 2005, 40, 349–358.
163. T. I. Oprea, Towards Drugs of the Future. Proceedings of the Solvay
Pharmaceuticals Symposium, C. G. Kruse, H. Timmerman, eds. IOS
Press, Amsterdam, 2008, 29–36.
CHAPTER 2
2.1 Introduction
In order to become an effective therapy against a disease, a compound must
exhibit potency versus a particular therapeutic target, combined with a degree of
selectivity over other targets; this drives to an appropriate safety margin. In
addition, the vast majority of drugs are delivered at sites which are remote from
the site of action. Over millions of years of human evolution, the body has
developed a host of defence mechanisms designed to protect against exogenous
substances that may cause harm. These mechanisms now form barriers to the
passage of therapeutic drugs from their site of administration to their site of
action. The extent to which a drug can avoid these barriers will to a great extent
determine the therapeutic potential of that particular compound. Successful drugs
need to strike a balance between three major determinants of therapeutic potential:
61
62 Chapter 2
3) Absorption, distribution, metabolism and excretion (ADME) to drive to
an acceptable dose and frequency of administration.
DG ¼ RT ln Kd ð2:1Þ
bioavailability of an intravenous dose is always 100% and all other dose routes
must be measured relative to an intravenous dose.
Intravenous administration gives complete bioavailability and the potential
to rapidly achieve therapeutic concentrations. A good example of a drug that is
given by intravenous infusion is lidocaine, which is used to treat severe neu-
ropathic pain in the clinical setting.9
There are several important considerations for a drug to be administered by
the intravenous route. First, it must be sufficiently soluble in the proposed
formulation (mainly saline with limited potential for organic co-solvents) to
deliver the whole dose in an acceptable volume. Secondly, the compound must
have prolonged chemical stability in the formulation, especially as the dose
must be taken through sterilisation procedures. Overall, the requirement for
venous cannulation means that intravenous administration tends to be limited
to severe and life-threatening indications, and usually under qualified medical
supervision.
Efflux
Transporter Liver
Paracellular
Absorption
Dissolution
Transcellular
Absorption
Degradation Drug in
Solution
Influx
Transporter Metabolism
Blood
Enterocyte
Plasma
Site of Measurement
Gut Metabolism
Excretion to Faeces
Figure 2.1 Summary of the barriers to drug delivery following oral administration.
2.2.1.1 Dissolution
To be absorbed a drug must be in solution in the gastrointestinal tract lumen.
The fluids in the lumen are predominantly aqueous in nature with a pH ranging
from acidic to physiological. Consequently, an important parameter for the
medicinal chemist to modulate within a chemical series is the aqueous
solubility.
Oral drugs tend to be administered in tablet or capsule formulation. The
formulation must first disintegrate before the active drug undergoes a dis-
solution process. These processes are governed in part by excipients in the
formulation and in part by the physical form and aqueous solubility of the
formulated drug.
Food may affect the absorption of drugs through the dissolution process.
The delay in gastric emptying with food after a meal may result in a longer time
to achieve maximal plasma concentrations, since the drug is held in the stomach
for longer than in the fasted state. The release of bile salts on ingestion of food
may result in enhanced dissolution of poorly soluble drugs. In addition, certain
compounds may form complexes with food resulting in a reduction in their
absorption.10
HO O
O
N
O Cefadroxil
OH
S N
H H
H2N
O O
N N
H2N N N H2N N N
H O H O
O
O OH
Valacyclovir Acyclovir
NH2
Cl
Cl
O N N
N NH2
N S
O
O
O N
O HN N
N
N
O S O
N
(a) Gut Wall Metabolism. The small intestine contains enzymes that can
metabolise drugs through both phase 1 (oxidative) and 2 (conjugation) reac-
tions.25,26 The involvement of the gut wall in the metabolism of drugs has
been extensively reviewed.27,28
A range of P450 enzymes have been identified in the human small intestine
including CYP3A4, 3A5, 2C9, 2C19, 2J2 and 2D6. CYP3A is the most highly
expressed CYP accounting for around 80% of the total CYP content.29 Even
72 Chapter 2
though the total mass of CYP3A in the intestine is only 1% of that in the liver,
clinical studies have demonstrated that intestinal metabolism can have a sig-
nificant effect on the overall first-pass metabolism of some drugs.30,31
The oral bioavailability of midazolam in humans has been estimated at
approximately 30%.32 This is significantly lower than would be expected from
the plasma clearance of midazolam (approximately 5 ml min1 kg1) with
respect to human hepatic blood flow, knowing that midazolam is completely
absorbed. When midazolam was administered to patients undergoing hepatic
transplantation,33 the gut wall first-pass extraction following intraduodenal
administration was calculated to be 0.43 (range 0.14 to 0.59). Thus, the oral
bioavailability of midazolam is a consequence of complete oral absorption and
loss of approximately 40% of the drug on first-pass through the gut wall, with
the remainder lost by hepatic first-pass extraction.
The gut wall first-pass extraction of drug molecules is thought to be enhanced
by the expression of CYP3A4 and P-gp together close to the apical membrane
of the enterocyte. It is thought that these two proteins can act as a concerted
barrier to the passage of their substrates across the enterocyte. The P-gp
intercepts its substrates on passage through the enterocyte and effluxes them
back into the gut lumen. Thus, the drug is cycled in and out of the enterocyte,
allowing CYP3A4 a number of opportunities for metabolism. In this way, it is
possible for a relatively low expression of enzyme to exert a significant gut wall
first-pass extraction.
(b) Hepatic. The entire blood supply of the upper gastrointestinal tract pas-
ses into the hepatic portal vein which flows directly to the liver. Conse-
quently, drug that has crossed the enterocyte and avoided gut wall
metabolism is also subject to first-pass extraction by the liver. Extraction of
drugs by the liver is discussed in detail later.
F ¼ Fa Fg Fh ð2:2Þ
Figure 2.6 Plasma concentration versus time curve for an intravenous dose and
corresponding schematic of drug distribution at early times following
intravenous administration.
to take place. This will occur by the law of mass action down a concentration
gradient. At this point, the rate into the tissues will be greater than the rate of
return into the blood. When tissue distribution and blood clearance (see below)
is taken into consideration, the concentrations in the blood will fall rapidly over
this period.
At some point following intravenous administration (C in Figure 2.6), the
unbound concentrations in the tissue will be in equilibrium with the unbound
78 Chapter 2
concentrations in the blood. This is termed the steady state of distribution.
After this point, tissue distribution is complete and plasma concentrations
decline solely as a result of clearance from the blood. Extrapolation of this
phase back to the concentration axis will give the concentration of the drug at
time zero (Co) or the blood concentration if all the drug was distributed
instantaneously following dosing. This is related to the distribution volume
(Vd) and the dose by eqn (2.3):
Co ¼ Dose=Vd ð2:3Þ
Figure 2.7 Relationship between logD(7.4) and Vd for acids, bases and neutral
compounds.
Absorption
Clearance
750
650
550
450
MW
350
250
H
N
Cl
NH N
H
N N N
S S Cl S
O O Cl
O O
OMe
SB 271046 699929
Brain:Blood = 0.05:1 Brain:Blood = 3:1
Css;avg CL t
Dose ¼ ð2:4Þ
F
where:
2.4.1 Clearance
The clearance of a molecule is defined as the amount of blood (blood clea-
rance) or plasma (plasma clearance) cleared of drug per unit time and body
weight. The units of clearance are ml min1 kg1. Clearance is dependent
upon the ability of organs to metabolise or excrete the compound, the
plasma protein binding of the molecule and the blood flow to the clearing
organ. The importance of clearance as a parameter is that, for a given dose, it
defines the exposure (or AUC) to a given compound as determined by
ADMET for the Medicinal Chemist 83
eqn (2.5):
Dose
Clearance ¼ ð2:5Þ
AUC
Thus, a compound with a low clearance will always exhibit a high exposure
at a given dose, relative to a compound with a higher clearance.
The free drug hypothesis suggests that therapeutic efficacy is driven by the
unbound concentrations of the drug. Thus, unbound clearance is the plasma
clearance corrected for the fraction unbound in plasma.
In addition, the potential amount of blood cleared of drug in the absence of
flow limitations and plasma protein binding considerations is defined as the
unbound intrinsic clearance (Clintu) of that particular molecule. For metabo-
lised compounds, this is an important parameter as it is related to the Vmax
and Km of the enzyme for metabolism of any particular compound. This is the
parameter that medicinal chemists attempt to reduce when required to modify
the plasma clearance of any particular molecule. Clintu is related to the clear-
ance by the well stirred model as described by eqn (2.6):
fu Clintu
Cl ¼ ð2:6Þ
fuðClintu þ QÞ
where fu is the fraction unbound in plasma and Q is the hepatic blood flow.
However, since many drugs are cleared by more than one organ and single
human organs cannot be studied in isolation, the clearance of a drug is always
calculated from the plasma concentration versus time curve using eqn (2.5).
84 Chapter 2
Clearance is most often calculated following intravenous administration
because the total dose of the drug is known to be delivered into the blood.
However, most drugs are delivered at sites remote from the blood, with the most
popular route being oral administration. As stated earlier, following oral admin-
istration the drug must overcome the barriers of dissolution, absorption, gut wall
metabolism and hepatic first-pass metabolism, dissolve in the gastrointestinal tract
lumen contents and cross the gastrointestinal tract enterocyte membrane.
Thus, at a number of points between the site of administration and the blood,
there is capacity for an oral drug to be lost. Consequently, the oral clearance
will often be higher than the intravenous clearance. The unifying factor is the
fraction bioavailable (F). Clearance is calculated using eqn (2.8).
DoseðoralÞ
ClearanceðoralÞ ¼ ð2:8Þ
AUCðoralÞ F
Drugs can be cleared by a variety of organs in the human body. However, the
major clearing organs are the liver and the kidney.
2.4.3 Metabolism
2.4.3.1 Cytochrome P450
The family of enzymes that metabolise the majority of drugs are the cytochrome
P450s (CYPs). These are a super-family of haem-containing enzymes that use
NADPH to catalyse the single electron oxidation of substrates. The overall
family consists of several hundred isoforms, with a variety of endogenous roles.
ADMET for the Medicinal Chemist 85
Mitochondria Sinusoidal
(MAO) Membrane Cytoplasm
Canalicular (Aldehyde
Bile
Membrane Oxidase,
Canaliculus
Sulfotransferases,
Amidases,
Esterases)
However, there are only a small number of major human drug metabolising
CYPs, with the most important being CYP3A4, CYP2C9 and CYP2D6.
The CYP enzymes are expressed in many tissues, but they are found in
highest concentrations in the hepatocyte. They are membrane-bound enzymes
found within the cell on the endoplasmic reticulum. Human liver microsomes
are preparations of endoplasmic reticulum made by differential centrifugation
of human liver homogenates. On preparation, these intracellular membranes
form into spheres that express CYPs on their inner and outer surfaces.
The substrate structure metabolism relationships for these major drug meta-
bolising CYPs have been extensively reviewed.49,66 CYP2D6 will tend to meta-
bolise compounds with a metabolically vulnerable site a certain distance away
from a basic centre. Since CYP2D6 is polymorphically expressed across the
human population (6–8% of the Caucasian population do not express CYP2D6),
compounds that are predominantly metabolised by this isoform will exhibit
significant variability in pharmacokinetic profile. For this reason, CYP2D6
substrates tend to be avoided in compound selection. CYP2C9 will tend to
metabolise acids or compounds with a significant degree of hydrogen bonding.
CYP3A4 tends to be more promiscuous, driven by a large active site, and will
metabolise relatively large molecules with no major preference for ionisation.
These CYP isoforms share a common requirement for lipophilic compounds.
In general, CYP metabolism is positively correlated with log D.67
Therefore, across a series of compounds, the most lipophilic analogue will tend
to be more rapidly metabolised than its analogues with lower log D. As oxidative
enzymes, CYPs will tend to oxidise lipophilic compounds at sites of significant
electron density. This points the way to medicinal chemistry strategies to address
CYP metabolism. Clearly, reducing log D is likely to lead to a reduction in
86 Chapter 2
intrinsic clearance observed in human liver microsomes. However, as previously
established, lipophilicity tends to be correlated positively with potency against
the pharmacological target. Therefore, a balance needs to be struck. Perhaps a
more informed strategy to reduce CYP metabolism would be to identify the sites
of metabolic attack and reduce their susceptibility to metabolism by blocking
with metabolically inert groups or by reducing electron density at that position.
In this way, lipophilicity to drive potency can be maintained whilst reducing
metabolic clearance due to CYP metabolism.
It is easily overlooked in compound optimisation (based on human liver
microsomal intrinsic clearance) that the liver contains a vast armoury of other
drug metabolising enzymes. Thus, it is not unusual to find that a successful
campaign to lower the CYP mediated clearance within a series has not reduced
the in vivo clearance as much as expected due to metabolism by other hepatic
enzymes. This has been exemplified by Williams et al.68 A review of the top 200
prescribed drugs suggests that metabolism is responsible for the clearance of
approximately two-thirds. Of these, approximately two-thirds are pre-
dominantly CYP metabolised, suggesting that the emphasis on CYP-mediated
metabolism is well placed. However, there is approximately a further third that
are substrates for metabolic clearance mediated by enzymes other than CYPs.
Of these, the most prevalent are UDP-glucuronosyltranferases and esterases,
representing approximately 8% and 5% of the metabolised drugs, respectively.
More minor contributions come from flavin monooxygenases (FMOs),
monoamine oxidases (MAOs) and N-acetyltransferases.
+
FMO3
N O-
N
N
N
Nicotine Nicotine N-oxide
H O COOH
N H O
S MAO N
S
O N O
H N
H
Sumatriptan Sumatriptan
indole acetic acid
OH
O O
N N
N Aldehyde Oxidase N
O O
N N
H H
O N O N
O O
2.4.3.5 Hydrolases
Hydrolases encompass a wide range of enzymes that use water to break-
down their substrates. In general, they catalyse the reaction exemplified by
Figure 2.17. Hydrolases comprise a huge family of enzymes (including estera-
ses, amidases and peptidases) that have been thoroughly reviewed in terms of
classification, mechanism of action and structure–activity relationship (SAR)
by Testa and Krämer.78
Probably the most important group of hydrolases from a drug metabolism
and prodrug point of view are the esterases including the cholinesterases
O O
HN HN
Cl Cl
Aldehyde Oxidase
Ziprasidone
N N
N N
N N
S
SH
O O
R H H R H
R X + O R OH + X
OH O
O H H
+ O +
OH OH
OH
O O
N N
HN O HN
N NAT2
N + CoA SH
N + CoA S N
N N N N
NH N
O
UK-469,413 N-acetylated UK-469,413
References
1. E. H. Kerns and L. Di, Drug Discov. Today, 2003, 8, 316.
2. C. A. Lipinski, F. Lombardo, B. W. Dominy and P. J. Feeney, Adv. Drug
Deliv. Rev, 1997, 23, 3.
ADMET for the Medicinal Chemist 95
3. K. Beaumont, E. Schmid and D. A. Smith, Bioorg. Med. Chem. Lett,
2005, 15, 3658.
4. D. A. Smith, Curr. Opin. Drug Discov. Dev, 2007, 10, 550.
5. G. C. Williams and P. J. Sinko, Adv. Drug Deliv. Rev, 1999, 39, 211.
6. K. Beaumont, R. Webster, I. Gardner and K. Dack, Curr. Drug Metab,
2003, 4, 461.
7. A. L. Hopkins, C. R. Groom and A. Alex, Drug Discov. Today, 2004, 9,
430.
8. P. D. Leeson and B. Springthorpe, Nat. Rev. Drug Discov, 2007, 6, 881.
9. F. M. Ferrante, J. Paggioli, S. Cherukuri and G. R. Arthur, Anesth.
Analg. (Baltimore), 1996, 82, 91.
10. B. N. Singh, Clin. Pharmacokinet, 1999, 37, 213.
11. Y.-L. He, S. Murby, G. Warhurst, L. Gifford, D. Walker, J. Ayrton,
R. Eastmond and M. Rowland, J. Pharm. Sci., 1998, 87, 626.
12. P. R. Reeves, D. J. Barnfield, S. Longshaw, D. A. D. McIntosh and M. J.
Winrow, Xenobiotica, 1978, 8, 305.
13. P. R. Reeves, J. McAinsh, D. A. D. McIntosh and M. J. Winrow,
Xenobiotica, 1978, 8, 313.
14. M. Brandsch, I. Knuetter and E. Bosse-Doenecke, J. Pharm. Pharmacol,
2008, 60, 543.
15. C. Hilgendorf, G. Ahlin, A. Seithel, P. Artursson, A.-L. Ungell and
J. Karlsson, Drug Metab. Dispos., 2007, 35, 1333.
16. B. Bretschneider, M. Brandsch and R. Neubert, Pharm. Res., 1999, 16, 55.
17. M. E. Ganapathy, W. Huang, H. Wang, V. Ganapathy and F. H.
Leibach, Biochem. Biophys. Res. Commun, 1998, 246, 470.
18. S. Weller, M. R. Blum, M. Doucette, T. Burnette, D. M. Cederberg, P. D.
Miranda and M. L. Smiley, Clin. Pharmacol. Ther, 1993, 54, 595.
19. F. Thiebaut, T. Tsuruo, H. Hamada, M. M. Gottesman, I. Pastan and
M. C. Willingham, Proc. Natl. Acad. Sci. U.S.A, 1987, 84, 7735.
20. A. Ayrton and P. Morgan, Xenobiotica, 2008, 38, 676.
21. K. Beaumont, A. Harper, D. A. Smith and S. Abel, Xenobiotica, 2000, 30,
627.
22. K. Beaumont, A. Harper, D. A. Smith and J. Bennett, Eur. J. Pharm. Sci,
2000, 12, 41.
23. S. Abel, K. C. Beaumont, C. L. Crespi, M. D. Eve, L. Fox, R. Hyland,
B. C. Jones, G. J. Muirhead, D. A. Smith, R. F. Venn and D. K. Walker,
Xenobiotica, 2001, 31, 665.
24. K. Westphal, A. Weinbrenner, T. Giessmann, M. Stuhr, G. Franke,
M. Zschiesche, R. Oertel, B. Terhaag, H. K. Kroemer and W. Siegmund,
Clin. Pharmacol. Ther, 2000, 68, 6.
25. T. Prueksaritanont, L. M. Gorham, J. H. Hochman, L. O. Tran and K. P.
Vyas, Drug Metab. Dispos, 1996, 24, 634.
26. J. H. Lin, M. Chiba and T. A. Baillie, Pharmacol. Rev., 1999, 51, 135.
27. K. Beaumont, in Methods and Principles in Medicinal Chemistry, ed.
H. van de Waterbeemd, H. Lennernäs and P. Artursson, 2003, vol. 18
(Drug Bioavailability), ch. 13, pp. 311–328.
96 Chapter 2
28. U. Fagerholm, J. Pharm. Pharmacol, 2007, 59, 1335.
29. M. F. Paine, H. L. Hart, S. S. Ludington, R. L. Haining, A. E. Rettie and
D. C. Zeldin, Drug Metab. Dispos, 2006, 34, 880.
30. J. Yang, T. Tucker Geoffrey and A. Rostami-Hodjegan, Clin Pharmacol
Ther, 2004, 76, 391.
31. M. F. Paine, M. Khalighi, E. M. Fisher, D. D. Shen, K. L. Kunze, C. L.
Marsh, J. D. Perkins and K. E. Thummel, J. Pharmacol. Exp. Ther, 1997,
283, 1552.
32. K. E. Thummel, S. D. O., M. F. Paine, D. D. Shen, K. L. Kunze and J. D.
Perkins, Clin. Pharmacol. Ther, 1996, 59, 491.
33. M. F. Paine, D. D. Shen, K. L. Kunze, J. D. Perkins, C. L. Marsh, J. P.
McVicar, D. M. Barr, B. S. Gillies and K. E. Thummel, Clin. Pharmacol.
Ther, 1996, 60, 14.
34. H. Kublik and M. T. Vidgren, Adv. Drug Deliv. Rev., 1998, 29, 157.
35. L. Illum, Trends Biotechnol., 1991, 9, 284.
36. M. Cazzola, R. Testi and M. G. Matera, Clin. Pharmacokinet, 2002, 41,
19.
37. J. S. Patton and P. R. Byron, Nat. Rev. Drug Discov, 2007, 6, 67.
38. P. W. Armstrong, J. A. Armstrong and G. S. Marks, Circulation, 1979,
59, 585.
39. C. T. Lombroso, Epilepsia, 1989, 30(Suppl 2), S11.
40. W. Jeal and P. Benfield, Drugs, 1997, 53, 109.
41. I. Power, Br. J. Anaesth, 2007, 98, 4.
42. R. J. Wills, Clin Pharmacokinet, 1990, 19, 390.
43. G. L. Trainor, Expert Opin. Drug Discov, 2007, 2, 51.
44. W. E. Lindup, Prog. Drug Metab, 1987, 10, 141.
45. P. T. Schoenemann, D. W. Yesair, J. J. Coffey and F. J. Bullock, Ann. N.
Y. Acad. Sci., 1973, 226, 162.
46. Z. Y. Wu, S. E. Cross and M. S. Roberts, J. Pharm. Sci., 1995, 84, 1020.
47. J. Krieglstein, Arzneim.-Forsch., 1973, 23, 1527.
48. P. H. Hinderling, Pharmacol. Rev., 1997, 49, 279.
49. D. A. Smith, B. C. Jones and D. K. Walker, Med. Res. Rev, 1996, 16, 243.
50. F. Lombardo, R. S. Obach, F. M. DiCapua, G. A. Bakken, J. Lu, D. M.
Potter, F. Gao, M. D. Miller and Y. Zhang, J. Med. Chem, 2006, 49, 2262.
51. F. Lombardo, R. S. Obach, M. Y. Shalaeva and F. Gao, J. Med. Chem,
2002, 45, 2867.
52. A. Petrauskas, P. Japertas, R. Didziapetris, K. Lanevskij and D. Bondarev,
Abstracts of Papers, 231st ACS National Meeting, Atlanta, GA, 26–30
March 2006, 2006, MEDI.
53. D. Alker, R. A. Burges, S. F. Campbell, A. J. Carter, P. E. Cross, D. G.
Gardiner, M. J. Humphrey and D. A. Stopher, J. Chem. Soc, Perkin
Trans. 2, 1992, 7, 1137.
54. D. A. Stopher, A. P. Beresford, P. V. Macrae and M. J. Humphrey,
J. Cardiovasc. Pharmacol., 1988, 12, S55.
55. M. S. Whitman and A. R. Tunkel, Infect. Control Hosp. Epidemiol, 1992,
13, 357.
ADMET for the Medicinal Chemist 97
56. P. Poulin and F. -P. Theil, J. Pharm. Sci., 2002, 91, 129.
57. G. E. Blakey, I. A. Nestorov, P. A. Arundel, L. J. Aarons and
M. Rowland, J. Pharmacokinet. Biopharm., 1997, 25, 277.
58. S. S. De Buck, V. K. Sinha, L. A. Fenu, M. J. Nijsen, C. E. Mackie and
R. A. H. J. Gilissen, Drug Metab. Dispos, 2007, 35, 1766.
59. J. P. Tillement, S. Urien, P. Chaumet-Riffaud, P. Riant, F. Bree, D. Morin,
E. Albengres and J. Barre, Fundam. Clin. Pharmacol, 1988, 2, 223.
60. L. A. Peletier, N. Benson and P. H. van der Graaf, J. Theor. Biol., 2009,
256, 253.
61. T. Obradovic, G. G. Dobson, T. Shingaki, T. Kungu and I. J. Hidalgo,
Pharm. Res, 2007, 24, 318.
62. P. H. Elsinga, N. H. Hendrikse, J. Bart, W. Vaalburg and A. Van Waarde,
Curr. Pharm. Des., 2004, 10, 1493.
63. N. Jung, C. Lehmann, A. Rubbert, M. Knispel, P. Hartmann, J. van
Lunzen, H.-J. Stellbrink, G. Faetkenheuer and D. Taubert, Drug Metab.
Dispos., 2008, 36, 1616.
64. M. Ahmed, M. A. Briggs, S. M. Bromidge, T. Buck, L. Campbell, N. J.
Deeks, A. Garner, L. Gordon, D. W. Hamprecht, V. Holland, C. N.
Johnson, A. D. Medhurst, D. J. Mitchell, S. F. Moss, J. Powles, J. T. Seal,
T. O. Stean, G. Stemp, M. Thompson, B. Trail, N. Upton, K. Winborn
and D. R. Witty, Bioorg. Med. Chem. Lett., 2005, 15, 4867.
65. J. C. Kalvass and T. S. Maurer, Biopharm. Drug Dispos., 2002, 23, 327.
66. D. A. Smith, M. J. Ackland and B. C. Jones, Drug Discov. Today, 1997, 2,
479.
67. H. van de Waterbeemd, D. A. Smith and B. C. Jones, J. Comput. Aided
Mol. Des., 2001, 15, 273.
68. J. A. Williams, R. Hyland, B. C. Jones, D. A. Smith, S. Hurst, T. C.
Goosen, V. Peterkin, J. R. Koup and S. E. Ball, Drug Metab. Disp., 2004,
32, 1201.
69. J. R. Cashman, Curr. Drug Metab., 2000, 1, 181.
70. J. R. Cashman and J. Zhang, Annual Rev. Pharmacol. Toxicol., 2006, 46,
65.
71. P. L. Dostert, B. M. Strolin and K. F. Tipton, Med. Res. Rev., 1989, 9, 45.
72. D. E. Edmondson, A. Mattevi, C. Binda, M. Li and F. Hubálek, Curr.
Med. Chem., 2004, 11, 1983.
73. K. F. Tipton, S. Boyce, S. J.O., G. P. Davey and J. Healy, Curr. Med.
Chem., 2004, 11, 1965.
74. C. M. Dixon, G. R. Park and M. H. Tarbit, Biochem. Pharmacol., 1994,
47, 1253.
75. B. Kaye, J. L. Offerman and J. L. Reid, Xenobiotica, 1984, 14, 935.
76. C. Beedham, S. E. Bruce, D. J. Critchley, T. Y. Al and D. J. Rance, Eur. J.
Drug Metab. Pharmacokinet., 1987, 12, 307.
77. C. Beedham, J. J. Miceli and R. S. Obach, J. Clin. Psychopharmacol.,
2003, 23, 229.
78. B. Testa and S. D. Krämer, Chem. Biodivers., 2007, 4, 2031.
79. F. M. Williams, Clin. Pharmacokinet., 1985, 10, 392.
98 Chapter 2
80. M. R. Redinbo and P. M. Potter, Drug Discov. Today, 2005, 10, 313.
81. T. Imai, Drug Metab. Pharmacokinet., 2006, 21, 173.
82. P. I. Mackenzie, K. W. Bock, B. Burchell, C. Guillemette, S. i. Ikushiro,
T. Iyanagi, J. O. Miners, I. S. Owens and D. W. Nebert, Pharmacogenet.
Genomics, 2005, 15, 677.
83. B. Burchell, D. J. Lockley, A. Staines, Y. Uesawa and M. W. H.
Coughtrie, Methods Enzymol., 2005, 400, 46.
84. N. Gamage, A. Barnett, N. Hempel, R. G. Duggleby, K. F. Windmill,
J. L. Martin and M. E. McManus, Toxicolog. Sci., 2006, 90, 5.
85. G. M. Pacifici and M. W. H. Coughtrie, Human Cytosolic Sulfo-
transferases, CRC Press, Boca Raton, FL, 2005.
86. E. Sim, K. Walters and S. Boukouvala, Drug Metab. Rev., 2008, 40, 479.
87. J. Rawal, R. Jones, A. Payne and I. Gardner, Xenobiotica, 2008, 38, 1219.
88. A. Ayrton and P. Morgan, Xenobiotica, 2001, 31, 469.
89. I. Pahlman, M. Edholm, S. Kankaanranta and M. L. Odell, Pharm.
Pharmacol. Commun., 1998, 4, 493.
90. U. Fagerholm, J. Pharm. Pharmacol., 2007, 59, 1463.
91. J. G. Riddell, D. W. Harron and R. G. Shanks, Clin. Pharmacokinet.,
1987, 12, 305.
92. K. W. Brammer, A. J. Coakley, S. G. Jezequel and M. H. Tarbit, Drug
Metab. Dispos., 1991, 19, 764.
93. S. G. Jezequel, J. Pharm. Pharmacol., 1994, 46, 196.
94. S. M. Grant, H. D. Langtry and R. N. Brogden, Drugs, 1989, 37, 801.
95. C. J. Roberts, Clin. Pharmacokinet., 1984, 9, 211.
96. M. V. S. Varma, B. Feng, R. S. Obach, M. D. Troutman, J. Chupka,
R. H. Miller and A. El-Kaman, J. Med. Chem., 2009, 52, 4844.
97. H. Takakusa, H. Masumoto, H. Yukinaga, C. Makino, S. Nakayama,
O. Okazaki and K. Sudo, Drug Metab. Dispos., 2008, 36, 1770.
98. B. Seguin and J. Uetrecht, Curr. Opin. Allergy Clin. Immunol., 2003, 3,
235.
99. N. Kaplowitz, Nat. Rev. Drug Discov., 2005, 4, 489.
100. J. L. Walgren, M. D. Mitchell and D. C. Thompson, Crit. Rev. Toxicol.,
2005, 35, 325.
101. E. C. Miller and J. A. Miller, Pharmacol. Rev., 1966, 18, 805.
102. J. R. Mitchell, D. J. Jollow, W. Z. Potter, D. C. Davis, J. R. Gillette and
B. B. Brodie, J. Pharmacol. Exp. Ther., 1973, 187, 185.
103. J. Uetrecht, Chem. Res. Toxicol., 2008, 21, 84.
CHAPTER 3
3.1 Introduction
Exposing the importance of the carboxylic acid functional group is best achieved
by examining the number of endogenous processes and individual molecules
which rely on the intrinsic chemical nature (e.g. pKa and hydrogen bonding
characteristics) of this functional group. From amino acid conjugation (peptide
synthesis - proteins) and post-translational protein acylation, to triglycerides,
bile acids, prostanoids, messenger molecules and hormone catabolites, it is evident
that the carboxylic acid represents a key functional group contributing to the
biochemistry critical to mammalian physiology. Not surprisingly then, there exists
an extensive number of drugs possessing the carboxylic acid functional group.
The compounds represent a heterogeneous group comprising, among others, non-
steroidal anti-inflammatory drugs (NSAIDs), b-lactam antibiotics, statins,
fibrates, and even food additives such as preservatives and flavouring agents; these
compounds range from hydrophilic to lipophilic organic compounds.
Over 450 drugs containing a free carboxylic acid group are marketed in var-
ious countries worldwide (see Table 3.1 for select examples). In addition to the
99
100
mone (1-84)
Delapril Gemfibrozil Proglumide Nateglinide Pazufloxacin Glycopin
Vigabatrin Clinofibrate Romurtide Ceftibuten Zofenoprilat Alvimopan
Cefminox Carprofen Sermorelin Limaprost Salmeterol Glycyrrhizinic acid
Gabapentin Cefmenoxime Cefalexin Gadopentetate Eprosartan Alatrofloxacin
Epalrestat Triflusal Pranoprofen Unoprostone Febuxostat Alitretinoin
Cefixime Levodopa Frusemide Tranexamic acid Angiotensin II Bexarotene
Alacepril Carbidopa Panipenem Temocapril Fudosteine Etacrynic acid
Enalaprilat Acetylcysteine Daptomycin Ibuprofen alpha-Lipoic acid Thymalfasin
Acrivastine Artesunic acid Tosufloxacin Flurbiprofen Trovafloxacin Enfuvirtide
Sarafloxacin Sulindac Trandolapril Levofloxacin Folinic acid Vedaprofen
Tropesin Tolmetin Acipimox Cetirizine Cefoselis Bendamustine
Ramipril Aspirin Ciprofibrate Gadoterate Sitafloxacin Gemifloxacin
Cefodizime Indomethacin Pseudomonic acid Loracarbef all-trans-Retinoic 4-Aminosalicylic
acid acid
Carboxylic Acids and their Bioisosteres
H-bond Zwitterion
H
O NH2 O NH3
O OH O OH
Figure 3.1 Example of intramolecular hydrogen bonding in the amino acid tyrosine
yielding a zwitterion intermediate.
HCOOH 3.77
(CH3)3CCOOH 5.05
ClCH2COOH 2.86
HOCH2COOH 3.83
Carboxylic Acids and their Bioisosteres 103
O O O O
F F
HO O
N N N N
NH2 NH2
Salicylic acid derivatives Anthranilic acid derivatives Acetic acid derivatives Propionic acid derivatives
O
R O OH O
O O H
CH3
H3CO OH O CH3
N Cl OH
OH CH3 H OH
N HO
N
H 3C
O
CH3 O
Salicylic acid: R = H O Cl Cl
OH O OH
OH O N F O CH3
H
CH3 N OH
N N CH3
H N S OH
N H
S CH3 N O O
S CH3 CH3 O
O O
O O S
O
OH OH O
H3C O O
CO2 OH (CH3CO)2O
OH
pressure
Scheme 3.1 Synthesis of aspirin, the first carboxylic acid containing NSAID.
1874 (Scheme 3.1), which led to the introduction of sodium salicylate in the
treatment for chronic rheumatoid arthritis and gout.
The search for a superior salicylic acid derivative was initiated at Bayer in
1895. Chemist Felix Hoffman, to whom the task was presented, also had a
personal reason for this endeavour; his father had been taking salicylic acid for
many years to treat arthritis and encountered emesis as a major side effect.
Hoffman found a way of acetylating the phenol group of salicylic acid to form
acetylsalicylic acid (aspirin) (see Scheme 3.1).4 After initial laboratory tests,
Hoffman’s father was given the drug; it was pronounced effective and later
confirmed as such in ‘impartial’ clinical trials. The drug was introduced in 1899
with a report suggesting that aspirin was a convenient way of delivering sal-
icylic acid to the body.5 That aspirin is a mere prodrug for salicylic acid has
been debated since its discovery, but as discussed below, aspirin clearly has
potent actions of its own that are not shared by salicylic acid.
It was not until 1971 that Vane and co-workers proposed that the anti-
inflammatory and analgesic properties of NSAIDs are due to inhibition of
prostaglandin biosynthesis, which is catalysed by the enzyme prostaglandin
endoperoxide synthase or cyclooxygenase (COX).6–8 COX catalyses the for-
mation of prostaglandins and thromboxane from the fatty acid substrate ara-
chidonic acid (itself derived from the cellular phospholipid bilayer by the action
of phospholipase A2).9
COX activity originates from two distinct and independently regulated
enzymes, termed COX-1 and COX-2.10,11 COX-1 is the constitutive isoform and
is mainly responsible for the synthesis of cytoprotective prostaglandins in the
gastrointestinal tract. COX-2 is inducible and short-lived; its expression is sti-
mulated in response to pro-inflammatory mediators and the isozyme plays a
major role in prostaglandin biosynthesis in inflammatory cells (monocytes/
macrophages).12 Classical NSAIDs act as non-selective inhibitors of COX-1 and
COX-2 isozymes.13 Inhibition of COX-1 is thought to be responsible for the
gastrointestinal liabilities associated with most NSAIDs, while inhibition of the
inducible COX-2 isozyme is thought to be responsible for the anti-inflammatory
effects.14 The hypothesis led to substantial research efforts towards the discovery
of selective COX-2 inhibitors and has resulted in the introduction of celecoxib,
valdecoxib and rofecoxib into the market as the next generation of NSAIDs
with reduced gastrointestinal liabilities.15
106 Chapter 3
k1 k2
E + I (E.I) (E.I)*
k-1
O
O
HO Ser530
O O
H3C O O HN
H3C O Ser530
OH
HN
OH
access channel with their carboxylic acid moiety ion paired to an active site
Arg120 residue. The Arg120 residue also ion pairs with the carboxylate of ara-
chidonic acid. Site-directed mutagenesis of the arginine residue in COX-1 to
glutamine or glutamate renders the protein resistant to inhibition by carbo-
xylic acid containing NSAIDs.25,26 Crystallisation of COX-1 acetylated by
bromoacetylsalicyclic acid not only confirms Ser530 acetylation but also reveals
a salicylate ion-paired to Arg120.22 Arg120 is part of a hydrogen bonding
network with Glu524 and Tyr355 which stabilises substrate–inhibitor interac-
tions and closes off the upper part of the COX active site from the spacious
opening at the base of the channel. Disruption of this hydrogen bonding
network opens the constriction and enables substrate–inhibitor binding and
release to occur.
It is important to note that selective COX-2 inhibitors are actually compe-
titive inhibitors of both COX-1 and COX-2, but exhibit selectivity for COX-2
in the time-dependent step by binding tightly at the active site and causing a
conformational change in the isozyme structure (see Chapter 5 for a detailed
description of sulfonamide-based selective COX-2 inhibitors).
H H H
R N S H R2 H
R1 N S
O N
O O N
O R3
COOH
COOH
CH2 OCH2
Penicillin V: R =
Cephalosporins: R2 = H; Cephamycins: R2 = OCH3
Penicillin G (Bicillin): R =
Cl
H2C NH2 R2 = H R3 = Cl
H2 C Cefaclor: R1 =
Cloxacillin (Cloxapen): R =
Oxacillin (Bactocill): R =
CH3 CH3 O
N O N O
Cefoxotin: R1 = CH2 R2 = OCH3 R3 = O NH2
S
CH
CH NH2
NH2
Amoxicillin: R = Ampicillin: R =
HO
O
O H
H3C N
H COOH
NH
O NH COOH O NH S CH3
H3C
peptidoglycan peptidoglycan
Acyl D -alanyl-D-alanine
motif in transpeptidase
O O
O
HO HN R
H
R N S HN O S
HN
O N Peptidoglycan transpeptidase O HN
O
COO COO
β-lactam antibiotic Acylated enzyme
NH2 O O
NH2 H H H
H H H O OH N S S
N S
N O N N
O N O O O
HO O COO K COOH
COOH
COOH
Ampicillin Sulbactam
Amoxicillin Clavulanate potassium
Augmentin Unasyn
cephalosporin structure were isolated from various organisms and were found
to be potent mechanism-based inactivators of b-lactamases (Figure 3.8).44,45
These compounds are used in combination with penicillins to destroy penicillin-
resistant strains of bacteria. For example, the combination of amoxicillin and
clavulanate (a b-lactamase inactivator) is sold as Augmentin, and ampicillin
plus sulbactam is sold as Unasyn (see Figure 3.8).46–48 The b-lactamase inhi-
bitors have no antibiotic activity, but they protect the penicillin from
destruction so that it can interfere with cell wall biosynthesis.
HO
Pravastatin
Mevastatin Lovastatin (Lactone-form) Lovastatin (hydroxyacid-form) Simvastatin (Lactone-form) Simvastatin free hydroxy acid
N OH OH OH
OH OH
COOH
COOH COOH
H N
N OH
N OH H3CO H3C
OH N N
O N O S
COOH CH3 F
F O
Atorvastatin Fluvastatin Cerivastatin Rosuvastatin
HO
HO HO COO
COO COO
H3C H3 C OH
O OH
H
S-CoA S-CoA
HMG-CoA Ring
Mevaldyl-CoA
transition state intermediate Statin pharmacophore
O O
O O
O OH
O O
O O
O O
O OH
Cl Cl
Cl Cl
Fenofibrate Fenofibric acid Clofibrate Clofibric acid
O
O
OH
CH3
O O
O
HN
OH O
OH
O Cl
CH3
Cl
Cl
H O
N
O HOOC
N N N
HOOC O H
OCH3 CF3
Figure 3.12 Novel PPARa agonists for the potential treatment of dyslipidemia.
O O O
OH R R
O N
H
O R = alkyl, aryl
OH O
O O O O
S
N R N
H H
R = alkyl, NH2, NH-R1
P4503A4 P4503A4
OH OH OH
HO HO HO
N N N
Cl
X
N O
O R
CH3
H3CO H
O N
H3CO
CH3
CH3 N
N CH3
O Cl Cl
O Cl
3.7 3.8: R = O 3.9: R = NH Meclofenamic acid: R = OH 3.10: R = NH(CH2)2OC6H5
IC50 (COX-2) > 33 μM IC50 (COX-2) ~ 0.05 μM IC50 (COX-2) ~ 0.05 μM IC50 (COX-2) ~ 0.05 μM IC50 (COX-2) ~ 0.15 μM
IC50 (COX-1) > 66 μM IC50 (COX-1) > 66 μM IC50 (COX-1) > 66 μM IC50 (COX-1) ~ 0.04 μM IC50 (COX-1) ~ 66 μM
Chapter 3
Figure 3.14 Neutralisation of the carboxylic acid group in non-selective COX inhibitors indomethacin and meclofenamic acid yields selective
COX-2 inhibitors.
Carboxylic Acids and their Bioisosteres 119
candidate compounds were also shown to possess oral anti-inflammatory
activity without the ulcerogenic effects associated with the parent NSAIDs.87 In
a manner similar to indomethacin, some secondary amide derivatives of the
fenamic acid NSAID, meclofenamic acid, also demonstrated potent and
selective COX-2 inhibition.89 The 2-phenoxyethylamide derivative 3.10 was the
most selective inhibitor, with a COX-2 selectivity ratio of B440 (Figure 3.14).
Unlike indomethacin SAR, only the amide derivatives of meclofenamic acid
demonstrated COX-2 selectivity. The esters were either inactive or non-selective
COX inhibitors. The reason(s) for this discrepancy is unclear. Finally, it is
interesting to note that simple derivatisation involving amidation and/or
esterification is not a universal strategy to convert all traditional carboxylic acid
NSAIDs into selective COX-2 inhibitors. For instance, amidation or ester-
ification of naproxen, sulfindac and/or ibuprofen yields inactive compounds
(A. S. Kalgutkar, unpublished observations).
4 4
O 3 3
N N
5 N 5 N
R R R
N NH
OH N1 2 1 N 2
H
(1H )-tetrazole (2H )-tetrazole
COOH
H
N
R COOH X
3.11: R = Cl; IC50 ~ 40 μM 3.13 (IC50 ~ 1.2 μM) O COOH
3.12: R = NO2; IC50 ~ 13 μM 3.14 (IC50 ~ 0.12 μM) 3.15 (X = O, CH2, NH, S)
Carboxylic Acids and their Bioisosteres
Cl Cl Cl
N N N
CYP Metabolism OH OH
N COOH N
N
N N
N HOOC
N
N NH N NH 3.16 (EXP7711) IC50 ~ 0.23 μM
Figure 3.16 Historical overview of the discovery of the tetrazole-based angiotensin II receptor antagonist, losartan.
121
122 Chapter 3
key in the discovery of the first non-peptide AT1 receptor antagonist losartan
(DuP 753) (3.17) with good oral bioavailability, vastly improved antagonist
potency (IC50 B5.5 nM) and a long duration of action.109–111 It is noteworthy
to point out that the carboxylic acid metabolite of losartan, i.e. EXP3174
(compound 3.18), is an active metabolite with greater AT1 receptor antagonism
than the parent compound.112,113
The value of the tetrazole motif in the discovery of novel, selective and orally-
active AT1 receptor antagonists is clearly evident from the fact that five out of
the six drugs in this class that are currently marketed for the treatment of
hypertension contain the tetrazole group. The list includes losartan, irbesartan,
olmesartan medoxomil (active metabolite: olmesartan), valsartan and cande-
sartan cilexetil (active metabolite: candesartan); telmosartan contains a car-
boxylic acid group instead of the tetrazole motif (Figure 3.17). It is important
to note that site-directed mutagenesis studies have provided evidence that the
tetrazole moiety in the non-peptide antagonists interacts with a protonated
lysine and histidine at the recognition site of the AT1 receptor in a manner
similar to the interaction of the carboxy terminus of the natural ligand
angiotensin with the receptor.114,115
N N N
N N
N N N
N N
N NH N NH N NH
N NH N NH
Carboxylic Acids and their Bioisosteres
O N H3 C
N
N COOH N O O N N
N
O HOOC
O N N
O
H3 C
O
O
N N N
N N N
N NH N NH N NH HOOC
Figure 3.17 Structures of the marketed angiotensin II receptor antagonists for treatment of hypertension.
123
124 Chapter 3
Table 3.3 Structure, dose, Caco-2 permeability and solubility characteristics
of NSAIDs.
High Papp (Caco-2)a Equilibrium solubilityb
dose Oral F
Compound MW (mg) (%) AB BA pH 1.2 pH 7.4
mg mL1.
b
O O RCOOH O CH2O O
O N
O
COOH
Ampicillin
NH2
H
R R R N
Esterases O OH OH S
O O R2
R=
O O O O N
R1 O R1
R2COOH O
3.19
3.8.3.1 Glucuronidation
From a quantitative perspective, glucuronidation is the most important route
of carboylic acid biotransformation yielding the corresponding b-1-O-acyl
glucuronides (also referred to as acyl glucuronides) (Scheme 3.5); these are
Carboxylic Acids and their Bioisosteres 127
1-O-β-glucuronide 3-O-β-glucuronide
HO2C
O HO2C HO2C OH
UGT O O HO
HO HO R O O
R OH HO O R R O OH OH
UDPGA OH OH O
Carboxylic acid O O NH2-protein
HX-protein
HO2C
X = S, O or NH OH protein
HO
HO2C R O N
O
O OH
HO protein
HO OH + R X O
OH
Schiff base
Amadori
Haptenization rearrangment
HO2C
OH protein
HO
Immune-mediated IADRs R O NH
O
O
more polar than the parent acids due to the hydrophilic nature of the linked
glucuronic acid moiety. The biotransformation is catalysed by the family of
uridine 5 0 -diphospho-glucuronosyl transferases (UDP-glucuronosyl trans-
ferases, UGT, EC 2.4.1.17) that require uridine-diphosphate glucuronic acid
(UDPGA) as a co-factor. Depending on the structural features, molecular
weight and recognition pattern for active uptake and/or efflux, acyl glucur-
onides can be eliminated via renal or biliary excretion. Following biliary
excretion, acyl glucuronides may also be hydrolysed to regenerate the parent
carboxylic acid (aglycone) that can be reabsorbed from the gut into the portal
circulation via a process referred to as enterohepatic recirculation.128,129 Acyl
glucuronide hydrolysis is usually catalysed by b-glucuronidase enzymes,
although non-specific esterases can also participate in this process.130 Particu-
larly noteworthy are the examples in which rearranged isomers of some acyl
glucuronides (vide infra) display resistance to glucuronidase-mediated hydro-
lysis (e.g. diflunisal–acyl glucuronide) and thus present variations in enter-
ohepatic recirculation.131 Provided the acyl glucuronide is released into the
systemic circulation, hydrolysis can also occur in plasma. Numerous carboxylic
acid containing drugs including members from the NSAID, statin and fibrate
classes of compounds are subject to some degree of acyl glucuronidation as a
component of their elimination mechanism.132
CH3
O O
OH OH
Ibufenac Ibuprofen
OH
O O
OH OH
O HN
O O
N
O
Gemcabene (CI-1027) RO O
Quinapril: R=CH3CH2
Quinaprilat: R=H
O
R
O
HN
N N
Cl
F N
HO2C
MLN8054: R = H O
MLN8054-glucuronide: R = HO
HO
OH
H H N O
N N P P N H
S O O O O N N
OH OH OH
O O H3C CH3
O O O
ATP AMP + PPi OH
P OH
Benzoic acid Acyl CoA thioester Hippuric acid
HO
Scheme 3.6 Mechanism of amino acid conjugation of carboxylic acids: conversion of benzoic acid to hippuric acid, the first biotransformation
reaction described in the scientific literature.
131
132 Chapter 3
As is usually the case with enzyme-catalysed reactions, the ability of
carboxylic acids to undergo amino acid conjugation depends on steric
hindrance around the carboxylic acid group and upon substituents on the
aromatic ring or aliphatic side chain. For instance, in rats, ferrets and
monkeys, the major pathway of phenylacetic acid biotransformation is
amino acid conjugation.152–155 However, due to steric hindrance, dipheny-
lacetic acid cannot be conjugated with an amino acid, so the major pathway
of diphenylacetic acid biotransformation in the same three species is acyl
glucuronidation.152,154,156
CoA-SH
H3 C CH3
Carboxylic Acids and their Bioisosteres
Acyl-glutathione conjugate
O COOH
H
Scheme 3.7 Acyl CoA thioester metabolites of carboxylic acid derivatives as electrophiles.
133
134 Chapter 3
and tolmetin with proteins and GSH has been proposed as a mechanism
for the idiosyncratic immune-mediated toxicity associated with these
drugs.157–160
A remarkable feature in the metabolism of NSAIDs such as ibuprofen
(see Figure 3.3) is the unidirectional chiral inversion from the pharmaco-
logically inactive (R)- to the active (S)-enantiomer. Such inversion has been
documented in several in vivo studies with 2-arylpropionic acid-based drugs
and xenobiotics.163 The mechanism of enantioselective inversion is believed
to involve the initial enantioselective formation of the acyl CoA thioester
followed by epimerisation by 2-arylpropionyl-CoA epimerase (this involves
the intermediacy of a symmetrical conjugated enolate anion), followed by
hydrolysis to regenerate the free acids. For each 2-arylpropionic acid drug
studied, almost no acyl CoA formation is observed for the S-enantiomers,
while the respective acyl CoA thioester derivatives are readily detected for
most R-enantiomers. The enantioselective covalent binding of the acyl
CoA thioester of R-2-phenylpropionic acid to hepatic tissue has been also
demonstrated.164
O O O O
+ CoA-SH,Thiolase H2O
CoA CoA
S S OH
O
3.23 3.24 3.25
CoA
H3C S
Acetyl-CoA
O O
H H N H N
H N OH N
OH
O O COOH
COOH
Lovastatin free hydroxy acid Atorvastatin
N
F COOH
F
OH OH N OH
OH
COOH
COOH
COOH
H3CO OH
H3CO F
N
N
F N
Cerivastatin Fluvastatin
COOH
HO
COOH
COOH COOH
OH COOH
O HO
O COOH COOH COOH
O CoASH, ATP OH + H2O OH NAD CoASH NADPH
OH
H OH
- H2O - H2 O
Unsubstituted
Simvastatin free hydroxy acid pentanoic acid
Scheme 3.9 Proposed mechanism for b-oxidation of statins leading to unsubstituted pentanoic acid metabolites: simvastatin as an example.
Chapter 3
O OH O OH O S
CoA O S O S O S
CoA CoA CoA
GSH
Carboxylic Acids and their Bioisosteres
O S O OH
CoA H2 O
GS GS
F F F
HO CoA-SH S S
P450 HO
CoA CoA
O O O O
3.30 3.31 3.32 3.33
Scheme 3.10 b-Oxidation of valproic to reactive, electrophilic intermediates: effect of 2-fluorine substitution.
139
140 Chapter 3
the allylic alcohol 3.28 intermediates (free acid forms) have been identified as
metabolites in perfused rat liver and in primates in vivo. The 3-oxo-D4-valproic
acid (3.29) is believed to be the reactive, electrophilic species that binds cova-
lently to the ketoacylthiolase protein resulting in its inactivation, while adducts
derived from the reaction of GSH and N-acetylcysteine with the diene 3.27 in
preclinical species and humans suggest a role for this reactive metabolite in the
hepatotoxic event.186–189
Of much interest in this aspect is the finding that substitution of the methine
hydrogen atoms on the C2 position in valproic acid with a fluorine atom yields
a non-heptatotoxic compound 3.30 that retains anticonvulsant activity of the
parent drug in mice.190 The fluorine atom prevents oxidation of the 4-ene-2-
fluoro valproic acid CoA intermediate 3.32 to the diene 3.33,191 although
subsequent experimental work argued that failure to form the acyl CoA
intermediate 3.32 prevents 4-ene-2-fluoro valproate from undergoing b-oxida-
tion (see Scheme 3.10).192
O F
N CH3
H
N O O COOH
O
O
CP-671,305
Figure 3.22 Chemical structure of a novel PDE4 inhibitor and carboxylate analog
which undergoes active hepatobiliary transport.
a four-fold decrease in clearance and a four-fold increase in the area under the
curve (AUC).220 Given the ability of cyclosporin A and rifampicin to inhibit
multiple drug transporters, the interactions of CP-671 305 with the major
human hepatic drug transporters, MDR1, MRP2, BCRP and OATPs, were
evaluated in vitro.220 CP-671 305 was identified as a substrate of MRP2 and
BCRP, but not MDR1. CP-671 305 was a high affinity substrate of human
OATP2B1, but not a substrate for human OATP1B1 or OATP1B3. Exam-
ination of the hepatobiliary transport of CP-671 305 in sandwich-cultured
hepatocytes indicated active uptake into hepatocytes followed by efflux into
bile canaliculi, consistent with the results from in vitro transporter studies. The
role of rat Mrp2 in the biliary excretion was also examined in TR (Mrp2-
deficient) rats, and the observations that CP-671 305 pharmacokinetics were
largely unaltered in TR rats were consistent with the finding that compro-
mised biliary clearance of CP-671 305 was compensated by increased urinary
clearance.220 As such, these in vitro and in vivo studies, which suggest an
important role for transport proteins in the hepatobiliary disposition of CP-671
305 in rat and human, could be valuable in the design of clinical DDI studies.
Troglitazone
OH SG
O S S
P450 GSH, H2O
S N C O NH2
C O
NH O O
O CO2
GSH, [O]
O OH
S
H
N SG
O O
O-demethylation
(P4502C9) Glucuronidation
O
OH
H3CO
O-demethylation
CH3 (P4503A4) O
N
N R
O Cl H3CO H
CH3
Indomethacin
N
O Cl
O-demethylation Oxidation
(P4503A4) (P4503A4)
O
3.37: R = F
N
H3CO H
CH3 3.38: R = F
N N
O Cl
3.36
Figure 3.24 Differences in the metabolic fate of carboxylic acid derivative indo-
methacin and its neutral amide derivatives.
150 Chapter 3
Table 3.7 Physiochemical and pharmacokinetic parameter comparison
between indomethacin and some of its neutral amide derivatives.
T1/2 (min)a
Fraction unbound CLp (mL/ Vdss
Compound MW cLogP (rat plasma) Rat Human min/kg) (L/kg) F (%)
Indomethacin 357 4.18 0.03 490 490 0.51 0.19 98
Amide 3.36 460 5.50 0.0045 1.0 1.5 155 5.3 o1
Amide 3.37 450 5.85 0.0040 8.5 25 26 2.3 20
Amide 3.38 451 5.05 0.0070 8.0 23 39 2.0 38
a
Liver microsomes.
HO
B OH
H2N OH
NH
O O H2N B H2N N
OH
N COOH O B
HO OH
feature that is linked to the increase in molecular weight and lipophilic char-
acter (Table 3.7). As such, the increase in molecular weight and lipophilicity
also results in a decrease in plasma free fraction (see Table 3.7). A comparison
of the pharmacokinetic attributes of indomethacin and neutral amides 3.36–
3.38 is depicted in Table 3.7 and reveals a dramatic difference in CLp and Vdss
between the parent carboxylic acid and its amide derivatives; the neutral amide
analogs are cleared at a more rapid rate than the free carboxylic acid derivative,
a phenomenon that ultimately results in a lower oral bioavailability for the
amides. Unlike indomethacin, which is susceptible to glucuronidation and
P4502C9-catalysed O-demethylation, indomethacin amides 3.36–3.38 under-
went oxidative O-demethylation and oxidation on the amide substituent, which
was almost exclusively mediated by P4503A4 (Figure 3.24).305,306
O OH O
H Human H
N N B Microsomes N N OH
N OH N
H H
O or r CYP O
N N
bortezomib deboronated
metabolites
CH3 COOH
O Cl
NH2 N
O Cl
COOH
COOH N
N
H
Br Cl
Cl O COOH
Benoxaprofen Bromfenac Clometacin Diclofenac
Hepatotoxicity, cholestasis, Hepatotoxicity, hepatic necrosis Hepatitis, renal injury Hepatitis, jaundice, skin rash,
skin rash, phototoxicity anaphylactic shock
H3 C
O COOH
O CH3
COOH N
COOH
COOH
Carboxylic Acids and their Bioisosteres
Cl
S CH3
O
Zomepirac
Fenbufen Ibufenac Suprofen
Hepatotoxicity, skin rash Hepatotoxicity, hepatic necrosis Anaphylactic shock, renal failure,
jaundice Renal injury hepatitis
O
F COOH COOH
O
H N N N COOH
Cl
F
H 2N
HN H Cl
S O
F
COOH Valproic acid
Amineptine Trovafloxacin Tienilic acid
(Antidepressant) (Antibiotic) (Diuretic) (Anticonvulsant)
Hepatotoxicity Hepatotoxicity Hepatotoxicity Hepatotoxicity, blood dyscrasia
jaundice, thrombocytopenia
153
Figure 3.26 Carboxylic acid based drugs withdrawn from the market or associated with significant incidences of idiosyncratic drug toxicity.
154 Chapter 3
toxicity caused by chemically reactive metabolites, it is difficult to conclude
whether specific carboxylic acid metabolites will ultimately cause toxicity.
From a safety risk mitigation perspective, additional considerations such as
the daily dose of the drug candidate may be a pivotal factor mitigating the risks
of IADRs. Examples of low dose drugs (o50 mg day1) that cause IADRs are
rare (whether or not these agents are prone to bioactivation).317,318 Atorvas-
tatin serves as the ideal example of this phenomenon; despite biotransforma-
tion to acyl glucuronide metabolites; there have been no instance of IADRs
with this blockbuster drug, a feature that can be linked with its low daily dose.
Likewise, it is important to note the differences in daily doses of troglitazone
(200–400 mg day1) when compared with the structurally related thiazolidine-
dione derivatives rosiglitazone and pioglitazone (10–40 mg day1). This feature
may offset the bioactivation liability associated with the thiazolidinedione ring
system in general resulting in an improved safety profile for the successor
agents relative to troglitazone, which has been withdrawn due numerous cases
of fatal hepatotoxicity.
References
1. J. S. Nowick, Org. Biomol. Chem., 2006, 4, 3869.
2. J. March, in Advanced Organic Chemistry: Reactions, Mechanisms, and
Structure, ed. J. March, Wiley Interscience, New York, 1985, p. 218–236.
3. M. E. Cavet, M. West and N. L. Simmons, Br. J. Pharmacol., 1997, 121,
1567.
4. J. R. Vane, R. J. Flower and R. M. Botting, Stroke, 1990, 21(Suppl IV),
IV–12.
5. H. Dreser, Pflugers Arch., 1899, 76, 306.
6. J. R. Vane, Nature New Biol., 1971, 231, 232.
7. R. J. Flower, Pharmacol. Rev., 1974, 26, 33.
8. S. H. Ferreira, S. Moncada and J. R. Vane, Nature, 1971, 231, 237.
9. C. J. Hawkey, Lancet, 1999, 353, 307.
10. D. L. DeWitt and W. L. Smith, Proc. Natl. Acad. Sci. U. S. A., 1988, 85,
1412.
11. T. Hla and K. Neilson, Proc. Natl. Acad. Sci. U. S. A., 1992, 89, 7384.
12. J. R. Vane, J. A. Mitchell, I. Appleton, A. Tomlinson, D. Bishop-Bailey,
J. Croxtall and D. A. Willoughby, Proc. Natl. Acad. Sci. U. S. A., 1994,
91, 2046.
13. C. J. Smith, Y. Zhang, C. M. Kobolt, J. Muhammad, B. S. Zweifel,
A. Shaffer, J. J. Talley, J. L. Masferrer, K. Seibert and P. C. Isakson,
Proc. Natl. Acad. Sci. U. S. A., 1998, 95, 13313.
14. T. D. Warner, F. Giuliano, I. Vojnovic, A. Bukasa, J. A. Mitchell and
J. R. Vane, Proc. Natl. Acad. Sci. U. S. A., 1999, 96, 7563.
15. L. J. Marnett, Annu. Rev. Pharmacol. Toxicol., 2009, 49, 265.
16. L. H. Rome and W. E. M. Lands, Proc. Natl. Acad. Sci. U. S. A., 1975, 72,
4863.
Carboxylic Acids and their Bioisosteres 155
17. R. A. Copeland, J. M. Williams, J. Giannaras, S. Nurnberg, M.
Covington, D. Pinto, S. Pick and J. M. Trzaskos, Proc. Natl. Acad. Sci.
U. S. A., 1994, 91, 11202.
18. F. J. Van Der Ouderaa, M. Buytenhek, D. H. Nugteren and D. A. Van
Dorp, Eur. J. Biochem., 1980, 109, 1.
19. G. J. Roth, E. T. Machuga and J. Ozols, Biochemistry, 1983, 22, 4672.
20. D. L. DeWitt and W. L. Smith, Proc. Natl. Acad. Sci. U. S. A., 1988, 85,
1412.
21. D. Picot, P. J. Loll and R. M. Garavito, Nature, 1994, 367, 243.
22. P. J. Loll, D. Picot and R. M. Garavito, Nature Struct. Biol., 1995, 2, 637.
23. P. J. Loll, D. Picot, O. Ekabo and R. M. Garavito, Biochemistry, 1996, 35,
7330.
24. R. G. Kurumbail, A. M. Stevens, J. K. Gierse, J. J. McDonald, R. A.
Stegeman, J. Y. Pak, D. Gildehaus, J. M. Miyashiro, T. D. Penning,
K. Seibert, P. C. Isakson and W. C. Stallings, Nature, 1996, 384, 644.
25. D. K. Bhattacharyya, M. Lecomte, C. J. Rieke, R. M. Garavito and W. L.
Smith, J. Biol. Chem., 1996, 271, 2179.
26. J. A. Mancini, D. Riendeau, J.-P. Falgueyret, P. J. Vickers and G. P.
O’Neill, J. Biol. Chem., 1995, 270, 29372.
27. H. L. Hirsh, Med. Ann. Dist. Columbia, 1948, 17, 7.
28. A. Dalhoff, Infection, 1979, 7, 294.
29. E. P. Abraham, Drugs, 1987, 34(Suppl. 2), 1.
30. W. K. Joklik, FASEB J., 1996, 10, 525.
31. T. N. Raju, Lancet, 1999, 353, 936.
32. B. L. Lignon, Semin. Pediatr. Infect. Dis., 2004, 15, 52.
33. R. Bentley, J. Ind. Microbiol. Biotechnol., 2009, 36, 775.
34. R. F. Pratt, Cell. Mol. Life Sci., 2008, 65, 2138.
35. J. A. Kelly, J. R. Knox, P. C. Moews, G. J. Hite, J. B. Bartolone, H. Zhao,
B. Boris, J. M. Frère and J. M. Ghuysen, J. Biol. Chem., 1985, 260, 6449.
36. J. A. Kelly, J. R. Knox and H. Zhao, J. Mol. Graph., 1989, 7, 87.
37. J. A. Kelly, J. R. Knox, H. Zhao, J. M. Frère and J. M. Ghuysen, J. Mol.
Biol., 1989, 209, 281.
38. A. D. Russell, Prog. Med. Chem., 1969, 6, 135.
39. N. Georgopapadakou, S. Hammerström and J. L. Strominger, Proc. Natl.
Acad. Sci. U. S. A., 1977, 74, 1009.
40. A. Marquet, J. M. Frère, J. M. Ghuysen and A. Loffet, Biochem. J., 1979,
177, 909.
41. R. R. Yocum, D. J. Waxman, J. R. Rasmussen and J. L. Strominger,
Proc. Natl. Acad. Sci. U. S. A., 1979, 76, 2730.
42. R. R. Yocum, J. R. Rasmussen and J. L. Strominger, J. Biol. Chem., 1980,
255, 3977.
43. S. O. Meroueh, G. Minasov, W. Lee, B. K. Shoichet and S. Mobasherry,
J. Am. Chem. Soc., 2003, 125, 9612.
44. H. Aoki and M. Okuhara, Annu. Rev. Microbiol., 1980, 34, 159–181.
45. J. K. Noguchi and M. A. Gill, Clin. Pharm., 1988, 7, 37.
46. B. R. Smith and J. L. LeFrock, Drug Intell. Clin. Pharm., 1985, 19, 415.
156 Chapter 3
47. G. E. Stein and M. J. Gurwith, Clin. Pharm., 1984, 3, 591.
48. D. M. Campoli-Richards and R. N. Brogden, Drugs, 1987, 33, 577.
49. J. Stamer, Arch. Surg., 1978, 113, 21.
50. M. S. Brown and J. L. Goldstein, Proc. Natl. Acad. Sci. U. S. A., 1974, 71,
788.
51. M. S. Brown, T. F. Deuel, S. K. Basu and J. L. Goldstein, J. Biol. Chem.,
1978, 253, 1121.
52. E. Z. Dajani, T. G. Shahwan and N. E. Dajani, J. Assoc. Acad. Minor
Phys., 2002, 13, 27.
53. A. M. Gotto Jr., Clin. Cardiol., 2003, 26(Suppl. 1), 121.
54. J. A. Tolbert, Nature Rev. Drug. Discov., 2003, 2, 517.
55. M. J. Pencina, R. B. D’Agostino Sr., M. G. Larson, J. M. Massaro and R.
S. Vasan, Circulation, 2009, 119, 3078.
56. A. Endo, M. Kuroda and Y. Tsujita, J. Antibiot., 1976, 29, 1346.
57. A. Endo, Y. Tsujita, M. Kuroda and K. Tanzawa, Eur. J. Biochem., 1977,
77, 31.
58. A. W. Alberts, J. Chen, G. Kuron, V. Hunt, J. Huff, C. Hoffman, J.
Rothrock, M. Lopez, H. Joshua, E. Harris, A. Patchett, R. Monaghan, S.
Currie, E. Stapley, G. Albert-Schonberg, O. Hensens, J. Hirschfield, K.
Hoogsteen, J. Liesch and J. Springer, Proc. Natl. Acad. Sci. U. S. A., 1980,
77, 3957.
59. E. S. Istvan and J. Deisenhofer, Science, 2001, 292, 1160.
60. K. Tanzawa and A. Endo, Eur. J. Biochem., 1979, 98, 195.
61. C. E. Nakamura and R. H. Abeles, Biochemistry, 1985, 24, 1364.
62. E. S. Istvan, Am. Heart J., 2002, 144(Suppl. 6), S27.
63. E. S. Istvan and J. Deisenhofer, Biochim. Biophys. Acta, 2000, 1529, 9.
64. E. S. Istvan, M. Palnitkar, S. K. Buchanan and J. Deisenhofer, EMBO J.,
2000, 19, 819.
65. D. L. Sprecher, Am. J. Cardiol., 2000, 86(Suppl), 46L.
66. Rader, Am. J. Cardiol., 2003, 91(Suppl), 18E.
67. I. Issemann and S. Green, Nature, 1990, 347, 645.
68. S. A. Kliewer, S. S. Sundseth, S. A. Jones, P. J. Brown, G. B. Wisely, C. S.
Koble, P. Devchand, W. Wahli, T. M. Willson, J. M. Lenhard and J. M.
Lehmann, Proc. Natl. Acad. Sci. U. S. A., 1997, 94, 4318.
69. B. Desvergne and W. Wahli, Endocr. Rev., 1999, 20, 649.
70. C. Duval, M. Muller and S. Kersten, Biochim. Biophys. Acta, 2007, 1771, 961.
71. Y. Xu, D. Mayhugh, A. Saeed, X. Wang, R. C. Thompson, S. J.
Dominianni, R. F. Kauffman, J. Singh, J. S. Bean, W. R. Bensch, R. J.
Barr, J. Osborne, C. Montrose-Rafizadeh, R. W. Zink, N. P. Yumibe,
N. Huang, D. Luffer-Atlas, D. Rungta, D. E. Maise and N. B. Mantlo,
J. Med. Chem., 2003, 46, 5121.
72. M. L. Sierra, V. Beneton, A. B. Boullay, T. Boyer, A. G. Brewster,
F. Donche, M. C. Forest, M. H. Fouchet, F. J. Gellibert, D. A. Grillot,
M. H. Lambert, A. Laroze, C. Le Grumelec, J. M. Linget, V. G.
Montana, V. L. Nguyen, E. Nicodeme, V. Patel, A. Penfornis, O. Pineau,
Carboxylic Acids and their Bioisosteres 157
D. Pohin, F. Potvain, C. B. Ruault, M. Saunders, J. Toum, H. E. Xu,
R. X. Xu and P. M. Pianetti, J. Med. Chem., 2007, 50, 685.
73. M. Nomura, T. Tanase, T. Ide, M. Tsunoda, M. Suzuki, H. Uchiki,
K. Murakami and H. Miyachi, J. Med. Chem., 2003, 46, 3581.
74. R. C. Desai, E. Metzger, C. Santini, P. T. Meinke, J. V. Heck, J. P. Berger,
K. L. MacNaul, T. Q. Cai, S. D. Wright, A. Agarwal, D. E. Moller and
S. P. Sahoo, Bioorg. Med. Chem. Lett., 2006, 16, 1673.
75. J. L. R. Barlow, R. E. Beitman and T. H. Tsai, Arzneim.-Forsch., 1982, 32,
1215.
76. D. McTavish, K. L. Goa and M. Ferrill, Drugs, 1990, 39, 552.
77. B. P. Monahan, C. L. Ferguson, E. S. Killeavy, B. K. Lloyd, J. Troy and
L. R. Cantilena Jr., JAMA, 1990, 264, 2788.
78. K. T. Kivistö, P. J. Neuvonen and U. Klotz, Clin. Pharmacokinet., 1994,
27, 1.
79. M. Jurima-Romet, K. Crawford, T. Cyr and T. Inaba, Drug Metab.
Dispos., 1994, 22, 849.
80. L. L. von Moltke, D. J. Greenblatt, S. X. Duan, J. S. Harmatz and R. I.
Shader, J. Clin. Pharmacol., 1994, 34, 1222.
81. K. H. Ling, G. A. Leeson, S. D. Burmaster, R. H. Hook, M. K. Reith and
L. K. Cheng, Drug Metab. Dispos., 1995, 23, 631.
82. M. G. Eller, B. J. Walker, P. A. Westmark, S. J. Ruberg, K. K. Antony,
B. E. McNutt and R. A. Okerholm, J. Clin. Pharmacol., 1992, 32, 267.
83. R. E. Benton, P. K. Honig, K. Zamani, L. R. Cantilena and R. L.
Woosley, Clin. Pharmacol. Ther., 1996, 59, 383.
84. P. K. Honig, D. C. Wortham, R. Hull, K. Zamani, J. E. Smith and L. R.
Cantilena, J. Clin. Pharmacol., 1993, 33, 1201.
85. A. Markham and A. J. Wagstaff, Drugs, 1998, 55, 269.
86. G. A. Patani and E. J. LaVoie, Chem. Rev., 1996, 96, 3147.
87. A. S. Kalgutkar, B. C. Crews, S. W. Rowlinson, A. B. Marnett, K. R.
Kozak, R. P. Remmel and L. J. Marnett, Proc. Natl. Acad. Sci. U. S. A.,
2000, 97, 925.
88. A. S. Kalgutkar, A. B. Marnett, B. C. Crews, R. P. Remmel and L. J.
Marnett, J. Med. Chem., 2000, 43, 2860.
89. A. S. Kalgutkar, S. W. Rowlinson, B. C. Crews and L. J. Marnett, Bioorg.
Med. Chem. Lett., 2002, 12, 521.
90. A. S. Kalgutkar, B. C. Crews, S. Saleh, D. Prudhomme and L. J. Marnett,
Bioorg. Med. Chem., 2005, 13, 6810.
91. R. N. Butler, Adv. Het. Chem., 1977, 21, 323.
92. C. W. Thornber, Chem. Soc. Rev., 1979, 8, 563.
93. H. Singh, A. S. Chawla, V. K. Kapoor, D. Paul and R. K. Malhotra,
Prog. Med. Chem., 1980, 17, 151.
94. A. Burger, Prog. Drug Res., 1991, 37, 287.
95. A. J. Albert, J. Chem. Soc. B, 1966, 427.
96. J. M. McManus and R. M. Herbst, J. Org. Chem., 1959, 24, 1643.
97. V. A. Ostrovskii and A. O. Koren, Heterocycles, 2000, 53, 1421.
158 Chapter 3
98. C. Hansch and L. Leo, in Exploring QSAR. Fundamentals and Applica-
tions in Chemistry and Biology, ed. C. Hansch and L. Leo, American
Chemical Society, Washington DC, 1995, ch. 13.
99. R. T. Eberhardt, R. M. Kevak, P. M. Kang and W. H. Frishman, J. Clin.
Pharmacol., 1993, 33, 1023.
100. D. H. Streeten, G. H. Anderson, J. M. Freiberg and T. G. Dalakos,
N. Engl. J. Med., 1975, 292, 657.
101. G. H. Anderson Jr., D. H. Streeten and T. G. Dalakos, Circ. Res., 1977,
40, 243.
102. A. T. Chiu, J. V. Duncia, D. E. McCall, P. C. Wong, W. A. Price Jr., M. J.
Thoolen, D. J. Carini, A. L. Johnson and P. B. Timmermans, J. Phar-
macol. Exp. Ther., 1989, 250, 867.
103. P. C. Wong, A. T. Chiu, W. A. Price, M. J. Thoolen, D. J. Carini, A. L.
Johnson, R. I. Taber and P. B. Timmermans, J. Pharmacol. Exp. Ther.,
1988, 247, 1.
104. K. Hsieh and G. R. Marshall, J. Med. Chem., 1986, 29, 1968.
105. P. C. Wong, W. A. Price Jr., A. T. Chiu, M. J. Thoolen, J. V. Duncia,
A. L. Johnson and P. B. Timmermans, Hypertension, 1989, 13, 489.
106. R. R. Wexler, D. J. Carini, J. V. Duncia, A. L. Johnson, G. J. Wells, A. T.
Chiu, P. C. Wong and P. B. Timmermans, Am. J. Hypertension, 1992, 5,
209S.
107. P. B. Timmermans, D. J. Carini and A. T. Chiu, Blood Vessels, 1990, 27,
295.
108. D. J. Carini, J. V. Duncia, P. E. Aldrich, A. T. Chiu, A. L. Johnson, M. E.
Pierce, W. A. Price, J. B. Santella, G. J. Well and R. R. Wexler, et al.,
J. Med. Chem., 1991, 34, 2525.
109. A. T. Chiu, D. E. McCall, W. A. Price Jr., P. C. Wong, D. J. Carini, J. V.
Duncia, R. R. Wexler, S. E. Yoo, A. L. Johnson and P. B. Timmermans,
Am. J. Hypertension, 1991, 4, 282S.
110. P. C. Wong, W. A. Price Jr., A. T. Chiu, J. V. Duncia, D. J. Carini, R. R.
Wexler, A. L. Johnson and P. B. Timmermans, Am. J. Hypertension,
1991, 4, 288S.
111. J. V. Duncia, D. J. Carini, A. T. Chiu, A. L. Johnson, W. A. Price, P. C.
Wong, R. R. Wexler and P. B. Timmermans, Med. Res. Rev., 1992, 12,
149.
112. R. A. Stearns, P. K. Chakravarty, R. Chen and S.-H. L. Chiu, Drug
Metab. Dispos., 1995, 23, 207.
113. M. W. Lo, M. R. Goldberg, J. B. McCrea, H. Lu, C. I. Furtek and T. D.
Bjornsson, Clin. Pharmacol. Ther., 1995, 58, 641.
114. K. Noda, Y. Saad, A. Kinoshita, T. P. Boyle, R. M. Graham, A. Husain
and S. S. Karnik, J. Biol. Chem., 1995, 270, 2284.
115. R. R. Wexler, W. J. Greenlee, J. D. Irvin, M. R. Goldberg, K. Prendergast,
R. D. Smith and P. B. M.W. M. Timmermans, J. Med. Chem., 1996, 39,
625.
116. M. Yazdanian, K. Briggs, C. Jankovsky and A. Hawi, Pharmaceutical
Res., 2004, 21, 293.
Carboxylic Acids and their Bioisosteres 159
117. P. L. Carl, P. K. Chakravarty and J. A. Katzenellenbogen, J. Med. Chem.,
1981, 24, 479.
118. W. Daehne, E. Frederiksen, E. Gundersen, F. Lund, P. Morch, H. J.
Petersen, K. Roholt, L. Tybring and W. O. Godtfredsen, J. Med. Chem.,
1970, 13, 607.
119. G. Paintaud, G. Alván, M. L. Dahl, A. Grahnén, J. Sjövall and J. O.
Svensson, Eur. J. Clin. Pharmacol., 1992, 43, 283.
120. F. Margarit, J. Moreno-Dalmau, R. Obach, C. Peraire and J. M.
Pla-Delfina, Eur. J. Drug Metab. Pharmacokinet., 1991, 3, 102.
121. Y. H. Zhao, J. Le, M. H. Abraham, A. Hersey, P. J. Eddershaw, C. N.
Luscombe, D. Butina, G. Beck, B. Sherborne, I. Cooper and J. A. Platts,
J. Pharm. Sci., 2001, 90, 749.
122. H. Lennernäs, L. Knutson, T. Knutson, A. Hussain, L. Lesko, T.
Salmonson and G. L. Amidon, Eur. J. Pharm. Sci., 2002, 15, 271.
123. R. T. Scheife and H. C. Neu, Pharmacotherapy, 1982, 2, 313.
124. M. Ehrnebo, S. O. Nilsson and L. O. Boréus, J. Pharmacokinet. Bio-
pharm., 1979, 7, 429.
125. N. O. Bodin, B. Ekström, U. Forsgren, L. P. Jalar, L. Magni, C. H.
Ramsay and B. Sjöberg, Antimicrob. Agents Chemother., 1975, 8, 518.
126. H. C. Neu, Rev. Infect. Dis., 1981, 3, 110.
127. R. S. Obach, F. Lombardo and N. J. Waters, Drug Metab. Dispos., 2008,
36, 1385.
128. G. M. Pollack and K. L. Brouwer, J. Pharmacokinet. Biopharm., 1991, 19,
189.
129. S. M. Pond and T. N. Tozer, Clin. Pharmacokinet., 1984, 9, 1.
130. B. C. Sallustio, L. Sabordo, A. M. Evans and R. L. Nation, Curr. Drug
Metab., 2000, 1, 163.
131. R. G. Dickinson and A. R. King, Biochem. Pharmacol., 1993, 46, 1175.
132. C. D. King, G. R. Rios, M. D. Green and T. R. Tephly, Curr. Drug
Metab., 2000, 1, 143.
133. A. F. McDonagh, L. A. Palma, J. J. Lauff and T. W. Wu, J. Clin. Invest.,
1984, 74, 763.
134. R. B. van Breemen and C. Fenselau, Drug Metab. Dispos., 1985, 13, 318.
135. R. Drew and K. Knights, Agents Actions Suppl., 1985, 17, 127.
136. P. C. Smith, A. F. McDonagh and L. Z. Benet, J. Clin. Invest., 1986, 77,
934.
137. M. Stogniew and C. Fenselau, Drug Metab. Dispos., 1982, 10, 609.
138. M. L. Hyneck, P. C. Smith, A. Munafo, A. F. McDonagh and L. Z.
Benet, Clin. Pharmacol. Ther., 1988, 44, 107.
139. A. Munafo, M. L. Hyneck and L. Z. Benet, Pharmacology, 1993, 47, 309.
140. L. Z. Benet, H. Spahn-Langguth and S. Iwakawa, Life Sci., 1993, 53,
L141.
141. A. Ding, J. C. Ojingwa, A. F. McDonagh, A. L. Burlingame and L. Z.
Benet, Proc. Natl. Acad. Sci. U.S.A., 1993, 90, 3797.
142. S. Bolze, N. Bromet, C. Gay-Feutry, F. Massiere, R. Boulieu and
T. Hulot, Drug Metab. Dispos., 2002, 30, 404.
160 Chapter 3
143. J. Wang, M. Davis, F. Li, F. Azam, J. Scatina and R. Talaat, Chem. Res.
Toxicol., 2004, 17, 1206.
144. G. S. Walker, J. Atherton, J. Bauman, C. Kohl, W. Lam, M. Reily,
Z. Lou and A. Mutlib, Chem. Res. Toxicol., 2007, 20, 876.
145. M. Castillo and P. C. Smith, J. Chromatogr., 1993, 614, 109.
146. M. Castillo and P. C. Smith, Drug Metab. Dispos., 1995, 23, 566.
147. H. Yuan, B. Feng, Y. Yu, J. Chupka, J. Y. Zheng, T. G. Heath and B. R.
Bond, J. Pharmacol. Exp. Ther., 2009, 330, 191.
148. C. Xia, L.-S. Gan, V. Kadambi, Y. Li, N. Liu, V. Uttamsingh, R. Gallegos,
M. Milton, J.-T. Wu, S. Prakash, C. Alden, F. Lee and S. Balani, Toxicol,
Pathol., 2009, 37, 123.
149. M. G. Manfredi, J. A. Ecsedy, K. A. Meetze, S. K. Balani, O. Burenkova,
W. Chen, K. M. Galvin, K. M. Hoar, J. J. Huck, P. J. LeRoy, E. T. Ray,
T. B. Sells, B. Stringer, S. G. Stroud, T. J. Vos, G. S. Weatherhead, D. R.
Wysong, M. Zhang, J. B. Bolen and C. F. Claiborne, Proc. Natl. Acad.
Sci. U. S. A., 2007, 104, 4106.
150. J. Liberg, Ann. Phys. Chem., 1829, 17, 389.
151. A. J. Hunt and J. Caldwell, in Conjugation Reactions in Drug Metabolism,
ed. G. J. Mulder, Taylor & Francis, London, 1990, pp. 273–305.
152. P. A. F. Dixon, J. Caldwell and R. L. Smith, Xenobiotica, 1977, 7, 727.
153. A. R. Jones, Xenobiotica, 1982, 12, 387.
154. P. A. F. Dixon, J. Caldwell, C. J. Woods and R. L. Smith, Biochem. Soc.
Trans., 1976, 4, 143–145.
155. M. O. James, R. L. Smith, R. T. Williams and M. Reidenberg, Proc.
Royal Soc. London, Series B: Biol. Sci., 1972, 182, 25.
156. P. A. F. Dixon, J. Caldwell and R. L. Smith, West African J. Biol. Appl.
Chem., 1977, 20, 21.
157. K. M. Knights, M. J. Sykes and J. O. Miners, Expert Opin. Drug Metab.
Toxicol., 2007, 3, 159.
158. M. P. Grillo and L. Z. Benet, Drug Metab. Dispos., 2002, 30, 55.
159. J. Olsen, C. Li, C. Skonberg, I. Bjornsdottir, U. Sidenius, L. Z. Benet and
S. H. Hansen, Drug Metab. Dispos., 2007, 35, 758.
160. M. P. Grillo and F. Hua, Drug Metab. Dispos., 2003, 31, 1429.
161. B. C. Sallustio, S. Nunthasomboon, C. J. Drogemuller and K. M.
Knights, Toxicol. Appl. Pharmacol., 2000, 163, 176.
162. U. Sidenius, C. Skonberg, J. Olsen and S. H. Hansen, Chem. Res. Toxi-
col., 2004, 17, 75.
163. J. M. Mayer, V. M. Roy-De, C. Audergon, B. Testa and J. C. Etter, Int. J.
Tiss. React., 1994, XVI, 59.
164. C. Li, L. Z. Benet and M. P. Grillo, Chem. Res. Toxicol., 2002, 15, 1480.
165. D. L. Nelson and M. M. Cox, in Lehninger Principles of Biochemistry, ed.
A. L. Lehninger, Worth Publishers, New York, 2000, pp. 598–625.
166. S. Vickers, C. A. Duncan, I. -W. Chen, A. Rosegay and D. E. Duggan,
Drug Metab. Dispos., 1990, 18, 138.
167. R. A. Halpin, E. H. Ulm, A. E. Till, P. H. Kari, K. P. Vyas, D. B.
Hunninghake and D. E. Duggan, Drug Metab. Dispos., 1993, 21, 1003.
Carboxylic Acids and their Bioisosteres 161
168. D. W. Everett, T. J. Chando, G. C. Didonato, S. M. Singhvi, H. Y. Pan
and S. H. Weinstein, Drug Metab. Dispos., 1991, 19, 740.
169. A. E. Black, M. E. Sinz, R. N. Hayes and T. F. Woolf, Drug Metab.
Dispos., 1998, 26, 755.
170. A. E. Black, R. N. Hayes, B. D. Roth, P. Woo and T. F. Woolf, Drug
Metab. Dispos., 1999, 27, 916.
171. J. G. Dain, E. Fu, J. Gorski, J. Nicoletti and T. J. Scallen, Drug Metab.
Dispos., 1993, 21, 567.
172. M. Boberg, R. Angerbauer, W. K. Kanhai, W. Karl, A. Kern, M. Radtke
and W. Steinke, Drug Metab. Dispos., 1998, 26, 640.
173. T. Prueksaritanont, B. Ma, X. Fang, R. Subramanian, J. Yu and J. H.
Lin, Drug Metab. Dispos., 2001, 29, 1251.
174. P. M. Haddad, A. Das, M. Ashfaq and A. Wieck, Expert Opin. Drug
Metab. Toxicol., 2009, 5, 539.
175. E. S. Zafrani and P. Berthelot, Hepatology, 1982, 2, 591.
176. H. Nau and W. Loscher, Epilepsia, 1984, 25(Suppl. 1), S14.
177. S. Russell, Curr. Opin. Pediatr., 2007, 19, 206.
178. S. A. Koenig, D. Buesing, E. Longin, R. Oehring, P. Häussermann,
G. Kluger, F. Lindmayer, R. Hanusch, I. Degen, H. Kuhn, K. Samii,
A. Jungck, R. Brückner, R. Seitz, W. Boxtermann, Y. Weber, R. Knapp,
H. H. Richard, B. Weidner, J. M. Kasper, C. A. Haensch, S. Fitzek,
M. Hartmann, P. Borusiak, A. Müller-Deile, V. Degenhardt, G. C.
Korenke, T. Hoppen, U. Specht and T. Gerstner, Epilepsia, 2006, 47,
2027.
179. N. Gerber, R. G. Dickinson, R. C. Harland and R. K. Lynn, J. Pediatr.
(St. Louis), 1979, 95, 142.
180. H. J. Zimmerman and K. G. Ishak, Hepatology, 1982, 2, 591.
181. T. A. Baillie, Chem. Res. Toxicol., 1988, 1, 195.
182. H. Schulz, Biochemistry, 1983, 22, 1827.
183. A. W. Rettenmeier, K. S. Prickett, W. P. Gordon, S. M. Bjorge, S.-L.
Chang, R. H. Levy and T. A. Baillie, Drug Metab. Dispos., 1985, 13, 81.
184. A. W. Rettenmeier, W. P. Gordon, K. S. Prickett, R. H. Levy, J. S.
Lockard, K. E. Thummel and T. A. Baillie, Drug Metab. Dispos., 1986,
14, 443.
185. A. W. Rettenmeier, W. P. Gordon, K. S. Prickett, R. H. Levy and T. A.
Baillie, Drug Metab. Dispos., 1986, 14, 454.
186. K. Kassahun, K. Farrell and F. Abbott, Drug Metab. Dispos., 1991, 19,
525.
187. K. Kassahun, P. Hu, M. P. Grillo, M. R. Davis, L. Jin and T. A. Baillie,
Chem. Biol. Interact., 1994, 90, 253.
188. W. Tang and F. S. Abbott, Chem. Res. Toxicol., 1996, 9, 517.
189. W. Tang and F. S. Abbott, J. Mass Spectrom., 1996, 31, 926.
190. W. Tang, J. Palaty and F. S. Abbott, J. Pharmacol. Exp. Ther., 1997, 282,
1163.
191. W. Tang, A. G. Borel, T. Fujiyama and F. S. Abbott, Chem. Res. Tox-
icol., 1995, 8, 671.
162 Chapter 3
192. M. P. Grillo, G. Chiellini, M. Tonelli and L. Z. Benet, Drug Metab.
Dispos., 2001, 29, 1210.
193. J. O. Minors and D. J. Birkett, Br. J. Clin. Pharmacol., 1998, 45.
194. C. R. Lee, J. A. Goldstein and J. A. Pieper, Pharmacogenetics, 2002, 12,
251.
195. M. A. Hamman, G. A. Thompson and S. D. Hall, Biochem. Pharmacol.,
1997, 54, 33.
196. A. Mancy, M. Antignac, C. Minoletti, S. Dijols, V. Mouries, N. T.
Duong, P. Battioni, P. M. Dansette and D. Mansuy, Biochemistry, 1999,
38, 14264.
197. T. S. Tracy, C. Marra, S. A. Wrighton, F. J. Gonzalez and K. P.
Korzekwa, Biochem. Pharmacol., 1996, 52, 1305.
198. W. Tassaneeyakul, D. J. Birkett, M. C. Pass and J. O. Miners, Br. J. Clin.
Pharmacol., 1996, 42, 774.
199. M. R. Wester, J. K. Yano, G. A. Schoch, C. Yang, K. J. Griffin, C. D.
Stout and E. F. Johnson, J. Biol. Chem., 2004, 279, 35630.
200. C. M. Mosher, M. A. Hummel, T. S. Tracy and A. E. Rettie, Biochem-
istry, 2008, 47, 11725.
201. J. U. Flanagan, L. A. McLaughlin, M. J. Paine, M. J. Sutcliffe, G. C.
Roberts and C. R. Wolf, Biochem. J., 2003, 370, 921.
202. A. Melet, N. Assrir, P. Jean, M. Pilar Lopez-Garcia, C. Marques-Soares,
M. Jaouen, P. M. Dansette, M. A. Sari and D. Mansuy, Arch. Biochem.
Biophys., 2003, 409, 80.
203. M. J. Sykes, R. A. McKinnon and J. O. Minors, J. Med. Chem., 2008, 51,
780.
204. S. Rao, R. Aoyama, M. Schrag, W. F. Trager, A. Rettie and J. P. Jones,
J. Med. Chem., 2000, 43, 2789.
205. M. J. De Groot, A. A. Alex and B. C. Jones, J. Med. Chem., 2002, 45,
1983.
206. I. Zamora, L. Afzelius and G. Cruciani, J. Med. Chem., 2003, 46, 2313.
207. J. Kirchheiner, I. Roots, M. Goldammer, B. Rosenkranz and J. Brock-
möller, Clin. Pharmacokinet., 2005, 44, 1209.
208. J. Kirchheiner and J. Brockmöller, Clin. Pharmacol. Ther., 2005, 77, 1.
209. H. Lennernas, Clin. Pharmacokinet., 2003, 42, 1141.
210. P. J. Neuvonen, J. T. Backman and M. Niemi, Clin. Pharmacokinet.,
2008, 47, 463.
211. Y. Shitara, T. Horie and Y. Sugiyama, Eur. J. Pharm. Sci., 2006, 27, 425.
212. R. B. Kim, Mol. Pharm., 2006, 3, 26.
213. Y. Shitara, T. Horie and Y. Sugiyama, Eur. J. Pharm. Sci., 2006, 27,
501.
214. T. Mikkaichi, T. Suzuki, M. Tanemoto, S. Ito and T. Abe, Drug Metab.
Pharmacokinet., 2004, 19, 171.
215. C. Funk, Expert Opin. Drug Metab. Toxicol., 2008, 4, 363.
216. M. Cvetkovic, B. Leake, M. F. Fromm, G. R. Wilkinson and R. B. Kim,
Drug Metab. Dispos., 1999, 27, 866.
Carboxylic Acids and their Bioisosteres 163
217. M. Shimizu, K. Fuse, K. Okudaira, R. Nishigaki, K. Maeda, H. Kusuhara
and Y. Sugiyama, Drug Metab. Dispos., 2005, 33, 1477.
218. M. A. Hamman, M. A. Bruce, B. D. Haehner-Daniels and S. D. Hall,
Clin. Pharmacol. Ther., 2001, 69, 114.
219. G. K. Dresser, R. B. Kim and D. G. Bailey, Clin. Pharmacol. Ther., 2005,
77, 170.
220. A. S. Kalgutkar, B. Feng, H. T. Nguyen, K. S. Frederick, S. D. Campbell,
H. L. Hatch, Y. A. Bi, D. C. Kazolias, R. E. Davidson, R. J. Mireles,
D. B. Duignan, E. F. Choo and S. X. Zhao, Drug Metab. Dispos., 2007,
35, 2111.
221. A. S. Kalgutkar, E. Choo, T. J. Taylor and A. Marfat, Xenobiotica, 2004,
34, 755.
222. B. Hsiang, Y. Zhu, Z. Wang, Y. Wu, V. Sasseville, W. P. Yang and T. G.
Kirchgessner, J. Biol. Chem., 1999, 274, 37161.
223. P. J. Neuvonen, J. T. Backman and M. Niemi, Clin. Pharmacokinet.,
2008, 47, 463.
224. Y. Sai and A. Tsuji, Drug Discov. Today, 2004, 9, 712.
225. C. W. Holtzman, B. S. Wiggins and S. A. Spinier, Pharmacotherapy, 2006,
26, 1601.
226. O. Ozdemic, M. Boran, V. Gokce, Y. Uzun, B. Kocak and S. Korkmaz,
Angiology, 2000, 51, 695.
227. T. K. Lau, D. R. Leachman and R. Lufschanowski, Tex. Heart Insti. J.,
2001, 28, 142.
228. W. Muck, I. Mai, L. Fritsche, K. Ochmann, G. Rohde, S. Unger,
A. Johne, S. Bauer, K. Budde and I. Roots, et al., Clin. Pharmacol. Ther.,
1999, 65, 251.
229. Y. Shitara, M. Hirano, H. Sato and Y. Sugiyama, J. Pharmacol. Exp.
Ther., 2004, 311, 228.
230. J. S. Wang, M. Neuvonen, X. Wen, J. T. Backman and P. J. Neuvonen,
Drug Metab. Dispos., 2002, 30, 1352.
231. B. W. Ogilvie, D. Zhang, W. Li, A. D. Rodrigues, A. E. Gipson,
J. Holasapple, P. Toren and A. Parkinson, Drug Metab. Dispos., 2006, 34,
191.
232. J. T. Backman, C. Kyrklund and K. T. Kivistö, Clin. Pharmacol. Ther.,
2000, 68, 122.
233. C. Kyrklund, J. T. Backman, K. T. Kivistö, M. Neuvonen, J. Laitila and
P. J. Neuvonen, Clin. Pharmacol. Ther., 2000, 68, 592.
234. C. Kyrkland, J. T. Backman, K. T. Kivistö, M. Neuvonen, J. Laitila and
P. J. Neuvonen, Clin. Pharmacol. Ther., 2001, 69, 340.
235. A. J. Bergman, G. Murphy, J. Burke, J. J. Zhao, R. Valesky, L. Liu, K. C.
Lasseter, W. He, T. Prueksaritanont, Y. Qiu, A. Hartford, J. M. Vega and
J. F. Paolini, J. Clin. Pharmacol., 2004, 44, 1054.
236. A. Asberg, Drugs, 2003, 63, 367.
237. H. C. Maltz, D. L. Balog and J. S. Cheigh, Ann. Pharmacother., 1999,
33, 1176.
164 Chapter 3
238. L. Gullestad, K. P. Nordal, K. J. Berg, H. Cheng, M. S. Schwartz and
S. Simonsen, Transplant Proc., 1999, 31, 2163.
239. M. F. Segaert, C. De Soete, I. Vandewiele and J. Verbanck, Nephrol. Dial.
Transplant., 1996, 11, 1846.
240. A. Lasocki, B. Vote, R. Fassett and E. Zamir, Ocul. Immunol. Inflamm.,
2007, 15, 345.
241. D. Williams and J. Feely, Clin. Pharmacokinet., 2002, 41, 343.
242. J. W. Deng, I. S. Song, H. J. Shin, C. W. Yeo, D. Y. Cho, J. H. Shon and
J. G. Shin, Pharmacogenet. Genomics, 2008, 18, 424.
243. M. K. Pasanen, H. Fredrikson, P. J. Neuvonen and M. Niemi, Clin.
Pharmacol. Ther., 2007, 82, 726.
244. I. Leiri, S. Suwannakul, K. Maeda, H. Uchimaru, K. Hashimoto,
M. Kimura, H. Fujino, M. Hirano, H. Kusuhara, S. Irie, S. Higuchi and
Y. Sugiyama, Clin. Pharmacol. Ther., 2007, 82, 541.
245. M. Niemi, K. A. Arnold, J. T. Backman, M. K. Pasanen, U. Gödtel-
Armburst, L. Wojnowski, U. M. Zanger, P. J. Neuvonen, M. Eichelbaum,
K. T. Kivistö and T. Lang, Pharmcogenet. Genomics, 2006, 16, 801.
246. G. D. Vladutiu and P. J. Isackson, N. Engl. J. Med., 2009, 360, 304.
247. SEARCH Collaborative Group, E. Link, S. Parish, J. Armitage,
L. Bowman, S. Heath, F. Matsuda, I. Gut, M. Lathrop and R. Collins,
N. Engl. J. Med., 2008, 359, 789.
248. E. Kamiyama, Y. Yoshigae, A. Kasuya, M. Takei, A. Kurihara and
T. Ikeda, Drug Metab. Pharmacokinet., 2007, 22, 267.
249. Z. H. Israili, J. Hum. Hypertension, 2000, 14(Suppl 1), S73.
250. D. A. Sica, T. W. B. Gehr and S. Ghosh, Clin. Pharmacokinet., 2005, 44,
797.
251. C. H. Gleiter and K. E. Mörike, Clin. Pharmacokinet., 2002, 41, 7.
252. L. J. Scott and P. L. McCormack, Drugs, 2008, 68, 1239.
253. M. Sharpe, B. Jarvis and K. L. Goa, Drugs, 2001, 61, 1501.
254. C. I. Furtek and M. W. Lo, J. Chromatogr., 1992, 573, 295.
255. F. Waldmeier, G. Flesch, P. Müller, T. Winkler, H. P. Kriemier,
P. Buhlmayer and M. De Gasparo, Xenobiotica, 1997, 27, 59.
256. J. Stangier, J. Schmid, D. Türck, H. Switek, A. Verhagen, P. A. Peeters,
S. P. Can Marle, W. J. Tamminga, F. A. Sollie and J. H. Jonkman, J.
Clin. Pharmacol., 2000, 40, 1312.
257. H. Stenhoff, P. O. Lagerström and C. Andersen, J. Chromatogr. B
Biomed. Sci. Appl., 1999, 731, 411.
258. D. Farthing, D. Sica, I. Fakhry, A. Pedro and T. W. Gehr, J. Chromatogr.
B Biomed. Sci. Appl., 1997, 704, 374.
259. A. Nohara, H. Kuriki, T. Ishiguro, T. Saijo, K. Ukawa, Y. Maki and
Y. Sanno, J. Med. Chem., 1979, 22, 290.
260. S. W. Huskey, G. A. Doss, R. R. Miller, W. R. Schoen and S. H. Chiu,
Drug Metab. Dispos., 1994, 22, 651.
261. R. A. Stearns, G. A. Doss, R. R. Miller and S. H. Chiu, Drug Metab.
Dispos., 1991, 19, 1160.
Carboxylic Acids and their Bioisosteres 165
262. R. A. Stearns, R. R. Miller, G. A. Doss, P. K. Chakravarty, A. Rosegay,
G. J. Gatto and S. H. Chiu, Drug Metab. Dispos., 1992, 20, 281.
263. H. Davi, C. Tronquet, G. Miscoria, L. Perrier, P. Dupont, J. Caix,
J. Simiand and Y. Berger, Drug Metab. Dispos., 2000, 28, 79.
264. T. J. Chando, D. W. Everett, A. D. Kahle, A. M. Stewart, N. Vachhar-
ajani, W. C. Shyu, K. J. Kripalani and R. H. Barbhaiya, Drug Metab.
Dispos., 1998, 26, 408.
265. T. Kondo, K. Yoshida, Y. Yoshimura, M. Motohashi and S. Tanayama,
J. Mass Spectrom., 1996, 31, 873.
266. G. D. Bowers, P. J. Eddershaw, S. Y. Hughes, G. R. Manchee and
J. Oxford, Rapid Commun. Mass Spectrom., 1994, 8, 217.
267. S. E. Huskey, R. R. Miller and S. H. L. Chiu, Drug Metab. Dispos., 1993,
21, 792.
268. A. Alonen, M. Finel and R. Kostiainen, Biochem. Pharmacol., 2008, 76, 763.
269. T. Unger and E. Kaschina, Drug Safety, 2003, 26, 707.
270. C. H. Yun, H. S. Lee, J. K. Rho, H. G. Jeong and F. P. Guengerich, Drug
Metab. Dispos., 1995, 23, 285.
271. U. Yasar, G. Tybring, M. Hidestrand, M. Oscarson, M. Ingelman-Sundberg,
M. L. Dahl and E. Eliasson, Drug Metab. Dispos., 2001, 29, 1051.
272. C. R. Lee, J. A. Pieper, A. L. Hinderliter, J. A. Blaisdell and J. A.
Goldstein, Pharmacotherapy, 2003, 23, 720.
273. U. Yasar, C. Forslund-Bergengren, G. Tybring, P. Dorado, A. Llerena,
F. Sjöqvist, E. Eliasson and M. L. Dahl, Clin. Pharmacol. Ther., 2002, 71,
89.
274. M. Bourrié, V. Meunier, Y. Berger and G. Fabre, Drug Metab. Dispos.,
1999, 27, 288.
275. S. Uchida, H. Watanabe, S. Nishio, H. Hashimoto, K. Yamazaki,
H. Hayashi and K. Ohashi, Clin. Pharmacol. Ther., 2003, 74, 505.
276. A. Nakashima, H. Kawashita, N. Masuda, C. Saxer, M. Niina, Y. Nagae
and K. Iwasaki, Xenobiotica, 2005, 35, 589.
277. A. Nishino, Y. Kato, T. Igarashi and Y. Sugiyama, Drug Metab. Dispos.,
2000, 28, 1146.
278. P. Laeis, K. Püchler and W. Kirch, J. Hypertens. Suppl., 2001, 19, S21.
279. E. Kamiyama, Y. Yoshigae, A. Kasuya, M. Takei, A. Kurihara and
T. Ikeda, Drug Metab. Pharmacokinet., 2007, 22, 267.
280. M. Takayanagi, N. Sano and H. Takikawa, J. Gastroenterol. Hepatol.,
2005, 20, 784.
281. R. Nakagomi-Hagihara, D. Nakai, K. Kawai, Y. Yoshigae, T. Tokui,
T. Abe and T. Ikeda, Drug Metab. Dispos., 2006, 34, 862.
282. A. Yamada, K. Maeda, E. Kamiyama, D. Sugiyama, T. Kondo,
Y. Shiroyanagi, H. Nakazawa, T. Okano, M. Adachi, J. D. Schuetz,
Y. Adachi, Z. Hu, H. Kusuhara and Y. Sugiyama, Drug Metab. Dispos.,
2007, 35, 2166.
283. W. Yamashiro, K. Maeda, M. Hirouchi, Y. Adachi, Z. Hu and
Y. Sugiyama, Drug Metab. Dispos., 2006, 34, 1247.
166 Chapter 3
284. K. Maeda, I. Ieiri, K. Yasuda, A. Fujino, H. Fujiwara, K. Otsubo,
M. Hirano, T. Watanabe, Y. Kitamura, H. Kusuhara and Y. Sugiyama,
Clin. Pharmacol. Ther., 2006, 79, 427.
285. N. Ishiguro, K. Maeda, W. Kishimoto, A. Saito, A. Harada, T. Ebner,
W. Roth, T. Igarashi and Y. Sugiyama, Drug Metab. Dispos., 2006, 34,
1109.
286. E. J. Murphy, T. J. Davern, A. O. Shakil, L. Shick, U. Masharani,
H. Chow, C. Freise, W. M. Lee and N. M. Bass, Dig. Dis. Sci., 2000, 45,
549.
287. J. Kohlroser, J. Mathai, J. Reichheld, B. F. Banner and H. L. Bonkovsky,
Am. J. Gastroenterol., 2000, 95, 272.
288. S. K. Herrine and C. Choudhary, Ann. Intern. Med., 1999, 130, 163.
289. P. J. Cox, D. A. Ryan, F. J. Hollis, A. M. Harris, A. K. Miller,
M. Vousden and H. Cowley, Drug Metab. Dispos., 2000, 28, 772.
290. M. Hanefeld, Int. J. Clin. Pract. Suppl., 2001, 121, 19.
291. K. Kassahun, P. G. Pearson, W. Tang, I. McIntosh, K. Leung, C. Elmore,
D. Dean, R. Wang, G. Doss and T. A. Baillie, Chem. Res. Toxicol., 2001,
14, 62.
292. K. He, R. E. Talaat, W. F. Pool, M. D. Reily, J. E. Reed, A. J. Bridges
and T. F. Woolf, Drug Metab. Dispos., 2004, 32, 639.
293. R. Alvarez-Sanchez, F. Montavon, T. Hartung and A. Pahler, Chem. Res.
Toxicol., 2006, 19, 1106.
294. L. Guo, L. Zhang, Y. Sun, L. Muskhelishvili, E. Blann, S. Dial, L. Shi,
G. Schroth and Y. P. Dragan, Mol. Divers., 2006, 10, 349.
295. Y. Masubuchi, S. Kano and T. Horie, Toxicology, 2006, 222, 233.
296. T. Jaakkola, J. Laitila, P. J. Neuvonen and J. T. Backman, Basic Clin.
Pharmacol. Toxicol., 2006, 99, 44.
297. S. J. Baldwin, S. E. Clarke and R. J. Chenery, Br. J. Clin. Pharmacol.,
1999, 48, 424.
298. C. M. Loi, M. Young, E. Randinitis, A. Vassos and J. R. Koup, Clin.
Pharmacokinet., 1999, 37, 91.
299. A. J. Scheen, Clin. Pharmacokinet., 2007, 46, 1.
300. M. Niemi, J. T. Backman, M. Granfors, J. Laitila, M. Neuvonen and
P. J. Neuvonen, Diabetologia, 2003, 46, 1319.
301. T. Jaakkola, J. T. Backman, M. Neuvonen and P. J. Neuvonen, Clin.
Pharmacol. Ther., 2005, 77, 404.
302. J. Y. Park, K. A. Kim, M. H. Kang, S. L. Kim and J. G. Shin, Clin.
Pharmacol. Ther., 2004, 75, 157.
303. B. W. Ogilvie, D. Zhang, W. Li, A. D. Rodrigues, A. E. Gipson, J.
Holsapple, P. Toren and A. Parkinson, Drug Metab. Dispos., 2006, 34, 191.
304. L. M. Moriarty, M. N. Lally, C. G. Carolan, M. Jones, J. M. Clancy and
J. F. Gilmer, J. Med. Chem., 2008, 51, 7991.
305. D. Boyer, J. N. Bauman, D. P. Walker, B. Kapinos, K. Karki and A. S.
Kalgutkar, Drug Metab. Dispos., 2009, 37, 999.
306. R. P. Remmell, B. C. Crews, K. R. Kozak, A. S. Kalgutkar and L. J.
Marnett, Drug Metab. Dispos., 2004, 32, 113.
Carboxylic Acids and their Bioisosteres 167
307. N. N. Kim, J. D. Cox, R. F. Baggio, F. A. Emig, S. K. Mistry, S. L.
Harper, D. W. Speicher, S. M. Morris Jr., D. E. Ash and A. Traish,
Biochemistry, 2001, 40, 2678.
308. J. M. Fevig, M. M. Abelman, D. R. Brittelli, C. A. Kettner, R. M. Knabb
and P. C. Weber, Bioorg. Med. Chem. Lett., 1996, 6, 295.
309. T. A. Kelly, J. Adams, W. W. Bachovchin, R. W. Barton, S. J. Campbell,
S. J. Coutts, C. A. Kennedy and R. J. Snow, J. Am. Chem. Soc., 1993, 115,
12637.
310. T. Pekol, J. S. Daniels, J. Labutti, I. Parsons, D. Nix, E. Baronas, F.
Hsieh, L. S. Gans and G. Miwa, Drug Metab. Dispos., 2005, 33, 771.
311. J. Labutti, I. Parsons, R. Huang, G. Miwa, L. S. Gans and J. S. Daniels,
Chem. Res. Toxicol., 2006, 19, 539.
312. O. M. Bakke, M. Manocchia, F. de Abajo, K. I. Kaitin and L. Lasagna,
Clin. Pharmacol. Ther., 1995, 58, 108.
313. M. Fung, A. Thornton and K. Mybeck, Drug Info J., 2001, 35, 293.
314. W. Tang, R. A. Stearns, R. W. Wang, S. H. Chiu and T. A. Baillie, Chem.
Res. Toxicol., 1999, 12, 192.
315. J. P. O’Donnell, D. K. Dalvie, A. S. Kalgutkar and R. S. Obach, Drug
Metab. Dispos., 2003, 31, 1369.
316. Q. Chen, G. A. Doss, E. C. Tung, W. Liu, Y. S. Tang, M. P. Braun, V.
Didolkar, J. R. Strauss, R. W. Wang, R. A. Stearns, D. C. Evans, T. A.
Baillie and W. Tang, Drug Metab. Dispos., 2006, 34, 145.
317. A. S. Kalgutkar, I. Gardner, R. S. Obach, C. L. Shaffer, E. Callegari, K.
R. Henne, A. E. Mutlib, D. K. Dalvie, J. S. Lee, Y. Nakai, J. P.
O’Donnell, J. Boer and S. P. Harriman, Curr. Drug Metab., 2005, 6, 161.
318. A. S. Kalgutkar and J. R. Soglia, Expert Opin. Drug Metab. Toxicol.,
2005, 1, 91.
CHAPTER 4
4.1 Introduction
Amines are one of the most well-represented functional groups amongst
small drug molecules. These compounds cover a wide range of therapeutic
applications and possess a broad spectrum of physicochemical properties.
In the majority of cases, the amine provides a positively charged function
that is important for interaction with the target receptor and thus provides
potency and selectivity. Whilst this ionic interaction may not be essential for
binding to the target, it clearly represents an opportunity to provide higher
affinity binding than may be achieved through weaker intermolecular interac-
tions. There are also a number of examples where the incorporation of an
amine confers pharmacokinetic advantages and some of these are discussed in
this chapter.
168
Primary, Secondary and Tertiary Amines and their Isosteres 169
4.1.1 Amines that Interact with Aminergic Receptors
Foremost amongst compounds where the target receptor favours binding to an
amine are the drugs targeting the aminergic receptors that play such a central
role in physiologic function. The natural ligands of these receptors (see
Figure 4.1) all contain amines and hence, for drugs to interact with these
receptors or their re-uptake mechanisms, an amine function is clearly indicated.
Noradrenaline, serotonin (5-hydroxytryptamine), dopamine and histamine
are primary amines, with adrenaline being the only secondary amine amongst
the group. Drugs that interact with these targets comprise a mixture of primary,
secondary and tertiary amines and are employed for a variety of pharmaco-
logical treatments. The discovery of early available antidepressants arose from
chance observations1 including the tricyclic antidepressants (e.g. imipramine,
chlorpromazine and amitryptiline) which are, in the majority, tertiary amines
and exert their pharmacological effect through interaction with multiple ami-
nergic receptor targets.2 Monoamine oxidase inhibitors (e.g. moclobemide,
tranylcypromine, selegiline and indeloxazine) are also used as antidepressants
and were also initially discovered by chance. These agents are inhibitors of
monoamine oxidases (MAO-A and MAO-B), the enzymes responsible for the
breakdown of serotonin, dopamine, adrenaline and noradrenaline. From the
1980s onward, there was a more deliberate design of antidepressants to target
specific single or combined pharmacological action. Hence there are a vast
number of amine containing drugs including:
OH OH HN
H
HO NH2 HO N
CH3
NH2
HO HO
OH
Noradrenaline Adrenaline Serotonin (5HT)
HO NH2
HN NH2
N
HO
Dopamine Histamine
Figure 4.1 Structure of aminergic transmitter substances that are drug targets.
170 Chapter 4
serotonin and noradrenaline re-uptake inhibitors (SNRIs), e.g.
venlafaxine;
dopamine and noradrenaline re-uptake inhibitors, e.g. bupropion;
serotonin and noradrenaline antagonists, e.g. mirtazapine.
O
CH3
CH3
H3C N+ O
H3C
Acetylcholine
R2 R3
N N
R3 R1 R5
R1 N N N
R4 R2 R4
Amidine Guanidine
R
H
N
N NH H2N R
Imidazoline Hydrazine
H H H CH3
Figure 4.4 Structures of ammonia and primary, secondary and tertiary amines.
..
N
H
O H
H
O
H3C H H
N
O
H H
Ammonia 9.21
Primary amines Methylamine 10.62
Ethylamine 10.63
Benzylamine 9.34
Ethanolamine 9.50
Hydrazine 8.10
Hydroxylamine 5.97
Acetamidine 12.52
Guanidine 13.71
Secondary amines Dimethylamine 10.64
Diethylamine 10.98
Benzylethylamine 9.68
Morpholine 8.36
Pyrolidine 11.27
Piperidine 11.22
Tertiary amines Trimethylamine 9.76
Triethylamine 10.65
Dimethylethylamine 9.99
Anilines Aniline 4.62
o-Hydroxyaniline 4.72
m-Hydroxyaniline 4.17
p-Hydroxyaniline 5.50
o-Chloroaniline 2.62
m-Chloroaniline 3.32
p-Chloroaniline 3.81
o-Methylaniline 4.38
m-Methylaniline 4.67
p-Methylaniline 5.07
OH OH
H
Me N
Me L-685434
Me O O
N OH OH
H
N N
N Indinavir
O
O NH
Me Me
Me
Figure 4.9 Structure of the protease inhibitors L-685434 and the amine containing
indinavir.
OH OH
O O O O
Br −
N N
+
Me Me Me
Me
Atropine Ipratropium
Figure 4.10 Structure of the tertiary and quaternary amine containing antimuscarinic
agents, atropine and ipratropium bromide.
1
f ¼ ð4:1Þ
1 þ 10ðpKapHÞ
Thus for an amine with a pKa of 8.0, the fraction in the unionised form in the
stomach (pH B2.0) will be only 0.0001% compared to about 25% in the ileum
(pH B7.5). Based on the theory that only the unionised form is able to pene-
trate biological membranes (pH partition theory), it can be expected that amine
drugs are better absorbed in the lower regions of the gastrointestinal tract with
higher pH values when a significant proportion will always be present in the
unionised form. This theory is nicely illustrated by the example above featuring
atropine; no compound is absorbed from the stomach whereas absorption
reaches 90% during passage to the distal part of the jejenum.17
The limitation to this theory derives from an assumption that only unionised
drug is able to be absorbed by passive diffusion when in fact the ionised form of
drug is able to be absorbed by the paracellular route, albeit at a relatively slow
rate. Paracellular absorption is particularly effective for small drug molecules
(MWo250) and can be the major route for some low molecular weight
hydrophilic compounds. Hence for a small, quaternary ammonium compound
such as pyridostigmine, which is permanently charged but is a very small
molecule (MW ¼ 181), oral bioavailability is as high as 15%;19 this is presumed
to be due to paracellular absorption. The clear difference in terms of physi-
cochemical properties between the two quaternary ammonium compounds,
pyridostigmine and ipratropium, is molecular size (MW values of 181 versus
332) and is reflected in the markedly higher oral bioavailability of the former
Primary, Secondary and Tertiary Amines and their Isosteres 179
due to its ability to pass through the restricted access afforded by the aqueous
pores (diameter typically 3–6 Å in humans).
Effective transcellular absorption requires molecules to possess sufficient
lipophilicity to partition into and through the lipophilic environment of the cell
membranes that line the gastrointestinal tract. The distribution coefficient (D)
is a convenient parameter for describing the lipophilicity of a molecule and
reflects the contribution of the degree of ionisation at the pH at which the
determination is carried out to the overall expressed lipid affinity of a mole-
cule.20 Generally speaking, a distribution coefficient greater than 1 (log D40)
will indicate adequate lipid affinity for a compound to diffuse into and through
cell membranes.21
Examination of the absorption properties of a series of b-adrenoceptor
antagonists demonstrates a number of factors that influence absorption.
Figure 4.11 shows the relationship between lipophilicity and the estimated oral
absorption for 16 b-adrenoceptor antagonists ranging from log D7.4 values of -
1.9 (atenolol) to 2.0 (carvedilol). The majority of these compounds show high
and essentially complete (480%) absorption with the exception of atenolol,
nadolol, xamoterol and talinolol. Absorption of the hydrophilic agents, sotalol
and practolol, is complete despite their log D7.4 values of 1.4 and 1.3,
respectively, and indicates that these molecules are sufficiently small (MW 272
and 266, respectively) to pass through the aqueous pores. Three other hydro-
philic members of this series show incomplete absorption in man with values of
9% for xamoterol, 30% for nadolol and 44% for atenolol. For these three
compounds, the increase in molecular size from atenolol (MW ¼ 266) to
60
talinolol
atenolol
40
nadolol
20
xamoterol
0
-2.5 -2 -1.5 -1 -0.5 0 0.5 1 1.5 2 2.5
Log D7.4
Figure 4.11 Relationship between lipophilicity (Log D7.4) and extent of absorption in
man for a series of 16 b-adrenoceptor antagonists.
180 Chapter 4
nadolol (MW ¼ 309) to xamoterol (MW ¼ 339) would appear to correlate with
their ability to enter the systemic circulation via the paracellular route. It should
be recognised that in vitro systems for permeability studies, such as Caco-2 cell
monolayers, may not provide permeability assessments that are representative
of in vivo permeability for compounds which are absorbed via the paracellular
pathway. For example, sotalol displays low permeability in the Caco-2 cell line
despite its greater than 90% absorption in vivo.22 This in vitro/in vivo difference
is considered to be due to differences between the tightness of the intercellular
junctions.
Vd ¼ Vp þ Vt :f u =f ut ð4:2Þ
Rifamycin Rifampicin
(R=H = Rifamycin SV)
(R=CH2COOH = Rifamycin B)
O Me
N
Me Me Me Me
Me Me
HO OH HO OH
Me OH Me O Me OH Me O
Me Me Me Me
O O N O O N
Me OH Me Me OH Me
O O O O
Me Me
O O
O O
Me Me Me Me
Me Me
OH OH
Erythromycin Azithromycin
O
N N
Me MeSO2NH
MeSO2NH
Dofetilide Compound I
pKa 8.0; Vd 4.0 L/kg pKa 8.2; Vd 3.9 L/kg
NHSO2Me
NHSO2Me
O O
N N
Me
MeSO2NH MeSO2NH
Figure 4.14 Structures of four potassium channel blockers with varying pKa values
and volumes of distribution.
O O
FMO / CYP2C19 / CYP2E1
Cl Cl
N N
+
N O N O
Moclobemide
O
N
O
N
N N
Cl
N
N
Cl
F
N
Cl
O
N-dealkylation
for enzyme affinity; CYP2D6 binding is via ionic interactions such as ion
pairing or hydrogen bonding, and is not predominately driven by
lipophilicity.21
It has been reported that CYP2D6 is involved in metabolism of 20% of all
drugs and its substrate binding specificity, with the requirement of a positively
charged centre, means that amine containing drugs are often substrates.73
CYP2D6 is polymorphic with approximately 10% of caucasians and less than
1% of asians being poor metabolisers, while 5% of caucasians are ultra-rapid
metabolisers. Therefore, the genotype of subjects dosed with CYP2D6 sub-
strates has a major impact on how they respond to therapy and if they
experience adverse events. For example, the tricyclic antidepressant ami-
triptyline is hydroxylated by CYP2D6. Poor metabolisers exhibit elevated
plasma levels of amitriptyline and may be at risk of adverse effects such as
nausea or hypotension.74 In general, the development as medicines of com-
pounds that are highly dependent on clearance by CYP2D6 is avoided in order
to reduce the risks associated with significant inter-subject variability. For
example, development of the calcium channel antagonist UK-84 149, which is a
CYP2D6 substrate, was abandoned because the variability in drug exposure
between poor and extensive metabolisers was considered too great.75
4.5.2.2 Deamination
Amines may be oxidatively deaminated to aldehydes or ketones by P450s or
amine oxidases. The aldehyde or ketone formed is often not the final metabolic
product as further oxidation to the corresponding alcohol or acid may follow.
The best characterised of the amine oxidases are the monoamine oxidases
Primary, Secondary and Tertiary Amines and their Isosteres 189
OH
N
HO OH
Sertraline
HO
CYP2B6 O O
CYP2C19 O
CYPC9 UGT2B7 N O
NH2 UGT1A3
CYP3A4
CYP2D6 UGT1A6
UGT2B4
Cl
Cl
CYP2C19
CYP3A4
MAO-A
Cl
Cl MAO-B
Cl
Cl O
CYP2C19
CYP2E1
CYP3A4
MAO-A
MAO-B
Cl
Cl
(MAOs). MAOs are located in the outer surface of the mitochondrial mem-
brane and are found in most mammalian tissues. In humans, the highest MAO
levels are found in liver and lowest levels in the spleen. The two major forms of
MAO are MAO-A and MAO-B.
MAOs have affinity for polar substrates such as the endogenous neuro-
transmitters, dopamine and serotonin, which are outside the lipophilicity range
of many P450 enzymes. To be a substrate for MAO, there must be an available
hydrogen on the a-alpha to the amine and thus aniline is not a substrate. There
is some overlap in specificity between MAOs and P450s, however, as exem-
plified by sertraline (Figure 4.17), which is deaminated to sertraline ketone by
CYP3A4, 2C19 and MAO-A and MAO-B.76 Other compounds such as
sumatriptan are deaminated by MAO-A only with no P450 involvement.77
P450-mediated oxidative dealkylation can activate secondary and tertiary
amines towards deamination by MAO. For example, zolmitriptan (Figure 4.18)
is converted to its indole acetic acid metabolite via a P450-mediated secondary
amine intermediate.78
The role of other mammalian amine oxidases in xenobiotic clearance is less
well characterised. For example, SSAO (semicarbazide-sensitive amine oxidase)
is a copper containing enzyme found predominantly in the plasma, aorta, lung
and duodenum, but exhibiting low hepatic activity.79 SSAO displays different
190 Chapter 4
O O
H H
N CYP1A2 N
NH NH
O O
Zolmitriptan
N NH
MAO-A
O O
H H
N N
NH NH
O O
OH O
O
H
N
NH
O
OH
4.5.2.3 N-Methylation
Amine N-methyltransferases are responsible for the biotransformations of
endogenous substrates; for example, the methylation of norepinephrine to
epinephrine and histamine and glycine to their methylated forms. They are also
involved in the metabolism of xenobiotics. This initially seems an illogical
metabolic route in that methylation often results in a less polar product
reducing the metabolites propensity to be excreted. When methylation intro-
duces a quaternary ammonium ion, however, the metabolite is more polar than
the parent. The methylation of nicotine (Figure 4.20) is an example of this.
The tricyclic antidepressants, imipramine and amitriptyline, are metabolised
by demethylation to give the active metabolise desipramine and nortriptyline
(Figure 4.21). However, a recent study has shown that 9–15% of subjects
showed methylation of desipramine and nortriptyline back to their parent
Primary, Secondary and Tertiary Amines and their Isosteres 191
N NH2
O N OH
O
O O O
O O
O O O O
Cl SSAO
Cl
Amolodipine
N N
+
N N
Nicotine
N
N
N
HN
Imipramine Desipramine
N N
H
Nortriptyline
Amitripyline
sulfonamides, aryl amines, amides, and acyclic and cyclic amines which
give secondary or tertiary amine glucuronides;
tertiary amines which give rise to quaternary N-glucuronides.
Man and the common preclinical species (rat, dog rabbit, guinea pig, rabbit
and non-human primates) exhibit the ability to form secondary and tertiary
N-glucuronides, the species differences tending to be quantitative rather than
qualitative.83 Sulfadimethoxine (Figure 4.22) is N-glucuronidated in primates
to the extent of 4–27% of the administered dose and 1–6% in non-primates.84
Glucuronidation of tertiary amines is seen in non-human primates and man,
but occurs to a lesser extent in other preclinical species. For example, chlor-
promazine is N-glucuronidated extensively in vivo in man, but not in rat or
dog.85
Where there are two possible sites of N-glucuronidation, both of similar
basicity, glucuronidation often occurs on the nitrogen with the least bulky
O
O O O
N
Primates and preclinical S
O O
N species N
S N
O
N OH
H N H2N
O O
H2N HO
OH
Sulfadimethoxine
O OH
Sulfadimethoxine N-glucuronide
N
N N
+ +
N N
OH OH
O O
HO HO
OH OH
O OH O OH
4.5.3.2 N-Sulfation
N-sulfates are formed by the cytosolic enzyme N-sulfotransferase (NST), which
transfers the sulfate group from its cofactor, 3 0 -phosphoadenosine 5 0 -phos-
phosulfate, to the substrate. A wide range of amines such as aniline and
octylamine89 have been shown be sulfated by NST in vitro. Reports of in vivo
xenobiotic substrates are limited, but desipramine (Figure 4.24) and 4-phe-
nylpiperazine,4-phenyl-1,2,3,6-tetrahydropyridine are both N-sulfated in vivo,
rat and man.90,91
4.5.3.3 N-Acetylation
The enzyme arylamine N-acetyltransferases (NAT) are cytosolic enzymes that
catalyse the transfer an acetyl group from acetyl coenzyme A to the nitrogen or
oxygen of primary aromatic amines, aryl hydroxyl amines or hydrazine.
Examples of N-acetylation of drugs in each these categories include hydrala-
zine, N-biphenyl-4-yl-hydroxylamine and procainamide.
194 Chapter 4
N N
O O
HN S
N
N N N N
H H
N N
N NAT2 N N
N N N
N N
N N
H
O
UK-469413
Major metabolite in rat (in vivo)
and human (cytosol).
Dibenzazepines, R1
e.g. clozapine, N R2
N
mianserin N
Agranulocytosis103
X
N
H R2
Aminothiazoles or R R
N
thiazoles, N
e.g. cefepime, Hepatotocicity
pramipexole
R S Neurotoxicity104
R
4-Substituted-N-
alkyl-tetra- A N R Neurotoxicity105
hydropyridine,
e.g. MPTP
Hydrazines, hydra- H
zides, e.g. phe- N Hepatotoxicity
R N R
nelzine, H Carcinogenic106
procarbazine
1,3-Disubstituted
R N
piperazine, e.g.
MB243 HN N Microsomal
binding107
O
Diazine-piperazine/ Y
X z
morpholine, e.g.
BAY-41-8543 Mutagenic108
N N
A
O O
O
O
N N HO N N
H2N H N N N H
N H
H
Procainamide
Cl O O O
Cl Cl
N N N N
H 2N H N N
HN H N H
O O O
HO O
Metoclopramine
FMO3
H
NH2 N
OH
Amphetamine N-hydroxy intermediate (toxic)
OH
+
N N OH
OH
Oxime
(not toxic)
CYP450
H
NH2 N
OH
2-Aminofluorene
NAT1
H
NH+ N O
O
Adducts with DNA
4.7 Zwitterions
Combining basic amine and carboxylic acid functions in the same molecule
provides zwitterions which have very different physicochemical properties to
molecules containing either functional group alone. These properties of
200 Chapter 4
O
Cl O Cl O
N OH N OH
N N
Hydroxyzine Cetirizine
Figure 4.29 Structures of the H1 antihistamines, the basic molecule hydroxyzine and
its zwitterionic metabolite cetirizine.
O O
CH3CONH S NHCOCH3 H2N S NH2
O O
+ 2 CH3COOH
OH OH
O O
O
OH N
Me O N H
H Ester
hydrolysis Me
Me Me
Me
Me Me Me Me
OH OH
Lactonisation
O
NH2
O
+
Me
Me
Me Me
OH
OH
Figure 4.31 Schematic for proposed conversion mechanism for esterase and redox-
sensitive double prodrugs of amines (ref. 120).
although these have not to date yielded viable agents. The strategy with such an
approach is to utilise the established esterase activation as a first pro-moiety
forming a chemically reactive intermediate that undergoes spontaneous con-
version to the active compound. This is illustrated by the example of esters
of chemically reactive hydroxy amides122 to liberate 4-methoxyaniline
(Figure 4.31). Formation of the amine was demonstrated to be mediated by
enzymatic catalysis by a serine esterase.
Alternative proposals for double prodrug approaches for amines have
involved N-acyloxyalkoxycarbonyl derivatives of primary and secondary
amines which undergo enzymatic hydrolysis of the ester moiety leading to a
(hydroxyalkoxy) carbonyl derivative that undergoes spontaneous decomposi-
tion to yield the parent amine via an unstable carbamic acid.123
H
N NH2
NH2
O
HO HO
OH OH
CH3 O O
N NH2
N H3C N N
H
O NH2
Cl O
Cl
Avizafone Diazepam
O O CH3
N
Cl O O CH3 Cl OH
O O
esterase
Cl Cl Cl Cl
Triclosan
Figure 4.34 An amine containing prodrug of the antimalarial agent triclosan with
enhanced cellular uptake.
References
1. S. H. Preskorn and R. Ross, in Handbook of Experimental Pharmacology,
ed. S. H. Preskorn, J. P. Feighner, C. Y. Stanga and R. Ross, Springer-
Verlag, Berlin, 2001, pp. 171–183.
2. J. Vetulani and I. Nalepa, Eur. J. Pharmacol., 2000, 405, 351.
3. V. Calderone, L. Testai, E. Martinotti, M. Del Tacca and M. C. Breschi,
J. Pharm. Pharmacol., 2005, 57, 151.
4. M. C. Sanguinetti and M. Tristani-Firouzi, Nature, 2006, 440, 463.
5. M. Foley and L. Tilley, Int. J. Parasitol., 1997, 27, 231.
6. S. Soloway and H. Lipschitz, J. Org. Chem., 1958, 23, 613.
7. A. Bryson, J. Am. Chem. Soc., 1960, 82, 4862.
8. H. K. Hall, J. Am. Chem. Soc., 1957, 79, 5441.
9. T. C. Bissot, R. W. Parry and D. H. Campbell, J. Am. Chem. Soc., 1957,
79, 796.
10. G. W. Stevenson and D. Williamson, J. Am. Chem. Soc., 1958, 80,
5943.
11. M. M. Tuckerman, J. R. Mayer and F. C. Nachod, J. Am. Chem. Soc.,
1959, 81, 92.
12. J. P. Vacca, B. D. Dorsey, W. A. Schleif, R. B. Levin, S. L. McDaniel,
P. L. Darke, J. Zugay, J. C. Quintero, O. M. Blahy, E. Roth, V. V.
Sardana, A. J. Schlabach, P. I. Graham, J. H. Condra, L. Gotlib, M. K.
Primary, Secondary and Tertiary Amines and their Isosteres 205
Holloway, J. Lin, I. W. Chen, K. Vastag, D. Ostovic, P. S. Anderson,
E. A. Emini and J. R. Huff, Proc. Natl. Acad. Sci. U. S. A., 1994, 91, 4096.
13. J. H. Lin, I. W. Chen, K. J. Vastag and D. Ostovic, Drug Metab. Dispos.,
1995, 23, 730.
14. C. Csajka, C. Marzolini, K. Fattinger, L. A. Decosterd, A. Telenti,
J. Biollaz and T. Buclin, Antimicrob. Agents Chemother., 2004, 48, 3226.
15. J. B. Dressman, G. L. Amidon, C. Reppas and V. P. Shah, Pharm. Res.,
1998, 15, 11.
16. A. T. M. Serajuddin, Adv. Drug Deliv. Rev., 2007, 59, 603.
17. B. Beerman, K. Hellstrom and A. Rosen, Clin. Sci., 1971, 40, 95.
18. K. Ensing, R. A. de Zeeuw, G. D. Nossent, G. H. Koeter and
C. Cornelissen, Eur. J. Clin. Pharmacol., 1989, 36, 189.
19. U. Breyer-Pfaff, U. Maier, A. M. Brinkmann and F. Schuman, Clin.
Pharmacol. Ther., 1985, 37, 495.
20. C. N. Manners, D. W. Payling and D. A. Smith, Xenobiotica, 1988, 18,
331.
21. D. A. Smith, B. C. Jones and D. K. Walker, Med. Res. Rev., 1996, 16,
243.
22. Y. Yang, P. J. Faustino, D. A. Volpe, C. D. Ellison, R. C. Lyon and L. X.
Yu, Mol. Pharm., 2007, 4, 608.
23. H. Spahn-Langguth, G. Baktir, A. Radschuweit, A. Okyar, B. Terhaag,
P. Ader, A. Hanafy and P. Langguth, Int. J. Clin. Pharmacol., 1998, 36,
16.
24. T. Gramatté and R. Oertel, Clin. Pharmacol. Ther., 1999, 66, 239.
25. A. Seelig, Int. J. Clin. Pharmacol. Ther., 1998, 36, 50.
26. S. Neuhoff, P. Langguth, C. Dressler, T. B. Andersson, C. G. Regardh
and H. Spahn-Langguth, Int. J. Clin. Pharmacol. Ther., 2000, 38, 168.
27. S. Doppenschmitt, H. Spahn-Langguth, C. G. Regardh and P. Langguth,
J. Pharm. Sci., 1999, 88, 1067.
28. D. K. Walker, S. Abel, P. Comby, G. J. Muirhead, A. N. R. Nedderman
and D. A. Smith, Drug Metab. Dispos., 2005, 33, 587.
29. K. Beaumont, A. Harper, D. A. Smith and J. Bennett, Eur. J. Pharm. Sci.,
2000, 12, 41.
30. S. Abel, K. C. Beaumont, C. L. Crespi, M. D. Eve, L. Fox, R. Hyland,
B. C. Jones, G. J. Muirhead, D. A. Smith, R. F. Venn and D. K. Walker,
Xenobiotica, 2001, 31, 665.
31. C. A. Lipinski, F. Lombardi, B. W. Dominy and P. J. Feeney, Adv. Drug
Deliv. Rev., 1997, 23, 3.
32. S. R. Grimsley and M. W. Jann, Clin. Pharm., 1992, 11, 930.
33. L. M. Tremaine, W. M. Welch and R. A. Ronfeld, Drug Metab. Dispos.,
1989, 17, 542.
34. N. Bergamini and G. Fowst, Arzneimittelforschung, 1965, 15, 951.
35. D. A. Smith, H. van de Waterbeemd and D. K. Walker, in Pharmaco-
kinetics and Metabolism in Drug Design, ed. D. A. Smith, H. van de
Waterbeemd and D. K. Walker, Wiley-VCH, Weinheim, 2nd edn., 2006,
pp. 55–65.
206 Chapter 4
36. W. J. Burman, K. Gallicano and C. Peloquin, Clin. Pharmacokinet., 2001,
40, 327.
37. G. Foulds, R. M. Shepard and R. B. Johnson, J. Antimicrob. Chemother.,
1990, 25, 73.
38. N. J. Lalak and D. L. Morris, Clin. Pharmacokinet., 1993, 25, 370.
39. D. K. Walker, K. C. Beaumont, D. A. Stopher and D. A. Smith, Xeno-
biotica, 1996, 26, 1101.
40. R. P. Mason, S. F. Campbell, S. D. Wang and L. G. Herbette, Mol.
Pharmacol., 1989, 36, 634.
41. D. Alker, R. A. Burges, S. F. Campbell, A. J. Carter, P. E. Cross, D. G.
Gardiner, M. J. Humphrey and D. A. Stopher, J. Chem. Soc. Perkin
Trans., 1992, 2, 1137.
42. G. R. Manchee, A. Barrow, S. Kulkarni, E. Palmer, J. Oxford, P. V.
Colthup, J. G. Maconopchie and M. H. Tarbit, Drug Metab. Dispos.,
1993, 21, 1022.
43. M. Cazzola, R. Testi and M. G. Matera, Clin. Pharmacokinet., 2002, 41,
19.
44. R. A. Coleman, M. Johnson, A. T. Nials and C. J. Varday, Trends Pharm.
Sci., 1996, 17, 324.
45. D. Jack, Br. J. Clin. Pharmacol., 1991, 31, 501.
46. K. Westphal, A. Weinbrenner, M. Zschiesche, G. Franke, M. Knoke,
R. Oertel, P. Fritz, O. von Richter, R. Warzok, T. Hachenberg, H. M.
Kauffmann, D. Schrenk, B. Terhaag, H. K. Kroemer and W. Siegmund,
Clin. Pharmacol. Ther., 2000, 68, 345.
47. N. Raghunand, B. P. Mahoney and R. J. Gillies, Biochem. Pharmacol.,
2003, 66, 1219.
48. P. Neyroz and M. Bonati, Experimentia, 1985, 41, 361.
49. K. C. Beaumont, A. G. Causey, P. E. Coates and D. A. Smith, Xeno-
biotica, 1996, 26, 459.
50. F. Zsila and Y. Iwao, Biochim. Biophys. Acta, 2007, 1770, 797.
51. K. I. Umehara, T. Nakamata, K. Suzuki, K. Noguchi, T. Usui and
H. Kamimura, Eur. J. Drug Metab. Pharmacokinet., 2008, 33, 117.
52. K. M. Piafsky, O. Borga, I. Odar-Cederlof, C. Johansson and F. Sjoqvist,
New Eng. J. Med., 1978, 299, 1435.
53. H. Wulf, P. Munstedt and C. Maier, Acta Anaesthesiol. Scand, 1991, 35,
129.
54. J. Kelder, P. D. J. Grootenhuis, D. M. Bayada, L. P. C. Delbressine and
J. P. Ploemen, Pharm. Res., 1999, 16, 1514.
55. P. R. Reeves, D. J. Barnfield, S. Longshaw, D. A. D. McIntosh and M. J.
Winrow, Xenobiotica, 1978, 8, 305.
56. A. Hayes and R. G. Cooper, J. Pharmacol. Exp. Ther., 1971, 176, 302.
57. D. K. Walker, S. J. Bowers, R. J. Mitchell, M. J. Potchoiba, C. M.
Schroeder and H. F. Small, Xenobiotica, 2008, 38, 1330.
58. P. Hlavica, Drug Metab. Rev., 2002, 34, 451.
59. D. M. Ziegler, Drug Metab. Rev., 2002, 34, 503.
60. B. Furnes and D. Schlenk, Toxicol. Sci., 2004, 78, 196.
Primary, Secondary and Tertiary Amines and their Isosteres 207
61. T. A. Mushiroda, R. Douya, E. Takahara and O. Nagata, Drug Met.
Dispos., 2000, 28, 1231.
62. M. S. Motika, J. Zhang and J. R. Cashman, M, Exp. Opin. Drug Metab.
Toxicol., 2007, 3, 831.
63. I. M. Hisamuddin and V. W. Yang, Pharmacogen., 2007, 8, 635.
64. J. R. Cashman and J. Zhang, Drug Metab. Dispos., 2002, 30, 1043.
65. E. Mayatepek, B. Flock and J. Zschocke Pharmacogen., 2004, 14, 775.
66. J. Hoskins, G. Shenfield, M. Murray and A. Gross, Xenobiotica, 2001, 31,
387.
67. B. Eiermann, G. Engel, I. Johansson, U. M. Zanger and L. Bertilsson, Br.
J. Clin. Pharmacol., 2003, 44, 439.
68. Y. Seto and P. J. Guengerich, Biol. Chem., 1993, 268, 9986.
69. P. Hlavica, Biochim. Biophys. Acta, 2006, 1764, 645.
70. A. L. Upthagrove and W. L. Nelson, Drug Metab. Dispos., 2001, 29, 1389.
71. D. A. Smith, in Computer-Assisted Lead Finding and Optimization,
ed. H. van de Waterbeemd, B. Testa and G. Folkers, Verlag Helvetica
Chimica Acta, Basel, 1997, pp. 267–276.
72. D. F. V. Lewis and M. Dickins, Drug Metab. Rev., 2003, 35, 1.
73. S. A. Islam, C. R. Wolf, M. S. Lennard and M. J. E. Sternberg, Carci-
nogen., 1991, 12, 2211.
74. J. Halling, P. Wehe and K. Brosen, Br. J. Clin. Pharmacol., 2007, 65, 134.
75. D. K. Walker, Br. J. Clin. Phamacol., 2004, 58, 601.
76. R. S. Obach, L. M. Cox and L. M. Tremaine, Drug Metab. Dispos., 2005,
33, 262.
77. C. M. Dixon, G. R. Park and M. H. Tarbit, Biochem. Pharmacol., 1994,
47, 1253.
78. M. J. Wild, Xenobiotica, 1999, 29, 847.
79. M. S. Benedetti, K. F. Tipton, R. Whomsley and E. Baltes, J. Neural
Transm., 2007, 114, 787.
80. B. Gong and P. J. Boor, Exp. Opin. Drug Metab. Toxicol., 2006, 2,
559.
81. A. P. Beresford, P. V. Macrae and D. A. Stopher, Xenobiotica, 1988, 18,
169.
82. M. P. Kurpius and B. Alexander, Pharmacotheraphy, 2006, 26, 505.
83. S. H. L. Chiu and S. E. W. Huskey, Drug Metab. Dispos., 1998, 26, 838.
84. R. H. Adamsom, J. W. Bridges, M. R. Kibby, S. R. Walker and R. T.
Williams, Biochem. J., 1970, 118, 41.
85. H. B. Hucker, S. C. Stauffer, A. J. Balletto, S. D. White, A. G. Zacchei
and B. H. Arison, Drug Metab. Dispos., 1978, 6, 659.
86. E. M. Hawes, Drug Metab. Dispos., 1998, 26, 830.
87. H. Schupke, C. Bauer, T. Kronbach, P. J. McNeilly, R. Hempel, J. M.
Strong, H. F. Kupferberg and J.. Engel, Pharmazie, 1997, 52, S22.
88. J. Kelder, C. Funke, T. de-Boer, L. Delbressine, D. Leysen and
V. Nickolson, J. Pharm. Pharmacol., 1997, 49, 403.
89. S. G. Ramaswamy and W. B. Jakoby, J. Biol. Chem., 1987, 262,
10039.
208 Chapter 4
90. K. Iwasaki, T. Shiraga, K. Noda, K. Tada and H. Noguchi, Xenobiotica,
1986, 16, 651.
91. K. O. Wong and K. P. Wong, Xenobiotica, 1996, 26, 17.
92. J. Rawal, R. Jones, A. Payne and I. Gardner, Xenobiotica, 2008, 38,
1219.
93. G. B. Scarfe, I. D. Wilson, M. A. Warne, E. Holmes, J. K. Nicholson and
J. C. Lindon, Xenobiotica, 2002, 32, 267.
94. H. B. Hughes, J. P. Biehl, A. P. Jones and L. H. Schmidt, Am. Rev.
Tuberc., 1954, 70, 266.
95. M. A. Payton and E. Sim, Biochem. Pharmacol., 1998, 55, 361.
96. M. K. Cho and I. S. Song, Drug Metab. Pharmacokinet., 2008, 23, 243.
97. H. Keosell, B. M. Schmitt and V. Goboulev, Rev. Physiol. Biochem.
Pharmacol., 2003, 150, 36.
98. K. I. Umehara, T. Watsububo, K. Noguchi and H. Kamimura, Xeno-
biotica, 2007, 37, 818.
99. M. Li, H. Yuan, N. Li, G. Song, Y. Zheng, M. Baratta, F. Hua,
A. Thurston, J. Wang and Y. Lai, Eur. J. Pharm. Sci., 2008, 35, 114.
100. B. Angelin, A. Arvidsson, R. Dahlqvist, A. Hedman and K. Schenck-
Gusatafsson, Eur. J. Clin. Invest., 1987, 17, 262.
101. A. Hedman, B. Angelin, A. Arvidsson, O. Beck, R. Dahlqvist, B. Nilsson,
M. Olsson and K. Schenck-Gustafsson, Clin. Pharmacol. Ther., 1991, 49,
256.
102. G. Sabbioni and O. Sepai, Chimica, 1995, 49, 374.
103. M. Hummer, M. Kurz, I. Kurtzhaler, H. Oberbauer, C. Miller and W. W.
Fleischhacker, J. Clin. Pharmacol., 1997, 17, 314.
104. D. K. Dalvie, A. S. Kalgutkar, S. C. Khojasteh-Bakht, R. S. Obach and
J. P. O’Donnell, Chem. Res. Toxicol., 2002, 15, 269.
105. T. Obata, Toxicol Lett, 2002, 132, 83.
106. P. M Gannet and J. H. Powell, J. Environ. Pathol. Toxicol. Oncol., 2002,
21, 1.
107. G. A. Doss, Chem. Res. Toxicol., 2005, 18, 271.
108. A. Straub, Bioorg. Med. Chem., 2002, 10, 1711.
109. D. A. Smith, H. Van de Waterbeemd and D. K. Walker, in Pharmaco-
kinetics and Metabolism in Drug Design, ed. R. Mannhold, H. Kubinyi
and H. Timmerman, Wiley-VCH, Weinheim, 2001, pp. 121–152.
110. J. Uetrecht, Drug Metab. Rev., 2002, 34, 651.
111. S. Shibutani, N. Suzuki, Y. R. S. Laxmi, L. J. Schild, R. L. Divi, A. P.
Grollman and M. C. Poirier, Cancer Res., 2003, 63, 4402.
112. J. R. Cashman, Y. N. Xiong, L. Xu and A. Janowsky, J. Pharmacol. Exp.
Therap., 1999, 288, 1251.
113. B. Burchell and M. H. Coughtrie, Environ. Health Persp., 1997, 105, 739.
114. P. Meisel, Pharmocogen., 2002, 3, 349.
115. J. Morganroth, J. Electrocardiol., 2004, 37, 25.
116. S. Ekins, J. Pharmacol. Exp. Therap., 2002, 301, 427.
117. A. Cavalli, E. Poluzzi, F. De Ponti and M. Recanatini, J. Med. Chem.,
2002, 45, 3844.
Primary, Secondary and Tertiary Amines and their Isosteres 209
118. G. Plemper van Balen, G. Caron, G. Ermondi, A. Pagliara, T. Grandi,
G. Bouchard, R. Fruttero, P. A. Carrupt and B. Testa, Pharm. Res., 2001,
18, 694.
119. C. Chen, Curr. Med. Chem., 2008, 15, 2173.
120. T. Prueksaritanont, P. Deluna, L. M. Gorham, B. Ma, D. Cohn, J. Pang,
X. Xu, K. Leung and J. H. Lin, Drug Metab. Dispos., 1998, 26, 520.
121. P. E. Thompson, Int. J. Leprosy, 1967, 35, 605.
122. K. L. Amsberry, A. E. Gerstenberger and R. T. Borchardt, Pharm. Res.,
1991, 8, 455.
123. J. Alexander, R. Cargill, S. R. Michelso and H. Schwamm, J. Med.
Chem., 1988, 31, 318.
124. J. J. Kyncl, F. N. Minard and P. H. Jones, in Peripheral Dopaminergic
Receptors, ed. J.-L. Imbs and J. Schwartz, Pergamon Press, Oxford,
pp. 369–380.
125. T. C. Marrs, Toxicol. Rev., 2004, 23, 145.
126. G. Lallement, F. Renault, D. Baubichon, M. Peoc’h, M. F. Burckhart,
M. Gallonier, D. Clarencon and N. Jourdil, Arch. Toxicol., 2000, 74, 480.
127. S. Mishra, K. Karmodiya, P. Parasuraman, A. Surolia and N. Surolia,
Bioorg. Med. Chem., 2008, 16, 5536.
CHAPTER 5
Sulfonamide as an Essential
Functional Group in Drug
Design
AMIT S. KALGUTKAR,a RHYS JONESb AND AARTI
SAWANTa
a
Pharmacokinetics, Dynamics and Metabolism Department, Pfizer Global
Research and Development, Eastern Point Road, Groton, Connecticut,
06340, USA; b Pharmacokinetics, Dynamics and Metabolism Department,
Pfizer Global Research and Development, Ramsgate Road, Sandwich, Kent,
UK, CT139NJ
5.1 Introduction
The sulfonamide, acylsulfonamide/sulfonimide and sulfonylurea functionalities
are found within numerous marketed agents for a wide range of therapies. In
2008 there were 112 marketed drugs in the United States that contained a
sulfonamide group (Table 5.1). The drugs differ in chemical structure, mole-
cular weight (MW) and lipophilicity, and act at different receptors/enzymes via
distinct biochemical mechanisms of action. In some instances, the presence of
the sulfonamide fragment is merely a circumstantial occurrence but in the vast
majority of cases, these drugs can be categorised into distinct groups based on
the role of the sulfonamide motif in the primary pharmacology. By definition a
sulfonamide is a molecule containing a sulfonyl group attached to an amine.
This yields the possibility of a number of sulfonamide expressions by virtue of
210
Table 5.1 Drugs containing a sulfonamide group marketed in the USA, 2008.
Acetazolamide Clorexolone Glyburide Penflutizide Sulprostone
Acetohexamide Co-trimoxazolea Glymidine Piretanide Sultiame
Almotriptan Cyclopenthiazide Hydrochlorothiazide Piroxicam Sumatriptan
Althiazide Cyclothiazide Hydroflumethiazide Polythiazide Tamsulosin
Amprenavir Darunavir Ibutilide Probenecid Teclothiazide
Amsacrine Delavirdine mesylate Indapamide Quinethazone Tenoxicam
Argatroban Diazoxide Isoxicam Rosuvastatin Tezosentan
Avitriptan Dichlorphenamide Levosulpiride Sildenafil citrate Thiothixene
Bemetizide Dofetilide Lornoxicam Sotalol Tipranavir
Bendroflumethiazide Dorzolamide Mafenide Sulfacetamide Tirofiban
Benzthiazide Droxicam Mebutizide Sulfacytine Tolazamide
Benzylhydrochlorothiazide Epithiazide Mefruside Sulfadiazine Tolbutamide
Bosentan Ethoxzolamide Meloxicam Sulfadimethoxine Torsemide
Brinzolamide Fenquizone Methazolamide Sulfafurazole Trichlormethiazide
Bumetanide Flosulide Methyclothiazide Sulfamazone Tripamide
Butizide Fosamprenavir Meticrane Sulfameter Valdecoxiby
Carbutamide Furosemide Metolazone Sulfamethizole Vardenafil
Celecoxib Glibornuride Mezlocillin Sulfamethoxazole Xipamide
Sulfonamide as an Essential Functional Group in Drug Design
O COOH
COOH
OH OH OH OH OH N COOH
O O H 2N COOH H
N P P N N
N O O OH N N N N
H H
Sulfonamide as an Essential Functional Group in Drug Design
R1 R1 R1
2 O2 O AH2 O
R2 R2
Peroxidase
R2 COX O O
OOH OH
PGG2 PGH2
Arachidonic acid
R1 = CH2CH=CH(CH2)3COOH
R2 = C5H11
k1 k2
E + I (E.I) (E.I)* ð5:1Þ
k-1
H3C
O O
S
H3C O N CF3
O O
S N N
H
N Br O
CH3
N
H3C H2 N S
O S
H2 N S
O O
O O O O
Phenylbutazone F Valdecoxib Celecoxib
DuP 697 Rofecoxib
H3 C
SC-8092
Celecoxib R = NH2 H3CO R = SO2
CF3 IC (COX-2) = 0.04 µM Cl IC50 (COX-2) = 0.06 µM
50 R = SO2NH2 COX-2 selective
N N IC50 (COX-1) = 13 µM S IC50 (COX-1) > 100 µM
R = SO2NHCH3 Inactive
SC 58125 R = CH3 SC-8076 R = SO2N(CH3)2 Inactive
N R=S R = NO2, CF3, COOH Inactive
R IC50 (COX-2) = 0.1 µM R = OCH3
S IC50 (COX-1) > 100 µM H3C IC50 (COX-2) > 100 µM R COX-1 selective
R IC50 (COX-1) = 0.1 µM
O O
Figure 5.2 Pivotal role of the methylsulfone/sulfonamide group in selective COX-2 inhibition by diarylheterocycles.
Chapter 5
Sulfonamide as an Essential Functional Group in Drug Design 217
the sulfonamide group to a methanesulfonamido moiety in celecoxib results in
inactive compounds (Figure 5.2).25 Likewise, bioisosteric replacement of the
methylsulfone or sulfonamide groups with nitro, trifluoromethyl, methoxy or
carboxylic acid substituents either reverses isozyme specificity or results in
inactive compounds (see Figure 5.2).25,27
Crystal structures of complexes of sheep COX-1, mouse COX-2, and human
COX-2 with non-selective and selective inhibitors have been solved at 3–3.5 Å
resolution.28,29 Most traditional NSAIDs (e.g. diclofenac, ibuprofen, aspirin
and indomethacin) contain a free carboxylic acid group, which ion pairs to an
active site Arg120 residue. The Arg120 residue also ion pairs with the carboxylate
of the fatty acid substrate arachidonic acid. Site-directed mutagenesis of the
arginine residue in COX-1 to glutamine or glutamate renders the protein
resistant to inhibition by carboxylic acid containing NSAIDs.30,31 Arg120 is part
of a hydrogen bonding network with Glu524 and Tyr355 which stabilises sub-
strate–inhibitor interactions and closes off the upper part of the COX active site
from the spacious opening at the base of the channel referred to as the lobby.
Disruption of this hydrogen bonding network opens the constriction and
enables substrate–inhibitor binding and release to occur.
Co-crystal structures of COX-2 with methylsulfonyl- and/or sulfonamide-
diarylheterocycles have shown that the selective inhibitors bind to regions
accessible in COX-2 but not COX-1. For instance, solution of the COX-2
crystal structure co-complexed with the celecoxib derivative SC-558 (the 4-
methylphenyl group in celecoxib is replaced with the 4-bromophenyl sub-
stituent in SC-558) demonstrates that the sulfonamide moiety wedges into a
hydrophobic ‘side pocket’ of COX-2 bordered by a valine residue (Val523) and
hydrogen bonds with a neighbouring arginine residue (Arg513) and the peptide
bond of Phe518.29 A similar hydrophobic side pocket off the main channel in
COX-1 is not accessible because of the presence of an isoleucine instead of
valine at position 523, which sterically hinders inhibitor approach. The COX-2
mutant V523I is resistant to time-dependent inhibition by diarylheterocycles
but not carboxylic acid type NSAIDs.32,33 Conversely, the COX-1 mutant
I523V is sensitive to time-dependent inhibition by diarylheterocycles.33
Movement of Val523 and insertion of the sulfonamide or methylsulfone moiety
into the side pocket is thought to contribute to the time dependence of inhi-
bition by diarylheterocycles.
O O O O
H S H S
N CH3 N CH3
O O
R
for COX inhibition, both nimesulide and flosulide revealed their selective
COX-2 inhibitory properties.40
The common structural features of the alkylsulfonanilides are depicted in
Figure 5.3. The alkyl substituent is typically methyl, but halogenated methyl
substituents such as trifluoromethyl have been reported.41 The ortho-anilide
substituent typically includes aryl, heterocyclic or cycloalkyl ethers and thioe-
thers. The para-anilide substituent invariably bears an electron-withdrawing
group that may be incorporated as part of a ring; absence of the electron-
withdrawing group results in inactive compounds. A variety of methane-
sulfoanilides with different para-electron-withdrawing groups have been
evaluated as selective COX-2 inhibitors and as orally active anti-inflammatory
agents. Substituents include para-acetyl, para-cyano, para-carboxamido
(FR115068), nitro (NS-398 and nimesulide) and para-sulfonamido.42 Another
structural variant is the incorporation of the para-electron-withdrawing
group as part of a ring system as seen in flosulide and the sulfonanilide analog
T-614.43 The phenyl ring in the alkylsulfonanilide scaffold can be replaced with
the electron-withdrawing pyridine ring.42,43
The crucial requirement of the electron-withdrawing group in COX-2
inhibition by alkylsulfonanilides is most likely related to a lowering of the pKa
of the sulfonamide moiety in the range of the carboxylic acid group (the pKa of
nimesulide is B6.4–6.8, whereas that of the corresponding pyridinyl derivative
is B2.9844,45), which then allows for an efficient ion pairing interaction with a
complementary active site amino acid residue(s). This is demonstrated in the
Sulfonamide as an Essential Functional Group in Drug Design 219
NS-398-COX-2 crystal structure complex wherein the sulfonamide group ion
pairs to Arg120 in a manner similar to carboxylic acid containing NSAIDs
rather than inserting into the Val523 side pocket like the diarylheterocycles.46
Consistent with the crystal structure findings are the SAR findings that N-
methylation of the sulfonamide nitrogen in alkylsulfonanilides generates
inactive compounds.41 Unlike the methylsulfone- or sulfonamide-based dia-
rylheterocycles, the structural basis for COX-2 selectivity by NS-398 is not
evident from the COX-2 co-crystal structure.
O
+ H2O HCO3 + H ð5:2Þ
C
O
O H O R
Thr199 S
N O
N HN
H
O H Zn2+
H Thr199 O His119
H
O O H His94 His96
O
Glu106 Zn2+
O His119
His94 His96 O
Glu106
N H3 C N
O N O N N N SO2NH2
SO2NH2 SO2NH2
SO2NH2 S O
H3C N S H3 C N S
H O S O
Methazolamide Ethoxzolamide Sulthiame
Acetazolamide
SO2NH2 NH
NH
SO2NH2
SO2NH2
Cl SO2NH2 H3C S S H3CO N
S S
Cl O O
O O
Dichlorophenamide Dorzolamide Brinzolamide
SO2NH2 O
H
N Cl
O
N NH
HOOC SO2NH2 O S
O
O
Furosemide Zonisamide Saccharin
O O O O O O O
O O
S (CH2)3CH3 S (CH2)3CH3 S (CH2)3CH3
N N N N N N
H H H H H H
H2N H3C Cl
Carbutamide Tolbutamide Chlorpropamide
CH3
O O O O
O O
S H3 C S
O O N N H3C COOH
H H O O N N O O
H H
N N N
H N
Sulfonamide as an Essential Functional Group in Drug Design
H H
H3C
Glimepiride Glibenclamide meglitinide
H3 C
Cl Cl
Figure 5.6 Representative examples of first and second generation sulfonylurea-based hypoglycemic agents.
223
224 Chapter 5
the sulfonylureas, namely they close the KATP channels, although they are
presumed to act in different regions of the SUR1 receptor.
R
R R
N N
OH OH
(H2C)4 O O
NH
O O
CH3 OH HN CH3
S
O O
N 5.6: R = CH2CH(CH3)2 5.8: R = CH2CH(CH3)2
H 5.7: R = CH3 5.9: R = CH3
CH3
5.4: R = OH
5.5: R = NHSO2CH3
Figure 5.7 Successes and failures in the utility of the methanesulfonamide group as a
phenol bioisostere.
Sulfonamide as an Essential Functional Group in Drug Design 225
O
O
O CH3 O N
O N
N S
H3 C
N
N N
H N N
H
O
5.10 5.11
IC50 (antiviral) = 30 nM IC50 (antiviral) = 860 nM
O O
O O
H3C O O N N
O H3C O F O N
O N S
N S
N
N N
N N N
H N H
H
O N
5.12 5.13 N 5.14
IC50 (antiviral) = 8 nM IC50 (antiviral) = 9200 nM N IC50 (antiviral) = 7 nM
O
O O
N H
C5H11 N N
H O N R
(CH2)5 H
O NH O
O HO (CH2)2
X
5.15: R = OH N COOH
5.16: R = NHSO2C6H5 O
H
H
N COOH
O 5.17: R =
R
H
(CH2)3 N COOH
NC 5.18: R =
F
N
H
5.21: R = OH O O
5.22: R = NHSO2CH3 H
N
O N R
H
O CH3 O NH O
HO (CH2)2
N
R O N COOH
H
Cl O O O
H
Zomepirac: R = COOH N S
5.19: R = N
H
IC50 (COX-2) = 2.0 µM
IC50 (COX-1) = 0.3 µM O O O
I
H
H N S
N 5.20: R = N
S
5.23: R = O
H
O O
IC50 (COX-2) = 0.9 µM
IC50 (COX-1) = 242 µM
(5.22) for the carboxylic acid derivative (5.21) and CXCR2 receptor antagonist
improved antagonist potency.88
A noteworthy example of the impact on primary pharmacology upon bio-
isosteric replacement of the carboxylic acid OH group with a sulfonamide
moiety was evident during studies on the COX-1-selective NSAID zomepirac.
Conversion of COX-1-selective inhibitor zomepirac into a COX-2-selective
inhibitor was achieved by simply replacing its carboxylic acid group with an
sulfonimide bioisostere (5.23) (Figure 5.9).89 Although 5.23 has been co-
crystallised with COX-2, the crystal structure does not provide a rationale for
COX-2-selective inhibition by the bioisostere.89 The acylsulfonamide motif in
5.23 breeches the constriction at the mouth of the COX active site and projects
Sulfonamide as an Essential Functional Group in Drug Design 227
into the sterically unconjugated lobby region. The sulfonamide portion of the
inhibitor hydrogen bonds to the Arg120, Glu524 and Tyr355 in COX-2 in a
manner similar to the carboxylic acid-based NSAIDs. Because the three amino
acid residues are conserved in COX-1 and COX-2, their importance in deter-
mining the selectivity of the 5.23 for COX-2 inhibition remains uncertain. As
such, replacement of the carboxylic acid moiety in several NSAIDs (e.g.
indomethacin, sulindac, meclofenamic acid, etc.) with bioisosteres (e.g. esters
and amides) has provided a facile strategy for converting non-selective or COX-
1-selective inhibitors into potent and selective COX-2 inhibitors.90–92
lipid OH acid
OH
COOH
HO O
S H
O O HO O N
O N
C5H11 H2N O N N
peptide O
NH
O COOH
COOH
FPL 55712 Tomelukast
LTD4
O
H CH3
COOH
N N
N O O
H N CH3
N OH O H3C
N S
O O N
H H
(CH2)4 Cl N N S
O O
O
Pranlukast Montelukast
Zafirlukast
Figure 5.10 Evolution of cysteinyl leukotriene receptor antagonists for the treatment of asthma.
Chapter 5
CH3
N
O CH3
O H3 C
O N
H H
N S
O CH3
O O
C4H9 O Zafirlukast O
N N
H H
Lipid-like tail Indole backbone N S
O
O CH3 O
O
C 4 H9 O
N N CH3
H
O CH3 N
O CH3
COOH O O
O N N
Acid head group O N
H H H H
N S
5.24 N S
O O
Sulfonamide as an Essential Functional Group in Drug Design
O O
O O
Table 5.3 Structure, physiochemical properties, dose and Caco-2 cell permeability of carboxylic acid based NSAIDs and coxibs.
PCaco-2106 cm/sec Equilibrium solubility
Compound MW pKa Dose (mg) A to B B to A (pH 7.4) (mg/mL)
H2NO2S
N
N CF3
H3 C
H2NO2S
CH3
O
N
Chapter 5
Rofecoxib 314 – 12.5–25 24.1 19.7 0.0009
H3CO2S
COOH
Cl
H
N
Cl
Sulfonamide as an Essential Functional Group in Drug Design
COOH
CH3
233
234
COOH
H3CO
CH3
N
O Cl
CH3 COOH
H
N
CH3
Chapter 5
Piroxicam 331 1.8, 5.1 10–20 24.1 19.7 0.26
OH O
N N
H
N
S CH3
O O
OH O N
CH3
N S
H
N
S CH3
O O
Sulfonamide as an Essential Functional Group in Drug Design
235
236 Chapter 5
propensity to undergo first pass metabolism, which is apparent upon com-
parison of the oral clearance values (valdecoxib: CL/F ¼ 1.7 mL min1 kg1;
celecoxib: CL/F ¼ 9.5 mL min1 kg1).113–116
The relatively poor aqueous solubility of COX-2 inhibitors like valdecoxib
can be considerably improved by conversion of the primary sulfonamide to the
corresponding sulfonimide derivative 5.25 (Figure 5.12).117 Conversion to the
sulfonimide derivative imparts acidic character to the sulfonamide motif and
consequently provides the options of preparing salts to improve upon the
aqueous solubility. Compared with valdecoxib, whose aqueous solubility is
0.010 mg mL1, the solubility of 5.25 (as the corresponding sodium salt) is
44 mg mL1. Although 5.25 is inactive as a selective COX-2 inhibitor in vitro,
the compounds displays potent anti-inflammatory activity following oral or
intravenous administration in rats due to rapid amidase-mediated hydrolysis of
the sulfonimide moiety (T1/2 ¼ 15 min) to the active ingredient valdecoxib.
However, resistance of 5.25 towards hydrolytic cleavage in dogs, monkeys and
human liver preparations precluded further studies on this compound. SAR
studies revealed that extension of the alkyl group attached to the sulfonimide
nitrogen led to compounds (e.g. 5.26 and 5.27) (Figure 5.12) that retained the
aqueous solubility characteristics of 5.25 while undergoing facile hydrolytic
cleavage in animals and human liver preparations. Of these compounds, 5.26
(parecoxib sodium) was chosen for clinical development; a randomised, double-
blind, placebo-controlled, multiple-dose study in healthy human volunteers
demonstrated that parecoxib was rapidly absorbed (TmaxB30 min) and con-
verted to valdecoxib (elimination T1/2 ¼ 0.69 h).118 Steady state drug levels were
achieved within seven days and linear pharmacokinetics discerned for both the
prodrug and the active compound. Parecoxib sodium has been shown to be well
tolerated with no clinically significant adverse events observed in this study.
Parecoxib sodium is marketed in the European Union as Dynastat for treat-
ment of post-operative pain.119
N N
O O
CH3 Na CH3
H2N N
S R S
O O
O O
O
Valdecoxib 5.25: R = CH3
5.26: R = CH2CH3
5.27: R = (CH2)2CH3
5.2.2 Distribution
Because sulfonamide groups are often used as non-classical carboxylic acid
bioisosteres, the affinity of the functional group towards distribution in tissue
proteins relative to plasma proteins is often similar to that discerned with
carboxylic acid containing drugs. In other words, the distribution volumes (Vd)
of drugs containing the two functional groups in a bioisosteric relationship are
often comparable. Modulation of Vd for sulfonamide-containing drugs can be
influenced by lipophilicity (clogP and/or clogD) and acidity (pKa) in a manner
similar for compounds of all physiochemical classes (acids, bases and neu-
trals).123 Thus, increases in lipophilicity generally coincide with increases in Vd,
and a more acidic character of the sulfonamide group will generally lead to a
lower Vd owing to extensive plasma protein (e.g. serum albumin) as discerned
with carboxylic acid analogs. These principles were outlined in Chapter 2. The
pharmacokinetic attributes of structurally diverse sulfonamide derivatives124 in
humans are shown in Table 5.4 and provide a glimpse of the relationship
between physiochemical properties and pharmacokinetic parameters such as
plasma free fraction (fu), Vd and clearance.
Loop diuretics
Acetazolamide 222 1.13 0.48 0.04 0.65 0.37
N
N
SO2NH2
S
H3COCHN
COOH
O
H2NO2S N
H
Cl
OH SO2NH2
NH Cl
O
Chapter 5
Thiazides
Hydroflumethiazide 330 0.21 0.33 N/A 9.70 2.20
O2
H2NO2S S
NH
F 3C N
H
Sulfonylureas
Chlorpropamide 276 2.35 0.32 0.03 0.05 0.19
O O
O
S
N N
H H
Cl
O O
Sulfonamide as an Essential Functional Group in Drug Design
O
S
N N
H H
(H2C)2
O NH
N
N
239
240
Cl
Tolbutamide 270 2.50 0.39 0.05 1.90 0.12
O O O
S
N N
H H
H3C
Sulfanilamides
Dapsone 248 0.89 0.94 0.25 0.48 0.83
H 2N NH2
S
O O
Chapter 5
Sulfadiazine 250 0.10 0.66 0.44 0.55 0.29
H2 N
H
N N
S
O O N
H 2N
H
N N
S
O
O O
CH3
H 2N
CH 3
Sulfonamide as an Essential Functional Group in Drug Design
H
N
S CH 3
O O O N
241
242 Chapter 5
CH3
Cl N
HN O
HO O HN
O NH2 H O
S O N S S
S O NH O O NH2 O
H3CO
S S NH2 CH3 N N S O
O O O NH2
Cl
L-693,612 Chlorthalidone Acetazolamide Indapamide
Figure 5.13 Examples of sulfonamides with preferential red blood cell partitioning
due to carbonic anhydrase binding.
binding sites in red blood cells; consequently, the free fraction available in
plasma for elimination is much greater (blood to plasma ratio reduces to B6),
resulting in significant increases in blood clearance. Likewise, the observed
dose-dependent increases in Vd of L-693 612 are consistent with the hypothesis
that high-affinity binding to CA confines the compound largely to the blood
compartment at low doses, but saturation of CA binding sites at high doses
increases availability to peripheral tissues.
Preferential partitioning of sulfonamides into red blood cells renders them
susceptible to significant pharmacokinetic interactions with agents that may
compete for binding to CA. For instance, a significant drug–drug interaction
has been noted in the clinic with the CA inhibitors and anti-hypertensive agents
chlorthalidone and acetazolamide (Figure 5.13), wherein acetazolamide was
able to displace chlorthalidone from blood cells following administration of the
two medications in humans.127 Thus, intravenous administration of acet-
azolamide to humans who had previously received [14C]-chlorthalidone resul-
ted in a marked drop in red blood cell radioactivity, whereas that in plasma
increased. A likely explanation for the pharmacokinetic interaction is the
relatively higher affinity of acetazolamide (versus chlorthalidone) towards
binding to CA as reflected from the greater than six-fold higher blood to
plasma ratio of acetazolamide; the blood-to-plasma ratio of acetazolamide and
chlorthalidone has been estimated to be 467 and 70, respectively.128 Drug–drug
interactions due to drug displacement from red blood cells have also been noted
with the anti-hypertensive agents and CA inhibitors indapamide (Figure 5.13),
chlorthalidone and acetazolamide in the rat. Both chlorthalidone and acet-
azolamide were able to displace indapamide from rat erythrocytes due to
competition for CA binding sites in vivo.129
5.2.3 Metabolism
Presence of the sulfonamide moiety in a molecule in a secondary or tertiary
amide expression usually provides an inert, metabolically robust group that is
not typically vulnerable to phase I or phase II metabolising enzymes. Contrary
to facile enzyme-catalysed hydrolysis of esters and amides, sulfonamides are
generally resistant to cleavage by amidases. An exception to the rule is sulfo-
nimide (acylsulfonamide) derivatives such as valdecoxib, which can undergo
Sulfonamide as an Essential Functional Group in Drug Design 243
N
O O N
N
S
N N
H
NH
H3C O
N O
O O N
N N
S O O N
N N N
S
N N N
H3C O
HN
Sildenafil O
? HO S HO S
O O O
5.31 5.32
HOOC O O
HO O
HO O S
CYP UGT HO N
HN S OH H
N
O O
O 5.30
5.28
CH3
HOOC
H2 N S O H
UGT HO N
O O HO S
OH
Valdecoxib O O
5.29
Scheme 5.3 Metabolism of the sulfonamide group as illustrated with the selective
COX-2 inhibitor and anti-inflammatory agent valdecoxib.
discerned with esters and amides. Instead, the studies implicate N-hydroxyla-
tion of the sulfonamide group as the rate-limiting step in the formation of the
sulfinic and sulfonic acid metabolites.132–135 For example, the decomposition of
benzenesulfohydroxamic acid (Piloty’s acid) (Scheme 5.4) and some of its
derivatives under anaerobic and strong alkaline conditions has been reported to
yield benzenesulfinic acid and nitrous oxide (N2O).132,133 Under physiological
conditions (aerobic environment, phosphate buffer, pH 7.4), Piloty’s acid
decomposes to form benzenesulfinic acid and nitric oxide (NO), which can be
intercepted with a nitrone free radical trap.134 A mechanism, which depicts the
formation of NO in the oxidative decomposition of Piloty’s acid is shown
in Scheme 5.4.135 Consistent with the general instability of N-hydroxy-
sulfonamides under neutral aerobic conditions, attempts to isolate the N-
hydroxy metabolite of valdecoxib, i.e. 5.28, in pure form by chemical synthesis
were met with failure due to facile decomposition at room temperature to
afford the corresponding sulfinic acid derivative 5.31.136 However, work-up and
crystallisation of crude 5.28 using EDTA and ascorbic acid to remove sources
of chelating metal ion traces and/or oxidants has afforded a stable mono-
hydrate form of 5.28 whose water content remained constant at room tem-
peratures (under standard humidity conditions) over a period of two years.136
It is interesting to note that 5.28 is an active metabolite of valdecoxib; however,
the COX-2 inhibitory potency and selectivity of 5.28 is significantly lower than
that of the parent compound.
Finally, it is noteworthy to point out that the sulfonamide group in celecoxib
is resistant to N-hydroxylation or N-glucuronidation in animals and
humans.116,137 In valdecoxib, CYP mediated hydroxylation occurs on the sul-
fonamide nitrogen, the 5-isooxazylmethyl group as well on the unsubstituted
phenyl ring (Scheme 5.5); in contrast, celecoxib is exclusively metabolised via
hydroxylation of the para-tolyl methyl group (Scheme 5.5). The difference in
O
O OH O OH
O O O OH
O O S O S
S O S O N OH
S O N N
S OH N
N NO
H H
H
HOOC
O H
HO N
HO S
OH
O O
O
H3 C
N HO
UGT HO
O
CF3 CF3
CYP
OH N N CF3
N N
HO N N
N
O CYP
H2 N S H2 N S H2 N
CH3 S
O O Celecoxib O O
O O
H2 N S HO
O O HN
Valdecoxib S
O O
Scheme 5.5 Differences in metabolic profile of the selective COX-2 inhibitors and diarylheterocycle derivatives valdecoxib and celecoxib.
Chapter 5
Sulfonamide as an Essential Functional Group in Drug Design 247
CH3 CH3
O O N O O N
S S
N N CH3 N N O
H H
CH3
H2 N H2N
Sulfasomidine Sulfamethomidine
metabolic profile between the two compounds is fairly intriguing and suggests
difference(s) in CYP active site binding modes.
Besides valdecoxib, sulfanilamide anti-bacterial drugs are also prone to N-
glucuronidation on the sulfonamide nitrogen. The degree to which N-glucur-
onidation occurs is highly dependent on the functional group attached to the
sulfonamide nitrogen atom as well as the species under consideration. For
example, the process of N-glucuronidation on the sulfonamide nitrogen in
sulfamethomidine is not observed with sulfasomidine (Figure 5.14).138 The
major structural difference is that the pendant 2-methoxy-4-methylpyrimidine
substituent in sulfamethomidine is replaced by the 2,4-dimethylpyrimidine
motif in sulfasomidine. Furthermore, while sulfonamide glucuronidation con-
stituted a major route of metabolism of sulfamethomidine in humans and
primates, the N-glucuronide metabolite was not detected in rats and rabbits.138
R3
N N
R2
R1
% Metabolised by
Compound R1 R2 R3 CYP2C9 (1 mM)
Thiazides
Cl
H H H
Cl N Cl N Cl N Cl N
Cl
H2N NH H2N NH H2N NH H2N NH
S S S S S S S S
O O O O O O O O O O O O O O O O
Chlorothiazide Hydrochlorothiazide Cyclothiazide Trichloromethiazide
HN O
O Cl NH N NH2 H3C
O N N NH2
S O N
O S O
H2N H2N S O
S COOH S COOH H3 C N S
H H3C N O
O O O O
Figure 5.15 Sulfonamide-based substrates for the human renal OAT transporters.
O O N N O O N N O O N N
S S S CH3
N CH3 N CH3 N O
H H H
H2N H2N H2N
Sulfisomidine Sulfametomidine Sulfadimethoxine
CLR = 231 mL/min CLR = 21 mL/min CLR = 10.8 mL/min
O
HO SG S
HO O N HN G
NH2 NH N
CYP2C9
or
myeloperoxidase [O] GSH
NAT 2 [O]
O
NO2
H3C NH
O S
NH
O S O
NH
O N
N O
O CH3
CH3 5.39
Data from the 11 008 patients studied demonstrated that the incidence of
hypersensitivity with celecoxib was statistically indifferent from that observed
with placebo or active comparators (NSAIDs) when the entire cohort was
examined. Although skin reactions occurred with greater frequency in patients
with a history of sulfonamide functionality, the trend was consistent across all
three treatment groups (celecoxib, NSAIDs and placebo), indicating a general
patient susceptibility rather than a functional group-specific effect.
Furthermore, a prospective study assessed the incidence of cross-reactivity in
28 patients with a self-reported history of sulfonamide antibacterial allergy who
then received celecoxib.195 Sulfonamide allergy was assessed by either skin
prick, intradermal testing, and/or lymphocyte toxicity assay (LTA) using sul-
famethoxazole. Positive reactions occurred in four out of 28 patients admi-
nistered a skin test and two out of 10 patients evaluated by LTA. Patients with
a negative skin test received an oral challenge with sulfamethoxazole. All
patients underwent a low- and high-dose oral challenge test with celecoxib
without any adverse events. Follow-up in 25 patients showed that 15 (60%)
continued to tolerate therapy with celecoxib, five did not take celecoxib after
the oral challenge, and five discontinued celecoxib therapy secondary to
adverse effects. In the six patients with a positive skin test, four continued to
take celecoxib, one discontinued therapy secondary to gastrointestinal adverse
Sulfonamide as an Essential Functional Group in Drug Design 255
effects, and one did not take the drug after the oral challenge upon the advice of
the primary care physician. The authors concluded that the risk for cross-
reactivity of celecoxib in patients with a sulfonamide antibacterial allergy is
low.
O
O O
H O O HS N COOH
O N O H H
N O O H
S S CYP N N O NH
O S GSH S
O
CH3 CH3 O O GSH =
NO2 NH2 CH3 CH3
NH NH2
Nimesulide
5.42 5.43 SG
H2N COOH
Scheme 5.7 Bioactivation of the aryl sulfonamide derivative nimesulide to reactive metabolites.
Chapter 5
Sulfonamide as an Essential Functional Group in Drug Design 257
H CF3
N N H
CYP N N
X X GSH
R R X R=
N N R N NH
N CH3
H
GS H
N
5.44: X = CO 5.46: X = CO CH3
5.45: X = SO2 5.47: X = SO2
N O N O COOH N
GSH
S O S NH S COOH
S * NH2 H S O S O
35
SO2 HN HN
GSH NH3 5.49 5.50
5.48 CH3
35
*= S label COOH
H2N
HOOC
O
HO S S N
HO
OH SH
N
S
5.52 5.51
O HO O O
N O
O S
S
R NH2
O S NH2
R
5.53: R = H O O
O 5.58: R = CH3
S S 5.60: R = H
NH2
5.54: R = R
HO N O
HO O
O S
O
O NH2
S 5.57
R
O S NH2 5.59: R = CH3
5.61: R = H
5.55: R = H
O
5.56: R =
Figure 5.17 SAR studies on the in vitro reactivity of glutathione with 2-sulfonamido-
heterocycles.
OH OH
CN CF3
O O O O
HN HN
S N S N
O O O O
PNU-109112 PNU-140690
N OH
N C N
N
GS
GST/GS GS
GS +
H H N
HN N N S GS N O O
S N N S
O O O SO2
O O O NH2
PNU-109112
Scheme 5.10 SAR and mechanistic studies on the glutathione mediated nucleophilic displacement of pyridinyl-2-sulfonamides.
Chapter 5
Sulfonamide as an Essential Functional Group in Drug Design 261
N COOH
Cl N N H
Cl N N CH3 S NH
S GSH S
S O O
O pH 7.4
phosphate Cl
Cl buffer
5.62 5.63 COOH
H2 N
O H
H H GS N
R1 N N ? C GSH
N R2
S R2
Sulofenur : R1 = R2 = R2
O O O
Cl O R1 NH2
S
H3C O O 5.64: R2 = 5.65: R2 =
Cl Cl
Tolbutamide : R1 = R2 =
5.66: R2 = 5.67: R2 =
5.5 Conclusions
The juxtaposition of terminologies around the sulfonamide-containing drugs
and sulfonamide drugs can cause confusion around the issue of hypersensitivity
reactions (sulfa allergy). Historically, the term ‘sulfa’ refers to a derivative of
the antibacterial agent sulfanilamide. More recently, the term has been applied
generously to a diverse group of drugs, all of which contain the sulfonamide
motif. This is largely for historic reasons, as many of the drugs were approved
prior to any toxicity mechanism investigations. In addition, some non-anti-
bacterial sulfonamides (e.g. sulfonylureas and loop diuretics) are classified as
derivatives of the antibacterials because their discovery has been influenced in
large part by the polypharmacology of the original sulfa drugs.
The sulfa allergy association raises concerns for medicinal chemistry and the
use of the sulfonamide group in drug discovery. However, to date, a thorough
examination of the literature does not support the dogma of sulfonamide
antibacterial cross reactivity with non-antibacterial sulfonamides. In fact, rig-
orous studies examining cross reactivity between sulfonamide antibiotics and
sulfonamide non-antibiotics suggest that predisposition to allergic reactions is
the primary risk factor for the association, not a true cross reactivity.95,194
In retrospect, the sulfonamide group should remain an essential part of the
medicinal chemist’s arsenal in the search of new chemical entities with phar-
macological action. There are numerous examples of blockbuster drugs which
contain this motif and yet are not associated with any significant incidence of
sulfa allergy type adverse drug reactions. In the current climate, which is
focused on a heightened awareness of toxicophores (structural alerts) and
toxicity, it is recommended that the sulfonamide motif should not be treated as
a toxicophore, although appropriate caution needs to be exercised in terms of
novel bioactivation pathways involving this functional group.
Finally, the finding that metabolic elimination of many acidic sulfonamide
derivatives is principally mediated by CYP2C9 needs to be taken into con-
sideration from a drug design perspective, given the polymorphic expression of
this isozyme in humans and its contribution towards variability in pharmaco-
kinetic and pharmacodynamic response.
264 Chapter 5
References
1. G. Domagk, Deut. Med. Wochschr., 1935, 61, 250.
2. J. Tréfouel, J. Mme. Tréfouel, F. Nitti and D. Bovet, C. R. Soc. Bio.
Paris., 1935, 120, 756.
3. N. Anand, in Burger’s Medicinal Chemistry, 4th edn, ed. M. E. Wolff,
Wiley-Interscience, New York, 1979, Part II, Chapter 13, p. 13.
4. D. Woods, Brit. J. Exptl. Pathol., 1940, 21, 74.
5. D. Woods and P. Fildes, Chem. Ind., 1940, 59, 133.
6. P. H. Bell and R. O. Roblin Jr., J. Am. Chem. Soc., 1942, 64, 2903.
7. A. Burger, in Progress in Drug Research, ed. H. von E. Jucker,
Birkhäuser-Verlag, Basel, 1991, vol. 37, pp. 287–371.
8. C. J. Hawkey, Lancet, 1999, 353, 307.
9. D. L. DeWitt and W. L. Smith, Proc. Natl. Acad. Sci. U. S. A., 1988, 85,
1412.
10. T. Hla and K. Neilson, Proc. Natl. Acad. Sci. U. S. A., 1992, 89,
7384.
11. S. H. Lee, E. Soyoola, P. Chanmugam, S. Hart, W. Sun, H. Zhong,
S. Liou, D. Simmons and D. Hwang, J. Biol. Chem., 1992, 267,
25934.
12. D. A. Kujubu, B. S. Fletcher, B. C. Varnum, R. W. Lim and H. R.
Herschman, J. Biol. Chem., 1991, 266, 12866.
13. J. R. Vane, J. A. Mitchell, I. Appleton, A. Tomlinson, D. Bishop-Bailey,
J. Croxtall and D. A. Willoughby, Proc. Natl. Acad. Sci. U. S. A., 1994,
91, 2046.
14. C. J. Smith, Y. Zhang, C. M. Kobolt, J. Muhammad, B. S. Zweifel,
A. Shaffer, J. J. Talley, J. L. Masferrer, K. Seibert and P. C. Isakson,
Proc. Natl. Acad. Sci. U. S. A., 1998, 95, 13313.
15. T. D. Warner, F. Giuliano, I. Vojnovic, A. Bukasa, J. A. Mitchell and
J. R. Vane, Proc. Natl. Acad. Sci. U. S. A., 1999, 96, 7563.
16. L. J. Marnett, Annu. Rev. Pharmacol. Toxicol., 2009, 49, 265.
17. L. H. Rome and W. E. M. Lands, Proc. Natl. Acad. Sci. U. S. A., 1975, 72,
4863.
18. R. A. Copeland, J. M. Williams, J. Giannaras, S. Nurnberg, M.
Covington, D. Pinto, S. Pick and J. M. Trzaskos, Proc. Natl. Acad. Sci.
U. S. A., 1994, 91, 11202.
19. J. J. Talley, Prog. Med. Chem., 1999, 36, 201.
20. R. Flower, R. Gryglewski, K. Herbaczynska-Cedro and J. R. Vane,
Nature New Biol., 1972, 238, 104.
21. L. J. Marnett and A. S. Kalgutkar, in Milestones in Drug Therapy–COX-2
Inhibitors, ed. M. Pairet and J. van Ryn, Birkhäuser, Verlag, Basel, 2004,
pp. 15–40.
22. K. R. Gans, W. Galbraith, R. J. Roman, S. B. Haber, J. S. Kerr, W. K.
Schmidt, C. Smith, W. E. Hewes and N. R. Ackerman, J. Pharmacol. Exp.
Ther., 1990, 254, 180.
Sulfonamide as an Essential Functional Group in Drug Design 265
23. Y. Leblanc, J. Y. Gauthier, D. Ethier, J. Guay, J. Mancini, D. Riendeau,
P. Tagari, P. Vickers, E. Wong and P. Prasit, Bioorg. Med. Chem. Lett.,
1995, 5, 2123.
24. S. R. Bertenshaw, J. J. Talley, D. J. Rogier, M. J. Graneto, R. S. Rogers,
S. W. Kramer, T. D. Penning, C. M. Koboldt, A. W. Veenhuizen and
Y. Zhang, et al., Bioorg. Med. Chem. Lett., 1995, 5, 2919.
25. T. D. Penning, J. J. Talley, S. R. Bertenshaw, J. S. Carter, P. W. Collins,
S. Doctor, M. J. Graneto, L. F. Lee, J. W. Malecha and J. M. Miyashiro,
et al., J. Med. Chem., 1997, 40, 1347.
26. A. S. Kalgutkar, B. C. Crews and L. J. Marnett, Biochemistry, 1996, 35,
9076.
27. J. S. Carter, Expert Opin. Ther. Pat., 2000, 10, 1011.
28. D. Picot, P. J. Loll and R. M. Garavito, Nature, 1994, 367, 243.
29. R. G. Kurumbail, A. M. Stevens, J. K. Gierse, J. J. McDonald,
R. A. Stegeman, J. Y. Pak, D. Gildehaus, J. M. Miyashiro, T. D.
Penning, K. Seibert, P. C. Isakson and W. C. Stallings, Nature, 1996, 384,
644.
30. D. K. Bhattacharyya, M. Lecomte, C. J. Rieke, R. M. Garavito and W. L.
Smith, J. Biol. Chem., 1996, 271, 2179.
31. J. A. Mancini, D. Riendeau, J.-P. Falgueyret, P. J. Vickers and G. P.
O’Neill, J. Biol. Chem., 1995, 270, 29372.
32. J. K. Gierse, J. J. McDonald, S. D. Hauser, S. H. Rangwala, C. M.
Koboldt and K. Seibert, J. Biol. Chem., 1996, 271, 15810.
33. E. Wong, C. Bayly, H. L. Waterman, D. Riendeau and J. A. Mancini,
J. Biol. Chem., 1997, 272, 9280.
34. J. Harrington, J. E. Robertson, D. C. Kvam, R. R. Hamilton, K. T.
McGurran, R. J. Trancik, K. F. Swingle, G. G. Moore and J. F. Gerster,
J. Med. Chem., 1970, 13, 137.
35. K. F. Swingle, G. G. Moore and T. J. Grant, Arch. Int. Pharmacodyn.,
1976, 221, 132.
36. R. Davis and R. N. Brogden, Drugs, 1994, 48, 431.
37. J. L. Masferrer, B. S. Zweifel, P. T. Manning, S. D. Hauser, K. M. Leahy,
W. G. Smith, P. C. Isakson and K. Seibert, Proc. Natl. Acad. Sci. U. S. A.,
1994, 91, 3228.
38. N. Futaki, S. Takahashi, T. Kitagawa, Y. Yamakawa, M. Tanaka and
S. Higuchi, Inflamm. Res., 1997, 46, 496.
39. I. Wiesenberg-Bottcher, A. Schweizer, J. R. Green, Y. Seltenmeyer and
K. Muller, Agents Actions, 1989, 26, 240.
40. T. Vago, M. Bevilacqua and G. Norbiato, Arzneim.-Forsch., 1995, 45,
1096.
41. C. S. Li, W. C. Black, C. C. Chan, A. W. Ford-Hutchinson, J. Y. Gauthier,
R. Gordon, D. Guay, S. Kargman, C. K. Lau and J. Mancini, et al.,
J. Med. Chem., 1995, 38, 4897.
42. K. Tsuji, K. Nakamura, N. Konishi, H. Okumura and M. Matsuo, Chem.
Pharm. Bull., 1992, 40, 2399.
266 Chapter 5
43. K. Tanaka, T. Shimotori, S. Makino, Y. Aikawa, T. Inaba, C. Yoshida
and S. Takano, Arzneimittelforschung, 1992, 42, 935.
44. A. K. Singh, M. Chawla and A. Singh, J. Pharm. Pharmacol., 2000, 52,
467.
45. F. Julémont, X. de Leval, C. Michaux, J.-F. Renard, J.-Y. Winum, J.-L.
Montero, J. Damas, J.-M. Dogne and B. Pirotte, J. Med. Chem., 2004, 47,
6749.
46. L. J. Marnett, S. W. Rowlinson, D. C. Goodwin, A. S. Kalgutkar and
C. A. Lanzo, J. Biol. Chem., 1999, 274, 22903.
47. C. T. Supuran, Curr. Pharm. Design, 2008, 14, 603.
48. T. Stams and D. W. Christianson, in The Carbonic Anhydrases–New
Horizons, ed. W. R. Chegwidden, Y. Edwards and N. Carter, Birkhäuser
Verlag, Basel, 2000, pp. 159–174.
49. A. Di Fiore, S. M. Monti, M. Hilvo, S. Parkkila, V. Romano, A. Scaloni,
C. Pedone, A. Scozzafava, C. T. Supuran and G. De Simone, Proteins,
2009, 74, 164.
50. C. Temperini, A. Cecchi, A. Scozzafava and C. T. Supuran, Bioorg. Med.
Chem. Lett., 2008, 18, 2567.
51. C. Y. Kim, D. A. Whittington, J. S. Chang, J. Liao, J. A. May and D. W.
Christianson, J. Med. Chem., 2002, 45, 888.
52. D. A. Whittington, A. Waheed, B. Ulmasov, G. N. Shah, J. H. Grubb,
W. S. Sly and D. W. Christianson, Proc. Natl. Acad. Sci. U. S. A., 2001,
98, 9545.
53. S. K. Nair, J. F. Krebs, D. W. Christianson and C. A. Fierke, Biochem-
istry, 1995, 34, 3981.
54. K. Fridborg, K. K. Kannan, A. Liljas, J. Lundin, B. Strandberg,
R. Strandberg, B. Tilander and G. Wirén, J. Mol. Biol., 1967, 25, 505.
55. A. Liljas, K. K. Kannan, P. C. Bergstén, I. Waara, K. Fridborg,
B. Strandberg, U. Carlbom, L. Järup, S. Lövgren and M. Petef, Nature
New Biol., 1972, 235, 131.
56. B. W. Clare and C. T. Supuran, Expert Opin. Drug Metab. Toxicol., 2006,
2, 113.
57. F. Z. Smaine, F. Pacchiano, M. Rami, V. Barragan-Montero, D. Vullo,
A. Scozzafava, J. Y. Winum and C. T. Supuran, Bioorg. Med. Chem.
Lett., 2008, 18, 6332.
58. O. Guzel, A. Innocenti, A. Scozzafava, A. Salman, S. Parkkila, M. Hilvo
and C. T. Supuran, Bioorg. Med. Chem. Lett., 2008, 16, 9113.
59. K. D’Ambrosio, R. M. Vitale, J. M. Dogné, B. Masereel, A. Innocenti,
A. Scozzafava, G. De Simone and C. T. Supuran, J. Med. Chem., 2008,
51, 3230.
60. A. Thiry, S. Rolin, D. Vullo, A. Frankart, A. Scozzafava, J. M. Dogné,
J. Wouters, C. T. Supuran and B. Masereel, Eur. J. Med. Chem., 2008, 43,
2853.
61. C. T. Supuran and A. Scozzafava, Bioorg. Med. Chem., 2007, 15, 4336.
62. B. L. Wilkinson, L. F. Bornaghi, T. A. Houston, A. Innocenti, C. T.
Supuran and S. A. Poulsen, J. Med. Chem., 2006, 49, 6539.
Sulfonamide as an Essential Functional Group in Drug Design 267
63. J. Caprioli, in Aldler’s Physiology of the Eye–Clinical Application, ed.
R. A. Moses and W. M. Hart, The C. V. Mosby Company, St Louis,
1987, pp. 204–222.
64. P. R. Lichter, L. P. Newman, N. C. Wheeler and O. V. Beal, Am. J.
Ophthalmol., 1978, 85, 495.
65. D. L. Epstein and W. M. Grant, Arch. Ophthalmol., 1977, 95, 1378.
66. F. Mincione, A. Scozzafava and C. T. Supuran, Curr. Pharm. Des., 2008,
14, 649.
67. C. Temperini, A. Cecchi, A. Scozzafava and C. T. Supuran, Org. Biomol.
Chem., 2008, 6, 2499.
68. Y. Masuda and T. Karasawa, Arzneimittelforschung, 1993, 43, 416.
69. K. Köhler, A. Hillebrecht, J. Schulze-Wischeler, A. Innocente, A. Heine
and C. T. Supuran, Angew. Chem., Int. Ed. Engl., 2007, 46, 7697.
70. J.-M. Dogné, A. Thiry, D. Pratico, B. Masereel and C. T. Supuran, Curr.
Topics Med. Chem., 2007, 7, 885.
71. A. Di Fiore, C. Pedone, K. D’Ambrosio, A. Scozzafava, G. De Simone
and C. T. Supuran, Bioorg. Med. Chem. Lett., 2006, 16, 437.
72. A. Weber, A. Casini, A. Heine, D. Kuhn, C. T. Supuran, A. Scozzafava
and G. Klebe, J. Med. Chem., 2004, 47, 550.
73. C. T. Supuran, Nat. Rev. Drug Discov., 2008, 7, 168.
74. J. C. Henquin, Diabetologia, 1992, 35, 907.
75. M. M. Loubatières-Mariani, J. Soc. Biol., 2007, 201, 121–125.
76. H. Kleinsorge, Exp. Clin. Endocrinol. Diabetes, 1998, 106, 149.
77. A. Farret, L. Lugo-Garcia, F. Galtier, R. Gross and P. Petit, Fundam.
Clin. Pharmacol., 2005, 19, 647.
78. F. M. Gribble and F. M. Ashcroft, Metabolism, 2000, 49, 3.
79. S. Fukuen, M. Iwaki, A. Yasui, M. Makishima, M. Matsuda and
I. Shimomura, J. Biol. Chem., 2005, 280, 23653.
80. M. Scarsi, M. Podvinec, A. Roth, H. Hug, S. Kersten, H. Albrecht,
T. Schwede, U. A. Meyer and C. Rucker, Mol. Pharmacol., 2007, 71, 398.
81. G. R. Brown and A. J. Foubister, J. Med. Chem., 1984, 27, 79.
82. A. Dornhorst, Lancet, 2001, 358, 1709.
83. P. Lin, D. Marino, J. L. Lo, Y. T. Yang, K. Cheng, R. G. Smith, M. H.
Fisher, M. J. Wyratt and M. T. Goulet, Bioorg. Med. Chem. Lett., 2001,
11, 1073.
84. C. R. McCurdy, R. M. Jones and P. S. Portoghese, Org. Lett., 2000, 2,
819.
85. R.-J. Liu, J. A. Tucker, T. Zinevitch, O. Kirichenko, V. Konoplev,
S. Kuznetsova, S. Sviridov, J. Pickens, S. Tandel, E. Brahmachary and
Y. Yang, et al., J. Med. Chem., 2007, 50, 6535.
86. Y. K. Yee, P. R. Bernstein, E. J. Adams, F. J. Brown, L. A. Cronk, K. C.
Hebbel, E. P. Vacek, R. D. Krell and D. W. A. Snyder, J. Med. Chem.,
1990, 33, 2437.
87. M. Bäck, P. O. Johansson, F. Wangsell, F. Thorstensson, I. Kvarnström,
S. Ayesa, H. Wähling, M. Pelcman, K. Jansson, S. Lindström and
H. Wallberg, et al., Bioorg. Med. Chem. lett., 2003, 11, 2551.
268 Chapter 5
88. M. P. Winters, C. Crysler, N. Subasinghe, D. Ryan, L. Leong, S. Zhao,
R. Donatelli, E. Yurkow, M. Mazzulla, L. Boczon, C. L. Manthey,
C. Molloy, H. Raymond, L. Murray, L. McAlonan and B. Tomczuk,
Bioorg. Med. Chem. Lett., 2008, 18, 1926.
89. C. Luong, A. Miller, J. Barnett, J. Chow, C. Ramesha and M. F.
Browner, Nat. Struct. Biol., 1996, 3, 927.
90. A. S. Kalgutkar, B. C. Crews, S. W. Rowlinson, A. B. Marnett, K. R.
Kozak, R. P. Remmel and L. J. Marnett, Proc. Natl. Acad. Sci. U. S. A.,
2000, 97, 925.
91. A. S. Kalgutkar, A. B. Marnett, B. C. Crews, R. P. Remmel and L. J.
Marnett, J. Med. Chem., 2000, 43, 2860.
92. A. S. Kalgutkar, S. W. Rowlinson, B. C. Crews and L. J. Marnett, Bioorg.
Med. Chem. Lett., 2002, 12, 521.
93. R. C. Murphy, S. Hammarström and B. Samuelsson, Proc. Natl. Acad.
Sci. U.S.A., 1979, 76, 4275.
94. J. M. Drazen, K. F. Austen, R. A. Lewis, D. A. Clark, G. Goto,
A. Marfat and E. J. Corey, Proc. Natl. Acad. Sci. U.S.A., 1980, 77,
4345.
95. J. Augstein, J. B. Farmer, T. B. Lee, P. Sheard and M. L. Tattersall,
Nature (New Biol.), 1973, 245, 215.
96. F. J. Brown, P. R. Bernstein, L. A. Cronk, L. L. Dosset, K. C. Hebbel,
T. P. Maduskuie, H. S. Shapiro, E. P. Vacek, Y. K. Lee, A. K. Willard,
R. D. Krell and D. W. Snyder, J. Med. Chem., 1989, 32, 807.
97. R. W. Fuller, P. N. Black and C. T. Dollery, J. Allergy Clin. Immunol.,
1989, 83, 939.
98. M. L. Cloud, G. C. Enas, J. Kemp, T. Platts-Mills, L. C. Altman,
R. Townley, D. Tinkelman, T. King Jr, E. Middleton and A. L. Sheffer,
et al., Am. Rev. Respir. Dis., 1989, 140, 1336.
99. H. Nakai, M. Konno, S. Kosuge, S. Sakuyama, M. Toda, Y. Arai,
T. Obata, N. Katsube, T. Miyamoto, T. Okegawa and A. Kawasaki,
J. Med. Chem., 1988, 31, 84.
100. Y. Taniguchi, G. Tamura, M. Honma, T. Aizawa, N. Maruyama,
K. Shirato and T. Takishima, J. Allergy Clin. Immunol., 1993, 92, 507.
101. F. J. Brown, Y. K. Yee, L. A. Cronk, K. C. Hebbel, R. D. Krell and
D. W. Synder, J. Med. Chem., 1990, 33, 1771.
102. Y. K. Yee, P. R. Bernstein, E. J. Adams, F. J. Brown, L. A. Cronk, K. C.
Hebbel, E. P. Vacek, R. D. Krell and D. W. Synder, J. Med. Chem., 1990,
33, 2437.
103. V. G. Matassa, F. J. Brown, P. R. Bernstein, H. S. Shapiro, T. P.
Maduskuie Jr., L. A. Cronk, E. P. Vacek, Y. K. Yee, D. W. Synder, R. D.
Krell, C. L. Lerman and J. J. Maloney, J. Med. Chem., 1990, 33, 2621.
104. V. G. Matassa, T. P. Maduskuie Jr, H. S. Shapiro, B. Hesp, D. W.
Snyder, D. Aharony, R. D. Krell and R. A. Keith, J. Med. Chem., 1990,
33, 1781.
105. P. R. Bernstein, Am. J. Respir. Crit. Care Med., 1998, 157, S220.
Sulfonamide as an Essential Functional Group in Drug Design 269
106. C. A. Lipinski, F. Lombardo, B. W. Dominy and P. J. Feeny, Adv. Drug
Deliv. Rev., 1997, 23, 3.
107. Y. C. Martin, J. Med. Chem., 2005, 48, 3164.
108. K. Palm, P. Stenberg, K. Luthman and P. Artursson, Pharm. Res., 1997,
14, 568.
109. D. F. Veber, S. R. Johnson, H.-Y. Cheng, B. R. Smith, K. W. Ward and
K. D. Kopple, J. Med. Chem., 2002, 45, 2615.
110. P. D. Leeson and A. M. Davis, J. Med. Chem., 2004, 47, 6338.
111. M. Yazdanian, K. Briggs, C. Jankovsky and A. Hawi, Pharm. Res., 2004,
21, 293.
112. S. G. Vijaya Kumar and D. N. Mishra, Chem. Pharm. Bull. (Tokyo),
2006, 54, 1102.
113. N. Sarapa, M. R. Britto, M. B. Mainka and K. Parivar, Eur. J. Clin.
Pharmacol., 2005, 61, 247.
114. J. J. Yuan, D.-C. Yang, J. Y. Zhang, R. Bible Jr., A. Karim and J. W. A.
Findlay, Drug Metab. Dispos., 2002, 30, 1013.
115. N. M. Davies, A. J. McLachlan, R. O. Day and K. M. Williams, Clin.
Pharmacokinet., 2000, 38, 225.
116. S. K. Paulson, J. D. Hribar, N. W. K. Liu, E. Hajdu, R. H. Bible Jr.,
A. Piergies and A. Karim, Drug Metab. Dispos., 2000, 28, 308.
117. J. J. Talley, S. R. Bertenshaw, D. L. Brown, J. S. Carter, M. J. Graneto,
M. S. Kellog, C. M. Koboldt, J. Yuan, Y. Y. Zhang and K. Seibert,
J. Med. Chem., 2000, 43, 1661.
118. V. A. Mehta, R. Johnston, R. Cheung, A. Bello and R. M. Langford,
Clin. Pharmacol. Ther., 2008, 83, 430.
119. S. F. Barton, F. F. Langeland, M. C. Snabes, D. LeComte, M. E. Kuss,
S. S. Dhadda and R. C. Hubbard, Anesthesiology, 2002, 97, 306.
120. P. E. Golstein, A. Boom, J. van Geffel, P. Jacobs, B. Masereel and
R. Beauwens, Pflügers Arch–Eur. J. Physiol., 1999, 437, 652.
121. L. Payen, L. Delugin, A. Courtois, Y. Trinquart, A. Guillouzo and
O. Fardel, Br. J. Pharmacol., 2001, 132, 778.
122. C. Gedeon, J. Behravan, G. Koren and M. Piquette-Miller, Placenta,
2006, 27, 1096.
123. D. A. Smith, B. C. Jones and D. K. Walker, Med. Res. Rev., 1996, 16,
243.
124. R. S. Obach, F. Lombardo and N. J. Waters, Drug Metab. Dispos., 2008,
36, 1385.
125. W. Rupp, O. Christ and W. Heptner, Arzneimittelforschung, 1969, 19,
1428.
126. B. K. Wong, P. J. Bruhin and J. H. Lin, Pharm. Res., 1994, 11, 438.
127. B. Beerman, K. Hellström, B. Lindström and A. Rosén, Clin. Pharmacol.
Ther., 1975, 17, 424.
128. S. Yu, S. Li, H. Yang, F. Lee, J.-T. Wu and M. G. Qian, Rapid Commun.
Mass Spectrom., 2005, 19, 250.
129. J. T. Lettieri and S. T. Portelli, J. Pharmacol. Exp. Ther., 1983, 224, 269.
270 Chapter 5
130. D. K. Walker, M. J. Ackland, G. C. James, G. J. Muirhead, D. J. Rance,
P. Wastall and P. A. Wright, Xenobiotica, 1999, 29, 297.
131. J. Y. Zhang, J. J. Yuan, Y. F. Wang, R. H. Bible Jr. and A. P. Breau,
Drug Metab. Dispos., 2003, 31, 491.
132. P. C. Wilkins, H. K. Jacobs, M. D. Johnson and A. S. Gopalan, Inorg.
Chem., 2004, 43, 7877.
133. F. T. Bonner and Y. Ko, Inorg. Chem., 1992, 31, 2514.
134. R. Zamora, A. Grzesiok, H. Weber and M. Feelisch, Biochem. J., 1995,
312, 333.
135. A. Grzeslok, H. Weber, R. P. Zamora and M. Feelisch, in Biology of
Nitric Oxide, ed. M. Moncada, M. Feelisch, R. Busse and E. A. Higgs,
Portland Press, London, 1994, pp. 238–241.
136. P. Erdélyi, T. Fodor, A. K. Varga, M. Czugler, A. Gere and J. Fischer,
Bioorg. Med. Chem., 2008, 16, 5322.
137. S. K. Paulson, J. Y. Zhang, A. P. Breau, J. D. Hribar, N. W. K. Liu, S. M.
Jessen, Y. M. Lawal, J. N. Cogburn, C. J. Gresk, C. S. Markos, T. J.
Maziasz, G. L. Schoenhard and E. G. Burton, Drug Metab. Dispos., 2000,
28, 514.
138. J. W. Bridges, S. R. Walker and R. T. Williams, Biochem. J., 1969, 111,
173.
139. P. M. Vyas, S. Roychowdhury, F. D. Khan, T. E. Prisinzano, J. Lamba,
E. G. Schuetz, J. Blaisdell, J. A. Goldstein, K. L. Munson, R. N. Hines
and C. K. Svensson, J. Pharmacol. Exp. Ther., 2006, 319, 488.
140. J. M. Hutzler, D. Kolwankar, M. A. Hummel and T. S. Tracy, Drug
Metab. Dispos., 2002, 30, 1194.
141. J. H. Shon, Y. R. Yoon, M. J. Kim, K. A. Kim, Y. C. Lim, K. H. Liu,
D. H. Shin, C. H. Lee, I. J. Cha and J. G. Shin, Br. J. Clin. Pharmacol.,
2005, 59, 552.
142. D. J. Elliott Suharjono, B. C. Lewis, E. M. Gillam, D. J. Birkett, A. S.
Gross and J. O. Miners, Br. J. Clin. Pharmacol., 2007, 64, 450.
143. C. Tang, M. Shou, Q. Mei, T. H. Rushmore and A. D. Rodrigues,
J. Pharmacol. Exp. Ther., 2000, 293, 453.
144. K. D. Rainsford, in Milestones in Drug Therapy–COX-2 Inhibitors, ed.
M. Pairet and J. van Ryn, Birkhäuser Verlag, Basel, 2004, pp. 67–131.
145. M. L. Ecker, L. E. Visser, P. H. Trienekens, A. Hofman, R. H. van Schaik
and B. H. Stricker, Clin. Pharmacol. Ther., 2008, 83, 288.
146. J. Kirchheiner and J. Brockmoller, Clin. Pharmacol. Ther., 2005, 77,
1–16.
147. J. Kirchheiner, E. Störmer, C. Meisel, N. Steinbach, I. Roots and J.
Bröckmoller, Pharmacogenetics, 2003, 13, 473.
148. V. Kumar, J. L. Wahlstrom, D. A. Rock, C. J. Warren, L. A. Gorman
and T. S. Tracy, Drug Metab. Dispos., 2006, 34, 1966.
149. K. Komatsu, K. Ito, Y. Nakajima, S. Kanamitsu, S. Imaoka, Y. Funae,
C. E. Green, C. A. Tyson, N. Shimada and Y. Sugiyama, Drug Metab.
Dispos., 2000, 28, 475.
Sulfonamide as an Essential Functional Group in Drug Design 271
150. H. Malhi, B. Atac, A. K. Daly and S. Gupta, Postgrad. Med. J., 2004, 80,
107.
151. P. A. Williams, J. Cosme, A. Ward, H. C. Angove, D. Matak Vinkovic
and H. Jhoti, Nature, 2003, 424, 464.
152. Y. Yao, W. W. Han, Y. H. Zhou, Z. S. Li, Q. Li, X. Y. Chen and D. F.
Zhong, Eur. J. Med. Chem., 2009, 44, 854.
153. M. M. Ahlström, M. Ridderström, I. Zamora and K. Luthman, J. Med.
Chem., 2007, 50, 4444.
154. S. Y. Ahn and V. Bhatnagar, Curr. Opin. Nephrol. Hypertens., 2008, 17,
499.
155. E. babu, M. Takeda, S. Narikawa, Y. Kobayashi, A. Enomoto, A. Tojo,
S. H. Cha, T. Sekine, D. Sakthisekaran and H. Endou, Biochem. Biophys.
Acta, 2002, 1590.
156. H. E. Ives, in Basic and Clinical Pharmacology, ed. B. G. Katzung,
McGraw-Hill, New York, 2001, pp. 245–264.
157. J. B. Hook and H. E. Williamson, J. Pharmacol. Exp. Ther., 1965, 149,
404.
158. J. H. Gustafson and L. Z. Benet, J. Pharmacokinet. Biopharm., 1981, 9,
461.
159. P. Chennavasin, R. Seiwell, D. C. Brater and W. M. Liang, Kidney Int.,
1979, 16, 187.
160. K. R. Sweeney, D. J. Chapron, J. L. Brandt, I. H. Gomolin, P. U. Feig
and P. A. Kramer, Clin. Pharmacol. Ther., 1986, 40, 518.
161. H. Hasannejad, M. Takeda, K. Taki, H. J. Shin, E. Babu, P. Jutabha, S.
Khamdang, M. Aleboyeh, M. L. Onozato, A. Tojo, A. Enomoto, N.
Anzai, S. Narikawa, X.-L. Huang, T. Niwa and H. Endou, J. Pharmacol.
Exp. Ther., 2004, 308, 1021.
162. J. A. Taylor, Clin. Pharmacokinet., 1972, 13, 710.
163. P. Marchetti, R. Giannarelli, A. D. Carlo and R. Navalesi, Clin. Phar-
macokinet., 1991, 21, 308.
164. L. Balant, Clin. Pharmacokinet., 1981, 6, 215.
165. J. Judis, J. Pharm. Sci., 1973, 62, 1906.
166. M. J. Crooks and K. F. Brown, J. Pharm. Pharmacol., 1974, 26, 304.
167. Y. Uwai, H. Saito, Y. Hashimoto and K.-I. Inui, Eur. J. Pharmacol.,
2000, 398, 193.
168. T. B. Vree, Y. A. Hekster, J. E. Damsma, M. Tijhuis and W. T. Friesen,
Eur. J. Clin. Pharmacol., 1981, 20, 283.
169. T. B. Vree, Y. A. Hekster, J. E. Damsma, R. van Dalen, J. C. Haf-
kenscheid and W. T. Friesen, Ther. Drug Monit., 1981, 3, 129.
170. T. B. Vree, Y. A. Hekster, M. W. Tijhuis, M. Baakman, M. J. Oosterbaan
and E. F. Termond, Pharm. Weekbl. Sci., 1984, 6, 150.
171. A. Despoppulos and P. X. Callahan, Am. J. Physiol., 1962, 203, 19.
172. L. Z. Cooper, M. A. Madoff and L. Weinstein, Antibiot. Chemother.,
1962, 12, 618.
173. S. A. Tilles, South. Med. J., 2001, 94, 817.
272 Chapter 5
174. A. E. Cribb, B. L. Lee, L. A. Trepanier and S. P. Spielberg, Adverse Drug
React. Toxicol. Rev., 1996, 15, 9.
175. G. Choquet-Kastylevsky, T. Vial and J. Descotes, Curr. Allergy Asthma
Rep., 2002, 2, 16.
176. A. S. Kalgutkar, I. Gardner, R. S. Obach, C. L. Shaffer, E. Callegari,
K. R. Henne, A. E. Mutlib, D. K. Dalvie, J. S. Lee, Y. Nakai,
J. P. O’Donnell, J. Boer and S. P. Harriman, Curr. Drug Metab., 2005, 6,
161.
177. A. E. Cribb, S. P. Spielberg and G. P. Griffin, Drug Metab. Dispos., 1995,
23, 406.
178. A. Carr, B. Tindall, R. Penny and D. A. Cooper, Clin. Exp. Immunol.,
1993, 94, 21.
179. S. N. Lavergne, J. R. Kurian, S. U. Bajad, J. E. Maki, A. R. Yoder, M. V.
Guzinski, F. M. Graziano and L. A. Trepanier, Toxicology, 2006, 222, 25.
180. J. P. Sanderson, D. J. Naisbitt, J. Farrell, C. A. Ashby, M. J. Tucker, M.
J. Rieder, M. Pirmohamed, S. E. Clarke and B. K. Park, J. Immunol.,
2007, 178, 5533.
181. D. J. Naisbitt, S. F. Gordon, M. Pirmohamed, C. Burkhart, A. E. Cribb,
W. J. Pichler and B. K. Park, Br. J. Pharmacol., 2001, 133, 295.
182. D. J. Naisbitt, J. Farrell, S. F. Gordon, J. L. Maggs, C. Burkhart,
W. J. Pichler, M. Pirmohamed and B. K. Park, Mol. Pharmacol., 2002, 62,
628.
183. A. E. Cribb, M. Miller, J. S. Leeder, J. Hill and S. P. Spielberg, Drug
Metab. Dispos., 1991, 19, 900.
184. D. J. Naisbitt, P. M. Neill, M. Pirmohamed and B. K. Park, Bioorg. Med.
Chem. Lett., 1996, 6, 1511.
185. H. E. Callan, R. E. Jenkins, J. L. Maggs, S. N. Lavergne, S. E. Clarke,
D. J. Naisbitt and B. K. Park, Chem. Res. Toxicol., 2009, 22, 937.
186. M. J. Rieder, J. Uetrecht, N. H. Shear and S. P. Spielberg, J. Pharmacol.
Exp. Ther., 1988, 244, 724.
187. A. E. Cribb, C. E. Nuss, D. W. Alberts, D. B. Lamphere, D. M.
Grant, S. J. Grossman and S. P. Spielberg, Chem. Res. Toxicol., 1996, 9,
500.
188. P. M. Vyas, S. Roychowdhury, F. D. Khan, T. E. Prisinzano, J. Lamba,
E. G. Schuetz, J. Blaisdell, J. A. Goldstein, K. L. Munson, R. N. Hines
and C. K. Svensson, J. Pharmacol. Exp. Ther., 2006, 319, 488.
189. S. Roychowdhury, P. M. Vyas, T. P. Reilly, A. A. Gaspari and C. K.
Svensson, J. Pharmacol. Exp. Ther., 2005, 314, 43.
190. P. Wolkenstein, V. Carriere, D. Charue, S. Bastuji-Garin, J. Revuz, J. C.
Roujeau, P. Beaune and M. Bagot, Pharmacogenetics, 1995, 5, 255.
191. T. A. Baillie, Chem. Res. Toxicol., 2008, 21, 129.
192. C. Ju and J. P. Uetrecht, Curr. Drug Metab., 2002, 3, 367.
193. D. A. Smith and R. M. Jones, Curr. Opin. Drug Discov. Develop., 2008,
11, 72.
194. R. Patterson, A. E. Bello and J. Lefkowith, Clin. Ther., 1999, 21, 2065.
Sulfonamide as an Essential Functional Group in Drug Design 273
195. L. E. Shapiro, S. R. Knowles, E. Weber, M. G. Neuman and N. H. Shear,
Drug Saf., 2003, 26, 187.
196. D. C. Dahlin, G. T. Miwa, A. Y. Lu and S. D. Nelson, Proc. Natl. Acad.
Sci. U.S.A., 1984, 81, 1327.
197. H. Tan, W. M. C. Ong, S. H. Lai and W. C. Chow, Singapore Med. J.,
2007, 48, 582.
198. M. A. Macia, A. Carvajal, J. G. del Pozo, E. Vera and A. del Pino, Clin.
Pharmacol. Ther., 2002, 72, 596.
199. F. Li, M. D. Chordia, T. Huang and T. L. Macdonald, Chem. Res.
Toxicol., 2009, 22, 72.
200. D. P. Walker, F. C. Bi, A. S. Kalgutkar, J. N. Bauman, S. X. Zhao, J. R.
Soglia, G. E. Aspnes, D. W. Kung, J. Klug-McLeod, M. P. Zawistoski,
M. A. McGlynn, R. Oliver, M. Dunn, J. C. Li, D. T. Richter, B. A.
Cooper, J. C. Kath, C. A. Hulford, C. L. Autry, M. J. Luzzio, E. J. Ung,
W. G. Roberts, P. C. Bonnette, L. Buckbinder, A. Mistry, M. C. Griffor,
S. Han and A. Guzman-Perez, Bioorg. Med. Chem. Lett., 2008, 18,
6071.
201. H. Sun, R. Sharma, J. Bauman, D. P. Walker, M. P. Zawistoski, G. E.
Aspnes and A. S. Kalgutkar, Bioorg. Med. Chem. Lett., 2009, 19, 3177.
202. J. W. Clapp, J. Biol. Chem., 1956, 223, 207.
203. D. F. Colucci and D. A. Buyske, Biochem. Pharmacol., 1965, 14, 457.
204. O. W. Woltersdorf Jr., H. Schwam, J. B. Bicking, S. L. Brown, S. J.
deSolms, D. R. Fishman, S. L. Graham, P. D. Gautheron, J. M. Hoff-
man, R. D. Larson, W. S. Lee, S. R. Michelson, C. M. Robb, N. N. Share,
K. L. Shepard, A. M. Smith, R. L. Smith, J. M. Sondey, K. M. Stroh-
maier, M. F. Sugrue and M. P. Viader, J. Med. Chem., 1989, 32, 2486.
205. S. L. Graham, K. L. Shepard, P. S. Anderson, J. J. Baldwin, D. B. Best,
M. E. Christy, M. B. Freedman, P. Gautheron, C. N. Habecker, J. M.
Hoffman, P. A. Lyle, S. R. Michelson, G. S. Ponticello, C. M. Robb,
H. Schwam, A. M. Smith, R. L. Smith, J. M. Sondey, K. M. Strohmaier,
M. F. Sugrue and S. L. Varga, J. Med. Chem., 1989, 32, 2548.
206. G. D. Hartman, W. Halczenko, R. L. Smith, M. F. Sugrue, P. J. Mal-
lorga, S. R. Michelson, W. C. Randall, H. Schwam and J. M. Sondey,
J. Med. Chem., 1992, 35, 3822.
207. S. L. Graham, J. M. Hoffman, P. Gautheron, S. R. Michelson, T. H.
Scholz, H. Schwam, K. L. Shepard, A. M. Smith, R. L. Smith, J. M.
Sondey and M. F. Sugrue, J. Med. Chem., 1990, 33, 649.
208. K. A. Koeplinger, Z. Zhao, T. Peterson, J. W. Leone, F. S. Schwende, R. L.
Heinrikson and A. G. Tomasselli, Drug Metab. Dispos., 1999, 27, 986.
209. Z. Zhao, K. A. Koeplinger, T. Peterson, R. A. Conradi, P. S. Burton,
A. Suarato, R. L. Heinrikson and A. G. Tomasselli, Drug Metab. Dispos.,
1999, 27, 992.
210. B. Forouzesh, C. H. Takimoto, A. Goetz, S. Diab, L. A. Hammond,
L. Smetzer, G. Schwartz, R. Gazak, J. T. Callaghan, D. D. Von Hoff and
E. K. Rowinsky, Clin. Cancer Res., 2003, 9, 5540.
274 Chapter 5
211. C. B. Pratt, L. C. Bowman, N. Marina, A. Pappo, L. Avery, X. Luo and
W. H. Meyer, Invest. New Drugs, 1995, 13, 63.
212. A. Saili and M. S. Sarna, Indian Pediatr., 1991, 28, 936.
213. C. M. Jochheim, M. R. Davis, K. M. Baillie, W. J. Ehlhardt and T. A.
Baillie, Chem. Res. Toxicol., 2002, 15, 240.
214. X. Guan, M. R. Davis, C. Tang, C. M. Jochheim, L. Jin and T. A. Baillie,
Chem. Res. Toxicol., 1999, 12, 1138.
215. B. F. Hales and H. Brown, Teratology, 1991, 44, 251.
CHAPTER 6
6.1 Introduction
Aromatic rings are commonly found in most drugs. Historically, these rings have
been a classic structural feature of drugs ever since the discovery of aspirin in the
late 1800s. Since then several drug molecules that span all therapeutic areas have
been designed and contain a combination of these rings along with other side
chains and functional groups. The best selling drugs shown in Figure 6.1, as well
as other marketed drugs, contain at least one aromatic ring system.
Carbocyclic six-membered aromatic rings are widely present in both natural
products and endogenous ligands, and are therefore commonly incorporated
by medicinal chemists into new chemical entities. Aromatic rings are also
preferred replacements of linear and branched alkyl/cycloalkyl groups. Such a
replacement not only imparts greater rigidity to the molecule but can poten-
tially increase the non-covalent interactions between the macromolecule and
the small molecules. These interactions generally involve aromatic amino acid
side chains of the receptor and/or aromatic and heteroaromatic rings of the
ligand. Analysis of the ligand–protein complexes reveals that aromatic rings are
275
276 Chapter 6
O O
O O HN
O NH
N O
N
S O
Cl O
HO HO ClO NH2
O HO
Clopidrogel
OH Atorvastatin Amlodipine
F
O N
O N F
F
N
Aspirin HN
Cl NH F
N S O
H Cl
Olazapine Sertraline Fluoxetine
O O O O
O O S
H3C S N N N N
H H H H
H3C
The phenyl ring is a neutral moiety and therefore its presence in a molecule
does not affect its pKa. Further, in the absence of any hydrogen bonding
capabilities of the phenyl ring, the inclusion of this group in a molecule does not
have an effect on its polar surface area (PSA). However, addition of a phenyl
ring into a molecule addresses another important physicochemical property
such as lipophilicity and therefore influences the absorption, distribution and
clearance (excretion and metabolism) properties of a compound. Although less
lipophilic relative to cyclohexane ring (clogP of cyclohexane is 3.16 and clogP
of benzene is 2.18), the clogP of a phenyl ring is still sufficiently high to change
the lipophilicity of the whole molecule. For example, a hypothetical methyl-
cyclohexyl analog of tolbutamide (compound 6.1) has a clogP of 3.09 while
tolbutamide has a clogP of 2.36 (Figure 6.2).
Addition of a substituent or a functional group in an aromatic ring in place
of hydrogen atom can significantly perturb the electronic, steric, hydrophobic,
hydrophilic and hydrogen-bonding parameters of the ring depending on the
characteristics of that group. These changes can affect the interactions of the
compound with the receptors and therefore alter its biological activity. For
example, the replacement of an aniline group in carbutamide with a tolyl group
in tolbutamide results in separation of the antibacterial effect from the anti-
diabetic effect (Figure 6.3).2 Similarly chlorothiazide is an antihypertensive
agent with a strong diuretic effect. Replacement of the sulphonamide moiety in
chlorothiazide with hydrogen atom results in an antihypertensive without the
diuretic activity (Figure 6.3).2
An isosteric change of a functional group or a substituent on a phenyl ring
has a major effect on the lipophilicity of the molecule and can therefore
influence the disposition (absorption, distribution or excretion) of compounds.
In addition to lipophilicity, the change in polarity brought about by pertur-
bation of electron distribution of the aromatic ring due to inclusion of func-
tional groups can also influence the metabolism of the compound in the body.
Hammett was the first to explore the effect of substituents on reactivity of the
phenyl ring.3,4 He demonstrated that both the inductive and resonance effect of
a substituent could affect the electron density of the ring. Based on these studies,
an electronic parameter, (s), also called as the Hammett constant, was assigned
to each substituent. The s value for each substituent is dependent on the elec-
tronic property of that substituent (electron withdrawing or electron donating)
and the position of the substituent on the ring (para or meta relative to other
group). The more electron withdrawing a substituent, the more positive is its
278 Chapter 6
R
H H
N N
S
OO O
R = NH2; Carbutamide
R = CH3; Tolbutamide
Cl N N
NH NH
H2NO2S S Cl S
O O O O
Chlorothiazide Diazoxide
s value (relative to H, which is set at 0); conversely, the more electron donating,
the more negative is its s value. Table 6.1 shows the Hammett constants (smeta
and spara) for some commonly used groups. A list of Hammett constants for
several groups is given in the review by Hansch and co-workers.5
The Hammett relationship does hold true for the substituents at the ortho
position where it is complicated due to steric interactions and polar effects. As
the substituent becomes more electron withdrawing, it potentially reduces the
electron density of the phenyl ring, making the ring less reactive. In contrast,
the electron donating substituents enhance the reactivity of the ring. Thus,
substitution of the phenyl ring with a cyano or the nitro group deactivates the
phenyl ring and makes it less susceptible to oxidative metabolism, while
replacement of hydrogen atom with substituents such as the methyl group can
make the phenyl ring more prone to P450 catalysed oxidation.
Like the substituent constants derived by Hammett, Hansch and co-workers
derived constants for the contribution of individual functional groups to the
partition coefficient of the phenyl ring (and eventually to the logP of the overall
molecule).6,7 The lipophilicity substitution constant (p), dictates the influence of
a particular substituent on the partition coefficient (clogP) of the compound. The
positive value indicates that the substituent makes the phenyl ring more lipo-
philic and therefore increases the logP of the molecule. Likewise, the negative
value indicates that the substituent makes the molecule more polar by decreasing
its lipophilicity. Table 6.1 also shows the p values for various substituents that
are commonly incorporated in molecules. A comprehensive listing of all p values
for various substituents is depicted in the review by Hansch and co-workers.5
The third important parameter which addresses the interaction of a drug
with the receptor or an enzyme is the steric effect. The most widely used
parameter for steric substituent effects is the steric parameter, Es. This was
defined by Taft in much the same way as the Hammett’s electronic effects, but
was standardised to the methyl group (Es methyl ¼ 0.0) as opposed to the
hydrogen by Hammett.8,9 While the Hammett equation accounts for how field,
Table 6.1 Aromatic substituent constants of some commonly used functional groups
R
Molar refrac- No. of hydrogen No. of hydrogen
R Formula p sp sm tion (MR) bonddonors bond acceptors
R
Molar refrac- No. of hydrogen No. of hydrogen
R Formula p sp sm tion (MR) bonddonors bond acceptors
F
Influence of Aromatic Rings on ADME Properties of Drugs 285
hydrogen with a fluoro group can potentially favour the membrane penetration
and therefore an increased absorption and improved blood–brain barrier
penetration of a compound.
Fluorine also forms a strong bond with carbon. The C–F bond energy is
higher than the C–H or C–O bond (116, 99 and 85 kcal1, respectively).18 This
in turn increases the thermal and oxidative stability of the ring and makes the
C–F bond less sensitive to metabolic degradation.18 The electronic change also
makes the overall ring less reactive to P450 mediated oxidation.
6.3.1 Absorption
A majority of drugs are administered orally. Two main factors influencing
intestinal absorption are the solubility of a compound in the gastrointestinal
fluid and its permeability through the gastrointestinal wall.19,20 Both these
parameters are dependent on several physicochemical properties of a molecule
such the molecular size, lipophilicity, polar surface area (PSA) and hydrogen
bond donating and accepting capacity of the polar atoms.21 Although mole-
cular size impacts the transport of molecules through biological membranes,
lipophilicity has been proven to be the most important parameter affecting both
the permeability and aqueous solubility of compounds.
The logP value of a molecule primarily governs its ability to partition into
biological membranes. Since addition of a phenyl ring to a molecule increases
its lipophilicity, it can have a positive influence on the membrane permeability
of the compound. Modulation of the log P/log D value of a compound by
adding a substituent on the ring will further enhance or reduce its absorption.
Substituents such as alkyl and trifluoromethyl groups or the halogen atoms (Cl
or F), which commonly increase the logP value of a molecule, can increase its
membrane permeability. Alternatively, inclusion of the polar groups such as
carboxy, carboxamido, cyano groups (which have a negative p value) can
negatively influence the permeation of the compound through the gastro-
intestinal membrane.
Fichert and co-workers have explored structure–permeability relationships,
particularly the effect on cell permeability of some commonly used functional
groups attached to the drug-like core structures in lead optimisation efforts
286 Chapter 6
Table 6.4 Effect of commonly used functional groups on cell permeability.22
R
N N
PCaco2106 cm/sec Log D
H 71.1 2.91
mCN 65.8 2.47
PCO2H 3.19 0.61
PSO2NH2 23.4 1.04
pSO3H 0.08 1.78
mNH2 57.4 1.92
mNHSO2CH3 23.4 2.08
mCONH2 25.1 1.70
mCO2H 1.56 0.72
mCH2NH2 27.6 0.21
(Table 6.4).22 For this set of compounds, those with log D values within the
range 40 to o3 showed high permeability, while compounds with log D values
o0 display variable permeability.
For simple drugs like b-adrenoceptor antagonists, the log D7.4 values are
remarkably predictive of the absorption potential. All these compounds have a
molecular weight that ranges from 250 to 310; however, their log D values differ
depending on the substituents on the phenyl ring. As seen in Table 6.5,
incorporation of the cyclohexane diol or the methyl carboxamido group in the
molecule results in a negative clogD7.4 value, which in turn affects the
absorption of these two drugs through the gastrointestinal tract. On the other
hand, introduction of the alkyl or phenyl group in betaxolol and propranolol
increases the lipophilicity of these compounds, which results in their 100%
absorption following oral administration23(Table 6.5)
Aqueous solubility dictates the amount of drug in solution and hence the
amount of drug available for absorption from the gastrointestinal tract.
Compounds with low solubility may suffer from dissolution-limited absorption.
Although an increase in lipophilicity enhances the permeability, a higher logP
value has a deleterious effect on the aqueous solubility of compounds. In
general, aqueous solubility is inversely proportional to lipophilicity, and
compounds with a high logD7.4 value (43) may show poor dissolution and
resultant poor bioavailability. For instance, promazine (Figure 6.4) has been
classified as a high permeability high solubility drug in the biopharmaceutics
classifications system while chlorpromazine belongs to high permeability-low
solubility class of compounds.24 The only difference in the two compounds is
the replacement of one of the hydrogen atoms in promazine with a chloro
group in chlorpromazine (Figure 6.4). This subtle change in chlorpromazine
results high permeability but low aqueous solubility (Figure 6.4). The effect on
solubility is probably due to the increase in the clogP of chlorpromazine relative
to promazine.
Influence of Aromatic Rings on ADME Properties of Drugs 287
S S
N Cl N
Promazine
Chlorpromazine
clogP = 4.69 N N
clogD7.4 = 2.73 clogP = 5.18
cSol = 0.4 mg/ml clogD7.4 = 3.24
cSol = 0.09 mg/ml
Table 6.5 Physicochemical properties of representative beta-blockers and its influence on absorption.
Name Structure MW cLogP clogD7.4 PSA % Absorbed23
O
OH
OH
N
H
OH
Chapter 6
Betaxolol 307 2.53 0.43 50.72 100
O
O NH
OH
N
H
Influence of Aromatic Rings on ADME Properties of Drugs
289
290 Chapter 6
6.3.2 Distribution
Distribution of drugs entails passage of a compound from circulation into the
tissues. Like absorption, there is a trend for increasing tissue permeability with
increasing lipophilicity.21 Hence the incorporation of the phenyl rings in a
molecule is likely to increase its affinity for the tissues. As in the case of
absorption, the substituents that increase the logP value of the compounds will
enhance distribution into the tissues more than the polar groups. For example,
propranolol and alprenolol with a clogP value of 1.2 and 1.0, respectively, show
good penetration into the cerebrospinal fluid (CSF) from the plasma as shown
by the CSF–plasma ratio (Table 6.7), while nadolol and atenolol are not
permeable through the blood–brain barrier.28 Clearly, incorporation of
cyclohexanediol in the ring system makes the compound quite polar for it to
cross the blood–brain barrier, whereas very hydrophilic nature of atenolol
hinders its permeability into the brain as is the case in the absorption of these
two drugs.
A similar relationship between lipophilicity and transfer across the blood–
brain barrier has been observed among the benzodiazepines midazolam,
alprazolam and triazolam.29 As seen in Table 6.8, the increase in lipophilicity
increases the brain to serum ratio.
Plasma protein binding is also an important parameter that influences tissue
distribution. Only free drug in plasma can cross the membrane and equilibrate
with free drug in the tissues. Incorporation of neutral lipophilic substituents on
the phenyl ring tends to increase the binding of the drug to plasma proteins and
therefore reduce the free fraction that is available to transfer to the tissues.
Comparison of the brain to unbound serum concentration ratio of lorazepam,
in which one of the hydrogen atom on the phenyl ring is replaced by a chloro
group (Table 6.9), versus oxazepam suggested that extraction of the latter into
the brain was slightly enhanced.29 Since the change in lipophilicity is minimal
between the two compounds, the increase in brain exposure is attributed to
more unbound fraction of oxazepam.
It has been demonstrated that other factors such as the hydrogen bonding
capabilities of a compound can also affect its penetration across the blood–
brain barrier.30,31 Hence, the established relationship between permeability
across the blood–brain barrier, protein binding and lipophilicity may not
necessarily hold in all cases of lipophilic drugs. For example, in spite of
similar clogD and protein binding values of flunitrazepam compared to
diazepam (Table 6.10), its brain to serum ratio is much lower than diaze-
pam. This disconnect is probably attributed to the increase in PSA of
flunitrazepam (78.94 Å2) due to presence of a nitro group in one of its aromatic
rings.
The effect of hydrogen bond donor and polar functionality in the phenyl
ring on brain exposures was also demonstrated in the following example
(Table 6.11). Compound 6.5 was synthesised as one of the analogs of mGluR2
potentiators but exhibited poor brain exposure following administration to
rats.32 Replacement of the phenolic group with an ethoxy group (compound 6.6)
Table 6.6 Comparison of physicochemical properties of sulpiride and remoxipride.25,26
Name Structure MW clogD7.4 PSA Å2 CSolubility mg/mL %F
NH
O
N S
O
NH2
H
N
Br N
O O
291
292 Chapter 6
Table 6.7 Brain penetration of b-blockers and its relationship to lipophilicity.28
Compound Structure cLogP CSF/Plasma
Propranolol 1.2 1
HO
O
N
H
Alprenolol 1.0 1
O N
H
OH
Nadolol OH 1.5 ND
H
HO N
O
OH
Atenolol O 1.1 ND
H2N
O NH
OH
gave better brain levels and higher brain-to-plasma ratio (1.2:1). Detectable
levels for 6.6 were attributed to the drop in the polar surface area by 16 Å2
compared with compound 6.5, in addition to the increase in the log D of the
molecule.
Another factor that affects brain penetration is the efflux of the compound
from blood–brain barrier by P-gp transporter.27 In contrast to absorption, the
P-gp transporter plays an important role in brain uptake as it limits cellular
uptake of drugs from the blood circulation into the brain. As mentioned above,
the structure–activity relationships in terms of the role of the substituted phenyl
ring systems are not fully understood. However, an increase in brain exposure
via modulation of PSA, which in turn affects the P-gp efflux properties, has
been demonstrated by Roberts and co-workers.33 This group focused on the
reduction of polar surface to influence the efflux properties of compounds.
2-Imidazole a1A partial agonist (6.7) was synthesised for its potential utility in
treating stress urinary incontinence (Table 6.12). However, 6.7 showed poor
brain penetration (rat free plasma/CSF 1:0.08), which was attributed to high
P-gp-mediated efflux when assessed in MDCK MDR-1 cells (AB/BA ratio
of 7:29). The medicinal chemistry strategy focused on reducing the PSA of new
Influence of Aromatic Rings on ADME Properties of Drugs 293
Table 6.8 Comparison of brain to serum (B/S) ratio of midazolam, alprazo-
lam and triazolam and its physicochemical properties.29
Name Structure cLogD7.4 B/S ratioa
F
Cl N
N
N
CI N
N
N
N
Cl
Cl N
N
N
N
a
B/S ratio ¼ brain to unbound serum ratio.
analogs in order to reduce the potential for P-gp recognition and therefore
improve blood–brain barrier penetration.
As shown in Table 6.12, the replacement of the sulfonamide with dialky-
lethers such as the methoxymethyl ether 6.8 resulted in retention of the a1A
partial agonist potency with no evidence of P-gp efflux. Since the log D values
of all the analogs made were similar, the modulation of the P-gp mediated efflux
of 6.7 by addition of the ether functionalities was attributed to the decrease in
PSA of the molecule. The brain penetration was further enhanced by increasing
the lipophilicity of the compounds via incorporation of the chloro and fluoro
groups (Table 6.12).
Although incorporation of polar groups in phenyl rings has a negative
influence on the permeability across the blood–brain barrier, the strategy of
substituting the aromatic hydrogen with polar moieties can be used in pre-
venting serious central nervous system (CNS) side effects that may be caused by
some lipophilic compounds. This has been demonstrated in the following two
294 Chapter 6
Table 6.9 Comparison of the brain to unbound serum concentration and
plasma protein binding of lorazepam and oxazepam.29
Compound Structure clogD7.4 Unbound fraction B/S ratioa
N NH
Cl
Cl
N NH
Cl
a
B/S ratio ¼ brain to unbound serum ratio.
N N
Cl
N N
F
NO2
a
Total B/S ¼ total brain to serum ratio,
b
B/unbound S ¼ brain to unbound serum concentration.
O
R O
HO2C
HO2C
a
HBD ¼ hydrogen bond donor,
b
B/P ¼ brain to plasma ratio,
c
BQL ¼ below quantitation limits.
6.3.3 Clearance
Metabolism and excretion are the two primary routes of elimination of a drug
from the body. Like absorption and distribution processes, the physicochemical
296 Chapter 6
Table 6.12 Modulation of P-gp efflux by reducing total polar surface area.33
No. Structure MDCK-MDR1 ratio PSA (Å2) logD7.4
6.7 N NH
7/29 83 1.5
Cl
NH
H3CO2S
6.8 N NH
40/36 38 1.8
6.9 N NH
39/37 38 2.0
6.10 N NH
40/36 38 2.6
CI
properties of a drug and its structure can primarily dictate its route of elim-
ination.21 Among all the physicochemical properties, lipophilicity is the key
factor determining the ultimate route by which the drug will be eliminated.21 It
governs the accessibility of the drug to many of the drug metabolising enzymes
as well as the ability of a molecule to cross the renal tubular membrane. The
latter governs the urinary excretion of a compound. Since substituted aromatic
rings play a major role in modulating a drug’s lipophilicity, incorporation of a
Influence of Aromatic Rings on ADME Properties of Drugs 297
H H
N N O
O S
N N N
N
O O
Compound 6.11
Sulmazole 6.12
cLogP = 2.59
logP = 1.17
PSA = 60.03
PSA = 87.08
6.3.3.1 Excretion
Molecules that have a relatively small size (approximate molecular weight of
200–300) and are polar are generally excreted unchanged in the urine. Since
incorporation of substituted phenyl rings adds a certain degree of lipophilicity
to the molecule, most molecules with one or more phenyl rings generally
undergo metabolism. However, replacement of a hydrogen atom in an aro-
matic ring with polar functionalities can decrease the logD7.4 and can result in
excretion of the drug renally. For example, comparison of structures of various
beta blockers and their physicochemical properties (Table 6.14) indicates that
analogs with polar functionalities which increase the PSA and decrease the
298
OH
OH
Chapter 6
Metoprolol O 267 0.47 50.72 10
O NH
OH
OH
Influence of Aromatic Rings on ADME Properties of Drugs
299
300 Chapter 6
O
S O
O HN S
O O
NH
N
Dofetilide
clogD7.4 = 0.52 (clogP = 1.38)
2
PSA = 121.57 Å
clogD value (e.g. atenolol, sotalol and practolol) are more prone to undergo
renal excretion than those with lipophilic groups such as betaxolol and
metoprolol.36
The effect of polar functionality has also been demonstrated in the phar-
macokinetics of the neuroleptic sulpiride. Assessment of unchanged sulpiride in
the urine indicates that metabolism is a minor route of clearance for this
compound and about 70% of the dose is excreted unchanged in the urine.25 In
contrast, the analog of sulpiride, remoxipride is extensively metabolised and
only 25% of the dose is excreted in the urine.26 This can be correlated to the
physicochemical properties of these two drugs (Table 6.6). The highly polar
nature of sulpiride (clogD7.4 ¼ 1.07) is linked to the high renal clearance of the
drug relative to remoxipride.
Even though it is commonly believed that hydrophilic compounds are
excreted unchanged renally and that the log D value of renally excreted drugs is
generally o0, there are some discrepancies. For example, the class III antiar-
rhythmic agent, dofetilide, is an amine with a log D7.4 value of 0.52 (Figure 6.6).
The moderate lipophilicity of this compound suggests that it has the potential
to undergo oxidative metabolism by P450. However, assessment of the renal
clearance of dofetilide in humans indicates that 80% of the compound is
excreted unchanged in the urine.37 This is probably due to the presence of the
two phenyl sulfonamide groups in the molecule, which increase the hydrogen
bonding potential and therefore the PSA of the molecule (121.57 Å2). Taken
together, dofetilide more closely resembles atenolol or sotalol as reflected by the
renal clearance of the compound.
6.3.3.3 Metabolism
As mentioned earlier, metabolism is a primary route of clearance for most
compounds containing one or more aromatic rings. Lipophilic compounds are
susceptible to oxidation by liver enzymes (namely cytochrome P450) as well as
by other conjugative enzymes such as uridinylglucuronyltransferase (UGT).21
Incorporation of phenyl rings into a molecule affects its metabolism in two
ways. Firstly, the presence of unsubstituted or substituted aromatic rings in the
molecule increases its lipophilicity; this facilitates an easy transfer of the
molecule across the membrane and brings it into close proximity to the xeno-
biotic metabolising enzymes. Secondly, the aromatic rings themselves become a
site of metabolism. The latter, however, is dependent upon the reactivity of the
ring. As discussed in Section 6.2, an aromatic ring is a delocalised system with
6p electrons. Although aromatic molecules display enhanced stability due to
the delocalisation of these electrons, these rings can undergo electrophilic
aromatic substitution reactions that involve replacement of hydrogen atom in
the ring by an electrophilic substituent. In the case of CYP450 catalysed oxi-
dations, the hypervalent iron–oxene complex is widely accepted as ‘the oxidant
species’ that is ultimately responsible for the oxidation of compounds.40 This
high-energy oxo-species can add into the aromatic ring to produce a tetrahedral
intermediate that subsequently rearranges via an epoxide (an arene oxide) or a
ketone intermediate and ultimately results in an arenol (Figure 6.7).
Like electrophilic aromatic substitution reactions, the substituents on the
ring influence the oxidation of the phenyl ring.41 Hence, aromatic rings that are
activated (contain electron rich substituents) are more susceptible to aromatic
hydroxylation.42 For example, aniline containing analogs are hydroxylated
extensively. The final location of the oxygen is further explained by simple
organic principles. Ring opening of the arene oxide of compounds containing
electron rich rings is expected to occur in the direction that results in a more
stable carbocation. Thus, aniline derivatives are hydroxylated in the ortho and
para positions of the amino group as observed in atorvastatin (Figure 6.8).43,44
On the other hand, deactivated rings (rings with electron withdrawing
groups) are hydroxylated at a slower rate or not at all. For example, the uri-
cosuric agent, probenecid, undergoes no detectable aromatic hydroxylation
(Figure 6.9).45,46 The presence of two deactivating groups on the ring—the
carboxy group and the sulfonamide moiety—results in a significant decrease in
the electron density of the phenyl ring. The same principle applies to drugs that
contain two phenyl rings. In case of chlorpromazine or fluphenazine, the
302
Table 6.15 Correlation of biliary clearance of D2 prostaglandin receptor antagonists with their physicochemical properties.39
No. Concentration in rat bile (mM) clogD7.4 cLogP PSA MW pKa
H
O
H
a
R R Epoxide
Intermediate R
b H
O b
a Fe 3+ OH
O
b Arenol
Fe 4+
R
O
HH
Ketone
Intermediate
OH
OH
N O
H
F N OH
O
para-Hydroxyatorvastatin
HO
OH
OH
N O F
H
N OH
O
Atorvastatin OH
OH
N O
H
N OH
O
OH
ortho-Hydroxyatorvastatin
S S
HO2C
N N CI
S N Cl
O O
Probenecid
N
N
Chlorpromazine N Fluphenazine
HO
N N N N
H H
N N
O O F
CP-99994 L-733060 HN NH L-741671 HN NH Aprepitant
HN HN HN
Cl CI Cl Cl Cl
OH Cl
Diclofenac Hydroxydiclofenac Fenclofenac
OCH3 OH
OH
F
N N
O O
SCH 48461 Ezetimibe
OCH3 F
F
N N
N N
Cinnarizine Flunarizine
F
NH
Y
X N
H
X Y F (%) cLogP pKa
H H Poor 10.4
H F 18 8.5
F F 80
F
Cl
CF3
N N CF3
N N
H3CO2S
H2NO2S
6.21 6.22
t1/2 = 221 hr t1/2 = 117 hr
H3C H2 O
C C
HO HO
CF3
N N CF3 CF3
N N N N
H2NO2S
Celecoxib H2NO2S H2NO2S
t1/2 = 3.5 hr
O
N
O
Lidocaine
O O CI O
N N
O O
H2N H2N
Propoxycaine Chloroprocaine
6.4 Toxicity
Aromatic rings can play a major role in influencing the toxicity of drugs. As
discussed above, hydrogen atoms in the aromatic rings of compounds are
generally replaced with lipophilic functional groups to improve their ADME
properties and potency. These changes in the physicochemical properties of
compounds also enhance their interactions with targets and proteins, and result
in side effects. For example, amiodarone (Figure 6.15) is a lipophilic compound
belonging to this category. This antiarrhythmic agent is quite efficacious but
causes a number of side effects including pulmonary toxicity and phospholi-
pidosis.56,57 The toxicity is attributed to the presence of a diiodophenyl moiety
in the molecule which increases the lipophilicity of the compound (clogP
¼ 7.81).58 This enhances its permeation and subsequent accumulation into
several tissues and results in a very long half-life in humans (B9 to 77 days).59
In another example, the alkyl and cyclohexyl groups present in proxicromil, an
anti-allergy agent, allow this compound absorb readily through the gastro-
intestinal tract (Figure 6.15).60 However, this highly lipophilic, detergent-like
drug tends to accumulate in the biliary canaliculus, which may be a reason for
its hepatotoxicity.61
O OH O
I
O
N OH
O O
I O
Amiodarone Proxicromil
clogP = 7.81 clogP = 4.89
F
HO HO
N N O
Cl Cl
CF3
F F
Penfluridol Haloperidol
clogP = 6.95 clogP = 3.76
O O O
HN HN HN
NH NH NH
O O O
OH
Phenytoin O Hydroxyphenytoin
O
O O
H3C N
H3C N H3C N
O
O O
O OH
Phensuximide Hydroxyphensuximide
H H
H O N O O N O
O N O
HN HN
HN
O O
O
O OH
Phenobarbital Hydroxyphenobarbital
Covalent Binding
Hepatic Necrosis
which may trigger mechanisms that finally lead to cell necrosis, immune
responses or blood toxicities.
Metabolic activation of acetaminophen represents a classic example in this
category.72–74 Although safe at therapeutic doses, this widely used analgesic
causes fulminant hepatic necrosis when taken in overdose. The cause of
hepatotoxicity is ascribed to P450 catalysed 2-electron oxidation of acet-
aminophen to a quinone imine intermediate (NAPQI) and subsequent covalent
binding of this electrophilic intermediate with various proteins in the liver
(Figure 6.18). Some additional examples include bioactivation of amodiaquine
(Figure 6.18) and troglitazone (Figure 6.19).75–78 The hepatotoxicity and
agranulocytosis observed following administration of antimalarial amodia-
quine is triggered by formation of reactive quinone imine intermediate (similar
to that observed for acetaminophen) which is produced in the liver as well as
other sites in the body (Figure 6.18).75
Troglitazone was developed for the treatment of type 2 diabetes but was
withdrawn from the clinic due to rare cases of liver failure and deaths. The
Influence of Aromatic Rings on ADME Properties of Drugs 311
OH O OH
-
2 e Oxidation
R1 R1 Nu R1
Nu
R R R
Quinone Methide
OH O OH
-
R1 2 e Oxidation Nu
R1 R1
Nu
NH N NH
R R R
Quinone Imine
Figure 6.17 Two electron oxidation of phenolic compounds to quinone methides and
quinone imines and their reaction with nucleophiles.
O O
HN HN
UGT OR
O ST CO2H
O
O OH OSO3H
HN OH
HO
O O
OH N HN
Acetaminophen
CYP450 Protein-SH
[O] Protein
S
O OH
Quinone imine
OH
O
N
HN N
N
Cl N
Cl N
Amodiaquine Quinone imine Formation
O O
O HO
SG =
GS glutathione
Quinone Methide
S OH
O O O O
O NH
O O
O O O
HO OH
Troglitazone
O OH O
S S S
O HO O
NH N C O
O O O N SG
H
SG
S
HN2
O
HO O
OH O
O
DNA Covalent
Catechols o-Quinones Binding
HO
O
Estrone O
HO
O
HO
O
OH OH O O
HO HO
HO HO
Estradiol Equilin Equilenin
17α-Ethynylestradiol
CO2H CO2H
Cl Cl
H
Acyl Glucuronide N N
Covalent
HO Binding
CO2H Cl O Cl
Cl
H
N 2C9
Hydroxydiclofenac
Cl
CO2H
Diclofenac 3A4 Cl CO2H
H Cl
N
N Covalent
Binding
Cl OH
Cl O
Figure 6.21 P450 catalyzed bioactivation of diclofenac and tacrine to quinone imines
via their hydroxylated metabolites.
Nefazodone HO
Cl
Cl
N N
GS O GS
N N
OH
HO O O
HO
Cl Cl Cl
Cl
Possibly leads
to inactivation of
CYP3A4
Ar Ar Ar Ar
N
O
Ar Ar P450
HO OH O O
P450
Ar Ar Ar Ar
OH
HO
Tamoxifen P450
OH O
Direct 2e- oxidation
N O N
O O O
O O Inactivation of
CYP3A4
OH O
S O S
HO
O
OH
HN
F F
F
HO
OH
2-Fluoroestradiol 2, 6- Difluoroacetaminophen
F
SG
OH
OH
N
HN N
HN
SG = glutathione
Cl N
Cl N
6-Fluoroamodiaquine
N
HN
Cl N
Fluoroamodiaquine
O O O O
N N
N N
H H
OH
6.23
O O
N
3A4 Inactivation N
H
O
Quinone Methide
O O
N
N
H
6.24 F
(Figure 6.24), which had similar antimalarial potency to the parent drug but
was not bioactivated to toxic metabolites in vivo.18
The following example demonstrates modification of the lead compound to
prevent mechanism based inhibition by introducing fluorine into the molecule.
The lead compound (6.23) was designed to be a KCNQ2 potassium channel
opener and had excellent oral bioavailability and pharmacological activity
(Figure 6.25).96 However, 6.23 was an inactivator of CYP3A4. P450 inactiva-
tion was consistent with a bioactivation pathway involving the initial aromatic
hydroxylation ortho to the morpholine ring (or para to the benzylamine
methine). Further two electron oxidation of this initial metabolite by P450 can
result in the formation of the reactive quinone methide intermediates. The
potential formation of either reactive intermediate was avoided by the intro-
duction of the fluorine atom on the position ortho to the morpholine ring.
Compound 6.24 is not only devoid of P450 inactivation liability, but also
retains the pharmacological and pharmacokinetic properties of the prototype
compound 6.23.
Although fluorine atoms have been commonly used in blocking metabolism
and bioactivation (as described earlier), it is important to note that fluorination
of the aromatic ring does not necessarily prevent the compound from under-
going bioactivation or from inactivating the enzyme.18 More recently, dasati-
nib, a tyrosine kinase inhibitor, has been shown to inactivate CYP3A4 and this
inactivation proceeds through a reactive intermediate.97 The major mechanism
of inactivation proceeds through hydroxylation at the 4-position of the 2-
chloro-6-methylphenyl ring in the molecule followed by further oxidation
forming a reactive quinone-imine (Figure 6.26). Interestingly, blocking the para
318 Chapter 6
N N
N N
Cl N N [O] Cl N N
H N H N
N N N N
S H OH S H OH
O HO O
Dasatinib
[O]
[O]
Cl N
Cl N
N N
N N S H
S H
O O
O
Quinoneimine
Iminomethide
Inactivation of CYP3A4
N
N N N
N OH
H NH
N S
Ar
O
6.26 F No inactivation
Cl
Cl
position with fluorine did not prevent inactivation of CYP3A4 (Table 6.17).
The rates of inactivation were virtually identical and the KI for the fluorinated
analog was lower, implying that it is more efficiently bioactivated possibly by
virtue of increase in the clogP value due to the incorporation of fluorine.
Influence of Aromatic Rings on ADME Properties of Drugs 319
O O O O O O
N N
N N HO N
H N
H H
O OH OH
Br Br Br NCQ344
Remoxipride
O O
O N
N
Covalent Adducts H
with proteins Peroxidase O
Br
H H
F3C N F3C N
O O
F F
O CH3 OH
M1
Cl Cl
MaxiPost
H
H F3C N
F3C N O
O
NHLys-Proteins O
OH
Cl
Cl
Covalent binding ortho-Quinone Methide
to Lysine residue
OH OX
NH 2 NH NH NH 2
R R Nu
R R Nu
Hydroxylamine
X = Acetyl
X = SO3H
O
N Reaction with
Nucleophiles on
R the Macromolecule
Nitroso
Intermediate
NH2
O O CF3
N
N H
N N
H F3C O N
H2N H
O
Procainamide Flecainide
Nomifensine
X
X = SO2, Dapsone
H2N NH2 X = S, O, CH2, CO
Procainamide (Figure 6.30) is another drug where removal of the aniline moiety
translates into a markedly improved safety profile. Bone marrow aplasis and
lupus syndrome observed upon administration of procainamide is not observed
upon administration of flecainide, which is devoid of aniline moiety.107
Dapsone (Figure 6.30), an antibacterial, is also shown to induce agranulo-
cytosis, aplastic anaemia and cutaneous adverse drug reactions; all these
toxicities are associated with N-oxidation of aniline nitrogen.108–110 The
influence of electronics on bioactivation potential is evident in structure–toxi-
city assessment on dapsone analogues. Replacement of the sulfone group in
dapsone with sulfur, oxygen, methylene or carbonyl substituents significantly
reduce methemoglobinaemia in human erythrocytes.111 Although the reason
for the influence of these substituents is unclear, a good correlation between
haemotoxicity and the Hammett constant (sp) has been demonstrated in this
analysis indicating that the substituent in the 4-position can potentially affect
the rate of oxidation of the aniline nitrogen.
Compounds containing nitroaromatic rings, which can be metabolically
reduced to anilines, can also elicit toxicity. Tolcapone, a catechol-O-methyl-
transferase inhibitor (Figure 6.31), has been associated with a number of
problems including abnormalities in liver function tests and three cases of fatal
hepatotoxicity.112 These problems have led to withdrawal of the drug from the
market in some countries and the introduction of a black box warning and
intensive monitoring requirements in the United States. A significant portion of
tolcapone biotransformation proceeds via reduction of the nitrobenzene group
to the aniline derivative, which is transformed to the corresponding anilide by
N-acetyltransferase. Both the aniline and the anilide metabolites of tolcapone
undergo facile two electron oxidation to the corresponding quinone-imine
metabolites that are trapped with GSH (Figure 6.31).113 Interestingly, the
structural analog of tolcapone, entacapone, does not undergo similar bioacti-
vation in spite of the presence of the same structural motif. This is attributed to
minimal (or lack of) reduction of the nitro group to the aromatic amine in
322 Chapter 6
O SG O
HO HO
HO CH 3 HO CH 3
NHCOCH3 NHCOCH 3
NAT
O O O
HO HO HO
Reduction
HO CH 3 HO CH3 O CH3
NO 2 NH2 NH Quinone imine
Tolcapone
O O
SG O
HO O
N HO
CN O CH3
HO HO CH3
NO2 NH 2
NH2
Entacapone ortho-Quinone
SG = Glutathione
Figure 6.31 Metabolic activation of tolcapone via reduction of the nitro group.
References
1. R. K. Castellano, F. Diederich and E. A. Meyer, Angew. Chem., Int. Ed.,
2003, 42, 1210.
2. R. B. Silverman, in The Organic Chemistry of Drug Design and Drug
Action, Academic Press, San Diego, 2004, pp. 7–99.
3. L. P. Hammett, J. Am. Chem. Soc., 1937, 59, 96.
4. L. P. Hammett, Physical Organic Chemistry, McGraw-Hill, New York,
1940.
5. C. Hansch, A. Leo and R. W. Taft, Chem. Rev., 1991, 91, 165.
6. T. Fujita, J. Iwasa and C. Hansch, J. Am. Chem. Soc., 1964, 86, 5175.
7. C. Hansch and T. Fujita, J. Am. Chem. Soc., 1964, 86, 1616.
8. R. W. Taft, in Steric Effects in Organic Chemistry, ed. M. S. Neuman,
Wiley, New York, 1956, pp. 556–675.
9. S. H. Unger and C. Hansch, Prog. Phys. Org. Chem., 1976, 12, 91.
10. C. Hansch, A. Leo, S. H. Unger, K.-H. Kim, D. Xikaitani and E. J. Lien,
J. Med. Chem., 1973, 16, 1207.
11. W. Guba, W. Neidhart and M. Nettekoven, Bioorg. Med. Chem. Lett.,
2005, 15, 1599.
Influence of Aromatic Rings on ADME Properties of Drugs 323
12. S. Purser, P. R. Moore, S. Swallow and V. Gouverneur, Chem. Soc. Rev.,
2008, 37, 320.
13. H.-J. Böhm, D. Banner, S. Bendels, M. Kansy, B. Kuhn, K. Müller,
U. Obst-Sander and M. Stahl, ChemBioChem, 2004, 5, 637.
14. W. K. Hagmann, J. Med. Chem., 2008, 51, 4359.
15. K. Müller, C. Faeh and F. Diederich, Science, 2007, 317, 1881.
16. A. Bondi, J. Phys. Chem., 1964, 68, 441.
17. B. K. Park, N. R Kitteringham and P. M. O’Neill, Annu. Rev. Pharmacol.
Toxicol., 2001, 41, 443.
18. B. K. Park and N. R. Kitteringham, Drug Metab. Rev., 1994, 26, 605.
19. G. L. Amidon, H. Lennernas, V. P. Shah and J. R. Crison, Pharm. Res.,
1995, 12, 413.
20. O. H. Chan and B. H. Stewart, Drug Discov. Today, 1996, 1, 461.
21. D. A. Smith, B. C. Jones and D. K. Walker, Med. Res. Rev., 1996, 16, 243.
22. T. Fichert, M. Yazdanianb and J. R. Proudfoota, Bioorg. Med. Chem.
Lett., 2003, 13, 719.
23. R. Griffith, in Foye’s Principles of Medicinal Chemistry, ed. D.A. Williams
and T. L. Lemke, Lippincott Williams and Wilkins, Baltimore, 2002,
pp. 292–312.
24. J. M. Custodio, C.-Y. Wu and L. Z. Benet, Adv. Drug Deliv. Rev., 2008,
60, 717.
25. F.-A. Wiesel, G. Alfredsson, M. Ehrnebo and G. Sedvall, Eur. J. Clin.
Pharmacol., 1980, 17, 385.
26. G. Movin-Osswald and M. Hammarlund-Udenaes, Br. J. Clin. Pharma-
col., 1991, 32, 355.
27. J. H. Lin and M. Yamazaki, Drug Metab. Rev., 2003, 35, 417.
28. F. Atkinson, S. Cole, C. Green and H. van de Waterbeemd, Curr. Med.
Chem. Cent. Nerv. Syst. Agents, 2002, 2, 229.
29. R. M. Arendt, D. J. Greenblatt, D. C. Liebisch, M. D. Luu and S. M.
Paul, Psychopharmacology, 1987, 93, 72.
30. J. Kelder, P. D. J. Grootenhuis, D. M. Bayada, L. P. C. Delbressine and
J.-P. Ploemen, Pharm. Res., 1999, 16, 1514.
31. S. A. Hitchcock and L. D. Pennington, J. Med. Chem., 2006, 49, 7559.
32. R. V. Cube, J. -M. Vernier, J. H. Hutchinson, M. F. Gardner, J. K. James,
B. A. Rowe, H. Schaffhauser, L. Daggett and A. B. Pinkerton, Bioorg.
Med. Chem. Lett., 2005, 15, 2389.
33. L. R. Roberts, J. Bryans, K. Conlon, G. McMurray, A. Stobie and G. A.
Whitlock, Bioorg. Med. Chem. Lett., 2008, 18, 6437.
34. E. Kutter and V. Austel, Arzneim.-Forsch., 1981, 31, 135.
35. S. M. Rubenstein, V. Baichwal, H. Beckmann, D. L. Clark, W. Frank-
moelle, D. Roche, E. Santha, S. Schwender, M. Thoolen, Q. Ye and J. C.
Jaen, J. Med. Chem., 2001, 44, 3599.
36. D. A. Smith, H. van de Waterbeemd and D. K. Walker, Pharmacokinetics
and Metabolism in Drug Design, Wiley-VCH, Weinheim, 2006, p. 83.
37. D. A. Smith, H.S. Rasmussent, D. A. Stopher and D. K. Walker,
Xenobiotica, 1992, 22, 709.
324 Chapter 6
38. H. Kusuhara, H. Suzuki and Y. Sugiyama, J. Pharm. Sci., 1998, 87, 1025.
39. C. F. Sturino, G. O’Neill, N. Lachance, M. Boyd, C. Berthelette,
M. Labelle, L. Li, B. Roy, J. Scheigetz, N. Tsou, Y. Aubin, K. P. Bate-
man, N. Chauret, S. H. Day, J. L. Le 0 vesque, C. Seto, J. H. Silva, L. A.
Trimble, M. C. Carriere, D. Denis, G. Greig, S. Kargman, S. Lamon-
tagne, M. C. Mathieu, N. Sawyer, D. Slipetz, W. M. Abraham, T. Jones,
M. McAuliffe, H. Piechuta, D. A. Nicoll-Griffith, Z. Wang, R. Zamboni,
R. N. Young and K. M. Metters, J. Med. Chem., 2007, 50, 794.
40. B. Meunier, S. P. de Visser and S. Shaik, Chem. Rev., 2004, 104, 3947.
41. C. M. Bathelt, L. Ridder, A. J. Mulholland and J. N. Harvey, Org.
Biomol. Chem., 2004, 2, 2998.
42. L. K. Low and N. Castagnoli Jr., in Burgers Medicinal Chemistry: Basis
of Medicinal Chemistry, ed. M. E. Wolff, Part 1, J. Wiley and Sons,
Hoboken, NJ, USA, 1979, pp. 107–226.
43. D. V. Parke, Biochem. J., 1960, 77, 493.
44. W. Jacobsen, B. Kuhn, A. Soldner, G. Kirchner, K.-F. Sewing, P. A.
Kollman, L. Z. Benet and U. Christians, Drug Metab. Dispos., 2000, 28,
1369.
45. P. G. Dayton, J. M. Perel, R. F. Cummingham, Z. H. Israeli and I. M.
Weiner, Drug Metab. Dispos., 1973, 1, 742.
46. P. G. Dayton and J. M. Perel, Ann. N. Y. Acad. Sci., 1971, 179, 399.
47. M. W. Dysken, J. l. Javai, S. S. Chang, C. Schaffer, A. Shahid and J. M.
Davis, Psychopharmacology, 1981, 17, 205.
48. T. L. Perry, C. F. A. Culling, K. Berry and S. Hansen, Science, 1964, 146,
81.
49. L. Quartara and M. Altamura, Curr. Drug Targets, 2006, 7, 975.
50. R. K. Verbeck, J. L. Blackburn and G. R. Loewen, Clin. Pharmacokin.,
1983, 8, 297.
51. S. B. Rosenblum, T. Huynh, A. Afonso, H. R. Davis Jr, N. Yumibe, J. W.
Clader and D. A. Burnett, J. Med. Chem., 1998, 41, 973.
52. S. Kariya, S. Isozaki, K. Uchino, T. Suzuki and S. Narimatsu, Biol.
Pharm. Bull., 1996, 19, 1511.
53. M. Rowley, D. J. Hallett, S. Goodacre, C. Moyes, J. Crawforth, T. J.
Sparey, S. Patel, R. Marwood, S. Patel, S. Thomas, L. Hitzel,
D. O’Connor, N. Szeto, J. L. Castro, P. H. Hutson and A. M. Macleod,
J. Med. Chem., 2001, 44, 1603.
54. T. D. Penning, J. J. Talley, S. R. Bertenshaw, J. S. Carter, P. W. Collins, S.
Docter, M. J. Graneto, L. F. Lee, J. W. Malecha, J. M. Miyashiro,
R. S. Rogers, D. J. Rogier, S. S. Yu, G. D. Anderson, E. G. Burton, J. N.
Cogburn, S. A. Gregory, K. M. Koboldt, W. E. Perkins, K. Seibert, A. W.
Veenhuizen, Y. Y. Zhang and P. C. Isakson, J. Med. Chem., 1997, 40, 1347.
55. M. C. Lu, in Foye’s Principles of Medicinal Chemistry, ed. D.A. Williams
and T. L. Lemke, Lippincott Williams and Wilkins, Baltimore, 2002,
pp. 338–353.
56. D. K. Ernawati, L. Stafford and J. D. Hughes, Br. J. Clin. Pharmacol.,
2008, 66, 82.
Influence of Aromatic Rings on ADME Properties of Drugs 325
57. K. N. S. Sirajudeen, P. Gurumoorthy, H. Devaraj and S. N. Devaraj,
Drug Chem. Toxicol., 2002, 25, 247.
58. C. Lafuente-Lafuente, J. C. Alvarez, A. Leenhardt, S. Mouly, F. Extra-
miana, C. Caulin, C. Funck-Brentano and J. F. Bergmann, Br. J. Clin.
Pharmacol., 2009, 67, 511.
59. M. D. Freedman and J. C. Somberg, J. Clin. Pharmacol., 1991, 31, 1061.
60. B. Clark, D. A. Smith, C. T. Eason and D. V. Parke, Xenobiotica, 1982,
12, 147.
61. D. A. Smith, K. Brown and M. G. Neale, Drug Metab. Rev., 1985, 16,
365.
62. J. M. Grindel, B. H. Migdalof and W. A. Cressman, Drug Metab. Dispos.,
1979, 7, 325.
63. W. A. Cressman, J. R. Bianchine, V. B. Slotnick, P. C. Johnson and
J. Plostnieks, Euro. J. Clin. Pharmacol., 1974, 7, 99.
64. A. S. Kalgutkar and J. R. Soglia, Expert Opin. Drug Metab. Toxicol.,
2005, 1, 91.
65. J. C. Erve, Expert Opin. Drug Metab. Toxicol., 2006, 2, 923.
66. M. Cosman, C. D. L. Santos, R. Fiala, B. E. Hingertyt, S. B. Singht,
V. Ibanezt, L. A. Margulist, D. Live, N. E. Geacintov, S. Broyde and D. J.
Patel, Proc. Natl. Acad. Sci. U. S. A., 1992, 89, 1914.
67. A. Dipple, Carcinogenesis, 1995, 16, 437.
68. C. Heidelberger, Ann. Rev. Biochem., 1975, 44, 79.
69. F. Oesch, Biochem. Pharmacol., 1976, 25, 1935.
70. J. A. Hinson, N. R. Pumford and S. D. Nelson, Drug Metab. Rev., 1994,
26, 395.
71. J. A. Hinson and D. W. Roberts, Annu. Rev. Pharmacol. Toxicol., 1992,
32, 471.
72. W. Chen, L. L. Koenigs, S. J. Thompson, R. M. Peter, A. E. Rettie, W. F.
Trager and S. D. Nelson, Chem. Res. Toxicol., 1998, 11, 295.
73. J. A. Hinson, A. B. Reid, S. S. Mccullough and L. P. James, Drug Metab.
Rev., 2004, 36, 805.
74. B. K. Park, N. R. Kitteringham, J. L. Maggs, M. Pirmohamed and D. P.
Williams, Annu. Rev. Pharmacol. Toxicol., 2005, 45, 177.
75. H. Jewell, J. L. Maggs, A. C. Harrison, P. M. O’Neill, J. E. Ruscoe and B.
K. Park, Xenobiotica, 1995, 25, 199.
76. J. N. Tettey, J. L. Maggs, W. G. Rapeport, M. Pirmohamed and B. K.
Park, Chem. Res. Toxicol., 2001, 14, 965.
77. M. T. Smith, Chem. Res. Toxicol., 2003, 16, 679.
78. S. Prabhu, A. Fackett, S. Lloyd, H. A. McClellan, C. M. Terrell, P. M.
Silber and A. P. Li, Chem-Bio. Interact., 2002, 142, 83.
79. K. Kassahun, P. G. Pearson, W. Tang, I. McIntosh, K. Leung, C. Elmore,
D. Dean, R. Wang, G. Doss and T. A. Baillie, Chem. Res. Toxicol., 2001,
14, 62.
80. E. L. Cavalieri, K.-M. Li, N. Balu, M. Saeed, P. Devanesan, S. Higgin-
botham, J. Zhao, M. L. Gross and E. G. Rogan, Carcinogenesis, 2002, 23,
1071.
326 Chapter 6
81. E. Cavalieri, K. Frenkel, J. G. Liehr, E. Rogan and D. Roy, J Natl.
Cancer Inst. Monogr., 2000, 27, 75.
82. C. P. Martucci and J. Fishman, Pharmacol. Ther., 1993, 57, 237.
83. J. L. Bolton and G. R. J. Thatcher, Chem. Res. Toxicol., 2008, 21, 93.
84. J. L. Bolton, Adv. Mol. Toxicol., 2006, 1, 1.
85. W. Tang, Curr. Drug Metab., 2003, 4, 319.
86. W. Tang, R. A. Stearns, S. M. Bandiera, Y. Zhang, C. Raab, M. P.
Braun, D. C. Dean, J. Pang, K. H. Leung, G. A. Doss, J. R. Strauss, G. Y.
Kwei, T. H. Rushmore, S.-H. L. Chiu and T. A. Baillie, Drug Metab.
Dispos., 1999, 27, 365.
87. G. K. Poon, Q. Chen, Y. Teffera, J. S. Ngui, P. R. Griffin, M. P. Braun,
G. A. Doss, C. Freeden, R. A. Stearns, D. C. Evans, T. A. Baillie and W.
Tang, Drug Metab. Dispos., 2001, 29, 1608.
88. S. Madden, T. F. Woolf, W. F. Pool and B. K. Park, Biochem. Phar-
macol., 1993, 46, 13.
89. T. F. Woolf, W. F. Pool, S. M. Bjorge, T. Chang, O. P. Goel, C. F.
Purchase, M. C. Schroeder, K. L. Kunze and W. F. Trager, Drug Metab.
Dispos., 1993, 21, 874.
90. A. S. Kalgutkar, A. D. N. Vaz, M. E. Lame, K. R. Henne, J. Soglia, S. X.
Zhao, Y. A. Abramov, F. Lombardo, C. Collin, Z. S. Hendsch and C. E.
C. A. Hop, Drug Metab. Dispos., 2005, 33, 243.
91. P. W. Fan and J. L. Bolton, Drug Metab. Dispos., 2001, 29, 891.
92. Q. Chen, J. S. Ngui, G. A. Doss, R. W. Wang, X. Cai, F. P. DiNinno, T.
A. Blizzard, M. L. Hammond, R. A. Stearns, D. C. Evans, T. A. Baillie
and W. Tang, Chem. Res. Toxicol., 2002, 15, 907.
93. B. R. Baer, L. C. Wienkers and D. A. Rock, Chem. Res. Toxicol., 2007,
20, 954.
94. J. T. Pearson, J. L. Wahlstrom, L. J. Dickmann, S. Kumar, J. R. Halpert,
L. C. Wienkers, R. S. Foti and D. A. Rock, Chem. Res. Toxicol., 2007, 20,
1778.
95. J. G. Liehr, Mol. Pharmacol., 1983, 23, 278.
96. Y.-J. Wu, C. D. Davis, W. C. Dworetzky, W. C. Fitzpatrick, D. Harden,
H. He, R. J. Knox, A. E. Newton, T. Philip, C. Polson, D. V. Sivarao,
L.-Q. Sun, S. Tertyshnikova, D. Weaver, S. Yeola, M. Zoeckler and
M. W. Sinz, J. Med. Chem., 2003, 46, 3778.
97. X. Li, Y. He, C. H. Ruiz, M. Koenig and M. D. Cameron, Drug Metab.
Dispos., 2009, 37, 1242.
98. S. Ahlenius, E. Ericson, V. Hillegaart, L. B. Nilsson, P. Salmi and
A. Wijkström, J. Pharmacol. Exp. Ther., 1997, 283, 1356.
99. S. T. Laidlaw, J. A. Snowden and M. J. Brown, Lancet, 1993, 342,
1245.
100. J. C. Erve, M. A. Swensson, H. Von Euler-Chelpin and E. Klasson-
Wehler, Chem. Res. Toxicol., 2004, 17, 564.
101. D. Zhang, R. Krishna, L. Wang, J. Zeng, J. Mitroka, R. Dai, N.
Narasimhan, R. A. Reeves, N. R. Srinivas and L. J. Klunk, Drug Metab.
Dispos., 2005, 33, 83.
Influence of Aromatic Rings on ADME Properties of Drugs 327
102. P. Hlavica, I. Golly, M. Lehnerer and J. Schulze, Hum. Exp. Toxicol.,
1997, 16, 441.
103. R. B. Silverman, in The Organic Chemistry of Drug Design and Drug
Action, Academic Press, San Diego, 2004, pp. 405–495.
104. J. P. Sanderson, D. J. Naisbitt, J. Farrell, C. A. Ashby, M. J. Tucker,
M. J. Rieder, M. Pirmohamed, S. E. Clarke and B. K. Park, J. Immunol.,
2007, 178, 5533.
105. P. D. Stonier, Pharmacoepidemiol. Drug Saf., 1992, 1, 177.
106. R. L. Hare, B. Holcomb, O. C. Page and J. W. Stephens, N. Engl. J. Med.,
1957, 256, 74.
107. J. P. Uetrecht, J. Pharmacol. Exp. Ther., 1985, 232, 420.
108. Z. H. Israili, S. A. Cucinell, J. Vaught, E. Davis, J. M. Lesser and P. G.
Dayton, J. Pharmacol. Exp. Ther., 1973, 187, 138.
109. S. J. Grossman and D. J. Jollow, J. Pharmacol. Exp. Ther., 1988, 244, 118.
110. J. Uetrecht, N. Zahid, N. H. Shear and W. D. Biggar, J. Pharmacol. Exp.
Ther., 1988, 245, 274.
111. R. Mahmud, M. D. Tingle, J. L. Maggs, M. T. Cronin, J. C. Dearden and
B. K. Park, Toxicology, 1997, 117, 1.
112. N. Borges, Expert Opin. Drug Saf., 2005, 4, 69.
113. K. S. Smith, P. L. Smith, T. N. Heady, J. M. Trugman, W. D. Harman
and T. L. Macdonald, Chem. Res. Toxicol., 2003, 16, 123.
114. T. Wikberg, A. Vuorela, P. Ottoila and J. Taskinen, Drug Metab. Dispos.,
1993, 21, 81.
CHAPTER 7
Influence of Heteroaromatic
Rings on ADME Properties of
Drugs
DEEPAK DALVIE, PING KANG, CHO-MING LOI,
LANCE GOULET AND SAJIV NAIR
7.1 Introduction
Aromatic heterocycles play a critical role in medicinal chemistry and drug
design.1–3 More than half of known drugs contain at least one heterocyclic
component in their structure. New chemical entities are generally designed as
analogues of endogenous ligands that are vital to biochemical processes or, in
some cases, as simplified congeners of complex natural products. Since most of
these substrates contain heterocyclic rings, these rings by default become core
structures of the newly designed compounds. These drugs compete with
endogenous ligands and mimic (agonist) the ligands in their action. Alter-
natively, they block (antagonist) the active site in the receptor and therefore its
normal function. Some common endogenous ligands that have heteroaromatic
rings as their core structure include histamine, 5-hydroxytryptamine (sero-
tonin), adenosine 5 0 -triphosphate (ATP) and nucleic acid bases such as the
328
Influence of Heteroaromatic Rings on ADME Properties of Drugs 329
uracil, thymine, cytosine, guanine and adenine (Figure 7.1). Figure 7.2 repre-
sents some drugs that are structurally similar to these ligands.
Cimetidine, an antiulcer drug (Figure 7.2), was one of the first compounds
developed through rational drug design.4 This is a classic example of where the
endogenous ligand, histamine, was used as a lead to design a H2 receptor
NH2 NH2
N N
HO O O O
N NH2
O P O P O P O N N
N O
HN O O O
H
Histamine 5-Hydroxytryptamine (5HT)
HO OH
Adenosine5′-Triphosphate (ATP)
NH2
N
H H
H3C N NHCH3
N N
H3C S
S N
NCN O2
HN N N
N H N
H
Cimetidine Sumatriptan Rizatriptan
H2 Receptor Antagonist 5-HT1B and 5-HT1D Agonist 5-HT Agonist
Cl
O N
N N N HN
N
N O
N N S O
O HN Cl
N H
Gefitinib HO Dasatinib
F Tyrosine kinase Inhibitor
Tyrosine kinase Inhibitor
O
F
NH
N O
H
5-Fluorouracil
Anticancer Agent
HO O
N N
HN
O N
HN
N NH2
Cl N
Quinine Primaquine Chloroquine
O
N
O O
N
O
O
NH2
O
Cocaine Procaine
HO O F
O O
N N
S
O N
O O
H
OH
OH OH O
Lovastatin Rosuvastatin
Figure 7.3 Structures of some naturally occurring ligands and their synthetic analogs.
NH2
CH3
H3C O
N
CH3
O H3C N
Acetylcholine
Endogenous Ligand Tacrine
OH H
H O
HO N HO N
R
N O
HO
R = CH3, Adrenaline N N
R = H, Noradrenaline S Timolol
Figure 7.4 Examples endogenous ligands that lack heteroaromatic rings and their
structural mimics.
O CO2H O CO2H
NH2
NH2
N
N N
O N
HO N N
HO
O O
OH H OH H
Deoxycytidine Deoxyadenosine
NH2 NH2
N N N O
O N N N Cl F
HO HO NH
O O
N O
F H
OH H
OH F
Gemcitabine Cladribine
5-Fluorouracil
A nucleoside analog A nucleoside analog
Anticancer Agent
used in chemotherapy used in chemotherapy
O O
N O
S
NH HN
S
O
O O
Pioglitazone
Ciglitazone
N
Pyrimidine 80.1 0.26 1.78 25.78
N
N
1,2,4-Triazine 81.1 0.92 1.32 38.67
N
N
1,3,5-Triazine 81.1 0.73 1.59 38.67
N N
N
N
N
Tetrazine 82.1 1.47 1.6 51.56
N
N
Quinazoline 130.2 1 3.44 25.78
N
a
All values were calculated using ACD software (version 11.0).
334 Chapter 7
Although the presence of an electronegative nitrogen atom in the ring does not
alter the character of the p-orbitals of these rings, it distorts the distribution of
the p electrons and therefore lowers the energies of p-orbitals relative to the
phenyl ring. This coupled with the inductive effect of the electronegative nitrogen
atom makes it less reactive and renders properties that corroborate with con-
jugated imines or conjugated carbonyls.15 As the number of sp2 nitrogen atoms
in the ring increase, there is a further decrease in the reactivity of the rings. This
makes it less susceptible to P450 mediated oxidation. However, the lone pair of
electrons on the nitrogen atom that are planar to the ring provide a site for
protonation and other reactions such as N-oxidation (equivalent to a tertiary
amine), which is not analogous to the carbocyclic ring system.
The presence of sp2 nitrogen atoms in the ring also affects the physicochemical
properties of these rings. These rings are weak bases in comparison with their
alicyclic counterparts, piperidine or piperazines. As shown in Table 7.1, pyridine
and quinoline are the most basic of them all, with a pKa of B5. As the number of
nitrogen atoms in the ring increase, the pKa decreases. The pKa is also affected
by the position of the second nitrogen atom in the ring (Table 7.1). The repla-
cement of a methine group of the phenyl ring with nitrogen atoms also affects the
lipophilicity (log P value) and the polarity of heteroaromatic rings compared
with the benzenenoid rings. As the number of ring nitrogens increase, the log P of
the fragment decreases and the polarity increases (Table 7.1). Thus, introduction
of one nitrogen atom lowers the log P of a phenyl ring from 2.13 to 0.65, while
the log P value of pyrimidine and triazine ring is 0.3 and 0.7, respectively.
Five-membered heterocyclic rings form the second popular class of hetero-
aromatic ring systems that are commonly used as substructures in compound
synthesis. The more commonly used five-membered heterocyclic rings and
their benz-fused analogs are shown in Table 7.2 and Table 7.3, respectively.
Like six-membered rings, all five-membered rings and their benz-fused deri-
vatives share the same electronic structure of the phenyl ring, though the pre-
sence of only five atoms in the ring instead of six makes these rings smaller in
size relative to the carbocyclic and six-membered heteroaromatic rings.
Five-membered rings with a single heteroatom are the simplest ring systems
that belong to this class. The most common rings in this class are the pyrrole,
furan and the thiophene and their respective benz-fused analogs (indole, ben-
zofuran and benzothiophene); they contain a nitrogen, an oxygen and a sulfur
atom, respectively, in the ring system. Given the electron rich nature of these
rings, the five-membered rings are more reactive than the six-membered rings
and can undergo oxidative metabolism like the benzenoid rings (see Section
7.3.3). However, the degree of reactivity and the physicochemical properties of
these rings are dependent on the nature and the electronegativity of the het-
eroatom present in the ring (Table 7.2). Thus, thiophene and furan is the most
lipophilic of all with a log P value that is quite similar to benzene (1.74 and 1.95
versus 2.13) while the log P value of pyrrole is 0.85. The pyrrole moiety pos-
sesses an NH group that is typical of a secondary amine; however, the basicity
of pyrrole, pKa ¼ -3.8 for a conjugated acid, is much less than that of a sec-
ondary amine (pKa ¼ 10.87). This large difference is primarily due to the
Influence of Heteroaromatic Rings on ADME Properties of Drugs 335
Table 7.2 Physicochemical properties of commonly used five-membered rings.
Heterocycle Name MW LogP pKaa PSA pKab
N
4H-1,2,4-Triazole 69.1 0.89 2.7 41.57 10.2
H
N
NH 2H-1,2,4-Triazole 69.1 0.009 1.2 41.57 8.7
N
N
Indole 117 2.59 2.4 15.79
H
N
N
Indazole 118 1.77 1.26 28.68
H
N Benzisoxazole 119 2.02 2.03 26.03
O
incorporation of the non-bonding electron pair of the N-atom into the cyclic
conjugated system of the pyrrole molecule. The pyrrole ring is also a weak acid
(about the same strength as the simple alcohol) and the pKa of its conjugate
base is 17 in aqueous solution.
Influence of Heteroaromatic Rings on ADME Properties of Drugs 337
These three rings form a platform for rest of the heteroaromatic rings that
contain an additional sp2 nitrogen atom (Table 7.2). The variation in the
position of the extra nitrogen atom further adds to the structural diversity of
the groups especially for heterocycles containing the two or three heteroatoms.
Extra nitrogen atoms and their placement in the ring (relative to the first
heteroatom) also have an important effect on the properties of the ring system.
The additional nitrogen atom has lone pair of electrons that is in plane with the
ring and therefore has attributes that are quite similar to six-membered het-
eroaromatic rings. Like pyridine and its six-membered analogs, the increase in
number of nitrogen atoms causes a decrease in the energy levels of the p orbitals
so that the heterocycles are less electron rich compared with their counterparts
containing only one heteroatom. This makes them less reactive relative to the
pyrrole, thiophene and furan.
The effect of additional sp2 nitrogen atoms can also be seen on the acidity
and basicity of these heterocycles (Table 7.2). Even though all ‘azoles’ are weak
bases like the pyrrole, the lone pair of electrons adds sites of protonation and
therefore most ‘azoles’ are stronger bases than pyrrole (Table 7.2).
The position of the sp2 nitrogen atom relative to the first heteroatom also
influences the basicity of the rings. For example, the imidazole ring has a pKa
of around 7.1 while pyrazole (nitrogen atom at the two position relative to the
first nitrogen) has a pKa of 2.5. The azoles containing additional NH groups
(imidazole, pyrazole, triazoles and tetrazoles) are stronger acids than the pyr-
role. Thus, imidazoles and pyrazoles are stronger acids (pKa of 14) than pyr-
role (pKa B17), while the triazoles and the tetrazole have pKa values of 9.5
and 4.92 (which is equivalent to a carboxy group), respectively.
The addition of nitrogen atoms also affects the log P value of these rings. The
logP values of all rings are listed in Table 7.2. The additional nitrogen atom
adds polarity to the ring and hence lowers the lipophilicity of all these rings
relative to their one heteroatom analogs. However, the trend is similar to that
observed in the rings with single heteroatom, i.e. logPthiazole or isothiazole 4
logPoxazole or isoxazole 4 logPimidazole or pyrazole.
An illustration of the effect of incorporation of different heterocyclic rings on
physicochemical properties of compounds is depicted in Table 7.4. Replace-
ment of a pyridine ring in sulfapyridine with pyrazine or pyrimidine results
changes the clogP and PSA values significantly. Similarly, replacement of six-
membered rings with thiazole, isoxazole or thiadiazole in sulfathiazole, sulfa-
methoxazole and sulfamethizole affects the lipophilicity and the PSA of the
molecule depending upon the ring incorporated in the compound. Also, the
addition of the type of substituents and the position of attachment can some-
times affect the physicochemical properties of a compound.
In addition to monocyclic heterocycles and their benz-fused derivatives, the
‘hybrid’ bicyclic rings resulting from annulation of two aromatic heterocycles
have also become very popular among medicinal chemists.2 These rings are
comprised of either five-membered rings fused with five-membered rings or five-
membered rings fused with six-membered rings. Given the number of hetero-
cycles that belong in the five-membered rings and five-membered rings classes,
338 Chapter 7
Table 7.4 Effect of heteraromatic rings on the physicochemical properties of
sulfa drugs.
Compound Structure MW clogP clogD7.4 PSA
H2N N
O
Sulfapyridine S
249 0.47 0.41 93.46
N
O H
H2N N
O
Sulfadiazine N 250 0.07 0.99 106
S N
O H N
H2N O
Sulfapyrazine 250 0.05 1.68 106
S N N
H2N O H
O S
Sulfathiazole 255 0.05 0.33 121.7
S
N N
O H
O H
N
Sulfamethoxazole S 253 0.66 0.54 106
O N
H2N O
O H
N N
S
Sulfamethizole N 270 0.52 1.16 134
O S
H2N O H
N
S
Sulfisomidine O N 278 0.3 0.004 106
N
H2N
O H N
N
S
Sulfadimidine O N
278 0.58 1.0 106
H2N
H2N
O O
Sulfamoxzole S
267 0.68 0.32 106
N N
O H
several bicyclic rings can be envisioned in this hybrid class. Some members of
this group are commonly used in drug design and are considered bioisosteres of
the benz-fused six-membered and five-membered rings. Others are commonly
observed in the natural and endogenous ligands, such as the purine ring (imi-
dazopyrimidine) observed in DNA bases or the pteridine (pyrazinopyrimidine)
observed in folic acid. Drugs containing these rings are shown in Figure 7.7.
O
N
N N N
N N O N
N
O N N
N N O
H2N
N
Zaleplon Zolpidem Temozolomide
7.3.1 Absorption
For most molecules, the rate and extent of absorption is largely dependent on
the dissolution of the compound in the aqueous contents of the gastrointestinal
tract and the permeability through the membrane. These in turn depend upon
the molecular characteristics of a drug such as molecular weight/size, aqueous
solubility, lipophilicity and the hydrogen bonding capacity of the molecule.16
Since aqueous solubility dictates the amount of drug that will be available for
absorption, most efforts are often directed towards the optimisation of solu-
bility of molecules.17 One approach involves introduction of polar solubilising
groups in a molecule to modulate the lipophilicity of the compound and hence
increase its absorption. Replacement of phenyl rings in the molecule with
aromatic heterocycles does not change the molecular weight (MW) of the
compound to a significant extent. However, it can modify the log D and pKa of
compounds to improve the absorption properties of a molecule.
Pyridine ring is one of the heterocycles that is commonly used to modify the
absorption properties of molecules. Replacement of a phenyl ring with this ring
not only imparts unique ionising properties at the acidic pH (therefore
increasing solubility), but also provides a handle for the introduction of various
groups that help in modulating the solubility and lipophilicity characteristics of
new molecules, which can assist in absorption of the compound. Khanna and
coworkers exemplify this in their continued efforts to seek new diaryimidazoles
as COX-2 inhibitors with better oral activity.18 Replacement of a tolyl group in
compound 7.1 (Table 7.5) with a pyridine ring (compound 7.2) yielded het-
eroaromatic analogues with improved water solubility which resulted in
excellent bioavailability (490%) and high plasma levels in the rat following a
dose of 20 mg kg1 relative to 7.1. The key to enhanced absorption of 7.2 was
probably ascribed to improved physicochemical properties of this compound
340 Chapter 7
18
Table 7.5 Physicochemical properties of COX-2 inhibitors.
CF3 CF3
N N
N N
N
SO2CH3 SO2CH3
Property Compound 7.1 Compound 7.2
MW 380 367
ClogP 2.68 1.13
PSA 60.34 73.23
clogD6.5 2.68 1.13
Sol6.5 (mg/ml) 0.005 0.05
(Table 7.5). Replacement of tolyl ring in 7.1 with a polar bioisosteric pyridyl
ring in 7.2 lowered the lipophilicity of 7.2 (clogP of 7.1 ¼ 2.68 and clogP of
7.2 ¼ 1.13) sufficiently to increase its aqueous solubility and therefore it
bioavailability.
Incorporation of a pyridine ring has also been shown to alter the absorption
characteristics of the glitazones. Assessment of the pharmacokinetic properties
of rosiglitazone indicated complete absorption following oral administration of
[14C]rosiglitazone solution (8 mg) and rosiglitazone tablets (4 mg, B.I.D.),
resulting in an absolute bioavailability of 95% and 99%, respectively.19 In
contrast, the oral bioavailability of troglitazone was estimated to be 40–50%
following oral administration of a 400 mg dose to humans.20 In addition,
administration of troglitazone (400 mg) with high fat meal increased the plasma
AUC by 30% to 80%, suggesting that oral absorption was incomplete in the
fasted state. Comparison of the physicochemical properties of these two gli-
tazones is shown in Table 7.6. Enhanced absorption of rosiglitazone is surmised
in part to the improved solubility at the clinically relevant doses. The key
modification is the replacement of the chroman moiety in troglitazone with an
aminopyridine ring in rosiglitazone. This change resulted in lower clogP of
rosiglitazone relative to troglitazone (3.02 versus 4.69) and increase in solubility
(0.05 versus 0.003 mg/ml) (Table 7.6). It is important to note that the dose,
which to a degree is also dependent upon its potency, also affects the solubility
of the compound in the gastrointestinal tract. Thus, in addition to lipophilicity,
increased solubility of rosiglitazone in contrast to troglitazone can be ascribed
to a lower dose (4 mg B.I.D., or 8 mg QD) of rosiglitazone in contrast to
troglitazone (200–400 mg, QD). The primary difference between the two
structurally related glitazones is its affinity for the PPAR-g receptor (332 nM
for troglitazone and 36 nM for rosiglitazone).
Optimum pKa and logP value of imidazole rings also makes this ring very
attractive in imparting solubilising properties in a molecule and modulating the
lipophilicity. These rings with appropriate substituents are also commonly used
Influence of Heteroaromatic Rings on ADME Properties of Drugs 341
19,20
Table 7.6 Comparison of troglitazone and rosiglitazone.
O
S O
HN
S NH
O HO O
N O
O
N
O
MW 357.4 441.5
clogP 3.02 4.69
PSA 96.83 110.16
clogD6.5 2.35 4.31
Sol6.5 (mg/ml) 0.05 0.003
% Oral 95–99% 40–50%
bioavailability
H2N
N O
Caco-2 Papp
R MW clogP PSA cm sec1106
SO2CH3
7.3 559 3.1 142 o0.1
SO2CH3
7.4 560 3.0 154 0.54
N
N
7.5 485 1.9 117 4.6
N
N
Razaxaban 528 3.24 120 5.6
N N
which is used as a first line treatment for CML. In humans, the absolute oral
bioavailability of imatinib is 98%, indicating that absorption is almost com-
plete.22 In contrast, the systemic exposure of nilotinib is increased by 82% when
administered 30 minutes after a high fat meal, suggesting that oral absorption
of this drug in the fasted state is incomplete.23 The variation in their absorption
properties can be attributed to the differences in the lipophilicity and solubility
(at pH 6.5) between the two compounds (Table 7.8). Incorporation of a 4-
methylimidazole group in nilotinib increases the clogD6.5 to 4.36 and reduces
the solubility to 0.0004 mg ml1 relative to imatinib (clogD6.5 ¼ 1.77 and
solubility 0.11 mg ml1). Thus, replacement of the substituted imidazole with
the piperazine side chain appears to be associated with decreased lipophilicity,
increased solubility and an increase in the extent of oral absorption.
The tetrazole moiety is comparable in size and in acidity to the -COOH
group and is therefore commonly used as a bioisosteric replacement of car-
boxylic acids.24 Tetrazoles as isosteres of carboxylic acids are also discussed in
detail in Section 3.7.2. The tetrazole moiety is 10-fold more lipophilic and
Influence of Heteroaromatic Rings on ADME Properties of Drugs 343
Table 7.8 Comparison of nilotinib (Tasigna) and imatinib (Gleevec).
O O
HN NH
N
N NH NH
N N
F
N N N N
F N
F N
O
O N
NH
OH
S N N
Property Eprosartan Losartan
masks the polarity of the carboxylate group, and hence tends to increase the
bioavailability of compounds. The effect of replacing a carboxylic acid func-
tionality with a tetrazole ring on absorption has been demonstrated in angio-
tensin II converting receptor antagonists, eprosartan and losartan (Table 7.9).
Assessment of the pharmacokinetic properties of losartan and eprosartan
indicates that, even though the bioavailability of losartan in humans was only
37%, the drug is completely absorbed through the gut following oral admin-
istration.25 In contrast, only 15% of eprosartan was absorbed through the gut,
resulting in bioavailability of 13%.26,27 Low absorption is inherent in the
physicochemical properties of eprosartan rather than incomplete dissolution.
As shown in Table 7.9, the tetrazole ring in losartan and the carboxy group in
344 Chapter 7
HN N
N N
CO2H
O CO2H O CO2H
O H O H
H H
O N N O N N
N N
H O H O
O O
7.8 7.9
clogD7.4 = 0.76 clogD7.4 = 0.2
Papp = 5.1 x 10-7 cm/sec Papp = <1 x 10-7 cm/sec
eprosartan have pKa values of 6 and B4, respectively; hence both compounds
are almost completely ionised at the pH in the gastrointestinal tract (B6.5).
Although losartan is less soluble than eprosartan (0.04 mg/ml versus 4.6 mg/
ml), the negative clogD6.5 of eprosartan (log D6.5 of 0.53) hinders the per-
meability of the compound through the membrane. On the other hand,
replacement of dicarboxylic acid with a tetrazole moiety and a hydroxymethyl
substituent in losartan results in a clogD6.5 of 2.6, thereby enhancing the per-
meability through the gastrointestinal tract in spite of poorer solubility. In
addition to lipophilicity, the difference in the polar surface area of the two
compounds (92.5 Å2 for losartan versus 120 Å2 for eprosartan) may favour
permeability of losartan through the membranes compared to eprosartan.
In another example, Liljebris and co workers have reported the influence of
tetrazole ring on the bioavailability of small molecular weight peptidomimetics
that were designed and synthesised as competitive inhibitors of protein tyrosine
phosphatases 1B (PTP1B).28 Because of the low cell permeability of this com-
pound class, the possibility of replacing one or both of the remaining carboxyl
groups while maintaining PTP1B inhibitory activity was explored. An impor-
tant discovery was the ortho tetrazole analogue 7.8, which was equipotent to the
lead dicarboxylic acid analogue 7.9 (Figure 7.8). This novel monocarboxylic
acid analogue revealed higher Caco-2 cell permeability and this was attributed
to the increase in lipophilicity compared to the lead (Figure 7.8).
7.3.2 Distribution
Central nervous system (CNS) penetration is an important consideration in
drug discovery and development for many therapeutic areas. Thus, examples
that demonstrate the influence of aromatic heterocycles on the permeability of
compounds to cross the blood–brain barrier (BBB) are discussed in this section.
Like absorption, permeability of a drug across the blood–brain barrier is
affected by its PSA and logP, i.e. the polarity of a compound.16 Since the
Influence of Heteroaromatic Rings on ADME Properties of Drugs 345
Table 7.10 Comparison of physicochemical properties and ratio of the brain
to blood concentrations for mianserin and mirtazapine.29
N N
N N N
S
N N
O N N O N N
H H
Compound 7.10 Compound 7.11
N N
sulfur atom is a weak hydrogen bond acceptor, its presence in the ring can
affect the PSA of the molecule as observed in 7.11 (PSA for 7.10 is 37.39 versus
65.63 for 7.11). However, no significant change in the lipophicity is observed
with such a replacement (clogP for 7.10 is 3.75 and clogP for 7.11 is 3.93).
Determination of brain to blood ratio of compounds 7.10 and 7.11 shows a 1.8-
fold decrease when the aminopyridine group in 7.10 is replaced with ami-
nothiazole moiety in compound 7.11.29 Since the clogP for both compounds are
similar, the decrease in the brain to blood ratio in 7.11 is ascribed to the change
in PSA of the two compounds.
Recent studies have shown that the permeability across the blood–brain
barrier is primarily affected by the number of hydrogen bond donors than
hydrogen bond acceptors present in the molecule.31,32 Reduction in the number
of hydrogen bond donors in a molecule has been attempted by addition of
chlorothienyl moiety to the cyclic carbamate 7.12 (Table 7.12) by Tatsumi and
co-workers.33 Compound 7.12 was synthesised as a part of developing a7
neuronal nicotininc acetylcholine receptor agonists for the treatment of schi-
zophrenia. Addition of the chlorothienyl group led to an increase in brain to
Influence of Heteroaromatic Rings on ADME Properties of Drugs 347
O
N
OCH3
HN N
HN
HN O
HN O
N
N N
Figure 7.9 Intramolecular hydrogen bonding in the most polar compound increases
the brain penetration.
plasma ratio from 0.76 for 7.12 to 59 for 7.13. The change was rationalised in
part to the reduction in hydrogen bond donating capability of 7.13, resulting in
a slightly lower PSA relative to 7.12. This approach also led to an overall
improvement of clogD7.4 of the compound (clogD7.4 ¼ 0.57 for 7.13 versus -1.5
for 7.12). The calculated physicochemical parameters are shown in Table 7.12.
A group of oxerin-1 and 2 (OX1 and OX2) receptor antagonists, designed
and synthesised by Porter and his co-workers, represents an interesting example
where introduction of a polar napthyridine group in a molecule increased its
brain to plasma ratio in spite of lower log P and higher PSA relative to the
lead.34 The lead compound 7.14 (Figure 7.9) showed good affinity for the
receptors but exhibited poor blood–brain barrier permeability following an
intravenous infusion to rats. The compound was essentially undetectable in the
CNS even though it exhibited good lipophilicity (clogP, 4.4; clogD, 2.6) and a
moderate polar surface area of 59 Å2. Structure–activity relationship (SAR)
studies to optimise potency and selectivity resulted in compound 7.15 that
showed considerably higher brain to plasma ratio (0.4) relative to 7.14. Higher
ratios were observed despite the lower clogP than 7.14 (clogP 3.4) and sig-
nificantly higher PSA (102 Å2). So in theory, even though 7.15 has two
hydrogen bond donors like the lead and unfavourable PSA, potential intra-
molecular hydrogen bonding between the nitrogen atom at the 5-position of
napthyridine and the proton on the proximal urea nitrogen can mask its proton
donor ability and therefore reduce the PSA. This in turn can help in enhancing
BBB penetration.
The previous section describes the advantages of introducing tetrazolyl
moiety as a bioisostere of carboxylic acid in an effort to increase in the bioa-
vailability of losartan. However, introduction of this ring has been shown to
negatively affect the brain exposure of compounds. This has been observed with
some tetrazolyl analogs that were synthesised as CCK-B antagonists.35 In this
report, replacement of the methyl group on the phenyl urea (7.16) with a
348 Chapter 7
Table 7.13 Comparison of the physicochemical properties of CCK-B
antagonists (1 and 2) and mGlu2 modulators (3 and 4).35,36
H
N N
O O
N NH O O N N
NH N NH
N NH
Ph 7.16 N
7.17
Ph
OH O
OH O
O
O S
O
N
N N
7.18 7.19
N NH
Compound MW clogP clogD PSA HBD HBA
tetrazole moiety (7.17) (Table 7.13) resulted in poor brain exposure of 7.17.
Similarly, enhanced brain to plasma (B/P) ratios in rat were observed following
replacement of tetrazole terminus in 7.18 (B/P ¼ 0.01) with a 4-thiopyridyl
group (47.19) (B/P ¼ 1.1) for the modulators of mGlu2 receptor (Table 7.13).36
In both examples, replacement of the substituent with the tetrazole moiety
increased the number of hydrogen bond donors and hydrogen bond acceptors,
and therefore increased PSA. Although there was a decrease in the clogD in
both cases, the authors attributed the poor permeability to cross the blood–
brain barrier to a high PSA and an increase in hydrogen bond donors.
7.3.3 Metabolism
In addition to enhancing absorption properties, heterocyclic rings are often
incorporated into a molecule to improve the metabolic stability of newly syn-
thesised compounds. Replacement of a carbocyclic ring with an aromatic
heterocycle is a common strategy to reduce the metabolic liability and some-
times block the sites of metabolism.
Structure–metabolism relationship studies have shown that incorporation of
one or more heteroatoms in an aromatic ring of a compound modulates its
chemical and biochemical reactivity (in addition to changing lipophilicity), and
therefore alters its metabolism. The degree of influence of heteroaromatic rings
on the metabolism of a compound depends on the type of ring system that is
Influence of Heteroaromatic Rings on ADME Properties of Drugs 349
incorporated in a molecule and the number of nitrogen atoms present in the
ring. As described in Section 7.2, compounds containing six-membered ring
systems are less reactive than their corresponding carbocyclic analogs. Incor-
poration of one or more nitrogen atoms in the rings decreases the electron
density in the aromatic ring carbons, thereby decreasing P450-mediated aro-
matic oxidation. Thus, the lower reactivity of six-membered rings with multiple
nitrogen atoms makes these prime fragments to improve metabolic stability
such as higher resistance to chemical or enzymatic degradation.
Despite low reactivity, compounds containing these rings are not devoid of
P450 catalysed metabolic transformations. For example, voriconazole, an
antifungal agent containing a pyrimidine ring, primarily undergoes CYP2C19
or 3A4-mediated N-oxidation of the pyrimidine ring (Figure 7.10).37,38 Simi-
larly, compounds containing a pyridine ring such as indinavir or rosiglitazone
are subject to oxidative metabolism either on the pyridine ring or on the rest of
the molecule by P450 (Figure 7.11).19,39 Binding of these compounds to the
P450 active site is attributed to their overall lipophilicity (clogD7.4 for indina-
vir ¼ 3.4; clogD7.4 for rosiglitazone ¼ 2.5 and clogD7.4 for voriconazole ¼ 1.9).
Alternatively, the presence of the lone pair of electrons on the nitrogen atom
of the aromatic heterocycles provides a site for glucuronidation by UDP-glu-
curonyl transferases (UGT) for most six-membered rings. Pyridine containing
compounds are readily prone to glucuronidation. Indinavir, which undergoes
extensive oxidative metabolism, is also subject to N-glucuronidation (Figure
7.11).39,40 Other examples include compounds such as nicotine that form the N-
glucuronide in humans (Figure 7.12).41 Lamotrigine, an anti-epileptic agent
containing a triazene ring, is also metabolised via N-glucuronidation (Figure
7.12).42 This suggests the importance of nucleophilicity of the nitrogen atom
rather than the reactivity of the heteroaromatic ring in glucuronidation of the
molecule. Not much is known about the interactions of the substrates with
UGTs.
A unique metabolic pathway for compounds containing electron deficient
azaheterocycles is oxidation by molybdenum hydroxylases such as aldehyde
oxidase (AO) and xanthine oxidase (XO).43 Unlike P450s that promote the
N N
N N
N N
OH F OH F
CYP3A4
N F N F
F F
N N N-Oxide
Voriconazole O Metabolite
N
O N N-Oxide metabolite
HN
O
OH O
N
N N-Glucuronide
N OH N conjugate
H N CO2H
O
Indinavir N HO OH
OH
Other Oxidative
Metabolites
O
O
NH OH Glucuronides
HN
O S
N O S
N N
O N
O Sulfates
Rosiglitazone
CH3
N
CO2H
O OH
N N
N H3C HO
OH
Nicotine
N-Glucuronide of
Nicotine
HO2C
NH2 NH2 O OH
OH
N N N N HO
N N
NH2 H2N
Cl Cl N-Gluruconide
of Lamotrigine
Cl Cl
Lamotrigine
HN NH
N Br
O N N
N O N
N N
Metyrapone Quinine
Brominidine
HN
O
H O
N O O
N
N N
N
N N O
DACA N
N N
N-(2-(dimethylamino)ethyl)
Zaleplon CN O
acridine-4-carboxamide
Carbazeran
F HN NH2 NH2
N
N O
N
O N H2N N OH
H N N HN
5-fluoro-2-pyrimidone N N
Hydralazine O O
HO
Methotrexate
HN
O NH2 O
NH
HN N
N
N H2N N
N
Acyclovir
N
OH
Zoniporide
Figure 7.13 Aldehyde oxidase mediated oxidation of various drugs. Arrows indicate
the site of oxidation.
O NH2
NH2 NH2
O N N
N N OH N NH
OH
N Deacetylation AO
N N O
N
N N
O Famciclovir OH Deoxypenciclovir OH Penciclovir
O
NH2 NH2
O N O N
N N N N
O N O O N O
Prazosin O O
Terazosin
O OH OH
N NH O
O AO O O
S NH2 S NH2 S NH2
O O O
Zonisamide Ketimine Hydroxyphenyl Metabolite
F 3C F F3C F
H N H N
N N
N N
N N N N
O AO N N
O
H2N H2N
Razaxaban
N O NH OH
O O
HN HN
N AO N
N N
N S S-Methyl transferase NH S
Cl Cl CH3
Ziprasidone
O O N S N
O N N
N N N
CH3 CH3 CH3
Arecoline
H
N N N N N
N
N Cl N
N NH O
N N OH
Losartan Irbesartan
7.3.4 Excretion
As mentioned earlier, introduction of polar heterocyclic rings systems in
molecules can influence the routes of clearance. Substitution with appropriate
polar heterocyclic rings during drug design that allows the drug to get absorbed
and confers metabolic stability to the molecule can sometimes favour the drug
to be cleared renally.
Fluconazole (Diflucan) (Table 7.14) is a classic example in which introduc-
tion of heterocyclic rings has influenced the route of elimination of this drug.
Fluconazole belongs to a triazole class of antifungal agents used in the treat-
ment and prevention of superficial and systemic fungal infections. As described
before, triazoles are ring systems that impart significant polarity to a molecule
and make them resistant to metabolism. However, these rings are unionised at
physiological pH and therefore allow complete absorption of the drug in the
Influence of Heteroaromatic Rings on ADME Properties of Drugs 355
Table 7.14 Physicochemical properties of fluconazole, voriconazole, itraco-
nazole and posaconazole
N
N N
N N N
F
O O
N
N N Cl N
O O N N N
F HO Cl N
Fluconazole
Itraconazole
N
N N
N
OH F N N
O O
OH
F F N
N O O N N N
F F N
N Posaconazole
Voriconazole
Compound MW clogD7.4 % Excreted in urine
N F
N
F N N NH2 F
F O
F
Sitagliptin
MW 407.3
clogD7.4 1.93
clogP 2.06
substrate properties of compounds for some efflux transporters can also lead to
renal excretion of lipophilic drugs. Sitagliptin (Januvia) (Figure 7.19), an orally
active triazolopyrazine derivative and an inhibitor of the dipeptidyl peptidase-4
(DPP-4) enzyme, presents an interesting case.65
Despite modest lipophilicity (clogD7.4 ¼ 1.93) and molecular weight
(MW ¼ 407), approximately 79% of sitagliptin is excreted unchanged renally
and metabolism is a minor pathway of elimination.65 Following a [[14]C]sita-
gliptin oral dose, only 16% of the radioactivity was excreted as metabolites of
sitagliptin. Metabolic stability is achieved by the presence of triazolopyrazine
and trifluorophenyl group in the molecule. Reports indicate that sitagliptin is a
low affinity substrate of human organic anion transporter-3 (hOAT-3).66
Organic anion transport systems such as human organic anion transporters
(hOAT1 and 3) are predominantly expressed in the kidney and play an
important role in the transport of organic anions across the basolateral
membrane of human proximal tubules.67 Hence, clearance of sitagliptin pri-
marily involves its active tubular secretion from the body into the urine.
Like the kidney, many transporters are expressed on the canalicular mem-
brane of the liver. These transporters can mediate the excretion of selected
compounds into the bile. Lipophilic compounds and high molecular weight
compounds with a considerable hydrogen bonding functionality and which
show poor permeability can be potentially cleared via this active efflux process.
Incorporation of heterocyclic rings in molecules can alter the route of clearance
for high molecular weight lipophilic compounds from metabolism to active
efflux (and therefore biliary excretion of unchanged drug). The change is
dependent upon the hydrogen bonding capability that is incorporated into the
molecule by the newly introduced heterocyclic ring system. Gardner and co-
workers have demonstrated this for inhibitors of thromboxane A2 synthase
(7.20 and 7.21, Table 7.15).68 Both the compounds showed high hepatic
extraction ratio (E ¼ 0.9) in the isolated perfused rat liver. Comparison of their
physicochemical properties showed that both compounds had molecular
weight close to 500 (Table 7.15). Compound 7.20 was mainly eliminated from
the body by metabolism whereas 7.21 was excreted into the bile unchanged.
This suggested that 7.21 was a substrate for one of the hepatic efflux
Influence of Heteroaromatic Rings on ADME Properties of Drugs 357
Table 7.15 Effect of replacement of heterocyclic ring systems in a compound
on the route of elimination.68
Compound Structure PSA clogD7.4 clogP HBD HBA
N
HO2C
CH3
N
F
N
N
HO2C
CH3
N
transporters. The change in the route of clearance for 7.21 was attributed to its
high polarity (negative clogD7.4) relative to 7.20 and increase in the number of
hydrogen bond donors and acceptors in the molecule. It should be noted that
the above inference is merely speculative since the influence of the physico-
chemical properties on the SAR of efflux transporters is not fully understood.
7.23 1690
F
7.24 911
O N
Cl
7.25 303
O N
CF3
7.26 88
O N
Cl Cl Cl
S O
S OH N
O O
O
O S
OH
Tienilic Acid (-)-Suprofen Ticlopidine
O
O
N N
HO NH H
O S
N
N O N
S H
Ritonavir
Nu O
S R S R H O
Nu S R
O O
H+ OH
Nu S R Nu S R
Cl Cl
O O
R= O O
OH
and furans commonly fall into this category.46,71 Several therapeutic agents with
these rings have been withdrawn from the market or have warnings on their
labels due to specific organ toxicities and related idiosyncratic reactions, and are
inactivators of drug metabolising enzymes. Tienilic acid, suprofen and ticlopi-
dine (Figure 7.20) are examples of thiophene containing compounds that have
been associated with adverse drug reactions as a consequence of metabolic
activation.
Tienilic acid was withdrawn from the market because of hepatotoxicity. It
has been proposed that tienilic acid undergoes sulfoxidation to a reactive
electrophilic, tienilic acid-S-oxide by CYP2C9, which inactivates the enzyme
(Figure 7.21).72
Tienilic acid specific autoantibodies (called anti-LKM2) directed against
CYP2C9 have been detected in patients treated with tienilic acid and suffering
360 Chapter 7
with tienilic acid-induced hepatitis. More recently, the structural analog ()-
suprofen was shown to also be a mechanism-based inactivator of P450 2C9.73
Tienilic acid and ()-suprofen are reported to cause mechanism-based inacti-
vation by a similar mechanism, leading to covalent modification of the
CYP2C9 apoprotein within the active site.74 Although ticlopidine is still utilised
clinically as an inhibitor of ADP-induced platelet aggregation, its use is asso-
ciated with a relatively high incidence of agranulocytosis, aplastic anaemia and
thrombocytopenia. The thiophene ring in ticlopidine is known to undergo
CYP2C19 and 2B6 mediated catalysis to reactive intermediate(s) that causes
enzyme inactivation.72,75 Other examples of sulfur containing heterocycles
include the potent mechanism-based inactivator of CYP3A4 by the protease
inhibitor, ritonavir (Figure 7.20).72 Ritonavir has been speculated to undergo
bioactivation of one or both of its thiazole rings and therefore result in inac-
tivation of the enzyme.
Furan-containing compounds such as furosemide, ipomeanol and L-739010
(Figure 7.22) also cause hepatic and renal necrosis in mouse and humans or
develop potentially lethal pulmonary lesions in rat.46 Oxidative ring opening of
furan and irreversible protein binding of the corresponding substituted ketoe-
nals has been reported during metabolism studies on many furan-containing
biologically active compounds (Figure 7.23).76–79
Several compounds containing furan rings are inactivators of P450
enzymes.76,77 Evidence for P450 inactivation by reactive intermediate(s) derived
from furan ring scission has also been presented for the experimental HIV
Cl O O
S NH2 O CN
OH N
O O
N
H O O
O CO2H 4-Ipomeanol
L-739010
Furosemide O
R R R
O
O O O OH
H
Reactive
Intermediates
Covalent Binding R
To Macromolecules
O
O
N OH O
H OH N
O N N N P450
N O N
O Inactivation
O NH O NH
L-745,394
N OH OH
H
O N N
N
O
O NH
Compound 7.27
protease inhibitor L-745 394 which inactivates rat CYP2C11 and human
CYP3A4 enzymes (Figure 7.24).80,81 Removal of the furan ring or reduction of
2,3-double bond in L-754,394 led to compound 7.27 (Figure 7.24) which did
not inactivate CYP3A4, implicating that the furan ring was involved in the
bioactivation sequence leading to enzyme inactivation.82 For this reason furan,
thiophenes and thiazoles (especially the amino thiazole ring) are considered as
structural alerts and are generally excluded by the medicinal chemists when
considering the design of new drugs candidates.
Some six-membered heterocycles with substituted alkyl, amino or hydroxyl
substituents can also undergo metabolic activation. For instance, 2,3-diami-
nopyridine containing compounds (7.28) (Figure 7.25) that were designed as
bradykinin B1 receptor antagonists undergo CYP3A4-catalysed bioactivation
and covalent binding to liver microsomal proteins and glutathione.83 The
presence of two amino groups on the pyridine ring increases its susceptibility to
P450-mediated two-electron oxidation and consequent formation of the reac-
tive pyridine-2,3-diimine 7.29 (Figure 7.25). Reaction of this intermediate with
glutathione afforded the glutathione adduct 7.30. The bioactivation liability of
the 2,3-diaminopyridine moiety was addressed by replacement of the 2-amino
group on the pyridine nucleus with an oxygen atom (7.31) or addition of a
methyl group on the 2-aminopyridine (7.32) (Figure 7.25).
Compounds possessing a sterically unhindered nitrogen heterocycles are
known to act as reversible inhibitors of P450 enzymes. These compounds have
been shown to coordinate with the heme iron inside the CYP catalytic pocket.
Known as type-II ligands, these compounds can inhibit CYP by displacing a
sixth (weak) ligand, water, and stabilising the iron in its low spin state. This
spin state change is accompanied by an increase in the redox potential of P450,
which makes the P450 reduction (by NADPH P450 reductase) more difficult.
362 Chapter 7
Cl H
N
CN Cl Cl H
O N GS N
N NH CYP450 CN CN
GSH
O O
CO2CH3 N N N NH
F
Ar Ar
7.28 7.29 7.30
Cl CN
H
N
Cl O
H N NCH3
N
CN
O Ar
N O
7.32
Ar
7.31
N
N N
N Cl N N
N
O
Cl O
O O N N
Cl NC CN
Ketoconazole Anastozole
Clotrimazole
Figure 7.26 Structures of additional azole containing compounds that can inhibit
CYP450.
X
N OH OH
H
N N
O
O NH
Indinavir II 0.45
N
7.33 II 0.87
N
7.34 I 15.1
N
7.35 I 8.7
N
catalytic activities.85 The strength of the bond between their heteroatomic lone
pair electrons and the prosthetic heme iron governs the inhibitory character of
these compounds in addition to hydrophobicity.
Minor structural modifications dramatically change the CYP3A4 potency of
these compounds as illustrated by some analogs of indinavir (Table 7.17).86
Although addition of a gem-dimethyl group (7.33) did not change the
inhibitory characteristics of indinavir, incorporation of a methyl group on the
carbon atom adjacent to nitrogen atom of the pyridine ring (7.33 to compounds
7.34 or 7.35) dramatically decreased CYP3A4 inhibitory potency by 410-fold
relative to indinavir (Table 7.17).86 The decrease in enzyme inhibition
was presumably due to the change in the interaction of substrate and P450
active site.
In another example, Smith and co-workers have conducted SAR relationship
studies focused on bioisosteric replacements of 2-pyridyl group (7.36) mGlu5
receptor antagonists to reduce inhibition of CYP1A2 (Table 7.18).87 Compared
to 7.36, which inhibited CYP1A2 with an IC50 of 3.8 mM, the thiazolyl deri-
vative 7.37 showed an undesired increase in CYP1A2 inhibition. On the other
hand and more importantly, CYP1A2 inhibition was greatly reduced (IC50
414 mM) with 2-imidazolyl derivative 7.38 compared with 7.36. The change in
the potency correlated with the log P value of the three compounds. Thus, the
364 Chapter 7
Table 7.18 Influence of clogP on the inhibition of CYP1A2 by mGlu5
receptor antagonists.87
Compound Structure CYP1A2 IC50 (mM) clogP
CN
7.36 N N 3.8 2.16
N
N N
S
CN
7.37 N N 1 2.31
N
N N
NH
CN
7.38 N N 414 1.63
N
N N
least lipophilic compound with an imidazole ring was the least potent while the
incorporation of a thiazole ring increased the lipophilicity of the compound and
hence its affinity for P450.
Many N-substituted azoles also have the ability to induce hepatic micro-
somal mixed-function oxidases. Some N-substituted imidazoles such as clo-
trimazole have been characterised as high magnitude inducers of rat hepatic
CYP. Although the number of clinically used drugs which induce P450 enzymes
in humans is limited, studies with nilotinib (Table 7.8) have shown that this
drug may induce CYP2B6, 2C8 and 2C9 in humans and thereby decrease the
concentrations of drugs eliminated by these enzymes.
It is now well recognised that that adverse events (either idiosyncratic toxicity
or induction) caused by drugs correlates well with the dose administered.
Reducing the dose size either by increasing the potency of the compound or by
improving its pharmacokinetic properties can potentially reduce the toxic
events of a compound. Thus the impact of the administered dose should not be
underestimated.
A classic example of the impact of dose on the toxicity of compounds can be
observed with glitazones. As described earlier, troglitazone, which is adminis-
tered at a relatively high dose was withdrawn from the market due to hepa-
totoxicity. The toxicity was attributed to the ring opening of thiazolidinone ring
in the molecule. Furthermore, this drug was associated with induction of
CYP3A4. In contrast rosiglitazone, which has the same thiazolidinone func-
tionality and is administered at a much lower dose, is devoid of hepatotoxicity;
neither does it show evidence of enzyme induction.
7.5 Summary
This chapter illustrates the influence of heteroaromatic rings on the ADME
properties of a compound. The examples reviewed here demonstrate that
Influence of Heteroaromatic Rings on ADME Properties of Drugs 365
incorporation of heteroaromatic rings in a compound can have major effect on
its disposition in the body. However, it is important to note that such a cor-
relation is not straightforward. The disposition of a compound can be quite
complex and is affected not only by addition of heterocyclic rings but other
groups and substituents on the molecule. Further, other physiological factors
(e.g. protein binding and transporters) can also affect the ADME properties of
compounds. Nevertheless, it is hoped that this overview provides a flavour
of the impact that various aromatic heterocycles have on the ADME properties
of the compound.
References
1. H. B. Broughton and I. A. Watson, J. Mol. Graphics Modelling, 2005, 23,
51.
2. S.-W. Zhao, L. Liu, Y. Fu and Q.-X. Guo, J. Phys. Org. Chem., 2005, 18,
353.
3. J. B. Sperry and D. L. Wright, Curr. Opin. Drug Discov. Dev., 2005, 8,
723.
4. R. B. Silverman, in The Organic Chemistry of Drug Design and Drug
Action, Academic Press, San Diego CA, 1992, pp. 88–95.
5. P. P. A. Humphrey, Headache, 2008, 48, 685.
6. P. Cohen, Curr. Opin. Chem. Biol., 1999, 3, 459.
7. M. E. M. Noble, J. A. Endicott and L. N. Johnson, Science, 2004, 303,
1800.
8. S. P. Davies, H. Reddy, M. Caivano and P. Cohen, Biochem. J., 2000, 351,
95.
9. Y.-Z. Shu, J. Nat. Prod., 1998, 61, 1053.
10. L. M. Lima and E. J. Barreiro, Curr. Med. Chem., 2005, 12, 23.
11. P. H. Olesen, Curr. Opin. Drug Discov. Dev., 2001, 4, 471.
12. B. C. C. Cantello, M. A. Cawthome, D. Haigh, R. M. Hindley, S. A. Smith
and P. L. Thurlby, Bio. Med. Chem. Lett., 1994, 4, 1181.
13. J. A. Joule and K. Mills, in Heterocylic Chemistry, Blackwell Publishing,
Oxford, 2000, pp. 1–15.
14. T. Eicher, S. Hauptmann and A. Speicher, in The Chemistry of Hetero-
cycles, Wiley-VCH, Weinheim, 2003, pp. 5–16.
15. T. L. Gilchrist, Heterocyclic Chemistry, Longman Scientific & Technical,
Harlow, 1985, pp. 5–30.
16. D. A. Smith, B. C. Jones and D. K. Walker, Med. Res. Rev., 1996, 16,
243.
17. O. H. Chan and B. H. Stewart, Drug Discov. Today, 1996, 1, 461.
18. I. K. Khanna, Y. Yu, R. M. Huff, R. M. Weier, X. Xu, F. J. Koszyk, P. W.
Collins, J. N. Cogburn, P. C. Isakson, C. M. Koboldt, J. L. Masferrer,
W. E. Perkins, K. Seibert, A. W. Veenhuizen, J. Yuan, D.-C. Yang and
Y. Y. Zhang, J. Med. Chem., 2000, 43, 3168.
366 Chapter 7
19. P. J. Cox, D. A. Ryan, F. J. Hollis, A.-M. Harris, A. K. Miller,
M. Vousden and H. Cowley, Drug Metab. and Dispos., 2000, 28, 772.
20. C.-M. Loi, M. Young, E. Randinitis, A. Vassos and J. R. Koup, Clin.
Pharmacokinet., 1999, 37, 91.
21. M. L. Quan, P. Y. S. Lam, Q. Han, D. J. P. Pinto, M. Y. He, R. Li, C. D.
Ellis, C. G. Clark, C. A. Teleha, J.-H. Sun, R. S. Alexander, S. Bai, J. M.
Luettgen, R. M. Knabb, P. C. Wong and R. R. Wexler, J. Med. Chem.,
2005, 48, 1729.
22. B. Peng, C. Dutreix, G. Mehring, M. J. Hayes, M. Ben-Am, M. Seiberling,
R. Pokomy, R. Capdeville and P. Lloyd, J. Clin. Pharmacol., 2004, 44, 158.
23. Tasigna (Nilotinib) product label.
24. R. J. Herr, Bioorg. Med. Chem., 2000, 10, 3379.
25. M.-W. Lo, M. R. Goldberg, J. B. McCrea, H. Lu, C. I. Furtek and T. D.
Bjornsson, Clin. Pharmacol. Ther., 1995, 58, 641.
26. P. J. Cox, B. D. Bush, P. D. Gorycki, G. Y. Kuo, D. Kenworthy, D. W.
Law, D. Murphy, P. C. Shardlow, A. Taylor, J. W. Upward, C. H.
Compton and R. D. Murdoch, Exp. Tox. Pathol., 1996, 48, 75.
27. D. Teneroa, D. Martinb, B. Ilsonb, J. Jushchyshync, S. Boikeb, D.
Lundberga, N. Zariffab, D. Boylea and D. Jorkaskyb, Biopharm.
Drug Dispos., 1998, 19, 351.
28. C. Liljebris, S. D. Larsen, D. Ogg, B. J. Palazuk and J. E. Bleasdale,
J. Med. Chem., 2002, 45, 1785.
29. J. Kelder, P. D. J. Grootenhuis, D. M. Bayada, L. P. C. Delbressine and
J.-P. Ploemen, Pharm. Res., 1999, 16, 1514.
30. K. Palm, P. Sternberg, K. Luthman and P. Artursson, Pharm. Res., 1997,
14, 568.
31. K. M. Mahar Doan, J. E. Humphreys, L. O. Webster, S. A. Wring, L. J.
Shampine, C. J. Serabjit-Singh, K. K. Adkison and J. W. Polli, J. Phar-
macol. Exp. Ther., 2002, 303, 1029.
32. S. A. Hitchcock and L. D. Pennington, J. Med. Chem., 2006, 49, 7559.
33. R. Tatsumi, M. Fujio, H. Satoh, J. Katayama, S.-I. Takanashi,
K. Hashimoto and H. Tanaka, J. Med. Chem., 2005, 48, 2678.
34. R. A. Porter, W. N. Chan, S. Coulton, A. Johns, M. S. Hadley, K. Widdowson,
J. J. C. Jerman, S. J. Brough, M. Coldwell, D. Smart, F. Jewitt, P. Jeffrey and
N. Austin, Bioorg. Med. Chem. Lett., 2001, 11, 1907.
35. J. L. Castro, R. G. Ball, H. B. Broughton, M. G. N. Russell, D. Rathbone,
A. P. Watt, R. Baker, K. L. Chapman, A. E. Fletcher, S. Patel, A. J.
Smith, G. R. Marshall, W. Ryecroft and V. G. Matassa, J. Med. Chem.,
1996, 39, 842.
36. A. B. Pinkerton, R. V. Cube, J. H. Hutchinson, J. K. James, M. F.
Gardner, H. Schaffhauser, B. A. Rowe, L. P. Daggett and J.-M. Vernier,
Bioorg. Med. Chem. Lett., 2004, 14, 5867.
37. N. Murayama, N. Imai, T. Nakane, M. Shimizu and H. Yamazaki, Bio-
chem. Pharmacol., 2007, 73, 2020.
38. R. Hyland, B. C. Jones and D. A. Smith, Drug Metab. Dispos., 2003, 31,
540.
Influence of Heteroaromatic Rings on ADME Properties of Drugs 367
39. S. K. Balani, B. H. Arison, L. Mathai, L. Kauffman, R. R. Miller,
R. A. Steams, I.-W. Chen and J. H. Lin, Drug Metab. Dispos., 1995, 23,
266.
40. M. Chiba, M. Hensleigh and J. H. Lin, Biochem. Pharmacol., 1997, 53,
1187.
41. O. Ghosheh and E. M. Hawes, Drug Metab. Dispos., 2002, 30, 1478.
42. J. Magdalou, R. Herber, R. Bidault and G. Siest, J. Pharmacol. Exp. Ther.,
1992, 260, 1166.
43. S. Kitamura, K. Sugihara and S. Ohta, Drug Metab. Pharmacokinet., 2006,
21, 83.
44. M. R. Rashidi, J. A. Smith, S. E. Clarke and C. Beedham, Drug Metab.
Dispos., 1997, 25, 805.
45. W. W. Hall and T. A. Krenitsky, Arch. Biochem. Biophys., 1986, 251,
36.
46. D. K. Dalvie, A. S. Kalgutkar, S. C. Khojasteh-Bakht, R. S. Obach and J.
P. O’Donnell, Chem. Res. Toxicol., 2002, 15, 269.
47. D. C. Hobbs, T. M. Twomey and R. F. Palmer, J. Clin. Pharmacol., 1978,
18, 402.
48. R. Griffith, in Foye’s Principles of Medicinal Chemistry, ed. D. A. Williams
and T. L. Lemke, Lippincott Williams and Wilkins, Philadelphia, 2002,
pp. 292–312.
49. J. J. Kyncl, R. C. Sonders, W. D. Sperzel, M. Winn and J. H. Seely,
Cardiovasc. Drug Rev., 1986, 4, 1.
50. K. Sugihara, S. Kitamura and K. Tatsumi, Comp. Biochem. Physiol., 1996,
24, 1996.
51. C. Beedham, J. J. Miceli and R. S. Obach, J. Clin. Psychopharmacol., 2003,
23, 229.
52. Z. Miao, A. Kamel and C. Prakash, Drug Metab. Dispos., 2005, 33,
879.
53. D. Zhang, N. Raghavan, S.-Y. Chen, H. Zhang, M. Quan, L. Lecureux,
L. M. Patrone, P. Y. S. Lam, S. J. Bonacorsi, R. M. Knabb, G. L. Skiles
and K. He, Drug Metab. Dispos., 2008, 36, 303.
54. P. Sauerberg, P. H. Olesen, S. Nielsen, S. Treppendahl, M. J. Sheardown,
T. Honor, C. H. Mitch, J. S. Ward, A. J. Pike, F. P. Bymaster, B. D.
Sawyer and H. E. Shannon, J. Med. Chem., 1992, 35, 2274.
55. P. Sauerberg, J. W. Kindtler, L. Nielsen, M. J. Sheardown and T. Honor,
J. Med.Chem., 1991, 34, 687.
56. K. W. Brammer, A. J. Coakley, S. G. Jezequel and M. H. Tarbit, Drug
Metab. Dispos., 1991, 19, 764.
57. R. A. Stearns, R. R. Miller, G. A. Doss, P. K. Chakravarty, A. Rosegay,
G. J. Gatto and G.-H. Chiu, Drug Metab. Dispos., 1992, 20, 281.
58. S. W. Huskey, R. R. Miller and S.-H. Chiu, Drug Metab. Dispos., 1993, 21,
792.
59. T. J. Chando, D. W. Everett, A. D. Kahle, A. M. Starrett, N. Vachhar-
ajani, W. C. Shyu, K. J. Kripalani and R. H. Barbhaiya, Drug Metab.
Dispos., 1998, 26, 408.
368 Chapter 7
60. K. W. Brammer, P. R. Farrow and J. K. Faulkner, Rev. Infect. Dis., 1990,
12(Suppl 3), S318.
61. T. C. Hardin, J. R. Graybill, R. Fetchick, R. Woestenborghs, M. G.
Rinaldi and J. G. Kuhn, Antimicrob. Agents Chemother., 1988, 32, 1310.
62. P. Krieter, B. Flannery, T. Musick, M. Gohdes, M. Martinho and
R. Courtney, Antimicrob. Agents Chemother., 2004, 48, 3543.
63. D. Levêque, Y. Nivoix, F. Jehl and Raoul Herbrecht, Int. J. Antimicrob.
Agents, 2006, 27, 274.
64. S. J. Roffey, S. Cole, P. Comby, D. Gibson, S. G. Jezequel, A. N. R.
Nedderman, D. A. Smith, D. K. Walker and N. Wood, Drug Metab.
Dispos., 2003, 31, 731.
65. N. A. Thornberry and A. E. Weber, Curr. Topics Med. Chem., 2007, 7, 557.
66. X.-Y. Chu, K. Bleasby, J. Yabut, X. Cai, G. H. Chan, M. J. Hafey, S. Xu,
A. J. Bergman, M. P. Braun, D. C. Dean and R. Evers, J. Pharmacol. Exp.
Ther., 2007, 321, 673.
67. W. Lee and R. B. Kim, Annu. Rev. Pharmacol. Toxicol., 2004, 44, 137.
68. I. B. Gardner, D. K. Walker, M. S. Lennard, D. A. Smith and G. T.
Tucker, Xenobiotica, 1995, 25, 185.
69. K. Samuel, W. Yin, R. A. Stearns, Y. S. Tang, A. G. Chaudhary, J. P.
Jewell, T. Lanza, L. S. Lin, W. K. Hagman, D. C. Evans and S. Kumar,
J. Mass Spectrom., 2003, 38, 211.
70. Y.-Z. Shu, B. M. Johnson and T. J. Yang, AAPS J., 2008, 10, 178.
71. A. S. Kalgutkar, I. Gardner, R. S. Obach, C. L. Shaffer, E. Callegari, K. R.
Henne, A. E. Mutlib, D. K. Dalvie, J. S. Lee Y. Nakai, J. P. O’Donnell,
J. Boer and S. P. Harriman, Curr. Drug Metab., 2005, 6, 161.
72. A. S. Kalgutkar, R. S. Obach and T. S. Maurer, Curr. Drug Metab., 2007,
8, 407.
73. J. P. O’Donnell, D. K. Dalvie, A. S. Kalgutkar and R. S. Obach, Drug
Metab. Dispos., 2003, 31, 1369.
74. J. M. Hutzler, L. M. Balogh, M. Zientek, V. Kumar and T. S. Tracy, Drug
Metab. Dispos., 2009, 37, 59.
75. E. Fontana, P. M. Dansette and S. M. Poli, Curr. Drug Metab., 2005, 6, 413.
76. T. M. Alvarez-Diez and J. Zheng, Drug Metab. Dispos., 2004, 32, 1345.
77. T. M. Alvarez-Diez and J. Zheng, Chem. Res. Toxicol., 2004, 17, 150.
78. D. P Williams, D. J Antoine, P. J. Butler, R. Jones, L. Randle, A. Payne,
M. Howard, I. Gardner, J. Blagg and B. K. Park, J. Pharmacol. Exp. Ther.,
2007, 322, 1208.
79. Y. Sahali-Sahly, S. K. Balani, J. H. Lin and T. A. Baillie, Chem. Res.
Toxicol., 1996, 9, 1007.
80. J. H. Lin, I. W. Chen, M. Chiba, J. A. Nishime and F. A. Deluna, Drug
Metab. Dispos., 2000, 28, 460.
81. J. H. Lin, M. Chiba and I. W. Chen, J. Pharmacol. Exp. Ther., 1995, 274,
264.
82. M. Chiba, J. A. Nishime and J. H. Lin, J. Pharmacol. Exp. Ther., 1995,
275, 1527.
83. J. J. Chen and K. Biswas, Prog. Med. Chem., 2008, 46, 173.
Influence of Heteroaromatic Rings on ADME Properties of Drugs 369
84. W. Zhang, Y. Ramamoorthy, T. Kilicarslan, H. Nolte, R. F. Tyndale and
E. M. Sellers, Drug Metab. Dispos., 2002, 30, 314.
85. S. W. Grimm and M. C. Dyroff, Drug Metab. Dispos., 1997, 25, 598.
86. M. Chiba, L. Jin, W. Neway, J. P. Vacca, J. R. Tata, K. Chapman and
J. H. Lin, Drug Metab. Dispos., 2001, 29, 1.
87. N. D. Smith, S. F. Poon, D. Huang, M. Green, C. King, L. Tehrani, J. R.
Roppe, J. Chung, D. P. Chapman, M. Cramera and N. D. P. Cosforda,
Bio. Med. Chem. Lett., 2004, 14, 5481.
CHAPTER 8
8.1 Introduction
The human genome is projected to have 5000–10 000 potential drug targets, of
which it is speculated that 3000–6000 are amenable to small molecule drugs and
another 1500–3000 to biopharmaceuticals.1–3 Currently marketed drugs are
directed towards less than 500 targets encoded by the genome.3,4 Thus, it would
appear that a large pool of potential therapeutic targets remain untapped.
Much of this has to do with a significant lack in detailed understanding of the
biological roles played by most of these genome products. For some, where
roles have been established, their mechanisms of action can involve protein–
protein or peptide–protein interactions for which a detailed understanding of
the epitopes and specifics of their molecular interactions remain unknown. For
a few peptide–protein and protein–protein interactions, such molecular spe-
cifics have been resolved from crystal structures of complexes or approaches
using targeted mutagenesis to identify critical amino acids or peptide sequences
that are involved in an interaction. The interactions are generally based on
complementary structural features at the interaction site and can include
370
Peptidomimetics and Peptides as Drugs 371
hydrogen bonding, polar and hydrophobic effects between proximal amino
acids, and restricted conformations in the bound state. For systems where a
detailed understanding of the interactions is known, peptides or proteins as oral
pharmaceuticals have not been successfully compared with small molecule
pharmaceuticals due to their intrinsic properties which include molecular size,
absorption across biological membranes, and rapid systemic digestion by
proteolytic enzymes.
For over three decades, medicinal chemists have taken a peptidomimetic
approach to overcoming these limitations. Several definitions have been
ascribed to peptidomimetics. Giannis and Kolter5 gave a functional definition
of a peptidomimetic as: ‘a compound that, as the ligand of a receptor, can
imitate or block the biological effect of a peptide at the receptor level’. Wiley
and Rich6 defined peptidomimetics as ‘chemical structures designed to convert
the information contained in peptides into small non-peptide structures’ and
Gante7 defined a peptidomimetic as ‘a substance having a secondary structure
as well as other structural features analogous to that of the original peptide
from receptors or enzymes’. These definitions confine peptidomimetics to
design features of molecules related to optimal binding to a receptor or enzyme.
The progress of research in the area of peptidomimetics as pharmaceutical
molecules encompasses a broader definition to include physicochemical prop-
erties compatible with current understanding of the parameters that constitute
a drug molecule.
Here we define a ‘peptidomimetic’ as any small molecule whose structural
base is derived from the peptide and defined by the minimum number of
interactions that provide binding at a receptor or enzyme active site with equal
or better affinity than the original peptide sequence, effects the same physio-
logical response as the original peptide, and possesses drug-like properties for
absorption and pharmacokinetics consistent with an appropriate therapeutic
index.
In this chapter we have selected a few examples from the literature where drug
design originated from a peptidomimetic approach to yield therapeutic agents,
and examples where peptidomimetics show promise with in vitro systems but
need further testing for drug properties consistent with therapeutic agents. We
examine a few functional groups used as isosteres for the peptide bonds, amino
acid side chains and spatial interaction within binding sites, and finally the drug
properties of the resulting molecules. This chapter includes some topics, tran-
sition state analogues, which are also covered in Chapter 9, although here we
study the absorption, distribution, metabolism and excretion (ADME) prop-
erties from the viewpoint of the peptidomimetic properties of the drugs.
(a) NH2
S HN
O O O
H H H
N N N NH2
N N N
H H H
O O O O
OH
S
H2N O
P3 P2 P1 P1′ P2′ P3′
O − H
O OH H
(b)
N N
N H H
H non-hydrolyzable
hydrolysis
transition state amide bond isostere
(c) HN NH 2
Gly27′ Asp29′
Gly27 NH
Asp29 NH
Asp25 Asp25′ O N HN
O H
CO2- N
H
O
COOH −
OOC
O O
H H H
N N N NH2
N N N
H H H
O O O O
OH O NH
H
N O H2N O Gly48′
Gly48 H H
O
H H
O N N O
Ile50 Ile50′
Figure 8.1 Schematic representations of: (a) the substrate hexapeptide with the arrow
showing the scissile amide bond; (b) the transition state for hydrolysis of
the amide bond by aspartic acid proteases and the non-hydrolysable
methylene-amine amide bond isostere; and (c) the peptidomimetic
MVT101 that has the methylene-amine amide bond isostere between
residues P1 and P1 0 (boxed) bound to the enzyme, showing the flap bound
water and hydrogen bonds from protein residues to the peptide.
-
O HO O HO H
H
N
N
N H
H hydrolysis non-hydrolysable
transition state hydroxyethyl isostere
H H
O N S O N
O O
H
N N HO
N N N N
H H H H
O CONH2 OH OH
Saquanavir H Nelf inavir H
NH2 NH 2
O O O
O O O
S
S O N N O
O N N O H
H O
OH O P
O- Ca2+
O-
Amprenavir Fosamprenavir calcium salt
H NH2
O N O
O O O
O O
O S
O N NH O
O N N H
H H OH
OH
Darunavir
compound x H
N OH OH
H O
N N H
N N O
N N
O H
O NH O OH
N O
H
Indinavir Lopinavir
O O O OH O
H H H
N N N N N O
N N N O S
H H O N N
O OH H H
S N O O
Ritonavir Atazanavit
Figure 8.2 Hydroxyethyl transition state isostere for hydrolysis of an amide bond: ten
approved peptidomimetic HIV-1 protease inhibitors incorporating the
hydroxyethyl isostere.
OH
CF3
O O
HN
S
Tipranavir O O
Figure 8.3 Structure of the approved anti HIV-PR drug, tipranavir, designed by a
non-peptidomimetic strategy.
Peptidomimetics and Peptides as Drugs 377
primarily by CYP 3A4. Fosamprenavir is a prodrug for amprenavir with higher
solubility and is rapidly hydrolysed at the intestinal brush border to release
amprenavir, thus allowing formulation of higher strength tablets. Indinavir
shows high bioavailability and plasma exposure, probably due to the hydro-
xyindan and pyridyl-methylpiperazine motifs which offer greater solubility.
Indinavir is a substrate and inhibitor of CYP 3A4. Ritonavir is a potent CYP
3A4 inhibitor; consequently, its primary use in antiretroviral therapy has
evolved as a low dose co-drug to increase the plasma exposure of co-admi-
nistered antiretrovirals.17,19 Atazanavir has high bioavailability providing high
oral exposure and a longer in-vivo half-life that can support once daily dosing
for treatment. Darunavir is metabolised by CYP 3A4 but is a weak inhibitor of
3A4. It is co-administered with low dose ritonavir to achieve plasma exposure
necessary for therapy. The high affinity of darunavir for the HIV-PR and its
higher solubility than the other anti HIV-PR drugs allows for doses in the range
of 300–400 mg twice daily as opposed to doses of 800–1000 mg twice or thrice
daily for the other protease inhibitors.
As these anti HIV-PR drugs are generally metabolised by CYP3A4, a clear
understanding of drug interactions between co-administered antiretrovirals in
clinical treatment of HIV is critical where multi-drug therapy is considered the
standard of care. For example, darunavir/ritonavir co-administered with
saquinavir resulted in a significant decrease in steady state exposure of dar-
unavir, suggesting that these drugs should not be co-administered.20 The
development of anti HIV-PR antivirals is an excellent example of successful
application of the peptidomimetic approach to the discovery of new therapeutic
agents.
Renin is another example of an aspartic acid protease where peptidomimetic
approaches have been tried for the development of antihypertensives. However,
with renin, the peptidomimetic approach has not met with success comparable
to that with the HIV-PR. Renin cleaves angiotensinogen to the decapeptide
angiotensin I, which is further digested by a non-specific dipeptidyl carbox-
ypeptidase, agiotensin-converting enzyme (ACE), to the octapeptide angio-
tensin II which is a potent vasopressor. Renin and ACE are popular targets for
the development of antihypertensive drugs.
Figure 8.4 shows a commonly used retro inversion of an amide bond as a
strategy in peptidomimetic design to prevent protease digestion. This strategy
was used to design the three peptidomimetic inhibitors of renin shown in
Figure 8.4. Histidine is the only residue that was not changed in the inhibitors,
as this residue provides specific binding and hydrogen bonding interaction with
Ser233. The P4 prolyl residue was replaced with piperidinyl or morpholino
rings to better fit the S4 renin binding site.21 The P3 phenylalanine was replaced
by 2-methyl-naphthyl-succinic acid, with a retro inverso amide to the amino
group of piperidine or morpholine at P4 and the a-amino group of histidine.
The naphthyl aromatic ring was selected to prevent recognition of a Phe–His-
like amide bond by chymotrypsin. The P1 isoleucine was replaced with nor-
statine (3-amino-2-hydroxy propionic acid) having either an isobutyl or
methylcyclohexyl group at P3. Compound 8.3 (Figure 8.4) was found to be
378 Chapter 8
O H
N R2
R2 R1
R1 N
H O
retro inversion
O O O
H H
N N O O O
N N N H
H H N
O O N N O
N H
O OH
N
N
P4 P3 H P2 P1 P1′ 8.1 IC50 = 41 nM
Pro Phe His Ile Val N
H
O O O O O O
H H
N N
N N O N N O
H H
O O OH O O OH
N N
Figure 8.4 Retro inversion of the amide bond used in peptidomimetics to decrease
protease susceptibility of peptidomimetic inhibitors of renin derived from
the angiotensinogen pentapeptide sequence (arrow shows the scissile
amide bond).
O NH
H
HN NH2 N
NH N O
O OH O H
H H O
N N N
H N N N
N H H
HN O O O O
O
N
O
O O
N
CGP29287 H
H
N
N O OH
O OH H
H H H 2N N NH2
N N
S N
O H O O O
O O O
CGP60536 (Aliskiren)
O
CGP38560
HS
O O
H H
N N
H 2N N OH
H
O O
S
P4 P3 P2 P1
Cys Val Ile Met
CVIM IC50 340 nM
O S
HS HN HS HN
HN
H 2N O OH H2 N OH
O O
8.4 IC50 150 nM 8.5
O S
HS HN HS HN
HN
H 2N OH H 2N OH
8.6 IC50 0.6 nM O 8.7 IC50 15 nM O
8.3.1 Summary
As shown by the examples above, peptidomimetics hold promise as another
means of developing novel therapeutics based on a clear understanding of the
biochemical pathway and the mechanisms underlying peptide–protein or pro-
tein–protein interactions. As was shown for the development of HIV-PR
inhibitors, the structures of the final drug products had little in common
structurally with the peptide from which they were derived, and with exception
of the core pharmacophore, have little in common with each other. Further-
more, a potent antiretroviral, tirpanavir, was developed by a traditional
iterative structure–activity–drug properties screening approach from lead
matter identified in high throughput screening, without structural knowledge of
the binding to the active site but a robust assay for activity. At the core of
discovering new chemical matter for therapeutic use is a clear understanding of
the biochemical basis for target selection, clear design strategies for new che-
mical matter with high potency and early evaluation of drug properties for
absorption, pharmacokinetics, and metabolism, which lead to the selection of
optimal structures to progress through the drug discovery pipeline.
382 Chapter 8
2 4 Trp
Trp3
23 Tyr3 3 23
small
hydrophobic O N
H
pocket N
N
H O
O
Gl NH 2 CONH 2
n3
19
08 06
Thr3 y3
Glu314 Gl
P1 P2 P3 P4
Ala Val Pro Leu
O
O O
HN N HN
N N
N
H H
O N O NH
O H
O
8.8 Kd = 12 nM 8.9 Kd = 5 nM
O O
O O
HN HN N
N
N N
H H
O N O N
O H O H
9.0 Kd = 12 nM 9.1 Kd = 12 nM
S S
G IV E QC CT SIC SL YQ LE NY CN
S S
S S
Figure 8.8 Structure of human insulin and the structure modifications to create short-
and long-acting analogs.
384 Chapter 8
subcutaneous tissue with the resultant slow appearance in blood. Structural
characterization of human insulin showed that residues 26–30 (YTPKT) were
not critical for binding of insulin to the insulin receptor, but were critical for
aggregation.47 Inversion of proline–lysine at position 28–29 in the B chain of
human insulin results in an analog (insulin lispro) with comparable insulin
receptor affinity but decreased self association. This in turn results in faster
dissociation of aggregates into monomers and consequently faster absorption
from the subcutaneous deposition site, leading to an earlier and greater peak
serum level for a shorter duration of time compared with unmodified human
insulin. Pharmacokinetics of insulin lispro shows an onset of activity between
0.2 and 0.5 h, with maximal activity between 0.5 and 2 h and a duration of 3–
4 h.48,49 Insulin lispro was the first insulin analog approved for human use in
1996.
Insulin aspart was derived from human insulin by substituting proline 28 in
the B chain with aspartic acid. Like insulin lispro, this analog results in weak
dimeric and hexameric aggregation leading to comparable rapid absorption
into serum after subcutaneous injection. The pharmacokinetic behaviour of
insulin aspart is similar to insulin lispro50 and it was approved for human use
in 2000.
Insulin glulisine is the third approved rapid onset insulin analog approved for
human use. Insulin glulisine was obtained by substituting B chain lysine 29 and
aspargine 3 with glutamic acid and lysine, respectively. The pharmacokinetic
properties of insulin glulisine are comparable with those of the other two fast
acting insulin analogs, with rapid onset of activity within 20 minutes and
maximal activity in 1.5 h.51,52
The pharmacological properties of the insulins lispro and aspart are com-
parable to human insulin in terms of insulin receptor affinity, receptor off rate,
metabolic potency, insulin like growth factor-1 receptor affinity, and mitogenic
potency. Insulin glulisine has somewhat lower affinity for the human insulin
receptor and is significantly weaker in affinity to the insulin-like growth factor-1
receptor and mitogenic potency.53,54 In general, the three fast-acting insulin
analogs are very similar in pharmacokinetic and physiological function and
more closely mimic the normal physiological insulin response following pre-
prandial subcutaneous administration when compared to subcutaneously
administered human insulin.
Long-acting insulin analogs have been developed by shifting the isoelectric
point towards neutral pH and by increasing the hydrophobicity by covalent
modification with a fatty acid. Insulin glargine was developed by replacing
asparagine 21 in the A chain with glycine and adding two arginine residues at
position 30 in the B chain. These amino acid substitutions result in an iso-
electric point shift from pH 5.4 to 6.7.55 This analog is injected subcutaneously
as an acidic solution at pH 4.0 and forms a precipitate. The precipitated insulin
glargine dissolves slowly and is absorbed into the serum without any significant
peak. It has a duration of action of about 20 h at physiological doses and is
significantly longer acting than NPH insulin.56 Insulin glargine (lantus) was the
first long-acting insulin analog to be approved for human use.
Peptidomimetics and Peptides as Drugs 385
Another long-acting insulin analog approved for human use is insulin
determir. This analog is formed by removing threonine 30 in the B chain and
acylating the lysine at position 29 in the B chain with myristic acid. Insulin
determir forms soluble hexameric and di-hexameric complexes at the site of
injection via aggregation of the fatty acid chains. These self-associated com-
plexes are thought to equilibrate with serum albumin. The equilibrium between
self-association and albumin binding are thought to be the reason for the long
depot residence time. The binding to albumin is also thought to slow its dis-
tribution to peripheral tissues resulting in an increased duration of activity.57
Binding of insulin determir to serum albumin does not appear to compete with
other albumin bound compounds or to show kinetic differences in low albumin
states, suggesting that albumin may serve as a reservoir for insulin determir
forming a buffer for any rapid changes in insulin absorption.58,59
Long-acting insulin analogs show comparable reductions in HbA1c to NPH
insulin.60,61 However, these analogs show a significantly lower risks of noc-
turnal hypoglycemia.62–64 Combination therapies with long- and short-acting
insulin analogs, as well as long-acting insulin analogs and oral antidiabetic
drugs are being considered as therapeutic options to achieve the rhythm of
insulin observed in normal humans.
GLP-1 and amylin analogs have become therapeutic agents whereas CCK,
PYY and ghrelin receptors are still investigational targets for therapy.
386 Chapter 8
Most of the incretin effect on pancreatic function is derived from
GLP-1.66 Glucose homeostasis in the postprandial period is regulated by
GLP-1 through several mechanisms that include inhibition of glucagon secre-
tion, increase of insulin synthesis, delay in gastric emptying and promotion of
satiety.67 Type 2 diabetic patients have a decreased incretin effect due to a
decrease in secretion of GLP-1, which results in lowered insulin secretion
and disruption of glucose homeostasis.68 Administration of GLP-1 to type 2
diabetic patients results in an increase in insulin secretion and lowering of
postprandial blood glucose.69 GLP-1 is, however, rapidly degraded by dipep-
tidyl peptidase 4 (DPP-4), a protease that is widely distributed in multiple cell
types including the capillary bed of the gut mucosa and so rapidly inactivates
GLP-1.
Exendin-4 (Bayetta, Exenatide) a 39 amino acid peptide extracted from the
venom of the Gila monster (Heloderma suspectum) with a high affinity for
GLP-1 receptors and has 53% homology with GLP-1. Exenatide is the
first GLP-1 receptor agonist approved for adjunctive therapy for type 2
diabetic patients receiving sulfonyl ureas or methformin treatment but who
have not achieved adequate control. Clinical trials have demonstrated that
exenatide improves glycemic control when combined with methformin and
sulfonylureas. It may be an alternative to insulin glargine for patients who
require non-traditional therapy.
Amylin’s site of action is the area posterma—a hind brain circumventricular
organ with a porous blood–brain barrier. Peripheral administration of amylin
analogs have been shown to reduce food intake in non human primates.70
Pramlintide is a human amylin analog that has been approved for human
therapy co-administered with mealtime insulin for patients who have not
achieved glucose control. Clinical studies have demonstrated reductions in
HbA1c and body weight.71
Most currently used peptide-based drugs are currently administered par-
enterally by subcutaneous or intramuscular injection. This is a major limitation
to the administration of such medications. Efforts to develop inhaled forms of
such drugs have met with limited success. Peptides and peptide-based drugs
show no oral bioavailability due to gastric digestion as well as poor absorption
across the intestinal wall.
Some efforts to develop methods for oral delivery of such drugs including
protein drugs have focused on the dietary uptake mechanism of vitamin B12.72
In this approach, vitamin B12 is either conjugated directly with the protein–
peptide of interest for oral administration or is encapsulated in a vitamin B12
coated dextran nanocapsule. Success has been achieved using this technology
for some proteins and peptides including insulin and erythropoietin.73,74 The
development of such approaches including exploring other carrier-mediated
transport processes across the intestinal wall needs further research and tech-
nological innovations to circumvent the oral gastric barrier when delivering
peptide and protein drugs. Such developments will dramatically enhance
the use and development of peptide-based drug molecules based on protein–
protein or protein–peptide interactions.
Peptidomimetics and Peptides as Drugs 387
References
1. J. Drews, Science, 2000, 287, 1960.
2. A. L. Hopkins and C. R. Groom, Nat. Rev. Drug Discov., 2002, 1, 727.
3. P. Imming, Nat. Rev. Drug Discov., 2007, 5, 821.
4. J. Drews and S. Ryser, Nat. Biotechnol., 1997, 15, 1318.
5. A. Giannis and T. Kolter, Angew. Chem., Int. Ed., 1993, 32, 1244.
6. R. A. Wiley and D. H. Rich, Med. Res. Rev., 1993, 13, 327.
7. J. Gante, Angew. Chem., Int. Ed., 1994, 33, 1699.
8. A. Frankel and J. A. T. Young, Annu. Rev. Biochem., 1998, 67, 1.
9. R. Ishima, D. Freedberg, Y.-X. Wang, J. M. Louis and D. A. Torchia,
Structure, 1997, 7, 1047.
10. G. Lange-Savage, H. Berchtold, A. Liesum, K.-H. Budt, A. Peyman, J.
Knolle, J. Sedlacelk, M. Fabry and R. Hilgenfeld, Eur. J. Biochem., 1997,
249, 912.
11. S. Piana, P. Carloni and M. Paminello, J. Mol. Biol., 2002, 319, 567.
12. M. Miller, J. Schneider, B. K. Sathyanarayana, M. V. Toth, G. R. Mar-
shall, L. Clawson, L. Selk, S. B. H. Kent and A. Slodawer, Science, 1989,
246, 1149.
13. N. A. Roberts, J. A. Martin, D. Kinchington, A. V. Broadhurst, J. C.
Craig, I. B. Duncan, S. A. Galpin, B. K. Handa, J. Kay, A. Krohn, R. W.
Lambert, J. H. Merrett, J. S. Mills, K. E. B. Parkes, S. Redshaw, A. J.
Ritchie, D. L. Taylor, G. J. Thomas and P. J. Machin, Science, 1990, 248,
358.
14. N. M. King, M. Prabu-Jeyabalan, E. A. Nalivaika, P. Wigerinck, M.-P. de
Bethune and C. A. Schiffer, J. Virol., 2004, 78, 12012.
15. D. L. Surleraux, A. Tahri, W. G. Verschueren, G. M. E. Pille, H. A. de
Kock, T. H. M. Jonckers, A. Peeters, S. De Meyer, H. Azijn, R. Pauwels,
M.-P. de Bethune, N. M. King, M. Prabu-Jeyabalan, C. A. Schiffer and P.
Wigerinck, J. Med. Chem, 2005, 48, 1813.
16. S. De Meyer, H. Azijn, D. Surleraux, D. Jochmans, A. Tahri, R. Pauwels,
P. Wigerinck and M.-P. de Bethune, Antimicrob. Agents Chemother., 2005,
49, 2314.
17. M. Rittweger and K. Arastéh, Clin. Pharmacokinet., 2007, 46, 739.
18. S. M. Poppe, D. E. Slade, K.-T. Chong, R. R. Hinshaw, P. J. Pagano, M.
Markowitz, D. D. Ho, H. Mo, R. R. Gorman III, T. J. Dueweke, S.
Thaisrivongs and W. G. Tarpley, Antimicrob. Agents Chemother., 1997, 41,
1058.
19. F. M. Uckun and O. J. D’Cruz, Expert Opin. Ther. Pat., 2006, 16, 1354.
20. V. Sekar, E. Lefebvre, K. Marien, M. De Pauw, T. Vangeneugden and R.
M. W. Hoetelmans, Ther. Drug Monit., 2007, 29, 795.
21. R. Kato, O. Takahashi, Y. Kiso, I. Moriguchi and S. Hirono, Chem.
Pharm. Bull., 1994, 42, 176.
22. K. Iizuka, T. Kamijo, H. Harada, K. Akahane, T. Kubota, H. Umeyama,
T. Ishida and Y. Kiso, J. Med. Chem., 1990, 33, 2707.
23. N. C. Cohen, Chem. Biol. Drug Des., 2007, 70, 557.
388 Chapter 8
24. M. Barbacid, Ann. Rev. Biochem., 1987, 56, 779.
25. R. J. A. Grand and D. Owen, Biochem. J., 1991, 279, 609.
26. Y. Reiss, J. L. Goldstein, M. C. Seabra, P. J. Casey and M. S. Brown, Cell,
1990, 62, 81.
27. C. L. Strickland, W. T. Windsor, R. Syto, L. Wang, R. Bond, Z. Wu, J.
Schwartz, H. V. Le, L. S. Beese and P. C. Weber, Biochemistry, 1998, 37,
16601.
28. Y. Qian, M. A. Blaskovich, M. Saleem, C. M. Seong, S. P. Wathen, A. D.
Hamilton and S. M. Sebti, J. Biol. Chem., 1994, 269, 12410.
29. M. A. Kothare, J. Ohkanda, J. W. Lockman, Y. Qian, M. A. Blaskovich,
S. M. Sebtib and A. D. Hamilton, Tetrahedron, 2000, 56, 9833.
30. P. T. Lansbury, Nat. Rev. Neurosci., 2004, 5, S51.
31. I. Tamm, S. M. Kornblau, H. Segall, S. Krajewski, K. Welsh, S. Kitada, D.
A. Scudiero, G. Tudor, Y. H. Qui, A. Monks, M. Andreeff and J. C. Reed,
Clin. Cancer Res., 2000, 6, 1796.
32. E. N. Shiozaki and Y. Shi, Trends Biochem. Sci., 2004, 29, 486.
33. C. Du, M. Fang, Y. Li and X. Wang, Cell, 2000, 102, 33.
34. A. M. Verhagen, P. G. Ekert, M. Pakusch, J. Silke, L. M. Connolly, G. E.
Reid, R. L. Moritz, R. J. Simpson and D. L. Vauz, Cell, 2000, 102, 43.
35. G. Wu, J. Chai, T. L. Suber, J. Wu, C. Du, X. Wang and Y. Shi, Nature,
2000, 408, 1008.
36. Z. Liu, C. Sun, E. T. Olejniczak, R. P. Meadows, S. F. Betz, T. Oost, J.
Herrmann, J. C. Wu and S. W. Fesik, Nature, 2000, 408, 1004.
37. C. R. Arnt, M. V. Chiorean, M. P. Helderbrant, G. J. Gores and S. H.
Kaufmann, J. Biol. Chem., 2002, 277, 44236.
38. L. Yang, T. Mashima, S. Sato, M. Mochizuki, H. Sakamoto, T. Yamon, T.
Oh-Hara and T. Tsuruo, Cancer Res, 2003, 63, 831.
39. S. Fulda, W. Wick, M. Weller and K. M. Debalin, Nat. Med. (N. Y.),
2002, 8, 808.
40. F. G. Banting, C. H. Best, J. B. Collip, W. R. Campbell and A. A. Fletcher,
Can. Med. Assoc. J., 1922, 12, 141.
41. H. C. Hagedorn, B. N. Jensen, N. B. Krarup and I. Wodstrug, J. Am. Med.
Assoc., 1936, 106.
42. K. Hallas-Moller, Diabetes, 1956, 5, 7.
43. Diabetes Control and Complications Trial Research Group, N. Eng. J.
Med., 1993, 329, 977.
44. UK Prospective Diabetes Study Group, Lancet, 1998; 352, 837.
45. Y. T. Kruszynska, P. D. Home, I. Hanning and K. G. M. M. Alberti,
Diabetologia, 1987, 30, 16.
46. R. H. Roscamp and G. Park, Diabetes Care, 1999, 22, S2B109.
47. Z. Vajo, J. Fawcett and W. Duckworth, Endocr. Rev., 1994, 22, 706.
48. D. C. Howey, R. R. Bowsher, R. L. Brunelle and J. R. Woodworth,
Diabetes, 1994, 43, 396.
49. E. Torlone, C. Fanelli, A. M. Rambotti, G. Kassi, F. Modarelli, A. Di
Vincenzo, L. Epifano, M. Ciofetta, S. Pampanelli and P. Brunetti, Dia-
betologia, 1994, 37, 713.
Peptidomimetics and Peptides as Drugs 389
50. S. R. Mudaliar, F. A. Lindberg, M. Joyce, P. Beerdsen, P. Strange, A. Lin
and R. R. Henry, Diabetes Care, 1999, 22, 1501.
51. S. K. Garg, S. L. Ellis and H. Ulrich, Expert Opin. Pharmacother., 2005, 6,
643.
52. R. H. Becker, A. D. Frick, F. Burger, H. Scholtz and J. H. Potgieter, Exp.
Clin. Endocrinol. Diabetes, 2005, 113, 292.
53. P. Kurtzhals, L. Schaffer, A. Sorensen, C. Kristensen, I. Jonassen, C.
Schmid and T. Trub, Diabetes, 2000, 49, 999.
54. G. Seipke and I. Stammberger, Diabetes, 2005, 54(Suppl. 1), A339.
55. G. B. Bolli and D. R. Owens, Lancet, 2000, 356, 443.
56. T. Heise, S. Bott, K. Rave, A. Dressler, R. Rosskamp and L. Heinemann,
Diabetes Med., 2002, 19, 490.
57. S. Havelund, A. Plum, U. Ribel, I. Jonassen, A. Volund, J. Markussen and
P. Kurtzhals, Pharm. Res., 2004, 21, 1498.
58. P. Kurtzhals, S. Havelund, I. Jonassen and J. Markussen, J. Pharm. Sci.,
1997, 86, 1365.
59. P. Kurtzhals, Int. J. Obes., 2004, 28(Suppl. 2), S23.
60. K. Hermansen, M. Davies, T. Derezinski, R. G. Martinez, P. Clauson and
P. Home, Diabetes Care, 2006, 29, 1269.
61. T. Haak, A. Tiengo, E. Draeger, M. Suntum and W. Waldhausl, Diabetes
Obes. Metab., 2005, 7, 56.
62. S. R. Heller, S. A. Amiel and P. Mansell, Diabetes Care, 1999, 22, 1607.
63. S. R. Heller, S. Colagiuri, S. Vaaler, B. H. R. Wolffenbuttel,
K. Koelendorf, H. H. Friberg, K. Windfeld and A. Lindholm, Diabet.
Med., 2004, 21, 769.
64. D. Russell-Jones, H. Kim, S. Heller and P. Clauson, Diabetologia, 2005, 48,
A92.
65. M. A. Nauck, E. Homberger, E. G. Siegel, R. C. Allen, R. P. Eaton, R.
Ebert and W. Creutzfeldt, J. Clin. Endocrinol. Metab., 1986, 63, 492.
66. A. Barnett, Int. J. Clin. Pract., 2006, 60, 1454.
67. D. J. Drucker and M. A. Nauck, Lancet, 2006, 368, 1696.
68. I. Idris and R. Donnelly, Diabet. Obes. Metab., 2007, 9, 153.
69. C. F. Deacon, Diabetes, 2004, 53, 2181.
70. N. T. Bello, M. H. Kemm and T. H. Moran, Am. J. Physiol. Regul. Integr.
Comp. Physiol., 2008, 295, R76.
71. D. Singh-Franco, G. Robles and D. Gazze, Clin. Ther., 2007, 29, 535.
72. A. K. Petrus, T. J. Fairchild and R. P. Doyle, Angew. Chem., Int. Ed., 2009,
48, 1022.
73. G. J. Russell-Jones, S. W. Westwood and A. D. Habberfield, Bioconjug.
Chem., 1995, 6, 459.
74. K. B. Chalasani, G. J. Russell-Jones, A. K. Jain, P. V. Diwan and S. K.
Jain, J. Control. Release, 2007, 122, 141.
CHAPTER 9
Pharmacokinetics and
Metabolism of Compounds that
Mimic Enzyme Transition States
IAIN GARDNER,a CHRIS BARBER,b MARTIN
HOWARD,d AARTI SAWANTc AND KENNY WATSONa
a
Pharmacokinetics, Dynamics and Metabolism, Pfizer Global Research and
Development, Sandwich, Kent, CT13 9NJ, UK; b Worldwide Medicinal
Chemistry, Pfizer Global Research and Development, Sandwich, Kent, CT13
9NJ, UK; c Pharmacokinetics, Dynamics and Metabolism, Pfizer Global
Research and Development, Eastern Point Road, Groton, CT 06340, USA;
d
Drug Metabolism and Pharmacokinetics Discovery Research,
Pharmaceuticals Division F. Hoffmann-La Roche LTD, CH-4070 Basel,
Switzerland
390
Pharmacokinetics and Metabolism of Compounds that Mimic Enzyme 391
ES*
Free
Energy
(ΔG°)
E+S
ES
E+P
EP
Figure 9.1 Graphical representation of energy changes along the reaction coordinate
as enzyme (E) binds to substrate (S) forming an enzyme-substrate complex
(ES), followed by formation of a high energy transition state (ES*) with
eventual formation of an enzyme–product complex (EP) and release of
product from the enzyme (E+P).
are equal (Figure 9.1). In reality, enzymatic transition states are dynamic with
lifetimes (a fraction of a picosecond) defying direct physical observation.2
The catalytic power of enzymes can be viewed to lie in the very high affinity
that enzymes have for the transition state relative to the affinity with which they
bind substrate in the ground state.3 As an example, calf intestinal adenosine
deaminase has a dissociation constant for the transition state approximately
1012 times smaller than it has for the substrate adenosine.
Transition state analogue inhibitors (TSAIs) are a class of competitive enzyme
inhibitors that are designed to take advantage of this very high affinity interaction
between an enzyme and the transition state of the reaction it catalyses. By
mimicking the chemically unstable transition state, whilst being chemically stable
themselves, TSAIs can bind to the enzyme with much higher affinity than sub-
strate.4–8 Although in most cases substrates and transition state analogues bind in
a similar fashion, the transition state analogue can not be converted to a product
and as a result is a potent enzyme inhibitor. The theoretical potency of a perfect
transition state analogue inhibitor can be calculated as shown in eqn (9.1):6
(1) Deriving the free energy relationship for the interaction between inhi-
bitor and enzyme
(2) Kinetic isotope experiments where individual atoms close to the bonds
being broken are replaced with isotopes (e.g. H is replaced with D or T,
and 13C replaces 12C and 15N replaces 14N) and the rate of catalysis
determined
(3) X-ray structural analysis
(4) NMR.
The potential to inhibit enzymes with high affinity and specificity using
transition state analogue inhibitor approaches has attracted much attention
and TSAIs have been described for all major classes of enzymes and for
more than 130 individual enzymes.7,15 However, the number of TSAIs that
have been successfully tested in humans as potential therapeutic agents is
relatively small.
In this chapter we discuss the impact of the structural features needed to
mimic the transition state of an enzymatic reaction on the disposition of drugs
in the body and some of the strategies used to successfully develop TSAIs as
therapeutic agents. In considering the effects of chemical structure and physi-
cochemical properties on oral bioavailability (F), we have used eqns (9.2) and
(9.3) to separate the effects on absorption and first pass clearance/metabolism:
F ¼ Fa Fg Fh ð9:2Þ
where:
F ¼ bioavailability
Fa ¼ fraction absorbed
Fg ¼ fraction escaping gastrointestinal tract first-pass metabolism
Fh ¼ fraction escaping first-pass clearance in the liver.
where:
Figure 9.2 Molecular weight vs. polar surface area (PSA) for human drugs (triangles)
and the transition state analogues discussed in this chapter (squares).
0.06
0.05
(HBA+HBD)/mw
0.04
0.03
0.02
0.01
Figure 9.3 Sum of hydrogen bond acceptors (HBA) and donors (HBD) divided by
molecular weight vs. PSA for human drugs (white triangles) and TSAI
discussed in this chapter (black squares).
600
500
400
PSA
300
200
100
0
-8 -6 -4 -2 0 2 4 6 8
CLogP
Figure 9.4 Polar surface area (PSA) vs. clogP for human drugs (white triangles) and
TSAI discussed in this chapter (black squares).
396 Chapter 9
20
10
8
6
4
Vss (I/kg)
1
0.8
0.6
0.4
0.2
-6 -4 -2 0 2 4 6 8
clogP
Figure 9.5 Vss (L kg1) of transition state analogue inhibitors vs. clogP. Triangles
represent human data and squares rat data. Basic compounds are light
grey, neutral compounds (basic pKa o5; acid pKa 48) are dark grey and
acidic compounds are coloured black.
clogP values, whilst the basic and neutral TSAIs tend to have larger Vss than
acidic compounds. The basic compounds show a trend for increased Vss with
an increase in clogP whilst the trend is not as obvious for neutral compounds.
The other thing to note from Figure 9.5 is that there are relatively few acidic
TSAI compounds.
The widely differing physicochemical properties of transition state analogues
present vastly different challenges to the drug discovery scientist depending on
the physicochemical properties of the TSAI needed to potently inhibit the
enzyme target. These are illustrated in the next section where specific enzyme
targets that have been successfully targeted by TSAI therapeutics are
considered.
Figure 9.6 The first step in the catalytic mechanism is attack of the active site serine
onto the substrate peptide’s scissile bond, forming a tetrahedral inter-
mediate which is stabilised by further interactions by the protease prior to
hydrolysis to release amine and acid products.
Transition state
Classification Drug precursor
intermediate
R R
alcohol R′ R′
OH OH
Enz
O
aldehyde R′ R′
OH O
Enz
O
R R
activated R′ R′ R = CF3, het, carbonyl,
ketone O phosphonate
OH
Enz
O OH
boronate R′ B OH R′ B
OH OH
OH OH
phosphonic P R
acid R′ P R
O R′
O
high lipophilicity together with the high PSA (average 174 Å2) and high
hydrogen bond acceptor (HBA) and donor (HBD) count has a large impact on
the pharmacokinetics of the compounds. So whilst these TSAIs are highly
potent inhibitors of their enzyme target (e.g. the renin inhibitors discussed have
Ki in the 0.7–14 nM range and the HIV protease inhibitors have Ki’s of
o10 nM),20 achieving potency and oral bioavailability in the same molecule is
often challenging.
The pharmacokinetic behaviour of the peptidic renin inhibitors such as
remikiren is fairly typical for compounds with these physicochemical proper-
ties. Although oral administration of the peptidic renin inhibitors to pre-clinical
species often resulted in pharmacological efficacy, oral F% was usually low
requiring high doses to be used, plasma levels were often highly variable (for
instance the plasma levels of remikiren varied up to 100-fold in rats given the
same dose) and the concentration time profiles often showed secondary max-
ima peaks or were an irregular shape. In addition, there is evidence that the
renin inhibitor compounds could inhibit their own metabolism after oral
Pharmacokinetics and Metabolism of Compounds that Mimic Enzyme 399
R N NH2
R N
R1 R1 1′
O -O OH R
H O H
O H H O Asp O HO Asp +
Asp Asp
HO R
O O O O
O
OH
R R1 O
R R
R1 R R1
OH O OH R1
OH O O
R2
Figure 9.8 Proposed catalytic mechanism and transition state for aspartyl proteases
where the active site aspartic acids and substrate (RCONR1) are shown,
together with alcohol and enol groups that have been used to inhibit
aspartyl proteases.
Amprenavir HIV 506 3.3 140 Neutral 6–10 h 90–96 h B11–20 h (CL/F) 490 h
Refs. 20, 33, 270 R protease 3 (V/F)
N
O 4
O N N
O S
O O
O
OH
R = H amprenavir R= P O fosamprenavir
OH
Atazanavir O HIV 705 5.2 171 Neutral 86h 6–8 h (CL/F) 480 h
H
Refs 26, 271 O N protease 13 (4.7 B)
N
5
O N N
O HO
NH
O N
O
Indinavir Chiral HIV 614 3.7 118 Weak 2.2 r 70 r* 107 r 16–24 r 81 r
Refs. 31, 52, protease 7 base 0.7 d 70 d* 16 d 16 d (pH 6.5) 100 d
246, 263, N N O 4 (5.2) 1.5 mk 70 m* 36 mk 72 d (pH 2.5) 100 mk
273–275 N O 0.8 h 64h 10–18 h 19 mk 480 h
N
O
60–65 h
N O
Chapter 9
KNI-272 Chiral HIV 668 4.7 201 Weak base 5–21 r 98–99 h 18–48 r 12 hp 32–100 h
Refs. 276–278 protease 9 (4.9) 0.9 mk 42 mk 16–55 h
N
4 0.11 hp 11 hp
10–11 h
O
O N
O
S N
O
O O
N
N
S
Lopinavir HIV 629 6.1 120 Neutral 98–99 h 1.4–1.7 h 25 r 470 h
Ref 30 protease 5 (CL/F in
O O 4 presence of
N N N ritonavir)
O N
O O
Nelfinavir ABS HIV 568 5.8 127 Basic(7.5) 3–7 r 98–99 h 27–60 r 30–43 r 100 r
O
Refs 82, 279, H protease 6 1.7 – 4 d 22–60 d 29–43 r 100 d
280 N 4 2.7 mk 20–28 mk 30–59 d 63 mk
H 10 h (V/F) 10–12 h (CL/F) 9–42 mk 480 h
N
H H
N
HO OH
O
S
Ritonavir Chiral HIV 720 5.2 202 Neutral 1r 97–99 r 11–17 r 71–78 r 470 h
Refs 34, protease 11 0.3 d 99 d 4–5 d 21–100 d
281–284 S O O N 4 2.2 mk 96–99 mk 11–12 m 30–70 mk
N N N O 0.4–0.5 h 99–99.5 h 1–2.1 h (CL/F)
N N S
O O (V/F)
Saquinavir O HIV 670 4.7 167 Basic (7.6) 4–10 h 98–99 h 13–17 h 4 h hgc B30 hge
Pharmacokinetics and Metabolism of Compounds that Mimic Enzyme
H2N
Refs 34, 51, 246 protease 7 12 h sgc B100 sgc
O OH
H 6
N N N
N
H N
O
O H
SDZ-PRI-053 HIV 723 5.6 158 Weak base 2.7 mo 19 m 47–61 m 60–77 m
Ref 285 O O O protease 7 (6) 1.6 r 34 r 22–54 r 43–100 r
N 6 9.5 d (CL/F)
O N N
O N O
401
O
402
Table 9.1 (continued )
a
TPSA
HBA
Compound Structure Target MW clogP HBD pKa Vss (L/kg) PPB (%) CLp (ml/min/kg) F% Fa (%)
Tipranavir F HIV 603 7.8 114 Acid (4.5) 0.5 r 4 99.9 20 m 11 m 53–60 r
F
Refs 29, O F protease 5 0.13 d in all 2.8 r 28–30 r 418 h
286–288 N 2 0.11–0.14 h species 1.3 d 0.2–0.3 h 10 rab
S N
O O (V/F with (CL/F with 7–8d
O O ritonavir) ritonavir)
O
O
N
O
N
O N
O
BW-175 Renin 688 6.6 154 Basic (6.8) 1.5 r 22 r 3–10 r o15 c r
Ref. 23 O O
N
S N N
O O
O O O
N N
CGP-38560 N Renin 730 5.0 179 Base (6.7) o2 h
Ref. 291 N 7
O O 5
N N
S N
O O O O
Enalkiren O Renin 657 3.7 192 Basic (8.9) 0.2 h 94 h Not Only phar-
Ref. 291 9 measurable macologi-
8 cally active
O O O by IV route
N
N N N
O O
N
N
Remikiren Renin 631 3.6 170 Basic(6.7) 1.8 m 86–96 h 119 r 3r B30 b r
Ref. 22, 24, 7 4.4 r 48 ma 6 ma B56 b d
291,292 O 5 1.5 d 25 d 1.1 d o2 d mk
O O H OH
S N
2.7 ma 20 mk o0.3 mk
N 0.9 h 13 h o2 h
H
O OH
N
N
H
YM-21095 Renin 732 3.5 206 Basic (6.7) 3.9 d 22 r 3–10 r o15 c r
Ref. 293 9 37 d 0.2 d
4
O O N N
H
N N
N N S N
H
O O OH
N NH
Zankiren S Renin 706 5.6 189 Basic (6.2) 148 r 24 r 100b r
N
28 f 32 f
N O OH 10 d 53 d 72 c d
N N 27 mk 8 mk 25 c mk
S N
O O O OH
Pharmacokinetics and Metabolism of Compounds that Mimic Enzyme
d ¼ dog, f ¼ ferret, h ¼ human, hp ¼ human paediatric population, ma ¼ marmoset, mk ¼ monkey, mo ¼ mouse, r ¼ rat, rab ¼ rabbit, hgc ¼ hard gelatin capsule,
sgc ¼ soft gelatin capsule
a
Topological polar surface area
b
Based on urinary and biliary recovery of radioactivity
c
Assuming blood : plasma ratio ¼ 1
d
Assuming blood : plasma ratio ¼ 0.6 (as in human)
403
* ¼ blood binding.
404 Chapter 9
26–30
by food. The requirement for some compounds to be given with food and
others to fasted subjects complicates the usage of these compounds in the
multiple drug regimes needed to treat HIV infection. In addition to food, the
absorption of some compounds is very dependent on intestinal pH. For
instance the low solubility of atazanavir at high pH means that this compound
can not be co-administered with antacids or proton pump inhibitors that raise
the pH of the gastrointestinal tract26 and similar pH dependent effects were
seen when indinavir was dosed orally in the dog.31
The commercial formulations of the protease inhibitors often require organic
excipients to maintain the compounds in solution and to achieve sufficient
exposure after oral dosing; the oral solution formulation of lopinavir contains
40% v/v ethanol, tipranavir is formulated as a fatty emulsion29 and the com-
mercial dosage form of amprenavir requires the use of organic excipients
(including PEG 400 and propylene glycol) and a low drug loading (150 mg),
meaning that pill burden for the patient was high (eight capsules have to be
taken twice a day).32
To try and overcome the problems caused by in solubility, the use of more
soluble prodrugs has been investigated.32 The alcohol group that mimics the
tetrahedral transition state in amprenavir has been conjugated with phosphate
to give a prodrug (fosamprenavir) that is 10-fold more soluble (0.4 mg ml1)
than parent amprenavir (Table 9.1). The advantage of this increased solubility
was that fosamprenavir could be formulated at a higher drug loading (700 mg
tablet), reducing the pill burden for the patient (two capsules dosed once or
twice a day). Once dosed orally, fosamprenavir is converted almost entirely to
amprenavir at or near the intestinal epithelium by alkaline phosphatase in both
animals and humans; the amprenavir is then absorbed and equivalent systemic
exposures are achievable using fosamprenavir and the standard amprenavir
capsules.33 In contrast to amprenavir, fosamprenavir pharmacokinetics are
unaffected by food and this together with the lower pill burden mean that
fosamprenavir has been commercially successful.32 Another advantage for
fosamprenavir is that it should be possible to combine fosamprenavir in a single
tablet with other classes of HIV drugs; the organic excipients and low drug
loading meant it was not easy to do this with amprenavir.
Saquinavir also has low solubility and formulation changes have been used
to increase its bioavailability. Originally marketed as a hard gelatine capsule
(HGC) formulation, saquinavir had low bioavailability (F% 4) partly as a
result of incomplete absorption. When saquinavir was reformulated in a soft
gelatine capsule as a lipid formulation containing mono- and di-glycerol
medium chain fatty acids (SGC) the F% increased 3–4 fold (12–16%) and was
further increased in the presence of food.34 The SGC formulation presumably
increases the solubility and dispersion of saquinavir and in some ways mimics
the effect of food on the compound. However, it is also possible that the
excipients have other effects (e.g. changing intestinal leakiness, inhibiting P-gp)
that contribute to the higher F%. Some evidence to support this theory is that,
when boosted with ritonavir, the plasma levels from the HGC formulation are
higher than those with the SGC formulation;35 this is hard to explain if the
Pharmacokinetics and Metabolism of Compounds that Mimic Enzyme 405
difference between formulations is only due to solubility effects. Adding a basic
centre to molecules is another strategy that has been employed to try and
increase the solubility and hence absorption of HIV protease inhibitors (the
discovery of indinavir is discussed in Chapter 4).
Once in solution the absorption of compounds depends on the passive per-
meability across the intestine and the action of drug transporters that can act to
increase (influx) or decrease (efflux) the rate and extent of drug passage across
the intestine. Ideally absorption of compounds is estimated from a comparison
of intravenous and oral pharmacokinetics of the compound. Many of the
compounds in Table 9.1 are too lipophilic to formulate in a vehicle that can be
administered intravenously. For these compounds, an estimate of human
absorption has been made by assuming that the compounds are not metabo-
lised in the luminal contents of the gastrointestinal tract and calculating the Fa
from the recovery of unchanged drug in faeces after oral dosing. Although this
may underestimate absorption (as material absorbed and excreted unchanged
in bile or by exsorption of unchanged drug back into the intestine is considered
not to have been absorbed), it allows the absorption of more of the compounds
in Table 9.1 to be compared.
PSA is one factor known to limit passive membrane permeability and hence
oral absorption of compounds. A plot of Fa (%) against PSA for the com-
pounds in Table 9.1 is shown in Figure 9.9. Clearly there are compounds with
high PSA and low absorption in this dataset. For example aliskiren (the only
marketed renin inhibitor) is highly soluble (4350 mg ml1) at acidic pH but has
high molecular weight (552) and PSA (146 Å2). Despite low clearance in
humans (9 L hr1; B2 ml min1 kg1), the compound has low oral
100
80
60
FA
40
20
the HIV protease inhibitors can also be transported by human OATPs. To date
no information has been published showing transport by the major human liver
OATPs (1B1, 1B3 and 2B1),59 but saquinavir is transported by OATP 1A2
which is present in the central nervous system (CNS) and cholangiocytes in the
liver and HIV protease inhibitors also appear to be substrates of OATP 3A1
and 4A1.60
A number of HIV protease inhibitors including saquinavir (F% 4–12),
indinavir (F% 60–65) and nelfinavir (F% 45 fasted, 30% fed) have higher oral
F% than would be expected if all the systemic clearance of the compounds
occurred only in the liver, so it is possible that other organs are also involved in
the systemic clearance of these compounds. The low and variable F% and short
half-lives (mainly due to high clearance rather than low volume) of many of the
HIV protease inhibitors means that they need to be dosed two or three times a
day at high doses to have therapeutic efficacy. As well as being CYP 3A4
substrates, most of the HIV proteases are also potent inhibitors and often
inducers of the enzyme and, at the high doses used, these compounds inhibit
and/or induce their own metabolism (resulting in non-linear pharmacokinetics
with increasing dose and time) and alterations in the pharmacokinetics of many
410 Chapter 9
61,263
other compounds co-administered with them. Because of its potent CYP
3A inhibition, ritonavir has been used at low doses to enhance (boost) the
pharmacokinetics of other protease inhibitors allowing a lower pill burden and
less frequent administration of the inhibited compound in HIV patients.
This is currently the main clinical use of ritonavir as, at the high doses needed
for the compound to exert its antiviral effect, ritonavir has an unacceptable
safety and toleration profile. The changes in AUC when protease inhibitors are
co-dosed with ritonavir are shown in Table 9.2 and in some instances can be
quite dramatic; for example, the AUC of saquinavir can increase up to 100-fold
in the presence of ritonavir. Although plasma levels are higher when co-dosed
with ritonavir, the pharmacokinetics of saquinavir are still highly variable.
The pharmacokinetic boosting strategy for HIV proteases is so well accepted
that a number of protease inhibitors (e.g. lopinavir, darunavir, tipranavir) are
only used in combination with ritonavir as they do not achieve sufficiently high
plasma exposure alone. As well as increasing F%, ritonavir also increases the
half-life of a number of compounds including saquinavir and darunavir.27,28
Interestingly ritonavir has been shown to change the disposition of darunavir in
humans. When radiolabelled darunavir was dosed alone, it was rapidly meta-
bolised and recovery of unchanged drug was low (7% faeces, 1% urine), in the
presence of ritonavir excretion of unchanged darunavir was extensive (41% in
faeces and 8% in urine), suggesting that in the presence of ritonavir, trans-
porters play a role in darunavir clearance.27,264
Although ritonavir boosts the levels of all the HIV protease inhibitors, its
interactions with other drugs are more complex. In addition to inhibiting CYP
3A4, ritonavir also inhibits 2D6 and 2C9 and 2C19 and P-gp, and induces levels
of a number of CYP isozymes (including 2B6, 2C9, 2C19) and some UGT iso-
zymes with full induction requiring two weeks of ritonavir dosing.62 Depending
on the length and dose of ritonavir administration and the co-administered
substrate, administration with ritonavir can result in inhibition or induction.
The distribution of the alcohol TSAI of proteases within the body is largely
as would be expected on the basis of the physicochemical properties of the
compounds (Figure 9.5, Table 9.1). The high lipophilicity of these compounds
means that they tend to show extensive binding to plasma proteins. All the HIV
proteases apart from the least lipophilic compound, indinavir, show extensive
protein binding in humans (498%). Amprenavir, darunavir and nelfinavir
bind mainly to a1-acid glycoprotein (AAG).20,27,28,82 Amprenavir also changes
the levels of AAG on chronic dosing and so changes its own protein binding
with time. Atazanavir and lopinavir bind to both albumin and AAG.26,30
The distribution of alcohol transition state analogue inhibitor compounds
has been investigated using whole body autoradiography techniques. Imme-
diately following intravenous dosing of 14C-remikiren in the rat,83 high levels of
radioactivity were seen in the liver, kidney, stomach, small intestine, choroid
plexus, thyroid, pancreas, pituitary and myocardium of all three species studied
(rat, guinea pig and monkey). After one hour, tissue radioactivity levels were
still high but radioactivity was not detectable in the blood. At 24 h post-dose,
radioactivity was still seen in the gastrointestinal tract, kidney and some brain
Pharmacokinetics and Metabolism of Compounds that Mimic Enzyme 411
areas. The kidney appears to act as a reservoir for remikiren because further
analysis of radioactive material in the kidney after 24 h showed that the
majority of radioactivity was unchanged remikiren. A possible explanation for
this retention is the material enters the kidney (possibly via an uptake trans-
porter), but unlike in the liver, there is no comparable excretory transporter to
remove the compound from the kidney. As metabolic capacity for this com-
pound is low in the kidney, the compound remains in the kidney predominantly
as unchanged drug.
HIV protease inhibitors in general have limited and variable ability to enter
the cerebrospinal fluid (CSF) of humans treated with the drugs (Table 9.3). One
reason for this is that P-gp (and possibly other transporters such as MRP 1 and
2) limits the permeability of the compounds across the blood–brain bar-
rier.267,268 Whilst in the intestine compounds can reach high concentrations and
inhibit (and overcome) P-gp, the free systemic concentrations of these com-
pounds are much lower than those seen in the intestine and thus P-gp at the
blood–brain barrier is not inhibited.
The effect of P-gp on limiting CNS penetration of HIV protease inhibitors
has been extensively studied in mice. Increases of between 5–36 fold were seen
in the brains of P-gp knockout mice compared to wild-type mice dosed with
amprenavir, saquinavir, indinavir, ritonavir and nelfinavir .31,44,84,85 Increases
of brain concentrations of a similar magnitude are seen for these compounds
when wild-type mice are dosed with P-gp inhibitors such as GF-120918 or
cyclosporine.31,44 As well as influencing the CNS penetration of protease
inhibitors, transporters (including P-gp and MRP proteins) also play a role in
modulating the levels of HIV protease inhibitors in lymphocytes, a pre-
dominant site of HIV replication.89–92
Inhibition of transporters may play a role in some of the side effects seen with
HIV proteases inhibitors. Saquinavir (1.2 mM) and indinavir (6.8 mM) both
inhibit human OATP 1B1, which has a physiological role in bilirubin transport.
When the Fu in plasma and therapeutic plasma concentrations were considered
it was concluded that, in vivo, the inhibition of OATP-1B1 would be 40% for
indinavir and 7% for saquinavir, and that this may explain the unconjugated
hyperbilirubinemia seen in some patients dosed with indinavir.58 Ritonavir and
Table 9.4 Chemical structure, physicochemical properties and clearance, bioavailability and absorption estimates of aldehyde
containing TSAI.
Compound Structure Target MW clogP TPSAa pKa HBA HBD F%
TPSA 140 92
cpKa 14 12
HBD* 6 3
F
FK-706 Chiral Elastase 593 2.7 162 Acid (3.5) 86 ha
Ref. 104 (phase 2) 7 73 hsa
O 4
F
N F
N F
O
O
O N
N O
O
O
ONO-6818 Elastase 453 51 r
Ref. 102 31 d
N 18 mk
N O
H2N
O N N
H N
O O
O O O
Telaprevir H NS-3 pro- 680 5.4 180 Neutral 0.5–5.8 r 18–53 rb 25–35 r 445 rc
N
(VX-950) O O tease 8 1–1.8 d 42 d 41 d Complete d
Ref. 111, N N HN N (phase 3) 4
N N
265, 296, O O
O O
297
417
418
OH
OH H 2O
R B R B OH
OH OH
Figure 9.11 Conversion from neutral sp2 boron to tetrahedral anionic sp3 boron.
boronic acids typically have a pKa in the 4.5–8 range and butylboronic acid has
a pKa of 10.6.250
As amide (peptide) cleavage requires conversion of a neutral sp2 carbonyl
carbon to a tetrahedral sp3 carbon, boronic acid compounds can mirror this
transition and make the same interactions with the protease enzyme as the
tetrahedral transition state thereby acting as potent TSAI of a number of
protease enzymes.115 Boronic acid inhibitors of serine proteases were first
synthesised in the early 1970s.116,117 Since that time boronic acid compounds
have been shown to be effective inhibitors of a number of enzymes including
thrombin,118,119 Factor Xa,120 trypsin, chymotrypsin, pancreatic elastase, leu-
kocyte elastase, thrombin, b-lactamases,121–125 DPP IV and HCV NS3 pro-
tease,126,127 and have been explored for use as therapeutic agents in various
disease states.115,128
There are two different mechanisms by which the tetrahedral boronate anion
is formed. In the case of arginase inhibitors, the tetrahedral boronate anion is
formed by reaction with water.129 In contrast for other enzymes such as serine
proteases, the hydroxyl anion of the catalytic amino acid typically attacks the
boronic acid to form the boronate intermediate, resulting in the formation of a
covalent bond to the enzyme which, while reversible, typically gives slow off-
rates of hours or days thereby giving a significant pharmacodynamic advantage
to this class of inhibitor.130 Because of this requirement for boronic acids to
form boronates to mimic the transition state of protease enzymes, it has been
argued that boronic acids are not formal transition state analogues but should
be designated as either ‘reaction coordinate analogues’ or covalent inhibitors
depending on the mechanism of boronate formation (water vs. protein).
The boronic acid TSAI of proteases discussed further in this chapter are
shown in Table 9.6. The enzyme targets of these compounds include thrombin,
dipeptidylpeptidase IV (DPP-IV) and the 20S proteasome. DPP-IV cleaves
dipeptides from the N-terminus of numerous polypeptides including incretins
such as GLP-1 that have an important role in blood glucose homeostasis.
Small, polar (clogP 2.2 to 5.3) boronic acid containing compounds are
effective TSAIs of DPP-IV. Proteasomes cleave peptides through threonine-
dependent nucleophilic attack and transition state inhibitors have been
designed to inhibit this process. The boron containing TSAI of proteasomes
need higher lipophilicity (clogP 0.81.3) to effectively bind in the active site of
the proteasome. Bortezomib is a potent boronic acid inhibitor of the chymo-
tryptic-like site of the 20S proteasome (Ki 0.62 nM) and establishes a slowly
reversible (half-life B20 min) tetrahedral intermediate with the active site
N-terminal threonine residue.131
422
Table 9.6 Chemical structure, physicochemical properties and clearance, bioavailability and absorption estimates of boronic acid
containing TSAI.
TPSA
HBA Vss PPB CL
Compound Structure Target MW clogP HBD pKa (L/kg) (%) (ml/min/kg) F% % absorbed
O
Tri-50b Thrombin 608 5.0 115 Neutral 24 r
Ref 134
4
O O 2
N N B
O N O
O O
O
DuP-714 N N Thrombin 460 2.4 183 Weak acid
5 (9.6)
N 8
N
N
O
O N B
O O
Chapter 9
O
S-18326 Thrombin 488 1.3 181 Weak acid 27 dfa
Ref 299 N N (9.6)
6 6 dfe
N 8
O O
N O
N N B
O O
Bortezomib Proteasome 384 0.8 124 Weak acid B3–13 h 93 r 14 mo 11 mo
(approved) N (9.7) Vz B
Ref. 136–138, O 24–7 h
247 4 94 d 13–27 hsd
N O
N N B 4 83 h B3.5–8.5 hmd
O O
dfa ¼ dog fasted, dfe ¼ dog fed, hmd ¼ human multiple dose, hsd human single dose, r ¼ rat.
a
Assuming blood : plasma ¼ 1.
mo ¼ mouse.
Pharmacokinetics and Metabolism of Compounds that Mimic Enzyme
423
424 Chapter 9
In general the boronic acid compounds discussed have reasonable PSA
values for oral absorption (six out of seven compounds have PSA in the range
85–140 Å2). One reason for this is that the boronic acid compounds can be
considered as prodrugs as they are converted to the final, more polar tetra-
hedral transition state analogue inhibitor in the body.
Polar boronic acids are poorly absorbed transcellularly as would be expected
from their physicochemistry; for instance DUP-714, a thrombin inhibitor
containing both a guanidine and boronic acid group (clogP 2.4, PSA 183 Å2),
has low permeability through rat jejunum (Papp 1.5106 cm sec1, pH 7.4).
Some analogues of DUP-714 had significantly higher permeability (Papp
18106 cm sec1), which was reduced in the presence of known inhibitors of
the oligopeptide transporter,132 suggesting that while passive permeability was
not high this could be overcome if the compounds cross the membrane by
active influx mechanisms. Dutogliptin is a small (MW 241), basic, polar (clogP
2.2) DPP IV inhibitor. It is orally bioavailable (CL/F 10–30 ml min1 kg1) in
humans and Tmax was 3–5 h. Given the physicochemical properties of duto-
gliptin, it is likely that this compound is absorbed via the paracellular route.
The more lipophilic protease inhibitor CEP-18770 (clogP 1.6) has reached
phase I trials. In pre-clinical species it has high absorption and oral bioavail-
ability, low clearance (although the protein binding and hence CLu are high)
and a long half-life.133 The data with CEP-18770 demonstrate that the presence
of the boronic acid moiety per se does not preclude good transcellular
absorption.
In an effort to increase the absorption of poorly absorbed boronic acid
TSAIs, prodrug strategies using boronic esters (which decrease the number
of HBD by two) have been employed. TRI-50b is a pinanediol ester of
flovagatran which showed modest bioavailability in the rat and is cur-
rently in clinical trials—phase I (oral as prodrug) and phase II (iv as
flovagatran).12,134,135
The distribution of boronic acid compounds in vivo can be influenced by their
ability to form reversible covalent bonds. For example, bortezomib in addition
to having moderate binding to human plasma proteins, binds slowly to the
proteasome in red blood cells. This binding has been shown to compli-
cate the bioanalysis of bortezomib in blood with differential recoveries
between incurred and spiked standard samples being noted.251 Bortezomib has
marked multiphasic pharmacokinetics after intravenous dosing with a long
terminal elimination phase (half-life B9–15 h) and a large terminal
volume of distribution (410 L kg1).136–137,247 Some studies have reported
a change in volume on multiple dosing,136 whilst others have shown it to be
unchanged.138 Extensive distribution of bortezomib related radioactivity
into tissues (apart from the brain) was also seen in monkeys dosed with
radiolabelled bortezomib.
The elimination and clearance pathways of boronic acid compounds vary
from compound to compound. After intravenous dosing of flovagatran, 14%
of the dose was excreted unchanged in the urine; in contrast after intravenous
administration bortezomib is extensively metabolised with little excretion of
Pharmacokinetics and Metabolism of Compounds that Mimic Enzyme 425
unchanged bortezomib in the urine and none in the faeces of humans dosed
with the compound.139,252 The initial route of metabolism of bortezomib is
oxidative deboronation (none of the thirteen metabolites in human plasma
contain the boronic acid group), which removes the functional group
mimicking the transition state producing metabolites inactive against the
proteosome.139 Extensive deboronation is likely to be a major metabolic
pathway for any boronic acid compound lipophilic enough to undergo CYP
metabolism and is consistent with the finding that boronate compounds in
general are susceptible to undergoing Baeyer–Villiger-type oxidative debor-
onation reactions.139–142
In human liver, CYP 3A4 is the predominant isozyme metabolising borte-
zomib with quantitatively important contributions also from CYP 2C19 and
CYP 1A2.143 In vivo drug–drug interaction (DDI) studies showed that clear-
ance of bortezomib was decreased; AUC increased by a mean value of 35%
(CI90% 1.03 to 1.8) in the presence of ketoconazole (a CYP 3A4 inhibitor) but
not in the presence of omeprazole (a potent CYP 2C19 inhibitor).247 As well as
being metabolised by CYP 2C19, bortezomib is also a weak inhibitor of the
isozyme (IC50 ¼ 18 mM),144 although the inhibition is related to the scaffold of
the compound rather than to the presence of the boronic acid group per se.
Mechanistic studies145 suggest that human P450 can deboronate bortezomib via
multiple mechanisms including CYP mediated formation of reactive oxygen
species (e.g. superoxide) in addition to direct oxidation by the hypervalent oxo-
iron species [Fe(V) ¼ O].
Bortezomib and CEP-18770 can also theoretically cause CYP-mediated DDI
by inhibiting the degradation of CYP isozymes by the 20S proteasome (their
pharmacological target). In rats, bortezomib inhibits the degradation of CYP
2E1 following induction of the isozyme by ethanol but paradoxically decreased
the levels of CYP 3A1/2 and total CYP content.144 In humans, bortezomib
administration caused no changes in results from the erythromycin breath test
(an in vivo measure of CYP 3A4 activity).
The physical form and stability of boronic acid TSAIs can also pose a pro-
blem in their development as therapeutics. For example in solid form borte-
zomib exists in its cyclic anhydride form as a trimeric boroxine. To overcome
this, the bortezomib drug product is a mannitol ester which, when recon-
stituted, exists in equilibrium with the monomeric boronic acid form (solubility
B3 mg ml1).146 Bortezomib is also chemically unstable in a number of
matrices and acidification was needed to stabilise bortezomib during protein
precipitation extractions.139
Transition state
Carboxylic acid O OH
H B
OH O E
O
N H
O N O O
CH3 H O H
N R
O
..
Enalaprilat R′
Zn
H
O Y
Phosphonic acid OR O O OH E
O
P
N O H
O
H
O
+
R= H Fosinoprilat P
R= Fosinopril
O O
O Zn HO Y
Figure 9.13 Example of sulfhydryl, carboxylic and phosphonic acid ACE inhibitors.
(including the kidney) were responsible for about 50% of the systemic
clearance.147
OH OH
O OH O OR
O δ+ O
HO HO OH
OH
HN HN
HO OH HO OH
O O
Figure 9.14 Structure of sialic acid and the proposed transition state of sialic acid
within viral neuraminidase.
(Figure 9.14, Table 9.7). Although, the interactions of sialic acid, zanamivir and
oseltamivir acid with viral neuraminidase are similar, the TSAI bind to the
enzyme 0.5–1.5106 times more tightly than sialic acid and show time-depen-
dent and slowly reversible inhibition kinetics.150,152
Zanamivir is a highly polar zwitterionic compound and as a result has poor
membrane permeability, limited absorption and low bioavailability in rat and
human (Table 9.7). Zanamivir has higher F% when administered via the
intranasal (B10%) or inhaled routes (B15–20%) in humans.153 Inhaled
administration of zanamivir Tc99m (10 mg) in healthy volunteers showed that
4–17% of the dose was deposited in the lungs. After dry powder inhalation
administration, the Tmax of zanamivir in the systemic circulation was 1–2 h;
the systemic half-life of zanamivir was longer than after intravenous adminis-
tration (3.6–5 vs. 1.7 h) suggesting that absorption through the lung is slow
enough to become the rate limiting step in the elimination of zanamivir (flip-
flop kinetics).
Oseltamivir acid is 4 log units more lipophilic than zanamivir and has a PSA
more typical of an oral drug (o140 A2). However, oseltamivir acid is still quite
hydrophilic clogP 1.2 and in vitro has limited ability to cross cell monolayers
(Papp o2106 cm s1). When administered orally to rats or humans, osel-
tamivir acid has low bioavailability, primarily due to low absorption.154 To
overcome this limitation, the ethyl ester prodrug (oseltamivir) of oseltamivir
acid was prepared. Oseltamivir is about 3 log units more lipophilic than osel-
tamivir acid and has high aqueous solubility (4500 mg ml1) and a moderate
ability to cross cell monolayers in vitro ( Papp 6–12106 cm s1 in Caco-2 and
LLC-PK1 cell lines).155,156 Despite its relatively small size, oseltamivir is a
substrate for human P-gp.
In pre-clinical pharmacokinetic studies, the bioavailability of oseltamivir
acid following oral dosing of oseltamivir was 11–73% (Table 9.7). The lowest
F% was seen in the ferret which had a relative inability to cleave oseltamivir to
the acid metabolite suggesting that, even in this species, absorption of oselta-
mivir was good.157 The improvement in F% (o5 to B75–80%)154 of
Table 9.7 Chemical structure, physicochemical properties and clearance, bioavailability and absorption estimates of neur-
aminidase TSAI.
Compound Structure MW clogP TPSA pKa HBA HBD CL (ml/min/kg) Oral F%
O O
Zanamivir 332 5.6 198 3.8 A 10 9 1.4–32 r 3–4 r
Ref 153 N O O
O 11.3 B 1.6 h 2h
N N
O N O
O N
N
O
O O
Oseltamivir acid 284 1.2 102 2.2 A 5 4 25 r 4.3 r
Ref 154
9.7 B 5–6 h o5 h
O NH2
11 fa
N
30–73* m, r,
O
d, ma
75–80 ha
O O
GS–4116 326 1.7 138 4.3 A 7 6 22 4r
Ref 157
N N
12.8 B 2 ra
O
N
Pharmacokinetics and Metabolism of Compounds that Mimic Enzyme
N
O
Dimeric zanamivir 943 5.5 412 19 17 0.7 o1
O O
Ref 172
O O N N O O
O O O
O O
O
N O N
N N N N
429
O O
N N
d ¼ dog, f ¼ ferret, h ¼ human, m ¼ mouse, ma ¼ marmoset, r ¼ rat. aFollowing dosing of ethyl ester prodrug.
430 Chapter 9
oseltamivir acid seen in humans when dosed as oseltamivir is one of the highest
seen with an ester prodrug 158 and it is unusual to get such a big increase in F%
with such a simple modification. In a study to look at regional absorption of
oseltamivir in humans, the prodrug was well absorbed from the stomach,
jejunum and ileum.156 Oseltamivir is absorbed at least in part transcellularly, as
slow absorption was also seen from the colon (a region where tight junctions
have low porosity and paracellular absorption is minimal).
Although oseltamivir was a successful prodrug, attempts to ‘prodrug’ a
more polar analogue (GS-4116: clogP 1.7, PSA 138 Å2) were unsuccessful
(Table 9.7). The ethyl ester prodrug of GS-4116 showed no improvement
in oral bioavailability (F% B2) in the rat compared with oral administration of
the acid compound itself.157 As illustrated by this example, prodrugging of
compounds is unlikely to be successful if the PSA of the parent molecule is too
far from ideal oral drug space.159
The distribution of zanamivir and oseltamivir acid within the body is heavily
influenced by their polar physicochemical characteristics. Plasma protein
binding of zanamivir is low and that of oseltamivir negligible, whilst the more
lipophilic oseltamivir has moderate plasma protein binding in humans
(B42%). Distribution of zanamivir into red blood cells is also negligible
(BP ¼ 0.4–0.6).160,161 The steady-state volume of distribution (Vss) in humans
of both zanamivir and oseltamivir acid is low (0.2–0.4 L kg1) and similar to the
volume of extracellular fluid. The volume of distribution in dog was similar to
that in human (0.3 L kg1) but higher volumes of distribution (1, 3 and
9 L kg1) were seen in rat, mouse and ferret, respectively.157
Some concerns about neuropsychiatric side effects of oseltamivir arising in
Japan162 have led to a number of studies looking at concentrations of oselta-
mivir and oseltamivir acid in the CNS. In rats following oral dosing of oselta-
mivir, the brain: plasma AUC ratio was 0.15 for oseltamivir and 0.01 for
oseltamivir acid. In vitro studies showed that rat and human S9 brain homo-
genates had little ability to convert oseltamivir to oseltamivir acid.162 The lim-
ited brain penetration of oseltamivir is in part due to transporter-mediated
efflux. Oseltamivir is a substrate for both human and murine P-gp and the brain
to plasma ratio of oseltamivir was increased (B5-fold) in P-gp knockout mice or
in wild-type mice dosed with the P-gp inhibitor GF-120918. The brain pene-
tration of oseltamivir is also lower in older rats that are known to have increased
P-gp expression in the blood–brain barrier than younger animals. The limited
brain penetration of oseltamivir acid is largely due to the poor membrane
permeability of the compound as the brain to plasma ratio of oseltamivir acid
was unaffected by the P-gp inhibitor and was similar in P-gp knockout and wild-
type mice,155,163 although other transporters may have a role in limiting osel-
tamivir acid brain penetration. The penetration of oseltamivir and oseltamivir
acid into the CNS of humans was also low (CSF: free plasma concentration
ratio was o0.05) and CSF Cmax was delayed relative to that in the plasma
suggesting a barrier to free permeability of the compounds in the CNS.164
Zanamivir and oseltamivir acid are highly polar and are almost exclusively
(87–90%) excreted into urine as unchanged drug following intravenous
Pharmacokinetics and Metabolism of Compounds that Mimic Enzyme 431
160
dosing. The renal clearance of zanamivir was similar to the glomerular fil-
tration rate (GFR) in humans suggesting minimal reabsorption of zanamivir
(a highly polar; water soluble compound) occurs in the kidney presumably due
to its poor membrane permeability. Renal clearance of oseltamivir acid
(B4.5 ml min1 kg1) in humans is greater than GFR suggesting active secre-
tion into the urine occurs. In vitro studies have shown that oseltamivir acid is a
substrate and weak inhibitor (Ki 45 mM) of human renal transporter
hOAT1165 and renal Cl in humans in vivo is inhibited by probenecid (a known
hOAT inhibitor).166 In the presence of probenecid, the renal Cl of oseltamivir
acid is roughly equal to GFR suggesting that the active component of oselta-
mivir acid is completely inhibited by probenecid. Oseltamivir itself is also
actively secreted in to human urine (mean renal Cl 360–480 ml min1, B6 ml
min1 kg1), although the responsible transporter has not been identified.167
Oseltamivir acid is the only metabolite of oseltamivir identified in humans. In
rats, however, some additional minor metabolites have been identified of which
the R-omega (o) carboxylic acid metabolite was the most abundant (metabo-
lism occurs on the terminal carbon of the pentoxy group). Rat liver microsomes
in the presence of NADPH metabolise oseltamivir to the o-hydroxy metabo-
lite—a first step on the pathway to the acid.168
As with other ester prodrugs,158 the tissues hydrolysing oseltamivir to osel-
tamivir acid differ across species. The hydrolysis of oseltamivir is rapid in rat
plasma but relatively slow in ferret, dog and human plasma.157,169 Oseltamivir
is also hydrolysed to oseltamivir acid in human liver. There are two major
carboxylesterase isozymes with differing substrate specificity in human liver,
HCE-1 and 2.170 Hydrolysis of oseltamivir in human liver correlates with HCE-
1 activity and the Km in human liver microsomes (187 mM) and recombinant
HCE-1(177 mM) were similar. The rate of hydrolysis of oseltamivir was shown
to vary when some naturally occurring variants of HCE1 were assessed for
catalytic activity. HCE1S58N slightly increased the hydrolysis (B25%)
whereas variants HCE1C70F and HCE1R128H markedly decreased the
hydrolysis.171 To date the impact of these mutations on oseltamivir human
pharmacokinetics has not been reported.
Once formed from orally administered oseltamivir, the elimination half-life
of oseltamivir acid in humans (6–10 h) is considerably longer than that seen
following intravenous administration of oseltamivir acid itself (2 h);154 a similar
phenomenon is seen in pre-clinical species.154 As oseltamivir itself is rapidly
absorbed, the longer plasma half-life of oseltamivir acid seen after oral
administration of oseltamivir probably reflects the rate at which the acid
metabolite is formed by hydrolysis and crosses the sinusoidal membrane of the
liver to re-enter the blood. Given the polarity of oseltamivir acid and its poor
membrane permeability, it is likely that transporters play a role in the re-entry
of the compound in to the circulation from the liver.
Attempts have been made to increase the half-life and retention of zanamivir
in the lung using polymers of di-, tri- and tetrameric zanamivir molecules linked
by lipophilic spacer chains, and by making a dextran linked polymer of zana-
mivir. These modifications have resulted in molecules with greater potency and
432 Chapter 9
172
a longer residence time in in vivo models of influenza infection. The zana-
mivir dimer shown in Table 9.7 is 100 times more potent than zanamivir in a
mouse influenza model, and levels in rat lung one week after dosing are 100-
fold higher than monomeric zanamivir.172 In rat studies where the pharma-
cokinetics of zanamivir and the dimer were directly compared, the dimer had
a two-fold lower Cl than zanamivir and a larger although still small volume
(0.2 vs. 0.08 L kg1) and has high F% (67) after intratracheal dosing.
O O
N N N N
OH N OH
N
H
N
N
OH OH OH OH
Forodesine BCX-4208
Mouse a 63
Rat B7 B0.6 (Vz) 7–22
Dog 100
Monkey 5–10 3–13
Human B1.5 B1.3 30–50 32–53
a ¼ terminal volume.
OH
NH2 HO NH2
N
N N
N N
N
N N N O
O N O N OH
H N OH
OH
HO OH
HO OH
OH
Figure 9.17 Chemical structure of pentostatin, the substrate of ADA (adenosine) and
the proposed transition state intermediate of ADA.
436 Chapter 9
198
lower and half-life longer in patients with impaired renal function and a large
proportion of pentostatin is excreted unchanged in the urine in humans (esti-
mates vary from 20–66, 32–48, 50–73 up to 90% in different studies) and ani-
mals (mice 91–100%, rats 81–86%, dogs 54–83%).198,199,197,259
(i) The zinc atom in the active site activates water to OH, which subse-
quently attacks the C4 position of substrate.
(ii) The resulting tetrahedral transition state is stabilised by a strong
interaction with the carboxylate group of the enzyme present in the
vicinity.
(iii) Intramolecular proton transfer is mediated entirely by the carboxylate
group.
(iv) Stabilisation of the leaving NH3 by hydrogen bonds to the protein
backbone carbonyl oxygen of T127.206
NH2 O
H2N OH
N H2O
NH NH
N O N O N O
NH3
R R R
cytidine uridine
H OH
H OH
N + H2O NH
NH
N O N O N O
R R R
zebularine tetrahydrouridine
Like the transition state of CDA that they mimic, zebularine (MW 228, clogP
2.1) and THU (MW 248, 1.90) are small polar compounds. Both com-
pounds have low oral F% due to a combination of extensive first-pass meta-
bolism and limited passive permeability and poor absorption.201,214–216 In
contrast to the low oral bioavailability, subcutaneous bioavailability of THU
was high (F% 84) in humans.201,216 The pharmacokinetics of zebularine and
THU are summarised in Table 9.9.
Despite the compounds having similar physicochemistry, their clearance
mechanisms are quite different. THU is excreted in the urine predominantly as
unchanged drug (80–90%) after intravenous dosing in humans, rats, dogs and
monkeys.216,218 In the mouse renal excretion of unchanged drug is lower (55%)
and renal clearance (5.0–6.2 ml min1 kg1) is less than GFR in this species
(B10 ml min1 kg1)219 suggesting that there is some renal reabsorption of this
polar compound in the mouse following glomerular filtration.217 The clearance
of THU in mice is low compared to other pyrimidine analogues because the
compound is not a substrate for the enzymes (thymidine phosphorylase or
uridine phosphorylase) that typically cleave the ribose moiety of pyrimidine
analogues.217 In humans, no metabolites of THU were seen after intravenous
or subcutaneous dosing, but after oral dosing of 14C THU, the radiochemical
and biochemical assays did not give the same results suggesting that a sig-
nificant part of the dose was metabolised to inactive species. This may be
due to metabolism by intestinal microflora as well as to first-pass intestinal
metabolism.216
In contrast to THU, zebularine is extensively metabolised. Enzymatic
degradation of zebularine occurs in both rat and human plasma, although the
438
HO OH OH
HO
Uracil
Zebularine Uridine
Dihydropyrimidine DH
NH
CO2 + NH3 + β-alanine
N O
H
dihydrouracil
animals and humans. Zebularine increases (two-fold) the steady state plasma
levels of 5-aza-2 0 -deoxycytidine when the compounds are co-infused in mice,
whilst concomitant administration of THU markedly increases the myelosup-
pressive potency of Ara-C in monkeys.224 In humans, simultaneous oral
administration of THU (10–50 mg kg1) and Ara-C resulted in a two-fold
increase of blood levels of Ara-C at all time points over a four hour period and
a slight increase in the half-life of Ara-C. The inhibitory effect of THU on CDA
metabolism of Ara-C was also reflected in a considerably increased ratio of
THU/uracil arabinoside excretion in the urine.225 Although there is a theore-
tical risk of toxicity from zebularine administration due to the large amount of
uridine formed, in reality this does not seem to be an issue.226–230
N NH2
O N
O
OH
HO OH
HO
O HO2C O
HO2C H2
N O
NH2 O H C
+ HN
2 P
P CO2H O
CO2H O O O
O
PALA
HO2C O
N O
H O
H2N P
CO2H O
O
HO2C O
N O O
H + P
NH2 O
CO2H O
As would be expected for such a polar compound (clogP 2.2, clogD 8.9),
the pharmacokinetics in animals and humans are characterised by poor oral
bioavailability, elimination unchanged by the kidney, little metabolism and
distribution into extracellular water with little tissue affinity.244 In humans,
clearance was B1.6 ml min1 kg1 and renal excretion of unchanged drug was
80–85%. The renal CL of PALA approximates to the glomerular filtration rate,
suggesting that CL may be by passive filtration with no reabsorption (though
equivalent rates of active renal secretion and reabsorption can not be ruled
out). The volume of distribution was B0.3 L kg1 reflecting the limited dis-
tribution of PALA into tissues and the half-life in man was 5–12 h.
OH H
+
HO O
R
O
HO
HO
OH
OH
O
HO
OH OH OH
HO N O O
O
HO HO O OH
OH OH
HO
Acarbose
Figure 9.22 Chemical structure of acarbose and the putative transition state of the
glycosidase enzyme.
9.4 Conclusions
In general, the group of transition state analogues discussed in this chapter
have a relatively high PSA, particularly when this value is compared with the
PSA value of the top 200 selling drugs. This high polar surface area imparts a
number of pharmacokinetic properties to this group of compounds.
Compounds with low clog P and high TPSA typically had poor membrane
permeability and low transcellular oral absorption. This has led to a number of
444 Chapter 9
compounds either being dosed as prodrugs or being given by other routes such
as intravenous, subcutaneous or inhaled administration. Even when this type of
transition state analogue compound shows good absorption and high bioa-
vailability, this can often be attributed to paracellular or transport-mediated
absorption rather than to good transcellular permeability.
One strategy to improve oral absorption of polar transition state analogues is
to increase clogP significantly and there are examples such as the HIV protease
inhibitors where the compounds have both high clogP and high TPSA yet still
show good absorption. The downside is that the extra lipophilicity added to
increase absorption also makes this class of compounds susceptible to rapid
metabolic clearance and frequently F% is low because the higher absorption is
counteracted by higher Cl and first-pass extraction.
Other properties of this group of compounds such as clearance and volume
of distribution are as would be expected from their physicochemical properties.
Polar compounds tend to have a low volume of distribution and are often
excreted unchanged into the urine. More lipophilic compounds have relatively
higher volumes of distribution and tend to be rapidly cleared by metabolism. In
some cases (e.g. boronic acid and activated ketones), the metabolism involves
the functional groups mimicking the transition states whilst in many other
compounds the metabolism occurs on the molecular scaffold.
The high PSA of these compounds also suggests that attempting to inhibit
central (behind blood–brain barrier) targets with TSAI will be challenging. The
affinity of many of the compounds to P-glycoprotein will exacerbate the
challenge of inhibiting central targets.
Striking a balance between absorption and clearance is the main challenge
for scientists looking to develop transition state analogues as therapeutic
agents. However, the high potency (and often slow offset from the target) of
these compounds means that, if the balance between absorption and clearance
can be achieved, useful therapeutic agents can be discovered.
9.5 Abbreviations
ACE angiotensin converting enzyme
ACT aspartate carbamyl transferase
ADA adenosine deaminase
ADME absorption, distribution, metabolism and excretion
ANG I angiotensin I
ANG II angiotensin II
AO aldehyde oxidase
Ara-C cytosine arabinoside
BP blood : plasma ratio
Cav0t average concentration from 0 to time t
CDA cytidine deaminase
CLh hepatic blood clearance
CNS central nervous system
CSF cerebrospinal fluid
Pharmacokinetics and Metabolism of Compounds that Mimic Enzyme 445
CYP cytochrome P450
DDI drug–drug interaction
DPP-IV dipeptidyl peptidase IV
F bioavailability
Fa fraction absorbed
Fg fraction escaping first-pass metabolism in the gastrointestinal
tract
Fh fraction escaping first-pass clearance in the liver
GFR glomerular filtration rate
HBA hydrogen bond acceptor
HBD hydrogen bond donor
HCV Hepatitis C virus
HIV human immunodeficiency virus
IMPDH inosine 5-monophosphate dehydrogenase
MTAP methylthioadenosine phosphorylase
OATP organic anion polypeptide
P-gp P-glycoprotein
PALA phosphonoacetyl-L-acetate
PNP purine nucleoside phosphorylase
Q hepatic blood flow
THU 3,4,5,6-terahydrouridine
TPSA topological polar absorption area
TSAI(s) transition state analogue inhibitors(s)
Vss steady state volume of distribution
References
1. J. G. Robertson, Biochemistry, 2005, 44, 5561.
2. V. L. Schramm, Curr. Opin. Struct. Biol., 2005, 15, 604.
3. L. Pauling, Nature, 1948, 161, 707.
4. V. H. Lillelund, H. H. Jensen, X. Liang and M. Bols, Chem. Rev., 2002,
102, 515.
5. V. L. Schramm, Annu. Rev. Biochem., 1998, 67, 693.
6. R. Wolfenden, Nature, 1969, 223, 704.
7. R. Wolfenden, Annu. Rev. Biophys. Bioeng., 1976, 5, 271.
8. R. Wolfenden and M. J. Snider, Acc. Chem. Res., 2001, 34, 938.
9. B. A. Malcolm, R. Liu, F. Lahser, S. Agrawal, B. Belanger, N. Butkie-
wicz, R. Chase, F. Gheyas, A. Hart, D. Hesk, P. Ingravallo, C. Jiang,
R. Kong, J. Lu, J. Pichardo, A. Prongay, A. Skelton, X. Tong, S. Ven-
katraman, E. Xia, V. Girijavallabhan and F. G. Njoroge, Antimicrob.
Agents Chemother., 2006, 50, 1013.
10. S. Venkatraman, S. L. Bogen, A. Arasappan, F. Bennett, K. Chen, E. Jao,
Y. T. Liu, R. Lovey, S. Hendrata, Y. Huang, W. Pan, T. Parekh, P. Pinto,
V. Popov, R. Pike, S. Ruan, B. Santhanam, B. Vibulbhan, W. Wu,
W. Yang, J. Kong, X. Liang, J. Wong, R. Liu, N. Butkiewicz, R. Chase,
446 Chapter 9
A. Hart, S. Agrawal, P. Ingravallo, J. Pichardo, R. Kong, B. Baroudy,
B. Malcolm, Z. Guo, A. Prongay, V. Madison, L. Broske, X. Cui, K. C.
Cheng, Y. Hsieh, J. M. Brisson, D. Prelusky, W. Korfmacher, R. White,
S. Bogdanowich-Knipp, A. Pavlovsky, P. Bradley, A. K. Saksena, A.
Ganguly, J. Piwinski, V. Girijavallabhan and F. G. Njoroge, J. Med.
Chem., 2006, 49, 6074.
11. J. J. Deadman, S. Elgendy, C. A. Goodwin, D. Green, J. A. Baban,
G. Patel, E. Skordalakes, N. Chino, G. Claeson and V. V. Kakkar, et al.,
J. Med. Chem., 1995, 38, 1511.
12. P. C. Weber, S. L. Lee, F. A. Lewandowski, M. C. Schadt, C. W. Chang
and C. A. Kettner, Biochemistry, 1995, 34, 3750.
13. V. L. Schramm, Nucleic Acids Res. Suppl., 2003, 3, 107.
14. T. L. Amyes and J. P. Richard, ACS Chem. Biol., 2007, 2, 711.
15. A. Radzicka and R. Wolfenden, Methods Enzymo.l, 1995, 249, 284.
16. J. B. Cooper, Curr. Drug Targets, 2002, 3, 155.
17. T. Harrison and D. Beher, Prog. Med. Chem., 2003, 41, 99.
18. A. Brik and C. H. Wong, Org. Biomol. Chem., 2003, 1, 5.
19. M. Jaskolski, A. G. Tomasselli, T. K. Sawyer, D. G. Staples, R. L.
Heinrikson, J. Schneider, S. B. Kent and A. Wlodawer, Biochemistry,
1991, 30, 1600.
20. B. M. Sadler and D. S. Stein, Ann. Pharmacother., 2002, 36, 102.
21. Y. Etoh, M. Miyazaki, H. Saitoh and N. Toda, Jpn. J. Pharmacol., 1993,
63, 109.
22. P. Coassolo, W. Fischli, J. P. Clozel and R. C. Chou, Xenobiotica, 1996,
26, 333.
23. H. Morishima, Y. Koike, M. Nakano, S. Atsuumi, S. Tanaka,
H. Funabashi, J. Hashimoto, Y. Sawasaki, N. Mino and M. Nakano,
et al., Biochem. Biophys. Res. Commun., 1989, 159, 999.
24. C. H. Kleinbloesem, C. Weber, E. Fahrner, M. Dellenbach, H. Welker,
V. Schroter and G. G. Belz, Clin. Pharmacol. Ther., 1993, 53, 585.
25. C. Weber, H. Birnbock, J. Leube, I. Kobrin, C. H. Kleinbloesem and
P. Van Brummelen, Br. J. Clin. Pharmacol., 1993, 36, 547.
26. C. Le Tiec, A. Barrail, C. Goujard and A. M. Taburet, Clin. Pharmaco-
kinet., 2005, 44, 1035.
27. D. Back, V. Sekar and R. M. Hoetelmans, Antivir. Ther., 2008, 13, 1.
28. C. McCoy, Clin. Ther., 2007, 29, 1559.
29. B. J. Dong and J. M. Cocohoba, Ann. Pharmacother., 2006, 40, 1311.
30. R. S. Cvetkovic and K. L. Goa, Drugs, 2003, 63, 769.
31. J. H. Lin, Ernst Schering Res. Found. Workshop, 2002, 37, 33.
32. V. J. Stella and K. W. Nti-Addae, Adv. Drug Deliv. Rev., 2007, 59,
677.
33. M. B. Wire, M. J. Shelton and S. Studenberg, Clin. Pharmacokinet., 2006,
45, 137.
34. J. W. Beach, Clin. Ther., 1998, 20, 2.
35. G. L. Plosker and L. J. Scott, Drugs, 2003, 63, 1299.
36. M. Azizi, J. Mol. Med., 2008, 86, 647.
Pharmacokinetics and Metabolism of Compounds that Mimic Enzyme 447
37. F. Waldmeier, U. Glaenzel, B. Wirz, L. Oberer, D. Schmid, M. Seiberling,
J. Valencia, G. J. Riviere, P. End and S. Vaidyanathan, Drug Metab.
Dispos., 2007, 35, 1418.
38. J. E. John, Chem. Biol. Drug Des., 2009, 73, 367.
39. H. D. Kleinert, S. H. Rosenberg, W. R. Baker, H. H. Stein, V. Klinghofer,
J. Barlow, K. Spina, J. Polakowski, P. Kovar and J. Cohen, et al., Science,
1992, 257, 1940.
40. A. Seelig, Eur. J. Biochem., 1998, 251, 252.
41. S. F. Zhou, Xenobiotica, 2008, 38, 802.
42. G. C. Williams and P. J. Sinko, Adv. Drug Deliv. Rev., 1999, 39, 211.
43. S. J. Mouly, M. F. Paine and P. B. Watkins, J. Pharmacol. Exp. Ther.,
2004, 308, 941.
44. J. W. Polli, J. L. Jarrett, S. D. Studenberg, J. E. Humphreys, S. W.
Dennis, K. R. Brouwer and J. L. Woolley, Pharm. Res., 1999, 16, 1206.
45. S. Vaidyanathan, V. Jarugula, H. A. Dieterich, D. Howard and W. P.
Dole, Clin. Pharmacokinet., 2008, 47, 515.
46. V. J. Wacher, J. A. Silverman, Y. Zhang and L. Z. Benet, J. Pharm. Sci.,
1998, 87, 1322.
47. A. Galetin, L. K. Hinton, H. Burt, R. S. Obach and J. B. Houston, Curr.
Drug Metab., 2007, 8, 685.
48. M. Gertz, J. D. Davis, A. Harrison, J. B. Houston and A. Galetin, Curr.
Drug Metab., 2008, 9, 785.
49. J. Yang, M. Jamei, K. R. Yeo, G. T. Tucker and A. Rostami-Hodjegan,
Curr. Drug Metab., 2007, 8, 676.
50. M. E. Fitzsimmons and J. M. Collins, Drug Metab. Dispos., 1997, 25,
256.
51. H. H. Kupferschmidt, K. E. Fattinger, H. R. Ha, F. Follath and S.
Krahenbuhl, Br. J. Clin. Pharmacol., 1998, 45, 355.
52. K. C. Yeh, P. J. Deutsch, H. Haddix, M. Hesney, V. Hoagland, W. D. Ju,
S. J. Justice, B. Osborne, A. T. Sterrett, J. A. Stone, E. Woolf and S.
Waldman, Antimicrob. Agents Chemother., 1998, 42, 332.
53. N. C. Cohen, Chem. Biol. Drug Des., 2007, 70, 557.
54. R. Jauch, P. Schmid, W. Fischli, W. Meister, R. Maurer and G. Wendt,
Xenobiotica, 1996, 26, 285.
55. A. J. Parker and J. B. Houston, Drug Metab. Dispos., 2008, 36, 1375.
56. R. G. Tirona, B. F. Leake, A. W. Wolkoff and R. B. Kim, J. Pharmacol.
Exp. Ther., 2003, 304, 223.
57. Z. W. Ye, P. Augustijns and P. Annaert, Drug Metab. Dispos., 2008, 36,
1315.
58. S. D. Campbell, S. M. de Morais and J. J. Xu, Chem. Biol. Interact., 2004,
150, 179.
59. B. Hagenbuch and C. Gui, Xenobiotica, 2008, 38, 778.
60. O. Janneh, E. Jones, B. Chandler, A. Owen and S. H. Khoo, J. Anti-
microb. Chemother., 2007, 60, 987.
61. C. L. Cooper, R. P. van Heeswijk, K. Gallicano and D. W. Cameron,
Clin. Infect. Dis., 2003, 36, 1585.
448 Chapter 9
62. M. M. Foisy, E. M. Yakiwchuk and C. A. Hughes, Ann. Pharmacother.,
2008, 42, 1048.
63. B. M. Sadler, P. J. Piliero, S. L. Preston, P. P. Lloyd, Y. Lou and D. S.
Stein, Aids, 2001, 15, 1009.
64. R. Wood, J. Eron, K. Arasteh, E. Teofilo, C. Trepo, J. M. Livrozet, J.
Yeo, J. Millard, M. B. Wire and O. J. Naderer, Clin. Infect. Dis., 2004, 39,
591.
65. M. B. Wire, K. L. Baker, L. S. Jones, M. J. Shelton, Y. Lou, G. J. Thomas
and M. M. Berrey, Antimicrob. Agents Chemother., 2006, 50, 1578.
66. A. Winston, M. Bloch, A. Carr, J. Amin, P. W. Mallon, J. Ray, D.
Marriott, D. A. Cooper and S. Emery, J. Antimicrob. Chemother., 2005,
56, 380.
67. H. Khanlou, L. Bhatti and C. Farthing, J. Acquir. Immune Defic. Syndr.,
2006, 41, 124.
68. J. Molto, J. R. Santos, M. Valle, C. Miranda, J. Miranda, A. Blanco, E.
Negredo and B. Clotet, Ther. Drug. Monit., 2007, 29, 648.
69. V. J. Sekar, E. Lefebvre, M. De Pauw, T. Vangeneugden and R. M.
Hoetelmans, Br. J. Clin. Pharmacol., 2008, 66, 215.
70. D. M. Burger, R. E. Aarnoutse, J. P. Dieleman, I. C. Gyssens, J. Nouwen,
S. de Marie, P. P. Koopmans, M. Stek Jr. and M. E. van der Ende,
Antivir. Ther., 2003, 8, 455.
71. D. W. Haas, B. Johnson, J. Nicotera, V. L. Bailey, V. L. Harris, F. B.
Bowles, S. Raffanti, J. Schranz, T. S. Finn, A. J. Saah and J. Stone,
Antimicrob. Agents Chemother., 2003, 47, 2131.
72. A. Hsu, G. R. Granneman and R. J. Bertz, Clin. Pharmacokinet., 1998,
35, 275.
73. F. S. Rhame, S. L. Rawlins, R. A. Petruschke, T. A. Erb, G. A. Winchell,
H. M. Wilson, J. M. Edelman and M. A. Abramson, Antimicrob. Agents
Chemother., 2004, 48, 4200.
74. R. P. van Heeswijk, A. I. Veldkamp, R. M. Hoetelmans, J. W. Mulder, G.
Schreij, A. Hsu, J. M. Lange, J. H. Beijnen and P. L. Meenhorst, Aids,
1999, 13, F95.
75. M. Kurowski, B. Kaeser, A. Sawyer, M. Popescu and A. Mrozikiewicz,
Clin. Pharmacol. Ther., 2002, 72, 123.
76. A. Hsu, G. R. Granneman, G. Cao, L. Carothers, T. el-Shourbagy, P.
Baroldi, K. Erdman, F. Brown, E. Sun and J. M. Leonard, Clin. Phar-
macol. Ther., 1998, 63, 453.
77. C. Merry, M. G. Barry, F. Mulcahy, M. Ryan, J. Heavey, J. F. Tjia, S. E.
Gibbons, A. M. Breckenridge and D. J. Back, Aids, 1997, 11, F29.
78. N. Buss, P. Snell, J. Bock, A. Hsu and K. Jorga, Br. J. Clin. Pharmacol.,
2001, 52, 255.
79. C. J. la Porte, Y. Li, L. Beique, B. C. Foster, B. Chauhan, G. E. Garber,
D. W. Cameron and R. P. van Heeswijk, Clin. Pharmacol. Ther., 2007, 82,
389.
80. T. R. MacGregor, J. P. Sabo, S. H. Norris, P. Johnson, L. Galitz and S.
McCallister, HIV Clin. Trials, 2004, 5, 371.
Pharmacokinetics and Metabolism of Compounds that Mimic Enzyme 449
81. S. McCallister, H. Valdez, K. Curry, T. MacGregor, M. Borin, W.
Freimuth, Y. Wang and D. L. Mayers, J. Acquir. Immune Defic. Syndr.,
2004, 35, 376.
82. B. Jarvis and D. Faulds, Drugs, 1998, 56, 147.
83. W. F. Richter, B. R. Whitby and R. C. Chou, Xenobiotica, 1996, 26, 243.
84. R. B. Kim, M. F. Fromm, C. Wandel, B. Leake, A. J. Wood, D. M.
Roden and G. R. Wilkinson, J. Clin. Invest., 1998, 101, 289.
85. N. N. Salama, E. J. Kelly, T. Bui and R. J. Ho, J. Pharm. Sci., 2005, 94,
1216.
86. B. M. Best, S. L. Letendre, E. Brigid, D. B. Clifford, A. C. Collier, B. B.
Gelman, J. C. McArthur, J. A. McCutchan, D. M. Simpson, R. Ellis, E.
V. Capparelli and I. Grant, Aids, 2009, 23, 83.
87. H. E. Wynn, R. C. Brundage and C. V. Fletcher, CNS Drugs, 2002, 16,
595.
88. A. Yilmaz, A. Izadkhashti, R. W. Price, P. W. Mallon, M. De Meulder, P.
Timmerman and M. Gisslen, AIDS Res. Hum. Retroviruses, 2009, 25, 457.
89. J. Ford, M. Boffito, A. Wildfire, A. Hill, D. Back, S. Khoo, M. Nelson, G.
Moyle, B. Gazzard and A. Pozniak, Antimicrob. Agents Chemother., 2004,
48, 2388.
90. J. Ford, D. Cornforth, P. G. Hoggard, Z. Cuthbertson, E. R. Meaden, I.
Williams, M. Johnson, E. Daniels, P. Hsyu, D. J. Back and S. H. Khoo,
Antivir. Ther., 2004, 9, 77.
91. J. Ford, E. R. Meaden, P. G. Hoggard, M. Dalton, P. Newton, I. Wil-
liams, S. H. Khoo and D. J. Back, J. Antimicrob. Chemother., 2003, 52,
354.
92. O. Janneh, R. C. Hartkoorn, E. Jones, A. Owen, S. A. Ward, R. Davey,
D. J. Back and S. H. Khoo, Br. J. Pharmacol., 2008, 155, 875.
93. M. P. McRae, C. M. Lowe, X. Tian, D. L. Bourdet, R. H. Ho, B. F.
Leake, R. B. Kim, K. L. Brouwer and A. D. Kashuba, J. Pharmacol. Exp.
Ther., 2006, 318, 1068.
94. M. Caron, M. Auclair, H. Sterlingot, M. Kornprobst and J. Capeau, Aids,
2003, 17, 2437.
95. S. G. Clarke, Proc. Natl. Acad. Sci. U. S. A., 2007, 104, 13857.
96. C. Coffinier, S. E. Hudon, E. A. Farber, S. Y. Chang, C. A. Hrycyna, S.
G. Young and L. G. Fong, Proc. Natl. Acad. Sci. U. S. A., 2007, 104,
13432.
97. W. C. Earnshaw, L. M. Martins and S. H. Kaufmann, Annu. Rev. Bio-
chem., 1999, 68, 383.
98. O. E. Levy, J. E. Semple, M. L. Lim, J. Reiner, W. E. Rote, E. Dempsey,
B. M. Richard, E. Zhang, A. Tulinsky, W. C. Ripka and R. F. Nutt, J.
Med. Chem., 1996, 39, 4527.
99. J. E. Semple, D. C. Rowley, T. K. Brunck, T. Ha-Uong, N. K. Minami, T.
D. Owens, S. Y. Tamura, E. A. Goldman, D. V. Siev, R. J. Ardecky, S. H.
Carpenter, Y. Ge, B. M. Richard, T. G. Nolan, K. Hakanson, A.
Tulinsky, R. F. Nutt and W. C. Ripka, J. Med. Chem., 1996, 39, 4531.
100. S. D. Linton, Curr. Top. Med. Chem., 2005, 5, 1697.
450 Chapter 9
101. J. C. Randle, M. W. Harding, G. Ku, M. Schonharting and R. Kurrle,
Expert Opin. Investig. Drugs, 2001, 10, 1207.
102. H. Ohbayashi, Expert Opin. Investig. Drugs, 2002, 11, 965.
103. D. S. Yamashita, R. W. Marquis, R. Xie, S. D. Nidamarthy, H. J. Oh, J.
U. Jeong, K. F. Erhard, K. W. Ward, T. J. Roethke, B. R. Smith, H. Y.
Cheng, X. Geng, F. Lin, P. H. Offen, B. Wang, N. Nevins, M. S. Head, R.
C. Haltiwanger, A. A. Narducci Sarjeant, L. M. Liable-Sands, B. Zhao,
W. W. Smith, C. A. Janson, E. Gao, T. Tomaszek, M. McQueney, I. E.
James, C. J. Gress, D. L. Zembryki, M. W. Lark and D. F. Veber, J. Med.
Chem., 2006, 49, 1597.
104. F. Koizumi, M. Murakami, H. Kageyama, M. Katashima, M. Terakawa
and A. Ohnishi, Clin. Pharmacol. Ther., 1999, 66, 501.
105. J. C. Williams, R. L. Stein, R. E. Giles and R. D. Krell, Ann. N. Y. Acad.
Sci., 1991, 624, 230.
106. P. C. Davis, R. A. Wildonger and H. S. Veale, J. Pharm. Biomed. Anal.,
1993, 11, 549.
107. J. P. Burkhart, J. R. Koehl, S. Mehdi, S. L. Durham, M. J. Janusz, E. W.
Huber, M. R. Angelastro, S. Sunder, W. A. Metz and P. W. Shum, et al.,
J. Med. Chem., 1995, 38, 223.
108. M. R. Angelastro, L. E. Baugh, P. Bey, J. P. Burkhart, T. M. Chen, S. L.
Durham, C. M. Hare, E. W. Huber, M. J. Janusz and J. R. Koehl, et al., J.
Med. Chem., 1994, 37, 4538.
109. C. A. Veale, P. R. Bernstein, C. M. Bohnert, F. J. Brown, C. Bryant, J. R.
Damewood Jr., R. Earley, S. W. Feeney, P. D. Edwards, B. Gomes, J. M.
Hulsizer, B. J. Kosmider, R. D. Krell, G. Moore, T. W. Salcedo, A. Shaw,
D. S. Silberstein, G. B. Steelman, M. Stein, A. Strimpler, R. M. Thomas,
E. P. Vacek, J. C. Williams, D. J. Wolanin and S. Woolson, J. Med.
Chem., 1997, 40, 3173.
110. P. D. Edwards, D. W. Andisik, C. A. Bryant, B. Ewing, B. Gomes, J. J.
Lewis, D. Rakiewicz, G. Steelman, A. Strimpler, D. A. Trainor, P. A.
Tuthill, R. C. Mauger, C. A. Veale, R. A. Wildonger, J. C. Williams, D. J.
Wolanin and M. Zottola, J. Med. Chem., 1997, 40, 1876.
111. D. J. Kempf, C. Klein, H. J. Chen, L. L. Klein, C. Yeung, J. T. Randolph,
Y. Y. Lau, L. E. Chovan, Z. Guan, L. Hernandez, T. M. Turner, P. J.
Dandliker and K. C. Marsh, Antivir. Chem. Chemother., 2007, 18, 163.
112. J. D. Tyndall and D. P. Fairlie, Curr. Med. Chem., 2001, 8, 893.
113. S. Venkatraman, F. Velazquez, W. Wu, M. Blackman, K. X. Chen, S.
Bogen, L. Nair, X. Tong, R. Chase, A. Hart, S. Agrawal, J. Pichardo, A.
Prongay, K. C. Cheng, V. Girijavallabhan, J. Piwinski, N. Y. Shih and F.
G. Njoroge, J. Med. Chem., 2009, 52, 336.
114. K. X. Chen, L. Nair, B. Vibulbhan, W. Yang, A. Arasappan, S. L. Bogen,
S. Venkatraman, F. Bennett, W. Pan, M. L. Blackman, A. I. Padilla,
A. Prongay, K. C. Cheng, X. Tong, N. Y. Shih and F. G. Njoroge, J.
Med. Chem., 2009, 52, 1370.
115. W. Yang, X. Gao and B. Wang, Med. Res. Rev., 2003, 23, 346.
116. K. A. Koehler and G. E. Lienhard, Biochemistry, 1971, 10, 2477.
Pharmacokinetics and Metabolism of Compounds that Mimic Enzyme 451
117. M. Philipp and M. L. Bender, Proc. Natl. Acad. Sci. U.S.A., 1971, 68,
478.
118. J. M. Fevig, J. Buriak Jr., J. Cacciola, R. S. Alexander, C. A. Kettner, R.
M. Knabb, J. R. Pruitt, P. C. Weber and R. R. Wexler, Bioorg. Med.
Chem. Lett., 1998, 8, 301.
119. S. L. Lee, R. S. Alexander, A. Smallwood, R. Trievel, L. Mersinger, P. C.
Weber and C. Kettner, Biochemistry, 1997, 36, 13180.
120. J. Cacciola, J. M. Fevig, P. F. Stouten, R. S. Alexander, R. M. Knabb and
R. R. Wexler, Bioorg. Med. Chem. Lett., 2000, 10, 1253.
121. C. A. Kettner, R. Bone, D. A. Agard and W. W. Bachovchin, Biochem-
istry, 1988, 27, 7682.
122. C. A. Kettner and A. B. Shenvi, J. Biol. Chem., 1984, 259, 15106.
123. I. E. Crompton, B. K. Cuthbert, G. Lowe and S. G. Waley, Biochem. J.,
1988, 251, 453.
124. S. Ness, R. Martin, A. M. Kindler, M. Paetzel, M. Gold, S. E. Jensen, J.
B. Jones and N. C. Strynadka, Biochemistry, 2000, 39, 5312.
125. G. S. Weston, J. Blazquez, F. Baquero and B. K. Shoichet, J. Med. Chem.,
1998, 41, 4577.
126. M. R. Attwood, J. M. Bennett, A. D. Campbell, G. G. Canning, M. G.
Carr, E. Conway, R. M. Dunsdon, J. R. Greening, P. S. Jones, P. B. Kay,
B. K. Handa, D. N. Hurst, N. S. Jennings, S. Jordan, E. Keech, M. A.
O’Brien, H. A. Overton, J. King-Underwood, T. M. Raynham, K. P.
Stenson, C. S. Wilkinson, T. C. Wilkinson and F. X. Wilson, Antivir.
Chem. Chemother., 1999, 10, 259.
127. R. M. Dunsdon, J. R. Greening, P. S. Jones, S. Jordan and F. X. Wilson,
Bioorg. Med. Chem. Lett., 2000, 10, 1577.
128. M. P. Groziak, Am. J. Ther., 2001, 8, 321.
129. N. N. Kim, J. D. Cox, R. F. Baggio, F. A. Emig, S. K. Mistry, S. L.
Harper, D. W. Speicher, S. M. Morris Jr, D. E. Ash, A. Traish and D. W.
Christianson, Biochemistry, 2001, 40, 2678.
130. J. Das and S. D. Kimball, Bioorg. Med. Chem., 1995, 3, 999.
131. T. Mc Cormack, W. Baumeister, L. Grenier, C. Moomaw, L. Plamondon,
B. Pramanik, C. Slaughter, F. Soucy, R. Stein, F. Zuhl and L. Dick, J.
Biol. Chem., 1997, 272, 26103.
132. H. Saitoh and B. J. Aungst, Pharm. Res., 1999, 16, 1786.
133. B. D. Dorsey, M. Iqbal, S. Chatterjee, E. Menta, R. Bernardini, A.
Bernareggi, P. G. Cassara, G. D’Arasmo, E. Ferretti, S. De Munari, A.
Oliva, G. Pezzoni, C. Allievi, I. Strepponi, B. Ruggeri, M. A. Ator, M.
Williams and J. P. Mallamo, J. Med. Chem., 2008, 51, 1068.
134. J. Ruef and H. A. Katus, Expert Opin. Investig. Drugs, 2003, 12, 781.
135. A. Schwienhorst, Cell. Mol. Life Sci., 2006, 63, 2773.
136. E. C. Attar, D. J. De Angelo, J. G. Supko, F. D’Amato, D. Zahrieh, A.
Sirulnik, M. Wadleigh, K. K. Ballen, S. McAfee, K. B. Miller, J. Levine,
I. Galinsky, E. G. Trehu, D. Schenkein, D. Neuberg, R. M. Stone and P.
C. Amrein, Clin. Cancer Res., 2008, 14, 1446.
137. D. Leveque, M. C. Carvalho and F. Maloisel, In vivo, 2007, 21, 273.
452 Chapter 9
138. Y. Ogawa, K. Tobinai, M. Ogura, K. Ando, T. Tsuchiya, Y. Kobayashi,
T. Watanabe, D. Maruyama, Y. Morishima, Y. Kagami, H. Taji, H.
Minami, K. Itoh, M. Nakata and T. Hotta, Cancer Sci., 2008, 99, 140.
139. T. Pekol, J. S. Daniels, J. Labutti, I. Parsons, D. Nix, E. Baronas, F.
Hsieh, L. S. Gan and G. Miwa, Drug Metab. Dispos., 2005, 33, 771.
140. H. C. Brown, M. M. Midland and G. W. Kabalka, J. Am. Chem. Soc.,
1971, 93, 1024.
141. S. B. Mirviss, J. Org. Chem., 1967, 32, 1713.
142. S. Wu, W. Waugh and V. J. Stella, J. Pharm. Sci., 2000, 89, 758.
143. V. Uttamsingh, C. Lu, G. Miwa and L. S. Gan, Drug Metab. Dispos.,
2005, 33, 1723.
144. C. Lu, R. Gallegos, P. Li, C. Q. Xia, S. Pusalkar, V. Uttamsingh, D. Nix,
G. T. Miwa and L. S. Gan, Drug Metab. Dispos., 2006, 34, 702.
145. J. Labutti, I. Parsons, R. Huang, G. Miwa, L. S. Gan and J. S. Daniels,
Chem. Res. Toxicol., 2006, 19, 539.
146. P. F. Bross, R. Kane, A. T. Farrell, S. Abraham, K. Benson, M. E.
Brower, S. Bradley, J. V. Gobburu, A. Goheer, S. L. Lee, J. Leighton, C.
Y. Liang, R. T. Lostritto, W. D. McGuinn, D. E. Morse, A. Rahman, L.
A. Rosario, S. L. Verbois, G. Williams, Y. C. Wang and R. Pazdur, Clin.
Cancer Res., 2004, 10, 3954.
147. R. A. Morrison, S. M. Singhvi, A. E. Peterson, D. A. Pocetti and B. H.
Migdalof, Drug Metab. Dispos., 1990, 18, 253.
148. S. M. Singhvi, K. L. Duchin, R. A. Morrison, D. A. Willard, D. W.
Everett and M. Frantz, Br. J. Clin. Pharmacol., 1988, 25, 9.
149. I. R. McNicholl and J. J. McNicholl, Ann. Pharmacother., 2001, 35, 57.
150. K. Klumpp and B. J. Graves, Curr. Top. Med. Chem., 2006, 6, 423.
151. D. B. Mendel, C. Y. Tai, P. A. Escarpe, W. Li, R. W. Sidwell, J. H.
Huffman, C. Sweet, K. J. Jakeman, J. Merson, S. A. Lacy, W. Lew, M. A.
Williams, L. Zhang, M. S. Chen, N. Bischofberger and C. U. Kim,
Antimicrob. Agents Chemother., 1998, 42, 640.
152. W. M. Kati, A. S. Saldivar, F. Mohamadi, H. L. Sham, W. G. Laver and
W. E. Kohlbrenner, Biochem. Biophys. Res. Commun., 1998, 244, 408.
153. A. W. Peng, E. K. Hussey and K. H. Moore, J. Clin. Pharmacol., 2000, 40,
242.
154. G. He, J. Massarella and P. Ward, Clin. Pharmacokinet., 1999, 37, 471.
155. K. Morimoto, M. Nakakariya, Y. Shirasaka, C. Kakinuma, T. Fujita, I.
Tamai and T. Ogihara, Drug Metab. Dispos., 2008, 36, 6.
156. C. Oo, P. Snell, J. Barrett, A. Dorr, B. Liu and I. Wilding, Int. J. Pharm.,
2003, 257, 297.
157. W. Li, P. A. Escarpe, E. J. Eisenberg, K. C. Cundy, C. Sweet, K. J.
Jakeman, J. Merson, W. Lew, M. Williams, L. Zhang, C. U. Kim,
N. Bischofberger, M. S. Chen and D. B. Mendel, Antimicrob. Agents
Chemother., 1998, 42, 647.
158. K. Beaumont, R. Webster, I. Gardner and K. Dack, Curr. Drug Metab.,
2003, 4, 461.
159. D. A. Smith, Curr. Opin. Drug Discov. Devel., 2007, 10, 550.
Pharmacokinetics and Metabolism of Compounds that Mimic Enzyme 453
160. M. J. Daniel, J. M. Barnett and B. A. Pearson, Clin. Pharmacokinet.,
1999, 36(Suppl 1), 41.
161. T. Yamanaka, M. Yamada, K. Tsujimura, T. Kondo, S. Nagata, S.
Hobo, M. Kurosawa and T. Matsumura, J. Vet. Med. Sci., 2007, 69,
293.
162. S. Toovey, C. Rayner, E. Prinssen, T. Chu, B. Donner, B. Thakrar, R.
Dutkowski, G. Hoffmann, A. Breidenbach, L. Lindemann, E. Carey, L.
Boak, R. Gieschke, S. Sacks, J. Solsky, I. Small and D. Reddy, Drug Saf.,
2008, 31, 1097.
163. A. Ose, H. Kusuhara, K. Yamatsugu, M. Kanai, M. Shibasaki, T. Fujita,
A. Yamamoto and Y. Sugiyama, Drug Metab. Dispos., 2008, 36, 427.
164. S. S. Jhee, M. Yen, L. Ereshefsky, M. Leibowitz, M. Schulte, B. Kaeser,
L. Boak, A. Patel, G. Hoffmann, E. P. Prinssen and C. R. Rayner,
Antimicrob. Agents Chemother., 2008, 52, 3687.
165. G. Hill, T. Cihlar, C. Oo, E. S. Ho, K. Prior, H. Wiltshire, J. Barrett, B.
Liu and P. Ward, Drug Metab. Dispos., 2002, 30, 13.
166. M. Holodniy, S. R. Penzak, T. M. Straight, R. T. Davey, K. K. Lee, M. B.
Goetz, D. W. Raisch, F. Cunningham, E. T. Lin, N. Olivo and L. R.
Deyton, Antimicrob. Agents Chemother., 2008, 52, 3013.
167. K. McClellan and C. M. Perry, Drugs, 2001, 61, 263.
168. D. J. Sweeny, G. Lynch, A. M. Bidgood, W. Lew, K. Y. Wang and K. C.
Cundy, Drug Metab. Dispos., 2000, 28, 737.
169. N. Lindegardh, G. R. Davies, T. H. Tran, J. Farrar, P. Singhasivanon, N.
P. Day and N. J. White, Antimicrob. Agents Chemother., 2006, 50, 3197.
170. D. Yang, R. E. Pearce, X. Wang, R. Gaedigk, Y. J. Wan and B. Yan,
Biochem. Pharmacol., 2009, 77, 238.
171. D. Shi, J. Yang, D. Yang, E. L. LeCluyse, C. Black, L. You, F. Akhlaghi
and B. Yan, J. Pharmacol. Exp. Ther., 2006, 319, 1477.
172. S. J. Macdonald, R. Cameron, D. A. Demaine, R. J. Fenton, G. Foster,
D. Gower, J. N. Hamblin, S. Hamilton, G. J. Hart, A. P. Hill, G. G.
Inglis, B. Jin, H. T. Jones, D. B. McConnell, J. McKimm-Breschkin, G.
Mills, V. Nguyen, I. J. Owens, N. Parry, S. E. Shanahan, D. Smith, K. G.
Watson, W. Y. Wu and S. P. Tucker, J. Med. Chem., 2005, 48, 2964.
173. G. B. Evans, R. H. Furneaux, B. Greatrex, A. S. Murkin, V. L. Schramm
and P. C. Tyler, J. Med. Chem., 2008, 51, 948.
174. M. D. Erion, J. D. Stoeckler, W. C. Guida, R. L. Walter and S. E. Ealick,
Biochemistry, 1997, 36, 11735.
175. M. D. Erion, K. Takabayashi, H. B. Smith, J. Kessi, S. Wagner, S.
Honger, S. L. Shames and S. E. Ealick, Biochemistry, 1997, 36, 11725.
176. R. W. Miles, P. C. Tyler, R. H. Furneaux, C. K. Bagdassarian and V. L.
Schramm, Biochemistry, 1998, 37, 8615.
177. J. D. Stoeckler, A. F. Poirot, R. M. Smith, R. E. Parks Jr., S. E. Ealick, K.
Takabayashi and M. D. Erion, Biochemistry, 1997, 36, 11749.
178. S. Bantia, J. A. Montgomery, H. G. Johnson and G. M. Walsh, Immu-
nopharmacology, 1996, 35, 53.
179. V. L. Schramm, Biochim. Biophys. Acta, 2002, 1587, 107.
454 Chapter 9
180. R. G. Silva, J. E. Nunes, F. Canduri, J. C. Borges, L. M. Gava, F. B.
Moreno, L. A. Basso and D. S. Santos, Curr. Drug Targets, 2007, 8, 413.
181. S. Banti, P. J. Miller, C. D. Parker, S. L. Ananth, L. L. Horn, Y. S. Babu
and J. S. Sandhu, Int. Immunopharmacol., 2002, 2, 913.
182. S. Bantia, S. L. Ananth, C. D. Parker, L. L. Horn and R. Upshaw, Int.
Immunopharmacol., 2003, 3, 879.
183. S. Bantia, P. J. Miller, C. D. Parker, S. L. Ananth, L. L. Horn, J. M.
Kilpatrick, P. E. Morris, T. L. Hutchison, J. A. Montgomery and J. S.
Sandhu, Int. Immunopharmacol., 2001, 1, 1199.
184. J. P. Jenuth, J. E. Dilay, E. Fung, E. R. Mably and F. F. Snyder, Adv.
Exp. Med. Biol., 1991, 309B, 273.
185. M. L. Markert, Immunodefic. Rev., 1991, 3, 45.
186. E. A. Taylor Ringia and V. L. Schramm, Curr. Top. Med. Chem., 2005, 5,
1237.
187. G. B. Evans, R. H. Furneaux, T. L. Hutchison, H. S. Kezar, P. E. Morris
Jr., V. L. Schramm and P. C. Tyler, J. Org. Chem., 2001, 66, 5723.
188. G. B. Evans, R. H. Furneaux, A. Lewandowicz, V. L. Schramm and P. C.
Tyler, J. Med. Chem., 2003, 46, 3412.
189. G. A. Kicska, P. C. Tyler, G. B. Evans, R. H. Furneaux, K. Kim and V.
L. Schramm, J. Biol. Chem., 2002, 277, 3219.
190. A. Lewandowicz, E. A. Ringia, L. M. Ting, K. Kim, P. C. Tyler, G. B.
Evans, O. V. Zubkova, S. Mee, G. F. Painter, D. H. Lenz, R. H. Fur-
neaux and V. L. Schramm, J. Biol. Chem., 2005, 280, 30320.
191. V. L. Schramm, Arch. Biochem. Biophys., 2005, 433, 13.
192. A. Lewandowicz, P. C. Tyler, G. B. Evans, R. H. Furneaux and V. L.
Schramm, J. Biol. Chem., 2003, 278, 31465.
193. J. M. Kilpatrick, P. E. Morris, D. G. Serota Jr, D. Phillips, D. R. Moore,
J. C. Bennett and Y. S. Babu, Int. Immunopharmacol., 2003, 3, 541.
194. C. M. Galmarini, IDrugs, 2006, 9, 712.
195. V. Gandhi, J. M. Kilpatrick, W. Plunkett, M. Ayres, L. Harman, M. Du,
S. Bantia, J. Davisson, W. G. Wierda, S. Faderl, H. Kantarjian and D.
Thomas, Blood, 2005, 106, 4253.
196. H. S. Kezar 3rd, J. M. Kilpatrick, D. Phillips, D. Kellogg, J. Zhang and P.
E. Morris Jr., Nucleosides Nucleotides Nucleic Acids, 2005, 24, 1817.
197. B. J. Kane, J. G. Kuhn and M. K. Roush, Ann. Pharmacother., 1992, 26,
939.
198. C. Lathia, G. F. Fleming, M. Meyer, M. J. Ratain and L. Whitfield,
Cancer Chemother. Pharmacol., 2002, 50, 121.
199. J. D. Geiger, J. L. Lewis, C. J. MacIntyre and J. I. Nagy, Neuro-
pharmacology, 1987, 26, 1383.
200. S. B. Kaye, Br. J. Cancer, 1998, 78(Suppl 3), 1.
201. W. Kreis, T. M. Woodcock, M. B. Meyers, L. A. Carlevarini and I. H.
Krakoff, Cancer Treat. Rep., 1977, 61, 723.
202. S. Urien, K. Rezai and F. Lokiec, J. Pharmacokinet. Pharmacodyn., 2005,
32, 817.
203. S. Grant, K. Bhalla and C. McCrady, Leuk. Res., 1991, 15, 205.
Pharmacokinetics and Metabolism of Compounds that Mimic Enzyme 455
204. G. L. Neil, T. E. Moxley, S. L. Kuentzel, R. C. Manak and L. J. Hanka,
Cancer Chemother. Rep., 1975, 59, 459.
205. S. Xiang, S. A. Short, R. Wolfenden and C. W. Carter Jr, Biochemistry,
1997, 36, 4768.
206. C. W. Carter Jr, Biochimie, 1995, 77, 92.
207. W. M. Kati and R. Wolfenden, Science, 1989, 243, 1591.
208. H. Guo, N. Rao, Q. Xu and H. Guo, J. Am. Chem. Soc., 2005, 127, 3191.
209. C. B. Yoo, J. C. Cheng and P. A. Jones, Biochem. Soc. Trans., 2004, 32,
910.
210. L. Zhou, X. Cheng, B. A. Connolly, M. J. Dickman, P. J. Hurd and D. P.
Hornby, J. Mol. Biol., 2002, 321, 591.
211. R. M. Cohen and R. Wolfenden, J. Biol. Chem., 1971, 246, 7561.
212. J. Laliberte, V. E. Marquez and R. L. Momparler, Cancer Chemother.
Pharmacol., 1992, 30, 7.
213. N. Eliopoulos, D. Cournoyer and R. L. Momparler, Cancer Chemother.
Pharmacol., 1998, 42, 373.
214. J. L. Holleran, R. A. Parise, E. Joseph, J. L. Eiseman, J. M. Covey, E. R.
Glaze, A. V. Lyubimov, Y. F. Chen, D. Z. D’Argenio and M. J. Egorin,
Clin. Cancer Res., 2005, 11, 3862.
215. J. H. Beumer, E. Joseph, M. J. Egorin, R. S. Parker, Z. D’Argenio D, J.
M. Covey and J. L. Eiseman, Clin. Cancer Res., 2006, 12, 5826.
216. D. H. Ho, G. P. Bodey, S. W. Hall, R. S. Benjamin, N. S. Brown, E. J.
Freireich and T. L. Loo, J. Clin. Pharmacol., 1978, 18, 259.
217. J. H. Beumer, J. L. Eiseman, R. A. Parise, J. A. Florian Jr., E. Joseph, D.
Z. D’Argenio, R. S. Parker, B. Kay, J. M. Covey and M. J. Egorin,
Cancer Chemother. Pharmacol., 2008, 62, 457.
218. S. M. el-Dareer, V. White, F. P. Chen, L. B. Mellett and D. L. Hill,
Cancer Treat. Rep., 1976, 60, 1627.
219. Z. Qi, I. Whitt, A. Mehta, J. Jin, M. Zhao, R. C. Harris, A. B. Fogo and
M. D. Breyer, Am. J. Physiol. Renal Physiol., 2004, 286, F590.
220. R. W. Klecker, R. L. Cysyk and J. M. Collins, Bioorg. Med. Chem., 2006,
14, 62.
221. S. Yoshihara and K. Tatsumi, Arch. Biochem. Biophys., 1997, 338, 29.
222. C. Beedham, Prog. Med. Chem., 1987, 24, 85.
223. K. Nagai, K. Nagasawa and S. Fujimoto, Biochem. Biophys. Res. Com-
mun., 2005, 334, 1343.
224. P. P. Wong, V. E. Currie, R. W. Mackey, I. H. Krakoff, C. T. Tan, J. H.
Burchenal and C. W. Young, Cancer Treat. Rep., 1979, 63, 1245.
225. W. Kreis, T. M. Woodcock, C. S. Gordon and I. H. Krakoff, Cancer
Treat. Rep., 1977, 61, 1347.
226. G. J. Peters, C. J. van Groeningen, E. J. Laurensse, J. Lankelma, A. Leyva
and H. M. Pinedo, Cancer Chemother. Pharmacol., 1987, 20, 101.
227. C. J. van Groeningen, A. Leyva, I. Kraal, G. J. Peters and H. M. Pinedo,
Cancer Treat. Rep., 1986, 70, 745.
228. A. Leyva, C. J. van Groeningen, I. Kraal, H. Gall, G. J. Peters, J. Lan-
kelma and H. M. Pinedo, Cancer Res., 1984, 44, 5928.
456 Chapter 9
229. G. P. Connolly and J. A. Duley, Trends Pharmacol. Sci., 1999, 20, 218.
230. D. M. Becroft, L. I. Phillips and A. Simmonds, J. Pediatr., 1969, 75, 885.
231. H. Koyama and M. Tsuji, Biochem. Pharmacol., 1983, 32, 3547.
232. L. Gan, M. R. Seyedsayamdost, S. Shuto, A. Matsuda, G. A. Petsko and
L. Hedstrom, Biochemistry, 2003, 42, 857.
233. T. Inou, R. Kusaba, I. Takahashi, H. Sugimoto, K. Kuzuhara, Y.
Yamada, J. Yamauchi and O. Otsubo, Transplant Proc., 1981, 13, 315.
234. F. Marumo, M. Okubo, K. Yokota, H. Uchida, K. Kumano, T. Endo, K.
Watanabe and N. Kashiwagi, Transplant Proc., 1988, 20, 406.
235. K. Mita, N. Akiyama, T. Nagao, H. Sugimoto, S. Inoue, T. Osakabe, Y.
Nakayama, K. Yokota, K. Sato and H. Uchida, Transplant Proc., 1990,
22, 1679.
236. A. Tajima, M. Hata, N. Ohta, Y. Ohtawara, K. Suzuki and Y. Aso,
Transplantation, 1984, 38, 116.
237. M. Honda, H. Itoh, T. Suzuki and Y. Hashimoto, Biol. Pharm. Bull.,
2006, 29, 2460.
238. M. Okada, K. Suzuki, M. Nakashima, T. Nakanishi and N. Fujioka, Eur.
J. Pharmacol., 2006, 531, 140.
239. S. D. Patil, L. Y. Ngo, P. Glue and J. D. Unadkat, Pharm. Res., 1998, 15,
950.
240. S. A. Gruber, G. R. Erdmann, B. A. Burke, A. Moss, L. Bowers, W. J.
Hrushesky, R. J. Cipolle, D. M. Canafax and A. J. Matas, Transplanta-
tion, 1992, 53, 12.
241. R. Kusaba, O. Otubo, H. Sugimoto, I. Takahashi, Y. Yamada, J.
Yamauchi, N. Akiyama and T. Inou, Proc. Eur. Dia.l Transplant. Assoc.,
1981, 18, 420.
242. Y. Kokado, S. Takahara, M. Ishibashi and A. Okuyama, Transplant.
Proc., 1994, 26, 2111.
243. D. Stypinski, M. Obaidi, M. Combs, M. Weber, A. J. Stewart and H.
Ishikawa, Br. J. Clin. Pharmacol., 2007, 63, 459.
244. P. J. O’Dwyer, Pharmacol. Ther., 1990, 48, 371.
245. K. D. Collins and G. R. Stark, J. Biol. Chem., 1971, 246, 6599.
246. R. S. Obach, F. Lombardo and N. J. Waters, Drug Metab. Dispos., 2008,
36, 1385.
247. www.emea.europa.eu/humandocs/PDFs/EPAR/velcade/velcade.htm.
248. C. H. Gu, H. Li, J. Levons, K. Lentz, R. B. Gandhi, K. Raghavan and R.
L. Smith, Pharm. Res., 2007, 24, 1118.
249. N. Tam-Zaman, Y. K. Tam, S. Tawfik and H. Wiltshire, Pharm. Res.,
2004, 21, 436.
250. J. O. Baker, S. H. Wilkes, M. E. Bayliss and J. M. Prescott, Biochemistry,
1983, 22, 2098.
251. A. H. Brockman, P. Hatsis, M. Paton and J. T. Wu, Anal. Chem., 2007, 4,
1599.
252. Y. Wang, N. Mealy, N. Serradell, J. Bolos and E. Rosa, Drugs Future,
2007, 32, 310.
253. F. Fyhrquist and O. Saijonmaa, J. Intern. Med., 2008, 264, 224.
Pharmacokinetics and Metabolism of Compounds that Mimic Enzyme 457
254. A. G. Tzakos, S. Galanis, G. A. Spyroulias, P. Cordopatis, E. Manessi-
Zoupa and I. P. Gerothanassis, Prot. Engin., 2003, 16, 993.
255. C. Shu, H. Shen, U. Hopfer and D. E. Smith, Drug Metab. Dispos., 2001,
29, 1307.
256. I. Knutter, C. Wollesky, G. Kottra, M. G. Hahn, W. Fischer, K. Zebisch,
R. H. Neubert, H. Daniel and M. Brandsch, J. Pharmacol. Exp. Ther.,
2008, 327, 432.
257. R. W. Klecker, R. L. Cysyk and J. M. Collins, Bioorg. Med. Chem., 2006,
14, 62.
258. J. Murase, K. Mizuno, K. Kawai, S. Nishiumi, Y. Kobayashi, M.
Hayashi, T. Morino, T. Suzuki and S. Baba, Oyo Yakuri, 1978, 15, 829.
259. J. M. Kolesar, A. K. Morris and J. G. Kuhn, J. Oncol. Pharm. Pract., ,
1996, 2, 211.
260. E. M. del Amo, A. T. Heikkinen and J. Monkkonen, Eur. J. Pharm. Sci.,
2009, 36, 200.
261. H. H. Usansky, P. Hu and P. J. Sinko, Drug Metab. Dispos., 2008, 36,
863.
262. W. L. Chiou, S. M. Chung, T. C. Wu and C. Ma, Int. J. Clin. Pharmacol.
Ther., 2001, 39, 93.
263. K. C. Yeh, J. A. Stone, A. D. Carides, P. Rolan, E. Woolf and W. D. Ju,
J. Pharm. Sci., 1999, 88, 568.
264. M. Vermeir, S. Lachau-Durand, G. Mannens, F. Cuyckens, B. van Hoof
and A. Raoof, Drug Metab. Dispos., 2009, 37, 809.
265. C. Li, T. Liu, L. Broske, J. M. Brisson, A. S. Uss, F. G. Njoroge, R. A.
Morrison and K. C. Cheng, Biochem. Pharmacol., 2008, 76, 1757.
266. P. Revill, N. Serradell, J. Bolos and E. Rosa, Drugs Future, 2007, 32, 788–
798.
267. S. Letendre, J. Marquie-Beck, E. Capparelli, B. Best, D. Clifford, A. C.
Collier, B. B. Gelman, J. C. McArthur, J. A. McCutchan, S. Morgello, D.
Simpson, I. Grant and R. J. Ellis, Arch. Neurol., 2008, 65, 65.
268. L. Varatharajan and S. A. Thomas, Antiviral. Res., 2009, 82, A99.
269. W. C. Ripka, Curr. Opin. Chem. Biol., 1997, 1, 242.
270. O. Okusanya, A. Forrest, R. DiFrancesco, S. Bilic, S. Rosenkranz, M. F.
Para, E. Adams, K. E. Yarasheski, R. C. Reichman and G. D. Morse,
Antimicrob. Agents Chemother., 2007, 51, 1822.
271. D. R. Goldsmith and C. M. Perry, Drugs, 2003, 63, 1679.
272. M. Rittweger and K. Arasteh, Clin. Pharmacokinet., 2007, 46, 739.
273. J. H. Lin, M. Chiba, S. K. Balani, I. W. Chen, G. Y. Kwei, K. J. Vastag
and J. A. Nishime, Drug Metab. Dispos., 1996, 24, 1111.
274. J. H. Lin, M. Chiba, I. W. Chen, J. A. Nishime, F. A. deLuna, M.
Yamazaki and Y. J. Lin, Drug Metab. Dispos., 1999, 27, 1187.
275. J. H. Lin, M. Chiba, I. W. Chen, J. A. Nishime and K. J. Vastag, Drug
Metab. Dispos., 1996, 24, 1298.
276. A. Kiriyama, T. Nishiura, H. Yamaji and K. Takada, Biopharm. Drug
Dispos., 1999, 20, 199.
458 Chapter 9
277. R. W. Humphrey, K. M. Wyvill, B. Y. Nguyen, L. E. Shay, D. R. Kohler,
S. M. Steinberg, T. Ueno, T. Fukasawa, M. Shintani, H. Hayashi, H.
Mitsuya and R. Yarchoan, Antiviral Res., 1999, 41, 21.
278. B. U. Mueller, B. D. Anderson, M. Q. Farley, R. Murphy, J. Zuckerman,
P. Jarosinski, K. Godwin, C. L. McCully, H. Mitsuya, P. A. Pizzo and F.
M. Balis, Antimicrob. Agents Chemother., 1998, 42, 1815.
279. B. V. Shetty, M. B. Kosa, D. A. Khalil and S. Webber, Antimicrob. Agents
Chemother., 1996, 40, 110.
280. K. A. Jackson, S. E. Rosenbaum, B. M. Kerr, Y. K. Pithavala, G. Yuen
and M. N. Dudley, Antimicrob. Agents Chemother., 2000, 44, 1832.
281. http://www.accessdata.fda.gov/drugsatfda_docs/nda/99/20-945.pdf_Ritonovir_
Prntlbl.pdf..
282. D. J. Kempf, K. C. Marsh, J. F. Denissen, E. McDonald, S. Vasava-
nonda, C. A. Flentge, B. E. Green, L. Fino, C. H. Park and X. P. Kong
et al., Proc. Natl. Acad. Sci. U. S. A., 1995, 92, 2484.
283. J. F. Denissen, B. A. Grabowski, M. K. Johnson, A. M. Buko, D. J.
Kempf, S. B. Thomas and B. W. Surber, Drug Metab. Dispos., 1997, 25,
489.
284. R. Lledo-Garcia, A. Nacher, L. Prats-Garcia, V. G. Casabo and M.
Merino-Sanjuan, J. Pharm. Sci., 2007, 96, 633.
285. A. Billich, G. Fricker, I. Muller, P. Donatsch, P. Ettmayer, H. Gstach, P.
Lehr, P. Peichl, D. Scholz and B. Rosenwirth, Antimicrob. Agents Che-
mother., 1995, 39, 1406.
286. S. Thaisrivongs and J. W. Strohbach, Biopolymers, 1999, 51, 51.
287. J. D. Courter, J. E. Girotto and J. C. Salazar, Expert Rev. Anti Infect.
Ther., 2008, 6, 797.
288. S. R. Turner, J. W. Strohbach, R. A. Tommasi, P. A. Aristoff, P. D.
Johnson, H. I. Skulnick, L. A. Dolak, E. P. Seest, P. K. Tomich, M. J.
Bohanon, M. M. Horng, J. C. Lynn, K. T. Chong, R. R. Hinshaw, K. D.
Watenpaugh, M. N. Janakiraman and S. Thaisrivongs, J. Med. Chem.,
1998, 41, 3467.
289. H. D. Kleinert, H. H. Stein, S. Boyd, A. K. Fung, W. R. Baker, K. M.
Verburg, J. S. Polakowski, P. Kovar, J. Barlow and J. Cohen et al.,
Hypertension, 1992, 20, 768.
290. M. Azizi, R. Webb, J. Nussberger and N. K. Hollenberg, J. Hypertens.,
2006, 24, 243.
291. G. A. Rongen, J. W. Lenders, P. Smits and T. Thien, Clin. Pharmacoki-
net., 1995, 29, 6.
292. F. H. Derkx, A. H. van den Meiracker, W. Fischli, P. J. Admiraal, A. J.
Man in’t Veld, P. van Brummelen and M. A. Schalekamp, Am. J.
Hypertens., 1991, 4, 602.
293. M. Shibasaki, T. Usui, O. Inagaki, M. Asano and T. Takenaka, J. Pharm.
Pharmacol., 1994, 46, 68.
294. A. Yilmaz, L. Stahle, L. Hagberg, B. Svennerholm, D. Fuchs and M.
Gisslen, Scand. J. Infect. Dis., 2004, 36, 823.
Pharmacokinetics and Metabolism of Compounds that Mimic Enzyme 459
295. Y. Khaliq, K. Gallicano, S. Venance, S. Kravcik and D. W. Cameron,
Clin. Pharmacol. Ther., 2000, 68, 637.
296. R. B. Perni, S. J. Almquist, R. A. Byrn, G. Chandorkar, P. R. Chaturvedi,
L. F. Courtney, C. J. Decker, K. Dinehart, C. A. Gates, S. L. Harbeson,
A. Heiser, G. Kalkeri, E. Kolaczkowski, K. Lin, Y. P. Luong, B. G. Rao,
W. P. Taylor, J. A. Thomson, R. D. Tung, Y. Wei, A. D. Kwong and C.
Lin, Antimicrob. Agents Chemother., 2006, 50, 899.
297. P. Revill, N. Serradell, J. Bolos and E. Rosa, Drugs of the Future, 2007,
32, 310.
298. Y. Wang, N. Mealy, N. Serradell, J. Bolos and E. Rosa, Drugs of the
Future, 2007, 32, 310.
299. M. O. Vallez, A. Rupin, D. Versluys, G. De Nantuil, and T. Verbeuren, J.
Thromb. Haemost., 1999, Suppl, Abst 2284.
300. Y.-J. Chyan and L. Ming, Recent Patents on Endocrine, Metabolic and
Immune Drug Discovery, 2007, 1, 15–24.
CHAPTER 10
460
Alcohols and Phenols: Absorption, Distribution, Metabolism and Excretion 461
solubility. Also, by decreasing lipophilicity, the addition of a hydroxyl group
will generally decrease tissue partitioning. The increase in hydrophilicity and
hydrogen-bonding capacity can also decrease membrane penetration. If not
oriented toward direct interaction with substituents on the macromolecule
target, a hydroxyl group can reside in solvent-exposed space. With these diverse
capabilities, the hydroxyl group offers the medicinal chemist a structural
modification that can substantially alter the behaviour of a molecule at the
target and in the body (Table 10.1).
The presence of the hydroxyl group in drugs is high, with 40 of the 150 most
frequently prescribed medications containing an alcohol or phenol function,
most of which are taken orally (Table 10.2). These drugs span a range of
indications including those which require penetration into the central nervous
system (CNS), supporting the notion that a hydroxyl group does not preclude
penetration across membranes or the blood–brain barrier.
Table 10.1 Average physicochemical properties for a set of fifteen drugs and
their hydroxy analogies.
Property Without -OH With -OH
Table 10.2 Prevalence of drugs containing an OH group among the top 150
most frequently prescribed drugs.
Drug Indication Drug Indication
hydroxyl group will decrease lipophilicity and increase PSA (Table 10.1), and
thus it is expected that tissue distribution will decrease and plasma protein
binding will also decrease. Hence, the impact on VD will depend on the relative
effect that addition of the hydroxyl will have on these two opposing properties.
After correcting for differences in plasma protein binding for drugs and their
hydroxyl containing analogues, the free VD values for the hydroxyl containing
set are generally lower than their non-hydroxyl counterparts. This reflects the
notion that addition of hydroxyl will decrease tissue partitioning (Table 10.4).
CL and VD determine the half-life (t1/2). Since the addition of a hydroxyl
group can impact both these properties, one may not be able to necessarily
predict the impact on half-life when a hydroxyl group is added during drug
design. The addition of the hydroxyl group could direct a compound to drug
metabolising enzymes that ordinarily would not act upon the parent, while at
the same time decreasing the rate of metabolism by P450 enzymes. For any new
chemical scaffold being used in drug design efforts, this determination needs to
be made experimentally, although there are increasingly improving computa-
tional approaches to predicting VD and plasma protein binding, as well as
interactions with drug-metabolising enzymes. The impact of a hydroxyl group
on t1/2 for a number of drugs is listed in Table 10.3. Comparisons of CL, VD
and t1/2 for hydroxyl containing compounds and their counterparts are shown
in Figure 10.1. It should be noted that, while the impact on CL and VD is
varied, the addition of a hydroxyl more often leads to a decrease in t1/2.
100
(mL/min/kg)
0.1
0.1 1 10 100
CL for non -OH compounds (mL/min/kg)
VD for -OH containing analogues (L/kg)
100
10
0.1
0.1 1 10 100
VD for non -OH compounds (L/kg)
100
t1/2 for -OH containing analogues (hr)
10
0.1
0.1 1 10 100
t1/2 for non-OH containing compounds (hr)
Figure 10.1 Comparison of CL, VD, and t1/.2 for compounds possessing and lacking
a hydroxyl group. The median CL for the OH-containing is approxi-
mately the same as for the non-OH counterparts, albeit there may be a
trend for high CL vs. low CL compound pairs. For VD, the median value
for the OH-containing compounds is 85% of the value for the non-OH
compounds. For half-life, the median value for the OH-containing
compounds is 60% of that of the non-OH counterparts.
466 Chapter 10
pharmacokinetic characteristics of that compound and, together with the
specifics regarding the interaction of that drug and its intended target, will
determine its overall efficacy and usefulness for treating disease. Thus, during
the drug design process, consideration of the potential impact of the addition or
removal of a hydroxyl group on these interactions is important. The questions
that need to be asked include:
Ethambutol
H3C H3C
N N O
N O ADH/ALDH N O
OH OH
Hydroxyzine Cetirizine
OH
O NH2 Hydrolysis
O NH2
O NH2
O
O
ADH
Felbamate
H
H
-CO2, + NH2 O
O
CH 2 O NH 2
O
Atropaldehyde
CF3 CF3
CF3
N N
N HO HO N
N N
CYP2C9 ADH1/ADH2
O
O S O O S O
O S O
NH2 NH2
NH2
Celecoxib
OH H O HO O
N ADH N ADH N
N N N
N N N
H 2N N NH H 2N N H 2N N NH
NH
Abacvir
two-electron oxidation
P450/PGHS
PGHS PGHS
OH O O
2x semiquinone
GSH GSH
2x
NHCOCH3
NHCOCH3
NCOCH3
OH SG
OH
OH
O
NHCOCH3
OH
Cl Cl
N N
OH
N N
O O
Diazepam Temazepam
(i.e. T1/2 and CL are 4.0 h and 45 mL min1 kg1 for temazepam, and 1.1 h and
80 mL min1 kg1 for diazepam, respectively).41,42 In humans, although the
major metabolite of temazepam is also direct glucuronide, the excretion of the
glucuronide is mainly through urine. The enterohepatic circulation is minimal,
so the half-life is 8–10 h for temazepam and 32–33 h for diazepam.41,43
A single hydroxy group in the molecule makes significant difference in the
pharmacokinetics and pharmacodynamics of temazepam and diazepam. The
former is categorised as a short-acting benzodiazepine used mainly as a hyp-
notic, and the latter is the long-acting one used for the treatment of anxiety.
1) Catecholamin
HO NH2 O NH2
COMT
HO SAM SAH
HO
Dopamine 3-Methoxytyramine
2) Catechoestrogen
O
O
COMT
O
HO
SAM SAH
HO
HO
2-Hydroxyestrone 2-Methoxyestrone
3) Flavonoids
OH O
OH OH
COMT
HO O HO O
SAM SAH
OH OH
OH O OH O
Quercetin Isohamnetin
4) Coumarins
OH O
OH OH
COMT
SAM SAH
O O
Esculetin Scopoletin
O O
HO COMT O
OH OH
NH2 SAM SAH NH2
HO HO
OH
HO O
HO
HO
N
P450 COMT
N
OH OH N
OH OH
SAM SAH
OH OH
OH HO O
H O
N ADH ADH N
N
N N N
N N
N
H2N H2N N NH
N NH H2N N NH
hypersensitivity
O O O
HN OH HN OH HN OH
Sulfotransferase UDP Glucuronide
Transferase
OSO3- OH O-Glu
P450
O
O
N OH
HN OH
Glutathione
Covalent Binding
to SH groups
Transferase
S-glutathione
O
OH
N-acetylbenzoquinone imine
Cell death
NAPQI
O OMe O OMe
O OMe
N OH N O [O]
N GSH N OH
N
H H N
H
HO O
HBr HO
SG Br
Br
OH OH
OH
DNA Damage
P450 [O]
2) Protein Damage
HO O
HO
OH O
17-Estradiol
OMe OMe
OMe
OMe OMe
OMe
N N
N
3) P450 GSH
O OH
OH
O O
O OMe
OMe OMe*
MeO MeO
MeO
SG
N N
N
Dauricine
Cl Cl Cl
Cl
P450 [O] N GSH NH
NH
NH
Cl O Cl O Cl O
4) Cl O
OH OH OH
OH GS
OH O OH
Diclofenac
implicated as responsible for the idiosyncratic toxicity that led to the with-
drawal of this agent from clinical use65 (Figure 10.10). Other examples of drugs
and endobiotics that undergo bioactivation to quinoid structures include the
estrogens, dauricine and diclofenac.66–68 In these cases, the phenols are oxidised
to highly conjugated quinonemethides. Diclofenac is oxidised to 5-hydro-
xydiclofenac by CYP3A; this p-aminophenol metabolite can then undergo a
second oxidation to a reactive quinoneimine, which may be involved in the
hepatotoxicity of this drug.
480 Chapter 10
Conjugative metabolism can also bioactivate alcohols to electrophiles. Sul-
fation of benzylic and allylic alcohols is a well-established mechanism of
bioactivation. The sulfate ester is reactive by virtue of the excellent leaving
group properties of the sulfate group and reactions occur with tissue nucleo-
philes (e.g. DNA) via SN2 reactions. For example, tamoxifen is well-established
as a hepatic carcinogen and clinical data have shown a link between tamoxifen
therapy and an increased incidence of endometrial cancer.69 The major meta-
bolic pathway is N-demethylation. Tamoxifen also forms two hydroxy meta-
bolites; one metabolite, a-hydroxy-tamoxifen has an allylic hydroxyl group that
undergoes sulfation. The resulting sulfate is hypothesised to be responsible for
the formation of DNA adducts48,70 (Figure 10.11).
N N
N
O O
O
DNA Adducts
O OSO- +
OH O
N N
Cl N Cl N
Amodiaquine AQQI
GSH
NH
Blocked Bioactivation (0% of dose)
Cl N
4-deOH-4F-Amodiaquine
10.4 Conclusions
The hydroxyl group represents an important substituent in drug design. Its
specific hydrogen-bonding capability can enhance or disrupt interactions with
macromolecule targets. The physicochemical properties that it imparts to drug
molecules can have a large impact on dispositional behaviour and result in
alterations in the pharmacokinetics relative to analogues lacking this group.
Furthermore, inclusion of a hydroxyl group in a molecule offers a new handle
which can be acted upon by drug-metabolising enzymes and metabolism can be
shifted from cytochrome P450 enzymes to conjugative enzymes such as UGTs.
References
1. R. S. Obach, F. Lombardo and N. Waters, Drug Metab. Dispos., 2008, 36,
1385.
2. J. Yu and Y. W. Chien, Pharm. Dev. Technol., 2002, 7, 215.
3. M. Aravagiri and S. R. Marder, Psychopharmacology, 2002, 159, 424.
4. T. B. Ejsing, A. D. Pedersen and K. Linnet, Hum. Psychopharmacol., 2005,
20, 493.
5. L. E. C. van Beijsterveldt, R. J. F. Geerts, J. E. Leysen and A. A. H. P.,
Psychopharmacology, 1994, 114, 53.
6. K. M. Kirschbaum, S. Henken, C. Hiemke and U. Schmitt, Behav. Brain
Res., 2008, 188, 298.
7. A. Doran, R. S. Obach, B. J. Smith, N. A. Hosea, S. Becker, E. Callegari, C.
Chen, X. Chen, E. Choo, J. Cianfrogna, L. M. Cox, J. P. Gibbs, M. A.
Gibbs, H. Hatch, C. E. C. A. Hop, I. N. Kasman, J. LaPerle, J.-H. Liu, X.
Liu, M. Logman, D. Maclin, F. M. Nedza, F. Nelson, E. Olson, S. Rahe-
matpura, D. Raunig, S. Rogers, K. Schmidt, D. K. Spracklin, M. Szewc, M.
Troutman, E. Tseng, M. Tu, J. W. van Deusen, K. Venkatakrishnan, G.
Walens, E. Q. Wang, D. Wong, A. S. Yasgar and C. Zhang, Drug Metab.
Dispos., 2005, 33, 165.
8. B. Feng, J. B. Mills, R. E. Davidson, R. J. Mireles, J. S. Janiszewski, M. D.
Troutman and S. M. de Morais, Drug Metab. Dispos., 2008, 36, 268.
9. A. E. Reed-Hagen, M. Tsuchiya, K. Shimada, J. Wentland and R. S.
Obach, Biopharm. Drug Dispos., 1999, 20, 429.
10. L. Slordal, R. Jaeger, J. Kjaeveand and J. Aarbakke, Pharmacol. Toxicol.,
1988, 63, 81.
11. G. Duester, J. Farres, M. R. Felder, R. S. Holmes, J. O. Höög, X. Parés, B.
V. Papp, S. J. Yin and H. Jörnvall, Biochem. Pharmacol., 1999, 58, 3.
12. J. O. Hoog, P. Stromberg, J. J. Hedberg and W. J. Griffiths, Chem-Biol.
Interact., 2003, 143–144, 175.
Alcohols and Phenols: Absorption, Distribution, Metabolism and Excretion 483
13. M. Estonius, S. Svensson and J. O. Hoeoeg, FEBS Letters, 1996, 397, 338.
14. A. Parkinson, Biotransformation of Xenobiotics, in Casarett and Doull’s
Toxicology, ed. Curtis D. Klassen. McGraw-Hill., New York, 2001, Sixth
Edition, pp. 152–154.
15. M. Sandberg, Y. Uemit, P. Stromberg, J. O. Hoeoeg and E. Eliasson, Br. J.
Clin. Pharmacol., 2002, 54, 423–429.
16. M. Breda, M. Strolin-Benedetti, M. Bani, C. Pellizzoni, I. Poggesi, G.
Brianceschi, M. Rocchetti, L. Dolfi, D. Sassella and R. Rimoldi, Phar-
macol. Res., 1999, 40, 351.
17. M. Sandberg, U. Yasur, P. Stromberg, J. O. Hoog and E. Eliasson, Br. J.
Pharmacol., 2002, 54, 423.
18. J. S. Walsh, M. J. Reese and L. M. Thurmond, Chem.-Biol. Interact., 2002,
142, 135.
19. L. K. Low and N. Castangnoli, Metabolic Changes of Drugs and Related
Organic Compounds, Wilson and Gisvold’s Textbook of Organic Medicinal
and Pharmaceutical Chemistry, eds. J. N. Delgado and W. A. Remers, JB
Lippincott Company, PA., USA, 1991, 9th edition, pp. 83–84.
20. S. Hamitouche, J. Poupon, Y. Dreano, Y. Amet and D. Lucas, Toxicol.
Lett., 2006, 167, 221.
21. P. T. Manyike, E. D. Kharasch, T. F. Kalhorn and J. T. Slattery, Clin.
Pharmacol. Ther., 2000, 67, 275.
22. R. L. Rose and P. E. Levi, in A Textbook of Modern Toxicology, ed. E.
Hodgson, 2004, Wiley, Hoboken, NJ, Ch. 8, pp. 149–162.
23. R. A. Stearns, P. K. Chakravarty, R. Chen and S. H. L. Chiu, Drug Metab.
Dispos., 1995, 23, 207.
24. U. Yasar, G. Tybring, M. Hidestrand, M. Oscarson, M. Ingelman-Sund-
berg, M. L. Dahl and E. Eliasson, Drug Metab. Dispos., 2001, 29, 1051.
25. J. Turgeon, Prog. Exper. Cardiol., 1998, 2, 153.
26. U. Yasar, M. L. Dahl, M. Christensen and E. Eliasson, Br. J. Clin.
Pharmacol., 2002, 54, 183.
27. U. Yasar, C. Forslund-Bergengren, G. Tybring, P. Dorado, A. Lerena, F.
Sjoeqvist, E. Eliasson and M. L. Dahl, Clin. Pharmacol. Ther., 2002, 71, 89.
28. M. O. Babaoglu, U. Yasar, M. Sandberg, E. Eliasson, M. L. Dahl, S. O.
Kayaalp and A. Bozkurt, Eur. J. Clin. Pharmacol., 2004, 60, 337.
29. K. Sekino, T. Kubota, Y. Okada, Y. Yamada, K. Yamamoto, R. Horiuchi,
K. Kimura and T. Iga, Eur. J. Clin. Pharmacol., 2003, 59, 589.
30. M. Lajer, L. Tarnow, S. Andersen and H. H. Parving, Diabetic Med., 2007,
24, 323.
31. M. J. Schlosser, R. D. Shurina and G. F. Kalf, Environ. Health Perspect.,
1989, 82, 229.
32. R. A. Franklin, Xenobiotica, 1988, 18, 105.
33. G. R. Murray, G. M. Whiffin, R. A. Franklin and J. A. Henry, Xenobio-
tica, 1989, 19, 669.
34. M. Lu, C. Zhang, J. Hao and Z. Qiu, Bioorg. Med. Chem. Lett., 2005, 15,
2607.
35. T. Walle, Molec. Pharmaceutics, 2007, 4, 826.
484 Chapter 10
36. B. Burchell, Transformation Reactions: Glucuronidation, in Handbook of
Drug Metabolism, ed. T. F. Woolf, Marcel Dekker Inc., New York, 1999,
pp. 153–173.
37. D. W. Nebert, D. R. Nelson, B. Burchell and K. W. Bock, DNA Cell Biol.,
1991, 10, 487.
38. C. Albert, O. Barbier, M. Vallee, G. Beaudry, A. Belanger and D. W. Hum,
Endocrinology, 2000, 141, 2472.
39. H. J. Schwarz, Br. J. Clin. Pharmacol., 1979, 8, 23s.
40. J. Escoriaza, M. C. Dios-Vieitez, I. F. Troconiz, M. J. Renedo and D. Fos,
Chromatographia, 1997, 44, 169.
41. M. Mandelli, G. Tognoni and S. Garattini, Clin. Pharmacokinet., 1978, 3, 72.
42. F. L. S. Tse, F. Balliard and J. M. Jaffe, J. Pharm. Sci., 1983, 72, 31.
43. D. D. Breimer, R. Jochemsen and H. H. von Albert, Arzneim.-Forsch.,
1980, 30, 875.
44. Z. Gregus, H. J. Kim, C. Madhu, V. Liu, P. Rozman and C. D. Klaassen,
Drug Metab. Dispos., 1994, 22, 725.
45. R. P. Miller, R. J. Roberts and L. J. Fischer, Clin. Pharmacol. Ther., 1976,
19, 284.
46. A. Parkinson, in Casaret and Doull’s Toxicology: The Basic Science of
Poisons, ed. C. D. Klaassen, McGraw-Hill, New York, 1996, Ch. 6, pp.
113–186.
47. L. Yi, J. Dratter, C. Wang, J. A. Tunge and H. Desaire, Anal Bioanal
Chem., 2006, 386, 666.
48. E. Banoglu, Curr. Drug Metab, 2000, 1, 1.
49. B. T. Zhu, Curr. Drug Metab., 2002, 3, 321.
50. J. A. Roth, M. H. Grossman and M. Adolf, Biochem. Pharmacol., 1990, 40,
1151.
51. E. Nissinen, R. K. Tuominen, V. Perhoniemi and S. Kaakkola, Life Sci.,
1988, 42, 2609.
52. P. T. Männistö, I. Ulmanen, K. Lundström, J. Taskinen, J. Tenhunen, C.
Tilgmann and S. Kaakkola, Prog. Drug Res., 1992, 39, 291.
53. B. Meister, A. J. Bean and A. Aperia, Kidney Int., 1993, 44, 726.
54. T. Karhunen, C. Tilgmann, I. Ulmanen, I. Julkunen and P. Panula, J.
Histochem. Cytochem., 1994, 42, 1079.
55. C. De Santi, P. C. Giulianotti, A. Pietrabissa, F. Mosca and G. M. Pacifici,
Eur. J. Clin. Pharmacol., 1998, 54, 215.
56. L. J. Bryan-Lluka, Arch.Pharmacol., 1995, 351, 408.
57. Y. S. Ding, S. J. Gatley, J. S. Fowler, R. Chen, N. D. Volkow, J. Logan, C.
E. Shea, Y. Sugano and J. Koomen, Life Sci., 1996, 58, 195.
58. B. T. Zhu, U. K. Patel, M. X. Cai and A. H. Conney, Drug Metab. Dispos.,
2000, 28, 1024.
59. H. C. Guldberg and C. A. Marsden, Pharmacol. Rev., 1975, 27, 135.
60. C. Prakash, D. Cui, M. J. Potchoiba and T. Butler, Drug Metab. Dispos.,
2007, 35, 1350.
61. F. R. Fontaine, R. A. Dunlop, D. R. Petersen and P. C. Burcham, Chem.
Res. Toxicol., 2002, 15, 1051.
Alcohols and Phenols: Absorption, Distribution, Metabolism and Excretion 485
62. J. W. Martin, S. A. Mabury and P. J. O’Brien, Chem.-Biol. Int., 2005
155, 165.
63. J. S. Walsh, M. J. Reese and L. M. Thurmond, Chem.-Biol. Int., 2002, 142,
135.
64. D. C. Dahlin, G. T. Miwa, A. Y. H. Lu and S. D. Nelson, Proc. Natl. Acad.
Sci. U. S. A., 1984, 81, 1327.
65. J. C. L. Erve, M. A. Svensson, H. von Euler-Chelpin and E. Klasson-
Wehler, Chem. Res. Toxicol., 2004, 17, 564.
66. J. L. Bolton, in Advances in Molecular Toxicology, ed. J. C. Fishbein,
Elsevier, Amsterdam, 2006, vol. 1, pp. 1–23.
67. Y. Wang, D. Zhong, X. Chen and J. Zheng, Chem. Res. Toxicol., 2009, 22,
824.
68. U. A. Boelsterli, Toxicol. Appl. Pharmacol., 2003, 192, 307.
69. C. J. Cohen, Semin. Oncol., 1997, 24, S55.
70. L. M. Notley, K. H. Crewe, P. J. Taylor, M. S. Lennard and E. M. J.
Gillam, Chem. Res. Toxicol., 2005, 18, 1611.
71. B. K. Park and N. R. Kitteringham, Drug Metab. Rev., 1994, 26, 605.
72. K. Harada, J. Matulic-Adamic, R. W. Price, R. F. Schinazi, K. Watanabe
and J. J. Fox, J. Med. Chem., 1987, 30, 226.
73. H. G. Howell, P. R. Brodfuehrer, S. P. Brundidge, D. A. Benigni and
C. Sapino, J. Org. Chem., 1988, 53, 85.
74. W. M. Watkins, D. G. Sixsmith, H. C. Spencer, D. A. Boriga, D. M.
Kiriuki, T. Kipingor and D. K. Koech, Lancet, 1984, 1(8373), 357.
75. K. A. Neftel, W. Woodtly and M. Schmid, Br. Med. J., 1986, 292, 721.
76. A. C. Harrison, N. R. Kitteringham, J. B. Clarke and B. K. Park, Biochem.
Pharmacol., 1992, 43, 1421.
77. P. M. O’Neill, A. C. Harrison, R. C. Storr, S. R. Hawley, S. A. Ward and
B. K. Park, J. Med. Chem., 1994, 37, 1362.
CHAPTER 11
486
Future Targets and Chemistry and ADME Needs 487
the ‘whole-genome random shotgun’ approach. The aim of this effort was to
sequence the human genome in three years and make the information available
as a commercial product. In 2000, Celera scientists publicly announced their first
draft and, in 2001, it was published.4
At this stage, B83% of the human genome had been sequenced, but it was
not until 2003 and 2005 that a fuller picture emerged, with 499% of genome
sequenced. Estimates at the start of both approaches suggested that the genome
may contain up to two million genes, but on completion, the general consensus
appears to be much less at approximately 30000.
This landmark achievement in science created a huge opportunity for the
pharmaceutical industry, academia and governmental agencies to convert this
wealth of information into drugs to treat disease.
The second question then relates to how many of the 30 000 gene products
may be seen as ‘druggable’. Hopkins and Groom took the gene family approach
and focussed on proteins as the most significant druggable biomolecule com-
pared with the other major classes of biomolecules—DNA, polysaccharides and
lipids.7 They first examined the chemical substrate used to target current gene
families and assumed that the binding site architecture was conserved across a
gene family. Future chemical substrates were then assumed to also be similar.
An assessment of the druggability8 of the small molecule substrate within each
gene family was then defined using the ‘rule of five’ (Ro5). It emerged from this
analysis that, of the 30 000 genes in the human genome, approximately 3000
488 Chapter 11
genes encode proteins with binding sites suitable for binding molecules which
are likely to have acceptable ADME properties.
The estimate of 5000–10 000 disease-related genes and 3000 genes which
encode proteins that are druggable prompts the question ‘how many disease
targets reside at the intersection’. In reality, the answer is that we do not have
the tools to give a precise estimate. An estimate by Hopkins and Groom was
put at 600–1500, and was based on extrapolation from the number of anti-
fungal targets in the yeast genome.7
It is this intersection of druggable proteins with disease-modifying proteins
that the majority of drugs in the current pharmacopeia occupy. An analysis by
Drews in 20006 showed that the modern pharmacopoeia modulates 483 targets.
A different analysis by Hopkins and Groom7 in 2002 put the number of pro-
teins that the pharmacopeia modulates at only 120 drugs with properties
commensurate for acceptable pharmacokinetic properties. These estimates
suggest that as researchers we have only scratched the surface of druggable and
disease-modifying proteins which could lead to important new drugs.
600
Successful
Number of targets
500
Research
400
Total
300
200
100
0
Muscloskeletal
Nutritional and
subcutaneous
Infectious and
blood forming
Genitourinary
abnormalities
Inflammation
Respiratory
system and
Neoplasms
Circulatory
Congenital
Blood and
Endocrine
Injury and
poisoning
metabolic
Ill defined
disorders
disorders
disorders
Digestive
Immunity
Skin and
parasitic
systems
Nervous
dieases
system
system
system
system
Mental
Figure 11.1 Therapeutic targets and disease: successful marketed drugs, drugs in
development and total.
Future Targets and Chemistry and ADME Needs 489
GCPRs
26%
Ther 119
genefamilies and
singleton targets Kinases
46% 10%
Zn peptidases
4%
CYPs Nuclear hormone Rs
2% 3%
Gated ion channels Ser proteases
Cys proteases 3%
2%
Cation channels 2%
2%
A
Kinases
22%
Other 114
genefamilies and
singleton targets
40%
GPCRs
15%
Cation channels
Dehydrogenases/ 5%
reductases
Nuclear hormone Rs
2%
2%
Carboxylases Zn peptidases
2% Protein phosphatases 2%
CYPs Ser proteases
4%
2% 4%
B
Figure 11.2 Drug target families shown as A—the molecular targets of experimental
and marketed drugs with physicochemical properties compliant with
druggable and B—the druggable genome.
druggable gene families within the human genome, although smaller than pre-
viously expected. GPCRs have previously been rich hunting grounds for drug
discovery9,10 and protein kinases have begun to yield some success stories.
Alongside these important classes, the total of zinc metallopeptidases, serine
proteases, phosphodiesterases and nuclear hormone receptors make up nearly
half of the gene families within the druggable human genome. In contrast to
what was initially expected, proteases in particular have emerged as a significant
gene family.
490 Chapter 11
A conservative estimate by Russ and Lampel made in 2005 puts the number of
druggable proteases at B230 proteins.11 Together with kinases, GPCRs and
other families, the number of druggable targets reached in this analysis is just
over 3000, which is remarkably similar to that estimated by Hopkins and
Groom.7
N
O N O
O N
N O O
O
N N N N
N
N F
Cl N N
Imatinib (Novartis - 2001) Gefitinib (AstraZeneca - 2003) Erlotinib (OSI, Roche - 2004)
MW 494 MW 447 MW 393
clogP 4.5 clogP 5.6 clogP 4.3
TPSA 86Å2 TPSA 69Å2 TPSA 75Å2
Cl
O
O
N Cl S
N N O
N N
N
F N
O O N N N
F N O N
N F N
N
N Cl N N
N
N O O
O O N
S N
O F N N
Figure 11.3 Structures of marketed kinase inhibitors (TPSA¼topological polar surface area).
Future Targets and Chemistry and ADME Needs 493
O
O O N
O HN
NH
O
O N
O NH
HN N N N N N
N
OH N
develop new tools to better understand the properties of molecules within this
challenging physicochemical space and ultimately to increase confidence pre-
clinically that the required ADME profiles will be achieved.
Alongside this, a deeper and more rational understanding of the physical
processes that occur during oral absorption is required. These tools should
include improved in silico methods for prospective, property based design of
orally bioavailable compounds, as well as more relevant in vitro methods for
investigating membrane permeability (both gut and blood–brain barrier
penetration).
Analysis of the HIV protease inhibitor, atazanavir, and the first kinase inhi-
bitor, imatinib, illustrates these points well (Figure 11.4). Both are high mole-
cular weight, relatively lipophilic compounds. Atazanavir moves further out of
desirable ADME space due to a high hydrogen bond donor count and high polar
surface area (PSA), yet displays remarkably good human oral pharmacoki-
netics.22,23 Similarly, imatinib24 is highly orally bioavailable despite its high
molecular weight and lipophilic character. If property-based design principles25
were applied to these molecules, both would be at the high end of predicted risk
for poor oral pharmacokinetics. In reality, both are highly successful oral agents.
O
O
N N
O HN N
N N O HN
S N N
N S N
O N
O
O
O
Sildenafil Vardenafil
Figure 11.5 Chemical structures of the PDE5 inhibitors, sildenafil and vardenafil.
Future Targets and Chemistry and ADME Needs 495
building blocks for singleton and library syntheses is critical to the rapid
advance of drug discovery programmes. An analysis of commercially available
tri- and tetra-substituted phenyl rings reveals that these molecular building
blocks are under-represented in supplier catalogues relative to the numbers of
combinations and permutations theoretically possible with such highly sub-
stituted ring systems. A similar scenario exists for tri-substituted pyridine ring
systems (Table 11.1). This is presumably due to synthetic complexity and
expense, but again highlights where synthetic chemistry could add value to
advancing medicinal chemistry programmes.
Moving beyond the paradigm of chemical space exploration, that is largely
driven by commercial availability of suitable building blocks, is a must. To
reiterate a point made above, the design and synthesis of the ‘right’ molecule is
key even if this involves significant investment in synthesis.
The synthetic methodology required to covalently join functionalised phenyl
and heteroaromatic ring systems and derivatise them has advanced significantly
in the last 10–15 years. A comprehensive review of this synthetic methodology
is beyond the scope of this chapter, but it has been reviewed extensively else-
where.27 Many of these chemical reactions are now sufficiently robust to be
carried out by parallel synthesis protocols. Moreover, advances in catalysts
have enabled a greater variety in coupling partners which has broadened the
utility of many of the commercially available compounds analysed in Table
11.1. But to reiterate the point made previously, the major limitation in
accessing this chemical space remains with the availability of diverse and highly
substituted compatible building blocks.
With many drug targets favouring lipophilic chemotypes such as the kinase
drugs described in Figure 11.3, P450-mediated oxidative metabolism is often
an inevitable consequence. This can then lead to poor oral bioavailability
and undesirable pharmacokinetic half-lives. Fluorination as a means to block
P450-mediated metabolism, as well as offering the potential to increase
potency, is emerging as a favoured tactic.28,29 Fluorination is also a useful
strategy to modulate the pKa of basic drugs, which is often important for
designing compounds with reduced affinity for the HERG cardiac potassium
channel.
The DPP-4 inhibitor, sitagliptin, is an example of a highly fluorinated drug
that benefits from improved potency and oral bioavailability by judicious use of
its fluorine substituents. Again, fluorination strategies in the NK1 antagonist,
aprepitant led to an improved duration of action, potency and central nervous
system (CNS) penetration (Figure 11.6). The role of fluorine as an isostere is
discussed at length in Chapter 10 section 3.4. With fluorine substitution capable
of imparting such stark pharmacokinetic and pharmacodynamic advantages
for such a small molecular change (often H to F), synthetic methods to
introduce this atom into molecules of interest is therefore high on the wish list
of most medicinal chemists. Advances in synthetic methodology to introduce
fluorine have emerged in recent years, yet fluorine scans still represent an
enormous synthetic challenge and therefore improved methods and/or more
commercially available fluorinated building blocks would have a huge impact.
496
A H A H A N A
H H
A 1728 A 1574 A 278
A H A H H A
A A A H A N H
H H
H 897 H 595 A 305
A A A A A H
A A H H A N H
H A
a
A ¼ non-hydrogen.
Chapter 11
Future Targets and Chemistry and ADME Needs 497
F F O
F
F O N
F NH
F N N
N F O N
N N H
NH 2 O
F
F F
F Sitagliptin F Aprepitant
F
F
OH O
N
N O
O O
N N N
F N
H
NH2 Oseltamivir
Voriconazole
F
Glomerular filtration
Plasma
systems
Carrier
Carrier
systems
A Lipoidal
D diffusion
Metabolism
Liver Kidney
Gut
Carrier
systems
Bile Faeces
Figure 11.8 Simplified schematic for the processes of drug absorption and clearance.
The figure illustrates the role of passive and active transport, and enzy-
matic processes.
Log P = MW - PSA
500
ADME
space
MW
5
Lipophilicity
140
PSA
O
N N
O O
CH3 HO CH3
HO
O O
N S F N S F
O O
with molecular weight exceeding 500 Da and combining high lipophilicity and
high PSA (118–202 Å2). These drugs have much lower permeabilities with
highly polarised flux due to transporters. Although low flux is evident, rea-
sonable bioavailabilities can be achieved mainly due to the high local con-
centrations of drug present in the gastrointestinal tract after oral
administration and resultant saturation of the transporters.
The properties of the protease inhibitors in Table 11.3 (which are discussed in
detail in Chapter 10) move the drugs outside the boundaries of ADME space
(high PSA and high lipophilicities) and the drugs have all low permeability;
their disposition will be highly influenced by transporters as illustrated by the
flux across Caco-2 cells.
The low permeability also manifests itself in clearance since most transporter
systems work to influx the drug into the hepatocyte or kidney tubule (poor
permeability essentially making it a one-way process) and then efflux it into the
urine or bile. With poor permeability, the process of metabolism may be sec-
ondary to the process of systemic clearance.
Figure 11.10 shows the structure of a combined thromboxane receptor
antagonist/synthetase inhibitor and its major N-oxide metabolite. The physi-
cochemical properties of the molecule are a log P of 4.3 and a PSA of 106 Å2.
Future Targets and Chemistry and ADME Needs 503
The compound is therefore likely to be of lower permeability. As an acid with
lipophilic regions, it is likely also to be a substrate for organic acid transport
protein (OATP) transporters. The biliary excretion and clearance of the com-
pound was monitored in a series of experiments in the isolated perfused rat
liver.34 In the first series of experiments in the untreated rat liver, systemic
plasma clearance was around 15 ml min1 kg1 and the drug was excreted
alongside its N-oxide, which represented the majority portion of the excretion.
In a second series of experiments, clearance remained unchanged following
ketoconazole pretreatment (a CYP450 inhibitor), but the N-oxide metabolite
was absent and only the parent was detected in bile. This compound was
therefore cleared by hepatic uptake (influx) and metabolism was subsequent to
the event (post-clearance metabolism).
Further examples of the interplay between uptake and metabolism with low
permeability compounds are provided by studies with HIV protease inhibitors.
The intrinsic metabolic clearance of saquinavir, nelfinavir and ritonavir35 was
determined in both rat liver microsomes and fresh isolated rat hepatocytes in
suspension. In the absence of metabolism (achieved by pretreating hepatocytes
with a mechanism-based inhibitor of cytochrome P450), the protease inhibitors
were actively and rapidly taken up into hepatocytes; intracellular unbound
drug concentrations were 5- and 12-fold higher than extracellular unbound
concentrations. Comparison of the rate of uptake into hepatocytes with the rate
of metabolism in hepatocytes and microsomes indicates that the former is the
rate-limiting step at low concentrations. These findings explain the rapid
clearance of saquinavir, nelfinavir and ritonavir and also indicate that the rate
of uptake limits the metabolic clearance of these three drugs. Further studies
(using an hepatic cell model, Hep G2 and Xenopus laevis oocytes over-
expressing human OATP-A) strongly indicate that basolaterally located
OATP-A (influx transporter) in human hepatocytes is the influx transporter
and is acting in concert with apically located P-gp and/or MRP2 (efflux
transporters) for the vectorial transport and excretion of saquinavir into bile.36
1.6
CSF / Cp (free)
1.2
0.8
0.4
-0.4
-4 -2 0 2 4 6 8
Log D
Figure 11.11 Penetration across the blood–brain barrier for the drugs shown in Table
11.3. The graph plots penetration expressed as CSF to unbound drug in
plasma ratio against lipophilicity (log P) with the size of point indicative
PSA. Note that high molecular weight drugs have penetration
decreased by poor permeability due to polar surface area, despite high
apparent lipophilicities. These trends are illustrative of the concepts
depicted in Figure 11.2.
To access the brain and interact with many of the new drug targets is a
major challenge. Figure 11.12 illustrates the properties of a series of b-secretase
inhibitors. Ab peptide is associated with the damage caused during
Alzheimer’s disease and is formed by processing amyloid precursor protein.
b-secretase (BACE-1) is the first enzyme in the cascade; it is an aspartyl
protease (a target that is already challenging), made even more difficult by
being present inside the CNS. Other aspartyl protease inhibitors such as renin
and HIV protease have traditionally started with transition state isosteres, but
in this approach new templates were screened for with selectivity to the
enzyme.37 Considerable synthetic diversity and large increases in potency were
achieved from this lead. The initial direction was to incorporate a hydro-
xyethylamine transition state isostere in the molecule.38 This dramatically
improves potency but also increases PSA. Despite good cellular activity of
10 nM, brain penetration was negligible. Reduction of PSA39 was attempted by
truncation of the hydroxyethylamine isostere, initially to a primary alcohol and
then to an amine (Figure 11.12). Despite impressive potency, poor pharma-
cokinetics (including limited transfer across the blood–brain barrier) hamper
the series.
The fundamental problem of such targets is the need to maintain consider-
able hydrogen bonding functionality to achieve selectivity and potency. The
resultant high PSA (and consequent permeability issues) attenuates absorption
and access to the target, which may be due to high clearance through efflux.
506 Chapter 11
O O
S CH3
H3C N
Cl
N
H3C N N
N
O
O NH2
CH3 CH3
Figure 11.13 TC-1—a BACE inhibitor with good permeability properties. Although
PSA remains high (136 Å2), it is possible that the difference between
hydrogen bond donors and acceptors accounts for its permeability
characteristics.
$
A1 A2 Cl
N O N O
N N N
N N N N
H £ F F H
O O
F
N NH2
O
O
HN CH3
HN CH3
B1
B2
O
S O
O O S
O O
N
Figure 11.14 Stabilisation to oxidation using fluorine (and chlorine, $) and con-
jugation (removal of functionality, d) for A1 and A2, a thrombin
inhibitor, and B1 and B2, a bradykinin receptor antagonist.
References
1. International Human Genome Sequencing Consortium, Finishing the
euchromatic sequence of the human genome, Nature, 2004, 431, 931.
2. International Human Genome Sequencing Consortium et al., Initial
sequencing and analysis of the human genome, Nature, 2001, 409, 860.
3. International Human Genome Sequencing Consortium et al., A physical
map of the human genome, Nature, 2001, 409, 934.
4. J. C. Venter, et al., The sequence of the human genome, Science, 2001, 291,
1304.
5. J. Drews, Nat. Biotechnol., 1996, 14, 1516.
6. J. Drews, Science, 2000, 287, 1960.
7. A. L. Hopkins and C. R. Groom, Nat. Rev. Drug. Discov., 2002, 1, 727.
8. C. Lipinski, F. Lombardo, B. Dominy and P. Feeney, P. Adv. Drug Deliv.
Rev., 1997, 23, 3.
9. C. J. Zheng, l. Y. Han, C. W. Yap, Z. L. Ji, Z. W. Cao and Y. Z. Chen,
Pharmacol. Rev., 2006, 58, 259.
510 Chapter 11
10. K. Beaumont, E. Schmid and D. A. Smith, Bioorg. Med. Chem. Lett., 2005,
15, 3658.
11. A. P. Russ and S. Lampel, Drug Discov. Today, 2005, 10, 1607.
12. M. Szyf, Ann. Rev. Pharmacol. Toxicol., 2009, 49, 243.
13. J. A. Gerrardl, C. A. Hutton and M. A. Perugini, Mini Rev. Med. Chem.,
2007, 7, 151.
14. S. Shangary and S. Wang, Annu. Rev. Pharmacol. Toxicol., 2009, 49, 223.
15. G. W. Carlile, R. Renaud, Z. Donglei Zhang, K. A. Teske, Y. Luo, J. W.
Hanrahan and D. Y. Thomas, ChemBioChem., 2007, 8, 1012.
16. H. Takemasa, T. Nagatomo, H. Abe, K. Kawakami, T. Igarashi, T.
Tsurugi, N. Kabashima, M. Tamura, M. Okazaki, B. P. Delisle, C. T.
January and Y. Otsuji, Br. J. Pharmacol., 2008, 153, 439.
17. D. Thomas, J. Kiehn, H. A. Katus and C. A. Karle, Cardiovasc. Res., 2003,
60, 235.
18. J. S. Carew, S. T. Nawrocki, Y. V. Krupnik, K. Dunner, D. J. McConkey,
M. J. Keating and P. Huang, Blood, 2006, 107, 222.
19. F. van Goor, K. S. Straley, D. Cao, J. Gonzalez, S. Hadida, A. Hazlewood,
J. Joubran, T. Knapp, L. R. Makings, M. Miller, T. Neuberger, E. Olson,
V. Panchenko, J. Rader, A. Singh, J. H. Stack, R. Tung, P. D. Grootenhuis
and P. Negulescu, Am. J. Physiol., 2006, 290, L1117.
20. J. Zhang, P. L. Yang and N. S. Gray, Nat. Rev. Cancer, 2009, 9, 28.
21. G. V. Paolini, R. H. B. Shapland, W. P. van Hoorn, J. S Mason and A. L.
Hopkins, Nat. Biotechnol., 2006, 24, 805.
22. P. J. Piliero, Drugs Today, 2004, 40, 901.
23. M. J. Pérez-Elı́as, Expert Opin. Pharmacother., 2007, 8, 689.
24. Drugs@FDA: www.accessdata.fda.gov/scripts/cder/drugsatfda/index.cfm;
human oral bioavailability value taken from the drug label, available freely
at www.fda.gov.
25. H. van de Waterbeemd, D. A. Smith, K. Beaumont and D. K. Walker, J.
Med. Chem., 2001, 44, 1313.
26. W. R. Pitt, D. M. Parry, B. G. Perry and C. R. Groom, J. Med. Chem.,
2009, 52, 2952.
27. R. Larsen, Curr. Opin. Drug. Discov. Dev., 1999, 2, 651.
28. W. K. Hagmann, J. Med. Chem., 2008, 51, 4359.
29. K. Müller, C. Faeh and F. Diederich, Science, 2007, 317, 1881.
30. F. Lovering, J. Bikker and C. Humblet, J. Med. Chem., 2009, 52, 6752.
31. D. W. C. MacMillan, Nature, 2008, 455, 304.
32. B. M. Trost, PNAS, 2004, 101, 5348–5355.
33. M. V. S. Varma, K. Sateesh and R. Panchagnula, Mol. Pharm., 2005, 2, 12.
34. I. B. Gardner, D. K. Walker, M. S. Lennard, D. A. Smith and G. T.
Tucker, Xenobiotica, 1995, 25, 185.
35. A. J. Parker and J. B Houston, Drug Metab. Dispos., 2008, 36, 1375.
36. Y. Su, X. Zhang and P. J. Sinko, Mol. Pharm., 2004, 1, 49.
37. C. A. Coburn, S. J. Stachel, Y. -M. Li, D. M. Rush, T. G. Steele, E. Chen
Dodson, M. K. Holloway, M. Xu, Q. Huang, M.-T. Lai, J. DiMuzio, M.-
C. Crouthamel, X.-P. Shi, V. Sardana, Z. X. Chen, S. Munshi, L. Kuo, G.
Future Targets and Chemistry and ADME Needs 511
M. Makara, D. A. Annis, P. K. Tadikonda, H. W. Nash, J. P. Vacca and T.
Wang, J. Med. Chem., 2004, 47, 6117.
38. S. J. Stachel, C. A. Coburn, T. G. Steele, K. G. Jones, E. F. Loutzenhiser,
A. R. Grego, H. A. Rajapaske, M. T. Lai, M.-C. Crouthamel, M. Xu, K.
Tugusheva, J. E. Lineberger, B. L. Pietrak, A. S. Espeseth, X.-P. Shi, E.
Chen-Dodson, M. K. Holloway, S. Munshi, A. J. Simon, L. Kuo and J. P.
Vacca, J. Med. Chem., 2004, 47, 6447.
39. M. G. Stanton, S. R. Stauffer, A. R. Grego, M. Steinbeiser, P. Nantermet,
S. Sanmkaranarayanan, E. A. Price, G. Wu, M.-C. Crouthamel, J. Ellis,
M. Y. Lai, A. S. Espeseth, X. P. Shi, L. Jin, D. Colussi, B. Pietrk, Q.
Huang, M. Xu, A. J. Simon, S. L. Gaham, J. P. Vacca and H. Seinick, J.
Med. Chem., 2007, 50, 3431.
40. L. Cheng, L. Nair, T. Liu, F. Li, J. Pichardo, S. Agrawal, R. Chase, X.
Tong, A. S. Uss, S. Bogen, F. G. Njoroge, R. A. Morrison and K. C.
Cheng, Biochem. Pharmacol., 2008, 75, 1186.
41. S. Sankaranarayanan, M. A. Holahan, D. Colussi, M. C. Crouthamel, V.
Devanarayan, J. Ellis, A. Espeseth, A. T. Gates, S. L. Graham, A. R.
Gregro, D. Hazuda, J. Hochman, K. Holloway, L. Jin, J. Kahana, M.-T.
Lai, J. Lineberger, G. McGaughey, K. P. Moore, P. Nantermet, B. Pietrak,
E. A. Price, H. Rajapakse, S. Stauffer, M. A. Steinbeiser, G. Seabrook, H.
G. Selnick, X. P. Shi, M. G. Stanton, J. Swestock, K. Tugusheva, K. X.
Tyler, J. P. Vacca, J. Wong, G. Wu, M. Xu, J. J. Cook and A. J. Simon, J.
Pharm. Exper. Therap., 2009, 328, 131.
42. Q. Ma and A. Y. H. Lu, Curr. Drug Metab., 2008, 9, 374.
43. M. Ekroos and T. Sjoegren, Proc. Natl. Acad. Sci. U.S.A., 2006, 103,
13682.
44. C. S. Burgey, K. A. Robinson, T. A. Lyle, P. E. J. Sanderson, S. D. Lewis,
B. J. Lucas, J. A. Krueger, R. Singh, C. Miller-Stein, R. B. White, B.
Wong, E. A. Lyle, P. D. Williams, C. A. Coburn, B. D. Dorsey, J. C.
Barrow, M. T. Stranieri, M. A. Holahan, G. R. Sitko, J. J. Cook, D. R.
McMasters, C. M. McDonough, W. M. Sanders, A. A. Wallace, F. C.
Clayton, D. Bohn, Y. M. Leonard, T. J. Detwiler, J. J. Lynch, Y. Yan, Z.
Chen, L. Kuo, S. J. Gardell, J. A. Shafer and J. P. Vacca, J. Med. Chem.,
2003, 46, 461.
45. D.-S. Su, J. L. Lim, E. Tinney, B.-L. Wan, K. L. Murphy, D. R. Reiss, C.
M. Harrell, S. S. O’Malley, R. W. Ransom, R. S. L. Chang, D. J. Petti-
bone, J. Yu, C. Tang, T. Prueksaritanont, R. M. Freidinger, M. G. Bock
and N. J. Anthony, J. Med. Chem., 2008, 51, 3946.
Subject Index