Medical Pharmacology and Therapeutics 5th .pdf







Any screen.
Any time.
Anywhere.
Activate the eBook version
of this title at no additional charge.

Unlock your eBook today.
1 Visit studentconsult.inkling.com/redeem
2 Scratch off your code
3 Type code into “Enter Code” box
4 Click “Redeem”
5 Log in or Sign up
6 Go to “My Library”
It’s that easy!
Student Consult eBooks give you the power to browse and find content,
view enhanced images, share notes and highlights—both online and offline.

For technical assistance:
email studentconsult.help@elsevier.com
call 1-800-401-9962 (inside the US)
call +1-314-447-8200 (outside the US)

Scan this QR code to redeem your
eBook through your mobile device:

Use of the current edition of the electronic version of this book (eBook) is subject to the terms of the nontransferable, limited license granted on
studentconsult.inkling.com. Access to the eBook is limited to the first individual who redeems the PIN, located on the inside cover of this book, at studentconsult.inkling.com and
may not be transferred to another party by resale, lending, or other means.

Place Peel Off
Sticker Here

2015v1.0

MEDICAL
PHARMACOLOGY
&THERAPEUTICS
Fifth Edition

Dedication
To our families

Fifth Edition

Derek G. Waller
BSc (HONS), DM, MBBS (HONS), FRCP
Consultant Cardiovascular Physician, University Hospital Southampton,
Senior Clinical Lecturer, University of Southampton,
Southampton, United Kingdom
Senior Medical Adviser to the British National Formulary
Anthony P. Sampson
MA, PhD, FHEA, FBPhS
Associate Professor in Clinical Pharmacology, Faculty of Medicine,
University of Southampton, Southampton, United Kingdom
MEDICAL
PHARMACOLOGY
& THERAPEUTICS

Edinburgh London New York Oxford Philadelphia St Louis Sydney Toronto 2018

© 2018, Elsevier Limited. All rights reserved.
First edition 2001
Second edition 2005
Third edition 2010
Fourth edition 2014
Fifth edition 2018
The rights of Derek G. Waller and Anthony P. Sampson to be identified as authors of this work
have been asserted by them in accordance with the Copyright, Designs and Patents Act 1988.
No part of this publication may be reproduced or transmitted in any form or by any means,
electronic or mechanical, including photocopying, recording, or any information storage and
retrieval system, without permission in writing from the publisher. Details on how to seek
permission, further information about the Publisher’s permissions policies and our arrangements
with organizations such as the Copyright Clearance Center and the Copyright Licensing Agency,
can be found at our website: www.elsevier.com/permissions.
This book and the individual contributions contained in it are protected under copyright by the
Publisher (other than as may be noted herein).
Notices
Knowledge and best practice in this field are constantly changing. As new research and experience
broaden our understanding, changes in research methods, professional practices, or medical
treatment may become necessary.
Practitioners and researchers must always rely on their own experience and knowledge in
evaluating and using any information, methods, compounds, or experiments described herein. In
using such information or methods they should be mindful of their own safety and the safety of
others, including parties for whom they have a professional responsibility.
With respect to any drug or pharmaceutical products identified, readers are advised to check
the most current information provided (i) on procedures featured or (ii) by the manufacturer of each
product to be administered, to verify the recommended dose or formula, the method and duration
of administration, and contraindications. It is the responsibility of practitioners, relying on their own
experience and knowledge of their patients, to make diagnoses, to determine dosages and the
best treatment for each individual patient, and to take all appropriate safety precautions.
To the fullest extent of the law, neither the Publisher nor the authors, contributors, or editors,
assume any liability for any injury and/or damage to persons or property as a matter of products
liability, negligence or otherwise, or from any use or operation of any methods, products,
instructions, or ideas contained in the material herein.
ISBN: 978-0-7020-7167-6
Printed in China
Last digit is the print number: 9 8 7 6 5 4 3 2 1
Senior Content Strategist: Pauline Graham
Content Development Specialist: Carole McMurray
Project Manager: Dr. Atiyaah Muskaan
Design: Amy Buxton
Illustration Manager: Karen Giacomucci
Marketing Manager: Deborah Watkins

The
publisher’s
policy is to use
paper manufactured
from sustainable forests

Preface ...........................................................................vii
Drug dosage and nomenclature .....................................ix
SECTION 1 GENERAL
PRINCIPLES
1. Principles of pharmacology and
mechanisms of drug action.................................. 3
2. Pharmacokinetics................................................ 33
3. Drug discovery, safety and efficacy ................... 63
4. Neurotransmission and the peripheral
autonomic nervous system................................. 73
SECTION 2 THE
CARDIOVASCULAR SYSTEM
5. Ischaemic heart disease ..................................... 93
6. Systemic and pulmonary hypertension ............ 111
7. Heart failure....................................................... 131
8. Cardiac arrhythmias.......................................... 143
9. Cerebrovascular disease and dementia........... 161
10. Peripheral vascular disease.............................. 169
11. Haemostasis ..................................................... 175
SECTION 3 THE RESPIRATORY
SYSTEM
12. Asthma and chronic obstructive
pulmonary disease............................................ 193
13. Respiratory disorders: cough, respiratory
stimulants, cystic fibrosis, idiopathic
pulmonary fibrosis and neonatal respiratory
distress syndrome............................................. 211
SECTION 4 THE RENAL SYSTEM
14. Diuretics ............................................................ 219
15. Disorders of micturition .................................... 231
16. Erectile dysfunction........................................... 239

SECTION 5 THE NERVOUS
SYSTEM
17. General anaesthetics ........................................ 247
18. Local anaesthetics ............................................ 257
19. Opioid analgesics and the management
of pain ............................................................... 263
20. Anxiety, obsessive-compulsive disorder
and insomnia..................................................... 279
21. Schizophrenia and bipolar disorder.................. 287
22. Depression, attention deficit hyperactivity
disorder and narcolepsy ................................... 297
23. Epilepsy............................................................. 311
24. Extrapyramidal movement disorders
and spasticity.................................................... 325
25. Other neurological disorders: multiple
sclerosis, motor neuron disease and
Guillain–Barré syndrome................................... 337
26. Migraine and other headaches......................... 341

SECTION 6 THE
MUSCULOSKELETAL SYSTEM
27. The neuromuscular junction and
neuromuscular blockade .................................. 351
28. Myasthenia gravis ............................................. 359
29. Nonsteroidal antiinflammatory drugs................ 363
30. Rheumatoid arthritis, other inflammatory
arthritides and osteoarthritis............................. 373
31. Hyperuricaemia and gout ................................. 385

SECTION 7 THE
GASTROINTESTINAL SYSTEM
32. Nausea and vomiting........................................ 393
33. Dyspepsia and peptic ulcer disease ................ 401
34. Inflammatory bowel disease............................. 411
35. Constipation, diarrhoea and irritable
bowel syndrome................................................ 417
36. Liver disease ..................................................... 425
37. Obesity .............................................................. 433

Contents

vi Medical Pharmacology and Therapeutics

SECTION 8 THE IMMUNE
SYSTEM
38. The immune response and
immunosuppressant drugs ............................... 439
39. Antihistamines and allergic disease ................. 451

SECTION 9 THE ENDOCRINE
SYSTEM AND METABOLISM
40. Diabetes mellitus............................................... 459
41. The thyroid and control of metabolic rate........ 475
42. Calcium metabolism and metabolic
bone disease..................................................... 481
43. Pituitary and hypothalamic hormones.............. 491
44. Corticosteroids (glucocorticoids and
mineralocorticoids)............................................ 503
45. Female reproduction......................................... 513
46. Androgens, antiandrogens and
anabolic steroids............................................... 531
47. Anaemia and haematopoietic
colony-stimulating factors ................................ 537
48. Lipid disorders .................................................. 547

SECTION 10 THE SKIN AND
EYES
49. Skin disorders ................................................... 561
50. The eye ............................................................. 569
SECTION 11 CHEMOTHERAPY
51. Chemotherapy of infections.............................. 581
52. Chemotherapy of malignancy........................... 631
SECTION 12 GENERAL
FEATURES: DRUG TOXICITY
AND PRESCRIBING
53. Drug toxicity and overdose .............................. 659
54. Substance abuse and dependence.................. 675
55. Prescribing, adherence and information
about medicines ............................................... 689
56. Drug therapy in special situations .................... 695
Index............................................................................ 707

The fifth edition of Medical Pharmacology and Therapeutics
has been extensively revised and updated while preserving
the popular approach of the fourth edition. We particularly
wish to acknowledge here the contributions made to previous
editions of this book by Emeritus Professor Andrew Renwick
OBE BSc PhD DSc and Dr Keith Hillier BSc PhD DSc, Senior
Lecturer in Pharmacology, formerly our colleagues in the
University of Southampton Faculty of Medicine. As before,
a disease-based approach has been taken to explain clinical
pharmacology and therapeutics and the principles of drug
use for the management of common diseases. Medical
Pharmacology and Therapeutics provides key information on
basic pharmacology and other relevant disciplines sufficient to
underpin the clinical context. It provides information suitable
for all healthcare professionals who require a sound knowledge
of the basic science and clinical applications of drugs.
The text is structured to reflect the ways that drugs are
used in clinical practice. The chapters covering generic
concepts in pharmacology and therapeutics include sections
on how drugs work at a cellular level, drug metabolism and
pharmacokinetics, pharmacogenetics, drug development and
drug toxicity. The basic principles of prescribing have been
emphasized and new information provided on preparing for
the Prescribing Safety Assessment (PSA) in the UK. New
sections have been included on pharmacovigilance, on
so-called ‘legal highs’ and in all other areas, and the sections
on clinical management have been thoroughly revised and
updated.
Each chapter in this fifth edition retains the following
key features:
■ An up-to-date and succinct explanation of the major
pathogenic mechanisms of disease and consequent
clinical symptoms and signs, helping the reader to put
Preface

into context the actions of drugs and the consequences
of their therapeutic use.
■ A structured approach to the principles of disease
management, outlining core principles of drug choice
and planning a therapeutic regimen for many common
diseases.
■ A comprehensive review of major drug classes relevant to
the disease in question. Basic pharmacology is described
with clear identification of the molecular targets, clinical
characteristics, important pharmacokinetic properties
and unwanted effects associated with individual drug
classes. Example drugs are used to illustrate the common
pharmacological characteristics of their class and to
introduce the reader to drugs currently in widespread
clinical use.
■ To complement the information provided within each
chapter, a drug compendium is included describing the
main characteristics of all drugs in each class listed in
the British National Formulary.
■ A revised section of self-assessment questions for learning
and revision of the concepts and content in each chapter,
including one best answer (OBA), extended matching
items (EMI), true-false and case-based questions.
It is our intention that the fifth edition of this book will
encourage readers to develop a deeper understanding of the
principles of drug usage that will help them to become safe
and effective prescribers and to carry out basic and clinical
research and to teach. As medical science advances these
principles should underpin the life-long learning essential
for the maintenance of these skills.

DGW
APS

This page intentionally left blank

DRUG NOMENCLATURE
In the past, the nonproprietary (generic) names of some drugs
have varied from country to country, leading to potential
confusion. Progressively, international agreement has been
reached to rationalise these variations in names and a single
recommended International Non-proprietary Name (INN) given
to all drugs. Where the previously given British Approved
Name (BAN) and the INN have differed, the INN is now the
accepted name and is used throughout this book.

A special case has been made for two medicinal sub-
stances: adrenaline (INN: epinephrine) and noradrenaline (INN:

norepinephrine). Because of the clinical importance of these

substances and the widespread European use and understand-
ing of the terms adrenaline and noradrenaline, manufacturers

have been asked to continue to dual-label products adrenaline
(epinephrine) and noradrenaline (norepinephrine). In this book,

where the use of these agents as administered drugs is being
described, dual names are given. In keeping with European
convention, however, adrenaline and noradrenaline alone are
used when referring to the physiological effects of the naturally
occurring substances.
DRUG DOSAGES
Medical knowledge is constantly changing. As new information
becomes available, changes in treatment, procedures,
equipment and the use of drugs become necessary. The
authors and the publishers have taken care to ensure that
the information given in the text is accurate and up to date.
However, readers are strongly advised to confirm that the
information, especially with regard to drug usage, complies
with the latest legislation and standards of practice.
Drug dosage and nomenclature

This page intentionally left blank

General principles
1

This page intentionally left blank

1

Studying pharmacology 3
Finding drug information 4
Receptors and receptor-mediated
mechanisms 4
Actions of drugs at binding sites (receptors) 4
Major types of receptors 5
Other sites of drug action 12
Properties of receptors 13
Properties of drug action 15
Dose–response relationships 16
Types of drug action 17
Agonists 17
Antagonists 18
Partial agonists 18
Inverse agonists 18
Allosteric modulators 19
Enzyme inhibitors and activators 19
Nonspecific actions 19
Physiological antagonists 19
Tolerance to drug effects 19
Genetic variation in drug responses 20
Summary 21
Self-assessment 21
Answers 22
Further reading 22
Appendix: student formulary 28

STUDYING PHARMACOLOGY
Drugs are defined as active substances administered to prevent,
diagnose or treat disease, to alleviate pain and suffering, or to
extend life. Pharmacology is the study of the effects of drugs
on biological systems, with medical (or clinical) pharmacology
concerned with the drugs that doctors, and some other
healthcare professionals, prescribe for their patients. The
prescribing of drugs has a central role in therapeutics, and
gaining a good knowledge of pharmacology is essential for
health professionals to become safe and effective prescribers.
Drugs may be chemically synthesised or purified from
natural sources with or without further modification, but
their development and clinical use are based on rational
evidence of efficacy and safety derived from controlled
experiments and randomised clinical trials. Drugs can be
contrasted with placebos (placebo is Latin for ‘I will please’),
defined as inactive substances administered as though they
are drugs, but which have no therapeutic effects other than
potentially pleasing the patient, providing a sense of security
and progress. Pharmacology evolved on the principle of
studying known quantities of purified, active substances to
identify their specific mechanisms of action and to quantify
their effects in a reproducible manner, usually compared
with a placebo or other control substance.
Much of the success of modern medicine is based
on pharmacological science and its contribution to the
development of safe and effective pharmaceuticals. This
book is confined to pharmacology as it relates to human
medicine and aims to develop knowledge and understanding
of medical pharmacology and its application to therapeutics.
The objectives of learning about medical pharmacology and
therapeutics are:
■ to understand the ways that drugs work to affect human
systems, as a basis for safe and effective prescribing;
■ to appreciate that pharmacology must be understood
in parallel with related biological and clinical sciences,
including biochemistry, physiology and pathology;
■ to develop numerical skills for calculating drug doses
and dilutions, and to enable accurate comparison of the
relative benefits and risks of different drugs;
■ to comprehend and participate in pharmacological
research advancing the better treatment of patients.
The answer to the frequently asked question ‘What do I
need to know?’ will depend upon the individual requirements
of the programme of study being undertaken and the
examinations that will be taken. The depth and type of
Principles of pharmacology
and mechanisms of
drug action

1 4 Medical Pharmacology and Therapeutics

RECEPTORS AND
RECEPTOR-MEDIATED
MECHANISMS
Pharmacology describes how the physical interaction of drug
molecules with their macromolecular targets (‘receptors’)
modifies biochemical, immunological, physiological and
pathological processes to generate desired responses in
cells, tissues and organs. Drugs have been designed to
interact with many different types of macromolecules that
evolved to facilitate endogenous signalling between cells,
tissue and organs, or to play key roles in the normal cellular

and physiological processes that maintain controlled condi-
tions (homeostasis). Drugs may also target macromolecules

produced by pathogens, including viruses and bacteria. While
the term ‘receptor’ was originally applied in pharmacology to
describe any such drug target, more commonly a receptor
is now defined in biochemical terms as a molecule on the
surface of a cell (or inside it) that receives an external signal
and produces some type of cellular response.
The function of such a receptor can be divided typically
into three main stages:
1. The generation of a biological signal. Homeostasis is
maintained by communication between cells, tissues and
organs to optimise bodily functions and responses to
external changes. Communication is usually by signals in

the form of chemical messengers, including neurotransmit-
ter molecules, local mediators or endocrine hormones.

The signal molecule is termed a ligand, because it ligates
(ties) to the specialised cellular macromolecule. The cellular
macromolecule is a receptor because it receives the ligand.
2. Cellular recognition sites (receptors). The signal is
recognised by responding cells by its interaction with a
site of action, binding site or receptor, which may be in the
cell membrane, the cytoplasm or the nucleus. Receptors
in the cell membrane react with extracellular ligands that
cannot readily cross the cell membrane (such as peptides).
Receptors in the cytoplasm often react with lipid-soluble
ligands that can cross the cell membrane.
3. Cellular changes. Interaction of the signal and its site of
action in responding cells results in functional changes
within the cell that give rise to an appropriate biochemical
or physiological response to the original homeostatic
stimulus. This response may be cell division, a change in
cellular metabolic activity or the production of substances
that are exported from the cell.
Each of these three stages provides important targets
for drug action, and this chapter will outline the principles
underlying drug action mainly in stages 2 and 3.
ACTIONS OF DRUGS AT BINDING
SITES (RECEPTORS)
For very many drugs, the first step in producing a biological
effect is by interaction of the drug with a receptor, either on
the cell membrane or inside the cell, and it is this binding
that triggers the cellular response. Drugs may be designed to
mimic, modify or block the actions of endogenous ligands at
that receptor. The classified list of receptors at the end of this
chapter shows that cell-membrane and cytosolic receptors

knowledge required in different areas and topics will vary
when progressing through the programme; for example,
early in the course it may be important to know whether a
drug has a narrow safety margin between its wanted and
unwanted effects, and in the later years this may translate into
detailed knowledge of how the drug’s effects are monitored
in clinical use. Personal enthusiasm for medical pharmacology
is important and should be driven by the recognition that
prescribing medicines is the most common intervention
doctors (and increasingly other health professionals) use
to improve the health of their patients.
FINDING DRUG INFORMATION
Learning about medical pharmacology is best approached
using a variety of resources in a range of learning scenarios

and preferably in the context of basic science and therapeu-
tics, not from memorising lists of drug names. The following

provides a useful structure to organise the types of information
that you should aim to encounter:
■ the nonproprietary (generic) drug name (not the proprietary
or trade name);
■ the class or group to which the drug belongs;
■ the way the drug works (its mechanism of action and
its clinical effects), usually shared to variable extents by
other drugs in the same class;
■ the main clinical reasons for using the drug (its
indications);
■ any reasons why the drug should not be used in a
particular situation (its contraindications);
■ whether the drug is a prescription-only medicine (PoM)
or is available without prescription (over-the-counter
[OTC]);
■ how the drug is given (routes of administration);
■ how its effects are quantified and its doses modified if
necessary (drug monitoring);
■ how the drug is absorbed, distributed, metabolised and
excreted (ADME; its pharmacokinetics), particularly where
these show unusual characteristics;
■ the drug’s unwanted effects, including any interactions
with other drugs or foods;
■ whether there are nonpharmacological treatments that are
effective alternatives to drug treatment or will complement
the effect of the drug.
The Appendix at the end of this chapter provides a
formulary of core members of each major drug class to
give students in the early stages of training a manageable
list of the drugs most likely to be encountered in clinical
practice. At the end of later chapters, the Compendium
provides a classified listing and key characteristics of those
drugs discussed within the main text of each chapter and
also other drugs listed in the corresponding section of the
British National Formulary (BNF).
The BNF (www.evidence.nhs.uk) contains monographs
for nearly all drugs licensed for use in the United Kingdom
and is the key drug reference for UK prescribers. Students
should become familiar at an early stage with using the
BNF for reference. More detailed information on individual
drugs (the Summary of Product Characteristics [SPC]),
patient information leaflets (PIL) and contact details for
pharmaceutical companies is available from the electronic
Medicines Compendium (eMC; www.medicines.org.uk/emc/).

Principles of pharmacology and mechanisms of drug action 5

■ ion pumps and transporters, which transport specific
ions from one side of the membrane to the other in
an energy-dependent manner, usually against their
concentration gradient;
■ ion channels, which open to allow the selective, passive
transfer of ions down their concentration gradients.
Based on concentration gradients across the cell membrane:
■ both Na+
and Ca2+
ions will diffuse into the cell if their
channels are open, making the electrical potential of
the cytosol more positive and causing depolarisation of
excitable tissues;
■ K+
ions will diffuse out of the cell, making the electrical
potential of the cytosol more negative and inhibiting
depolarisation;
■ Cl−
ions will diffuse into the cell, making the cytosol more
negative and inhibiting depolarisation.
The two major families of ion channel are the ligand-gated
ion channels (LGICs) and the voltage-gated ion channels
(VGICs; also called ionotropic receptors). LGICs are opened
by the binding of a ligand, such as the neurotransmitter
acetylcholine, to an extracellular part of the channel. VGICs
in contrast are opened at particular membrane potentials by
voltage-sensing segments of the channel. Both channel types
can be targets for drug action. Both LGICs and VGICs can
control the movement of a specific ion, but a single type of
ion may flow through more than one type of channel, including
both LGIC and VGIC types. This evolutionary complexity can
be seen in the example of the multiple types of K+
channel

listed in Table 8.1.
LGICs include nicotinic acetylcholine receptors,
γ-aminobutyric acid (GABA) receptors, glycine receptors
and serotonin (5-hydroxytryptamine) 5-HT3 receptors. They
are typically pentamers, with each subunit comprising four
transmembrane helices clustering around a central channel or
pore. Each peptide subunit is orientated so that hydrophilic
chains face towards the channel and hydrophobic chains
towards the membrane lipid bilayer. Binding of an active
ligand to the receptor causes a conformational change in
the protein and results in extremely fast opening of the
ion channel. The nicotinic acetylcholine receptor is a good
example of this type of structure (Fig. 1.1). It requires the
binding of two molecules of acetylcholine for channel opening,
which lasts only milliseconds because the ligand rapidly
dissociates and is inactivated. Drugs may modulate LGIC
activity by binding directly to the channel, or indirectly by
acting on G-protein-coupled receptors (GPCRs; discussed
later), with the subsequent intracellular events then affecting
the status of the LGIC.
VGICs include Ca2+
, Na+
and K+
channels. The K+
channels
consist of four distinct peptide subunits, each of which
has between two and six transmembrane helices; in Ca2+
and Na+
channels there are four domains, each with six
transmembrane helices, within a single large protein. The
pore-forming regions of the transmembrane helices are largely
responsible for the selectivity of the channel for a particular
ion. Both Na+
and K+
channels are inactivated after opening;
this is produced by an intracellular loop of the channel,
which blocks the open channel from the intracellular end.
The activity of VGICs may thus be modulated by drugs acting
directly on the channel, such as local anaesthetics which
maintain Na+
channels in the inactivated site by binding at
an intracellular site (Chapter 18). Drugs may also modulate

tend to occur in different families (receptor types), reflecting
their evolution from common ancestors. Within any one family
of receptors, different receptor subtypes have evolved to
facilitate increasingly specific signalling and distinct biological
effects. As might be expected, different receptor families
have different characteristics, but subtypes within each family
retain common family traits.
In pharmacology, the perfect drug would be one that binds
only to one type or subtype of receptor and consistently
produces only the desired biological effect without the
unwanted effects that can occur when drugs bind to a
related receptor. Although this ideal is impossible to attain,
it has proved possible to develop drugs that bind avidly
to their target receptor to produce their desired effect and
have very much less (but not zero) ability to bind to other
receptors, even ones within the same family, which might
produce unwanted effects.
Where a drug binds to one type of receptor in preference
to another, it is said to show selectivity of binding or selectivity
of drug action. Selectivity is never absolute but is high with
some drugs and lower with others. A drug with a high degree
of selectivity is likely to show a greater difference between
the dose required for its biological action and the dose that
produces unwanted actions at other receptor types. Even
a highly selective drug may produce unwanted effects if
its target receptors are also found in tissues and organs
other than those in which the drug is intended to produce
its therapeutic effect.
MAJOR TYPES OF RECEPTORS
Despite the great structural diversity of drug molecules, most
act on the following major types of receptors to bring about
their pharmacological effects:
■ Transmembrane ion channels. These control the passage
of ions across membranes and are widely distributed.
■ Seven-transmembrane (7TM) (heptahelical) receptors.
This is a large family of receptors, most of which signal
via guanine nucleotide-binding proteins (G-proteins).
Following activation by a ligand, second messenger
substances are formed inside the cell, which can bring
about cellular molecular changes, including the opening
of transmembrane ion channels.
■ Enzyme-linked transmembrane receptors. This is a
family of transmembrane receptors with an integral or
associated enzymic component, such as a kinase or
phosphatase. Activation of these enzymes produces
changes in cells by phosphorylating or dephosphorylating
intracellular proteins, including the receptor itself, thereby
altering their activity.
■ Intracellular (nuclear) receptors. These receptors are
found in the nucleus or translocate to the nucleus from the
cytosol to modify gene transcription and the expression
of specific cellular proteins.
Transmembrane ion channels
Transmembrane ion channels that create pores across
phospholipid membranes are ubiquitous and allow the
transport of ions into and out of cells. The intracellular
concentrations of ions are controlled by a combination of
two types of ion channel:

1 6 Medical Pharmacology and Therapeutics

7TM receptors and they are the targets of over 30% of current
drugs. The function of over a hundred 7TM receptors is still
unknown. The structure of a hypothetical 7TM receptor is
shown in Fig. 1.2; the N-terminal region of the polypeptide
chain is on the extracellular side of the membrane, and
the polypeptide traverses the membrane seven times with
helical regions, so that the C terminus is on the inside of
the cell. The extracellular loops provide the receptor site
for an appropriate agonist (a natural ligand or a drug), the
binding of which alters the three-dimensional conformation
of the receptor protein. The intracellular loops are involved in
coupling this conformational change to the second messenger
system, usually via a heterotrimeric G-protein, giving rise to
the term GPCR.
The G-protein system
The heterotrimeric G-protein system (Fig. 1.3) consists of
α, β and γ subunits.
■ The α-subunit. More than 20 different types have been
identified, belonging to four families (αs, αi

, αq and α12/13).
The α-subunit is important because it binds guanosine

VGICs indirectly via intracellular signals from other receptors.
For example, L-type Ca2+

channels are inactivated directly
by calcium channel blockers, but also indirectly by drugs
which reduce intracellular signalling from the β1 subtype of
adrenoceptors (see Fig. 5.5).
The ability of highly variable transmembrane subunits to
assemble in a number of configurations leads to the existence
of many different subtypes of channels for a single ion.
For example, there are many different voltage-gated Ca2+
channels (L, N, P/Q, R and T types).
Seven-transmembrane receptors
Also known as 7TM receptors, heptahelical receptors and
serpentine receptors, this family is an extremely important
group, as the human genome has about 750 sequences for

ACh
C

N

ACh
α

α
γ

δ
β
M1 M2 M3 M4

Extracellular
Na+

Intracellular

A

B
Fig. 1.1 The acetylcholine nicotinic receptor, a typical
ligand-gated transmembrane ion channel. (A) The receptor
is constructed from subunits with four transmembrane
regions (M1–M4). (B) Five subunits are assembled into the ion
channel, which has two sites for acetylcholine binding, each
formed by the extracellular domains of two adjacent subunits.
On acetylcholine binding, the central pore undergoes
conformational change that allows selective Na+
ion flow
down its concentration gradient into the cell. N, amino
terminus; C, carboxyl terminus.

Extracellular

Intracellular
Lipid layer

Agonist
binding site

Cell membrane

Carboxyl end

Amino end

Fig. 1.2 Hypothetical seven-transmembrane (7TM)
receptor. The 7TM receptor is a single polypeptide chain
with its amino (N-) terminus outside the cell membrane
and its carboxyl (C-) terminus inside the cell. The chain
is folded such that it crosses the membrane seven times,
with each hydrophobic transmembrane region shown
here as a thickened segment. The hydrophilic extracellular
loops create a confined three-dimensional environment in
which only the appropriate ligand can bind. Other potential
ligands may be too large for the site or show much weaker
binding characteristics. Selective ligand binding causes
conformational change in the three-dimensional form of the
receptor, which activates signalling proteins and enzymes
associated with the intracellular loops, such as G-proteins
and nucleotide cyclases.

Principles of pharmacology and mechanisms of drug action 7

Cyclic nucleotide system
This system is based on cyclic nucleotides, such as:
■ Cyclic adenosine monophosphate (cAMP), which is
synthesised from adenosine triphosphate (ATP) by adenylyl
cyclase. cAMP induces numerous cellular responses by
activating protein kinase A (PKA), which phosphorylates
proteins, many of which are enzymes. Phosphorylation
can either activate or suppress cell activity.
■ Cyclic guanosine monophosphate (cGMP), which is
synthesised from GTP by guanylyl cyclase. cGMP exerts
most of its actions through protein kinase G, which, when
activated by cGMP, phosphorylates target proteins.
There are 10 isoforms of adenylyl cyclase; these
show different tissue distributions and could be important
sites of selective drug action in the future. The cyclic
nucleotide second messenger (cAMP or cGMP) is inactivated
by hydrolysis by phosphodiesterase (PDE) isoenzymes to
give AMP or GMP. There are 11 different families of PDE
isoenzymes (Table 1.1), some of which are the targets of
important drug groups, including selective PDE4 inhibitors
used in respiratory disease and PDE5 inhibitors used in
erectile dysfunction.
The phosphatidylinositol system
The other second messenger system is based on
inositol 1,4,5-triphosphate (IP3) and diacylglycerol (DAG),

diphosphate (GDP) and guanosine triphosphate (GTP)
in its inactive and active states, respectively; it also has
GTPase activity, which is involved in terminating its own
activity. When an agonist binds to the receptor, GDP
(which is normally present on the α-subunit) is replaced
by GTP. The active α-subunit–GTP dissociates from the
βγ-subunits and can activate enzymes such as adenylyl
cyclase. The α-subunit–GTP complex is inactivated when
the GTP is hydrolysed back to GDP by the GTPase.
■ The βγ-complex. There are many different isoforms
of β- and γ-subunits that can combine into dimers,
the normal function of which is to inhibit the α-subunit
when the receptor is unoccupied. When the receptor is
occupied by a ligand, the βγ-complex dissociates from the
α-subunit and can itself activate cellular enzymes, such
as phospholipase C. The α-subunit–GDP and βγ-subunit
then recombine with the receptor protein to give the
inactive form of the receptor–G-protein complex.
Second messenger systems
Second messengers are the key distributors of an external
signal, as they are released into the cytosol as a consequence
of receptor activation and are responsible for affecting a wide
variety of intracellular enzymes, ion channels and transporters.
There are two complementary second messenger systems:
the cyclic nucleotide system and the phosphatidylinositol
system (Fig. 1.4).

β
Agonist
binding

Replacement
of GDP

E β γ E E α β E by GTP E α β E

γ γ

α γ γ

E E
α
Inactive receptor
GDP GDP GTP
Recombination of
GDP α- and βγ-subunits
with transmembrane
receptor

GTP
hydrolysis E

Dissociation

GDP

Intracellular
effects

Intracellular
effects

Intracellular
effects
β
E

α
GTP

Fig. 1.3 The functioning of G-protein subunits. Ligand (agonist) binding results in replacement of GDP on the α-subunit by
guanosine triphosphate (GTP) and the dissociation of the α- and βγ-subunits, each of which can affect a range of intracellular
systems (shown as E in the figure) such as second messengers (e.g. adenylyl cyclase and phospholipase C), or other enzymes
and ion channels (see Figs 1.4 and 1.5). Hydrolysis of GTP to GDP inactivates the α-subunit, which then recombines with the
βγ-dimer to reform the inactive receptor.

1 8 Medical Pharmacology and Therapeutics

consequences of GPCR dimerisation and its implications
for drug therapy are unclear.
Protease-activated receptors
Protease-activated receptors (PARs) are GPCRs stimulated
unusually by a ‘tethered ligand’ located within the N terminus
of the receptor itself, rather than by an independent ligand.
Proteolysis of the N-terminal sequence by serine proteases
such as thrombin, trypsin and tryptase enables the residual
tethered ligand to bind to the receptor within the second
extracellular loop (Fig. 1.6). To date, four protease-activated
receptors (PAR 1–4) have been identified, each with distinct
N-terminal cleavage sites and different tethered ligands. The
receptors appear to play roles in platelet activation and
clotting (Chapter 11), and in inflammation and tissue repair.
Most of the actions of PAR are mediated by Gi

, Gq and G12/13.
Enzyme-linked transmembrane
receptors
Enzyme-linked receptors, most notably the receptor
tyrosine kinases, are similar to the GPCRs in that they
have a ligand-binding domain on the surface of the cell
membrane; they traverse the membrane; and they have an
intracellular effector region (Fig. 1.7). They differ from GPCRs
in their extracellular ligand-binding domain, which is very
large to accommodate their polypeptide ligands (including
hormones, growth factors and cytokines), and in having
only one transmembrane helical region. Importantly, their
intracellular action requires a linked enzymic domain, most
commonly an integral kinase which activates the receptor
itself or other proteins by phosphorylation. Activation of
enzyme-linked receptors enables binding and activation of
many intracellular signalling proteins, leading to changes

which are synthesised from the membrane phospholipid
phosphatidylinositol 4,5-bisphosphate (PIP2) by phospholipase
C (see Fig. 1.4). There are a number of isoenzymes of
phospholipase C, which may be activated by the α-subunit–
GTP or βγ-subunits of G-proteins. The main function of IP3
is to mobilise Ca2+

in cells. With the increase in Ca2+
brought
about by IP3, DAG is able to activate protein kinase C (PKC)
and phosphorylate target proteins. IP3 and DAG are then
inactivated and converted back to PIP2.
Which second messenger system is activated when a
GPCR binds a selective ligand depends primarily on the
nature of the Gα-subunit, as illustrated in Fig. 1.5:
■ Gs: Stimulation of adenylyl cyclase (increases cAMP),
activation of Ca2+
channels

■ Gi/o: Inhibition of adenylyl cyclase (reduces cAMP),
inhibition of Ca2+

channels, activation of K+
channels
■ Gq/11: Activation of phospholipase C, leading to DAG and
IP3 signalling
■ G12/13: Activation of cytoskeletal and other proteins via the
Rho family of GTPases, which influence smooth muscle
contraction and proliferation
The βγ-complex also has signalling activity: it can activate
phospholipases and modulate some types of K+
and Ca2+

channels.
Activation of these second messenger systems by
G-protein subunits thus affects many cellular processes
such as enzyme activity (either directly or by altering gene
transcription), contractile proteins, ion channels (affecting
depolarisation of the cell) and cytokine production. The many
different isoforms of Gα, Gβ and Gγ proteins may represent
important future targets for selective drugs.
It is increasingly recognised that GPCRs may assemble
into dimers of identical 7TM proteins (homodimers) or into
heterodimers of different receptor proteins; the functional

P

P P P

P

P

P P

P P
Phospholipase C

DAG
Inactivation by
phosphorylation

P

Hydrolysis
to inositol
IP3

adenosine ribose
ATP
Adenylyl
cyclase
adenosine ribose
cAMP
Inactivation by
phosphodiesterase
adenosine ribose
5¢- AMP

acyl glycerol inositol
acyl

PIP2

acyl
acyl glycerol

inositol

Fig. 1.4 Second messenger systems. Stimulation of G-protein-coupled receptors produces intracellular changes by
activating or inhibiting cascades of second messengers. Examples are cyclic adenosine monophosphate (cAMP), and
diacylglycerol (DAG) and inositol triphosphate (IP3) formed from phosphatidylinositol 4,5-bisphosphate (PIP2). See also Fig. 1.5.

Principles of pharmacology and mechanisms of drug action 9

receptors and signalling via the JAK/Stat pathways to
affect inflammatory gene expression.
■ Receptor serine-threonine kinase family. Activation of
these phosphorylates serine and threonine residues in
target cytosolic proteins; everolimus is a serine-threonine
kinase inhibitor used in renal and pancreatic cancer.
■ Receptor guanylyl cyclase family. Members of this family
catalyse the formation of cGMP from GTP via a cytosolic
domain.
Intracellular (nuclear) receptors
Many hormones act at intracellular receptors to produce
long-term changes in cellular activity by altering the genetic
expression of enzymes, cytokines or receptor proteins. Such
hormones are lipophilic to facilitate their movement across the
cell membrane. Examples include the thyroid hormones and
the large group of steroid hormones, including glucocorticoids,
mineralocorticoids and the sex steroid hormones. Their
actions on DNA transcription are mediated by interactions
with intracellular receptors (Table 1.2) located either in the
cytoplasm (type 1) or the nucleus (type 2).

in gene transcription and in many cellular functions. There
are five families of enzyme-linked transmembrane receptors:
■ Receptor tyrosine kinase (RTK) family. Ligand binding
causes receptor dimerisation and transphosphorylation of
tyrosine residues within the receptor itself and sometimes
in associated cytoplasmic proteins. Up to 20 classes of
RTK include receptors for growth factors, many of which
signal via proteins of the mitogen-activated protein (MAP)
kinase cascade, leading to effects on gene transcription,
apoptosis and cell division. Constitutive overactivity of
an RTK called Bcr-Abl causes leucocyte proliferation in
chronic myeloid leukaemia, which is treated with imatinib,
a drug that blocks the uncontrolled RTK activity. Several
other RTKs are also the targets of anticancer drugs.

■ Tyrosine phosphatase receptor family. These dephos-
phorylate tyrosines on other transmembrane receptors

or cytoplasmic proteins; they are particularly common in
immune cells.
■ Tyrosine kinase-associated receptor family (or
nonreceptor tyrosine kinases). These lack integral
kinase activity but activate separate kinases associated
with the receptor; examples include inflammatory cytokine
Table 1.1 Isoenzymes of phosphodiesterase
Enzyme Main substrate Main site(s) Examples of
inhibitors

Therapeutic potential

PDE1 cAMP + cGMP Heart, brain, lung, lymphocytes,
vascular smooth muscle

– Atherosclerosis?

PDE2 cAMP + cGMP Adrenal gland, brain, heart, lung,
liver, platelets, endothelial
cells

– Involved in memory?

PDE3 cAMP + cGMP Heart, lung, liver, platelets,
adipose tissue, inflammatory
cells, smooth muscle

Aminophylline
Enoximone
Milrinone
Cilostazol

Asthma (Chapter 12)
Congestive heart failure (Chapter 7)
Peripheral vascular disease
(Chapter 10)

PDE4 cAMP Sertoli cells, endothelial cells,
kidney, brain, heart, liver,
lung, inflammatory cells

Aminophylline
Roflumilast

Asthma, COPD (Chapter 12)
Inflammation
IBD?

PDE5 cGMP Smooth muscle, endothelium,
neurons, lung, platelets

Sildenafil
Tadalafil
Vardenafil
Dipyridamole

Erectile dysfunction (Chapter 16)
Pulmonary hypertension
(Chapter 6)

PDE6 cGMP Photoreceptors, pineal gland Dipyridamole
Sildenafil (weak)
Undefined

PDE7 cAMP Skeletal muscle, heart, kidney,
brain, pancreas, spinal cord,
T-lymphocytes

– Inflammation (combined with PDE4

inhibitor)?
Spinal cord injury?

PDE8 cAMP Testes, eye, liver, skeletal
muscle, heart, kidney, ovary,
brain, T-lymphocytes

– Undefined

PDE9 cGMP Kidney, liver, lung, brain – Undefined
PDE10 cAMP + cGMP Testes, brain, thyroid – Schizophrenia?
PDE11 cAMP + cGMP Skeletal muscle, prostate,
kidney, liver, pituitary and
salivary glands, testes

Tadalafil (weak) Undefined

cAMP, Cyclic adenosine monophosphate; cGMP, cyclic guanosine monophosphate; COPD, chronic obstructive pulmonary disease; IBD,
inflammatory bowel disease, PDE, phosphodiesterase.

1 10 Medical Pharmacology and Therapeutics

Adenylyl cyclase/
guanylyl cyclase

cAMP/
cGMP

Protein kinases
(e.g., A, G)

Receptor-
activated

G-protein
Gi
Gs

Gq

Phospholipase C

DAG

IP3
Protein kinase C
Release of Ca2+
from sarcoplasmic
reticulum

Intracellular enzymes
Ion channels (Ca2+ and K+)
Contractile proteins

+
+

Fig. 1.5 The intracellular consequences of receptor activation. The second messengers cyclic adenosine monophosphate
(cAMP), cyclic guanosine monophosphate (cGMP), diacylglycerol (DAG) and inositol 1,4,5-triphosphate (IP3) produce a number
of intracellular changes, either directly or indirectly via actions on protein kinases (which phosphorylate other proteins) or
by actions on ion channels. The pathways can be activated or inhibited depending upon the type of receptor and G-protein
and the particular ligand stimulating the receptor. The effect of the same second messenger can vary depending upon the
biochemical functioning of cells in different tissues.

G-protein G-protein G-protein

Protease
hydrolysis

Inactive receptor Protease activation Active receptor
Amino acid sequence with agonist activity

Second
messengers

Fig. 1.6 Protease-activated receptors. These G-protein-coupled receptors are activated by proteases such as thrombin
which hydrolyse the extracellular peptide chain to expose a segment that acts as a tethered ligand (shown in red) and activates
the receptor. The receptor is inactivated by phosphorylation of the intracellular (C-terminal) part of the receptor protein.

Principles of pharmacology and mechanisms of drug action 11

into the nucleus. Via their DNA-binding domain, the active
hormone–receptor complexes can interact with hormone
response elements (HRE) at numerous sites in the genome.
Binding to the HRE usually activates gene transcription, but
sometimes it silences gene expression and decreases mRNA
synthesis.
Translocation and binding to DNA involves a variety of
different chaperone, co-activator and co-repressor proteins,
and the system is considerably more complex than indicated

The intracellular receptor typically includes a highly con-
served DNA-binding domain with zinc-containing loops and

a variable ligand-binding domain (Table 1.3). The sequence

of hormone binding and action for type 1 intracellular recep-
tors is shown in Fig. 1.8. Type 1 receptors are typically found

in an inactive form in the cytoplasm linked to chaperone
proteins such as heat-shock proteins (HSPs). Binding of the
hormone induces conformational change in the receptor;
this causes dissociation of the HSP and reveals a nuclear
localisation sequence (or NLS) which enables the hormone–
receptor complex to pass through nuclear membrane pores

OH

OH
Ligand

OPO3
Ligand

2–
Ligand

OPO3
2–
OH
Ligand

Single
inactive
receptor

Activation
of
intracellular
enzymes

Ligand-binding site

Extracellular

Intracellular
Tyrosine
residue
Ligand binding
to two receptors

Mutual phosphorylation
and activation

Fig. 1.7 Enzyme-linked transmembrane receptors. This
receptor tyrosine kinase has a large extracellular domain,
a single transmembrane segment and an integral kinase
domain. Ligand binding causes phosphorylation of tyrosine
residues on the receptor and on other target proteins, leading

to intracellular changes in cell behaviour. Other enzyme-
linked receptors have tyrosine phosphatase, serine-threonine

kinase or guanylyl cyclase enzymic activity.

Table 1.2 Some families of intracellular receptors
Subtypes

Type 1 (cytoplasmic)
Oestrogen receptors ER (α, β)
Progesterone receptors PR (A, B)
Androgen receptors AR (A, B)
Glucocorticoid receptor GR
Mineralocorticoid receptor MR
Type 2 (nuclear)
Thyroid hormone receptors TR (α1, β1, β2)
Vitamin D receptor VDR
Retinoic acid receptors RAR (α, β, γ)
Retinoid X receptors RXR (α, β, γ)
Liver X (oxysterol) receptors LXR (α, β)
Peroxisome proliferator-activated
receptors

PPAR (α, β/δ, γ)

Table 1.3 The structure of steroid hormone
receptors
Section
of protein

Domain Role
A/B N-terminal
variable domain

Regulates transcriptional
activity
C DNA-binding
domain (DBD)

Highly conserved; binds
receptor to hormone
response element in

DNA by two zinc-
containing regions

D Hinge region Enables intracellular
translocation to the
nucleus
E Ligand-binding
domain (LBD)

Moderately conserved;
enables specific ligand
binding; contains
nuclear localisation
sequence (NLS); also
binds chaperone
proteins and facilitates
receptor dimerisation

F C-terminal
domain

Highly variable;
unknown function

1 12 Medical Pharmacology and Therapeutics

fibrate class of lipid-lowering drugs act on specific members
of the PPAR family of type 2 receptors.
OTHER SITES OF DRUG ACTION
Probably every protein in the human body has the
potential to have its structure or activity altered by foreign
compounds. Traditionally, all drug targets were described
pharmacologically as ‘receptors’, although many drug targets
would not be defined as receptors in biochemical terms; in
addition to the receptor types discussed previously, drugs
may act at numerous other sites.
■ Cell-membrane ion pumps. In contrast to passive
diffusion, primary active transport of ions against their
concentration gradients occurs via ATP-dependent ion
pumps, which may be drug targets. For example, Na+
/

K+
-ATPase in the brain is activated by the anticonvulsant
drug phenytoin, whereas in cardiac tissue it is inhibited by
digoxin; K+
/H+
-ATPase in gastric parietal cells is inhibited
by proton pump inhibitors such as omeprazole.
■ Transporter (carrier) proteins. Secondary active transport
involves carrier proteins, which transport a specific ion or
organic molecule across a membrane; the energy for the
transport derives not from a coupled ATPase but from the
co-transport of another molecule down its concentration
gradient, either in the same direction (symport) or in the
opposite direction (antiport). Examples include:
■ Na+
/Cl−
co-transport in the renal tubule, which is
blocked by thiazide diuretics (Chapter 14);
■ the reuptake of neurotransmitters into nerve terminals
by a number of transporters selectively blocked by
classes of antidepressant drugs (Chapter 22).
■ Enzymes. Many drugs act on the intracellular or
extracellular enzymes that synthesise or degrade the
endogenous ligands for extracellular or intracellular
receptors, or which are required for growth of bacterial,
viral or tumour cells. Table 1.4 provides examples of drug
groups that act on enzyme targets. The PDE isoenzymes
that regulate second messenger molecules are important
drug targets and are listed in Table 1.1. In addition to being
sites of drug action, enzymes are involved in inactivating
many drugs, while some drugs are administered as inactive
precursors (pro-drugs) that are enzymatically activated
(Chapter 2).
■ Adhesion molecules. These regulate the cell-surface
interactions of immune cells with endothelial and other
cells. Natalizumab is a monoclonal antibody directed
against the α4-integrin component of vascular cell
adhesion molecule (VCAM)-1 and is used to inhibit the
autoimmune activity of lymphocytes in acute relapsing
multiple sclerosis. Other monoclonal antibody-based
therapies are targeted at cellular and humoral proteins,
including cytokines and intracellular signalling proteins
to suppress inflammatory cell proliferation, activity and
recruitment in immune disease.
■ Organelles and structural proteins. Examples include
some antimicrobials that interfere with the functioning
of ribosomal proteins in bacteria, and some types of
anticancer drugs that interrupt mitotic cell division by
blocking microtubule formation.
The sites of action of some drugs remain unknown
or poorly understood. Conversely, many receptors have

in Fig. 1.8. Co-activators are transcriptional cofactors that
also bind to the receptor and increase the level of gene
induction; an example is histone acetylase, which facilitates
transcription by increasing the ease of unravelling of DNA from
histone proteins. Co-repressors also bind to the receptor and
repress gene activation; an example is histone deacetylase,
which prevents further transcription by tightening histone
interaction with the DNA.
Type 2 intracellular receptors, such as the thyroid hormone
receptors (TR) and the peroxisome proliferator-activated
receptor (PPAR) family (see Table 1.2), are found within the
nucleus bound to co-repressor proteins, which are liberated
by ligand binding without a receptor translocation step from
the cytoplasm. PPAR nuclear receptors function as sensors
for endogenous fatty acids, including eicosanoids (Chapter
29), and regulate the expression of genes that influence
metabolic events.
Intracellular receptors are the molecular targets of 10–15%
of marketed drugs, including steroid drugs acting at type 1
receptors and other drugs acting at type 2 receptors. Steroids
show selectivity for different type 1 intracellular receptors
(ER, PR, AR, GR, MR; see Table 1.2), which determine the
spectrum of gene expression that is affected (Chapters 14,
44, 45 and 46). Steroid effects are also determined by the
differential expression of these receptors in different tissues.
Intracellular hormone–receptor complexes typically dimerise
to bind to their HRE sites on DNA. Steroid receptors form
homodimers (e.g. ER–ER), while most type 2 receptors
form heterodimers, usually with RXR (e.g. RAR–RXR). The
thiazolidinedione drugs used in diabetes mellitus and the
Cell membrane

Nuclear
membrane

mRNA mRNA

Increased synthesis
of cytokines,
enzymes, receptors Decreased synthesis
of cytokines,
enzymes, receptors

HSP90
HR
HSP90
HR ST
HR ST

ST

HRE Gene

Fig. 1.8 The activation of intracellular hormone
receptors. Steroid hormones (ST) are lipid-soluble
compounds which readily cross cell membranes and bind
to their intracellular receptors (HR). This binding displaces a
chaperone protein called heat-shock protein (HSP90) and the
hormone–receptor complex enters the nucleus, where it can
increase or decrease gene expression by binding to hormone
response elements (HRE) on DNA. Intracellular receptors for
many other ligands are activated in the nucleus itself.

Principles of pharmacology and mechanisms of drug action 13

ligand and its receptor does not involve covalent chemical
bonds but weaker, reversible forces, such as:
■ ionic bonding between ionisable groups in the ligand
(e.g. NH3
+
) and the receptor (e.g. COO−
);

■ hydrogen bonding between amino-, hydroxyl-, keto- and
other groups in the ligand and the receptor;
■ hydrophobic interactions between lipid-soluble sites in
the ligand and receptor;
■ van der Waals forces, which are very weak interatomic
attractions.
The receptor protein is not a rigid structure: binding of
the ligand alters the conformation and biological properties
of the protein, enabling it to trigger intracellular signalling
pathways. There is growing recognition that different ligands
may stabilise different conformational states of the same
receptor that are distinct from those produced by the
endogenous ligand. Rather than simply switching a receptor
between inactive and active states, a ‘biased’ ligand may
produce preferential receptor signalling via specific G-protein
pathways or by non-G-protein effectors, such as the family
of arrestin proteins, leading to different cellular behaviours.
Drugs may therefore be developed with functional selectivity

been discovered for which the natural ligands are not yet
recognised; these orphan receptors may represent targets for
novel drugs when their pharmacology is better understood.
PROPERTIES OF RECEPTORS
Receptor binding
The binding of endogenous ligands and most drugs to their
receptors is normally reversible; consequently, the intensity
and duration of the intracellular changes are dependent on
repeated ligand–receptor interactions that continue for as
long as the ligand molecules remain in the local environment
of the receptors. The duration of activity of a reversible drug
therefore depends mainly on its distribution and elimination
from the body (pharmacokinetics), which typically requires
hours or days (Chapter 2), not on the duration of binding
of a drug molecule to its receptor, which may last only a
fraction of a second. For a reversible drug, the extent of drug
binding to the receptor (receptor occupancy) is proportional
to the drug concentration; the higher the concentration, the
greater the occupancy. The interaction between a reversible
Table 1.4 Examples of enzymes as drug targets
Enzyme Drug class or use Examples
Acetylcholinesterase (AChE) AChE inhibitors (Chapter 27) Neostigmine, edrophonium,
organophosphates
Angiotensin-converting enzyme (ACE) ACE inhibitors (Chapter 6) Captopril, perindopril, ramipril
Antithrombin (AT)III Heparin anticoagulants (enhancers of
antithrombin III) (Chapter 10)

Enoxaparin, dalteparin

Carbonic anhydrase Carbonic anhydrase inhibitors
(Chapters 14, 50)

Acetazolamide

Cyclo-oxygenase (COX)-1 Nonsteroidal anti-inflammatory drugs (NSAIDs)

(Chapter 29)

Ibuprofen, indometacin,
naproxen
Cyclo-oxygenase (COX)-2 Selective COX-2 inhibitors (Chapter 29) Celecoxib, etoricoxib
Dihydrofolate reductase Folate antagonists (Chapters 51, 52) Trimethoprim, methotrexate
DOPA decarboxylase Peripheral decarboxylase inhibitors (PDIs)

(Chapter 24)

Carbidopa, benserazide
Coagulation factor Xa Direct inhibitors of Factor Xa (Chapter 11) Rivaroxaban
HMG-CoA reductase Statins (HMG-CoA reductase inhibitors)

(Chapter 48)

Atorvastatin, simvastatin

Monoamine oxidases (MAOs) A and B MAO-A and MAO-B inhibitors
(Chapters 22, 24)

Moclobemide, selegiline
Phosphodiesterase (PDE) isoenzymes PDE inhibitors (Chapters 12, 16) Theophylline, sildenafil
(see Table 1.1)

Reverse transcriptase (RT) Nucleos(t)ide and nonnucleoside RT inhibitors

(Chapter 51)

Zidovudine, efavirenz
Ribonucleotide reductase Ribonucleotide reductase inhibitor (Chapter 52) Hydroxycarbamide (hydroxyurea)
Thrombin Direct thrombin inhibitors (Chapter 11) Dabigatran
Viral proteases HIV/hepatitis protease inhibitors (Chapter 51) Saquinavir, boceprevir
Vitamin K epoxide reductase Coumarin anticoagulants (Chapter 11) Warfarin
Xanthine oxidase Xanthine oxidase inhibitors (Chapter 31) Allopurinol

1 14 Medical Pharmacology and Therapeutics

to generate different cell responses from the same receptor,
in contrast to the classical concept of different responses
being generated by drugs acting at different receptors.
Receptor selectivity
There are numerous possible extracellular and intracellular
signals produced in the body, which can affect many different
processes. Therefore a fundamental property of a useful
ligand–receptor interaction is its selectivity; that is, the
extent to which the receptor can recognise and respond to
the correct signals, represented by one ligand or group of
related ligands. Some receptors show high selectivity and
bind a single endogenous ligand (e.g. acetylcholine is the
only endogenous ligand that binds to N1 nicotinic receptors;
see Chapter 4), whereas other receptors are less selective
and will bind a number of related endogenous ligands (e.g.
the β1-adrenoceptors on the heart will bind noradrenaline,
adrenaline and to some extent dopamine, all of which are
catecholamines).
The ability of receptors to recognise and bind the
appropriate ligand depends on the intrinsic characteristics
of the chemical structure of the ligand. The formulae of
a few ligand families that bind to different receptors are
shown in Fig. 1.9; differences in structure that determine
selectivity of action between receptors may be subtle,
such as the those illustrated between the structures of
testosterone and progesterone, which nevertheless have
markedly different hormonal effects on the body due to their
receptor selectivity. Receptors are protein chains folded into
tertiary and quaternary structures such that the necessary
arrangement of specific binding centres is brought together
in a small volume – the receptor site (Fig. 1.10). Receptor
selectivity occurs because the three-dimensional organisation
of the different sites for reversible binding (such as anion
and cation sites, lipid centres and hydrogen-bonding sites)
corresponds better to the three-dimensional structure of the
endogenous ligand than to that of other ligands.
There may be a number of subtypes of a receptor, all of
which can bind the same common ligand but which differ
in their ability to recognise particular variants or derivatives
of that ligand. The different characteristics of the receptor
subtypes therefore allow a drug (or natural ligand) with a
particular three-dimensional structure to show selective
actions by recognising one receptor preferentially, with
fewer unwanted effects from the stimulation or blockade
of related receptors. For example, α1-, α2-, β1-, β2- and
β3-adrenoceptors all bind adrenaline, but isoprenaline, a
synthetic derivative of adrenaline, binds selectively to the three
β-adrenoceptor subtypes rather than the two α-adrenoceptor
subtypes (Chapter 4). As the adrenoceptor subtypes occur to
a different extent in different tissues, and produce different
intracellular changes when stimulated or blocked, drugs
can be designed that have highly selective and localised
actions. The cardioselective β-adrenoceptor antagonists such
as bisoprolol are selective blockers of the β1-adrenoceptor
subtype that predominates on cardiac smooth muscle, with
much less binding to the β2-adrenoceptors that predominate
on bronchial smooth muscle. Although ligands may have a
much higher affinity for one receptor subtype over another,
this is never absolute, so the term selective is preferred
over specific.

CHCH2NH2

OH
OH
OH

CH2CH2NH2

OH
OH
HO CH2CH2NH2

N
H

Dopamine Noradrenaline

5-Hydroxytryptamine (5-HT; serotonin)

CH2CH2NH2 N
HN

A Histamine
NH2
CH2
COOH

NH2
CH2CH2CH
COOH

O

HO

CH2CH2CH2
NH2

HO
NH2 O

C

COOH

O

HO
CH2 CH
Glycine Glutamate

Aspartate - Aminobutyrate
(GABA)
γ
C

C

B

O

C O
CH3

Progesterone
O
Testosterone
OH

C
Fig. 1.9 Groups of related chemicals that show
selectivity for different receptor subtypes in spite of
similar structure. (A) Biogenic amines; (B) amino acids; (C)
steroids.

Principles of pharmacology and mechanisms of drug action 15

mode of treatment and selection of the optimal drug and
dosage (personalised medicine).
Drug stereochemistry and activity
The three-dimensional spatial organisation of receptors
means that the ligand must have the correct configuration
to fit the receptor, analogous to fitting a right hand into
a right-handed glove. Drugs and other organic molecules
show stereoisomerism if they contain four different chemical
groups attached to a single carbon atom, or one or more
double bonds, with the result that compounds with the same
molecular formula can exist in different three-dimensional
configurations. If a drug is an equal (racemic) mixture of two
stereoisomers, the stereoisomers may show different receptor
binding characteristics and biological properties. Most often,
one stereoisomer is pharmacologically active while the other
is inactive, but in some cases the inactive isomer may be
responsible for the unwanted effects of the racemic mixture.
Alternatively the two isomers may be active at different
receptor subtypes and have synergistic or even opposing
actions. The different isomers may also show different rates
of metabolism. As a consequence, there has been a trend
for the development of single stereoisomers of drugs for
therapeutic use; one of the earliest examples was the use of
levodopa, the levo-isomer of dihydroxyphenylalanine (DOPA)
in Parkinson’s disease (Chapter 24).
Receptor numbers
The number of receptors present in, or on the surface of, a cell
is not static. There is usually a high turnover of receptors being
formed and removed continuously. Cell-membrane receptor
proteins are synthesised in the endoplasmic reticulum
and transported to the plasma membrane. Regulation of
functional receptor numbers in the membrane occurs both by
transport to the membrane (often as homo- or heterodimers)
and by removal by internalisation. The number of receptors
within the cell membrane may be altered by the drug being
used for treatment, with either an increase in receptor
number (upregulation) or a decrease (downregulation) and
a consequent change in the ability of the drug to effect
the desired therapeutic response. This change may be an
unwanted loss of drug activity contributing to tolerance
to the effects of the drug (e.g. opioids; Chapter 19). As
a result, increased doses may be needed to maintain the
same activity. Alternatively, the change in receptor number
may be an important part of the therapeutic response itself.
One example is tricyclic antidepressants (Chapter 22);
these produce an immediate increase in the availability of
monoamine neurotransmitters, but the therapeutic response
is associated with a subsequent, adaptive downregulation in
monoamine receptor numbers occurring over several weeks.
PROPERTIES OF DRUG ACTION
Drug actions can show a number of important properties:
■ dose–response relationship,
■ selectivity,
■ potency,
■ efficacy.

Traditionally, receptor subtypes were discovered pharma-
cologically when a new agonist or antagonist compound was

found to alter some but not all of the activities of a currently
known receptor class. Developments in molecular biology,
including the Human Genome Project, have accelerated
the recognition and cloning of new receptors and receptor
subtypes, including orphan receptors for which the natural
ligands are unknown. These developments are important in
guiding development of new drugs with greater selectivity
and fewer unwanted effects. Based on such information it is
recognised that there are multiple types of most receptors,
and that there is genetic variation among individuals in the
structures, properties and abundance of these receptors,

which can lead to differences in drug responses (pharma-
cogenetic variation; discussed later). Greater understanding

of genetic differences underlying human variability in drug
responses offers the potential for individualisation of the

NH2
OH

HO
HO

R
+

+ O
N

O
OH OH HO

HO
HO

C OH
O O–

IV

II
I
VII VI

V

III
H-bonding Ionic centre

R centre

Aromatic
centre

H-bonding


Adrenoceptor

Muscarinic receptor

Fig. 1.10 Receptor ligand-binding sites. The coloured
areas are schematic representations of the regions of
the adrenoceptor (top) and muscarinic receptor (bottom)
responsible for binding their respective catecholamine
and acetylcholine ligands. In the muscarinic receptor,
cross-sections of the seven transmembrane segments are
labelled I–VII. Different segments provide different properties
(hydrogen bonding, anionic site, etc.) to make up the active
binding site.

1 16 Medical Pharmacology and Therapeutics

because the dose then produces the maximal response of
which the biological tissue is capable.
Potency
The potency of a drug in vitro is largely determined by the
strength of its binding to the receptor, which is a reflection of
the receptor affinity, and by the inherent ability of the drug/
receptor complex to elicit downstream signalling events.
The more potent a drug, the lower the concentration needed
to give a specified response. In Fig. 1.12, drug A1 is more
potent than drug A2 because it produces a specified level of
response at a lower concentration. It is important to recognise
that potencies of different drugs are compared using the
doses required to produce (or block) the same response
(often chosen arbitrarily as 50% of the maximal response).
The straight-line portions of log dose–response curves are
usually parallel for drugs that share a common mechanism
of action, so the potency ratio is broadly the same at most
response values – for example, 20%, 50% or 80%, but
not at 100% response. A drug concentration sufficient to
produce half of the greatest response achievable by that

DOSE–RESPONSE RELATIONSHIPS
Using a purified preparation of a single drug, it is possible to
define accurately and reproducibly the relationship between
the doses of drug administered (or concentrations applied) and
the biological effects (responses) at each dose. The results
for an individual drug can be displayed on a dose–response
curve. In many biological systems, the typical relationship
between an increasing drug dose (or concentration in plasma)
and the biological response is a hyperbola, with the response
curve rising with a gradually diminishing slope to a plateau,
which represents the maximal biological response. Plotting
instead the logarithm of the dose (or concentration) against
the response (plotted on a linear scale) generates a sigmoid
(S-shaped) curve. The sigmoid shape of the curve provides
a number of advantages for understanding the relationship
between drug dose and response: a very wide range of doses
can be accommodated easily on the graph, the maximal
response plateau is clearly defined, and the central portion
of the curve (between about 15% and 85% of maximum)
approximates to a straight line, allowing the collection of
fewer data points to delineate the relationship accurately.
Fig. 1.11 shows the log dose–response relationship
between a drug and responses it produces at two types of
adrenoceptors. In each case, the upward slope of the curve
to the right reflects the chemical law that a greater number of
reversible molecular interactions of a drug (D) with its receptor
(R), due in this case to increasing drug dose, leads to more
intracellular signalling by active drug–receptor complexes (DR)
and hence a greater response of the cell or tissue (within
biological limits). This principle is diametrically opposed
to the principle of homeopathy, which argues that serially
diluting a drug solution until essentially no drug molecules
remain enhances its activity, a belief that is not supported
theoretically or experimentally.
Selectivity
As drugs may act preferentially on particular receptor types
or subtypes, such as β1- and β2-adrenoceptors, it is important
to be able to quantify the degree of selectivity of a drug.
For example, it is important in understanding the therapeutic
efficacy and unwanted effects of the bronchodilator drug
salbutamol to recognise that it is approximately 10 times
more effective in stimulating the β2-adrenoceptors in the
airway smooth muscle than the β1-adrenoceptors in the heart.
In pharmacological studies, selectivity is likely to
be investigated by measuring the effects of the drug in
vitro on different cells or tissues, each expressing only
one of the receptors of interest. Comparison of the two
log dose–response curves in Fig. 1.11 shows that for a
given response, smaller doses of the drug being tested
are required to stimulate the β1-adrenoceptor compared
with those required to stimulate the β2-adrenoceptor; the
drug is therefore said to have selectivity of action at the
β1-adrenoceptor. An example might be dobutamine, which is
used to selectively stimulate β1-adrenoceptors on the heart
in heart failure. The degree of receptor selectivity is given
by the ratio of the doses of the drug required to produce
a given level of response via each receptor type. It is clear
from Fig. 1.11 that the ratio is highly dose-dependent and
that the selectivity disappears at extremely high drug doses,
% Maximum response

Increasing dose of β-adrenoceptor agonist
(logarithmic scale)

100

50

0 D1 D2 D3
Drug action at
β1-adrenoceptor

Drug action at
β2-adrenoceptor

Fig. 1.11 Dose–response relationship and receptor
selectivity. Each curve shows the responses (expressed as
percentage of maximum on a linear vertical axis) produced
by a hypothetical β-adrenoceptor agonist drug at a range
of doses shown on a logarithmic horizontal axis. Plotting
the logarithmic dose allows a wide range of doses to be
shown on the same axes and transforms the dose–response
relationship from a hyperbolic curve to a sigmoid curve, in
which the central portion is close to a straight line. The two
curves illustrate the relative selectivity of the same drug for
the β1-adrenoceptor compared with the β2-adrenoceptor. At
most doses the drug produces β1-adrenoceptor stimulation
with less effect on β2-adrenoceptors. If dose D1 is 10 times
lower than dose D2, the selectivity of the drug for the β1-
adrenoceptor is 10-fold higher. This selectivity diminishes
at the higher end of the log dose–response curve and is
completely lost at a dose (D3) that produces a maximum
response on both β1- and β2-adrenoceptors.

Principles of pharmacology and mechanisms of drug action 17

obtained with a more efficacious drug, while a more potent
drug may merely allow a smaller dose to be given for the
same clinical benefit. In turn, efficacy and potency need
to be balanced against drug toxicity to produce the best
balance of benefit and risk for the patient. Drug toxicity and
safety are discussed in Chapters 3 and 53.
TYPES OF DRUG ACTION
Drugs can be classified by their receptor action as:
■ agonists,
■ antagonists,
■ partial agonists,
■ inverse agonists,
■ allosteric modulators,
■ enzyme inhibitors or activators,
■ nonspecific,
■ physiological antagonists.
AGONISTS
An agonist, whether a therapeutic drug or an endogenous
ligand, binds to the receptor or site of action and changes
the conformation of the receptor to its active state, leading
to signalling via second messenger pathways. An agonist
shows both affinity (the strength of binding for the receptor)
and intrinsic activity (the extent of conformational change
imparted to the receptor leading to receptor signalling). Drugs
differ in their affinity and intrinsic activity at the same receptor,
as well as between different receptors.
Agonists are traditionally divided into two main groups
(Fig. 1.12):
■ full agonists (curves A1 and A2), which give an increase
in response with an increase in concentration until the
maximum possible response is obtained for that system;
■ partial agonists (curve A3), which also give an increase
in response with increase in concentration, but cannot
produce the maximum possible response in the system.
The reasons for this difference, and also a third group of
agonists (inverse agonists), are described as follows.
Affinity and intrinsic activity
The affinity of a drug is related to the aggregate strength of
the atomic interactions between the drug molecule and its
receptor site of action, which determines the relative rates
of drug binding and dissociation. The higher the affinity, the
lower the drug concentration required to occupy a given
fraction of receptors. Affinity therefore determines the drug
concentration necessary to produce a certain response and
is directly related to the potency of the drug. In Fig. 1.12,
drug A1 is more potent than drug A2, but both are capable of
producing a maximal response (they have the same efficacy
as they are full agonists).
Intrinsic activity describes the ability of the bound drug
to induce the conformational changes in the receptor that
induce receptor signalling. Although affinity is a prerequisite
for binding to a receptor, a drug may bind with high affinity
but have low intrinsic activity. A drug with zero intrinsic
activity is an antagonist (as discussed later).

drug is described as its EC50 (the effective concentration
for 50% of the maximal response). The EC50 (or ED50 if drug
dose is considered) is a convenient way to compare the
potencies of similar drugs; the lower the EC50 (or ED50), the
more potent the drug.
In vivo the potency of a drug, defined as the dose of the
drug required to produce a desired clinical effect, depends not
only on its affinity for the receptor, the receptor number and
the efficiency of the stimulus–response mechanism, but also
on pharmacokinetic variables that determine the delivery of
the drug to its site of receptor action (Chapter 2). Therefore
the relative potencies of related drugs in vivo may not directly
reflect their in vitro receptor-binding properties.
Efficacy
The efficacy of a drug is its ability to produce the maximal
response possible for a particular biological system and
relates to the extent of functional change that can be imparted
to the receptor by the drug, based on its affinity for the
receptor and its ability to induce receptor signalling (discussed
later). Drug efficacy is arguably of greater clinical importance
than potency because a greater therapeutic benefit may be
A2 + RA

A2 + IA

Response (%)

Concentration of agonist (logarithmic scale)
100

50

0
A1 A2

A3

Fig. 1.12 Concentration–response curves for agonists
in the absence and presence of competitive and
noncompetitive antagonists. Responses are plotted at
different concentrations of two different full agonists (A1
being more potent than A2) and also a partial agonist (A3),
which is unable to produce a maximal response even at
high concentrations. Responses are also shown for the
full agonist A2 in the presence of a fixed concentration
of a competitive (reversible) antagonist (RA) or a fixed
concentration of a noncompetitive (irreversible) antagonist
(IA). The competitive antagonist reduces the potency of
agonist A2 (the curve is shifted parallel to the right), but high
concentrations of A2 are able to surmount the effects of the
competitive antagonist and produce a maximal response.
A noncompetitive antagonist reduces agonist activity either
by irreversibly blocking the agonist binding site, or by
changing its conformation by binding reversibly or irreversibly
at an allosteric site. Unlike competitive antagonists, a
noncompetitive antagonist reduces the maximal response
even at high agonist concentrations, as shown in the curve A2
+ IA compared with A2 alone.

1 18 Medical Pharmacology and Therapeutics

β1-adrenoceptor antagonists lower the heart rate markedly
only when it is already elevated by endogenous agonists such
as adrenaline and noradrenaline. The reversible binding of
competitive antagonists means that the receptor blockade can
be overcome (surmounted) by an increase in the concentration
of an agonist. Therefore competitive antagonist drugs move
the dose–response curve for an agonist in a parallel fashion
to the right but do not alter the maximum possible response
at high agonist concentrations (as shown in curve A2 + RA
when compared with A2 alone in Fig. 1.12).
Noncompetitive antagonists either bind to the receptor
irreversibly (covalently) at the ligand binding site, denying
access to the agonist, or they change the conformation
of the receptor by binding reversibly or irreversibly at
another (allosteric) site, producing conformational changes
that impede the ability of the agonist to access its binding
site or block the conformational changes in the receptor
needed for intracellular signalling. In either case, the
effects of noncompetitive antagonists cannot be negated
(surmounted) by competition from the agonist, so they
reduce the magnitude of the maximum response that can
be produced by any concentration of agonist (as shown by
curve A2 + IA in Fig. 1.12). A noncompetitive antagonist will
also cause a rightward shift of the agonist log dose–response
curve if there is no reserve of spare receptors.
Like agonists, antagonists exhibit varying degrees of
selectivity of action. For example, phenoxybenzamine
is an antagonist which blocks the ligand binding site of
α-adrenoceptors, but not that of β-adrenoceptors. Conversely,
propranolol is an antagonist of β-adrenoceptors, but not
α-adrenoceptors. Bisoprolol is further selective for the
β1-adrenoceptor subtype, and has less blocking action at
β2-adrenoceptors (or α-adrenoceptors).
PARTIAL AGONISTS
An agonist that is unable to produce a maximal response
is a partial agonist (e.g. drug A3 in Fig. 1.12). Even maximal
occupancy of all available receptors produces only a
submaximal response due to low intrinsic activity of the partial
agonist, for example because of incomplete amplification of
the receptor signal via the G-proteins. Despite their name,
partial agonists can be considered to have both agonist
and antagonist properties, depending on the presence and
type of other ligands. A partial agonist usually shows weak
agonist activity in the absence of another ligand, and such
partial agonism can be blocked by an antagonist. But in the
presence of a full agonist, a partial agonist will behave as a
weak antagonist because it prevents access to the receptor
of a molecule with higher intrinsic ability to initiate receptor
signalling; this results in a reduced response. Partial agonism
is responsible for the therapeutic efficacy of several drugs,
including buspirone, buprenorphine, pindolol and salbutamol.
These drugs can act as stabilisers of the variable activity of
the natural ligand, as they enhance receptor activity when
the endogenous ligand levels are low or zero, but block
receptor activity when endogenous ligand levels are high.
INVERSE AGONISTS
The previously provided definitions of agonists, partial
agonists and antagonists reflect the classical model of

It should be noted that the rate of binding and rate
of dissociation of a reversible drug at its receptor are of
negligible importance in determining its rate of onset or
duration of effect in vivo, because these depend mainly on
the rates of delivery of the drug to, and removal from, the
target organ; that is, on the overall absorption, distribution
and elimination rates of the drug from the body (Chapter 2).
Spare receptors
Some full agonists that have relatively low intrinsic activity
may have to occupy all of the available receptors to produce a
maximal response. However, many full agonists have sufficient
affinity and intrinsic activity that the maximal response can be
produced even though many receptors remain unoccupied;
that is, there may be spare receptors (or a receptor reserve).
The concept of spare receptors does not imply a distinct pool
of permanently redundant receptors, only that a proportion
of the receptor population is unoccupied at a particular
point in time. Spare receptors may function to enhance the
speed of cellular response, because an excess of available
receptors reduces the distance and therefore the time that
a ligand molecule needs to diffuse to find an unoccupied
receptor; an example is the excess of acetylcholine nicotinic
N2 receptors that contributes to fast synaptic transmission
in the neuromuscular junction (Chapter 27).
The concept of spare receptors is also helpful when
considering changes in receptor numbers during chronic
treatment, particularly receptor downregulation. As maximal
responses are often produced at drug concentrations that
do not attain 100% receptor occupancy, the same maximal
response may still be produced when receptor numbers
are downregulated, but only with higher percentage
occupancy of the reduced number of receptors. If receptors
are downregulated still further, the number remaining may
be insufficient to generate a maximal response. Receptor
downregulation may therefore contribute to a decline in
responsiveness to some drugs during chronic treatment
(drug tolerance).
ANTAGONISTS
Pharmacological antagonists (often called ‘blockers’) reduce
the activity of an agonist at the same receptor, and can
be contrasted with physiological antagonists (discussed
later) that act at another type of receptor or at other sites
of action to oppose the physiological response to the
agonist. Pharmacological antagonists can be competitive
(surmountable) or noncompetitive (nonsurmountable).
A competitive antagonist binds reversibly to the ligand
binding site of a receptor, either alone or in competition with
a drug agonist or natural ligand. It therefore must have affinity
for the ligand binding site (which may be as high as that
of any agonist), but it has zero intrinsic activity. It therefore
cannot cause the conformational change that converts the
receptor to its active state and induces intracellular signalling.
The antagonist will, however, competitively impair access
of agonist molecules to the ligand binding site and thereby
reduce receptor activation. The presence of a competitive
antagonist may only be detectable by its impairment of
agonist activity, and the extent of antagonism will depend on
the relative amounts of agonist and antagonist. For example,

Principles of pharmacology and mechanisms of drug action 19

an allosteric modulator may not bind to the allosteric site
(or only bind poorly) in the absence of the agonist, but its
allosteric binding increases when binding of the agonist to
the orthosteric site alters receptor conformation. An example
of allosteric modulators is the family of benzodiazepine
anxiolytic drugs, which allosterically alter the affinity of
chloride channels for the neurotransmitter ligand GABA
and enhance its inhibitory activity on neurons (Chapter 20).
ENZYME INHIBITORS AND ACTIVATORS
The site of action of many drugs is an enzyme, which may
be an intracellular or cell-surface enzyme or one found in
plasma or other body fluids. Such drugs act reversibly or
irreversibly either on the catalytic site or at an allosteric
site on the enzyme to modulate its catalytic activity; most
often the effect is inhibition, and important examples of
enzyme inhibitors are shown in Table 1.4. An example of
an enzyme activator is heparin, which enhances the activity
of antithrombin III, a protease that regulates the activity of
the coagulation pathway.
NONSPECIFIC ACTIONS
A few drugs produce their desired therapeutic outcome
without interaction with a specific site of action on a protein;
for example, the diuretic mannitol exerts an osmotic effect in
the lumen of the kidney tubule, which reduces reabsorption
of water into the blood (Chapter 14).
PHYSIOLOGICAL ANTAGONISTS
Physiological antagonism is said to occur when a drug
has a physiological effect opposing that of an agonist
but without binding to the same receptor. The increase
in heart rate produced by a β1-adrenoceptor agonist, an
effect which mimics the action of the sympathetic autonomic
nervous system, can be blocked pharmacologically with
an antagonist at β1-adrenoceptors or physiologically by a
muscarinic receptor agonist, which mimics the opposing
(parasympathetic) autonomic nervous system. The site of
action of the physiological antagonist may be on a different
cell, tissue or organ than that of the agonist.
TOLERANCE TO DRUG EFFECTS
Tolerance to drug effects is defined as a decrease in
response to repeated doses, often necessitating an increase
in dosage to maintain an adequate clinical response.
Tolerance may occur through pharmacokinetic changes in
the concentrations of a drug available at the receptor or
through pharmacodynamic changes at the drug receptor.
Pharmacokinetic effects are discussed in Chapter 2; some
drugs stimulate their own metabolism, so they are eliminated
more rapidly on repeated dosing, and lower concentrations
of the drug are available to produce a response.
Most clinically important examples of tolerance arise
from pharmacodynamic changes in receptor numbers and in
concentration–response relationships. Desensitisation is used
to describe both long-term and short-term changes arising

drug–receptor interactions, in which an unoccupied receptor
has no signalling activity. It is now recognised that many
GPCRs show constitutive signalling independently of an
agonist. Inverse agonists were first recognised when some
compounds were found to show negative intrinsic activity: they
acted alone on unoccupied receptors to produce a change
opposite to that caused by an agonist. Inverse agonists
shift the receptor equilibrium towards the inactive state,
thereby reducing the level of spontaneous receptor activity.
An inverse agonist can be distinguished from the typical
antagonists discussed previously, which, on their own, bind
to the receptor without affecting receptor signalling, as they
have zero intrinsic activity (‘neutral’ or ‘silent’ antagonists).
The action of a neutral antagonist depends on depriving
the access of agonists to the receptor; a neutral antagonist
can therefore block the effects of either a positive or inverse
agonist at a receptor with spontaneous signalling activity.
The role of inverse agonism in the therapeutic effects of
drugs remains to be fully elucidated, but a number of drugs
exhibit this type of activity (Table 1.5). The same drug may
even show a mixed pattern of full or partial agonism, inverse
agonism or antagonism at different receptors. Some drugs
(e.g. some β-adrenoceptor antagonists) can act as neutral
antagonists at a receptor in one tissue and as inverse agonists
when the same receptor is expressed in a different tissue,
probably due to association of the receptor with different
G-proteins.
ALLOSTERIC MODULATORS
Allosteric modulation has been described previously in the
context of one type of noncompetitive antagonist, which
does not compete directly with an agonist for access to the
ligand binding site (also called the orthosteric site), but binds
to a different (allosteric) site. Allosteric modulation changes
receptor activity by altering the conformation of the orthosteric
binding site or of sites involved in intracellular signalling.
Allosteric modulators can also enhance the binding of the
natural ligand or other drugs to the receptor or enhance their
propensity to induce receptor signalling. In some cases,
Table 1.5 Examples of drugs with inverse agonist
activity
Receptor Drugs
α1-Adrenoceptor Prazosin, terazosin
β1-Adrenoceptor Metoprolol, carvedilol,
propranolol
Angiotensin II receptor (AT1) Losartan, candesartan,
irbesartan
Cysteinyl-leukotriene (CysLT1) Montelukast
Dopamine (D2) Haloperidol, clozapine,
olanzapine
Histamine (H1) Cetirizine, loratadine
Histamine (H2) Cimetidine, ranitidine,
famotidine
Muscarinic (M1) Pirenzepine
Opioid (μ, MOR) Naloxone

1 20 Medical Pharmacology and Therapeutics

The phosphorylated receptor protein is endocytosed
and may undergo intracellular dephosphorylation prior
to re-entering the cytoplasmic membrane.
Downstream modulation of the signal may also occur
through feedback mechanisms or simply through depletion
of some essential cofactor. An example of the latter is the
depletion of the thiol (-SH or sulphydryl) groups necessary
for the generation of nitric oxide during chronic administration
of organic nitrates (Chapter 5).

GENETIC VARIATION IN
DRUG RESPONSES
Biological characteristics, including responses to drug
administration, vary among individuals, and genetic differences
can contribute to these inter-individual variations. For most
drugs, the nature of the response is broadly similar in different
individuals, but the magnitude of the response to the same
dose can differ markedly, at least partly due to genetic
factors. Such variability creates the need to individualise
drug dosages for different people.
Drug responses may follow a unimodal (Gaussian)
distribution, reflecting the sum of many small genetic variations
in receptors, enzymes or transporters that respond to or
handle the drug (Fig. 1.13A). Genetic variation may also give
rise to discrete subpopulations of individuals in which a drug
shows distinctly different responses (see Fig. 1.13B), such
that some individuals may have no response to a standard
dose, while others show toxicity. Understanding genetic
variation is of increasing importance in drug development
(see Chapter 3) because it allows the possibility of genetic

from a decrease in response of the receptor. Desensitisation
can occur by a number of mechanisms:
■ decreased receptor numbers (downregulation), due
to decreased transcriptional expression or receptor
internalisation;
■ decreased receptor binding affinity;
■ decreased G-protein coupling;
■ modulation of the downstream response to the initial
signal.
GPCRs can show rapid desensitisation (within minutes)
during continued activation, which occurs through three
mechanisms:
■ Homologous desensitisation. The enzymes activated

following selective binding of an agonist to its receptor–G-
protein complex include G protein-coupled receptor

kinases (GRKs), which interact with the βγ-subunit of the
G-protein and inactivate the occupied receptor protein by
phosphorylation; a related peptide, arrestin-2, enhances
the GRK-mediated desensitisation of the GPCR and may
itself activate distinct cell signalling pathways.

■ Heterologous desensitisation. Also known as cross-
desensitisation, this occurs when an agonist at one

receptor causes loss of sensitivity to other agonists.
The agonist increases intracellular cAMP which activates

protein kinase A or C; these phosphorylate the cross-
desensitised receptors (whether occupied or not), and

members of the arrestin family prevent them from coupling
with G-proteins. Other mechanisms of heterologous
desensitisation exist.
■ Receptor internalisation. Internalisation can occur
within minutes when constant activation of a GPCR
makes the receptor unavailable for further agonist
action by uncoupling the G-protein from the receptor.

Gaussian distribution of response Polymorphic distribution of response

Response Response

Number of subjects

Number of subjects

A B

Fig. 1.13 Inter-individual variation in response. The graphs show the numbers of individual subjects in a population
plotted against their varying levels of response to a single dose of a drug. (A) In the unimodal distribution most individuals
show a middling response and the overall shape is a normal (Gaussian) distribution. Part of this variability may result from
polymorphism in multiple genes encoding drug receptors and proteins involved in the drug’s absorption and elimination.
(B) The bimodal distribution shows discrete responder and nonresponder subgroups, possibly due to a single genetic
polymorphism in a drug receptor or drug-metabolising enzyme.

Principles of pharmacology and mechanisms of drug action 21

screening to optimise drug and dosage selection (personalised
or individualised medicine).
Pharmacogenetics has been defined as the study
of genetic variation that results in differing responses to
drugs. Such variation may arise from genetic factors that
alter the structure, expression or regulation of drug targets
(pharmacodynamic effects) or that change the metabolic fates
of drugs in the body, usually by altering proteins involved in
their absorption, distribution or elimination (pharmacokinetic
effects, discussed in Chapter 2). Pharmacogenetic research
has been undertaken for many decades, largely in relation
to variability in vivo, and has often used classic genetic
techniques such as studies of patterns of inheritance in twins.
Pharmacogenomics has been defined as the investigation
of variation in DNA and RNA characteristics related to
drug response, and the term refers mainly to genome-wide
approaches that define the presence of single-nucleotide
polymorphisms (SNPs) which affect the activity of the gene
product. Molecular biological techniques have predicted more
than 3 million SNPs in the human genome. SNPs can be:
■ in the upstream regulatory sequence of a coding gene,
which can result in increased or decreased expression
of the gene product (this product remains identical to
the normal or ‘wild-type’ gene product);
■ in the coding region of the gene resulting in a gene product
with an altered amino acid sequence (this may have higher
activity, although this is unlikely; similar activity; lower
activity or no activity at all, compared with the wild-type
protein);
■ inactive, because they are in noncoding or nonregulatory
regions of the genome, or, if in a coding region, because
the base change does not alter the amino acid encoded,
due to the redundancy of the genetic code.
There is still a major challenge in defining the functional

consequences of the large numbers of identified SNPs (func-
tional genomics), particularly in the context of combinations of

genetic variants (haplotypes). Such studies often require very
large numbers of subjects to allow comparison of function
in multiple, small haplotype subgroups.
Rapid advances in molecular biology have allowed analysis
of interindividual differences in the sequences of many genes
encoding drug receptors and proteins involved in drug
metabolism and transport. Polymorphism in the latter is likely
to have the greatest impact on dosage selection (Chapter 2),
while polymorphism in drug targets may be more important
in determining the optimal drug for a particular condition.
For example, genetic variation in angiotensin AT1 receptors,
β1-adrenoceptors and Ca2+

ion channels may determine the
relative effectiveness of angiotensin II receptor antagonists,
β-adrenoceptor antagonists (β-blockers) and calcium channel
blockers in the treatment of essential hypertension.
In practice, although genetic polymorphism has been
reported in many receptor types and these have been a
major focus of research in relation to the aetiology of disease,
relatively few studies to date have demonstrated a clear
influence on drug responses. Common polymorphisms have
been identified in the human β2-adrenoceptor gene ADRB2,
and certain variants have been associated with differences
in receptor downregulation and loss of therapeutic response
in people with asthma while using β2-adrenoceptor agonist
inhalers (Chapter 12). The clinical response in people with
asthma to treatment with leukotriene modulator drugs is

influenced by genetic polymorphism in enzymes of the
leukotriene (5-lipoxygenase) pathway. Variants in the
epidermal growth factor receptor (EGFR), an RTK, have been
reported to predict tumour response to the EGFR inhibitor
gefitinib in individuals with nonsmall-cell lung cancer. Such
examples may support genotyping to target drug treatments
to those individuals most likely to respond.
Conversely, pharmacogenetic information may be used
to avoid a particular treatment in people likely to experience
serious adverse reactions to a specific drug. Variation in
human leucocyte antigen (HLA) genes has been associated
with adverse skin and liver reactions to several drugs,
including abacavir, an antiretroviral drug used in HIV infection.
Compared with pharmacodynamic targets, genetic

variation has been more extensively characterised in drug-
metabolising enzymes, particularly in cytochrome P450

isoenzymes and others involved in glucuronidation, acetylation
and methylation. Gene variations in drug-metabolising
enzymes are discussed at the end of Chapter 2. Detailed
information on human genotypic variation can be found in
the Online Mendelian Inheritance in Man (OMIM) database
(www.ncbi.nlm.nih.gov/omim). Therapeutic exploitation of
genotypic differences will require specific information about
individuals based on detailed genetic testing. Until such
genetic information is routinely incorporated in clinical trials,
careful monitoring of clinical response will remain the best
guide to successful treatment.
SUMMARY
The therapeutic benefits of drugs arise from their ability to
interact selectively with target receptors, most of which are
regulatory molecules involved in the control of cellular and
systemic functions by endogenous ligands. Drugs may also
cause unwanted effects; judging the balance of benefit and
risk is at the heart of safe and effective prescribing. Increasing
knowledge of the complexity of receptor pharmacology and
improvements in drug selectivity offer the promise of safer
drugs in the future, especially when information on genetic
variation is more routinely available.
SELF-ASSESSMENT
True/false questions
1. Clinical pharmacology is the study of drugs that doctors
use to treat disease.
2. Drugs act at receptors only on the external surface of
cells.
3. Diluting drugs enhances their pharmacological effects.
4. Drugs produce permanent biochemical changes in their
receptors.
5. Plotting drug dose (or plasma concentration) against
response usually produces a sigmoid curve.
6. The EC50 is the concentration of drug that produces a
half-maximal response.
7. On a log dose–response plot, the drug with a curve to
the right is more potent than a drug with a curve on
the left.

1 22 Medical Pharmacology and Therapeutics

reversible; irreversible drugs may act by covalent
chemical bonding.
5. False. Plotting drug dose or plasma concentration
against response typically produces a hyperbola; a
sigmoid (S-shaped) curve is produced by plotting the
logarithm of dose or concentration against the response.
6. True. The EC50 (or ED50) is the concentration (or dose)
effective in producing 50% of the maximal response
and is a convenient way of comparing drug potencies.
7. False. A drug with its log dose–response curve to the
left is the more potent, as it produces a given level of
response at a lower dose.
8. False. A full (‘neutral’ or ‘silent’) antagonist must have
affinity to bind to its receptor, but it has zero intrinsic
ability to activate the receptor. Partial agonists can
also have an antagonist effect in the presence of a
full agonist. Receptors with inherent signalling activity,
even when unoccupied, can be antagonised by inverse
agonists.
9. True. A fixed dose of a competitive antagonist shifts the
log dose–response curve of the agonist to the right in a
parallel fashion; it can be surmounted by increasing the
dose of agonist, so that the same maximal response
can be achieved.
10. True. A partial agonist has low intrinsic ability to induce
conformational change in the receptor so it does not
elicit a maximal response even with full receptor
occupancy.
11. False. Many full agonists are able to elicit a maximal
response when less than 100% of receptors are
occupied; the unoccupied receptors are termed ‘spare
receptors’.
12. True. Tolerance may be caused by desensitisation,
internalisation or downregulation of receptors, requiring
higher drug doses to maintain the same response.
Tolerance also often results from enhanced drug
elimination that alters the concentrations of drugs
available to interact with the receptor.

8. A receptor antagonist is defined as a drug with zero
affinity for the receptor.
9. A competitive antagonist shifts the log dose–response
curve of an agonist to the right, without affecting the
maximal response.
10. A partial agonist is one that, even at its highest dose,
cannot achieve the same maximal response as a full
agonist at the same receptor.
11. A full agonist achieves a maximal response when all
its receptors are occupied.
12. Changes in receptor numbers can cause tolerance to
drug effects.
ANSWERS
True/false answers
1. True. Clinical pharmacology also deals with drugs
used in disease prevention and diagnosis, and in the
alleviation of pain and suffering.
2. False. While many types of receptors are found in cell
membranes, including ion channels, GPCRs and tyrosine
kinase receptors, other drug targets, including steroid
receptors and many enzymes (e.g. cyclo-oxygenase,
PDE), are intracellular and others are humoral, such as
thrombin in plasma.

3. False. The relationship between the dose or concentra-
tion of a drug and the response may be complex but

is typically dependent on the number of interactions
between the drug molecules and their molecular target,
a consequence of the Law of Mass Action, and so are
usually greater at higher drug concentrations, within
biological limits.
4. False. Molecular interactions between most drugs
and their receptors are typically transient, and the
conformational changes induced in the receptor are

FURTHER READING
Alexander, S.P.H., Davenport, A.P., Kelly, E., et al., 2015. IUPHAR/
BPS concise guide to pharmacology 2015–16. Br. J. Pharmacol.
172, 5729–6202.
Baker, E.H., Pryce Roberts, A., Wilde, K., et al., 2011. Development
of a core drug list towards improving prescribing education and
reducing errors in the UK. Br. J. Clin. Pharmacol. 71, 190–198.
Katritch, V., Cherezov, V., Stevens, R.C., 2012. Diversity and
modularity of G protein-coupled receptor structures. Trends
Pharmacol. Sci. 33, 17–27.
Kenakin, T., 2004. Principles: receptor theory in pharmacology.
Trends Pharmacol. Sci. 25, 186–192.
Keravis, T., Lugnier, C., 2012. Cyclic nucleotide phosphodiesterase
(PDE) isozymes as targets of the intracellular signalling network:
benefits of PDE inhibitors in various diseases and perspectives
for future therapeutic developments. Br. J. Pharmacol. 165,
1288–1305.

Khilnani, G., Khilnani, A.J., 2011. Inverse agonism and its
therapeutic significance. Indian J. Pharmacol. 43, 492–501.
Luttrell, L.M., 2014. More than just a hammer: ligand ‘bias’ and
pharmaceutical discovery. Mol. Endocrinol. 28, 281–294.
Maxwell, S., Walley, T., 2003. Teaching safe and effective
prescribing in UK medical schools: a core curriculum for
tomorrow’s doctors. Br. J. Clin. Pharmacol. 55, 496–503.
Rosenbaum, D.M., Rasmussen, S.G.F., Kobilka, B.K., 2009. The
structure and function of G-protein-coupled receptors. Nature
459, 356–363.
Schöneberg, T., Kleinau, G., Brüser, A., 2016. What are they
waiting for? – Tethered agonism in G protein-coupled receptors.
Pharmacol. Res. http://dx.doi.org/10.1016/j.phrs.2016.03.027.
Wang, L., McLeod, H.L., Weinshilboum, R.M., 2011. Genomics and
drug response. N. Engl. J. Med. 364, 1144–1153.

Principles of pharmacology and mechanisms of drug action 23

Examples of cell surface receptor families and their properties
This is a reference list of members of important families of GPCRs, LGICs and VGICs, many of which are therapeutic drug
targets. Examples of agonists and antagonists are also shown; these include endogenous ligands and some drugs not
currently in clinical use. For further information see the relevant sections of Alexander, S.P.H., et al., 2015. IUPHAR/BPS
concise guide to pharmacology 2015–16. Br. J. Pharmacol. 172, 5729–6202 (available at http://guidetopharmacology.org).
For examples of important intracellular receptors and enzymes targeted by therapeutic drugs, see Tables 1.2 and 1.4.
Type Typical
location(s)

Principal
transduction
mechanism

Major biological
actions

Agonists Antagonists

G-protein-coupled receptors (GPCRs)
Cholinergic
Muscarinic
M1 CNS, salivary,
gastric; minor role in
autonomic ganglia

Gq Neurotransmission
in CNS, gastric
secretion

Nonselective for
all M receptors:
carbachol

Pirenzepine
Nonselective for
all M receptors:
atropine,
ipratropium,
oxybutynin,
tolterodine

M2 Heart, CNS Gi Bradycardia,
smooth muscle
contraction (GI tract,
airways, bladder)

M3 Smooth muscles,
secretory glands,
CNS

Gq Contraction,
secretion

I Darifenacin,
tiotropium

M4 CNS Gi Unclear
M5 CNS Gq Unclear
Adrenergic
α-Adrenoceptors
α1 (α1A, α1B,
α1D)

CNS, postsynaptic
in sympathetic
nervous system,
human prostate (α1A)

Gq Contraction of
arterial smooth
muscle, decrease
in contractions of
gut, contraction of
prostate tissue

Phenylephrine,
methoxamine, NA
≥ Adr

Prazosin,
indoramin
(tamsulosin α1A)

α2 (α2A, α2B,
α2C)

Presynaptic (in both
α- and β-adrenergic
neurons)

Gi Decreased NA
release

Clonidine, Adr > NA
(oxymetazoline α2A)

Yohimbine

β-Adrenoceptors
β1 CNS, heart (nodes
and myocardium),
kidney

Gs Increased force
and rate of cardiac
contraction, renin
release

Dobutamine, NA
> Adr

Atenolol,
metoprololol

β2 Bronchial smooth
muscle, also
widespread

Gs Bronchodilation,
decrease in
contraction of gut,
glycogenolysis

Salbutamol,
salmeterol,
terbutaline, Adr >
NA

Butoxamine

β3 Adipocytes, bladder Gs Lipolysis, bladder
emptying

Adr = NA –

Cannabinoids
CB1 Cortex,
hippocampus,
amygdala, basal
ganglia, cerebellum

Gi/o Behaviour,
pain, nausea,
stimulation of
appetite, addiction,
depression,
hypotension

Tetrahydrocannabinol,
anandamide,
2-arachidonylglycerol

Rimonabant

1 24 Medical Pharmacology and Therapeutics

CB2 Leucocytes,
osteocytes

Gi/o Immunity, bone
growth

Tetrahydrocannabinol

Dopamine
D1 CNS (N, O, P,
S – see footnote for
key to CNS areas),
kidney, heart

Gs Vasodilation in
kidney

Fenoldepam Chlorpromazine

D2 CNS (C, N, O, SN),
pituitary gland,
chemoreceptor
trigger zone,
gastrointestinal tract

Gi Cognition
(schizophrenia),
prolactin secretion,
nigrostrial control of
movement, memory

Cabergoline,
pramipexole,
ropinirole, rotigotine

Butyrophenones,
chlorpromazine
domperidone,
metoclopramide,
sulpiride

D3 CNS (F, Me, Mi)
(limbic system)

Gi Cognition, emotion Cabergoline,
pramipexole,
ropinirole, rotigotine

Chlorpromazine,
sulpiride

D4 CNS, heart Gi Cognition
(schizophrenia)

Cabergoline,
ropinirole, rotigotine

Chlorpromazine,
clozapine

D5 CNS (Hi, Hy) Gs Similar to D1
5-Hydroxytryptamine (5-HT, serotonin)
5-HT1A CNS, blood vessels Gi Anxiety, appetite,
mood, sleep

Buspirone

5-HT1B CNS, blood vessels Gi Vasoconstriction
presynaptic
inhibition

Sumatriptan,
eletriptan

5-HT1D CNS, blood vessels Gi Anxiety,
vasoconstriction

Sumatriptan,
eletriptan

Metergoline

5-HT1E CNS, blood vessels Gi
5-HT1F CNS Gi Sumatriptan,
eletriptan

5-HT2A CNS, GI tract,
platelets, smooth
muscle

Gq Schizophrenia,
platelet aggregation,
vasodilation/
vasoconstriction

LSD, psilocybin Ketanserin,
atypical
antipsychotics,
e.g. olanzapine

5-HT2B CNS, GI tract,
platelets

Gq Contraction,
morphogenesis

5-HT2C CNS, GI tract,
platelets

Gq Satiety

5-HT4 CNS, myenteric
plexus, smooth
muscle

Gs Anxiety, memory,
gut motility

Metoclopramide,
renzapride

5-HT5a CNS Gi Anxiety, memory,
mood
5-HT6 CNS Gs Anxiety, memory,
mood

5-HT7 CNS, GI, blood
vessels

Gs Anxiety, memory,
mood

LSD

Histamine
H1 CNS, endothelium,
smooth muscle

Gq Sedation,
sleep, vascular
permeability,
inflammation

Cetirizine,
desloratadine

Examples of cell surface receptor families and their properties (cont’d)
Type Typical
location(s)

Principal
transduction
mechanism

Major biological
actions

Agonists Antagonists

Principles of pharmacology and mechanisms of drug action 25

H2 CNS, cardiac
muscle, stomach

Gs Gastric acid
secretion

Dimaprit Cimetidine,
ranitidine

H3 CNS (presynaptic),
myenteric plexus

Gi Appetite, cognition Thioperamide

H4 Eosinophils,
basophils, mast
cells

Gi 4-Methylhistamine

Gamma-aminobutyric acid receptor type B (GABAB)
GABAB Brain neurons,
glial cells, spinal
motor neurons and
interneurons

Gi Inhibition of
neurotransmission
in brain and spinal
cord

Baclofen

Peptides
Angiotensin II
AT1 Blood vessels,
adrenal cortex, brain

Gq/Go Vasoconstriction,
salt retention,
aldosterone
synthesis, increased
noradrenergic
activity, cardiac
hypertrophy

Candesartan,
losartan,
valsartan

AT2 Blood vessels,
endothelium,
adrenal cortex, brain

Gi/o, tyrosine
and ser/thr
phosphatases

Weak vasodilation
(endothelial nitric
oxide release),
foetal development,
vascular growth

Bradykinin
B1
(induced)

Widespread
(induced by injury,
cytokines)

Gq Acute inflammation;
stimulates nitric
oxide synthesis

ACE inhibitors
(indirect, by
blocking bradykinin
breakdown)

B2
(constitutive)

Gq Chronic
inflammation.
Most kinin actions
(vasodilation, pain)

Icatibant

Endothelin
ETA Endothelium Gq Vasoconstriction,
angiogenesis

Bosentan

ETB Endothelium Gq, Gi Indirect vasodilation
(nitric oxide
release), direct
vasoconstriction,
natriuresis

Bosentan

Opioids
DOP (δ),
KOP (κ),
MOP (μ),
nociception

Brain, spinal cord,
peripheral sensory
neurons

Gi Analgesia, sedation,
respiratory
depression

Endogenous
opioids, opioid
drugs (morphine)

Naloxone,
naltrexone

Protease-activated receptors
PAR1,
PAR2,
PAR3, PAR4

Platelets, endothelial
cells, epithelial cells,
myocytes, neurons

Gq, Gi Activated by
proteolytic cleavage

Trypsin, thrombin,
tryptase
Examples of cell surface receptor families and their properties (cont’d)
Type Typical
location(s)

Principal
transduction
mechanism

Major biological
actions

Agonists Antagonists

1 26 Medical Pharmacology and Therapeutics

Vasopressin and oxytocin
Vasopressin
V1a

Brain, uterus, blood
vessels, platelets

Gq Vasoconstriction,
platelet aggregation

Desmopressin Conivaptan,
demeclocycline

Vasopressin
V1b

Pituitary, brain Gq Modulates ACTH
secretion

Desmopressin Conivaptan,
demeclocycline

Vasopressin
V2

Kidney Gs Antidiuretic effect
on collecting duct
and ascending limb
of loop of Henle

Desmopressin Conivaptan
demeclocycline,
tolvaptan

Oxytocin
OXT

Brain, uterus Gq, Gi Lactation, uterine
contraction, CNS
actions (mood)

Oxytocin > arginine-
vasopressin

Atosiban

Purinergic receptors (purinoceptors)
Adenosine
A1

Heart, lung Gi Cardiac depression,
vasoconstriction,
bronchoconstriction

Methylxanthines

Adenosine
A2A

Widespread Gs Vasodilation,
inhibition of platelet
aggregation,
bronchodilation

Regadenoson Methylxanthines

Adenosine
A2B

Leucocytes Gs Bronchoconstriction Methylxanthines

Adenosine
A3

Leucocytes Gi Inflammatory
mediator release

Methylxanthines

Purinergic
P2Y family
(P2Y1,
P2Y2, P2Y4,
P2Y6,
P2Y11–
P2Y14)

Widespread Gq, Gs or Gi Depends upon
G-protein coupling

ATP, ADP, UTP,
UDP, UDP-glucose
P2Y12:
clopidogrel,
ticlopidine

Ligand-gated ion channels (LGICs)
Nicotinic N1 Autonomic ganglia LGIC Ganglionic
neurotransmission

Carbachol, nicotine Trimetaphan,
mecamylamine

Nicotinic N2 Neuromuscular
junction

LGIC Skeletal muscle
contraction

Nicotine,
suxamethonium
(depolarising)

Gallamine,
vecuronium,
atracurium

Serotonin
5-HT3

CNS (A), enteric
nerves, sensory
nerves

Ligand-gated
Na+
/K+
channel
Emesis Granisetron,
ondansetron,
metoclopramide

GABAA Brain neurons,
spinal motor
neurons and
interneurons

Ligand-gated
Cl−
channel
(open)

Inhibition of
neurotransmission
in brain and spinal
cord

Muscimol,
barbiturates,
benzodiazepines,
zolpidem

Picrotoxin,
flumazenil
(benzodiazepine
antagonist)

Glycine
GlyR

Brain neurons,
spinal motor
neurons and
interneurons

Ligand-gated
Cl−
channel
(open)

Inhibition of
neurotransmission
in brain and spinal
cord

Intravenous
anaesthetics,
alanine, taurine

Strychnine,
caffeine,
tropisetron,
endocannabinoids

Ionotropic
glutamate
(NMDA)
receptor

CNS (B, C, sensory
pathways)

Ligand-gated
Ca2+
channel
(slow)

Synaptic plasticity,
excitatory
transmitter release;
excessive amounts
may cause neuronal
damage

NMDA Ketamine,
phencyclidine,
memantine

Examples of cell surface receptor families and their properties (cont’d)
Type Typical
location(s)

Principal
transduction
mechanism

Major biological
actions

Agonists Antagonists

Principles of pharmacology and mechanisms of drug action 27

Ionotropic
glutamate
(kainate)
receptor

CNS (Hi) Ligand-gated
Ca2+
channel
(fast)

Synaptic plasticity,
transmitter release

Kainate Topiramate

Ionotropic
glutamate
(AMPA)
receptor

CNS (similar to
NMDA receptors)

Ligand-gated
Ca2+
channel
(fast)

Synaptic plasticity,
transmitter release

AMPA Topiramate

Purinergic
P2X family
(P2X1–P2X7)

CNS, autonomic
nervous system
(P2X2), smooth
muscle (P2X1),
leucocytes

LGICs (Na+
,
Ca2+
, K+
)
Neuronal
depolarisation, influx
of Na+
and Ca2+
,
efflux of K+

ATP Suramin

Voltage-gated ion channels (VGICs)
Epithelial
sodium
channel
(ENaC)

Renal tubule,
airways, colon
Na+
channel,
tonically open
Sodium
reabsorption in

aldosterone-
sensitive distal

tubule and
collecting duct

Expression
upregulated by
aldosterone

Amiloride,
triamterene

L-type
calcium
channels
(Cav1.1–1.4)

Widespread Voltage-gated
Ca2+
channels

(dihydropyridine-
sensitive)

Vascular and
cardiac smooth
muscle contraction,
prolong cardiac
action potential

Nifedipine,
amlodipine,
diltiazem,
verapamil

Ryanodine
(RyR1,
RyR2,
RyR3)

Skeletal muscle
(RyR1), heart
(RyR2), widespread
(RyR3)

Ca2+
channels Calcium-induced
Ca2+
release (CICR)

Cytosolic Ca2+
, ATP,
ryanodine, caffeine

Dantrolene

Abbreviations: ACE, Angiotensin-converting enzyme; ACTH, adrenocorticotropic hormone (corticotropin); Adr, adrenaline; AMPA, α-amino-
3-hydroxy-5-methyl-4-isoxazole propionic acid; CNS, central nervous system; GI, gastrointestinal; LSD, lysergic acid diethylamide; NA,

noradrenaline; NMDA, N-methyl D-aspartate.
Key to CNS areas: A, Area postrema; B, basal ganglia; C, caudate putamen; F, frontal cortex; Hi, hippocampus; Hy, hypothalamus; Me,
medulla; Mi, midbrain: N, nucleus accumbens; O, olfactory tubercle; P, putamen; S, striatum; SN, substantia nigra.
Examples of cell surface receptor families and their properties (cont’d)
Type Typical
location(s)

Principal
transduction
mechanism

Major biological
actions

Agonists Antagonists

1 28 Medical Pharmacology and Therapeutics

Appendix: student formulary
This student formulary used for educational purposes at Southampton University Faculty of Medicine is adapted from
the formulary originally described by Maxwell and Walley (Br. J. Clin. Pharmacol. 2003; 55:496–503). See also the lists of
commonly prescribed drugs in Baker et al. (Br. J. Clin. Pharmacol. 2011; 71:190–198) and the World Health Organisation
(WHO) Model List of Essential Medicines (www.who.int/medicines/publications/essentialmedicines/en/).
Therapeutic problem Core drugs
Gastrointestinal system
Emergency treatment of poisoning Adsorbant: activated charcoal

Paracetamol antidotes, e.g. acetylcysteine, methionine
Opioid antagonist, e.g. naloxone
Organophosphate antidote, e.g. pralidoxime

Dyspepsia, GORD and gastric
ulcer healing

Antacids, e.g. magnesium salts
Compound alginates, e.g. Gaviscon
Proton pump inhibitors, e.g. omeprazole, lansoprazole
H2 receptor antagonists, e.g. ranitidine, cimetidine
Helicobacter pylori antibiotics: clarithromycin, amoxicillin, metronidazole
Motility stimulants, e.g. metoclopramide
Others: misoprostol, sulcralfate

Inflammatory bowel disease
(ulcerative colitis, Crohn’s
disease)

Corticosteroids, e.g. prednisolone
Aminosalicylates, e.g. sulfasalazine, mesalazine
Cytokine inhibitors, e.g. infliximab

Antibiotic-associated colitis Antibiotics for Clostridium difficile, e.g. metronidazole, vancomycin
Diarrhoea Oral rehydration therapy

Opiate antimotility drugs, e.g. loperamide
Constipation Bulk-forming laxatives, e.g. ispaghula, methylcellulose
Stimulant laxatives, e.g. senna, docusate
Osmotic laxatives, e.g. magnesium hydroxide, lactulose

Antispasmodics Antimuscarinics, e.g. atropine, hyoscine

Others: mebeverine

Cardiovascular system
Hypertension β-Adrenoceptor antagonists, e.g. atenolol
α-Adrenoceptor antagonists, e.g. doxazosin
Centrally acting drugs, e.g. clonidine
Angiotensin-converting enzyme (ACE) inhibitors, e.g. captopril, ramipril, perindopril
Angiotensin receptor antagonists, e.g. candesartan, losartan
Thiazide diuretics, e.g. bendroflumethazide, indapamide
Loop diuretics, e.g. furosemide
Potassium-sparing diuretics, e.g. amiloride, spironolactone
Compound potassium-sparing diuretic: co-amilofruse
Calcium channel antagonists, e.g. nifedipine verapamil
Potassium channel openers, e.g. minoxidil, nicorandil
Heart failure Many of the above plus the following positive inotropic drugs:

Cardiac glycosides, e.g. digoxin
PDE inhibitors, e.g. milrinone
β-Adrenoceptor antagonist: bisoprolol

Acute coronary syndrome (angina,
myocardial infarction)

Many drugs listed under hypertension plus the following:
Inhibitors of platelet aggregation, e.g. low-dose aspirin, dipyridamole, clopidogrel,
abciximab
Thrombolytics, e.g. streptokinase, tenecteplase
Heparin (unfractionated)
Heparins (low molecular weight), e.g. enoxaparin
Oral anticoagulants, e.g. warfarin

Hyperlipidaemia Statins, e.g. simvastatin, atorvastatin, pravastatin, rosuvastatin

Fibrates, e.g. gemfibrozil, fenofibrate

Arrhythmias Antiarrhythmic drugs, e.g. digoxin, adenosine, amiodarone, lidocaine, β-adrenoceptor

antagonists, calcium channel blockers

Principles of pharmacology and mechanisms of drug action 29

Respiratory system
Asthma, COPD, respiratory failure Oxygen

β2-Adrenoceptor agonists, e.g. salbutamol, salmeterol
Antimuscarinics, e.g. ipratropium, tiotropium
Methylxanthines, e.g. theophylline, aminophylline
PDE type 4 inhibitor, e.g. roflumilast
Leukotriene antagonists, e.g. montelukast
Antiallergic drugs, e.g. cromoglicate, nedocromil
Magnesium sulphate
Inhaled corticosteroids, e.g. beclometasone, fluticasone
Oral corticosteroid, e.g. prednisolone
β2-Agonist/corticosteroid coformulations, e.g. Seretide
Allergy, anaphylaxis Antihistamines, e.g. cetirizine, loratadine

Adrenaline, e.g. Epipen
Cough suppression Dextromethorphan, codeine
Central nervous system
Insomnia, anxiety Benzodiazepines, e.g. temazepam, diazepam
Z-drugs, e.g. zopiclone, zolpidem
Others, e.g. buspirone, propranolol

Schizophrenia, mania Classical antipsychotics, e.g. chlorpromazine, haloperidol, flupentixol
Atypical antipsychotics, e.g. olanzapine, risperidone, quetiapine
Depot preparations, e.g. fluphenazine decanoate
Mood stabilisers, e.g. lithium

Depression Tricyclic antidepressants (TCAs), e.g. amitriptyline

Selective serotonin reuptake inhibitors (SSRIs), e.g. fluoxetine, citalopram, sertraline
Serotonin-noradrenaline reuptake inhibitors, e.g. venlafaxine
Monoamine oxidase-B inhibitors, e.g. moclobemide
Others, e.g. mirtazepine

Analgesia Nonsteroidal anti-inflammatory drugs (NSAIDs): see section on musculoskeletal

disease
Compound analgesics, e.g. co-codamol, co-dydramol
Moderately potent opioid analgesics, e.g. tramadol
Potent opioid analgesics, e.g. codeine, morphine, fentanyl
Nausea and vertigo Dopamine antagonists, e.g. metoclopramide
Serotonin receptor antagonists, e.g. ondansetron
Muscarinic receptor antagonists, e.g. hyoscine
Others: betahistine

Migraine Acute: 5-HT1 receptor agonists, e.g. sumatriptan
Prophylaxis: β-adrenoceptor antagonists, e.g. propranolol

Epilepsy Anticonvulsant drugs, e.g. diazepam, carbamazepine, clonazepam, ethosuximide,

gabapentin, lamotrigine, phenytoin, tiagabine, valproate

Parkinson’s disease Levodopa/DOPA decarboxylase coformulations, e.g. co-careldopa, co-beneldopa

Dopamine receptor agonists, e.g. ropinirole
COMT inhibitors, e.g. entacapone
MAO-B inhibitors: selegiline, rasagiline
Antimuscarinic drugs, e.g. procyclidine
Dementia (Alzheimer’s) Anticholinesterases, e.g. donepezil
NMDA receptor antagonists, e.g. memantine

Appendix: student formulary (cont’d)
Therapeutic problem Core drugs

1 30 Medical Pharmacology and Therapeutics

Infectious diseases
Community- and hospital-acquired
infections

Penicillins, e.g. benzylpenicillin
Penicillinase-resistant penicillins, e.g. flucloxacillin
Broad-spectrum penicillins, e.g. amoxicillin, co-amoxiclav
Cephalosporins, e.g. cefalexin, cefuroxime
Tetracyclines, e.g. oxytetracycline
Folate inhibitors, e.g. trimethoprim
Aminoglycosides, e.g. gentamicin
Vancomycin (for C. difficile)
Macrolides, e.g. erythromycin
Chloramphenicol
Quinolones, e.g. ciprofloxacin
Metronidazole (for anaerobes and protozoans)
Antituberculosis drugs, e.g. isoniazid, rifampicin, ethambutol
Antifungal drugs, e.g. amphotericin, fluconazole, miconazole, terbinafine
Viral DNA polymerase inhibitors, e.g. aciclovir (herpes), ganciclovir (CMV)
Neuraminidase inhibitors, e.g. zanamivir (influenza)
Nucleoside reverse transcriptase inhibitors, e.g. zidovudine, abacavir
Nonnucleotide reverse transcriptase inhibitors, e.g. efavirenz
HIV protease inhibitors, e.g. saquinavir
Antimalarial drugs, e.g. mefloquine, proguanil, malarone

Endocrine system
Diabetes mellitus, thyroid disease
and hypothalamo pituitary
hormones

Insulins (long- and short-acting)
Secretagogues
Sulfonylureas, e.g. gliclazide
Meglitinides, e.g. repaglinide
DPP4 inhibitors, e.g. saxagliptin
GLP agonists, e.g. exenatide
Sensitisers:
Biguanides, e.g. metformin
Thiazolidinediones, e.g. pioglitazone
Thyroid disease, e.g. levothyroxine, carbimazole
ADH mimetics, e.g. desmopressin
LHRH, e.g. gonadorelin
Human growth hormone, e.g. somatropin

Osteoporosis Calcium, vitamin D, calcitonin, parathyroid hormone, teriparatide
Bisphosphonates, e.g. alendronic acid, risedronate
Selective estrogen receptor modulators (SERMs), e.g. clomifene

Genitourinary system
Urinary retention, benign prostatic
hypertrophy and prostate cancer

α1-Adrenoceptor antagonists, e.g. doxazosin
5α-Reductase inhibitors, e.g. finasteride
Antiandrogens, e.g. flutamide

Urinary frequency/incontinence Antimuscarinic drugs, e.g. darifenacin, fesoterodine
Erectile dysfunction PDE5 inhibitors, e.g. sildenafil, tadalafil
Obstetrics and gynaecology
Steroidal contraception Combined hormonal contraceptives (oral, transdermal patch)
Progestogen-only contraceptives (oral, subdermal implant)
Progestogen-containing intrauterine device
Emergency contraception, e.g. levonorgestrel, ulipristal
Injectable contraception, e.g. medroxyprogesterone acetate

Dysmenorrhoea` Combined oral hormonal contraceptives
NSAIDs, e.g. mefenamic acid
Menorrhagia Antifibrinolytic agent, e.g. tranexamic acid
Progestogen-containing intrauterine device

Endometriosis Progestins

Gonadorelin analogues, e.g. goserelin
Danazol
Appendix: student formulary (cont’d)
Therapeutic problem Core drugs

Principles of pharmacology and mechanisms of drug action 31

Induction of labour Oxytocics, e.g. oxytocin

Prostaglandin analogues, e.g. gemeprost

Prevention of pre-term labour
(tocolysis)

Calcium channel blockers, e.g. nifedipine
β-Adrenoceptor agonists, e.g. terbutaline

Induction of abortion Oxytocics, mifepristone
Antiprogestogen, e.g. mifepristone
Prostaglandin analogues, e.g. gemeprost

Postpartum haemorrhage Oxytocics, ergometrine
Menopause Oestrogens (with progestins)
Others: tibolone, raloxifene

Malignant disease and immunosuppression
Cancer and immunosuppression Alkylating agents, e.g. cyclophosphamide
Cytotoxic antibiotics, e.g. doxorubicin
Antimetabolites, e.g. methotrexate, fluorouracil
Vinca alkaloids, e.g. vinblastine
Taxanes, e.g. paclitaxel
Tyrosine kinase inhibitors, e.g. gefitinib, imatinib
Topoisomerase I and II inhibitors, e.g. irinotecan, etoposide
Other cytotoxic drugs, e.g. crisantaspase, cisplatin
Antioestrogens, e.g. tamoxifen, anastrazole
Immunosuppressant drugs, e.g. azathioprine, corticosteroids, ciclosporin
Immunobiologicals, e.g. rituximab (anti-B-lymphocyte CD20), interferon alfa,
bevacizumab (anti-VEGF), trastuzumab (anti-HER2)

Musculoskeletal disease
Rheumatoid arthritis NSAIDs, e.g. indometacin, diclofenac
Corticosteroids, e.g. prednisolone
Immunosuppressants, e.g. methotrexate, azathioprine, leflunomide
Cytokine (TNFα) modulators, e.g. adalimumab, etanercept, golimumab, infliximab
Others: gold, penicillamine, sulfasalazine, hydroxychloroquine

Myasthenia gravis Anticholinesterases, e.g. pyridostigmine
Spasticity Skeletal muscle relaxants, e.g. baclofen, dantrolene
Gout Acute: colcichine; chronic: allopurinol
Ophthalmology
Glaucoma β-Adrenoceptor antagonists, e.g. timolol
Prostaglandin analogues, e.g. latanoprost
Sympathomimetics (α2-agonists), e.g. brimonidine
Carbonic anhydrase inhibitors, e.g. acetazolamide
Miotics, e.g. pilocarpine

Conjunctivitis Topical antibiotics, e.g. chloramphenicol
Tear deficiency Ocular lubricants, e.g. hypromellose
Others Mydriatics, e.g. phenylephrine

Mydriatics/cycloplegics, e.g. atropine, tropicamide
Topical formulations (eye drops) of many drugs, including anti-inflammatory
corticosteroids (e.g. betamethasone), antivirals and local anaesthetics (e.g. tetracaine)

Surgery, anaesthetics and intensive care
Surgery, anaesthetics and
intensive care

Many drugs used are listed in other sections, including opioid analgesics,
sympathomimetics and antiemetics, plus the following:
Intravenous (induction) anaesthetics, e.g. thiopental, propofol
Inhalation (maintenance) anaesthetics, e.g. isoflurane
Muscle relaxants, e.g. suxamethonium, atracurium
Antimuscarinics, e.g. atropine, glycopyrronium
Anticholinesterases, e.g. neostigmine
Local anaesthetics, e.g. lidocaine, bupivacaine

Abbreviations: ADH, Antidiuretic hormone; CMV, cytomegalovirus; COMT, catechol-O-methyltransferase; COPD, chronic obstructive pulmonary
disease; DPP4, dipeptidyl peptidase 4; GLP, glucagon-like peptide; GORD, gastro-oesophageal reflux disease; HER2, human epidermal
growth factor receptor 2; LHRH, luteinising hormone-releasing hormone; MAO-B, monoamine oxidase B; NMDA, N-methyl D-aspartate; NSAID,
nonsteroidal anti-inflammatory drug; PDE, phosphodiesterase; TNFα, tumour necrosis factor alpha; VEGF, vascular endothelial growth factor.
Appendix: student formulary (cont’d)
Therapeutic problem Core drugs

This page intentionally left blank

2 Pharmacokinetics

The biological basis of pharmacokinetics 33
General considerations 34
Absorption 38
Absorption from the gut 38
Absorption from other routes 39
Distribution 39
Reversible protein binding 40
Irreversible protein binding 41
Distribution to specific organs 41
Elimination 42
Metabolism 42
Excretion 45
The mathematical basis
of pharmacokinetics 48
General considerations 48
Absorption 49
Rate of absorption 49
Extent of absorption 50
Distribution 51
Rate of distribution 51
Extent of distribution 52
Elimination 53
Rate of elimination 53
Extent of elimination 54
Chronic administration 54
Time to reach steady state 55
Plasma concentration at steady state 55
Oral administration 55
Loading dose 56
Pharmacokinetics of biological drugs 56
Genetic variation and pharmacokinetics 56
Self-assessment 57
Answers 60
Further reading 62

Pharmacokinetics refers to the movement of drugs into,
through and out of the body. The type of response of an
individual to a particular drug depends on the inherent
pharmacological properties of the drug at its site of action.
However, the speed of onset, the intensity and the duration
of the response usually depend on parameters such as:
■ the rate and extent of uptake of the drug from its site of
administration;
■ the rate and extent of distribution of the drug to different
tissues, including the site of action;
■ the rate of elimination of the drug from the body.
Overall, the response of an individual to a drug depends
upon a combination of the effects of the drug at its site
of action in the body called pharmacodynamics (or ‘what the
drug does to the body’) and the way the body influences drug
delivery to its site of action (pharmacokinetics, or ‘what the
body does to the drug’; Fig. 2.1). Both pharmacodynamic
and pharmacokinetic aspects are subject to a number of
variables, which affect the dose–response relationship.
Pharmacodynamic aspects are determined by processes
such as drug–receptor interaction and are specific to the class
of the drug (see Chapter 1). These aspects are determined
by general processes, such as transfer across membranes,
metabolism and renal elimination, which apply irrespective
of the pharmacodynamic properties.
Pharmacokinetics may be divided into four basic processes,
sometimes referred to collectively as ‘ADME’:
■ Absorption. The transfer of the drug from its site of
administration to the general circulation
■ Distribution. The transfer of the drug from the general
circulation into the different organs of the body
■ Metabolism. The extent to which the drug molecule is
chemically modified in the body
■ Excretion. The removal of the parent drug and any
metabolites from the body; metabolism and excretion
together account for drug elimination.
This chapter will first describe each of these processes
qualitatively in biological terms, and then in mathematical
terms, which determine many of the quantitative aspects
of drug prescribing.
THE BIOLOGICAL BASIS OF
PHARMACOKINETICS
Most drug structures bear little resemblance to normal dietary
constituents such as carbohydrates, fats and proteins, and
they are handled in the body by different processes. Drugs
that bind to the same receptor as an endogenous ligand rarely
resemble the natural ligand sufficiently closely in chemical
structure to share the same carrier processes or metabolising

1 34 Medical Pharmacology and Therapeutics

enzymes. Consequently, the movement of drugs in the tissues
is mostly by simple passive diffusion rather than by specific
transporters, whereas metabolism is usually by enzymes of
low substrate specificity that can handle a wide variety of
drug substrates and other xenobiotics (foreign substances).
GENERAL CONSIDERATIONS
Passage across membranes
With the exception of intravenous or intra-arterial injections,
a drug must cross at least one membrane in its movement
from the site of administration into the general circulation.
Drugs acting at intracellular sites must also cross the cell
membrane to exert an effect. The main mechanisms by
which drugs can cross membranes (Fig. 2.2) are:
■ passive diffusion through the lipid layer,
■ diffusion through pores or ion channels,
■ carrier-mediated processes,
■ pinocytosis.
Passive diffusion
All drugs can move passively down a concentration gradient.
To cross the phospholipid bilayer directly (see Fig. 2.2),
a drug must have a degree of lipid solubility, such as
ethanol or steroids. Eventually a state of equilibrium will
be reached in which equal concentrations of the diffusible
form of the drug are present in solution on each side of the
membrane. The rate of diffusion is directly proportional to
the concentration gradient across the membrane, and to
the area and permeability of the membrane, but inversely
proportional to its thickness (Fick’s law). In the laboratory,
transient water-filled pores can be created in the phospholipid

Pharmacology

Pharmacodynamics

Dose–response relationship
Pharmacokinetics

Specific to drug or
drug class
Interaction with cellular
component, e.g. receptor
or target site
Effects at the site of action
Concentration–effect
relationship
Reduction in symptoms
Modification of disease
progression
Unwanted effects
Drug interactions
Inter- and intraindividual
differences

Absorption from the site
of administration
Delivery to the site of
action
Elimination from the body
Time to onset of effect
Duration of effect
Accumulation on repeat
dosage
Drug interactions
Inter- and intraindividual
differences
Nonspecific, general
processes

Fig. 2.1 Factors determining the response of an
individual to a drug.
Extracellular fluid D

D

D

D

D

D

D

D

Closed
ion channel
Passive diffusion
through lipid

bilayer (lipid-
soluble drugs)

Diffusion
through open
ion channel
(small molecules,
water-soluble drugs)

Facilitated
diffusion
(SLC transporters)
(nutrients, e.g.,
glucose, and some
amine
neurotransmitters)

Active
transport
(ABC transporters)
(numerous drugs,
see Table 2.1)
ATP
ADP
NBD
TMD TMD TMD

NBD

Fig. 2.2 The passage of drugs across membranes. Molecules can cross the membrane by simple passive diffusion
through the lipid bilayer or via a channel, or by facilitated diffusion, or by ATP-dependent active transport. ABC, ATP-binding
cassette superfamily of transport proteins; D, drug; NBD, nucleotide-binding domain; SLC, solute carrier superfamily of
transporters; TMD, transmembrane domain (see Table 2.1).

Pharmacokinetics 35

5-HT), noradrenaline, histamine and agmatine; although some
drugs are substrates for the transporters (see Table 2.1),
many basic drugs act as inhibitors of the transporters.
Members of the ABC and SLC transporter families are
therefore important in many of the processes by which
drugs are absorbed in the gut, distributed into tissues and
eliminated in the liver or kidney. Interactions between drugs
at the same transporter, or between drugs and the natural
substrates of the transporter, may contribute to variation in
kinetic parameters for individual drugs between patients and
over time. Genetic variation in the expression and functioning
of transporters may also contribute to drug toxicity.
Pinocytosis
This can be regarded as a form of carrier-mediated entry
into the cell cytoplasm. Pinocytosis is normally concerned
with the uptake of endogenous macromolecules and may be
involved in the uptake of recombinant therapeutic proteins;
drugs can also be incorporated into a lipid vesicle or liposome
for pinocytotic uptake (e.g. amphotericin and doxorubicin;
see Chapter 51).
Drug ionisation and membrane
diffusion
Ionisation is a fundamental property of most drugs that are
either weak acids, such as aspirin, or weak bases, such as
propranolol. The presence of an ionisable group(s) is essential
for the mechanism of action of most drugs, because ionic
forces represent a key part of ligand–receptor interactions
(see Chapter 1). The extent of ionisation may also influence
the extent of absorption of a drug, its distribution into organs
such as the brain or adipose tissue, and the mechanism and
route of its elimination from the body.
Drugs with ionisable groups exist in equilibrium between
charged (ionised) and uncharged (un-ionised) forms (Fig. 2.3).
The extent of ionisation of a drug depends on the strength
of the ionisable group and the pH of the solution. The extent
of ionisation is given by the acid dissociation constant, Ka.

Ka
conjugate base H
conjugate acid =
+ [ ][ ]
[ ] (2.1)
The term conjugate acid refers to a form of the drug able
to release a proton, such as:
■ an un-ionised acidic drug (Drug–COOH), or
■ an ionised basic drug (Drug–NH3
+
).

The conjugate base is the corresponding equilibrium form
of the drug that has lost a proton, such as:
■ an ionised acidic drug (Drug–COO−
), or
■ an un-ionised basic drug (Drug–NH2).
The value of Ka is normally much less than 1, so it is
easier to compare compounds using the negative logarithm
of Ka, which is called pKa. For example, a Ka of 10−5
becomes
pKa 5, and a Ka of 10−10 becomes pKa 10. Based on the
equation given previously, a strong acid (such as an −SO3H
functional group) that readily donates its H+

ion will have a

relatively high Ka value (e.g. 10−1
or 10−2
) and hence a low
pKa (i.e. 1 or 2), whereas weakly acidic groups, which donate
their H+
ion less readily, have a pKa of 4–5. Conversely, for
basic functional groups, the stronger the base, the greater

bilayer by applying a strong external electric field, and this
process (electroporation) is used to introduce large or charged
molecules, such as DNA, drugs and probes into live cells
in suspension.
Passage through membrane pores or
ion channels
Movement through channels occurs down a concentration
gradient and is restricted to extremely small water-soluble
molecules (<100 Da), such as ions. This is applicable to
therapeutic ions such as lithium and also to radioactive iodine.
Water itself crosses membranes rapidly via a ubiquitous
family of aquaporins.
Carrier-mediated processes
Two carrier-mediated processes are of widespread importance
in the transport of drugs, particularly those with low lipid
solubility, across membranes.
Active transport utilises energy (adenosine triphosphate
[ATP]) and transports drugs into or out of cells against their
concentration gradient. It is performed particularly by a family
of nonspecific carriers termed the ATP-binding cassette (ABC)
superfamily of membrane transporters (see Fig. 2.2 and Table
2.1). In humans, the ABC active-transporter superfamily
contains 49 members organised into seven subfamilies
(A–G) based on their relative sequence homology. Interest
in this area has expanded rapidly since the discovery of
P-glycoprotein (P-gp), also known as multidrug resistance
1 (MDR1) or ABCB1 transporter. P-gp transports a wide
range of drug substrates, including anticancer drugs, steroids
and immunosuppressive agents, from the cytoplasm to the
extracellular side of the cell membrane, and therefore acts as
an efflux transporter. Verapamil increases the concentrations
of anticancer drugs at their intracellular sites of action by
inhibiting P-gp (see Chapter 52). ABCB transporter proteins
contain two hydrophobic transmembrane domains, which
consist of different numbers of membrane-spanning α-helices
(12 in P-gp), and two hydrophilic nucleotide (ATP)-binding
domains, which bind and hydrolyse intracellular ATP. The
transporter is on the apical surface and acts as an efflux pump
that transports substrates from the cell into the interstitial
fluid, plasma, bile, urine or gut lumen. Examples of other
ABC transporters are given in Table 2.1.
Facilitated transport of a molecule by a carrier either aids
its passive movement down its own concentration gradient,
or uses the electrochemical gradient of a co-transported
solute to transport the molecule against its own concentration
gradient; in neither case is the use of ATP required. The
major examples are members of the solute carrier (SLC)
superfamily of transporters (see Fig. 2.2 and Table 2.1).
The SLC superfamily comprises over 300 types of organic
anion transporters (OATs), organic anion-transporting
polypeptides (OATPs), organic cation transporters (OCTs),
organic cation/carnitine transporters (OCTNs), and members
of other transporter families (see Table 2.1). OAT1 to OAT4
are present in various tissues; OAT1 is the classic organic
anion transporter in the kidney, which secretes urate and
penicillins and is blocked by probenecid (see Chapter
31). Organic cation transporters (OCT1, OCT2 and OCT3)
effect facilitated diffusion and can transport cations in both
directions across the membrane. Substrates common to all
three OCT transporters are serotonin (5-hydroxytryptamine,

1 36 Medical Pharmacology and Therapeutics

and most ionised in acid solutions (low pH). In either case,
the ionised form of the molecule can generally be regarded
as the water-soluble form and the un-ionised form as the
lipid-soluble form. The ease with which an ionisable drug
can diffuse across a lipid bilayer is determined by the lipid
solubility of its un-ionised form (Fig. 2.4).
The pH of body fluids is controlled by the buffering
capacity of the ionic groups present in endogenous molecules

its ability to retain the H+

, resulting in low Ka and high pKa
values. Thus strongly basic groups have a pKa of 10–11,
while weakly basic groups have a pKa of 7–8.
Drugs are 50% ionised when the pH of the solution equals
the pKa of the drug. Acidic drugs (low pKa values) are least
ionised in acidic solutions (low pH) and most ionised in
alkaline solutions (high pH). Conversely, basic drugs (high
pKa values) are least ionised in alkaline solutions (high pH),
Table 2.1 Examples of carrier molecules involved in drug transport
Transporter Typical substrates Sites in the body
ABC superfamily ATP-binding cassette superfamily of transport proteins. All use ATP hydrolysis and function
as active efflux or uptake transporters. Although there are a number of transporters in each
family, the four ABC transporters listed here can explain multidrug resistance in most cells
analysed to date.

MDR1 or
P-glycoprotein
(ABCB1)

Hydrophobic and cationic (basic)
molecules; numerous drugs,
including anticancer drugs

Apical surface of membranes of epithelial cells of
intestine, liver, kidney, blood–brain barrier, testis,
placenta and lungs

MRP1 (ABCC1) Numerous, including anticancer
drugs, glucuronide and glutathione
conjugates

Basolateral surface of membranes of most cell types
with high levels in lung, testis and kidney and in
blood–tissue barriers

MRP2 (ABCC2) Numerous, including anticancer
drugs, glucuronide and glutathione
conjugates

Apical surface of membranes; mainly in liver, intestine
and kidney tubules

BCRP (ABCG2)
Breast cancer
resistance protein

Anticancer, antiviral drugs,
fluoroquinolones, flavonoids

Apical surface of breast ducts and lobules, small
intestine, colon epithelium, liver, placenta, brain barrier
and lung

SLC superfamily Solute carrier superfamily of transporters. Comprises organic anion transporters (OATs),
organic anion-transporting polypeptides (OATPs), organic cation transporters (OCTs), organic
cation/carnitine transporters (OCTNs) and many other families. Solute carriers do not use ATP
hydrolysis and most function as uptake transporters

OAT1 (SLC22A6) Anionic drugs, acyclovir, adefovir,
NSAIDs, penicillins, diuretics and
phase 2 drug metabolites

Kidney (basolateral), brain, placenta, smooth muscle

OAT2 (SLC22A7) Anionic drugs, acyclovir, salicylate,
acetylsalicylate, PGE2, dicarboxylates

Kidney (basolateral), liver

OAT3 (SLC22A8) Similar to OAT1 Kidney (basolateral), liver, brain, smooth muscle, testis
OAT4 (SLC22A11) Methotrexate, pravastatin, sulfated

sex steroids

Kidney (apical), placenta

OATP1B1 (SLCO1B1) Pravastatin, rosuvastatin Liver
OATP1B3 (SLCO1B3) Methotrexate, rosuvastatin Liver
OCT1 (SLC22A1) Cationic drugs, serotonin,
noradrenaline, histamine, agmatine,
aciclovir, ganciclovir, metformin

Mainly in the liver, but also in kidney, small intestine,
heart, skeletal muscle and placenta

OCT2 (SLC22A2) Cationic drugs, serotonin,
noradrenaline, histamine, agmatine,
amantadine, metformin, cimetidine

Mainly in the kidney, but also in placenta, adrenal gland,
neurons and choroid plexus

OCT3 (SLC22A3) Cationic drugs, serotonin,
noradrenaline, histamine, agmatine,
metformin

Liver, kidney, intestine, skeletal and smooth muscle,
heart, lung, spleen, neurons, placenta and the choroid
plexus
OCTN1 (SLC22A4) Carnitine, acetylcholine Kidney, intestine
OCTN2 (SLC22A5) Carnitine, choline Kidney, skeletal muscle
ABC, ATP-binding cassette; NSAIDs, nonsteroidal anti-inflammatory drugs; PGE2, prostaglandin E2; SLC, solute carrier.
For details of other ABC and SLC transporters, see Nigam, S.K., 2015. What do drug transporters really do? Nat. Rev. Drug. Discov. 14, 29–44.

Pharmacokinetics 37

it in the plasma. The opposite situation prevails with basic
drugs, which are enabled to diffuse from the plasma into
urine, where they become trapped.
After drug overdose, when the aim is to enhance drug
elimination, alkalinisation of the urine (using intravenous
sodium bicarbonate) can be used to reduce reabsorption of
acidic drugs, such as aspirin, leading to their faster elimination
in the urine. Acidification of the urine (with oral ammonium
chloride) can enhance ionisation and renal elimination of
basic drugs, such as dexamfetamine.
The pH difference between gastric contents (pH 1–2)
and plasma (pH 7.4) affects the absorption of many oral
drugs. The acidity of stomach contents means that an acidic
drug is present largely in its un-ionised (protonated) form,
allowing it to pass into plasma where its ionised form becomes

such as phosphate ions and proteins. When the fluids on
each side of a membrane have the same pH value, there will
be equal concentrations of both the diffusible (un-ionised)
form and the nondiffusible (ionised) form of the drug on each
side of the membrane at equilibrium (see Fig. 2.4).
When the fluids on each side of a membrane are at
different pH values, the concentrations of the un-ionised form
on each side of the membrane at equilibrium will remain equal,
as it can diffuse reversibly across the membrane, but the
concentrations of the ionised form will be determined by the
pH of the solution. This results in pH-dependent differences
in total drug concentration on each side of a membrane (pH
trapping or partitioning), with the total drug concentration
being higher on the side of the membrane on which it is most
ionised. This is exemplified by the pH difference between
urine (pH 5–7) and plasma (pH 7.4), which can influence
renal elimination of drugs (Fig. 2.5). The relatively low pH
of the urine forces an acidic drug to become predominantly
un-ionised, allowing its reabsorption into the plasma, while
the higher pH in plasma (7.4) converts the drug to the ionised
form, preventing it diffusing back and trapping (partitioning)
Extracellular fluid Intracellular fluid
Administration

Ionisation

Redistribution
to other tissues
Protein
binding

Metabolism Excretion

Protein
binding

Dissolution
in fat
D D D Ionisation

Fig. 2.3 The effect of pH on drug ionisation. Acidic conditions (low pH, high H+

concentrations) push the equilibrium of
acidic drugs towards their un-ionised (protonated) form, and basic drugs towards their ionised form. Basic conditions (high pH)
have the opposite effect.

Acid–

Base-H+ Base
Acid-H

Low pH
High pH

(e.g. – COO–) (e.g. – COOH)

(e.g. – NH3
+) (e.g. – NH2)

Low pH
High pH

Ionised

water-
soluble form

Un-ionised

lipid-
soluble form

Fig. 2.4 Passive diffusion and the factors that affect
drug concentrations in equilibrium between un-ionised
and ionised forms. In this case, the pH is assumed to be
the same on each side of the membrane; see Fig. 2.5 for
drug partitioning when there is a pH gradient across the
membrane.

For an acidic
drug (DH)
Urine (pH 6) Membrane Plasma (pH 7.4)

Overall

Overall
For a basic drug (D)

+ DH D D D + DH
– D DH DH DH D

Fig. 2.5 Partitioning of acidic and basic drugs across
a pH gradient. Only the un-ionised forms (DH and D) are
able to diffuse across the membrane. In urine (pH 6), the
un-ionised acidic drug (DH) can be readily reabsorbed
into the plasma, where its ionised form (D−
) becomes
concentrated, while the ionised basic drug (D+
) is trapped
within the urine. Alkalinising the urine would reduce
reabsorption of the acid drug and enhance that of the basic
drug.

1 38 Medical Pharmacology and Therapeutics

of drug from the formulation determines the rate of absorption.
In modified-release (i.e. slow-release) formulations the drug
is either incorporated into a complex matrix from which
it diffuses slowly or in a crystallised form that dissolves
slowly. Dissolution of a tablet in the stomach can also be
prevented by coating it in an acid-insoluble layer, producing
enteric-coated formulations. This is useful for drugs such
as omeprazole (see Chapter 31), which is unstable in an
acid environment, and allows delivery of the intact drug to
the duodenum.
Gastric emptying
The rate of gastric emptying determines how soon a drug
taken orally is delivered to the small intestine, the major site
of absorption. Delay between oral drug ingestion and the
drug being detected in the circulation is usually caused by
delayed gastric emptying. Drugs that slow gastric emptying
(e.g. antimuscarinics) can delay the absorption of other drugs
taken at the same time.
Food has complex effects on drug absorption; it slows
gastric emptying and delays drug absorption, and it can also
bind drugs and reduce the total amount of drug absorbed.
First-pass metabolism
Metabolism of drugs can occur before and during their
absorption, and this can limit the amount of parent compound
that reaches the general circulation. Drugs taken orally have
to pass four major metabolic barriers before they reach the
general circulation. If there is extensive metabolism of a
drug at one or more of the sites listed in the sections that
follow, only a fraction of the original oral dose reaches the
general circulation as the parent compound. This process
is known as first-pass metabolism because it occurs at the
first passage through the organ.
Intestinal lumen
The intestinal lumen contains digestive enzymes secreted
by the mucosal cells and pancreas that are able to split
peptide, ester and glycosidic bonds. Intestinal proteases
prevent the oral administration of peptide drugs, such as
insulin and other products of molecular biological approaches
to drug development. In addition, the lower bowel contains
large numbers of aerobic and anaerobic bacteria that are
capable of performing a range of metabolic reactions on
drugs, especially hydrolysis and reduction.
Intestinal wall
The walls of the upper intestine are rich in cellular enzymes
such as monoamine oxidase (MAO), aromatic L-amino acid
decarboxylase, cytochrome P450 isoenzymes (e.g. CYP3A4)
and the enzymes responsible for phase 2 conjugation
reactions described later. In addition, the luminal membrane
of the intestinal cells (enterocytes) contains efflux transporters
such as P-gp (noted previously), which may limit the
absorption of a drug by transporting it back into the intestinal
lumen. Drug molecules that enter the enterocyte may thus
undergo three possible fates – that is, diffusion into the
hepatic portal circulation, metabolism within the cell or
transport back into the gut lumen (by P-gp). The substrate

partitioned. In contrast, basic drugs are highly ionised in
the stomach, and absorption is negligible until the stomach
empties and the drug can be absorbed from the more alkaline
lumen of the duodenum (pH ~8).
Drugs that are fixed in their ionised form at all pH values,
such as the quaternary amine compound suxamethonium
(see Chapter 27), cross cell membranes extremely slowly
or not at all; they are given by injection (because of lack of
absorption from the gastrointestinal tract) and have limited
effects on the brain (because of lack of entry).
ABSORPTION
Absorption is the process of transfer of the drug from the
site of administration into the general or systemic circulation.
ABSORPTION FROM THE GUT
The easiest and most convenient route of administration of
medicines is orally by tablets, capsules or syrups. The large
surface area of the small intestine combined with its high
blood flow can give rapid and complete absorption of oral
drugs. However, this route presents a number of obstacles
for a drug before it reaches the systemic circulation.
Drug structure
Drug structure is a major determinant of absorption. Drugs
need to be lipid-soluble to be absorbed from the gut. Highly
polar acids and bases tend to be absorbed only slowly
and incompletely, with much of the unabsorbed dose being
voided in the faeces. High polarity may, however, be useful
for delivery of the drug to a site of action in the lower bowel
(see Chapter 34). The structures of some drugs can make
them unstable either at the low pH of the stomach (e.g.
benzylpenicillin) or in the presence of digestive enzymes
(e.g. insulin). Such compounds have to be given by injection,
but administration by other routes may be possible (e.g.
inhalation for insulin).
Drugs that are weak acids or bases undergo pH
partitioning between the gut lumen and mucosal cells.
Acidic drugs will be least ionised in the stomach lumen,
and most absorption would be expected at this site, but
absorption in the stomach is limited by its relatively low
surface area (compared with the small intestine) and the
presence of a zone of neutral pH on the immediate surface of
the gastric mucosal cells (the mucosal bicarbonate layer; see
Chapter 33). As a consequence, the bulk of the absorption of
drugs, even weak acids such as aspirin, occurs in the small
intestine.
Drug formulation
A drug cannot be absorbed when it is taken in a tablet
or capsule until the vehicle disintegrates and the drug is
dissolved in the gastrointestinal contents to form a molecular
solution. Most tablets disintegrate and dissolve quickly and
completely, and the whole dose rapidly becomes available
for absorption. However, some formulations are designed to
disintegrate slowly so that the rate of release and dissolution

Pharmacokinetics 39

Absorption of drugs from the injection site can be prolonged
intentionally by incorporation of the drug into a lipophilic
vehicle, such as flupentixol decanoate (see Chapter 21),
creating a depot formulation from which the drug is released
over days or weeks.
Intranasal administration
The nasal mucosa provides a good surface area for absorption,
combined with low levels of proteases and drug-metabolising
enzymes compared with the gastrointestinal tract. As a
consequence, intranasal administration is used for the
administration of some drugs, such as triptan drugs for
migraines (see Chapter 26) and desmopressin (see Chapter
43), as well as drugs designed to produce local effects,
such as nasal decongestants and topical corticosteroids
(see Chapter 39).
Inhalation
Although the lungs possess the characteristics of a good site
for drug absorption (a large surface area and extensive blood
flow), inhalation is rarely used to produce systemic effects.
The principal reasons for this are the difficulty of delivering
nonvolatile drugs to the alveoli and the potential for local
toxicity to alveolar membranes. Therefore drug administration
by inhalation is largely restricted to:
■ volatile compounds, such as general anaesthetics (see
Chapter 17);
■ locally acting drugs, such as bronchodilators and
corticosteroids used in the treatment of airway disease
such as asthma and chronic obstructive pulmonary disease
(see Chapter 12).
Drugs in the latter group are not volatile and have to be
given either as aerosols containing droplets of dissolved
drug or as fine particles of the solid drug (dry powder; see
Chapter 12). Particles greater than 10 μm in diameter settle
out in the pharynx and upper airways, which are poor sites
for absorption, and the drug then passes back up the airways
via ciliary motion and is eventually swallowed. Particles less
than 1 μm in diameter are not deposited in the airways and
are immediately exhaled. It was shown some years ago
that only 4–6% of an inhaled dose is deposited in the small
airways, although the percentage of deposited drug may
be higher with modern inhaler devices delivering particle
sizes closer to the optimum for airways deposition (2–5 μm).
Minor routes
Drugs may be applied to any body surface or orifice to produce
a local effect. Absorption from the site of administration may
be important in limiting the duration of local action and the
production of unwanted systemic actions.
DISTRIBUTION
Distribution is the process by which the drug is transferred
reversibly from the general circulation into the tissues as the
concentrations in blood increase, and then returns from the

specificities of CYP3A4 and P-gp overlap, and for common
substrates their combined actions can prevent most of
the oral dose of some drugs reaching the hepatic portal
circulation.
Liver
Blood from the intestine is delivered by the splanchnic
circulation directly to the liver, which is the major site of
drug metabolism in the body. Hepatic first-pass metabolism
can be avoided by administering the drug to a region of the
gut from which the blood does not drain into the hepatic
portal vein (e.g. the buccal cavity or rectum); a good example
of this is the buccal administration of glyceryl trinitrate (see
Chapter 5).
Lung
Cells of the lungs have high affinities for many basic drugs
and are the main site of metabolism for many local hormones
via monoamine oxidase or peptidase activity.
ABSORPTION FROM OTHER ROUTES
Percutaneous (transcutaneous)
administration
The human epidermis (especially the stratum corneum) is
an effective permeability barrier to water loss and to the
transfer of water-soluble compounds. Although lipid-soluble
drugs are able to cross this barrier, the rate and extent
of entry are very limited. As a consequence, this route is
only effective for use with potent, nonirritant drugs, such as
glyceryl trinitrate (see Chapter 5) or fentanyl (see Chapter
19), or to produce a local effect. The slow and continued
absorption from dermal administration (e.g. via adhesive
patches) can be used to produce low but relatively constant
blood concentrations of some drugs (e.g. nicotine replacement
therapy; see Chapter 54).
Intradermal and
subcutaneous injection
Intradermal or subcutaneous injection avoids the barrier
presented by the stratum corneum, and entry into the general
circulation is limited mainly by the rate of blood flow to the
site of injection. However, these sites generally only allow
the administration of small volumes of drugs and tend to be
used mostly for local effects, such as local anaesthesia, or to
deliberately limit the rate of drug absorption (e.g. insulin; see
Chapter 40). Subdermal implants are increasingly used for
long-term hormonal contraception; the implants are flexible
polymer rods or tubes inserted under the skin of the upper
arm that slowly release the hormone for up to 3 years, with
contraception being reversible by removal of the implant
(see Chapter 45).
Intramuscular injection
The rate of absorption from an intramuscular injection
depends on two variables: the local blood flow and the
water solubility of the drug. An increase in either of these
factors enhances the rate of removal from the injection site.

1 40 Medical Pharmacology and Therapeutics

REVERSIBLE PROTEIN BINDING
Many drugs show an affinity for sites on nonreceptor proteins,
resulting in reversible binding:

Drug + − protein D  rug protein complex
Such binding occurs with plasma proteins, most commonly
with albumin, which binds many acidic or steroidal drugs, and
α1-acid glycoprotein, which binds many basic or neutral drugs
(Table 2.3). Drugs may also bind reversibly with intracellular

Table 2.2 Relative organ perfusion rates in a typical
adult at rest
Organ Proportion
of cardiac
output (%)

Blood flow
(mL/min
per 100 g
of tissue)

Well-perfused organs
Lung 100 1000
Adrenals 0.5 200
Kidneys 15 350
Thyroid 1.5 500
Liver 27 110
Heart 4 100
Gastrointestinal tract 15 300
Brain 12 56
Placenta (full term) — 10–15
Poorly perfused organs
Skin 5 12
Skeletal muscle 12 4
Bone, connective
tissue

3 3
Adipose (fat) 4 3

Plasma
Concentration

Concentration Concentration

Time

Time
Poorly perfused tissues
Well-perfused tissues

A
B

A
B

A
B
Time

Fig. 2.6 A simplified scheme for the redistribution of
drugs between tissues. The initial decrease in plasma
concentrations results from uptake into well-perfused tissues,
which essentially reaches equilibrium at point A. Between
points A and B, the drug continues to enter poorly perfused
tissues, resulting in a decrease in the concentrations in both
plasma and well-perfused tissues. At point B all tissues are
in equilibrium. The additional presence of an elimination
process would produce a decrease from point B (shown as a
dashed line), which would be parallel in all tissues.
Table 2.3 Examples of drugs that undergo
extensive binding to plasma protein
Bound to albumin Bound to α1-acid glycoprotein
Dexamethasone Chlorpromazine
Digitoxin Erythromycin
Furosemide Lidocaine
Ibuprofen Methadone
Indometacin Propranolol
Phenytoin Quinine
Salicylates Tricyclic antidepressants
Sulphonamides
Thiazides
Tolbutamide
Warfarin
tissues into blood when the blood concentrations decrease.
For most lipid-soluble drugs this occurs by passive diffusion
across cell membranes (see Fig. 2.2). Once equilibrium is
reached, any process that removes the drug from one side
of the membrane results in movement of drug across the
membrane to re-establish the equilibrium (see Fig. 2.4). Drugs
that are less lipid-soluble can penetrate tissues by diffusing
through intercellular junctions.
After an intravenous injection of the drug, there is a high
initial plasma concentration and the drug may rapidly enter
well-perfused tissues such as the brain, liver and lungs (Table
2.2). This may be so rapid that these tissues can be assumed
to equilibrate instantaneously with plasma and represent part
of the ‘central’ compartment (described later). However, the
drug will continue to enter poorly perfused tissues, and this
will lower the plasma concentration. The high concentrations
in the rapidly perfused tissues then decrease in parallel with
the decreasing plasma concentration, resulting in a transfer
of drug back into the plasma (Fig. 2.6). This redistribution is
important for terminating the action of some drugs given as a
rapid intravenous injection or bolus. For example, intravenous
thiopental produces rapid anaesthesia, but effects in the brain
are short-lived because continued uptake into muscle lowers
the concentrations in the blood and therefore indirectly in
the brain (see Figs 2.6 and 17.2).
The processes of elimination (such as metabolism and
excretion) are of major importance and are discussed in
detail in the upcoming sections. Elimination results in a net
transfer of the drug from other tissues via the circulation
to the organ(s) of elimination (see the dashed lines in
Fig. 2.6).

Pharmacokinetics 41

re-enter the circulation. In contrast, some covalent binding
may be slowly reversible, such as the formation of disulphide
bridges by captopril with its target (angiotensin-converting
enzyme [ACE]) and with plasma proteins (see Chapter 6); the
covalently bound drug will not dissociate rapidly in response
to a decrease in the concentration of unbound drug, and such
binding represents a slowly equilibrating reservoir of drug.
DISTRIBUTION TO SPECIFIC ORGANS
Two systems require more detailed consideration of drug
distribution: the brain, because of the difficulty of drug entry,
and the fetus, because of the potential for drug toxicity.
Brain
Lipid-soluble drugs readily pass from the blood into the
brain, and for such drugs the brain represents a typical
well-perfused tissue (see Table 2.2). In contrast, the entry of
water-soluble drugs into the brain is much slower than into
other well-perfused tissues, giving rise to the concept of a
blood–brain barrier as a highly selective permeability barrier
that separates the blood from the extracellular fluid of the
brain. The functional basis of the barrier to water-soluble drugs
(Fig. 2.7) is reduced permeability of brain capillaries owing to:
■ tight junctions between adjacent endothelial cells
(capillaries are composed of an endothelial layer a single
cell thick, with no smooth muscle),
■ smaller size and lower number of pores in the endothelial
cell membranes,
■ the presence of a contiguous surrounding layer of
astrocytes.

proteins. The drug–protein binding interaction resembles
the drug–receptor interaction since it is rapid, reversible
and saturable, and different ligands can compete for the
same site. It does not result in a pharmacological effect but
lowers the free concentration of the drug available to act at
receptors; the amounts of drug remaining available may be
only a minute fraction of the total body load. Proteins such
as albumin can therefore act as depots, rapidly releasing
the bound drug when the free drug is distributed to other
compartments or eliminated.
Competition for binding to proteins in plasma or inside
cells can occur between different drugs (drug interaction; see
Chapter 56) and between drugs and endogenous ligands.
A highly protein-bound drug such as aspirin can displace
other drugs such as warfarin from their binding sites on
plasma proteins; the increase in unbound drug concentration
can increase the biological activity of the displaced drug.
Relatively few such interactions have important clinical
consequences, although an example of drug interaction
with an endogenous ligand is the displacement of bilirubin
from albumin by sulphanilamide drugs, causing a potentially
dangerous increase in the bilirubin concentration in plasma,
leading to kernicterus.
IRREVERSIBLE PROTEIN BINDING
Certain drugs, because of chemical reactivity of the parent
compound or a metabolite, undergo covalent binding to
plasma or tissue components such as proteins or nucleic
acids. When the binding is irreversible (e.g. the interaction
of some cytotoxic drugs with DNA), this can be considered
as equivalent to elimination, because the parent drug cannot

Tight junctions between
endothelial cells

Endothelial cell
Mitochondrion

Nontight
junction

Astrocyte foot
projection
Mitochondrion
Carrier
system

Basal
lamina

Endothelial cell Nucleus
Tight
junction

Active transport

Astrocyte
Foot process of astrocyte

Fig. 2.7 The blood–brain barrier. The barrier arises from the low number of membrane pores, the tight junctions between
adjacent cells and the presence of efflux transporters that remove any drug that enters the endothelial cell. The presence of
astrocytes is the stimulus for these changes in endothelial structure and function. Astrocytes are one of the several types of
cells found in the central nervous system that make up the glia; they have numerous sheet-like processes and may provide
nutrients to neurons.

1 42 Medical Pharmacology and Therapeutics

■ Increase in activity. Sometimes metabolites may be more
potent than the parent drug; for example a prodrug is
an inactive compound that is converted by metabolism
into an active molecular species.
■ Change in the nature of the activity. The metabolite shows
qualitatively different pharmacological or toxicological
properties.
Drug metabolism can be divided into two phases (Fig.
2.8). Although many compounds undergo both phases of
metabolism, it is possible for a drug to undergo only a phase
1 or a phase 2 reaction, or for a proportion of the drug
to be excreted unchanged. Phase 1 metabolism (often an
oxidation, reduction or hydrolysis reaction) is sometimes
described as preconjugation, because it produces a molecule
that is a suitable substrate for a phase 2 or conjugation
reaction. The enzymes involved in these reactions have low
substrate specificities and can metabolise a wide range of
drug substrates and other xenobiotics.
Phase 1
Cytochrome P450 is a superfamily of membrane-bound
haemoprotein isoenzymes (Table 2.4). They are present in the
smooth endoplasmic reticulum of cells (Fig. 2.9), particularly in
the liver, which is the major site of drug oxidation; the amounts
in other tissues are low in comparison. The cytochrome P450
families CYP1 to CYP4 are involved in drug metabolism; the
specific isoenzymes CYP2C9, CYP2D6 and CYP3A4 are
involved in the phase 1 metabolism of approximately 10%,
24% and 55% of drugs, respectively.
Oxidation reactions (Table 2.5) are the most important of
the phase 1 reactions and can occur at carbon, nitrogen or
sulphur atoms within the drug structure. In most cases, an
oxygen atom is retained in the metabolite, although some
reactions, such as dealkylation, result in loss of the oxygen
atom in a small fragment of the original molecule. Oxidation
reactions are catalysed by a diverse group of enzymes, of
which the cytochrome P450 system is the most important.
The cytochrome P450 isoenzyme binds both the drug and
molecular oxygen (Fig. 2.10), and catalyses the transfer of
one oxygen atom to the substrate, while the other oxygen
atom is reduced to water:
RH + + O NADPH + → H ROH + + H O NADP + + 2 2

In addition, efflux transporters in the endothelial cells
are an important part of the blood–brain barrier and return
drug molecules back into the circulation, thereby preventing
their entry into the brain and reducing effects in the central
nervous system.
Water-soluble endogenous compounds needed for normal
brain functioning, such as carbohydrates and amino acids,
enter the brain via specific uptake transporters of the SLC
superfamily (see Table 2.1). Some drugs (e.g. levodopa) may
enter the brain using these transport processes, and in such
cases the rate of transport of the drug will be influenced by
the concentrations of competitive endogenous substrates.
There is limited drug-metabolising ability in the brain,
and drugs leave by diffusion back into plasma, by active
transport processes in the choroid plexus or by elimination
in the cerebrospinal fluid. Organic acid transporters of the
SLC superfamily (see Table 2.1) are important in removing
polar neurotransmitter metabolites from the brain.
Fetus
Lipid-soluble drugs can readily cross the placenta and enter
the fetus. The placental blood flow is low compared with that
in the liver, lung and spleen (see Table 2.2); consequently,
the fetal drug concentration equilibrates slowly with that
in the maternal circulation. Highly polar and very large
molecules (such as heparin; see Chapter 11) do not readily
cross the placenta. The fetal liver has only low levels of
drug-metabolising enzymes, so it is mainly the maternal
elimination that clears the fetal circulation of drugs.
After delivery, the neonate may show effects from drugs
given to the mother close to delivery (such as pethidine for
pain control; see Chapter 19). Such effects may be prolonged
because the neonate now has to rely on his/her own immature
elimination processes (see Chapter 56).
ELIMINATION
Elimination is the removal of a drug’s activity from the body
and may involve metabolism, by which the drug molecule
is transformed into a different molecule, and/or excretion,
in which the drug molecule is expelled in the body’s liquid,
solid or gaseous ‘waste’.
METABOLISM
A degree of lipid solubility is a useful property of most
drugs, since it allows the compound to cross lipid barriers
and hence to be given via the oral route. Metabolism is
necessary for the elimination of lipid-soluble drugs from the
body because it converts a lipid-soluble molecule into a
water-soluble species capable of rapid elimination in the
urine. A lipid-soluble molecule filtered into the kidney tubule
would otherwise be reabsorbed from the tubule into the
circulation (discussed later).
Metabolism of the drug produces one or more new
chemical entities, which may show different pharmacological
properties from the parent compound:
■ Decrease in biological activity. The most common result
of drug metabolism, arising from reduced receptor binding
due to the changed molecular structure.

Phase
1
Benzene
0%
Phenol
0.3%
Phenyl sulfate
99.9%+
Phase
2

Percentage
ionised
at pH 7.4

OH O – SO3

Fig. 2.8 The two phases of drug metabolism. Benzene
is shown as a simple example of a lipid-soluble substance
undergoing phase 1 and phase 2 metabolism. In Phase 1
benzene is oxidised to phenol and in Phase 2 the phenol
metabolite is conjugated with sulfate to produce a highly
hydrophilic metabolite-conjugate (phenyl sulfate).

Pharmacokinetics 43

Table 2.4 Cytochrome P450 superfamily
Isoenzyme Comments
CYP1A Important for
methylxanthines and
paracetamol; induced by
smoking
CYP2A Limited number of
substrates; significant
interindividual variability
CYP2B Limited number of
substrates
CYP2C CYP2C9 is an important
isoform; CYP2C19 shows
genetic polymorphism
CYP2D Metabolises numerous
drugs; CYP2D6 shows
genetic polymorphism
CYP2E Metabolises alcohol
CYP3A Main isoform in liver and
intestine; metabolises
50–60% of current drugs
CYP4 Metabolises fatty acids
Human liver contains at least 20 isoenzymes of cytochrome P450.
Families CYP1–4 are involved in drug metabolism.

D
Cytochrome
P450
Active
site

D

D in
cytosol
M

M

MS
D in
cytosol
M
MS
M
MGA

P450
Active
site
UDPGT

Phospholipid
bilayer

Fig. 2.9 Drug metabolism in the smooth endoplasmic reticulum. The lipid-soluble drug (D) partitions into the lipid bilayer
of the endoplasmic reticulum. The cytochrome P450 oxidises the drug to a metabolite (M) that is more water-soluble and
diffuses out of the lipid layer. The metabolite may undergo a phase 2 (conjugation) reaction catalysed by UDP-glucuronyl
transferase (UDPGT) in the endoplasmic reticulum to give a glucuronide conjugate (MGA) or with sulfate in the cytosol to give a
sulfate conjugate (MS).

Table 2.5 Examples of oxidation, reduction and
hydrolytic reactions
Oxidation
Alkyl groups RCH3 → RCH2OH → RCHO → RCOOH
Deamination RCH2NH2 → RCHO + NH3
Amines R′-NH-R → R′-N(OH)-R
Reduction
Aldehydes RCHO → RCH2OH
Disulfides R–S–S–R′ → RSH + HSR′
Hydrolysis
Esters RCO⋅OR′ → RCOOH + HOR′
Amides RCO⋅NHR′ → RCOH + H2NR′
R, R′, Aliphatic groups.

1 44 Medical Pharmacology and Therapeutics

Hydrolysis and hydration reactions (see Table 2.5) involve
the addition of water to the drug molecule. In hydrolysis, the
molecule is then split by the addition of water. A number of
ubiquitous enzymes are able to hydrolyse ester and amide
bonds in drugs. Intestinal bacteria are also important for the
hydrolysis of esters and amides, and of drug conjugates
eliminated in the bile (discussed later). In hydration reactions,
the water molecule is retained in the drug metabolite.
Hydration of an epoxide ring by epoxide hydrolase is an
important reaction in the metabolism and toxicity of a number
of aromatic drugs (e.g. carbamazepine; see Chapter 23).
Phase 2
Phase 2 (conjugation) reactions involve the formation of a
covalent bond between the drug, or its phase 1 metabolite,
and an endogenous substrate. Table 2.6 shows the types

The reaction involves initial binding of the drug substrate
to the ferric (Fe3+

) form of cytochrome P450, followed by
reduction (via a specific cytochrome P450 reductase) and then
binding of molecular oxygen. Further reduction is followed
by molecular rearrangement, with release of the reaction
products (drug metabolite and water) and regeneration of
ferric cytochrome P450.
Oxidations at nitrogen and sulphur atoms are frequently
performed by a second enzyme of the endoplasmic reticulum,
the flavin-containing mono-oxygenase, which also requires
molecular oxygen and NADPH. A number of other enzymes,
such as alcohol dehydrogenase, aldehyde oxidase and MAO,
may be involved in the oxidation of specific functional groups.
Reduction reactions (see Table 2.5) are less common
than oxidation reactions, but occur at unsaturated carbon
atoms and at nitrogen and sulfur centres by the actions of
cytochrome P450 and cytochrome P450 reductase (and also
by the intestinal microflora).

ROH
H2O Fe3+

Fe3+–RH

Fe2+–RH

Fe2+–RH

Fe3+–RH
Fe3+–RH

O2
_

O2

O2
e–
RH

e–
O2
2–
2H+

From reduced
cytochrome b5
or cytochrome
P450 reductase

From NADPH-
cytochrome P450

reductase

Fig. 2.10 The oxidation of substrate (RH) by cytochrome P450. Fe3+

, the active site of cytochrome P450 in its ferric state;
RH, drug substrate; ROH, oxidised metabolite. Cytochrome b5 is present in the endoplasmic reticulum and can transfer an
electron to cytochrome P450 as part of its redox reactions. NADPH, nicotinamide adenine dinucleotide phosphate.
Table 2.6 Major conjugation reactions
Reaction Functional group Activated species Products
Glucuronidation –OH
–COOH
–NH2

Uridine-diphosphate
glucuronic acid (UDPGA)

Glucuronide conjugates

Sulfation –OH
–NH3

3′-Phosphoadenosine-5′-
phosphosulfate (PAPS)

–O–SO3H
–NH–SO3H
Acetylation –NH2 Acetyl-CoA –NH–COCH3
Methylation –OH
–NH2
–SH

S-Adenosyl methionine –OCH3
–NHCH3
–SCH3
Amino acid conjugation –COOH Drug-CoA CO-NH⋅CHR⋅COOH
Glutathione conjugation Various – Glutathione conjugates

Pharmacokinetics 45

few days, during which the inducer interacts with nuclear
receptors to increase mRNA transcription of genes coding
for cytochrome P450. The increased amounts of the enzyme
last for a few days after the removal of the inducing agent,
and are removed by normal protein turnover. Environmental
contaminants such as organochlorine compounds (e.g.
dioxins) and polycyclic aromatic hydrocarbons (e.g. benzo[a]
pyrene in cigarette smoke) induce CYP1A. Therapeutic drugs
can induce members of the CYP2 and CYP3 families. Chronic
consumption of alcohol induces CYP2E.
In contrast, inhibition of cytochrome P450 by drugs
occurs by direct reversible competition for the enzyme site,
not a change in enzyme expression, so the time course
closely follows the absorption and elimination of the inhibitor
substance. Examples of inhibitors are the histamine H2 receptor
antagonist cimetidine (see Chapter 33) and components of
grapefruit juice (see Table 2.7).
The activity of drug-metabolising enzymes is also
dependent on the delivery of their drug substrates by the
circulation. The metabolism of many drugs is affected
significantly by lower hepatic blood flow in the very young
and in the elderly (see Chapter 56). Genetic variation in
drug-metabolising enzymes is discussed at the end of this
chapter.
EXCRETION
Drugs and their metabolites may be eliminated via circulation
by various routes:
■ In fluids (urine, bile, sweat, tears, breast milk, etc.).
Important for low-molecular-weight polar compounds;

of phase 2 reactions, the functional group necessary in the
drug molecule and the activated species needed for the
reaction. The products of conjugation reactions are usually
highly water-soluble and lack biological activity.
The activated endogenous substrate for glucuronide
synthesis is uridine-diphosphate glucuronic acid (UDPGA).
UDP-glucuronyl transferases in the endoplasmic reticulum
close to the cytochrome P450 system (see Fig. 2.9) transfer
glucuronate to the drug. Glucuronide conjugation in the gut
wall and liver is important in the first-pass metabolism of
substrates such as simple phenols.
Sulfate conjugation is performed by a cytosolic enzyme,
which utilises high-energy sulphate (3′-phosphoadenosine-5′-
phosphosulfate or PAPS) as the rate-limiting endogenous
substrate. Saturation of sulfate conjugation contributes to
the hepatotoxic consequences of overdose with paracetamol
(acetaminophen; see Chapter 53).
Acetylation and methylation reactions often decrease
polarity because they block an ionisable functional group
(see Table 2.6), but they mask active groups such as amino
and catechol moieties. These reactions are primarily involved
in inactivation of neurotransmitters such as noradrenaline
and local hormones such as histamine.
The conjugation of drug carboxylic acid groups with amino
acids is unusual because the drug is converted to a high-energy
form (a Coenzyme A derivative) prior to the formation of the
conjugate bond by transferase enzymes. Conjugation with the
tripeptide glutathione (GSH or L-α-glutamyl-L-cysteinylglycine)
is catalysed by a family of transferases which covalently bind
the drug to the thiol group in the cysteine (Fig. 2.11). The
substrates are often reactive drugs or activated metabolites,
which are inherently unstable (see Chapter 53), and the reaction
can also occur nonenzymatically. The glutathione conjugate
then undergoes further metabolic reactions.
Glutathione conjugation is a detoxification process in which
glutathione acts as a scavenging agent to protect the cell
from toxic damage. Glutathione conjugates and endogenous
cysteine conjugates, such as the cysteinyl-leukotriene (LT)
C4, are transported out of cells by the MRP1 transporter
(see Table 2.1).
The complex array of biotransformation reactions typically
involved in drug metabolism is illustrated by the anxiolytic
drug diazepam (see Chapter 20), which is metabolised
to biologically active intermediates before undergoing
conjugation with glucuronide (Fig. 2.12).
Factors affecting drug metabolism:
inducers and inhibitors
The liver is the main site of drug metabolism; the large surface
area of the sinusoids, combined with high levels of enzyme
activity in hepatocytes, can result in very rapid drug uptake and
metabolism as the blood flows through the liver (see Chapter
56 for normal sinusoid architecture and the effects of liver
disease on hepatic drug uptake). Environmental influences
including chemical contaminants and therapeutic drugs may
induce or inhibit the activity of hepatic drug-metabolising
enzymes, particularly cytochrome P450 (Table 2.7). This can
affect both the bioavailability and the elimination of other drugs
undergoing hepatic elimination (discussed later).
Inducing agents increase the cellular expression of
cytochrome P450 enzymes. This occurs over a period of a

RX

R S CYS
GLU
GLY
Hydrolysis
R S Cysteine
Lyase
R SH
Further metabolism

Excretory
product
N-acetylation

GLU
CYS
GLY
HS

Glutathione transferase
or spontaneous reaction
Glutathione Unstable drug or
reactive metabolite

Fig. 2.11 The formation and further metabolism
of glutathione conjugates. There are multiple types
of glutathione transferase that detoxify substances by
glutathione conjugation.

1 46 Medical Pharmacology and Therapeutics

compounds free in the plasma enter the filtrate on each pass.
Plasma proteins and protein-bound drugs are not filtered, so
the efficiency of glomerular filtration for a drug is influenced
by the extent of plasma protein binding.
Reabsorption
The glomerular filtrate contains numerous constituents
that the body cannot afford to lose. Most of the water is
reabsorbed, and there are specific tubular uptake processes
for carbohydrates, amino acids, vitamins and so on (see
Chapter 14). A few drugs also pass from the tubule back
into the plasma, as they are substrates for these specific
uptake processes. The urine is concentrated on its passage
down the renal tubule; as the tubule-to-plasma concentration
gradient increases, only the most polar molecules remain in
the urine. Because of extensive reabsorption, lipid-soluble
drugs are not eliminated via the urine, but are returned to
the circulation until they are metabolised to water-soluble
products, which are then efficiently removed from the body
by renal excretion. The pH of urine is usually less than that
of plasma; consequently, pH partitioning between urine (pH
5–6) and plasma (pH 7.4) may increase or decrease the
tendency of the compound to be reabsorbed (see Fig. 2.5).
Tubular secretion
The renal tubule also has secretory transporters (see Table
2.1) on both the basolateral and apical membranes

urine is the major route; milk is important because of
the potential for exposure of the breastfed infant.
■ In solids (faeces, hair, etc.). Faecal elimination is
most important for high-molecular-weight compounds
excreted in bile; the sequestration of drugs into hair is
not quantitatively important, but the distribution of a drug
along the hair shaft can indicate the history of drug intake
during the preceding weeks.
■ In gases (expired air). Important only for volatile
compounds
Excretion via the urine
There are three processes involved in the handling of drugs
and their metabolites in the kidney: glomerular filtration,
reabsorption and tubular secretion. The total urinary excretion
of a drug depends on the balance of these three processes:
Total excretion glomerular filtration
tubular secretion re
=
+ − absorption

Glomerular filtration
All molecules less than about 20 kDa undergo filtration under
positive hydrostatic pressure through pores of 7–8 nm diameter
in the glomerular membrane. The glomerular filtrate comprises
about 20% of the flow of plasma to the glomeruli, and
hence about 20% of all water-soluble, low-molecular-weight
Cytochrome P450

Cytochrome P450
Ring oxidation

N-demethylation
N
N
CH3
O

Cl

Diazepam

UDPGT

Cytochrome P450

Water-soluble
glucuronide
conjugate

Water-soluble
glucuronide
conjugate

N-demethylation
N
N
CH3
O

Cl

Temazepam

N
N
H
O

Cl

Oxazepam

Cytochrome P450
N
N
H
O

Cl

Desmethyldiazepam
(nordiazepam)

OH

UDPGT
OH

Ring oxidation

Fig. 2.12 Complex pathways of metabolism in humans. This figure illustrates that a single drug, in this case diazepam,
may generate a number of active metabolites before phase 2 conjugation terminates the activity of the parent drug and
metabolites.

Pharmacokinetics 47

Excretion via the faeces
Uptake into hepatocytes and subsequent elimination in
bile is the principal route of elimination of larger molecules
(molecular weight >500 Da). Conjugation with glucuronic
acid increases the molecular weight of the substrate by
almost 200 Da, so bile is an important route for eliminating
glucuronide conjugates. Once the drug or its conjugate has
entered the intestinal lumen via the bile (Fig. 2.13), it passes
down the gut and may eventually be eliminated in the faeces.
However, some drugs may be reabsorbed from the lumen of

for compounds that are acidic (OATs) or basic (organic
cation transporters [OCTs]). Drugs and their metabolites,
especially the glucuronic acid and sulphate conjugates, may
undergo an active carrier-mediated elimination, primarily by
OATs but also by multidrug-resistance-associated proteins
(MRPs). Because tubular secretion rapidly lowers the plasma
concentration of unbound drug, there will be a rapid
dissociation of any drugs bound to proteins in the plasma.
As a result, even highly protein-bound drugs may be cleared
almost completely from the blood in a single passage through
the kidney.
Table 2.7 Examples of common substrates, inhibitors and inducers of cytochrome P450 isoenzymes
Isoenzyme Substrates Inhibitors Inducers
CYP1A2 Caffeine, clozapine,
haloperidol, naproxen,
olanzapine, paracetamol,
theophylline, verapamil

Amiodarone, cimetidine,
efavirenz, grapefruit juice

Carbamazepine, chargrilled
meat, cigarette smoke,
rifampicin

CYP2A6 Coumarin, halothane,
nicotine

Grapefruit juice,
ketoconazole,
tranylcypromine

Dexamethasone,
phenobarbital, rifampicin

CYP2B6 Bupropion,
cyclophosphamide,
efavirenz, ifosfamide,
ketamine, methadone,
propofol, selegiline

Orphenadrine, ticlopidine,
voriconazole

Artemisinin, carbamazepine,
phenobarbital, phenytoin,
rifampicin

CYP2C8 Montelukast, repaglinide Gemfibrozil, montelukast Phenobarbital, rifampicin
CYP2C9 Celecoxib, diclofenac,
glibenclamide, glipizide,
ibuprofen, losartan,
tolbutamide, S-warfarin

Amiodarone, efavirenz,
fluconazole, isoniazid,
metronidazole, paroxetine,
voriconazole

Carbamazepine,
phenobarbital, rifampicin, St
John’s wort

CYP2C19 Citalopram, clopidogrel,
diazepam, omeprazole,
pantoprazole, phenytoin,
proguanil

Cimetidine, fluoxetine,
fluvoxamine, ketoconazole,
lansoprazole,
moclobemide, omeprazole,
oral contraceptives,
pantoprazole, voriconazole

Efavirenz, rifampicin,
ritonavir, St John’s wort

CYP2D6 Amitriptyline, bisoprolol,
codeine, desipramine,
encainide, many SSRIs,
metamfetamine, metoprolol,
ondansetron, propranolol,
risperidone

Amiodarone, bupropion,
cimetidine, duloxetine,
fluoxetine, paroxetine,
quinidine

Carbamazepine,
phenobarbital, phenytoin,
rifampicin

CYP2E Chlorzoxazone, ethanol,
halothane (and other
inhalation anaesthetics),
paracetamol

Disulfiram Ethanol

CYP3A4 Numerous drugs of many
different classes, e.g.
amlodipine, alfentanil,
atorvastatin, carbamazepine,
ciclosporin, diazepam,
diltiazem, erythromycin,
fluconazole, lidocaine,
midazolam, nifedipine,
saquinavir, sildenafil,
tamoxifen, terfenadine

Antivirals (indinavir,
nelfinavir, ritonavir),
clarithromycin, erythromycin,
grapefruit juice, itraconazole,
ketoconazole

Carbamazepine, efavirenz,
phenobarbital, phenytoin,
pioglitazone, rifampicin, St
John’s wort

SSRIs, Selective serotonin re-uptake inhibitors.
Information taken mainly from Flockhart, D.A., 2007. Drug interactions: cytochrome P450 drug interaction table. Indiana University School of
Medicine. <http://medicine.iupui.edu/clinpharm/ddis/clinical-table> (accessed 16.04.2016).

1 48 Medical Pharmacology and Therapeutics

■ the dosage adjustment that may be necessary in hepatic
and renal disease,
■ the calculation of dosages for vulnerable patient groups.
Calculation of drug doses may be required in the UK
Prescribing Safety Assessment (PSA) (see Chapter 55).
GENERAL CONSIDERATIONS
The processes of drug absorption, distribution, metabolism
and excretion (ADME) are described in this section in
mathematical terms, as it is important to quantify the rate
and extent to which the drug undergoes each process.
For nearly all physiological and metabolic processes,
the rate of reaction is not uniform but proportional to the
amount of substrate (drug) available: this is described
as a first-order reaction. Diffusion down a concentration
gradient, glomerular filtration and enzymatic hydrolysis are
examples of first-order reactions. At higher concentrations,
more drug diffuses or is filtered or hydrolysed than at
lower concentrations. Protein-mediated reactions, such as
metabolism and active transport, are also first order, because
if the concentration of the substrate is doubled, then the rate
of formation of product is also doubled. However, as the
substrate concentration increases, the enzyme or transporter
can become saturated with substrate, and the rate of reaction
cannot respond to a further increase in concentration. The
process then occurs at a fixed maximum rate independent
of substrate concentration, and the reaction is described as
a zero-order reaction; rare examples are the metabolism of
ethanol (see Chapter 54) and phenytoin (see Chapter 23).
When the substrate concentration has decreased sufficiently
for protein sites to become available again, then the reaction
will revert to first order.
Zero-order reactions
If a drug is being processed (absorbed, distributed or
eliminated) according to zero-order kinetics, then the change
in concentration with time (dC/dt) is a fixed amount (mass)
of the drug per time, independent of concentration:

d
d
C
t = −k (2.2)
The units of k (the reaction rate constant) are therefore a
mass per unit time (e.g. mg/min). A graph of concentration
against time will produce a straight line with a slope of −k
(Fig. 2.14A).
First-order reactions
In first-order reactions, the change in concentration at any
time (dC/dt) is proportional to the concentration present
at that time:
d
d
C
t = −kC (2.3)
The rate of change will be high at high drug concentrations
but low at low concentrations (see Fig. 2.14B), and a graph
of concentration against time will produce an exponential

the gut and re-enter the hepatic portal vein. As a result, the
drug is recycled between the gut lumen, hepatic portal vein,
liver, bile and back to the gut lumen; this is described as
enterohepatic circulation. Some of the reabsorbed drug may
escape hepatic extraction and proceed into the hepatic vein,
maintaining the drug concentrations in the general circulation.
Highly polar glucuronide conjugates of drugs or their
oxidised metabolites that are excreted into the bile undergo
little reabsorption in the upper intestine. However, the bacterial
flora of the lower intestine may hydrolyse the conjugate, so
the original, lipid-soluble drug or its metabolite is liberated and
can be reabsorbed and undergo enterohepatic circulation.
THE MATHEMATICAL BASIS
OF PHARMACOKINETICS
The use of mathematics to describe the fate of a drug in
the body can be complex and daunting for undergraduates.
Nevertheless, a basic knowledge is essential for understanding
many aspects of drug handling and the rational prescribing
of drugs:
■ why oral and intravenous treatments may require different
doses,
■ the calculation of dosages and dose intervals during
chronic therapy,
■ why a loading dose may be needed,
General circulation

Bile

Drug

Drug

Drug

Conjugate
Liver

Conjugate
Conjugate
Conjugate
Drug

Drug

Small
intestine

Colon/rectum
Bacterial hydrolysis
Fig. 2.13 Enterohepatic circulation of drugs. Drug
molecules may circulate repeatedly between the bile, gut,
portal circulation, liver and general circulation, particularly if
the drug conjugate is hydrolysed by the gut flora.

Pharmacokinetics 49

ABSORPTION
The mathematics of absorption apply to all nonintravenous
routes (e.g. oral, inhalation, percutaneous) and are illustrated
by absorption from the gut lumen.
RATE OF ABSORPTION
The rate of absorption after oral administration is determined
by the rate at which the drug is able to pass from the gut
lumen into the systemic circulation. Following oral doses
of some drugs, particularly lipid-soluble drugs with very
rapid absorption, it may be possible to see three distinct
phases in the plasma concentration–time curve, which reflect
distinct phases of absorption, distribution and elimination
(Fig. 2.16A). However, for most drugs slow absorption masks
the distribution phase (see Fig. 2.16B). A number of factors
can influence this pattern.
■ Gastric emptying. Basic drugs undergo negligible
absorption from the stomach, so there can be a delay
of up to an hour between drug administration and the
detection of drug in the general circulation.
■ Food. Food in the stomach slows drug absorption and
also slows gastric emptying.
■ Decomposition or first-pass metabolism before or
during absorption. This will reduce the amount of drug
that reaches the general circulation but will not affect the
rate of absorption, which is usually determined by lipid
solubility.
■ Modified-release formulation. If a drug is eliminated
rapidly, the plasma concentrations will show rapid
fluctuations during regular oral dosing, and it may be
necessary to take the drug at very frequent intervals
to maintain a therapeutic plasma concentration. The
frequency with which a drug is taken can be reduced
by giving a modified-release formulation that releases
drug at a slower rate. The plasma concentration then
becomes more dependent on the rate of absorption than
the rate of elimination.

decrease. Such a curve can be described by an exponential
equation:

C C kt = − 0e (2.4)
where C is the concentration at time t and C0 is the initial
concentration (when time = 0). This equation may be written
more simply by taking natural logarithms:

lnC C= − 0 kt (2.5)
and a graph of lnC against time will produce a straight line
with a slope of −k and an intercept of lnC0 (see Fig. 2.14C).
The units of the rate constant k (1/time, e.g. per hour)
may be regarded as the proportional change per unit of time
but are difficult to use practically, so the rate of a first-order
reaction is usually described in terms of its half-life (t1/2),
which is the time taken for a concentration to decrease by
one-half. In the next half-life, the drug concentration falls
again by one-half, to a quarter of the original concentration,
and then to one-eighth in the next half-life, and so on. The
half-life is therefore independent of concentration and is
a characteristic for a particular first-order process and a
particular drug. The intravenous drug shown in Fig. 2.15
has a t1/2 of 1 hours.
The relationship between the half-life and the rate constant
is derived by substituting C0 = 2 and C = 1 into the previous
equation, when the time interval t will be one half-life (t1/2),
giving:

ln ln
.
. .
1 2
0 0 693
0 693 0 693
1 2
1 2
1 2 1 2
= −
= −
= =
kt
kt
t k or k t

(2.6)

(Note: 0.693 = ln2.)
A half-life can be calculated for any first-order process
(e.g. for absorption, distribution or elimination). In practice,
the ‘half-life’ normally reported for a drug is the half-life for
the elimination rate from plasma (the slowest, terminal phase
of the plasma concentration–time curve; discussed later).
C
Slope = –k
(units = mass/time)

Time Time Time
Slope = –k
(units = 1/time)
InC0

Zero order First order

C

In
C

A B C
Fig. 2.14 Zero- and first-order kinetics. (A) The zero-order reaction is a uniform change in concentration C over time,
representing the same amount (mass) of drug being removed per unit of time. (B) The first-order reaction is an exponential
curve in which concentrations fall fastest when they are highest; the curve reflects the same proportion of drug being removed
per unit of time. (C) Plotting the natural logarithm of the concentration (lnC) in a first-order reaction against time generates a
straight line with slope −k (where k is the rate constant) and the intercept gives the concentration at time zero, C0.

1 50 Medical Pharmacology and Therapeutics

The bioavailability of a drug has important therapeutic
implications, because it is the major factor determining
the drug dosage for different routes of administration.
For example, if a drug has an oral bioavailability of 0.1,
the oral dose needed for therapeutic effectiveness will
need to be 10 times higher than the corresponding
intravenous dose.
The bioavailability is a characteristic of the drug and
independent of dose, providing that absorption and
elimination are not saturated. Bioavailability is normally
determined by comparison of plasma concentration data
obtained after oral administration (when the fraction F of
the parent drug enters the general circulation) with data
following intravenous administration (when, by definition,

EXTENT OF ABSORPTION
Bioavailability (F) is defined as the fraction of the administered
dose that reaches the systemic circulation as the parent drug
(unaltered, not as metabolites). For intravenous administration
the bioavailability (F) is therefore 1, as 100% of the parent
drug enters the general circulation. For oral administration,
incomplete bioavailability (F < 1) may result from:
■ incomplete absorption and loss in the faeces, because
either the molecule is too polar to be absorbed or the
tablet did not release all of its contents;
■ first-pass metabolism in the gut lumen, during passage
across the gut wall or by the liver before the drug reaches
the systemic circulation.
100
80
60
40
20
0
0 1 2 3 4

5
4
3
2
1
0
0 1 2 3 4
Time (h) Time (h)

C

In (concentration)

Slope = –k

Fig. 2.15 The elimination half-life of a drug in plasma. Here the drug concentration C decreases by 50% every hour (i.e.
the half-life is 1 hour).

Rate of absorption > rate of distribution

Time after dosage
Plasma drug
concentration

Absorption phase
Distribution phase
Elimination phase

Rate of absorption < rate of distribution

Time after dosage
Plasma drug
concentration
Absorption phase
Distribution phase
Elimination phase

A B
Fig. 2.16 Plasma concentration–time profiles after oral administration of drugs with different rates of absorption.
The processes of distribution and elimination start as soon as some of the drug has entered the general circulation. (A) A clear
distribution phase is seen if the rate of absorption is so rapid as to be essentially complete before distribution is finished. (B)
For most drugs, the rate of absorption is slower and masks the distribution phase.

Pharmacokinetics 51

DISTRIBUTION
Distribution concerns the rate and extent of movement
of the parent drug from the blood into the tissues after
administration and its return from the tissues into the blood
during elimination.
RATE OF DISTRIBUTION
Because a distinct distribution phase is not usually seen
when a drug is taken orally (see Fig. 2.16B), the rate of
distribution is normally measured following an intravenous
bolus dose. Some intravenous drugs reach equilibrium
between blood and tissues very rapidly, and a distinct
distribution phase is not apparent. In Fig. 2.17A the slope
of plasma concentration against time therefore mainly reflects
elimination of the drug; this is described as a one-compartment
model.
Most intravenous drugs, however, take a finite time to
distribute into the tissues; the initial distribution out of the
plasma, combined with underlying elimination, produces a
steep initial slope (slope A–B in Fig. 2.17B), followed by
a slower terminal phase (slope B–C) in which distribution
has been largely completed and elimination predominates.
Back-extrapolation of this terminal elimination phase to

F = 1). The amount in the circulation cannot be compared
at a single time point, because intravenous and oral dosing
show different concentration–time profiles, so instead
the total area under the plasma concentration–time curve
(AUC) from t = 0 to t = infinity is used, as this reflects
the total amount of drug that has entered the general
circulation. If the oral and intravenous (iv) doses administered
are equal:

F = AUC
AUC
oral
iv

(2.7)

or if different doses are used:
F = ×
×
AUC Dose
AUC Dose
oral iv
iv oral

(2.8)
This calculation assumes that the elimination is first order.
An alternative method to calculate F is to measure
the total urinary excretion of the parent drug (Aex)
following oral and intravenous administration of identical
doses:

F = Aex
Aex
oral
iv

(2.9)

A

B

In
C

In
C

Time
Slope = –k

Model Model

Dose V Elimination

One-compartment Two-compartment
Instantaneous distribution

A

B

Time

Dose V1
Slower distribution

C

V2

Elimination
D

A B

Fig. 2.17 Plasma concentration–time curves for the distribution of intravenous drugs into one- and two-compartment
models. (A) When distribution of an intravenous drug bolus into tissues is so rapid as to be essentially instantaneous, the slope
of the plasma concentration–time curve mainly reflects the rate of elimination (one-compartment model). (B) When distribution
is slower, the initial fall in concentration (slope A–B) is due to simultaneous distribution and elimination followed by the terminal
elimination phase (two-compartment model; slope B–C). Back-extrapolating to D at time zero allows the contribution of
distribution during A–B to be distinguished from the underlying contribution of elimination.

Comments

Search This Blog

Archive

Show more

Popular posts from this blog

TRIPASS XR تري باس

CELEPHI 200 MG, Gélule

ZENOXIA 15 MG, Comprimé

VOXCIB 200 MG, Gélule

Kana Brax Laberax

فومي كايند

بعض الادويه نجد رموز عليها مثل IR ، MR, XR, CR, SR , DS ماذا تعني هذه الرموز

NIFLURIL 700 MG, Suppositoire adulte

Antifongiques مضادات الفطريات

Popular posts from this blog

علاقة البيبي بالفراولة بالالفا فيتو بروتين

التغيرات الخمس التي تحدث للجسم عند المشي

إحصائيات سنة 2020 | تعداد سكَان دول إفريقيا تنازليا :

ما هو الليمونير للأسنان ؟

ACUPAN 20 MG, Solution injectable

CELEPHI 200 MG, Gélule

الام الظهر

VOXCIB 200 MG, Gélule

ميبستان

Popular posts from this blog

TRIPASS XR تري باس

CELEPHI 200 MG, Gélule

Popular posts from this blog

TRIPASS XR تري باس

CELEPHI 200 MG, Gélule

ZENOXIA 15 MG, Comprimé

VOXCIB 200 MG, Gélule

Kana Brax Laberax

فومي كايند

بعض الادويه نجد رموز عليها مثل IR ، MR, XR, CR, SR , DS ماذا تعني هذه الرموز

NIFLURIL 700 MG, Suppositoire adulte

Antifongiques مضادات الفطريات

Popular posts from this blog

Kana Brax Laberax

TRIPASS XR تري باس

PARANTAL 100 MG, Suppositoire بارانتال 100 مجم تحاميل

الكبد الدهني Fatty Liver

الم اسفل الظهر (الحاد) الذي يظهر بشكل مفاجئ bal-agrisi

SEDALGIC 37.5 MG / 325 MG, Comprimé pelliculé [P] سيدالجيك 37.5 مجم / 325 مجم ، قرص مغلف [P]

نمـو الدمـاغ والتطـور العقـلي لـدى الطفـل

CELEPHI 200 MG, Gélule

أخطر أنواع المخدرات فى العالم و الشرق الاوسط

Archive

Show more