You are on page 1of 38

2

Clinical pharmacology
Michael Rawlins, Munir Pirmohamed

Introduction 17
Why do patients need drugs? 17
The choice of drug 17
The dose 18
Pharmacokinetics 19
Pharmacodynamics 20
Affordability and cost-effectiveness 20
Prescribing in special populations 21
Monitoring drug therapy 22
Adverse drug reactions 22
Drug interactions 24
Inter-individual variability in drug response 25
Evidence-based medicine 26
Statistical analyses 29
Information sources 30

Introduction
Prescribing is a key skill required by all doctors, at least at some stage in their
careers. This chapter provides an introduction to the principles of rational
therapeutics. The 1984 ‘Nairobi Declaration’ emphasized that prescribing should
be to the right patient with the right drug at the right dose, and at an affordable
cost. Thus, the prescription of a medicine needs to take into account the potential
benefits and harms of that medicine, and ensure that every step is taken to
maximize the benefit : harm ratio. The key principles of prescribing are
described in Box 2.1.

 Box 2.1
Key principles of prescribing
• Be clear about the reasons for prescribing. Obtain an accurate diagnosis
and have clear aims of what the benefits will be from prescribing the drug.
• Obtain an accurate drug history. Patients should be asked about their
current medications, over-the-counter herbal medicines, and illicit drug
usage. A past history of intolerances, including true allergic reactions, is also
required.
• Obtain an accurate history of other factors that might affect the benefit : 
harm ratio of drugs. Take note of liver and renal impairment, age, whether
patient is pregnant or breast-feeding, and co-morbidities.
• Establish what the patient expects from the drug, and deal with any
concerns. Provide good information for the patient and answer any
questions clearly.
• Select the most clinically effective, cost-effective and safe medicine for the
patient, based on an assessment of the individual. Optimize the benefit : 
harm balance by choosing the right dose and frequency, and the right drug,
in the the best formulation, route of administration and duration of
treatment. Prescribe within the licence of the medicine, except when no
alternative is available.
• Adhere to guidelines and local and national formularies. Based on the
needs of the individual patient, use the most reliable information to identify
the medicine most suited to the patient.
• Write prescriptions legibly using the correct documents, or prescribe
electronically, depending on availability. Take care to avoid medication
errors.
• Monitor the patient for both efficacy and safety after starting the drug.
This is essential to optimize dose, identify adverse reactions, and determine
when to stop the medicine. Report adverse reactions using spontaneous
reporting schemes, if appropriate.
• Communicate clearly with other healthcare professionals and patients
about the reasons for prescribing decisions. Communication is essential to
optimize the benefit : harm ratio of drugs, particularly when there are
multiple prescribers for a single patient.
• Prescribe within your competencies. Prescribing should be within the limits
of your knowledge and experience. Do not be afraid to ask for advice, and
ensure that complex prescriptions are checked (e.g. when calculating doses).
(Adapted from British Pharmacological Society;
http://www.bps.ac.uk/SpringboardWebApp/userfiles/bps/file/Clinical/BPSPrescribingStatement03Feb201
0.pdf.)

Why Do Patients Need Drugs?


There are many reasons for prescribing drugs to patients. These are as follows:

Disease treatment
A reliable diagnosis is crucial before starting treatment, which may either cure
or, more often, control a disease process. An accurate diagnosis ensures that a
patient is not exposed, unnecessarily, to the hazards or costs of a particular
intervention. However, a prior diagnosis may not be possible in every
circumstance; for example, in infection, the initiation of ‘blind’ antimicrobial
therapy is justified when delay would expose a patient to hazard or discomfort.
The choice of the drug often depends on clinical factors (such as age,
concomitant disease and concurrent therapy), pharmaceutical factors (the
availability of other medicines and relative cost-effectiveness) and, increasingly,
individual (host) factors. For instance, trastuzumab is only of value in women
with breast cancer whose malignant cells express the HER2 epidermal growth
factor receptor. Tailoring or targeting treatment, dependent on the use of
biomarkers (e.g. genetic polymorphisms or gene expression pattern), is
increasingly being used. This promising approach has become known as
‘personalized medicine’.

Symptom relief
Drugs are frequently used in the relief of symptoms: for example, in the
treatment of severe pain, constipation or pruritus. The issue here is to monitor
the patient so that the effectiveness of symptom relief can be judged and, if it is
inadequate, either the dose or drug can be altered. Ensure that the drug is stopped
when no longer indicated. Doctors are good at starting drugs but very bad at
stopping them!

Prevention
Medicines are also given to otherwise healthy individuals. In such
circumstances, there must be a very clear imperative to ensure that the benefits
to the individual outweigh the harm. Examples include:
• immunization against serious microbial infections (e.g. influenza vaccination)
• the reduction of individual risk factors to prevent later disease (e.g. the use of
antihypertensive or lipid-lowering agents to reduce the chances of ischaemic
heart disease and stroke)
• oral contraceptives in sexually active women wishing to avoid pregnancy.

The Choice of Drug


Selecting the right drug involves three elements:
• the drug's clinical efficacy for the proposed use
• the balance between the drug's efficacy and safety
• patient preference.
The most common approach to assessing a drug's efficacy is the randomized
controlled trial (RCT), although other approaches (see p. 28) can be informative.
The demonstration of absolute efficacy (against placebo) may, itself, be
insufficient. Where there is more than one treatment for the same indication,
these should be compared with one another, taking account of the magnitude of
their benefits, their individual adverse reaction profiles, and their costs. There
are also ethical concerns to the use of placebos; these should only be used when
there are no other effective treatments, when withholding drugs poses no serious
risk of harm, and when there are methodological reasons (but where there is no
risk of harm from withholding treatment).
Direct comparisons of one treatment versus another are particularly useful but
are often unavailable. In such cases, indirect techniques can be used. For
example, none of the novel oral anticoagulants (dabigatran, rivaroxaban,
apixaban and edoxaban) has been tested against the others, but comparative
effectiveness can be modelled using indirect comparison of the individual RCTs
in which each drug was compared with warfarin.
Patients' own preferences should be discussed to enable them to be equal
partners in decision-making about whether, and how, they wish to be treated.
Moreover, a full understanding of the reasons for considering treatment, the
likely benefits and the possible adverse reactions has repeatedly been shown to
improve ‘concordance’ with treatment regimens.

Further reading
Millum J, Grady C. The ethics of placebo-controlled trials: methodological
justifications. Contemp Clin Trials 2013; 36:510–514.
Skjoth F, Larsen TB, Rasmussen LH. Indirect comparison studies – are they
useful? Insights from the novel oral anticoagulants for stroke prevention in
atrial fibrillation. Thromb Haemost 2012; 108:405–406.

The Dose
The dose administered to the patient is crucial in determining both efficacy and
safety of medicines. Appropriate drug dosages will usually have been
determined from the results of so-called ‘dose-ranging’ studies during the
original development programme.
In some cases, drug doses are fixed, with all patients being given the same
dose: for example, levonorgestrel for emergency contraception. However, these
situations are unusual. For example, although the majority of patients are given
75 mg per day of aspirin, doses of up to 300 mg per day have been used for the
secondary prevention of myocardial infarction.
More commonly, doses are titrated. For many drugs, there are wide inter-
individual variations in response. As a consequence, whilst a particular dose
may, in one person, lack any therapeutic effect, the same dose in another may
cause serious toxicity. Prescribers should start at a low dose and titrate to the
most effective dose, which should, in all cases, determine the lowest effective
dose. This is necessary because of the inter-individual variability in the dose–
response curve, whereby there is an increase in response with an increase in
dose, until a plateau effect is reached (Fig. 2.1).
FIGURE 2.1 Dose–response curve, showing the variable response of the same dose
in different patients.

The reasons for this variability in response can be divided into two broad
areas:
• Pharmacokinetic factors, where the dose administered does not necessarily
equate with systemic or tissue drug concentrations: that is, exposure. This may
be due to differences in the rates of drug absorption, metabolism and/or
excretion.
• Pharmacodynamic factors, where there are differences in the sensitivity of the
drug target that generally fall into the categories of receptors, enzymes, ion
channels, transporters, DNA or transcription factors.

Pharmacokinetics
Pharmacokinetics is the study of what the body does to a drug. This can be
divided into four different processes: absorption, distribution, metabolism and
excretion.

Absorption
Drugs can be administered by various different routes (Box 2.2) but oral
administration is the most common. The main determinants of a drug's plasma
concentration after oral administration are its bioavailability (Box 2.3) and its
rate of systemic clearance (by hepatic metabolism or renal excretion). A drug's
oral bioavailability depends on the extent to which it is:
• Destroyed in the gastrointestinal tract, especially by gastric acid.
• Unable to cross the gastrointestinal epithelium. Drugs can cross either by
passive diffusion or by active uptake by transporters that are abundant in the
gut epithelial barrier. Such transporters can be either influx transporters
(pumping drugs from the gut lumen into the cell, such as the PEPT1 transporter
for the absorption of penicillins) or efflux transporters (pumping drugs from
cells back into the gut lumen, such as the P-glycoprotein that can limit the
absorption of many drugs).
• Metabolized by the liver before reaching the systemic circulation (so-called pre-
systemic or ‘first-pass’ metabolism). First-pass metabolism can be avoided by
the intravascular, intramuscular or sublingual routes.

 Box 2.2
Routes of drug administration

Route Description Example


Buccal Rapid absorption, avoids first-pass metabolism Midazolam for status epilepticus
Epidural Used for analgesia Epidural anaesthetic for pain relief during
labour
Inhaled Achieves high concentrations at site of disease with limited Salbutamol for acute asthma relief
systemic absorption
Intramus Easier than intravenous route but absorption is less Adrenaline (epinephrine) in anaphylaxis
cular predictable
Intranas Often used for local delivery Intranasal corticosteroids in allergic
al conditions
Intrathec Used to achieve high local levels Methotrexate in acute lymphoblastic
al leukaemia
Intraven 100% of dose enters systemic circulation but invasive and Intravenous teicoplanin for treatment of
ous inconvenient invasive MRSA infections
Oral Most common route, convenient, but subject to first-pass Ramipril for hypertension
metabolism for certain drugs
Subcuta Good absorption, can be self-administered Insulin for diabetes
neous
Sublingu Rapid absorption, avoids first-pass metabolism GTN for acute relief of angina chest pain
al
Topical Used for local application of cutaneous or joint diseases Corticosteroids for eczema
Transder Absorption through the skin via a patch Nicotine replacement therapy for smoking
mal cessation

GTN, glyceryl trinitrate; MRSA, meticillin-resistant Staphylococcus aureus


 Box 2.3
Key pharmacokinetic parameters
• Area under the curve: A plasma concentration time curve and a measure of
the actual body exposure to a drug after administration of a dose of the drug;
expressed in mg/L x h.
• Bioavailability: The fraction or proportion of a drug that enters the systemic
circulation. Thus, while intravenous formulations have 100% bioavailability,
administration via other routes will have <100% bioavailability.
• Clearance: Volume of plasma cleared of the drug per unit time.
• Cmax: Peak plasma concentration after drug administration.
• Concentration: Amount of drug in a given volume. Even when the dose is
similar, concentrations (and thus exposure) will vary between individuals.
• Dose: Amount of drug administered. This may be micrograms, milligrams or
grams. Occasionally, drugs are administered per body weight or body
surface area.
• Elimination half-life: The time required for the drug concentration to reach
half its original value.
• Tmax: Time to reach Cmax.
• Vd: Volume of distribution – the apparent volume in which a drug is
distributed. Drugs with low Vd are predominantly in the circulation, while
drugs with a high Vd have diffused into the peripheral tissue compartment.

Distribution
This is the process by which drugs are distributed from the bloodstream to
organs and cells. Most drugs will circulate in the bloodstream bound to plasma
proteins, most commonly albumin. Free drug, which is the active drug, is usually
in dynamic equilibrium with protein-bound drug, with the latter being released
as the free drug concentration decreases.

Metabolism
Metabolism takes place in many different organs, the liver being the most
common site. The function of metabolism is to convert lipid-soluble drugs to
water-soluble drugs that can then be excreted by the kidneys. Metabolism
usually inactivates a drug, but in some cases, it can convert a prodrug into an
active drug (e.g. enalapril to enalaprilat), or lead to the formation of toxic
metabolites (e.g. the formation of N-acetyl-para-benzoquinoneimine in
paracetamol overdose). Drug metabolism occurs in two stages:
• Phase I is the modification of a drug, by oxidation, reduction or hydrolysis. Of
these, oxidation is the most frequent route and is largely undertaken by a family
of isoenzymes known as the cytochrome P450 system. Inhibition or induction
of cytochrome P450 isoenzymes is a major cause of drug interactions (see p.
24).
• Phase II involves conjugation with glucuronate, sulphate, acetate or other
substances to render the drug more water-soluble and therefore able to be
excreted in the urine.

Excretion
Excretion of drugs and their metabolites occurs most commonly via the kidneys,
although faeces, breath and sweat represent other routes of excretion. In the
kidney, free drug is filtered through the glomerulus and excreted in the urine as
long as it is water-soluble. Lipid-soluble drugs are usually reabsorbed in the
renal tubules. Some drugs, such as benzylpenicillin, are actively excreted by
cells in the proximal convoluted tubule. Renal impairment reduces the
elimination of water-soluble drugs, and thus reduction of the maintenance dose
must be undertaken to avoid toxicity.

Pharmacodynamics
Pharmacodynamics is the study of what the drug does to the body.
Pharmacodynamic sources of variation in drug action can be due to changes in
the expression of the drug target, their affinity or their selectivity. This can be
caused by genetic factors (see p. 25) and also by disease. For example, in renal
impairment, pharmacodynamic factors determine the sensitivity to a drug.
Pharmacodynamic tests can be used in clinical practice to target therapy or to
avoid undue sensitivity. Examples include the following:
• Expression of oestrogen and HER2 receptors in women with breast cancer are
used as markers of responsiveness to anti-oestrogens and trastuzumab
(respectively).
• The V600E BRAF somatic gene mutation in patients with malignant melanoma
is used to determine whether vemurafenib will be effective (see p. 25).
• Pseudocholinesterase (also known as butyrylcholinesterase) inactivates the
neuromuscular blocking agent suxamethonium. Patients with mutations in the
pseudocholinesterase gene can develop prolonged paralysis after the
administration of suxamethonium.
Pharmacodynamic sources of variation in drug response have not been as well
studied as pharmacokinetic factors. The prospect for ‘personalized prescribing’
will be enhanced when the interplay between pharmacokinetics and
pharmacodynamics is better understood, which will then permit drug selection
and dosing to become much more precise.

Further reading
Mcleod HL. Cancer pharmacogenomics: early promise, but concerted effort
needed. Science 2013; 339:1563–1566.

Affordability and Cost-Effectiveness


All healthcare systems try to provide their populations with the highest standards
of care within the resources they have at their disposal. The money countries
spend on healthcare, however, varies widely (Fig. 2.2). The differences between
countries' expenditure on healthcare are almost entirely determined by their
wealth, as reflected in per capita gross domestic product (GDP), with richer
countries spending more on healthcare than poorer ones. Consequently, an
intervention that might be affordable in the USA, Norway or Sweden is not
necessarily affordable in China, India or Indonesia. Increasing use is therefore
being made of formal analyses of costs and benefits, by both developed and, to a
lesser extent, lower- to middle-income countries, to help decide on healthcare
priorities.
FIGURE 2.2 Total healthcare expenditure (US$) in various countries, as US$ per head
of the population (data for 2009).

There are a number of different approaches to the economic evaluation of


healthcare interventions but the main one is known as ‘cost–utility analysis’. The
main features are as follows:
• The costs are reflected in the price of the product, together with additional sums
associated with its administration (including extra costs of any associated
hospital care, and additional investigations, for the care of patients with adverse
drug reactions). Any savings that may accrue from the use of an intervention
are also included in the calculation.
• The benefits are assessed from the increase in quality of life (sometimes known
as the gain in health utility) as a result of using an intervention, in comparison
with alternative forms of treatment (which, in the absence of an active
comparator, may be ‘best supportive care’). The utility gain is expressed on a
scale of 0 (dead) to 1 (perfect health). This utility gain is then multiplied by the
number of years for which it is ‘enjoyed’, to provide the quality-adjusted life
years (QALY) gained.
• The incremental cost-effectiveness ratio (ICER) is estimated by dividing the
total net costs (which includes any savings) by the QALY gained. In England,
ICERs of an intervention costing less than £20 000 per QALY gained are likely
to be regarded as cost-effective; as the ICER rises, so does the likelihood of the
intervention being rejected for use in the National Health Service on grounds of
cost-ineffectiveness.
The virtue of cost–utility analysis, in the evaluation of both new and
established products, is that it allows comparisons to be made between differing
treatments for differing conditions. Examples of the ICERs of some of the
interventions examined by the National Institute for Health and Care Excellence
(NICE) are shown in Box 2.4.

 Box 2.4
Incremental cost-effectiveness ratios (ICERs),
expressed as £ per quality-adjusted life years
(QALY) gained, for selected pharmaceutical products

Most plausible
Product Indication Comparator
ICERa

Alimtuzum Relapsing–remitting multiple sclerosis Glatiramer acetate 26 000


ab
Alteplase Acute ischaemic stoke (0–4.5 h post onset) Best supportive care 4500
Apixaban Prevention of stroke and systemic embolism in atrial Warfarin 13 000
fibrillation
Dabigatra Prevention of stoke and systemic embolism in atrial Warfarin 18 000
n fibrillation
Imatinib Chronic myeloid leukaemia (chronic phase) Interferon-alpha 26 000
Rituximab Chronic lymphatic leukaemia (first-line) Fludarabine +  22 500
cyclophosphamide
Rivaroxab Prevention of stroke and systemic embolism in atrial Warfarin 18 800
an fibrillation
Trastuzum Early-stage breast cancer (after surgery and Best supportive care 18 000
ab chemotherapy)
a
Based on estimates published by the National Institute for Health and Care Excellence (NICE).

Prescribing in Special Populations


Neonates, infants, children and adolescents
The use of drugs in young people poses special problems. Extrapolating from
adult dosage regimens, merely adjusting for weight, leads to excessive (and
potentially toxic) doses because:
• the rates of hepatic metabolism and renal excretion of drugs are reduced in
neonates and infants
• premature babies have approximately 1% of their body weight as fat (compared
with 20% in adults), leading to a marked increase in plasma drug levels of fat-
soluble drugs.
There are other difficulties in prescribing for children:
• Many treatments have never been subject to formal trials in either children or
adolescents and their benefits and risks have not, therefore, been appropriately
assessed in these age groups. In order to redress this, regulatory agencies, such
as the European Medicines Agency, now require manufacturers to provide
Paediatric Investigation Plans if a new drug is licensed for adults only.
• For many drugs, there are no paediatric preparations or formulations. Instead,
adult products are used.
• Precise oral dosing is often impossible in babies, who spit out unpleasant-
tasting products!
• Adverse effect profiles of medicines may be different in children compared
with adults (e.g. Reye syndrome in children given aspirin, suicidal ideas in
depressed adolescents treated with selective serotonin reuptake inhibitors).

The elderly
The demography of our population is changing with the numbers of elderly
people increasing. It is therefore essential to be aware of the potential issues that
arise in prescribing for the elderly:
• Renal function declines with age at the rate of approximately 1 mL/min after
the age of 40 years. Extrapolation of drug dosages, from those appropriate in
younger adults, may therefore lead to toxic plasma levels in older patients.
• Changes in drug distribution due to changes in body composition, and the
preferential distribution of the cardiac output to the brain, may also predispose
to toxicity.
• Co-morbidity, often associated with polypharmacy, leads to increased
opportunities for disease–drug and drug–drug interactions.
• Concordance with treatment regimens diminishes as the number of prescribed
drugs increases, and is especially poor in the face of cognitive impairment.
• Exaggerated pharmacodynamic effects of drugs acting on the central nervous,
cardiovascular and gastrointestinal systems are common.
Examples of common problems encountered in the use of drugs amongst older
people are shown in Box 2.5.
 Box 2.5
Common adverse reactions to drugs in the elderly

Drug or drug class Effect


Beta-blockers (including eye drops) Bradycardia
Digoxin
Nitrates Postural hypotension
α-Adrenoceptor-blockers
Loop diuretics
Diuretics (thiazides) Glucose intolerance, gout
Antimuscarinic drugs Confusion, cognitive dysfunction
Tricyclic antidepressants
Neuroleptics
Anticonvulsants
Hypnotics
Opioids
Bisphosphonates (mainly alendronic acid) Oesophageal ulceration and stricture formation
Non-steroidal anti-inflammatory drugs Gastric erosions
(NSAIDs) Small bowel and colonic lesions (less frequently than
gastric)
Upper gastrointestinal bleeding
Perforated peptic ulcer
Salt and water retention
Renal impairment

Pregnant women
Clinicians should be extremely cautious about prescribing drugs to pregnant
women and only essential treatments should be given. When a known teratogen
is needed during pregnancy (e.g. an anticonvulsant drug or lithium), the potential
adverse effects should be discussed with the parents, preferably before
conception. Teratogens are best known for causing structural malformations: for
example, congenital heart disease. However, teratogens can also lead to
neurodevelopmental problems without any obvious structural damage. For
example, exposure to sodium valproate in utero can lead to developmental
neurotoxicity, which is manifested as cognitive and behavioural problems. Some
known human teratogens are shown in Box 2.6.

 Box 2.6
Some human teratogens

Drug or drug class Effect


ACE inhibitors/antagonists Oligohydramnios
Retinoids, e.g. isotretinoin Multiple abnormalities
Carbimazole Neonatal hypothyroidism
Abnormalities of bone growth
Fluconazole Fallot's tetralogy
Antiepileptics: carbamazepine, lamotrigine, phenytoin, Cleft palate, facial abnormalities, neural tube
valproate defect
Neural tube defects, neurodevelopmental delay
NSAIDs Delayed closure of the ductus arteriosus
Cytotoxic drugs Most are presumed teratogens
Lithium Ebstein's anomaly
Misoprostol Moebius syndrome (rare neurological condition)
Thalidomide (and possibly lenalidomide) Phocomelia

Note: All drugs should be avoided in pregnancy unless benefit clearly outweighs the risk.

Breast-feeding women
Although most drugs can be detected in breast milk, the quantity is generally
small. This is because, for most drugs, the concentration in milk is in equilibrium
with plasma water (i.e. the non-protein-bound fraction). A few drugs (e.g.
aspirin, carbimazole, methotrexate) may, however, cause harm to the infant if
ingested in breast milk. The relevant drug literature should be consulted when
prescribing for nursing mothers.

Monitoring Drug Therapy


The combination of pharmacokinetic and pharmacodynamic causes of variability
makes monitoring of the effects of treatment essential. Three approaches are
used.

Pre-treatment dose selection


In patients who have known, or suspected, impaired renal function, it is usually
possible to predict dose requirements from their estimated glomerular filtration
rate (eGFR), usually based on serum creatinine levels. If treatment needs to be
started before the eGFR is available, in patients who have very advanced renal
impairment or if renal function is fluctuating, then treatment can be started with
conventional doses but the prescriber should be prepared to make dose
adjustments within 24 hours.

Measuring plasma drug concentrations


For a few drugs, dosages can be effectively monitored by reference to their
plasma concentrations (Box 2.7). This technique is only useful, however, if both
of the following criteria are fulfilled:
• There is a reliable and available drug assay.
• Plasma concentrations correlate well with both therapeutic efficacy and
toxicity.

 Box 2.7
Drugs for which therapeutic drug monitoring is used

Drug Therapeutic plasma concentration range Toxic level Optimum post-dose sampling time (h)

Carbamazepin 20–50 µmol/L >50 µmol/L >8


e
Ciclosporin 50–200 µg/L >200 µg/L Pre-dose
Digoxin 1.3–2.6 nmol/L >2.6 nmol/L >8
Gentamicin Trough <2 mg/L >14 mg/L Pre-dose
Peak 5–10 mg/L >12 mg/L >1
Lithium 0.6–1.0 mmol/L >1.5 mmol/L >10
Phenytoin 40–80 µmol/L >80 µmol/L >10
Theophylline 55–110 µmol/L >110 µmol/L >4
Vancomycin 15–20 mg/L – Pre-dose

Measuring drug effects


For many drugs, dosage adjustments are made in line with patients' responses.
Monitoring can involve dose titration against a therapeutic end-point or a toxic
effect. Objective measures (such as monitoring antihypertensive therapy by
measuring blood pressure, or cytotoxic therapy with serial white blood cell
counts) are most helpful, but subjective ones are necessary in many instances (as
with antipsychotic therapy in people with schizophrenia).
Adverse Drug Reactions
Adverse drug reactions (ADRs), defined as ‘the unwanted effects of drugs
occurring under normal conditions of use’, are a significant cause of morbidity
and mortality. The burden of ADRs on hospitals is huge in both adults and
children (Box 2.8). ADRs also account for about 2.5% of emergency room visits,
and about 1 in 4 patients in primary care will have an adverse reaction, but
fortunately these are mild in most cases. Unwanted effects of drugs are 5–6
times more likely in the elderly than in young adults, and the risk of an ADR
rises sharply with the number of drugs administered. ADRs are also more
common in women than in men.

 Box 2.8
Burden of adverse drug reactions (ADRs) in
hospitals

Study Setting Number of patients studied Frequency of ADRs

Adults
Davies et ala Patients in hospital 3695 14.7%
Children
Gallagher et alb Patients requiring hospital admission 8345 2.9%
c
Thiesen et al Inpatients 16 601 17.7%
a
Davies EC, Green CF, Taylor S et al. Adverse drug reactions in hospital in-patients: a
prospective analysis of 3695 patient episodes. PLoS One 2009; 4:e4439.
b
Gallagher RM, Mason JR, Bird KA et al. Adverse drug reactions causing admission to a
paediatric hospital. PLoS One 2012; 7:e50127.
c
Thiesen S, Conroy EJ, Bellis JR et al. Incidence, characteristics and risk factors of adverse
drug reactions in hospitalized children – a prospective observational cohort study of 6,601
admissions. BMC Medicine 2013; 11:237.

Classification
Two types of ADR are recognized.

Type A (augmented) reactions (Box 2.9)


These:
• are qualitatively normal, but quantitatively abnormal, manifestations of a drug's
pharmacological or toxicological properties
• are predictable from its known pharmacological or toxicological actions
• generally show a clear dose–response relationship
• are usually common
• are only occasionally serious.

 Box 2.9
Examples of adverse drug reactions

Type of reaction and drug (or drug class) Adverse reaction

Type A (augmented)
ACE inhibitors, e.g. ramipril Hypotension
Chronic cough
Angiotensin receptor antagonists, e.g. losartan Hypotension
Anticoagulants, e.g. warfarin Gastrointestinal bleeding
Intracerebral haemorrhage
Antipsychotics, e.g. haloperidol Acute dystonia/dyskinesia
Parkinsonian symptoms
Tardive dyskinesia
Cytotoxic agents, e.g. 5-fluorouracil Bone marrow dyscrasias
Cancer
Diuretics, e.g. furosemide Dehydration, renal impairment
Glucocorticosteroids, e.g. prednisolone Hypoadrenalism
Osteoporosis
Insulin Hypoglycaemia
Tricyclic antidepressants, e.g. amitriptyline Dry mouth
Type B (idiosyncratic)
Benzylpenicillin Anaphylaxis
Radiological contrast media
Broad-spectrum penicillins. e.g. co-amoxiclav Maculopapular rash
Carbamazepine Toxic epidermal necrolysis
Lamotrigine Stevens–Johnson syndrome
Phenytoin
Sulphonamides
Carbamazepine Hepatotoxicity
Diclofenac
Isoniazid
Phenytoin
Rifampicin
Isotretinoin Depression
SSRIsa, e.g. fluoxetine Suicidal ideation
a
In children and adolescents.
ACE, angiotensin-converting enzyme; SSRIs, selective serotonin reuptake inhibitors.

Whilst some reactions, such as hypotension with angiotensin-converting


enzyme (ACE) inhibitors, may occur after a single dose, others may develop
only after months (pulmonary fibrosis with amiodarone) or years (second
malignancies with anticancer drugs).

Type B (idiosyncratic) reactions (see Box 2.9)


These have no resemblance to the recognized pharmacological or toxicological
effects of the drug. They:
• are qualitatively abnormal responses to the drug
• are unpredictable from its known pharmacological or toxicological actions
• show a dose–response relationship that is complex and not easily discernible
• are usually rare
• are often serious.

Diagnosis
All ADRs mimic some naturally occurring disease, and the distinction between
an iatrogenic aetiology and an event unrelated to the drug is often difficult. The
patient should be asked not only about prescription drugs but also about over-
the-counter medicines, herbal medicines and illicit drugs. Although some effects
are obviously iatrogenic (e.g. acute anaphylaxis occurring a few minutes after
intravenous penicillin), many are less so.
Causality assessment of a suspected ADR can be difficult, particularly in
patients on multiple drugs. The following simple rules (remembered by the
mnemonic TREND) can help in assessing causality:
• Temporal relationship. The time interval between the administration of a drug
and the suspected adverse reaction should be appropriate. Acute anaphylaxis
usually occurs within a few minutes of administration, whilst aplastic anaemia
will only become apparent after a few weeks (because of the lifespan of
erythrocytes). Drug-induced malignancy, however, will take years to develop.
• Rechallenge. Re-institution of the suspected drug, to see if the event recurs, is
often regarded as an absolute diagnostic test. This is, in many instances, correct
but there are two caveats. First, it is rarely justifiable to subject a patient to
further hazard. Second, some ADRs develop because of particular
circumstances, which may not necessarily be replicated on rechallenge (e.g.
hypoglycaemia with an antidiabetic agent).
• Exclusion of other causes. An ADR is a diagnosis of exclusion, and other
causes should be excluded by taking a careful clinical history and undertaking
relevant investigations. For example, in drug-induced liver injury, it may be
necessary to exclude viral, autoimmune and metabolic causes by blood tests,
and cholelithiasis by ultrasound. In a few instances, the diagnosis of an adverse
reaction can be inferred from the plasma concentration (see Box 2.7 ).
• Nature and novelty of the reaction. Many ADRs have been reported before,
which can help in assessing causality. Some conditions (e.g. angio-oedema or
toxic epidermal necrolysis) are so typically iatrogenic that an ADR is very
likely. Where an event is a manifestation of the known pharmacological
property of the drug, it can be recognized as a type A ADR (e.g. hypotension
with an antihypertensive agent, or hypoglycaemia with an antidiabetic drug).
• Dechallenge. Failure of remission when the drug is withdrawn (i.e.
‘dechallenge’) is unlikely to be an ADR. The diagnostic reliability of
dechallenge is not, however, absolute; if the ADR has caused irreversible organ
damage (e.g. malignancy), then dechallenge will result in a false-negative
response.

Management
As a general rule, type A reactions can usually be managed by a reduction in
dosage, whilst type B reactions almost invariably require the drug to be
withdrawn (and never re-instituted). Most ADRs resolve spontaneously when the
drug is withdrawn, but specific therapy is sometimes required for ADRs such as
bleeding with warfarin (vitamin K), acute dystonias (benztropine) or acute
anaphylaxis (see Emergency Box 8.24 ). Desensitization is sometimes carried
out for immune-mediated reactions but this should be done by a specialist with
appropriate resuscitation facilities.
When a patient suffers an ADR, this should be reported to the regulatory
agency: for example, through the UK's yellow card reporting scheme to the
Medicines Healthcare Products Regulatory Agency. The criteria for reporting
vary across the world but, in general, all ADRs to new drugs should be reported,
while only the serious ADRs to established drugs (i.e. those that have been on
the market for more than 2 years) should be reported.

Further reading
Pirmohamed M. Anticipating, investigating and managing the adverse effects
of drugs. Clin Med 2005; 5:23–26.

Drug Interactions
Drugs can interact with each other (drug–drug interactions; DDIs), with food
(drug–food interactions) and with herbal medicines (drug–herbal interactions)
(Box 2.10). These interactions, in general, can be either pharmacokinetic
(affecting the processes of absorption, distribution, metabolism and excretion),
pharmacodynamic (synergistic or antagonistic) or mixed. Approximately 1% of
all ADRs that lead to hospital admission are due to DDIs. A major risk factor for
DDIs is polypharmacy, which is becoming more common because of the co-
morbidities associated with an ageing population. About 57% of patients above
the age of 65 years are on more than five drugs, with 12% being on more than
ten drugs. The use of multiple drugs in a patient is necessary in some cases, but
regular review of medicines is essential in order to optimize doses and stop those
drugs that are no longer required. Prescribing additional drugs in a patient
already on medicines should only be done after a careful history, if it is clinically
indicated, and with adjustment of dose through regular monitoring (clinical
and/or laboratory, as appropriate).

 Box 2.10
Some examples of drug interactions

Implicated Affected
Mechanism Clinical effect
drug(s) drug(s)

Pharmacokinetic: absorption
Colestyramine Digoxin Binding of colestyramine to Reduced digoxin plasma concentration and
digoxin reduced efficacy
Pharmacokinetic: enzyme induction
Carbamazepine Oral CYP3A4 induction Contraceptive failure
Phenytoin contracept CYP2C9 induction Thrombosis
Phenobarbital ive CYP3A4 and P-glycoprotein Allograft rejection
Rifampicin Warfarin induction
St John's Wort Ciclosporin
Pharmacokinetic: enzyme inhibition
Amiodarone Warfarin CYP2C9 inhibition Haemorrhage
Cimetidine Warfarin CYP2C9 inhibition Haemorrhage
Ciprofloxacin Theophylline CYP1A2 inhibition Arrhythmias
Grapefruit juice Simvastatin CYP3A4 inhibition Myopathy
Itraconazole Verapamil CYP3A4 inhibition Bradycardia
Ritonavir Fluticasone CYP3A4 inhibition Cushing syndrome
Pharmacokinetic: excretion
Naproxen Methotrexate Inhibition of methotrexate Methotrexate toxicity
excretion
Bendroflumethi Lithium Inhibition of lithium excretion Lithium toxicity
azide
Pharmacodynamic: synergistic interaction
Verapamil Atenolol Atrioventricular node block Bradycardia, heart block
Ramipril Spironolactone Reduced excretion of Hyperkalaemia
potassium
Pharmacodynamic: antagonistic interaction
Salbutamol Atenolol Opposing effects at the β2- Lack of bronchodilatation
receptor
Furosemide Digoxin Hypokalaemia Digoxin toxicity

Food–drug interactions are also becoming increasingly recognized. A


common example is grapefruit juice and drugs metabolized by CYP3A4 (e.g.
simvastatin – see Box 2.10, verapamil and tacrolimus, to name a few).
Grapefruit juice inhibits CYP3A4 in the intestinal wall, increasing the
bioavailability of these drugs and resulting in higher exposure, and toxicity.
Herbal medicines are taken by almost 20% of the population but rarely asked
about when a patient's clinical history is taken. Many different herbal–drug
interactions have been described, the best recognized of which is with St John's
Wort (SJW), which is an inducer of P450 enzymes and drug transporters such as
P-glycoprotein. SJW can lead to increased drug metabolism, reduction in drug
concentration, and failure of therapeutic effect. An example is interaction with
immunosuppressants such as prednisolone and ciclosporin, where the use of
SJW has led to allograft rejection.

Further reading
Bailey DG, Dresser G, Arnold JM. Grapefruit–medication interactions:
forbidden fruit or avoidable consequences? CMAJ 2013; 185:309–316.
Magro L, Moretti U, Leone R. Epidemiology and characteristics of adverse
drug reactions caused by drug–drug interactions. Expert Opin Drug Saf 2012;
11:83–94.

Inter-Individual Variability in Drug Response


Patients vary in the way they respond to drugs, in terms of both efficacy and
safety. In some circumstances, this may be due to clinical and environmental
factors. Age, body weight, renal and hepatic function, concomitant drugs, co-
morbidities, diet, smoking and alcohol are all known to affect the
pharmacokinetics and sometimes the pharmacodynamics of drugs. Genetic
factors affecting both the pharmacokinetics and pharmacodynamics of drugs also
play a role in determining how individuals respond to drugs.

Genetic causes of altered pharmacokinetics


Both pre-systemic hepatic metabolism and the rate of systemic hepatic clearance
may vary markedly between healthy individuals. Variability in the genes that
encode drug-metabolizing enzymes (Box 2.11) is a major determinant of the
inter-individual differences in the therapeutic and adverse responses to drug
treatment. Genetic polymorphisms can affect both phase I and phase II drug-
metabolizing enzymes.

 Box 2.11
Genetic polymorphisms affecting drug response

Organ or system Associated Pharmacokinetic/phar


Drug/drug response phenotype
involved gene/allele macodynamic

Blood
Red blood cells G6PD Primaquine and others Pharmacodynamic
TPMT*2 Azathioprine/6MP-induced neutropenia Pharmacokinetic
Neutrophils UGT1A1*28 Irinotecan-induced neutropenia Pharmacokinetic
Platelets CYP2C19*2 Stent thrombosis Pharmacokinetic
Coagulation CYP2C9*2, *3, Warfarin dose requirement Mixed
VKORC1
Brain and peripheral nervous system
CNS depression CYP2D6*N Codeine-related sedation and respiratory Pharmacokinetic
depression
Anaesthesia Butyrylcholines Prolonged apnoea Pharmacokinetic
terase
Peripheral nerves NAT-2 Isoniazid-induced peripheral neuropathy Pharmacokinetic
Drug hypersensitivity and liver injury
HLA-B*57:01 Abacavir hypersensitivity Pharmacodynamic
HLA-B*15:02 Carbamazepine-induced Stevens–Johnson Pharmacodynamic
syndrome (in some Asian groups)
HLA-A*31:01 Carbamazepine-induced hypersensitivity in Pharmacodynamic
Caucasians and Japanese
HLA-B*58:01 Allopurinol-induced serious cutaneous reactions Pharmacodynamic
HLA-B*57:01 Flucloxacillin hepatotoxicity Pharmacodynamic
Malignancy
Breast cancer CYP2D6 Response to tamoxifen Pharmacokinetic
Chronic myeloid BCR-ABL Imatinib and other tyrosine kinase inhibitors Pharmacokinetic
leukaemia
Colon cancer KRAS Cetuximab efficacy Pharmacokinetic
Gastrointestinal c-kit Imatinib efficacy Pharmacokinetic
stromal tumours
Lung cancer EGFR Gefitinib efficacy Pharmacokinetic
EML4-ALK Crizotinib efficacy Pharmacokinetic
Malignant BRAF V600E Vemurafenib efficacy Pharmacodynamic
melanoma
Muscle
General anaesthetics Ryanodine Malignant hyperthermia Pharmacodynamic
receptor
Statins SLCO1B1 Myopathy/rhabdomyolysis Pharmacokinetic

(Adapted from Pirmohamed M. Pharmacogenetics: past, present and future. Drug Discov Today
2011; 16:852–861.)

Approximately 25% of all medicines currently in use are substrates for


CYP2D6. The frequencies of the variant alleles show racial variation and a small
proportion of individuals may have three or more copies of the active gene. The
phenotypic consequences of the defective CYP2D6 include the increased risk of
toxicity with those antidepressants or antipsychotics that undergo metabolism by
this pathway. Additionally, reduced conversion of a prodrug to the active
metabolite may also compromise efficacy in some patients; for example, poor
metabolizers of CYP2D6 may be at increased risk of breast cancer relapse with
tamoxifen. Conversely, in individuals with multiple copies of the active gene
(ultra-rapid metabolizers), there are extremely rapid rates of metabolism and
therapeutic failure at conventional doses. Ultra-rapid metabolizers may also be at
increased risk of toxicity with the conversion of prodrugs to active metabolites;
this has been shown with codeine, which is metabolized to morphine and has
caused respiratory depression, particularly in children.
Warfarin is predominantly metabolized by CYP2C9. In most populations,
between 2% and 10% are homozygous for an allele that results in low enzyme
activity. Such individuals will therefore metabolize warfarin more slowly,
leading to higher plasma levels, a greater risk of bleeding, and a requirement for
lower doses if the international normalized ratio (INR) is to be maintained within
the therapeutic range. When combined with the genetic polymorphisms in the
vitamin K epoxide reductase complex (VKORC1) genes (a pharmacodynamic
variation), and age and body mass index (BMI), over 50% of the variation in
individual daily dose requirement can be predicted. Pre-prescription genotyping
for CYP2C9 and VKORC1 has been shown to improve anticoagulation control.
Individual differences in the activity of thiopurine methyltransferase (TPMT),
a phase II enzyme, are used to determine the appropriate doses of
mercaptopurine and azathioprine. Testing for TMPT activity is therefore
undertaken routinely in children undergoing treatment for acute lymphatic
leukaemia and people with Crohn's disease. Slow acetylators (deficient in N-
acetyltransferase type 2) have an increased risk of hepatotoxicity with isoniazid,
while individuals with deficient UGT1A1 activity are at increased risk of toxi​city
from irinotecan.

Genetic causes of altered pharmacodynamics


Genetic variation also affects pharmacodynamic targets, which can, in turn, lead
to variation in both the efficacy and the safety of drugs. For example, variation
in the β2-adrenoceptor gene can affect response to salbutamol (albuterol).
Glucose-6-phosphate dehydrogenase deficiency, the most common enzyme
deficiency in the world, can predispose patients to acute red cell haemolysis with
certain drugs such as primaquine, dapsone, sulphonamides and rasburicase. The
biggest clinical advances in this area have occurred in two therapeutic areas:
• Cancer. Sequencing of the cancer genome has identified novel driver
mutations. This has led to the development of targeted therapies, which have
been remarkably successful in the treatment of some malignancies, even when
they are at an advanced stage (e.g. melanoma, see p. 20).
• Drug safety. Some immune-mediated reactions involving the skin and liver can
now be predicted by genotyping for certain human leucocyte antigen (HLA)
polymorphisms. For example, pre-prescription genotyping for HLA-B*57:01
has been shown to be clinically effective and cost-effective in preventing
hypersensitivity to the anti-human immunodeficiency virus (HIV) drug
abacavir. Similarly, HLA-B*15:02 predisposed patients from South-east Asia to
Stevens–Johnson syndrome with carbamazepine.

Future perspective
As we learn more about the human genome, and the fact that many drugs
undergo metabolism via many pathways and act on a number of targets, most of
which show polymorphic expression, it may be possible in the future to utilize
this information to personalize drug therapy to improve both efficacy and safety
of drugs. This represents one component of the drive towards personalized
medicine.

Further reading
Pirmohamed M. The applications of pharmacogenetics to prescribing: what is
currently known? Clin Med 2009; 9:493–495.
Pirmohamed M, Burnside G, Eriksson N et al. A randomized trial of
genotype-guided dosing of warfarin. N Engl J Med 2013; 369:2294–2303.
Wang L, McLeod JL, Weinshilboum RM. Genomics and drug response. N
Engl J Med 2011; 364:1144–1153.
Woodcock J, Lesko LJ. Pharmacogenetics – tailoring treatment for outliers.
N Engl J Med 2009; 360:811–813.

Evidence-Based Medicine
Clinical practice should, as far as possible, be based on formal evidence of
benefit rather than theoretical speculation, anecdote or authoritative
pronouncements.
One of the main applications of evidence-based medicine has been in
therapeutics. Treatments should only be introduced into routine clinical care if
they have been demonstrated to be effective in appropriate studies. Three
approaches are used:
1. randomized controlled trials
2. controlled observational trials
3. uncontrolled observational studies.

Randomized controlled trials (RCTs)


In this type of study, patients with a particular condition are allocated – at
random – to one of two (and sometimes more) treatments. At the end of the
study, the outcomes in the groups are compared. The purpose of the RCT is to
minimize bias and confounding. In order to minimize patient bias, the
participants themselves are generally unaware of their treatment allocations (a
‘single-blind’ trial); and in order to reduce doctor bias, treatment allocations are
also withheld from the investigators (a ‘double-blind’ trial). To recruit sufficient
numbers of patients, and to examine the effects of treatment in different settings,
it is often necessary to conduct the trial at several locations (a ‘multicentre’ trial).
RCTs may be designed to assess whether one treatment is better than another
(a ‘superiority’ trial), or whether one treatment is similar to another (an
‘equivalence’ trial). There are a number of other variants, including crossover
trials, cluster randomized controlled trials, inferiority trials and futility trials (see
‘Further reading’).

Assessing randomized controlled trials


In assessing the relevance and reliability of an RCT, a number of features in its
design and analysis need to be taken into account.

Randomization
The method of randomization should be robust and, in particular, the investigator
should be unaware of which treatment a patient entering a trial will receive. This
avoids selection bias.

Maintaining blindness
Ideally, in RCTs, neither the investigator nor the patient should be aware of the
treatment allocations until the end of the study. This, though, is not always
possible. Adverse drug reactions, for example, may make it obvious which
treatment a patient has been given. Nevertheless, maintaining ‘blindness’ is
necessary where the outcome is subjective (e.g. relief of pain, alleviation of
depression) if bias is to be avoided.

Were the treatment groups comparable?


The key questions to ask are:
• Were the patients in each group similar in their ‘baseline’ characteristics?
• Were they, for example, of similar age, with similar severity and duration of
illness?
• If not, are the differences likely to influence the results?
• Has the statistical analysis (using analysis of co-variance or Cox's proportional
hazards model – see below) tried to adjust for them?
Box 2.12 shows some of the baseline characteristics of a trial comparing
prednisolone with placebo in the treatment of Bell's palsy.

 Box 2.12
Summary of a multicentre randomized placebo-
controlled trial of prednisolone (25 mg twice daily,
for 10 days) in the treatment of Bell's palsy

Placebo (n = 245) Prednisolone (n = 251)

n (%) n (%)

Baseline characteristics
Gender
 Male 118 48.2 135 53.8
 Female 127 51.8 116 46.2
Age (years) (mean ± SD) 44.9 ± 16.6 43.2 ± 16.2
a
Score on House–Brackmann scale (mean ± SD) 3.8 ± 1.3 3.5 ± 1.2
Time between onset of symptoms and start of treatment
 Within 24 h 147 60.0 120 47.8
 >24 to <48 h 64 26.1 95 37.8
 >48 to <72 h 18 7.3 25 10.0
 Unknown (but <72 h) 16 6.5 11 4.4
Results at 3 and 12 months
Grade 1 on House–Brackmann scalea
 At 3 monthsb 152/239 63.1 205/247 83.0
b
 At 12 months 200/245 81.6 237/251 94.4
a
The House–Brackmann scale scores facial nerve function from 1 = normal to 6 = complete
paralysis.
b
p <0.001.
SD, standard deviation.
(From Sullivan FM, Swan IRC, Donnan PT et al. Early treatment with prednisolone or acyclovir
in Bell's palsy. N Engl J Med 2007; 357:1598–1607, with permission.)

Outcomes
There are two ways to analyse the outcomes of an RCT:
• Per protocol analysis: this includes only those who completed the study.
• Intention-to-treat analysis: this includes all patients from the time of
randomization.
Ideally, there should be no difference in the results of these two types of
analysis but, in reality, the results of a per protocol analysis are usually more
advantageous to a treatment than an intention-to-treat analysis. The reason is that
the intention-to-treat analysis will take account of patients who have withdrawn
from the trial because of adverse reactions. It is therefore a much more robust
approach. The results of the intention-to-treat analysis, in the trial of
prednisolone in Bell's palsy, are shown in Box 2.12. The trial results indicate,
with a high probability, that treatment of Bell's palsy with prednisolone will
increase the chances of a full recovery of facial nerve function.

Are the results generalizable?


Were the patients enrolled into the study a reasonable reflection of those likely to
be treated in routine clinical practice? Or were they a selected population that
excluded significant patient groups (such as the elderly)? If the latter is true,
view the results with caution.

Analysis of a superiority trial


The analysis of a superiority trial is based on the premise – the ‘null hypothesis’
– that there is no difference between the treatments. The null hypothesis is
rejected if the probability of the observed result occurring by chance, the p
value, is less than 1 in 20 (i.e. p <0.05). There are three caveats, though:
• Any difference may still be due to chance; and it is often better to await the
results of at least two independent studies before adopting a new treatment.
• A trial may show no ‘statistically significant’ difference, when one in fact
exists, because too few patients have been included. In other words, the trial
lacked sufficient ‘power’. The ‘power’ of a study (the number of patients
needed in each treatment group to detect a pre-defined difference) should have
been defined at the outset. If the study was under-powered, the results of the
study should be interpreted with extreme care.
• A statistically significant difference may not necessarily be clinically relevant.
Scrutiny of the magnitude of the effect, and its 95% confidence intervals (CI),
is a far better guide than the p value.

Effect size.
The results of the well-designed trial in Box 2.12 show, very convincingly, that
the treatment of Bell's palsy with prednisolone increases the chances of complete
recovery of facial nerve function, at 12 months, from 81.6% to 94.4%. This is a
far more convincing description of the benefits of treatment than the p value.

Number needed to treat (NNT).


Another expression of the benefit of a treatment can be derived from the NNT.
This is an estimate of the numbers of patients who need to be treated with a drug
to achieve one positive result. In the study shown in Box 2.12, the NNT to
enable one patient with Bell's palsy to regain normal facial nerve function, after
prednisolone treatment, is 8.

Analysis of an equivalence trial


The aim of an equivalence trial is to determine whether two (or possibly more)
treatments produce similar benefits. During the design of such trials, it is
necessary to decide what difference is unimportant from a clinical perspective
and then to calculate the number of patients needed in order to have an 80% or
90% chance of showing this. In equivalence trials, such power calculations show
that the number of patients required is invariably greater than those needed for
superiority trials. In equivalence studies, the comparator itself must, of course,
already have been shown to be effective.

Meta-analysis
It is possible to summate the results of all the controlled trials that have been
performed in the treatment of a particular condition so as to refine the estimate
of effectiveness. This technique minimizes random error in the assessment of the
effect size of a treatment because more patients are included than could be
accommodated in any single trial. A meta-analysis should be performed (and
interpreted) because of the heterogeneity of the individual studies used in it.
Meta-analyses are most often based on the results of the summary statistics
reported in the original reports. Occasionally (but more often in the past few
years), meta-analyses use the data from the individual patients involved in the
original studies. This technique is time-consuming and requires the original
triallists to agree to providing the necessary data.

Controlled observational trials


Three types of observational study have been used to test the clinical
effectiveness of therapeutic interventions:
• historical controlled trials
• case–control studies
• before-and-after studies.

Historical controlled trials


Despite the value of the prospective RCT, there are many treatments that have
never been subjected to this technique; their efficacy, however, is unquestioned.
Examples include insulin in the treatment of diabetic ketoacidosis, thyroxine for
hypothyroidism, vitamin B12 in pernicious anaemia and defibrillation for
ventricular fibrillation. In a historical controlled trial, the outcome in patients
treated with the study drug is compared to that of previously untreated people
with the same disease. Treatments can be accepted into routine use on the basis
of favourable comparisons with historical controls when the following criteria
are met:
• There should be a biologically plausible basis for the observed benefits.
• There should be no appropriate treatment that could be reasonably used as a
control.
• The condition should have a known and predictable natural history.
• The treatment should not be expected to have adverse effects that would
compromise its potential benefits.
• There should be a reasonable expectation that the magnitude of the therapeutic
effects of treatment will be large enough to make the interpretation of its
benefits unambiguous.

Case–control studies
This type of study design compares people with a particular condition (the
‘cases’) with those without (the ‘controls’). The approach has been used
predominantly to identify epidemiological ‘risk factors’ for specific conditions
such as lung cancer (smoking) or sudden infant death syndrome (lying prone), as
well as in the evaluation of potential ADRs (such as deep venous thrombosis
with oral contraceptives).
A case–control design allows an estimation of the odds ratio (OR), which is
the ratio of the probability of an event occurring to the probability of the event
not occurring (Box 2.13). An OR that is significantly greater than unity indicates
a statistical association that may be causal. The ORs for deep venous thrombosis
and the current use of oral contraceptives varies from 2 to 4 (depending on the
preparation); this indicates that the risk of developing a deep venous thrombosis
on oral contraceptives is between 2 and 4 times greater than the background rate.

 Box 2.13
Estimation of odds ratio

Cases Controls

Risk factor present a c


Risk factor absent b d
The odds ratio (OR) = (a ÷ b) / (c ÷ d)

In some studies, the OR for a particular observation has been found to be


significantly less than unity, suggesting ‘protection’ from the condition under
study. Some studies of women with myocardial infarction indicated protection in
those using hormone-replacement therapies but it has subsequently been shown
that the result was due to bias. On the other hand, case–control studies have
consistently shown that aspirin and other non-steroidal anti-inflammatory drugs
(NSAIDs) are associated with a reduced risk of colon cancer. This seems to be a
causal effect.
Case–control studies claiming to demonstrate the efficacy of a drug need to be
interpreted with great care; the possibility of bias and confounding is substantial,
as in the studies of hormone-replacement therapy and myocardial infarction.
Confirmation from one or more RCT is usually essential.

Before-and-after studies
It has sometimes been inferred that observed improvements seen in patients
before, and after, the use of a particular treatment is evidence of efficacy. Such
an approach is fraught with difficulties; the combination of a placebo effect, as
well as regression to the mean, is likely to negate most studies using this type of
design. Nevertheless, there are some circumstances where genuine efficacy can
be confidently concluded with such designs: the benefits of hip replacement and
cataract surgery are good examples. Such instances can be regarded as special
examples of the use of implicit historical controls.

Uncontrolled observational studies


Uncontrolled case series cannot provide primary evidence for efficacy, unless
they are undertaken in circumstances that are virtually those of historical
controlled trials. Case series can, however, sometimes be of value in
demonstrating the generalizability of the results of RCTs.

Evaluation of new drugs


New drugs are subjected to a vigorous programme of pre-clinical and clinical
testing before they are licensed for general use (Box 2.14). At the time new
products are licensed there can be confidence in their efficacy for their agreed
indications. Conclusions about their safety must, at this point, be provisional,
which is why emphasis is placed on post-marketing.

 Box 2.14
Evaluation of new drugs
Phase I: Healthy human subjects (usually men)
• First use in humans
• Evaluation of safety and toxicity
• Pharmacokinetic assessment
• Sometimes pharmacodynamic assessment
• Approximately 100 subjects
Phase II: First assessment in patients
• Safety and toxicity evaluated
• Dose range identified
• Pharmacokinetic and pharmacodynamic monitoring
• Approximately 500 subjects
Phase III: Use in wider patient population
• Efficacy main objective
• Safety and toxicity also carefully monitored
• Often multicentre trials
• Approximately 2000 patients involved
Phase IV: Post-marketing surveillance
• All patients prescribed drug monitored
• Efficacy, safety and toxicity measured
• Quantification of unusual drug adverse effects
• Yellow Card and spontaneous reports of adverse reactions
• Often very large numbers of patients observed

Statistical Analyses
The relevance of statistics is not confined to those who undertake research but
extends to anyone who wants to understand the relevance of research studies to
their clinical practice.

The average
Clinical studies may describe, quantitatively, the value of a particular variable
(e.g. height, weight, blood pressure, haemoglobin) in a sample from a defined
population. The ‘average’ value (or ‘central tendency’, in statistical language)
can be expressed as the mean, median or mode, depending on the circumstances
(Fig. 2.3).
• The mean is the average of a distribution of values that are grouped
symmetrically around the central tendency.
• The median is the middle value of a sample. It is used, particularly, where the
values in a sample are asymmetrically distributed around the central tendency.
• The mode is the interval in a frequency distribution of values that contains
more values than any other.
FIGURE 2.3 Averages. In symmetrical data, the mean, mode and median are the same.

In a symmetrically distributed population, the mean, median and mode are the
same. The average value of a sample, on its own, is of only modest interest. Of
equal (and often greater) relevance is the confidence we can place on the sample
average as truly reflecting the average value of the population from which it has
been drawn. This is most often expressed as a confidence interval, which
describes the probability of a sample mean being a certain distance from the
population mean. If, for example, the mean systolic blood pressure of 100
undergraduates is 124 mmHg, with a 95% confidence interval of ± 15 mmHg, we
can be confident that if we replicated the study 100 times, the value of the mean
would be within the range 109–139 mmHg on 95 occasions. It is intuitively
obvious that the larger the sample, the smaller will be the size of the confidence
interval.

Correlation
In clinical studies, two or more independent variables may be measured in the
same individuals in a sample population (e.g. weight and blood pressure). The
degree of correlation between the two can be investigated by calculating the
correlation coefficient (often abbreviated to ‘r’).
The correlation coefficient measures the degree of association between the
two variables and may range from 1 to −1:
• If r = 1, there is complete and direct concordance between the two variables.
• If r = −1, there is complete but inverse concordance.
• If r = 0, there is no concordance.
Statistical tables are available to inform investigators as to the probability that
r is due to chance. As in other areas of statistics, if the probability is less than 1
in 20 (p <0.05), then by custom and practice it is regarded as statistically
significant. There are, however, two caveats:
• The 1 in 20 rule is a convention and does not exclude the possibility that a
presumed association is due to chance.
• The fact that there is an association between two variables does not necessarily
mean that it is causal. For example, a correlation between blood pressure and
weight, with r = 0.75 and p <0.05, does not mean that weight has a direct effect
on blood pressure (or vice versa).
Correlation analyses can become complicated. The simplest (least squares
regression analysis) presumes a straight-line relationship between the two
variables. More complicated techniques can be used to estimate r where a non-
linear relationship is presumed (or assumed); where the distributions deviate
from normal; where the scales of one or both variables are intervals or ranks; or
where a correlation between three or more variables is sought.
Survival analyses
In studies where individuals are observed over a long(ish) period of time, and in
which it is unreasonable (or erroneous) to assume that event rates are constant,
the technique of survival analysis is used. This is most commonly reported as the
hazard ratio (HR) with its 95% confidence interval. The HR is the probability
that, if an event in question has not already occurred, it will happen in the next
(short) time interval.

Continuous outcomes
Studies such as that in Box 2.12 may report outcomes using one or more
continuous scales. In this study of the effects of prednisolone in the treatment of
Bell's palsy, the House–Brackmann measure of facial nerve function was used as
the outcome measure. Conventional tests of statistical significance using
Student's t-test, for example, can be calculated to assess whether the null
hypothesis can be rejected.

Number needed to treat (NNT)


As discussed earlier, the NNT is an estimate of the number of patients that need
to be treated for one to benefit compared to no treatment. If the probabilities of
the end-points with the active drug and no treatment (i.e. placebo) are
respectively pactive and pno treatment, then the NNT can be calculated thus:

An analogous measure – the number needed to harm (NNH) – is the number


of patients that need to be treated with a drug to cause one patient to be subject
to a specific harm.

Other statistical techniques


Statisticians have developed a range of sophisticated methods to handle a wide
variety of biomedical problems. Unless an investigator is supremely (and usually
unwisely) confident, it is sensible to seek professional advice in analysing
numerical data that look complicated. In doing so, it is invariably wiser to do so
at the time the study is being designed rather than after the results have been
generated!
Further reading
Jadad AR, Enkin MW. Randomized Controlled Trials: Questions, Answers
and Musings, 2nd edn. London: Blackwell; 2007.
Matthews JN. Introduction to Randomized Controlled Trials, 2nd edn.
London: Chapman & Hall; 2006.
Morris S, Devlin N, Parkin D. Economic Analysis in Health Care.
Chichester: John Wiley & Sons; 2007.
Rawlins MD. Therapeutic Evidence and Decision-Making. London:
Hodder; 2011.

Information Sources
Pharmacotherapy moves at a very rapid pace and it is impossible for anyone to
keep up with contemporary advances. Reliable prescribing advice can be found
in:
• The Summary of Product Characteristics (SmPCs) produced by manufacturers
and vetted by drug regulatory authorities (for the UK, these are the Medicines
and Healthcare Products Regulatory Agency and the European Medicines
Agency).
• The relevant national formulary. Many countries have their own formularies:
for example, the UK has the British National Formulary (BNF), produced
jointly by the British Medical Association and the Royal Pharmaceutical
Society.
• Guidance produced by national health technology agencies such as the UK's
National Institute for Health and Care Excellence (NICE).
• Advice on the management of individual conditions, available in the form of
clinical guidelines (systematically developed statements to assist practitioner
and patient decisions about appropriate healthcare for specific clinical
circumstances). Some of these are published by national medical societies. In
the UK, NICE is the main developer of clinical guidelines.
Patient information leaflets are supplied with all prescribed medication.

Significant websites
http://www.nice.org.uk National Institute for Clinical Excellence (NICE)
clinical guidelines

You might also like