Download as pdf or txt
Download as pdf or txt
You are on page 1of 28

P1: SDA/MKV P2: SDA/VKS QC: SDA

February 18, 1997 11:30 Annual Reviews AR027-11 AR27-011N

Annu. Rev. Pharmacol. Toxicol. 1997. 37:269–96


Copyright c 1997 by Annual Reviews Inc. All rights reserved

MOLECULAR MECHANISMS OF
GENETIC POLYMORPHISMS
OF DRUG METABOLISM
Annu. Rev. Pharmacol. Toxicol. 1997.37:269-296. Downloaded from www.annualreviews.org

Urs A. Meyer
Biozentrum of the University of Basel, CH-4056 Basel, Switzerland
by University of Sussex on 11/05/12. For personal use only.

Ulrich M. Zanger
Dr. Margarete Fischer-Bosch-Institute for Clinical Pharmacology,
D-70376 Stuttgart, Germany

KEY WORDS: Genetic polymorphism, CYP2D6, CYP2C19, N-acetyltransferases, drug meta-


bolism, pharmacogenetics

ABSTRACT
One of the major causes of interindividual variation of drug effects is genetic
variation of drug metabolism. Genetic polymorphisms of drug-metabolizing en-
zymes give rise to distinct subgroups in the population that differ in their ability to
perform certain drug biotransformation reactions. Polymorphisms are generated
by mutations in the genes for these enzymes, which cause decreased, increased, or
absent enzyme expression or activity by multiple molecular mechanisms. More-
over, the variant alleles exist in the population at relatively high frequency. Ge-
netic polymorphisms have been described for most drug metabolizing enzymes.
The molecular mechanisms of three polymorphisms are reviewed here.
The acetylation polymorphism concerns the metabolism of a variety of ary-
lamine and hydrazine drugs, as well as carcinogens by the cytosolic N-acetyltrans-
ferase NAT2. Seven mutations of the NAT2 gene that occur singly or in combi-
nation define numerous alleles associated with decreased function.
The debrisoquine-sparteine polymorphism of drug oxidation affects the meta-
bolism of more than 40 drugs. The poor metabolizer phenotype is caused by
several “loss of function” alleles of the cytochrome P450 CYP2D6 gene. On the
other hand, “ultrarapid” metabolizers are caused by duplication or amplification
of an active CYP2D6 gene. Intermediate metabolizers are often heterozygotes or
carry alleles with mutations that decrease enzyme activity only moderately.
The mephenytoin polymorphism affects the metabolism of mephenytoin and
several other drugs. Two mutant alleles of CYP2C19 have so far been identified
to cause this polymorphism.

269
0362-1642/97/0415-0269$08.00
P1: SDA/MKV P2: SDA/VKS QC: SDA
February 18, 1997 11:30 Annual Reviews AR027-11 AR27-011N

270 MEYER & ZANGER

These polymorphisms show recessive transmission of the poor or slow metab-


olizer phenotype, i.e. two mutant alleles define the genotype in these individuals.
Simple DNA tests based on the primary mutations have been developed to predict
the phenotype. Analysis of allele frequencies in different populations revealed
major differences, thereby tracing the molecular history and evolution of these
polymorphisms.
Annu. Rev. Pharmacol. Toxicol. 1997.37:269-296. Downloaded from www.annualreviews.org

INTRODUCTION
The term genetic polymorphism defines monogenic traits that exist in the normal
population in at least two phenotypes, neither of which is rare (1, 2). Genetic
polymorphisms in drug metabolism are common occurrences. Mutations in the
genes for numerous drug-metabolizing enzymes cause enzyme variants with
by University of Sussex on 11/05/12. For personal use only.

higher or lower activity or lead to the partial or total absence of an enzyme


protein.
Genetic polymorphisms have been described for a wide variety of drug and
xenobiotic metabolizing enzymes. Many of these variations were first iden-
tified by the occurrence of adverse reactions after normal doses of drugs in
patients or volunteers. The association between decreased drug clearance and
decreased activity of a drug-metabolizing enzyme, the inherited nature of the
deficiency, and its frequency and clinical importance was then evaluated. Phe-
notyping methods involving administration of probe drugs and measurement
of metabolites in body fluids or determination of enzyme activity in families
and populations were used for this purpose. During the past 10 years, several
of these polymorphisms have been studied at the protein and gene level, and
simple DNA tests have been developed to predict the phenotype.
In this review, we summarize studies in our laboratory and in others of the
molecular mechanisms of three genetic polymorphisms, the acetylation (or
NAT2) polymorphism, the debrisoquine (or CYP2D6) polymorphism, and the
mephenytoin (or CYP2C19) polymorphism. We restrict our review to these
polymorphisms because their occurrence in different large populations is well
established, they involve clinically important drugs, and they represent the best
studied examples of these variations in regard to mutations of the genes and
their functional and clinical consequences.
In addition to the three polymorphisms described here, genetic polymor-
phisms are known for several other human cytochrome P450 genes that code for
drug-metabolizing enzymes, including CYP1A1, CYP2A6, CYP2C9, CYP2E1,
and for a number of conjugating enzymes, such as glutathion S-transferases,
methyl-transferases, and sulfotransferases. However, the importance of these
polymorphisms for clinically used drugs is less obvious at this time, or their
P1: SDA/MKV P2: SDA/VKS QC: SDA
February 18, 1997 11:30 Annual Reviews AR027-11 AR27-011N

GENETIC POLYMORPHISM 271

genes have not been studied in detail yet. Recent reviews and books have been
published that cover various aspects of genetic polymorphisms and their clinical
or toxicologic relevance (3–11). Because of space limitations, we sometimes
refer to these reviews, rather than citing original articles.

THE N-ACETYLTRANSFERASE POLYMORPHISM


Discovery and Clinical Relevance
Variability in drug acetylation was discovered nearly 40 years ago when the fate
Annu. Rev. Pharmacol. Toxicol. 1997.37:269-296. Downloaded from www.annualreviews.org

of isoniazid was studied in patients who had tuberculosis. Marked interindi-


vidual variation in the elimination of isoniazid was first described in 1953 by
Bönicke & Reif (12). Family studies revealed that this variation was geneti-
cally controlled and that “slow acetylators” of isoniazid were homozygous for
by University of Sussex on 11/05/12. For personal use only.

a recessive gene, whereas “rapid acetylators” were either homozygous or het-


erozygous for the normal or wild-type gene. In the following years, a number
of sulfonamides (sulfadiazine, sulfamethazine, sulfapyridine, sulfameridine,
and sulfadoxine), numerous other drugs (aminoglutethimide, amonafide, am-
rinone, dapsone, dipyrone, endralazine, hydralazine, isoniazid, prizidilol, and
procainamide), and metabolites of clonazepam and caffeine were discovered to
be polymorphically acetylated. Recent reviews have discussed the acetylation
polymorphism (6, 13–18).
The proportions of rapid and slow acetylator phenotypes vary remarkably in
populations of different ethnic or geographic origin. Most populations in Europe
and North America have 40 to 70% slow acetylators, whereas populations
along the Pacific Asian littoral (e.g. Japanese, Chinese, Korean, and Thai) have
only 10 to 30% slow acetylators. The molecular aspect of interethnic and
geographic variation of the acetylation polymorphism is discussed later in this
review.
The association of the acetylation polymorphism with drug toxicity, such
as sulfonamide hypersensitivity reactions (19), and its role as a risk factor for
certain diseases, in particular bladder cancer, has received much attention and
has been extensively studied (6, 13, 20–22). A discussion of these associations
and their presumed mechanisms is beyond the scope of this review.
Molecular Mechanism
STUDIES IN HUMAN LIVER Early studies clearly established that the observed
variation in acetylation of isoniazid was caused by a difference in the activity
of a cytosolic N-acetyltransferase in the liver that involved transfer of acetate
from acetyl coenzyme A (CoASAc) to the substrate (CoASAc; NAT, EC2.3.1.5)
Some arylamines, such as p-aminobenzoate and (PABA) and p-aminosalicylate
(PAS), are equally well acetylated by rapid and slow acetylators and therefore
P1: SDA/MKV P2: SDA/VKS QC: SDA
February 18, 1997 11:30 Annual Reviews AR027-11 AR27-011N

272 MEYER & ZANGER

were called “monomorphic” in contrast to the “polymorphic” substrates iso-


niazid, sulfamethazine, and others (13, 23). Jenne (24) performed the first
detailed biochemical study of drug acetylation in liver tissue of rapid and slow
acetylators and suggested that two different enzymes were responsible for the
metabolism of monomorphic and polymorphic arylamines, one being quanti-
tatively decreased in livers of slow acetylators.
Major progress in the purification of a human liver acetyltransferase that
can acetylate the substrate sulfamethazine and the development of polyclonal
antibodies that recognize this protein was achieved in 1989 (25). In liver biopsy
Annu. Rev. Pharmacol. Toxicol. 1997.37:269-296. Downloaded from www.annualreviews.org

specimens from surgical patients who had been phenotyped with the use of
the caffeine test, slow acetylation was associated with markedly reduced but
still detectable levels of immunoreactive NAT protein (26), which suggests a
quantitative rather than a qualitative difference of the human enzyme protein,
and confirms the initial proposal of Jenne (24). Subsequent studies (23) revealed
by University of Sussex on 11/05/12. For personal use only.

that an early unstable peak of NAT-activity on anion-exchange chromatography


with low affinity for sulfamethazine, but high affinity for the monomorphic
substrates PABA and PAS, had been overlooked. This fraction corresponds to
what we now recognize as the distinct but related enzyme NAT1, whereas the
initially purified protein (25) corresponds to the polymorphic NAT2.
In New Zealand white rapid acetylator rabbits, NAT protein and its mRNA
are more abundant. This finding allowed Andres et al (27) to prepare a highly
purified enzyme. Information from peptide sequences of this enzyme were used
by Blum et al (28) to synthesize an oligonucleotide probe and, in combination
with antibodies prepared against purified rabbit NAT, to clone a rabbit cDNA
encoding an enzyme with high specific activity for sulfamethazine acetylation.
Immunochemical studies suggested complete absence of this protein in slow
acetylator rabbits, in contrast to the only partial reduction of NAT2 protein
in livers of human slow acetylators. The absence of the NAT protein in slow
acetylator rabbits has now been explained by a gene deletion mechanism (28).

N-ACETYLTRANSFERASE GENES NAT1 AND NAT2 Human genes for acetyltrans-


ferases were cloned by Blum et al (29), who used the rabbit cDNA probe from
genomic DNA of an obligate heterozygous rapid acetylator individual identi-
fied by a pedigree phenotyping study. The recombinant clones define three
different, but closely related, human NAT genes: NAT1, NAT2, and the pseudo-
gene NATP. According to a recently proposed nomenclature, NAT1 corresponds
to NAT1∗ 3, and NAT2 corresponds to NAT2∗ 4 (18). The product of the NAT2
(NAT2∗ 4) gene expressed in COS cells showed protein immunoreactivity and ki-
netic properties identical to those of human NAT2 (NAT2∗ 4) (23, 26, 29), whose
content in human liver varies with the acetylator phenotype. This established
the NAT2 locus as the site of the classic human acetylation polymorphism. The
P1: SDA/MKV P2: SDA/VKS QC: SDA
February 18, 1997 11:30 Annual Reviews AR027-11 AR27-011N

GENETIC POLYMORPHISM 273

NAT1 (NAT1∗ 3) gene gives rise to a NAT protein with different electrophoretic
and kinetic properties. This protein corresponds to the monomorphic NAT1
(NAT1 3) enzyme with high affinity for substrates such as PABA and PAS (23).
Each of the two genes NAT1 and NAT2 has a single, intronless protein-coding
exon with an open reading frame of 870 bp that produces a 290 amino acid (33–
34 kDa) protein. Whereas the NAT1 gene produces its entire transcript from a
single exon, the NAT2 mRNA is derived from both the protein-coding exon and
a second, untranslated exon of 100 bp that is located about 8 kb upstream of
the translation start site (29–30). The nucleotide homology between NAT1 and
Annu. Rev. Pharmacol. Toxicol. 1997.37:269-296. Downloaded from www.annualreviews.org

NAT2 is 87% in the coding region. There are only 55 amino acid differences,
only 28 of which are nonconservative changes. The homologies to the rabbit
cDNA rnat are 82% for both NAT1 and NAT2 (29). The pseudogene NATP
has high sequence homology to NAT1 and NAT2 (79%) but contains multiple
frameshifts and stop codons. NAT1 and NAT2 were mapped to chromosome
by University of Sussex on 11/05/12. For personal use only.

8p21.3–23.1 (29, 31).


The discovery of 2 separate genes that encode NAT1 and NAT2 has re-
solved the old question of monomorphic and polymorphic substrates (23).
Although NAT2 and NAT1 also have overlapping arylamine substrate speci-
ficities, e.g. for 2-aminofluorene and sulfomethoxazole (32), they also have
unique activities. Only NAT1 can metabolize PABA and the folate catabolite p-
aminobenzoyl-glutamate (33, 34). There is a conserved cysteine residue across
all vertebrate NATs that corresponds to cysteine 68 (18, 35, 36), whereas the
different substrate specificity between NAT1 and NAT2 seems to be related
to differences in the C-terminal portion. Recent studies suggest that NAT1
function also may be genetically variable in human populations (see below).
GENETIC VARIABILITY OF NAT2 At the protein level, the slow acetylator pheno-
type is associated with a reduction of 10 to 20% of the quantity of NAT2 in liver
cytosol (26), whereas the mRNA content remains unchanged (37). Several “de-
creased function alleles” of NAT2 have been identified that cause this decrease
in NAT2 protein and, consequently, the recessively inherited slow acetylator
phenotype (37–41) reviewed by Vatsis et al (18).
Decreased function alleles Mutant alleles of NAT2 were first identified by re-
striction fragment analysis of genomic DNA samples from healthy individuals
whose acetylator phenotype had been established by caffeine phenotyping (37)
or isoniazid (42) and of DNA from human liver samples with known NAT2 en-
zyme activity (37). Restriction fragment length polymorphisms (RFLPs) with
NAT2 probes generated by Mspl, Kpnl, and Bam HI segregated with acetylator
phenotype and defined three mutant alleles of NAT2, initially called M1, M2,
and M3 (37). To explore the molecular mechanism of slow acetylation, NAT2
genes were cloned from genomic libraries (37) or characterized as cDNAs (38),
P1: SDA/MKV P2: SDA/VKS QC: SDA
February 18, 1997 11:30 Annual Reviews AR027-11 AR27-011N

274 MEYER & ZANGER

Table 1 Mutations of the NAT2 gene

Consequence for
Amino acid N-acetylation activity
Mutation change in vivoa in vitrob Reference

G191A R64Q Decreased Decreased a 41, b 140

C282T Silent (Y94) Decreased Normal a 44, b 37, b 39, b 52

T341C I114T Decreased Decreased a 44, b 5, b 52

C481T Silent (L161) Unknown Normal b 37, b 51

G590A R197Q Unknown Decreased b 37, b 52


Annu. Rev. Pharmacol. Toxicol. 1997.37:269-296. Downloaded from www.annualreviews.org

A803G K268R Normal Normal a 45, b 51, b 52

G857A G286E Unknown Controversial b Agundez and Meyer, unpublished;


b 39, b 52

a
The in vivo effect of the mutation is derived from the acetylator phenotype in homozygous carriers of this
mutation. No homozygous carriers have been reported for the C481T, G590A, and G857A mutation.
by University of Sussex on 11/05/12. For personal use only.

b
Data from expression of recombinant NAT2 alleles with site-specific mutations and chimeric genes in various
systems, COS-1 cells (37), CHO cells (39, 51) or in E. coli (52, 140).

and their sequence determined and compared with the wild-type gene sequence.
The M1 allele had mutations at positions 341 and 481 (Tables 1 and 2), M2 at
positions 282 and 590 (37), and M3 at position 857 (38). According to a recent
nomenclature (18), M1 corresponds to NAT2∗ 5A, M2 to NAT2∗ 6A, and M3 to
NAT2∗ 7A (Table 2). Initially, only one mutation was reported at position 857
for NAT2∗ 7A (38). More recent studies now reveal two mutations in most if
not all NAT ∗ 7 alleles, one at position 857, and the other at position 282 (39).
These data were confirmed and extended by Vatsis et al (43) and by Hickman
et al (40), who sequenced amplified DNA from livers and lymphocytes of slow
acetylators by using the sequence information of Blum et al (29). Vatsis and co-
workers (43) observed an allele with mutations at positions 341, 481, and 803
(NAT2∗ 5B). Hickman and co-workers (40) observed an allele with mutations at
positions 341 and 803 (NAT2∗ 5C). Moreover, Bell et al (41) described a muta-
tion at position 191 that occurs almost exclusively in blacks. The mutations of
NAT2 and their functional consequences are summarized in Table 1.
Of the seven point mutations in the coding region, five cause amino acid
changes, and two are silent. Three mutations (G191A, C282T, and T341C),
when present on both alleles (homozygosity), demonstrably cause the slow
acetylator phenotype in vivo (41, 44), whereas homozygous carriers of a sin-
gle mutation A803G are rapid acetylators (45). No homozygous carriers of
mutations C418T, G590A, and G857A have been identified, even in large epi-
demiologic studies totaling more than 2000 DNA samples (44, 46–48).
The common slow acetylator alleles carry multiple mutations, as summa-
rized in Table 2. The NAT2∗ 5 A, B, C alleles share mutation T341C, whereas
P1: SDA/MKV P2: SDA/VKS QC: SDA
February 18, 1997 11:30 Annual Reviews AR027-11 AR27-011N

GENETIC POLYMORPHISM 275

Table 2 Common NAT2 alleles and their frequencies in caucasian populationsa

Mutations Allele designationb Associated phenotype Frequency (whites)

None NAT2∗ 4 (wild-type) Rapid 22.7c /21.6d


T341C, C481T, A803G NAT2∗ 5B (r3 , S1a) Slow 38.3/41.6
C282T, G590A NAT2∗ 6A (M2) Slow 27.8/23.6
T341C, C481T NAT2∗ 5A (M1) Slow 4.2/1.5
T341C, A803G NAT2∗ 5C (S1c) Slow 4.0/0.8
C282T, G857A NAT2∗ 7B (M3) Slow 1.3/1.2
C282T NAT2∗ 13 (S4) Slow 1.5/1.9
Annu. Rev. Pharmacol. Toxicol. 1997.37:269-296. Downloaded from www.annualreviews.org

G590A NAT2∗ 6B (R2) Unknown <0.1/2.0


G191A NAT2∗ 14A (M4) Slow <0.1/0.6e
A803G NAT2∗ 12A Rapid <0.1/2.5
a
Compiled from published genotyping/phenotyping results, the presumed inactivating mutation (Table 1) is
underlined.
by University of Sussex on 11/05/12. For personal use only.

b
The nomenclature of NAT2 alleles is summarized in Ref. 18.
c
From data on 1688 alleles (Germany) (44)
d
From data on 1080 alleles (Spain) (46)
e
Allele 14A is found in frequencies of 7 to 9% in US blacks (41) and 5 and 9% in 2 native African populations
(50).

NAT2∗ 6A, ∗ 7B, and ∗ 13 share mutation C282T. Together they account for more
than 99% of mutant alleles in white slow acetylators (16, 44, 46, 47). Addi-
tional rare alleles are NAT2∗ 12A and the extremely rare NAT2∗ 5D, ∗ 5E, ∗ 12B,

12C, ∗ 14B, ∗ 14C, ∗ 14D, and ∗ 14E (44–48), at least in whites. New point
mutations (A434C, A845C) and additional alleles have been reported (40, 46,
48, 49). The knowledge of practically all mutations of NAT2 that occur with
frequencies of more than 0.1% and the information on their in vivo relevance in
causing the slow acetylator phenotype has led to the development of genotyp-
ing tests of high efficiency and accuracy, either by allele-specific amplification
by polymerase chain reaction (PCR), by restriction analysis, or ideally by a
combination of the 2 procedures (37, 43–47, 49, 50).

Functional characterization of mutant alleles It is not entirely clear yet how


these mutations cause a decreased amount of NAT2 protein in livers of slow
acetylators. Initial studies in liver tissue (37) indicated that the genotype NAT2

5A/ ∗ 5A and ∗ 6A/ ∗ 6A produce normal levels of NAT2 mRNA in the liver,
whereas the quantity of NAT2 immunoreactive protein was markedly reduced
or absent. None of these alleles appears to result in a total absence of NAT2
protein in the liver (26). Expression of wild-type–variant chimeric genes in
COS cells (37) or in CHO cells (51), however, produced controversial results.
In the study of Blum et al (37), both mutations of ∗ 5A were required. In the study
of Abe et al (51), however, the T341C mutation alone reduced enzyme protein.
The mutation G590A causing Arg197Glu in NAT2∗ 6A and the mutation G857A
P1: SDA/MKV P2: SDA/VKS QC: SDA
February 18, 1997 11:30 Annual Reviews AR027-11 AR27-011N

276 MEYER & ZANGER

causing Gly286Glu of NAT2∗ 7B apparently result in proteins with reduced


stability (37, 52). Additional studies by Hein et al (52) with recombinant NAT2
allotypic variants expressed in Escherichia coli provide additional in vitro data.
The expressed products of NAT2∗ 5A, ∗ 5B and ∗ 5C had lower activities than

6A and ∗ 7B, whereas ∗ 13 (C282T) had normal activity. Similarly, in vitro
expressions in mammalian cell systems suggested normal activity of expressed
recombinant NAT2 with C282T (37, 39, 52). These in vitro results are in
contrast to recent in vivo data from comparisons of caffeine metabolic ratios
with genotypes by Cascorbi et al (44). NAT2∗ 13 was clearly a slow acetylator
Annu. Rev. Pharmacol. Toxicol. 1997.37:269-296. Downloaded from www.annualreviews.org

allele in homozygous carriers, and ∗ 5B had an apparently higher activity than



6A. Thus, heterologous expression of NAT2 alleles in E. coli, COS, or CHO
cells apparently does not consistently reflect the effects of mutations on stability
or affinity for substrates of the protein in human liver (Table 1).
by University of Sussex on 11/05/12. For personal use only.

ALLELES OF NAT1 Recent studies by Cribb et al (53), Ward et al (54), and


Vatsis & Weber (55) have provided evidence that NAT1 function is also widely
variable in human populations. PAS acetylation (by urinary metabolic ratio) in
130 individuals was highly variable and suggested a bimodal distribution (56).
The comparison of sequences of NAT1 genes from 13 individuals revealed
two variants: allele V2 (NAT1∗ 10), with two mutations in the 30 UTR, and
allele V3 (NAT1∗ 11), with five point mutations and a nonanucleotide deletion
in the 30 UTR. These alleles were present in 50% of the individuals and showed
Mendelian inheritance in a small family. Sequence differences were also noted
between the wild-type (V1 ) allele and a previously published NAT1 sequence
derived from a Japanese individual (38).
The functional significance of these variants has not been tested. It is there-
fore not known whether the described variants are responsible for the apparent
catalytic differences seen with NAT1 specific substrates. These recent findings,
however, far from being resolved, put the term “monomorphic” for the NAT1
locus in question.

Interethnic Variability and Evolution


The frequency of the slow acetylator phenotype varies considerably among
ethnic groups. Genotyping data of the common NAT2 alleles now allows us to
trace the molecular history of the acetylation polymorphism (Table 3). Two
groups can be distinguished by the incidence of mutations T341C/C481T and
G857A. Whites and Africans are very similar and have high frequencies of
T341C/C481T (>28%) and low frequencies of G857A (≤5%). Japanese,
Chinese, Korean, and other Asian populations have a low incidence of T341C/
C481T mutations (≤7%) and a higher incidence of G857A mutation (10 to
18%). The difference in the incidence of the slow acetylator phenotype be-
P1: SDA/MKV P2: SDA/VKS QC: SDA
February 18, 1997 11:30 Annual Reviews AR027-11 AR27-011N

GENETIC POLYMORPHISM 277

Table 3 Interethnic comparison of NAT2 allelesa

Rapid
No. of acetylator
individuals alleles (%)b Slow acetylator alleles (%)b
Ethnic group analyzed NAT2∗ 4 (wt) NAT2∗ 5A/B/C NAT2∗ 6A/B NAT2∗ 7A/B NAT2∗ 14A/B

Whites, US 421 24 43 31 2 0
Whites, Europe 434 26 46 26 2 0
Whites, Spain 504 22 44 26 1 1
African Americans 214 35 30 23 5 8
Annu. Rev. Pharmacol. Toxicol. 1997.37:269-296. Downloaded from www.annualreviews.org

Native Africans 102 27 40 22 2 9


(Gabonese)
Hispanics 148 40 28 18 14 1
Japanese 224 67 1 22 10 0
Chinese 254 53 5 30 12 0
by University of Sussex on 11/05/12. For personal use only.

Koreans 85 68 2 18 11 1
Filipino 100 40 7 36 18 0
Aborigines (Australia) 49 41 2 17 40 —
Amerindians
Ngawbe 71 74 2 4 24 0
Embera 101 65 10 18 21 0
a
Compiled and modified from Refs. 41, 44, 46–48, 50, 57, 142.
b
The sum of all alleles is not necessarily 100% because rare alleles were not considered in all studies.

tween Whites/Africans (40 to 70%) and Asian populations (10 to 30%) is thus
mostly the result of the low incidence of the T341C/C481T alleles (NAT2∗ 5A,
B, C) in Asians. Amerindians appear to be similar to Asian populations regard-
ing their high proportion of rapid acetylators and the T341C/C481T to G857A
relationship (57). Hispanics appear as intermediate between the White/African
and the Asian groups (47, 48). The data are consistent with the concept that the
acetylation polymorphism has an ancient and common African origin, before
spreading of human populations (58). The practical absence of the NAT2∗ 12
and ∗ 14 alleles (A803G and G191A) in whites as compared with the >8% in-
cidence of G191A in Africans suggests that these mutations occurred after the
African/Non-African split ∼90,000 years ago.

THE DEBRISOQUINE-SPARTEINE POLYMORPHISM


Discovery and Clinical Relevance
In 1977, physicians at St. Mary’s Hospital Medical School in London made the
serendipitous observation that a volunteer’s hypotensive response to debriso-
quine, a sympatholytic antihypertensive drug, was markedly increased because
of impaired metabolism (59). At about the same time, a group of physicians
P1: SDA/MKV P2: SDA/VKS QC: SDA
February 18, 1997 11:30 Annual Reviews AR027-11 AR27-011N

278 MEYER & ZANGER

in Bonn, Germany, independently observed increased side effects associated


with decreased oxidative metabolism of sparteine, an alkaloid drug with antiar-
rhythmic actions (60). Family studies revealed that both oxidative metabolic
reactions are under monogenic control and that poor metabolizers are homozy-
gous for a recessive allele. A simple test, the urinary metabolic ratio (MR), after
administration of a probe drug can identify the extensive metabolizer (EM) and
the poor metabolizer (PM) phenotype. This test, which has been applied to
thousands of individuals, has documented that 5 to 10% of individuals in white
populations are of the PM phenotype, compared with only 1 to 2% in Asian
Annu. Rev. Pharmacol. Toxicol. 1997.37:269-296. Downloaded from www.annualreviews.org

populations (6, 61).


The clinical importance of this polymorphism was initially questioned be-
cause the drugs leading to its discovery were soon either obsolete or of small
distribution. It soon became known, however, that many other drugs are inef-
ficiently metabolized in PM individuals. More than 40 clinically used drugs
by University of Sussex on 11/05/12. For personal use only.

are now identified whose metabolism in vivo cosegregates with that of debriso-
quine and sparteine. They include antiarrhythmics, such as N-propylajmaline,
flecainide, propafenone, and mexiletine; antidepressants, such as amitriptyline,
clomipramine, desipramine, fluoxetine, imipramine, maprotiline, mianserin,
paroxetine, and nortriptyline; neuroleptics, such as haloperidol, perphenazine,
and thioridazine; antianginals, such as perhexiline; opioids, such as dextrome-
torphan and codeine; and amphetamines, such as methamphetamine and
methylenedioxymethamphetamine (ecstasy). Moreover, a large number of
β-adrenergic receptor antagonists are influenced in their elimination by this
polymorphism (6, 10, 62, 65, 66).
Numerous case reports and clinical studies have demonstrated that for some
of these drugs, the polymorphic oxidation has therapeutic consequences, which
lead either to a higher propensity to develop adverse reactions at conventional
doses (as demonstrated by its discovery) or to decreased drug effects, e.g. ab-
sence of the analgesic effect of codeine in PMs (63) or therapeutic failure at
normal antidepressant doses in ultrarapid metabolizers (64). Moreover, epi-
demiologic studies have indicated that PMs and EMs may be at higher risk
for certain diseases, including Parkinson’s disease and some forms of cancer,
although many of these associations are disputed. Detailed reviews of the clin-
ical and toxicologic implications of the debrisoquine polymorphism have been
published (4–6, 62, 65, 66).

Molecular Mechanism
STUDIES IN HUMAN LIVER The identification of the target P450 isozyme took
several years of research by different groups, as has been reviewed in detail
(67). Early studies in human liver microsomes had suggested the deficiency of
a minor P450 enzyme with debrisoquine 4-hydroxylase activity in the liver of
P1: SDA/MKV P2: SDA/VKS QC: SDA
February 18, 1997 11:30 Annual Reviews AR027-11 AR27-011N

GENETIC POLYMORPHISM 279

PM individuals. However, purification of the respective P450 from human liver


was extremely difficult, and antibodies obtained against these fractions were
not specific enough to allow conclusions regarding the mechanism of deficient
metabolism (68, 69). With improved enzyme assays of high specificity and sen-
sitivity, the deficient enzyme was characterized in liver biopsy specimens from
donors who had known in vivo phenotype as a stereoselective (+)-bufuralol 10 -
hydroxylase (70, 71). Polyclonal antibodies against rat P450db1, which defined
the 2D subfamily in rats (72), were then shown to cross-react specifically with
human CYP2D6, and the levels of CYP2D6 determined in a large collection of
Annu. Rev. Pharmacol. Toxicol. 1997.37:269-296. Downloaded from www.annualreviews.org

human liver specimens were highly correlated to bufuralol 10 -hydroxylase ac-


tivity (71, 73). The lack of immunoreactivity observed in several liver samples
of PMs suggested absence of CYP2D6 protein as cause for the metabolic defi-
ciency (71, 73). This result was further confirmed with monoclonal antibodies
(71, 74) and human LKM1 autoantibodies (75).
by University of Sussex on 11/05/12. For personal use only.

By using the polyclonal antibody against rat P450db1, a full length human
cDNA was isolated. Sequence analysis revealed a distinct, novel member of
the CYP2D subfamily (73). In addition to the biochemical evidence, one other
observation strongly suggested CYP2D6 as the target gene of the debrisoquine-
sparteine polymorphism. Eichelbaum and colleagues (76) found strong linkage
of polymorphic sparteine oxidation to the P1 blood group, which had been
mapped to the long arm of chromosome 22. The CYP2D locus (22q13.1) could
subsequently be mapped by using the CYP2D6 cDNA (77, 78).

CYP2D GENES The CYP2D6 wild-type locus in human beings is comprised


of the three highly homologous genes, CYP2D8P, CYP2D7P, and CYP2D6,
which are located in this orientation (50 to 30 ) on a contiguous region of about
45 kb (79–81). Like other genes of the CYP2 family, CYP2D genes consist of
nine exons and eight introns. CYP2D8P is a pseudogene that contains multiple
deletions and insertions and causes a highly disrupted open reading frame (79,
81). The CYP2D7P gene is more similar to CYP2D6 than to CYP2D8P, and
its coding sequence indicates only a single inactivating mutation, an insertion
of T226 in the first exon. However, no specific mRNA product was detected
in RNA from eight human livers, which indicates that it also is a pseudogene
(79). Other sequenced CYP2D7 genes also contain the T226 mutation (81–
83). Hence, CYP2D6 is the only CYP2D protein product in human beings, in
contrast to rodents, which express most or all of their five functional CYP2D
genes.

GENETIC VARIABILITY Basic information on the structural variability of the


CYP2D6 locus was obtained in an extended RFLP analysis, in which 20 re-
striction enzymes were used to identify phenotype-associated polymorphisms
P1: SDA/MKV P2: SDA/VKS QC: SDA
February 18, 1997 11:30 Annual Reviews AR027-11 AR27-011N

280 MEYER & ZANGER

in leukocyte DNA of PM and EM individuals and their families (84). The


fragmentation patterns obtained with Xba I initially defined four allelic forms
of the CYP2D6 gene, represented by fragments of 29-kb, 44-kb, 11.5-kb, or
16 + 9-kb length, respectively (84, 85). However, only the alleles represented
by 44-kb and 11.5-kb fragments were clearly associated with the PM pheno-
type, thereby accounting for about one half the alleles in the PM group. The
remaining half was of the noninformative 29-kb type, which was also the most
frequent haplotype in EMs and thus represented the major polymorphic allele
(84). The 16 + 9-kb pattern was later also shown to be associated with the PM
Annu. Rev. Pharmacol. Toxicol. 1997.37:269-296. Downloaded from www.annualreviews.org

phenotype (85).

Null alleles Different approaches were taken in several parallel investigations


to uncover primary mutations at the DNA level that could explain the PM pheno-
type (80, 82, 86, 87). By complete sequence analysis of a CYP2D6 allele (44M)
by University of Sussex on 11/05/12. For personal use only.

isolated from a PM individual characterized as 44-kb/11.5-kb by Xba I RFLP


analysis, several mutations were identified in introns and in exons, four of which
cause amino acid substitutions (82). One intronic mutation, G1934A, appeared
to be the most consequential difference to the wild-type gene in that it affected
the consensus acceptor splice site of the third intron. The mechanism proposed
for this mutation is outlined in Figure 1. The G residue within the consensus AG
dinucleotide of the 30 acceptor splice site of the normal gene is changed into an A
in the mutant allele. As the first exon residue is a G, a new consensus site forms

Figure 1 Mechanism of the common CYP2D6 B-mutation present in CYP2D6∗ 4 alleles as pro-
posed by Hanioka et al (82) and Gough et al (87).
P1: SDA/MKV P2: SDA/VKS QC: SDA
February 18, 1997 11:30 Annual Reviews AR027-11 AR27-011N

GENETIC POLYMORPHISM 281

one base downstream, thereby predicting an mRNA with a single base deletion
to be the spliced gene product (82). Indeed, variant cDNAs with this deletion
were isolated from libraries prepared from low in vitro activity livers (73, 87).
At the protein level, this mechanism predicts premature translation termination
to occur at codon 182 (Figure 1), thus explaining the absence of immunoreactive
P450 protein and enzyme activity in PMs homozygous for this mutant allele.
The major polymorphic allele, characterized as a 29-kb Xba I fragment
(84), was studied by Kagimoto et al (86), who constructed genomic libraries
from leukocyte DNA of three different PM individuals with a homozygous
Annu. Rev. Pharmacol. Toxicol. 1997.37:269-296. Downloaded from www.annualreviews.org

29-/29-kb Xba I RFLP genotype. Sequence analysis of all exons and exon/intron
junctions and comparison to the CYP2D6 wild-type sequence (79) of sev-
eral clones revealed that an allele with multiple mutations, termed 29-B (or a
very similar allele, 29-B0 ; Figure 2) was present in all three individuals. The
by University of Sussex on 11/05/12. For personal use only.

Figure 2 Distribution and effect of the most common mutations of the CYP2D6 gene. Some
rare alleles are not included (see Ref. 92 for a more complete list). Positions of mutations are
numbered according to Kimura et al (79). Mutations are represented by open diamonds (silent
mutations), closed diamonds (mutations causing amino acid changes), and closed circles (loss-of-
function mutations) arranged along vertical lines that represent the alleles indicated on top. The
dotted line for CYP2D6∗ 5 indicates that the CYP2D6 gene is deleted. Frequencies for the whole
population (white) and for PM populations are approximate values (in %) compiled and modified
from Refs. 84, 89, 91, 95, 96, 99, 102, 103, 137, 138, 141, 143.
P1: SDA/MKV P2: SDA/VKS QC: SDA
February 18, 1997 11:30 Annual Reviews AR027-11 AR27-011N

282 MEYER & ZANGER

seven mutations found in the sequenced areas were the same as those of the
44M allele and included the G1934A transition (86). COS cell expression of
chimeric CYP2D6 genes and mutant cDNAs that contain individual and groups
of mutations by Kagimoto et al (86) demonstrated that the G1934A mutation
completely abolishes expression of protein and activity. In further agreement
with the mechanistic model, mutant CYP2D6 mRNA observed in transfected
COS cells had the same size as wild-type mRNA (86). Diagnostic DNA tests
then confirmed the high frequency of this mutation among PM individuals (10,
88–91).
Annu. Rev. Pharmacol. Toxicol. 1997.37:269-296. Downloaded from www.annualreviews.org

The other mutations present in the “B-type” alleles do not severely affect
protein expression and activity, with the exception of C188T (P34S), which
dramatically decreased enzyme activity but not protein expression in transfected
COS cells (86). This mutation, as well as a mutation in exon 9 (4268), is also
present in several functional alleles (Figure 2) and is discussed separately below.
by University of Sussex on 11/05/12. For personal use only.

According to a recently proposed CYP2D6 nomenclature system, alleles with


multiple mutations are grouped according to common detrimental mutations
(92). Group ∗ 4 comprises all alleles with the G1934A mutation, to which the
29-B (∗ 4A) and 29-B0 (∗ 4B), as well as the 44M alleles (∗ 4A) belong (92)
(Figure 2).
The second most frequent nonfunctional or null allele in whites is represented
by the 11.5-kb Xba I fragment, which was shown to be the result of a large
deletion mutation at the CYP2D locus, designated D-allele (80). The structure
of this allele was analyzed in genomic clones from a PM individual homozygous
for the XbaI 11.5-kb haplotype and was shown to comprise the pseudogenes
CYP2D8P and CYP2D7P, whereas the entire CYP2D6 coding sequence is
deleted.
Figure 2 summarizes the currently known loss of function mutations, most of
which can be detected with simple DNA tests (88, 93, 94). The more recently
reported alleles (∗ 6 to ∗ 16) were usually detected in PM individuals who had
none of the known mutations of alleles A, B, and D. The two most frequent
of these (∗ 6 and ∗ 3) also result in premature translation termination because of
single base deletions. In contrast, the single mutation of allele ∗ 7 predicts an
amino acid substitution, H324P, in exon six (95). Expression of CYP2D6(324P)
in COS cells and in insect cells resulted in comparable amounts of mutant and
wild-type protein, but the 324P mutant lacked heme and was enzymatically
inactive (96).

Functional alleles Among EM individuals, the MR for debrisoquine varies


up to 1000-fold, from as low as 0.01 up to the antimode of 12.6 for debrisoquine
(64, 97) and to a similar extent for other substrates. The terms “intermediate
metabolizer” (for MR1-12.6) and “ultrarapid metabolizer” [for MR < 0.1, (64)]
P1: SDA/MKV P2: SDA/VKS QC: SDA
February 18, 1997 11:30 Annual Reviews AR027-11 AR27-011N

GENETIC POLYMORPHISM 283

have been used to describe these phenotypic subpopulations. Some recently


identified mutations can be correlated to increased or decreased metabolic ca-
pacity and thus appear to contribute to the extreme variability among EMs.
In two human liver samples with low CYP2D6 enzymatic activity, trace
amounts of a cross-reactive protein with slightly higher apparent Mr were de-
tected (74). Subsequent cDNA cloning revealed a CYP2D6-C variant (∗ 9) with
a deleted codon 281, coding for lysine (98). The kinetic parameters determined
with several substrates for the mutant enzyme expressed by recombinant vac-
cinia virus in HepG2 cells were comparable to those of the wild-type. Based
Annu. Rev. Pharmacol. Toxicol. 1997.37:269-296. Downloaded from www.annualreviews.org

on the in vitro analyses of liver microsomes, it was initially thought that the
allele could be associated with the PM phenotype, but a family study with sev-
eral members of the C/D (∗ 5/∗ 9) genotype then allowed the correlation of this
allele to intermediate debrisoquine MRs (99). The overall C-allele frequency
in whites is about 1% (97). Interestingly, however, a higher frequency of this
by University of Sussex on 11/05/12. For personal use only.

mutation was reported in Spain (100).


The C188T mutation, which causes a Pro to Ser amino acid change at position
34, was first identified as one of several amino acid substitutions in the B-type
alleles (82, 86) and was shown by expression in COS cells to impair CYP2D6
enzyme function, but not expression, dramatically (86). Because of this effect,
alleles with this mutation but absent detrimental G1934A transition, as first
observed in Asian populations and termed J (for Japanese) or Ch (for Chinese)
(83, 101, 136), are grouped in family ∗ 10 (Figure 2). In whites, these alleles
are much less frequent than in Asians, but may contribute to the intermediate
metabolizer phenotype, as the presence of this mutation was correlated to higher
MR of EMs in several studies (83, 101, 102, 136). As noted earlier (82), Pro34
is part of a proline-rich region that is highly conserved among microsomal
P450s and may function as a hinge between the hydrophobic membrane anchor
and the globular heme-binding portion of the enzyme. Enzymatic activity
of COS cell–expressed P34S mutant CYP2D6 was not detectable (86). In
combination with a mutation in exon 9 (allele ∗ 10; Figure 2), it was estimated
to be about 5% of the wild-type activity (83). Nevertheless, individuals who
have a combined genotype of one ∗ 10 and one nonfunctional allele are always
EMs, which indicates that the low amounts of functional enzyme produced by

10 still provide sufficient metabolic capacity to cause metabolic ratios within
the EM phenotype range.
At the opposite side of the spectrum of EMs, the “ultrarapid metabolizer phe-
notype” was also shown to have a genetic basis. It was first observed in a clinical
setting in which two patients required antidepressants at unusually high doses
(“megaprescribing”) to attain therapeutic plasma concentrations (64). Both
patients were subsequently shown to carry one normal and one novel allele
characterized by a 42-kb Xba I fragment, which was distinct from the 44-kb
P1: SDA/MKV P2: SDA/VKS QC: SDA
February 18, 1997 11:30 Annual Reviews AR027-11 AR27-011N

284 MEYER & ZANGER

haplotype (64). The extremely rapid metabolism in the two patients was sug-
gested to be the consequence of a gene duplication, which led to two functional
copies of CYP2D6∗ 2 on the 42-kb fragment, thus causing more enzyme to be
expressed (64, 97, 103). In another family with ultrarapid metabolism of de-
brisoquine in three members (MR < 0.02), molecular analysis demonstrated
dominant inheritance of a 13-fold amplified CYP2D6∗ 2 gene from the father
to two children (103). The mechanism of this excessive gene amplification re-
mains speculative and restricted to a single observation, whereas the duplicated
allele (CYP2D6∗ 2 × 2) was shown to occur with allele frequencies of ∼3.5%
Annu. Rev. Pharmacol. Toxicol. 1997.37:269-296. Downloaded from www.annualreviews.org

in Spanish (104) and ∼1% in Swedish populations, where alleles with tripli-
cated CYP2D6∗ 2 were also observed (97). As much as 13% of an Ethiopian
population seemed to carry this gene duplication (105).
The implication of these findings is that each of the multiple CYP2D6∗ 2
copies on the variant amplified alleles is expressed and contributes to expression
by University of Sussex on 11/05/12. For personal use only.

of a functional protein. Although it was not explicitly shown that each of the
multiple copies represents a complete functional gene, the inverse correlation
found between the number of copies and the MR for debrisoquine strongly
suggests such a mechanism (97, 103).
Multiple copies and amplification of CYP2D6 apparently involves only the
variant CYP2D6∗ 2, which is distinct from the wild-type gene (79) at several
positions, including a gene conversion with CYP2D7 in the first intron and
two point mutations that cause amino acid substitutions R296C and S486T
(103). Being present also in a significant proportion of the genes of the com-
mon 29-kb haplotype, it represents the second most frequent functional allele
in whites, with an overall frequency of approximately 30% (103, 106). Corre-
lation studies suggest that this variant codes for a protein with normal function,
although position 296 maps to a putative substrate recognition sequence of
cytochrome P450.

Interethnic Variability and Evolution


As recently reviewed (6, 9), significant ethnic differences exist in the frequency
of the PM phenotype, which appears to be rare (1% to 2%) in most popula-
tions with the exception of Caucasians (5% to 10%). Interethnic variability of
CYP2D6 alleles is less completely studied than the NAT2 polymorphism, but
the existing data allow some conclusions on the origin of the most common
alleles (Table 4).
The most extreme variabilities are observed for the two allele families,
CYP2D6 ∗ 4 and ∗ 10 (Figure 2). The practical absence of the common “white”
B-allele from East Asian populations explains the very low prevalence of the
PM trait in the latter. A similarly low frequency of this allele was described
for two African populations (105, 107). In most African population studies,
P1: SDA/MKV P2: SDA/VKS QC: SDA
February 18, 1997 11:30 Annual Reviews AR027-11 AR27-011N

GENETIC POLYMORPHISM 285

Table 4 Interethnic comparison of CYP2D6 allele frequencies

No. of Nonfunctional
individuals Functional alleles (%) alleles (%)
Ethnic group analyzed ∗1 ∗2 ∗2 × 2 ∗ 10 ∗3 ∗4 ∗5

Whitesa Compiled 34 33 1.5–3.5 3 1 20 5


Chineseb Compiled 23 20 1 50 0 <1 5
Japanesec Compiled 42 12 0 33 0 0 13
Koreand 21 33 24 2 36 0 0 5
Ngawbe 30 17 0
Annu. Rev. Pharmacol. Toxicol. 1997.37:269-296. Downloaded from www.annualreviews.org

Amerindianse
Sinhalesef 77 0 9
African 127 2.4 8.5 5.5
Americansg
Ethiopiansh 115 13 10 0 1.3
Zimbabweani
by University of Sussex on 11/05/12. For personal use only.

114 0 1.8 3.9


Data compiled and modified from references. The sum of all alleles is not necessarily 100%.
a
89, 91, 97, 103, 104, 138.
b
83, 101, 102, 139.
c
136, 139.
d
139.
e
108.
f
144.
g
6.
h
105.
i
107.

PM frequencies were also found to be low, although dissociations have been


observed for some probe drugs (6, 9). A possible explanation for these findings
is that the B-allele appeared relatively recently in Caucasoids after their sepa-
ration from Asian/Amerindian groups about 35,000 years ago (58, 107). The
intermediate frequency of ∗ 4 alleles in American blacks is most likely caused by
penetrance of Caucasian-derived genes (6). However, a high frequency (17%)
was observed in an Amerindian tribes by Jorge et al (108), an observation that
challenges the above hypothesis. The ∗ 3 allele (A-mutation) also appears to
be rare except in whites, whereas the deletion-allele, CYP2D6∗ 5, occurs with
comparable frequency in most populations.
Another major interethnic difference is a shift in the metabolic ratio distri-
bution to higher values in Chinese populations as compared with whites (9).
The molecular basis for this phenotypic difference is a very high prevalence
in Asians of ∗ 10 alleles with lower enzyme activity due to the C188T (P34S)
mutation. Its frequency in whites may have been overestimated initially, as it
appears to be rather rare according to recent data (91).
The two major functional ∗ 1 and ∗ 2 alleles appear to be frequent throughout
all populations, but one should keep in mind that the frequencies for these al-
leles are determined by the absence of other mutations. Therefore, undiscovered
P1: SDA/MKV P2: SDA/VKS QC: SDA
February 18, 1997 11:30 Annual Reviews AR027-11 AR27-011N

286 MEYER & ZANGER

mutations may change the picture considerably in the future. Compared with
whites, a considerably higher incidence of gene duplication and amplification
alleles was reported for Ethiopians (105). The authors discussed the interest-
ing possibility that duplicated and amplified L-alleles (CYP2D6∗ 2 × N) may
have originated in Africa and transmitted to whites rather recently during the
immigration of Arabs into Spain, where a higher frequency is observed than in
the rest of Europe (104). There are no data yet on CYP2D6 alleles in Pacific
populations.
Annu. Rev. Pharmacol. Toxicol. 1997.37:269-296. Downloaded from www.annualreviews.org

THE MEPHENYTOIN POLYMORPHISM


Discovery and Clinical Relevance
A genetic polymorphism of another cytochrome P450 isozyme was revealed
by the discovery of deficient 40 -hydroxylation of mephenytoin (3-methyl-5-
by University of Sussex on 11/05/12. For personal use only.

phenyl-5 ethylhydantoin; Mesantoin), a now rarely used anticonvulsant drug


(109, 110). The deficiency is restricted to one of the two major metabolic
pathways of mephenytoin disposition, namely stereoselective hydroxylation of
S-mephenytoin in the p-phenyl position to 40 -OH-mephenytoin. The other main
reaction, N-demethylation of R-mephenytoin to 5-phenyl-5-ethylhydantoin
(PEH; Nirvanol) remains unaffected. This deficiency is inherited as an au-
tosomal recessive trait (111, 112) and occurs with a frequency of 2.5 to 6% in
whites, but with a much higher frequency in Japanese (18 to 23%) and Chinese
(15 to 17%) individuals (6, 113). Recent reviews of this polymorphism have
been published (6, 113, 114). Individuals who have the PM phenotype are
predisposed to have the central adverse effects of mephenytoin, eg., sedation
after administration of a single dose of 100 mg for phenotyping purposes (115,
116). Additional substrates for the enzyme metabolizing S-mephenytoin to 40 -
OH-mephenytoin have been tentatively identified by in vitro inhibition studies
and some of these have been confirmed by cosegregation in populations. Thus,
the metabolism of the N-demethylated metabolite of mephenytoin (nirvanol),
mephobarbital, and hexobarbital, the side-chain oxidation of propanolol, the
demethylation of imipramin, and the metabolism of diazepam and desmethyl-
diazepam cosegregate with the 20 -hydroxylation of S-mephenytoin (6, 113,
114). More recent studies also suggest that the metabolism of the antidepres-
sant citalopram, the proton-pump inhibitor omeprazol (117), and probably the
related drugs pantoprazole and lansoprazole (UA Meyer, unpublished data), as
well as the antimalarial prodrugs proguanyl and chlorproguanyl (118), coseg-
regate with polymorphic mephenytoin metabolism. The clinical consequences
of impaired mephenytoin hydroxylation have not been studied in detail. There
is some evidence that the frequency of adverse effects of mephenytoin, mepho-
barbital, and hexobarbital may be higher in poor metabolizers of mephenytoin
P1: SDA/MKV P2: SDA/VKS QC: SDA
February 18, 1997 11:30 Annual Reviews AR027-11 AR27-011N

GENETIC POLYMORPHISM 287

(115, 116). Of potential significance would be a decreased efficacy of pro-


phylaxis for chloroquine-resistant P. falciparum malaria in PMs because of
the decreased formation of the active antiplasmodial metabolite cycloguanil or
chlorocycloguanil. Initial studies in Tanzania, however, did not support such
a relationship (119). May et al (120) have described an association of the PM
phenotype with scleroderma, particularly when it occurs in combination with
slow dapsone hydroxylation.
Molecular Mechanism
Annu. Rev. Pharmacol. Toxicol. 1997.37:269-296. Downloaded from www.annualreviews.org

STUDIES IN HUMAN LIVER Initial studies in microsomes of liver biopsy spec-


imens from EM and PM individuals suggested the deficiency of a specific
P450 isozyme (121). Studies in several laboratories during the next few years,
however, failed to identify unequivocally the gene and the enzyme affected
by this polymorphism. Proteins were purified from human liver by Gut et al
by University of Sussex on 11/05/12. For personal use only.

(122) and Shimada et al (123) by monitoring S-mephenytoin 40 hydroxyla-


tion activity. Antibodies against the purified fractions inhibited S-mephenytoin
40 -hydroxylation, but there was no correlation between immunochemically rec-
ognized levels of these proteins and catalytic activity (124, 125). The issue has
now been resolved in that a protein with the N-terminus of CYP2C19, identified
by cloning studies (126), was purified by Wrighton et al (127). Immunoblots of
this protein correlated with S-mephenytoin 40 -hydroxylase activity. These data
were confirmed by Goldstein et al (128) by expression of a CYP2C19 cDNA in
yeast and refuted previous assumptions that variants of CYP2C9 may contribute
significantly to hepatic S-mephenytoin 40 -hydroxylase activity. CYP2C19 ap-
pears to be virtually absent on immunoblots of liver microsomes of PMs of
mephenytoin (129).
CYP2C GENES At least seven genes or pseudogenes of the CYP2C subfamily
appear to be localized on chromosome 10 (124, 130). The complete CYP2C18
gene and part of the CYP2C9 gene have been isolated and both are ∼55 kb in
size and have nine exons, as have other genes of family 2 (131). The CYP2C19
gene has not been studied so far, but it is closely related to other genes of the
2C subfamily, e.g. CYP2C9, with 95% similarity in the upstream region and in
several exons (114). According to recent nomenclature concepts, the wild-type
CYP2C19 gene may also be designated as CYP2C19∗ 1.
ALLELES OF CYP2C19 Reverse transcription and amplification of mRNA from
liver samples with low microsomal S-mephenytoin 40 -hydroxylase activity re-
vealed a 2C19 cDNA fragment with a 40-bp deletion at the beginning of exon 5,
which included a SmaI restriction site. This results in the deletion of 13 amino
acids (215–227) and shifts the reading frame, thereby producing a stop codon
20 amino acids downstream and consequently a truncated protein (234 amino
P1: SDA/MKV P2: SDA/VKS QC: SDA
February 18, 1997 11:30 Annual Reviews AR027-11 AR27-011N

288 MEYER & ZANGER

acids), which would be devoid of a heme binding site and therefore be inactive.
Sequencing of amplified genomic DNA fragments from PM livers revealed a G
to A substitution in the coding sequence of exon 5 that corresponded to bp 681
of the cDNA. This mutation, designated CYP2C19m1 (or CYP2C19∗ 2) creates
an aberrant splice site that is apparently preferred exclusively over the normal
splice site. The abolition of a SmaI site by this mutation led to the development
of a convenient genotyping test. This mutation accounted for 75% and 83% of
the defective allele in a small sample of Japanese and white PM individuals,
respectively (129). A second mutation, defined as CYP2C19m2 (CYP2C19∗ 3),
Annu. Rev. Pharmacol. Toxicol. 1997.37:269-296. Downloaded from www.annualreviews.org

has recently been discovered in DNA of a Japanese PM who did not have the
m1 mutation. CYP2C19m2 involves a G → A transition at bp 636 in exon 4,
which also creates a premature stop codon and destroys a BamHI restriction
site. This change would result in a truncated 211 amino acid polypeptide (132).
A PCR test for this mutation revealed that CYP2C19m2 accounts for the re-
by University of Sussex on 11/05/12. For personal use only.

maining defective alleles in Japanese or Chinese PMs. PCR-based tests for


the CYP2C19m1 and m2 mutations have been described (129, 132) and have
been evaluated in population and family studies to be of high sensitivity and
selectivity. However, whereas virtually all PMs in Japan had the genotypes
m1/m1, m1/m2, or m2/m2, the m2 mutation appears to be exceedingly rare in
whites.
A large family study (43 relatives of families of 11 PM probands) in Denmark
(133) revealed only one m2 mutation, and random testing of 800 alleles from
whites revealed no m2 alleles (114). In conclusion, the two CYP2C19 mutations
m1 and m2 explain virtually all PMs in Japanese and Chinese populations, but
additional mutations of CYP2C19 must exist in whites.

Interethnic Variability and Evolution


The ethnic and geographic distribution of the mephenytoin polymorphism has
been less well studied than that of the acetylation and debrisoquine polymor-
phism. A summary of 18 phenotyping studies that has recently been published
(134) includes Saudi Arabians, Filipinos, and Indians. An interesting relation-
ship between longitude and the frequency of PMs was observed, which suggests
a cline containing a step that occurs between Riyadh and Bombay. The Saudi
population resembled Europeans in the frequency of PMs for mephenytoin but
was similar to Asians in the frequency of PMs for debrisoquine.

DIRECTIONS FOR FUTURE RESEARCH


The initial observations of unexpected toxicity and the large number of com-
pounds metabolized by NAT2, CYP2D6, and CYP2C19 have resulted in a
considerable research effort to understand the molecular mechanisms of these
P1: SDA/MKV P2: SDA/VKS QC: SDA
February 18, 1997 11:30 Annual Reviews AR027-11 AR27-011N

GENETIC POLYMORPHISM 289

polymorphisms. For example, the unraveling of the complex and highly poly-
morphic CYP2D gene cluster on chromosome 22 has thus resulted in the char-
acterization of a total of 48 mutations and 53 alleles of the CYP2D6 locus alone
(91). In contrast to the detailed knowledge of molecular mechanisms, how-
ever, the clinical relevance of these polymorphisms is poorly documented and
is deduced largely from anecdotal reports and epidemiologic studies in small
numbers of patients. There is an obvious discrepancy between detailed molec-
ular information and only rudimentary clinical evaluation of the consequences
of these polymorphisms. The availability of genotyping tests should help de-
Annu. Rev. Pharmacol. Toxicol. 1997.37:269-296. Downloaded from www.annualreviews.org

sign prospective clinical studies of adverse drug reactions and drug toxicity and
should ultimately serve to help in the adjustment of drug doses to individual
metabolic capacity.
The associations of these polymorphisms with individual risk to develop cer-
tain diseases suggest that the polymorphic enzymes may activate or inactivate
by University of Sussex on 11/05/12. For personal use only.

environmental carcinogens, other xenobiotics, or endobiotics. With the ex-


ception of amine carcinogens as substrates for acetyltransferases, however, the
search for chemical carcinogens that may be activated or inactivated by CYP2D6
or CYP2C19 has not been successful. Finally, the potential role of the discussed
enzymes in the metabolism of endogenous substances, for instance endogenous
codeine and morphine by CYP2D6 (135) or hormones and neurotransmitters,
as inferred from indirect observations, has not been evaluated (62).

ACKNOWLEDGMENTS
We thank Drs. F. Broly, J. Agundez, and M. Ingelman-Sundberg for sending
us data and copies of articles that were in press. The research in the laboratory
of UAM is supported by the Swiss National Research Foundation. Ulrich M.
Zanger is supported by the Robert Bosch Foundation, Stuttgart, Germany.
Visit the Annual Reviews home page at
http://www.annurev.org.

Literature Cited

1. Vogel F, Motulsky AG, eds. 1986. Hu- Drug Metabolism. New York: Pergamon
man Genetics. Problems and Approaches. 5. Meyer UA. 1992. Drugs in special patient
New York: Springer-Verlag. 435 pp. groups: clinical importance of genetics in
2. Meyer UA. 1991. Genotype or phenotype: drug effects. In Clinical Pharmacology:
the definition of a pharmacogenetic poly- Basic Principles of Therapeutics, ed. KF
morphism. Pharmacogenetics 1:66–67 Melmon, HF Morelli, pp. 875–94. New
3. Meyer UA, Zanger UM, Grant D, Blum York: McGraw-Hill
M. 1990. Genetic polymorphisms of drug 6. Evans DAP. 1993. Genetic Factors in
metabolism. In Advances in Drug Re- Drug Therapy. Clinical and Molecular
search, ed. B Testa, pp. 197–41. London: Pharmacogenetics. Cambridge: Cam-
Academic bridge Univ. Press
4. Kalow W. 1992. Pharmacogenetics of 7. Meyer UA. 1994. Commentary. Pharma-
P1: SDA/MKV P2: SDA/VKS QC: SDA
February 18, 1997 11:30 Annual Reviews AR027-11 AR27-011N

290 MEYER & ZANGER

cogenetics: The slow, the rapid, and the polymorphism and low-level environ-
ultrarapid. Proc. Natl. Acad. Sci. USA 91: mental exposure to carcinogens. Nature
1983–84 369:154–56
8. Meyer UA. 1994. The molecular basis of 22. Risch A, Wallace DMA, Bathers S, Sim
genetic polymorphisms of drug metabo- E. 1995. Slow N-acetylation genotype
lism. J. Pharm. Pharmacol. 46:409–15 is a susceptibility factor in occupational
9. Bertilsson L. 1995. Geographical/inter- and smoking related bladder cancer. Hum.
racial differences in polymorphic drug Mol. Genet 4:231–36
oxidation. Clin. Pharmacokinet. 29:192– 23. Grant DM, Beer M, Blum M, Meyer UA.
99 1991. Monomorphic and polymorphic
10. Daly AK. 1995. Molecular basis of poly- human arylamine N-acetyltransferases: a
morphic drug metabolism. J. Mol. Med. comparison of liver isozymes and ex-
73:539–53 pressed products of two cloned genes.
Annu. Rev. Pharmacol. Toxicol. 1997.37:269-296. Downloaded from www.annualreviews.org

11. Pacifici GM, Fracchia GN. 1995. Medi- Mol. Pharmacol. 39:184–91
cine and health. In Advances in Drug 24. Jenne JW. 1965. Partial purification and
Metabolism in Man. EUR 15439. Luxem- properties of the isoniazid transacetylase
bourg: Office for Official Publications of in human liver. Its relationship to the
the European Communities acetylation of p-aminosalicyclic acid. J.
12. Bönicke R, Reif W. 1953. Enzymatic in- Clin. Invest. 44:1992–2002
activation of isonicotinic acid hydrazide 25. Grant DM, Lottspeich F, Meyer UA.
by University of Sussex on 11/05/12. For personal use only.

in humans and animals. Arch. Exp. Pathol. 1989. Evidence for two closely re-
Pharmakol. 220:321–33 lated isozymes of arylamine N-acetyl-
13. Weber WW. 1987. The Acetylator Genes transferase in human liver. FEBS. Lett.
and Drug Response. New York: Oxford 244:203–7
Univ. Press 26. Grant DM, Mörike K, Eichelbaum M,
14. Evans DAP. 1989. N-Acetyltransferase. Meyer UA. 1990. Acetylation pharma-
Pharmacol. Ther. 42:157–234 cogenetics. The slow acetylator pheno-
15. Grant DM. 1993. Molecular genetics type is caused by decreased or absent
of the N-acetyltransferases. Pharmacoge- arylamine N-acetyltransferase in human
netics 3:45–50 liver. J. Clin. Invest. 85:968–72
16. Meyer UA, Blum M, Grant D, Heim M, 27. Andres HH, Vogel RS, Tarr GE, John-
Broly F, et al. 1993. Acetylation pharma- son L, Weber WW. 1987. Purification,
cogenetics. In Human Drug Metabolism. physicochemical, and kinetic properties
From Molecular Biology to Man, ed. of liver acetyl-CoA: arylamine N-acetyl-
EH Jeffery, pp. 117–24. Boca Ra- transferase from rapid acetylator liver.
ton/AnnArbor/London: CRC Press Mol. Pharmacol. 31:446–56
17. Vatsis KP, Weber WW. 1994. Human 28. Blum M, Grant DM, Demierre A, Meyer
N-acetyltransferases. In Conjugation-De- UA. 1989. N-acetylation pharmacogenet-
conjugation. Reactions in Drug Meta- ics: A gene deletion causes absence of
bolism and Toxicity. Handbook of Ex- arylamine N-acetyltransferase in liver of
perimental Pharmacology, ed. FC Kauff- slow acetylator rabbits. Proc. Natl. Acad.
mann, pp. 109–30. Berlin/Heidelberg: Sci. USA 86:9554–57
Springer-Verlag 29. Blum M, Grant DM, McBride W, Heim
18. Vatsis KP, Wendell WW, Douglas AB, M, Meyer UA. 1990. Human arylamine
Dupret J-M, Price Evans DA, et al. 1995. N-acetyltransferase genes: isolation,
Nomenclature for N-acetyltransferases. chromosomal localization, and functional
Pharmacogenetics 5:1–17 expression. DNA Cell Biol. 9:193–203
19. Wolkenstein P, Carrière V, Charue D, 30. Ebisawa T, Deguchi T. 1991. Structure
Bastuji-Garin S, Revuz J. 1995. A slow and restriction fragment length polymor-
acetylator genotype is a risk factor phism of genes for human liver arylamine
for sulphonamide-induced toxic epider- N-acetyltransferases. Biochem. Biophys.
mal necrolysis and Stevens-Johnson syn- Res. Commun. 177:1252–57
drome. Pharmacogenetics 5:255–58 31. Hickman D, Risch A, Buckle V, Spurr
20. Hayes RB, Bi W, Rothman N, Broly F, NK, Jeremiah SJ, et al. 1994. Chromoso-
Caproaso N, et al. 1993. N-Acetylation mal localization of human genes for ary-
phenotype and genotype and risk of blad- lamine N -acetyltransferase. Biochem. J.
der cancer in benzidine-exposed workers. 297:441–45
Carcinogenesis 14:675–78 32. Cribb AE, Nakamura H, Grant DM,
21. Vineis P, Bartsch H, Caporaso N, Harring- Miller MA, Spielberg SP. 1993. Role of
ton AM, Kadlubar FF, et al. 1994. Geneti- polymorphic and monomorphic human
cally based N-acetyltransferase metabolic arylamine N -acetyltransferases in de-
P1: SDA/MKV P2: SDA/VKS QC: SDA
February 18, 1997 11:30 Annual Reviews AR027-11 AR27-011N

GENETIC POLYMORPHISM 291

termining sulfamethoxazole metabolism. human gene for polymorphic N-acetyl-


Biochem. Pharmacol. 45:1277–82 transferase. Proc. Natl. Acad. Sci. USA
33. Minchin RF. 1995. Acetylation of p- 88:6333–37
aminobenzoylglutamate, a folic acid 44. Cascorbi I, Drakoulis N, Brockmöller J,
catabolite, by recombinant human ary- Maurer A, Sperling K, Roots I. 1995. Ary-
lamine N -acetyltransferase and U937 lamine N -acetyltransferase (NAT2) mu-
cells. Biochem. J. 307:1–3 tations and their allelic linkage in unre-
34. Ward A, Summer MJ, Sim E. 1995. lated Caucasian individuals: correlation
Purification of recombinant human N - with phenotypic activity. Am. J. Hum.
acetyltransferase type 1 (NAT1) ex- Genet. 57:581–92
pressed in E. coli and characterization 45. Cascorbi I, Brockmöller J, Bauer S, Reum
of its potential role in folate metabolism. T, Roots I. 1996. NAT2∗ 12A (803A-G)
Biochem. Pharmacol. 49:1759–67 codes for rapid arylamine N -acetylation
Annu. Rev. Pharmacol. Toxicol. 1997.37:269-296. Downloaded from www.annualreviews.org

35. Dupret JM, Grant DM. 1992. Site directed in human. Pharmacogenetics 6:257–59
mutagenesis of recombinant human ary- 46. Agundez JAG, Olivera M, Martinez C,
lamine N-acetyltransferase expressed in Ladero JM, Benitez J. 1996. Identification
E. coli: evidence for direct involvement and prevalence study of 17 allelic variants
of cys 68 in the catalytic mechanism of of the human NAT2 gene in a white pop-
polymorphic human NAT2. J. Biol. Chem. ulation. Pharmacogenetics 6:423–28
267:7381–85 47. Lin HJ, Han C-Y, Lin BK, Hardy S.
by University of Sussex on 11/05/12. For personal use only.

36. Watanabe M, Sofuni T, Nohmi TJ. 1992. 1993. Slow acetylator mutations in the
Involvement of cys69 residue in the cat- human polymorphic N-acetyltransferase
alytic mechanism of N-hydroxylamine gene in 786 Asians, Blacks, Hispanics
O-acetyltransferase of Salmonella ty- and Whites: application to metabolic epi-
phimurium: sequence similarity at the demiology. Am. J. Hum. Genet. 52:827–
amino acid level suggests a common cat- 34
alytic mechanism of acetyltransferase for 48. Lin HJ, Han C-Y, Lin BK, Hardy S.
S. typhimurium and higher organisms. J. 1994. Ethnic distribution of slow acety-
Biol. Chem. 267:8427–36 lator mutations in the polymorphin N -
37. Blum M, Demierre A, Grant DM, Heim acetyltransferase (NAT2) gene. Pharma-
M, Meyer UA. 1991. Molecular mecha- cogenetics 4:125–34
nism of slow acetylation in man. Proc. 49. Martinez C, Agundez JAG, Olivera M,
Natl. Acad. Sci. USA 88:5237–41 Martin R, Ladero JM, Benitez J. 1995.
38. Ohsako S, Deguchi T. 1990. Cloning Lung cancer and mutations at the poly-
and expression of cDNAs for polymor- morphic NAT2 gene locus. Pharmacoge-
phic and monomorphic arylamine N- netics 5:207–14
acetyltransferases from human liver. J. 50. Deloménie C, Sica L, Grant DM,
Biol. Chem. 265:4630–34 Krishnamoorthy R, Dupret J-M. 1996.
39. Deguchi T. 1992. Sequences and expres- Genotyping of the polymorphic N -
sion of alleles of polymorphic arylamine acetyltransferase (NAT2∗ ) gene locus in
N-acetyltransferase of human liver. J. two native African populations. Pharma-
Biol. Chem. 267:18140–47 cogenetics 6:177–85
40. Hickman D, Risch A, Camilleri JP, Sim 51. Abe M, Deguchi T, Suzuki T. 1993. The
E. 1992. Genotyping human polymorphic structure and characteristics of a fourth al-
arylamine N -acetyltransferase: identifi- lele of polymorphic N -acetyltransferase
cation of new slow allotypic variants. gene found in the Japanese popula-
Pharmacogenetics 2:217–26 tion. Biochem. Biophys. Res. Commun.
41. Bell DA, Taylor JA, Butler MA, Stephens 191:811–16
EA, Wiest J, et al. 1993. Gentoype/- 52. Hein DW, Ferguson RJ, Doll MA, Rustan
phenotype discordance for human ary- TD, Gray K. 1994. Molecular genetics of
lamine N -acetyltransferase (NAT2) re- human polymorphic N-acetyltransferase:
veals a new slow acetylator allele com- enzymatic analysis of 15 recombinant
mon in African-Americans. Carcinogen- wild-type, mutant, and chimeric NAT2 al-
esis 14:1689–92 lozymes. Hum. Mol. Genet. 3:729–34
42. Deguchi T, Mashimo M, Suzuki T. 1990. 53. Cribb AE, Grant DM, Miller MA, Spiel-
Correlation between acetylator pheno- berg SP. 1991. Expression of monomor-
types and genotypes of polymorphic phic arylamine N-acetyltransferase
arylamine N-acetyltransferase in human (NAT1) in human leucocytes. J. Pharma-
liver. J. Biol. Chem. 265:12757–60 col. Exp. Ther. 259:1241–46
43. Vatsis KP, Martel KJ, Weber WW. 54. Ward A, Hickman D, Gordon JW, Sim E.
1991. Diverse point mutations in the 1992. Arylamine N -acetyltransferase in
P1: SDA/MKV P2: SDA/VKS QC: SDA
February 18, 1997 11:30 Annual Reviews AR027-11 AR27-011N

292 MEYER & ZANGER

human red blood cells. Biochem. Phar- mechanisms. Pharmacol. Ther. 46:297–
macol. 44:1099–104 308
55. Vatsis KP, Weber WW. 1993. Struc- 68. Gut J, Catin T, Dayer P, Kronbach T,
tural heterogeneity of Caucasian N - Zanger U, Meyer UA. 1986. Debri-
acetyltransferase at the NAT1 gene lo- soquine/sparteine-type polymorphism of
cus. Arch. Biochem. Biophys. 301:71– drug oxidation. Purification and charac-
76 terization of two functionally different
56. Grant DM, Vohra P, Avis Y, Ima A. 1992. human liver cytochrome P-450 isozymes
Detection of a new polymorphism of hu- involved in impaired hydroxylation of
man arylamine N-acetyltransferase NAT1 the prototype substrate bufuralol. J. Biol.
using p-aminosalicylic acid as an in vivo Chem. 261:11734–43
probe. Clin. Physiol. Pharmacol. 3:244 69. Distlerath LM, Reilly PE, Martin MV,
57. Arias TD, Jorge LF, Griese E-U, In- Davis GG, Wilkinson GR, Guengerich FP.
Annu. Rev. Pharmacol. Toxicol. 1997.37:269-296. Downloaded from www.annualreviews.org

aba TEM. 1993. Polymorphic N -acetyl- 1985. Purification and characterization of


transferase (NAT2) in amerindian popu- the human liver cytochromes P-450 in-
lations of Panama and Colombia: high volved in debrisoquine 4-hydroxylation
frequencies of M3 mutation. Pharmaco- and phenacetin O-deethylation, two pro-
genetics 3:328–31 totypes for genetic polymorphism in ox-
58. Cavalli-Sforza LL, Menozzi P, Piazza A. idative drug metabolism. J. Biol. Chem.
1994. The History and Geography of 260:9057–67
by University of Sussex on 11/05/12. For personal use only.

Human Genes. Princeton, NJ: Princeton 70. Dayer P, Kronbach T, Eichelbaum M,


Univ. Press Meyer UA. 1987. Enzymatic basis of the
59. Mahgoub A, Idle JR, Dring LG, Lancester debrisoquine/sparteine-type genetic poly-
R, Smith RL. 1977. Polymorphic hydrox- morphism of drug oxidation. Character-
ylation of debrisoquine in man. Lancet ization of bufuralol 10 -hydroxylation in
ii:584–86 liver microsomes of in vivo phenotyped
60. Eichelbaum M, Spannbrucker N, Ste- carriers of the genetic deficiency. Bio-
incke B, Dengler JJ. 1979. Defective N- chem. Pharmacol. 36:4145–52
oxidation of sparteine in man: a new har- 71. Zanger UM, Vilbois F, Hardwick JP,
macogenetic defect. Eur. J. Clin. Pharma- Meyer UA. 1988. Absence of hepatic
col. 16:183–87 cytochrome P450bufI causes genetically
61. Alvan G, Bechtel P, Iselius L, Gundert- deficient debrisoquine oxidation in man.
Remy U. 1990. Hydroxylation polymor- Biochemistry 27:5447–54
phisms of debrisoquine and mephenytoin 72. Gonzalez FJ, Matsunaga T, Nagata K,
in European populations. Eur. J. Clin. Meyer UA, Nebert DW, et al. 1987. Debri-
Pharmacol. 39:533–37 soquine 4-hydroxylase: characterization
62. Kroemer HK, Eichelbaum M. 1995. “It’s of a new P450 gene subfamily, regula-
the genes, stupid.” Molecular bases and tion, chromosomal mapping, and molec-
clinical consequences of genetic cy- ular analysis of the DA rat polymorphism.
tochrome P450 2D6 polymorphism. Life DNA 6:149–61
Sci. 56:2285–98 73. Gonzalez FJ, Skoda RC, Kimura S,
63. Sindrup SH, Brosen K. 1995. The parma- Umeno M, Zanger UM, et al. 1988. Char-
cogenetics of codeine hypoalgesia. Phar- acterization of the common genetic de-
macogenetics 5:335–46 fect in humans deficient in debrisoquine
64. Bertilsson L, Dahl ML, Sjoqvist F, Aberg metabolism. Nature 331:442–46
Wistedt A, Humble M, et al. 1993. Molec- 74. Zanger UM, Meyer UA. 1989. Absent,
ular basis for rational megaprescribing in decreased or variant P450db1 in liv-
ultrarapid hydroxylators of debrisoquine. ers with poor capacity for debrisoquine
Lancet 341:63 metabolism. In Cytochrome P450: Bio-
65. Brosen K, Gram LF, Sindrup S, Skjelbo chemistry and Biophysics, ed. I Schuster,
E, Nielsen KK. 1992. Pharmacogenetics pp. 568–71. London: Taylor & Francis
of tricyclic and novel antidepressants: re- 75. Zanger UM, Hauri HP, Loeper J,
cent developments. Clin. Neuropharma- Homberg JC, Meyer UA. 1988. Antibod-
col. 15 (suppl 1):80A–81A ies against human cytochrome P-450db1
66. Eichelbaum M, Gross AS. 1990. The in autoimmune hepatitis type II. Proc.
genetic polymorphism of debrisoquine/- Natl. Acad. Sci. USA 85:8256–60
sparteine metabolism-clinical aspects. 76. Eichelbaum M, Baur MP, Dengler
Pharmacol. Ther. 46:377–94 HJ, Osikowska Evers BO, Tieves G,
67. Meyer UA, Skoda RC, Zanger UM. 1990. et al. 1987. Chromosomal assignment
The genetic polymorphism of debri- of human cytochrome P-450 (debriso-
soquine/sparteine metabolism-molecular quine/sparteine type) to chromosome 22.
P1: SDA/MKV P2: SDA/VKS QC: SDA
February 18, 1997 11:30 Annual Reviews AR027-11 AR27-011N

GENETIC POLYMORPHISM 293

Br. J. Clin. Pharmacol. 23:455–58 tions of the human cytochrome P450IID6


77. Gonzalez FJ, Vilbois F, Hardwick JP, gene (CYP2D6) in poor metabolizers of
McBride OW, Nebert DW, et al. 1988. debrisoquine. Study of the functional sig-
Human debrisoquine 4-hydroxylase nificance of individual mutations by ex-
(P450IID1): cDNA and deduced amino pression of chimeric genes. J. Biol. Chem.
acid sequence and assignment of the 265:17209–14
CYP2D locus to chromosome 22. 87. Gough AC, Miles JS, Spurr NK, Moss JE,
Genomics 2:174–79 Gaedigk A, et al. 1990. Identification of
78. Gough AC, Smith CA, Howell SM, Wolf the primary gene defect at the cytochrome
CR, Bryant SP, Spurr NK. 1993 . Local- P450 CYP2D locus. Nature 347:773–76
ization of the CYP2D gene locus to hu- 88. Heim M, Meyer UA. 1990. Genotyping
man chromosome 22q13.1 by polymerase of poor metabolisers of debrisoquine by
chain reaction, in situ hybridization, and allele-specific PCR amplification. Lancet
Annu. Rev. Pharmacol. Toxicol. 1997.37:269-296. Downloaded from www.annualreviews.org

linkage analysis. Genomics 15:430–32 336:529–32


79. Kimura S, Umeno M, Skoda RC, Meyer 89. Broly F, Gaedigk A, Heim M, Eichel-
UA, Gonzalez FJ. 1989. The human de- baum M, Morike K, Meyer UA. 1991. De-
brisoquine 4-hydroxylase (CYP2D) lo- brisoquine/sparteine hydroxylation geno-
cus: sequence and identification of the type and phenotype: analysis of com-
polymorphic CYP2D6 gene, a related mon mutations and alleles of CYP2D6 in
gene, and a pseudogene. Am. J. Hum. a European population. DNA Cell Biol.
by University of Sussex on 11/05/12. For personal use only.

Genet. 45:889–904 10:545–58


80. Gaedigk A, Blum M, Gaedigk R, Eichel- 90. Broly F, Marez D, Sabbagh N, Legrand
baum M, Meyer UA. 1991. Deletion of the M, Millecamps S, et al. 1995. An efficient
entire cytochrome P450 CYP2D6 gene strategy for detection of known and new
as a cause of impaired drug metabolism mutations of the CYP2D6 gene using sin-
in poor metabolizers of the debriso- gle strand conformation polymorphism
quine/sparteine polymorphism. Am. J. analysis. Pharmacogenetics 5:373–84
Hum. Genet. 48:943–50 91. Marez D, Legrand M, Sabbagh N, Lo
81. Heim MH, Meyer UA. 1992. Evolution of Guidice J-M, Spire C, et al. 1996. Po-
a highly polymorphic human cytochrome lymorphism of the cytochrome P450
P450 gene cluster: CYP2D6. Genomics CYP2D6 gene in a European population:
14:49–58 characterization of 48 mutations and 53
82. Hanioka N, Kimura S, Meyer UA, Gonza- alleles, their frequencies and evolution.
lez FJ. 1990. The human CYP2D locus as- Pharmacogenetics. In press
sociated with a common genetic defect in 92. Daly AK, Brockmöller J, Broly F, Eichel-
drug oxidation: a G1934-A base change baum M, Evans WE, Gonzalez FJ, et al.
in intron 3 of a mutant CYP2D6 allele re- 1996. Nomenclature for human CYP2D6
sults in an aberrant 30 splice recognition alleles. Pharmacogenetics 6:193–201
site. Am. J. Hum. Genet. 47:994–1001 93. Steen VM, Andreassen OA, Daly AK,
83. Johansson I, Oscarson M, Yue QY, Tefre T, Borresen AL, et al. 1995. De-
Bertilsson L, Sjoqvist F, Ingelman Sund- tection of the poor metabolizer-associated
berg M. 1994. Genetic analysis of the Chi- CYP2D6(D) gene deletion allele by
nese cytochrome P4502D locus: char- long-PCR technology. Pharmacogenetics
acterization of variant CYP2D6 genes 5:215–23
present in subjects with diminished ca- 94. Stüven T, Griese E-U, Kroemer HK,
pacity for debrisoquine hydroxylation. Eichelbaum M, Zanger UM. 1996. Rapid
Mol. Pharmacol. 46:452–59 detection of CYP2D6 null alleles by long
84. Skoda RC, Gonzalez FJ, Demierre A, distance- and multiplex-polymerase chain
Meyer UA. 1988. Two mutant alleles of reaction. Pharmacogenetics 6:417–21
the human cytochrome P-450db1 gene 95. Evert B, Griese EU, Eichelbaum M. 1994.
(P450C2D1) associated with genetically A missense mutation in exon 6 of the
deficient metabolism of debrisoquine and CYP2D6 gene leading to a histidine 324
other drugs. Proc. Natl. Acad. Sci. USA to proline exchange is associated with the
85:5240–43 poor metabolizer phenotype of sparteine.
85. Evans WE, Relling MV. 1990. Xbal 16- Naunyn. Schmiedebergs. Arch. Pharma-
plus 9-kilobase DNA restriction frag- col. 350:434–39
ments identify a mutant allele for de- 96. Evert B, Eichelbaum M, Haubruck H,
brisoquin hydroxylase: report of a family Zanger UM. 1997. Functional prop-
study. Mol. Pharmacol. 37:639–42 erties of CYP2D6 1 (wild-type) and
86. Kagimoto M, Heim M, Kagimoto K, Zeu- CYP2D6 7 (His324 Pro) expressed by re-
gin T, Meyer UA. 1990. Multiple muta- combinant baculovirus in insect cells.
P1: SDA/MKV P2: SDA/VKS QC: SDA
February 18, 1997 11:30 Annual Reviews AR027-11 AR27-011N

294 MEYER & ZANGER

Naunyn-Schmiedeberg’s Arch. Pharma- sive metabolisers. Hum. Genet. 94:401–6


col. In press 107. Masimirembwa CM, Johansson I, Hasler
97. Dahl ML, Johansson I, Bertilsson L, In- JA, Ingelman Sundberg M. 1993. Ge-
gelman Sundberg M, Sjoqvist F. 1995. netic polymorphism of cytochrome P450
Ultrarapid hydroxylation of debrisoquine CYP2D6 in Zimbabwean population.
in a Swedish population. Analysis of the Pharmacogenetics 3:275–80
molecular genetic basis. J. Pharmacol. 108. Jorge LF, Arias TD, Griese U, Nebert
Exp. Ther. 274:516–20 DW, Eichelbaum M. 1993. Evolution-
98. Tyndale RF, Aoyama T, Broly F, Mat- ary pharmacogenetics of CYP2D6 in
sunaga T, Inaba T, et al. 1991. Identifica- Ngawbe Guaymi of Panama: allele-
tion of a new variant CYP2D6 allele lack- specific PCR detection of the CYP2D6B
ing the codon encoding Lys 281: possi- allele and RFLP analysis. Pharmacoge-
ble association with the poor metabolizer netics 3:231–38
Annu. Rev. Pharmacol. Toxicol. 1997.37:269-296. Downloaded from www.annualreviews.org

phenotype. Pharmacogenetics 1:26–32 109. Küpfer APR, Ward S, Schenker S, Preisig


99. Broly F, Meyer UA. 1993. Debriso- R, Branch RA. 1984. Stereoselective
quine oxidation polymorphism: pheno- metabolism and pharmacogenetic con-
typic consequences of a 3-base-pair dele- trol of 5-phenyl-5-ethylhydantoin (Nir-
tion in exon 5 of the CYP2D6 gene. vanol) in humans. J. Pharmacol. Exp.
Pharmacogenetics 3:123–30 Ther. 230:28–33
100. Agundez JA, Martinez C, Ladero JM, 110. Wedlund PJ, Aslanian WS, Jacqz E,
by University of Sussex on 11/05/12. For personal use only.

Ledesma MC, Ramos JM, et al. 1994. De- McAllister CB, Branch RA, Wilkin-
brisoquin oxidation genotype and suscep- son GR. 1985. Phenotypic differences
tibility to lung cancer. Clin. Pharmacol. in mephenytoin pharmacokinetis in nor-
Ther. 55:10–14 mal subjects. J. Pharmacol. Exp. Ther.
101. Wang SL, Huang JD, Lai MD, Liu BH, 234:662–69
Lai ML. 1993. Molecular basis of ge- 111. Inaba T, Jurima M, Kalow W. 1986. Fam-
netic variation in debrisoquin hydroxyla- ily studies of mephenytoin hydroxylation
tion in Chinese subjects: polymorphism deficiency. Am. J. Hum. Genet. 38:768–72
in RFLP and DNA sequence of CYP2D6. 112. Ward SA, Goto F, Nakamura K, Jacqz
Clin. Pharmacol. Ther. 53:410–18 E, Wilkinson GR, Branch RA. 1987.
102. Armstrong M, Fairbrother K, Idle JR, S-Mephenytoin 40 -hydroxylase is inher-
Daly AK. 1994. The cytochrome P450 ited as an autosomal-recessive trait in
CYP2D6 allelic variant CYP2D6J and re- Japanese families. Clin. Pharmacol. Ther.
lated polymorphisms in a European pop- 42:96–99
ulation. Pharmacogenetics 4:73–81 113. Wilkinson GR, Guengerich FP, Branch
103. Johansson I, Lundqvist E, Bertilsson L, RA. 1992. Genetic polymorphism of S-
Dahl ML, Sjoqvist F, Ingelman Sundberg mephenytoin hydroxylation. In Pharma-
M. 1993. Inherited amplification of an ac- cogenetics of Drug Metabolism, ed. W
tive gene in the cytochrome P450 CYP2D Kalow. pp. 657–85. New York: Pergamon
locus as a cause of ultrarapid metabolism 114. Goldstein JA, de Morais SMF. 1994. Bio-
of debrisoquine. Proc. Natl. Acad. Sci. chemistry and molecular biology of the
USA 90:11825–29 human CYP2C subfamily. Pharmacoge-
104. Agundez JA, Ledesma MC, Ladero JM, netics 4:285–99
Benitez J. 1995. Prevalence of CYP2D6 115. Küpfer A, Preisig R. 1984. Pharmacoge-
gene duplication and its repercussion on netics of mephenytoin: a new drug hy-
the oxidative phenotype in a white popu- droxylation polymorphism in man. Eur.
lation. Clin. Pharmacol. Ther. 57:265–69 J. Clin. Pharmacol. 26:753–59
105. Aklillu E, Persson I, Bertilsson L, Rod- 116. Sohn D-R, Kusaka M, Ishizaki T, Shin
riques F, Ingelman-Sundberg M. 1996. S-G, Jang I-J, Chiba K. 1992. Inci-
Frequent distribution of ultrarapid me- dence of S-mephenytoin hydroxylation
tabolizers of debrisoquine in a black deficiency in a Korean population and
African population carrying duplicated the interphenotypic differences in di-
and multiduplicated function CYP2D6 al- azepam pharmacokinetics. Clin. Pharma-
leles detection. J. Pharmacol. Exp. Ther. col. Ther. 52:160–69
278:441–46 117. Andersson T, Regårdh C-G, Lou Y-
106. Panserat S, Mura C, Gerard N, Vin- C, Zhang Y, Dahl M-L, Bertilsson L.
cent Viry M, Galteau MM, et al. 1994. 1992. Polymorphic hydroxylation of S-
DNA haplotype-dependent differences in mephenytoin and omeprazole metabolism
the amino acid sequence of debriso- in Caucasian and Chinese subjects. Phar-
quine 4-hydroxylase (CYP2D6): evi- macogenetics. 2:25–31
dence for two major allozymes in exten- 118. Wright JD, Helsby NA, Ward SA. 1995.
P1: SDA/MKV P2: SDA/VKS QC: SDA
February 18, 1997 11:30 Annual Reviews AR027-11 AR27-011N

GENETIC POLYMORPHISM 295

The role of S-mephenytoin hydroxylase hydroxylation. Arch. Biochem. Biophys.


(CYP2C19) in the metabolism of the an- 306:240–45
timalarial biguanides. Br. J. Clin. Phar- 128. Goldstein JA, Faletto MB, Romkes-
macol. 39:441–44 Sparks M, Sullivan T, Raucy JL, et al.
119. Skjelbo E, Mutabingwa TK, Bygbjerg I, 1994. Evidence for a role for 2C19 in
Nielsen KK, Gram LG, Brosen K. 1996. metabolism of S-mephenytoin in humans.
Chloroguanide metabolism in relation to Biochemistry 33:1743–52
the efficacy in malaria prophylaxis and the 129. de Morais SMF, Wilkinson GR, Blais-
S-mephenytoin oxidation in Tanzanians. dell J, Nakamura K, Meyer UA, Goldstein
Clin. Pharmacol. Ther. 59:304–11 JA. 1994. The major defect responsible
120. May DG, Black CM, Olsen NJ, Csuka for the polymorphism of S-mephenytoin
ME, Tanner SB, et al. 1990. Sclero- metabolism in humans. J. Biol. Chem.
derma is associated with differences in 269:15419–22
Annu. Rev. Pharmacol. Toxicol. 1997.37:269-296. Downloaded from www.annualreviews.org

individual routes of drug metabolism. 130. Meehan RR, Gosden JR, Rout D, Has-
A study with dapsone, debrisoquine, tle ND, Friedberg T. 1988. Human cy-
and mephenytoin. Clin. Pharmacol. Ther. tochrome P-450 PB-1: a multigene fam-
48:286–95 ily involved in mephenytoin and steroid
121. Meier UT, Dayer P, Male PJ, Kronbach T, oxidations that maps to chromosome 10.
Meyer UA. 1985. Mephenytoin hydroxy- Am. J. Hum. Genet. 42:26–37
lation polymorphism: characterization of 131. de Morais SMF, Schweikl H, Blaisdell J,
by University of Sussex on 11/05/12. For personal use only.

the enzymatic deficiency in liver micro- Goldstein JA. 1993. Gene structure and
somes of poor metabolizers phentoyped upstream regulatory regions of human
in vivo. Clin. Pharmacol. Ther. 38:488– CYP2C9 and CYP2C18. Biochem. Bio-
94 phys. Res. Commun. 194:194–201
122. Gut J, Meier UT, Catin T, Meyer UA. 132. de Morais SMF, Wilkinson GR, Blaisdell
1986. Mephenytoin-type polymorphism J, Meyer UA, Nakamura K, Goldstein JA.
of drug oxidation: purification and char- 1994. Identification of a new genetic de-
acterization of a human liver cytochrome fect responsible for the polymorphism of
P-450 isozyme catalyzing microsomal S-mephenytoin metabolism in Japanese.
mephenytoin hydroxylation. Biochem. Mol. Pharmacol. 46:594–98
Biophys. Acta 884:435–47 133. Brosen K, de Morais SMF, Meyer UA,
123. Shimada T, Misono KS, Guengerich Goldstein JA. 1995. A multifamily study
FP. 1986. Human liver microsome cy- on the relationship between CYP2C19
tochrome P-450 mephenytoin 4-hydro- genotype and S-mephenytoin oxidation
xylase, a prototype of genetic polymor- phenotype. Pharmacogenetics 5:312–17
phism in oxidative drug metabolism: pu- 134. Evans DAP, Krahn P, Narayanan N. 1995.
rification and characterization ot two sim- The mephenytoin (cytochrome P450 2C
ilar forms involved in the reaction. J. Biol. 19) and dextromethorphan (cytochrome
Chem. 261:909–21 P450 2D6) polymorphisms in Saudi Ara-
124. Ged C, Umbenhauer DR, Bellew TM, bians and Filipinos. Pharmacogenetics
Bork RW, Srivastave P, et al. 1994. Char- 5:64–71
acterization of cDNAs, mRNAs, and pro- 135. Mikus G, Bochner F, Eichelbaum M, Ho-
teins related to human liver microso- rak P, Somogyi AA, Spector S. 1994.
mal cytochrome P-450 (S)-mephenytoin Endogenous codeine and morphine in
40 -hydroxylase. Biochemistry 27:6929– poor and extensive metabolisers of the
40 CYP2D6 (Debrisoquine/Sparteine) poly-
125. Meier UT, Meyer UA. 1987. Genetic po- morphism. J. Pharmacol. Exp. Ther. 268:
lymorphism of human cytochrome P450 546–51
S-mephenytoin-4-hydroxylase. Biochem- 136. Yokota H, Tamura S, Furuya H, Kimura
istry 26:8466–74 S, Watanabe M, et al. 1993. Evi-
126. Romkes M, Faletto MB, Blaisdell JA, dence for a new variant CYP2D6 allele
Raucy JL, Goldstein JA. 1991. Cloning CYP2D6J in a Japanese population asso-
and expression of complementary DNAs ciated with lower in vivo rates of sparteine
for multiple members of the human cy- metabolism. Pharmacogenetics 3:256–
tochrome P450IIC subfamily. Biochem- 63
istry 30:3247–55 137. Broly F, Marez D, Lo Guidice JM, Sab-
127. Wrighton SA, Stevens JC, Becker GW, bagh N, Legrand M, et al. 1995. A non-
VandenBranden M. 1993. Isolation and sense mutation in the cytochrome P450
characterization of human liver cy- CYP2D6 gene identified in a Caucasian
tochrome P4502C19: correlation be- with an enzyme deficiency. Hum. Genet.
tween 2C19 and S-mephenytoin 40 - 96:601–3
P1: SDA/MKV P2: SDA/VKS QC: SDA
February 18, 1997 11:30 Annual Reviews AR027-11 AR27-011N

296 MEYER & ZANGER

138. Brosen K, Nielsen PN, Brusgaard K, splice recognition site associated with the
Gram LF, Skjodt K. 1994. CYP2D6 geno- poor metabolizer phenotype. Pharmaco-
type determination in the Danish popula- genetics 5:305–11
tion. Eur. J. Clin. Pharmacol. 47:221–25 142. Ilett KF, Chiswell GM, Spargo RM, Platt
139. Dahl ML, Yue QY, Roh HK, Johansson E, Minchin RF. 1993. Acetylation phe-
I, Sawe J, et al. 1995. Genetic analysis of notype and genotype in Aboriginal lep-
the CYP2D locus in relation to debriso- rosy patients from the north-west region
quine hydroxylation capacity in Korean, of Western Australia. Pharmacogenetics
Japanese and Chinese subjects. Pharma- 3:264–69
cogenetics 5:159–64 143. Saxena R, Shaw GL, Relling MV, Frame
140. Ferguson RJ, Doll MA, Rustan TD, Gray JN, Moir DT, et al. 1994. Identification of
K, Hein DW. 1994. Cloning expres- a new variant CYP2D6 allele with a sin-
sion, and functional characterization of gle base deletion in exon 3 and its associa-
Annu. Rev. Pharmacol. Toxicol. 1997.37:269-296. Downloaded from www.annualreviews.org

two mutant (NAT2191 and NAT2341/803 ) tion with the poor metabolizer phenotype.
and wild-type human polymorphic N - Hum. Mol. Genet. 3:923–26
acetyltransferase (NAT2) alleles. Drug 144. Weerasuriya K, Jayakody RL, Smith CA,
Metab. Disp. 22:371–376 Wolf CR, Tucker GT, Lennard S. 1994.
141. Marez D, Sabbagh N, Legrand M, Lo Debrisoquine and mephenytoin oxidation
Guidice JM, Boone P, Broly F. 1995. A in Sinhalese: a population study. Br. J.
novel CYP2D6 allele with an abolished Clin.Pharmacol. 38:466–70
by University of Sussex on 11/05/12. For personal use only.

You might also like