Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2013 Mar 1.
Published in final edited form as: Pharmacogenet Genomics. 2012 Mar;22(3):219–228. doi: 10.1097/FPC.0b013e32834eb313

PharmGKB summary: very important pharmacogene information for G6PD

Ellen M McDonagh a, Caroline F Thorn a, José M Bautista c, Ilan Youngster d, Russ B Altman a,b, Teri E Klein a
PMCID: PMC3382019  NIHMSID: NIHMS373489  PMID: 22237549

Introduction

Glucose-6-phosphate dehydrogenase (G6PD) was one of the first genes found to be associated with variable drug response. It is also very polymorphic, with G6PD deficiency found in more than 300 million people worldwide [1]. Here, we provide an overview of G6PD as a very important pharmacogene, and detail genetic variants and haplotypes associated with drug response (Although most G6PD variants are caused by single nucleotide polymorphisms (SNPs) in the coding region of the G6PD gene at the X chromosome, due to the heterogeneity of alleles causing G6PD deficiency; here, we use the term ‘haplotype’ to define the set of linked SNPs in a G6PD variant that are inherited together and that may or may not produce G6PD deficiency). A more in-depth report, with interactive links to individual literature annotations, can be found at http://www.pharmgkb.org/gene/PA28469.

Molecular structure and function

The G6PD enzyme is conserved throughout evolution, with human G6PD sharing approximately 93% amino acid identity with rat and 37% with Escherichia coli [2,3]. G6PD is encoded by a gene on the X chromosome (Xq28) [4-6], contrary to an early report describing the G6PD enzyme as a fusion protein encoded by genes on chromosomes 6 and X [7]. The G6PD gene is around 18 kb in length and consists of 13 exons and 12 introns, and was originally cloned in 1986 [8-10].

The promoter region of the G6PD gene has some sequence homology with other housekeeping genes, and contains elements for tissue-specific expression that regulate transcription in response to oxidative stress, hormones, nutrients, and growth factors [3,10]. Alternative transcriptional start sites and mRNA splice variants have been described [8,10-13]. The G6PD mature peptide of 514 amino acids in length (59 kDa) is active as a dimer or tetramer, and one molecule of NADP + is bound per protein subunit [13-17]. The binding of NADP + is thought to be integral to the enzyme’s stability, and thus, its function, as point mutations close to the NADP + and dimer interface result in severe G6PD deficiency, revealed by the crystal structure of the Canton variant [14] and site-directed mutagenesis studies [18].

G6PD is a cytoplasmic protein and has two main roles within the cell: the production of NADPH and ribose-5-phosphate (reviewed in [19,20]). Both are synthesized by steps within the pentose phosphate pathway (PPP), also known as the hexose monophosphate shunt (for example [21]); reviewed in [19,22]). NADPH is essential to maintain the redox state of the cell and relieves oxidative stress through the reduction of glutathione, which in turn reduces hydrogen peroxide and oxidative free radicals (reviewed in [3,19,20,22,23]). Ribose-5-phosphate is required for glycolysis and for DNA and RNA biosynthesis (reviewed in [3,19,20,22,23]). Alternative pathways can be utilized for the biosynthesis of nucleic acids, but G6PD is essential for a cell’s ability to cope with oxidative stress [24]. Tumor suppressor protein p53 has been shown to regulate the PPP by binding to G6PD, preventing dimer formation and thus NADP + binding, inhibiting NADPH production [25]. Several p53 mutants associated with tumors were shown to lack this inhibitory property, and therefore disregulation of G6PD in cancer cells may result in increased cell growth through unregulated glucose biosynthesis and the production of NADPH [20,25].

G6PD is expressed in all cells, but its role is particularly important in red blood cells (RBCs), which do not have mitochondria and are therefore dependent on G6PD as the only source of NADPH to relieve oxidative stress and protect the hemoglobin β chain from oxidation (reviewed in [19,22,23]). In addition, enzyme levels decrease during the RBC lifespan [23]. When the required levels of NADPH cannot be maintained, the amount of reduced glutathione decreases, resulting in oxidative damage, which can ultimately lead to lysis of RBCs (reviewed in [1,19,22]). Under normal conditions, G6PD activity in RBCs is only around 2% of its capacity, inhibited through a negative feedback loop with NADPH (reviewed in [23,26]). However, under oxidative pressure, oxidation of NADPH releases the inhibitory effect, and G6PD enzyme activity increases, enabling enhanced reducing activity to deal with the additional stress (reviewed in [23,26]). In G6PD-deficient RBCs, where enzyme activity can be below 10% of the normal value, homeostasis can be maintained and most G6PD-deficient individuals remain asymptomatic [23]. However, the deficiency becomes apparent under oxidative stress conditions when an increased demand in NADP/NADPH turnover cannot be met (reviewed by the WHO Working Group [23]).

G6PD as an important pharmacogene

We have known for more than 2000 years that the ingestion of fava beans can have adverse consequences in some individuals, and could indeed be why Pythagoras imposed abstinence from beans among his followers [27,28]. However, it was not until the 20th century that a deficiency in the G6PD enzyme was discovered to be the underlying cause of ‘Favism’, and the connection that agents other than fava beans can cause similar adverse events in G6PD-deficient individuals (discussed in Cappellini and colleagues [22,29]). In the 1950s, it was observed that a subset of African-American soldiers were more likely to develop an adverse reaction to the antimalarial drug primaquine compared with their Caucasian counterparts [30,31]. This susceptibility to primaquine-induced intravascular hemolysis led to the discovery of a deficiency in G6PD enzyme activity in RBCs [32].

More than 400 variations of the G6PD enzyme have now been described on the basis of clinical manifestations and biochemical properties, and G6PD deficiency is the most prevalent enzyme deficiency in the world (reviewed in [19,22,33-35]), affecting an estimated 4.9% of the world’s population (more than 300 million people) [1]. Polymorphic variants in G6PD are those of significant frequencies (1–70%) in specific human populations: these fall into World Health Organization (WHO) classes II and III (see Table 1) [19,23,34,46]. Different polymorphic variants have arisen in different geographical areas; for example, the Canton variant is predominantly found in Chinese and South-East Asian populations [35,40,47,48], reviewed in [19]. Higher population densities of humans resulting from the development of agriculture may have facilitated malaria endemics, and introduced a selective pressure for the spread of G6PD variants (reviewed in [19]). Nevertheless, G6PD-deficient variants can also occur sporadically due to de-novo mutations (reviewed in [19]). Most of the genetic variants tend to be single-point mutations, and the lack of large or out-of-frame deletions may indicate that the total absence of enzyme activity is fatal [19,23,33,34,38]. In-frame deletions are usually associated with the most severe clinical manifestations (class I; reviewed in [19,33,34]), such as the novel G6PD Tondela variant, a six-amino acid deletion identified in a heterozygous woman with chronic hemolytic anemia (see Table 1) [36].

Table 1.

Classification of G6PD variants

World Health Organization class Enzyme activity Associated phenotype Variant example Genotyping reference
I Severe deficiency Congenital nonspherocytic hemolytic anemia (CNSHA) Tondela [36]
Palermo [37]
II < 10% severely deficient Risk of acute hemolytic anemia Mediterranean [38]
Canton [39,40]
III 10–60% Moderate deficiency Risk of acute hemolytic anemia A- Haplotype [41,42]
Asahi [42]
IV 60–150% Normal activity No clinical manifestations B (wild type)
A [43]
V > 150% Enhanced activity Hektoen [44,45]

Table based on [22,23,33,34,46] reviews.

The mechanism underlying the G6PD-deficient phenotype may vary depending on the location of the mutation in the enzyme’s three-dimensional structure, and include alterations to protein assembly, dimer formation and stability, interaction with substrates, and protein turnover (reviewed in [49]). Studies suggest that most G6PD variants result in G6PD enzyme instability (reviewed in [19,46]), and variants that cause the most severe deficiency (resulting in congenital nonspherocytic hemolytic anemia [CNSHA]) are found predominantly at or near the dimer interface of the G6PD protein in exon 10 [50], although not exclusively. Different gene variants can confer similar effects on enzyme function and clinical manifestations, and conversely, the same genetic variation can result in different molecular and clinical phenotypes [33,35,51]. Therefore, G6PD variants based on enzyme biochemistry may have been characterized differently and yet have the same underlying genetic mutation [41,52,53]. The term ‘haplotype’ is used in this review to define a set of linked alleles in G6PD that are inherited together. The B haplotype is considered the ‘wild-type’ precursor sequence, as alignment of human and chimpanzee DNA sequences have shown [41]. Variants that differ from the B wild type are often named with the region where the population in which the variant was found [35,54]. Although attempts have been made to standardize nomenclature since 1967 [54], there are still instances where haplotype names have been used to refer to more than one set of variations (see text on haplotype A-). It should be noted that many studies do not screen the whole G6PD gene, and therefore some genetic variants may be unreported and may be a factor underlying differences in biochemical properties.

Exogenous agents can trigger hemolytic anemia in G6PD-deficient individuals by inducing oxidative stress in RBCs (reviewed in [19,22,55]). These include certain food items, therapeutic drugs, infections, and exposure to chemicals (e.g. hair dye containing naphthol) [19,22,37,55]. G6PD variants have been classified into five WHO categories according to the severity of clinical manifestation resulting from the genotype (see Table 1), with classes II and III being the most common type of polymorphic G6PD-deficient variant [23,33]. Due to lower RBC G6PD activity, patients carrying class I sporadic variants (associated with CNSHA) are highly susceptible to hemolytic anemia caused by the same drugs that can induce adverse reactions in carriers of polymorphic G6PD variants (reviewed in [19]).

Testing for G6PD deficiency

G6PD variants that result in enzyme deficiency confer a G6PD-deficient phenotype in hemizygous men (with one copy of the G6PD gene) and homozygous women (for example [56,57]). To diagnose a phenotype of G6PD deficiency in heterozygous women is more difficult, as the extent of enzyme deficiency activity can vary considerably within heterozygous individuals, due to X-linked mosaicism [34,57,58]. This pattern of gene inactivation is random; therefore, female heterozygotes will have G6PD-deficient RBCs combined with those expressing normal G6PD activity, and the population sizes of these cells can vary from 50 : 50 to minimal levels or a majority of G6PD-deficient cells [23,34,58]. Genotyping is therefore essential to establish heterozygosity in women; however, this can make prediction of drug response difficult without phenotypic information of G6PD enzyme activity levels. For example, 75% of women genetically heterozygous for the Mediterranean variant had normal G6PD activity, whereas 25% were enzyme deficient, as assessed by a colorimetric test [56]. Testing for both genotype and enzyme function is the ideal method; however, due to various factors, including the impracticalities and costs of genotyping in the field, many studies and clinics have solely tested G6PD enzyme activity, making the association of causative variants difficult. In a meta-analysis of 280 studies, less than 8% used DNA analysis to assess G6PD deficiency [1].

G6PD and therapeutic drug response

The WHO recommends testing of drugs to predict the risk of hemolysis in G6PD-deficient individuals if the drugs are to be prescribed in areas of high prevalence of G6PD deficiency [23]. As a consequence of adverse reactions in individuals with G6PD deficiency, the Food and Drug Administration (FDA) has introduced warnings or precautions on the drug labeling of primaquine, chloroquine, dapsone, rasburicase, avandaryl tablets (glimepiride + rosiglitazone maleate), and glucovance tablets (metformin + glibenclamide) [59]. These highlight the possible risk of hemolytic anemia in G6PD-deficient individuals upon drug intake [59]. Numerous factors can contribute to drug-induced hemolytic anemia in G6PD-deficient individuals, including high dosage, other drugs taken in combination, concurrent infections, and other genetic variants, as discussed in Cappellini and colleagues [22,61,62]. Therefore, it may be that drugs (including aspirin, vitamin C, and chloroquine) reported to cause hemolysis can actually be taken safely by G6PD-deficient individuals. These drugs, however, should be administered with caution, especially in combination with other drugs or at high doses, potentially with monitoring of RBC or hemoglobin levels [61,62]. Readdressing the safety of drugs in these individuals could make effective therapeutics available (as discussed in Youngster et al. [62]). More information and comprehensive advice on unsafe drugs in G6PD-deficient individuals can be found at www.favism.org [63]. Several drug subsets and the possible consequences of drug intake for G6PD-deficient individuals are outlined below.

G6PD deficiency and antimalarial drugs

It is hypothesized that variants associated with G6PD deficiency have remained prevalent in the human population due to positive selective pressures, in particular resistance to uncomplicated and severe malaria (reviewed in [19,22]). At the same time, G6PD deficiency confers susceptibility to hemolytic anemia triggered by some antimalarial drug treatments [1]. The prevalence of G6PD deficiency in malaria endemic regions where antimalarial drugs are required is an important public health issue [1,64]. Hemolysis is also a phenotype induced by malaria infection; therefore, distinguishing a disease from a drug-induced effect may be difficult (as discussed in Alloueche and colleagues [65,66]). In areas where malaria is endemic, the financial and practical costs of G6PD screening are important considerations [22,56]. However, the wide-scale drug-based eradication of malaria is likely to require G6PD deficiency testing [64], and thus the application of pharmacogenomics in the treatment of G6PD-deficient individuals.

Race-specific differences in sensitivity to hemolytic anemia after treatment with aminoquinolines have been reported since the 1920s (reviewed in [29]). In 1956, a deficiency in the G6PD enzyme was determined as the underlying cause of ‘primaquine sensitivity’ [30-32]. Metabolism of primaquine mediated by CYP proteins is thought to be one of the mechanisms behind the drug sensitivity, as the resulting metabolites induce the formation of methemoglobin and reactive oxygen intermediates [67]. Alternatives to primaquine have been sought for use in areas where G6PD deficiency is prevalent. However, unfortunately, numerous antimalarial drugs can also induce hemolysis in G6PD-deficient individuals, including dapsone which has FDA drug labeling precautions [59,66,68,69] (reviewed in [62]). The WHO released guidelines after a technical consultation in 2004 for the use of Lapdap (a combination of chloroproguanil and dapsone) due to safety concerns in G6PD-deficient individuals, recommending use only if malaria infection is confirmed, testing for G6PD deficiency and avoiding drug treatment in these individuals [70], as discussed in Luzzatto et al. [71].

FDA labeling of chloroquine advises that caution should be taken when administering this drug to G6PD-deficient individuals [59], although it is not contra-indicated. Treatment of normal RBCs with chloroquine in vitro does not boost the PPP, and does not reduce the survival of G6PD-deficient RBCs transferred into wild-type recipients [72,73]. However, a high dose of chloroquine (600 mg) for the prophylaxis of malaria was reported to induce severe hemolytic anemia in 50 soldiers, all G6PD deficient [74]. G6PD deficiency may also increase the risk of side affects such as pruritus induced by chloroquine [75]. Chloroquine treatment combined with primaquine for Plasmodium vivax infection has been associated with significantly decreased hemocrit levels [76], although the development of hemolysis in G6PD-deficient individuals has been affiliated with primaquine administration rather than chloroquine [77]. Severe hemolysis has been reported in several children treated with combinations of chloroquine, chloramphenicol, aspirin, and primaquine [78]. In contrast, chloroquine treatment combined with methylene blue or elubaquine does not induce severe adverse effects in individuals with G6PD deficiency [76,79,80]. To conclude, the safety of chloroquine in G6PD-deficient individuals is unclear, and seems to depend on numerous factors including dosage, concurrent drugs and infections, as discussed in Beutler and colleagues [61,62].

When comparing G6PD-deficient and G6PD ‘normal’ children with uncomplicated malaria infection in a trial of artesunate and amodiaquine or artemether-lumefantrine (also known as artemisinin-based combination therapy ACT), no significant differences in adverse events were observed, indicating that these drug combinations are a promising alternative (see PharmGKB Artemisinin and Derivatives Pharmacokinetics Pathway http://www.pharmgkb.org/pathway/PA165378192 and PharmGKB Amodiaquine Pharmacokinetics Pathway http://www.pharmgkb.org/pathway/PA165815256) [81].

G6PD deficiency and cancer therapeutics

Several agents involved in the treatment of cancer patients have the potential to result in severe adverse side effects in G6PD-deficient individuals, due to induction of oxidative stress in RBCs. Carmustine [3-bis(chloroethyl)-1-nitrosourea (BCNU)] treatment results in a deficiency in glutathione reductase in erythrocytes, platelets, and leukocytes [21,82]. This lowers the levels of reduced glutathione and leads to insufficient removal of hydrogen peroxide and susceptibility to oxidative hemolysis, particularly in RBCs that are G6PD deficient [21,82]. Doxorubicin (adriamycin) stimulates the PPP, but to a lower extent in G6PD-deficient RBCs, and results in oxidative stress through the accumulation of hydrogen peroxide due to an inability to increase G6PD activity [see the PharmGKB Doxorubicin Pathway (Cancer Cell), Pharmacodynamics http://www.pharmgkb.org/pathway/PA165292163] [21,83]. This response can also be mirrored in ‘normal’ RBCs when cells are pretreated with carmustine before doxorubicin treatment, resulting in diminished glutathione stability and enhanced susceptibility to oxidative stress [21]. A number of cases of methemoglobinemia and acute hemolysis after treatment with rasburicase in G6PD-deficient individuals have been described [84-86]. The FDA labeling has contraindicated rasburicase for G6PD-deficient individuals [59,60].

The anti-leukemia drug daunorubicin is metabolized into the less potent form daunorubinol, a process dependent on NADPH via G6PD [87]. This biotransformation was considerably reduced in RBCs from A- or Mediterranean G6PD-deficient individuals, and thus may have implications in the clinic, raising the issue of whether or not the drug is more effective in these individuals due to prolonged exposure to the more active form and possible toxicity concerns [87]. NADPH inhibitors, including primaquine aldehyde, were also shown to reduce daunorubicin metabolism and daunorubinol appearance [87].

G6PD deficiency and Aspirin

A child carrying the Mediterranean variant who was diagnosed with systemic arthritis and prescribed a daily dose of 100 mg/kg of aspirin subsequently developed severe hemolytic anemia, with no sign of viral or bacterial infection [88]. Hemolysis was reported in a male administered 1.5 g of aspirin for a fever (likely a viral infection), who was later found to be G6PD deficient and had a family history of Jaundice [89]. However, in 22 healthy G6PD-deficient individuals, normal therapeutic doses of aspirin for 4 days (50 mg/kg daily) had no effect on the RBC count or the hemoglobin levels [90]. Therefore, the effect of aspirin on G6PD-deficient individuals may be dependent on the variant type and pre-existing clinical conditions (such as an infection or inflammatory disease) [91]. Dosage is also likely to contribute, as exemplified in the above studies and demonstrated by in-vitro studies in which higher levels of aspirin were required to observe a drop in reduced glutathione levels in the blood from sensitive patients [89].

G6PD deficiency and diabetes mellitus treatment

Glibenclamide (glyburide) has been shown to induce acute hemolysis in diabetic patients carrying the A- haplotype or the Mediterranean variant [92,93], although it should be noted that these seem to be sporadic case studies, as discussed in Youngster et al. [62]. FDA labeling of glucovance tablets, which contain glyburide and metformin HCl, cautions use in G6PD-deficient individuals as sulfonylurea agents can result in hemolytic anemia, and advise using a non-sulfonylurea alternative [60].

G6PD variants and haplotypes

A selection of the most studied G6PD variants and known haplotypes are discussed below, and their association with drug response is summarized in Table 2.

Table 2.

Polymorphic G6PD variants and haplotypes associated with drug response

Variants genotyped Drug or treatment Associated response Referencea
A-Haplotype
 202A/376G (rs1050828 and rs1050829) Glibenclamide Acute hemolysis [93] (case study)
 202A (rs1050828) Sulfadoxine-pyrimethamine and artemisinin and primaquine Increased risk of developing moderate anemia [66]b n = 562 total population genotyped, 8.4% heterozygous, 3.9% homozygous/ hemizygous.
 202A (rs1050828) Chlorproguanil-dapsone (CD) Increased risk of a drop in hemoglobin levels, compared with sulfadoxine-pyrimethamine (SP) treatment [65] (n = 1480 total study group treated with CD, n = 370 treated with SP. n = 237 treated with CD had a > 20 g/L decrease in hemoglobin and of these 35% were carriers of this variant, defined as G6PD deficient), compared with 24% treated with SP
 202A (rs1050828) Chlorproguanil-dapsone-artesunate Severe decreases in hemoglobin levels and increased risk of blood transfusion [69] 13% were carriers of this variant and defined as A- G6PD deficient, in n = 343 total genotyped.
 202A (rs1050828) Sulphadoxine-pyrimethamine coadministered with amodiaquine Increased risk of requiring a blood transfusion [69] 11% were carriers of this variant and defined as A- G6PD deficient in n = 359 total genotyped.
 202A (rs1050828) Chlorproguanil-dapsone-artesunate Severe reduction in hemoglobin levels and an increased risk of requiring a blood transfusion [68] n = 800 genotyped. G6PD-deficient individuals were defined as A- hemizygous men (17% of n = 388), and homozygous A-/ A- women (4% of n = 412)
 376G (rs1050829)
 680T (rs137852328)
 968C (rs76723693)
 542G (rs5030872)
 NS Rasburicase Hemolytic anemia [84] (case study)
 NS Daunorubicin Reduced drug metabolism [87] (in vitro)
 NS Methylene blue Hemolysis in an individual with methemoglobinemia [94] (case study)
 NS Vitamin C (high dose of 80 g intravenously, 2 days) Hemolysis [95] (case study)
Mediterranean variant
 563T (rs5030868) Glibenclamide Acute hemolysis [92] (case study)
 NS Aspirin (high dose of 100 mg/kg daily) Severe hemolytic anemia in a child with systemic arthritis [88] (case study)
 NS Daunorubicin Reduced drug metabolism [87] (in vitro)
 NS Rasburicase Severe G6PD deficiency was revealed during treatment [86] (case study)

NS, not specified.

a

For each reference, details of whether the study was a single case study, or total study numbers and percentage of individuals carrying the indicated G6PD allele, are given in the reference column – see www.pharmgkb.org website for further annotations of these studies. The G6PD gene is found on the minus chromosomal strand. Please note that for standardization, the PharmGKB presents all allele base pairs on the positive chromosomal strand, and therefore, the alleles on the website will differ (in a complementary manner) from those in this review that are given on the minus strand as reported in the literature.

b

Please note that in this study, heterozygous 202A individuals were considered as G6PD A, and hemizygous/ homozygous 202A individuals were classified as G6PD A-.

Canton (rs72554665)

The Canton variant was first identified in 1966 [96] and is encoded by a G to T nucleotide change at position 1376 (NM_001042351.1:c), resulting in the amino acid change Arg 459 Leu (NP_001035810.1:p.) [35,39]. The same genetic variant is also known as Taiwan-Hakka, Gifu-like, and Agrigento-like [34,35], all G6PD enzyme variants that have been reported to display differences in thermostability, enzyme activity, and substrate utilization but share the same nucleotide change 1376T [40]. G6PD Canton is commonly found in China, with one study reporting incidence of the variant in 2% of men and 0.9% of women, and is also found in South-East Asia [35,40,47,48]. G6PD Canton is a category II variant, associated with hemolytic anemia and Favism, resulting in less than 10% enzyme activity and reduced thermostability (see Table 1) [33,40,97,98]. This variant was used to determine the first crystal structure of G6PD [14].

A (rs1050829)

The A variant is an A > G change at nucleotide position 376, resulting in an Asn to Asp missense mutation at amino acid 126 [35,38,43,99]. The A variant results in normal thermostability and enzyme activity levels, between 60 and 150%, and is a class IV phenotype according to the WHO classification, not associated with acute hemolytic anemia (see Table 1) [23,38,57]. The polymorphic A variant is found at frequencies between 0.1 and 0.34 in Sub-Saharan African populations (as reviewed in [100]) and is part of the A- haplotype.

A- haplotype

The A- haplotype is a combination of two SNPs, both of which confer missense mutations in the protein sequence [12,38,41]. This is complicated by the fact that what was previously defined as the A- G6PD enzyme form by biochemical properties is encoded by three possible different haplotypes (a set of linked SNP alleles that are inherited together) [35,41]. Each is made up of two variants: the A variant coupled with a second SNP to yield the A- haplotype [35,41]. Here, to avoid confusion, we have distinguished the three different haplotypes according to the second variant nucleotide position:

  1. A-202A/376G: is a G > A nucleotide change at position 202 (rs1050828) resulting in the missense mutation Val to Met at amino acid 68, combined with the A variant.

  2. A-680T/376G: is a G > T nucleotide change at position 680 (rs137852328) resulting in the missense mutation Arg to Leu at amino acid 227, combined with the A variant.

  3. A-968C/376G: is a T > C nucleotide change at position 968 (rs76723693) resulting in the missense mutation Leu to Pro at amino acid 323, combined with the A variant.

Haplotype A-202A/376G is found at frequencies up to 0.24 in African populations, and has also been described as Distrito Federal, Matera, Betica, Castilla, Alabama, Tepic, Ferrara, Laghouat, and Kabyle, found in populations from around the world [35,38,41,51,57,100,101]. The A-968C/376G haplotype has also been named Betica, Guantanamo, and Selma [35,41].

The Santamaria haplotype (542T/376G) is an A > T nucleotide change at position 542 (rs5030872) resulting in a missense mutation Asp to Val at amino acid 181, combined with the A variant [102,103]. Although associated with G6PD deficiency, the Asp to Val does not result in a change of charge and hence the electrophoretic properties of the Santamaria protein are not the same as those defined for the G6PD A- enzyme [102,103]; therefore, this is not included here under the A-haplotype description (as discussed in Beutler et al. [102]). The Sierra Leone haplotype (311A/376G) is another example where the A variant is found to be combined with a second variant, resulting in G6PD deficiency [104].

The G6PD A- enzyme has a class III phenotype according to the WHO classification, conferring moderate enzyme deficiency of between 10 and 60% activity, and has been associated with hemolytic anemia [23,34,57]. Because of differences in both enzyme kinetics and genetic sequence, it is predicted that the G6PD A- haplotype derived from the A variant, which originally derived from the B wild type [12,105,106]. The A- enzyme is more likely to be found in a tetramer form than the A or B enzyme, which may explain a faster rate of enzyme inactivation [105]. The synergistic interaction of the two mutations spatially close together in the three-dimensional structure of the A- enzyme has been shown to result in defective folding of the subunits, which decreases the stability and thus results in lower levels active enzyme within RBCs [107]. The B, A, and A- G6PD enzymes also have differences in the Km for glucose-6-phosphate (G6P) at pH7.2 (intracellular erythrocyte conditions) [105].

Numerous associations with G6PD A- and drug response have been described, although the underlying genotype may remain unknown (see Table 2). Some studies do not mention genotyping and only refer to the A- enzyme phenotype, as determined by electrophoretic or biochemical properties. In these cases, we have used the terms A- G6PD-deficient individuals or A- enzyme. In studies which genotype for both variants of the A- haplotype, we indicate the location of the second variation if all three were not screened for. Other studies do not genotype at both loci, screening for only the 202A variant, and yet use the description A- due to the strong linkage disequilibrium that exists between this and the 376G variant in African populations [106] (see Table 2).

The metabolism of daunorubicin into the less potent form daunorubinol was considerably diminished in RBCs from A- or Mediterranean G6PD-deficient individuals [87]. Methylene blue treatment of toxin-induced methemoglobinemia resulted in hemolysis in a patient carrying the A- G6PD deficiency [94]. Large amounts of vitamin C were reported to induce hemolysis in an A- male individual [95]. A woman diagnosed with type 2 diabetes developed acute hemolysis after prescription of glibenclamide and metformin, and was subsequently genotyped homozygous for the A-202A/376G haplotype [93]. Glibenclamide treatment was therefore ended, and hemoglobin levels rose on metformin treatment alone [93]. Combined chlorproguanil-dapsone-artesunate (CDA) treatment is unsafe in carriers of the A- haplotype, resulting in severe reductions in the hemoglobin level, and an increased risk of requiring a blood transfusion [ 68].

Studies revealing strong linkage disequilibrium between the SNPs 202A and 376G, and haplotype analysis with other polymorphic sites, provide evidence that the G6PD A- variant occurred more recently in evolution, originating from the A variant and expanding in African populations due to selection [12,100,105,106,108]. The pharmacogenetic studies described below only genotype for the 202A variant, but report individuals as carriers of the G6PD A- enzyme (also see Table 2). This is due to the evidence which supports the assumption that in African populations those with the 202A variant also carry the A variant (376G), and thus are defined as having the A- haplotype. A study of artemisinin-based combination therapy (ACT) combined with primaquine revealed that carriers of the G6PD A- haplotype are more likely to have a significant reduction in hemoglobin and are at a higher risk of developing moderate anemia compared with those with the wild-type genotype, although a significant reduction in hemoglobin was found in all individuals taking the combination therapy compared to placebo control [66]. Chlorproguanil-dapsone (CD) combination therapy is more efficacious than sulfadoxine-pyrimethamine (SP) treatment, resulting in fewer new malaria infections; however, G6PD-deficient individuals are at a higher risk of a drop in hemoglobin levels when taking the CD combination than SP [65]. Combined chlorproguanil-dapsone-artesunate (CDA) treatment is unsafe in A- G6PD-deficient individuals, resulting in severe decreases in hemoglobin levels and an increased risk of a blood transfusion compared to wild type [69]. A- G6PD-deficient individuals are also more likely to require a blood transfusion when treated with antimalarial SP coadministered with amodiaquine [69].

Although A- G6PD deficiency does not confer full protection from Plasmodium falciparum (P. falciparum) malaria infection, several studies have shown a protective effect of A- G6PD from severe malaria in hemizygous men [109,110]. Contradictory results exist for females carrying A- G6PD, with some studies finding protection from severe and mild malaria in heterozygous females and lower levels of P. falciparum parasitaemia [110,111], and yet other studies find no protection from severe malaria [109]. No significant association with protection from severe malaria was found in carriers of the A-202A/376G haplotype; however, when pooled together with data from carriers of A-968C/376G and the Santamaria haplotype, a significant association with reduced risk of severe P. falciparum malaria in both men and women was observed [112]. Therefore, the methodology and the extent of alleles examined likely contribute to these mixed reports [112,113]. Increased carbonylation of numerous host membrane proteins involved in oxidative stress and transport of parasite proteins is seen in malaria-infected A- deficient erythrocytes compared with RBCs from G6PD B individuals, and may underlie the mechanism that is thought to confer protection [114].

Asahi (rs1050828)

The Asahi variant is a G > A nucleotide change at position 202, which confers a missense mutation Val to Met in the G6PD protein sequence at position 68 [35,42]. This was originally identified in a Japanese boy admitted to the Asahi hospital with jaundice and anemia, who had been previously diagnosed with neonatal jaundice [42]. Because of its association with acute hemolysis, the Asahi variant is classified as a class III phenotype (see Table 1) [42]. Sequencing of the whole G6PD gene in this child also revealed a single base pair deletion in intron 5 [42], which could also contribute to the risk of hemolysis. Asahi has also been reported to occur in Southern China (a single female in the cohort) [47]. Asahi is thought to be a rare variant, occurring independently from the A- haplotype common in Africa, as the A variant and other polymorphic sites found to be in linkage disequilibrium with A-202A/376G were not found in this individual [42,106].

Mediterranean (rs5030868)

The Mediterranean variant is a C > Tchange at position 563 in exon 6, resulting in an amino acid change Ser to Phe at position 188 [38]. This confers less than 10% residual enzyme activity and is associated with acute hemolytic anemia, and therefore has a class II WHO classification (see Table 1). Compared with the G6PD B wild type, the Mediterranean variant has higher affinity for G6P and reduced thermostability [38]. This variant is also known as Dallas, Birmingham, Sassari, Cagliari, or Panama, on the basis of the place of discovery [35,52,101,115]. An association with protection against Plasmodium vivax malaria infection in carriers of the Mediterranean variant has been reported [56], and may be one of the contributing factors why the variant has been retained in areas correlating with endemic regions in human history (as discussed in Leslie et al. [56]).

This variant is associated with acute hemolytic anemia in response to therapeutic drugs and fava beans (reviewed in [22]). Sporadic case studies have shown glibenclamide-induced acute hemolysis [92] and severe hemolytic anemia in response to high doses of aspirin [88] in individuals with the Mediterranean variant. This variant has also been shown to have a role in pharmacokinetics of anthracyclines in vitro, likely due to NADPH requirements of the metabolizing enzymes (see PharmGKB Doxorubicin Pathway, Pharmacokinetics http://www.pharmgkb.org/pathway/PA165292177) [83]. The metabolism of daunorubicin to daunorubinol was considerably diminished in erythrocytes from A- or Mediterranean G6PD-deficient individuals potentially influencing drug efficacy and toxicity [87].

Mediterranean haplotype

In three separate studies, a second SNP was found in all individuals expressing the Mediterranean variant, conferring a C > T nucleotide change at position 1311 which results in a silent mutation of Tyr to Tyr at amino acid position 437 (rs2230037) [52,101,116]. The Mediterranean variant 563T may have arisen in an individual carrying the silent 1311T SNP (as discussed in Vives-Corrons et al. [116]). The G6PD563T/1311T haplotype is found in populations of Mediterranean and Middle Eastern countries, but not in Asian individuals with the Mediterranean variant, suggesting the Mediterranean variant 563T arose independently in the two geographical locations [115].

Conclusion

G6PD is a historically important pharmacogene, with variants conferring G6PD enzyme deficiency found in around 4.9% of the world’s population [1]. The essential role of the enzyme in preventing oxidative stress in RBCs means that numerous drugs have been reported to induce hemolytic anemia in G6PD-deficient individuals. G6PD variants have also been reported to influence pharmaco-kinetics, by a requirement for NADPH in drug metabolism. However, the wide variety of different G6PD genetic variants, enzyme phenotypes, and clinical manifestations has made it difficult to determine the precise influence of individual genetic variants on the drug response. Confounding factors also include a lack of genotyping data in many studies, and the influence of other exogenous agents such as infection. Teasing apart the genotype–phenotype–drug interaction by applying modern technologies as well as examining more extensive haplotypes could help identify possible effective therapeutics in certain populations, as well as avoid adverse reactions in others.

Acknowledgments

The authors thank Li Gong, Joan M. Hebert, and Michelle Whirl-Carrillo for critical reading of the manuscript. They also thank Tom Vulliamy for discussions on G6PD variants. This work is supported by the National Institutes of Health/National Institutes of General Medical Sciences (R24 GM61374), the Spanish Ministry of Science and Innovation [Ministerio de Ciencia e Innovación BIO2010-17039], and from the Santander-Universidad Complutense de Madrid Consolidated Research Groups [920267].

Footnotes

Conflicts of interest

R.B.A. serves as a founder and consultant for Personalis and a consultant to 23andMe. There are no conflicts of interest.

References

  • 1.Nkhoma ET, Poole C, Vannappagari V, Hall SA, Beutler E. The global prevalence of glucose-6-phosphate dehydrogenase deficiency: a systematic review and meta-analysis. Blood Cells Mol Dis. 2009;42:267–278. doi: 10.1016/j.bcmd.2008.12.005. [DOI] [PubMed] [Google Scholar]
  • 2.Jeffery J, Barros-Soderling J, Murray L, Wood I, Hansen R, Szepesi B, Jornvall H. Glucose-6-phosphate dehydrogenase: characteristics revealed by the rat liver enzyme structure. Eur J Biochem. 1989;186:551–556. doi: 10.1111/j.1432-1033.1989.tb15242.x. [DOI] [PubMed] [Google Scholar]
  • 3.Kletzien RF, Harris PK, Foellmi LA. Glucose-6-phosphate dehydrogenase: a “housekeeping” enzyme subject to tissue-specific regulation by hormones, nutrients, and oxidant stress. FASEB J. 1994;8:174–181. doi: 10.1096/fasebj.8.2.8119488. [DOI] [PubMed] [Google Scholar]
  • 4.Pai GS, Sprenkle JA, Do TT, Mareni CE, Migeon BR. Localization of loci for hypoxanthine phosphoribosyltransferase and glucose-6-phosphate dehydrogenase and biochemical evidence of nonrandom X chromosome expression from studies of a human X-autosome translocation. Proc Natl Acad Sci USA. 1980;77:2810–2813. doi: 10.1073/pnas.77.5.2810. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Beutler E, Gelbart T, Kuhl W. Human red cell glucose-6-phosphate dehydrogenase: all active enzyme has sequence predicted by the X chromosome-encoded cDNA. Cell. 1990;62:7–9. doi: 10.1016/0092-8674(90)90231-3. [DOI] [PubMed] [Google Scholar]
  • 6.Mason PJ, Bautista JM, Vulliamy TJ, Turner N, Luzzatto L. Human red cell glucose-6-phosphate dehydrogenase is encoded only on the X chromosome. Cell. 1990;62:9–10. doi: 10.1016/0092-8674(90)90232-4. [DOI] [PubMed] [Google Scholar]
  • 7.Kanno H, Huang IY, Kan YW, Yoshida A. Two structural genes on different chromosomes are required for encoding the major subunit of human red cell glucose-6-phosphate dehydrogenase. Cell. 1989;58:595–606. doi: 10.1016/0092-8674(89)90440-6. [DOI] [PubMed] [Google Scholar]
  • 8.Persico MG, Viglietto G, Martini G, Toniolo D, Paonessa G, Moscatelli C, et al. Isolation of human glucose-6-phosphate dehydrogenase (G6PD) cDNA clones: primary structure of the protein and unusual 5′ non-coding region. Nucleic Acids Res. 1986;14:2511–2522. doi: 10.1093/nar/14.6.2511. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Takizawa T, Huang IY, Ikuta T, Yoshida A. Human glucose-6-phosphate dehydrogenase: primary structure and cDNA cloning. Proc Natl Acad Sci USA. 1986;83:4157–4161. doi: 10.1073/pnas.83.12.4157. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Martini G, Toniolo D, Vulliamy T, Luzzatto L, Dono R, Viglietto G, et al. Structural analysis of the X-linked gene encoding human glucose 6-phosphate dehydrogenase. EMBO J. 1986;5:1849–1855. doi: 10.1002/j.1460-2075.1986.tb04436.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Hirono A, Beutler E. Alternative splicing of human glucose-6-phosphate dehydrogenase messenger RNA in different tissues. J Clin Invest. 1989;83:343–346. doi: 10.1172/JCI113881. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Hirono A, Beutler E. Molecular cloning and nucleotide sequence of cDNA for human glucose-6-phosphate dehydrogenase variant A(-) Proc Natl Acad Sci USA. 1988;85:3951–3954. doi: 10.1073/pnas.85.11.3951. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Kanno H, Kondoh T, Yoshida A. 5′ Structure and expression of human glucose-6-phosphate dehydrogenase mRNA. DNA Cell Biol. 1993;12:209–215. doi: 10.1089/dna.1993.12.209. [DOI] [PubMed] [Google Scholar]
  • 14.Au SW, Gover S, Lam VM, Adams MJ. Human glucose-6-phosphate dehydrogenase: the crystal structure reveals a structural NADP(+) molecule and provides insights into enzyme deficiency. Structure. 2000;8:293–303. doi: 10.1016/s0969-2126(00)00104-0. [DOI] [PubMed] [Google Scholar]
  • 15.Cohen P, Rosemeyer MA. Subunit interactions of glucose-6-phosphate dehydrogenase from human erythrocytes. Eur J Biochem. 1969;8:8–15. doi: 10.1111/j.1432-1033.1969.tb00488.x. [DOI] [PubMed] [Google Scholar]
  • 16.Camardella L, Damonte G, Carratore V, Benatti U, Tonetti M, Moneti G. Glucose 6-phosphate dehydrogenase from human erythrocytes: identification of N-acetyl-alanine at the N-terminus of the mature protein. Biochem Biophys Res Commun. 1995;207:331–338. doi: 10.1006/bbrc.1995.1192. [DOI] [PubMed] [Google Scholar]
  • 17.Shreve DS, Levy HR. On the molecular weight of human glucose 6-phosphate dehydrogenase. Biochem Biophys Res Commun. 1977;78:1369–1375. doi: 10.1016/0006-291x(77)91443-7. [DOI] [PubMed] [Google Scholar]
  • 18.Scopes DA, Bautista JM, Naylor CE, Adams MJ, Mason PJ. Amino acid substitutions at the dimer interface of human glucose-6-phosphate dehydrogenase that increase thermostability and reduce the stabilising effect of NADP. Eur J Biochem. 1998;251(1–2):382–388. doi: 10.1046/j.1432-1327.1998.2510382.x. [DOI] [PubMed] [Google Scholar]
  • 19.Mason PJ, Bautista JM, Gilsanz F. G6PD deficiency: the genotype-phenotype association. Blood Rev. 2007;21:267–283. doi: 10.1016/j.blre.2007.05.002. [DOI] [PubMed] [Google Scholar]
  • 20.Furuta E, Okuda H, Kobayashi A, Watabe K. Metabolic genes in cancer: their roles in tumor progression and clinical implications. Biochim Biophys Acta. 2010;1805:141–152. doi: 10.1016/j.bbcan.2010.01.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Sagone AL, Jr, Burton GM. The effect of BCNU and adriamycin on normal and G6PD deficient erythrocytes. Am J Hematol. 1979;7:97–106. doi: 10.1002/ajh.2830070202. [DOI] [PubMed] [Google Scholar]
  • 22.Cappellini MD, Fiorelli G. Glucose-6-phosphate dehydrogenase deficiency. Lancet. 2008;371:64–74. doi: 10.1016/S0140-6736(08)60073-2. [DOI] [PubMed] [Google Scholar]
  • 23.WHO Working Group. Glucose-6-phosphate dehydrogenase deficiency. Bull World Health Organ. 1989;67:601–611. [PMC free article] [PubMed] [Google Scholar]
  • 24.Pandolfi PP, Sonati F, Rivi R, Mason P, Grosveld F, Luzzatto L. Targeted disruption of the housekeeping gene encoding glucose 6-phosphate dehydrogenase (G6PD): G6PD is dispensable for pentose synthesis but essential for defense against oxidative stress. EMBO J. 1995;14:5209–5215. doi: 10.1002/j.1460-2075.1995.tb00205.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Jiang P, Du W, Wang X, Mancuso A, Gao X, Wu M, Yang X. p53 regulates biosynthesis through direct inactivation of glucose-6-phosphate dehydrogenase. Nat Cell Biol. 2011;13:310–316. doi: 10.1038/ncb2172. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Spolarics Z. Endotoxemia, pentose cycle, and the oxidant/antioxidant balance in the hepatic sinusoid. J Leukoc Biol. 1998;63:534–541. doi: 10.1002/jlb.63.5.534. [DOI] [PubMed] [Google Scholar]
  • 27.Brumbaugh RS, Schwartz J. Pythagoras and beans: a medical explanation. The Classical World. 1980;73:421–422. Available at http://www.jstor.org/stable/4349235. [Google Scholar]
  • 28.Pirmohamed M. Pharmacogenetics and pharmacogenomics. Br J Clin Pharmacol. 2001;52:345–347. doi: 10.1046/j.0306-5251.2001.01498.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Beutler E. The hemolytic effect of primaquine and related compounds: a review. Blood. 1959;14:103–139. [PubMed] [Google Scholar]
  • 30.Hockwald RS, Arnold J, Clayman CB, Alving AS. Toxicity of primaquine in Negroes. J Am Med Assoc. 1952;149:1568–1570. doi: 10.1001/jama.1952.72930340027010c. [DOI] [PubMed] [Google Scholar]
  • 31.Clayman CB, Arnold J, Hockwald RS, Yount EH, Jr, Edgcomb JH, Alving AS. Toxicity of primaquine in Caucasians. J Am Med Assoc. 1952;149:1563–1568. doi: 10.1001/jama.1952.72930340022010b. [DOI] [PubMed] [Google Scholar]
  • 32.Alving AS, Carson PE, Flanagan CL, Ickes CE. Enzymatic deficiency in primaquine-sensitive erythrocytes. Science. 1956;124:484–485. doi: 10.1126/science.124.3220.484-a. [DOI] [PubMed] [Google Scholar]
  • 33.Vulliamy T, Beutler E, Luzzatto L. Variants of glucose-6-phosphate dehydrogenase are due to missense mutations spread throughout the coding region of the gene. Hum Mutat. 1993;2:159–167. doi: 10.1002/humu.1380020302. [DOI] [PubMed] [Google Scholar]
  • 34.Beutler E. G6PD deficiency. Blood. 1994;84:3613–3636. [PubMed] [Google Scholar]
  • 35.Beutler E, Vulliamy TJ. Hematologically important mutations: glucose-6-phosphate dehydrogenase. Blood Cells Mol Dis. 2002;28:93–103. doi: 10.1006/bcmd.2002.0490. [DOI] [PubMed] [Google Scholar]
  • 36.Manco L, Pereira J, Relvas L, Rebelo U, Crisostomo AI, Bento C, Ribeiro ML. Chronic hemolytic anemia is associated with a new glucose-6-phosphate dehydrogenase in-frame deletion in an older woman. Blood Cells Mol Dis. 2011;46:288–293. doi: 10.1016/j.bcmd.2011.02.001. Epub 2011 Mar 11. [DOI] [PubMed] [Google Scholar]
  • 37.Rigano P, Fabiano C, Pojero F, Niceta M, Pecoraro A, Maggio A, Sammarco P. Glucose 6-phosphate dehydrogenase Palermo R257M: a novel variant associated with chronic non-spherocytic haemolytic anaemia. Br J Haematol. 2010;149:296–297. doi: 10.1111/j.1365-2141.2009.08044.x. [DOI] [PubMed] [Google Scholar]
  • 38.Vulliamy TJ, D’Urso M, Battistuzzi G, Estrada M, Foulkes NS, Martini G, et al. Diverse point mutations in the human glucose-6-phosphate dehydrogenase gene cause enzyme deficiency and mild or severe hemolytic anemia. Proc Natl Acad Sci USA. 1988;85:5171–5175. doi: 10.1073/pnas.85.14.5171. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Stevens DJ, Wanachiwanawin W, Mason PJ, Vulliamy TJ, Luzzatto L. G6PD canton a common deficient variant in South East Asia caused by a 459 Arg–Leu mutation. Nucleic Acids Res. 1990;18:7190. doi: 10.1093/nar/18.23.7190. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Chiu DT, Zuo L, Chen E, Chao L, Louie E, Lubin B, et al. Two commonly occurring nucleotide base substitutions in Chinese G6PD variants. Biochem Biophys Res Commun. 1991;180:988–993. doi: 10.1016/s0006-291x(05)81163-5. [DOI] [PubMed] [Google Scholar]
  • 41.Beutler E, Kuhl W, Vives-Corrons JL, Prchal JT. Molecular heterogeneity of glucose-6-phosphate dehydrogenase A- Blood. 1989;74:2550–2555. [PubMed] [Google Scholar]
  • 42.Hirono A, Kawate K, Honda A, Fujii H, Miwa S. A single mutation 202G > A in the human glucose-6-phosphate dehydrogenase gene (G6PD) can cause acute hemolysis by itself. Blood. 2002;99:1498. doi: 10.1182/blood.v99.4.1498. [DOI] [PubMed] [Google Scholar]
  • 43.Takizawa T, Yoneyama Y, Miwa S, Yoshida A. A single nucleotide base transition is the basis of the common human glucose-6-phosphate dehydrogenase variant A (+) Genomics. 1987;1:228–231. doi: 10.1016/0888-7543(87)90048-6. [DOI] [PubMed] [Google Scholar]
  • 44.Dern RJ, McCurdy PR, Yoshida A. A new structural variant of glucose-6-phosphate dehydrogenase with a high production rate (G6PD Hektoen) J Lab Clin Med. 1969;73:283–290. [PubMed] [Google Scholar]
  • 45.Yoshida A. Amino acid substitution (histidine to tyrosine) in a glucose-6-phosphate dehydrogenase variant (G6PD Hektoen) associated with over-production. J Mol Biol. 1970;52:483–490. doi: 10.1016/0022-2836(70)90414-6. [DOI] [PubMed] [Google Scholar]
  • 46.Yoshida A, Beutler E, Motulsky AG. Human glucose-6-phosphate dehydrogenase variants. Bull World Health Organ. 1971;45:243–253. [PMC free article] [PubMed] [Google Scholar]
  • 47.Yan T, Cai R, Mo O, Zhu D, Ouyang H, Huang L, et al. Incidence and complete molecular characterization of glucose-6-phosphate dehydrogenase deficiency in the Guangxi Zhuang autonomous region of southern China: description of four novel mutations. Haematologica. 2006;91:1321–1328. [PubMed] [Google Scholar]
  • 48.Iwai K, Hirono A, Matsuoka H, Kawamoto F, Horie T, Lin K, et al. Distribution of glucose-6-phosphate dehydrogenase mutations in Southeast Asia. Hum Genet. 2001;108:445–449. doi: 10.1007/s004390100527. [DOI] [PubMed] [Google Scholar]
  • 49.Bautista JM, Luzzatto L. Glucose 6-phosphate dehydrogenase. In: Swallow DM, Edwards YH, editors. Protein dysfunction in human genetic disease. Oxford: BIOS Scientific Publishers; 1997. pp. 33–56. [Google Scholar]
  • 50.Naylor, Rowland P, Basak AK, Gover S, Mason PJ, Bautista JM, et al. Glucose 6-phosphate dehydrogenase mutations causing enzyme deficiency in a model of the tertiary structure of the human enzyme. Blood. 1996;87:2974–2982. [PubMed] [Google Scholar]
  • 51.Cappellini MD, Sampietro M, Toniolo D, Carandina G, Martinez di Montemuros F, Tavazzi D, Fiorelli G. G6PD Ferrara I has the same two mutations as G6PD A(-) but a distinct biochemical phenotype. Hum Genet. 1994;93:139–142. doi: 10.1007/BF00210598. [DOI] [PubMed] [Google Scholar]
  • 52.De Vita G, Alcalay M, Sampietro M, Cappelini MD, Fiorelli G, Toniolo D. Two point mutations are responsible for G6PD polymorphism in Sardinia. Am J Hum Genet. 1989;44:233–240. [PMC free article] [PubMed] [Google Scholar]
  • 53.Hirono A, Kuhl W, Gelbart T, Forman L, Fairbanks VF, Beutler E. Identification of the binding domain for NADP + of human glucose-6-phosphate dehydrogenase by sequence analysis of mutants. Proc Natl Acad Sci USA. 1989;86:10015–10017. doi: 10.1073/pnas.86.24.10015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.The World Health Organization Scientific Group on Standardization of Procedures for the Study of Glucose-6-Phosphate Dehydrogenase Nomenclature of glucose-6-phosphate dehydrogenase in man. Am J Hum Genet. 1967;19:757–761. [PMC free article] [PubMed] [Google Scholar]
  • 55.Luzzatto L. Glucose 6-phosphate dehydrogenase deficiency: from genotype to phenotype. Haematologica. 2006;91:1303–1306. [PubMed] [Google Scholar]
  • 56.Leslie T, Briceno M, Mayan I. The impact of phenotypic and genotypic G6PD deficiency on risk of Plasmodium vivax infection: a case-control study amongst Afghan refugees in Pakistan. PLoS Med. 2010;7:e1000283. doi: 10.1371/journal.pmed.1000283. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.May J, Meyer CG, Grossterlinden L, Ademowo OG, Mockenhaupt FP, Olumese PE, et al. Red cell glucose-6-phosphate dehydrogenase status and pyruvate kinase activity in a Nigerian population. Trop Med Int Health. 2000;5:119–123. doi: 10.1046/j.1365-3156.2000.00529.x. [DOI] [PubMed] [Google Scholar]
  • 58.Beutler E, Yeh M, Fairbanks VF. The normal human female as a mosaic of X-chromosome activity: studies using the gene for C-6-PD-deficiency as a marker. Proc Natl Acad Sci USA. 1962;48:9–16. doi: 10.1073/pnas.48.1.9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59. [April 2011];FDA website with a table of Pharmacogenomic Biomarkers. Available at: http://www.fda.gov/Drugs/ScienceResearch/ResearchAreas/Pharmacogenetics/ucm083378.htm.
  • 60. [April 2011];FDA MedWatch. Available at: http://www.fda.gov/Safety/MedWatch/default.htm.
  • 61.Beutler E. Glucose-6-phosphate dehydrogenase deficiency. N Engl J Med. 1991;324:169–174. doi: 10.1056/NEJM199101173240306. [DOI] [PubMed] [Google Scholar]
  • 62.Youngster I, Arcavi L, Schechmaster R, Akayzen Y, Popliski H, Shimonov J, et al. Medications and glucose-6-phosphate dehydrogenase deficiency: an evidence-based review. Drug Saf. 2010;33:713–726. doi: 10.2165/11536520-000000000-00000. [DOI] [PubMed] [Google Scholar]
  • 63.The Association is affiliated with Italian Federation for Rare Diseases, which in turn is a member of European Organization for Rare Diseases. Copyright 1996–2011 by Associazione Italiana Favismo – Deficit di G6PD. Available at: www.favism.org
  • 64.Barnwell JW, Bell D, Hanson K, Mendis K, Moonen B, Newman RD, et al. A research agenda for malaria eradication: diagnoses and diagnostics. PLoS Med. 2011;8:e1000396. doi: 10.1371/journal.pmed.1000396. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Alloueche A, Bailey W, Barton S, Bwika J, Chimpeni P, Falade CO, et al. Comparison of chlorproguanil-dapsone with sulfadoxine-pyrimethamine for the treatment of uncomplicated falciparum malaria in young African children: double-blind randomised controlled trial. Lancet. 2004;363:1843–1848. doi: 10.1016/S0140-6736(04)16350-2. [DOI] [PubMed] [Google Scholar]
  • 66.Shekalaghe SA, ter Braak R, Daou M, Kavishe R, van den Bijllaardt W, van den Bosch S, et al. In Tanzania, hemolysis after a single dose of primaquine coadministered with an artemisinin is not restricted to glucose-6-phosphate dehydrogenase-deficient (G6PD A-) individuals. Antimicrob Agents Chemother. 2010;54:1762–1768. doi: 10.1128/AAC.01135-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Ganesan S, Tekwani BL, Sahu R, Tripathi LM, Walker LA. Cytochrome P(450)-dependent toxic effects of primaquine on human erythrocytes. Toxicol Appl Pharmacol. 2009;241:14–22. doi: 10.1016/j.taap.2009.07.012. [DOI] [PubMed] [Google Scholar]
  • 68.Premji Z, Umeh RE, Owusu-Agyei S, Esamai F, Ezedinachi EU, Oguche S, et al. Chlorproguanil-dapsone-artesunate versus artemether-lumefantrine: a randomized, double-blind phase III trial in African children and adolescents with uncomplicated Plasmodium falciparum malaria. PLoS One. 2009;4:e6682. doi: 10.1371/journal.pone.0006682. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Fanello CI, Karema C, Avellino P, Bancone G, Uwimana A, Lee SJ, et al. High risk of severe anaemia after chlorproguanil-dapsone + artesunate antimalarial treatment in patients with G6PD (A-) deficiency. PLoS One. 2008;3:e4031. doi: 10.1371/journal.pone.0004031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.World Health Organization. Review of the safety of Chlorproguanil-dapsone in the treatment of uncomplicated falciparum malaria in Africa: report of a technical consultation convened by WHO, 1–2 July 2004, Geneva, 2005. [April 2011]; Available at: http://www.who.int/malaria/publications/atoz/who_htm_mal_2005_1106/en/index.html.
  • 71.Luzzatto L. The rise and fall of the antimalarial Lapdap: a lesson in pharmacogenetics. Lancet. 2010;376:739–741. doi: 10.1016/S0140-6736(10)60396-0. [DOI] [PubMed] [Google Scholar]
  • 72.Gaetani GD, Mareni C, Ravazzolo R, Salvidio E. Haemolytic effect of two sulphonamides evaluated by a new method. Br J Haematol. 1976;32:183–191. doi: 10.1111/j.1365-2141.1976.tb00921.x. [DOI] [PubMed] [Google Scholar]
  • 73.Chan TK, Todd D, Tso SC. Red cell survival studies in glucose 6 phosphate dehydrogenase deficiency. Bull Hong Kong Med Assoc. 1974;26:41–48. [Google Scholar]
  • 74.Sicard D, Kaplan JC, Labie D. Haemoglobinopathies and G.-6-P.D. deficiency in Laos. Lancet. 1978;2:571–572. doi: 10.1016/s0140-6736(78)92899-4. [DOI] [PubMed] [Google Scholar]
  • 75.Ademowo OG, Sodeinde O. Certain red cell genetic factors and prevalence of chloroquine-induced pruritus. Afr J Med Med Sci. 2002;31:341–343. [PubMed] [Google Scholar]
  • 76.Krudsood S, Wilairatana P, Tangpukdee N, Chalermrut K, Srivilairit S, Thanachartwet V, et al. Safety and tolerability of elubaquine (bulaquine, CDRI 80/53) for treatment of Plasmidium vivax malaria in Thailand. Korean J Parasitol. 2006;44:221–228. doi: 10.3347/kjp.2006.44.3.221. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Ramos Junior WM, Sardinha JF, Costa MR, Santana MS, Alecrim MG, Lacerda MV. Clinical aspects of hemolysis in patients with P. vivax malaria treated with primaquine, in the Brazilian Amazon. Braz J Infect Dis. 2010;14:410–412. doi: 10.1590/s1413-86702010000400017. [DOI] [PubMed] [Google Scholar]
  • 78.Choudhry VP, Ghafary A, Zaher M, Qureshi MA, Fazel I, Ghani R. Drug-induced haemolysis and renal failure in children with glucose-6-phosphate dehydrogenase deficiency in Afghanistan. Ann Trop Paediatr. 1990;10:335–338. doi: 10.1080/02724936.1990.11747454. [DOI] [PubMed] [Google Scholar]
  • 79.Mandi G, Witte S, Meissner P, Coulibaly B, Mansmann U, Rengelshausen J, et al. Safety of the combination of chloroquine and methylene blue in healthy adult men with G6PD deficiency from rural Burkina Faso. Trop Med Int Health. 2005;10:32–38. doi: 10.1111/j.1365-3156.2004.01356.x. [DOI] [PubMed] [Google Scholar]
  • 80.Meissner PE, Mandi G, Witte S, Coulibaly B, Mansmann U, Rengelshausen J, et al. Safety of the methylene blue plus chloroquine combination in the treatment of uncomplicated falciparum malaria in young children of Burkina Faso [ISRCTN27290841] Malar J. 2005;4:45. doi: 10.1186/1475-2875-4-45. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Owusu-Agyei S, Asante KP, Owusu R, Adjuik M, Amenga-Etego S, Dosoo DK, et al. An open label, randomised trial of artesunate + amodiaquine, artesunate+chlorproguanil-dapsone and artemether-lumefantrine for the treatment of uncomplicated malaria. PLoS One. 2008;3:e2530. doi: 10.1371/journal.pone.0002530. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Frischer H, Ahmad T. Severe generalized glutathione reductase deficiency after antitumor chemotherapy with BCNU” [1,3-bis(chloroethyl)-1-nitrosourea] J Lab Clin Med. 1977;89:1080–1091. [PubMed] [Google Scholar]
  • 83.Thorn CF, Oshiro C, Marsh S, Hernandez-Boussard T, McLeod H, Klein TE, Altman RB. Doxorubicin pathways: pharmacodynamics and adverse effects. Pharmacogenet Genomics. 2011;21:440–446. doi: 10.1097/FPC.0b013e32833ffb56. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Bain BJ. A ghostly presence-G6PD deficiency. Am J Hematol. 2010;85:271. doi: 10.1002/ajh.21660. [DOI] [PubMed] [Google Scholar]
  • 85.Browning LA, Kruse JA. Hemolysis and methemoglobinemia secondary to rasburicase administration. Ann Pharmacother. 2005;39:1932–1935. doi: 10.1345/aph.1G272. [DOI] [PubMed] [Google Scholar]
  • 86.Joly P, Bon C, Francina A, Gelineau MC, Lacan P, Orfeuvre H. A severe G6PD deficiency revealed during a chemotherapy protocol including rasburicase. Ann Biol Clin (Paris) 2009;67:432–436. doi: 10.1684/abc.2009.0353. [DOI] [PubMed] [Google Scholar]
  • 87.Amitai Y, Bhooma T, Frischer H. Glucose-6-phosphate dehydrogenase deficiency severely restricts the biotransformation of daunorubicin in human erythrocytes. J Lab Clin Med. 1996;127:588–598. doi: 10.1016/s0022-2143(96)90150-x. [DOI] [PubMed] [Google Scholar]
  • 88.Meloni T, Forteleoni G, Ogana A, Franca V. Aspirin-induced acute haemolytic anaemia in glucose-6-phosphate dehydrogenase-deficient children with systemic arthritis. Acta Haematol. 1989;81:208–209. doi: 10.1159/000205562. [DOI] [PubMed] [Google Scholar]
  • 89.Szeinberg A, Kellermann J, Adam A, Sheba C, Ramot B. Haemolytic jaundice following aspirin administration to a patient with a deficiency of glucose-6-phosphate dehydrogenase in erythrocytes. Acta Haematol. 1960;23:58–64. doi: 10.1159/000206201. [DOI] [PubMed] [Google Scholar]
  • 90.Glader BE. Evaluation of the hemolytic role of aspirin in glucose-6-phosphate dehydrogenase deficiency. J Pediatr. 1976;89:1027–1028. doi: 10.1016/s0022-3476(76)80626-9. [DOI] [PubMed] [Google Scholar]
  • 91.Stockman JA, III, Lubin B, Oski FA. Aspirin-induced hemolysis: the role of concomitant oxidant (H2O2) challenge. Pediatr Res. 1978;12:927–931. doi: 10.1203/00006450-197809000-00009. [DOI] [PubMed] [Google Scholar]
  • 92.Meloni G, Meloni T. Glyburide-induced acute haemolysis in a G6PD-deficient patient with NIDDM. Br J Haematol. 1996;92:159–160. doi: 10.1046/j.1365-2141.1996.275810.x. [DOI] [PubMed] [Google Scholar]
  • 93.Vinzio S, Andres E, Perrin AE, Schlienger JL, Goichot B. Glibenclamide-induced acute haemolytic anaemia revealing a G6PD-deficiency. Diabetes Res Clin Pract. 2004;64:181–183. doi: 10.1016/j.diabres.2003.11.006. [DOI] [PubMed] [Google Scholar]
  • 94.Rosen PJ, Johnson C, McGehee WG, Beutler E. Failure of methylene blue treatment in toxic methemoglobinemia: association with glucose-6-phosphate dehydrogenase deficiency. Ann Intern Med. 1971;75:83–86. doi: 10.7326/0003-4819-75-1-83. [DOI] [PubMed] [Google Scholar]
  • 95.Campbell GD, Jr, Steinberg MH, Bower JD. Letter: ascorbic acid-induced hemolysis in G-6-PD deficiency. Ann Intern Med. 1975;82:810. doi: 10.7326/0003-4819-82-6-810_1. [DOI] [PubMed] [Google Scholar]
  • 96.McCurdy PR, Kirkman HN, Naiman JL, Jim RT, Pickard BM. A Chinese variant of glucose-6-phosphate dehydrogenase. J Lab Clin Med. 1966;67:374–385. [PubMed] [Google Scholar]
  • 97.Chan TK, Todd D. Characteristics and distribution of glucose-6-phosphate dehydrogenase-deficient variants in South China. Am J Hum Genet. 1972;24:475–484. [PMC free article] [PubMed] [Google Scholar]
  • 98.Shibuya A, Hirona A, Ishii S, Fujii H, Miwa S, et al. Hemolytic crisis after excessive ingestion of fava beans in a male infant with G6PD Canton. Int J Hematol. 1999;70:233–235. [PubMed] [Google Scholar]
  • 99.Yoshida A. A single amino acid substitution (asparagine to aspartic acid) between normal (b+) and the common negro variant (a+) of human glucose-6-phosphate dehydrogenase. Proc Natl Acad Sci USA. 1967;57:835–840. doi: 10.1073/pnas.57.3.835. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Manco L, Botigue LR, Ribeiro ML, Abade A. G6PD deficient alleles and haplotype analysis of human G6PD locus in Sao Tome e Principe (West Africa) Hum Biol. 2007;79:679–686. doi: 10.1353/hub.2008.0013. [DOI] [PubMed] [Google Scholar]
  • 101.Viglietto G, Montanaro V, Calabro V, Vallone D, D’Urso M, Persico MG, Battistuzzi G. Common glucose-6-phosphate dehydrogenase (G6PD) variants from the Italian population: biochemical and molecular characterization. Ann Hum Genet. 1990;54(Pt 1):1–15. doi: 10.1111/j.1469-1809.1990.tb00355.x. [DOI] [PubMed] [Google Scholar]
  • 102.Beutler E, Kuhl W, Saenz GF, Rodriguez W. Mutation analysis of glucose-6-phosphate dehydrogenase (G6PD) variants in Costa Rica. Hum Genet. 1991;87:462–464. doi: 10.1007/BF00197169. [DOI] [PubMed] [Google Scholar]
  • 103.Saenz GF, Chaves M, Berrantes A, Elizondo J, Montero AG, Yoshida A. A glucose-6-phosphate dehydrogenase variant, Gd(-) Santamaria found in Costa Rica. Acta Haematol. 1984;72:37–40. doi: 10.1159/000206354. [DOI] [PubMed] [Google Scholar]
  • 104.Jalloh A, Jalloh M, Gamanga I, Baion D, Sahr F, Gbakima A, et al. G6PD deficiency assessment in Freetown. Sierra Leone, reveals further insight into the molecular heterogeneity of G6PD A. J Hum Genet. 2008;53:675–679. doi: 10.1007/s10038-008-0294-y. [DOI] [PubMed] [Google Scholar]
  • 105.Babalola AO, Beetlestone JG, Luzzatto L. Genetic variants of human erythrocyte glucose-6-phosphate dehydrogenase: kinetic and thermodynamic parameters of variants A, B, and A- in relation to quaternary structure. J Biol Chem. 1976;251:2993–3002. [PubMed] [Google Scholar]
  • 106.Vulliamy TJ, Othman A, Town M, Nathwani A, Falusi AG, Mason PJ, Luzzatto L. Polymorphic sites in the African population detected by sequence analysis of the glucose-6-phosphate dehydrogenase gene outline the evolution of the variants A and A. Proc Natl Acad Sci USA. 1991;88:8568–8571. doi: 10.1073/pnas.88.19.8568. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Gomez-Gallego F, Garrido-Pertierra A, Bautista JM. Structural defects underlying protein dysfunction in human glucose-6-phosphate dehydrogenase A(-) deficiency. J Biol Chem. 2000;275:9256–9262. doi: 10.1074/jbc.275.13.9256. [DOI] [PubMed] [Google Scholar]
  • 108.Saunders MA, Slatkin M, Garner C, Hammer MF, Nachman MW. The extent of linkage disequilibrium caused by selection on G6PD in humans. Genetics. 2005;171:1219–1229. doi: 10.1534/genetics.105.048140. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Guindo A, Fairhurst RM, Doumbo OK, Wellems TE, Diallo DA. X-linked G6PD deficiency protects hemizygous males but not heterozygous females against severe malaria. PLoS Med. 2007;4:e66. doi: 10.1371/journal.pmed.0040066. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Ruwende C, Khoo SC, Snow RW, Yates SN, Kwiatkowski D, Gupta S, et al. Natural selection of hemi- and heterozygotes for G6PD deficiency in Africa by resistance to severe malaria. Nature. 1995;376:246–249. doi: 10.1038/376246a0. [DOI] [PubMed] [Google Scholar]
  • 111.Bienzle U, Ayeni O, Lucas AO, Luzzatto L. Glucose-6-phosphate dehydrogenase and malaria: greater resistance of females heterozygous for enzyme deficiency and of males with non-deficient variant. Lancet. 1972;1:107–110. doi: 10.1016/s0140-6736(72)90676-9. [DOI] [PubMed] [Google Scholar]
  • 112.Clark TG, Fry AE, Auburn S, Campino S, Diakite M, Green A, et al. Allelic heterogeneity of G6PD deficiency in West Africa and severe malaria susceptibility. Eur J Hum Genet. 2009;17:1080–1085. doi: 10.1038/ejhg.2009.8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Johnson MK, Clark TD, Njama-Meya D, Rosenthal PJ, Parikh S. Impact of the method of G6PD deficiency assessment on genetic association studies of malaria susceptibility. PLoS One. 2009;4:e7246. doi: 10.1371/journal.pone.0007246. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Mendez D, Linares M, Diez A, Puyet A, Bautista JM. Stress response and cytoskeletal proteins involved in erythrocyte membrane remodeling upon Plasmodium falciparum invasion are differentially carbonylated in G6PD A(-) deficiency. Free Radic Biol Med. 2011;50:1305–1313. doi: 10.1016/j.freeradbiomed.2011.02.024. Epub 2011 Mar 1. [DOI] [PubMed] [Google Scholar]
  • 115.Beutler E, Kuhl W. The NT 1311 polymorphism of G6PD: G6PD Mediterranean mutation may have originated independently in Europe and Asia. Am J Hum Genet. 1990;47:1008–1012. [PMC free article] [PubMed] [Google Scholar]
  • 116.Vives-Corrons JL, Kuhl W, Pujades MA, Beutler E. Molecular genetics of the glucose-6-phosphate dehydrogenase (G6PD) Mediterranean variant and description of a new G6PD mutant, G6PD Andalus1361A. Am J Hum Genet. 1990;47:575–579. [PMC free article] [PubMed] [Google Scholar]

RESOURCES