Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2019 Feb 5.
Published in final edited form as: Mol Biotechnol. 2010 Feb;44(2):152–167. doi: 10.1007/s12033-009-9220-6

Pharmacogenetics of Membrane Transporters: An Update on Current Approaches

Tristan M Sissung 1, Caitlin E Baum 2, C Tyler Kirkland 3, Rui Gao 4, Erin R Gardner 5, William D Figg 6
PMCID: PMC6362991  NIHMSID: NIHMS1008786  PMID: 19950006

Abstract

This review provides an overview of the pharmacogenetics of membrane transporters including selected ABC transporters (ABCB1, ABCC1, ABCC2, and ABCG2) and OATPs (OATP1B1 and OATP1B3). Membrane transporters are heavily involved in drug clearance and alters drug disposition by actively transporting substrate drugs between organs and tissues. As such, polymorphisms in the genes encoding these proteins may have significant effects on the absorption, distribution, metabolism and excretion of compounds, and may alter pharmacodynamics of many agents. This review discusses the techniques used to identify substrates and inhibitors of these proteins and subsequently to assess the effect of genetic mutation on transport, both in vitro and in vivo. A comprehensive list of substrates for the major drug transporters is included. Finally, studies linking transporter genotype with clinical outcomes are discussed.

Keywords: ABCBl, ABCCl, ABCC2, ABCG2, OATP1B3, OATPlB1, Pharmacogenetics, Membrane transporters

Background

The fate of a drug in vivo is dictated by a variety of physiochemical properties including size, lipophilicity, and charge. These properties determine how a drug is absorbed, distributed throughout the body, metabolized, and eventually eliminated. While movement of a drug molecule can occur through simple diffusion, there are many transporter proteins expressed on cell membranes to assist with influx or efflux via active transport. As such, these transporters can significantly affect drug disposition. For example, influx of a drug from the blood to the liver, where it is subsequently metabolized and excreted, may increase the rate of elimination. These proteins and the genes that encode them are essential to drug uptake, bioavailability, targeting, efficacy, toxicity, and clearance. The genes encoding these transporters are polymorphic, phenotypically resulting in transporters with different expression levels and transport efficiency. Consequently, polymorphisms in transporters often contribute to variability in drug pharmacokinetics and response to treatment.

Many drugs undergo transport mediated by the ATP­binding cassette (ABC) family of transporters. There are a total of 49 known ABC genes including ABCB1 (P-glycoprotein, MDR-1), ABCC1 (MRPl), and ABCG2(BCRP, MXR, ABCP), all of which utilize ATP to move substrates across membranes [14]. These transporters generally counteract uptake through the intestinal wall, efflux substrates out of tissues into the systemic circulation, and eventually mediate the clearance of drugs from the body. Proteins in the ABC family are primarily known to be efflux transporters, moving substrates across the cell membrane and out of the cell. The most characterized polymorphic transporters to date are ABCB1 and ABCG2 [5]. Many current FDA approved drugs are substrates of these transporters, although both transporters also efflux a plethora of other compounds including naturally occurring toxins. ABCB1 and ABCG2 are expressed in enterocytes, the colon, the intestinal epithelium, the canalicular plasma membrane of hepatocytes, and the proximal renal tubule [610]. As such, ABCB1 and ABCG2 often mediate bioavailability and exposure to their substrate drugs mentioned in Table 1 [3, 11]. In addition, they have been shown to be expressed in hematologic tissues including hematopoietic stem cells and endothelial cells composing blood–tissue barriers of the brain, heart, nerves, testes, and placenta, where they efflux substrates out of these tissues into the systemic circulation [7, 1217]. An exception includes the expression of ABCB 1in the choroid plexus where it transports molecules from the circulation into the cerebrospinal fluid [16, 18]. It is believed that the evolutionary role of these transporters is to limit the penetration of toxic molecules into critical organs, thereby serving a protective role in blood–tissue barriers.

Table 1.

Substrates of ABCB1, ABCG2, ABCC1, ABCC2, OATP1B1, OATP1B3

Substrates Inhibitors
ABCB1
Antibiotics Actinomycin D
Erythromycin
Gramicidin D
Rifampin
Salinomycin
Sparfloxacin
Valinomycin
Anti-cancer drugs Bisantrene Sunitinib
Daunorubicin Tamoxifen
Diflomotecan
Docetaxel
Doxorubicin
Epirubicin
Etoposide
Gefitinib
Imatinib
Irinotecan
Mitoxantrone
Paclitaxel
Romidepsin
Teniposid
Tipifarnib
Vinblastine
Vincristine
Antifungals Itraconazole Ketoconazole
Ketoconazol
Antihistamines Certirizine
Fexofenadine
Loratadine
Terfenadine
Antihypertensive drugs Losartin Nicardipine
Talinolol Quinidine
Verapami
CNS drugs Chlorpromazine
Clozapine
Fluphenazine
Olanzapine
Quetiapine
Risperidone
Flavonoids Biochanin A
Genistein
Oroxylin A
Heart medications Digoxin Gallopamil
Diltiazem
Ouabain
Quinidine
Verapamil
HIV-1 protease inhibitors Abacavir
Amprenavir
Aquinavir
Darunavir
Indinavir
Lopinavir
Nelfinavir
Ritonavir
Saquinavir
Immunosuppressants Cyclosporin A Cyclosporin A
Dexamethasone Valspodar
D-penicillamine enkephalin
FK 506
Hydrocortisone
Prednisolone
Rapamycin
Tacrolimus
Triamcinolone
Sedatives Midazolam
Statins Atorvastatin
Cerivastatin
Lovastatin
Miscellaneous Asimadoline Dexverapamil
Cimetidine Emopamil
Colchicine JAI-51
Domperidine Quinacrine
Eletriptan Tariquidar
Flesinoxan
Glabridin
Ivermectin
Loperamide
Ondansetron
Quinacrine
Ranitidine
Topiramate
ABCG2
Antibiotics Ciprofloxacin Novobiocin
Erythromycin Rapamycin
Nitrofurantoin
Norfloxacin
Ofloxacin
Anti-cancer drugs 9-Aminocamptothecin Biricodar
Bisantrene Diethyls tilbestrol
Cladribine Elacridar
Daunorubicin Fumitremorgin
Diflomotecan Gefitinib
Doxorubicin Ginsenoside
Epirubicin Ortataxel
Erlotinib Sunitinib
Etoposide Tamoxifen
Flavopiridol Tryprostatin
Gefitinib Vandetanib
Gimatecan
Homocamptothecin
Imatinib
Methotrexate
Mitoxantrone
SN-38 (irinotecan metabolite)
Teniposide
Tomudex
Topotecan
Antihypertensive drugs Olmesartan Dihydropyridine
Dipyridamole
Reserpine
Anti-inflammatory drugs Chrysin
Curcumin
Antiplatelets Dipyridamole
Calcium channel blockers Azidopine Nicardapine
Dipyridamole Nimodipine
Nitrendipine Nitrendipine
Flavonoids Seravastatin Acacetin
Apigenin
Genistein
Naringenin
Quercetin
Silymarin
Techochrysin
HIV-1 protease inhibitors Abacavir Abacavir
Lamivudine Amprenavir
Nelfinavir Atazanavir
Zidovudine (AZT) Delavirdine
Efavirenz
Lopinavir
Nelfinavir
Ritonavir
Saquinavir
Immunosuppressants Cyclosporin A Cyclosporin A
Lefunomide Sirolimus
Sirolimus Tacrolimus
Sulfasalazine
Tacrolimus
Specific inhibitors GF120918
Ko143
Tariquidar (XR9576)
Statins Pitavastatin Rosuvastatin
Posuvastatin
Seravastatin
Miscellaneous Glyburide Pantoprazole
Protoporphyrin
ABCC1
Antibiotics Berberine
Ciprofloxacin
Difloxacin
Grepafloxacin
Pirarubicin
Anti-cancer drugs Apicidin Vandetanib
Camptothecin
Chlorambucil
Chlorambucil
Cyclophosphamide
Daunorubicin
Depsipeptide (FK228)
Doxorubicin
Edatrexate
Epirubicin
Etoposide
Flutamide
Hydroxyflutamide
Idarubicin
Irinotecan
Melphalan
Methotrexate
Paclitaxel
SN-38 (irinotecan metabolite)
Vinblastine
Vincristine
ZD1694
Antihypertensive drugs Verapamil
Anitiinflammatory drugs Indomethacin
Quercetin
Sulindac
Flavonoids Biochanin A
Genistein
HIV-1 protease inhibitors Indinavir Kaempferol
Ritonavir
Saquinavir
Immunosuppressants Cyclosporin A
Miscellaneous Probenecid
Sulfinpyrazone
ABCC2
Antibiotics Ampicillin Azithromycin
Azithromycin
Cefodizime
Ceftriaxone
Grepafloxacine
Anti-cancer drugs Camptothecin BTK
Cisplatin Lonafarnib
Doxorubicin
Etoposide
Irinotecan
Methotrexate
Mitoxantrone
Vinblastine
Vincristine
Antihypertensive drugs Olmesartan
Antiflammatory drugs Curcumin
Blood-glucode lowering drugs Glibenclamide
HIV-1 protease inhibitors Adefovir
Cidofovir
Indinavir
Lopinavir
Nelfinavir
Ritonavir
Saquinavir
Immunosuppressants Cyclosporin A
Statins Pravastatin
Miscellaneous Temocaprilate MK-571
Valproate Furosemide
PAK-104P
Phenobarbital
Probenecid
OATP1B1
Antibiotics Benzylpenicilliin Clarithromycin
Rifampin Erythromycin
Hyperforin
Rapamycin
Rifampin
Rifamycin SV
Roxithromycin
Telithromycin
Anti-cancer drugs ACU-154 Antamanide
Atrasentan Ketoconazole
Bamet-R2 Paclitaxel
Bamet-UD2 PKI-166
Demethylphalloin SN-38
Dihydromicrocystin-LR
Irinotecan
Methotrexate
SN-38
Anti-diabetics Glibenclamide
Pioglitazone
Rosiglitazone
Antifungals Caspofungin Clotrimazole
Antihistamines Fexofenadine
Antihypertensive drugs Bosentan Telmisartan
Enalapril
Olmesartan
Temocapril
Valsartan
Anti-inflammatory drugs D-penicillamin encephalin Troglitazone
Troglitazone sulfate Troglitazone sulfate
Blood-glucose lowering drugs Repaglinide
Fibrates Gemfibrozil
Gemfibrozil-1-O-glucuronide
Flavonoids Biochanin A
Heart medications Digoxin
HIV-1 protease inhibitors Indinavir
Nelfinavir
Ritonavir
Saquinavir
Immunosuppressants Cyclosporin A
Tacrolimus
Statins Atorvastatin Atorvastatin
Cerivastatin BMS-241423 (atorvastatin analog)
Fluvastatin BMS-243887 (atorvastatin analog)
Pitavastatin Lovastatin
Pravastatin Lovastatin acid
Rosuvastatin Lovastatin lactone
Simvastatin acid Pravastatin
Simvastatin
Simvastatin lactone
Miscellaneous BQ-123 Carbamazepine
Bromosulphophthalein Glycyrrhizin
Metyrapone
Mifepristone
Sildenafil
OATP1B3
Antibiotics Rifampin Clarithromycin
Erythromycin
Hyperforin
Rifampin
Rifamycin
Roxithromycin
Telithromycin
Anti-cancer drugs Demethylphalloin
Dihydromicrocystin-LR
Docetaxel
Imatinib
Irinotecan
Methotrexate
Paclitaxel
SN-38
Antihistamines Fexofendadine
Antihypertensives Bosentan
Enalapril
Olmesartan
Telmisartan
Valsartan
Anti-inflammatory drugs D-penicillamine Troglitazone sulfate
enkephalin
Blood-glucose lowring drugs Repaglinide
Heart medications Digoxin
Ouabain
Immunosuppressants Cyclosporin A
Statins Fluvastatin Pravastatin
Pitavastatin
Pravastatin
Rosuvastatin
Miscellaneous BQ-123 Bromosulphophthalein
Bromosulphophthalein Glycyrrhizin

Two other efflux transporters, ABCC1 (MRP-1) and ABCC2 (MRP-2) are also involved in drug disposition. ABCC1 is expressed ubiquitously and is localized to the basolateral, rather than apical, membranes of epithelial cells. Due to its basolateral localization, ABCC1 pumps drugs into the body rather than into the bile, urine, or intestine. For this reason, it is thought to serve mainly as a protective barrier in epithelial cells of tissues rather than a classic drug efflux pump [19]. ABCC2 is similar in function to ABCB1. It is expressed on the apical domain of epithelial cells and is involved in luminal excretion in organs such as the liver, the intestine, and the kidney, but it also plays a role in blood-tissue barriers [20]. Both ABCC1 and ABCC2 primarily secrete drugs that have undergone phase II metabolism into glutathione, glucuronide, or sulfate conjugates, but both efflux a wide range of drugs [21].

There are also several classes of “influx” or “uptake “ transporters that mediate the cellular uptake and reabsorption of drugs by moving substrates against a concentration gradient. The main uptake transporters are the organic anion transporting proteins (OATPs), organic cation transporters (OCTs), concentrative nucleoside transporters (CNT), dipeptide transporters (PEPT), and mono carboxylate transporters (MCT) [22]. For the sake of brevity, we will discuss only two members of the OATP1B family of proteins as these are well-characterized influx transporters. OATP1B1 and OATP1B3 are expressed in liver tissues and are responsible for hepatocellular uptake of drugs from blood across the basolateral membrane [23]. It was previously thought that these transporters were primarily involved in uptake of substrates into the liver where metabolism occurs [23]. However, more recent evidence suggests that OATP1B3 is overexpressed in several tumor types such as prostate, colon, and liver [2426]. Thus, since OATP1B3 influences drug treatment with docetaxel, paclitaxel and irinotecan (along with active metabolite SN-38), it is possible that those tumors will be more sensitive to OATP1B3 substrate drugs. Therefore, the OATP1B family is important in regulating the pharmacokinetics and potentially the response to several substrate drugs.

There is significant variation in the genes encoding all of the aforementioned transporters. Several of these genetic variants result in alterations in mRNA expression levels (e.g., promoter variants), translational efficiency (e.g., alterations in mRNA folding), and protein function (e.g., coding polymorphisms). Such genetic variability in transporters often explains a component of the inter-individual variability in drug disposition, ultimately resulting in differences in clinical endpoints including toxicity and response. The field of transporter pharmacogenetics is concerned with elucidating the mechanisms by which genetic variation in transporters determines individual differences in drug transport with a goal of eventually personalizing treatment with substrate drugs based on genotype. This review will provide an overview of the methods by which investigators have discovered and characterized such associations in the ABCB1, ABCG2, ABCC1, ABCC2, OATP1B1, and OATP1B13 transporters. This methodology could be readily applied to the study of many additional transporters.

Genetic Variation and Genotyping Methods

More than 50 polymorphisms, three insertions/deletions, and several promoter alterations that modify gene transcription have been described in the ABCB1 gene [27]. There are three single-nucleotide polymorphisms (SNPs) that are common in most ethnic groups and demonstrate strong linkage disequilibrium: the synonymous transition at nucleotide 1236C>T (Gly411Gly) in exon 12, the non­synonymous tri-allelic transition 2677G>T/A (Ala893Ser/Thr) in exon 21, and the synonymous transition 3435C>T (Ile 1145Ile) in exon 26. Of these SNPs, only the 2677G>T/A (Ala893Ser/Thr) transition causes an amino acid change within a structurally important transmembrane domain of the translated protein, and the effects of this transition are controversial and drug-specific [15, 22, 2830]. The 3435C>T SNP is associated with decreased mRNA stability and lower expression levels [31]. Recently, Kimchi­Sarfaty et al. showed that synonymous polymorphisms in ABCB1 may be responsible for altered protein conformations due to the phenomenon of ribosomal stalling [32]. As the genetic code is degenerate and relative frequencies of codons vary, there is occasion for frequent-to-rare synonymous codon substitutions to appear. The substitution of a rarer codon can lead to pauses in ribosomal translation, during which the protein can adopt different secondary structures that may result in functional changes. The mechanism of this phenomenon, which may apply to other transporters in addition to ABCB1, is well described in the 2008 review by Tsai et al. [33]. When compared to single polymorphisms, haplotype combinations of these SNPs can result in greater protein function differences because each SNP has at least an additive effect on ABCB1 conformational changes. The 3435C>T polymorphism, which results in a synonymous protein change, has been shown to confer differences in ABCBl transport characteristics when combined with the 2677G>T/A and 1236C>T alleles as compared to the 2677G>T/A and 1236C>T alleles alone [32]. While this evidence is compelling, linkage between SNPs should be studied for confounding factors. For example, the 1236C>T polymorphism is in ~90% D’ linkage with the 2677G>T/A polymorphism in several populations, and by virtue of that linkage may be only artificially associated with inter-individual ABCB1 transport alterations.

While there are many polymorphisms in ABCG2, the common ABCG2 421C>A allele in exon 5 is by far the most well characterized. This SNP results in an amino acid change of Gln to Lys at codon 141 and has been shown in Flp-In-293 cells to have half the protein expression of the wildtype [34]. The variant alleles (i.e., 421A and 141K) have also been associated with lower ATPase activity as compared with the wild-type ABCG2 [35]. Thus, the ABCG2 421C>A SNP, much like the ABCB1 2677G>T/A allele, may alter both expression and activity of the encoded protein. The frequency of this mutation varies significantly by race; it occurs at 35% frequency in Chinese populations, whereas the mutation is very rare in African Americans (1%) [36]. Another SNP exists at nucleotide 34, resulting in a V12M amino acid change. This mutation results in poor localization of the ABCG2 protein [35], but does not change protein expression levels [37]. Surprisingly, this mutation does not appear to modify substrate transport [38]. Furthermore, mutations at R482 which result in non-synonymous protein changes have been identified in numerous cancer cell lines (presumably a mechanism of multidrug resistance) but have never been found in humans. This mutation affects both transport and substrate specificity [3942].

There are several polymorphisms in ABCC1, many of which are non-synonymous. Those studied include C43S, T73I, S92F, Tl 17M, R230Q, V353M, R433S, R633Q, G671V, R723Q, A989T, C1047S, R1058Q, A1337T, and S1512L. The majority of these SNPs do not alter the functionality of the expressed protein and are unlikely to significantly influence the expression [43]. However, it has been noted that C43S, R433S, and A989T result in decreased ABCB1 function [43]. Others have evaluated non-synonymous polymorphisms to assess their impact on mRNA expression, but have found no significant results [44]. The ABCC2 gene also contains several polymorphisms. In particular, patients with Dubin Johnson Syndrome (DJS) commonly have the 2302 C>T A768W polymorphism [20]. For the most part, however, ABCC2 polymorphisms have not been significantly associated with any differences in functionality or expression with respect to drug transport [20].

There are many polymorphisms in SLCO1B1 (which encodes OATP1B1) that have been associated with a decreased transport phenotype toward several drugs (see Table 1) and endogenous substrates [23]. In vitro assays have consistently validated altered transport efficiency in at least 13 synonymous and non-synonymous polymorphisms. At least three of these SNPs, the −11187G>A, the 388A>G (SLCO1B1*1b), and the 521T>C (SLCO1B1*5), have been shown to influence clinical outcome, although the allele frequencies differ between races. For example, while the SLCO1B1*5 polymorphism is present in approximately 14% of the Caucasian population [45], only 1% of Japanese subjects carry this allele [46]. For this reason, studies evaluating associations between SLCO1B1*5 and clinical outcome in Caucasians have been more statistically powered and have resulted in clearer clinical outcomes [45, 47, 48]. The SLCO1B3 gene has four polymorphisms (334T>G, 699G>A, 1564G>T, 1748G>A) that have been associated with altered transport and differences. It was recently determined that a common haplotype consisting of the 334T>G (Sl 12A) and 699G>A (M2331) SNPs was related to altered OATP1B3 transport characteristics in COS-7 cells, while no differences in the transport of cells transfected were observed with either variant alone [24]. However, this observation may be substrate- or assay-specific given that paclitaxel transport was not altered based on any of the SNPs (334T>G, 699G>A, 1564G>T) or haplotype combinations thereof in Xenopus oocytes [49].

Many of the recent publications regarding transporter genotyping have utilized restriction fragment length polymorphism (RFLP) analysis or direct sequencing, although several other methods of genotyping are available such as resequencing, allele-specific PCR, TaqMan PCR, Fluorescence Resonance Energy Transfer (FRET), etc. Table 2 includes polymerase chain reaction (PCR) and direct sequencing PCR primers that are currently used in the field to amplify polymorphic regions of DNA.

Table 2.

Primers used to amplify selected polymorphisms in the ABCB1, ABCG2, OATP1B1, and OATP1B3 genes

Transporter polymorphism Forward primer Reverse primer Ref.
ABCB1 1236C>T
 PCR amplification 5′-GGCACAAACCAGATAATATTAAGG-3′ 5′-TATCCTGTCCATCAACACTGACC-3′ #
 Nested PCR 5′-GTTCACTTCAGTTACCCATCTCG-3′ 5′-TCCTGTCCATCAACACTGACCTG-3′ #
 Sequencing PCR 5′-GTCAGTTCCTATATCCTGTGTCTG-3′ 5′-TCGCATGGGTCATCTCACCATC-3′ #
ABCB1 2677G>T/A (A893S/T)
 PCR amplification 5′-AGGCTATAGGTTCCAGGCTTGC-3′ 5′-AGAACAGTGTGAAGACAATGGCC-3′ #
 Nested PCR 5′-CCCATCATTCGAAT AGCAGGAG-3′ 5′-GAACAGTGTGAAGACAATGGCCT-3′ #
 Sequencing PCR 5′-ATCCTTCATCTATGGTTGGCAAC-3′ 5′-TGAGTCCAAGAACTGGCTTTGC-3′ #
ABCB1 3435C>T
 PCR amplification 5′-ATCTCACAGTAACTTGGCAGTTTC-3′ 5′-AACCCAAACAGGAAGTGTGGCC-3′ #
 Sequencing PCR 5′-GCTGGTCCTGAAGTTGATCTGTG-3′ 5′-AAACAGGAAGTGTGGCCAGATGC-3′ #
ABCG2 421C>T
 PCR amplification 5′-TGGCAAATCCTTGTATGAAGCAG-3′ 5′-TTCACGTACAACACCACATTGCC-3′ #
 Sequencing PCR 5′-GCAGGTTCATCA TTAGCTAGAAC-3′ 5′-CCTACTTATGCTGATCATGAGC-3′ #
OATP1B1*1b
 PCR amplification 5′-GCAAATAAAGGGGAATATTTCTC-3′ 5′-AGAGATGTAATTAAATGTATAC-3′ [45]
 RFLP enzyme ClaI
OATP1B1*5
 PCR amplification 5′-GTTAAATTTGTAATAGAAATGC-3′ 5′-GTAGACAAAGGGAAAGTGATCATA-3′ [45]
 Allele specific PCR primers WT 5′-CATACATGTGGATATATGT-3′ N/A [45]
MT 5′-CATACATGTGGATATATGC-3′D
OATP1B3 334T>G (S112A)
 PCR amplification 5′-CCTTCACAGTTAAATTACATGGTC-3′ 5′-TATTCATTTCATATAAAACTGTATACC-3′ #
 Sequencing PCR 5′-GGGCATATTTGCATTCATTTGGG-3′ 5′-CATGATAAATAAAGAAATACATGATG-3′ #
OATP1 B3 699G>A (M2331)
 PCR amplification 5′-TCCTTGTATTTAGGT AACGTACAG-3′ 5′-TCAAGTTTGGTTA TTTTGGATCAAG-3′ #
 Sequencing PCR 5′-GATCTACCCTTGAAATAATAATGTC-3′ 5′GTAAAAGCAAAGTATAAATAGGAGC-3′ #
OATP1B3 l564G>T (G522C)
 PCR amplification 5′-ATATACAGAATTTCATACACTAATTTC-3′ 5′-AATTCTAAGAAAATGCATTCTCAAAG-3′ #
 Sequencing PCR 5′-TATTTTGCCTTCACTATTAAGCAAC-3′ 5′-AATATGAATTTGAGCTCAAAATACAG-3′ #

Primers are used for direct sequencing following amplification from genomic DNA, unless otherwise indicated

#

Denotes previously unpublished primer

Genetic sequence variation may provide useful information to assist in making clinical decisions about drug treatment. Like all potential prognostic markers, the effect of polymorphisms on clinical endpoints must be validated through numerous preclinical and clinical processes that will be mentioned in the following subsections. Studies on the pharmacogenetics of transporters should (1) establish in silico and experimental evidence that a transporter polymorphism is associated with inter-individual variability in drug treatment, (2) establish drug interaction with a transporter, (3) establish that a polymorphism results in differential drug transport in vitro, (4) verify that transporter function is potentially important in vivo using animal models, (5) validate that the polymorphism is associated with clinical inter-individual variability of drug treatment in specific drug-treated populations with specific measurement methods of specific endpoints, and (6) validate the precision, reproducibility, and accuracy for clinical endpoint measurement [50]. Finally, established causative variants require CUA-certified, FDA-approved genotyping methods and applications to therapy in order to be useful to the population at large. Only then these polymorphisms useful in predicting clinical outcome in the general public, and the utility of these methods are only relevant to informed physicians that can apply the results toward changes in therapy.

Examples of established SNPs and genotyping methods can be found in those variants in TPMT, UGT1A9, CYP2D6, CYP2C9, and VKORC1 that are relevant for an a priori assessment of the starting dose and/or clinical outcome of 6-mercaptopurine, irinotecan, tamoxifen, and warfarin (CYP2C9 and VKORC1) treatment, respectively. Genotyping for SNPs in these genes is recommended by the FDA and can be achieved by CLIA-certified genotyping services, many of which use the AmpliChip P450 or the CodeLink P450 genotyping platforms. However, to our knowledge, no genetic variation in a drug transporter has yet been evaluated by the FDA, and no CUA-certified genotyping platform has been developed to genotype transporter variations. Although FDA approval and CUA certification remain to be worked out, the drug metabolizing enzyme transporter (DMET) platform may provide a basis to evaluate hundreds of polymorphisms in drug transporters and factors that regulate transporter expression (i.e., PXR) in future clinical trials. A brief overview of these genotyping platforms is reviewed in [51].

Substrate Identification

ABCB1 and ABCG2 substrates (see Table 1) are typically hydrophobic molecules such as lipids, peptides, steroids, and xenobiotics, and include anticancer, HIV, atypical antipsychotic, and immunosuppressant drugs. There is often broad overlap between ABCB1 and ABCG2 substrates. The ABCC proteins are multispecific anion transporters. ABCC1 is known to be involved in anthracycline transport [52], but ABCC2 effluxes a wider range of drugs such as cyclosporin A, cisplatin, vinblastine, and camptothecan derivatives [19, 53]. OATP1B1 and OATP1B3 also interact with a wide range of substrates (not only organic anions as the nomenclature implies) including bilirubin, bile acids [54], peptides, eicosanoids, hormones, flavonoids [55], and prescribed drugs including fexofenadine [56]. However, each transporter has distinct substrate specificity, so some compounds are transported by one transporter but not by another in the same family.

Substrates are usually identified using transfected MDCK, LLC-PK1, Caco-2, or endothelial cell lines expressing the transporter of interest. Often following selection for cells expressing the transporter of interest, these cell types are grown in a monolayer on a membrane separating two chambers of culture medium (i.e., the Transwell Cell Culture Assay, Corning Costar Corp., Cambridge, MA). Drug is administered into one chamber, and drug transport across the monolayer is evaluated by sampling from the other chamber. The experiment is then repeated by applying drug to the opposite chamber. Due to the directionality of the transporters, these experimental systems allow investigators to assess the basolateral to apical and apical to basolateral transport of drug. If the drug is a substrate for the transporter, the A:B and B:A ratios will differ significantly. Several other methods also exist to evaluate the transport capabilities toward individual drugs such as ATPase assays, and the transport of fluorescent or radiolabeled compounds into and out of cells that are native expressing, drug-selected expressing, or transfected with a transporter of interest [57, 58].

Assessing Functional Significance of Polymorphisms In Vitro

Cell-Based Assays

Polymorphic efflux of ABCB1 substrates was initially evaluated using flow cytometry, although such assays are limited in that only fluorescent compounds can be assayed and differences in polymorphic transporter expression and function are not made clear. To date, the influx of Rhodamine 123, JAI-51, calceine, doxorubicin, and daunorubicin have been evaluated using such methods and are still used in drug–drug interaction studies [59]. The same technique has been used with mitoxantrone to assess transport by, and inhibition of, ABCG2. Such assays were initially used in the field of transporter pharmacogenetics to show that Rhodamine 123 transport is lower in 3435TT human CD56+ cells [60]. As the pharmacokinetics of many other drugs could potentially be altered based on polymorphic ABCB1 expression and function, many have evaluated ABCB1 efflux using other in vitro assays. Some have used transfected cell lines to evaluate the functional significance of non-synonymous polymorphisms in ABCB1 and have demonstrated that differences in activity exist between proteins carrying a single amino acid difference brought on by these SNPs. For example, using this technique, it was found that the 2677G>T/A (893S>T/A) polymorphism results in activity differences toward vincristine such that Vmax 893T>893S>893A, while Km 893S>893T/A [61]. Other investigators have employed ATPase assays to evaluate the ATP-dependent active transport of substrates. In this assay, vesicles obtained from Sf9 cells transfected with ABCB 1 variants have been studied and have validated the previously mentioned finding with ABCB1 [62]. The effect of different polymorphisms on substrate transport by ABCG2 has been assessed using stably transfected HEK293 cells [63]. Following incubation of the cells with the drug, concentrations can be measured via flow cytometry [41], liquid scintillation counting (if radiolabeled drug is available) [64], or LC-MS [65]. In vitro analyses of OATP1B1 functional polymorphisms were evaluated similarly [45, 46, 6669]. Interestingly, the above assays have also been employed to address the functional consequences of polymorphisms in the ABCC family of transporters, but no notable alterations in transport capacity were found for many commonly studied SNPs in ABCC1 [43]. It seems that while ABCC transporters contain several potentially important polymorphisms and are important in drug transport overall, functional variability is actually quite low. This is perhaps the reason for the multiple negative studies that have assessed ABCC polymorphisms as they relate to drug bioavailability [20].

Assessing the Cause of Phenotypic Differences

Polymorphic differences that result in altered transporter kinetics and possible subsequent changes in drug disposition can effect this change via multiple mechanisms, including modulated tissue expression. For example, the ABCB1 2677TT genotype was associated with decreased mRNA expression in several human tissues as compared to the wild-type allele [60, 70, 71]; thus, the functional consequences of the 2677G>T/A polymorphism may be explained by expression alterations alone and not necessarily by altered substrate binding or transport efficiency of the protein. Some postulate that polymorphisms encoding rarer codons for the same amino acid (a synonymous or silent mutation) result in decreased translation efficiency of the mRNA, resulting in lower protein levels. Possible alterations in polymorphic mRNA secondary structure could also result in inefficient translation leading to alterations in protein folding [32]. This mechanism has been suggested as one possible explanation for the effects seen with the synonymous ABCB1 3435C>T transition that nevertheless is associated with differential drug efflux capability. An alternate, though not mutually exclusive, explanation has also been proposed ; the 3435C>T SNP is in linkage with the non-synonymous 2677G>T (893T>S) transition, and therefore may be associated with a protein product with attenuated efflux capacity through lowered efflux efficiency.

The first hypothesis has been evaluated using mRNA expression measurements in human tissues, and researchers found that ABCB1 is generally expressed at higher levels with the 3435C allele [60, 7072]. These observations were replicated with co-transfection of equal amounts of plasmid, and it was concluded that the 3435T allele lowers mRNA stability and is therefore responsible for decreased efflux capacity [73]. In the case of ABCG2, the effect of the 421C>A polymorphism has been debated. Originally, the resulting amino acid change was believed to reduce protein expression due to instability [37], but this finding was not confirmed by human intestinal samples [74]. Subsequently, it has been shown that the transport efficiency of the protein is decreased. This was demonstrated by measuring ATPase activity in wild-type and mutant cells, normalizing for expression [35].

When OATP1B1 variants were expressed in HeLa cells, it was noted that SLCO1B1*2, *3, *5, *6, *9, *12, and *13 alleles were associated with reduced transport of OATP1B1 substrates [45]. Others noted that when the SLCO1B1*15 variant was expressed in HEK293 cells and Xenopus laevis oocytes, these cells also had reduced transport capability [66, 68]. The reasons for the reduced transport capacity of these alleles was made clear after it was demonstrated that the plasma membrane localization of many of these polymorphic transporters was impaired due to a cell surface trafficking defect [45]. It was also shown that some polymorphisms encode for impaired protein maturation that results in intracellular retention of the OATP1B1 protein [67]. A single study was published with regard to the 334T>G (S112A), the 699G>A (M233I), and the 1564G>T (G522C) in SLCO1B3 (encoding OATP1B3) [75]. No differences were observed between wild-type and variant alleles at 334T>G or 699G>A with regard to protein localization, but the 1564G>T SNP conferred a phenotype where the protein was retained intracellularly in MDCKII cells [75].

Despite the encouraging results of these investigations, not all studies using the above experimental systems have consistently validated these observations in other tissues and cell types. For example, associations between genotype and expression seem to be tissue-specific, as lymphocytes and the small intestine both express ABCB1. However, expression levels were not associated with polymorphic variants, and it is often the case that reports evaluating the same tissues conflict [5]. Furthermore, some tissues such as cardiac endothelium actually express ABCB1 at greater levels in patients carrying variant alleles, which differs from data generated in other tissues [72]. Others have used non-human in vitro expression systems in an attempt to validate the effect of ABCB1 polymorphisms. However, transfected variant alleles do not seem to influence ABCBl transport in some of these experimental systems—perhaps due to differences in mRNA processing membranes in different cell lines and between species [76].

Assessing Functional Significance of Polymorphisms In Vivo

Mice carry two homologs of ABCB1 (Abcb1a, Abcb1b) and viable single (Abcb1a−/−) and double (Abcbla/b−/−) knockout mice are commercially available (Taconic Laboratories). Triple knockout (TKO) mice have recently become available in which homologous genes encoding ABCB1 and ABCC family members (covered later) have been removed from the mouse genome. An Abcg2 (the mouse homolog of ABCG2) knockout mouse has also recently become commercially available from Taconic Laboratories, in addition to a triple knockout, null for Abcb1a, Abcb1b, and Abcg2. Many have utilized such mice to evaluate the influence of ABC transporters on the pharmacokinetics and toxicity of drugs. Based on data obtained from these mice, ABCB1 has been shown to play a major role in detoxification and serves as a protective barrier against the toxic effects of xenobiotics [77]. These knockouts have been used as animal models of compromised blood–brain barrier function [9, 78], intestinal drug absorption [79], fetal drug exposure [80], and drug-induced damage to testicular tubules, choroid plexus epithelium [18], oropharyngeal mucosa [16], and peripheral nervous tissues [17].

Mice lacking the expression of a transporter generally have decreased ability to eliminate substrate drugs, except in cases where compensatory pathways are upregulated that circumvent transporter-mediated clearance [81, 82]. Alterations in plasma pharmacokinetics result from the lack of transporter expression in gut, liver, and renal tissues where several transporters are involved in the elimination of substrate drugs through hepatobilliary pathways and glomerular filtration. Such mice generally also demonstrate increased uptake of oral substrate drugs, as efflux transporters are involved in the excretion of toxic substances back into the gut lumen in normal mice. As such, bioavailability and exposure are usually increased in knockout mice, while clearance is decreased. This can have both positive and negative effects and can allow translational researchers to make clinical decisions based on the outcome of these drug-treated mouse models. However, this is not necessarily always the case. Compounds that are highly bioavailable in wild-type mice are unlikely to show great increases in absorption when the transporter protein is impaired, though decreased elimination may be observed. Also, as mentioned previously, many drugs have alternate routes of elimination, which may become more important when the primary transport mechanism is not functioning. As such, it is critical that in vivo testing is carried out for each compound, rather than assuming that because a drug is a substrate, it will be greatly affected by these polymorphisms. Furthermore, it is essential that agents which are transported by multiple transporters be tested in models representative of this physiology. For example, brain:­ plasma ratios of imatinib, a substrate for both Abcb1 and Abcg2, were only marginally higher than control in Abcg2(−/−) mice, yet 4-fold higher in Abcb1a/b(−/−) mice and 28-fold higher in double knockouts, suggesting that there is some synergy between the transporters [83].

Mice that do not express a specific transporter are generally more likely to experience benefit from treatment with a substrate drug because greater exposure to the drug is typically equated with improved efficacy. Lack of transporter function may also allow penetration into tissues that were previously impermeable to the agent. For example, Abcb1 knockout mice with brain metastases can be successfully treated with drugs that otherwise would not penetrate the blood–brain barrier such as paclitaxel [84]. Abcb1a−/− mice also showed 10 times higher brain–serum ratios of both risperidone and its active metabolite, 9-hydroxyrisperidone, than control mice [85], and many central nervous drugs showed 1.1- to 2.6-fold greater brain-to-plasma ratios in double knockout mice compared to wild-type mice [86].

Although the efficacy of drug treatment may increase, this is counterbalanced by increases in toxicity through routes other than increased plasma concentrations as blood–tissue barriers are disrupted allowing increased penetration of drugs into organs–especially the brain, where ABCB1 is an important mediator of drug exposure. In drugs with a narrow therapeutic window (e.g., many anticancer agents), the toxicity can outweigh the beneficial aspects of drug treatment. Following the above example, Abcb1 knockout mice treated with paclitaxel are more susceptible to treatment-related peripheral neuropathy due to increases in drug concentrations in nerve cells [17].

Transporter Genetics in Clinical Pharmacology

Initially, investigators determined that the ABCB1 3435C>T transition was associated with lowered ABCB1 expression and higher digoxin levels in vivo [70]. The association was stronger when the ABCB1 2677G>T/A and 3435C>T polymorphisms were evaluated together as a haplotype — those patients variant at both alleles having both the lowest ABCB1 expression and the highest digoxin AUC [87, 88]. Since then, many investigators have found similar associations between these polymorphisms and plasma concentrations of several other drugs, although these observations have not been consistently confirmed [13, 20, 89]. Often a polymorphism is found to be potentially important to drug treatment based on in vitro and in vivo evidence, but in clinical studies the polymorphism is not significantly associated with clinical parameters. Table 3 provides examples of transporter polymorphisms that have had an effect on certain substrates in vitro and/or in vivo that have or have not explained inter-individual variability toward these drugs in clinical pharmacology. While it is exceedingly difficult to assign a true mechanism behind the clinical findings, the in vitro and in vivo data provide powerful tools to ascertain potential mechanisms and provide the basis on which to form specific hypotheses that can be tested in the clinic. Nonetheless, one cannot say with certainty whether or not the proposed mechanism is actually occurring, and this is a major limitation of pharmacogenetics studies in general. A polymorphism that has no effect on a substrate drug in the clinic does not preclude a potential gene–drug interaction with another substrate drug.

Table 3.

Selected examples of studies that have either verified, or detracted from the importance of transporter polymorphisms on drug treatment after in vitro or in vivo validation

Transporter Drug In vitro or in vivo verification of polymorphic effects Ref. Clinical verification of inter-individual variability Ref.
ABCB1
Fexofenadine Increased transport with 893Ser
Increased transport with 893Thr
[98] Decreased AUC in patients with 2677TT and 2677AA genotypes [30, 99]
Docetaxel Decreased expression of ABCB1 in survival unchanged. [70, 71] Docetaxel AUC and overall liver with 3435TT genotype [100102]
Docetaxel is transported by ABCB1 [103]
ABCG2
Topotecan Increased accumulation of drug in cells transfected with 421A (Q141K) [104] Increased bioavailability in heterozygous C421A patients [104]
Imatinib Increased accumulation of drug in cells transfected with 421A (Q141K) [105] No significant difference in AUC or Cmax in heterozygous C421 patients [105]
ABCC1
N/A None to date
ABCC2
N/A None to date
OATP1B1
Pravastatin Reduced membrane expression of OATP1B1*5 variant [45] Increased AUC in OATP1B1*5 [94, 106]
OATP1B3 Leuprolide and goserelin Reduced transport of testosterone with the 334G/699A haplotype [24] Increased progression-free survival in patients with 334G/699A haplotype [97]

These relationships are often dependent on route of administration, drug dosage, and schedule and can also be largely dependent on drug metabolism. For example, polymorphisms in the cytochrome P450s (CYPs) are often found to be associated with pharmacokinetics of transporter substrate drugs whereas ABCB1 polymorphisms are not. This is believed to occur because metabolism by the CYPs may be the rate-limiting step in drug clearance while variation in ABCB1 expression levels plays a lesser role in inter-individual variability. For example, if an impaired transporter limits transport out of the liver, then it is also possible that more metabolism occurs. However, the mechanistic relationship between transporter polymorphisms and drug plasma levels remains largely unclear, and the reasons that several drugs are more or less associated with ABCB1 polymorphisms across multiple studies in various racial populations also remains unclear.

Recent evidence suggests that polymorphic ABCB1 expression not only influences plasma pharmacokinetics, but also the degree to which drugs are able to penetrate into tissues that express ABCB1 (e.g., tumors, brain, HIV­infected cells, etc.) [90, 91]. As previously mentioned, drug penetration into tissues can be both efficacious (i.e., by increasing therapeutic effect) and deleterious (i.e., by increasing toxicity). ABCB1 is also deterministic of the intracellular concentration of drugs as it efftuxes drugs from the cytoplasm into the extracellular matrix in several cell types. Based on the above findings, the hypothesis can be formed that ABCB1 expression levels are associated with pharmacokinetic parameters. Since the efficacy and toxicity of drugs is ultimately determined by plasma pharmacokinetic parameters and by the degree to which drugs are able to penetrate into tissues, studies investigating ABCB1 polymorphisms as they relate to drug administration are likely to become increasingly important in the clinical setting. The effect of ABCG2 polymorphisms on clinical pharmacology has only recently begun to be evaluated. Thus far, associations between the 421C>A mutation and plasma pharmacokinetics have been evaluated for several drugs including topotecan, irinotecan, and imatinib. This polymorphism was shown to increase bioavailability of topotecan [17] and more than double exposure of rosuvastatin, a statin commonly used in the treatment high cholesterol [92]. Findings for other drugs were less deterministic. Increasingly, investigators have sought to correlate genotype with toxicity. It has been shown that variant haplotypes for ABCB1 (3435C/T, 1236C/T, 2677G/T) or ABCG2 (−15622C/T, 1143C/T) increased the likelihood of sunitinib toxicity, including hand-foot syndrome [93].

The clinical consequences of OATPs are still being investigated, although several studies have confirmed that SLCO polymorphisms are associated with inter-individual variability in drug treatment [23]. This is especially true for pravastatin, which reduces cholesterol biosynthesis. It has been suggested that reduced hepatic uptake of pravastatin resulting from decreased transport through OATP1B1 is responsible for the increased plasma AUC of pravastatin [94], the unfavorable plasma pharmacokinetics of cholesterol synthesis biomarkers [95], and the resulting decreased efficacy of the drug. Interestingly, an SLCO1B1 allele that was associated with lower pravastatin AUC (and presumably greater liver uptake) was also associated with increased efficacy [96], although in vitro assays do not confirm an altered transport efficiency [68]. An indirect gene–drug interaction between OATP1B3 and androgen­deprivation therapeutics (ADT; e.g., leuprolide and goserelin) was recently proposed where OATP1B3 may scavenge low levels of testosterone during ADT, thereby opposing the effects of ADT on androgen receptor signaling [97]. Thus, endogenous processes affected by transporters may also alter drug treatment in some cases.

Transporter Genetics in Clinical Endpoint Analysis

The ultimate research goal of transporter pharmacogenetics is to further our understanding of the ways in which transporter genetics influences clinical endpoints so that current drug treatment can be made safer and more efficacious and that investigational therapies can be better developed. The literature consists of a multitude of studies that have evaluated drug efficacy and toxicity and have made associations between these parameters and polymorphisms in drug transporters [5, 20, 23]. The FDA recommends several endpoints to evaluate specific diseases which should be used when making associations between a genetic variation and the treatment of diseases with drugs (see www.fda.gov/cder/guidance; last accessed September 16, 2009). In pharmacogenetic studies, these endpoints should be evaluated in a standard fashion in similar populations in order to establish the predictive value of a polymorphism. The genotypes should also be evaluated using a consistent set of SNPs in the relevant transporters. Unfortunately, the literature has not typically been consistent mainly due to the availability of samples for analysis, and perhaps this is the reason that transporter polymorphisms have not been consistently validated. However, the advent of genotyping technologies such as DMET may offer standardization of major pharmacogenetics studies in the future. Thus far, all studies linking pharmacogenomics of membrane transporters with clinical outcome have been retrospective and have taken place in eclectic populations with relatively low statistical power. It is essential that prospective studies are conducted, prior to any treatment modification, in order to assess the true effects of these polymorphisms and to determine whether the effect is drug-specific or disease related.

Acknowledgments

This project has been funded in whole or in part with federal funds from the National Cancer Institute, National Institutes of Health, under contract NOl-C0–12400 and HHSN261200800001E. * The content of this publication does not necessarily reflect the views or policies of the Department of Health and Human Services, nor does mention of trade names, commercial products, or organizations imply endorsement by the U.S. Government (*ERG). This work was supported by the Intramural Research Program of the NIH, National Cancer Institute, Center for Cancer Research.

Contributor Information

Tristan M. Sissung, Clinical Pharmacology Program, Medical Oncology Branch, Center for Cancer Research, National Cancer Institute, 9000 Rockville Pike, Building 10, Room 5A01, Bethesda, MD 20892, USA, Sissungt@mail.nih.gov

Caitlin E. Baum, Molecular Pharmacology Section, Medical Oncology Branch, Center for Cancer Research, National Cancer Institute, 9000 Rockville Pike, Building 10, Room 5A01, Bethesda, MD 20892, USA, baumcai@mail.nih.gov

C. Tyler Kirkland, Molecular Pharmacology Section, Medical Oncology Branch, Center for Cancer Research, National Cancer Institute, 9000 Rockville Pike, Building 10, Room 5A01, Bethesda, MD 20892, USA, kirklandct@mail.nih.gov.

Rui Gao, Molecular Pharmacology Section, Medical Oncology Branch, Center for Cancer Research, National Cancer Institute, 9000 Rockville Pike, Building 10, Room 5A01, Bethesda, MD 20892, USA, gaor@mail.nih.gov.

Erin R. Gardner, Clinical Pharmacology Program, SAIC-Frederick, NCI-Frederick, Frederick, MD 21702, USA, gardnerer@mail.nih.gov

William D. Figg, Clinical Pharmacology Program, Medical Oncology Branch, Center for Cancer Research, National Cancer Institute, 9000 Rockville Pike, Building 10, Room 5A01, Bethesda, MD 20892, USA; Molecular Pharmacology Section, Medical Oncology Branch, Center for Cancer Research, National Cancer Institute, 9000 Rockville Pike, Building 10, Room 5A01, Bethesda, MD 20892, USA

References

  • 1.Borst P, Evers R, Kool M, & Wijnholds J (2000). A family of drug transporters: The multidrug resistance-associated proteins. Journal of the National Cancer Institute, 92, 1295–1302. [DOI] [PubMed] [Google Scholar]
  • 2.Dean M, Rzhetsky A, & Allikmets R (2001). The human ATP-binding cassette (ABC) transporter superfamily. Genome Research, 11, 1156–1166. [DOI] [PubMed] [Google Scholar]
  • 3.Gottesman MM, & Ambudkar SV (2001). Overview: ABC transporters and human disease. Journal of Bioenergetics and Biomembranes, 33, 453–458. [DOI] [PubMed] [Google Scholar]
  • 4.Vasiliou V, Vasiliou K, & Nebert DW (2009). Human ATP-binding cassette (ABC) transporter family. Human Genomics, 3, 281–290. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Lepper ER, Nooter K, Verweij J, Acharya MR, Figg WD, & Sparreboom A (2005). Mechanisms of resistance to anticancer drugs: The role of the polymorphic ABC transporters ABCB1 and ABCG2. Pharmacogenomics, 6, 115–138. [DOI] [PubMed] [Google Scholar]
  • 6.Fojo AT, Shen DW, Mickley LA, Pastan I, & Gottesman MM (1987). Intrinsic drug resistance in human kidney cancer is associated with expression of a human multi­ drug-resistance gene. Journal of Clinical Oncology, 5, 1922–1927. [DOI] [PubMed] [Google Scholar]
  • 7.Maliepaard M, Scheffer GL, Faneyte IF, van Gastelen MA, Pijnenborg AC, Schinkel AH, et al. (2001). Subcellular localization and distribution of the breast cancer resistance protein transporter in normal human tissues. Cancer Research, 61, 3458–3464. [PubMed] [Google Scholar]
  • 8.Schellens JH, Malingre MM, Kruijtzer CM, Bardelmeijer HA, van Tellingen O, Schinkel AH, et al. (2000). Modulation of oral bioavailability of anticancer drugs: from mouse to man. European Journal of Pharmaceutical Science, 12, 103–110. [DOI] [PubMed] [Google Scholar]
  • 9.Schinkel AH, Mayer U, Wagenaar E, Mol CA, van Deemter L, Smit JJ, et al. (1997). Normal viability and altered pharmacokinetics in mice lacking mdrl-type (drug­ transporting) P-glycoproteins. Proceedings of the National Academy of Sciences USA, 94, 4028–4033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Thiebaut F, Tsuruo T, Hamada H, Gottesman MM, Pastan I, & Willingham MC (1987). Cellular localization of the multidrug-resistance gene product P-glycoprotein in normal human tissues. Proceedings of the National Academy of Sciences USA, 84, 7735–7738. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Xiao JJ, Foraker AB, Swaan PW, Liu S, Huang Y, Dai Z, et al. (2005). Efflux of depsipeptide FK228 (FR90l228, NSC-630 l76) is mediated by P-glycoprotein and multidrug resistance-associated protein l. Journal of Pharmacology and Experimental Therapeutics, 313, 268–276. [DOI] [PubMed] [Google Scholar]
  • 12.Chaudhary PM, & Roninson IB (1991). Expression and activity of P-glycoprotein, a multidrug efflux pump, in human hematopoietic stem cells. Cell, 66, 85–94. [DOI] [PubMed] [Google Scholar]
  • 13.Eichelbaum M, Fromm MF, & Schwab M (2004). Clinical aspects of the MDR l (ABCB1) gene polymorphism. Therapeutic Drug Monitoring, 26, 180–185. [DOI] [PubMed] [Google Scholar]
  • 14.Fromm MF (2004). Importance of P-glycoprotein at blood­tissue barriers. Trends in Pharmacological Sciences, 25, 423–429. [DOI] [PubMed] [Google Scholar]
  • 15.Meissner K, Sperker B, Karsten C, Zu Schwabedissen HM, Seeland U, Bohm M, et al. (2002). Expression and localization of P-glycoprotein in human heart: Effects of cardiomyopathy. Journal of Histochemistry and Cytochemistry, 50, 1351–1356. [DOI] [PubMed] [Google Scholar]
  • 16.Rao VV, Dahlheimer JL, Bardgett ME, Snyder AZ, Finch RA, Sartorelli AC, et al. (1999). Choroid plexus epithelial expression of MDRl P glycoprotein and multidrug resistance-associated protein contribute to the blood-cerebrospinal-fluid drug-permeability barrier. Proceedings of the National Academy of Sciences USA, 96, 3900–3905. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Saito T, Zhang ZJ, Ohtsubo T, Noda I, Shibamori Y, Yamamoto T, et al. (2001). Homozygous disruption of the mdrla P-glycoprotein gene affects blood-nerve barrier function in mice administered with neurotoxic drugs. Acta Oto-Larynologica, 121, 735–742. [DOI] [PubMed] [Google Scholar]
  • 18.Wijnholds J, deLange EC, Scheffer GL, van den Berg DJ, Mol CA, van der Valk M, et al. (2000). Multidrug resistance protein 1 protects the choroid plexus epithelium and contributes to the blood-cerebrospinal fluid barrier. Journal of Clinical Investigation, 105, 279–285. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Borst P, Evers R, Kool M, & Wijnholds J (1999). The multidrug resistance protein family. Biochimica et Biophysica Acta, 1461, 347–357. [DOI] [PubMed] [Google Scholar]
  • 20.Cascorbi I (2006). Role of pharmacogenetics of ATP-binding cassette transporters in the pharmacokinetics of drugs. Pharmacology and Therapeutics, 112, 457–473. [DOI] [PubMed] [Google Scholar]
  • 21.Deeley RG, & Cole SP (2006). Substrate recognition and transport by multidrug resistance protein 1 (ABCC l). FEBS Letters, 580, l103–1111. [DOI] [PubMed] [Google Scholar]
  • 22.Ho RH, & Kim RB (2005). Transporters and drug therapy: Implications for drug disposition and disease. Clinical Pharmacology and Therapeutics, 78, 260–277. [DOI] [PubMed] [Google Scholar]
  • 23.Smith NF, Figg WD, & Sparreboom A (2005). Role of the liver-specific transporters OATP1B1 and OATP1B3 in governing drug elimination. Expert Opinion on Drug Metabolism and Toxicology, 1, 429–445. [DOI] [PubMed] [Google Scholar]
  • 24.Hamada A, Sissung T, Price DK, Danesi R, Chau CH, Sharifi N, et al. (2008). Effect of SLCOlB3 haplotype on testosterone transport and clinical outcome in caucasian patients with androgen-independent prostatic cancer. Clinical Cancer Research, 14, 3312–3318. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Lee W, Belkhiri A, Lockhart AC, Merchant N., Glaeser H, Harris EI, et al. (2008). Overexpression of OATP1B3 confers apoptotic resistance in colon cancer. Cancer Research, 68, 10315–10323. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Narita M, Hatano E, Arizono S, Miyagawa-Hayashino A, Isoda H, Kitamura K, et al. (2009). Expression of OATP1B3 determines uptake of Gd-EOB-DTPA in hepatocellular carcinoma. Journal of Gastroenterology, 44, 793–798. [DOI] [PubMed] [Google Scholar]
  • 27.Sharom FJ (2008). ABC multidrug transporters: Structure, function and role in chemoresistance. Pharmacogenomics, 9, 105–127. [DOI] [PubMed] [Google Scholar]
  • 28.Kurata Y, Ieiri I, Kimura M, Morita T, Irie S, Urae A, et al. (2002). Role of human MDRl gene polymorphism in bioavailability and interaction of digoxin, a substrate of P-glycoprotein. Clinical Pharmacology and Therapeutics, 72, 209–219. [DOI] [PubMed] [Google Scholar]
  • 29.Tanabe M, Ieiri I, Nagata N, Inoue K, Ito S, Kanamori Y, et al. (2001). Expression of P-glycoprotein in human placenta: Relation to genetic polymorphism of the multidrug resistance (MDR)-l gene. Journal of Pharmacology and Experimental Therapeutics, 297, 1137–1143. [PubMed] [Google Scholar]
  • 30.Yi SY, Hong KS, Lim HS, Chung JY, Oh DS, Kim JR, et al. (2004). A variant 2677A allele of the MDR l gene affects fexofenadine disposition. Clinical Pharmacology and Therapeutics, 76, 418–427. [DOI] [PubMed] [Google Scholar]
  • 31.Sun J, He ZG, Cheng G, Wang SJ, Hao XH, & Zou MJ (2004). Multidrug resistance P-glycoprotein: Crucial significance in drug disposition and interaction. Medical Science Monitor, 10, RA5–RA14. [PubMed] [Google Scholar]
  • 32.Kimchi-Sarfaty C, Oh JM, Kim IW, Sauna ZE, Calcagno AM, Ambudkar SV, et al. (2007). A “silent” polymorphism in the MDRl gene changes substrate specificity. Science, 315, 525–528. [DOI] [PubMed] [Google Scholar]
  • 33.Tsai CJ, Sauna ZE, Kimchi-Sarfaty C, Ambudkar SV, Gottesman MM, & Nussinov R (2008). Synonymous mutations and ribosome stalling can lead to altered folding pathways and distinct minima. Journal of Molecular Biology, 383, 281–291. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Tamura A, Wakabayashi K, Onishi Y, Takeda M, Ikegami Y, Sawada S, et al. (2007). Re-evaluation and functional classification of non-synonymous single nucleotide polymorphisms of the human ATP-binding cassette transporter ABCG2. Cancer Science, 98, 231–239. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Mizuarai S, Aozasa N., & Kotani H (2004). Single nucleotide polymorphisms result in impaired membrane localization and reduced atpase activity in multidrug transporter ABCG2. International Journal of Cancer, 109, 238–246. [DOI] [PubMed] [Google Scholar]
  • 36.de Jong FA, Marsh S, Mathijssen RH, King C, Verweij J, Sparreboom A, et al. (2004). ABCG2 pharmacogenetics: Ethnic differences in allele frequency and assessment of influence on irinotecan disposition. Clinical Cancer Research, 10, 5889–5894. [DOI] [PubMed] [Google Scholar]
  • 37.Imai Y, Nakane M, Kage K, Tsukahara S, Ishikawa E, Tsuruo T, et al. (2002). C421A polymorphism in the human breast cancer resistance protein gene is associated with low expression of Ql41K protein and low-level drug resistance. Molecular Cancer Therapeutics, 1, 611–616. [PubMed] [Google Scholar]
  • 38.Kondo C, Suzuki H, Itoda M, Ozawa S, Sawada J, Kobayashi D, et al. (2004). Functional analysis of SNPs variants of BCRP/ABCG2. Pharmaceutical Research, 21, 1895–1903. [DOI] [PubMed] [Google Scholar]
  • 39.Allen JD, Jackson SC, & Schinkel AH (2002). A mutation hot spot in the Bcrp I (Abcg2) multidrug transporter in mouse cell lines selected for Doxorubicin resistance. Cancer Research, 62, 2294–2299. [PubMed] [Google Scholar]
  • 40.Honjo Y, Hrycyna CA, Yan QW, Medina-Perez WY, Robey RW, van de Laar A, et al. (2001). Acquired mutations in the MXR/BCRP/ABCP gene alter substrate specificity in MXR/BCRP/ABCP-overexpressing cells. Cancer Research, 61, 6635–6639. [PubMed] [Google Scholar]
  • 41.Robey RW, Honjo Y, Morisaki K, Nadjem TA, Runge S, Risbood M, et al. (2003). Mutations at amino-acid 482 in the ABCG2 gene affect substrate and antagonist specificity. British Journal of Cancer, 89, 1971–1978. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Robey RW, Steadman K, Polgar O,Morisaki K, Blayney M, Mistry P, et al. (2004). Pheophorbide a is a specific probe for ABCG2 function and inhibition. Cancer Research, 64, 1242–1246. [DOI] [PubMed] [Google Scholar]
  • 43.Letourneau IJ, Deeley RG, & Cole SP (2005). Functional characterization of non-synonymous single nucleotide polymorphisms in the gene encoding human multidrug resistance protein I (MRP1/ABCC1). Pharmacogenetics Genomics, 15, 647–657. [DOI] [PubMed] [Google Scholar]
  • 44.Oselin K, Mrozikiewicz PM, Gaikovitch E, Pahkla R, & Roots I (2003). Frequency of MRPI genetic polymorphisms and their functional significance in Caucasians: Detection of a novel mutation G816A in the human MRPI gene. European Journal of Clinical Pharmacology, 59, 347–350. [DOI] [PubMed] [Google Scholar]
  • 45.Tirona RG, Leake BF, Merino G, & Kim RB (2001). Polymorphisms in OATP-C: Identification of multiple allelic variants associated with altered transport activity among European- and African-Americans. Journal of Biological Chemistry, 276, 35669–35675. [DOI] [PubMed] [Google Scholar]
  • 46.Nozawa T, Nakajima M, Tamai I, Noda K, Nezu J, Sai Y, et al. (2002). Genetic polymorphisms of human organic anion transporters OATP-C (SLC21A6) and OATP-B (SLC21A9): Allele frequencies in the Japanese population and functional analysis. Journal of Pharmacology and Experimental Therapeutics, 302, 804–813. [DOI] [PubMed] [Google Scholar]
  • 47.Ho RH, Choi L, Lee W, Mayo G, Schwarz UI, Tirona RG, et al. (2007). Effect of drug transporter genotypes on pravastatin disposition in European- and African-American participants. Pharmacogenetics Genomics, 17, 647–656. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Mwinyi J, Kopke K, Schaefer M, Roots I, & Gerloff T (2008). Comparison of SLCO lB I sequence variability among German, Turkish, and African populations. European Journal of Clinical Pharmacology, 64, 257–266. [DOI] [PubMed] [Google Scholar]
  • 49.Smith NF, Marsh S, Scott-Horton TJ, Hamada A, Mielke S, Mross K, et al. (2007). Variants in the SLCOIB3 gene: interethnic distribution and association with paclitaxel pharmacokinetics. Clinical Pharmacology and Therapeutics, 81, 76–82. [DOI] [PubMed] [Google Scholar]
  • 50.Colburn WA (2003). Biomarkers in drug discovery and development: From target identification through drug marketing. Journal of Clinical Pharmacology, 43, 329–341. [DOI] [PubMed] [Google Scholar]
  • 51.Deeken J (2009). The Affymetrix DMET platform and pharmacogenetics in drug development. Current Opinion in Molecular Therapy, 11, 260–268. [PubMed] [Google Scholar]
  • 52.Cole SP, Bhardwaj G, Gerlach JH, Mackie JE, Grant CE, Almquist KC, et al. (1992). Overexpression of a transporter gene in a multidrug-resistant human lung cancer cell line. Science, 258, 1650–1654. [DOI] [PubMed] [Google Scholar]
  • 53.Hesselink DA, van Schaik RH, van der Heiden IP, van der Werf M, Gregoor PJ, Lindemans J, et al. (2003). Genetic polymorphisms of the CYP3A4, CYP3A5, and MDR-1 genes and pharmacokinetics of the calcineurin inhibitors cyclosporine and tacrolimus. Clinical Pharmacology and Therapeutics, 74, 245–254. [DOI] [PubMed] [Google Scholar]
  • 54.Noe B, Hagenbuch B, Stieger B, & Meier PJ (1997). Isolation of a multispecific organic anion and cardiac glycoside transporter from rat brain. Proceedings of the National Academy of Sciences USA, 94, 10346–10350. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Wang X, Wolkoff AW, & Morris ME (2005). Flavonoids as a novel class of human organic anion-transporting polypeptide OATP1B1 (OATP-C) modulators. Drug Metabolism and Disposition, 33, 1666–1672. [DOI] [PubMed] [Google Scholar]
  • 56.Cvetkovic M, Leake B, Fromm MF, Wilkinson GR, & Kim RB (1999). OATP and P-glycoprotein transporters mediate the cellular uptake and excretion of fexofenadine. Drug Metabolism and Disposition, 27, 866–871. [PubMed] [Google Scholar]
  • 57.Aszalos A (2004). P-glycoprotein-based drug–drug interactions: Preclinical methods and relevance to clinical observations. Archives of Pharmacal Research, 27, 127–135. [DOI] [PubMed] [Google Scholar]
  • 58.Smith NF, Acharya MR, Desai N, Figg WD, & Spar­ reboom, A. (2005). Identification of OATPIB3 as a high-affinity hepatocellular transporter of paclitaxel. Cancer Biology & Therapy, 4, 815–818. [DOI] [PubMed] [Google Scholar]
  • 59.Boumendjel A, McLeer-Florin A, Champelovier P, Allegro D, Muhammad D, Souard F, et al. (2009). A novel chalcone derivative which acts as a microtubule depolymerising agent and an inhibitor of P-gp and BCRP in in vitro and in vivo glioblastoma models. BMC Cancer, 9, 242. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Hitzl M, Drescher S, van der Kuip H, Schaffeler E, Fischer J, Schwab M, et al. (2001). The C3435T mutation in the human MDRI gene is associated with altered efflux of the P­ glycoprotein substrate rhodamine 123 from CD56+ natural killer cells. Pharmacogenetics, 11, 293–298. [DOI] [PubMed] [Google Scholar]
  • 61.Schaefer M, Roots I, & Gerloff T (2005). In vitro transport characteristics discriminate wildtype mdr l (abcb I) from ala893ser and ala893thr polymorphisms. European Journal of Clinical Pharmacology, 61, 718. [DOI] [PubMed] [Google Scholar]
  • 62.Ishikawa T, Sakurai A, Kanamori Y, Nagakura M, Hirano H, Takarada Y, et al. (2005). High-speed screening of human ATP-binding cassette transporter function and genetic polymorphisms: New strategies in pharmacogenomics. Methods in Enzymology, 400, 485–510. [DOI] [PubMed] [Google Scholar]
  • 63.Morisaki K, Robey RW, Ozvegy-Laczka C, Honjo Y, Polgar O, Steadman K, et al. (2005). Single nucleotide polymorphisms modify the transporter activity of ABCG2. Cancer Chemotherapy and Pharmacology, 56, 161–172. [DOI] [PubMed] [Google Scholar]
  • 64.Zhang Y, Gupta A, Wang H, Zhou L, Vethanayagam RR, Unadkat JD, et al. (2005). BCRP transports dipyridamole and is inhibited by calcium channel blockers. Pharmaceutical Research, 22, 2023–2034. [DOI] [PubMed] [Google Scholar]
  • 65.Nakamura Y, Oka M, Soda H, Shiozawa K, Yoshikawa M, Itoh A, et al. (2005). Gefitinib (“Iressa “, ZDl839), an epidermal growth factor receptor tyrosine kinase inhibitor, reverses breast cancer resistance protein/ ABCG2-mediated drug resistance. Cancer Research, 65, 1541–1546. [DOI] [PubMed] [Google Scholar]
  • 66.lwai M, Suzuki H, Ieiri I, Otsubo K, & Sugiyama Y (2004). Functional analysis of single nucleotide polymorphisms of hepatic organic anion transporter OATP1B1 (OATP-C). Pharmacogenetics, 14, 749–757. [DOI] [PubMed] [Google Scholar]
  • 67.Michalski C, Cui Y, Nies AT, Nuessler AK, Neuhaus P, Zanger UM, et al. (2002). A naturally occurring mutation in the SLC2lA6 gene causing impaired membrane localization of the hepatocyte uptake transporter. Journal of Biological Chemistry, 277, 43058–43063. [DOI] [PubMed] [Google Scholar]
  • 68.Nozawa T, Minami H, Sugiura S, Tsuji A, & Tamai I (2005). Role of organic anion transporter OATP1B1 (OATP-C) in hepatic uptake of irinotecan and its active metabolite, 7-ethyl­10-hydroxycamptothecin: in vitro evidence and effect of single nucleotide polymorphisms. Drug Metabolism and Disposition, 33, 434–439. [DOI] [PubMed] [Google Scholar]
  • 69.Tirona RG, Leake BF, Wolkoff AW, & Kim RB (2003). Human organic anion transporting polypeptide-C (SLC2lA6) is a major determinant of rifampin-mediated pregnane X receptor activation. Journal of Pharmacology and Experimental Therapeutics, 304, 223–228. [DOI] [PubMed] [Google Scholar]
  • 70.Hoffmeyer S, Burk O, von Richter O, Arnold HP, Brockmoller J, Johne A, et al. (2000). Functional polymorphisms of the human multidrug-resistance gene: Multiple sequence variations and correlation of one allele with P-glycoprotein expression and activity in vivo. Proceedings of the National Academy of Sciences USA, 97, 3473–3478. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Song P, Lamba JK, Zhang L, Schuetz E, Shukla N, Meibohm B, et al. (2006). G2677T and C3435T genotype and haplotype are associated with hepatic ABCB1 (MDR 1) expression. Journal of Clinical Pharmacology, 46, 373–379. [DOI] [PubMed] [Google Scholar]
  • 72.Meissner K, Jedlitschky G, Meyer zu Schwabedissen H, Dazert P, Eckel L, Vogelgesang S, et al. (2004). Modulation of multidrug resistance P-glycoprotein l (ABCB l) expression in human heart by hereditary polymorphisms. Pharmacogenetics, 14, 381–385. [DOI] [PubMed] [Google Scholar]
  • 73.Wang D, Johnson AD, Papp AC, Kroetz DL, & Sadee W (2005). Multidrug resistance polypeptide I (MDRl, ABCB1) variant 3435C>T affects mRNA stability. Pharmacogenetics Genomics, 15, 693–704. [PubMed] [Google Scholar]
  • 74.Zamber CP, Lamba JK, Yasuda K, Farnum J, Thummel K, Schuetz JD, et al. (2003). Natural allelic variants of breast cancer resistance protein (BCRP) and their relationship to BCRP expression in human intestine. Pharmacogenetics, 13, 19–28. [DOI] [PubMed] [Google Scholar]
  • 75.Letschert K, Keppler D, & Konig J (2004). Mutations in the SLCO 1B3 gene affecting the substrate specificity of the hepatocellular uptake transporter OATP1B3 (OATP8). Pharmacogenetics, 14, 441–452. [DOI] [PubMed] [Google Scholar]
  • 76.Kimchi-Sarfaty C, Gribar JJ, & Gottesman MM (2002). Functional characterization of coding polymorphisms in the human MDR1l gene using a vaccinia virus expression system. Molecular Pharmacology, 62, l–6. [DOI] [PubMed] [Google Scholar]
  • 77.Lin JH, & Yamazaki M (2003). Role of P-glycoprotein in pharmacokinetics: Clinical implications. Clinical Pharmacokinetics, 42, 59–98. [DOI] [PubMed] [Google Scholar]
  • 78.Schinkel AH, Smit JJ, van Tellingen O, Beijnen JH, Wagenaar E, van Deemter L, et al. (1994). Disruption of the mouse mdr la P-glycoprotein gene leads to a deficiency in the blood-brain barrier and to increased sensitivity to drugs. Cell, 77, 491–502. [DOI] [PubMed] [Google Scholar]
  • 79.Sparreboom A, van Asperen J, Mayer U, Schinkel AH, Smit JW, Meijer DK, et al. (1997). Limited oral bioavailability and active epithelial excretion of paclitaxel (Taxol) caused by P-glycoprotein in the intestine. Proceedings of the National Academy of Sciences USA, 94, 2031–2035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Smit JW, Huisman MT, van Tellingen O, Wiltshire HR, & Schinkel AH (1999). Absence or pharmacological blocking of placental P-glycoprotein profoundly increases feta! drug exposure. Journal of Clinical Investigation, 104, 1441–1447. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Allen JD, Brinkhuis RF, van Deemter L, Wijnholds J, & Schinkel AH (2000). Extensive contribution of the multidrug transporters P-glycoprotein and Mrp1 to basal drug resistance. Cancer Research, 60, 5761–5766. [PubMed] [Google Scholar]
  • 82.Allen JD, Brinkhuis RF, Wijnholds J, & Schinkel AH (1999). The mouse Bcrpl/Mxr/Abcp gene: Amplification and overexpression in cell lines selected for resistance to topotecan, mitoxantrone, or doxorubicin. Cancer Research, 59, 4237–4241. [PubMed] [Google Scholar]
  • 83.Zhou L, Schmidt K, Nelson FR, Zelesky V, Troutman MD, & Feng B (2009). The effect of breast cancer resistance protein and P-glycoprotein on the brain penetration of flavopiridol, imatinib mesylate (Gleevec), prazosin, and 2-methoxy- 3-(4-(2-(5-methyl-2-pheny loxazol-4-y l)ethoxy)phenyl)propa­ noic acid (PF-407288) in mice. Drug Metabolism and Disposition, 37, 946–955. [DOI] [PubMed] [Google Scholar]
  • 84.Gallo JM, Li S, Guo P, Reed K, & Ma J (2003). The effect of P-glycoprotein on paclitaxel brain and brain tumor distribution in mice. Cancer Research, 63, 5114–5117. [PubMed] [Google Scholar]
  • 85.Ejsing TB, Pedersen AD, & Linnet K (2005). P-glycoprotein interaction with risperidone and 9–0H-risperidone studied in vitro, in knock-out mice and in drug-drug interaction experiments. Human Psychopharmacolo gy, 20, 493–500. [DOI] [PubMed] [Google Scholar]
  • 86.Doran A, Obach RS, Smith BJ, Hosea NA, Becker S, Callegari E, et al. (2005). The impact of P-glycoprotein on the disposition of drugs targeted for indications of the central nervous system: Evaluation using the MDR1N1B knockout mouse model. Drug Metabolism and Disposition, 33, 165–174. [DOI] [PubMed] [Google Scholar]
  • 87.Johne A, Kopke K, Gerloff T, Mai I, Rietbrock S, Meisel C, et al. (2002). Modulation of steady-state kinetics of digoxin by haplotypes of the P-glycoprotein MDR l gene. Clinical Pharmacology and Therapeutics, 72, 584–594. [DOI] [PubMed] [Google Scholar]
  • 88.Verstuyft C, Schwab M, Schaeffeler E, Kerb R, Brink­mann U, Jaillon P, et al. (2003). Digoxin pharmacokinetics and MDR1 genetic polymorphisms. European Journal of Clinical Pharmacology, 58, 809–812. [DOI] [PubMed] [Google Scholar]
  • 89.Sakaeda T (2005). MDR l genotype-related pharmacokinetics: Fact or fiction ? Drug Metabolism and Pharmacokinetics, 20, 391–414. [DOI] [PubMed] [Google Scholar]
  • 90.Lin SK, Su SF, & Pan CH (2006). Higher plasma drug concentration in clozapine-treated schizophrenic patients with side effects of obsessive/compulsive symptoms. Therapeutic Drug Monitoring, 28, 303–307. [DOI] [PubMed] [Google Scholar]
  • 91.Sissung TM, Mross K, Steinberg SM, Behringer D, Figg WD, Sparreboom A, et al. (2006). Association of ABCB1 genotypes with paclitaxel-mediated peripheral neuropathy and neutropenia. European Journal of Cancer, 42, 2893–2896. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Keskitalo JE, Zolk O, Fromm MF, Kurkinen KJ, Neuvonen PJ, & Niemi M (2009). ABCG2 polymorphism markedly affects the pharmacokinetics of atorvastatin and rosuvastatin. Clinical Pharmacology and Therapeutics, 86, 197–203. [DOI] [PubMed] [Google Scholar]
  • 93.van Erp NP, Eechoute K, van der Veldt AA, Haanen JB, Reyners AK, Mathijssen RH, et al. (2009). Pharmacogenetic pathway analysis for determination of sunitinib-induced toxicity. Journal of Clinical Oncology, 27, 4406–4412. [DOI] [PubMed] [Google Scholar]
  • 94.Niemi M, Schaeffeler E, Lang T, Fromm MF, Neuvonen M, Kyrklund C, et al. (2004). High plasma pravastatin concentrations are associated with single nucleotide polymorphisms and haplotypes of organic anion transporting polypeptide-C (OATP-C, SLCO1B1). Pharmacogenetics, 14, 429–440. [DOI] [PubMed] [Google Scholar]
  • 95.Niemi M, Neuvonen PJ, Hofmann U, Backman JT, Schwab M, Lutjohann D, et al. (2005). Acute effects of pravastatin on cholesterol synthesis are associated with SLCO1B1 (encoding OATP1B1) haplotype * 17. Pharmacogenetics Genomics, 15, 303–309. [DOI] [PubMed] [Google Scholar]
  • 96.Tachibana-Iimori R, Tabara Y, Kusuhara H, Kohara K, Kawamoto R, Nakura J, et al. (2004). Effect of genetic polymorphism of OATP-C (SLCO IB l) on lipid-lowering response to HMG-CoA reductase inhibitors. Drug Metabolism and Pharmacokinetics, 19, 375–380. [DOI] [PubMed] [Google Scholar]
  • 97.Sharifi N, Hamada A, Sissung T, Danesi R, Venzon D, Baum C, et al. (2008). A polymorphism in a transporter of testosterone is a determinant of androgen independence in prostate cancer. BJU International, 102, 617–621. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Schaefer M, Roots I, & Gerloff T (2006). In vitro transport characteristics discriminate wild-type ABCB1 (MDR1) from ALA893SER and ALA893THR polymorphisms. Pharmacogenetics Genomics, 16, 855–861. [DOI] [PubMed] [Google Scholar]
  • 99.Kim RB, Leake BF, Choo EF, Dresser GK, Kubba SV, Schwarz UI, et al. (2001). Identification of functionally variant MDRl alleles among European Americans and African Americans. Clinical Pharmacolog y and Therapeutics, 70, 189–199. [DOI] [PubMed] [Google Scholar]
  • 100.Goh BC, Lee SC, Wang LZ, Fan L, Guo JY, Lamba J, et al. (2002). Explaining interindividual variability of docetaxel pharmacokinetics and pharmacodynamics in Asians throughphenotyping and genotyping strategies. Journal of Clinical Oncology, 20, 3683–3690. [DOI] [PubMed] [Google Scholar]
  • 101.Isla D, Sarries C, Rosell R, Alonso G, Domine M, Taron M, et al. (2004). Single nucleotide polymorphisms and outcome in docetaxel-cisplatin-treated advanced non-small-cell lung cancer. Annals of Oncology, 15, 1194–1203. [DOI] [PubMed] [Google Scholar]
  • 102.Puisset F, Chatelut E, Dalene F, Busi F, Cresteil T, Azema J, et al. (2004). Dexamethasone as a probe for docetaxel clearance. Cancer Chemotherapy and Pharmacology, 54, 265–272. [DOI] [PubMed] [Google Scholar]
  • 103.Wils P, Phung-Ba V, Warnery A, Lechardeur D, Raeissi S, Hidalgo IJ, et al. (1994). Polarized transport of docetaxel and vinblastine mediated by P-glycoprotein in human intestinal epithelial cell monolayers. Biochemical Pharmacology, 48, 1528–1530. [DOI] [PubMed] [Google Scholar]
  • 104.Sparreboom A, Loos WJ, Burger H, Sissung TM, Verweij J, Figg WD, et al. (2005). Effect of ABCG2 genotype on the oral bioavailability of topotecan. Cancer Biology & Therapy, 4, 650–658. [DOI] [PubMed] [Google Scholar]
  • 105.Gardner ER, Burger H, van Schaik RH, van Oosterom AT, de Bruijn EA, Guetens G, et al. (2006). Association of enzyme and transporter genotypes with the pharmacokinetics of imatinib. Clinical Pharmacology and Therapeutics, 80, 192–201. [DOI] [PubMed] [Google Scholar]
  • 106.Mwinyi J, Johne A, Bauer S, Roots I, & Gerloff T (2004). Evidence for inverse effects of OATP-C (SLC21A6) 5 and lb haplotypes on pravastatin kinetics. Clinical Pharmacology and Therapeutics, 75, 415–421. [DOI] [PubMed] [Google Scholar]

RESOURCES