Skip to main content
Chinese Journal of Cancer logoLink to Chinese Journal of Cancer
. 2012 Feb;31(2):73–99. doi: 10.5732/cjc.011.10320

Breast cancer resistance protein (BCRP/ABCG2): its role in multidrug resistance and regulation of its gene expression

Takeo Nakanishi 1, Douglas D Ross 2
PMCID: PMC3777471  PMID: 22098950

Abstract

Breast cancer resistance protein (BCRP)/ATP-binding cassette subfamily G member 2 (ABCG2) is an ATP-binding cassette (ABC) transporter identified as a molecular cause of multidrug resistance (MDR) in diverse cancer cells. BCRP physiologically functions as a part of a self-defense mechanism for the organism; it enhances elimination of toxic xenobiotic substances and harmful agents in the gut and biliary tract, as well as through the blood-brain, placental, and possibly blood-testis barriers. BCRP recognizes and transports numerous anticancer drugs including conventional chemotherapeutic and targeted small therapeutic molecules relatively new in clinical use. Thus, BCRP expression in cancer cells directly causes MDR by active efflux of anticancer drugs. Because BCRP is also known to be a stem cell marker, its expression in cancer cells could be a manifestation of metabolic and signaling pathways that confer multiple mechanisms of drug resistance, self-renewal (sternness), and invasiveness (aggressiveness), and thereby impart a poor prognosis. Therefore, blocking BCRP-mediated active efflux may provide a therapeutic benefit for cancers. Delineating the precise molecular mechanisms for BCRP gene expression may lead to identification of a novel molecular target to modulate BCRP-mediated MDR. Current evidence suggests that BCRP gene transcription is regulated by a number of trans-acting elements including hypoxia inducible factor 1α, estrogen receptor, and peroxisome proliferator-activated receptor. Furthermore, alternative promoter usage, demethylation of the BCRP promoter, and histone modification are likely associated with drug-induced BCRP overexpression in cancer cells. Finally, PI3K/AKT signaling may play a critical role in modulating BCRP function under a variety of conditions. These biological events seem involved in a complicated manner. Untangling the events would be an essential first step to developing a method to modulate BCRP function to aid patients with cancer. This review will present a synopsis of the impact of BCRP-mediated MDR in cancer cells, and the molecular mechanisms of acquired MDR currently postulated in a variety of human cancers.

Keywords: BCRP, ABCG2, multidrug resistance (MDR), transporter, gene expression, tyrosine kinase inhibitors, cancer stem cells


Multidrug resistance (MDR) is a phenomenon in which cancer cells simultaneously become resistant to structurally unrelated chemotherapeutic agents when exposed to a single chemotherapeutic drug. The development of MDR in the course of chemotherapy has been considered as a major obstacle in cancer treatment. For the last three decades, the biological causes underlying MDR have been extensively studied and attributed to diverse molecular mechanisms. Active efflux mediated by drug efflux pumps has been described in a wide variety of cancer cells since MDR1, which encodes the membrane transport protein P-glycoprotein (P-gp), was isolated from KB cells selected with vinblastine in 1986[1]. P-gp was the first human ATP-binding cassette (ABC) transporter protein to be identified and is classified as the first member of the B subfamily in the ABC transporter superfamily according to the Human Genome Organization. In 1992, a second MDR-causing transporter named multidrug resistance-associated protein 1 (MRP1) was reported[2] and found to cause resistance to xenobiotics and anticancer agents. MRP1 belongs to ABC subfamily C, which consists of 13 members. Although P-gp and MRP1 could impart MDR to cancer cells, neither P-gp nor MRP1 accounted for all of the transport-based drug resistance observed in blast cells from patients with acute leukemia when measured by functional efflux assays[3]. Multidrug-resistant MCF- 7/AdrVp cells lacking expression of P-gp or MRP1 were described following selection of MCF-7 human breast cancer cells with doxorubicin and verapamil[4]. In 1998, Doyle et al. [5] isolated a novel ABC transporter from MCF-7/AdrVp cells and showed by transfection of MCF-7 cells that forced overexpression of the transporter reproduced the MDR phenotype of MCF-7/AdrVp cells. The new transporter was named breast cancer resistance protein (BCRP) because it was isolated from multidrug-resistant human breast cancer cells [5]. Two other groups reported finding a similar cDNA at approximately the same time, which were designated as ABCP [6] or MXR [7], respectively. Some review articles describing the function of BCRP in health and disease are available [8][10].

This review discusses recent progress in understanding BCRP as an MDR transporter and focuses on BCRP substrate drugs, particularly novel small molecules developed for targeted therapies, which have impact on cancer treatment. The review also focuses on the molecular regulation of BCRP gene expression and summarizes recently proposed mechanisms underlying BCRP overexpression in MDR cancer cells and cancer stem cells.

Functional Configuration of BCRP

According to the Human Gene Nomenclature Committee, BCRP is classified as the second member of the G subfamily of the ABC transporter superfamily (ABCG2). ABC transporters are distinguished by the use of ATP hydrolysis for transporter function and exhibit extensive conservation of the ATP-binding domains throughout evolution across a large number of functionally diverse transmembrane proteins[11]. The typical ABC transporter consists of two highly conserved ATP-binding domains and two transmembrane domains. A smaller group of ABC transporters, including BCRP/ABCG2, are termed half-transporters. BCRP consists of 655 amino acids and possesses six transmembrane helices and one ATP-binding site (Figure 1). Because BCRP is a half-transporter, current evidence suggests that homodimerization or multimerization is required for transporter activity as illustrated in Figure 1. Our laboratory studied the effect of co-expression of wild-type and dominant-negative BCRP on BCRP-mediated transport in Xenopus oocytes[12]. We observed that BCRP-mediated transport of daunorubicin was significantly reduced in a manner dependent on the amount of dominant-negative mutant (S187T) cRNA injected into the oocytes, strongly suggesting that it is essential for BCRP to at least homodimerize to function. Similar observations were made in cultured cells transduced with wild-type and mutant forms of BCRP[13]. Further biochemical analysis using gel-filtration chromatography suggests that BCRP exists as a homotetramer that may act only to regulate the level of functional homodimerized BCRP transporters[14]. Although disulfide bond formation (particularly at cysteine 603) has been postulated to participate in dimer/multimer formation[15],[16], in vivo studies in intact cells using fluorescence resonance energy transfer techniques recently showed that cysteine 603 is not essential for dimer/oligomer formation[17]. These findings provide a basis for structural and mechanistic analysis of BCRP and related ABC transporters.

Figure 1. Summary of BCRP function, tissue distribution, and mechanism of overexpression in drug-resistant cancer cells.

Figure 1.

BCRP consists of 6 transmembrane helices and homodimerizes to function at the plasma membranes. It pumps natural substrates, including folate, steroid hormones, and urate; toxic xenobiotics; and anticancer agents, including conventional chemotherapeutics and tyrosine kinase inhibitors. NBD, nucleotide-binding domain to which ATP can bind.

Furthermore, to date, mutant forms of BCRP in which amino acid arginine at codon 482 is substituted with threonine or glycine have been reported in various cancer cells when cells were selected with a BCRP substrate chemotherapeutic drug such as doxorubicin[18]. To the best of our knowledge, expression of these mutants has not been reported in clinical specimens[19][21]. Because these mutations alter BCRP substrate specificity, interactions between chemotherapeutic agents and wild-type as well as mutant BCRPs have been extensively studied. These studies are summarized in the “Role of BCRP in MDR” section of this review.

Physiological Function of BCRP

As an efflux transporter for xenobiotics and unwanted toxic compounds, BCRP has been characterized as an important part of self-defense systems in organisms. BCRP substrates are listed in Table 1. This is particularly true at polarized cells in normal tissues, such as placental syncytiotrophoblasts, hepatocytes, and intestinal mucosal cells, where apically expressed BCRP protects organisms by eliminating substances to the maternal circulation, bile ducts, or intestinal lumen, respectively[8]. In brain microvasculature, BCRP is located on the luminal surface of microvessel endothelium[22] and hence, may constitute an important component of the blood-brain barrier. The tissue distribution pattern of BCRP expression reflects its major role in protecting cells from potentially toxic xenobiotics and in assisting the clearance of xenobiotics from the organisms.

Table 1. Drug resistance profile of K562-imatinib cells.

Substrate Transport
Inhibition
Km (µmol/L) Assay systema Ref IC50 (µmol/L) Ki (µmol/L) Substrate used Assay systema Ref
Natural substrates/dyes
 2-Amino-1-methyl-6-phenylimidazo[4,5-b]pyridine (PhIP) MDCK/BCRP1 (TP) [23]
 2-Amino-3-methylimidazo [4,5-f]quinolone (IQ) MDCK/BCRP (TP) [24]
 Benzo[a]pyrene (BP) 3-sulfate HCT116 NRI (TP) [25]
 Aflatoxin B1 MDCK/BCRP (TP) [24]
 Protoporphyrin IX Bcrp1−/− BM, AML/BCRP (AC) [26]
 Hematoporphyrin 17.8 Sf9 vesicle (AC) [27]
 Folic acidb HEK/BCRP vesicle (AC) [31]
 Estrone 3-sulfate 6.8 K562/BCRP vesicle (AC) [33]
16.6 P388/BCRP vesicle (AC) [32]
 Estradiol-17β-D-glucuronide 44.2 HEK/BCRP vesicle (AC) [31]
 Dehydroepiandrosterone sulfate (DHEAS) P388/BCRP vesicle (AC) [32] 55 E3S P388/BCRP vesicle (AC) [32]
 17β-Estradiol L. lactis/BCRP (AC) [34] H33342 L. lactis/BCRP vesicle (AC) [34]
4.8 TPT K562/BCRP (AC) [36]
 Dihydrotestosterone Mx-RPE (EF) [35]
 Uric acid X. Oocyte [38]
Fluorescent substrates/probes
 Prazosin-BODIPY Selected Cancer Cells [119]
HEK/BCRP (AC) [86]
 Hoechst 33342 A5449/BCRP (SP analysis) [40]
 LysoTracker®b HEK/BCRP (AC) [86]
 Rhodamine 123b HEK/BCRP (AC) [86]
X. Oocyte (AC) [12]
 Pheophorbide a HEK/BCRP, Selected Cancer Cells (AC) [29]
 Sulfasalazine 0.7 SB-BCRP-M-VT (AC) [122]
 D-luciferin HEK/BCRP (AC) [125]
Chemotherapeutlcs
 Mitoxantrone (MX) 7.0 Sf9 vesicle (ATP) [85] 61 E3S P388/BCRP vesicle (AC) [32]
X. Oocyte (EF, AC) [12]
 SN-38 4.0 PC-6/SN2-5H vesicles (AC) [94] 1.6 E3S HEK/BCRP vesicle (AC) [117]
 SN-38-glucuronide 26.0 PC-6/SN2 -5H vesicles (AC) [94]
 9-Aminocamptothecin HEK/BCRP [91]
 Indolocarbazole topoisomerase I inhibitors (eg. NB-506, J-107088) PC-13/BCRP (AC, EF) [97]
 Belotecan >500 MDCKII/BCRP (TP) [95]
 Doxorubicin (DOX)b 5.0 Sf9 vesicles (ATP) [85]
 Daunorubicinb 2.5 Sf9 vesicles (ATP) [85]
HEK/BCRP (EF) [86]
X. Oocyte (AC, EF) [12]
 Epirubicinb [86]
 Bisantreneb MCF-7/AdrVp (AC) [119]
 Methotrexatec 681 MCF-7/MX vesicle (AC) [30]
1340 HEK/BCRP vesicle (AC) [31]
1410 HEK/BCRP vesicle (AC) [131]
 Imatinib 0.15 HiFive vesicles (ATP) [130] 0.9 H33342 [112]
HEK/BCRP (AC) [76]
 Nilotinib 0.008 HiFive vesicles (ATP) [130] 0.69 MTX HEK/BCRP vesicle (AC) [131]
K562/BCRP [132] 0.05 H33342 K562/BCRP (AC) [132]
 Dasatinib Mef/BCRP (AC) [136] >2.0 H33342 K562/BCRP (AC) [132]
 CI1033 (Canertinib) MDA-MB231/BCRP (AC) [139]
 Gefitinib (Iressa, ZD1839) HEK293/BCRP R-5 (AC) [144] 0.4 H33342 [112]
 Erlotinib hydrochloride (Tarceva, OSI-774, CP-358774) HEK293/BCRP R-5 (AC) [144] 0.13 E3S K562/BCRP vesicle (AC) [107]
 Lapatinib ditosylate (Tykerb, GW572016) Bcrp1−/− mice (BD) [152]
 Sunitinib malate (Sutent, SU11248) 0.18 HiFive vesicles (ATP) [157]
MDCK/BCRP (TP) [159]
MDCK/P-gp&BCRP (TP) [158]
 Sorafenib (Nexavar) MDCK/BCRP (TP) [161]
 Axitinib (AG013736) MCDK/BCRP (TP) [164]
 Flavopiridol (Alvocidib) X. Oocytes (AC) [12] MX X. Oocyte (AC) [12]

aExperimental condition used to evaluate BCRP-mediated transport of each compound: BCRP, with forced expression of BCRP; TP, transport assay across monolayer of cells; AC, intracellular accumulation study; ATP, ATPase activity assay; EF, efflux study (retention assay); BD, brain distribution in vivo in animals. bReported substrates for mutant BCRP (R482T and R482G). cReported substrates for wild-type BCRP (R482). IC50, 50% inhibition concentration; E3S, estrone 3-sulfate; TPT, topotecan; MTX, methotrexate; H33342, Hoechst 33342 dye; DOX, doxorubicin; MX, mitoxantrone; BM, bone marrow.

Naturally occurring toxic xenobiotics against which BCRP may play a protective role include dietary mutagens and carcinogens such as heterocyclic amines and polycyclic aromatic hydrocarbons. Area under the plasma concentration-time curve (AUC) was observed to be higher in Bcrp1−/− mice compared to wild-type mice following oral or intravenous administration of 2-amino-1-methyl-6-phenylimidazo[4,5-b]pyridine[23] and 2-amino-3-methylimidazo[4,5-f]quinolone[24], which are heterocyclic amines abundantly contained in cooked meat and fish. Polycyclic aromatic hydrocarbons such as benzo[α] pyrene, which endogenously exists as benzo [α]pyrene sulfate[25], and micotoxins including aflatoxin B1[24] are substrates of BCRP. BCRP is also involved in the elimination of photosensitive protoporphyrin IX [26] and hematoporphyrin[27]. Thus, reduced BCRP function may increase the risk for developing protoporphyria and diet-dependent phototoxicity[26],[28]. The photosensitizer pheophorbide a, a breakdown product of chlorophyll found in mouse chow, was also reported to be a BCRP-specific substrate[29]. In addition to these substances, BCRP transports folic acid and its polyglutamate conjugates and may play a role in cellular folate homeostasis[30],[31].

BCRP transports conjugates of steroid hormones, such as estrone 3-sulfate (E3S), dehydroepiandrosterone sulfate, and, to lesser extent, estradiol-17β-D-glucuronide[32],[33]. Although there is less evidence, 17β-estradiol (E2)[34] and dihydrotestosterone [35] have been indicated as substrates of BCRP. E2 was also shown to effectively inhibit BCRP-mediated transport[34],[36]. These findings suggest a role for BCRP in hormone metabolism and regulation. More recently, Dehghan et al.[37] identified an association of the non-synonymous coding SNP (Q141K) in exon 5 of the BCRP gene with a risk of gout based on a genome-wide association study. This led to the discovery that BCRP plays a role in secretory elimination of uric acid in the proximal tubular cells[38]. Moreover, to date, numerous natural or physiological substrates for BCRP have been identified and have been well reviewed elsewhere[8],[39].

BCRP is known to be a marker for pluripotent hematopoietic and tissue stem cells. A characteristic finding on flow cytometric analysis of cells stained with Hoechst 33342 dye is that a side population (SP) of cells possesses low Hoechst 33342 dye accumulation[40], which has been shown to be enriched for stem and progenitor cells [41]. Because the SP was absent in bone marrow from Bcrp1−/− mice[42], BCRP has been recognized as a critical determinant of the SP phenotype in not only hematopoietic cells, but also in various normal tissues, including the liver[43], lung[44], heart[45], mammary glands[46], skeletal muscle[47], neurons[48], and corneal stroma[49], and cancer cells[50][57].

BCRP Expression in Human Cancers

Acute myelogenous leukemia (AML)

Acute myelogenous leukemia (AML) is a malignant neoplastic disease in which BCRP expression has been the best characterized. Ross et al.[58] initially detected BCRP mRNA at relatively high levels in approximately one third of 20 AML samples, suggesting that BCRP plays a role in resistance to conventional chemotherapies. In the same set of 20 samples, BCRP expression seemed to be correlated with MDR1 expression (r = 0.66) although the relationship was not statistically significant. Subsequently, we showed that BCRP mRNA expression correlated with in vitro resistance to the cyclin-dependent kinase (CDK) inhibitor flavopiridol in blast cells from patients with AML[19]. In a study with paired AML blast cell samples collected before treatment and at the time of relapse or refractoriness, van den Heuvel-Eibrink et al.[59] found that BCRP mRNA was higher in the relapsed or refractory samples. Benderra et al. [60] found that pre-treatment high BCRP mRNA level was a prognostic factor for achieving a complete remission (CR) and associated with lower disease-free survival in 149 patients. Furthermore, in a study of 40 patients with newly diagnosed AML, Abbott et al.[61] found that BCRP mRNA expression was higher in AML samples (78%) than in normal bone marrow; however, only 7% of patients had “functional” levels of BCRP mRNA. In that study, there was no correlation between BCRP expression and patient outcomes or clinical characteristics, but the data suggested the presence of a small subpopulation of “primitive leukemic stem cells with intrinsic drug efflux capacity” in which BCRP mRNA expression was confined. This notion was supported by the findings of van der Kolk et al.[62], who detected BCRP protein expression in subpopulations of cells with an immature phenotype (CD 34+) but found no increase in these BCRP+ subpopulati ons at time of relapse in 20 AML patients. Suvannasankha et al.[20] also found that BCRP mRNA expression correlated poorly to BCRP protein expression but observed that BCRP mRNA may be limited to small subpopulations of blast cells pretreatment.

There is compelling evidence, accumulated over the past decade, that BCRP is expressed in premature AML blast cell subpopulations.[9] The highest BCRP expression and function were found in stem-like CD34+/CD38 cells in both normal and AML marrows [63]. Although blocking BCRP function somewhat reversed drug resistance of CD34+/CD38 AML blast cells to mitoxantrone, it had little effect on the intracellular accumulation of mitoxantrone. Thus, additional factors are likely involved in the drug resistance exhibited by these cells[63]. An evaluation of the mRNA expression of 47 human ABC transporters in blast cells from 18 AML patients who achieved CR and 13 AML patients who were refractory to induction chemotherapy revealed no difference in ABC transporter expression between the CR and refractory groups; however, the refractory group had significantly higher expression of BCRP and/or MDR1 in CD34+/CD38cells[64]. Moreover, a study with 26 bone marrow samples from patients with de novo AML done by de Figueriedo-Pontes et al.[65] showed that leukemia stem cell population, defined as CD34+/CD38/CD123+, showed higher BCRP and P-gp expression than other subsets of cell populations, implying that co-expression of both BCRP and P-gp in leukemia stem cells may augment MDR synergistically. Thus, more recent studies of BCRP expression in AML likely confirm that BCRP is often co-expressed with P-gp and connotes a worse prognosis [64][68]. Furthermore, BCRP and P-gp expression appear to be associated with subpopulations of cells with primitive characteristics, such as expression of CD34 but not CD38 [63],[65],[68],[69], although two studies showed no association of BCRP with CD34 expression (but not CD38)[70],[71].

Acute lymphoblastic leukemia(ALL) and chronic myelogenous leukemia (CML)

There is limited information available for BCRP expression in acute lymphoblastic leukemia (ALL). To date, BCRP expression is suggested to be detected at the highest level in B-lineage ALL, although no definite trends were shown between BCRP expression and prognosis of ALL[72],[73]. A study of BCRP expression in childhood ALL showed that BCRP mRNA was not increased in specimens from children in relapse[73]. No relationship was found between BCRP mRNA expression and relapse-free survival in this study. Our study of BCRP function and mRNA and protein expression in blast cells from 30 adult ALL cases by Suvannasankha et al.[21] revealed a relatively high frequency of positive staining (37% to 47% of cases) with a variety of anti-BCRP antibodies; however, there was poor concordance of antibody staining, mRNA expression, and functional assays in this study. Positive staining with the BXP21 mouse anti-human monoclonal antibody to BCRP was predictive of a shorter disease-free survival.

In case of chronic myeloid leukemia (CML), previous studies have indicated a high expression of BCRP in premature CML cell populations. Jordanides et al.[74] detected aberrant overexpression of BCRP protein in CD34+ cells derived from CML patients. BCRP expression was detected in over 25% of cells from 5 out of 7 individual subjects, and BCRP function was confirmed by active efflux of mitoxantrone out of the cells. Moreover, quantitative gene expression analysis indicated a strong differentiation-associated decrease of BCRP mRNA expression in mononuclear cells from patients with CML[75]. In this study, BCRP was more highly expressed in lin/CD34+/38 cells derived from CML than in normal bone marrow cells. Although imatinib was shown to be a substrate of BCRP [76],[77], BCRP-mediated active efflux of imatinib has not yet been observed in CD34+ CML cells[74],[78]. These findings strongly suggest that BCRP expression in CML is confined in primitive cell populations, but the functionality of BCRP in primitive cells is not completely understood.

Solid tumors

BCRP has been reported to be overexpressed in solid tumors. Immunohistochemical studies with BXP-21 antibody demonstrated a high frequency of BCRP immunoreactivity among a panel of 150 untreated human solid tumors comprising 21 tumor types[79]. Specificity of immunohistochemistry was confirmed by the detection of a 72 kDa band representing BCRP on Western blots. According to this article, BCRP expression was widely present in untreated human solid tumors. Kawabata et al.[80] used reverse transcription-polymerase chain reaction (RT-PCR) and detected high levels of BCRP mRNA expression in 6 of 8 non-small cell lung cancer (NSCLC) cell lines and 5 (22%) of 23 non-small cell lung tumor tissues. In these lung cancer cell lines, topotecan efflux was correlated with the levels of BCRP mRNA. In in vitro models, BCRP overexpression is commonly observed in cells derived from various types of human solid tumors selected with drugs, such as mitoxantrone[5],[7],[81],[82], topoisomerase inhibitor topotecan[82], flavopiridol[83], and imatinib[84], and seems to be independent of P-gp or MRP expression.

Role of BCRP in MDR

Conventional chemotherapies known as BCRP substrates

BCRP functions as an efflux pump for a wide variety of xenobiotics inc luding many approved cancer chemotherapies (Table 1). Therefore, emergence of BCRP renders cancer cells resistant to structurally unrelated drugs simultaneously, fulfilling the definition of MDR. BCRP is also known as mitoxantrone resistance protein (MXR) because it was isolated from human colon carcinoma cells (S1M1 80) after in vitro selection with mitoxantrone[7]. Mitoxantrone is shown to be a substrate of BCRP in in vitro assays[12],[85]. Strong cellular resistance to mitoxantrone is attributable to BCRP-mediated efflux, which reduces its intracellular accumulation[5],[7],[12],[86]. Following selection with mitoxantrone, BCRP overexpression was observed not only in colon carcinoma cell lines (S1M1 80, HT29RNOV, KM12/MX) but also in a variety of cancer cell lines including breast carcinoma (MCF-7/MX8, MCF-7/Mitox, MDA-MB-231RNOV), gastric carcinoma (EPG85-257RNOV), fibrosarcoma (EPF86-079RNOV), NSCLC (H460/MX), glioblastoma (SF295/MX), and myeloma (8226/MR20)[81],[87],[88]. Thus, BCRP is a component of the cellular defense mechanism elicited in response to mitoxantrone.

Accumulated evidence on MDR shows that BCRP plays a critical role in the development of resistance to irinotecan-based therapy. Topotecan is a type I topoisomerase inhibitor derived from camptothecin. Selection with topotecan in a human ovarian cancer cell line (IGROV1/T8) [82] and a breast cancer cell line (MCF-7/TPT3000)[89] was found to induce overexpression of BCRP, resulting in acquired resistance to 10-hydroxy-7-ethylcamptothecin (SN-38), the active metabolite of irinotecan; 9-aminocamptothecin; and, to a lesser extent, irinotecan[82],[89],[90]. Overexpression of both the wild-type and R482T forms of BCRP conferred resistance to 9-aminocamptothecin, which was associated with reduced intracellular drug accumulation, but did not confer resistance to 9-nitrocamptothecin. These observations, in addition to the knowledge that BCRP confers resistance to topotecan and SN-38, suggest that polar residues at the 9 or 10 position of camptothecin may be important to facilitate interactions with BCRP[91]. BCRP was also induced in cancer cells by selection with SN-38[92],[93], and further transport study showed that both SN-38 and SN-38-glucuronide are substrates of BCRP[94]. BCRP mRNA levels were higher in hepatic metastases obtained from patients after irinotecan-based chemotherapy than in irinotecan-naive metastases, indicating that BCRP was involved in irinotecan resistance in vivo[93]. Belotecan, a semi-synthetic topoisomerase I inhibitor, was shown to be transported by BCRP[95]. More recently, novel effective camptothecin analogues of the 7-oxyiminomethyl compound, ST1968 and ST1978, were identified as substrates of BCRP[96].

Although BCRP expression was enhanced in MCF-7 cells after being selected with either doxorubicin (MCF-7/AdrVp cells) or mitoxantrone (MCF-7/MX8 cells), a high level of efflux was noted with rhodamine 123 in MCF-7/AdrVp cells, whereas such efflux was not seen in MCF-7/MX8 cells[87]. This was later explained by the presence of a mutation at codon 482 in the mRNA sequence of BCRP, with replacement of the wild-type arginine by threonine or glycine (R482T or R482G)[18],[97]. Compared to wild-type BCRP, overexpression of either the R482T or R482G BCRP mutants conferred greater resistance to anthracyclines, including doxorubicin, daunorubicin and epirubicin, and bisantrene[5],[86],[98]. Because intracellular accumulation of these substrate drugs was significantly lowered in cells expressing the mutant protein[12],[86], mutation at codon 482 has a major impact on substrate specificity of BCRP. Similar findings were made for codon 482 of murine Bcrp1/Abcg2, except that the mutations were to methionine or serine[99]. Active efflux of hydrophilic antifolates is also affected by the mutations. Wild-type BCRP is capable of transporting methotrexate [30],[31],[100] and its polyglutamate form [30],[31], whereas the mutants likely lack the ability to efflux methotrexate. However, one report described that human embryonic kidney cells stably transfected with R482G-BCRP cDNA displayed resistance to hydrophilic antifolates, including methotrexate, raltitrexed, and GW1843, relative to parental cells[101]. This discrepancy needs to be addressed further. In addition, hydrophobic antifolates may be substrates of both mutants, but not wild-type (R482) BCRP. These mutants have been shown to confer high-level of resistance to lipophilic antifolate inhibitors of dihydrofolate reductase (e.g., trimetrexate and piritrexim) and thymidylate synthase (e.g., AG337 and AG377)[102]. Indeed, the lipophilic antifolate analogue chromophore tetramethylrosamine has been identified as a substrate of both BCRP R482T and R482G but not wild-type BCRP[102]. Codon 482 mutation to methionine in human BCRP has also been reported[103]. To date, however, these mutations have only been found in cultured drug-selected cell lines. Similarly, no BCRP mutations have been observed in clinical samples of blast cells obtained from leukemia patients[19],[20],[72].

Compared to control transfectants, BCRP-transfected cells have been shown to display only minimal resistance to etoposide and depsipeptide[86] but no resistance to cisplatin, paclitaxel, or vincristine[5]. Recently, the prostate cancer cell line LNCaP was shown to display overexpression of phosphorylated BCRP after in vitro selection in the presence of docetaxel, resulting in enhanced decetaxel resistance. This suggests that post-translational modification (PTM) in response to drug may be involved in substrate specificity or functionality of BCRP[104].

Inhibitors for modulation of BCRP function

Because BCRP plays a role in MDR, specifically modulating BCRP function to sensitize BCRP-overexpressed cancer cells to chemotherapies is of pharmacological interest. To date, several specific inhibitors of BCRP have been reported, and some are currently undergoing clinical trials or are available to treat patients. Among them, fumitremorgin C (FTC) is one of the most specific inhibitors for BCRP and has been frequently used to test BCRP activity in a wide variety of experiments. FTC, a mycotoxin isolated from Aspergillus fumigates, was the first reported specific inhibitor for BCRP[105],[106] and is active at micromolar concentrations [31],[87],[107]. Ko143 is a derivative of FTC and approximately 10 times more potent at BCRP inhibition than FTC[108],[109]. GF120918 (also known as Elacridar) was first developed as a P-gp inhibitor and has been shown to block BCRP function. Indeed, GF120918 inhibited BCRP-mediated mitoxantrone transport with an IC50 value of 0.31 µmol/L[110]. A pharmacokinetic study demonstrates the effectiveness of GF120918 in vivo, showing co-administration of GF120918 can increase the oral bioavailability of topotecan, a BCRP substrate drug, from 30% to 90% [111]. Other BCRP inhibitors include gefitinib (Iressa, ZD1839)[112], flavopiridol[12], 6-prenylchrysin (a hydrophobic flavone) [110], reserpine [26],[113], taxane derivatives (i.e., ortataxel and tRA96023)[114], estrone and antiestrogens (e.g., tamoxifen and TAG-related compounds) [115], imatinib [112],[116], certain HIV protease inhibitors[109], and 3-hydroxymethylglutaryl coenzyme A reductase inhibitors[117]. Recently, synthetic compounds related to piperazinopyranones and phenalkylaminobenzopyranones were identified to be strong inhibitors of BCRP comparable to FTC[118]. BCRP inhibitors are summarized in Table 2.

Table 2. Inhibitors for BCRP.

Inhibitor IC50 (µmol/L) Ki (µmol/L) Substrate used Assay systema Ref
FTC 1.0 MX S1M1 80 vesicle (ATP) [87]
0.30 MTX HEK/BCRP vesicle (AC) [31]
0.25 0.55 E3S K562/BCRP vesicle (AC) [107]
0.47 PhA MDCK/BCRP (TP) [109]
Ko143 0.01 PhA MDCK/BCRP [109]
Elacridar (GF120918) 0.31 MX Sf9 vesicles [110]
6-Prenylchrysin 0.29 MX Sf9 vesicles [110]
Reserpine <10 H33342 Saos2/BCRP (AC) [113]
Ortataxel <10 MX 8226/MR20 [114]
tRA96023 <10 MX 8226/MR20 [114]
Tamoxifen and antiestrogens (TAG compounds) TPT K562/BCRP (AC) [115]
Phenalkylamine derivatives (5b, 5c) ∼ 1 MX HCT116/R (AC) [118]
Piperazine derivatives (4c) ∼10 MX HCT116/R (AC) [118]
Daunomycin 59 E3S P388/BCRP vesicle (AC9 [32]
Anti-HIV therapeutics
 Amprenavir 181 PhA MDCK/BCRP (TP) [109]
 Atazanavir 69.1 PhA MDCK/BCRP (TP) [109]
 Lopinavir 7.66 PhA MDCK/BCRP (TP) [109]
 Nelfinavir 13.5 PhA MDCK/BCRP (TP) [109]
 Saquinavir 27.4 PhA MDCK/BCRP (TP) [109]
 Delavirdine 18.7 PhA MDCK/BCRP (TP) [109]
 Efavirenz 20.6 PhA MDCK/BCRP (TP) [109]
HMG-CoA reductase inhibitors
 Atorvastatin 14.3 E3S HEK/BCRP vesicle (AC) [117]
 Cerivastatin 18.1 E3S HEK/BCRP vesicle (AC) [117]
 Fluvastatin 5.43 E3S HEK/BCRP vesicle (AC) [117]
 Pitavastatin 2.92 E3S HEK/BCRP vesicle (AC) [117]
 Rosuvastatin 15.4 E3S HEK/BCRP vesicle (AC) [117]
 Simvastatin acid 18.0 E3S HEK/BCRP vesicle (AC) [117]

Footnotes as in Table 1.

Specific probe substrates to characterize BCRP function

To date, several probe substrates for BCRP have been reported. These agents are useful for evaluating the effect of drugs on BCRP-mediated transport. The fluorescence-labeled compound, BODIPY-prazosin, is one of the most frequently used probe substrates for BCRP [87],[119]. LysoTracker Green® and rhodamine 123 were described as good substrates for both codon 482 mutants of BCRP but not for wild-type BCRP[12],[87]. Mitoxantrone is often used as a fluorescent substrate in flow cytometry-based functional assays[7],[86], but it is also a weak substrate for P-gp. Robey et al.[29] demonstrated that pheophorbide a is a fluorescent substrate specific to wild-type and mutant BCRP, and that using a flow cytometric assay with this compound is convenient to determine expression of functional BCRP. Earlier pharmacokinetic studies showed that the transport of sulfasalazine across a monolayer of Caco-2 human colon carcinoma cells from the basolateral to apical side was significantly reduced by FTC but not by cyclosporine A, a typical inhibitor for P-gp, indicating that sulfasalazine is a substrate of BCRP[120]. In Bcrp1−/− mice, the AUC of sulfasalazine was over 100-fold greater than that in wild-type mice[121]. Affinity of sulfasalazine to BCRP was estimated at approximately 70 µmol/L by in vitro membrane vesicle study[122]. Furthermore, clinical studies revealed that measured plasma concentrations in vivo correlate with single nucleotide polymorphisms in the BCRP gene (c.421C>A)[123]. Selection of T-lymphocytes by sulfasalazine caused BCRP overexpression[124]. These findings suggest that sulfasalazine can be used as a probe substrate to evaluate BCRP function. More recently, intracellular accumulation of D-luciferin was shown to be modulated by FTC [125]. D-luciferin bioluminescent output in vivo was substantially influenced by BCRP function within the regions of interest. Furthermore, FTC had no effect on the bioluminescent output in cells overexpressing human P-gp, MRP1, or MRP2 genes; therefore D-luciferin-based bioluminescent imaging has been suggested as a new high-throughput method for identifying modulators of BCRP function[126].

BCRP and Molecular Targeted Therapy

Tyrosine kinase inhibitors (TKIs) for BCR-ABL

Previously, Özvegy-Laczka et al. [112] described a clear interaction between tyrosine kinase inhibitors, including imatinib mesylate (originally known as Gleevec, STI571) and BCRP, which modulate BCRP transport and stimulate BCRP-specific ATPase activity. There are conflicting reports on whether imatinib is a substrate transported by BCRP. Imatinib is shown to be a substrate for BCRP in human embryonic kidney HEK293 cells transfected with BCRP[76], whereas it is not in Soas2 cells with forced expression of BCRP[116] or primitive CML CD34+ cells aberrantly overexpressing BCRP[74]. The bioavailability, pharmacokinetics, and disposition of imatinib are shown to be influenced by BCRP, suggesting that BCRP functions as an imatinib transporter in vivo[127]. Subsequently, chronic exposure of Caco-2 human colon carcinoma cells to imatinib is found to result in induction of BCRP expression[84], representing a novel potential mechanism of acquired drug resistance in cancer patients treated with imatinib over a long time. Although BCRP did not confer a survival advantage to imatinib-treated Saos2 cells compared to mock-transfected ones[116], this could be because Saos2 are not growth- or survival-dependent upon the intracellular targets of imatinib, including BCR-ABL, c-KIT, or platelet-derived growth factor receptor (PDGFR). We examined the effects of BCRP expression on cellular resistance to imatinib in human CML-derived K562 cells, whose growth is dependent on BCR-ABL, and demonstrated that forced expression of BCRP in these cells caused resistance to imatinib cytotoxicity that was overcome by the BCRP-specific inhibitor FTC[128]. Interestingly, this effect was attenuated by imatinib-mediated inhibition of BCR-ABL, which in turn down-regulated BCRP expression post- transcriptionally via the PI3K/AKT pathway. This imatinib-mediated down-regulation of BCRP was further confirmed by Dohse et al.[129]; however, detailed mechanisms should be further examined.

Nilotinib (Tasigna, AMN107) is a second-generation BCR-ABL TKI used to surmount resistance or intolerance to imatinib in patients with Ph (+) CML. Photoaffinity labeling assays[130] and kinetic analyses[131] have shown that nilotinib competitively inhibits BCRP-mediated drug efflux. The Ki value of nilotinib for BCRP-mediated methotrexate transport was 0.69 µmol/L, suggesting that nilotinib is a highly potent inhibitor of BCRP[131]. Nilotinib was also shown to be a high-affinity substrate of BCRP and P-gp[132]. Because BCRP is highly expressed on CML CD34+ cells[74], several studies have been conducted to determine whether the inhibition of BCRP activity potentiates the effect of nilotinib on apoptosis, BCR-ABL inhibition, or CML CD34+ cell proliferation. So far, it is unlikely that BCRP and P-gp will have any effect on the clinical response to this drug[133]. BODIPY-labeled nilotinib was also found to be a substrate for P-gp and BCRP and has been proposed as a potentially useful probe for functional analysis of both transporters in cancer cells and other preclinical studies[134].

Dasatinib (Sprycel, BMS-354825) is another second-generation dual kinase inhibitor of ABL and SRC. In in vitro assays, dasatinib is approximately 300 times more active than imatinib and effective against most imatinib-resistant BCR-ABL active-site mutants, except the T315I mutation[135]. Ko143 significantly increased dasatinib uptake in BCRP-overexpressing cell lines and reduced the IC50 of dasatinib[136]. In vivo brain distribution studies showed that brain-to-plasma concentration ratio of dasatinib was significantly higher in Mdr1a/b−/−Bcrp1−/− than in Mdr1a/b−/− mice and increased in wild-type mice with co-administration with Ko143[137]. These findings are a clear demonstration that dasatinib is a potent substrate for BCRP.

Bosutinib (SKI-606), a third-generation TKI inhibitor, is useful in patients whose leukemia is resistant to both first- and second-generation TKI inhibitors. Unlike imatinib, bosutinib inhibits both SRC and ABL[138]. Current evidence regarding the interaction of bosutinib with ABC transporters is limited. In in vitro growth assays, neither P-gp nor BCRP induced resistance to bosutinib, although their transport activities were inhibited by bosutinib at relatively high concentrations (> 2 µmol/L)[132].

TKIs for epidermal growth factor receptor (EGFR, HER1) and HER2

CI1033 (also named canertinib) was the first EGFR TKI that was shown to be a substrate of BCRP. Intracellular accumulation of CI1033 was diminished in BCRP-expressing T98G glioblastoma and HCT8 colon cancer cells[139]. This was further confirmed by a trans-cellular transport study using MDCK/BCRP cells[140].

Gefitinib (Iressa, ZD1839) is an orally active, small-molecule inhibitor of EGFR and a first-generation TKI. Gefitinib, a strong inhibitor of BCRP, modulated ATPase activity of BCRP in a plasma membrane vesicle study[112] and reversed BCRP-mediated resistance to SN38[141],[142] and topotecan[143]. Gefitinib accumulation was lowered in HEK293 cells transfected with BCRP, suggesting that gefitinib is a substrate of BCRP[144]. Another study showed that BCRP conferred resistance to gefitinib[141],[145] and that chronic exposure of colorectal carcinoma cells to gefitinib induced BCRP expression[146]. Hence, BCRP has been postulated to play a critical role in cellular resistance to and pharmacokinetic disposition of gefitinib. The systemic exposure of gefitinib was greater in patients heterozygous at the BCRP c.421C>A locus than in patients with wild-type BCRP[144]. In an animal pharmacokinetic study, brain accumulation of gefitinib increased approximately 70-fold in Mdr1a/b−/−Bcrp1−/− mice compared to wild-type mice, suggesting that the distribution of gefitinib to the brain is highly restricted by both P-gp and BCRP [147]. Because concomitant administration of gefitinib with topotecan could enhance topotecan penetration to the central nervous system, co-administration of gefitinib with a BCRP substrate drug may provide a potential benefit for patients with brain tumors.

Erlotinib hydrochloride (Tarceva, OSI-774, CP-358774), another orally-active EGFR TKI, is similar to gefitinib in structure. Erlotinib reverses P-gp- and BCRP-mediated MDR in cancer cells through direct inhibition of their drug efflux function[148]. Erlotinib competitively inhibited BCRP-mediated estrone 3-sulfate (E3S) transport with an IC50 value of 0.13 µmol/L, which is comparable to FTC, suggesting that erlotinib is one of the most powerful inhibitors for BCRP among the TKIs[107]. An in vitro accumulation study in HEK293 cells transfected with BCRP showed that erlotinib is a substrate for BCRP[144]. Furthermore, in vivo systemic exposure of orally administered erlotinib was significantly increased in Mdr1a/b−/−/Bcrp1−/− mice compared with wild-type mice, and brain distribution of erlotinib was observed to be restricted by both BCRP and P-gp[149]. According to the possible clinical consequences of interactions between P-gp/BCRP-inhibiting substrates and erlotinib, erlotinib pharmacokinetics in vivo may be affected by drug-drug interaction on P-gp or BCRP, which needs to be addressed in patients.

Lapatinib ditosylate (Tykerb, GW572016) is an orally active, dual tyrosine kinase inhibitor that interrupts both EGFR and HER2 signaling and is used primarily to treat breast cancer[150]. Lapatinib sensitized MDR cells by inhibiting BCRP transport activity directly[151]. The brain-to-plasma ratio of lapatinib was over 10-fold higher in Mdr1a/b−/−/Bcrp1−/− mice compared to Mdr1a/b−/− mice, suggesting that lapatinib is a substrate for BCRP[152]. Synergistic interactions were seen with lapatinib and SN38, and this effect was attributed to increased intracellular accumulation of SN38 as a result of the inhibitory effect of lapatinib on BCRP-mediated SN38 effluxi[153].

Vandetanib (Zactima, ZD6474) is a multitargeted TKI that affects EGFR and VEGFR2, and is RET kinase. Chronic exposure of HT-29 colorectal cancer cells to vandetanib reportedly induced BCRP expression[146]; however, BCRP overexpression did not confer resistance to vandetanib in drug-resistant S1M1 80 cells[154]. Vandetanib treatment significantly increased the intracellular accumulation of BCRP substrate drugs in a dose-dependent manner[154], suggesting that vandetanib functions as an MDR sensitizer.

TKIs for Class III kinases (c-KIT and PDGFR)

Sunitinib malate (Sutent, SU11248) is a multitargeted small-molecule TKI used to treat renal cell carcinoma (RCC) and imatinib-resistant gastrointestinal stromal tumors[155]. Sunitinib was observed to inhibit BCRP transport activity and stimulate its ATPase activity in a concentration-dependent manner[156],[157]. Direct binding of sunitinib to BCRP was demonstrated by photoaffinity labeling and antibody binding assays[157]. A recent in vitro transport study showed that sunitinib is a substrate for both BCRP and P-gp[158]. An in vivo pharmacokinetic analysis revealed a 23-fold increase in brain accumulation of sunitinib in Mdr1a/b−/−/Bcrp1−/− mice, whereas there was only a 2.3-fold increase in Mdr1a/b−/−mice and no increase in Bcrp1−/− mice, suggesting that sunitinib is cooperatively restricted by both P-gp and Bcrp1 at the blood-brain barrier[159].

Sorafenib tosilate (Nexavar) is another multitargeted TKI used to treat hepatocellular carcinoma (HCC) and RCC[160]. Like sunitinib, sorafenib was shown to potently inhibit BCRP-mediated drug efflux[156]. A recent transcellular study showed that sorafenib is efficiently transported by BCRP rather than P-gp[161]. In an in vivo pharmacokinetic analysis, brain accumulation of oral sorafenib increased in Bcrp1−/− mice compared with wild-type mice, suggesting that its distribution to the brain is primarily restricted by BCRP[161][163].

TKIs for VEGFR

Axitinib (AG013736) is a small-molecule indazole derivative and newly developed effective TKI, which inhibits vascular endothelial growth factor receptor (VEGFR)-1, -2, and -3 as well as PDGFR and cKIT. A transport study using monolayers of MDCK II cells expressing BCRP showed that axitinib is a moderate substrate of BCRP[164]. The combination of axitinib and SN38 was found to exert a strong synergistic effect on antiproliferative and proapoptotic activities in endothelial and other cancer cells. This effect could be due to increased intracellular accumulation of SN38 by axitinib-mediated inhibition of BCRP and ATP7A expression, which is the result of extracellular signal-regulated kinase (ERK)1/2 deactivation and AKT phosphorylation in human pancreatic cancer cell lines[165]. However, such a synergistic interaction was not observed with SN38 and a similar TKI, pazopanib (Votrient), suggesting that an interaction of pazopanib with BCRP is unlikely[153].

Apatinib (YN968D1), another promising inhibitor of receptor tyrosine kinases including VEGFR2, is currently in phase III clinical trials in China to determine its efficacy in treating gastric cancer and NSCLC [166]. Apatinib reversed P-gp- and BCRP-mediated MDR by inhibiting their transport function, but not by suppressing the AKT or ERK1/2 pathway. Use of these TKIs with other conventional chemotherapeutics may provide a clinical benefit for patients to circumvent MDR to other conventional antitumer drugs.

Cyclin-dependent kinase (CDK) inhibitors

Flavopiridol (Alvocidib, HMR1275, also known as L86-8275) is a synthetic N-methylpiperidinyl chlorophenyl flavone and the first CDK3 inhibitor that is currently in clinical trials[167],[168]. In one study, flavopiridol induced overexpression of ABCG2 but not P-gp or MRP1 in MCF-7 cells [83]. Furthermore, resistance to flavopiridol was observed in selected BCRP-overexpressing cell lines but not in P-gp- or MRP1-overexpressing cell lines[83]. Flavopiridol was shown as a substrate for wild-type and mutant BCRP (482T) in Xenopus oocytes injected with BCRP cRNA[12]. We also found that BCRP mRNA expression correlated proportionally with cell viability in the presence of 250 nmol/L flavopiridol (r = 0.86, P = 0.003) and negatively with apoptosis induced by flavopiridol in blast cell specimens derived from patients with AML, suggesting that BCRP plays a role in leukemia cellular resistance to flavopiridol[19].

Another conventional CDK inhibitor that was described to interact with BCRP is purvalanol A[169]. An et al.[169] showed that purvalanol A can effectively block the SRC signaling and cell cycle progression. These investigators studied the effect of a series of CDK inhibitors on BCRP-mediated transport by photosensitivity assay. Among compounds tested, purvalanol A inhibited BCRP-mediated hematoporphyrin most potently with an IC50 value of 3.5 µmol/L. WHP-180 is also a CDK inhibitor that was shown to inhibit BCRP function, but the rest of the agents tested, including bohemine, seliciclib (roscovitine), and olomoucine, showed only minimal effect on BCRP-mediated transport. According to the findings, planar structure of these CDK inhibitors may be an important factor for interactions with the active site of BCRP.

Interaction of BCRP with new CDK inhibitors under development has been reported. JNJ-7706621 (3,5-diamino-1,2,4-triazole) is an effective inhibitor for both CDK and aurora kinases[170]. Selection of human cervical carcinoma HeLa cells with JNJ-7706621 caused overexpression of BCRP, resulting in MDR in these cells (HeLa-6621)[171]. In the same study, the AUC of oral JNJ-7706621 was approximately 3-fold higher in Bcrp1−/− mice compared with wild-type mice, implying that this agent is transported by BCRP.

Molecular Mechanisms for BCRP Gene Expression

Transcriptional regulation of BCRP under physiological conditions

The human BCRP gene is localized to chromosome 4(4q22), spans over 66 kb, and consists of 16 exons and 15 introns. The exon-intron organization of BCRP gene was originally described by Bailey-Dell et al.[172]. The translation start site is located in exon 2(Ex2), and the translation termination site is located in Ex16. A putative transcription start site (TSS, +1), whose nucleotide position was defined as previously described (GenBank AF15130.1), was found 529 bp upstream of the Ex1 and Ex2 junction and an active proximal promoter region was identified at nucleotides -300 to -50 relative to the TSS in cell lines of human choriocarcinoma (JAR, JEG-3, BeWo cells) and breast cancer (MCF-7)[172]. In the same study, the promoter region was characterized as TATA-less and as having a CAAT box and several Sp1 sites downstream from a putative CpG island (Figure 2). This is similar to the promoters reported for the MDR1 (ABCBI) and MRP1 (ABCC1) genes, which lack TATA boxes and have multiple Sp1 sites[173],[174].

Figure 2. Cis-regulatory elements in BCRP promoter and splice variants of BCRP transcripts.

Figure 2.

Identified cis-acting elements are shown in the promoter region of BCRP gene with identified splice variants of BCRP mRNA. A putative transcription start site (TSS, +1), defined as previously described (GenBank AF15130.1), was found 529 bp upstream of the Ex1 and 2 junction, and 18899 bp upstream of Ex2. An active proximal promoter region was identified at nucleotides -300 to -50 relative to the TSS. The same Ex2 acceptor is used for all 5′-UTR exons (E1U, E1A, E1B, E1C, E1E, and E1D). Variable TSSs are found for E1U, E1B, and E1C. Figure is not completely to scale and all nucleotide positions are shown relative to the TSS (+1). XRE sites with superscript “*” are also identified as core sequences for DRE. PPAR, peroxisome proliferator-activated receptor; XRE, xenobiotic response element; DRE, dioxin response element; XBBF, X-box binding factor; ARE, antioxidant response element; HRE, hypoxia response element; PRE, progesterone response element; MED, multiple start site element downstream; iMED, inverted MED.

To date, several cis-acting elements have been reported in the proximal promoter region of BCRP by many groups, as summarized in Figure 2. Krishnamurthy et al.[26] found that BCRP gene transcription is activated by binding of the hypoxia-inducible factor 1α (HIF1α)/ARNT heterodimer to a hypoxia response element (HRE) under low oxygen conditions. Because BCRP interacts with heme and other porphyrins, its regulation by hypoxia suggests that BCRP plays a role in protecting cells/tissue from protoporphyrin accumulation under hypoxia. Moreover, hormonal regulation of BCRP transcription has been reported during pregnancy in regard to its protective role for the fetus at the placental barrier[175][177]. Indeed, E2-mediated up-regulation of BCRP mRNA expression was shown in estrogen receptor (ER)-positive breast cancer T47D cells[178] and primary placental trophoblast cells[179]. Ee et al.[178] found that this stimulatory effect is mediated through the classical pathway that involves binding of E2/ER complexes on the estrogen response element in the BCRP promoter region. Subsequent studies observed estrogen-mediated transcriptional and post-transcriptional down-regulation of BCRP in BeWo choriocarcinoma cells[177] and in breast cancer cells[180], respectively. However, Wang et al.[175] reported that BCRP mRNA and protein expression was further increased by progesterone (P4) in combination with E2 in BeWo cells, suggesting that E2 and P4 are synergistically involved in up-regulation of BCRP expression. Although the precise mechanism is under investigation, these investigators demonstrated that P4-mediated up-regulation of BCRP could be triggered by direct binding of progesterone receptor B to a progesterone response element identified in the BCRP promoter[181]. Core sequences of that PRE were located at the same position as the ER element core sequence[178],[181],[182]. Therefore, BCRP transcription may be regulated by both ERs and PRs in a complex manner.

Other nuclear receptors are likely involved in transcriptional regulation of BCRP. BCRP is found to be transcriptionally up-regulated via activation of peroxisome proliferator-activated receptor (PPAR) γ in human myeloid dendritic cells[183]. A 150-bp long, conserved enhancer region containing three functional PPAR response elements was identified upstream of human BCRP, and binding of the PPARγ/retinoid X receptor (RXR) heterodimer to this region was confirmed. Moreover, BCRP was transcriptionally up- regulated by PRARα-specific agonists (e.g. Wy14643 or GW7674) in the intestines and liver of wild-type mice but not in PPARα-null mice. Hence, BCRP transcription has been suggested to be regulated in a PPARα-dependent manner[184].

Previously, aryl hydrocarbon receptor (AHR) agonists, such as polycyclic aromatics (e.g. BP, indolo [3,2-b]carbazole and benzo [k]fluoranthene) and phytochemicals (e.g. quercetin, chrysin, and flavone), were shown to induce BCRP expression in Caco-2 cells, indicating an involvement of AHR in BCRP transcription[25],[185]. BCRP mRNA expression was recently found to be enhanced by binding of the AHR/ARNT heterodimer in proximal and distal regions upstream of the BCRP TSS that contain AHR elements [AHREs, also known as xenobiotic response elements (XREs)][186],[187].

Nrf2, a nuclear factor-erythroid 2-related transcription factor, plays a critical role in transcriptional up-regulation of many metabolizing enzymes and transporters that rescue cells from oxidative stress and/or electrophilic stress mediated by extracellular stimuli such as xenobiotics[188]. Nrf2 was found to function as a transcription factor for BCRP gene expression in hepatocytes: silencing Nrf2 gene expression abolished induction of BCRP mRNA expression mediated by the Nrf2 inducer, tert-butylhydroquinone[189]. Detailed promoter analysis using luciferase reporter assays revealed an antioxidant response element (ARE) critical for the Nrf2-mediated expression in lung cancer cells[190]. These findings provide a rationale for up-regulation of functional BCRP by xenobiotics.

Furthermore, growth factors are also likely involved in BCRP gene transcription. Epidermal growth factor (EGF)-mediated activation of BCRP mRNA expression has been observed in ovarian[191] and breast[175] cancer cell lines. In breast cancer cells, this effect could involve ERK1/2 and c-jun N-terminal kinase, suggesting that BCRP mRNA expression may be under control of the MEK-ERK signaling pathway[175]. This notion is supported by the evidence that activation of EGF receptors (e.g. EGFR and HER2) increased the side population (SP) fraction in head and neck squamous cell carcinoma[192] and breast cancer cells[53]. However, the downstream transcription factors and trans-acting effectors have not been identified for EGF-mediated signaling pathways. Further studies are warranted to understand the mechanism underlying BCRP up-regulation by growth factors.

Alternative promoter utilization of BCRP

In general, alternative splice variants of 5′-untranslated leader exons are associated with alternative promoter usage. Such usage can result in diversity of gene expression in a variety of ways, including tissue- or cell type-specific gene expression and changes in the efficiency with which the mRNA with different leader exons is translated. We discovered at least three 5′-UTR splice variants of Ex1 (designated as E1A, E1B, and E1C) that are alternatively spliced to the common Ex2, compliant with the GT-AG rule. E1A, E1B, and E1C start at nucleotides -393, +244, and +344 and end at nucleotides -291, +338, and +529, respectively (Figure 2)[77]. These alternative splice variants are expressed in a tissue-specific manner among human normal tissues, suggesting transcriptional regulation by tissue-specific alternative promoter usage. In addition to the splice variants, our GenBank search predicted two more mRNA variants: one with an Ex1 located approximately 73 kb upstream from the TSS (designated E1U) and one that lacks Ex1 (that starts from Ex2, designated as E1-). This was further confirmed by Poonkuzhali et al. [193], who performed an extensive review of the expressed sequence tag (EST) database and genome assembly using Aceview (http://ncbi.nih.gov/IEB/Research/Acembly) and the University of California Santa Cruz genome browser (http://genome.ucsc.edu/). At least three different alternative first exons were found for BCRP: Ex1a (corresponding to E1C[77]), Ex1b (E1A), and Ex1c (E1U) [193]. Interestingly, liver samples that generated a BCRP transcript using Ex1 b had significantly lower BCRP mRNA levels. Zong et al.[194] reported three isoforms of mouse Bcrp1 mRNA that differ in their 5′-UTR (also denoted as E1a, E1b, and E1c) like human BCRP. Importantly, they found that the expression of mouse Bcrp1 during hematopoiesis is transcriptionally regulated by alternative use of multiple leader exons and promoters in a developmental stage-specific manner, indicating a promising mechanism for the lineage-specific expression of BCRP in hematopoiesis in both humans and mice.

To date, a set of cis- and trans-acting control elements distinct for each alternative promoter has not yet been identified in humans or mice. Such information may help to explain how BCRP is expressed in a tissue-specific manner. Nevertheless, some light has been shed on this subject recently. Natarajan et al.[195] found that Bcrp1 E1b mRNA spice variant expression, which is predominant in mouse intestine, is regulated by binding of phosphorylated cAMP-response element binding protein (p-CREB) to its cis site on the mouse E1b promoter region. Possibly, p-CREB may be involved in transcriptional regulation of the human BCRP mRNA variant expressed in the small intestine, E1C. This is currently under investigation in our laboratory.

BCRP alternative promoter usage was recently studied in bone marrow samples derived from pediatric patients with M7 AM L by Campbell et al. [196]. They identified, similar to our prediction from a GenBank search, the leader exon E1U to be located approximately 73 kb upstream from TSS and reported small exons (designated E1U2-E1U5) between E1U and E1A (Figure 2). They also found two other Ex1s (designated E1D and E1E) located between E1C and Ex2 and reported that these Ex1s were expressed less frequently in bone marrow samples. Although the function of the extensive variability in the BCRP 5′-UTR is not yet fully understood, further investigations may clarify the role of alternative promoters in the tissue- and cell type-specific regulation of BCRP expression.

Post-translational regulation of BCRP function

Post-translational modifications (PTMs) of BCRP, including N-glycosylation and phosphorylation, have been reported in normal and cancer cells. The first described BCRP PTM was N-glycosylation of asparagine 596[197]. Although this modification did not affect BCRP function, it may be involved in expression of functional BCRP. N-glycosylation status was shown to be important for apical sorting in polarized cells such as hepatocytes[198]. We found that the proteasome is involved, at least in part, in the post-translational down-regulation of BCRP [128]. Nakagawa et al. [199],[200] showed that N-glycosylation status may be important for BCRP post-translational down-regulation induced by ubiquitin-mediated, ER-associated degradation during the trafficking process. Once BCRP is shuttled to the plasma membrane, it is known to be degraded by the endosome-lysosome pathway after remaining in the plasma membrane domain for a certain period [201],[202]. Translocation of BCRP is also regulated by PTMs. Xie et al. [104] found that BCRP is phosphorylated by the serine/threonine kinase Pim-1 (Pim-1L, 44kDa) at threonine 362, which is essential for its translocation to the plasma membrane, in human prostate cancer LNCaP cells. LNCaP cells transfected with Pim-1L cDNA became resistant to docetaxel, which was thought not to be substrate for BCRP. These findings suggest that phosphorylation at T362 is critical for translocation and substrate specificity of BCRP. Furthermore, disrupting the Pim-1 signaling to reverse BCRP-mediated MDR may be a potential therapeutic approach[104].

Post-translational regulation of BCRP has been extensively studied in relation to the PI3K/AKT signaling axis in a variety of cells. The first report showed that impaired PI3K/AKT signaling (in Akt1−/− mice) caused loss of the SP fraction in the bone marrow, as well as localization of BCRP in the endoplasmic reticulum and absence of BCRP in the active plasma membrane location [203]. To date, similar observations have been made in tumors including glioma[204] and HCC[205]. The PI3K/AKT signaling pathway may be involved in previously reported E2-induced post-transcriptional down-regulation in breast and choriocarcinoma cell lines [177],[180]. Although the precise mechanism remains unclear, Hartz et al. [206] showed that BCRP transport function was promptly reduced in isolated brain capillaries treated with E2. This was further explained by internalization of BCRP protein after short-term exposure to E2 [207]. Under long-term exposure conditions, E2 signals through ERβ to activate the PTEN/PI3K/AKT/GSK axis, resulting in BCRP degradation through a process in which ubiquitination and proteasomal proteolysis may be involved.

In neurospheres produced from gliomas, loss of PTEN increased the SP fraction, and this increase was abolished with a PI3K inhibitor, LY294002. These results suggest that the PTEN/PI3K/AKT pathway is critical for BCRP function[204],[208]. In BCR-ABL-dependent CML cells, we noted that inhibition of BCR-ABL by imatinib blocks the PI3K/AKT pathway, leading to post-transcriptional down-regulation of BCRP surface expression, which could attenuate the BCRP-mediated resistance to TKIs and confound experiments characterizing the interaction of TKIs with BCRP[128]. Similar observations were made with nilotinib and dasatinib in BCR-ABL-dependent model systems showing a concomitant down-regulation of phosphorylated Crkl, which undergoes phosphorylation by BCR-ABL as a substrate[129]. However, similar post-transcriptional down-regulation was also observed following imatinib treatment in head and neck squamous cell carcinoma in which BCR-ABL is not expressed[209]. This effect may be mediated by blocking tyrosine kinases other than BCR-ABL, such as c-KIT or SRC; hence, the precise mechanism needs to be addressed further.

Folate deprivation was found to induce BCRP expression associated with mitoxantrone resistance in Caco-2 cells. Interestingly, BCRP expression was associated with a cytoplasmic compartment in these cells, possibly resulting in intracellular drug sequestration as a mechanism of resistance, rather than the classical mechanism of efflux of drugs from cells[210].

Regulation of BCRP expression in side population and other stem cells

Since the first report of a distinct SP in neuroblastoma cells[50], SP cells have been reported in various human cancer cell lines, including gastrointestinal[51], lung[52], breast[53],[54] hepatic[55], esophageal[56], and pancreatic[57] cancer cell lines. SP cells have been shown to be highly tumorigenic in immunodeficient mice and to possess a potent ability for colony formation and proliferation. Accordingly, SP cells are enriched for cancer stem cells (CSCs). The low Hoechst dye accumulation that is characteristic of the SP fraction is mainly mediated by expression of functional BCRP[40],[53]; therefore, SP cells have been thought to be highly resistant to conventional chemotherapeutic agents. Modulation of BCRP function in the CSC population may provide a promising way to overcome MDR by efficiently eradicating CSCs. To date, molecular mechanisms underlying BCRP expression in CSCs are not fully understood. Krishnamurthy et al.[26] showed that, under hypoxia, progenitor cells from Bcrp1−/− mice had reduced ability to form colonies as compared with those from Bcrp1+/+ mice. Inhibiting heme biosynthesis rescued Bcrp1−/− progenitor cells. Thus, BCRP has been postulated to play a critical role in protecting progenitors/stem cells from intracellular accumulation of heme-related molecules (e.g. porphyrin) to enable cell survival under conditions of hypoxia. In this regard, BCRP transcription may be regulated in stem cells via binding of the HIF1/ARNT heterodimer to the HRE in the promoter region of BCRP upon hypoxic demand.

The PI3K/AKT pathway may be essential for expression of functional BCRP in stem cells. Recently, we found that HER2 expression was significantly correlated with the presence of an SP in luminal types of breast cancer cell lines and primary cells from breast cancer patients[53]. The occurrence of SP and in vivo tumorigenicity of SP cells decreased when cells were treated with HER2 signaling inhibitor AG825 or trastuzumab, which reduced phosphorylation of HER3 and AKT. These results suggest that PI3K/AKT signaling triggered by HER2/HER3 heterodimers contributes to expression of functional BCRP in SP cells. Accordingly, HER2-induced enhancement of SP cells may be a new rationale for the aggressive phenotype of HER2+ breast tumors. How BCRP functionality is augmented by the HER2 signaling is poorly understood, and further study is warranted to clarify this molecular mechanism.

The pluripotent Oct4 gene may play a role in maintenance of cancer stem cell-like properties[211]. Oct4 induces transcription of T-cell leukemia/lymphoma 1A (TCL1A), resulting in activation of AKT in embryonic stem cells[212]. Wang et al.[213] found that overexpression of Oct4 resulted in induction of BCRP expression in chemoresistant HCC cell lines and tumor samples from patients with HCC, suggesting that BCRP is under control of the Oct4/TCL1/AKT signaling pathway in HCCs. Because the AKT signaling may modulate BCRP function through regulating its intracellular localization[203],[208] and post-transcriptionally down-regulating BCRP expression in BCR-ABL+ CML cells [128], these findings suggest there is likely a link between upstream effectors of AKT and BCRP in stem cell populations. However, how the AKT signaling regulates BCRP function through translocation, up-regulation of BCRP expression, or other mechanisms still remains unknown.

Sal-like protein 4 (SALL4), a zinc finger transcription factor essential for histogenesis, was recently shown to activate BCRP expression indirectly, suggesting that SALL4 regulates the function of BCRP in SP cells[214]. The homeobox gene MSX2 is an inducer of epithelial-mesenchymal transition. In vitro promoter analysis suggests that MSX2 requires the SP1-binding site in the BCRP promoter region so that MX2 and SP1 may cooperatively regulate BCRP transcription in pancreatic cancer stem cells[215]. BCRP expression can be transcriptionally silenced by TGF-β through direct binding of its downstream targets, Smad2/3, to the BCRP promoter/enhancer, resulting in a decrease in the number of SP cells. This observation suggests a role for TGF-β in negative regulation to maintain the cancer-initiating cells within gastric cancer[216]. In addition to such complex transcriptional regulation, post-transcriptional regulation of BCRP mediated by microRNAs (miRs) has been reported in some cancer cells. In the stem-like (BCRP+) cell population fractionated from RB143 human retinoblastoma cells, miR-328, miR-519c, and miR-520h levels are found to be lowered 9-, 15-, and 3-fold, respectively, suggesting that the high level of BCRP expression is regulated by these miRs [217]. Other mechanisms may relate to the methylation status of the CpG island region of the BCRP promoter. In human prostate cancer, BCRP is reportedly up-regulated in response to the DNA- demethylating reagent 5-aza-dC in cancer-initiating cells marked by CD117+/BCRP+, suggesting that BCRP transcription is negatively regulated by hypermethylation of CpG sites in the promoter[218].

Putative Mechanisms for Overexpression of BCRP in Response to Drug Treatments

Previously reported mechanisms for differential expression of BCRP in cells selected with chemotherapeutic agents compared to parental, drug-sensitive cells include gene amplification, histone modification, and miR-mediated regulation. Recent progresses are illustrated in Figure 3.

Figure 3. Putative mechanisms for BCRP overexpression in drug-resistant cells.

Figure 3.

A, in drug-sensitive cells, BCRP transcription is regulated by histone 3 trimethylated at lysine 9 (H3K9me3) and the proximal promoter region is reported to be methylated in cells prior to drug selection or treatment. Synthesized mRNA is negatively regulated by possible candidate miRs, including miR-519c, miR-328, and miR-520h, which are purported to bind miR response elements in 3′-UTR of BCRP mRNA. Thus, BCRP expression is transcriptionally and post-transcriptionally regulated. B, in drug-resistant cells, several possible mechanisms are hypothesized based on current evidence. BCRP gene amplification is observed in some drug-selected cancer cells. Transcription factors become more accessible because of histone 3 modulations, including acetylation at lysine 9 and 14 (H3K9ac, H3K14ac), methylation at lysine 4 (H3K4me), and phosphorylation at serine 10 (H3S10p). Multiple TSSs are likely used for induction of BCRP. Demethylation of CpG islands around promoter regions may contribute to overexpression of BCRP in response to drugs. Once BCRP mRNA is synthesized, its 3′-UTR becomes truncated, resulting in deletion of miR response elements, which in turn increases BCRP mRNA stability and levels. M: methylation; A: acetylation; P: phosphorylation.

Gene amplification

Comparative genomic hybridization and cDNA hybridization studies demonstrated a high level of BCRP amplification in cancer cell lines treated with mitoxantrone (MCF-7/MX) and, to a less extent, with doxorubicin in the presence of verapamil (MCF-7/AdrVp3000)[219]. These cell lines, produced by high selective pressure, highly overexpress BCRP. However, no amplification was observed in another cancer cell line selected with mitoxantrone (S1M1 80)[219]. This result was further confirmed by Volk et al.[80], who showed that not all cancer cell lines selected with mitoxantrone had BCRP gene amplification. We examined amplification of BCRP in MDR cancer cells, MCF-7 breast cancer cell sublines (MCF/MX and MCF/AdrVp), and human ovarian cancer IGROV1 sublines (IGROV1/MX and IGROV1/T8) by quantitative PCR. BCRP gene copy number was 15-fold greater in MCF-7/MX (selected with mitoxantrone) than in parental cells or MCF-7/AdrVp cells (selected with doxorubicin with verapamil)[77]. No BCRP amplification occurred in IGROV1/MX3 (selected with mitoxantrone) or IGROV1/T8 (selected with topotecan) cells relative to their respective parental cells. These findings suggest that gene amplification may be induced in response to drugs; however, gene amplification does not necessarily occur in all drug-selected MDR cancer cells. The reason for this differential effect is poorly understood.

Multiple transcription start sites (TSSs) and alternative promoter usage

In MCF-7 and IGROV1 cells, each splice variant (E1A, E1B, or E1C) generally has a distinct major TSS, whereas the TSS for these Ex1 variants in drug-resistant MCF-7/AdrVp and IGROV1/T8 cells are more heterogeneous. This is particularly true for the E1C variant, such that no single TSS can be assigned for a given exonal variant[77]. Moreover, our study showed that the proportion of E1C BCRP transcripts increased to up to 47% of the total BCRP mRNA transcripts in MCF-7/AdrVp cells, whereas this transcript variant was only 23% of the BCRP transcripts in parental MCF-7 cells. In contrast, proportion of the E1A transcripts to total transcripts was much less (approximately 6%) than that of E1C in MCF-7/AdrVp cells; however, the E1A transcript was predominant over E1C in MCF-7 cells. Similar observations were made in sublines of drug-selected IGROV1 cells, implying that alternative promoters using different TSSs are used to induce BCRP transcription. Heterogeneity of the TSS for Ex1 observed in drug-resistant MCF-7/AdrVp cells reflects the previous report by Ince et al.[220] that multiple TSSs in Mdr1 were often observed in actinomycin D-selected sublines of Chinese hamster lung DC-3F cells, which overexpress Mdr1 mRNA without a concomitant increase in Mdr1 gene amplification. The same investigators later described that this multiple TSSs could be associated with the response element, termed as MED-1 (multiple start site element downstream), in the Mdr1 promoter region for RNA polymerase II, which can initiate transcription at distinctive sites. In human MDR1 gene, response element termed inverted MED (iMED) was indicated to act as a cis-activator for the MDR1 gene[221],[222]. Indeed, MED-1 and iMED elements are located in the BCRP promoter region as we reported [77] (Figure 2). Further study of the role that these response elements play in BCRP expression is reasonable to identifying trans-acting factors for each promoter region.

Demethylation

Methylation of the 5′-carbon of cytosine in the CpG island of a gene promoter is a well-known epigenetic modification that silences gene expression. Methylation has been reported in the CpG island within the proximal BCRP promoter region in RCC cell lines such as UOK121 and UOK143. Furthermore, BCRP expression was up-regulated when these cells were treated with the DNA-demethylating agent, 5-aza-2′-deoxycitidine (5-aza-dC)[223]. Other investigators found that DNA demethylation is a molecular mechanism by which BCRP is overexpressed in response to drug exposure. Turner et al.[224] found that BCRP transcription is up-regulated because of demethylation of the CpG island in the promoter in human multiple myeloma 8226 cells selected with mitoxantrone (8226/MR). BCRP promoter methylation appeared to be present not only in CCRF-CEM and Jurkat leukemia cells and primary T-ALL lymphoblasts obtained from ALL patients, but also in IGROV1 ovarian carcinoma cells[225]. In solid tumors, we examined the methylation status of the BCRP promoter in drug-resistant MCF-7/AdrVp breast cancer cells and IGROV1/MX3 and IGROV1/T8 ovarian cancer cells and found no demethylation compared to parental cells [77]. However, Bram et al.[225] described prominent BCRP promoter demethylation in IGROV1/MX3 and IGROV1/T8 cells compared with their parental IGROV1 cells. Because no BCRP gene amplification was found in these cells, demethylation of the promoter region could be the cause for BCRP up-regulation in response to drugs.

Histone modifications

Recently, it has become increasingly evident that histone modulation affects BCRP expression in response to drug exposure. Histone deacetylase inhibitors (HDACi), including vorinostat and romidepsin, are in clinical use[226]. Indeed, BCRP activation has been observed in malignant peripheral blood mononuclear cells and S1 colorectal cancer cells treated with romidepsin, which increases acetylated histone H3[227]. Another HDACi phenylbutyrate[228] induced BCRP expression in S1 and KG-1a leukemia cells. Histone hyperacetylation resulting from a single-step selection with anticancer agents (e.g. doxorubicin) was associated with overexpression of BCRP in cancer cell lines MCF-7, IGROV1, and S1[229]. Furthermore, Hauswald et al.[230] demonstrated that BCRP and P-gp could be induced in mononuclear leukemic cells derived from a patient with relapsed AML when treated with HDACi (e.g. phenylbutyrate), with histone being hyperacetylated in the promoter region of BCRP. These findings show that exposure of cancer cells to HDACi induces a drug resistance phenotype that might negatively affect treatment effectiveness. To et al. [226] hypothesized an association of the BCRP promoter with histone H3 trimethylated at Lys 9 (H3K9me3) keeps BCRP expression low in drug-sensitive cells. According to the hypothetical model, once the cells are treated with drugs, H3 modifications, including acetylation at Lys 9 and 14 (H3K9ac, H3K14ac), trimethylation at Lys 4 (H3K4me3), and phosphorylation at Ser 10 (H3S10ph), induce recruitment of a chromatin remodeling factor (Brg-1) and RNA polymerase IIs, resulting in withdrawal of class I HDACs from the BCRP promoter. This modification functions as a switch to open the chromatin configuration to enhance BCRP transcription. Thus, H3 modulation seems to be a relevant rationale for manipulating BCRP expression; however, this needs to be examined further because HDACi such as romidepsin has not induced BCRP mRNA expression in all cell lines studied so far[226].

Regulation of BCRP gene expression by microRNA (miR)

To date, several miR response elements that might target BCRP mRNA 3′-UTR have been identified using multiple algorithms[231],[232]. Among them, miR-519c may be involved in BCRP overexpression in drug-resistant cancer cells. To et al.[232] found that BCRP mRNA with a truncated 3′-UTR is expressed in drug-resistant S1M1 80 cells but not in their parental S1 cells. Because the shorter 3′-UTR lacked the miR response element target site(s) for miR-519c, BCRP mRNA became stabilized, resulting in BCRP overexpression in the drug-resistant cells. Shortening of the BCRP mRNA 3′-UTR has been reported in various BCRP-overexpressing resistant cell lines, including MCF-7/FLV1000, SF295/MX2000, H460/MX20, and A549/Beca250[233]. Thus, instability of BCRP mRNA 3′-UTR in response to drugs may be one of the mechanisms for induction of BCRP expression in selected cancer cells. miR-328 is also involved in post-transcriptional regulation of BCRP because miR-328 negatively regulates both mRNA and protein expression of BCRP and is underexpressed in MDR cancer cells such as MCF-7/MX100 cells[231]. Another miR postulated to affect BCRP expression is miR-520h. Although BCRP expression was unchanged in MCF-7/MX100 cells with forced expression of miR-520h [217], miR-520h is likely involved in BCRP expression during differentiation of hematopoietic stem cells. miR-520h is enriched in hematopoietic stem cells but underexpressed in CD34+ cells, thereby maintaining BCRP expression in the stem cell population[234]. This was also noted in pancreatic cancer PANC-2 cells because introduction of miR-520h resulted in cell migration and invasion as well as reduction of the side population[235].

BCRP Polymorphisms and Cancer Treatment Outcomes

Many single nucleotide polymorphisms (SNPs) have been observed in the BCRP gene[236]. Honjo et al.[18] first described 7 SNPs in the intron sequences, 3 non-coding SNPs in the UTR, 2 synonymous SNPs, and 3 non-synonymous SNPs [c.34G>A (V12M, Ex2; rs2231137), c.421C>A (Q141K, Ex5; rs2231142), and c. 1858G>A (D620N, Ex16; rs34783571)]. Transport activity of BCRP variant proteins V12M and Q141K have been studied the most frequently and have been found to decrease BCRP function significantly when transfected into cultured cells[237][239]. Reduction of the activity of the V12M variant resulted from its decreased expression at the apical membranes[238]. Tamura et al.[240],[241] reanalyzed the transport activity of BCRP protein when 7 known BCRP variants due to SNPs, including V12M, Q141K, F208S (c.623T>C, Ex6; rs1061018), and S248P (c. 742T>C, Ex7; rs3116448), were expressed in Flp-ln-293 cells. The results indicated that relative to wild-type BCRP-expressing cells, the IC50 value of SN38 against Q141K-expressing Flp-ln-293 cells was reduced approximately 50%, but it was not reduced in V12M-expressing cells. Interestingly, the IC50 values of SN38, mitoxantrone, doxorubicin, daunorubicin, and etoposide for Flp-ln-293 cells expressing other variants were significantly lower than those for cells expressing wild-type BCRP. Notably, F208S and S441N were not expressed in the cells, suggesting that the rest of the variants (S248P, F431L, and F489L) may also impart impaired function of BCRP [241]. Another study showed that the P269S (C.805OT, Ex7; rs34678167) and Q126Stop (C.376OT, Ex4; rs72552713) alleles were found among Korean subjects, and that the transport activity of the P296S variant decreased based on methotrexate uptake by vesicles expressing this variant[242]. Not all BCRP SNPs are associated with diminished transporter activity, however. The I206L (c.616A>C, Ex6; rs12721643) variant was found to have high transporter activity but low protein expression when transfected into HEK cells, whereas the N590Y (c.1768A>T, Ex15; rs34264773) and D620N had higher expression but lower activity[243].

The effects of promoter and non-coding region SNPs on BCRP expression in the liver, intestines and lymphoblasts were recently investigated by Poonkuzhali et al.[193]. Forty-one SNPs were found in the promoter region and 49 in the introns. Promoter and intron 1 alleles were found to be associated with altered BCRP mRNA expression. Furthermore, a decrease in BCRP expression may be caused by -30477C>G (rs2127861), -15622C>T (in upstream region), and 1143G>A (in intron 1; rs2622604), and an increase in BCRP expression may be caused by -15994C>T (rs7699188), -15846A>C (in upstream region), and 12283T>C and 16702C>T (in intron 1; rs2046134). Interestingly, these investigators showed that individuals with the -15994C>T polymorphism had significantly higher clearance of oral imatinib.

Previous pharmacokinetic studies have shown that people with BCRP polymorphisms can experience significant alterations in the absorption, distribution, metabolism, or elimination of BCRP substrate drugs. The oral bioavailability of topotecan increased significantly in patients heterozygous for the c.421C>A allele [244]. Another work showed that cancer patients heterozygous for the c.421C>A (Q141K) allele had approximately 3-fold higher plasma levels of intravenous diflomotecan (a 10,11-difluoro-homocamptothecin), a new promising topoisomerase I inhibitor with enhanced plasma stability and superior preclinical anti-tumor activity, than those harboring the wild-type allele, implying that diflomotecan can be transported by BCRP[245]. The heterozygous c.421C>A allele also did not appear to affect the pharmacokinetics of irinotecan or its metabolite SN38[246] despite that BCRP transports SN38 and its glucuronide [90],[94]. Similarly, the c.34G>A or c.421C>A polymorphisms of BCRP were not associated with alterations in the pharmacokinetics of irinotecan, SN38, or SN38 glucuronide, nor were they associated with the tumor response rate or toxicity in Korean patients with advanced NSCLC treated with irinotecan[247]. Recent progress in research of both functional SNPs and BCRP inhibitory agents that modulate the in vivo pharmacokinetics and pharmacodynamics of BCRP substrate drugs are reviewed elsewhere[248].

In a study of 200 Japanese patients with non-papillary RCC and age- and sex-matched controls, Korenaga et al.[249] investigated the c.421C>A (Q141K) allele and found that individuals with the wild-type C/C genotype had a higher risk for developing non-papillary RCC. These authors concluded that BCRP may be a candidate RCC susceptibility gene. Increased susceptibility and shorter survival of patients with diffuse large B-cell lymphoma was observed in subjects with the c.34G>A or c.421C>A alleles[250]. A recent study by Hahn et al.[251] showed that patients with the c.421C>A (C/A) allele and hormone-refractory prostate cancer had significantly longer survival beyond 15 months compared to those with the wild-type (C/C) genotype when treated with docetaxel-based combination chemotherapy. Similarly, Muller et al. [252] found shorter overall survival after treatment with platinum-based regimens in a large series of patients with small cell lung cancer and NSCLC who carried the 421A allele of BCRP. Because platinum-based drugs are not identified as substrates for BCRP, the precise reason for this phenotype is not clear. Further study is needed to address this issue.

Conclusions

BCRP plays a significant role in pharmacokinetics and contributes to MDR in cancer. Therefore, to overcome MDR, it is important to understand how this transporter exerts its function and how its expression is regulated, as this information may provide clues for developing compounds that modulate BCRP activity. As we reviewed here, many mechanisms are involved in BCRP expression, suggesting that a complex overall mechanism underlies expression among a variety of tissues. BCRP function is likely regulated by the PI3K/AKT signaling pathway in a transcriptional and post-translational manner; however, the precise molecular mechanism used by this signaling cascade has not yet been determined. Because this signaling pathway is also critical for cancer cell survival and proliferation, further investigations may clarify common molecular targets that modulate BCRP expression and activity as well as cell proliferation and growth. Cancer stem cells have been recently the subject of increased scrutiny, and BCRP expression in such cell populations is garnering attention because of its potential to confer drug resistance. Current evidence does not explain sufficiently how BCRP is regulated in undifferentiated cells. Therefore, it is worthwhile to address this issue in terms of circumventing transporter-mediated MDR acquired by cancer cells by modulating activity of MDR transporters such as BCRP to sensitize cancer stem cells to conventional cancer chemotherapeutic agents and to efficiently eradicate them from tumor mass, thereby providing a clinical benefit for patients.

Acknowledgments

We thank Dr. Ikumi Tamai of Kanazawa University for helpful comments and suggestions.

References

  • 1.Roninson IB, Chin JE, Choi KG, et al. Isolation of human MDR DNA sequences amplified in multidrug-resistant KB carcinoma cells. Proc Natl Acad Sci USA. 1986;83:4538–4542. doi: 10.1073/pnas.83.12.4538. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Cole S, Bhardwaj G, Gerlach J, et al. Overexpression of a transporter gene in a multidrug-resistant human lung cancer cell line. Science. 1992;258:1650–1654. doi: 10.1126/science.1360704. [DOI] [PubMed] [Google Scholar]
  • 3.Ross DD, Doyle L, Schiffer C, et al. Expression of multidrug resistance-associated protein (MRP) mRNA in blast cells from acute myeloid leukemia (AML) patients. Leukemia. 1996;10:48–55. [PubMed] [Google Scholar]
  • 4.Chen YN, Mickley LA, Schwartz AM, et al. Characterization of adriamycin-resistant human breast cancer cells which display overexpression of a novel resistance-related membrane protein. J Biol Chem. 1990;265:10073–10080. [PubMed] [Google Scholar]
  • 5.Doyle LA, Yang W, Abruzzo LV, et al. A multidrug resistance transporter from human MCF-7 breast cancer cells. Proc Natl Acad Sci U S A. 1998;95:15665–15670. doi: 10.1073/pnas.95.26.15665. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Allikmets R, Schriml LM, Hutchinson A, et al. A human placenta-specific ATP-binding cassette gene (ABCP) on chromosome 4q22 that is involved in multidrug resistance. Cancer Res. 1998;58:5337–5339. [PubMed] [Google Scholar]
  • 7.Miyake K, Mickley L, Litman T, et al. Molecular cloning of cDNAs which are highly overexpressed in mitoxantrone-resistant cells. Cancer Res. 1999;59:8–13. [PubMed] [Google Scholar]
  • 8.Doyle LA, Ross DD. Multidrug resistance mediated by the breast cancer resistance protein BCRP (ABCG2) Oncogene. 2003;22:7340–7358. doi: 10.1038/sj.onc.1206938. [DOI] [PubMed] [Google Scholar]
  • 9.Ross DD, Nakanishi T. Clifton: Springer; 2010. Impact of breast cancer resistance protein on cancer treatment outcomes. [DOI] [PubMed] [Google Scholar]
  • 10.Nakanishi T. Drug transporters as targets for cancer chemotherapy. Cancer Genomics Proteomics. 2007;4:241–254. [PubMed] [Google Scholar]
  • 11.Higgins CF. The ABC of channel regulation. Cell. 1995;82:693–696. doi: 10.1016/0092-8674(95)90465-4. [DOI] [PubMed] [Google Scholar]
  • 12.Nakanishi T, Doyle LA, Hassel B, et al. Functional characterization of human breast cancer resistance protein (BCRP, ABCG2) expressed in the oocytes of Xenopus laevis. Mol Pharmacol. 2003;64:1452–1462. doi: 10.1124/mol.64.6.1452. [DOI] [PubMed] [Google Scholar]
  • 13.Kage K, Tsukahara S, Sugiyama T, et al. Dominant-negative inhibition of breast cancer resistance protein as drug efflux pump through the inhibition of S-S dependent homodimerization. Int J Cancer. 2002;97:626–630. doi: 10.1002/ijc.10100. [DOI] [PubMed] [Google Scholar]
  • 14.Xu J, Liu Y, Yang Y, et al. Characterization of oligomeric human half-ABC transporter ATP-binding cassette G2. J Biol Chem. 2004;279:19781–19789. doi: 10.1074/jbc.M310785200. [DOI] [PubMed] [Google Scholar]
  • 15.Kage K, Fujita T, Sugimoto Y. Role of Cys-603 in dimer/oligomer formation of the breast cancer resistance protein BCRP/ABCG2. Cancer Sci. 2005;96:866–872. doi: 10.1111/j.1349-7006.2005.00126.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Henriksen U, Fog JU, Litman T, et al. Identification of intra-and intermolecular disulfide bridges in the multidrug resistance transporter ABCG2. J Biol Chem. 2005;280:36926–36934. doi: 10.1074/jbc.M502937200. [DOI] [PubMed] [Google Scholar]
  • 17.Ni Z, Mark ME, Cai X, et al. Fluorescence resonance energy transfer (FRET) analysis demonstrates dimer/oligomer formation of the human breast cancer resistance protein (BCRP/ABCG2) in intact cells. Int J Biochem Mol Biol. 2010;1:1–11. [PMC free article] [PubMed] [Google Scholar]
  • 18.Honjo Y, Morisaki K, Huff LM, et al. Single-nucleotide polymorphism (SNP) analysis in the ABC half-transporter ABCG2 (MXR/BCRP/ABCP1) Cancer Biol Ther. 2002;1:690–696. doi: 10.4161/cbt.322. [DOI] [PubMed] [Google Scholar]
  • 19.Nakanishi T, Karp JE, Tan M, et al. Quantitative analysis of breast cancer resistance protein and cellular resistance to flavopiridol in acute leukemia patients. Clin Cancer Res. 2003;9:3320–3328. [PubMed] [Google Scholar]
  • 20.Suvannasankha A, Minderman H, O'Loughlin KL, et al. Breast cancer resistance protein (BCRP/MXR/ABCG2) in acute myeloid leukemia: discordance between expression and function. Leukemia. 2004;18:1252–1257. doi: 10.1038/sj.leu.2403395. [DOI] [PubMed] [Google Scholar]
  • 21.Suvannasankha A, Minderman H, O'Loughlin KL, et al. Breast cancer resistance protein (BCRP/MXR/ABCG2) in adult acute lymphoblastic leukaemia: frequent expression and possible correlation with shorter disease-free survival. Br J Haematol. 2004;127:392–398. doi: 10.1111/j.1365-2141.2004.05211.x. [DOI] [PubMed] [Google Scholar]
  • 22.Cooray HC, Blackmore CG, Maskell L, et al. Localisation of breast cancer resistance protein in microvessel endothelium of human brain. Neuroreport. 2002;13:2059–2063. doi: 10.1097/00001756-200211150-00014. [DOI] [PubMed] [Google Scholar]
  • 23.van Herwaarden AE, Jonker JW, Wagenaar E, et al. The breast cancer resistance protein (BCRP1/ABCG2) restricts exposure to the dietary carcinogen 2-amino-1-methyl-6-phenylimidazo [4,5-b]pyridine. Cancer Res. 2003;63:6447–6452. [PubMed] [Google Scholar]
  • 24.van Herwaarden AE, Wagenaar E, Karnekamp B, et al. Breast cancer resistance protein (BCRP1/ABCG2) reduces systemic exposure of the dietary carcinogens aflatoxin B1, IQ and Trp-P-1 but also mediates their secretion into breast milk. Carcinogenesis. 2006;27:123–130. doi: 10.1093/carcin/bgi176. [DOI] [PubMed] [Google Scholar]
  • 25.Ebert B, Seidel A, Lampen A. Identification of BCRP as transporter of benzo [a]pyrene conjugates metabolically formed in Caco-2 cells and its induction by Ah-receptor agonists. Carcinogenesis. 2005;26:1754–1763. doi: 10.1093/carcin/bgi139. [DOI] [PubMed] [Google Scholar]
  • 26.Krishnamurthy P, Ross DD, Nakanishi T, et al. The stem cell marker BCRP/ABCG2 enhances hypoxic cell survival through interactions with heme. J Biol Chem. 2004;279:24218–24225. doi: 10.1074/jbc.M313599200. [DOI] [PubMed] [Google Scholar]
  • 27.Tamura A, Watanabe M, Saito H, et al. Functional validation of the genetic polymorphisms of human ATP-binding cassette (ABC) transporter ABCG2: identification of alleles that are defective in porphyrin transport. Mol Pharmacol. 2006;70:287–296. doi: 10.1124/mol.106.023556. [DOI] [PubMed] [Google Scholar]
  • 28.Jonker JW, Buitelaar M, Wagenaar E, et al. The breast cancer resistance protein protects against a major chlorophyll-derived dietary phototoxin and protoporphyria. Proc Natl Acad Sci USA. 2002;99:15649–15654. doi: 10.1073/pnas.202607599. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Robey RW, Steadman K, Polgar O, et al. Pheophorbide a is a specific probe for ABCG2 function and inhibition. Cancer Res. 2004;64:1242–1246. doi: 10.1158/0008-5472.can-03-3298. [DOI] [PubMed] [Google Scholar]
  • 30.Volk EL, Schneider E. Wild-type breast cancer resistance protein (BCRP/ABCG2) is a methotrexate polyglutamate transporter. Cancer Res. 2003;63:5538–5543. [PubMed] [Google Scholar]
  • 31.Chen ZS, Robey RW, Belinsky MG, et al. Transport of methotrexate, methotrexate polyglutamates, and 17β-estradiol 17-(β-D-glucuronide) by ABCG2: effects of acquired mutations at R482 on methotrexate transport. Cancer Res. 2003;63:4048–4054. [PubMed] [Google Scholar]
  • 32.Suzuki M, Suzuki H, Sugimoto Y, et al. ABCG2 transports sulfated conjugates of steroids and xenobiotics. J Biol Chem. 2003;278:22644–22649. doi: 10.1074/jbc.M212399200. [DOI] [PubMed] [Google Scholar]
  • 33.Imai Y, Asada S, Tsukahara S, et al. Breast cancer resistance protein exports sulfated estrogens but not free estrogens. Mol Pharmacol. 2003;64:610–618. doi: 10.1124/mol.64.3.610. [DOI] [PubMed] [Google Scholar]
  • 34.Janvilisri T, Venter H, Shahi S, et al. Sterol transport by the human breast cancer resistance protein (ABCG2) expressed in Lactococcus lactis. J Biol Chem. 2003;278:20645–20651. doi: 10.1074/jbc.M301358200. [DOI] [PubMed] [Google Scholar]
  • 35.Huss WJ, Gray DR, Greenberg NM, et al. Breast cancer resistance protein-mediated efflux of androgen in putative benign and malignant prostate stem cells. Cancer Res. 2005;65:6640–6650. doi: 10.1158/0008-5472.CAN-04-2548. [DOI] [PubMed] [Google Scholar]
  • 36.Imai Y, Tsukahara S, Ishikawa E, et al. Estrone and 17beta-estradiol reverse breast cancer resistance protein-mediated multidrug resistance. Jpn J Cancer Res. 2002;93:231–235. doi: 10.1111/j.1349-7006.2002.tb02162.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Dehghan A, Köttgen A, Yang Q, et al. Association of three genetic loci with uric acid concentration and risk of gout: a genome-wide association study. Lancet. 2008;372:1953–1961. doi: 10.1016/S0140-6736(08)61343-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Woodward OM, Kottgen A, Coresh J, et al. Identification of a urate transporter, ABCG2, with a common functional polymorphism causing gout. Proc Natl Acad Sci USA. 2009;106:10338–10342. doi: 10.1073/pnas.0901249106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Polgar O, Robey RW, Bates SE. ABCG2: structure, function and role in drug response. Expert Opin Drug Metab Toxicol. 2008;4:1–15. doi: 10.1517/17425255.4.1.1. [DOI] [PubMed] [Google Scholar]
  • 40.Scharenberg CW, Harkey MA, Torok-Storb B. The ABCG2 transporter is an efficient Hoechst 33342 efflux pump and is preferentially expressed by immature human hematopoietic progenitors. Blood. 2002;99:507–512. doi: 10.1182/blood.v99.2.507. [DOI] [PubMed] [Google Scholar]
  • 41.Goodell M, Brose K, Paradis G, et al. Isolation and functional properties of murine hematopoietic stem cells that are replicating in vivo. J Exp Med. 1996;183:1797–1806. doi: 10.1084/jem.183.4.1797. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Zhou S, Schuetz JD, Bunting KD, et al. The ABC transporter BCRP1/ABCG2 is expressed in a wide variety of stem cells and is a molecular determinant of the side-population phenotype. Nat Med. 2001;7:1028–1034. doi: 10.1038/nm0901-1028. [DOI] [PubMed] [Google Scholar]
  • 43.Nakauchi H. Isolation and clonal characterization of hematopoietic and liver stem cells. Cornea. 2004;23(8):S2–S7. doi: 10.1097/01.ico.0000136664.18129.4f. [DOI] [PubMed] [Google Scholar]
  • 44.Summer R, Kotton DN, Sun X, et al. Side population cells and Bcrp1 expression in lung. Am J Physiol Lung Cell Mol Physiol. 2003;285:L97–L104. doi: 10.1152/ajplung.00009.2003. [DOI] [PubMed] [Google Scholar]
  • 45.Martin CM, Meeson AP, Robertson SM, et al. Persistent expression of the ATP-binding cassette transporter, ABCG2, identifies cardiac SP cells in the developing and adult heart. Dev Biol. 2004;265:262–275. doi: 10.1016/j.ydbio.2003.09.028. [DOI] [PubMed] [Google Scholar]
  • 46.Alvi AJ, Clayton H, Joshi C, et al. Functional and molecular characterisation of mammary side population cells. Breast Cancer Res. 2003;5:R1–R8. doi: 10.1186/bcr547. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Meeson AP, Hawke TJ, Graham S, et al. Cellular and molecular regulation of skeletal muscle side population cells. Stem Cells. 2004;22:1305–1320. doi: 10.1634/stemcells.2004-0077. [DOI] [PubMed] [Google Scholar]
  • 48.Mouthon M-A, Fouchet P, Mathieu C, et al. Neural stem cells from mouse forebrain are contained in a population distinct from the ‘side population’. J Neurochem. 2006;99:807–817. doi: 10.1111/j.1471-4159.2006.04118.x. [DOI] [PubMed] [Google Scholar]
  • 49.Du Y, Funderburgh ML, Mann MM, et al. Multipotent stem cells in human corneal stroma. Stem Cells. 2005;23:1266–1275. doi: 10.1634/stemcells.2004-0256. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Hirschmann-Jax C, Foster AE, Wulf GG, et al. A distinct “side population” of cells with high drug efflux capacity in human tumor cells. Proc Natl Acad Sci USA. 2004;101:14228–14233. doi: 10.1073/pnas.0400067101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Haraguchi N, Utsunomiya T, Inoue H, et al. Characterization of a side population of cancer cells from human gastrointestinal system. Stem Cells. 2005;24:506–513. doi: 10.1634/stemcells.2005-0282. [DOI] [PubMed] [Google Scholar]
  • 52.Ho MM, Ng AV, Lam S, et al. Side population in human lung cancer cell lines and tumors is enriched with stem-like cancer cells. Cancer Res. 2007;67:4827–4833. doi: 10.1158/0008-5472.CAN-06-3557. [DOI] [PubMed] [Google Scholar]
  • 53.Nakanishi T, Chumsri S, Khakpour N, et al. Side-population cells in luminal-type breast cancer have tumour-initiating cell properties, and are regulated by HER2 expression and signalling. Br J Cancer. 2010;102:815–826. doi: 10.1038/sj.bjc.6605553. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Christgen M, Ballmaier M, Bruchhardt H, et al. Identification of a distinct side population of cancer cells in the Cal-51 human breast carcinoma cell line. Mol Cell Biochem. 2007;306:201–212. doi: 10.1007/s11010-007-9570-y. [DOI] [PubMed] [Google Scholar]
  • 55.Shi GM, Xu Y, Fan J, et al. Identification of side population cells in human hepatocellular carcinoma cell lines with stepwise metastatic potentials. J Cancer Res Clin Oncol. 2008;134:1155–1163. doi: 10.1007/s00432-008-0407-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Huang D, Gao Q, Guo L, et al. Isolation and identification of cancer stem-like cells in esophageal carcinoma cell lines. Stem Cell Dev. 2008;18:465–474. doi: 10.1089/scd.2008.0033. [DOI] [PubMed] [Google Scholar]
  • 57.Wang Y, Li F, Luo B, et al. A side population of cells from a human pancreatic carcinoma cell line harbors cancer stem cell characteristics. Neoplasma. 2009;56:371–378. doi: 10.4149/neo_2009_05_371. [DOI] [PubMed] [Google Scholar]
  • 58.Ross DD, Karp JE, Chen TT, et al. Expression of breast cancer resistance protein in blast cells from patients with acute leukemia. Blood. 2000;96:365–368. [PubMed] [Google Scholar]
  • 59.van den Heuvel-Eibrink MM, Wiemer EA, Prins A, et al. Increased expression of the breast cancer resistance protein (BCRP) in relapsed or refractory acute myeloid leukemia (AML) Leukemia. 2002;16:833–839. doi: 10.1038/sj.leu.2402496. [DOI] [PubMed] [Google Scholar]
  • 60.Benderra Z, Faussat AM, Sayada L, et al. Breast cancer resistance protein and P-glycoprotein in 149 adult acute myeloid leukemias. Clin Cancer Res. 2004;10:7896–7902. doi: 10.1158/1078-0432.CCR-04-0795. [DOI] [PubMed] [Google Scholar]
  • 61.Abbott BL, Colapietro AM, Barnes Y, et al. Low levels of ABCG2 expression in adult AML blast samples. Blood. 2002;100:4594–4601. doi: 10.1182/blood-2002-01-0271. [DOI] [PubMed] [Google Scholar]
  • 62.van der Kolk DM, Vellenga E, Scheffer GL, et al. Expression and activity of breast cancer resistance protein (BCRP) in de novo and relapsed acute myeloid leukemia. Blood. 2002;99:3763–3770. doi: 10.1182/blood.v99.10.3763. [DOI] [PubMed] [Google Scholar]
  • 63.Raaijmakers MHGP, de Grouw EPLM, Heuver LHH, et al. Breast cancer resistance protein in drug resistance of primitive CD34+38− cells in acute myeloid leukemia. Clin Cancer Res. 2005;11:2436–2444. doi: 10.1158/1078-0432.CCR-04-0212. [DOI] [PubMed] [Google Scholar]
  • 64.Ho MM, Hogge DE, Ling V. MDR1 and BCRP1 expression in leukemic progenitors correlates with chemotherapy response in acute myeloid leukemia. Exp Hematol. 2008;36:433–442. doi: 10.1016/j.exphem.2007.11.014. [DOI] [PubMed] [Google Scholar]
  • 65.de Figueiredo-Pontes LL, Pintaõ MCT, Oliveira LCO, et al. Determination of P-glycoprotein, MDR-related protein 1, breast cancer resistance protein, and lung-resistance protein expression in leukemic stem cells of acute myeloid leukemia. Cytometry B Clin Cytom. 2008;74B:163–168. doi: 10.1002/cyto.b.20403. [DOI] [PubMed] [Google Scholar]
  • 66.Damiani D, Tiribelli M, Calistri E, et al. The prognostic value of P-glycoprotein (ABCBI) and breast cancer resistance protein (ABCG2) in adults with de novo acute myeloid leukemia with normal karyotype. Haematologica. 2006;91:825–828. [PubMed] [Google Scholar]
  • 67.Galimberti S, Guerrini F, Palumbo GA, et al. Evaluation of BCRP and MDR-1 co-expression by quantitative molecular assessment in AML patients. Leuk Res. 2004;28:367–372. doi: 10.1016/j.leukres.2003.09.002. [DOI] [PubMed] [Google Scholar]
  • 68.van den Heuvel-Eibrink M, van der Holt B, Burnett A, et al. CD34-related coexpression of MDR1 and BCRP indicates a clinically resistant phenotype in patients with acute myeloid leukemia (AML) of older age. Ann Hematol. 2007;86:329–337. doi: 10.1007/s00277-007-0269-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Ho RH, Choi L, Lee W, et al. Effect of drug transporter genotypes on pravastatin disposition in European- and African-American participants. Pharmacogenet Genomics. 2007;17:647–656. doi: 10.1097/FPC.0b013e3280ef698f. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Benderra Z, Faussat AM, Sayada L, et al. MRP3, BCRP, and P-glycoprotein activities are prognostic factors in adult acute myeloid leukemia. Clin Cancer Res. 2005;11:7764–7772. doi: 10.1158/1078-0432.CCR-04-1895. [DOI] [PubMed] [Google Scholar]
  • 71.Shman T, Fedasenka U, Savitski V, et al. CD34+ leukemic subpopulation predominantly displays lower spontaneous apoptosis and has higher expression levels of Bcl-2 and MDR1 genes than CD34− cells in childhood AML. Ann Hematol. 2008;87:353–360. doi: 10.1007/s00277-008-0439-2. [DOI] [PubMed] [Google Scholar]
  • 72.Plasschaert SLA, van der Kolk DM, de Bont ESJM, et al. The role of breast cancer resistance protein in acute lymphoblastic leukemia. Clin Cancer Res. 2003;9:5171–5177. [PubMed] [Google Scholar]
  • 73.Sauerbrey A, Sell W, Steinbach D, et al. Expression of the BCRP gene (ABCG2/MXR/ABCP) in childhood acute lymphoblastic leukaemia. Br J Haematol. 2002;118:147–150. doi: 10.1046/j.1365-2141.2002.03550.x. [DOI] [PubMed] [Google Scholar]
  • 74.Jordanides NE, Jorgensen HG, Holyoake TL, et al. Functional ABCG2 is overexpressed on primary cml CD34+ cells and is inhibited by imatinib mesylate. Blood. 2006;108:1370–1373. doi: 10.1182/blood-2006-02-003145. [DOI] [PubMed] [Google Scholar]
  • 75.Jiang X, Zhao Y, Smith C, et al. Chronic myeloid leukemia stem cells possess multiple unique features of resistance to BCR-ABL targeted therapies. Leukemia. 2007;21:926–935. doi: 10.1038/sj.leu.2404609. [DOI] [PubMed] [Google Scholar]
  • 76.Burger H, van Tol H, Boersma AWM, et al. Imatinib mesylate (STI571) is a substrate for the breast cancer resistance protein (BCRP)/ABCG2 drug pump. Blood. 2004;104:2940–2942. doi: 10.1182/blood-2004-04-1398. [DOI] [PubMed] [Google Scholar]
  • 77.Nakanishi T, Bailey-Dell KJ, Hassel BA, et al. Novel 5′ untranslated region variants of BCRP mRNA are differentially expressed in drug–selected cancer cells and in normal human tissues: implications for drug resistance, tissue-specific expression, and alternative promoter usage. Cancer Res. 2006;66:5007–5011. doi: 10.1158/0008-5472.CAN-05-4572. [DOI] [PubMed] [Google Scholar]
  • 78.Engler JR, Frede A, Saunders VA, et al. Chronic myeloid leukemia CD34+ cells have reduced uptake of imatinib due to low OCT-1 activity. Leukemia. 2010;24:765–770. doi: 10.1038/leu.2010.16. [DOI] [PubMed] [Google Scholar]
  • 79.Diestra JE, Scheffer GL, Catala I, et al. Frequent expression of the multi-drug resistance-associated protein BCRP/MXR/ABCP/ABCG2 in human tumours detected by the bxp-21 monoclonal antibody in paraffin-embedded material. J Pathol. 2002;198:213–219. doi: 10.1002/path.1203. [DOI] [PubMed] [Google Scholar]
  • 80.Kawabata S, Oka M, Soda H, et al. Expression and functional analyses of breast cancer resistance protein in lung cancer. Clin Cancer Res. 2003;9:3052–3057. [PubMed] [Google Scholar]
  • 81.Ross DD, Yang W, Abruzzo LV, et al. Atypical multidrug resistance: Breast cancer resistance protein messenger RNA expression in mitoxantrone-selected cell lines. J Natl Cancer Inst. 1999;91:429–433. doi: 10.1093/jnci/91.5.429. [DOI] [PubMed] [Google Scholar]
  • 82.Maliepaard M, van Gastelen MA, de Jong LA, et al. Overexpression of the BCRP/MXR/ABCP gene in a topotecan-selected ovarian tumor cell line. Cancer Res. 1999;59:4559–4563. [PubMed] [Google Scholar]
  • 83.Robey RW, Medina-Perez WY, Nishiyama K, et al. Overexpression of the ATP-binding cassette half-transporter, ABCG2 (MXR/BCRP/ABCP1), in flavopiridol-resistant human breast cancer cells. Clin Cancer Res. 2001;7:145–152. [PubMed] [Google Scholar]
  • 84.Burger H, van Tol H, Brok M, et al. Chronic imatinib mesylate exposure leads to reduced intracellular drug accumulation by induction of the ABCG2 (BCRP) and ABCBI (MDR1) drug transport pumps. Cancer Biol Ther. 2005;4:747–752. doi: 10.4161/cbt.4.7.1826. [DOI] [PubMed] [Google Scholar]
  • 85.Özvegy C, Litman T, Szakacs G, et al. Functional characterization of the human multidrug transporter, ABCG2, expressed in insect cells. Biochem Biophys Res Commun. 2001;285:111–117. doi: 10.1006/bbrc.2001.5130. [DOI] [PubMed] [Google Scholar]
  • 86.Robey RW, Honjo Y, Morisaki K, et al. Mutations at amino-acid 482 in the ABCG2 gene affect substrate and antagonist specificity. Br J Cancer. 2003;89:1971–1978. doi: 10.1038/sj.bjc.6601370. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Robey RW, Honjo Y, van de Laar A, et al. A functional assay for detection of the mitoxantrone resistance protein, MXR (ABCG2) Biochim Biophys Acta. 2001;1512:171–182. doi: 10.1016/s0005-2736(01)00308-x. [DOI] [PubMed] [Google Scholar]
  • 88.Volk EL, Farley KM, Wu Y, et al. Overexpression of wild-type breast cancer resistance protein mediates methotrexate resistance. Cancer Res. 2002;62:5035–5040. [PubMed] [Google Scholar]
  • 89.Yang CH, Schneider E, Kuo ML, et al. BCRP/MXR/ABCP expression in topotecan-resistant human breast carcinoma cells. Biochem Pharmacol. 2000;60:831–837. doi: 10.1016/s0006-2952(00)00396-8. [DOI] [PubMed] [Google Scholar]
  • 90.Schellens JH, Maliepaard M, Scheper RJ, et al. Transport of topoisomerase I inhibitors by the breast cancer resistance protein. Potential clinical implications. Ann N Y Acad Sci. 2000;922:188–194. doi: 10.1111/j.1749-6632.2000.tb07037.x. [DOI] [PubMed] [Google Scholar]
  • 91.Rajendra R, Gounder MK, Saleem A, et al. Differential effects of the breast cancer resistance protein on the cellular accumulation and cytotoxicity of 9-aminocamptothecin and 9-nitrocamptothecin. Cancer Res. 2003;63:3228–3233. [PubMed] [Google Scholar]
  • 92.Kawabata S, Oka M, Shiozawa K, et al. Breast cancer resistance protein directly confers SN-38 resistance of lung cancer cells. Biochem Biophys Res Commun. 2001;280:1216–1223. doi: 10.1006/bbrc.2001.4267. [DOI] [PubMed] [Google Scholar]
  • 93.Candeil L, Gourdier I, Peyron D, et al. ABCG2 overexpression in colon cancer cells resistant to SN38 and in irinotecan-treated metastases. Int J Cancer. 2004;109:848–854. doi: 10.1002/ijc.20032. [DOI] [PubMed] [Google Scholar]
  • 94.Nakatomi K, Yoshikawa M, Oka M, et al. Transport of 7-ethyl-10-hydroxycamptothecin (SNr-38) by breast cancer resistance protein ABCG2 in human lung cancer cells. Biochem Biophys Res Commun. 2001;288:827–832. doi: 10.1006/bbrc.2001.5850. [DOI] [PubMed] [Google Scholar]
  • 95.Li H, Jin HE, Kim W, et al. Involvement of P-glycoprotein, multidrug resistance protein 2 and breast cancer resistance protein in the transport of belotecan and topotecan in Caco-2 and MDCK II cells. Pharmaceutical Res. 2008;25:2601–2612. doi: 10.1007/s11095-008-9678-0. [DOI] [PubMed] [Google Scholar]
  • 96.De Cesare M, Beretta GL, Tinelli S, et al. Preclinical efficacy of ST1976, a novel camptothecin analog of the 7-oxyiminomethyl series. Biochem Pharmacol. 2007;73:656–664. doi: 10.1016/j.bcp.2006.11.004. [DOI] [PubMed] [Google Scholar]
  • 97.Komatani H, Kotani H, Hara Y, et al. Identification of breast cancer resistant protein/mitoxantrone resistance/placenta-specific, ATP-binding cassette transporter as a transporter of NB-506 and J-107088, topoisomerase I inhibitors with an indolocarbazole structure. Cancer Res. 2001;61:2827–2832. [PubMed] [Google Scholar]
  • 98.Honjo Y, Hrycyna CA, Yan QW, et al. Acquired mutations in the MXR/BCRP/ABCP gene alter substrate specificity in MXR/BCRP/ABCP-overexpressing cells. Cancer Res. 2001;61:6635–6639. [PubMed] [Google Scholar]
  • 99.Allen JD, van Dort SC, Buitelaar M, et al. Mouse breast cancer resistance protein (BCRP1/ABCG2) mediates etoposide resistance and transport, but etoposide oral availability is limited primarily by P-glycoprotein. Cancer Res. 2003;63:1339–1344. [PubMed] [Google Scholar]
  • 100.Mitomo H, Kato R, Ito A, et al. A functional study on polymorphism of the ATP-binding cassette transporter ABCG2: critical role of arginine-482 in methotrexate transport. Biochem J. 2003;373:767–774. doi: 10.1042/BJ20030150. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Shafran A, Ifergan I, Bram E, et al. ABCG2 harboring the Gly482 mutation confers high-level resistance to various hydrophilic antifolates. Cancer Res. 2005;65:8414–8422. doi: 10.1158/0008-5472.CAN-04-4547. [DOI] [PubMed] [Google Scholar]
  • 102.Bram E, Ifergan I, Shafran A, et al. Mutant Gly482 and Thr482 ABCG2 mediate high-level resistance to lipophilic antifolates. Cancer Chemother Pharmacol. 2006;58:826–834. doi: 10.1007/s00280-006-0230-9. [DOI] [PubMed] [Google Scholar]
  • 103.Wang X, Furukawa T, Nitanda T, et al. Breast cancer resistance protein (BCRP/ABCG2) induces cellular resistance to HIV-1 nucleoside reverse transcriptase inhibitors. Mol Pharmacol. 2003;63:65–72. doi: 10.1124/mol.63.1.65. [DOI] [PubMed] [Google Scholar]
  • 104.Xie Y, Xu K, Linn DE, et al. The 44-kda Pim-1 kinase phosphorylates BCRP/ABCG2 and thereby promotes its multimerization and drug-resistant activity in human prostate cancer cells. J Biol Chem. 2008;283:3349–3356. doi: 10.1074/jbc.M707773200. [DOI] [PubMed] [Google Scholar]
  • 105.Rabindran SK, He H, Singh M, et al. Reversal of a novel multidrug resistance mechanism in human colon carcinoma cells by fumitremorgin c. Cancer Res. 1998;58:5850–5858. [PubMed] [Google Scholar]
  • 106.Rabindran SK, Ross DD, Doyle LA, et al. Fumitremorgin c reverses multidrug resistance in cells transfected with the breast cancer resistance protein. Cancer Res. 2000;60:47–50. [PubMed] [Google Scholar]
  • 107.Noguchi K, Kawahara H, Kaji A, et al. Substrate-dependent bidirectional modulation of P-glycoprotein-mediated drug resistance by erlotinib. Cancer Sci. 2009;100:1701–1707. doi: 10.1111/j.1349-7006.2009.01213.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Allen JD, Schinkel AH. Multidrug resistance and pharmacological protection mediated by the breast cancer resistance protein (BCRP/ABCG2) Mol Cancer Ther. 2002;1:427–434. [PubMed] [Google Scholar]
  • 109.Weiss J, Rose J, Storch CH, et al. Modulation of human BCRP (ABCG2) activity by anti-HIV drugs. J Antimicrob Chemother. 2007;59:238–245. doi: 10.1093/jac/dkl474. [DOI] [PubMed] [Google Scholar]
  • 110.Ahmed-Belkacem A, Pozza A, Muñoz-Martínez F, et al. Flavonoid structure-activity studies identify 6-prenylchrysin and tectochrysin as potent and specific inhibitors of breast cancer resistance protein ABCG2. Cancer Res. 2005;65:4852–4860. doi: 10.1158/0008-5472.CAN-04-1817. [DOI] [PubMed] [Google Scholar]
  • 111.Kruijtzer CM, Beijnen JH, Rosing H, et al. Increased oral bioavailability of topotecan in combination with the breast cancer resistance protein and P-glycoprotein inhibitor GF120918. J Clin Oncol. 2002;20:2943–2950. doi: 10.1200/JCO.2002.12.116. [DOI] [PubMed] [Google Scholar]
  • 112.Özvegy-Laczka C, Hegedüs T, Varady G, et al. High-affinity interaction of tyrosine kinase inhibitors with the ABCG2 multidrug transporter. Mol Pharmacol. 2004;65:1485–1495. doi: 10.1124/mol.65.6.1485. [DOI] [PubMed] [Google Scholar]
  • 113.Wierdl M, Wall A, Morton CL, et al. Carboxylesterase-mediated sensitization of human tumor cells to CPT-11 cannot override ABCG2-mediated drug resistance. Mol Pharmacol. 2003;64:279–288. doi: 10.1124/mol.64.2.279. [DOI] [PubMed] [Google Scholar]
  • 114.Minderman H, Brooks T, O'Loughlin K, et al. Broad-spectrum modulation of ATP-binding cassette transport proteins by the taxane derivatives ortataxel (IDN-5109, BAY 59-8862) and tRA96023. Cancer Chemother Pharmacol. 2004;53:363–369. doi: 10.1007/s00280-003-0745-2. [DOI] [PubMed] [Google Scholar]
  • 115.Sugimoto Y, Tsukahara S, Imai Y, et al. Reversal of breast cancer resistance protein-mediated drug resistance by estrogen antagonists and agonists. Mol Cancer Ther. 2003;2:105–112. [PubMed] [Google Scholar]
  • 116.Houghton PJ, Germain GS, Harwood FC, et al. Imatinib mesylate is a potent inhibitor of the ABCG2 (BCRP) transporter and reverses resistance to topotecan and SN-38 in vitro. Cancer Res. 2004;64:2333–2337. doi: 10.1158/0008-5472.can-03-3344. [DOI] [PubMed] [Google Scholar]
  • 117.Hirano M, Maeda K, Matsushima S, et al. Involvement of BCRP (ABCG2) in the biliary excretion of pitavastatin. Mol Pharmacol. 2005;68:800–807. doi: 10.1124/mol.105.014019. [DOI] [PubMed] [Google Scholar]
  • 118.Boumendjel A, Nicolle E, Moraux T, et al. Piperazinobenzo-pyranones and phenalkylaminobenzopyranones: potent inhibitors of breast cancer resistance protein (ABCG2) J Med Chem. 2005;48:7275–7281. doi: 10.1021/jm050705h. [DOI] [PubMed] [Google Scholar]
  • 119.Litman T, Brangi M, Hudson E, et al. The multidrug-resistant phenotype associated with overexpression of the new ABC half-transporter, MXR (ABCG2) J Cell Sci. 2000;113:2011–2021. doi: 10.1242/jcs.113.11.2011. [DOI] [PubMed] [Google Scholar]
  • 120.Wang Q, Strab R, Kardos P, et al. Application and limitation of inhibitors in drug-transporter interactions studies. Int J Pharm. 2008;356:12–18. doi: 10.1016/j.ijpharm.2007.12.024. [DOI] [PubMed] [Google Scholar]
  • 121.Zaher H, khan A, Palandra J, et al. Breast cancer resistance protein (BCRP/ABCG2) is a major determinant of sulfasalazine absorption and elimination in the mouse. Mol Pharm. 2006;3:55–61. doi: 10.1021/mp050113v. [DOI] [PubMed] [Google Scholar]
  • 122.Jani M, Szabó P, Kis E, et al. Kinetic characterization of sulfasalazine transport by human ATP-binding cassette G2. Biol Pharm Bull. 2009;32:497–499. doi: 10.1248/bpb.32.497. [DOI] [PubMed] [Google Scholar]
  • 123.Yamasaki Y, leiri I, Kusuhara H, et al. Pharmacogenetic characterization of sulfasalazine disposition based on NAT2 and ABCG2 (BCRP) gene polymorphisms in humans. Clin Pharmacol Ther. 2008;84:95–103. doi: 10.1038/sj.clpt.6100459. [DOI] [PubMed] [Google Scholar]
  • 124.van der Heijden J, de Jong MC, Dijkmans BAC, et al. Development of sulfasalazine resistance in human T cells induces expression of the multidrug resistance transporter ABCG2 (BCRP) and augmented production of TNFα. Ann Rheum Dis. 2004;63:138–143. doi: 10.1136/ard.2002.005249. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Zhang Y, Bressler JP, Neal J, et al. ABCG2/BCRP expression modulates D-luciferin based bioluminescence imaging. Cancer Res. 2007;67:9389–9397. doi: 10.1158/0008-5472.CAN-07-0944. [DOI] [PubMed] [Google Scholar]
  • 126.Zhang Y, Byun Y, Ren YR, et al. Identification of inhibitors of ABCG2 by a bioluminescence imaging-based high-throughput assay. Cancer Res. 2009;69:5867–5875. doi: 10.1158/0008-5472.CAN-08-4866. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Breedveld P, Pluim D, Cipriani G, et al. The effect of BCRP1 (ABCG2) on the in vivo pharmacokinetics and brain penetration of imatinib mesylate (Gleevec): implications for the use of breast cancer resistance protein and P-glycoprotein inhibitors to enable the brain penetration of imatinib in patients. Cancer Res. 2005;65:2577–2582. doi: 10.1158/0008-5472.CAN-04-2416. [DOI] [PubMed] [Google Scholar]
  • 128.Nakanishi T, Shiozawa K, Hassel BA, et al. Complex interaction of BCRP/ABCG2 and imatinib in BCR-ABL-expressing cells: BCRP-mediated resistance to imatinib is attenuated by imatinib-induced reduction of BCRP expression. Blood. 2006;108:678–684. doi: 10.1182/blood-2005-10-4020. [DOI] [PubMed] [Google Scholar]
  • 129.Dohse M, Scharenberg C, Shukla S, et al. Comparison of ATP-binding cassette transporter interactions with the tyrosine kinase inhibitors imatinib, nilotinib, and dasatinib. Drug Metab Dispos. 2010;38:1371–1380. doi: 10.1124/dmd.109.031302. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Brendel C, Scharenberg C, Dohse M, et al. Imatinib mesylate and nilotinib (AMN107) exhibit high-affinity interaction with ABCG2 on primitive hematopoietic stem cells. Leukemia. 2007;21:1267–1275. doi: 10.1038/sj.leu.2404638. [DOI] [PubMed] [Google Scholar]
  • 131.Tiwari AK, Sodani K, Wang SR, et al. Nilotinib (AMN107, Tasigna) reverses multidrug resistance by inhibiting the activity of the ABCB1/Pgp and ABCG2/BCRP/MXR transporters. Biochem Pharmacol. 2009;78:153–161. doi: 10.1016/j.bcp.2009.04.002. [DOI] [PubMed] [Google Scholar]
  • 132.Hegedüs C, Özvegy-Laczka C, Apáti Á, et al. Interaction of nilotinib, dasatinib and bosutinib with ABCBI and ABCG2: implications for altered anti-cancer effects and pharmacological properties. Br J Pharmacol. 2009;158:1153–1164. doi: 10.1111/j.1476-5381.2009.00383.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Davies A, Jordanides NE, Giannoudis A, et al. Nilotinib concentration in cell lines and primary CD34+ chronic myeloid leukemia cells is not mediated by active uptake or efflux by major drug transporters. Leukemia. 2009;23:1999–2006. doi: 10.1038/leu.2009.166. [DOI] [PubMed] [Google Scholar]
  • 134.Shukla S, Skoumbourdis AP, Walsh MJ, et al. Synthesis and characterization of a BODIPY conjugate of the BCR-ABL kinase inhibitor Tasigna (nilotinib): evidence for transport of Tasigna and its fluorescent derivative by ABC drug transporters. Mol Pharm. 2011;8:1292–1302. doi: 10.1021/mp2001022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.O'Hare T, Walters DK, Stoffregen EP, et al. In vitro activity of BCR-ABL inhibitors AMN107 and BMS-354825 against clinically relevant imatinib-resistant ABL kinase domain mutants. Cancer Res. 2005;65:4500–4505. doi: 10.1158/0008-5472.CAN-05-0259. [DOI] [PubMed] [Google Scholar]
  • 136.Hiwase DK, Saunders V, Hewett D, et al. Dasatinib cellular uptake and efflux in chronic myeloid leukemia cells: therapeutic implications. Clin Cancer Res. 2008;14:3881–3888. doi: 10.1158/1078-0432.CCR-07-5095. [DOI] [PubMed] [Google Scholar]
  • 137.Chen Y, Agarwal S, Shaik NM, et al. P-glycoprotein and breast cancer resistance protein influence brain distribution of dasatinib. J Pharmacol Exp Ther. 2009;330:956–963. doi: 10.1124/jpet.109.154781. [DOI] [PubMed] [Google Scholar]
  • 138.Goias JM, Arndt K, Etienne C, et al. SKI-606, a 4-anilino-3-quinolinecarbonitrile dual inhibitor of Src and Abl kinases, is a potent antiproliferative agent against chronic myelogenous leukemia cells in culture and causes regression of K562 xenografts in nude mice. Cancer Res. 2003;63:375–381. [PubMed] [Google Scholar]
  • 139.Erlichman C, Boerner SA, Hallgren CG, et al. The HER tyrosine kinase inhibitor CI1033 enhances cytotoxicity of 7-ethyl-10-hydroxycamptothecin and topotecan by inhibiting breast cancer resistance protein-mediated drug efflux. Cancer Res. 2001;61:739–748. [PubMed] [Google Scholar]
  • 140.Xiao Y, Davidson R, Smith A, et al. A 96-well efflux assay to identify ABCG2 substrates using a stably transfected MDCK II cell line. Mol Pharm. 2006;3:45–54. doi: 10.1021/mp050088t. [DOI] [PubMed] [Google Scholar]
  • 141.Yanase K, Tsukahara S, Asada S, et al. Gefitinib reverses breast cancer resistance protein-mediated drug resistance. Mol Cancer Ther. 2004;3:1119–1125. [PubMed] [Google Scholar]
  • 142.Stewart CF, Leggas M, Schuetz JD, et al. Gefitinib enhances the antitumor activity and oral bioavailability of irinotecan in mice. Cancer Res. 2004;64:7491–7499. doi: 10.1158/0008-5472.CAN-04-0096. [DOI] [PubMed] [Google Scholar]
  • 143.Nakamura Y, Oka M, Soda H, et al. Gefitinib (“Iressa”, ZD1839), an epidermal growth factor receptor tyrosine kinase inhibitor, reverses breast cancer resistance protein/ABCG2-mediated drug resistance. Cancer Res. 2005;65:1541–1546. doi: 10.1158/0008-5472.CAN-03-2417. [DOI] [PubMed] [Google Scholar]
  • 144.Li J, Cusatis G, Brahmer J, et al. Association of variant ABCG2 and the pharmacokinetics of epidermal growth factor receptor tyrosine kinase inhibitors in cancer patients. Cancer Biol Ther. 2007;6:432–438. doi: 10.4161/cbt.6.3.3763. [DOI] [PubMed] [Google Scholar]
  • 145.Elkind NB, Szentpétery Z, Apáti Á, et al. Multidrug transporter ABCG2 prevents tumor cell death induced by the epidermal growth factor receptor inhibitor Iressa (ZD1839, Gefitinib) Cancer Res. 2005;65:1770–1777. doi: 10.1158/0008-5472.CAN-04-3303. [DOI] [PubMed] [Google Scholar]
  • 146.Azzariti A, Porcelli L, Xu JM, et al. Prolonged exposure of colon cancer cells to the epidermal growth factor receptor inhibitor gefitinib (Iressatm) and to the antiangiogenic agent ZD6474: cytotoxic and biomolecular effects. World J Gastroenterol. 2006;12:5140–5147. doi: 10.3748/wjg.v12.i32.5140. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Agarwal S, Sane R, Gallardo JL, et al. Distribution of gefitinib to the brain is limited by P-glycoprotein (ABCBI) and breast cancer resistance protein (ABCG2)-mediated active efflux. J Pharmacol Exp Ther. 2010;334:147–155. doi: 10.1124/jpet.110.167601. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Shi Z, Peng XX, Kim IW, et al. Erlotinib (Tarceva, OSI-774) antagonizes ATP-binding cassette subfamily B member 1 and ATP-binding cassette subfamily G member 2-mediated drug resistance. Cancer Res. 2007;67:11012–11020. doi: 10.1158/0008-5472.CAN-07-2686. [DOI] [PubMed] [Google Scholar]
  • 149.Marchetti S, de Vries NA, Buckle T, et al. Effect of the ATP-binding cassette drug transporters ABCBI, ABCG2, and ABCC2 on erlotinib hydrochloride (Tarceva) disposition in in vitro and in vivo pharmacokinetic studies employing Bcrp1−/−/Mdr1a/1b−/− (triple-knockout) and wild-type mice. Mol Cancer Ther. 2008;7:2280–2287. doi: 10.1158/1535-7163.MCT-07-2250. [DOI] [PubMed] [Google Scholar]
  • 150.Burris HA. Dual kinase inhibition in the treatment of breast cancer: initial experience with the EGFR/ErbB-2 inhibitor lapatinib. Oncologist. 2004;9:10–15. doi: 10.1634/theoncologist.9-suppl_3-10. [DOI] [PubMed] [Google Scholar]
  • 151.Dai CL, Tiwari AK, Wu CP, et al. Lapatinib (Tykerb, GW572016) reverses multidrug resistance in cancer cells by inhibiting the activity of ATP-binding cassette subfamily B member 1 and G member 2. Cancer Res. 2008;68:7905–7914. doi: 10.1158/0008-5472.CAN-08-0499. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Polli JW, Olson KL, Chism JP, et al. An unexpected synergist role of p-glycoprotein and breast cancer resistance protein on the central nervous system penetration of the tyrosine kinase inhibitor lapatinib (n-{3-chloro-4-[(3-fluorobenzyl)oxy]phenyl}-6-[5-({[2-(methylsulfonyl)ethyl]amino}methyl)-2-furyl]-4-quinazolina-mine; GW572016) Drug Metab Dispos. 2009;37:439–442. doi: 10.1124/dmd.108.024646. [DOI] [PubMed] [Google Scholar]
  • 153.Perry J, Ghazaly E, Kitromilidou C, et al. A synergistic interaction between lapatinib and chemotherapy agents in a panel of cell lines is due to the inhibition of the efflux pump BCRP. Mol Cancer Ther. 2010;9:3322–3329. doi: 10.1158/1535-7163.MCT-10-0197. [DOI] [PubMed] [Google Scholar]
  • 154.Zheng LS, Wang F, Li YH, et al. Vandetanib (Zactima, ZD6474) antagonizes ABCC1- and ABCG2-mediated multidrug resistance by inhibition of their transport function. PLoS One. 2009;4:e5172. doi: 10.1371/journal.pone.0005172. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Faivre S, Delbaldo C, Vera K, et al. Safety, pharmacokinetic, and antitumor activity of SU11248, a novel oral multitarget tyrosine kinase inhibitor, in patients with cancer. J Clin Oncol. 2006;24:25–35. doi: 10.1200/JCO.2005.02.2194. [DOI] [PubMed] [Google Scholar]
  • 156.Hu S, Chen Z, Franke R, et al. Interaction of the multikinase inhibitors sorafenib and sunitinib with solute carriers and ATP-binding cassette transporters. Clin Cancer Res. 2009;15:6062–6069. doi: 10.1158/1078-0432.CCR-09-0048. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Shukla S, Robey RW, Bates SE, et al. Sunitinib (Sutent, SU11248), a small-molecule receptor tyrosine kinase inhibitor, blocks function of the ATP-binding cassette (ABC) transporters P-glycoprotein (ABCBI) and ABCG2. Drug Metab Dispos. 2009;37:359–365. doi: 10.1124/dmd.108.024612. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Poller B, Wagenaar E, Tang S, et al. Double-transduced MDCKII cells to study human P-glycoprotein (ABCBI) and breast cancer resistance protein (ABCG2) interplay in drug transport across the blood-brain barrier. Mol Pharm. 2011;8:571–582. doi: 10.1021/mp1003898. [DOI] [PubMed] [Google Scholar]
  • 159.Tang SC, Lagas JS, Lankheet NAG, et al. Brain accumulation of sunitinib is restricted by P-glycoprotein (ABCBI) and breast cancer resistance protein (ABCG2) and can be enhanced by oral elacridar and sunitinib coadministration. Int J Cancer. 2011 doi: 10.1002/ijc.26000. [DOI] [PubMed] [Google Scholar]
  • 160.Wilhelm SM, Adnane L, Newell P, et al. Preclinical overview of sorafenib, a multikinase inhibitor that targets both RAF and VEGF and PDGF receptor tyrosine kinase signaling. Mol Cancer Ther. 2008;7:3129–3140. doi: 10.1158/1535-7163.MCT-08-0013. [DOI] [PubMed] [Google Scholar]
  • 161.Lagas J, van Waterschoot R, Sparidans R, et al. Breast cancer resistance protein and P-glycoprotein limit sorafenib brain accumulation. Mol Cancer Ther. 2010;9:319–326. doi: 10.1158/1535-7163.MCT-09-0663. [DOI] [PubMed] [Google Scholar]
  • 162.Agarwal S, Sane R, Ohlfest JR, et al. The role of the breast cancer resistance protein (ABCG2) in the distribution of sorafenib to the brain. J Pharmacol Exp Ther. 2011;336:223–233. doi: 10.1124/jpet.110.175034. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Asakawa C, Ogawa M, Kumata K, et al. [11C]Sorafenib: radiosynthesis and preliminary PET study of brain uptake in P-gp/BCRP knockout mice. Bioorg Med Chem Lett. 2011;21:2220–2223. doi: 10.1016/j.bmcl.2011.03.002. [DOI] [PubMed] [Google Scholar]
  • 164.Poller B, lusuf D, Sparidans RW, et al. Differential impact of P-glycoprotein (ABCBI) and breast cancer resistance protein (ABCG2) on axitinib brain accumulation and oral plasma pharmacokinetics. Drug Metab Dispos. 2011;39:729–735. doi: 10.1124/dmd.110.037317. [DOI] [PubMed] [Google Scholar]
  • 165.Canu B, Fioravanti A, Orlandi P, et al. Irinotecan synergistically enhances the antiproliferative and proapoptotic effects of axitinib in vitro and improves its anticancer activity in vivo. Neoplasia. 2011;13:217–229. doi: 10.1593/neo.101334. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Mi YJ, Liang YJ, Huang HB, et al. Apatinib (YN968D1) reverses multidrug resistance by inhibiting the efflux function of multiple ATP-binding cassette transporters. Cancer Res. 2010;70:7981–7991. doi: 10.1158/0008-5472.CAN-10-0111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Senderowicz AM, Headlee D, Stinson SF, et al. Phase I trial of continuous infusion flavopiridol, a novel cyclin-dependent kinase inhibitor, in patients with refractory neoplasms. J Clin Oncol. 1998;16:2986–2999. doi: 10.1200/JCO.1998.16.9.2986. [DOI] [PubMed] [Google Scholar]
  • 168.Wang LM, Ren DM. Flavopiridol, the first cyclin-dependent kinase inhibitor: recent advances in combination chemotherapy. Mini Rev Med Chem. 2010;10:1058–1070. doi: 10.2174/1389557511009011058. [DOI] [PubMed] [Google Scholar]
  • 169.An R, Hagiya Y, Tamura A, et al. Cellular phototoxicity evoked through the inhibition of human ABC transporter ABCG2 by cyclin-dependent kinase inhibitors in vitro. Pharm Res. 2009;26:449–458. doi: 10.1007/s11095-008-9738-5. [DOI] [PubMed] [Google Scholar]
  • 170.Emanuel S, Rugg CA, Gruninger RH, et al. The in vitro and in vivo effects of JNJ-7706621: a dual inhibitor of cyclin-dependent kinases and aurora kinases. Cancer Res. 2005;65:9038–9046. doi: 10.1158/0008-5472.CAN-05-0882. [DOI] [PubMed] [Google Scholar]
  • 171.Seamon JA, Rugg CA, Emanuel S, et al. Role of the ABCG2 drug transporter in the resistance and oral bioavailability of a potent cyclin-dependent kinase/aurora kinase inhibitor. Mol Cancer Ther. 2006;5:2459–2467. doi: 10.1158/1535-7163.MCT-06-0339. [DOI] [PubMed] [Google Scholar]
  • 172.Bailey-Dell KJ, Hassel B, Doyle LA, et al. Promoter characterization and genomic organization of the human breast cancer resistance protein (ATP-binding cassette transporter G2) gene. Biochim Biophys Acta. 2001;1520:234–241. doi: 10.1016/s0167-4781(01)00270-6. [DOI] [PubMed] [Google Scholar]
  • 173.Ueda K, Pastan I, Gottesman M. Isolation and sequence of the promoter region of the human multidrug-resistance (P-glycoprotein) gene. J Biol Chem. 1987;262:17432–17436. [PubMed] [Google Scholar]
  • 174.Zhu Q, Center MS. Cloning and sequence analysis of the promoter region of the MRP gene of HL60 cells isolated for resistance to adriamycin. Cancer Res. 1994;54:4488–4492. [PubMed] [Google Scholar]
  • 175.Meyer zu Schwabedissen HE, Grube M, Dreisbach A, et al. Epidermal growth factor-mediated activation of the map kinase cascade results in altered expression and function of ABCG2 (BCRP) Drug Metab Dispos. 2006;34:524–533. doi: 10.1124/dmd.105.007591. [DOI] [PubMed] [Google Scholar]
  • 176.Yasuda S, Itagaki S, Hirano T, et al. Expression level of ABCG2 in the placenta decreases from the mid stage to the end of gestation. Biosci Biotechnol Biochem. 2005;69:1871–1876. doi: 10.1271/bbb.69.1871. [DOI] [PubMed] [Google Scholar]
  • 177.Wang H, Zhou L, Gupta A, et al. Regulation of BCRP/ABCG2 expression by progesterone and 17beta-estradiol in human placental BeWo cells. Am J Physiol Endocrinol Metab. 2006;290:E798–807. doi: 10.1152/ajpendo.00397.2005. [DOI] [PubMed] [Google Scholar]
  • 178.Ee PLR, Kamalakaran S, Tonetti D, et al. Identification of a novel estrogen response element in the breast cancer resistance protein (ABCG2) gene. Cancer Res. 2004;64:1247–1251. doi: 10.1158/0008-5472.can-03-3583. [DOI] [PubMed] [Google Scholar]
  • 179.Evseenko DA, Paxton JW, Keelan JA. Independent regulation of apical and basolateral drug transporter expression and function in placental trophoblasts by cytokines, steroids, and growth factors. Drug Metab Dispos. 2007;35:595–601. doi: 10.1124/dmd.106.011478. [DOI] [PubMed] [Google Scholar]
  • 180.Imai Y, Ishikawa E, Asada S, et al. Estrogen-mediated post transcriptional down-regulation of breast cancer resistance protein/ABCG2. Cancer Res. 2005;65:596–604. [PubMed] [Google Scholar]
  • 181.Wang H, Lee E-W, Zhou L, et al. Progesterone receptor (PR) isoforms PRA and PRB differentially regulate expression of the breast cancer resistance protein in human placental choriocarcinoma BeWo cells. Mol Pharmacol. 2008;73:845–854. doi: 10.1124/mol.107.041087. [DOI] [PubMed] [Google Scholar]
  • 182.Vore M, Leggas M. Progesterone acts via progesterone receptors A and B to regulate breast cancer resistance protein expression. Mol Pharmacol. 2008;73:613–615. doi: 10.1124/mol.107.044289. [DOI] [PubMed] [Google Scholar]
  • 183.Szatmari I, Vamosi G, Brazda P, et al. Peroxisome proliferator-activated receptor γ-regulated ABCG2 expression confers cytoprotection to human dendritic cells. J Biol Chem. 2006;281:23812–23823. doi: 10.1074/jbc.M604890200. [DOI] [PubMed] [Google Scholar]
  • 184.Hirai T, Fukui Y, Motojima K. PPARα agonists positively and negatively regulate the expression of several nutrient/drug transporters in mouse small intestine. Biol Pharm Bull. 2007;30:2185–2190. doi: 10.1248/bpb.30.2185. [DOI] [PubMed] [Google Scholar]
  • 185.Ebert B, Seidel A, Lampen A. Phytochemicals induce breast cancer resistance protein in Caco-2 cells and enhance the transport of benzo [a]pyrene-3-sulfate. Toxicol Sci. 2007;96:227–236. doi: 10.1093/toxsci/kfl147. [DOI] [PubMed] [Google Scholar]
  • 186.Tan KP, Wang B, Yang M, et al. Aryl hydrocarbon receptor is a transcriptional activator of the human breast cancer resistance protein (BCRP/ABCG2) Mol Pharmacol. 2010;78:175–185. doi: 10.1124/mol.110.065078. [DOI] [PubMed] [Google Scholar]
  • 187.Tompkins LM, Li H, Li L, et al. A novel xenobiotic responsive element regulated by aryl hydrocarbon receptor is involved in the induction of BCRP/ABCG2 in LST74T cells. Biochem Pharmacol. 2010;80:1754–1761. doi: 10.1016/j.bcp.2010.08.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Dhakshinamoorthy S, Long DN, Jaiswal A. Antioxidant regulation of genes encoding enzymes that detoxify xenobiotics and carcinogens. Curr Top Cell Regul. 2000;36:201–216. doi: 10.1016/s0070-2137(01)80009-1. [DOI] [PubMed] [Google Scholar]
  • 189.Adachi T, Nakagawa H, Chung I, et al. Nrf2-dependent and -independent induction of ABC transporters ABCC1, ABCC2, and ABCG2 in HepG2 cells under oxidative stress. J Exp Ther Oncol. 2007;6:335–348. [PubMed] [Google Scholar]
  • 190.Singh A, Wu H, Zhang P, et al. Expression of ABCG2 (BCRP) is regulated by Nrf2 in cancer cells that confers side population and chemoresistance phenotype. Mol Cancer Ther. 2010;9:2365–2376. doi: 10.1158/1535-7163.MCT-10-0108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.Nakanishi T, Shiozawa K, Hamburger AW, et al. Bcrp expression is functionally upregulated by epidermal growth factor receptor (EGFR, ERBB1) mediated signaling in human ovarian cancer cell lines, but not in human breast cancer cell lines. AACR Meeting Abstracts. 2006:146. [Google Scholar]
  • 192.Chen JS, Pardo FS, Wang-Rodriguez J, et al. EGFR regulates the side population in head and neck squamous cell carcinoma. Laryngoscope. 2006;116:401–406. doi: 10.1097/01.mlg.0000195075.14093.fb. [DOI] [PubMed] [Google Scholar]
  • 193.Poonkuzhali B, Lamba J, Strom S, et al. Association of breast cancer resistance protein/ABCG2 phenotypes and novel promoter and intron 1 single nucleotide polymorphisms. Drug Metab Dispos. 2008;36:780–795. doi: 10.1124/dmd.107.018366. [DOI] [PubMed] [Google Scholar]
  • 194.Zong Y, Zhou S, Fatima S, et al. Expression of mouse Abcg2 mRNA during hematopoiesis is regulated by alternative use of multiple leader exons and promoters. J Biol Chem. 2006;281:29625–29632. doi: 10.1074/jbc.M606314200. [DOI] [PubMed] [Google Scholar]
  • 195.Natarajan K, Xie Y, Nakanishi T, et al. Identification and characterization of the major alternative promoter regulating Bcrp1/Abcg2 expression in the mouse intestine. Biochim Biophys Acta. 2011;1809:295–305. doi: 10.1016/j.bbagrm.2011.06.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 196.Campbell PK, Zong Y, Yang S, et al. Identification of a novel, tissue-specific ABCG2 promoter expressed in pediatric acute megakaryoblastic leukemia. Leuk Res. 2011;35:1321–1329. doi: 10.1016/j.leukres.2011.05.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197.Diop NK, Hrycyna CA. N-linked glycosylation of the human ABC transporter ABCG2 on asparagine 596 is not essential for expression, transport activity, or trafficking to the plasma membrane. Biochemistry. 2005;44:5420–5429. doi: 10.1021/bi0479858. [DOI] [PubMed] [Google Scholar]
  • 198.Draheim V, Reichel A, Weitschies W, et al. N-glycosylation of ABC transporters is associated with functional activity in sandwich-cultured rat hepatocytes. Eur J Pharm Sci. 2010;41:201–209. doi: 10.1016/j.ejps.2010.06.005. [DOI] [PubMed] [Google Scholar]
  • 199.Nakagawa H, Wakabayashi-Nakao K, Tamura A, et al. Disruption of N-linked glycosylation enhances ubiquitin-mediated proteasomal degradation of the human ATP-binding cassette transporter ABCG2. FEBS J. 2009;276:7237–7252. doi: 10.1111/j.1742-4658.2009.07423.x. [DOI] [PubMed] [Google Scholar]
  • 200.Sugiyama T, Shuto T, Suzuki S, et al. Posttranslational negative regulation of glycosylated and non-glycosylated BCRP expression by derlin-1. Biochem Biophys Res Commun. 2011;404:853–858. doi: 10.1016/j.bbrc.2010.12.074. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 201.Peng H, Dong Z, Qi J, et al. A novel two mode-acting inhibitor of ABCG2-mediated multidrug transport and resistance in cancer chemotherapy. PLoS One. 2009;4:e5676. doi: 10.1371/journal.pone.0005676. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 202.Wakabayashi-Nakao K, Tamura A, Furukawa T, et al. Quality control of human ABCG2 protein in the endoplasmic reticulum: Ubiquitination and proteasomal degradation. Adv Drug Deliv Rev. 2009;61:66–72. doi: 10.1016/j.addr.2008.08.008. [DOI] [PubMed] [Google Scholar]
  • 203.Mogi M, Yang J, Lambert JF, et al. Akt signaling regulates side population cell phenotype via Bcrp1 translocation. J Biol Chem. 2003;278:39068–39075. doi: 10.1074/jbc.M306362200. [DOI] [PubMed] [Google Scholar]
  • 204.Bleau A-M, Hambardzumyan D, Ozawa T, et al. PTEN/PI3K/AKT pathway regulates the side population phenotype and ABCG2 activity in glioma tumor stem-like cells. Cell Stem Cell. 2009;4:226–235. doi: 10.1016/j.stem.2009.01.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 205.Hu C, Li H, Li J, et al. Analysis of ABCG2 expression and side population identifies intrinsic drug efflux in the HCC cell line MHCC-97L and its modulation by AKT signaling. Carcinogenesis. 2008;29:2289–2297. doi: 10.1093/carcin/bgn223. [DOI] [PubMed] [Google Scholar]
  • 206.Hartz AMS, Mahringer A, Miller DS, et al. 17-β-estradiol: a powerful modulator of blood-brain barrier BCRP activity. J Cereb Blood Flow Metab. 2010;30:1742–1755. doi: 10.1038/jcbfm.2010.36. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 207.Hartz AMS, Madole EK, Miller DS, et al. Estrogen receptor beta signaling through phosphatase and tensin homolog/phosphoinositide 3-kinase/Akt/glycogen synthase kinase 3 down-regulates blood-brain barrier breast cancer resistance protein. J Pharmacol Exp Ther. 2010;334:467–476. doi: 10.1124/jpet.110.168930. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 208.Takada T, Suzuki H, Gotoh Y, et al. Regulation of the cell surface expression of human BCRP/ABCG2 by the phosphorylation state of AKT in polarized cells. Drug Metab Dispos. 2005;33:905–909. doi: 10.1124/dmd.104.003228. [DOI] [PubMed] [Google Scholar]
  • 209.Chu TS, Chen JS, Lopez JP, et al. Imatinib-mediated inactivation of AKT regulates ABCG2 function in head and neck squamous cell carcinoma. Arch Otolaryngol Head Neck Surg. 2008;134:979–984. doi: 10.1001/archotol.134.9.979. [DOI] [PubMed] [Google Scholar]
  • 210.Lemos C, Kathmann I, Giovannetti E, et al. Folate deprivation induces BCRP (ABCG2) expression and mitoxantrone resistance in Caco-2 cells. Int J Cancer. 2008;123:1712–1720. doi: 10.1002/ijc.23677. [DOI] [PubMed] [Google Scholar]
  • 211.Hochedlinger K, Yamada Y, Beard C, et al. Ectopic expression of Oct-4 blocks progenitor-cell differentiation and causes dysplasia in epithelial tissues. Cell. 2005;121:465–477. doi: 10.1016/j.cell.2005.02.018. [DOI] [PubMed] [Google Scholar]
  • 212.Matoba R, Niwa H, Masui S, et al. Dissecting Oct3/4-regulated gene networks in embryonic stem cells by expression profiling. PLoS One. 2006;1:e26. doi: 10.1371/journal.pone.0000026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 213.Wang XQ, Ongkeko WM, Chen L, et al. Octamer 4 (Oct4) mediates chemotherapeutic drug resistance in liver cancer cells through a potential Oct4-AKT-ATP-binding cassette G2 pathway. Hepatology. 2010;52:528–539. doi: 10.1002/hep.23692. [DOI] [PubMed] [Google Scholar]
  • 214.Jeong HW, Cui W, Yang Y, et al. SALL4, a stem cell factor, affects the side population by regulation of the ATP-binding cassette drug transport genes. PLoS One. 2011;6:e18372. doi: 10.1371/journal.pone.0018372. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 215.Hamada S, Satoh K, Hirota M, et al. The homeobox gene MSX2 determines chemosensitivity of pancreatic cancer cells via the regulation of transporter gene ABCG2. J Cell Physiol. 2012;227:729–738. doi: 10.1002/jcp.22781. [DOI] [PubMed] [Google Scholar]
  • 216.Ehata S, Johansson E, Katayama R, et al. Transforming growth factor-β decreases the cancer-initiating cell population within diffuse-type gastric carcinoma cells. Oncogene. 2011;30:1693–1705. doi: 10.1038/onc.2010.546. [DOI] [PubMed] [Google Scholar]
  • 217.Li X, Pan YZ, Seigel GM, et al. Breast cancer resistance protein BCRP/ABCG2 regulatory microRNAs (hsa-miR-328, -519c and -520h) and their differential expression in stem-like ABCG2+ cancer cells. Biochem Pharmacol. 2011;81:783–792. doi: 10.1016/j.bcp.2010.12.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 218.Liu T, Xu F, Du X, et al. Establishment and characterization of multi-drug resistant, prostate carcinoma-initiating stem-like cells from human prostate cancer cell lines 22Rv1. Mol Cell Biochem. 2010;340:265–273. doi: 10.1007/s11010-010-0426-5. [DOI] [PubMed] [Google Scholar]
  • 219.Knutsen T, Rao VK, Ried T, et al. Amplification of 4q21-q22 and the MXR gene in independently derived mitoxantrone-resistant cell lines. Genes Chromosomes Cancer. 2000;27:110–116. [PubMed] [Google Scholar]
  • 220.Ince TA, Scotto KW. Differential utilization of multiple transcription start points accompanies the overexpression of the p-glycoprotein-encoding gene in Chinese hamster lung cells. Gene. 1995;156:287–290. doi: 10.1016/0378-1119(94)00907-a. [DOI] [PubMed] [Google Scholar]
  • 221.Ince TA, Scotto KW. A conserved downstream element defines a new class of RNA polymerase II promoters. J Biol Chem. 1995;270:30249–30252. doi: 10.1074/jbc.270.51.30249. [DOI] [PubMed] [Google Scholar]
  • 222.Labialle S, Gayet L, Marthinet E, et al. Transcriptional regulation of the human MDR1 gene at the level of the inverted MED-1 promoter region. Ann N Y Acad Sci. 2002;973:468–471. doi: 10.1111/j.1749-6632.2002.tb04685.x. [DOI] [PubMed] [Google Scholar]
  • 223.To KKW, Zhan Z, Bates SE. Aberrant promoter methylation of the ABCG2 gene in renal carcinoma. Mol Cell Biol. 2006;26:8572–8585. doi: 10.1128/MCB.00650-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 224.Turner JG, Gump JL, Zhang C, et al. ABCG2 expression, function, and promoter methylation in human multiple myeloma. Blood. 2006;108:3881–3889. doi: 10.1182/blood-2005-10-009084. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 225.Bram E, Stark M, Raz S, et al. Chemotherapeutic drug-induced ABCG2 promoter demethylation as a novel mechanism of acquired multidrug resistance. Neoplasia. 2009;11:1359–1370. doi: 10.1593/neo.91314. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 226.To KKW, Polgar O, Huff LM, et al. Histone modifications at the ABCG2 promoter following treatment with histone deacetylase inhibitor mirror those in multidrug-resistant cells. Mol Cancer Res. 2008;6:151–164. doi: 10.1158/1541-7786.MCR-07-0175. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 227.Robey RW, Zhan Z, Piekarz RL, et al. Increased MDR1 expression in normal and malignant peripheral blood mononuclear cells obtained from patients receiving depsipeptide (FR901228, FK228, NSC630176) Clin Cancer Res. 2006;12:1547–1555. doi: 10.1158/1078-0432.CCR-05-1423. [DOI] [PubMed] [Google Scholar]
  • 228.To KKW, Robey R, Zhan Z, et al. Upregulation of ABCG2 by romidepsin via the aryl hydrocarbon receptor pathway. Mol Cancer Res. 2011;9:516–527. doi: 10.1158/1541-7786.MCR-10-0270. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 229.Calcagno AM, Fostel JM, To KKW, et al. Single-step doxorubicin-selected cancer cells overexpress the ABCG2 drug transporter through epigenetic changes. Br J Cancer. 2008;98:1515–1524. doi: 10.1038/sj.bjc.6604334. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 230.Hauswald S, Duque-Afonso J, Wagner MM, et al. Histone deacetylase inhibitors induce a very broad, pleiotropic anticancer drug resistance phenotype in acute myeloid leukemia cells by modulation of multiple ABC transporter genes. Clin Cancer Res. 2009;15:3705–3715. doi: 10.1158/1078-0432.CCR-08-2048. [DOI] [PubMed] [Google Scholar]
  • 231.Pan YZ, Morris ME, Yu AM. MicroRNA-328 negatively regulates the expression of breast cancer resistance protein (BCRP/ABCG2) in human cancer cells. Mol Pharmacol. 2009;75:1374–1379. doi: 10.1124/mol.108.054163. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 232.To KKW, Zhan Z, Litman T, et al. Regulation of ABCG2 expression at the 3′ untranslated region of its mRNA through modulation of transcript stability and protein translation by a putative microRNA in the S1 colon cancer cell line. Mol Cell Biol. 2008;28:5147–5161. doi: 10.1128/MCB.00331-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 233.To KKW, Robey RW, Knutsen T, et al. Escape from hsa-miR-519c enables drug-resistant cells to maintain high expression of ABCG2. Mol Cancer Ther. 2009;8:2959–2968. doi: 10.1158/1535-7163.MCT-09-0292. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 234.Liao R, Sun J, Zhang L, et al. MicroRNAs play a role in the development of human hematopoietic stem cells. J Cell Biochem. 2008;104:805–817. doi: 10.1002/jcb.21668. [DOI] [PubMed] [Google Scholar]
  • 235.Wang F, Xue X, Wei J, et al. Hsa-miR-520h downregulates ABCG2 in pancreatic cancer cells to inhibit migration, invasion, and side populations. Br J Cancer. 2010;103:567–574. doi: 10.1038/sj.bjc.6605724. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 236.Iida A, Saito S, Sekine A, et al. Catalog of 605 single-nucleotide polymorphisms (SNPs) among 13 genes encoding human ATP-binding cassette transporters: ABCA4, ABCA7, ABCA8, ABCD1, ABCD3, ABCD4, ABCE1, ABCF1, ABCG1, ABCG2, ABCG4, ABCG5, and ABCG8. J Hum Genet. 2002;47:285–310. doi: 10.1007/s100380200041. [DOI] [PubMed] [Google Scholar]
  • 237.Imai Y, Nakane M, Kage K, et al. C421A polymorphism in the human breast cancer resistance protein gene is associated with low expression of Q141K protein and low-level drug resistance. Mol Cancer Ther. 2002;1:611–616. [PubMed] [Google Scholar]
  • 238.Mizuarai S, Aozasa N, Kotani H. Single nucleotide polymorphisms result in impaired membrane localization and reduced ATPase activity in multidrug transporter ABCG2. Int J Cancer. 2004;109:238–246. doi: 10.1002/ijc.11669. [DOI] [PubMed] [Google Scholar]
  • 239.Zamber CP, Lamba JK, Yasuda K, et al. Natural allelic variants of breast cancer resistance protein (BCRP) and their relationship to BCRP expression in human intestine. Pharmacogenetics. 2003;13:19–28. doi: 10.1097/00008571-200301000-00004. [DOI] [PubMed] [Google Scholar]
  • 240.Tamura A, Wakabayashi K, Onishi Y, et al. Genetic polymorphisms of human ABC transporter ABCG2: development of the standard method for functional validation of SNPs by using the Flp recombinase system. J Exp Ther Oncol. 2006;6:1–11. [PubMed] [Google Scholar]
  • 241.Tamura A, Wakabayashi K, Onishi Y, et al. Re-evaluation and functional classification of non-synonymous single nucleotide polymorphisms of the human ATP-binding cassette transporter ABCG2. Cancer Sci. 2007;98:231–239. doi: 10.1111/j.1349-7006.2006.00371.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 242.Lee SS, Jeong HE, Yi JM, et al. Identification and functional assessment of BCRP polymorphisms in a Korean population. Drug Metab Dispos. 2007;35:623–632. doi: 10.1124/dmd.106.012302. [DOI] [PubMed] [Google Scholar]
  • 243.Vethanayagam RR, Wang H, Gupta A, et al. Functional analysis of the human variants of breast cancer resistance protein: I206L, N590Y, and D620N. Drug Metab Dispos. 2005;33:697–705. doi: 10.1124/dmd.105.003657. [DOI] [PubMed] [Google Scholar]
  • 244.Sparreboom A, Loos W, Burger H, et al. Effect of ABCG2 genotype on the oral bioavailability of topotecan. Cancer Biol Ther. 2005;4:650–653. doi: 10.4161/cbt.4.6.1731. [DOI] [PubMed] [Google Scholar]
  • 245.Sparreboom A, Gelderblom H, Marsh S, et al. Diflomotecan pharmacokinetics in relation to ABCG2 421C>A genotype. Clin Pharmacol Ther. 2004;76:38–44. doi: 10.1016/j.clpt.2004.03.003. [DOI] [PubMed] [Google Scholar]
  • 246.de Jong FA, Marsh S, Mathijssen RHJ, et al. ABCG2 pharmacogenetics. Clin Cancer Res. 2004;10:5889–5894. doi: 10.1158/1078-0432.CCR-04-0144. [DOI] [PubMed] [Google Scholar]
  • 247.Han JY, Lim HS, Yoo YK, et al. Associations of ABCBI, ABCC2, and ABCG2 polymorphisms with irinotecan-pharmacokinetics and clinical outcome in patients with advanced non-small cell lung cancer. Cancer. 2007;110:138–147. doi: 10.1002/cncr.22760. [DOI] [PubMed] [Google Scholar]
  • 248.Noguchi K, Katayama K, Mitsuhashi J, et al. Functions of the breast cancer resistance protein (BCRP/ABCG2) in chemotherapy. Adv Drug Deliv Rev. 2009;61:26–33. doi: 10.1016/j.addr.2008.07.003. [DOI] [PubMed] [Google Scholar]
  • 249.Korenaga Y, Naito K, Okayama N, et al. Association of the BCRP C421A polymorphism with nonpapillary renal cell carcinoma. Int J Cancer. 2005;117:431–434. doi: 10.1002/ijc.21187. [DOI] [PubMed] [Google Scholar]
  • 250.Hu LL, Wang XX, Chen X, et al. BCRP gene polymorphisms are associated with susceptibility and survival of diffuse large B-cell lymphoma. Carcinogenesis. 2007;28:1740–1744. doi: 10.1093/carcin/bgm113. [DOI] [PubMed] [Google Scholar]
  • 251.Hahn NM, Marsh S, Fisher W, et al. Hoosier oncology group randomized phase II study of docetaxel, vinorelbine, and estramustine in combination in hormone-refractory prostate cancer with Pharmacogenetic survival analysis. Clin Cancer Res. 2006;12:6094–6099. doi: 10.1158/1078-0432.CCR-06-1188. [DOI] [PubMed] [Google Scholar]
  • 252.Moller PJ, Dally H, Klappenecker CN, et al. Polymorphisms in ABCG2, ABCC3 and CNT1 genes and their possible impact on chemotherapy outcome of lung cancer patients. Int J Cancer. 2009;124:1669–1674. doi: 10.1002/ijc.23956. [DOI] [PubMed] [Google Scholar]

Articles from Chinese Journal of Cancer are provided here courtesy of BMC

RESOURCES