Skip to main content
Drug Metabolism and Disposition logoLink to Drug Metabolism and Disposition
. 2014 Apr;42(4):611–622. doi: 10.1124/dmd.113.055806

Uptake Carriers and Oncology Drug Safety

Jason A Sprowl 1, Alex Sparreboom 1,
PMCID: PMC3965905  PMID: 24378324

Abstract

Members of the solute carrier (SLC) family of transporters are responsible for the cellular influx of a broad range of endogenous compounds and xenobiotics in multiple tissues. Many of these transporters are highly expressed in the gastrointestinal tract, liver, and kidney and are considered to be of particular importance in governing drug absorption, elimination, and cellular sensitivity of specific organs to a wide variety of oncology drugs. Although the majority of studies on the interaction of oncology drugs with SLC have been restricted to the use of exploratory in vitro model systems, emerging evidence suggests that several SLCs, including OCT2 and OATP1B1, contribute to clinically important phenotypes associated with those agents. Recent literature has indicated that modulation of SLC activity may result in drug-drug interactions, and genetic polymorphisms in SLC genes have been described that can affect the handling of substrates. Alteration of SLC function by either of these mechanisms has been demonstrated to contribute to interindividual variability in the pharmacokinetics and toxicity associated with several oncology drugs. In this report, we provide an update on this rapidly emerging field.

Introduction

The cellular membrane lipid bilayer is an effective and complex barrier that prevents the movement of many molecules, including drugs used in oncology, into cells. For this reason, evolutionary processes have developed membrane transport proteins to regulate movement of endogenous compounds that are required to maintain cellular function. This movement is commonly reliant on the action of a class of membrane proteins known as solute carriers (SLCs), which have received a great deal of attention due to accruing evidence that many drugs can accumulate inside cells by “hitchhiking” on these transporters. In fact, the contribution of transporter-mediated uptake of xenobiotics is now believed to be the predominant mechanism of intracellular accumulation (Dobson and Kell, 2008). Based on this principle, it is likely that many oncology drugs require specific transporter proteins to gain intracellular access in tumor cells to generate desired therapeutic effects, and thus that interindividual differences in the expression and/or function of transporters can contribute to variability in response to treatment. Similarly, SLCs also likely regulate accumulation of oncology drugs into normal healthy tissues and thereby directly contribute to drug-induced toxicity. Moreover, the ability of drugs to compete for the natural substrates of these transporters can potentially lead to altered cellular function and trigger unwanted adverse reactions. Considering that the clinical use of virtually all currently used oncology drugs is associated with toxic side effects that limit the dose that can be safely administered and, in some cases, cause life-threatening toxicities associated with organ damage, knowledge of specific transporters and the extent of oncology drug substrate specificity can theoretically contribute to the development of improved strategies or the design of cotherapies that ameliorate the incidence and/or severity of these effects.

Currently, there are still only a few reports demonstrating that specific SLCs can modulate cellular accumulation of oncology drugs (Supplemental Table 1), and most studies have been performed in cell-based model systems involving mammalian or amphibian cells that are engineered to overexpress a single or, at best, a limited number of transporters. The continual identification of oncology substrates for SLCs using these heterologous in vitro expression systems provides valuable information for predicting drug-drug and drug-protein interactions. However, the well recognized limitation of these preliminary determinations is that they do not indicate the true relevance of a transporter in handling a substrate in the context of whole-body disposition. Indeed, to attribute an abnormality in normal physiology or oncology drug-induced phenotypes to transporter perturbation, the relevance of the transporter to the tissue-specific distribution of a drug must first be determined in vivo. The availability of rodent models for many SLCs that have been associated with the transport of oncology drugs is now providing an opportunity to close the in vitro–in vivo knowledge gap. In the present article, we provide an overview of this rapidly emerging field for commonly used oncology drugs, emphasize recently explored translational approaches, and discuss strategies that can be used to avoid drug-induced damage to healthy tissues.

Platinum Chemotherapeutics

Cisplatin is among the most widely used chemotherapeutic drugs and has significantly improved outcome in various human malignancies, such as those in patients with head and neck (Chitapanarux et al., 2010), testicular (Nichols and Kollmannsberger, 2011), lung (Ardizzoni et al., 2007), and ovarian cancer (Matei et al., 2009). Additionally, use of oxaliplatin has led to significantly improved outcome in patients with advanced or metastatic colorectal cancer (Goldberg et al., 2004). Both platinum drugs share structural similarities and are thought to primarily exert their antitumor properties by their ability to form inter- or intrastrand cross-links with DNA (Zwelling et al., 1979). Unfortunately, use of these agents is limited by debilitating off-target effects that vary significantly in both severity and time of onset between individual patients. Patients treated with cisplatin are at a high risk of toxicities originating from severe renal tubular damage (up to 41% of patients; the major dose limiting toxicity) (de Jongh et al., 2003), as well as from irreversible bilateral hearing loss (22–70% of patients) (Ruggiero et al., 2013), and a chronic neurotoxicity characterized by numbness and tingling, paresthesia, reduced deep-tendon reflexes, and leg weakness (up to 50% of patients) (van den Bent et al., 2002). Despite structural similarities, the use of oxaliplatin is almost exclusively limited by the onset of peripheral neurotoxicity, which can manifest as an acute and/or a chronic form. The acute neurotoxicity occurs in up to 92% of patients but quickly resolves, whereas the chronic form is similar to that seen in patients treated with cisplatin (McWhinney et al., 2009). Although the mechanisms responsible for these toxicities remain poorly understood, accumulating evidence has shown that particular SLCs play dominant roles in facilitating movement of these agents into the affected organs.

Several decades ago, investigations revealed that cisplatin preferentially accumulates in human renal cortex against a concentration gradient and could competitively inhibit uptake of the cation tetraethylammonium in mouse kidney slices or basolateral membrane vesicles of the renal cortex (Weiner and Jacobs, 1983; Nelson et al., 1984; Williams and Hottendorf, 1985), as well as inhibit the renal clearance of organic cations from the basolateral site in isolated perfused rat kidneys (Miura et al., 1987). After further study, it was established that the organic cation transporter 2 (OCT2; encoded by the gene SLC22A2) is highly expressed at the basolateral membrane of proximal tubules (Fig. 1), and that both the murine and human homologs of this protein could mediate cellular uptake of cisplatin in vitro (Yokoo et al., 2007; Tanihara et al., 2009; Burger et al., 2010). Following these results, it was shown that OCT2 mediates cisplatin-induced renal damage in vivo using mice deficient in both Oct2 and the redundant murine transporter Oct1. These animals have decreased elimination of cisplatin via urinary excretion and are also protected from cisplatin-induced nephrotoxicity (Filipski et al., 2009). These findings were further expanded to show that patients with a reduced functional variant of OCT2 (c.808G>T) are protected from cisplatin-induced nephrotoxicity. This observation has since been replicated using an independent cohort of patients (Iwata et al., 2012). The implications of these findings have since generated promising clinical applications, given that expression of OCT2 appears to be absent in most tumors (Franke et al., 2010b; Sprowl et al., 2013a). Collectively, it therefore appears possible to decrease cisplatin-induced nephrotoxicity via pharmacological inhibition of OCT2 function without sacrificing the antitumor efficacy of cisplatin. In fact, promising studies have already been conducted demonstrating that concurrent administration of cisplatin with the OCT2 inhibitor cimetidine (Zhang et al., 1998; Urakami et al., 2002; Franke et al., 2010b) can offer at least partial protection against renal toxicity (Sleijfer et al., 1987; Ciarimboli et al., 2010) without altering antitumor efficacy in either murine models or humans (Katsuda et al., 2010; Sprowl et al., 2013b). Nonetheless, it is important to point out that cimetidine has an inhibition constant (Ki) that is comparable to OCT2 for two solute carriers expressed on the luminal membrane of renal tubular cells, the multidrug and toxin extrusion protein 1 (MATE1) and MATE2-K (Koepsell et al., 2007; Matsushima et al., 2009; Ohta et al., 2009; Tsuda et al., 2009). Although a genetic polymorphism of MATE1 (rs2289669 G>A) associated with reduced function has been reported to have no impact on cisplatin-induced nephrotoxicity in patients (Iwata et al., 2012), loss of Mate1 function in mice has led to increased cisplatin-induced renal tubular damage (Nakamura et al., 2010). Likewise, the 5-hydroxytryptamine-3 receptor antagonist ondansetron, which has been shown to inhibit the function of OCT2 and MATEs, can potentiate cisplatin-induced nephrotoxicity in mice by preferentially inhibiting Mate1 function and increasing accumulation of cisplatin into proximal tubular cells (Li et al., 2013). Regardless, the use of cimetidine currently remains a viable option as the previous in vivo studies indicate protective effects rather than potentiating toxicity, although the use of other inhibitors could be explored to circumvent this problem. In fact, the tyrosine kinase inhibitor (TKI) imatinib has been identified as a potent, noncompetitive inhibitor of OCT2, and its use appears to protect rats from cisplatin-induced nephrotoxicity by preventing platinum accumulation into the kidney (Tanihara et al., 2009; Franke et al., 2010b).

Fig. 1.

Fig. 1.

Solute carriers known to regulate accumulation of cisplatin or oxaliplatin into tissues associated with drug toxicity. The organic cation transporter OCT2 is expressed in the cochlea, dorsal root ganglia (DRG), and kidney, and mediates uptake while the multidrug and toxin extrusion proteins MATE1 and MATE2-K remove cisplatin and oxaliplatin from proximal tubule cells.

In addition to the connection of OCT2 with cisplatin-induced nephrotoxicity, there is current evidence to suggest that this transporter is also involved in the etiology of other toxicities associated with platinum drugs. For example, OCT2 is expressed in the inner ear, and mice lacking Oct1 and Oct2 or wild-type mice receiving cimetidine as a transporter inhibitor are protected from cisplatin-induced hearing loss (Ciarimboli et al., 2010). The clinical relevance of these findings remains currently unknown. Similarly, a recent study demonstrated that OCT2 is also a modulator of platinum-induced neurotoxicity, in particular, the neurotoxicity associated with oxaliplatin. In patients, oxaliplatin does not induce renal or ototoxic damage to a significant extent. However, like cisplatin, oxaliplatin accumulates at relatively high levels within dorsal root ganglia of the peripheral nervous system, which is the initial trigger leading to peripheral neuropathy (Holmes et al., 1998; Screnci et al., 2000; McDonald et al., 2005). Oxaliplatin is a substrate for OCT2, and this SLC was recently found to be expressed at high levels in both murine and human dorsal root ganglia (Yonezawa et al., 2006; Burger et al., 2010; Sprowl et al., 2013a). In fact, genetic or pharmacological inhibition of Oct2 function protects mice from acute oxaliplatin-induced neuropathy and, like cisplatin, decreases urinary elimination without altering systemic concentrations (Sprowl et al., 2013a). Further studies are currently underway to assess whether inhibition of OCT2 function will ameliorate chronic forms of cisplatin- and oxaliplatin-induced neuropathy, as well as to assess the clinical applications of this strategy. It is worth pointing out that probenecid, an organic anion that does not inhibit OCT2, can significantly reduce the tubular secretion of cisplatin without affecting glomerular filtration, and that this leads to reduced nephrotoxicity (Ross and Gale, 1979; Jacobs et al., 1984, 1991; Klein et al., 1991). Although the underlying mechanism for this interaction is unclear, it is conceivable that anionic platinum-sulfhydryl complexes formed from cisplatin in the γ-glutamyltranspeptidase pathway accumulate in kidney cells through an organic anion transport mechanism that is probenecid-sensitive, possibly organic anion transporter 1 (OAT1) or OAT3.

There are other SLCs that are possibly involved in side effects associated with platinum-based oncology drugs (Table 1). In particular, the copper transporter 1 (CTR1; SLC31A1) has been implicated in the cellular uptake of all three approved platinum compounds, namely, cisplatin, oxaliplatin, and carboplatin (Larson et al., 2009). Expression of CTR1 has been reported in dorsal root ganglia of rats (Liu et al., 2009), the basolateral membrane of the proximal tubular cells (Pabla et al., 2009), and murine cochlea, although generally the expression levels are low in comparison with OCT2 (Ciarimboli et al., 2010). In line with expression patterns, studies performed in mice have demonstrated that reduced function of CTR1 is associated with amelioration of cisplatin-induced cell death in proximal tubules (Pabla et al., 2009), and that administration of copper sulfate, a competitive inhibitor of CTR1, with cisplatin protects against cisplatin-induced ototoxicity (More et al., 2010). Moreover, cells within the dorsal root ganglia that express high levels of CTR1 appear to be more susceptible to atrophy by platinum agents (Liu et al., 2009; Ip et al., 2010). However, more recent findings indicate that CTR1 may not be an important contributor to the neurotoxicity induced by oxaliplatin (Liu et al., 2013), and despite the previous promising findings, a direct clinical demonstration of CTR1 involvement in modulating platinum-induced toxicity remains lacking. Moreover, recent experimental evidence has been generated that questions the ability of CTR1 to regulate cisplatin uptake (Ivy and Kaplan, 2013). Clearly, additional investigation into the role of CTR1 and platinum-induced toxicity is required. In this context, it should be noted that CTR1 is expressed in multiple solid tumors, and that pharmacological inhibition of this transporter will possibly impair antitumor efficacy (Larson et al., 2009).

TABLE 1.

Solute carriers associated with cellular accumulation of platinum chemotherapeutics

Drug [Alias] Mechanism of Action Principle Side Effects Transporters References
Carboplatin [Paraplatin] Platination; DNA cross-linking Nausea and vomiting SLC28A1 (CTR1)a (Larson et al., 2009)
Myelosuppression
Neurotoxicity
Electrolyte abnormalities
Hypersensitivity reaction
Cisplatin [Platinol] Platination; DNA cross-linking Nausea and vomiting SLC22A1 (OCT1)a,b SLC22A2 (OCT2)a,b,c SLC31A1 (CTR1)a,b SLC47A1 (MATE1)a,b (Yokoo et al., 2007; Filipski et al., 2009; Larson et al., 2009; Tanihara et al., 2009; Nakamura et al., 2010)
Myelosuppression
Neurotoxicity
Electrolyte abnormalities
Hypersensitivity reaction
Renal toxicity
Ototoxicity
Oxaliplatin [Eloxatin] Platination; DNA cross-linking Nausea and vomiting SLC22A1 (OCT1)aSLC22A2 (OCT2)a,b SLC22A3 (OCT3)a SLC22A4 (OCTN1)a SLC22A5 (OCTN2)a SLC31A1 (CTR1)a SLC47A1 (MATE1)a SLC47A2 (MATE2-K)a (Yonezawa et al., 2006; Yokoo et al., 2007; Larson et al., 2009; Burger et al., 2010; Sprowl et al., 2013a)
Myelosuppression
Neurotoxicity
Hypersensitivity reaction
Diarrhea
a

Evidence of drug uptake being regulated by solute carriers in vitro.

b

Evidence of drug uptake being regulated by solute carriers in vivo.

c

Evidence of drug uptake being regulated by solute carriers in patient populations.

Several other transporters require additional exploration with in vivo models, such as organic cation transporter novel type 1 (OCTN1) and OCTN2, which are expressed at detectable levels in the dorsal root ganglia and have been implicated in the transport of oxaliplatin in some (Jong et al., 2011) but not all studies (Yonezawa et al., 2005; Haschke et al., 2010; Lancaster et al., 2010). The main endogenous substrate of OCTN2 is carnitine (vitamin BT; β-hydroxy-γ-trimethylaminobutyric acid), a hydrophilic zwitterion that plays an essential role in the transfer of long-chain fatty acids inside mitochondria for β-oxidation, and as such is critical in the production of cellular energy (Longo et al., 2006). Carnitine is normally maintained at a steady level in the blood via a rescue mechanism that involves reabsorption in the kidney of filtered carnitine by the solute carrier OCTN2 (SLC22A5) (Ohtani et al., 1984). The existence of this reabsorption mechanism for carnitine, as well as the occurrence of mutations in OCTN2 that lead to primary systemic carnitine deficiency, strongly supports the critical physiologic importance of OCTN2 (Scaglia et al., 1998). Indeed, metabolic abnormalities similar to human primary systemic carnitine deficiency have been described in juvenile visceral steatosis mice (Koizumi et al., 1988) and demonstrated to be caused by a null mutation (1114T>G; L352R) in the mouse Octn2 gene, Slc22a5 (Lu et al., 1998). Furthermore, heterozygosity for OCTN2 mutations in humans is associated with an intermediate carnitine-deficiency phenotype, suggesting that even partial loss of transporter function may be detrimental. In this context, it is particularly noteworthy that, in a previous study involving a small cohort of adult patients, treatment with cisplatin was associated with a significant increase in the urinary excretion of carnitine, which eventually normalized about 7 days after discontinuing therapy (Heuberger et al., 1998). Similar observations have been made in patients undergoing combination chemotherapy with ifosfamide-doxorubicin-cisplatin (Hockenberry et al., 2009), paclitaxel-carboplatin, or vinorelbine-carboplatin (Mancinelli et al., 2007). In the case of cisplatin treatment, the increases in urinary carnitine were hypothesized to be due to inhibition of active tubular reabsorption of carnitine and acylcarnitines (Heuberger et al., 1998), and direct inhibition of OCTN2 has been reported for several oncology drugs, including actinomycin D (Ohashi et al., 1999), vinblastine (Diao et al., 2009), and etoposide (Hu et al., 2012). Unlike other agents previously linked with hypocarnitinemia, however, cisplatin-induced carnitine wasting is associated with an unusual regulatory mechanism likely involving deactivation of the peroxisome proliferator-activated receptor alpha (PPAR-α) transcription factor and subsequent downregulation of OCTN2 (Lancaster et al., 2010). Additional studies are clearly warranted to further understand the pharmacodynamic and toxicological implications of oncology drugs affecting the carnitine-wasting phenomenon.

Methotrexate

Methotrexate (MTX) is an antifolate agent commonly used in the treatment of acute lymphoblastic leukemia, non-Hodgkin lymphoma, and osteosarcoma that acts by inhibiting the synthesis of folic acid required for nucleotide synthesis and subsequent cellular replication. When given at high doses, the clinical use of MTX becomes limited by the onset of a number of side effects, including neurotoxicity, hepatotoxicity, nephrotoxicity, and gastrointestinal mucositis, all of which are associated with accumulation of MTX and subsequent damage to normal tissues. Due to the frequency at which MTX is used as a chemotherapeutic agent, there have been numerous investigations aimed at evaluating the role of SLCs in these toxicities.

Although multiple transporters have been identified in regulating the uptake of MTX into cells (Fig. 2; Table 2), its movement across cellular membranes is most commonly linked to the reduced folate carrier (RFC; SLC19A1). Considering that MTX is an antifolate, it is not surprising that cellular uptake is mediated by RFC, and that activity of RFC has been associated with treatment efficacy (Fischer, 1962; Gorlick et al., 1996; Sowers et al., 2011). Unfortunately, studies are limited in demonstrating a role of RFC in regulating MTX toxicity, although it is known that the gene encoding RFC is located on chromosome 21q22.3 and children with trisomy 21 are more susceptible to adverse events in response to therapy with MTX (Garre et al., 1987; Dixon et al., 1994; Williams and Flintoff, 1995; Yang-Feng et al., 1995; Buitenkamp et al., 2010). Whether these events are a result of a gene dosage effect leading to increased activity in RFC and cellular uptake of MTX has yet to be demonstrated. The c.80G>A variant, which is thought to increase function of RFC, has been associated with increased incidences of MTX-induced toxicity, including hepatotoxicity in patients with psoriasis (Campalani et al., 2007) and with a higher degree of bone marrow and liver toxicity in patients with acute lymphoblastic leukemia (Gregers et al., 2010). However, multiple studies have also reported that this variant has no effect on toxicity after MTX treatment (Kishi et al., 2003; Shimasaki et al., 2006; Huang et al., 2008; Chandran et al., 2010; Faganel Kotnik et al., 2010; Chiusolo et al., 2012), suggesting that additional comprehensive studies are required to definitively determine the role of RFC for this drug.

Fig. 2.

Fig. 2.

Solute carriers known to regulate accumulation of methotrexate into tissues associated with drug toxicity. The organic anion transporting polypeptides OATP1B1 and OATP1B3 are expressed in human hepatocytes. The organic anion transporters OAT1 and OAT2 are expressed at the basolateral membrane of proximal tubule cells within the kidney and are likely key regulators of methotrexate clearance and nephrotoxicity.

TABLE 2.

Solute carriers associated with cellular accumulation of methotrexate

Drug [Alias] Mechanism of Action Principle Side Effects Transporters References
Methotrexate [Trexall] Interferes with folate metabolism Myelosuppression SLC22A6 (OAT1)a SLC22A7 (OAT2)a SLC22A8 (OAT3)a,b SLC22A9 (OAT4)a SLCO1A2 (OATP1A2)a,b SLCO1B1 (OATP1B1)a,b,c SLCO1B3 (OATP1B3)a,b SLCO4C1 (OATP4C1)a SLC19A1 (RFC)a,c SLC46A1 (PCFT)a (Abe et al., 1999; Sun et al., 2001; Takeda et al., 2002; Mikkaichi et al., 2004; Mori et al., 2004; Badagnani et al., 2006; van de Steeg et al., 2009, 2013; Zhao et al., 2009; Sowers et al., 2011)
Mucositis
Dermatological toxicity (skin rash and/or nail changes)
Renal toxicity
Neurotoxicity
a

Evidence of drug uptake being regulated by solute carriers in vitro.

b

Evidence of drug uptake being regulated by solute carriers in vivo.

c

Evidence of drug uptake being regulated by solute carriers in patient populations.

Regardless of the importance of RFC, there are several other transporters that modulate cellular accumulation of MTX, including the organic anion transporters OAT1, OAT2, and OAT3 (Sun et al., 2001; Mori et al., 2004). Renal elimination is the major route of MTX clearance, and since OAT1 and OAT3 are both expressed along the basolateral membrane of proximal tubule cells, they are likely regulators of renal clearance and MTX-induced nephrotoxicity (Sweet et al., 1997, 2003). This notion is consistent with in vivo evidence showing that mice with an OAT3 deficiency have significantly delayed clearance of methotrexate (VanWert and Sweet, 2008), although a direct connection with toxicity remains uncertain. In addition to OAT1 and OAT3, several hepatic organic anion transporting polypeptides (OATPs), including OATP1B1 (SLCO1B1), OATP1B3 (SLCO1B3), and OATP1A2 (SLCO1A2) (Abe et al., 1999; Badagnani et al., 2006), are involved in cellular uptake of MTX. Of these transporters, OATP1A2 in the liver is expressed in cholangiocytes and not in hepatocytes (Lee et al., 2005), and the direct contribution of this transporter to the disposition of MTX is unclear (van de Steeg et al., 2013). Unlike OATP1B1 and OATP1B3, OATP1A2 is also expressed in the epithelial cells at the blood-brain barrier and on the distal tubule cells of the kidney, where it may contribute to MTX-related toxicities observed in the central nervous system and kidney (Bronger et al., 2005; Lee et al., 2005). Follow-up in vivo studies are required to address these hypotheses.

Among the transporters associated with MTX, the role of OATP1B1 in its disposition profile has been most clearly established. The first in vivo studies demonstrating a role of OATP1B1 in modulating MTX disposition were performed in humanized transgenic mice expressing human OATP1B1, and these animals have reduced systemic levels of MTX and increased liver-to-plasma ratios (van de Steeg et al., 2009, 2013). Interestingly, similar results were observed in mice overexpressing human OATP1B3. These findings are consistent with the notion that several inherited variants in these SLCs are associated with altered clearance of MTX in patients, as well as with increased gastrointestinal toxicity (Trevino et al., 2009; Lopez-Lopez et al., 2011). The association of the same variants with altered MTX clearance and toxicity has since been replicated in follow-up studies (Radtke et al., 2013; Ramsey et al., 2013). Additional rare variants of this SLC, including those with damaging function, have also been associated with further decreased clearance of MTX (Ramsey et al., 2012). However, since hepatobiliary excretion of MTX in humans is only a minor pathway of elimination, the overall contribution of OATP1B-transporter genotypes to explaining interindividual variability in MTX clearance is relatively modest (<15%). Therefore, the main contributing factors underlying genetic predisposition to low MTX clearance in humans are currently unclear, indicating that regular monitoring of MTX plasma concentrations remains warranted to prevent the occurrence of unwanted toxicities. Moreover, further studies are warranted to assess the physiologic roles of other transporters that have been associated with cellular uptake of methotrexate in vitro, such as OAT4, OATP4C1, and the proton-coupled folate transporter (PCFT) (Table 2) (Takeda et al., 2002; Mikkaichi et al., 2004; Zhao et al., 2009).

Nucleoside Analogs

Analogs of nucleosides have been used as a common therapeutic strategy for the treatment of various types of cancers. These agents include cytarabine (Ara-C), used primarily in the treatment of acute myeloid leukemia, as well as gemcitabine, used in the treatment of pancreatic cancer. The clinical use of Ara-C has been associated with severe adverse effects, some of which can prove fatal, including anemia, neutropenia, thrombocytopenia, renal impairment, stomatitis, and acute pulmonary syndrome (Reykdal et al., 1995; Cole and Gibson, 1997; Briasoulis and Pavlidis, 2001). Moreover, high-dose Ara-C treatment is known to initiate peripheral neuropathy, encephalopathy and cerebellar dysfunction, paraplegia, as well as ataxia, nystagmus, and dysarthrias, which typically occur within 3–7 days after treatment (Lazarus et al., 1981; Winkelman and Hines, 1983; Herzig et al., 1987). Gemcitabine has also been associated with numerous adverse events, including the onset of anemia, neutropenia, thrombocytopenia, pulmonary toxicity, as well as hepatic and renal impairment (Koczor et al., 2012).

Not surprisingly, expression of the human equilibrative nucleoside transporter 1 (ENT1; SLC29A1) is known to regulate cellular uptake of Ara-C. In fact, ENT1 is reported to regulate up to 80% of Ara-C uptake into target cells at plasma concentrations of 0.5–1 µM, which are readily achieved at even standard low-dose regimens (Sundaram et al., 2001; Clarke et al., 2002). Acute myeloid leukemia patients with blasts negative for ENT1 expression experience a shorter disease-free survival (7.7 vs. 13 months) and are at an 8-fold increased risk of early relapse (Galmarini et al., 2002). Moreover, high expression of ENT1 is associated with increased survival in patients with pancreatic ductal adenocarcinoma and cholangiocarcinoma treated with gemcitabine (Kobayashi et al., 2012; Marechal et al., 2012). In addition to associations with antitumor effects, coadministration of Ara-C with an ENT1 inhibitor has been shown to decrease toxicity in peripheral blood mononuclear cells (Parmar et al., 2011). ENT1 has low genetic variability with currently no known functional variants found in the coding region (Osato et al., 2003), although there is one particular variant located in the transcription factor binding region at −706G>C, which has an allelic frequency of 0.2 in Caucasians and is thought to increase mRNA expression in peripheral blood cells (Myers et al., 2006). Unfortunately, patients with this variant have only a modest 1.1-fold increase in ENT1 expression, and there appears to be no correlation with Ara-C–induced hematologic toxicity (Parmar et al., 2011).

ENT1 is also known to regulate cellular uptake of gemcitabine, and recent studies have identified another genetic variant of ENT1 (913C>T) as being associated with severe (grade 3–4) neutropenia in patients with locally advanced pancreatic cancer following treatment with gemcitabine (Tanaka et al., 2010). Interestingly, this result was not observed in an earlier study by the same investigators (Okazaki et al., 2010), indicating that further studies are required to confidently assess the role of ENT1 in modulating nucleoside analog safety in patients.

In addition to ENT1, there are other nucleoside transporters that have been implicated in regulating cellular uptake of nucleoside analogs as well as modulating response and the onset of adverse events (Table 3). The concentrating nucleoside transporter 1 (CNT1; SLC28A1) has been shown to mediate transport of gemcitabine and Ara-C (Graham et al., 2000). In fact, patients with an SLC28A1 c.1561G>A genetic variant who received gemcitabine experienced significantly higher rates of neutropenia and thrombocytopenia (Soo et al., 2009), which is likely due to an increased intracellular influx of gemcitabine in bone marrow progenitors. CNT3 has also been identified in regulating nucleoside analog uptake, although to a lesser degree compared with ENT1 (Ritzel et al., 2001). Nonetheless, it has been shown that 42.3% of patients with a certain CNT3 genetic variant experienced grade 3–4 neutropenia following treatment with gemcitabine, compared with only 25.7% of patients with the wild-type or heterozygous genotype (P = 0.035) (Tanaka et al., 2010). Unfortunately, considering the lack of replication in multiple studies as well as the lack of in vitro analysis to assess the impact that these variants have on nucleoside transporter function, there is still much work to be done in this field.

TABLE 3.

Solute carriers associated with accumulation of nucleoside analogs

Drug [Alias] Mechanism of Action Principle Side Effects Transporters References
Cytarabine [Cytosar; Ara-C, cytosine arabinoside] Incorporated into DNA; inhibits DNA polymerase; terminates DNA chain elongation Myelosuppresion SLC28A1 (CNT1)a SLC29A1 (ENT1)a (Graham et al., 2000; Sundaram et al., 2001)
Nausea and vomiting
Mucositis
Gastrointestinal (ulcers, pancreatitis)
Hyperuricemia
Dermatological toxicity (skin rash and/or nail changes)
Hand-foot syndrome
Peripheral edema
Conjunctivitis
Neurotoxicity
Hepatotoxicity
Gemcitabine [Gemzar] Incorporated into DNA; inhibits DNA polymerase; terminates DNA chain elongation; inhibits ribonucleotide reductase (prodrug) Myelosuppresion SLC28A1 (CNT1)a,b SLC29A1 (ENT1)a,b SLC28A3 (CNT3)a,b (Graham et al., 2000; Ritzel et al., 2001; Soo et al., 2009; Tanaka et al., 2010)
Nausea and vomiting
Mucositis
Diarrhea
Fever
Flu-like symptoms
Dermatological toxicity (skin rash and/or nail changes)
Neurotoxicity
Peripheral edema
Hepatotoxicity
Renal toxicity
a

Evidence of drug uptake being regulated by solute carriers in vitro.

b

Evidence of drug uptake being regulated by solute carriers in patient populations.

Taxanes

Paclitaxel, docetaxel, and cabazitaxel are members of the taxane family of agents used in the treatment of many malignancies, including breast, lung, ovarian, prostate, and gastric cancers. Taxanes interact with tubulin and prevent their dissociation, which in turn prevents chromosomal segregation, leading to a cell cycle arrest in G2/M and subsequent promotion of apoptosis (Ringel and Horwitz, 1991; Chazard et al., 1994). Clinical use of all three approved taxanes is limited by a relatively narrow therapeutic window that, when exceeded, can lead to the onset of various side effects, including peripheral neuropathy and severe bone marrow depression, in particular neutropenia. In fact, neutropenia is considered the major dose-limiting toxicity, and severe (grade 3 or 4) neutropenia occurs in up to 88% of patients who receive docetaxel (Rougier et al., 2000; Ishimoto et al., 2006), depending on the patient’s disease type. Due to their common use in the clinic, there have been many studies conducted to investigate potential transporters that can modulate uptake of taxanes. Some correlations of hematologic and neurologic toxicities with transporter expression or function have been made, although these studies have been mostly limited to members of the ATP-binding cassette protein family (Sissung et al., 2006, 2008).

Despite intense efforts, there have been no direct correlations made between SLCs and the incidence and severity of neurologic toxicities, which demonstrates that this field remains in its infant stages. Nonetheless, strong evidence exists to demonstrate that specific SLCs can indirectly influence drug-induced hematologic toxicities by modulating the systemic clearance of taxanes. Both paclitaxel and docetaxel have been identified as a substrate of OAT2 (SLC22A7), OATP1B1, and OATP1B3 (Fig. 3; Table 4) (Smith et al., 2005; Franke et al., 2010a; Svoboda et al., 2011; van de Steeg et al., 2011; de Graan et al., 2012). Studies performed in transfected cells have shown that paclitaxel and docetaxel are also transported substrates of the functionally related mouse transporter Oatp1b2 (Slco1b2), but this was not noted under the same experimental conditions for cabazitaxel. The lack of transport of cabazitaxel by Oatp1b2 is somewhat surprising considering its structural similarity with docetaxel, both having a 10-deacetylbaccatin III backbone (Vrignaud et al., 2013), and because cabazitaxel has been reported to inhibit OATP1B1 and OATP1B3, albeit at relatively high concentrations (Jevtana prescribing information; see http://products.sanofi.us/jevtana/jevtana.pdf).

Fig. 3.

Fig. 3.

Solute carriers known to regulate accumulation of taxanes into tissues associated with drug accumulation. The organic anion transporting polypeptides OATP1B1 and OATP1B3 are expressed in human hepatocytes. OATP1B1 and OATP1B3 modulate uptake of paclitaxel and docetaxel and loss of function increases systemic exposure, which can lead to induction of devastating adverse events. The organic anion transporter OAT2 is also expressed in human hepatocytes and loss of function or expression can also lead to increased systemic exposure of taxanes.

TABLE 4.

Solute carriers associated with cellular accumulation of taxanes

Drug [Alias] Mechanism of Action Principle Side Effects Transporters References
Docetaxel [Taxotere] Microtubule depolarization inhibitor Myelosuppression SLC22A7 (OAT2)a,b SLCO1B1 (OATP1B1)a,b,c SLCO1B3 (OATP1B3)a,b,c (Franke et al., 2010a; de Graan et al., 2012)
Hypersensitivity reaction
Alopecia
Neurotoxicity
Dermatological toxicity (skin rash and/or nail changes)
Mucositis
Paclitaxel [Taxol; Abraxane] Microtubule depolarization inhibitor Myelosuppression SLC22A7 (OAT2)a SLCO1B1 (OATP1B1)a,b SLCO1B3 (OATP1B3)a,b (Smith et al., 2005; Svoboda et al., 2011; van de Steeg et al., 2011)
Hypersensitivity reaction
Alopecia
Neurotoxicity
Dermatological toxicity (skin rash and/or nail changes)
Mucositis
Cardiotoxicity
a

Evidence of drug uptake being regulated by solute carriers in vitro.

b

Evidence of drug uptake being regulated by solute carriers in vivo.

c

Evidence of drug uptake being regulated by solute carriers in patient populations.

Since hepatic metabolism is the dominant elimination mechanism of docetaxel (mediated by CYP3A4) and paclitaxel (mediated by CYP2C8 and CYP3A4), it is not surprising that OATP1B-type transporters, which are exclusively expressed at the basolateral membrane of hepatocytes, were identified as SLCs regulating their liver uptake. So far, only one study has identified OATP1B1 as a transporter of paclitaxel in which it was demonstrated that SLC overexpressing Xenopus laevis oocytes accumulated higher levels of paclitaxel, and that OATP1B1-overexpressing SKOV-3 cells were more sensitive to paclitaxel-induced cytotoxic effects compared with cells transfected with an empty control vector (Svoboda et al., 2011). However, these findings are in conflict with other reports in which uptake of paclitaxel was evaluated in similar heterologous expression models. Studies performed in transgenic mice expressing OATP1B1 in hepatocytes have confirmed the relatively minor contribution of this SLC to the disposition of paclitaxel, and these findings collectively support the possibility that OATP1B1 is a high-affinity transporter that becomes saturated at plasma concentrations associated with even relatively low doses of the drug (Smith et al., 2005; van de Steeg et al., 2011). Similar discrepancies in OATP1B1 transporter affinity have been reported for docetaxel (Baker et al., 2009; de Graan et al., 2012).

Based on recently performed in vitro uptake studies (de Graan et al., 2012), multiple functionally different haplotypes, including OATP1B1*5 and OATP1B1*15, were found to have a detrimental impact on docetaxel transport. This finding is consistent with previous studies showing substantially diminished transport activity of several OATP1B1 substrates by these particular variants when transfected into mammalian cells (Kameyama et al., 2005), and these variants have also been associated with altered systemic exposure and toxicity in response to multiple substrate drugs (Niemi et al., 2011). Interestingly, the relevance of these genetic variants in OATP1B1 could not be confirmed in a pharmacogenetic-association study performed in Caucasian patients (de Graan et al., 2012). It is possible that additional rare genetic variants or haplotypes in OATP1B1 of importance to the transport docetaxel in this population are yet to be discovered, and that much larger numbers of patients are then needed to more precisely quantify genotype-phenotype associations. In the same study, several genetic variants in OATP1B3 were also not significantly associated with the pharmacokinetics of docetaxel. This is in line with an earlier report (Baker et al., 2009), although the findings are at odds with several other investigations performed in patients of Asian descent. For example, homozygosity (GG) for rs11045585 was associated with reduced clearance of docetaxel, compared with patients carrying the AA or AG genotypes (Chew et al., 2011). In another study, a particular OATP1B3 genotype combination comprising the reference allele at IVS4+76G>A (rs4149118) and variant alleles at c.699G>A (rs7311358), IVS12+5676A>G (rs11045585), and *347-*348insA (rs3834935) was also linked with reduced clearance of docetaxel (Chew et al., 2012). It is possible that differences in outcome with the study by de Graan et al. are associated with the fact that such variants may occur at different frequencies between Asians and Caucasians, and/or on different, ethnicity-dependent haplotype structures.

Regardless of any potential ethnic considerations, the existence of at least two potentially redundant uptake transporters in the human liver with similar affinity for agents such as paclitaxel and docetaxel supports the possibility that functional defects in both of these proteins may be required to confer substantially altered disposition phenotypes such as those seen in Oatp1b-deficient mice. Although complete functional deficiency of either OATP1B1 or OATP1B3 has been recorded to occur (Kim et al., 2007), deficiency of both transporters is extremely rare, with an estimated frequency in the human population of about 1 in a million (van de Steeg et al., 2012). It can thus be postulated that intrinsic physiologic and environmental variables influencing OATP1B1- or OATP1B3-mediated uptake of taxanes into hepatocytes may have a more profound influence on clearance than do defective genetic variants. This recognition is particularly relevant in the context of the recent guidelines offered by The International Transporter Consortium regarding preclinical criteria needed to trigger the conduct of clinical studies to evaluate drug-transporter interactions (Giacomini et al., 2010). Indeed, it is conceivable that instances of idiosyncratic hypersensitivity to taxanes are the result of currently unrecognized drug-drug interactions at the level of hepatocellular uptake mechanisms.

To add further to the complexity of the disposition properties of taxanes, the solubilizers Kolliphor EL (formerly Cremophor EL; used for paclitaxel) and polysorbate 80 (used for docetaxel and cabazitaxel) were recently found to strongly inhibit the uptake of OATP1B1 and OATP1B3 substrates into cells overexpressing the transporters (de Graan et al., 2012; Engel et al., 2012). Although further investigation is required to confirm direct involvement of solubilizer-mediated inhibition of OATP1B1 and OATP1B3 as the primary mechanistic basis for the apparent lack of observed in vivo effects of genetic variants, it is of note that similarly altered, solubilizer-dependent hepatic uptake has also been described for colchicine in the presence of Solutol HS15 (Bravo Gonzalez et al., 2004). These observations suggest that the impact of reduced-function variants of OATP1B1 and OATP1B3 on the clearance of taxanes may be much more pronounced for solubilizer-free formulations of the drugs, such as nab-paclitaxel (ABI-007; Abraxane) and nab-docetaxel (ABI-008).

In this context, it is of interest to point out that OATP1B-type transporters, in particular OATP1B3, are expressed at relatively high levels in tumors of the colon, endometrium, esophagus, lung, prostate, stomach, testis, and bladder (Lancaster et al., 2013), and that both OATP1B1 and OATP1B3 contribute to the in vitro cytotoxicity of taxanes in ovarian cancer cells (Svoboda et al., 2011). Although systemic exposure to free (unbound) taxanes is believed to be the dominant driver of drug-induced cytotoxicity at tumor sites (Ait-Oudhia et al., 2012), it is conceivable that these transporters can contribute directly to tumoral drug uptake, and that this process can be inhibited by solubilizers such as Kolliphor EL and polysorbate 80, leading to diminished antitumor activity. This possibility would be consistent with available clinical data on the comparative efficacy of the various paclitaxel formulations (reviewed in Ait-Oudhia et al., 2012), and with preclinical findings suggesting that the absorption rate constant of paclitaxel uptake into tumors is dramatically decreased in the presence of Kolliphor EL compared with nab-paclitaxel (Desai et al., 2006), and that a tumor-delivery mechanism exists for nab-paclitaxel that is independent of secreted protein acidic and rich in cysteine (SPARC) (Neesse et al., 2013), a matricellular glycoprotein produced by tumors and/or neighboring stroma that facilitates the intracellular accumulation of intact albumin nanoparticles (Yardley, 2013).

In addition to the OATP1B-transporter family, docetaxel and paclitaxel are known substrates of the organic anion transporter OAT2 (Kobayashi et al., 2005; Baker et al., 2009), which may also play a role in hepatic uptake of these drugs. Using a rat model, castration has been shown to increase OAT2 expression in the liver and increase hepatic uptake of docetaxel, leading to an increased elimination. This observation is in line with findings in human patients with prostate cancer, where chemical or surgical castration is associated with dramatically increased clearance of docetaxel by a mechanism that is unrelated to altered activity of hepatic CYP3A4 (Franke et al., 2010a). This suggests that the decreased incidence of neutropenia in castrated patients with prostate cancer following docetaxel-based chemotherapy (Petrylak et al., 2004; Tannock et al., 2004; Hussain et al., 2005; Taplin et al., 2006; Rathkopf et al., 2008) is possibly due to increased hepatic uptake of docetaxel, leading to a decrease in systemic exposure. This strongly suggests that castrated patients with prostate cancer undergoing docetaxel treatment should be capable of tolerating elevated doses to offset their intrinsically increased ability to clear the drug.

Tyrosine Kinase Inhibitors

There are currently 22 different TKIs marketed for the treatment of various cancer types. One of the key challenges with TKIs in the field of oncology involves the identification and pairing of the right patient with the right drug. However, even if the right drug is identified, important issues still remain regarding its optimal dose. Despite the revolutionary changes in the various stages of drug development used for TKIs (Humphrey et al., 2011), the standard strategy still in use for dose selection is to establish a therapeutic dose in phase II trials and subsequently, at best, modify it for individual differences in body surface area in children. This is an unsustainable situation since failure to deliver the right dose in routine practice may have detrimental implications for both the efficacy and safety of TKIs (Di Gion et al., 2011; Scheffler et al., 2011). Indeed, a wealth of clinical data have become available over the last few years indicating that clinical outcome, in particular decreased toxicity, can be optimized if dosing strategies for TKIs take into consideration their unique pharmacokinetic profiles. Moreover, although TKIs possibly offer a number of important theoretical advantages over conventional cytotoxic chemotherapy, they are still afflicted by some of the same problems, including an extensive interindividual pharmacokinetic variability and the existence of a rather narrow therapeutic window (Baker and Hu, 2009). Although many of these targeting agents have been identified as substrates of various ATP-binding cassette protein transporters, the identity of SLCs that regulate absorption and disposition, as well as their uptake into tumor cells, remains essentially unknown.

Model-based simulations have indicated that a critical determinant of pharmacokinetic variability observed with TKIs is associated with absorption-related processes and to a lesser extent with clearance (Dai et al., 2008), and this has led to several studies exploring associations of intestinal SLCs with transport of TKIs. For example, the TKI imatinib has been identified as a substrate of OATP1A2 (SLCO1A2), OCTN1, and OCTN2, which are expressed in duodenal enterocytes (Glaeser et al., 2007; Hu et al., 2008), using heterologous expression models. Accordingly, investigators have reported preliminary results indicating that genetic variants in OATP1A2 are associated with altered pharmacokinetics of imatinib in cancer patients (Yamakawa et al., 2011) in a manner that is possibly dependent on racial ancestry (Eechoute et al., 2011). Recent reports have also been promising in regards to OCTN1 and OCTN2 in that genetic polymorphisms of these genes were associated with time to progression in patients with unresectable gastrointestinal stromal tumors treated with imatinib, as well as with a reduction in transcripts of the breakpoint cluster - Abelson (Bcr-Abl) fusion product (Angelini et al., 2013a,b), the main kinase target of imatinib.

There have been many studies claiming that OCT1 activity is associated with the accumulation of imatinib in leukemia cells and with responsiveness to treatment (Thomas et al., 2004; Crossman et al., 2005; Clark et al., 2008; Wang et al., 2008), although imatinib is a rather weak substrate of OCT1, and the causal connection of this transporter with the reported phenotypes remains uncertain (Hu et al., 2008; Wang et al., 2008). Nonetheless, since imatinib has high oral bioavailability and undergoes extensive hepatic metabolism, it is tempting to speculate that OCT1 is possibly involved in regulating hepatocellular entry of the drug. Since multiple TKIs, including imatinib, are transported substrates of OATP1B1 and/or OATP1B3 (Hu et al., 2008; Zimmerman et al., 2013), it is possible that multiple, potentially redundant SLCs are involved in the liver uptake process. Preliminary evidence supporting this thesis comes from a study in which genetic variation in OATP1B3 was shown to be linked with altered concentrations of imatinib in circulating cells as well as with altered oral clearance of the drug (Nambu et al., 2011). However, this same variant was not associated with response to treatment (Bedewy and El-Maghraby, 2013), and there is no evidence that this transporter is connected with imatinib-induced adverse events that would be either directly associated with uptake in a target tissue or indirectly through an effect on systemic drug levels.

Although clinically actionable data indicating that SLCs can regulate trafficking of TKIs and modulate toxicity or response remain lacking, some recent studies have revealed that TKIs can inhibit the activity of some transporters that are relevant to maintain homeostasis of natural endogenous substrates, via an off-target mechanism. For example, clinical safety profile investigations of pazopanib have determined that this drug causes an elevation in total bilirubin levels in patients without any evidence of direct liver damage. Since OATP1B1 is an important transporter that regulates levels of bilirubin (Johnson et al., 2009), the hyperbilirubinemia occurring in patients receiving pazopanib could be an artifact related to the TKI’s inhibitory effects on OATP1B1 rather than represent liver injury (Xu et al., 2010). Another example of an interaction of endogenous SLC substrates with drugs is demonstrated by the ability of TKIs to inhibit OCT2, which facilitates transport of serum creatinine into the renal tubule (Ciarimboli et al., 2012). Serum creatinine is used in routine practice to estimate glomerular filtration rate to assess kidney function, and since creatinine undergoes tubular secretion, this approach can overestimate glomerular filtration rate by as much as 20% (Breyer and Qi, 2010). Previous studies have indicated that competitive substrates of OCT2 can artificially increase serum creatinine, which could be mistakenly interpreted as evidence of renal injury (Dollery, 2013). A similar observation has been made with vandetanib, a TKI that markedly increases serum creatinine due to an inhibitory effect on OCT2-mediated transport of creatinine (Caprelsa prescribing information; see http://www1.astrazeneca-us.com/pi/caprelsa.pdf). The inhibition of OCT2 by vandetanib has also been proposed as a principle mechanism by which the TKI can increase the systemic exposure to cisplatin (Blackhall et al., 2010). Given that many TKIs can potently inhibit OCT2 activity, this event is likely not restricted to vandetanib (Morrow et al., 2010; Minematsu and Giacomini, 2011). Moreover, vandetanib also potently inhibits MATE1 and MATE2-K, further demonstrating that the function of multiple SLCs can be affected by off-target effects of this clinically important drug class (Shen et al., 2013).

Conclusion and Future of Field

The SLC transporters have an established role in regulating the pharmacokinetic profile of several oncology drugs, as well as in mediating the intrinsic cellular sensitivity of both healthy organs and malignant cells. Several polymorphic variants in SLC genes relevant to oncology drugs have been described in recent years, of which some, including those of the OATP1B-type transporter genes, may be of particular relevance to treatment outcome in patients with cancer. The actual in vivo effects of many SLCs in relation to their phenotypical consequences are still debatable, as contradictory results have been reported. Most studies published to date suffer from a lack of clinical evidence, or from small sample sizes in relation to the detected phenotypes in vivo, as well as from a host of potentially confounding factors that influence their outcome. Most important among these are environmental and physiologic factors that may affect expression or function of the SLCs, and confounding links to other genes or variants of putative relevance for drug absorption and/or disposition pathways. Additional research is clearly needed on the role of SLC in the handling of oncology drugs and how to use transporters to improve treatment by minimizing toxicity. However, considering the sheer magnitude of the number of SLCs identified thus far, it is evident that alternative strategies are required to identify additional, potentially important oncology drug-SLC pairs. Such future studies will hopefully clarify existing discrepant findings, and contribute to improving the quality of life and treatment outcome of a wide variety of malignant diseases.

Supplementary Material

Data Supplement

Abbreviations

Ara-C

cytarabine

CNT1/3

concentrating nucleoside transporter 1 or 3

CTR1

copper transporter 1

ENT1

equilibrative nucleoside transporter 1

MATE1/2-K

multidrug toxin and extrusion protein 1 or 2-K

MTX

methotrexate

OAT1/2/3

organic anion transporter 1, 2, or 3

OATP1B1/2/3

organic anion transporting polypeptides family 1B1, 2, or 3

OCT1/2

organic cation transporter 1 or 2

OCTN1/2

organic cation transporter novel type 1 or 2

RFC

reduced folate carrier

SLC

solute carrier

TKI

tyrosine kinase inhibitor

Author Contributions

Wrote or contributed to the writing of the manuscript: Sprowl, Sparreboom.

Footnotes

The preparation of this manuscript was supported in part by the American Lebanese Syrian Associated Charities and the National Institutes of Health [grant R01-CA151633].

Inline graphicThis article has supplemental material available at dmd.aspetjournals.org.

References

  1. Abe T, Kakyo M, Tokui T, Nakagomi R, Nishio T, Nakai D, Nomura H, Unno M, Suzuki M, Naitoh T, et al. (1999) Identification of a novel gene family encoding human liver-specific organic anion transporter LST-1. J Biol Chem 274:17159–17163 [DOI] [PubMed] [Google Scholar]
  2. Ait-Oudhia S, Straubinger RM, Mager DE. (2012) Meta-analysis of nanoparticulate paclitaxel delivery system pharmacokinetics and model prediction of associated neutropenia. Pharm Res 29:2833–2844 [DOI] [PubMed] [Google Scholar]
  3. Angelini S, Pantaleo MA, Ravegnini G, Zenesini C, Cavrini G, Nannini M, Fumagalli E, Palassini E, Saponara M, Di Battista M, et al. (2013a) Polymorphisms in OCTN1 and OCTN2 transporters genes are associated with prolonged time to progression in unresectable gastrointestinal stromal tumours treated with imatinib therapy. Pharmacol Res 68:1–6 [DOI] [PubMed] [Google Scholar]
  4. Angelini S, Soverini S, Ravegnini G, Barnett M, Turrini E, Thornquist M, Pane F, Hughes TP, White DL, Radich J, et al. (2013b) Association between imatinib transporters and metabolizing enzymes genotype and response in newly diagnosed chronic myeloid leukemia patients receiving imatinib therapy. Haematologica 98:193–200 [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Ardizzoni A, Boni L, Tiseo M, Fossella FV, Schiller JH, Paesmans M, Radosavljevic D, Paccagnella A, Zatloukal P, Mazzanti P, et al. CISCA (CISplatin versus CArboplatin) Meta-analysis Group (2007) Cisplatin- versus carboplatin-based chemotherapy in first-line treatment of advanced non-small-cell lung cancer: an individual patient data meta-analysis. J Natl Cancer Inst 99:847–857 [DOI] [PubMed] [Google Scholar]
  6. Badagnani I, Castro RA, Taylor TR, Brett CM, Huang CC, Stryke D, Kawamoto M, Johns SJ, Ferrin TE, Carlson EJ, et al. (2006) Interaction of methotrexate with organic-anion transporting polypeptide 1A2 and its genetic variants. J Pharmacol Exp Ther 318:521–529 [DOI] [PubMed] [Google Scholar]
  7. Baker SD, Hu S. (2009) Pharmacokinetic considerations for new targeted therapies. Clin Pharmacol Ther 85:208–211 [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Baker SD, Verweij J, Cusatis GA, van Schaik RH, Marsh S, Orwick SJ, Franke RM, Hu S, Schuetz EG, Lamba V, et al. (2009) Pharmacogenetic pathway analysis of docetaxel elimination. Clin Pharmacol Ther 85:155–163 [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Bedewy AM, El-Maghraby SM. (2013) Do SLCO1B3 (T334G) and CYP3A5*3 polymorphisms affect response in Egyptian chronic myeloid leukemia patients receiving imatinib therapy? Hematology 18:211–216 [DOI] [PubMed] [Google Scholar]
  10. Blackhall FH, O’brien M, Schmid P, Nicolson M, Taylor P, Milenkova T, Kennedy SJ, Thatcher N. (2010) A phase I study of Vandetanib in combination with vinorelbine/cisplatin or gemcitabine/cisplatin as first-line treatment for advanced non-small cell lung cancer. J Thorac Oncol 5:1285–1288 [DOI] [PubMed] [Google Scholar]
  11. Bravo González RC, Boess F, Durr E, Schaub N, Bittner B. (2004) In vitro investigation on the impact of Solutol HS 15 on the uptake of colchicine into rat hepatocytes. Int J Pharm 279:27–31 [DOI] [PubMed] [Google Scholar]
  12. Breyer MD, Qi Z. (2010) Better nephrology for mice—and man. Kidney Int 77:487–489 [DOI] [PubMed] [Google Scholar]
  13. Briasoulis E, Pavlidis N. (2001) Noncardiogenic pulmonary edema: an unusual and serious complication of anticancer therapy. Oncologist 6:153–161 [DOI] [PubMed] [Google Scholar]
  14. Bronger H, König J, Kopplow K, Steiner HH, Ahmadi R, Herold-Mende C, Keppler D, Nies AT. (2005) ABCC drug efflux pumps and organic anion uptake transporters in human gliomas and the blood-tumor barrier. Cancer Res 65:11419–11428 [DOI] [PubMed] [Google Scholar]
  15. Buitenkamp TD, Mathôt RA, de Haas V, Pieters R, Zwaan CM. (2010) Methotrexate-induced side effects are not due to differences in pharmacokinetics in children with Down syndrome and acute lymphoblastic leukemia. Haematologica 95:1106–1113 [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Burger H, Zoumaro-Djayoon A, Boersma AW, Helleman J, Berns EM, Mathijssen RH, Loos WJ, Wiemer EA. (2010) Differential transport of platinum compounds by the human organic cation transporter hOCT2 (hSLC22A2). Br J Pharmacol 159:898–908 [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Campalani E, Arenas M, Marinaki AM, Lewis CM, Barker JN, Smith CH. (2007) Polymorphisms in folate, pyrimidine, and purine metabolism are associated with efficacy and toxicity of methotrexate in psoriasis. J Invest Dermatol 127:1860–1867 [DOI] [PubMed] [Google Scholar]
  18. Chandran V, Siannis F, Rahman P, Pellett FJ, Farewell VT, Gladman DD. (2010) Folate pathway enzyme gene polymorphisms and the efficacy and toxicity of methotrexate in psoriatic arthritis. J Rheumatol 37:1508–1512 [DOI] [PubMed] [Google Scholar]
  19. Chazard M, Pellae-Cosset B, Garet F, Soares JA, Lucidi B, Lavail Y, Lenaz L. (1994) [Taxol (paclitaxel), first molecule of a new class of cytotoxic agents: taxanes]. Bull Cancer 81:173–181 [PubMed] [Google Scholar]
  20. Chew SC, Sandanaraj E, Singh O, Chen X, Tan EH, Lim WT, Lee EJ, Chowbay B. (2012) Influence of SLCO1B3 haplotype-tag SNPs on docetaxel disposition in Chinese nasopharyngeal cancer patients. Br J Clin Pharmacol 73:606–618 [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Chew SC, Singh O, Chen X, Ramasamy RD, Kulkarni T, Lee EJ, Tan EH, Lim WT, Chowbay B. (2011) The effects of CYP3A4, CYP3A5, ABCB1, ABCC2, ABCG2 and SLCO1B3 single nucleotide polymorphisms on the pharmacokinetics and pharmacodynamics of docetaxel in nasopharyngeal carcinoma patients. Cancer Chemother Pharmacol 67:1471–1478 [DOI] [PubMed] [Google Scholar]
  22. Chitapanarux I, Tharavichitkul E, Lorvidhaya V, Sittitrai P, Pattarasakulchai T. (2010) Induction chemotherapy with paclitaxel, ifosfamide, and cisplatin followed by concurrent chemoradiotherapy for unresectable locally advanced head and neck cancer. Biomed Imaging Interv J 6:e23. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Chiusolo P, Giammarco S, Bellesi S, Metafuni E, Piccirillo N, De Ritis D, Marietti S, Federica S, Laurenti L, Fianchi L, et al. (2012) The role of MTHFR and RFC1 polymorphisms on toxicity and outcome of adult patients with hematological malignancies treated with high-dose methotrexate followed by leucovorin rescue. Cancer Chemother Pharmacol 69:691–696 [DOI] [PubMed] [Google Scholar]
  24. Ciarimboli G, Deuster D, Knief A, Sperling M, Holtkamp M, Edemir B, Pavenstädt H, Lanvers-Kaminsky C, am Zehnhoff-Dinnesen A, Schinkel AH, et al. (2010) Organic cation transporter 2 mediates cisplatin-induced oto- and nephrotoxicity and is a target for protective interventions. Am J Pathol 176:1169–1180 [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Ciarimboli G, Lancaster CS, Schlatter E, Franke RM, Sprowl JA, Pavenstädt H, Massmann V, Guckel D, Mathijssen RH, Yang W, et al. (2012) Proximal tubular secretion of creatinine by organic cation transporter OCT2 in cancer patients. Clin Cancer Res 18:1101–1108 [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Clark RE, Davies A, Pirmohamed M, Giannoudis A. (2008) Pharmacologic markers and predictors of responses to imatinib therapy in patients with chronic myeloid leukemia. Leuk Lymphoma 49:639–642 [DOI] [PubMed] [Google Scholar]
  27. Clarke ML, Mackey JR, Baldwin SA, Young JD, Cass CE. (2002) The role of membrane transporters in cellular resistance to anticancer nucleoside drugs. Cancer Treat Res 112:27–47 [DOI] [PubMed] [Google Scholar]
  28. Cole N, Gibson BE. (1997) High-dose cytosine arabinoside in the treatment of acute myeloid leukaemia. Blood Rev 11:39–45 [DOI] [PubMed] [Google Scholar]
  29. Crossman LC, Druker BJ, Deininger MW, Pirmohamed M, Wang L, Clark RE. (2005) hOCT 1 and resistance to imatinib. Blood 106:1133–1134, author reply 1134 [DOI] [PubMed] [Google Scholar]
  30. Dai G, Pfister M, Blackwood-Chirchir A, Roy A. (2008) Importance of characterizing determinants of variability in exposure: application to dasatinib in subjects with chronic myeloid leukemia. J Clin Pharmacol 48:1254–1269 [DOI] [PubMed] [Google Scholar]
  31. de Graan AJ, Lancaster CS, Obaidat A, Hagenbuch B, Elens L, Friberg LE, de Bruijn P, Hu S, Gibson AA, Bruun GH, et al. (2012) Influence of polymorphic OATP1B-type carriers on the disposition of docetaxel. Clin Cancer Res 18:4433–4440 [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. de Jongh FE, van Veen RN, Veltman SJ, de Wit R, van der Burg ME, van den Bent MJ, Planting AS, Graveland WJ, Stoter G, Verweij J. (2003) Weekly high-dose cisplatin is a feasible treatment option: analysis on prognostic factors for toxicity in 400 patients. Br J Cancer 88:1199–1206 [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Desai N, Trieu V, Yao Z, Louie L, Ci S, Yang A, Tao C, De T, Beals B, Dykes D, et al. (2006) Increased antitumor activity, intratumor paclitaxel concentrations, and endothelial cell transport of cremophor-free, albumin-bound paclitaxel, ABI-007, compared with cremophor-based paclitaxel. Clin Cancer Res 12:1317–1324 [DOI] [PubMed] [Google Scholar]
  34. Diao L, Ekins S, Polli JE. (2009) Novel inhibitors of human organic cation/carnitine transporter (hOCTN2) via computational modeling and in vitro testing. Pharm Res 26:1890–1900 [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Di Gion P, Kanefendt F, Lindauer A, Scheffler M, Doroshyenko O, Fuhr U, Wolf J, Jaehde U. (2011) Clinical pharmacokinetics of tyrosine kinase inhibitors: focus on pyrimidines, pyridines and pyrroles. Clin Pharmacokinet 50:551–603 [DOI] [PubMed] [Google Scholar]
  36. Dixon KH, Lanpher BC, Chiu J, Kelley K, Cowan KH. (1994) A novel cDNA restores reduced folate carrier activity and methotrexate sensitivity to transport deficient cells. J Biol Chem 269:17–20 [PubMed] [Google Scholar]
  37. Dobson PD, Kell DB. (2008) Carrier-mediated cellular uptake of pharmaceutical drugs: an exception or the rule? Nat Rev Drug Discov 7:205–220 [DOI] [PubMed] [Google Scholar]
  38. Dollery CT. (2013) Intracellular drug concentrations. Clin Pharmacol Ther 93:263–266 [DOI] [PubMed] [Google Scholar]
  39. Eechoute K, Sparreboom A, Burger H, Franke RM, Schiavon G, Verweij J, Loos WJ, Wiemer EA, Mathijssen RH. (2011) Drug transporters and imatinib treatment: implications for clinical practice. Clin Cancer Res 17:406–415 [DOI] [PubMed] [Google Scholar]
  40. Engel A, Oswald S, Siegmund W, Keiser M. (2012) Pharmaceutical excipients influence the function of human uptake transporting proteins. Mol Pharm 9:2577–2581 [DOI] [PubMed] [Google Scholar]
  41. Faganel Kotnik B, Dolžan V, Grabnar I, Jazbec J. (2010) Relationship of the reduced folate carrier gene polymorphism G80A to methotrexate plasma concentration, toxicity, and disease outcome in childhood acute lymphoblastic leukemia. Leuk Lymphoma 51:724–726 [DOI] [PubMed] [Google Scholar]
  42. Filipski KK, Mathijssen RH, Mikkelsen TS, Schinkel AH, Sparreboom A. (2009) Contribution of organic cation transporter 2 (OCT2) to cisplatin-induced nephrotoxicity. Clin Pharmacol Ther 86:396–402 [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Fischer GA. (1962) Detective transport of amethopterin (methotrexate) as a mechanism of resistance to the antimetabolite in L5178Y leukemic cells. Biochem Pharmacol 11:1233–1234 [DOI] [PubMed] [Google Scholar]
  44. Franke RM, Carducci MA, Rudek MA, Baker SD, Sparreboom A. (2010a) Castration-dependent pharmacokinetics of docetaxel in patients with prostate cancer. J Clin Oncol 28:4562–4567 [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Franke RM, Kosloske AM, Lancaster CS, Filipski KK, Hu C, Zolk O, Mathijssen RH, Sparreboom A. (2010b) Influence of Oct1/Oct2-deficiency on cisplatin-induced changes in urinary N-acetyl-beta-D-glucosaminidase. Clin Cancer Res 16:4198–4206 [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Galmarini CM, Thomas X, Calvo F, Rousselot P, Rabilloud M, El Jaffari A, Cros E, Dumontet C. (2002) In vivo mechanisms of resistance to cytarabine in acute myeloid leukaemia. Br J Haematol 117:860–868 [DOI] [PubMed] [Google Scholar]
  47. Garré ML, Relling MV, Kalwinsky D, Dodge R, Crom WR, Abromowitch M, Pui CH, Evans WE. (1987) Pharmacokinetics and toxicity of methotrexate in children with Down syndrome and acute lymphocytic leukemia. J Pediatr 111:606–612 [DOI] [PubMed] [Google Scholar]
  48. Giacomini KM, Huang SM, Tweedie DJ, Benet LZ, Brouwer KL, Chu X, Dahlin A, Evers R, Fischer V, Hillgren KM, et al. International Transporter Consortium (2010) Membrane transporters in drug development. Nat Rev Drug Discov 9:215–236 [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Glaeser H, Bailey DG, Dresser GK, Gregor JC, Schwarz UI, McGrath JS, Jolicoeur E, Lee W, Leake BF, Tirona RG, et al. (2007) Intestinal drug transporter expression and the impact of grapefruit juice in humans. Clin Pharmacol Ther 81:362–370 [DOI] [PubMed] [Google Scholar]
  50. Goldberg RM, Sargent DJ, Morton RF, Fuchs CS, Ramanathan RK, Williamson SK, Findlay BP, Pitot HC, Alberts SR. (2004) A randomized controlled trial of fluorouracil plus leucovorin, irinotecan, and oxaliplatin combinations in patients with previously untreated metastatic colorectal cancer. J Clin Oncol 22:23–30 [DOI] [PubMed] [Google Scholar]
  51. Gorlick R, Goker E, Trippett T, Waltham M, Banerjee D, Bertino JR. (1996) Intrinsic and acquired resistance to methotrexate in acute leukemia. N Engl J Med 335:1041–1048 [DOI] [PubMed] [Google Scholar]
  52. Graham KA, Leithoff J, Coe IR, Mowles D, Mackey JR, Young JD, Cass CE. (2000) Differential transport of cytosine-containing nucleosides by recombinant human concentrative nucleoside transporter protein hCNT1. Nucleosides Nucleotides Nucleic Acids 19:415–434 [DOI] [PubMed] [Google Scholar]
  53. Gregers J, Christensen IJ, Dalhoff K, Lausen B, Schroeder H, Rosthoej S, Carlsen N, Schmiegelow K, Peterson C. (2010) The association of reduced folate carrier 80G>A polymorphism to outcome in childhood acute lymphoblastic leukemia interacts with chromosome 21 copy number. Blood 115:4671–4677 [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Haschke M, Vitins T, Lüde S, Todesco L, Novakova K, Herrmann R, Krähenbühl S. (2010) Urinary excretion of carnitine as a marker of proximal tubular damage associated with platin-based antineoplastic drugs. Nephrol Dial Transplant 25:426–433 [DOI] [PubMed] [Google Scholar]
  55. Herzig RH, Hines JD, Herzig GP, Wolff SN, Cassileth PA, Lazarus HM, Adelstein DJ, Brown RA, Coccia PF, Strandjord S, et al. (1987) Cerebellar toxicity with high-dose cytosine arabinoside. J Clin Oncol 5:927–932 [DOI] [PubMed] [Google Scholar]
  56. Heuberger W, Berardi S, Jacky E, Pey P, Krähenbühl S. (1998) Increased urinary excretion of carnitine in patients treated with cisplatin. Eur J Clin Pharmacol 54:503–508 [DOI] [PubMed] [Google Scholar]
  57. Hockenberry MJ, Hooke MC, Gregurich M, McCarthy K. (2009) Carnitine plasma levels and fatigue in children/adolescents receiving cisplatin, ifosfamide, or doxorubicin. J Pediatr Hematol Oncol 31:664–669 [DOI] [PubMed] [Google Scholar]
  58. Holmes J, Stanko J, Varchenko M, Ding H, Madden VJ, Bagnell CR, Wyrick SD, Chaney SG. (1998) Comparative neurotoxicity of oxaliplatin, cisplatin, and ormaplatin in a Wistar rat model. Toxicol Sci 46:342–351 [DOI] [PubMed] [Google Scholar]
  59. Hu S, Franke RM, Filipski KK, Hu C, Orwick SJ, de Bruijn EA, Burger H, Baker SD, Sparreboom A. (2008) Interaction of imatinib with human organic ion carriers. Clin Cancer Res 14:3141–3148 [DOI] [PubMed] [Google Scholar]
  60. Hu C, Lancaster CS, Zuo Z, Hu S, Chen Z, Rubnitz JE, Baker SD, Sparreboom A. (2012) Inhibition of OCTN2-mediated transport of carnitine by etoposide. Mol Cancer Ther 11:921–929 [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Huang L, Tissing WJE, de Jonge R, van Zelst BD, Pieters R. (2008) Polymorphisms in folate-related genes: association with side effects of high-dose methotrexate in childhood acute lymphoblastic leukemia. Leukemia 22:1798–1800 [DOI] [PubMed] [Google Scholar]
  62. Humphrey RW, Brockway-Lunardi LM, Bonk DT, Dohoney KM, Doroshow JH, Meech SJ, Ratain MJ, Topalian SL, Pardoll DM. (2011) Opportunities and challenges in the development of experimental drug combinations for cancer. J Natl Cancer Inst 103:1222–1226 [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Hussain A, Dawson N, Amin P, Engstrom C, Dorsey B, Siegel E, Guo C. (2005) Docetaxel followed by hormone therapy in men experiencing increasing prostate-specific antigen after primary local treatments for prostate cancer. J Clin Oncol 23:2789–2796 [DOI] [PubMed] [Google Scholar]
  64. Ip V, Liu JJ, Mercer JF, McKeage MJ. (2010) Differential expression of ATP7A, ATP7B and CTR1 in adult rat dorsal root ganglion tissue. Mol Pain 6:53. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Ishimoto O, Sugawara S, Inoue A, Ishida T, Munakata M, Koinumaru S, Hasegawa Y, Suzuki T, Miki H, Saijo Y, et al. (2006) Phase II study of carboplatin combined with biweekly docetaxel for advanced non-small cell lung cancer. J Thorac Oncol 1:979–983 [PubMed] [Google Scholar]
  66. Ivy KD, Kaplan JH. (2013) A re-evaluation of the role of hCTR1, the human high-affinity copper transporter, in platinum-drug entry into human cells. Mol Pharmacol 83:1237–1246 [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Iwata K, Aizawa K, Kamitsu S, Jingami S, Fukunaga E, Yoshida M, Yoshimura M, Hamada A, Saito H. (2012) Effects of genetic variants in SLC22A2 organic cation transporter 2 and SLC47A1 multidrug and toxin extrusion 1 transporter on cisplatin-induced adverse events. Clin Exp Nephrol 16:843–851 [DOI] [PubMed] [Google Scholar]
  68. Jacobs C, Coleman CN, Rich L, Hirst K, Weiner MW. (1984) Inhibition of cis-diamminedichloroplatinum secretion by the human kidney with probenecid. Cancer Res 44:3632–3635 [PubMed] [Google Scholar]
  69. Jacobs C, Kaubisch S, Halsey J, Lum BL, Gosland M, Coleman CN, Sikic BI. (1991) The use of probenecid as a chemoprotector against cisplatin nephrotoxicity. Cancer 67:1518–1524 [DOI] [PubMed] [Google Scholar]
  70. Johnson AD, Kavousi M, Smith AV, Chen MH, Dehghan A, Aspelund T, Lin JP, van Duijn CM, Harris TB, Cupples LA, et al. (2009) Genome-wide association meta-analysis for total serum bilirubin levels. Hum Mol Genet 18:2700–2710 [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Jong NN, Nakanishi T, Liu JJ, Tamai I, McKeage MJ. (2011) Oxaliplatin transport mediated by organic cation/carnitine transporters OCTN1 and OCTN2 in overexpressing human embryonic kidney 293 cells and rat dorsal root ganglion neurons. J Pharmacol Exp Ther 338:537–547 [DOI] [PubMed] [Google Scholar]
  72. Kameyama Y, Yamashita K, Kobayashi K, Hosokawa M, Chiba K. (2005) Functional characterization of SLCO1B1 (OATP-C) variants, SLCO1B1*5, SLCO1B1*15 and SLCO1B1*15+C1007G, by using transient expression systems of HeLa and HEK293 cells. Pharmacogenet Genomics 15:513–522 [DOI] [PubMed] [Google Scholar]
  73. Katsuda H, Yamashita M, Katsura H, Yu J, Waki Y, Nagata N, Sai Y, Miyamoto K. (2010) Protecting cisplatin-induced nephrotoxicity with cimetidine does not affect antitumor activity. Biol Pharm Bull 33:1867–1871 [DOI] [PubMed] [Google Scholar]
  74. Kim SR, Saito Y, Sai K, Kurose K, Maekawa K, Kaniwa N, Ozawa S, Kamatani N, Shirao K, Yamamoto N, et al. (2007) Genetic variations and frequencies of major haplotypes in SLCO1B1 encoding the transporter OATP1B1 in Japanese subjects: SLCO1B1*17 is more prevalent than *15. Drug Metab Pharmacokinet 22:456–461 [DOI] [PubMed] [Google Scholar]
  75. Kishi S, Griener J, Cheng C, Das S, Cook EH, Pei D, Hudson M, Rubnitz J, Sandlund JT, Pui CH, et al. (2003) Homocysteine, pharmacogenetics, and neurotoxicity in children with leukemia. J Clin Oncol 21:3084–3091 [DOI] [PubMed] [Google Scholar]
  76. Klein J, Bentur Y, Cheung D, Moselhy G, Koren G. (1991) Renal handling of cisplatin: interactions with organic anions and cations in the dog. Clin Invest Med 14:388–394 [PubMed] [Google Scholar]
  77. Kobayashi H, Murakami Y, Uemura K, Sudo T, Hashimoto Y, Kondo N, Sueda T. (2012) Human equilibrative nucleoside transporter 1 expression predicts survival of advanced cholangiocarcinoma patients treated with gemcitabine-based adjuvant chemotherapy after surgical resection. Ann Surg 256:288–296 [DOI] [PubMed] [Google Scholar]
  78. Kobayashi Y, Ohshiro N, Sakai R, Ohbayashi M, Kohyama N, Yamamoto T. (2005) Transport mechanism and substrate specificity of human organic anion transporter 2 (hOat2 [SLC22A7]). J Pharm Pharmacol 57:573–578 [DOI] [PubMed] [Google Scholar]
  79. Koczor CA, Torres RA, Lewis W. (2012) The role of transporters in the toxicity of nucleoside and nucleotide analogs. Expert Opin Drug Metab Toxicol 8:665–676 [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Koepsell H, Lips K, Volk C. (2007) Polyspecific organic cation transporters: structure, function, physiological roles, and biopharmaceutical implications. Pharm Res 24:1227–1251 [DOI] [PubMed] [Google Scholar]
  81. Koizumi T, Nikaido H, Hayakawa J, Nonomura A, Yoneda T. (1988) Infantile disease with microvesicular fatty infiltration of viscera spontaneously occurring in the C3H-H-2(0) strain of mouse with similarities to Reye’s syndrome. Lab Anim 22:83–87 [DOI] [PubMed] [Google Scholar]
  82. Lancaster CS, Hu C, Franke RM, Filipski KK, Orwick SJ, Chen Z, Zuo Z, Loos WJ, Sparreboom A. (2010) Cisplatin-induced downregulation of OCTN2 affects carnitine wasting. Clin Cancer Res 16:4789–4799 [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Lancaster CS, Sprowl JA, Walker AL, Hu S, Gibson AA, Sparreboom A. (2013) Modulation of OATP1B-type transporter function alters cellular uptake and disposition of platinum chemotherapeutics. Mol Cancer Ther 12:1537–1544 [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Larson CA, Blair BG, Safaei R, Howell SB. (2009) The role of the mammalian copper transporter 1 in the cellular accumulation of platinum-based drugs. Mol Pharmacol 75:324–330 [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Lazarus HM, Herzig RH, Herzig GP, Phillips GL, Roessmann U, Fishman DJ. (1981) Central nervous system toxicity of high-dose systemic cytosine arabinoside. Cancer 48:2577–2582 [DOI] [PubMed] [Google Scholar]
  86. Lee W, Glaeser H, Smith LH, Roberts RL, Moeckel GW, Gervasini G, Leake BF, Kim RB. (2005) Polymorphisms in human organic anion-transporting polypeptide 1A2 (OATP1A2): implications for altered drug disposition and central nervous system drug entry. J Biol Chem 280:9610–9617 [DOI] [PubMed] [Google Scholar]
  87. Li Q, Guo D, Dong Z, Zhang W, Zhang L, Huang SM, Polli JE, Shu Y. (2013) Ondansetron can enhance cisplatin-induced nephrotoxicity via inhibition of multiple toxin and extrusion proteins (MATEs). Toxicol Appl Pharmacol 273:100–109 [DOI] [PMC free article] [PubMed] [Google Scholar]
  88. Liu JJ, Jamieson SM, Subramaniam J, Ip V, Jong NN, Mercer JF, McKeage MJ. (2009) Neuronal expression of copper transporter 1 in rat dorsal root ganglia: association with platinum neurotoxicity. Cancer Chemother Pharmacol 64:847–856 [DOI] [PubMed] [Google Scholar]
  89. Liu JJ, Kim Y, Yan F, Ding Q, Ip V, Jong NN, Mercer JF, McKeage MJ. (2013) Contributions of rat Ctr1 to the uptake and toxicity of copper and platinum anticancer drugs in dorsal root ganglion neurons. Biochem Pharmacol 85:207–215 [DOI] [PubMed] [Google Scholar]
  90. Longo N, Amat di San Filippo C, Pasquali M. (2006) Disorders of carnitine transport and the carnitine cycle. Am J Med Genet C Semin Med Genet 142C:77–85 [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Lopez-Lopez E, Martin-Guerrero I, Ballesteros J, Piñan MA, Garcia-Miguel P, Navajas A, Garcia-Orad A. (2011) Polymorphisms of the SLCO1B1 gene predict methotrexate-related toxicity in childhood acute lymphoblastic leukemia. Pediatr Blood Cancer 57:612–619 [DOI] [PubMed] [Google Scholar]
  92. Lu Km, Nishimori H, Nakamura Y, Shima K, Kuwajima M. (1998) A missense mutation of mouse OCTN2, a sodium-dependent carnitine cotransporter, in the juvenile visceral steatosis mouse. Biochem Biophys Res Commun 252:590–594 [DOI] [PubMed] [Google Scholar]
  93. Mancinelli A, D’Iddio S, Bisonni R, Graziano F, Lippe P, Calvani M. (2007) Urinary excretion of L-carnitine and its short-chain acetyl-L-carnitine in patients undergoing carboplatin treatment. Cancer Chemother Pharmacol 60:19–26 [DOI] [PubMed] [Google Scholar]
  94. Marechal R, Bachet JB, Mackey JR, Dalban C, Demetter P, Graham K, Couvelard A, Svrcek M, Bardier-Dupas A, Hammel P, et al. (2012) Levels of gemcitabine transport and metabolism proteins predict survival times of patients treated with gemcitabine for pancreatic adenocarcinoma. Gastroenterology 143:664–674 [DOI] [PubMed] [Google Scholar]
  95. Matei D, Miller AM, Monahan P, Gershenson D, Zhao Q, Cella D, Champion VL, Williams SD. (2009) Chronic physical effects and health care utilization in long-term ovarian germ cell tumor survivors: a Gynecologic Oncology Group study. J Clin Oncol 27:4142–4149 [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Matsushima S, Maeda K, Inoue K, Ohta KY, Yuasa H, Kondo T, Nakayama H, Horita S, Kusuhara H, Sugiyama Y. (2009) The inhibition of human multidrug and toxin extrusion 1 is involved in the drug-drug interaction caused by cimetidine. Drug Metab Dispos 37:555–559 [DOI] [PubMed] [Google Scholar]
  97. McDonald ES, Randon KR, Knight A, Windebank AJ. (2005) Cisplatin preferentially binds to DNA in dorsal root ganglion neurons in vitro and in vivo: a potential mechanism for neurotoxicity. Neurobiol Dis 18:305–313 [DOI] [PubMed] [Google Scholar]
  98. McWhinney SR, Goldberg RM, McLeod HL. (2009) Platinum neurotoxicity pharmacogenetics. Mol Cancer Ther 8:10–16 [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Mikkaichi T, Suzuki T, Onogawa T, Tanemoto M, Mizutamari H, Okada M, Chaki T, Masuda S, Tokui T, Eto N, et al. (2004) Isolation and characterization of a digoxin transporter and its rat homologue expressed in the kidney. Proc Natl Acad Sci USA 101:3569–3574 [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Minematsu T, Giacomini KM. (2011) Interactions of tyrosine kinase inhibitors with organic cation transporters and multidrug and toxic compound extrusion proteins. Mol Cancer Ther 10:531–539 [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Miura K, Goldstein RS, Pasino DA, Hook JB. (1987) Cisplatin nephrotoxicity: role of filtration and tubular transport of cisplatin in isolated perfused kidneys. Toxicology 44:147–158 [DOI] [PubMed] [Google Scholar]
  102. More SS, Akil O, Ianculescu AG, Geier EG, Lustig LR, Giacomini KM. (2010) Role of the copper transporter, CTR1, in platinum-induced ototoxicity. J Neurosci 30:9500–9509 [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Mori S, Ohtsuki S, Takanaga H, Kikkawa T, Kang YS, Terasaki T. (2004) Organic anion transporter 3 is involved in the brain-to-blood efflux transport of thiopurine nucleobase analogs. J Neurochem 90:931–941 [DOI] [PubMed] [Google Scholar]
  104. Morrow CJ, Ghattas M, Smith C, Bönisch H, Bryce RA, Hickinson DM, Green TP, Dive C. (2010) Src family kinase inhibitor Saracatinib (AZD0530) impairs oxaliplatin uptake in colorectal cancer cells and blocks organic cation transporters. Cancer Res 70:5931–5941 [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Myers SN, Goyal RK, Roy JD, Fairfull LD, Wilson JW, Ferrell RE. (2006) Functional single nucleotide polymorphism haplotypes in the human equilibrative nucleoside transporter 1. Pharmacogenet Genomics 16:315–320 [DOI] [PubMed] [Google Scholar]
  106. Nakamura T, Yonezawa A, Hashimoto S, Katsura T, Inui K. (2010) Disruption of multidrug and toxin extrusion MATE1 potentiates cisplatin-induced nephrotoxicity. Biochem Pharmacol 80:1762–1767 [DOI] [PubMed] [Google Scholar]
  107. Nambu T, Hamada A, Nakashima R, Yuki M, Kawaguchi T, Mitsuya H, Saito H. (2011) Association of SLCO1B3 polymorphism with intracellular accumulation of imatinib in leukocytes in patients with chronic myeloid leukemia. Biol Pharm Bull 34:114–119 [DOI] [PMC free article] [PubMed] [Google Scholar]
  108. Neesse A, Frese KK, Chan DS, Bapiro TE, Howat WJ, Richards FM, Ellenrieder V, Jodrell DI, Tuveson DA. (2013) SPARC independent drug delivery and antitumour effects of nab-paclitaxel in genetically engineered mice. Gut DOI: 10.1136/gutjnl-2013-305559 [published ahead of print]. [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Nelson JA, Santos G, Herbert BH. (1984) Mechanisms for the renal secretion of cisplatin. Cancer Treat Rep 68:849–853 [PubMed] [Google Scholar]
  110. Nichols C, Kollmannsberger C. (2011) First-line chemotherapy of disseminated germ cell tumors. Hematol Oncol Clin North Am 25:543–556, viii viii [DOI] [PubMed] [Google Scholar]
  111. Niemi M, Pasanen MK, Neuvonen PJ. (2011) Organic anion transporting polypeptide 1B1: a genetically polymorphic transporter of major importance for hepatic drug uptake. Pharmacol Rev 63:157–181 [DOI] [PubMed] [Google Scholar]
  112. Ohashi R, Tamai I, Yabuuchi H, Nezu JI, Oku A, Sai Y, Shimane M, Tsuji A. (1999) Na(+)-dependent carnitine transport by organic cation transporter (OCTN2): its pharmacological and toxicological relevance. J Pharmacol Exp Ther 291:778–784 [PubMed] [Google Scholar]
  113. Ohta KY, Inoue K, Yasujima T, Ishimaru M, Yuasa H. (2009) Functional characteristics of two human MATE transporters: kinetics of cimetidine transport and profiles of inhibition by various compounds. J Pharm Pharm Sci 12:388–396 [DOI] [PubMed] [Google Scholar]
  114. Ohtani Y, Nishiyama S, Matsuda I. (1984) Renal handling of free and acylcarnitine in secondary carnitine deficiency. Neurology 34:977–979 [DOI] [PubMed] [Google Scholar]
  115. Okazaki T, Javle M, Tanaka M, Abbruzzese JL, Li D. (2010) Single nucleotide polymorphisms of gemcitabine metabolic genes and pancreatic cancer survival and drug toxicity. Clin Cancer Res 16:320–329 [DOI] [PMC free article] [PubMed] [Google Scholar]
  116. Osato DH, Huang CC, Kawamoto M, Johns SJ, Stryke D, Wang J, Ferrin TE, Herskowitz I, Giacomini KM. (2003) Functional characterization in yeast of genetic variants in the human equilibrative nucleoside transporter, ENT1. Pharmacogenetics 13:297–301 [DOI] [PubMed] [Google Scholar]
  117. Pabla N, Murphy RF, Liu K, Dong Z. (2009) The copper transporter Ctr1 contributes to cisplatin uptake by renal tubular cells during cisplatin nephrotoxicity. Am J Physiol Renal Physiol 296:F505–F511 [DOI] [PMC free article] [PubMed] [Google Scholar]
  118. Parmar S, Seeringer A, Denich D, Gärtner F, Pitterle K, Syrovets T, Ohmle B, Stingl JC. (2011) Variability in transport and biotransformation of cytarabine is associated with its toxicity in peripheral blood mononuclear cells. Pharmacogenomics 12:503–514 [DOI] [PubMed] [Google Scholar]
  119. Petrylak DP, Tangen CM, Hussain MH, Lara PN, Jr, Jones JA, Taplin ME, Burch PA, Berry D, Moinpour C, Kohli M, et al. (2004) Docetaxel and estramustine compared with mitoxantrone and prednisone for advanced refractory prostate cancer. N Engl J Med 351:1513–1520 [DOI] [PubMed] [Google Scholar]
  120. Radtke S, Zolk O, Renner B, Paulides M, Zimmermann M, Möricke A, Stanulla M, Schrappe M, Langer T. (2013) Germline genetic variations in methotrexate candidate genes are associated with pharmacokinetics, toxicity, and outcome in childhood acute lymphoblastic leukemia. Blood 121:5145–5153 [DOI] [PubMed] [Google Scholar]
  121. Ramsey LB, Bruun GH, Yang W, Treviño LR, Vattathil S, Scheet P, Cheng C, Rosner GL, Giacomini KM, Fan Y, et al. (2012) Rare versus common variants in pharmacogenetics: SLCO1B1 variation and methotrexate disposition. Genome Res 22:1–8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  122. Ramsey LB, Panetta JC, Smith C, Yang W, Fan Y, Winick NJ, Martin PL, Cheng C, Devidas M, Pui CH, et al. (2013) Genome-wide study of methotrexate clearance replicates SLCO1B1. Blood 121:898–904 [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. Rathkopf D, Carducci MA, Morris MJ, Slovin SF, Eisenberger MA, Pili R, Denmeade SR, Kelsen M, Curley T, Halter M, et al. (2008) Phase II trial of docetaxel with rapid androgen cycling for progressive noncastrate prostate cancer. J Clin Oncol 26:2959–2965 [DOI] [PMC free article] [PubMed] [Google Scholar]
  124. Reykdal S, Sham R, Kouides P. (1995) Cytarabine-induced pericarditis: a case report and review of the literature of the cardio-pulmonary complications of cytarabine therapy. Leuk Res 19:141–144 [DOI] [PubMed] [Google Scholar]
  125. Ringel I, Horwitz SB. (1991) Studies with RP 56976 (taxotere): a semisynthetic analogue of taxol. J Natl Cancer Inst 83:288–291 [DOI] [PubMed] [Google Scholar]
  126. Ritzel MW, Ng AM, Yao SY, Graham K, Loewen SK, Smith KM, Hyde RJ, Karpinski E, Cass CE, Baldwin SA, et al. (2001) Recent molecular advances in studies of the concentrative Na+-dependent nucleoside transporter (CNT) family: identification and characterization of novel human and mouse proteins (hCNT3 and mCNT3) broadly selective for purine and pyrimidine nucleosides (system cib). Mol Membr Biol 18:65–72 [DOI] [PubMed] [Google Scholar]
  127. Ross DA, Gale GR. (1979) Reduction of the renal toxicity of cis-dichlorodiammineplatinum(II) by probenecid. Cancer Treat Rep 63:781–787 [PubMed] [Google Scholar]
  128. Rougier P, Adenis A, Ducreux M, de Forni M, Bonneterre J, Dembak M, Clouet P, Lebecq A, Baille P, Lefresne-Soulas F, et al. (2000) A phase II study: docetaxel as first-line chemotherapy for advanced pancreatic adenocarcinoma. Eur J Cancer 36:1016–1025 [DOI] [PubMed] [Google Scholar]
  129. Ruggiero A, Trombatore G, Triarico S, Arena R, Ferrara P, Scalzone M, Pierri F, Riccardi R. (2013) Platinum compounds in children with cancer: toxicity and clinical management. Anticancer Drugs 24:1007–1019 [DOI] [PubMed] [Google Scholar]
  130. Scaglia F, Wang Y, Singh RH, Dembure PP, Pasquali M, Fernhoff PM, Longo N. (1998) Defective urinary carnitine transport in heterozygotes for primary carnitine deficiency. Genet Med 1:34–39 [DOI] [PubMed] [Google Scholar]
  131. Scheffler M, Di Gion P, Doroshyenko O, Wolf J, Fuhr U. (2011) Clinical pharmacokinetics of tyrosine kinase inhibitors: focus on 4-anilinoquinazolines. Clin Pharmacokinet 50:371–403 [DOI] [PubMed] [Google Scholar]
  132. Screnci D, McKeage MJ, Galettis P, Hambley TW, Palmer BD, Baguley BC. (2000) Relationships between hydrophobicity, reactivity, accumulation and peripheral nerve toxicity of a series of platinum drugs. Br J Cancer 82:966–972 [DOI] [PMC free article] [PubMed] [Google Scholar]
  133. Shen H, Yang Z, Zhao W, Zhang Y, Rodrigues AD. (2013) Assessment of vandetanib as an inhibitor of various human renal transporters: inhibition of multidrug and toxin extrusion as a possible mechanism leading to decreased cisplatin and creatinine clearance. Drug Metab Dispos 41:2095–2103 [DOI] [PubMed] [Google Scholar]
  134. Shimasaki N, Mori T, Samejima H, Sato R, Shimada H, Yahagi N, Torii C, Yoshihara H, Tanigawara Y, Takahashi T, et al. (2006) Effects of methylenetetrahydrofolate reductase and reduced folate carrier 1 polymorphisms on high-dose methotrexate-induced toxicities in children with acute lymphoblastic leukemia or lymphoma. J Pediatr Hematol Oncol 28:64–68 [DOI] [PubMed] [Google Scholar]
  135. Sissung TM, Baum CE, Deeken J, Price DK, Aragon-Ching J, Steinberg SM, Dahut W, Sparreboom A, Figg WD. (2008) ABCB1 genetic variation influences the toxicity and clinical outcome of patients with androgen-independent prostate cancer treated with docetaxel. Clin Cancer Res 14:4543–4549 [DOI] [PMC free article] [PubMed] [Google Scholar]
  136. Sissung TM, Mross K, Steinberg SM, Behringer D, Figg WD, Sparreboom A, Mielke S. (2006) Association of ABCB1 genotypes with paclitaxel-mediated peripheral neuropathy and neutropenia. Eur J Cancer 42:2893–2896 [DOI] [PMC free article] [PubMed] [Google Scholar]
  137. Sleijfer DT, Offerman JJ, Mulder NH, Verweij M, van der Hem GK, Schraffordt Koops HS, Meijer S. (1987) The protective potential of the combination of verapamil and cimetidine on cisplatin-induced nephrotoxicity in man. Cancer 60:2823–2828 [DOI] [PubMed] [Google Scholar]
  138. Smith NF, Acharya MR, Desai N, Figg WD, Sparreboom A. (2005) Identification of OATP1B3 as a high-affinity hepatocellular transporter of paclitaxel. Cancer Biol Ther 4:815–818 [DOI] [PubMed] [Google Scholar]
  139. Soo RA, Wang LZ, Ng SS, Chong PY, Yong WP, Lee SC, Liu JJ, Choo TB, Tham LS, Lee HS, et al. (2009) Distribution of gemcitabine pathway genotypes in ethnic Asians and their association with outcome in non-small cell lung cancer patients. Lung Cancer 63:121–127 [DOI] [PubMed] [Google Scholar]
  140. Sowers R, Wenzel BD, Richardson C, Meyers PA, Healey JH, Levy AS, Gorlick R. (2011) Impairment of methotrexate transport is common in osteosarcoma tumor samples. Sarcoma 2011:834170. [DOI] [PMC free article] [PubMed] [Google Scholar]
  141. Sprowl JA, Ciarimboli G, Lancaster CS, Giovinazzo H, Gibson AA, Du G, Janke LJ, Cavaletti G, Shields AF, Sparreboom A. (2013a) Oxaliplatin-induced neurotoxicity is dependent on the organic cation transporter OCT2. Proc Natl Acad Sci USA 110:11199–11204 [DOI] [PMC free article] [PubMed] [Google Scholar]
  142. Sprowl JA, van Doorn L, Hu S, van Gerven L, de Bruijn P, Li L, Gibson AA, Mathijssen RH, Sparreboom A. (2013b) Conjunctive therapy of cisplatin with the OCT2 inhibitor cimetidine: influence on antitumor efficacy and systemic clearance. Clin Pharmacol Ther 94:585–592 [DOI] [PMC free article] [PubMed] [Google Scholar]
  143. Sun W, Wu RR, van Poelje PD, Erion MD. (2001) Isolation of a family of organic anion transporters from human liver and kidney. Biochem Biophys Res Commun 283:417–422 [DOI] [PubMed] [Google Scholar]
  144. Sundaram M, Yao SY, Ingram JC, Berry ZA, Abidi F, Cass CE, Baldwin SA, Young JD. (2001) Topology of a human equilibrative, nitrobenzylthioinosine (NBMPR)-sensitive nucleoside transporter (hENT1) implicated in the cellular uptake of adenosine and anti-cancer drugs. J Biol Chem 276:45270–45275 [DOI] [PubMed] [Google Scholar]
  145. Svoboda M, Wlcek K, Taferner B, Hering S, Stieger B, Tong D, Zeillinger R, Thalhammer T, Jäger W. (2011) Expression of organic anion-transporting polypeptides 1B1 and 1B3 in ovarian cancer cells: relevance for paclitaxel transport. Biomed Pharmacother 65:417–426 [DOI] [PubMed] [Google Scholar]
  146. Sweet DH, Chan LM, Walden R, Yang XP, Miller DS, Pritchard JB. (2003) Organic anion transporter 3 (Slc22a8) is a dicarboxylate exchanger indirectly coupled to the Na+ gradient. Am J Physiol Renal Physiol 284:F763–F769 [DOI] [PubMed] [Google Scholar]
  147. Sweet DH, Wolff NA, Pritchard JB. (1997) Expression cloning and characterization of ROAT1. The basolateral organic anion transporter in rat kidney. J Biol Chem 272:30088–30095 [DOI] [PubMed] [Google Scholar]
  148. Takeda M, Khamdang S, Narikawa S, Kimura H, Hosoyamada M, Cha SH, Sekine T, Endou H. (2002) Characterization of methotrexate transport and its drug interactions with human organic anion transporters. J Pharmacol Exp Ther 302:666–671 [DOI] [PubMed] [Google Scholar]
  149. Tanaka M, Javle M, Dong X, Eng C, Abbruzzese JL, Li D. (2010) Gemcitabine metabolic and transporter gene polymorphisms are associated with drug toxicity and efficacy in patients with locally advanced pancreatic cancer. Cancer 116:5325–5335 [DOI] [PMC free article] [PubMed] [Google Scholar]
  150. Tanihara Y, Masuda S, Katsura T, Inui K. (2009) Protective effect of concomitant administration of imatinib on cisplatin-induced nephrotoxicity focusing on renal organic cation transporter OCT2. Biochem Pharmacol 78:1263–1271 [DOI] [PubMed] [Google Scholar]
  151. Tannock IF, de Wit R, Berry WR, Horti J, Pluzanska A, Chi KN, Oudard S, Théodore C, James ND, Turesson I, et al. TAX 327 Investigators (2004) Docetaxel plus prednisone or mitoxantrone plus prednisone for advanced prostate cancer. N Engl J Med 351:1502–1512 [DOI] [PubMed] [Google Scholar]
  152. Taplin ME, Xie W, Bubley GJ, Ernstoff MS, Walsh W, Morganstern DE, Regan MM. (2006) Docetaxel, estramustine, and 15-month androgen deprivation for men with prostate-specific antigen progression after definitive local therapy for prostate cancer. J Clin Oncol 24:5408–5413 [DOI] [PubMed] [Google Scholar]
  153. Thomas J, Wang L, Clark RE, Pirmohamed M. (2004) Active transport of imatinib into and out of cells: implications for drug resistance. Blood 104:3739–3745 [DOI] [PubMed] [Google Scholar]
  154. Treviño LR, Shimasaki N, Yang W, Panetta JC, Cheng C, Pei D, Chan D, Sparreboom A, Giacomini KM, Pui CH, et al. (2009) Germline genetic variation in an organic anion transporter polypeptide associated with methotrexate pharmacokinetics and clinical effects. J Clin Oncol 27:5972–5978 [DOI] [PMC free article] [PubMed] [Google Scholar]
  155. Tsuda M, Terada T, Ueba M, Sato T, Masuda S, Katsura T, Inui K. (2009) Involvement of human multidrug and toxin extrusion 1 in the drug interaction between cimetidine and metformin in renal epithelial cells. J Pharmacol Exp Ther 329:185–191 [DOI] [PubMed] [Google Scholar]
  156. Urakami Y, Akazawa M, Saito H, Okuda M, Inui K. (2002) cDNA cloning, functional characterization, and tissue distribution of an alternatively spliced variant of organic cation transporter hOCT2 predominantly expressed in the human kidney. J Am Soc Nephrol 13:1703–1710 [DOI] [PubMed] [Google Scholar]
  157. van den Bent MJ, van Putten WL, Hilkens PH, de Wit R, van der Burg ME. (2002) Retreatment with dose-dense weekly cisplatin after previous cisplatin chemotherapy is not complicated by significant neuro-toxicity. Eur J Cancer 38:387–391 [DOI] [PubMed] [Google Scholar]
  158. van de Steeg E, Stránecký V, Hartmannová H, Nosková L, Hřebíček M, Wagenaar E, van Esch A, de Waart DR, Oude Elferink RP, Kenworthy KE, et al. (2012) Complete OATP1B1 and OATP1B3 deficiency causes human Rotor syndrome by interrupting conjugated bilirubin reuptake into the liver. J Clin Invest 122:519–528 [DOI] [PMC free article] [PubMed] [Google Scholar]
  159. van de Steeg E, van der Kruijssen CM, Wagenaar E, Burggraaff JE, Mesman E, Kenworthy KE, Schinkel AH. (2009) Methotrexate pharmacokinetics in transgenic mice with liver-specific expression of human organic anion-transporting polypeptide 1B1 (SLCO1B1). Drug Metab Dispos 37:277–281 [DOI] [PubMed] [Google Scholar]
  160. van de Steeg E, van Esch A, Wagenaar E, Kenworthy KE, Schinkel AH. (2013) Influence of human OATP1B1, OATP1B3, and OATP1A2 on the pharmacokinetics of methotrexate and paclitaxel in humanized transgenic mice. Clin Cancer Res 19:821–832 [DOI] [PubMed] [Google Scholar]
  161. van de Steeg E, van Esch A, Wagenaar E, van der Kruijssen CM, van Tellingen O, Kenworthy KE, Schinkel AH. (2011) High impact of Oatp1a/1b transporters on in vivo disposition of the hydrophobic anticancer drug paclitaxel. Clin Cancer Res 17:294–301 [DOI] [PubMed] [Google Scholar]
  162. VanWert AL, Sweet DH. (2008) Impaired clearance of methotrexate in organic anion transporter 3 (Slc22a8) knockout mice: a gender specific impact of reduced folates. Pharm Res 25:453–462 [DOI] [PMC free article] [PubMed] [Google Scholar]
  163. Vrignaud P, Sémiond D, Lejeune P, Bouchard H, Calvet L, Combeau C, Riou JF, Commerçon A, Lavelle F, Bissery MC. (2013) Preclinical antitumor activity of cabazitaxel, a semisynthetic taxane active in taxane-resistant tumors. Clin Cancer Res 19:2973–2983 [DOI] [PubMed] [Google Scholar]
  164. Wang L, Giannoudis A, Lane S, Williamson P, Pirmohamed M, Clark RE. (2008) Expression of the uptake drug transporter hOCT1 is an important clinical determinant of the response to imatinib in chronic myeloid leukemia. Clin Pharmacol Ther 83:258–264 [DOI] [PubMed] [Google Scholar]
  165. Weiner MW, Jacobs C. (1983) Mechanism of cisplatin nephrotoxicity. Fed Proc 42:2974–2978 [PubMed] [Google Scholar]
  166. Williams FM, Flintoff WF. (1995) Isolation of a human cDNA that complements a mutant hamster cell defective in methotrexate uptake. J Biol Chem 270:2987–2992 [DOI] [PubMed] [Google Scholar]
  167. Williams PD, Hottendorf GH. (1985) Effect of cisplatin on organic ion transport in membrane vesicles from rat kidney cortex. Cancer Treat Rep 69:875–880 [PubMed] [Google Scholar]
  168. Winkelman MD, Hines JD. (1983) Cerebellar degeneration caused by high-dose cytosine arabinoside: a clinicopathological study. Ann Neurol 14:520–527 [DOI] [PubMed] [Google Scholar]
  169. Xu CF, Reck BH, Xue Z, Huang L, Baker KL, Chen M, Chen EP, Ellens HE, Mooser VE, Cardon LR, et al. (2010) Pazopanib-induced hyperbilirubinemia is associated with Gilbert’s syndrome UGT1A1 polymorphism. Br J Cancer 102:1371–1377 [DOI] [PMC free article] [PubMed] [Google Scholar]
  170. Yamakawa Y, Hamada A, Shuto T, Yuki M, Uchida T, Kai H, Kawaguchi T, Saito H. (2011) Pharmacokinetic impact of SLCO1A2 polymorphisms on imatinib disposition in patients with chronic myeloid leukemia. Clin Pharmacol Ther 90:157–163 [DOI] [PubMed] [Google Scholar]
  171. Yang-Feng TL, Ma YY, Liang R, Prasad PD, Leibach FH, Ganapathy V. (1995) Assignment of the human folate transporter gene to chromosome 21q22.3 by somatic cell hybrid analysis and in situ hybridization. Biochem Biophys Res Commun 210:874–879 [DOI] [PubMed] [Google Scholar]
  172. Yardley DA. (2013) nab-Paclitaxel mechanisms of action and delivery. J Control Release 170:365–372 [DOI] [PubMed] [Google Scholar]
  173. Yokoo S, Yonezawa A, Masuda S, Fukatsu A, Katsura T, Inui K. (2007) Differential contribution of organic cation transporters, OCT2 and MATE1, in platinum agent-induced nephrotoxicity. Biochem Pharmacol 74:477–487 [DOI] [PubMed] [Google Scholar]
  174. Yonezawa A, Masuda S, Nishihara K, Yano I, Katsura T, Inui K. (2005) Association between tubular toxicity of cisplatin and expression of organic cation transporter rOCT2 (Slc22a2) in the rat. Biochem Pharmacol 70:1823–1831 [DOI] [PubMed] [Google Scholar]
  175. Yonezawa A, Masuda S, Yokoo S, Katsura T, Inui K. (2006) Cisplatin and oxaliplatin, but not carboplatin and nedaplatin, are substrates for human organic cation transporters (SLC22A1-3 and multidrug and toxin extrusion family). J Pharmacol Exp Ther 319:879–886 [DOI] [PubMed] [Google Scholar]
  176. Zhang L, Brett CM, Giacomini KM. (1998) Role of organic cation transporters in drug absorption and elimination. Annu Rev Pharmacol Toxicol 38:431–460 [DOI] [PubMed] [Google Scholar]
  177. Zhao R, Min SH, Wang Y, Campanella E, Low PS, Goldman ID. (2009) A role for the proton-coupled folate transporter (PCFT-SLC46A1) in folate receptor-mediated endocytosis. J Biol Chem 284:4267–4274 [DOI] [PMC free article] [PubMed] [Google Scholar]
  178. Zimmerman EI, Hu S, Roberts JL, Gibson AA, Orwick SJ, Li L, Sparreboom A, Baker SD. (2013) Contribution of OATP1B1 and OATP1B3 to the disposition of sorafenib and sorafenib-glucuronide. Clin Cancer Res 19:1458–1466 [DOI] [PMC free article] [PubMed] [Google Scholar]
  179. Zwelling LA, Anderson T, Kohn KW. (1979) DNA-protein and DNA interstrand cross-linking by cis- and trans-platinum(II) diamminedichloride in L1210 mouse leukemia cells and relation to cytotoxicity. Cancer Res 39:365–369 [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Data Supplement

Articles from Drug Metabolism and Disposition are provided here courtesy of American Society for Pharmacology and Experimental Therapeutics

RESOURCES