Feature Review

Impact of Membrane Drug

Transporters on Resistance to

Small-Molecule Tyrosine

Kinase Inhibitors

1 1 2

Claudia Neul, Elke Schaeffeler, Alex Sparreboom,

3 1,4,5, 1

Stefan Laufer, Matthias Schwab, * and Anne T. Nies

Small-molecule inhibitors of tyrosine kinases (TKIs) are the mainstay of treat-

Trends

ment for many malignancies and represent novel treatment options for other

Small-molecule TKIs are the backbone

fi fi

diseases such as idiopathic pulmonary brosis. Twenty- ve TKIs are currently of cancer therapy. Despite profound

success of imatinib in improving out-

FDA-approved and >130 are being evaluated in clinical trials. Increasing evi-

come of patients with chronic myeloid

dence suggests that drug exposure of TKIs may significantly contribute to drug

leukemia, survival benefits of TKIs in

resistance, independently from somatic variation of TKI target . Mem- other settings are varying and

moderate.

brane transport may limit the amount of TKI reaching the target cells.

This review highlights current knowledge on the basic and clinical pharmacol-

In addition to genetic variation of the

ogy of membrane transporters involved in TKI disposition and their contribution targeted kinases, and diverse meta-

bolic and cellular escape mechanisms,

to drug efficacy and adverse drug effects. In addition to non-genetic and

insufficient TKI exposure and intracel-

epigenetic factors, genetic variants, particularly rare ones, in transporter genes

lular accumulation mediated by mem-

are promising novel factors to explain interindividual variability in the response brane uptake and ef ux transporters

are important for interindividual varia-

to TKI therapy.

bility of TKI response and the occur-

rence of drug resistance.

TKIs: Response Encounters Resistance While in vitro studies and studies using

Protein kinases are key players in signal transduction networks mediating fundamental knockout mice clearly demonstrate TKI

transport by efflux transporters, the

cellular processes including cell differentiation, proliferation, apoptosis, transcription, metab-

picture is less clear in pharmacogenetic

olism, and intercellular communication [1–3]. During the past 15 years it has become evident

association studies in humans.

that many cancers, as well as metabolic disorders and immunological, neurological, and Furthermore, the role of currently inves-

inflammatory diseases, may originate from dysregulation of these signaling networks [2,4,5]. tigated uptake transporters appears to

be limited.

Protein kinases, including tyrosine kinases, have thereby emerged as the most intensively

studied target structures. Concurrently, drug development has shifted towards small

Novel techniques and strategies are

fi ‘

molecules that speci cally block protein kinases crucial to disease biology. These molecu- needed to better predict the relevance

larly targeted therapies’ are now routinely used in the therapy of many cancers as first-line of membrane transporters for in vivo

TKI disposition.

therapy but are also essential in the treatment of drug-resistant malignancies as well as

of other diseases such as idiopathic fibrosis and rheumatoid arthritis [5,6]. In 2001 the first

TKI, imatinib, received accelerated approval by the USA FDA for the treatment of Phila-

1

delphia positive chronic myeloid leukemia (CML) [7]. Since then, 24 further Dr. Margarete Fischer-Bosch Institute

of Clinical Pharmacology, Stuttgart,

TKIs have been approved, mostly for the treatment of various cancers (Figure 1A). A plethora

and University of Tübingen, Germany

of >3000 novel agents inhibiting diverse protein kinases are currently being 2

Division of Pharmaceutics, College of

explored preclinically, covering not only a broad range of cancers but also ophthalmic Pharmacy, Ohio State University,

Columbus, OH, USA and central nervous system disorders, osteoporosis, as well as disease-related complica-

3

Department of Pharmaceutical

tions [2]. At present, more than 130 novel TKIs are being evaluated in clinical trials

Chemistry, University of Tübingen,

(Figure 1B). Tübingen, Germany

904 Trends in Pharmacological Sciences, November 2016, Vol. 37, No. 11 http://dx.doi.org/10.1016/j.tips.2016.08.003

© 2016 Elsevier Ltd. All rights reserved.

4

Department of Clinical Pharmacology,

Contrary to the profound improvement of outcomes for CML patients treated with imatinib as first-

Institute of Experimental and Clinical

line therapy [8], the success rate of TKIs in other diseases is varying, and only moderate survival

Pharmacology and Toxicology,

benefits are achieved for some disease entities [6,9,10]. Poor initial response (i.e., endogenous University Hospital, Tübingen,

Germany

resistance) or disease relapse (i.e., acquired resistance) are being discussed as major explan- 5

Department of Pharmacy and

ations. Of course, the development of mutations in the target receptor tyrosine kinases (RTK) or

Biochemistry, University of Tübingen,

intracellular non-receptor tyrosine kinases (nonRTK) (Figure 2, Key Figure) is crucial for target- Tübingen, Germany

associated drug resistance, thereby preventing kinase inhibition by the TKI in the target cells. For

example, patients with non-small cell lung cancer (NSCLC) harboring an activating mutation in the

*Correspondence:

RTK epidermal growth factor receptor (EGFR) initially respond very well to erlotinib and ge tinib; [email protected]

however, the acquisition of secondary somatic mutations, such as the gatekeeper mutation (M. Schwab).

Thr790Met, leads to drug resistance. The development of second- and third-generation TKIs,

such as afatinib and osimertinib, respectively, aimed to overcome this resistance [3,11–13].

Another example is provided by CML patients who become refractory to imatinib therapy as

a result of point mutations in BCR–ABL, the constitutively-active fusion kinase driving CML.

Because imatinib-resistant CML patients frequently acquire the same somatic mutations in BCR–

ABL, the second-generation TKIs dasatinib, nilotinib, and bosutinib have been developed and

approved by the USA FDA for treatment of CML patients carrying these mutations. For CML

patients with the BCR–ABL Thr315Ile gatekeeper mutation, the third-generation TKI ponatinib has

been FDA-approved [3,13,14]. In addition to this target-associated resistance, the activation of

new compensatory intracellular signaling pathways is an alternative explanation (so-called bypass

resistance) in which target cells evade the signaling pathway inhibited by TKI as first-line therapy.

Several recent reviews excellently summarize these mechanisms of resistance [3,5,6,9–13].

For decades multidrug resistance has been well known as a cause of poor response to cancer

chemotherapy [15], and therefore might also contribute to TKI-related molecularly targeted drug

resistance [9,10]. This indicates that, in addition to acquired mutations in the target kinases,

ADME (absorption, disposition, metabolism, excretion) processes may play an important role in

the pathophysiology of drug non-response. There is an increasing body of evidence that

intracellular concentrations of TKI in target cells (Figure 2) vary substantially between patients

even after identical dosages [16–22], with consequences for drug response. By analogy, drug

resistance in HIV-infected patients receiving antiretroviral drug therapy has been significantly

linked to interindividual variability of drug levels of protease inhibitors in peripheral blood

mononuclear cells (PBMC) [23].

Membrane transport proteins, in other words solute carrier transporters (SLC) mediating drug

uptake and ATP-dependent (ATP-binding cassette, ABC) transporters mediating drug efflux, are

involved in several ADME processes [24] and thus are promising candidates to explain TKI-

related non-response in cancer patients as well as drug toxicity. Membrane transporters are

necessary for TKIs to enter the target cells before they can interact with the intracellular ATP-

binding sites of the tyrosine kinases. Various mechanisms, including genetic variation, epigenetic

factors, non-genetic factors such as transporter-mediated drug–drug interaction, as well as

regulation processes, are thought to explain interindividual variability in the expression and

function of transport proteins even in cancer cells, resulting in clinical consequences (e.g., non-

response) [25–27].

This review highlights current knowledge on membrane transporters and their potential impact

on TKI disposition. Current limitations and challenges in transporter research are summarized

and strategies for future research directions are proposed.

Mechanism of Action of TKIs

The family of human tyrosine kinases consists of 90 members: 58 RTKs and 32 intracellular

nonRTKs [1]. They share the same catalytic mechanism of using ATP to phosphorylate tyrosine

Trends in Pharmacological Sciences, November 2016, Vol. 37, No. 11 905

residues on target proteins [2,3,13]. This phosphorylation modifies target protein activities and

leads to the activation of different downstream signaling pathways. While tyrosine kinases differ

in their amino acid sequences, the 3D structures of their kinase domains and particularly the

ATP-binding pockets are highly conserved. The kinase domains consist of an N-terminal lobe, a

C-terminal lobe, and a hinge region linking the two lobes and enabling them to move relative to

each other [3,13]. ATP binds in the active site between these two lobes of the ATP-binding

pocket. Until the 1980s the ATP-binding pockets were considered too conserved to permit the

development of kinase-specific small-molecule inhibitors. The synthesis of compounds selec-

tively inhibiting EGFR [28] was seminal to the establishment of the new therapeutic class of small-

molecule protein kinase inhibitors [5]. Twenty-two of the currently approved TKIs reversibly bind

to the ATP-binding pockets of the respective tyrosine kinase, either to their active or their inactive

conformations. Only afatinib, ibrutinib, and osimertinib irreversibly bind to the ATP-binding

pockets. A detailed description of the molecular characteristics of TKIs and their modes of

inhibition can be found in recent reviews [3,5,13].

Factors Affecting Responsiveness to Oral TKIs

Drug Exposure as a Determinant of Clinical Outcome

All currently approved TKIs are administered orally. Most of those have a relatively long

elimination half-time (Table 1) resulting in a once- or twice-daily dosing interval with beneficial

consequences for adherence to TKI use. Systemic TKI plasma concentrations are affected by

various parameters: (i) the fraction of TKI absorbed in the intestine, (ii) TKI binding to serum

proteins, (iii) TKI uptake into the hepatocytes, (iv) hepatic metabolism primarily by cytochrome

P450 enzyme CYP3A4/A5, and finally (v) biliary excretion (Figure 2B). All these factors may

contribute to the large interpatient variability of TKI exposure, which is reflected by the area under

the plasma concentration–time curve (AUC) and coefficients of variations of up to 77% (e.g.,

axitinib, Table 1). Intestinal absorption may be impaired by concomitant food intake or the use of

antacids (reviewed in [29]). Genetic variation in drug-metabolizing enzymes has been identified

as a major source of variation in the pharmacokinetics and response of many drugs [30]. It is

beyond the scope of this article to comprehensively review the contribution of genetic variation in

metabolizing enzymes to TKI pharmacokinetics and response. Only three examples for the

various effects of genetic variation on TKI response and toxicity should be given: (i) imatinib

pharmacokinetics appears to be affected only to a minor degree by genetic variation in CYP3A4/

5, CYP2C9, CYP2C19, or CYP2D6 [31–33], and had no effect on clinical response in CML

patients [33]; (ii) early sorafenib-induced severe toxicity has been associated with a genetic

variant in UGT1A9 encoding an enzyme involved in glucuronidation of sorafenib, but not with

genetic variation of CYP3A5 [34]; (iii) sunitinib-induced toxicity and hypertension are associated

with genetic variants in CYP3A5 and CYP3A4, respectively [35,36]. In addition to metabolizing

enzymes, membrane uptake and efflux transporters are important determinants of drug plasma

concentrations [24].

All TKIs are characterized by large AUC coefficients of variation (Table 1), and it has been

suggested that this wide pharmacokinetic variability of TKI exposure can affect both TKI efficacy

and toxicity [37]. Indeed, a large number of clinical studies have been conducted correlating

pharmacokinetic parameters such as the AUC or the minimum TKI steady-state plasma

concentrations with TKI response and/or adverse reactions. Selected significant correlations

are summarized in Table 2. For example, in patients with gastrointestinal stromal tumors (GIST)

the probability of a complete/partial response and toxicity is increased 2.6- and 2.2-fold,

respectively, for every doubling of exposure to free imatinib (AUC0–24) [38]. Furthermore, a

minimum imatinib steady-state plasma concentration of >1000 ng/ml was consistently associ-

ated in five independent studies with an increased probability of complete cytogenetic or major

molecular response in CML patients (summarized in [39]). Therefore, several clinical recom-

mendations have been published suggesting therapeutic monitoring of plasma TKI levels to

906 Trends in Pharmacological Sciences, November 2016, Vol. 37, No. 11 (A)

Non-small cell lung cancer Medullary thyroid cancer Erlonib (Tarceva, 2005) Cabozannib (Cometriq, 2012) Lapanib (Tykerb/Tyverb, 2007) Vandetanib (Caprelsa, 2012) Gefinib (Iressa, 2009) Sorafenib (Nexavar, 2013) Crizonib (Xalkori, 2012) Lenvanib (Lenvima, 2015) Afanib (Gilotrif/Giotrif, 2013) Cerinib (Zykadia, 2014) Nintedanib (Vargatef, 2014) Breast cancer Alecnib (Alecensa, 2015) Lapanib (Tykerb/Tyverb, 2008) Osimernib (Tagrisso, 2015)

Idiopathic pulmonary fibrosis Renal cell carcinoma Nintedanib (Ofev, 2014) Sorafenib (Nexavar, 2005) Suninib (Sutent, 2006) Hepatocellular carcinoma Pazopanib (Votrient, 2010) Axinib (Inlyta, 2012) Sorafenib (Nexavar, 2007)

Polycythemia vera Colorectal carcinoma Ruxolinib (Jakafi/Jakavi, 2012) Regorafenib (Svarga, 2012)

Mantle cell lymphoma Chronic myeloid leukemia Ibrunib (lmbruvica, 2013) Imanib (Gleevec/Glivec, 2001) Dasanib (Sprycel, 2006) Nilonib (Tasigna, 2007) Rheumatoid arthris Bosunib (Bosulif, 2013) Tofacinib (Xeljanz, 2012)

Ponanib (Iclusig, 2013) (B) Other neoplasms 20 Lung 40 10

20

Non-tumor al

0

diseases Phase I Phase II Phase III 0

5 Phase I Phase II Phase III Phase IV

0 Phase I Phase II Blood and

Solid tumors and bone marrow sarcom as 30 20 20 10 10

0

0 Phase I Phase II Phase III

Phase I Phase II Liver and kidney

10

Gastrointes nal

tract Female reprod ucve 20 system and breast

20 0 Phase I Phase II 10 10

0 0

Phase I Phase II Phase I Phase II

Figure 1. Small-Molecule Tyrosine Kinase inhibitors (TKIs) Approved (A) and in Clinical Trials (B). (A) Small-molecule TKIs approved by the US FDA and the

European Medicines agency (EMA) as of May 2016 according to their indications. Following the drug name, tradenames of the FDA/EMA are given. Only one tradename

(Figure legend continued on the bottom of the next page.)

Trends in Pharmacological Sciences, November 2016, Vol. 37, No. 11 907

adjust dosages, subsequently leading to improved efficacy or a reduced rate of adverse drug

reactions [20–22,29,39–42]. However, the concept of monitoring serum levels is limited by the

fact that plasma and intracellular TKI concentrations in target cells may differ [17], thereby

supporting the pharmacological rationale that determination of intracellular TKI concentrations is

likely to be a more reliable indicator [40].

Drug Transporters as Major Contributors to Drug Exposure

Membrane transporters are well recognized key determinants of therapeutic response contrib-

uting to drug resistance and drug toxicity [24,43–45]. The encodes >400

membrane transporters that belong to two major groups, the ABC and the SLC superfamilies,

which mediate the cellular uptake and efflux of small molecules. About 20 transporters have

been particularly implicated in drug transport and have been intensively studied with respect to

their function, localization, and regulation. Because these are located in the enterocytes of the

intestine, hepatocytes of the liver, proximal tubule cells of the kidney, and endothelial cells of

tissue barriers (e.g., blood–brain barrier), they are directly involved in the absorption, distribution,

and elimination of drugs, and also indirectly influence metabolism by controlling access to drug-

metabolizing enzymes (Figure 2B–F). Transporter-mediated uptake into non-target cells may

also contribute to drug toxicity, as for example recently demonstrated for OAT6 (SLC22A20),

which regulates entry of sorafenib into keratinocytes and contributes to sorafenib-induced skin

toxicity [46]. Finally, the presence of membrane transporters in the plasma membrane of cancer

cells, either of the primary tumor or metastases, is a major determinant of drug resistance

because decreased transporter-mediated TKI influx or increased TKI efflux may lead to insuffi-

cient intracellular TKI concentrations (Figure 2A).

Table 3 summarizes currently available functional data on TKIs as substrates for uptake and

efflux transporters. While most transporters have only been tested incidentally, almost complete

functional data are available for the efflux transporters ABCB1 (MDR1 P-glycoprotein) and

ABCG2 (BCRP) as well as for the hepatic uptake transporters SLCO1B1 (OATP1B1), SLCO1B3

(OATP1B3), and SLC22A1 (OCT1). ABCB1 and ABCG2 have both been extensively charac-

terized over the past decades and have also been implicated in the multidrug resistance

phenotype against several cytotoxic chemotherapies [15,47,48]. They are recommended by

the US FDA to be tested during the drug development process [24]. Moreover, the organic anion

transporters OATP1B1 and OATP1B3 and organic cation transporter OCT1 are important

hepatocellular drug uptake transporters [24,49–51]. Because almost all TKIs must enter the

hepatocytes for metabolism and biliary elimination (Table 1), several studies have investigated

the involvement of these transporters in TKI uptake.

In Vitro Studies for the Identification and Assessment of Transporters

A common and well-established method to identify whether TKIs are substrates for ABC

transporters is to use polarized cell lines recombinantly expressing the respective ABC trans-

porter in the apical membrane and to measure bidirectional TKI transport [52]. The cut-off for

significant transport is typically a transport ratio of 2 (amount of apically directed transport

divided by the amount of basolaterally directed transport) [52]. Accordingly, all TKIs, except for

cabozantinib, ibrutinib, ruxolitinib and vandetanib, are substrates for at least ABCB1 or ABCG2

(Table 3). However, methodological issues, notably those associated with cell-based assays,

can significantly alter in vitro study results [53], and hence there is still debate in the literature,

indicates that the same tradename is current in the USA and Europe with the following exceptions: (i) EMA refused the marketing authorization of tofacitinib on 25 July

2013; (ii) alectinib has not yet been approved by the EMA and has only recently been approved by the FDA (December 2015). The date indicates the FDA registration date,

except for nintedanib, which is only approved by the EMA for the treatment of lung cancer. (B) More than 130 novel TKIs are currently being evaluated in different clinical

phases. The pie chart shows their distribution to the neoplasms of the different organs and tissues (percentage of total). In the bar charts, the number of new TKIs in the

different clinical study phases is given. All data were taken from http://clinicaltrials.gov using the keyword ‘tyrosine kinase inhibitors’.

908 Trends in Pharmacological Sciences, November 2016, Vol. 37, No. 11

Key Figure

Tyrosine kinase inhibitor (TKI) resistance mechanisms in cancer cells (A)

and the location of membrane transporters involved in absorption, dis-

tribution, and excretion of TKIs (B–F)

(A) Mutaons TKI Lysosomal in RTK sequestraon

Decreased influx Mutaons in Acvaon of TKI intracellular compensatory CYP kinase signaling pathways TKI metabolites Increased efflux

(B) (C) Oral dose Entero cyte

Blood SLC22 A4

SLCO2B1 Othe r SLCs TKI Enterocyte ABCC3 TKI TKI ABCB1 ABCC2 ABCG2

CYP Portal vein TKI metabolites

(D) Hepatocyte Blood SLCO1B1 TKI SLCO1B3

Bile SLCO2B1

CYP SLC22 A1

SLC22 A3 TKI metabolites To systemic Bile SLC22 A7 circulaon ABCB1 and target cells

Hepatocyte ABCC2

ABCG2 ABCC4 SLC47 A1 ABCC6

To feces

(E) Renal proximal (F) Brain

tubule cell endo thelial cell

Blood Blood

SLC22 A2

SLC22 A6 SLCO1A2

SLC22 A7 SLCO2B1 SLC22 A8

ABCB1

ABCB1 SLC47 A1 ABCC4 ABCC2 SLC47 A2 ABCG2 ABCC4

ABCG2

(See figure legend on the bottom of the next page.)

Trends in Pharmacological Sciences, November 2016, Vol. 37, No. 11 909

particularly regarding bosutinib, nilotinib, and sorafenib, on whether these TKIs are indeed

transported by ABCB1. Therefore, standardization of cell-based transporter assays is urgently

required to better predict the relevance of ABC transporters for in vivo TKI disposition.

A common method to study function of uptake transporters is to determine substance accu-

mulation by transfected versus parental or nontransfected cells. A twofold-increased accu-

mulation by transfected cells is typically considered to demonstrate significant transport [52]. In

general, accumulation ratios for TKIs are

considered to be substrates for uptake transporters according to recent recommendations [52].

Accumulation ratios 2.0 were only measured for OATP1B1 with axitinib and regorafenib, for

OATP1B3 with axitinib, and for OCT1 with dasatinib and nintedanib (Table 3). As noted above for

the in vitro analysis of ABC transporters, it appears essential to standardize cell-based uptake

transporter assays.

Furthermore, the currently investigated SLC transporters do not explain cellular TKI uptake,

suggesting that the major TKI uptake mechanisms have not yet been identified. However,

there is considerable debate as to whether substances, particularly uncharged molecules, can

cross biomembranes by passive diffusion or whether membrane transporters are required for

this process [54–57]. This discussion also applies to TKIs because several of them are

uncharged at the physiological pH of 7.4 (Table 3). Moreover, irrespective of their charge,

some TKIs, such as imatinib and sorafenib, have a high permeability in intestinal Caco-2 cells

[58,59]. Putative TKI uptake transporters are still elusive, but they should display the following

characteristics, as recently suggested for highly-permeable drugs [56]: they should be

equilibrative, have Km values in the low millimolar range, and should have high expression

levels in the apical and basolateral membranes of polarized cells. Novel strategies and

experimental techniques [45] will be necessary to discover additional transporters engaged

in cellular TKI uptake.

In addition to transporter-transfected cells, cancer cell lines are often used to study drug

transport processes [60] (Figure 3). However, the use of these cell models may be limited

because drug transporter expression can be downregulated as a result of DNA hyper-

methylation [61,62] and may not reflect expression in the primary tumor or in metastases

[62]. For example, the drug uptake transporter OCT2, which is localized in the basolateral

membrane of renal proximal tubule cells (Figure 2E), and whose expression is preserved

in primary clear cell renal cell carcinoma, cannot be studied in commonly used renal cell

lines because of a lack of OCT2 protein expression [62]. As a further example, human

hepatocellular carcinoma cell lines such as HepG2 lack OCT1, OATP1B1, and OATP1B3

protein expression [64,65], and their use in the preclinical drug development process is also

limited.

Figure 2. (A) Resistance mechanisms in target cells. Target cells are the tumor cells themselves or endothelial cells of tumor

vessels, either in the primary tumor or in metastases. Resistance mechanisms can be grouped into different categories:

(i) resistance by insufficient drug accumulation, for example by decreased influx or increased efflux mediated by membrane

transporters. In addition, sequestration in lysosomes has been recognized as novel mechanism that determines intracellular

levels of, for example, nilotinib, imatinib, and sunitinib [108,112,113]. (ii) Resistance by gatekeeper mutations in receptor

tyrosine kinases (RTKs) and intracellular non-RTKs is considered as the major reason that TKIs fail to inhibit their target

kinases. (iii) Resistance by the response of the cancer cell by upregulating and/or activating compensatory signaling

pathways, thereby bypassing the signaling pathway inhibited by the administered TKI. (B) Absorption, distribution,

metabolism, and excretion of TKIs. All currently approved TKIs are orally administered and therefore need to be taken

up by enterocytes and transported into portal blood. TKIs are then taken up into hepatocytes, where they are primarily

metabolized by cytochrome P450 enzymes (CYPs) and eliminated in the bile. (C–F) Location of major membrane

transporters contributing to drug uptake and efflux in enterocytes, hepatocytes, renal proximal tubule cells, and brain

endothelial cells. See Table 3 for alias transporter names.

910 Trends in Pharmacological Sciences, November 2016, Vol. 37, No. 11 Table 1. Targeted Kinases and Pharmacokinetics of Approved TKIsa b d e f g h TKI Targeted Kinases Protein Dosage t1/2 (h) AUC (ng.h/ml) Metabolizing Enzyme Elimination (%) Bindingc

Major Minor Fecal Renal

Afatinib EGFR, HER2, HER4 95 40 mg single 37 3240–24 (69%) Covalent adducts to proteins, minimal metabolism 85 4

Alectinib ALK, RET >99 600 mg bid 33 74300–12 (46%) CYP3A4 98

Axitinib VEGFR1–3, PDGFR, KIT >99 5 mg bid 3–6 2650–24 (77%) CYP3A4/5 CYP1A2, CYP2C19, 41 23 UGT1A1

Bosutinib BCR–ABL, SRC, LYN, HCK >94 500 mg single 22.5 3650 (425) CYP3A4 91 3

Cabozantinib RET, MET, VEGFR1–3, KIT, 99.7 140 mg qd 55 37 8500–24 (43%) CYP3A4 54 27 TRKB, FLT3, AXL, TIE2

Ceritinib EML4–ALK, IGF-1R, INSR, 97 750 mg single 41 33900–24 (121) CYP3A 92 1 ROS1

Crizotinib EML4–ALK, MET, ROS1, 91 250 mg single 42 2192–2946 (27–31%) CYP3A4/5 63 22 MST1R

Dasatinib BCR–ABL, SRC, LCK, YES, 96 70 mg bid, 3.4 1390–24 (74%) CYP3A4 FMO3, UGT 85 4 FYN, KIT, EPHA2, PDGFRb after 1 day

Erlotinib EGFR 93 150 mg qd 36.2 24 9000–24 CYP3A4 CYP1A2, CYP1A1 83 8

Gefitinib EGFR 90 525 mg qd 37 14 300–15 7000–24 CYP3A4 CYP2D6 86 4 (35–59%)

Ibrutinib BTK 97.3 560 mg qd, 6.9 10520–24 (583) CYP3A CYP2D6 80 10 after 1 day Trends Imatinib BCR–ABL, KIT, PDGFR 95 400 mg qd 15 36 3410–24 (16 572) CYP3A4 CYP1A2, CYP2D6, 68 13 CYP2C9, CYP2C19 in

Pharmacological Lapatinib EGFR, ERBB2 >99 1250 mg qd 24 36 2000–t CYP3A4/5 CYP2C19, CYP2C8 Fecal <2 (23 400–56 000)

Lenvatinib VEGFR1–3, PDGFR/, 98–99 24 mg qd 28 3692steady-state CYP3A, AO 64 25 FGFR1–4, KIT, RET

Sciences, Nilotinib BCR–ABL, KIT, CSF1R, 98 400 mg bid, 17 53300–t (2210) CYP3A4 CYP2C8, CYP2J2, 94 5 PDGFR, DDR1 after 1 day CYP1A2

– / b November Nintedanib VEGFR1 3, PDGFR / , 97.8 150 mg single 9.5 0.8190–12,norm (68%) Esterases UGT1A1, UGT1A7, 93 1 FGFR1–3, FLT3 UGT1A8, UGT1A10

Osimertinib EGFR-T970M 99 80 mg single 48 6709steady-state (54%) CYP3A 68 14

2016, Pazopanib VEGFR1–3, PDGFR//b, >99 800 mg qd 30.9 1 037 5000–t (34%) CYP3A4 CYP1A2, CYP2C8 82 3

Vol. FGFR1/3, KIT, LCK, CSF1R,

37, ITK

No. Ponatinib BCR–ABL, BCR–ABL-T315I, 99 45 mg qd 24 12530–24 (73%) CYP3A4 CYP2C8, CYP2D6, 87 5

11 SRC, FLT3, FGFR, VEGFR, CYP3A5 PDGFR, KIT, RET, TIE2, EPH 911 912 Table 1. (continued)

TKI Targeted Kinasesb Protein Dosaged t (h)e AUCf (ng.h/ml) Metabolizing Enzymeg Elimination (%)h Trends 1/2 Bindingc in Major Minor Fecal Renal

Pharmacological Regorafenib FGFR1/2, PDGFR//b, 99.5 160 mg single 28 70 400 (35%) CYP3A4, UGT1A9 71 19 VEGFR1–3, KIT, RET, RAF1, BRAF, BRAF-V600E, ABL1, TIE2, EPH2A, MAPK11, FRK,

Sciences, NTRK1

Ruxolitinib JAK1/2 97 25 mg bid 1.9 1 335 0780–t ( 632 196) CYP3A4 CYP2C9 22 74

November Sorafenib VEGFR1–3, BRAF, BRAF- 99.5 400 mg bid 25–48 67 3000–12,steady-state (57%) CYP3A4, UGT1A9 77 19 V600E, RAF1, KIT, FLT3, RET, PDGFRb

2016, Sunitinib VEGFR1–3, PDGFR//b, KIT, 95 50 mg qd 40–60 1035–17060–24 (2–56%) CYP3A4 61 16

Vol. FLT3, CSF-1R, RET

37, Tofacitinib JAK1/3 40 5 mg bid, 3 1610–t (86) CYP3A4 CYP2C19 70 30

No. after 1 day

11 Vandetanib EGFR, RET, VEGFRs, PTK6, 90 300 mg qd 456 17 926–38 6110–24 CYP3A4, FMO1, 44 25 TIE2, EPHRs, SRCs (15–58%) FMO3

aAll data were extracted from the Labels and/or the Clinical Pharmacology Biopharmaceutics Reviews of the FDA homepage (www.accessdata.fda.gov/scripts/cder/drugsatfda/). AUC values are from reported studies on patients. bAbbreviations of kinases: ABL1, ABL proto-oncogene 1, non-receptor tyrosine kinase; ALK, anaplastic lymphoma kinase; AXL, AXL receptor tyrosine kinase; BCR–ABL, breakpoint cluster region–Abelson murine leukemia viral oncogene homolog; BRAF, B-Raf proto-oncogene, serine/threonine kinase; BTK, Bruton's tyrosine kinase; CSF1R, colony stimulating factor 1 receptor; DDR1, discoidin domain receptor 1; EGFR, epidermal growth factor receptor; EML4–ALK, echinoderm microtubule-associated protein-like 4–anaplastic lymphoma kinase; EPHA2, ephrin type-A receptor 2; EPH, ephrin receptors; ERBB2, Erb-b2 receptor tyrosine kinase 2; FGFR1–4, fibroblast growth factor receptors 1–4; FLT3, Fms-related tyrosine kinase 3; FRK, Fyn-related Src family tyrosine kinase; FYN, proto-oncogene tyrosine-protein kinase Fyn; HCK, hemopoetic cell kinase proto-oncogene, Src family tyrosine kinase; HER1–4, human epidermal growth factor receptors 1–4; IGF-1R, insulin-like growth factor 1 receptor; INSR, insulin receptor kinase; ITK, IL2 inducible T-cell kinase; JAK1–3, janus kinase 1–3; KIT, mast/stem cell growth factor receptor; LCK, Lck proto-oncogene, Src family tyrosine kinase; LYN, Lyn proto-oncogene, Src family tyrosine kinase; MAPK11, mitogen-activated protein kinase 11; MET, hepatocyte growth factor receptor; MST1R, macrophage stimulating 1 receptor; NTRK1, neurotrophic receptor tyrosine kinase 1; PDGFR, platelet-derived growth factor receptor; PTK6, protein tyrosine kinase 6; RAF1, rapidly accelerated fibrosarcoma 1 proto-oncogene, serine/threonine kinase; RET, Ret proto-oncogene receptor tyrosine kinase; ROS1, Ros proto-oncogene 1, receptor tyrosine kinase; SRC, Src proto-oncogene, non-receptor tyrosine kinase; TIE2, TEK receptor tyrosine kinase; TRKB, tropomyosin receptor kinase B; VEGFR1–3, vascular endothelial growth factor receptors 1–3; YES, Yes proto-oncogene 1, Src family tyrosine kinase. cProtein binding in %. dqd, once daily; bid, twice daily. e t1/2 , elimination half-time. fAUC, area under the plasma concentration-time curve; 0–12, from time 0 to 12 h; 0–24, from time 0 to 24 h; 0–t, from time 0 to next dosing; norm; dose-normalized; variability is given as coefficient of variation (%), as SD, or as range. gAO, aldehyde oxidase; CYP, cytochrome P450; FMO, flavin-containing monooxygenase; UGT, UDP-glucuronosyl transferase. hElimination measured as % radioactivity recovered in feces or urine after a single oral dose.

Table 2. Examples of Significant Correlations of Pharmacokinetic Parameters with the Efficacy and/or

a

Toxicity of Selected TKIs

b c

TKI PK Parameter Cancer Type Outcome

Efficacyd Toxicity

Imatinib Ctrough CML Hematological, Hematological toxicity,

cytogenetic, rash, fluid retention,

molecular response nausea, muscoskeletal pain

Ctrough GIST TTP, PFS Hematological toxicity

Ctrough GIST with KIT OOBR

exon 11 mutation

AUC0–24 free GIST CR, PR

AUC0–24 CML, GIST Number of adverse effects

Sorafenib Ctrough Solid tumors SUVmax

AUCmax Melanoma Tumor control,

PR + SD, PFS

cumulative AUC Solid tumors Grade 3 toxicity

Serum conc. >5.8 mg/ml RCC, HCC Grade 2 hand/foot

skin reaction

Serum conc. >4.8 mg/ml RCC, HCC Grade 2 hypertension

Ctrough Solid tumors Tumor response

Sunitinib Ctrough mRCC, GIST, Hypertension

solid tumors

AUC0–24 mRCC, GIST, TTP, OS, ORR, SD Neutropenia

solid tumors

a

Data extracted from [20,22,39,40] and references therein.

b

AUC, area under the plasma concentration–time curve; 0–24, from time 0 to 24 h; Ctrough, minimum (trough) concentration.

c

CML, chronic myeloid leukemia; GIST, gastrointestinal stromal tumor; HCC, hepatocellular carcinoma; KIT, mast/stem cell

growth factor receptor; mRCC, metastatic renal clear cell carcinoma; RCC, renal clear cell carcinoma.

d

CR, complete response; OOBR, overall objective benefit rate; ORR, objective response rate; OS, overall survival; PFS,

progression-free survival; PR, partial response; SD, stable disease; SUVmax, maximum standardized uptake value; TTP,

time to progression.

Figure 3. Protein Expression of

Tis sue Selected Uptake and Ef ux Trans-

porters in Normal Human Liver and

Normal Tumor Metastasis Tumor cell lines

Kidney Compared to Respective

Tumor and Metastases as well as

Liver

OATP1B1 OATP1B1 OATP1B1X Tumor Cell Lines. White boxes indicate

OATP1B3 OATP1B3 ↓ OATP1B3X expression in normal tissue or unchanged

expression in comparison to normal tis-

OCT1 OCT1 ↓ ? OCT1

X sue. Black boxes and down-arrows indi-

ABCB1 ABCB1 ABCB1 cate reduced expression compared to

normal tissue. Black boxes with red

Kidney OCT2 OCT2 OCT2 OCT2

X crosses indicate lack of expression. Liver

ABCB1 ABCB1 ↓ ? ABCB1 data are from [64,65,91,150] and kidney

data from [62,151,152].

Knockout Mouse Studies To Assess the In Vivo Relevance of ABC

Transporters

A valuable tool to assess the in vivo relevance of ABC transporters are studies with mice

genetically deficient for either Abcb1a/1b or Abcg2, or both Abcb1 and Abcg2 (Table 3). Abcb1

and Abcg2 are not only localized in enterocytes, hepatocytes, and renal proximal tubule cells,

Trends in Pharmacological Sciences, November 2016, Vol. 37, No. 11 913 914 fl

Table 3. TKIs as Substrates of Membrane Uptake and Ef ux Transporters

Trends TKI Charge at 7.4a ABC Transporterb,c

ABCB1 Abcb1a/b ABCG2 Abcg2 Abcb1/g2 ABCB11 ABCC1 ABCC2 ABCC4 in

In Vitro KO Mice in vitro KO mice KO Mice in vitro in vitro in vitro in vitro Pharmacological Afatinib +1 1.4 Y

Alectinib +0.3 <2 20 (4 h oral) <2

Axitinib 0 5.3 7.9 (4 h oral) 1.3 1.1 (4 h oral) 42.4 (4 h oral)

Sciences, Bosutinib +0.9 Y/N* N*

Cabozantinib 0 N N NN

November Ceritinib +1 22.6 38 (3 h oral) 2.3 0.5 (3 h oral) 90 (3 h oral)

Crizotinib +1 15.4 23.1 (4 h oral) 1 0.9 (4 h oral) 26.9 (4 h oral)

Dasatinib +0.3 27.7 5.7 (6 h oral) Y* 0.8 (6 h oral) 26.4 (6 h oral) 3.0 2016, Erlotinib 0 >3 1 (microd) Y* 5 (microd) 3 (microd) 1 N*

Vol. Gefitinib +0.2 8.4 4 (1.5 h oral) Y* 1 (1.5 h oral) 18 (1.5 h oral)

37,

Ibrutinib +0.1 0.1 No.

11 Imatinib +0.8 4.9 4.5 (2 h sc) 3.6 2.4* 0.9 (2 h sc) 2.5 (2 h iv) 64 (2 h sc) (2 h iv) 11.2 (1 h oral)

Lapatinib +0.3 15 6.4 (2 h iv) 2.6 1.3 (infus) 40–42 (infus) 3–4 (infus)

Lenvatinib 0 Y Y N

Nilotinib +0.1 1.0–5.6 Y*

Nintedanib +0.6 Y N N

Osimertinib +1 Y Y

Pazopanib 0 19.5

Ponatinib +0.8 Y Y

Regorafenib 0 1.3 1 (2 h oral) 2.7 3.7 (2 h oral) 7.9 (2 h oral)

Ruxolitinib 0 <1

Sorafenib 0 1.4–6.8 1.9 (4 h oral) 0.9 3.7 4.2 (6 h oral) 3.8 (infus) 8.7–27.8 (6 h oral) N* N* (6 h oral) 1.2 (infus) 9.7 (infus) Table 3. (continued) TKI Charge at 7.4a ABC Transporterb,c

ABCB1 Abcb1a/b ABCG2 Abcg2 Abcb1/g2 ABCB11 ABCC1 ABCC2 ABCC4 In Vitro KO Mice in vitro KO mice KO Mice in vitro in vitro in vitro in vitro

Sunitinib +1 2 2.9 (4 h oral) 2 1.3 (6 h oral) 23.4 (6 h oral) N* Y* 2.3 (6 h oral)

Tofacitinib +0.3 Y N

Vandetanib +1 1

TKI SLC Transporter (In Vitro)b,d Refse

SLCO1A2 SLCO1B1 SLCO1B3 SLCO2B1 SLC22A1 SLC22A2 SLC22A3 SLC22A4 SLC22A5 SLC22A6 SLC22A7 SLC22A8 SLC22A20 SLC47A1

Afatinib FDA

Alectinib FDA [114]

Axatinib Y Y FDA [115]

Bosutinib 1.0 [116,117]

Cabozantinib N N N N N N [118]

Ceritinib 1.3 1 1 1.4 FDA [119]

Crizotinib 1.2 1.3 [77,120]

Dasatinib 1.1 1.3 2.5 N [66,77,101, 116,121–125]

Erlotinib 1 1 1 1.5 0.6 1 0.9 0.8 1 1.6 [126–129]

Gefitinib 1.3 1.6 1.0–1.2 1 [77,129–132]

Ibrutinib FDA

Imatinib 1.5 1.2 1.6 1.0–1.7 1 0.9 1.1–1.6 1.0–1.4 0.9 0.8 0.9 1.4 [63,67,68,70, 123,133–139] Trends

Lapatinib [140,141] in

Lenvatinib N N N N N N FDA Pharmacological Nilotinib 1.2–1.3 1.6–2.0 1 N [77,101,116, 128,142,143]

Nintedanib NNNYN FDA

Osimertinib N N FDA Sciences,

November

2016,

Vol.

37,

No.

11

915 916 Table 3. (continued)

TKI SLC Transporter (In Vitro)b,d Refse Trends SLCO1A2 SLCO1B1 SLCO1B3 SLCO2B1 SLC22A1 SLC22A2 SLC22A3 SLC22A4 SLC22A5 SLC22A6 SLC22A7 SLC22A8 SLC22A20 SLC47A1 in

– –

Pharmacological Pazopanib 1.0 1.3 1.0 1.3 N FDA [77, 101,128]

Ponatinib N N N FDA

Regorafenib 2.5 1.2 [144,145]

Sciences, Ruxolitinib FDA

Sorafenib 0.9 1.0–1.5 1.0–1.5 1.1–1.3 1 1.3 1 1.2 Y* [46,59,66,74, 76,77,146–148]

November Sunitinib 1.0 1.0–1.4 0.9–1.0 1 1.1 1.1 1.2 1.2 FDA [76,77, 148,149]

Tofacitinib FDA 2016,

Vandetanib 1.1–1.3 1.4 N FDA [77,

Vol. 101,128]

37, a

Charge at pH 7.4 was calculated with MarvinSketch 15.9.14 using structures downloaded as SMILES from the PubChem compound library (http://www.ncbi.nlm.nih.gov/pccompound/). No. bAlternative/alias names of transporters: ABCB1, MDR1 P-glycoprotein; ABCG2, breast cancer resistance protein BCRP; ABCB11, bile salt export pump BSEP; ABCC1, multidrug resistance protein MRP1; ABCC2,

11 multidrug resistance protein MRP2; ABCC4, multidrug resistance protein MRP4; SLCO1A2, organic anion transporting polypeptide OATP1A2; SLCO1B1, organic anion transporting polypeptide OATP1B1; SLCO1B3, organic anion transporting polypeptide OATP1B3; SLCO2B1, organic anion transporting polypeptide OATP2B1; SLC22A1, organic cation transporter OCT1; SLC22A2, organic cation transporter OCT2; SLC22A3, organic cation transporter OCT3; SLC22A4, organic cation/zwitterion transporter OCTN1; SLC22A5, organic cation/zwitterion transporter OCTN2; SLC22A6, organic anion transporter OAT1; SLC22A7, organic anion transporter OAT2; SLC22A8, organic anion transporter OAT3; SLC22A20, organic anion transporter OAT6; SLC47A1, multidrug and toxin extrusion MATE1. cValues of in vitro studies are transport ratios (amount of apically directed transport divided by the amount of basolaterally directed transport) from bidirectional transport assays in polarized cell monolayers with transporter-transfected cells or Caco2 cells with transport inhibitors. Values of 2 are considered to reflect significant transport [52]. Y, yes but no transport ratios given; N, no but no transport ratios given; Y*, yes from accumulation studies with transporter-transfected cells; N*, no from accumulation studies with transporter-transfected cells. Values of knockout (ko) mice studies are fold TKI increase in brain accumulation or brain penetration in comparison to wild-type mice at indicated time points after oral administration (oral), intravenous injection (iv), subcutaneous application (sc), continuous infusion (infus) or microdialysis (microd). All brain accumulation data are for the respective parent TKI compound, except for the data of imatinib accumulation 2 h iv, which result from measurement of total radioactivity of imatinib and its metabolites [134]. dData are fold stimulation in uptake caused by the transporter. Values of 2 are considered to reflect significant transport [52]. Studies on knockout mice are only available for Slco1b and Slc22a1 and are discussed in the main text in the section ‘In Vitro and Knockout Mouse Studies To Assess the Role of OCT1 as a Transporter of Imatinib and Sorafenib’. eFDA: data from Clinical Pharmacology Biopharmaceutics Reviews downloaded from the FDA homepage (www.accessdata.fda.gov/scripts/cder/drugsatfda/).

but also in the luminal membrane of brain endothelial cells, where they efflux TKIs back into the

blood, thereby limiting TKI entry into the brain (Figure 2B–F). Therefore, in Abc transporter

knockout mice, the increase in brain accumulation of the investigated TKI in comparison to wild-

type mice reflects the contribution of the respective Abc transporters to TKI disposition.

Notably, these in vivo studies mainly support the results obtained by cell-based assays

(Table 3). Abcb1 appears to be more important than ABCG2/Abcg2 for brain disposition of

most TKIs, except for erlotinib, regorafenib, and sorafenib. Of interest, the combined knockout

of Abcb1 and Abcg2 often leads to a much higher TKI brain accumulation than anticipated from

the single knockouts of Abcb1 and Abcg2. This is particularly observed for axitinib, ceritinib,

dasatinib, gefitinib, imatinib, lapatinib, sorafenib, and sunitinib, indicating that deficiency of one

transporter may be compensated by the other. Notably, while complete knockout of Abcb1

and Abcg2 leads to 58-, 28- and 20-fold higher dasatinib, sorafenib and sunitinib brain

concentrations, respectively, brain accumulation is only 2.9-, 1.5-, and 2.4-fold, respectively,

in heterozygous mice that have 50% of the levels of Abcb1 and Abcg2 relative to wild-type mice

[66]. This indicates that the transport activity of each Abc transporter is apparently large enough

to sufficiently efflux either drug.

In Vitro and Knockout Mouse Studies To Assess the Role of OCT1 as a

Transporter of Imatinib and Sorafenib

OCT1 has been discussed most intensely and controversially as a transporter of imatinib [63,67–

71], and is given here as an example to demonstrate the difficulty of reconciling conflicting data

and the urgent need for standardized cell-based assays. Several in vitro studies using OCT1-

transfected HEK cells, which express functional OCT1 at high levels, detected only a 1.0- to 1.2-

fold increase of cellular imatinib uptake [63,67,69,70], indicating that OCT1 is not involved in

imatinib uptake. This is supported by in vivo studies in knockout mice in which the genetic

deficiency of Oct1 did not affect imatinib disposition [70]. Even so, another study detected a

1.7-fold increase of imatinib accumulation when OCT1 was overexpressed in the CML cell line

KCL22, which was interpreted as OCT1-mediated uptake [68]. However, according to recent

recommendations [52] this should not be considered as significant transport. In addition,

imatinib uptake assays using ‘specific’ inhibitors to inhibit OCT1 function [72,73] should be

interpreted cautiously because those inhibitors are not specific for OCT1 and may also interfere

with non-OCT1-dependent imatinib uptake mechanisms [69,70].

OCT1 has also been suggested as uptake transporter for sorafenib in some studies [74,75], but

not in another study [76]. As for imatinib, accumulation ratios in OCT1-transfected HEK cells or

OCT1-expressing oocytes did not exceed 1.6, demonstrating that OCT1 does not transport

sorafenib. The accumulation ratio of sorafenib by OATP1B1- and OATP1B3-transfected cells is

also <1.5, and, consistent with this in vitro result, OATP1B1 and OATP1B3 had no role in in vivo

disposition of sorafenib in transgenic mice expressing human OATP1B1 and OATP1B3 [76,77].

By contrast, OATP1B1 and OATP1B3 both transport the sorafenib glucuronide, and therefore

are involved in the elimination of this inactive metabolite [77,78]. When OATP1B function is

compromised, higher sorafenib glucuronide levels may result, leading to an increased risk for

toxicity [77,78].

Factors Affecting Interindividual Variability of Transporter Function and

Expression in Humans

The interindividual variability of transporter function and expression is the consequence of

multiple factors, including epigenetic factors, genetic variants, and regulatory processes, as

well as non-genetic factors such as sex, age, organ function, and drug treatments [79–81]

(Figure 4). It is well documented that the concomitant administration of two or more drugs may

lead to transporter-mediated drug–drug interactions, thereby leading to altered transporter

function or expression [42,82,83]. For example, vandetanib inhibits ABCB1 and OCT2 and leads

Trends in Pharmacological Sciences, November 2016, Vol. 37, No. 11 917 Sources of interindividual variability in drug response Non-c factors Genec variaon Epigenecs Regulaon Histone Lipid Weight Very rare Transcriponal Age Diet modificaon miRNA environment Circadian Sex Starvaon rhythm Pregnancy Disease Exercise Renal Rare Post- funcon Smoking Cardiov. transcriponal funcon Occupaonal Common Trafficking G.I. Drugs Post- funcon Hepac DNA methylaon funcon translaonal

Transporter funcon and expression Transporter

expression Drug uptake CC CT TT Drug concentraon Transporter Immuno- Tissue genotype localizaon microarray

Pharmacokinecs Concentraon Time

Drug response Drug efficacy Drug resistance Drug toxicity

Figure 4. Factors Affecting the Interindividual Variability of Transporter Function and Expression. Variability of transporter function and expression can be

explained by several different factors and processes, such as (i) non-genetic factors, including age, sex, organ function, underlying disease, and concomitant medications

potentially leading to drug–drug interactions; (ii) genetic variation, whereby very rare variants with minor-allele frequencies <0.1% account for approximately 83% of

variants, rare variants with minor-allele frequencies <1% account for 10%, and common variants with minor-allele frequencies >1% account for only 7% of total variants

[27]; (iii) epigenetic factors including DNA methylation, miRNA, and histone modifications; and (iv) regulatory processes affecting protein expression at different levels.

Altered transporter function and expression may lead to altered drug pharmacokinetics and thereby drug response. Abbreviation: G.I., gastrointestinal.

to higher plasma concentrations of simultaneously administered drugs whose elimination

depends on ABCB1, such as dabigatran or digoxin, or on OCT2, such as metformin [42].

Role of Genetic Transporter Variants in TKI Pharmacokinetics and Response in Humans

Pharmacokinetic studies with Abcb1 and Abcg2 knockout mice clearly show an important role

of these transporters for distribution of most TKIs (Table 3). Therefore, many studies have been

performed to elucidate the impact of genetic variants in human ABCB1 and ABCG2 on TKI

pharmacokinetics and TKI outcome by analyzing variants that have been studied in detail before

and that have been associated with drug response or altered expression [84] ((Table S1 in the

supplemental information online): for ABCB1, these are the three coding-sequence variants

1236C > T (rs1128503, p.G412G), 2677G > T/A (rs2032582, p.A893S/T), and 3435C > T

(rs1045642, p.I1145I), which occur at high allele frequencies in different ethnicities and create

a common haplotype [47,84]; for ABCC2, it is the promoter variant 24C > T (rs717620)

918 Trends in Pharmacological Sciences, November 2016, Vol. 37, No. 11 Table 4. Effect of Genetic Variants on Pharmacokinetics and Outcome for Selected TKIs Transporter Genetic Variants Number of Ethnicity Effects on Drug PKa Significant Effects on Drug Significant Refs Tested (Common Patients Association Treatment Outcomea Association Name) with PK with Outcome Imatinib

ABCB1 1236C > T 87 Caucasian $ MMR No [153]

1236C > T 33 Caucasian $ response No [154]

2677G > T/A 96 Egyptian 2677TT: Yes [155] " CCR, MR

3435C > T 82 Caucasian $ Oral Cl No [31]

3435C > T 70 Iranian 3435CC: Yes [156] # CR

1236C > T 82 Korean $ Cmin No $ MMR No [33] 3435C > T

2677G > T/A 84 Caucasian 3435TT: Yes [90] 3435C > T " time to MMR

2677G > T/A 111 Indian $ Cmin No $ Thrombocytopenia No [157] 3435C > T

1236C > T 22 Australian Haplotype TTT: Yes [158] 2677G > T/A " Cl 3435C > T

1236C > T 90 Caucasian 1236TT: Yes 1236TT: " MMR Yes [159] 2677G > T/A " Cmin 2677G: # response 3435C > T Haplotype CGC:

Trends # MMR #

1236C > T 229 Mainly Caucasian 3435TT: OS Yes [160] in

Pharmacological 2677G > T/A 3435C > T

1236C > T 52 Chinese 1236TT: Yes [161] 2677G > T/A " resistance

3435C > T 2677GT: Sciences, " resistance 3435CC:

# resistance November

1236C > T 67 Japanese $ Cmin No $ MMR No [32] 2677G > T/A

2016, 3435C > T

Vol. 1236C > T 46 Caucasian 1236TT, 1236CT: Yes [162] " 2677G > T/A resistance 37,

3435C > T 2677TT: No. " resistance

11

3435TT: " 919 resistance

65 Caucasian Yes [163] 920 Table 4. (continued)

Transporter Genetic Variants Number of Ethnicity Effects on Drug PKa Significant Effects on Drug Significant Refs Trends Tested (Common Patients Association Treatment Outcomea Association Name) with PK with Outcome in

Pharmacological 1236C > T Haplotype TGC: 2677G > T/A " resistance 3435C > T

1236C > T 34 Japanese 3435CC: " Cl Yes [164]

2677G > T/A Sciences, 3435C > T

1236C > T 118 Brazilian Haplotype 1236CT/ Yes [165]

November 2677G > T/A 2677GT/3435CT: 3435C > T " MMR " 1236C > T 189 Caucasian 3435CC: CMR Yes [166] 2016, 2677G > T/A

Vol. 3435C > T

37, 1236C > T 215 Malaysian Haplotype CGC: Yes [167]

No. 2677G > T/A " resistance

11 3435C > T

1236C > T 60 Caucasian $ Cl No $ EFS No [89] 2677G > T/A 3435C > T

1236C > T 284 Korean $ Cmin No $ PFS No [168] 2677G > T/A 3435C > T

1236C > T 48 Chinese 1236TT: Yes [169] 2677G > T/A # CR 3435C > T

1236C > T 38 Asian Effect on Cl Yes $ Response No [170] 2677G > T/A 3435C > T Haplotype with OCT1 variants

ABCC2 24C > T 67 Japanese $ Cmin No $ MMR No [32] 24C > T 215 Malaysian Haplotype TGT: Yes [171] 1249G > A # CR, MR 3972C > T

ABCG2 421C > A 82 Caucasian $ Oral Cl No [31]

421C > A 67 Japanese 421CA/AA: " Cmin Yes $ MMR No [32] 421C > A 229 Mainly Caucasian " MMR Yes [172]

421C > A 15 Japanese $ Conc. in leukocytes No [17]

421C > A 34 Japanese $ Cl No [164]

421C > A 111 Indian $ Cmin No $ Thrombocytopenia No [157] Table 4. (continued) Transporter Genetic Variants Number of Ethnicity Effects on Drug PKa Significant Effects on Drug Significant Refs Tested (Common Patients Association Treatment Outcomea Association Name) with PK with Outcome

421C > A 100 Korean $ Cmin No $ CR No [33] 34G > A 229 Mainly Caucasian 34GG: # CCR Yes [160] 421C > A 421CA/CC: # MMR, CMR

34G > A 189 Caucasian $ MCR, CCR, No [166] 421C > A MMR, CMR

34G > A 118 Brazilian $ Response No [173] 421C > A

34G > A 215 Malaysian Haplotype AA: Yes [167] 421C > A " response

34G > A 284 Korean $ Cmin No 421 CC/CA: Yes [168] 421C > A # PFS

rs12505410 T > G 105 Caucasian Haplotype GG: Yes [174] rs2725252 T > G " rate of MMR (intron variants)

OCT1/SLC22A1 181C > T 32 Caucasian No $ MMR No [175]

480G > C 229 Caucasian 480GG: Yes [160] " rate of LOR

480G > C 33 Caucasian $ Response No [154]

1022C > T 15 Japanese $ Conc. in No [17] Trends leukocytes, Css $

1022C > T 34 Japanese Cl No [164] in

Pharmacological 1022C > T 82 Asian $ Cmin No $ CCR, MMR No [33] 1201G > A 132 Caucasian 1201GA: " MMR Yes [176]

480G > C 118 Brazilian $ Response No [173]

1222A > G Sciences,

480G > C 84 Caucasian $ Cmin No $ Response No [90] 1222A > G

November 156T > C 67 Japanese $ Cmin No 1222GG: " MMR Yes [32] 480G > C

1022C > T 2016, 1222A > G

Vol. 181C > T 74 Caucasian $ Cl No [67]

37, 1393G > A

No. 181C > T 167 Caucasian 1222AA/AG: Yes [177]

11 480G > C # median EFS, OS

921 922 Table 4. (continued)

Transporter Genetic Variants Number of Ethnicity Effects on Drug PKa Significant Effects on Drug Significant Refs Trends Tested (Common Patients Association Treatment Outcomea Association Name) with PK with Outcome in

Pharmacological 1222A > G 1503G > A

181C > T 136 Caucasian $ Response No [178] 262T > C 659G > T

Sciences, 1022C > T 1201G > A

1222A > G November 1258delATG 1393G > A

2016, 181C > T 65 Caucasian 1795AA: " overall Yes [163] 1222A > G inadequate response

Vol. 1258delATG

37, 1795G > A

No. 181C > T 189 Caucasian Combination of Yes [166]

11 480G > C variants: " MMR 848C > T 859C > G 1022C > T 1258delATG

181C > T 60 Caucasian 480CG/GG: # Cl, " Cmin Yes 480CG/GG: # EFS Yes [89] 480G > C 848C > T 859C > G 1258delATG

23 variants including 181C > T 336 Caucasian 1258delATG: " failure Yes [179] 480G > C 1222A > G 1258delATG

1022C > T 111 Indian $ Cmin No $ Thrombocytopenia No [157] 1222A > G 1386C > A

1201A > G 153 Caucasian 1275_1276 + 6T: Yes [180] 1222A > G " time to MMR 1239G > A 1239delATG 1275_1276 + 6T

Haplotype including 1222A > G 38 Asian effect on Cl Yes $ Response No [170] and ABCB1 variants

OCTN1/SLC22A4 1507C > T 189 Caucasian 1507TT: # MMR, $ CMR Yes [166]

1507C > T 54 Caucasian 1507CC/CT: " TTP Yes [181]

54 Caucasian 207GC/GG: " TTP Yes [181] Table 4. (continued) Transporter Genetic Variants Number of Ethnicity Effects on Drug PKa Significant Effects on Drug Significant Refs Tested (Common Patients Association Treatment Outcomea Association Name) with PK with Outcome 207C > G 2087G > C

OCTN2/SLC22A5 38T > C 94 Caucasian $ Css No [137] 516A > C

OATP1A2/SLCO1A2 361G > A 34 Japanese -361GG: # reduced Cl Yes [182]

516A > C 189 Caucasian $ MMR No [166]

361G > A 118 Brazilian $ CMR, MMR No [183] 516A > C

OATP1B3/SLCO1B3 334T > G 67 Japanese $ Cmin No $ MMR No [32] 334T > G 15 Japanese 334TT: " conc. in leukocytes Yes [17]

334T > G 86 Egyptian $ HR, CR, MR No [184]

334T > G 118 Brazilian 334TT: # CMR, $ MMR Yes [183]

Sorafenib

ABCB1 3435C > T 33 Japanese " Risk for grade 3 skin rash Yes [185]

2677G > T/A 54 Caucasian $ AUC No " Risk for grade 3 toxicity Yes [34] 3435C > T

ABCC2 24C > T 33 Japanese "Risk for grade 3 skin rash Yes [185]

ABCG2 34G > A 54 Caucasian $ AUC No " Risk for grade 3 toxicity Yes [34] 421C > A Trends 1143C > T

" in SLC15A2 rs2257212 C > T 174 Koreans PFS for CT and TT vs CC Yes [186]

Pharmacological Sunitinib

ABCB1 3435C > T 31 Asian 3435CC: " Cmin Yes 3435CC: " risk for rash, Yes [187] mucositis, disease progression

Sciences, 2677G > T/A 92 Caucasian $ AUC No [188] 3435C > T

November 1236C > T 333 Mainly Caucasian Haplotype CGT: Yes [35] 2677G > T/A " risk for HFS, " PFS 3435C > T

2016, 1236C > T 136 Caucasian Haplotype TCG: " PFS Yes [189]

Vol. 2677G > T/A

37, 3435C > T

No. 1236C > T 19 Japanese Haplotype TTT: No [190]

11 2677G > T/A $ AUC

3435C > T 923 101 Caucasian $ PFS, OS, toxicity No [191] 924 Table 4. (continued)

Transporter Genetic Variants Number of Ethnicity Effects on Drug PKa Significant Effects on Drug Significant Refs Trends Tested (Common Patients Association Treatment Outcomea Association Name) with PK with Outcome in

Pharmacological 1236C > T 2677G > T/A 3435C > T

1236C > T 88 Mainly Caucasian 1236TT: # PFS, OS Yes [192]

2677G > T/A Sciences, 3435C > T

1236C > T 96 Mainly Caucasian 1236TT: " time to Yes [193]

November 2677G > T/A dose reduction 3435C > T $ 1236C > T 65 Korean Risk for grade 3 toxicity No [194] 2016, 2677G > T/A

Vol. 3435C > T

37, 1236C > T 114 Mainly Caucasian 3435TT: " Cl Yes [195]

No. 2677G > T/A

11 3435C > T

1236C > T 97 Asian 1236T, 2677T, 3435T, haplotype Yes [196] 2677G > T/A TTT: # risk for neutropenia 3435C > T haplotype TTT: # PFS, OS

ABCG2 34G > A 136 Caucasian $ PFS No [189]

421C > A 19 Japanese 421CA, 421AA: Yes [190,197] " higher AUC, oral Cl

421C > A 101 Caucasian $ PFS, OS, toxicity No [191]

421C > A 65 Korean 421AA: " risk for grade 3 toxicity Yes [194]

421C > A 97 Asian 421AA: # risk for neutropenia Yes [196]

421C > A 219 Japanese 421A: " risk for thrombocytopenia Yes [198]

15622C > T 1143C > T 219 Caucasian Haplotype TT: " risk for Yes [199] grade 2 toxicity

15622C > T 333 Mainly Caucasian 421AA: " risk for hypertension Yes [35] 421C > A

15622C > T 114 Mainly Caucasian $ Cl No [195] 421 C > A 1143 C > T

34G > A 92 Caucasian 421AA: " higher AUC Yes [188] 421C > A 1143C > T

a Symbols and abbreviations: $, no effect; ", increased; #, decreased; C, concentration; Cmin , minimum concentration; Css, concentration at steady-state; CCR, complete cytogenetic response; Cl, clearance; CMR, complete molecular response; CR, cytogenetic response; EFS, event-free survival; HR, hematological response; LOR loss of response; MCR, major cytogenetic response; MMR, major molecular response; MR, molecular response; OS, overall survival; PK, pharmacokinetics; TTP: time to progression.

[84,85]; for ABCG2, the intron variants 15622C > T (rs559306529) and 1143C > T

(rs2622604), and the two coding sequence variants 34G > A (rs2231137, p.V12 M),

421C > A (rs2231142, p.Q141K), have been intensely studied [84,86].

Table 4 summarizes studies investigating the effects of transporter genetic variants on the

pharmacokinetics and outcome of the selected TKIs imatinib, sorafenib, and sunitinib. With

respect to the investigated ABC transporters, the clinical relevance of common genetic variants

in humans for TKI pharmacokinetics and TKI resistance appears to be limited despite the

convincing functional evidence of ABC transporters in TKI transport in vitro and in knockout

mouse experiments. Based on recent studies [27,87] it is likely that rare rather than common

transporter genetic variants will have much larger effects on TKI disposition, and their analyses

should therefore be the focus of future research. The reasons for the difficulty in translating the

preclinical in vitro and knockout mouse studies on ABC transporters into the clinical context are

manifold. Interpretation of pharmacogenetic association studies can be difficult owing to the

complex variables involved, such as the small sample sizes in most studies, different ethnicities,

different genetic variants tested from one study to another, studying only common genetic

variants, and further confounders such as underlying disease and the presence of transporters

with redundant function. In addition, as detailed in the next paragraph, differential regulation of

ABC transporter expression may be an important reason for variable TKI responses.

Notably, despite the limited role of uptake transporters in the in vitro uptake of imatinib, in

addition to the other TKIs (Table 3), several pharmacogenetic association studies have been

performed to investigate the impact of common and rare OCT1/SLC22A1 genetic variants on

imatinib pharmacokinetics and response, as well as few studies on OCTN1/SLC22A4, OCTN2/

SLC22A5, OATP1A2/SLCO1A2, and OATP1B3/SLCO1B3 (Table 4). Here again, most studies

have focused on variants that were previously shown to have consequences for the transport of

prototypic substrates as well as for transporter expression and drug disposition [50,71,88]

(Table S1). Similarly to findings for ABC transporters, the results of these clinical studies are

ambiguous and sometimes contradictory. For example, one study detected a significant

association between the genetic variant c.480G > C (Leu160Phe) in OCT1/SLC22A1 and

minimum imatinib plasma concentrations [89], while others did not [32,90]. No further variants

were associated with imatinib pharmacokinetic parameters. Based on the notion that cellular

imatinib uptake is independent from OCT1 (Table 3), correlations between OCT1/SLC22A1

genetics and clinical outcome of imatinib therapy are surprising, and suggest that the tested

genetic variants are not the primary cause underlying imatinib failure. Instead, it can be

postulated that those OCT1/SLC22A1 variants may be genetically linked to variants in other

genes relevant for imatinib uptake and subsequent action [70].

Role of Transporter Regulation for TKI Resistance

Transporter expression in non-tumor tissue has been investigated in great detail; however, data

on transporter expression, particularly at the protein level, in primary tumor or metastatic tissue

are sparse. Transporter expression may change during progression from normal tissue to tumor

tissue (Figure 3). For example, in histologically normal human liver, genetic variants in SLC22A1/

OCT1 cause interindividual differences in OCT1 expression [49]. However, in hepatocellular

carcinoma, epigenetic regulation of OCT1 is the predominant mechanism, and DNA methylation

of the SLC22A1/OCT1 gene leads to significant downregulation of OCT1 mRNA and protein in

the tumor tissue [91]. Furthermore, epigenetic regulation by miRNAs has been identified as an

important regulatory mechanism of ABC transporter expression in hepatocellular carcinoma and

other cancer types [92–94].

In addition to genetic variation and epigenetics, numerous studies have identified various

regulatory mechanisms that affect transporter expression at different levels (Figure 4).

Trends in Pharmacological Sciences, November 2016, Vol. 37, No. 11 925

For example, at the transcriptional level, many drugs and xenobiotics, such as rifampin, Outstanding Questions

phenobarbital, and carbamazepine, induce hepatic expression of ABCB1 and ABCG2, as well Which membrane transporters medi-

ate the uptake of the currently

as of OATP1B1 and OATP1B3, by binding as ligands to the nuclear receptors pregnane X

approved TKIs?

receptor (PXR) and/or constitutive androstane receptor (CAR) [95–97]. Moreover, several

pathophysiological conditions such as inflammation or hypoxia have been shown to affect

Which membrane transporters are

ABCB1 gene expression by activating stress-induced transcription factors [97,98]. At the post- involved in the uptake and efflux of

transcriptional level, miRNAs have been recognized as important regulators of ABC transporter the >130 novel TKIs currently being

evaluated in clinical trials for cancer

expression [93,94]. It may be speculated that treatment with a drug leads to a distinct cellular

and other disease entities?

miRNA expression pattern and a drug-specific ABC transporter expression pattern [94]. For

example, earlier studies showed that chronic TKI exposure may lead to induction of drug efflux

What innovative techniques in preclini-

transporters ABCB1 and ABCG2, and thereby reduced intracellular accumulation, as shown for cal and early clinical drug testing are

necessary to enhance the study of

imatinib in the target CML cells [99]. Recent studies indeed identified an association between the

drug transport processes and TKI

downregulation of particular miRNAs and concomitant upregulation of ABCG2 expression by

resistance?

imatinib treatment of leukemic K562 cells in vitro [100].

How can systems-medicine

In addition to the above-described mechanisms affecting transporter expression, rapid regula- approaches and omics technologies

be used to identify the factors which

tion of transporter activity may occur by post-translational modifications within minutes. For

affect drug transporter expression and

example, a recent study showed that phosphorylation of OCT2 by the Src family kinase Yes1 is

function and which predict tumor

essential for OCT2 transport function [101]. Inhibition of Yes1 by TKIs including dasatinib

response or resistance?

resulted in reduced OCT2 function in vitro and in vivo [101]. Other post-translational modifica-

tions include glycosylation, which is apparently required for proper trafficking of ABCB1 to the How can an improved understanding

of the molecular basis of drug resis-

plasma membrane or for protein stability, without affecting transport function [102]. Moreover,

tance help to devise rational drug com-

interacting proteins may be required for targeting of transporters to the plasma membrane [102].

binations for selected patient

Finally, the lipid environment of the membrane, in which the transporters are embedded, and populations or individual patients?

transporter–lipid interactions have been identified as important determinants of ABCB1 and

ABCG2 expression and function [103].

These numerous studies demonstrate the variety of regulatory processes by which transporter

expression and function can be modulated. However, much more work will be necessary to fully

explore all these different aspects of regulation, not only in non-tumor but also in tumor tissues,

and their role in TKI resistance.

Concluding Remarks and Future Directions

Small-molecule TKIs have emerged as the backbone of cancer therapy and the treatment of

other diseases. However, in contrast to the remarkable success of imatinib in the improvement

of outcome of patients with CML, the degree of success of other TKIs is variable, and most

achieve only moderate survival benefits. Similar to the situation when using conventional

cytotoxic chemotherapy, the occurrence of drug resistance often counteracts successful

treatment. In addition to somatic variation of the targeted tyrosine kinases, insufficient drug

exposure may significantly contribute to drug resistance. Drug uptake and efflux transporters in

enterocytes, hepatocytes, proximal tubule kidney epithelial cells, and not least in the cancer cells

themselves, may preclude sufficient intracellular TKI accumulation.

While the role of ABC efflux transporters in in vitro transport and in vivo disposition in knockout

mice has been established for most TKIs, their role in TKI disposition and occurrence of drug

resistance in humans – as assessed by pharmacogenetic association studies – is less clear. As a

consequence, future research is warranted to include not only rare variants but also a compre-

hensive definition of clinical phenotypes based on representative and valid sample sizes [104].

Moreover, the currently identified and investigated uptake transporters play apparently only a

minor role in TKI uptake (see Outstanding Questions). Therefore, novel techniques such as

thermal proteomics profiling [105,106] will be necessary to identify novel transporters engaged in

TKI uptake. Once these are identified, the contributions of non-genetic, epigenetic, and common

926 Trends in Pharmacological Sciences, November 2016, Vol. 37, No. 11

and rare genetic variants in these transporter genes on the function and expression of the

transporters need to be assessed. In this context, it will be important to study transporter

expression not only in the primary tumor but also in metastases because therapeutic responses

to TKI may differ. As recently demonstrated for clear cell renal cell carcinoma, drug transporter

expression was shared between primary tumors and metastases, indicating comparable drug

transport processes and drug effects in tumors and metastases [62]. However, this may be

different for other tumor entities and needs to be investigated.

Because cell lines often do not correspond to the conditions in primary tumors or metastases in

vivo [62], other innovative techniques such as patient-derived tumor grafts (PDx) may be used to

study drug transport processes [107]. Furthermore, novel imaging techniques such as hyper-

spectral stimulated Raman scattering [108] and PET imaging with TKIs [109] will help to track

and quantify intracellular TKI concentrations in vivo to improve TKI efficacy.

Moreover, owing to the large number of factors (non-genetic, genetic, epigenetic, regulatory)

that may affect transporter expression and function, systems-medicine approaches and omics

technologies [79] will be required for identifying novel mechanisms of drug resistance, including

drug transporters, as well as molecular signatures and genotypes that may predict tumor

response or resistance. This improved understanding of the molecular basis of drug resistance

will foster the design of rational drug combinations for selected patient populations. This is of

particular importance in view of the fact that currently >130 novel TKIs are being evaluated in

clinical trials (Figure 1B), not only for the treatment of cancer but also for other disease entities.

It can be expected that these strategies of identifying drug transporters and resistance mecha-

nisms will also be important when inhibitors for kinases other than tyrosine kinases are clinically

developed [110], as well as for novel classes of anticancer agents such as small-molecule checkpoint inhibitors [111].

Acknowledgments

This work was supported in part by the Robert-Bosch Foundation, Stuttgart, Germany, the Interfaculty Centre for

Pharmacogenomics and Pharma Research (ICEPHA) Grant Tübingen–Stuttgart, Germany, the Bundesministerium für

Bildung und Forschung, Germany (LiSyM 031L0037) and European Commission Horizon 2020 UPGx grant (668353). The

authors thank Bernd Borstel for drafting the figures.

Supplemental Information

Supplemental information associated with this article can be found online at http://dx.doi.org/10.1016/j.tips.2016.08.003.

References

1. Manning, G. et al. (2002) The protein kinase complement of the 9. Hanahan, D. (2014) Rethinking the war on cancer. Lancet 383,

human genome. Science 298, 1912–1934 558–563

2. Rask-Andersen, M. et al. (2014) Advances in kinase targeting: 10. Holohan, C. et al. (2013) Cancer drug resistance: an evolving

current clinical use and clinical trials. Trends Pharmacol. Sci. 35, paradigm. Nat. Rev. Cancer 13, 714–726

604–620

11. Juchum, M. et al. (2015) Fighting cancer drug resistance: Oppor-

3. Roskoski, R., Jr (2015) A historical overview of protein kinases tunities and challenges for mutation-specific EGFR inhibitors.

and their targeted small molecule inhibitors. Pharmacol. Res. Drug Resist. Updat. 20, 12–28

100, 1–23

12. Günther, M. et al. (2016) Lung cancer: EGFR inhibitors with

4. Hanahan, D. and Weinberg, R.A. (2011) Hallmarks of cancer: the low nanomolar activity against a therapy-resistant L858R/

next generation. Cell 144, 646–674 T790M/C797S mutant. Angew. Chem. Int. Ed. Engl. 55,

10890–10894

5. Levitzki, A. (2013) Tyrosine kinase inhibitors: views of selectivity,

sensitivity, and clinical performance. Annu. Rev. Pharmacol. Tox- 13. Wu, P. et al. (2015) FDA-approved small-molecule kinase inhib-

icol. 53, 161–185 itors. Trends Pharmacol. Sci. 36, 422–439

6. Murray, B.W. and Miller, N. (2015) Durability of kinase-directed 14. Baccarani, M. et al. (2015) A review of the European Leukemia-

therapies – a network perspective on response and resistance. Net recommendations for the management of CML. Ann. Hem-

Mol. Cancer Ther. 14, 1975–1984 atol. 94 (Suppl. 2), S141–S147

7. Druker, B.J. (2008) Translation of the Philadelphia chromosome 15. Gillet, J.P. and Gottesman, M.M. (2010) Mechanisms of multi-

into therapy for CML. Blood 112, 4808–4817 drug resistance in cancer. In Methods in Molecular Biology

(Zhou, J., ed.), pp. 47–76, Springer

8. Talpaz, M. et al. (2013) Re-emergence of interferon-alpha in the

treatment of chronic myeloid leukemia. Leukemia 27, 803–812

Trends in Pharmacological Sciences, November 2016, Vol. 37, No. 11 927

16. Mahon, F.X. (2009) Pharmacologic monitoring and determinants 39. Yu, H. et al. (2014) Practical guidelines for therapeutic drug

of intracytoplasmic drug levels. Best Pract. Res. Clin. Haematol. monitoring of anticancer tyrosine kinase inhibitors: focus on

22, 381–386 the pharmacokinetic targets. Clin. Pharmacokinet. 53, 305–325

17. Nambu, T. et al. (2011) Association of SLCO1B3 polymorphism 40. Widmer, N. et al. (2014) Review of therapeutic drug monitoring of

with intracellular accumulation of imatinib in leukocytes in patients anticancer drugs part two – targeted therapies. Eur. J. Cancer

with chronic myeloid leukemia. Biol. Pharm. Bull. 34, 114–119 50, 2020–2036

18. White, D.L. et al. (2013) Proton pump inhibitors significantly 41. Terada, T. et al. (2015) Management of dose variability and side

increase the intracellular concentration of nilotinib, but not ima- effects for individualized cancer pharmacotherapy with tyrosine

tinib in target CML cells. Leukemia 27, 1201–1204 kinase inhibitors. Pharmacol. Ther. 152, 125–134

19. Bouchet, S. et al. (2013) From in vitro to in vivo: intracellular 42. Wulkersdorfer, B. et al. (2016) Pharmacokinetic aspects of vas-

determination of imatinib and nilotinib may be related with clinical cular endothelial growth factor tyrosine kinase inhibitors. Clin.

outcome. Leukemia 27, 1757–1759 Pharmacokinet. 55, 47–77

20. Drenberg, C.D. et al. (2013) Integrating clinical pharmacology 43. Hediger, M.A. et al. (2013) The ABCs of membrane transporters

concepts in individualized therapy with tyrosine kinase inhibitors. in health and disease (SLC series): Introduction. Mol. Aspects

Clin. Pharmacol. Ther. 93, 215–219 Med. 34, 95–107

21. Josephs, D.H. et al. (2013) Clinical pharmacokinetics of tyrosine 44. Nigam, S.K. (2015) What do drug transporters really do? Nat.

kinase inhibitors: implications for therapeutic drug monitoring. Rev. Drug Discov. 14, 29–44

Ther. Drug Monit. 35, 562–587

45. Cesar-Razquin, A. et al. (2015) A call for systematic research on

22. de Wit, D. et al. (2015) Individualized dosing of tyrosine kinase solute carriers. Cell 162, 478–487

inhibitors: are we there yet? Drug Discov. Today 20, 18–36

46. Zimmerman, E.I. et al. (2016) Multi-kinase inhibitors induce cuta-

23. Bazzoli, C. et al. (2010) Intracellular pharmacokinetics of antire- neous toxicity through OAT6-mediated uptake and MAP3K7-

troviral drugs in HIV-infected patients, and their correlation with driven cell death. Cancer Res. 76, 117–126

drug action. Clin. Pharmacokinet. 49, 17–45

47. Wolking, S. et al. (2015) Impact of genetic polymorphisms of

24. Giacomini, K.M. et al. (2010) Membrane transporters in drug ABCB1 (MDR1, P-glycoprotein) on drug disposition and potential

development. Nat. Rev. Drug Discov. 9, 215–236 clinical implications: update of the literature. Clin. Pharmacokinet.

54, 709–735

25. Cascorbi, I. and Schwab, M. (2016) Epigenetics in drug

response. Clin. Pharmacol. Ther. 99, 468–470 48. Durmus, S. et al. (2015) Apical ABC transporters and cancer

chemotherapeutic drug disposition. Adv. Cancer Res. 125, 1–41

26. Ingelman-Sundberg, M. and Cascorbi, I. (2016) Pharmacoge-

nomic or -epigenomic biomarkers in drug treatment: 49. Nies, A.T. et al. (2009) Expression of organic cation transporters

Two sides of the same medal? Clin. Pharmacol. Ther. 99, OCT1 (SLC22A1) and OCT3 (SLC22A3) is affected by

478–480 genetic factors and cholestasis in human liver. Hepatology 50,

1227–1240

27. Kozyra, M. et al. (2016) Rare genetic variants in cellular trans-

porters, metabolic enzymes, and nuclear receptors can be 50. Nies, A.T. et al. (2011) Organic cation transporters (OCTs,

important determinants of interindividual differences in MATEs), in vitro and in vivo evidence for the importance in drug

drug response. Genet. Med. http://dx.doi.org/10.1038/gim. therapy. Handb. Exp. Pharmacol. 201, 105–167

2016.33

51. Nies, A.T. et al. (2013) Genetics is a major determinant of expres-

28. Yaish, P. et al. (1988) Blocking of EGF-dependent cell prolifera- sion of the human hepatic uptake transporter OATP1B1, but not

tion by EGF receptor kinase inhibitors. Science 242, 933–935 of OATP1B3 and OATP2B1. Genome Med. 5, 1

29. Herbrink, M. et al. (2015) Variability in bioavailability of small 52. Brouwer, K.L. et al. (2013) In vitro methods to support trans-

molecular tyrosine kinase inhibitors. Cancer Treat. Rev. 41, porter evaluation in drug discovery and development. Clin. Phar-

412–422 macol. Ther. 94, 95–112

30. Zanger, U.M. and Schwab, M. (2013) Cytochrome P450 53. Eadie, L.N. et al. (2014) Interaction of the efflux transporters

enzymes in drug metabolism: regulation of gene expression, ABCB1 and ABCG2 with imatinib, nilotinib, and dasatinib. Clin.

enzyme activities, and impact of genetic variation. Pharmacol. Pharmacol. Ther. 95, 294–306

Ther. 138, 103–141

54. Mendes, P. et al. (2015) Fitting transporter activities to cellular

31. Gardner, E.R. et al. (2006) Association of enzyme and transporter drug concentrations and fluxes: why the bumblebee can fly.

genotypes with the pharmacokinetics of imatinib. Clin. Pharma- Trends Pharmacol. Sci. 36, 710–723

col. Ther. 80, 192–201

55. Matsson, P. et al. (2015) Quantifying the impact of transporters on

32. Takahashi, N. et al. (2010) Influence of CYP3A5 and drug trans- cellular drug permeability. Trends Pharmacol. Sci. 36, 255–262

porter polymorphisms on imatinib trough concentration and

56. Matsson, P. et al. (2016) The need for speed-kinetic limits of drug

clinical response among patients with chronic phase chronic

transporters. Trends Pharmacol. Sci. 37, 243–245

myeloid leukemia. J. Hum. Genet. 55, 731–737

57. Mendes, P. et al. (2016) Response to ‘The Need for Speed’, by

33. Seong, S.J. et al. (2013) Influence of enzyme and transporter

Matsson et al. Trends Pharmacol. Sci. 37, 245–246

polymorphisms on trough imatinib concentration and clinical

58. Graber-Maier, A. et al. (2010) A new intestinal cell culture model

response in chronic myeloid leukemia patients. Ann. Oncol.

to discriminate the relative contribution of P-gp and BCRP

24, 756–760

on transport of substrates such as imatinib. Mol. Pharm. 7,

34. Boudou-Rouquette, P. et al. (2012) Early sorafenib-induced tox- 1618–1628

icity is associated with drug exposure and UGTIA9 genetic

59. Gnoth, M.J. et al. (2010) In vitro to in vivo comparison of the

polymorphism in patients with solid tumors: a preliminary study.

substrate characteristics of sorafenib tosylate toward P-glyco-

PLoS ONE 7, e42875

protein. Drug Metab. Dispos. 38, 1341–1346

35. Diekstra, M.H. et al. (2015) CYP3A5 and ABCB1 polymorphisms

60. Wilding, J.L. and Bodmer, W.F. (2014) Cancer cell lines for drug

as predictors for sunitinib outcome in metastatic renal cell carci-

discovery and development. Cancer Res. 74, 2377–2384

noma. Eur. Urol. 68, 621–629

61. Varley, K.E. et al. (2013) Dynamic DNA methylation across

36. Diekstra, M.H. et al. (2016) Sunitinib-induced hypertension in

diverse human cell lines and tissues. Genome Res. 23, 555–567

CYP3A4 rs4646437 A-allele carriers with metastatic renal cell

62. Winter, S. et al. (2016) Methylomes of renal cell lines and tumors

carcinoma. Pharmacogenomics J. http://dx.doi.org/10.1038/

tpj.2015.100 or metastases differ signi cantly with impact on pharmacogenes.

Sci. Rep. 6, 29930

37. Baker, S.D. and Hu, S. (2009) Pharmacokinetic considerations

63. Hendrickx, R. et al. (2013) Identification of novel substrates and

for new targeted therapies. Clin. Pharmacol. Ther. 85, 208–211

structure activity relationship of cellular uptake mediated by the

38. Widmer, N. et al. (2008) Relationship of imatinib-free plasma

human organic cation transporters 1 and 2 (hOCT1 and hOCT2).

levels and target genotype with efficacy and tolerability. Br. J.

J. Med. Chem. 56, 7232–7242

Cancer 98, 1633–1640

928 Trends in Pharmacological Sciences, November 2016, Vol. 37, No. 11

64. Cui, Y. et al. (2003) Detection of the human organic anion trans- 88. Eechoute, K. et al. (2011) Drug transporters and imatinib

porters SLC21A6 (OATP2) and SLC21A8 (OATP8) in liver and treatment: implications for clinical practice. Clin. Cancer Res.

hepatocellular carcinoma. Lab. Invest. 83, 527–538 17, 406–415

65. Wisniewski, J.R. et al. (2016) In-depth quantitative analysis and 89. Di Paolo, A. et al. (2014) The c.480C > G polymorphism of

comparison of the human hepatocyte and hepatoma cell line hOCT1 influences imatinib clearance in patients affected by

HepG2 proteomes. J. Proteomics 136, 234–247 chronic myeloid leukemia. Pharmacogenomics J. 14, 328–335

66. Tang, S.C. et al. (2013) Impact of P-glycoprotein (ABCB1) and 90. Vine, J. et al. (2014) Polymorphisms in the human organic cation

breast cancer resistance protein (ABCG2) gene dosage on transporter and the multidrug resistance gene: correlation with

plasma pharmacokinetics and brain accumulation of dasatinib, imatinib levels and clinical course in patients with chronic myeloid

sorafenib, and sunitinib. J. Pharmacol. Exp. Ther. 346, 486–494 leukemia. Leuk. Lymphoma 55, 2525–2531

67. Hu, S. et al. (2008) Interaction of imatinib with human organic ion 91. Schaeffeler, E. et al. (2011) DNA methylation is associated with

carriers. Clin. Cancer Res. 14, 3141–3148 downregulation of the organic cation transporter OCT1

(SLC22A1) in human hepatocellular carcinoma. Genome Med.

68. Wang, L. et al. (2008) Expression of the uptake drug transporter

3, 82

hOCT1 is an important clinical determinant of the response to

imatinib in chronic myeloid leukemia. Clin. Pharmacol. Ther. 83, 92. Borel, F. et al. (2012) Adenosine triphosphate-binding cassette

258–264 transporter genes up-regulation in untreated hepatocellular

carcinoma is mediated by cellular microRNAs. Hepatology 55,

69. Burger, H. et al. (2013) Can ‘specific’ OCT1 inhibitors be used to

821–832

determine OCT1 transporter activity toward imatinib? Blood 121,

4965–4966 93. Chen, Z. et al. (2016) Mammalian drug efflux transporters of the

ATP binding cassette (ABC) family in multidrug resistance: A

70. Nies, A.T. et al. (2014) Cellular uptake of imatinib into leukemic

review of the past decade. Cancer Lett. 370, 153–164

cells is independent of human organic cation transporter 1

(OCT1). Clin. Cancer Res. 20, 985–994 94. Haenisch, S. et al. (2014) MicroRNAs and their relevance to ABC

transporters. Br. J. Clin. Pharmacol. 77, 587–596

71. Watkins, D.B. et al. (2015) OCT1 and imatinib transport in CML: is

it clinically relevant? Leukemia 29, 1960–1969 95. Burk, O. (2009) Nuclear receptor-mediated regulation of drug

transporters. In Nuclear Receptors in Drug Metabolism (Xie, W.,

72. Thomas, J. et al. (2004) Active transport of imatinib into and out of

ed.), pp. 111–146, John Wiley & Sons

cells: implications for drug resistance. Blood 104, 3739–3745

96. Klaassen, C.D. and Aleksunes, L.M. (2010) Xenobiotic, bile acid,

73. White, D.L. et al. (2006) OCT-1-mediated influx is a key determi-

and cholesterol transporters: function and regulation. Pharmacol.

nant of the intracellular uptake of imatinib but not nilotinib

Rev. 62, 1–96

(AMN107): reduced OCT-1 activity is the cause of low in vitro

sensitivity to imatinib. Blood 108, 697–704 97. Miller, D.S. (2015) Regulation of ABC transporters blood-brain

barrier: the good, the bad, and the ugly. Adv. Cancer Res. 125,

74. Swift, B. et al. (2013) Sorafenib hepatobiliary disposition: mech-

43–70

anisms of hepatic uptake and disposition of generated metab-

olites. Drug Metab. Dispos. 41, 1179–1186 98. Staud, F. et al. (2010) Expression and function of p-glycoprotein

in normal tissues: effect on pharmacokinetics. Methods Mol. Biol.

75. Herraez, E. et al. (2013) The expression of SLC22A1 variants may

596, 199–222

affect the response of hepatocellular carcinoma and cholangio-

carcinoma to sorafenib. Hepatology 58, 1065–1073 99. Burger, H. et al. (2005) Chronic imatinib mesylate exposure leads

to reduced intracellular drug accumulation by induction of the

76. Hu, S. et al. (2009) Interaction of the multikinase inhibitors sor-

ABCG2 (BCRP) and ABCB1 (MDR1) drug transport pumps.

afenib and sunitinib with solute carriers and ATP-binding cassette

Cancer Biol. Ther. 4, 747–752

transporters. Clin. Cancer Res. 15, 6062–6069

100. Turrini, E. et al. (2012) MicroRNA profiling in K-562 cells under

77. Zimmerman, E.I. et al. (2013) Contribution of OATP1B1 and

imatinib treatment: influence of miR-212 and miR-328 on ABCG2

OATP1B3 to the disposition of sorafenib and sorafenib-glucuro-

expression. Pharmacogenet. Genomics. 22, 198–205

nide. Clin. Cancer Res. 19, 1458–1466

101. Sprowl, J.A. et al. (2016) A phosphotyrosine switch regulates

78. Vasilyeva, A. et al. (2015) Hepatocellular shuttling and recircula-

organic cation transporters. Nat. Commun. 7, 10880

tion of sorafenib-glucuronide is dependent on Abcc2 Abcc3, and

Oatp1a/1b. Cancer Res. 75, 2729–2736 102. Hong, M. (2016) Biochemical studies on the structure–function

relationship of major drug transporters in the ATP-binding cas-

79. Meyer, U.A. et al. (2013) Omics and drug response. Annu. Rev.

sette family and . Adv. Drug Deliv. Rev. http://

Pharmacol. Toxicol. 53, 475–502

dx.doi.org/10.1016/j.addr.2016.06.003

80. Ivanov, M. et al. (2014) Epigenetic mechanisms of importance for

103. Hegedus, C. et al. (2015) Lipid regulation of the ABCB1 and

drug treatment. Trends Pharmacol. Sci. 35, 384–396

ABCG2 multidrug transporters. Adv. Cancer Res. 125, 97–137

81. Fisel, P. et al. (2016) DNA methylation of ADME genes. Clin.

104. Zakim, D. and Schwab, M. (2015) Data collection as a barrier to

Pharmacol. Ther. 99, 512–527

personalized medicine. Trends Pharmacol. Sci. 36, 68–71

82. Mandery, K. et al. (2012) Interaction of innovative small molecule

105. Franken, H. et al. (2015) Thermal proteome profiling for unbiased

drugs used for cancer therapy with drug transporters. Br. J.

identification of direct and indirect drug targets using multiplexed

Pharmacol. 165, 345–362

quantitative mass spectrometry. Nat. Protoc. 10, 1567–1593

83. König, J. et al. (2013) Transporters and drug-drug interactions:

106. Huber, K.V. et al. (2015) Proteome-wide drug and metabolite

important determinants of drug disposition and effects. Pharma-

interaction mapping by thermal-stability profiling. Nat. Methods

col. Rev. 65, 944–966

12, 1055–1057

84. Bruhn, O. and Cascorbi, I. (2014) Polymorphisms of the drug

107. Cassidy, J.W. et al. (2015) Maintaining tumor heterogeneity in

transporters ABCB1, ABCG2, ABCC2 and ABCC3 and their

patient-derived tumor xenografts. Cancer Res. 75, 2963–2968

impact on drug bioavailability and clinical relevance. Expert Opin.

Drug Metab. Toxicol. 10, 1337–1354 108. Fu, D. et al. (2014) Imaging the intracellular distribution of tyrosine

kinase inhibitors in living cells with quantitative hyperspectral

85. Nies, A.T. and Lang, T. (2014) Multidrug resistance proteins of

stimulated Raman scattering. Nat. Chem. 6, 614–622

the ABCC subfamily. In Drug Transporters: Molecular Charac-

terization and Role in Drug Disposition (You, G. and Morris, M., 109. Slobbe, P. et al. (2012) PET imaging with small-molecule tyrosine

eds), pp. 161–185, John Wiley & Sons kinase inhibitors: TKI-PET. Drug Discov. Today 17, 1175–1187

86. Poonkuzhali, B. et al. (2008) Association of breast cancer resis- 110. Elkins, J.M. et al. (2016) Comprehensive characterization of the

tance protein/ABCG2 phenotypes and novel promoter and intron published kinase inhibitor set. Nat. Biotechnol. 34, 95–193

1 single nucleotide polymorphisms. Drug Metab. Dispos. 36,

111. Adams, J.L. et al. (2015) Big opportunities for small molecules in

780–795

immuno-oncology. Nat. Rev. Drug Discov. 14, 603–622

87. Ramsey, L.B. et al. (2012) Rare versus common variants in

112. Gotink, K.J. et al. (2011) Lysosomal sequestration of sunitinib: a

pharmacogenetics: SLCO1B1 variation and methotrexate dispo-

novel mechanism of drug resistance. Clin. Cancer Res. 17,

sition. Genome Res. 22, 1 8 7337–7346

Trends in Pharmacological Sciences, November 2016, Vol. 37, No. 11 929

113. Burger, H. et al. (2015) Lysosomal sequestration determines 134. Breedveld, P. et al. (2005) The effect of Bcrp1 (Abcg2) on the in

intracellular imatinib levels. Mol. Pharmacol. 88, 477–487 vivo pharmacokinetics and brain penetration of imatinib mesylate

(Gleevec): implications for the use of breast cancer resistance

114. Aikawa, H. et al. (2016) Visualizing spatial distribution of alectinib

protein and P-glycoprotein inhibitors to enable the brain pene-

in murine brain using quantitative mass spectrometry imaging.

tration of imatinib in patients. Cancer Res. 65, 2577–2582

Sci. Rep. 6, 23749

135. Houghton, P.J. et al. (2004) Imatinib mesylate is a potent inhibitor

115. Poller, B. et al. (2011) Differential impact of P-glycoprotein

of the ABCG2 (BCRP) transporter and reverses resistance to

(ABCB1) and breast cancer resistance protein (ABCG2) on axi-

topotecan and SN-38 in vitro. Cancer Res. 64, 2333–2337

tinib brain accumulation and oral plasma pharmacokinetics. Drug

Metab. Dispos. 39, 729–735 136. Burger, H. et al. (2004) Imatinib mesylate (STI571) is a substrate

for the breast cancer resistance protein (BCRP)/ABCG2 drug

116. Hegedus, C. et al. (2009) Interaction of nilotinib, dasatinib and

pump. Blood 104, 2940–2942

bosutinib with ABCB1 and ABCG2: implications for altered anti-

cancer effects and pharmacological properties. Br. J. Pharmacol. 137. Eechoute, K. et al. (2011) Environmental and genetic factors

158, 1153–1164 affecting transport of imatinib by OATP1A2. Clin. Pharmacol.

Ther. 89, 816–820

117. Redaelli, S. et al. (2015) In vitro and in vivo identification of ABCB1

as an efflux transporter of bosutinib. J. Hematol. Oncol. 8, 81 138. Zhou, L. et al. (2009) The effect of breast cancer resistance

protein and P-glycoprotein on the brain penetration of flavopir-

118. Lacy, S. et al. (2015) Metabolism and disposition of cabozantinib

idol, imatinib mesylate (Gleevec), prazosin, and 2-methoxy-3-(4-

in healthy male volunteers and pharmacologic characterization of

(2-(5-methyl-2-phenyloxazol-4-yl)ethoxy)phenyl)propanoic acid

its major metabolites. Drug Metab. Dispos. 43, 1190–1207

(PF-407288) in mice. Drug Metab. Dispos. 37, 946–955

119. Kort, A. et al. (2015) Brain accumulation of the EML4–ALK

139. Schmidt-Lauber, C. et al. (2012) Transport mechanisms and their

inhibitor ceritinib is restricted by P-glycoprotein (P-GP/ABCB1)

pathology-induced regulation govern tyrosine kinase inhibitor

and breast cancer resistance protein (BCRP/ABCG2). Pharma-

delivery in rheumatoid arthritis. PLoS ONE 7, e52247

col. Res. 102, 200–207

140. Polli, J.W. et al. (2008) The role of efflux and uptake transporters

120. Tang, S.C. et al. (2014) Increased oral availability and

in [N-{3-chloro-4-[(3-fluorobenzyl)oxy]phenyl}-6-[5-({[2-(methyl-

brain accumulation of the ALK inhibitor crizotinib by coadminis-

sulfonyl)ethyl]amino}methyl)-2-furyl]-4-quinazolinamine

tration of the P-glycoprotein (ABCB1) and breast cancer resis-

(GW572016, lapatinib) disposition and drug interactions. Drug

tance protein (ABCG2) inhibitor elacridar. Int. J. Cancer. 134,

1484–1494 Metab. Dispos. 36, 695 701

141. Polli, J.W. et al. (2009) An unexpected synergist role of P-glyco-

121. Hiwase, D.K. et al. (2008) Dasatinib cellular uptake and efflux in

protein and breast cancer resistance protein on the central

chronic myeloid leukemia cells: therapeutic implications. Clin.

nervous system penetration of the tyrosine kinase inhibitor lapa-

Cancer Res. 14, 3881–3888

tinib (N-{3-chloro-4-[(3-fluorobenzyl)oxy]phenyl}-6-[5-({[2-(meth-

122. Furmanski, B.D. et al. (2013) Contribution of ABCC4-mediated

ylsulfonyl)ethyl]amino}methyl)-2-furyl]-4-quinazolinamine;

gastric transport to the absorption and efficacy of dasatinib. Clin.

GW572016). Drug Metab. Dispos. 37, 439–442

Cancer Res. 19, 4359–4370

142. Davies, A. et al. (2009) Nilotinib concentration in cell lines and

123. Giannoudis, A. et al. (2008) Effective dasatinib uptake may occur +

primary CD34 chronic myeloid leukemia cells is not mediated by

without human organic cation transporter 1 (hOCT1): implica-

active uptake or efflux by major drug transporters. Leukemia 23,

tions for the treatment of imatinib resistant chronic myeloid

1999–2006

leukemia. Blood 112, 3348–3354

143. Shukla, S. et al. (2011) Synthesis and characterization of a

124. Chen, Y. et al. (2009) P-glycoprotein and breast cancer resis-

BODIPY conjugate of the BCR-ABL kinase inhibitor Tasigna

tance protein influence brain distribution of dasatinib. J. Pharma-

(nilotinib): evidence for transport of Tasigna and its fluorescent

col. Exp. Ther. 330, 956–963

derivative by ABC drug transporters. Mol. Pharm. 8, 1292–1302

125. Lagas, J.S. et al. (2009) Brain accumulation of dasatinib is

144. Kort, A. et al. (2015) Brain and testis accumulation of regorafenib

restricted by P-glycoprotein (ABCB1) and breast cancer resis-

is restricted by breast cancer resistance protein (BCRP/ABCG2)

tance protein (ABCG2) and can be enhanced by elacridar treat-

and P-glycoprotein (P-GP/ABCB1). Pharm. Res. 32, 2205–2216

ment. Clin. Cancer Res. 15, 2344–2351

145. Ohya, H. et al. (2015) Regorafenib is transported by the organic

126. Marchetti, S. et al. (2008) Effect of the ATP-binding cassette drug

anion transporter 1B1 and the multidrug resistance protein 2.

transporters ABCB1, ABCG2, and ABCC2 on erlotinib hydro-

Biol. Pharm. Bull. 38, 582–586

chloride (Tarceva) disposition in in vitro and in vivo pharmacoki-

/ / 146. Lagas, J.S. et al. (2010) Breast cancer resistance protein and P-

netic studies employing Bcrp1 /Mdr1a/1b (triple-knockout)

glycoprotein limit sorafenib brain accumulation. Mol. Cancer

and wild-type mice. Mol. Cancer Ther. 7, 2280–2287

Ther. 9, 319–326

127. Elmeliegy, M.A. et al. (2011) Role of ATP-binding cassette and

147. Agarwal, S. et al. (2011) The role of the breast cancer resistance

solute carrier transporters in erlotinib CNS penetration and intra-

protein (ABCG2) in the distribution of sorafenib to the brain. J.

cellular accumulation. Clin. Cancer Res. 17, 89–99

Pharmacol. Exp. Ther. 336, 223–233

128. Khurana, V. et al. (2014) Role of OATP-1B1 and/or OATP-1B3 in

148. Poller, B. et al. (2011) Double-transduced MDCKII cells to study

hepatic disposition of tyrosine kinase inhibitors. Drug Metabol.

human P-glycoprotein (ABCB1) and breast cancer resistance

Drug Interact. 29, 179–190

protein (ABCG2) interplay in drug transport across the blood-

129. Li, J. et al. (2007) Association of variant ABCG2 and the

brain barrier. Mol. Pharm. 8, 571–582

pharmacokinetics of epidermal growth factor receptor tyrosine

149. Tang, S.C. et al. (2012) Brain accumulation of sunitinib is

kinase inhibitors in cancer patients. Cancer Biol. Ther. 6,

432–438 restricted by P-glycoprotein (ABCB1) and breast cancer resis-

tance protein (ABCG2) and can be enhanced by oral elacridar

130. Agarwal, S. et al. (2010) Distribution of gefitinib to the brain is

and sunitinib coadministration. Int. J. Cancer 130, 223–233

limited by P-glycoprotein (ABCB1) and breast cancer resistance

150. Namisaki, T. et al. (2014) Differential expression of drug uptake

protein (ABCG2)-mediated active efflux. J. Pharmacol. Exp. Ther.

and efflux transporters in Japanese patients with hepatocellular

334, 147–155

carcinoma. Drug Metab. Dispos. 42, 2033–2040

131. Galetti, M. et al. (2010) Functional characterization of gefitinib

151. Haenisch, S. et al. (2007) Influence of polymorphisms of ABCB1

uptake in non-small cell lung cancer cell lines. Biochem. Phar-

and ABCC2 on mRNA and protein expression in normal and

macol. 80, 179–187

18 cancerous kidney cortex. Pharmacogenomics J. 7, 74

132. Vlaming, M.L. et al. (2015) PET-CT imaging with F-gefitinib to

152. Mizutani, K. et al. (2015) Effects of cabazitaxel in renal cell

measure Abcb1a/1b (P-gp) and Abcg2 (Bcrp1) mediated drug–

carcinoma cell lines. Anticancer Res. 35, 6671–6677

drug interactions at the murine blood–brain barrier. Nucl. Med.

Biol. 42, 833–841 153. Marin, D. et al. (2010) Adherence is the critical factor for achieving

molecular responses in patients with chronic myeloid leukemia

133. Dai, H. et al. (2003) Distribution of STI-571 to the brain is limited

who achieve complete cytogenetic responses on imatinib. J.

by P-glycoprotein-mediated efflux. J. Pharmacol. Exp. Ther. 204,

1085–1092 Clin. Oncol. 28, 2381 2388

930 Trends in Pharmacological Sciences, November 2016, Vol. 37, No. 11

154. Gromicho, M. et al. (2013) Instability of mRNA expression sig- 174. Delord, M. et al. (2013) High imatinib dose overcomes insufficient

natures of drug transporters in chronic myeloid leukemia patients response associated with ABCG2 haplotype in chronic myelog-

resistant to imatinib. Oncol. Rep. 29, 741–750 enous leukemia patients. Oncotarget. 4, 1582–1591

155. Elghannam, D.M. et al. (2014) Association of MDR1 gene poly- 175. Zach, O. et al. (2008) OCT1 (SLC22A1) R61C polymorphism and

morphism (G2677T) with imatinib response in Egyptian chronic response to imatinib treatment in chronic myeloid leukemia

myeloid leukemia patients. Hematology. 19, 123–128 patients. Leuk. Lymphoma 49, 2222–2223

156. Salimizand, H. et al. (2016) Concurrent effects of ABCB1 176. Bazeos, A. et al. (2010) hOCT1 transcript levels and single

C3435T, ABCG2 C421A, and XRCC1 Arg194Trp genetic poly- nucleotide polymorphisms as predictive factors for response

morphisms with risk of cancer, clinical output, and response to to imatinib in chronic myeloid leukemia. Leukemia 24, 1243–

treatment with imatinib mesylate in patients with chronic myeloid 1245

leukemia. Tumour. Biol. 37, 791–798

177. Koren-Michowitz, M. et al. (2014) OCT1 genetic variants are

157. Francis, J. et al. (2015) A study to explore the correlation of associated with long term outcomes in imatinib treated chronic

ABCB1, ABCG2, OCT1 genetic polymorphisms and trough level myeloid leukemia patients. Eur. J. Haematol. 92, 283–288

concentration with imatinib mesylate-induced thrombocytopenia

178. White, D.L. et al. (2010) OCT-1 activity measurement provides

in chronic myeloid leukemia patients. Cancer Chemother. Phar-

a superior imatinib response predictor than screening for

macol. 76, 1185–1189

single-nucleotide polymorphisms of OCT-1. Leukemia 24,

158. Gurney, H. et al. (2007) Imatinib disposition and ABCB1 (MDR1, 1962–1965

P-glycoprotein) genotype. Clin. Pharmacol. Ther. 82, 33–40

179. Giannoudis, A. et al. (2013) The hOCT1 SNPs M420del and

159. Dulucq, S. et al. (2008) Multidrug resistance gene (MDR1) poly- M408 V alter imatinib uptake and M420del modifies clinical out-

morphisms are associated with major molecular responses to come in imatinib-treated chronic myeloid leukaemia. Blood 121,

standard-dose imatinib in chronic myeloid leukemia. Blood 112, 628–637

2024–2027

180. Grinfeld, J. et al. (2013) A common novel splice variant of

160. Kim, D.H. et al. (2009) Clinical relevance of a pharmacogenetic SLC22A1 (OCT1) is associated with impaired responses to

approach using multiple candidate genes to predict response imatinib in patients with chronic myeloid leukaemia. Br. J. Hae-

and resistance to imatinib therapy in chronic myeloid leukemia. matol. 163, 631–639

Clin. Cancer Res. 15, 4750–4758

181. Angelini, S. et al. (2013) Polymorphisms in OCTN1 and OCTN2

161. Ni, L.N. et al. (2011) Multidrug resistance gene (MDR1) polymor- transporters genes are associated with prolonged time to pro-

phisms correlate with imatinib response in chronic myeloid leu- gression in unresectable gastrointestinal stromal tumours treated

kemia. Med. Oncol. 28, 265–269 with imatinib therapy. Pharmacol. Res. 68, 1–6

162. Deenik, W. et al. (2010) Polymorphisms in the multidrug resis- 182. Yamakawa, Y. et al. (2011) Pharmacokinetic impact of SLCO1A2

tance gene MDR1 (ABCB1) predict for molecular resistance in polymorphisms on imatinib disposition in patients with chronic

patients with newly diagnosed chronic myeloid leukemia receiv- myeloid leukemia. Clin. Pharmacol. Ther. 90, 157–163

ing high-dose imatinib. Blood 116, 6144–6145

183. Lima, L.T.de et al. (2015) Relationship between SLCO1B3 and

163. Maffioli, M. et al. (2011) Correlation between genetic polymor- ABCA3 polymorphisms and imatinib response in chronic myeloid

phisms of the hOCT1 and MDR1 genes and the response to leukemia patients. Hematology. 20, 137–142

imatinib in patients newly diagnosed with chronic-phase chronic

184. Bedewy, A.M. and El-Maghraby, S.M. (2013) Do SLCO1B3

myeloid leukemia. Leuk. Res. 35, 1014–1019

(T334G) and CYP3A5*3 polymorphisms affect response in Egyp-

164. Yamakawa, Y. et al. (2011) Association of genetic polymor- tian chronic myeloid leukemia patients receiving imatinib ther-

phisms in the influx transporter SLCO1B3 and the efflux trans- apy? Hematology. 18, 211–216

porter ABCB1 with imatinib pharmacokinetics in patients with

185. Tsuchiya, N. et al. (2013) Risk factors for sorafenib-induced high-

chronic myeloid leukemia. Ther. Drug Monit. 33, 244–250

grade skin rash in Japanese patients with advanced renal cell

165. Vivona, D. et al. (2012) ABCB1 haplotype is associated with carcinoma. Anticancer Drugs. 24, 310–314

major molecular response in chronic myeloid leukemia patients

186. Lee, Y.S. et al. (2015) SLC15A2 genomic variation is associated

treated with standard-dose of imatinib. Blood Cells Mol. Dis. 48,

with the extraordinary response of sorafenib treatment: whole-

132–136

genome analysis in patients with hepatocellular carcinoma.

166. Angelini, S. et al. (2013) Association between imatinib trans- Oncotarget 6, 16449–16460

porters and metabolizing enzymes genotype and response in

187. Teo, Y.L. et al. (2016) Effect of the CYP3A5 and ABCB1 geno-

newly diagnosed chronic myeloid leukemia patients receiving

type on exposure, clinical response and manifestation of toxic-

imatinib therapy. Haematologica 98, 193–200

ities from sunitinib in Asian patients. Pharmacogenomics J. 16,

167. Au, A. et al. (2014) Association of genotypes and haplotypes of 47–53

multi-drug transporter genes ABCB1 and ABCG2 with clinical

188. Narjoz, C. et al. (2015) Role of the lean body mass and of

response to imatinib mesylate in chronic myeloid leukemia

pharmacogenetic variants on the pharmacokinetics and phar-

patients. Biomed. Pharmacother. 68, 343–349

macodynamics of sunitinib in cancer patients. Invest. New Drugs

168. Koo, D.H. et al. (2015) Association of ABCG2 polymorphism with 33, 257–268

clinical efficacy of imatinib in patients with gastrointestinal stromal

189. van der Veldt, A.A. et al. (2011) Genetic polymorphisms associ-

tumor. Cancer Chemother. Pharmacol. 75, 173–182

ated with a prolonged progression-free survival in patients with

169. Yin, C.X. et al. (2016) Association between the concentration of metastatic renal cell cancer treated with sunitinib. Clin. Cancer

imatinib in bone marrow mononuclear cells, mutation status of Res. 17, 620–629

ABCB1 and therapeutic response in patients with chronic mye-

190. Mizuno, T. et al. (2012) Impact of genetic variation in breast

logenous leukemia. Exp. Ther. Med. 11, 2061–2065

cancer resistance protein (BCRP/ABCG2) on sunitinib pharma-

170. Singh, O. et al. (2012) SLC22A1-ABCB1 haplotype profiles cokinetics. Drug Metab. Pharmacokinet. 27, 631–639

predict imatinib pharmacokinetics in asian patients with chronic

191. Garcia-Donas, J. et al. (2011) Single nucleotide polymorphism

myeloid leukemia. PLoS ONE 7, e51771

associations with response and toxic effects in patients with

171. Au, A. et al. (2014) Clinical impact of ABCC1 and ABCC2 gen- advanced renal-cell carcinoma treated with first-line sunitinib:

otypes and haplotypes in mediating imatinib resistance among a multicentre, observational, prospective study. Lancet Oncol.

chronic myeloid leukaemia patients. J. Clin. Pharm. Ther. 39, 12, 1143–1150

685–690

192. Beuselinck, B. et al. (2013) Single-nucleotide polymorphisms

172. Kim, D.H. et al. (2010) The IFNG (IFN-gamma) genotype predicts associated with outcome in metastatic renal cell carcinoma

cytogenetic and molecular response to imatinib therapy in treated with sunitinib. Br. J. Cancer 108, 887–900

chronic myeloid leukemia. Clin. Cancer Res. 16, 5339–5350

193. Beuselinck, B. et al. (2014) Efflux pump ABCB1 single nucleo-

173. Lima, L.T. and de et al. (2014) Reduced ABCG2 and increased tide polymorphisms and dose reductions in patients with met-

SLC22A1 mRNA expression are associated with imatinib astatic renal cell carcinoma treated with sunitinib. Acta Oncol.

response in chronic myeloid leukemia. Med. Oncol. 31, 851 53, 1413–1422

Trends in Pharmacological Sciences, November 2016, Vol. 37, No. 11 931

194. Kim, H.R. et al. (2013) Pharmacogenetic determinants associ- 197. Mizuno, T. et al. (2014) The effect of ABCG2 genotype on the

ated with sunitinib-induced toxicity and ethnic difference in population pharmacokinetics of sunitinib in patients with renal cell

Korean metastatic renal cell carcinoma patients. Cancer Chemo- carcinoma. Ther. Drug Monit. 36, 310–316

ther. Pharmacol. 72, 825–835

198. Low, S.K. et al. (2016) Association study of a functional variant on

195. Diekstra, M.H. et al. (2014) Association analysis of genetic poly- ABCG2 gene with sunitinib-induced severe adverse drug reac-

morphisms in genes related to sunitinib pharmacokinetics, spe- tion. PLoS ONE 11, e0148177

cifically clearance of sunitinib and SU12662. Clin. Pharmacol.

199. van Erp, N.P. et al. (2009) Pharmacogenetic pathway analysis for

Ther. 96, 81–89

determination of sunitinib-induced toxicity. J. Clin. Oncol. 27,

196. Chu, Y.H. et al. (2015) Association of ABCB1 and FLT3 poly- 4406–4412

morphisms with toxicities and survival in Asian patients receiving

sunitinib for renal cell carcinoma. PLoS ONE 10, e0134102

932 Trends in Pharmacological Sciences, November 2016, Vol. 37, No. 11