multiple mechanisms of ligand interaction with the human ......g of salmon sperm, 18 g of pog44, and...

16
Multiple Mechanisms of Ligand Interaction with the Human Organic Cation Transporter, OCT2 Item Type text; Electronic Thesis Authors Harper, Jaclyn Nicole Publisher The University of Arizona. Rights Copyright © is held by the author. Digital access to this material is made possible by the University Libraries, University of Arizona. Further transmission, reproduction or presentation (such as public display or performance) of protected items is prohibited except with permission of the author. Download date 11/05/2021 15:11:10 Link to Item http://hdl.handle.net/10150/297644

Upload: others

Post on 18-Dec-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Multiple Mechanisms of Ligand Interaction with the Human ......g of salmon sperm, 18 g of pOG44, and 2 g of pcDNA5/FRT/ V5-His TOPO containing an OCT2 (hOCT2) construct. The CHOhOCT2

Multiple Mechanisms of Ligand Interaction withthe Human Organic Cation Transporter, OCT2

Item Type text; Electronic Thesis

Authors Harper, Jaclyn Nicole

Publisher The University of Arizona.

Rights Copyright © is held by the author. Digital access to this materialis made possible by the University Libraries, University of Arizona.Further transmission, reproduction or presentation (such aspublic display or performance) of protected items is prohibitedexcept with permission of the author.

Download date 11/05/2021 15:11:10

Link to Item http://hdl.handle.net/10150/297644

Page 2: Multiple Mechanisms of Ligand Interaction with the Human ......g of salmon sperm, 18 g of pOG44, and 2 g of pcDNA5/FRT/ V5-His TOPO containing an OCT2 (hOCT2) construct. The CHOhOCT2
Page 3: Multiple Mechanisms of Ligand Interaction with the Human ......g of salmon sperm, 18 g of pOG44, and 2 g of pcDNA5/FRT/ V5-His TOPO containing an OCT2 (hOCT2) construct. The CHOhOCT2
Page 4: Multiple Mechanisms of Ligand Interaction with the Human ......g of salmon sperm, 18 g of pOG44, and 2 g of pcDNA5/FRT/ V5-His TOPO containing an OCT2 (hOCT2) construct. The CHOhOCT2

Abstract

OCT2 is the entry step for organic cation secretion by renal proximal tubules. Although many

drugs inhibit OCT2 activity, neither the mechanistic basis of their inhibition nor their transport

status is generally known. Using representatives of several structural classes of OCT2 inhibitory

ligands described by Kido et al, we determined the kinetic basis of their inhibition of 1-methyl-4-

phenylpyridinium (MPP) transport into CHO cells that stably expressed hOCT2. The 'Cluster II'

inhibitors, metformin and cimetidine, interacted competitively with MPP. However, other

Cluster II compounds, including tetraethylammonium (TEA), diphenidol and phenyltoloxamine,

were mixed-type inhibitors of MPP transport. Cluster III representative, adrenosterone, and

Cluster I representative, carvedilol, displayed noncompetitive inhibitory profiles. Competitive

Counterflow (CCF) was used to determine if the inhibitory ligands served as substrates of

hOCT2. Carvedilol and adrenosterone did not support CCF, consistent with the prediction that

members of these structural classes are likely to be nontransported inhibitors of OCT2. Cluster II

representatives MPP, metformin, cimetidine and TEA all supported CCF, consistent with

independent assessments of their OCT2-mediated transport. Other Cluster II representatives,

diphenidol and phenyltoloxamine, failed to support CCF, suggesting that neither compound is

transported by OCT2. The results underscore the caution required for development of predictive

models of ligand interaction with multidrug transporters.

Page 5: Multiple Mechanisms of Ligand Interaction with the Human ......g of salmon sperm, 18 g of pOG44, and 2 g of pcDNA5/FRT/ V5-His TOPO containing an OCT2 (hOCT2) construct. The CHOhOCT2

Multiple mechanisms of ligand interaction with the human organic cationtransporter, OCT2

Jaclyn N. Harper and Stephen H. WrightDepartment of Physiology, University of Arizona, Tucson, Arizona

Submitted 26 August 2012; accepted in final form 28 September 2012

Harper JN, Wright SH. Multiple mechanisms of ligand interac-tion with the human organic cation transporter, OCT2. Am J PhysiolRenal Physiol 304: F56–F67, 2013. First published October 3, 2012;doi:10.1152/ajprenal.00486.2012.—OCT2 is the entry step for or-ganic cation (OC) secretion by renal proximal tubules. Although manydrugs inhibit OCT2 activity, neither the mechanistic basis of theirinhibition nor their transport status is generally known. Using repre-sentatives of several structural classes of OCT2-inhibitory ligandsdescribed recently (Kido Y, Matsson P, Giacomini KM. J Med Chem54: 4548–4558, 2011), we determined the kinetic basis of theirinhibition of 1-methyl-4-phenylpyridinium (MPP) transport into Chi-nese hamster ovary cells that stably expressed hOCT2. The “clusterII” inhibitors (which contain known OCT2 substrates) metformin andcimetidine interacted competitively with MPP. However, other clusterII compounds, including tetraethylammonium (TEA), diphenidol andphenyltoloxamine, were mixed-type inhibitors of MPP transport (i.e.,decreasing Jmax and increasing Kt). A cluster III (neutral steroid)representative, adrenosterone, and a cluster I (large, flexible cation)representative, carvedilol, displayed noncompetitive inhibitory pro-files. Competitive counterflow (CCF) was used to determine whetherthe inhibitory ligands served as substrates of hOCT2. Carvedilol(cluster I) and adrenosterone (cluster III) did not support CCF,consistent with the prediction that members of these structural classesare likely to be nontransported inhibitors of OCT2. The cluster IIrepresentatives MPP, metformin, cimetidine, and TEA all supportedCCF, consistent with independent assessments of their OCT2-medi-ated transport. However, the other cluster II representatives, dipheni-dol and phenyltoloxamine, failed to support CCF, suggesting thatneither compound is transported by OCT2. An independent assess-ment of diphenidol transport (using liquid chromatography withtandem mass spectroscopy) confirmed this observation. The resultsunderscore the caution required for development of predictive modelsof ligand interaction with multidrug transporters.

transport; proximal tubule; organic cation; multidrug

IN A RECENT ASSESSMENT OF the physicochemical determinants ofrenal drug clearance (51), the kidney was found to contribute�50% of total body clearance for �30% of the �400 com-pounds surveyed. Renal secretion, mediated principally byproximal tubule cells, plays a central role in this process, ahallmark of which is the structurally diverse array of bothexogenous (xenobiotic), and endogenous, organic compoundsit handles (43). A consequence of the structural diversity ofligand interaction that is a functional hallmark of a compara-tively small set of “multidrug” transport processes is the highprobability of unwanted drug-drug-interactions (DDIs) at thepoint of renal secretion, interactions that can significantlyinfluence a drug’s clinical or toxicological profile. In addition,inhibition of renal drug transport can influence drug accumu-

lation within the kidney, thereby contributing to the nephro-toxic impact associated with use of some pharmaceuticalagents. Unwanted DDIs are typically detrimental, so predictingand, ideally, preempting their occurrence is a major focus ofcurrent work on renal drug transport.

Our focus is on the role of organic cation transporter 2(OCT2; SLC22A2) in renal drug secretion. Approximately40% of prescribed drugs carry a net positive charge at or nearphysiological pH (1) and, so, are collectively referred to as“organic cations” (OCs), and the mechanism of OC secretionby renal proximal tubule cells has received considerable atten-tion. The basic cellular model of renal OC secretion in renalproximal tubule (RPT) cells, proposed first by Holohan andRoss (45), includes the sequential activity of 1) a basolateral“entry step,” from blood to cell, that involves an electrogenicorganic cation transporter; and 2) an apical “exit step,” fromcell to tubular filtrate mediated by electroneutral OC/H� ex-change. Following the cloning of a series of electrogenicorganic cation transporters (19, 23, 40), there is now broadconsensus that, in the human kidney, the basolateral step in thisprocess is dominated by activity of OCT2 (36, 43).

The increasing attention given to the clinical impact ofunwanted DDIs has lead to development of several predictivemodels of ligand interaction with OCT2 (24, 49, 56) and itshepatic homolog, OCT1 (1, 6, 35). All of these studies corre-lated inhibition of OCT-mediated transport produced by sets oftest ligands with various structural and/or physicochemicaldescriptors of the inhibitory agents, as the basis for develop-ment of two dimensional (2D)- or 3D-quantitative structureactivity relationships (QSARs) of ligand/transporter interac-tion. Although details vary, 2D-QSARs for both OCTs consis-tently show that inhibitory effectiveness is correlated with1) cationic charge and 2) various measures of ligand lipophi-licity; and 3D-QSARs (“pharmacophores”) generally include a“cationic feature,” one or more “hydrophobes,” and, fre-quently, one or more hydrogen bond donor/acceptor features.In the most extensive study of ligand interaction with OCT2 todate, Kido et al. (24) screened 900 compounds and the effec-tive inhibitors of transport (244 compounds) were distributedinto 3 structural classes, referred to as “clusters,” based onselected molecular properties. Cluster I inhibitors are generallylong, flexible cations; cluster II inhibitors are smaller, globular,cations and, interestingly, included the known substrates ofOCT2; and cluster III inhibitors are noncharged heterotypicring structures (including many sterols).

Despite the increase in understanding of physical and struc-tural factors that influence ligand interaction with the OCTs,the work to date has largely overlooked two issues we suggestare central to understanding the molecular basis of OCTactivity and successful development of predictive models oftheir influence on OC clearance. Specifically, existing studies

Address for reprint requests and other correspondence: S. H. Wright, Univ.of Arizona College of Medicine, Dept. of Physiology, PO Box 245051,Tucson, AZ 85724 (e-mail: [email protected]).

Am J Physiol Renal Physiol 304: F56–F67, 2013.First published October 3, 2012; doi:10.1152/ajprenal.00486.2012.

1931-857X/13 Copyright © 2013 the American Physiological Society http://www.ajprenal.orgF56

at University of A

rizona Health S

ciences Library on January 20, 2013http://ajprenal.physiology.org/

Dow

nloaded from

Page 6: Multiple Mechanisms of Ligand Interaction with the Human ......g of salmon sperm, 18 g of pOG44, and 2 g of pcDNA5/FRT/ V5-His TOPO containing an OCT2 (hOCT2) construct. The CHOhOCT2

have largely ignored 1) the mechanistic/kinetic basis of ob-served inhibition of OCT transport activity and 2) whether aninhibitory ligand is itself a transported substrate. The first ofthese issues is central to development of pharmacophores thatreflect the possibility, frequently alluded to but not directlyaddressed, that OCTs may have multiple binding sites and thatsubstrates and inhibitors may interact with one or more of thesesites, perhaps simultaneously (e.g., Refs. 15 and 34). There isa large extant literature on ligand interaction with OCTs (seeRef. 39 for a thorough review of these data). Unfortunately,large variability in quantitative data, even from studies exam-ining the same substrates/inhibitors using similar (if not iden-tical) experimental systems, precludes any effort to draw broadconclusions about the kinetic basis of OCT/ligand interactions.

The second issue noted above, i.e., the need to understandthe structural basis of ligand interaction that can result insuccessful transport, is critical to development of “physiolog-ically based pharmacokinetic models.” Whereas hundreds ofcompounds have been shown to inhibit OCT transport activity(e.g., Refs. 1 and 24), evidence for transport is limited to fewerthan 30 compounds (39, 46).

The present study provides an initial examination of both ofthese issues for human OCT2. Within the structurally diverseset of compounds studied, which included representatives ofthe three structural classes of OCT2 inhibitors defined by Kidoet al. (24), examples of “competitive,” “noncompetitive,” and“mixed-type” inhibition were noted. We also describe amethod, “competitive counterflow” (CCF), that distinguishedbetween nontransported inhibitors and those that also serve astransported substrates of OCT2.

MATERIALS AND METHODS

Reagents

Chinese hamster ovary (CHO) cells containing a single integratedFlp recombination target site (CHO Flp-In), Flp recombinase expres-sion plasmid (pOG44), platinum high-fidelity DNA polymerase, hy-gromyosin, zeocin, and mammalian expression vector pcDNA5/FRT/V5-His TOPO were acquired from Invitrogen (Carlsbad, CA). Ham’sF-12 Nutrient Mixture, 1-methyl-4-phenylpyridinium (MPP), cimeti-dine, tetraethyl ammonium chloride (TEA), imipramine hydrochlo-ride, pyrimethamine, adrenosterone, carvedilol, and phenyltoloxaminewere obtained from Sigma-Aldrich (St. Louis, MO). Metformin wasobtained from AK Scientific (Mountain View, CA). Tacrine hydro-chloride was obtained from Cayman Chemical (Ann Arbor, MI).Diphenidol hydrochloride was obtained from Santa Cruz Biotechnol-ogy (Santa Cruz, CA). N,N,N-trimethyl-2-[methyl(7-nitrobenzo[c][l,2,5]oxadiazol-4-yl)amino]ethanaminium iodide (NBD-MTMA) andN1-ethylacridinium were synthesized by the Synthesis Core of theSouthwest Environmental Health Sciences Center/Department ofChemistry at the University of Arizona (Tucson, AZ).

Cell Culture and Stable Expression of hOCT2

CHO cells containing the Flp recombination target site were grownin Ham’s F-12 Nutrient Mixture with 10% fetal bovine serum andsupplemented with 100 �g/ml zeocin. Cells that stably expressedhOCT2 were prepared using methods described previously (42).Briefly, cells (5 � 106) were transfected by electroporation (BTXECM 630, San Diego; 260 V with time constant of �25 ms) with 10�g of salmon sperm, 18 �g of pOG44, and 2 �g of pcDNA5/FRT/V5-His TOPO containing an OCT2 (hOCT2) construct. TheCHOhOCT2 cells were grown and maintained under hygromycin pres-sure (100 �g/ml) in plastic cell culture flasks at 37°C in a humidified

atmosphere with 5% CO2. Cells were passed every 3–4 days andseeded into 48-well plates at 550,000 cells/ml (24-h confluence) or275,000 cells/ml (48-h confluence).

Transport Studies

CHO cells grown to confluence in 48-well plates were rinsed twicewith Waymouth buffer (WB; 135 mM NaCl, 13 mM HEPES, 2.5 mMCaCl2·2H2O, 1.2 mM MgCl2, 0.8 mM MgSO4·7H2O, 5 mM KCl, and28 mM D-glucose). A loading buffer containing [3H]MPP (80 Ci/mmol; typically 10–20 nM), plus the desired concentration of a testcompound, was added to each well, removed after a set amount oftime, and washed three times with cold WB to stop transport. Thecells were then solubilized in 200 �l of 0.5 N NaOH with 1%SDS/well and shaken for 15 min. The solubilized cells were neutral-ized with 100 �l of 1 N HCl, and 250 �l were placed in a scintillationvial. The radioactivity in each sample was determined using scintil-lation spectroscopy (Beckman model LS6000IC). Individual transportexperiments were conducted in triplicate and typically repeated atleast three times. Experiments were conducted using cells betweenpassages 4 and 35 with no consistent difference in transport ratesbetween earlier and later passages.

Mass Spectroscopy

Analyses were performed using a TSQ Quantum Ultra triple-quadrupole mass spectrometer in conjunction with a Finnigan Sur-veyor Autosampler and Finnigan Surveyor quaternary HPLC pump(Thermo Finnigan, San Jose, CA).

Diphenidol. The method was adapted from that of Marcelin-Jimanez et al. (32). The mass spectrometer was operated in thepositive ionization mode using electrospray (ESI) ionization under thefollowing conditions: spray voltage of 4,800 V, sheath gas (nitrogen)flow of 35 (arbitrary units), capillary temperature of 370°C with argonas the collision gas at a pressure of 0.8 mtorr. Detection used selectivereaction monitoring (SRM), the parent mass/charge ratio for dipheni-dol was 310.2, and the fragment monitored was 292.3. The collisionenergy was 24 eV. HPLC separation was achieved using a Kinetex2.6� C-18 50 � 2.1-mm column (Phenomenex, Torrance, CA).Mobile phase A was 5:95 acetonitrile:water with 0.1% formic acid(vol/vol); mobile phase B was 95:5, acetonitrile:water with 0.1%formic acid (vol/vol). Mobile phase gradient was 80% A from 0 to 4min, increasing to 100% B by 4.10 min, and maintaining 100% B until8 min, followed by reequilibration at 80% A from 8.10 min until 12min. Flow rate was 200 �l/min, and injection volume was 5 �l. Totalanalysis time was 12 min; diphenidol eluted at �2.4 min.

Metformin. The mass spectrometer was operated in the positiveionization mode using atmospheric pressure chemical ionization(APCI) under the following conditions: spray voltage of 3,500 V,sheath gas (nitrogen) flow of 21 (arbitrary units), auxiliary gas(nitrogen) flow of 5, vaporizer temperature of 395°C, and capillarytemperature of 200°C with argon as the collision gas at a pressure of0.8 mtorr. Detection used SRM. The parent mass/charge ratio formetformin was 130.1, and the fragment monitored was 71.1. Thecollision energy was 33 eV. HPLC separation was achieved using aBDS HypersilCyano3� 50 � 4.6-mm column (Thermo Scientific,Bellefonte, PA). The mobile phase was 60:20:20, 10 mM ammoniumacetate:acetonitrile:methanol (vol/vol), at a flow rate of 650 �l/min,and injection volume was 20 �l. Total analysis time was 5 min;metformin eluted at �3.6 min.

Statistical Analysis

Data are expressed as means � SE, with calculations of SEs basedon the number of separate experiments conducted on cells at adifferent passage number. One-way ANOVA was used to test theeffect of multiple treatments and was followed by the Student-Newman-Keuls test for pairwise comparisons. Comparison of sample

F57MECHANISMS OF LIGAND INTERACTION WITH HUMAN OCT2

AJP-Renal Physiol • doi:10.1152/ajprenal.00486.2012 • www.ajprenal.org

at University of A

rizona Health S

ciences Library on January 20, 2013http://ajprenal.physiology.org/

Dow

nloaded from

Page 7: Multiple Mechanisms of Ligand Interaction with the Human ......g of salmon sperm, 18 g of pOG44, and 2 g of pcDNA5/FRT/ V5-His TOPO containing an OCT2 (hOCT2) construct. The CHOhOCT2

means was done using an unpaired Student’s t-test. All statisticalanalyses were performed with Prism 5 (GraphPad Software, La Jolla,CA) or Excel 2007 (Microsoft, Redmond, WA) and deemed signifi-cant when P � 0.05.

RESULTS

Kinetics of MPP Transport

The hOCT2-mediated uptake of [3H]MPP was linear for atleast 90 s (data not shown), so 1-min uptakes were used for allsubsequent studies of transport kinetics. To establish the kinet-ics of hOCT2-mediated MPP transport, the rate of transport of12 nM [3H]MPP into CHO cells that stably expressed thetransporter was measured in the presence of increasing con-centrations of unlabeled MPP (Fig. 1). The relationship wasadequately described by the Michaelis-Menten equation forcompetitive interaction of labeled and unlabeled substrate (31)

J * �Jmax[S*]

Kt � [S*] � [S]� Dns[S*] (1)

where J* is the rate of transport of the radiolabeled substrate(in this case, [3H]MPP) from a concentration of the labeledsubstrate equal to [S*]; Jmax is the maximal rate of mediatedsubstrate transport; Kt is the Michaelis constant of the trans-ported substrate; [S] is the concentration of unlabeled sub-strate; and Dns is a first-order rate constant that describes thenonsaturable component of labeled substrate accumulation(reflecting the combined influence of diffusion, nonspecificbinding, and incomplete rinsing of labeled substrate from thecell culture well). In 22 separate experiments Kt and Jmax

values for hOCT2-mediated MPP transport were 11.1 � 1.2�M and 16.0 � 1.4 pmol·cm�2·min�1, respectively. Expressedper milligram of membrane protein, Jmax was (approximately)400 pmol·mg�1·min�1.

Ligand Inhibition of hOCT2

Thirteen structurally diverse OCs (Fig. 2) were tested for theirinhibition of OCT2-mediated [3H]MPP transport (Table 1).The rationale for selection of these compounds varied: cime-tidine, imipramine, and tacrine IC50 values were reported byKido et al. (24) in their study on ligand interaction with

hOCT2; and carvedilol, diphenidol, phenyltoloxamine, andadrenosterone were cited as representatives of the three cate-gories of OC structures they defined as OCT2 inhibitors. Theremaining compounds were chosen either because of their use,either as substrates (metformin, TEA) (6, 25, 49) or inhibitors(pyrimethamine, corticosterone) (15, 27) in studies of OCT activ-ity; or their use by us in previous studies of renal OC transport(N1-ethylacridinium, and NBD-MTMA) (7, 53). Figure 3 showsthe inhibitory kinetic profiles of several representative test

Fig. 1. Kinetics of 1-methyl-4-phenylpyridinium (MPP) transport into Chinesehamster ovary (CHO) cells that stably express human organic cation trans-porter 2 (hOCT2). Each point is the mean (�SE) of the 1-min uptake of 12 nM[3H]MPP, measured in the presence of increasing concentrations of unlabeledMPP, determined in 22 separate experiments (each in triplicate). Kinetic valueswere determined using the relationship outlined in Eq. 1. Jmax, maximal rate ofmediated substrate transport; Kt, Michaelis constant of the transported sub-strate.

Fig. 2. Structural formulas of the test compounds used in this study.

F58 MECHANISMS OF LIGAND INTERACTION WITH HUMAN OCT2

AJP-Renal Physiol • doi:10.1152/ajprenal.00486.2012 • www.ajprenal.org

at University of A

rizona Health S

ciences Library on January 20, 2013http://ajprenal.physiology.org/

Dow

nloaded from

Page 8: Multiple Mechanisms of Ligand Interaction with the Human ......g of salmon sperm, 18 g of pOG44, and 2 g of pcDNA5/FRT/ V5-His TOPO containing an OCT2 (hOCT2) construct. The CHOhOCT2

agents: N1-ethylacridinium, tacrine, diphenidol, and met-formin. For these (and all the other) compounds tested, uptakeof [3H]MPP was inhibited by increasing concentrations of thetest agent, although the concentrations that inhibited 50% ofthe mediated (i.e., blockable) component of total MPP uptake(IC50) varied from ��1 �M to ��100 �M (Table 1), ac-cording to the relationship

J* �Jmapp[S*]

IC50 � [I]� Dns[S*] (2)

where IC50 is the concentration of the inhibitor [I] that reducedmediated (i.e., blockable) [3H]MPP transport by 50%; Jmapp isthe product of the maximum rate of S* (i.e., [3H]MPP) uptake(Jmax) and the ratio of the IC50 of the test agent and Kt for MPPtransport. Table 1 presents the average IC50 values for theseand the other test agents included in this study.

N1-ethylacridinium was the most potent inhibitor ofhOCT2-mediated MPP transport (IC50 of 0.26 � 0.16 �M);metformin was the least effective inhibitor (IC50 of 545 � 45 �M).

As we are about to discuss a more detailed view of thekinetic basis of ligand interaction with hOCT2, it is relevant tocompare, where available, our results to those obtained byothers using similar methods. Figure 4 plots the data presentedin Table 1, comparing the IC50 values we obtained for inhibi-tion of hOCT2-mediated MPP transport in CHO cells to thosereported in other studies. We restricted the comparison tostudies of expression of human OCT2 in cultured mammaliancells (typically HEK 293 cells) and indicate the identity of thetransported substrate used as the measure of OCT2 activity.Using these selection criteria, we found comparable IC50 datafor 9 of our 13 test ligands, including 5 studies that used MPPas the transported substrate. In general, there was a strongcorrelation between the values we measured and those reportedby others. We found it particularly intriguing that the correla-tion appeared strongest when limited to the IC50 values thatused MPP as the transported probe; the several IC50 valuesobtained using ASP as the test probe were consistently lowerthan the values obtained using MPP. The correlation betweenour results, and the technical methods we employed, with thoseobtained by others lends credence, we suggest, to the dataobtained in the more detailed kinetic analyses described below.

Kinetic Basis of Inhibition

The inhibitory profiles evident in Fig. 3, and the IC50 valuesreported in Table 1, provided no insight by themselves into themechanistic basis of the decreases in OCT2-mediated MPPtransport produced by the test ligands; although consistent with“competition” between [3H]MPP and the test agent, qualita-tively similar profiles are expected if the interactions werenoncompetitive, uncompetitive, or a “mixed type” (47). Toestablish the kinetic basis of observed inhibition of hOCT2-mediated transport, we measured the kinetics of MPP uptake inthe absence and presence of a fixed concentration of inhibitorequal to approximately two times its IC50 value (Fig. 5).Inhibition was considered competitive if its presence resulted

Table 1. IC50 values for inhibition of hOCT2-mediatedtransport

IC50, �M

Compound Means � SE; n 3 Other studies

N1-ethylacridinium 0.26 � 0.16 0.09TEA (49)Tacrine 3.1 � 1.5 0.7ASP (24)Imiprimine 3.3 � 1.3 6MPP (56)

0.6ASP (24)Phenyltoloxamine 4.2 � 2.8Adrenosterone 7.0 � 2Corticosterone 17 � 2.9 34MPP (20)Pyrimethamine 24 � 5.4 10Met (27)Diphenidol 35 � 12Carvedilol 55 � 7.3 63MPP (56)NBD-MTMA 63 � 12Cimetidine 110 � 32 126MPP (56)

23ASP (24)70TEA (49)

TEA 269 � 190 189TEA (33)Metformin 545 � 45 398MPP (56)

336TEA (49)

Means � SE values were determined in the present study using the methodsoutlined for the studies in Fig. 3 (each value is the mean of 3 separateexperiments). TEA, tetraethylammonium. The IC50 values for inhibition ofhOCT2 activity presented in the right-hand column were reported in theindicated references (in parentheses), and the superscripts refer to the humanorganic cation transporter 2 (hOCT2) substrate used to generate each value(note: all measurements reflect transport into cultured mammalian cells trans-fected with hOCT2).

Fig. 3. Kinetics of inhibition of hOCT2-mediated [3H]MPP transport producedby increasing concentrations of the indicated test agents. Data points are mean1-min uptakes (�SE) of 12 nM [3H]MPP, normalized to the rate of transportmeasured in the absence of inhibitor, determined in 3 separate experiments (fordiphenidol, n 4), each performed in triplicate.

Fig. 4. Comparison of IC50 values, listed in Table 1, for the inhibition ofhOCT2-mediated transport produced by selected test agents. Filled circlescompare values determined in the present study, using [3H]MPP as the testsubstrate, with those determined in other studies that also used radiolabeledMPP as the test substrate. Open symbols compare values determined in thepresent study (using [3H]MPP) with those determined in other studies using theindicated compound (TEA, ASP, or metformin) as the test substrate. The dashedline is the line of identity (equal IC50 values).

F59MECHANISMS OF LIGAND INTERACTION WITH HUMAN OCT2

AJP-Renal Physiol • doi:10.1152/ajprenal.00486.2012 • www.ajprenal.org

at University of A

rizona Health S

ciences Library on January 20, 2013http://ajprenal.physiology.org/

Dow

nloaded from

Page 9: Multiple Mechanisms of Ligand Interaction with the Human ......g of salmon sperm, 18 g of pOG44, and 2 g of pcDNA5/FRT/ V5-His TOPO containing an OCT2 (hOCT2) construct. The CHOhOCT2

in an increase in the apparent Kt for MPP transport but nochange in Jmax. Moreover, because the inhibitor concentrationsused were each about twice the measured IC50 value, apparentKt values were expected to increase about threefold (47). Basedon these criteria, cimetidine’s inhibition of OCT2-mediatedMPP transport was competitive: in three experiments, thepresence of 200 �M cimetidine increased Kt from 7.4 � 2.5 to24.9 � 7.9 �M (P � 0.01, n 6), while Jmax remainedunchanged (8.6 � 4.9 vs. 11.4 � 4.1 pmol·cm�2·min�1; Table 2).Metformin’s inhibition of OCT2-mediated MPP transport alsodisplayed a competitive profile (Table 2). Carvedilol, in con-trast, was a noncompetitive inhibitor of MPP transport, reduc-ing the Jmax of transport from 16.7 � 4.3 to 3.8 � 3.7pmol·cm�2·min�1 (P � 0.05, n 3), without significantlyinfluencing the Kt (13.0 � 4.0 vs. 8.5 � 2.4 �M; Fig. 5, Table2). Adrenosterone produced a similar, noncompetitive inhibi-tory profile (Table 2). Interestingly, TEA, a well-characterizedsubstrate of OCT2 (e.g., Refs. 5 and 40), proved to be amixed-type, rather than competitive, inhibitor of MPP trans-

port; 500 �M TEA produced a decrease in Jmax (from 10.6 �2.5 to 5.8 � 1.1 pmol·cm�2·min�1; P � 0.05, n 3) and anincrease in apparent Kt (from 7.6 � 1.0 to 18.5 � 0.6 �M;P � 0.01) (Fig. 5, Table 2). Diphenidol and phenyltoloxaminealso produced mixed-type inhibitory profiles (Table 2).

Indirect Assessment of Inhibitor Transport:Competitive Counterflow

Knowledge of the kinetic basis of ligand interaction withOCT2 cannot support conclusions on whether an inhibitor oftransport is itself a transported substrate of OCT2. Addressingthe issue of transport directly can be a challenge; unfortunately,most of the inhibitory ligands used in this (and other) studiesare not routinely available in radiolabeled form (and customsyntheses can be prohibitively expensive). Furthermore, al-though the transport of most of these compounds, if not all, canbe determined employing various other analytic methods (e.g.,HPLC and/or mass spectroscopy), their use in a study attempt-

Fig. 5. Kinetics of [3H]MPP uptake in CHOhOCT2 cells inthe absence (control) and presence of the indicated inhibi-tory ligands: 200 �M cimetidine (n 6; A); 100 �Mcarvedilol (n 3; B); 500 �M TEA (n 3; C); 70 �Mdiphenidol (n 3; D). Data points represent mean uptakes(1 min; �SE) under the indicated condition from severalseparate experiments (as indicated by n values), each per-formed in triplicate. Uptakes were corrected for the non-saturable component of transport as determined by thecalculated value of Dns for each experiment (see Eq. 1).

Table 2. Kinetic impact of inhibitory ligands on hOCT2-mediated MPP transport

InhibitorMPP Transport, Jmax,

pmol·cm�2·min�1Jmax in the Presence of

Inhibitor, pmol·cm�2·min�1 MPP Transport Kt, �MKt in the Presence of

Inhibitor, �M

CompetitiveMetformin (1 mM) 12.8 � 1.8 14.0 � 3.5 8.1 � 1.9 22.9 � 4.3†Cimetidine (200 �M) 8.6 � 4.9 11.4 � 4.1 7.4 � 2.5 24.9 � 7.9†

NoncompetitiveAdrenosterone (15 �M) 13.9 � 3.9 2.6 � 0.7† 11.6 � 2.6 7.3 � 1.7Carvedilol (100 �M) 16.7 � 4.3 3.8 � 3.7* 13.0 � 4.0 8.5 � 2.4

Mixed TypeTEA (500 �M) 10.6 � 2.5 5.8 � 1.1* 7.6 � 1.0 18.5 � 0.6†Diphenidol (70 �M) 21.1 � 7.6 13.1 � 3.0* 15.8 � 6.6 34.1 � 4.1†Phenyltoloxamine (15 �M) 19.3 � 4.6 8.0 � 3.7* 17.6 � 2.4 30.1 � 6.9*

Values are means � SE of 3–6 separate, paired experiments (each performed in triplicate). MPP, 1-methyl-4-phenylpyridinium. The maximal rate of mediatedsubstrate transport (jmax) and the Michaelis constant of the transported substrate (Kt) values were measured in the absence (control) and presence of the indicatedconcentration of each inhibitory ligand. *,†: Differences at the P � 0.05 (*) or 0.01 (†) level.

F60 MECHANISMS OF LIGAND INTERACTION WITH HUMAN OCT2

AJP-Renal Physiol • doi:10.1152/ajprenal.00486.2012 • www.ajprenal.org

at University of A

rizona Health S

ciences Library on January 20, 2013http://ajprenal.physiology.org/

Dow

nloaded from

Page 10: Multiple Mechanisms of Ligand Interaction with the Human ......g of salmon sperm, 18 g of pOG44, and 2 g of pcDNA5/FRT/ V5-His TOPO containing an OCT2 (hOCT2) construct. The CHOhOCT2

ing to screen transport of a large number of ligands can also beprohibitively expensive (in both time and money). In light ofthese issues, we sought to develop a method that is both costeffective and yet still capable of providing information con-cerning the “transportability” of inhibitory ligands.

The approach we developed is a modification of methodsintroduced for the study of glucose transport in red cells byRosenberg and Wilbrandt (“counterflow”) (44) and Lacko andBurger (“competitive exchange diffusion”) (28), as discussedby Stein (48). We refer to the method as CCF as it combinedelements of the aforementioned studies to determine whethertwo compounds share the same transport process. Briefly, cellsexpressing the transporter of interest (e.g., hOCT2) are allowedto accumulate a probe substrate to steady state. The cells arethen rinsed of the extracellular solution and immediately ex-posed to an experimental solution containing either 1) the sameconcentration of the labeled substrate or 2) the same concen-tration of labeled substrate plus a test compound to determinewhether the presence of the test compound stimulates efflux ofthe intracellular compound. In the first condition, cell contentof the radiolabeled substrate should remain unchanged becausethe cells have already achieved a steady-state distributionarising from equal rates of substrate entry and exit. However,under the second condition, the presence of the extracellularinhibitory ligand reduces the rate of influx of the labeledsubstrate. If binding of the test ligand still permits turnover ofthe transporter (i.e., if the inhibitor is a substrate), then theintracellular labeled substrate can still exit and cell content ofthe radiolabel will decrease with time. Importantly, if bindingof an inhibitory ligand prevents effective turnover of thetransporter, then mediated efflux of the label is eliminated,indicating that the test compound is a nontransported inhibitor.

As a “proof-of-concept” for our modified CCF protocol, wedetermined the influence of extracellular unlabeled MPP on theefflux of preloaded [3H]MPP. Figure 6 shows experiments thatestablished the time and concentration dependence of [3H]MPPtransport using the CCF protocol. Figure 6A shows that steady-state accumulation of [3H]MPP (�12 nM) into hOCT2-ex-pressing cells was achieved within �30 min. In the experiment

summarized in Fig. 6B, cells allowed to accumulate [3H]MPPto steady state (30 min) were exposed to media containingeither 1) 12 nM [3H]MPP (identical to the loading solution) or2) 12 nM [3H]MPP plus 100 �M unlabeled MPP. When thesolution contained only [3H]MPP (the “control” condition),cell content of labeled MPP remained unchanged over 1 min,reflecting maintenance of the steady-state condition. The addi-tion of unlabeled MPP in the extracellular solution was asso-ciated with a rapid net loss of label from the cells thatapproached a new steady state after 60 s. This was expectedbecause OCTs support mediated exchange of an OC for an OC(10); i.e., unlabeled MPP should compete with labeled MPP atthe extracellular binding surface of hOCT2 to 1) preventunidirectional influx of [3H]MPP while 2) supporting exchangefor the intracellular [3H]MPP. The rapid decline in the rate ofloss of label from the cells reflected the anticipated inhibitoryimpact of increasing competition between labeled MPP and theincreasing concentration of unlabeled substrate at the cytoplas-mic face of OCT2. It also underscores the advantage of usinga labeled substrate with a comparatively high affinity for thetransporter, such as MPP for hOCT2, because it delays theinhibitory influence of the accumulated test compound onprobe efflux. Figure 6C shows the effect of increasing theconcentration of extracellular unlabeled MPP on the rate of[3H]MPP efflux from preloaded cells. Neither 1 or 3 �M MPP,concentrations well below the Kt for MPP uptake (Fig. 1),induced a significant loss of [3H]MPP from the cells (Fig. 6).However, the presence of 10 �M MPP, equal to the Kt forOCT2-mediated MPP uptake (Fig. 1), and all concentrationsgreater than that, did stimulate a net loss of labeled MPP (Fig.6). This trans-effect was consistent with the fact that MPP is asubstrate for hOCT2. Efflux increased up to 100 �M, butconcentrations �100 �M had no further stimulatory effect. Inother words, increasing extracellular MPP beyond �100 �Mwas incapable of stimulating more efflux, due to the effectivesaturation of available OCT2 transporters by extracellularMPP.

Figure 7 shows the effect of several additional known OCT2substrates, TEA, metformin, and cimetidine, on the efflux of

Fig. 6. Competitive counterflow (CCF) profiles. A: time course of 12 nM [3H]MPP uptake in OCT2 and wild-type (WT) CHO cells. Each point is the mean oftriplicate determinations of cell content of [3H]MPP into WT CHO cells or CHO cells that stably expressed hOCT2 (lines fit by eye). Data shown are from asingle representative experiment. B: time course of [3H]MPP efflux from hOCT2-expressing CHO cells. Before the measurement of efflux, cells were exposedto 12 nM [3H]MPP for 30 min. At time 0, the medium was rinsed and replaced with medium containing either 12 nM [3H]MPP (control) or 12 nM [3H]MPPplus 100 �M MPP. Each point is the mean (�SE) of cell [3H]MPP content (normalized to content at time 0) measured in 3 wells from a representative experiment.C: effect of increasing extracellular concentrations of unlabeled MPP on efflux of [3H]MPP from hOCT2-expressing CHO cells. Before the measurement ofefflux, cells were exposed to 12 nM [3H]MPP for 30 min. At time 0, the medium was rinsed and replaced with medium containing either 12 nM [3H]MPP (control)or 12 nM [3H]MPP plus the indicated concentration of unlabeled MPP. The height of each bar represents triplicate determinations of the [3H]MPP remainingin cells after a 60-s efflux period. Data shown are from a single representative experiment. Bars marked with * were different from the control (P � 0.05).

F61MECHANISMS OF LIGAND INTERACTION WITH HUMAN OCT2

AJP-Renal Physiol • doi:10.1152/ajprenal.00486.2012 • www.ajprenal.org

at University of A

rizona Health S

ciences Library on January 20, 2013http://ajprenal.physiology.org/

Dow

nloaded from

Page 11: Multiple Mechanisms of Ligand Interaction with the Human ......g of salmon sperm, 18 g of pOG44, and 2 g of pcDNA5/FRT/ V5-His TOPO containing an OCT2 (hOCT2) construct. The CHOhOCT2

[3H]MPP from OCT2-expressing CHO cells. The cells wereexposed to a concentration of each agent equal to �10 timesthe IC50 value for that compound as an inhibitor of OCT2-mediated MPP uptake (Table 1), concentrations that, for eachcompound, may be expected to occupy �90% of availablebinding sites at the extracellular face of the transporter. Asexpected, each compound supported CCF because each hasbeen shown to be a transported substrate of OCT2, in additionto being an inhibitory ligand (39). Figure 7 also shows that, asexpected, compounds that do not interact with OCT2 (glucose,mannitol, sucrose, and succinic acid) do not support CCF.

CCF of Inhibitory Ligands

CCF was used to determine whether ligands selected fromthe list of representatives for each OCT2 inhibitory clusteridentified by Kido et al. (24) are also transported by OCT2.Carvedilol was the cluster I representative (long, flexible cat-ions), and adrenosterone was the cluster III representative(noncharged heterotypic ring structures). Known transportedsubstrates of OCT2 have the structural characteristics of thecluster II compounds (smaller, globular, cations), so we testedtwo representatives of this group of inhibitory ligands: dipheni-dol and phenyltoloxamine. The concentration of the unlabeledtest compound was generally 10 times its IC50 as an inhibitorof OCT2-mediated MPP uptake (because of solubility issues,the carvedilol concentration was limited to 4 times its IC50

value). Figure 8 shows that none of these compounds sup-ported a net loss of [3H]MPP through CCF from OCT2-expressing cells. These data support the contention that theseinhibitors of OCT2, including diphenidol and phenyltoloxam-ine, both members of the cluster II series of compounds

described by Kido et al. (24), are not transported substrates ofOCT2.

Chemical Assessment of OCT2-Mediated Metforminand Diphenidol Transport

The failure of phenyltoloxamine and diphenidol to supportCCF was unexpected; both have structural characteristics com-monly found in compounds shown to be substrates for OCT2(small, mildly hydrophobic, and cationic at physiological pH).Indeed, as noted earlier, both were included in this studybecause they were explicitly listed as “cluster II-inhibitoryligands,” i.e., the class in which most OCT substrates are found(24). The failure to support CCF is, of course, a “negativeresult.” Therefore, it was desirable to get independent confir-mation whether a “CCF-negative” ligand can, nevertheless, bea transported substrate. Unfortunately, neither compound isavailable in radiolabeled form, nor are they fluorescent. Con-sequently, we used direct chemical detection (using liquidchromatography with tandem mass spectroscopy) to determinethe extent to which hOCT2 supports accumulation of dipheni-dol. To provide a “positive control,” accumulation of met-formin, a known substrate of OCT2 that does support CCF(Fig. 7), was determined in parallel with the measurement ofdiphenidol accumulation. OCT2-expressing and wild-typeCHO cells were exposed for 5 min to either metformin ordiphenidol, in the presence or absence of 1 mM MPP. Theconcentration of each test substrate was set at its IC50 value,which should be sufficient to drive OCT2-mediated transport at�50% of its maximal value. As shown in Fig. 9, OCT2supported a robust accumulation of metformin that wasblocked by 1 mM MPP to a level equal to that measured inwild-type CHO cells. In contrast, there was no inhibitable,OCT2-specific accumulation of diphenidol.

Fig. 7. CCF as elicited by known OCT2 substrates and by compounds that haveno interaction with OCT2. Before the measurement of CCF, cells wereexposed to 12 nM [3H]MPP for 30 min. At time 0, the medium was rinsed andreplaced with medium containing either 12 nM [3H]MPP (control) or 12 nM[3H]MPP plus the indicated compound. For the compounds known to besubstrates of OCT2, the concentration of unlabeled compound was �10 timesthe IC50 measured in the present study (MPP, 100 �M; TEA, 1 mM;metformin, 10 mM; cimetidine, 2 mM). Each of the nontransported compoundswas present at a concentration 10 mM. The height of each bar represents themean (�SE) of [3H]MPP remaining in cells after a 60-s efflux period,normalized to control content. Data shown are from 2–7 separate experiments(each run in triplicate). Bars marked with * were different from the control(P � 0.05).

Fig. 8. CCF as elicited by known OCT2 inhibitory ligands. Before measure-ment of CCF, cells were exposed to 12 nM [3H]MPP for 30 min. At time 0, themedium was rinsed and replaced with medium containing either 12 nM[3H]MPP (control) or 12 nM [3H]MPP plus the indicated compound (MPP,100 �M; adrenosterone, 100 �M ; carvedilol, 225 �M; diphenidol, 70 �M ;phenyltoloxamine, 60 �M). The height of each bar represents the mean (�SE)of [3H]MPP remaining in cells after a 60-s efflux period, normalized to controlcontent. Data shown are from 3–5 separate experiments (each run in triplicate).Bar marked with * (MPP) refers to the only compound that supportedsignificant CCF (P � 0.05).

F62 MECHANISMS OF LIGAND INTERACTION WITH HUMAN OCT2

AJP-Renal Physiol • doi:10.1152/ajprenal.00486.2012 • www.ajprenal.org

at University of A

rizona Health S

ciences Library on January 20, 2013http://ajprenal.physiology.org/

Dow

nloaded from

Page 12: Multiple Mechanisms of Ligand Interaction with the Human ......g of salmon sperm, 18 g of pOG44, and 2 g of pcDNA5/FRT/ V5-His TOPO containing an OCT2 (hOCT2) construct. The CHOhOCT2

DISCUSSION

Between 1964 and 1999, �8% of the drugs approved by theUS Food and Drug Administration were subsequently with-drawn from the US market (38), underscoring the value ofpredicting unwanted DDIs as early as possible in the drugdiscovery process. Indeed, the influence of drug interactions atthe level of membrane transport led regulatory agencies toissue guidance on conducting in vitro and clinical studiesduring drug development (30), which, in turn, has led todevelopment of decision trees to determine whether a drugcandidate may be a substrate or inhibitor of key transporterscommonly reported to be involved in DDIs (14). The accuracyof such tools depends, ultimately, on understanding the under-lying mechanism(s) of ligand interaction with the processesinvolved in drug transport.

Here, we assessed the mechanism of inhibitory interaction ofselected ligands with the human ortholog of OCT2, the firststep in renal secretion of many cationic drugs. In the absenceof contrary evidence, it is frequently assumed, either explicitly(e.g., Refs. 2 and 50) or tacitly (e.g., Ref. 37), that theinhibition of OCT-mediated transport produced by small cat-ions is competitive. However, our observations indicate thatsuch assumptions are unwarranted; we found examples of threedistinct kinetic mechanisms of ligand interaction with OCT2:competitive, noncompetitive, and mixed-type.

Metformin and cimetidine are cluster II substrates of hOCT2(e.g., Refs. 50 and 56) and were, in fact, competitive inhibitorsof hOCT2-mediated MPP transport (Fig. 4; Table 2). Indeed,one of the predictions of the subcluster analysis is that clusterII ligands should display “competitive binding” with hOCT2

(24). Despite the substantial attention given to studies of OCTselectivity in recent years, there are comparatively few otherexamples of kinetically competitive interactions of ligandswith OCT2 as assessed by determining the effect of the testagent on Kt and Jmax of the transported probe. This criterion isimportant to establish because the other commonly applied testfor competition between a transported substrate and an inhib-itory ligand, the Dixon plot (1/J vs. [I]) (e.g., Ref. 29 and 55),while capable of discriminating between competitive and non-competitive interactions, does not discriminate between com-petitive and certain types of mixed-type inhibitory interactions(47) (we expand on this in an upcoming section). With thiscaveat in mind, cimetidine and the antimalarial drug pyrimeth-amine have been shown to be competitive inhibitors ofhOCT2-mediated metformin transport (21, 27); cimetidine hasbeen shown to competitively inhibit hOCT2-mediated TEAtransport (5); and NMN has been shown to competitivelyinhibit rOCT2-mediated TEA transport (3). Each of theseindividual interactions indicate that the ligand pairs in questionshare a common binding site or, alternatively, a set of mutuallyexclusive binding sites on a larger binding surface within thetranslocation pathway of OCT2.

The noncompetitive interactions seen here between adren-osterone and carvedilol and hOCT2-mediated MPP transport(Fig. 5; Table 2) were, arguably, consistent with their presencewithin structural subclusters identified by Kido et al. (24) asbeing effectively free of known transported substrates ofOCT2. Adrenosterone, an endogenous steroid hormone thathas been promoted as a dietary supplement capable of reducingbody fat and increasing muscle mass (8), is a representativemember of cluster III, which consists primarily of neutralsterols. It is relevant to note, therefore, that previous studies ofcorticosterone’s inhibition of OCT2-mediated transport describedbehavior inconsistent with a competitive mechanism, includingevidence of a noncompetitive mode of interaction (5, 52).

Carvedilol was also a noncompetitive inhibitor of hOCT2-mediated MPP transport. Carvedilol, a nonselective blocker,is a cluster I-inhibitory ligand (24). These compounds aredistinguished from the cluster II and cluster III ligands by theirsize (MW generally �400) and flexibility (as characterized bytheir asphericity and comparatively large geometrical radii ofgyration and second-order principal static moments). Of 18known transported substrates of OCT2, none falls into thecluster I category, leading to speculation that cluster I ligandsmay act by occluding the substrate binding site (24). Thatdescription, however, implies that cluster I compounds andOCT2 substrates may occupy mutually exclusive binding sites,a profile inconsistent with the noncompetitive profile observedfor carvedilol.

The molecular basis of the noncompetitive inhibitory pro-files observed here for adrenosterone and carvedilol is notclear. Adrenosterone is not charged, and carvedilol’s pKa of�8 would result in a substantial fraction (�25%) being un-charged at physiological pH. Consequently, it is conceivablethese compounds could diffuse across the membrane and exertan inhibitory interaction at the cytoplasmic (trans) face ofOCT2 that could not readily be displaced by simply increasingthe extracellular (cis) concentration of MPP. Just such a sce-nario proved to be the basis of the apparent noncompetitiveprofile of quinine’s inhibition of rOCT2-mediated TEA trans-port (3).

Fig. 9. Uptake of metformin and diphenidol by OCT2 measured by liquidchromatography with tandem mass spectrometry. WT and hOCT2-expressingCHO cells were exposed for 5 min to buffers containing either metformin ordiphenidol. The concentration of each approximated its IC50 value for inhibi-tion of OCT2-mediated MPP transport (500 �M for metformin; 35 �M fordiphenidol). In parallel, cells were exposed to these same substrate concentra-tions plus 1 mM MPP (100 times its Kt value). The height of each bar reflectsthe mean (�SE) accumulation of test substrate in OCT2-expressing cells,expressed relative to content measured in WT cells. The data presented arefrom 1 of 2 separate experiments, each measured in triplicate. Absolute WTaccumulations in the 2 experiments were WT metformin, 35.2 � 14.4 pmol/cm; WT metformin w/block, 27.0 � 14.3 pmol/cm; WT diphenidol, 79.1 �14.5 pmol/cm; WT diphenidol w/block, 102 � 15 pmol/cm.

F63MECHANISMS OF LIGAND INTERACTION WITH HUMAN OCT2

AJP-Renal Physiol • doi:10.1152/ajprenal.00486.2012 • www.ajprenal.org

at University of A

rizona Health S

ciences Library on January 20, 2013http://ajprenal.physiology.org/

Dow

nloaded from

Page 13: Multiple Mechanisms of Ligand Interaction with the Human ......g of salmon sperm, 18 g of pOG44, and 2 g of pcDNA5/FRT/ V5-His TOPO containing an OCT2 (hOCT2) construct. The CHOhOCT2

TEA is a prototypic cluster II substrate of hOCT2 (11, 16),so it was somewhat unexpected to find that, along with di-phenidol and phenyltoloxamine, TEA’s inhibition of hOCT2-mediated MPP transport displayed a mixed-type, rather than acompetitive profile (Fig. 5; Table 2). As small, globularcharged ligands, all three of these compounds would fall intocluster II, along with the other known substrates of OCT2 (24).As noted earlier, the cluster II ligands cimetidine and NMNinteract competitively with OCT2-mediated TEA transport,consistent with the view that these three compounds interact ata common binding site or, as noted earlier, a set of mutuallyexclusive binding sites within a larger binding surface. Basedon the observation that a site-directed mutation in rOCT1(D475E) changed the kinetics of transport of some (e.g., TEAand choline), but not all (MPP) substrates, Gorboulev et al.(17) suggested “that rOCT1 contains a large cation-bindingpocket with several interaction domains.” We subsequentlyfound that a single site-directed replacement (E447L) alteredthe interaction of TEA and cimetidine with rbOCT2 but had noeffect on the interaction MPP had with the transporter (54).These observations, along with others (see Ref. 26), supportthe view that (some) ligands can bind simultaneously to OCT2at sites that are spatially distinct. When these sites overlap oneanother, competitive (mutually exclusive) interactions are ex-pected, consistent with the MPP-metformin and MPP-cimeti-dine interactions noted here, and with the metformin-cimeti-dine, metformin-pyrimethamine, TEA-cimetidine, and TEA-NMN interactions noted previously (3, 5, 21, 27). However,ligand binding sites that are spatially distinct, but which exertan allosteric influence on one another, may display mixed-typeinhibitor profiles of the type observed here for inhibition ofhOCT2-mediated MPP transport by TEA, diphenidol, andphenyltoloxamine. Indeed, several studies have presented evi-dence supporting the presence of multiple binding sites forselected OCT ligands (18, 34).

Mixed-type inhibition of the type seen here can reflectseveral types of ligand interaction, but they all involve simul-taneous binding of substrate and inhibitor (or second sub-strate). The simplest interaction is one in which 1) binding ofthe inhibitor results in a complex that has reduced (althoughfinite) affinity for substrate, but 2) the resulting transporter/substrate/inhibitor complex does not support transport.1 Theequilibria describing this type of interaction (linear mixed-typeinhibition or LMT) (47) are shown below

To + S1o ToS1o Ti + S1i

S2o

++

S2o

ToS2o + S1o

Ti + S2i

ToS2oS1o

Kt

αKtKi αKi

To and Ti indicate the transporter in its outwardly vs. inwardlyfacing conformations; S1 and S2 are two transported substrates(e.g., MPP and TEA); Kt and Ki are the Michaelis constants forinteraction of S1 and S2, respectively, with the transporter; and� is the factor by which Kt (or Ki) changes (via allostery) whenthe other substrate binds to the transporter. In linear mixed-type inhibition, � � 1 � � (if � 1, inhibition is “noncom-petitive”; if � �, inhibition is “competitive”). Examinationof this scheme reveals that during exposure to S1, if S2 ispresent, some of the transporter will be in the nontransportingToS2oS1o form, even at infinitely high concentrations of S1, soJmax-i will be less than Jmax. Also, any concentration of S2 willresult in a fraction of the transporters available for interactionwith S1 being in the lower affinity ToS2o form, resulting in anincrease in apparent Kt. A key prediction of LMT inhibition isthat increasing concentrations of S2 will result in completeinhibition of S1 transport; other schemes of mixed-type inhi-bition (e.g., hyperbolic mixed-type) or partial competitiveinhibition are characterized by the inability of the inhibitionligand to completely block transport of S1 (47). Thus it isrelevant to emphasize that concentrations of TEA, diphenidol,and phenyltoloxamine �20-times larger than their respectiveIC50 values reduced hOCT2-mediated [3H]MPP transport tolevels not different from that produced by similar concentra-tions of unlabeled MPP. In other words, hOCT2 does notappear to support “cotransport” of two bound substrates; whentwo substrates bind, translocation of one requires dissociationof the other.

Our evidence suggests that ligand interaction with hOCT2can involve multiple, nonoverlapping sites, which has impor-tant implications with respect to the development of predictivemodels of drug-drug interaction with OCT2 and its role as theentry step in renal OC secretion. Structure-activity relation-ships are typically based on the degree to which a structurallydiverse array of test compounds inhibits biological activity(e.g., transport). However, if ligands can interact with multiplesites it introduces the potential for “substrate-dependent” in-hibitory profiles. Therefore, the profile of inhibitory interac-tions with hOCT2 reported here using MPP as a substratemight be expected to differ markedly had another substrate(e.g., TEA, or ASP) been used. Consequently, models of thespatial arrangements of structural features required for ligandinteraction with multidrug transporters, i.e., pharmacophores,are likely to vary depending on the substrate used. This hasbeen shown to be the case for P-gp (e.g., Refs. 13 and 41), andwe recently presented evidence suggesting this is true forMATE1 (4). The present study argues for a more concertedeffort to evaluate the kinetic mechanism of ligand interaction

1 The stipulation that the simultaneous binding of two ligands results in anontransporting complex runs counter to the observation of Koepsell andcolleagues (18, 34) that OCTs have high- and low-affinity binding sites forselected ligands and that transport may occur when both are occupied. Wesuggest, however, that the high-affinity site(s) characterized in these studies,which display KD values in the picomolar range, place them in a differentkinetic category from the multiple sites in the present model. As noted by theseauthors (34), dissociation of ligands from these high-affinity sites can beexpected to have half-times of many hours and, consequently, these sites aremore likely to play a modulatory role to the transport process. The modeloutlined here suggests an interaction of ligands with multiple sites from whichtransport may occur.

F64 MECHANISMS OF LIGAND INTERACTION WITH HUMAN OCT2

AJP-Renal Physiol • doi:10.1152/ajprenal.00486.2012 • www.ajprenal.org

at University of A

rizona Health S

ciences Library on January 20, 2013http://ajprenal.physiology.org/

Dow

nloaded from

Page 14: Multiple Mechanisms of Ligand Interaction with the Human ......g of salmon sperm, 18 g of pOG44, and 2 g of pcDNA5/FRT/ V5-His TOPO containing an OCT2 (hOCT2) construct. The CHOhOCT2

with OCT2, and other multidrug transporters, in efforts todevelop robust, predictive models of drug binding to thiscritical element in renal OC secretion.

The present results showed that possession of cluster IIstructural characteristics is insufficient to ensure that a ligandwill have a competitive interaction with other cluster II com-pounds. Importantly, they also showed that possession of suchcharacteristics is insufficient to support predictions concerningthe extent to which the ligand is itself a substrate for transport.As noted by Kido et al. (24), known hOCT2 substrates arecongregated in cluster II. Diphenidol and phenyltoloxaminewere included in our study because of their designation ascluster II-inhibitory ligands of hOCT2 (24), and we weresurprised to find that neither supported CCF. Adrenosteroneand corticosterone also appear to be nontransported inhibitorsof OCT2, but their cluster III characteristics are consistent withthis observation. The antiretroviral drug nelfinavir is also anon-transported, high-affinity (IC50 of 13 �M) inhibitor ofhOCT2-mediated MPP transport (22), but its large size (MW567) and weak charge (pKa of �7) make it a poor fit for acluster II designation. Diphenidol and phenyltoloxamine, incontrast, are models of the structural characteristics of clusterII (24). Their failure to support CCF was sufficiently unex-pected that we used liquid chromatography with tandem massspectroscopy to measure directly the effect of expression ofhOCT2 on accumulation of diphenidol (Fig. 9), with the resultssupporting the same conclusion: diphenidol (and, by extension,phenyltoloxamine) is not a transported substrate of hOCT2,despite general structural features that would seem to predictthat it be.

CCF offers an important advantage over the “trans-stimula-tion” strategy more commonly used to test indirectly whethera compound can serve as a transported substrate. The trans-assay as commonly applied involves preloading the cells witha suspected substrate and then determining whether the pres-ence of this outwardly directed gradient stimulates uptake of alabeled substrate (the alternative approach is to preload thecells with the labeled substrate and then measure efflux of thelabel following exposure to the test agent in the extracellularsolution). In either case, a trans-stimulation of flux, althoughnot “proof” that the test compound is a substrate (it could be an“allosteric stimulator”), is generally considered strong support-ive evidence that it is (e.g., Ref. 12). However, failure toobserve a trans-effect is not sufficient to conclude that the testcompound is a not a substrate. A uniporter like OCT2 canundergo the conformational changes associated with translo-cation either when occupied by substrate or when “empty”(thereby accounting for the documented electrogenicity ofOCT-mediated OC transport) (9). Trans-stimulation only oc-curs if the “occupied” transporter undergoes these conforma-tional changes more readily than the empty transporter. Al-though that may be “probable,” such behavior is not assured.CCF makes no such assumptions. If an extracellular test agentcan be transported, it will induce a net loss of substrate if thesubstrate is distributed at steady state.

There is a limitation to the use of CCF as an indirect measureof transport: it cannot provide a precise measure of the rate oftest agent transport (because the rapid increase in intracellularconcentration of the inhibitory agent can be expected to beginto reduce the rate of labeled substrate efflux with kinetics thatare ill defined). With this caveat, we suggest that CCF can

provide a means to establish a high-throughput assay capableof collecting data required to develop a predictive model ofstructural features that permit the ligand/transporter complex toundergo the conformational changes that result in successfulsubstrate translocation.

In summary, ligand interaction with hOCT2 involves acomplex array of kinetic mechanisms, including competitive,noncompetitive, and mixed-type mechanisms that are not ap-parent in selectivity assays based on simple inhibition oftransport activity. In addition, using CCF, several compoundspredicted to be transported substrates of hOCT2 based on theirinhibition of hOCT2 were shown to be nontransported inhibi-tors. CCF is a comparatively simple assay capable of deter-mining whether an inhibitory ligand is also capable of servingas a transported substrate, providing a potential means fordeveloping models of the structural determinants required tosupport effective transporter turnover.

ACKNOWLEDGMENTS

The authors gratefully acknowledge the technical contributions of MarkMorales and Alexandra Cooke, and useful discussions with Dr. WilliamDantzler.

GRANTS

This work was supported by National Institutes of Health (NIH) award5R01DK58251 and NIH Grants 5P30E006694 and 5T32HL07249.

DISCLOSURES

No conflicts of interest, financial or otherwise, are declared by the authors.

AUTHOR CONTRIBUTIONS

Author contributions: J.N.H. performed experiments; J.N.H. and S.H.W.analyzed data; J.N.H. and S.H.W. interpreted results of experiments; J.N.H.and S.H.W. prepared figures; J.N.H. drafted manuscript; S.H.W. providedconception and design of research; S.H.W. edited and revised manuscript;S.H.W. approved final version of manuscript.

REFERENCES

1. Ahlin G, Karlsson J, Pedersen JM, Gustavsson L, Larsson R, MatssonP, Norinder U, Bergstrom CA, Artursson P. Structural requirements fordrug inhibition of the liver specific human organic cation transport protein.J Med Chem 51: 5932–5942, 2008.

2. Amphoux A, Vialou V, Drescher E, Bruss M, La Cour CM, Rochat C,Millan MJ, Giros B, Bonisch H, Gautron S. Differential pharmacolog-ical in vitro properties of organic cation transporters and regional distri-bution in rat brain. Neuropharmacology 50: 941–952, 2006.

3. Arndt P, Volk C, Gorboulev V, Budiman T, Popp C, Ulzheimer-Teuber I, Akhoundova A, Koppatz S, Bamberg E, Nagel G, KoepsellH. Interaction of cations, anions, and weak base quinine with rat renalcation transporter rOCT2 compared with rOCT1. Am J Physiol RenalPhysiol 281: F454–F468, 2001.

4. Astorga B, Ekins S, Morales M, Wright SH. Molecular determinants ofligand selectivity for the human multidrug and toxin extrusion proteins,MATE1 and MATE-2K. J Pharmacol Exp Ther 341: 743–755, 2012.

5. Barendt WM, Wright SH. The human organic cation transporter(hOCT2) recognizes the degree of substrate ionization. J Biol Chem 277:22491–22496, 2002.

6. Bednarczyk D, Ekins S, Wikel JH, Wright SH. Influence of molecularstructure on substrate binding to the human organic cation transporter,hOCT1. Mol Pharm 63: 489–498, 2003.

7. Bednarczyk D, Mash EA, Aavula BR, Wright SH. NBD-TMA: a novelfluorescent substrate of the peritubular organic cation transporter of renalproximal tubules. Pflügers Arch 440: 184–192, 2000.

8. Brooker L, Parr MK, Cawley A, Flenker U, Howe C, Kazlauskas R,Schanzer W, George A. Development of criteria for the detection ofadrenosterone administration by gas chromatography-mass spectrometryand gas chromatography-combustion-isotope ratio mass spectrometry fordoping control. Drug Test Anal 1: 587–595, 2009.

F65MECHANISMS OF LIGAND INTERACTION WITH HUMAN OCT2

AJP-Renal Physiol • doi:10.1152/ajprenal.00486.2012 • www.ajprenal.org

at University of A

rizona Health S

ciences Library on January 20, 2013http://ajprenal.physiology.org/

Dow

nloaded from

Page 15: Multiple Mechanisms of Ligand Interaction with the Human ......g of salmon sperm, 18 g of pOG44, and 2 g of pcDNA5/FRT/ V5-His TOPO containing an OCT2 (hOCT2) construct. The CHOhOCT2

9. Budiman T, Bamberg E, Koepsell H, Nagel G. Mechanism of electro-genic cation transport by the cloned organic cation transporter 2 from rat.J Biol Chem 275: 29413–29420, 2000.

10. Busch AE, Quester S, Ulzheimer JC, Waldegger S, Gorboulev V,Arndt P, Lang F, Koepsell H. Electrogenic properties and substratespecificity of the polyspecific rat cation transporter rOCT1. J Biol Chem271: 32599–32604, 1996.

11. Dresser MJ, Gray AT, Giacomini KM. Kinetic and selectivity differ-ences between rodent, rabbit, and human organic cation transporters(OCT1). J Pharmacol Exp Ther 292: 1146–1152, 2000.

12. Dresser MJ, Xiao G, Leabman MK, Gray AT, Giacomini KM. Inter-actions of n-tetraalkylammonium compounds and biguanides with a hu-man renal organic cation transporter (hOCT2). Pharm Res 19: 1244–1247,2002.

13. Garrigues A, Loiseau N, Delaforge M, Ferte J, Garrigos M, Andre F,Orlowski S. Characterization of two pharmacophores on the multidrugtransporter P-glycoprotein. Mol Pharmacol 62: 1288–1298, 2002.

14. Giacomini KM, Huang SM, Tweedie DJ, Benet LZ, Brouwer KL, ChuX, Dahlin A, Evers R, Fischer V, Hillgren KM, Hoffmaster KA,Ishikawa T, Keppler D, Kim RB, Lee CA, Niemi M, Polli JW,Sugiyama Y, Swaan PW, Ware JA, Wright SH, Wah YS, Zamek-Gliszczynski MJ, Zhang L. Membrane transporters in drug development.Nat Rev Drug Discov 9: 215–236, 2010.

15. Gorboulev V, Shatskaya N, Volk C, Koepsell H. Subtype-specificaffinity for corticosterone of rat organic cation transporters rOCT1 andrOCT2 depends on three amino acids within the substrate binding region.Mol Pharmacol 67: 1612–1619, 2005.

16. Gorboulev V, Ulzheimer JC, Akhoundova A, Ulzheimer-Teuber I,Karbach U, Quester S, Baumann C, Lang F, Busch AE, Koepsell H.Cloning and characterization of two human polyspecific organic cationtransporters. DNA Cell Biol 16: 871–881, 1997.

17. Gorboulev V, Volk C, Arndt P, Akhoundova A, Koepsell H. Selectivityof the polyspecific cation transporter rOCT1 is changed by mutation ofaspartate 475 to glutamate. Mol Pharmacol 56: 1254–1261, 1999.

18. Gorbunov D, Gorboulev V, Shatskaya N, Mueller T, Bamberg E,Friedrich T, Koepsell H. High-affinity cation binding to transporterOCT1 induces movement of helix 11 and blocks transport after mutationsin a modelled interaction domain between two helices. Mol Pharmacol 73:50–61, 2008.

19. Gründemann D, Gorboulev V, Gambaryan S, Veyhl M, Koepsell H.Drug excretion mediated by a new prototype of polyspecific transporter.Nature 372: 549–552, 1994.

20. Hayer-Zillgen M, Bruss M, Bonisch H. Expression and pharmacologicalprofile of the human organic cation transporters hOCT1, hOCT2 andhOCT3. Br J Pharmacol 136: 829–836, 2002.

21. Ito S, Kusuhara H, Yokochi M, Toyoshima J, Inoue K, Yuasa H,Sugiyama Y. Competitive inhibition of the luminal efflux by MATEs, butnot basolateral uptake by OCT2, is the likely mechanism underlying thepharmacokinetic drug-drug interactions caused by cimetidine in the kid-ney. J Pharmacol Exp Ther 340: 393–403, 2012.

22. Jung N, Lehmann C, Rubbert A, Knispel M, Hartmann P, vanLunzen J, Stellbrink HJ, Faetkenheuer G, Taubert D. Relevance of theorganic cation transporters 1 and 2 for antiretroviral therapy in humanimmunodeficiency virus infection. Drug Metab Dispos 36: 1616–1623,2008.

23. Kekuda R, Prasad PD, Wu X, Wang H, Fei YJ, Leibach FH, Ganapa-thy V. Cloning and functional characterization of a potential-sensitive,polyspecific organic cation transporter (OCT3) most abundantly expressedin placenta. J Biol Chem 273: 15971–15979, 1998.

24. Kido Y, Matsson P, Giacomini KM. Profiling of a prescription druglibrary for potential renal drug-drug interactions mediated by the organiccation transporter 2. J Med Chem 54: 4548–4558, 2011.

25. Kimura N, Masuda S, Tanihara Y, Ueo H, Okuda M, Katsura T, InuiK. Metformin is a superior substrate for renal organic cation transporterOCT2 rather than hepatic OCT1. Drug Metab Pharmacokinet 20: 379–386, 2005.

26. Koepsell H. Substrate recognition and translocation by polyspecific or-ganic cation transporters. Biol Chem 392: 95–101, 2011.

27. Kusuhara H, Ito S, Kumagai Y, Jiang M, Shiroshita T, Moriyama Y,Inoue K, Yuasa H, Sugiyama Y. Effects of a MATE protein inhibitor,pyrimethamine, on the renal elimination of metformin at oral microdoseand at therapeutic dose in healthy subjects. Clin Pharmacol Ther 89:837–844, 2011.

28. Lacko L, Burger M. Common carrier system for sugar transport in humanred cells. Nature 191: 881–882, 1961.

29. Lee WK, Reichold M, Edemir B, Ciarimboli G, Warth R, Koepsell H,Thevenod F. Organic cation transporters OCT1, 2, and 3 mediate high-affinity transport of the mutagenic vital dye ethidium in the kidneyproximal tubule. Am J Physiol Renal Physiol 296: F1504–F1513, 2009.

30. Lepist EI, Ray AS. Renal drug-drug interactions: what we have learnedand where we are going. Expert Opin Drug Metab Toxicol 8: 433–448,2012.

31. Malo C, Berteloot A. Analysis of kinetic data in transport studies: newinsights from kinetic studies of Na�-D-glucose cotransport in humanintestinal brush-border membrane vesicles using a fast sampling, rapidfiltration apparatus. J Membr Biol 122: 127–141, 1991.

32. Marcelin-Jimenez G, Morales-Martinez M, Angeles-Moreno AP,Mendoza-Morales L. Ultra-fast chromatographic micro-assay for quan-tification of diphenidol in plasma: application in an oral multi-doseswitchability trial. Biomed Chromatogr 22: 1143–1148, 2008.

33. Ming X, Ju W, Wu H, Tidwell RR, Hall JE, Thakker DR. Transport ofdicationic drugs pentamidine and furamidine by human organic cationtransporters. Drug Metab Dispos 37: 424–430, 2009.

34. Minuesa G, Volk C, Molina-Arcas M, Gorboulev V, Erkizia I, ArndtP, Clotet B, Pastor-Anglada M, Koepsell H, Martinez-Picado J. Trans-port of lamivudine [(�)-beta-L-2=,3=-dideoxy-3=-thiacytidine] and high-affinity interaction of nucleoside reverse transcriptase inhibitors withhuman organic cation transporters 1, 2, and 3. J Pharmacol Exp Ther 329:252–261, 2009.

35. Moaddel R, Ravichandran S, Bighi F, Yamaguchi R, Wainer IW.Pharmacophore modelling of stereoselective binding to the human organiccation transporter (hOCT1). Br J Pharmacol 151: 1305–1314, 2007.

36. Motohashi h Sakurai Y, Saito H, Masuda S, Urakami Y, Goto M,Fukatsu A, Ogawa O, Inui K. Gene expression levels and immunolo-calization of organic ion transporters in the human kidney. J Am SocNephrol 13: 866–874, 2002.

37. Muller J, Lips KS, Metzner L, Neubert RH, Koepsell H, Brandsch M.Drug specificity and intestinal membrane localization of human organiccation transporters (OCT). Biochem Pharmacol 70: 1851–1860, 2005.

38. Nassar AE, Talaat RE, Tokuno H. Drug interactions: concerns andcurrent approaches. IDrugs 10: 47–52, 2007.

39. Nies AT, Koepsell H, Damme K, Schwab M. Organic cation transporters(OCTs, MATEs), in vitro and in vivo evidence for the importance in drugtherapy. Handb Exp Pharmacol 201: 105–167, 2011.

40. Okuda M, Saito H, Urakami Y, Takano M, Inui KI. cDNA cloning andfunctional expression of a novel rat kidney organic cation transporter,OCT2. Biochem Biophys Res Commun 224: 500–507, 1996.

41. Pascaud C, Garrigos M, Orlowski S. Multidrug resistance transporterP-glycoprotein has distinct but interacting binding sites for cytotoxic drugsand reversing agents. Biochem J 333: 351–358, 1998.

42. Pelis RM, Dangprapai Y, Wunz TM, Wright SH. Inorganic mercuryinteracts with cysteine residues (C451 and C474) of hOCT2 to reduce itstransport activity. Am J Physiol Renal Physiol 292: F1583–F1591, 2007.

43. Pelis RM, Wright SH. Renal transport of organic anions and cations.Compr Physiol 1: 1795–1835, 2011.

44. Rosenberg T, Wilbrandt W. Uphill transport induced by counterflow. JGen Physiol 41: 289–296, 1957.

45. Ross CR, Holohan PD. Transport of organic anions and cations inisolated renal plasma membranes. Annu Rev Pharmacol Toxicol 23:65–85, 1983.

46. Schomig E, Lazar A, Grundemann D. Extraneuronal monoamine trans-porter and organic cation transporters 1 and 2: a review of transportefficiency. Handb Exp Pharmacol 175: 151–180, 2006.

47. Segel IH. Enzyme Kinetics. New York: Wiley, 1975, p. 161–192.48. Stein WD. The Movement of Molecules Across Cell Membranes. New

York: Academic, 1967, p. 143–148.49. Suhre WM, Ekins S, Chang C, Swaan PW, Wright SH. Molecular

determinants of substrate/inhibitor binding to the human and rabbit renalorganic cation transporters, hOCT2 and rbOCT2. Mol Pharmacol 67:1067–1077, 2005.

50. Tahara H, Kusuhara H, Endou H, Koepsell H, Imaoka T, Fuse E,Sugiyama Y. A species difference in the transport activities of h2 receptorantagonists by rat and human renal organic anion and cation transporters.J Pharmacol Exp Ther 315: 337–345, 2005.

F66 MECHANISMS OF LIGAND INTERACTION WITH HUMAN OCT2

AJP-Renal Physiol • doi:10.1152/ajprenal.00486.2012 • www.ajprenal.org

at University of A

rizona Health S

ciences Library on January 20, 2013http://ajprenal.physiology.org/

Dow

nloaded from

Page 16: Multiple Mechanisms of Ligand Interaction with the Human ......g of salmon sperm, 18 g of pOG44, and 2 g of pcDNA5/FRT/ V5-His TOPO containing an OCT2 (hOCT2) construct. The CHOhOCT2

51. Varma MV, Feng B, Obach RS, Troutman MD, Chupka J, Miller HR,El-Kattan A. Physicochemical determinants of human renal clearance. JMed Chem 52: 4844–4852, 2009.

52. Volk C, Gorboulev V, Budiman T, Nagel G, Koepsell H. Differentaffinities of inhibitors to the outwardly and inwardly directed substratebinding site of organic cation transporter 2. Mol Pharmacol 64: 1037–1047, 2003.

53. Wright SH, Wunz TM, Wunz TP. Structure and interaction of inhibitorswith the TEA/H� exchanger of rabbit renal brush border membranes.Pflügers Arch 429: 313–324, 1995.

54. Zhang X, Shirahatti NV, Mahadevan D, Wright SH. A conservedglutamate residue in transmembrane helix 10 influences substrate speci-ficity of rabbit OCT2 (SLC22A2). J Biol Chem 280: 34813–34822, 2005.

55. Zolk O, Solbach TF, Konig J, Fromm MF. Functional characterizationof the human organic cation transporter 2 variant p.270Ala�Ser. DrugMetab Dispos 37: 1312–1318, 2009.

56. Zolk O, Solbach TF, Konig J, Fromm MF. Structural determinants ofinhibitor interaction with the human organic cation transporter OCT2(SLC22A2). Naunyn Schmiedebergs Arch Pharmacol 379: 337–348,2009.

F67MECHANISMS OF LIGAND INTERACTION WITH HUMAN OCT2

AJP-Renal Physiol • doi:10.1152/ajprenal.00486.2012 • www.ajprenal.org

at University of A

rizona Health S

ciences Library on January 20, 2013http://ajprenal.physiology.org/

Dow

nloaded from