Skip to main content
Log in

How Drugs Interact with Transporters: SGLT1 as a Model

  • Published:
Journal of Membrane Biology Aims and scope Submit manuscript

Abstract

Drugs are transported by cotransporters with widely different turnover rates. We have examined the underlying mechanism using, as a model system, glucose and indican (indoxyl-β-d-glucopyranoside) transport by human Na+/glucose cotransporter (hSGLT1). Indican is transported by hSGLT1 at 10% of the rate for glucose but with a fivefold higher apparent affinity. We expressed wild-type hSGLT1 and mutant G507C in Xenopus oocytes and used electrical and optical methods to measure the kinetics of glucose (using nonmetabolized glucose analogue α-methyl-d-glucopyranoside, αMDG) and indican transport, alone and together. Indican behaved as a competitive inhibitor of αMDG transport. To examine protein conformations, we recorded SGLT1 capacitive currents (charge movements) and fluorescence changes in response to step jumps in membrane voltage, in the presence and absence of indican and/or αMDG. In the absence of sugar, voltage jumps elicited capacitive SGLT currents that decayed to steady state with time constants (τ) of 3–20 ms. These transient currents were abolished in saturating αMDG but only slightly reduced (10%) in saturating indican. SGLT1 G507C rhodamine fluorescence intensity increased with depolarizing and decreased with hyperpolarizing voltages. Maximal fluorescence increased ∼150% in saturating indican but decreased ∼50% in saturating αMDG. Modeling indicated that the rate-limiting step for indican transport is sugar translocation, whereas for αMDG it is dissociation of Na+ from the internal binding sites. The inhibitory effects of indican on αMDG transport are due to its higher affinity and a 100-fold lower translocation rate. Our results indicate that competition between substrates and drugs should be taken into consideration when targeting transporters as drug delivery systems.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10

Similar content being viewed by others

References

  • Bezanilla F (2000) The voltage sensor in voltage-dependent ion channels. Physiol Rev 80:555–592

    PubMed  CAS  Google Scholar 

  • Birnir B, Loo DDF, Wright EM (1991) Voltage-clamp studies of the Na+/glucose cotransporter cloned from rabbit small intestine. Pfluegers Arch 418:79–85

    Article  CAS  Google Scholar 

  • Briasoulis E, Judson I, Pavlidis N, Beale P, Wanders J, Groot Y, Veerman G, Schuessler M, Niebch G, Siamopoulos K, Tzamakou E, Rammou D, Wolf L, Walker R, Hanauske A (2000) Phase I trial of 6-hour infusion of glufosfamide, a new alkylating agent with potentially enhanced selectivity for tumors that overexpress transmembrane glucose transporters: a study of the European Organization for Research and Treatment of Cancer Early Clinical Studies Group. J Clin Oncol 18:3535–3544

    PubMed  CAS  Google Scholar 

  • Díez-Sampedro A, Lostao MP, Wright EM, Hirayama BA (2000) Glycoside binding and translocation in Na+-dependent glucose cotransporters: comparison of SGLT1 and SGLT3. J Membr Biol 176:111–117

    Article  PubMed  Google Scholar 

  • Díez-Sampedro A, Wright EM, Hirayama BA (2001) Residue 457 controls sugar binding and transport in the Na+/glucose cotransporter. J Biol Chem 276:49188–49194

    Article  PubMed  CAS  Google Scholar 

  • Eskandari S, Wright EM, Loo DDF (2005) Kinetics of the reverse mode of the Na+/glucose cotransporter. J Membr Biol 204:23–32

    Article  PubMed  CAS  Google Scholar 

  • Falk S, Guay A, Chenu C, Patil SD, Berteloot A (1998) Reduction of an eight-state mechanism of cotransport to a six-state model using a new computer program. Biophys J 74:816–830

    Article  PubMed  CAS  Google Scholar 

  • Hazama A, Loo DDF, Wright EM (1997) Pre-steady-state currents of the Na+/glucose cotransporter (SGLT1). J Membr Biol 155:175–186

    Article  PubMed  CAS  Google Scholar 

  • Hirayama BA, Díez-Sampedro A, Wright EM (2001) Common mechanisms of inhibition for the Na+/glucose (hSGLT1) and Na+/Cl-/GABA (hGAT1) cotransporters. Br J Pharmacol 134:484–495

    Article  PubMed  CAS  Google Scholar 

  • Hirayama BA, Loo DDF, Díez-Sampedro A, Leung DW, Meinild A-K, Lai-Bing M, Turk E, Wright EM (2007) Na-dependent reorganization of the sugar-binding site of SGLT1. Biochemistry 46:13391–13406

    Article  PubMed  CAS  Google Scholar 

  • Hirayama BA, Loo DDF, Wright EM (1997) Cation effects on protein conformations and transport in the Na+/glucose cotransporter. J Biol Chem 272:2110–2115

    Article  PubMed  CAS  Google Scholar 

  • Hirayama BA, Lostao MP, Panayotova-Heiermann M, Loo DDF, Wright EM (1996) Kinetic and specificity differences between rat, human and rabbit Na+-glucose transporter (SGLT-1). Am J Physiol 270:G919–G926

    PubMed  CAS  Google Scholar 

  • King AE, Ackley MA, Cass CE, Young JD, Baldwin SA (2006) Nucleoside transporters: from scavengers to novel therapeutic targets. Trends Pharmacol Sci 27:416–425

    Article  PubMed  CAS  Google Scholar 

  • Knutter I, Hartrodt B, Toth G, Keresztes A, Kottra G, Mrestani-Klaus C, Born I, Daniel H, Neubert K, Brandsch M (2007) Synthesis and characterization of a new and radiolabeled high-affinity substrate for H+/peptide cotransporters. FEBS J 274:5905–5914

    Article  PubMed  CAS  Google Scholar 

  • Larrayoz IM, Casado FJ, Pastor-Anglada M, Lostao MP (2004) Electrophysiological characterization of the human Na+/nucleoside cotransporter 1 (hCNT1) and role of adenosine on hCNT1 function. J Biol Chem 279:8999–9007

    Article  PubMed  CAS  Google Scholar 

  • Loewen SK, Yao SY, Slugoski MD, Mohabir NN, Turner RJ, Mackey JR, Weiner JH, Gallagher MP, Henderson PJ, Baldwin SA, Cass CE, Young JD (2004) Transport of physiological nucleosides and anti-viral and anti-neoplastic nucleoside drugs by recombinant Escherichia coli nucleoside-H+ cotransporter (NupC) produced in Xenopus laevis oocytes. Mol Membr Biol 21:1–10

    Article  PubMed  CAS  Google Scholar 

  • Loo DDF, Hazama A, Supplisson S, Turk E, Wright EM (1993) Relaxation kinetics of the Na+/glucose cotransporter. Proc Natl Acad Sci USA 90:5767–5771

    Article  PubMed  CAS  Google Scholar 

  • Loo DDF, Hirayama BA, Gallardo E, Lam J, Turk E, Wright EM (1998) Conformational changes couple Na+ and glucose transport. Proc Natl Acad Sci USA 95:7789–7794

    Article  PubMed  CAS  Google Scholar 

  • Loo DDF, Eskandari S, Hirayama BA, Wright EM (2002) A kinetic model for secondary active transport. In: Layton HE, Weinstein AM (eds), Membrane transport and renal physiology, the IMA volumes in mathematics and its applications, vol 129, Springer-Verlag, New York, pp 65–83

    Google Scholar 

  • Loo DDF, Hirayama BA, Cha A, Bezanilla F, Wright EM (2005) Perturbation analysis of the voltage-sensitive conformational changes of the Na+/glucose cotransporter. J Gen Physiol 125:13–36

    Article  PubMed  CAS  Google Scholar 

  • Loo DDF, Hirayama BA, Karakossian MH, Meinild A-K, Wright EM (2006) Conformational dynamics of hSGLT1 during Na+/glucose cotransport. J Gen Physiol 128:701–720

    Article  PubMed  CAS  Google Scholar 

  • Loo DDF, Hirayama BA, Meinild A-K, Chandy G, Zeuthen T, Wright EM (1999) Passive water and ion transport by cotransporters. J Physiol 518:195–202

    Article  PubMed  CAS  Google Scholar 

  • Lostao MP, Hirayama BA, Loo DDF, Wright EM (1994) Phenylglucosides and the Na+/glucose cotransporter (SGLT1): analysis of interactions. J Membr Biol 142:161–170

    PubMed  CAS  Google Scholar 

  • Mackenzie B, Loo DDF, Wright EM (1998) Relationships between Na+/glucose cotransporter (SGLT1) currents and fluxes. J Membr Biol 162:101–106

    Article  PubMed  CAS  Google Scholar 

  • Mackey JR, Yao SYM, Smith KM, Karpinski E, Baldwin SA, Cass CE, Young JD (1999) Gemcitabine transport in Xenopus oocytes expressing recombinant plasma membrane mammalian nucleoside transporters. J Natl Cancer Inst 91:1876–1881

    Article  PubMed  CAS  Google Scholar 

  • Meinild A-K, Hirayama BA, Wright EM, Loo DDF (2002) Fluorescence studies of ligand-induced conformational changes of the Na+/glucose cotransporter. Biochemistry 41:1250–1258

    Article  PubMed  CAS  Google Scholar 

  • Parent L, Supplisson S, Loo DDF, Wright EM (1992) Electrogenic properties of the cloned Na+/glucose cotransporter: II. A transport model under nonrapid equilibrium conditions. J Membr Biol 125:63–79

    PubMed  CAS  Google Scholar 

  • Quick M, Loo DDF, Wright EM (2001) Neutralization of a conserved amino acid residue in the human Na+/glucose transporter (hSGLT1) generates a glucose-gated H+ channel. J Biol Chem 276:1728–1734

    Article  PubMed  CAS  Google Scholar 

  • Sala-Rabanal M, Loo DDF, Hirayama BA, Turk E, Wright EM (2006) Molecular interactions between dipeptides, drugs and the human intestinal H+-oligopeptide cotransporter hPEPT1. J Physiol 574:149–166

    Article  PubMed  CAS  Google Scholar 

  • Sala-Rabanal M, Loo DDF, Hirayama BA, Wright EM (2008) Molecular mechanism of dipeptide and drug transport by the human renal H+/oliopeptide cotransporter, hPEPT2. Am J Physiol 294:F1422–F1432

    Article  CAS  Google Scholar 

  • Smith KM, Ng AML, Yao SYM, Labedz KA, Knaus EE, Wiebe LI, Cass CE, Baldwin SA, Chen X-Z, Karpinski E, Young JD (2004) Electrophysiological characterization of a recombinant human Na+-coupled nucleoside transporter (hCNT1) produced in Xenopus oocytes. J Physiol 558:807–823

    Article  PubMed  CAS  Google Scholar 

  • Veenstra M, Lanza S, Hirayama BA, Turk E, Wright EM (2004) Local conformational changes in the Vibrio Na+/galactose cotransporter. Biochemistry 43:3620–3627

    Article  PubMed  CAS  Google Scholar 

  • Veyhl M, Wagner K, Volk C, Gorboulev V, Baumgarten K, Weber WM, Schaper M, Bertram B, Wiessler M, Koepsell H (1998) Transport of the new chemotherapeutic agent beta-d-glucosylisophosphoramide mustard (D-19575) into tumour cells is mediated by the Na+-d-glucose cotransporter SAAT1. Proc Natl Acad Sci USA 95:2914–2919

    Article  PubMed  CAS  Google Scholar 

  • Yao SYM, Ng AML, Ritzel MWL, Gati WP, Cass CE, Young JD (1996) Transport of adenosine by recombinant purine- and pyrimidine-selective sodium/nucleoside cotransporters from rat jejunum expressed in Xenopus laevis oocytes. Mol Pharmacol 50:1529–1535

    PubMed  CAS  Google Scholar 

  • Zampighi GA, Kreman M, Boorer KJ, Loo DDF, Bezanilla F, Chandy G, Wright EM (1995) A method for determining the unitary functional capacity of cloned channels and transporters in Xenopus laevis oocytes. J Membr Biol 148:65–78

    PubMed  CAS  Google Scholar 

Download references

Acknowledgements

We thank Teresa Ku for the preparation, injection and care of oocytes. This research was supported by National Institutes of Health grant DK19567. M. S.-R. was supported by a postdoctoral scholarship from the Ministerio de Educacion y Ciencia (government of Spain).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Donald D. F. Loo.

Appendix

Appendix

This focuses on the formulation and simulation of the kinetic model for SGLT1 in the presence of indican and αMDG, alone and together.

Kinetic Model

The model is based on extensive experimental data (Parent et al. 1992; Loo et al. 1993, 1998, 2005, 2006; Hirayama et al. 1997, 2007; Meinild et al. 2002; Eskandari et al. 2005; Mackenzie et al. 1998, Zampighi et al. 1995). The transporter has eight kinetic states: ligand-free (empty), Na+-bound, Na+- and sugar-bound, and intermediates (Ca, Cb) between the inward- and outward-facing empty conformations (Fig. 8). The empty protein has a valence of −2, and the voltage-sensitive steps are the binding of external Na+ (C1 to C2Na2) and translocation of the empty carrier between the external and internal sides of the membrane (C1–C6).

Transitions between states Ci and Cj are represented by first-order (or pseudo-first-order in case of Na+ and sugar binding) rate constants k ij: Ci → Cj. Eyring rate theory is used to describe the dependence of rate constants on membrane potential. k ij = k oij exp(−εij F/RT), where k oij is a voltage-independent rate, εij is the equivalent charge movement (up to the transition state between Ci → Cj) and F, R and T have their usual physicochemical meanings (Parent et al. 1992). Na+ and sugar binding on external and internal membrane surfaces is represented by pseudo-rate constants k 12 = k o 12 [Na] 2 o exp(−ε12 FV/RT), k 23 = k o 23 [αMDG]o, k 27 = k o 27 [indican] o , k 65 = k o 65 [Na] 2 i and k 54 k o 54 [αMDG] i , k 85 = k o85 [indican] i . Na+ ions bind to two identical binding sites in one step before glucose (C1 + 2Na+ ⇄ C2Na2), and this is considered to be valid at high [Na+] o with high cooperativity between Na+ binding sites (Falk et al. 1998).

Pre-steady-state currents (hSGLT1 capacitive currents) are associated with the reorientation of the empty carrier between outward- and inward-facing conformations (C1 ⇄ Ca ⇄ Cb ⇄ C6) and external Na+ binding/dissociation (C1 ⇄ C2Na2). There are four predicted components of pre-steady-state currents (region I, Figs. 1 and 8): C2Na2 ⇄ C1, C1 ⇄ Ca, Ca ⇄ Cb, and Cb ⇄ C6. The current (I ij) associated transition Ci ⇄ Cj is given by I ij = eij + εji)(k ijCi k jiCj), where e is elementary charge (Loo et al. 2005; Parent et al. 1992).

Total current (I) associated with SGLT1 is

$$ \begin{aligned}{} & I{\text{ = }}N_{{\text{T}}} {\text{ (}}I_{{{\text{12}}}} + I_{{{\text{1a}}}} + I_{{{\text{ab}}}} + I_{{{\text{b6}}}} {\text{)}} \\ & {\text{ = }}N_{{\text{T}}} {\text{e[(}}\varepsilon _{{{\text{12}}}} + \varepsilon _{{{\text{21}}}} {\text{)(}}k_{{{\text{12}}}} {\text{C}}_{{\text{1}}} {-k}_{{{\text{21}}}} {\text{C}}_{{\text{2}}} {\text{Na}}_{{\text{2}}} ) + (\varepsilon _{{{\text{1a}}}} + \varepsilon _{{{\text{a1}}}} {\text{)(}}k_{{{\text{1a}}}} {\text{C}}_{{\text{1}}} {-k}_{{{\text{a1}}}} {\text{C}}_{{\text{a}}} ) + (\varepsilon _{{{\text{ab}}}} + \varepsilon _{{{\text{ba}}}} {\text{)(}}k_{{{\text{ab}}}} {\text{C}}_{{\text{a}}} {-k}_{{{\text{ba}}}} {\text{C}}_{{\text{b}}} ) + \\ & {\text{(}}\varepsilon _{{{\text{b6}}}} + \varepsilon _{{{\text{6b}}}} {\text{)(}}k_{{{\text{b6}}}} {\text{C}}_{{\text{b}}} {-k}_{{{\text{6b}}}} {\text{C}}_{{{\text{6}}}} {\text{)]}} \\ & \\ \end{aligned} $$
(4)

where N T is the total number of transporters in the oocyte plasma membrane.

Changes of fluorescence intensity (ΔF) associated with SGLT1 with step jumps in membrane voltage are assumed to be due to changes in occupancy probabilities:

$$ \Delta F \approx {\text{ }}qy_{{\text{1}}} \Delta {\text{C}}_{{\text{1}}} + qy_{{\text{2}}} \Delta {\text{C}}_{{\text{2}}} + qy_{a} \Delta {\text{C}}_{{\text{a}}} + qy_{b} \Delta {\text{C}}_{{\text{b}}} + qy_{3} \Delta {\text{C}}_{{\text{3}}} + qy_{4} \Delta {\text{C}}_{{\text{4}}} + qy_{5} \Delta {\text{C}}_{{\text{5}}} + qy_{6} \Delta {\text{C}}_{{\text{6}}} $$
(5)

where qy j is the apparent quantum yield of the fluorophore (TMR6M) when SGLT1 is in conformation Cj (Loo et al. 2006).

Simulation of SGLT1

In the presence of two different substrates the model has 10 states (Fig. 8), with the differential equation (equation 6):

$$ d/dt\,{\left[ {\begin{array}{*{20}c} {{C1}} \\ {{C2}} \\ {{C3}} \\ {{C4}} \\ {{C5}} \\ {{C6}} \\ {{C7}} \\ {{C8}} \\ {{Ca}} \\ {{Cb}} \\ \end{array} } \right]}\, = \,{\left[ {\begin{array}{*{20}c} {{ - (k1a + k12)}} & {{k21}} & {0} & {0} & {0} & {0} & {0} & {0} & {{ka1}} & {0} \\ {{k12}} & {{ - (k21 + k23 + k25 + k27)}} & {{k32}} & {0} & {{k52}} & {0} & {{k72}} & {0} & {0} & {0} \\ {0} & {{k23}} & {{ - (k32 + k34)}} & {{k43}} & {0} & {0} & {0} & {0} & {0} & {0} \\ {0} & {0} & {{k34}} & {{ - (k43 + k45)}} & {{k54}} & {0} & {0} & {0} & {0} & {0} \\ {0} & {{k25}} & {0} & {{k45}} & {{ - (k52 + k54 + k56 + k58)}} & {{k65}} & {0} & {{k85}} & {0} & {0} \\ {0} & {0} & {0} & {0} & {{k56}} & {{ - (k65 + k6b)}} & {0} & {0} & {0} & {{kb6}} \\ {0} & {{k27}} & {0} & {0} & {0} & {0} & {{ - (k72 + k78)}} & {{k87}} & {0} & {0} \\ {0} & {0} & {0} & {0} & {{k58}} & {0} & {{k78}} & {{ - (k85 + k87)}} & {0} & {0} \\ {{k1a}} & {0} & {0} & {0} & {0} & {0} & {0} & {0} & {{ - (ka1 + kab)}} & {{kba}} \\ {0} & {0} & {0} & {0} & {0} & {{k6b}} & {0} & {0} & {{kab}} & {{ - (kba + kb6)}} \\ \end{array} } \right]}\,{\left[ {\begin{array}{*{20}c} {{C1}} \\ {{C2}} \\ {{C3}} \\ {{C4}} \\ {{C5}} \\ {{C6}} \\ {{C7}} \\ {{C8}} \\ {{Ca}} \\ {{Cb}} \\ \end{array} } \right]} $$
(6)

Ci is the occupancy probability in state i and C1+C2+C3 +C4+C5+C6+C7+C8+Ca+Cb=1. In case of a single substrate, equation 6 reduces to equation 4 of Loo et al. (2006).

Computer simulations were performed using Berkeley Madonna 8.0.1 (Loo et al. 2006). The voltage pulse protocol was simulated by determining the occupancy probabilities at holding potential (V h = −50 mV). At each test voltage (V t ranging between + 50 and −150 mV), the time course of the occupancy probabilities was obtained by numerically integrating equation 6. Cotransporter currents and fluorescence intensity changes (ΔF) were calculated using equations 4 and 5. Steady-state kinetic parameters (I max, K 0.5) were simulated by generating the I–V relations for the sugar-coupled current as functions of [Na+] o and [sugar] o . At each V m, the I vs. [sugar] o and I vs. [Na+] o curves were fitted to equation 1.

For pre-steady-state simulations, the predicted transient cotransporter currents for ON and OFF responses at each voltage (V) were integrated to obtain the charge (Q). QV relations were fitted with the Boltzmann relation (equation 3) to obtain Q max, Q and \( {\text{V}}^{Q}_{{{\text{0}}{\text{.5}}}} \). Model predictions on relaxation time constants were obtained by fitting the simulated pre-steady-state currents to

$$ I{\text{(t) = }}I_{{{\text{fast}}}} {\text{ exp}(-t/}\tau _{{{\text{fast}}}} ) + I_{{{\text{med}}}} {\text{ exp}(-t/}\tau_{\text{med}}) + I_{{{\text{slow}}}} {\text{ exp}(-t/}\tau _{{{\text{slow}}}} ) + I_{{{\text{ss}}}} $$
(7)

Ifastexp(−t/τfast) is the fast submillisecond component and is beyond the resolution of the two-electrode voltage clamp (Loo et al. 2005).

Estimating Kinetic Parameters

In two-electrode voltage-clamp experiments, the rate constants for sugar translocation (C3Na2S1 ⇄ C4Na2S1) and internal ligand binding (C4Na2S1 ⇄ C5Na2 + S1; C5Na2 ⇄ C6 + 2Na+) cannot be uniquely determined. This requires the kinetics of outward currents, using, e.g., excised patches (Eskandari et al. 2005). Here, we used kinetic parameters obtained from experiments where individual partial reactions were isolated; i.e., (1) kinetic parameters for voltage-dependent reactions (Fig. 1, region I) were obtained from fitting the pre-steady-state currents in the absence of sugar (Loo et al. 2005) and (2) kinetic parameters for αMDG on the internal membrane surface (Fig. 1, region III) were obtained from our previous study on reverse sugar transport using the excised giant patch (Eskandari et al. 2005).

Simulations were performed by varying the rate constants (initially obtained by trial and error) until a global fit (by eye) was obtained for the experimentally obtained steady-state and pre-steady-state kinetics (steady-state I–V curves, \( {\text{K}}^{{{\text{Na}}}}_{{0.5}} \), \( {\text{K}}^{{\alpha {\text{MDG}}}}_{{0.5}} \), \( {\text{K}}^{{{\text{indican}}}}_{{0.5}} \), time course of pre-steady-state currents, QV and τV curves) and fluorescence intensity changes (ΔFV). The objective was to obtain a single set of parameters that fit the global data set rather than those required to obtain an optimal fit to one type of experiment. Each rate was varied in turn to determine the sensitivity of the simulations to that value.

The simulations shown in Figs. 24, 6, 7, 9 and 10 were carried out for hSGLT1 and TMR6M-labeled mutant G507C at 20°C using the parameters of Table 1 with [Na+] o  = 100 mm, [Na+] i  = 5 mm, [αMDG] i  = 0 and NT ∼1011–1012 transporters. All simulations were performed at 0–100 mm [Na+]o, 0–10 mm [αMDG] o and 0–10 mm [indican] o .

Model Simulations

The kinetic parameters obtained for indican and αMDG are summarized in Table 1. Simulations with one set of parameters account for the steady-state and pre-steady-state kinetics of hSGLT1 in the presence of αMDG and/or indican. This is seen from the goodness of fit for the steady-state and pre-steady-state kinetics (e.g., Figs. 2D, E, 3B, 4D–F, 6D–F and 7D–F). The sensitivity of the fit was tested by systematically changing each parameter in turn; e.g., k 27 and k 72 can only be varied by threefold (keeping the ratio k 72/k 27 constant) to account for the indican K 0.5 and the effect of indican on the time constants for the pre-steady-state currents. On the other hand, global fits are insensitive to variations in k 85 from 100 to 1,000 s−1: studies of outward currents as a function of internal [indican] are required to place limits on k 85 (k 58 is constrained by microscopic reversibility).

In order to account for the natural variation in kinetics from experiment to experiment, we determined the range in each parameter that is required (Table 1). The experimental parameters that vary include the following: (1) the V 0.5 for charge movement (\( {\text{V}}^{Q}_{{{\text{0}}{\text{.5}}}} \)) ranges from −33 to −70 mV (Loo et al. 2005), and this is accounted for by variation of k 12 of 50,000–140,000 m −2s−1 (Table 1); (2) the K 0.5 for indican is ∼60–100 μm, and this can be accounted for by variations in k 27 of ∼1000,000–300,000 m −1s−1 and in k 72 of ∼12–35 s−1. As previously noted, there is a poor fit of the capacitive currents recorded in the submillisecond range, and this is probably due to our assumption about the simultaneous binding of the two external Na+ ions to hSGLT1 (Loo et al. 2005).

The most striking differences between the kinetics of indican and glucose transport are in the turnover numbers, the competition between αMDG and indican and the effects of the sugars on the pre-steady-state kinetics. The fact that these differences can readily be accommodated may be taken as additional support for our model of Na+/sugar cotransport by hSGLT1 (Fig. 8).

Specifically, the simulations permit the following:

  1. 1.

    Isolation of conformation C 7 Na 2 S2: Namely, undersaturating external Na+ and indican concentrations at large negative membrane potentials, the transporter is predominantly in the C7Na2S2 conformation (Po = 0.8, Fig. 9B) due to the low rate of indican translocation (k78 = 0.5 s−1). This is in sharp contrast to that for αMDG transport, where the transporter is predominantly in the C5Na2 conformation (Po = 0.7, Fig. 9C) due to the higher rate for sugar translocation (k34 = 50 s−1) and low rate of Na+ dissociation (k56 = 5 s−1). The high population of state C7Na2S2 in the presence of indican also accounts for (a) the charge transfer on depolarizing the membrane potential, i.e., the shift in conformation from C7Na2S2 to C6 (C7Na2S2 → C2Na2 → Ca → Cb → C6; Fig. 9B, E), and (b) the increase in time constant for both charge movement and fluorescence changes (Figs. 4 and 9). On the other hand, saturating αMDG eliminates charge movement—depolarizing the membrane potential shifts C5Na2 to C6 (Fig. 9C, F), an electroneutral step (see Loo et al. 2006). Implicit in the model is that the Na+/sugar cotransport step (C3Na2S1 → C4Na2S1 and C7Na2S2 → C8Na2S2) is electroneutral and that the only voltage-sensitive steps in the transport cycle are those involving external Na+ binding (C1 ⇄ C2Na2) and the reorientation of the ligand-free transporter (C1 ⇄ Ca ⇄ Cb ⇄ C6, see Fig. 8).

  2. 2.

    Evidence for a conformational change after sugar binding: The conformational transitions monitored by fluorescence change (ΔF) when Vm is stepped from −50 mV (Vh) to + 50 mV differed in the presence of saturating concentrations of sugars: in saturating αMDG, ΔF is associated with transition from C5Na2 → C6, whereas in saturating indican it arises from C7Na2S2 → C6 (C7Na2S2 → C2Na2 → C1 → Ca → Cb → C6, Fig. 9B, C). The increase in maximal fluorescence change (ΔFmax) in saturating indican compared to NaCl buffer alone indicates that the (apparent) quantum yield of C7Na2S2 is lower than that of C2Na2 (qy7 = 0.7, qy2 = 1). This difference in quantum yields between C2Na2 and C7Na2S2 provides evidence for a conformational change of SGLT1 induced by sugar binding, prior to the sugar translocation step.

  3. 3.

    Substrate competition: This is simply a consequence of the large difference in the transporter turnover number for the two substrates. In the case of SGLT1, where the turnover for indican is only 10% of that for αMDG, the addition of indican causes a reduction in αMDG transport. Simulations demonstrate that this is caused by an increase in the probability of the sugar-binding site being occupied by indican (see Fig. 10) and the concomitant reduction in translocation; k34 is reduced from 50 to 0.5 s−1 (k78). The model accurately predicts the apparent inhibition constant for indican, 210 μm (Fig. 3), and the difference between this Ki and the K0.5 for indican transport in the absence of other substrates (0.1 mm). Simulations also predict the changes in the kinetics of αMDG transport at fixed concentrations of indican; i.e., a fixed concentration of indican produces the expected decrease in apparent affinity (increase in K0.5) of αMDG with no change in maximum velocity, or classical competitive inhibition. Irrespective of the fixed indican concentration, a sufficiently high αMDG concentration can be obtained, in theory at least, to outcompete indican binding.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Loo, D.D.F., Hirayama, B.A., Sala-Rabanal, M. et al. How Drugs Interact with Transporters: SGLT1 as a Model. J Membrane Biol 223, 87–106 (2008). https://doi.org/10.1007/s00232-008-9116-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00232-008-9116-6

Keywords

Navigation