Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

Neurotransmitter and psychostimulant recognition by the dopamine transporter

Abstract

Na+/Cl-coupled biogenic amine transporters are the primary targets of therapeutic and abused drugs, ranging from antidepressants to the psychostimulants cocaine and amphetamines, and to their cognate substrates. Here we determine X-ray crystal structures of the Drosophila melanogaster dopamine transporter (dDAT) bound to its substrate dopamine, a substrate analogue 3,4-dichlorophenethylamine, the psychostimulants d-amphetamine and methamphetamine, or to cocaine and cocaine analogues. All ligands bind to the central binding site, located approximately halfway across the membrane bilayer, in close proximity to bound sodium and chloride ions. The central binding site recognizes three chemically distinct classes of ligands via conformational changes that accommodate varying sizes and shapes, thus illustrating molecular principles that distinguish substrates from inhibitors in biogenic amine transporters.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Dopamine occupies central binding site.
Figure 2: Amphetamines bind to central site.
Figure 3: DCP induces partial occlusion of central site.
Figure 4: Multivalent binding of cocaine.
Figure 5: Plasticity confers versatile recognition.

Similar content being viewed by others

Accession codes

Accessions

Protein Data Bank

Data deposits

The coordinates for the structure have been deposited in the Protein Data Bank under the accession codes 4XP1, 4XP9, 4XP6, 4XPA, 4XP4, 4XP5, 4XPB, 4XPF, 4XPG, 4XPH, 4XPT (see Supplementary Table 1 for details).

References

  1. Carlsson, A., Lindqvist, M. & Magnusson, T. 3,4-Dihydroxyphenylalanine and 5-hydroxytryptophan as reserpine antagonists. Nature 180, 1200 (1957)

    Article  ADS  CAS  PubMed  Google Scholar 

  2. Twarog, B. M. & Page, I. H. Serotonin content of some mammalian tissues and urine and a method for its determination. Am. J. Physiol. 175, 157–161 (1953)

    Article  CAS  PubMed  Google Scholar 

  3. von Euler, U. Sympathin in adrenergic nerve fibres. J. Physiol. (Lond.) 105, 26 (1946)

    CAS  Google Scholar 

  4. Hertting, G. & Axelrod, J. Fate of tritiated noradrenaline at the sympathetic nerve-endings. Nature 192, 172–173 (1961)

    Article  ADS  CAS  PubMed  Google Scholar 

  5. Iversen, L. L. Role of transmitter uptake mechanisms in synaptic neurotransmission. Br. J. Pharmacol. 41, 571–591 (1971)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Iversen, L. L. & Kravitz, E. A. Sodium dependence of transmitter uptake at adrenergic nerve terminals. Mol. Pharmacol. 2, 360–362 (1966)

    CAS  PubMed  Google Scholar 

  7. Kristensen, A. S. et al. SLC6 neurotransmitter transporters: structure, function, and regulation. Pharmacol. Rev. 63, 585–640 (2011)

    Article  CAS  PubMed  Google Scholar 

  8. Pramod, A. B., Foster, J., Carvelli, L. & Henry, L. K. SLC6 transporters: structure, function, regulation, disease association and therapeutics. Mol. Aspects Med. 34, 197–219 (2013)

    Article  CAS  PubMed  Google Scholar 

  9. Penmatsa, A., Wang, K. H. & Gouaux, E. X-ray structure of dopamine transporter elucidates antidepressant mechanism. Nature 503, 85–90 (2013)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  10. Koldsø, H., Christiansen, A. B., Sinning, S. & Schiøtt, B. Comparative modeling of the human monoamine transporters: similarities in substrate binding. ACS Chem Neurosci 4, 295–309 (2013)

    Article  PubMed  Google Scholar 

  11. Yamashita, A., Singh, S. K., Kawate, T., Jin, Y. & Gouaux, E. Crystal structure of a bacterial homologue of Na+/Cl-dependent neurotransmitter transporters. Nature 437, 215–223 (2005)

    Article  ADS  CAS  PubMed  Google Scholar 

  12. Beuming, T., Shi, L., Javitch, J. A. & Weinstein, H. A comprehensive structure-based alignment of prokaryotic and eukaryotic neurotransmitter/Na+ symporters (NSS) aids in the use of the LeuT structure to probe NSS structure and function. Mol. Pharmacol. 70, 1630–1642 (2006)

    Article  CAS  PubMed  Google Scholar 

  13. Gu, H., Wall, S. C. & Rudnick, G. Stable expression of biogenic amine transporters reveals differences in inhibitor sensitivity, kinetics, and ion dependence. J. Biol. Chem. 269, 7124–7130 (1994)

    CAS  PubMed  Google Scholar 

  14. Kilty, J. E., Lorang, D. & Amara, S. G. Cloning and expression of a cocaine-sensitive rat dopamine transporter. Science 254, 578–579 (1991)

    Article  ADS  CAS  PubMed  Google Scholar 

  15. Kurian, M. A. et al. Homozygous loss-of-function mutations in the gene encoding the dopamine transporter are associated with infantile parkinsonism-dystonia. J. Clin. Invest. 119, 1595–1603 (2009)

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Giros, B., Jaber, M., Jones, S. R., Wightman, R. M. & Caron, M. G. Hyperlocomotion and indifference to cocaine and amphetamine in mice lacking the dopamine transporter. Nature 379, 606–612 (1996)

    Article  ADS  CAS  PubMed  Google Scholar 

  17. Mergy, M. A. et al. The rare DAT coding variant Val559 perturbs DA neuron function, changes behavior, and alters in vivo responses to psychostimulants. Proc. Natl Acad. Sci. USA 111, E4779–E4788 (2014)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Zaczek, R., Culp, S., Goldberg, H., McCann, D. J. & De Souza, E. B. Interactions of [3H]amphetamine with rat brain synaptosomes. I. Saturable sequestration. J. Pharmacol. Exp. Ther. 257, 820–829 (1991)

    CAS  PubMed  Google Scholar 

  19. Bönisch, H. The transport of (+)-amphetamine by the neuronal noradrenaline carrier. Naunyn Schmiedebergs Arch. Pharmacol. 327, 267–272 (1984)

    Article  PubMed  Google Scholar 

  20. Eshleman, A. J., Henningsen, R. A., Neve, K. A. & Janowsky, A. Release of dopamine via the human transporter. Mol. Pharmacol. 45, 312–316 (1994)

    CAS  PubMed  Google Scholar 

  21. Goodwin, J. S. et al. Amphetamine and methamphetamine differentially affect dopamine transporters in vitro and in vivo. J. Biol. Chem. 284, 2978–2989 (2009)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Sitte, H. H. et al. Carrier-mediated release, transport rates, and charge transfer induced by amphetamine, tyramine, and dopamine in mammalian cells transfected with the human dopamine transporter. J. Neurochem. 71, 1289–1297 (1998)

    Article  CAS  PubMed  Google Scholar 

  23. Sonders, M. S., Zhu, S. J., Zahniser, N. R., Kavanaugh, M. P. & Amara, S. G. Multiple ionic conductances of the human dopamine transporter: the actions of dopamine and psychostimulants. J. Neurosci. 17, 960–974 (1997)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Wall, S. C., Gu, H. & Rudnick, G. Biogenic amine flux mediated by cloned transporters stably expressed in cultured cell lines: amphetamine specificity for inhibition and efflux. Mol. Pharmacol. 47, 544–550 (1995)

    CAS  PubMed  Google Scholar 

  25. Beuming, T. et al. The binding sites for cocaine and dopamine in the dopamine transporter overlap. Nature Neurosci. 11, 780–789 (2008)

    Article  CAS  PubMed  Google Scholar 

  26. Bisgaard, H. et al. The binding sites for benztropines and dopamine in the dopamine transporter overlap. Neuropharmacology 60, 182–190 (2011)

    Article  MathSciNet  CAS  PubMed  Google Scholar 

  27. Sørensen, G. et al. Neuropeptide Y Y5 receptor antagonism attenuates cocaine-induced effects in mice. Psychopharmacology (Berl.) 222, 565–577 (2012)

    Article  Google Scholar 

  28. Richelson, E. & Pfenning, M. Blockade by antidepressants and related compounds of biogenic amine uptake into rat brain synaptosomes: most antidepressants selectively block norepinephrine uptake. Eur. J. Pharmacol. 104, 277–286 (1984)

    Article  CAS  PubMed  Google Scholar 

  29. Sørensen, L. et al. Interaction of antidepressants with the serotonin and norepinephrine transporters: mutational studies of the S1 substrate binding pocket. J. Biol. Chem. 287, 43694–43707 (2012)

    Article  PubMed  PubMed Central  Google Scholar 

  30. Dahal, R. A. et al. Computational and biochemical docking of the irreversible cocaine analog RTI 82 directly demonstrates ligand positioning in the dopamine transporter central substrate binding site. J. Biol. Chem. 289, 29712–29727 (2014)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Pörzgen, P., Park, S. K., Hirsh, J., Sonders, M. S. & Amara, S. G. The antidepressant-sensitive dopamine transporter in Drosophila melanogaster: a primordial carrier for catecholamines. Mol. Pharmacol. 59, 83–95 (2001)

    Article  PubMed  Google Scholar 

  32. Abdul-Hussein, S., Andrell, J. & Tate, C. G. Thermostabilisation of the serotonin transporter in a cocaine-bound conformation. J. Mol. Biol. 425, 2198–2207 (2013)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Wang, H. et al. Structural basis for action by diverse antidepressants on biogenic amine transporters. Nature 503, 141–145 (2013)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  34. Huang, X. & Zhan, C. G. How dopamine transporter interacts with dopamine: insights from molecular modeling and simulation. Biophys. J. 93, 3627–3639 (2007)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  35. Kitayama, S. et al. Dopamine transporter site-directed mutations differentially alter substrate transport and cocaine binding. Proc. Natl Acad. Sci. USA 89, 7782–7785 (1992)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  36. Han, D. D. & Gu, H. H. Comparison of the monoamine transporters from human and mouse in their sensitivities to psychostimulant drugs. BMC Pharmacol. 6, 6 (2006)

    Article  PubMed  PubMed Central  Google Scholar 

  37. Newman, A. H., Zou, M. F., Ferrer, J. V. & Javitch, J. A. [3H]MFZ 2–12: a novel radioligand for the dopamine transporter. Bioorg. Med. Chem. Lett. 11, 1659–1661 (2001)

    Article  CAS  PubMed  Google Scholar 

  38. Kazmier, K. et al. Conformational dynamics of ligand-dependent alternating access in LeuT. Nature Struct. Mol. Biol. 21, 472–479 (2014)

    Article  CAS  Google Scholar 

  39. Krishnamurthy, H. & Gouaux, E. X-ray structures of LeuT in substrate-free outward-open and apo inward-open states. Nature 481, 469–474 (2012)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  40. Hong, W. C. & Amara, S. G. Membrane cholesterol modulates the outward facing conformation of the dopamine transporter and alters cocaine binding. J. Biol. Chem. 285, 32616–32626 (2010)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Zhao, Y. et al. Substrate-modulated gating dynamics in a Na+-coupled neurotransmitter transporter homologue. Nature 474, 109–113 (2011)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Malinauskaite, L. et al. A mechanism for intracellular release of Na+ by neurotransmitter/sodium symporters. Nature Struct. Mol. Biol. 21, 1006–1012 (2014)

    Article  CAS  Google Scholar 

  43. Perez, C., Koshy, C., Yildiz, O. & Ziegler, C. Alternating-access mechanism in conformationally asymmetric trimers of the betaine transporter BetP. Nature 490, 126–130 (2012)

    Article  ADS  CAS  PubMed  Google Scholar 

  44. Weyand, S. et al. Structure and molecular mechanism of a nucleobase-cation-symport-1 family transporter. Science 322, 709–713 (2008)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  45. Loland, C. J. et al. Relationship between conformational changes in the dopamine transporter and cocaine-like subjective effects of uptake inhibitors. Mol. Pharmacol. 73, 813–823 (2008)

    Article  CAS  PubMed  Google Scholar 

  46. Lee, S. H. et al. Importance of valine at position 152 for the substrate transport and 2β-carbomethoxy-3β-(4-fluorophenyl)tropane binding of dopamine transporter. Mol. Pharmacol. 57, 883–889 (2000)

    CAS  PubMed  Google Scholar 

  47. Chen, J. G., Sachpatzidis, A. & Rudnick, G. The third transmembrane domain of the serotonin transporter contains residues associated with substrate and cocaine binding. J. Biol. Chem. 272, 28321–28327 (1997)

    Article  CAS  PubMed  Google Scholar 

  48. Singh, S. K., Piscitelli, C. L., Yamashita, A. & Gouaux, E. A competitive inhibitor traps LeuT in an open-to-out conformation. Science 322, 1655–1661 (2008)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  49. Claxton, D. P. et al. Ion/substrate-dependent conformational dynamics of a bacterial homolog of neurotransmitter:sodium symporters. Nature Struct. Mol. Biol. 17, 822–829 (2010)

    Article  CAS  Google Scholar 

  50. Schmitt, K. C., Rothman, R. B. & Reith, M. E. Nonclassical pharmacology of the dopamine transporter: atypical inhibitors, allosteric modulators, and partial substrates. J. Pharmacol. Exp. Ther. 346, 2–10 (2013)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Reeves, P. J., Callewaert, N., Contreras, R. & Khorana, H. G. Structure and function in rhodopsin: high-level expression of rhodopsin with restricted and homogeneous N-glycosylation by a tetracycline-inducible N-acetylglucosaminyltransferase I-negative HEK293S stable mammalian cell line. Proc. Natl Acad. Sci. USA 99, 13419–13424 (2002)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  52. Goehring, A. et al. Screening and large-scale expression of membrane proteins in mammalian cells for structural studies. Nature Protocols 9, 2574–2585 (2014)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Otwinowski, Z. &. Minor, W. Processing of X-ray diffraction data collected in oscillation mode. Methods Enzymol. 276, 307–326 (1997)

    Article  CAS  PubMed  Google Scholar 

  54. Kabsch, W. XDS. Acta Crystallogr. D 66, 125–132 (2010)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. McCoy, A. J. et al. Phaser crystallographic software. J. Appl. Crystallogr. 40, 658–674 (2007)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Afonine, P. V. et al. Towards automated crystallographic structure refinement with phenix.refine. Acta Crystallogr. D 68, 352–367 (2012)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Emsley, P. & Cowtan, K. Coot: model-building tools for molecular graphics. Acta Crystallogr. D 60, 2126–2132 (2004)

    Article  PubMed  Google Scholar 

  58. Quick, M. & Javitch, J. A. Monitoring the function of membrane transport proteins in detergent-solubilized form. Proc. Natl Acad. Sci. USA 104, 3603–3608 (2007)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  59. Eshleman, A. J. et al. Metabolism of catecholamines by catechol-O-methyltransferase in cells expressing recombinant catecholamine transporters. J. Neurochem. 69, 1459–1466 (1997)

    Article  CAS  PubMed  Google Scholar 

  60. Dehnes, Y. et al. Conformational changes in dopamine transporter intracellular regions upon cocaine binding and dopamine translocation. Neurochem. Int. 73, 4–15 (2014)

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank J. Coleman, C.-H. Lee, S. Mansoor and other Gouaux laboratory members for helpful discussions, L. Vaskalis for assistance with figures and H. Owen for help with manuscript preparation. We acknowledge the staff of the Berkeley Center for Structural Biology at the Advanced Light Source and the Northeastern Collaborative Access Team at the Advanced Photon Source for assistance with data collection. This work was supported by an NIMH Ruth Kirschstein postdoctoral fellowship and Brain and Behavior Research Foundation Young Investigator research award (K.H.W.), a postdoctoral fellowship from the American Heart Association (A.P.) and by the NIH (E.G) and the Methamphetamine Abuse Research Center of OHSU (P50DA018165 to E.G). E.G. is an investigator with the Howard Hughes Medical Institute.

Author information

Authors and Affiliations

Authors

Contributions

A.P., K.H.W. and E.G. designed the project. A.P. and K.H.W. performed protein purification, crystallography and biochemical assays. A.P., K.H.W. and E.G. wrote the manuscript.

Corresponding author

Correspondence to Eric Gouaux.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Extended data figures and tables

Extended Data Figure 1 Design of the minimal functional construct.

a, Thermostabilizing (ts) mutations V275A, V311A, G538L were removed. Modification of the EL2 deletion from 164–206 to 162–202, which recovered transport activity. The del 162–201 construct has robust dopamine uptake activity. b, Structural organization of EL2 regions. Organization of dDATcryst with a deletion of region 164–206 depicted as green surface. c, EL2 structure in dDATmfc with the deletion 162–202 depicted as cyan surface showing contacts between EL2 and EL6. d, EL2 organization in the construct with a deletion from 162–201 depicted as magenta surface. e, Fab 9D5 interferes with the interaction between EL2 and EL6 in the crystal lattice, with loops depicted as magenta and cyan surfaces, respectively. Fab disrupts the EL organization in all structures. The del 162–201 subB structure is shown.

Extended Data Figure 2 Measurement of dissociation constants using purified dDATmfc protein, and dopamine uptake in whole cells.

a, dDATmfc binds [3H]nisoxetine with a Kd of 36 ± 3 nM (s.e.m.). b, dDATmfc with subB mutations binds [3H]nisoxetine with a Kd of 10 ± 1 nM (s.e.m.). c, d, Michaelis–Menten plots of [14C]dopamine uptake by HEK293S cells expressing dDATwt or dDATmfc, respectively, which yielded a KM of 2.1 ± 0.7 μM and Vmax of 4.5 ± 0.4 pmol min−1 per 106 cells for dDATwt and a KM of 8.2 ± 2.3 μM and Vmax of 2.4 ± 0.2 pmol min−1 per 106 cells for dDATmfc (s.e.m.). One representative plot of total and background counts (in the presence of 10 μM nortriptyline) is shown of two experimental trials as squares and triangles, respectively. Data points and error bars show the average and standard deviation, respectively, of technical replicates (n = 3). Welch’s t-test indicates that the specific uptake signal at each concentration of dopamine is significant with a two-tailed P value < 0.02. e, Eadie–Hofstee plot of specific dopamine uptake shown in Fig. 1a and panels c and d of this figure. Data for dDATwt and dDATmfc are shown as squares and triangles, respectively, and error bars denote s.d. of technical replicates (n = 3). f, The thermal melting curve of dDATmfc solubilized from HEK293S membranes in the presence of 100 nM [3H]nisoxetine exhibits a melting temperature of 48 ± 2 °C (s.e.m.). The fraction bound describes the signal remaining after incubation at the specified temperature for 10 min, normalized to the signal at 4 °C. Data points show the mean values for one experimental trial, and error bars show the s.d. of technical replicates (n = 3).

Extended Data Figure 3 Mutagenesis and effects of DAT subsite B.

a, Sequence alignment of subsite B regions for dDAT and human NSS orthologues. b, 2Fo − Fc density contoured at 0.9σ around the vicinity of the D121G (TM3) and S426M mutations (TM8). c, Abrogation of dopamine transport activity by dDATwt bearing both subsite B mutations in infected HEK293S cells. Data show the average uptake and error bars show the data range of technical duplicates for a single trial. Reactions were performed without and with 100 μM desipramine in black and grey bars, respectively.

Extended Data Figure 4 Measurement of inhibition constants using purified dDAT protein.

a–c, Inhibition of [3H]nisoxetine binding to dDATmfc (squares) and dDATmfc subB (triangles). Ki inhibition constants for dDATmfc and dDATmfc subB are, respectively, 98 ± 4 μM, and 1.4 ± 0.1 μM, (a, β-CFT), 371 ± 25 nM and 271 ± 59 nM (b, RTI-55), and 4.5 ± 0.3 μM, and 267 ± 20 nM (c, DCP). All errors are s.e.m. One representative trial of two is shown for all experiments in panels ac, and data points and error bars denote the average values for fraction bound and standard deviation, respectively, for technical replicates (n = 3). d, Inhibition of [3H]nisoxetine (50 nM) binding to dDATdel by 1 and 10 μM unlabelled compound (grey and black bars, respectively). Error bars show the data range of technical replicates (n = 2). Abbreviations: 3-BrPE, 3-bromophenethylamine; 4-BrPE, 4-bromophenethylamine; 2-pTE, 2-(pTolyl)ethylamine, 4-ClPE, 4-chlorophenethylamine; DCP, 3,4-dichlorophenethylamine.

Extended Data Figure 5 Fo − Fc densities for ligands complexed with dDAT.

a, d-amphetamine (2.4σ); b, (+)-methamphetamine (1.8σ); c, DCP (2.2σ); d, cocaine (2.2σ); e, β-CFT (2.2σ); f, RTI-55 (2.6σ).

Extended Data Figure 6 Helical movements in dDATmfc upon binding to substrate analogue DCP (orange) and inhibitor nortriptyline (grey).

ad, Helices undergoing maximal shifts are a, TM1b; b, TM2; c, TM6a; d, TM11. Arrows in black represent direction of shift. e, Table comparing angular shifts between nortriptyline–dDATcryst (PDB ID 4M48) and DCP–dDATmfc structures in column one, and between the outward-open Trp–LeuT (PDB ID 3F3A) and outward-occluded Leu–LeuT (PDB ID 2A65) structures in column two. f, Superposition of the outward open state of nortriptyline–dDATcryst (PDB ID 4M48) and DCP–dDATmfc structures in grey and orange ribbon, respectively. Extracellular gating TMs 1b and 6a are shown as cylinders. Arrows in red indicate inward movement of TMs 1b and 6a. g, Superposition of the occluded state of LeuT (PDB ID 2A65) and DCP–dDATmfc structures in grey and orange ribbon, respectively.

Extended Data Figure 7 Cholesterol binding sites in DAT.

a, Cholesterol binding sites seen on the dDAT surface corresponding to the inner leaflet of the plasma membrane, with a second novel cholesterol site into which a cholesteryl hemisuccinate (CHS) could be modelled. Fo − Fc densities for cholesterol contoured at 2.0σ. b, Close-up view of cholesterol site II at the junction of TM2, TM7 and TM11 interacting with multiple hydrophobic residues. Asterisk denotes thermostabilizing mutant V74A. c, Effect of CHS concentration on [3H]nisoxetine binding to DATmfc construct. Graph depicts one representative trial of two independent experiments, and total and background counts were measured using technical replicates (n = 3) for each binding curve at each CHS concentration. Arrow represents increasing concentration of CHS. Error bars represent s.d.

Extended Data Figure 8 Analogues of cocaine and binding site comparisons.

a, The position of RTI-55 in the binding pocket with anomalous difference density for iodide displayed as purple mesh and contoured at 4σ. b, Superposition of cocaine, β-CFT, and RTI-55 using the RTI-55-dDATmfc structure. Ligands are shown as sticks and coloured yellow (cocaine), pink (β-CFT), and teal (RTI-55). Sodium ions are shown as purple spheres. c–f, Residues that line the binding pocket are superposed between the nortriptyline–dDATcryst (magenta, PDB ID 4M48) and those of c, DA–dDATmfc (cyan), d, DCP–dDATmfc (marine), e, cocaine–dDATmfc (yellow). f, Organization of S1 binding site in complex with nortriptyline (PDB ID 4M48). Black arrows describe the change in rotamers and positions of D46, F319, and F325 compared to the nortriptyline-bound structure.

Extended Data Table 1 Superposition statistics of dDAT structures
Extended Data Table 2 a, Ligand surface and interface areas*; b, crystallization conditions for ligand–DAT complexes

Supplementary information

Supplementary Tables

This file contains Supplementary Table 1. (PDF 209 kb)

Ligand-induced conformational rearrangements illustrate structural plasticity

Shown is a morph between selected structural states of the Drosophila dopamine transporter beginning with the previously published nortriptyline-bound outward-open state (magenta, PDB. id 4M48) followed by structural changes associated with the binding of cocaine (yellow, PDB. ids 4XP4, 4XPB) wherein Phe 325 and the TM6a-6b linker move into the binding site to interact with the benzyloxy aromatic group of cocaine. Transition from the cocaine bound state is followed by the D-amphetamine-bound (light orange, PDB. ids 4XP9) state in which additional ‘contraction’ of the TM6a-6b linker retains aromatic π-π interactions with Phe 325. The dopamine-bound conformation (cyan, PDB. id 4XP1) follows the D-amphetamine-bound state, illustrating the rotameric shift in the side chain of Asp 46, allowing it to interact with the primary amine of the neurotransmitter. Upon binding of the dopamine analogue DCP (teal, PDB. ids 4XPA, 4XPH) there are shifts in TM helices 1b (deep red) and 6a (green) and a rotameric reorientation of Phe 319 that blocks the substrate binding site from solvent access. Scaffold helices TMs 3 and 8 are colored in orange and cyan respectively. (MPG 22010 kb)

PowerPoint slides

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wang, K., Penmatsa, A. & Gouaux, E. Neurotransmitter and psychostimulant recognition by the dopamine transporter. Nature 521, 322–327 (2015). https://doi.org/10.1038/nature14431

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nature14431

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing