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:

Crystal structure of a SLC11 (NRAMP) transporter reveals the basis for transition-metal ion transport

Abstract

Members of the SLC11 (NRAMP) family transport iron and other transition-metal ions across cellular membranes. These membrane proteins are present in all kingdoms of life with a high degree of sequence conservation. To gain insight into the determinants of ion selectivity, we have determined the crystal structure of Staphylococcus capitis DMT (ScaDMT), a close prokaryotic homolog of the family. ScaDMT shows a familiar architecture that was previously identified in the amino acid permease LeuT. The protein adopts an inward-facing conformation with a substrate-binding site located in the center of the transporter. This site is composed of conserved residues, which coordinate Mn2+, Fe2+ and Cd2+ but not Ca2+. Mutations of interacting residues affect ion binding and transport in both ScaDMT and human DMT1. Our study thus reveals a conserved mechanism for transition-metal ion selectivity within the SLC11 family.

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: Transition-metal ion binding and transport.
Figure 2: ScaDMT structure.
Figure 3: Ion coordination.
Figure 4: Transition-metal ion selectivity.
Figure 5: Cd2+ binding to ScaDMT binding-site mutants.
Figure 6: Functional properties of DMT1 binding-site mutants.
Figure 7: Transport mechanism.

Similar content being viewed by others

Accession codes

Primary accessions

Protein Data Bank

References

  1. Nevo, Y. & Nelson, N. The NRAMP family of metal-ion transporters. Biochim. Biophys. Acta 1763, 609–620 (2006).

    Article  CAS  Google Scholar 

  2. Montalbetti, N., Simonin, A., Kovacs, G. & Hediger, M.A. Mammalian iron transporters: families SLC11 and SLC40. Mol. Aspects Med. 34, 270–287 (2013).

    Article  CAS  Google Scholar 

  3. Cellier, M.F., Bergevin, I., Boyer, E. & Richer, E. Polyphyletic origins of bacterial Nramp transporters. Trends Genet. 17, 365–370 (2001).

    Article  CAS  Google Scholar 

  4. Vidal, S.M., Malo, D., Vogan, K., Skamene, E. & Gros, P. Natural resistance to infection with intracellular parasites: isolation of a candidate for Bcg. Cell 73, 469–485 (1993).

    Article  CAS  Google Scholar 

  5. Illing, A.C., Shawki, A., Cunningham, C.L. & Mackenzie, B. Substrate profile and metal-ion selectivity of human divalent metal-ion transporter-1. J. Biol. Chem. 287, 30485–30496 (2012).

    Article  CAS  Google Scholar 

  6. Vidal, S.M., Pinner, E., Lepage, P., Gauthier, S. & Gros, P. Natural resistance to intracellular infections: Nramp1 encodes a membrane phosphoglycoprotein absent in macrophages from susceptible (Nramp1 D169) mouse strains. J. Immunol. 157, 3559–3568 (1996).

    CAS  PubMed  Google Scholar 

  7. Gunshin, H. et al. Cloning and characterization of a mammalian proton-coupled metal-ion transporter. Nature 388, 482–488 (1997).

    Article  CAS  Google Scholar 

  8. Beaumont, C. et al. Two new human DMT1 gene mutations in a patient with microcytic anemia, low ferritinemia, and liver iron overload. Blood 107, 4168–4170 (2006).

    Article  CAS  Google Scholar 

  9. Johnson, E.E. & Wessling-Resnick, M. Iron metabolism and the innate immune response to infection. Microbes Infect. 14, 207–216 (2012).

    Article  CAS  Google Scholar 

  10. Shawki, A., Knight, P.B., Maliken, B.D., Niespodzany, E.J. & Mackenzie, B.H. H+-coupled divalent metal-ion transporter-1: functional properties, physiological roles and therapeutics. Curr. Top. Membr. 70, 169–214 (2012).

    Article  CAS  Google Scholar 

  11. Mackenzie, B., Ujwal, M.L., Chang, M.H., Romero, M.F. & Hediger, M.A. Divalent metal-ion transporter DMT1 mediates both H+ -coupled Fe2+ transport and uncoupled fluxes. Pflugers Arch. 451, 544–558 (2006).

    Article  CAS  Google Scholar 

  12. Shawki, A. & Mackenzie, B. Interaction of calcium with the human divalent metal-ion transporter-1. Biochem. Biophys. Res. Commun. 393, 471–475 (2010).

    Article  CAS  Google Scholar 

  13. Au, C., Benedetto, A. & Aschner, M. Manganese transport in eukaryotes: the role of DMT1. Neurotoxicology 29, 569–576 (2008).

    Article  CAS  Google Scholar 

  14. Bressler, J.P., Olivi, L., Cheong, J.H., Kim, Y. & Bannona, D. Divalent metal transporter 1 in lead and cadmium transport. Ann. NY Acad. Sci. 1012, 142–152 (2004).

    Article  CAS  Google Scholar 

  15. Guerinot, M.L. Microbial iron transport. Annu. Rev. Microbiol. 48, 743–772 (1994).

    Article  CAS  Google Scholar 

  16. Makui, H. et al. Identification of the Escherichia coli K-12 Nramp orthologue (MntH) as a selective divalent metal ion transporter. Mol. Microbiol. 35, 1065–1078 (2000).

    Article  CAS  Google Scholar 

  17. Czachorowski, M., Lam-Yuk-Tseung, S., Cellier, M. & Gros, P. Transmembrane topology of the mammalian Slc11a2 iron transporter. Biochemistry 48, 8422–8434 (2009).

    Article  CAS  Google Scholar 

  18. 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  CAS  Google Scholar 

  19. Cellier, M.F. Nramp: from sequence to structure and mechanism of divalent metal import. Curr. Top. Membr. 69, 249–293 (2012).

    Article  CAS  Google Scholar 

  20. Cellier, M.F. Nutritional immunity: homology modeling of Nramp metal import. Adv. Exp. Med. Biol. 946, 335–351 (2012).

    Article  CAS  Google Scholar 

  21. Geertsma, E.R. & Dutzler, R. A versatile and efficient high-throughput cloning tool for structural biology. Biochemistry 50, 3272–3278 (2011).

    Article  CAS  Google Scholar 

  22. Pardon, E. et al. A general protocol for the generation of Nanobodies for structural biology. Nat. Protoc. 9, 674–693 (2014).

    Article  CAS  Google Scholar 

  23. Schulze, S., Koster, S., Geldmacher, U., Terwisscha van Scheltinga, A.C. & Kuhlbrandt, W. Structural basis of Na+-independent and cooperative substrate/product antiport in CaiT. Nature 467, 233–236 (2010).

    Article  CAS  Google Scholar 

  24. Ressl, S., Terwisscha van Scheltinga, A.C., Vonrhein, C., Ott, V. & Ziegler, C. Molecular basis of transport and regulation in the Na+/betaine symporter BetP. Nature 458, 47–52 (2009).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  26. Faham, S. et al. The crystal structure of a sodium galactose transporter reveals mechanistic insights into Na+/sugar symport. Science 321, 810–814 (2008).

    Article  CAS  Google Scholar 

  27. Gao, X. et al. Structure and mechanism of an amino acid antiporter. Science 324, 1565–1568 (2009).

    Article  CAS  Google Scholar 

  28. Fang, Y. et al. Structure of a prokaryotic virtual proton pump at 3.2 Å resolution. Nature 460, 1040–1043 (2009).

    Article  CAS  Google Scholar 

  29. 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  CAS  Google Scholar 

  30. Lam-Yuk-Tseung, S., Govoni, G., Forbes, J. & Gros, P. Iron transport by Nramp2/DMT1: pH regulation of transport by 2 histidines in transmembrane domain 6. Blood 101, 3699–3707 (2003).

    Article  CAS  Google Scholar 

  31. Edward, R.A., Whittaker, M.M., Whittaker, J.W., Jameson, G.B. & Baker, E.N. Distinct metal environment in Fe-substituted manganese superoxide dismutatse provides a structural basis of metal specificity. J. Am. Chem. Soc. 120, 9684–9685 (1998).

    Article  CAS  Google Scholar 

  32. Qi, W. & Cowan, J.A. Structural, mechanistic and coordination chemistry of relevance to the biosynthesis of iron-sulfur and related iron cofactors. Coord. Chem. Rev. 255, 688–699 (2011).

    Article  CAS  Google Scholar 

  33. Freisinger, E. & Vasak, M. Cadmium in metallothioneins. Met. Ions. Life Sci. 11, 339–371 (2013).

    Article  CAS  Google Scholar 

  34. Hoch, E. et al. Histidine pairing at the metal transport site of mammalian ZnT transporters controls Zn2+ over Cd2+ selectivity. Proc. Natl. Acad. Sci. USA 109, 7202–7207 (2012).

    Article  CAS  Google Scholar 

  35. Cellier, M. et al. Nramp defines a family of membrane proteins. Proc. Natl. Acad. Sci. USA 92, 10089–10093 (1995).

    Article  CAS  Google Scholar 

  36. Courville, P. et al. Solute carrier 11 cation symport requires distinct residues in transmembrane helices 1 and 6. J. Biol. Chem. 283, 9651–9658 (2008).

    Article  CAS  Google Scholar 

  37. Haemig, H.A. & Brooker, R.J. Importance of conserved acidic residues in mntH, the Nramp homolog of Escherichia coli. J. Membr. Biol. 201, 97–107 (2004).

    Article  CAS  Google Scholar 

  38. Haemig, H.A., Moen, P.J. & Brooker, R.J. Evidence that highly conserved residues of transmembrane segment 6 of Escherichia coli MntH are important for transport activity. Biochemistry 49, 4662–4671 (2010).

    Article  CAS  Google Scholar 

  39. Xu, H., Jin, J., DeFelice, L.J., Andrews, N.C. & Clapham, D.E. A spontaneous, recurrent mutation in divalent metal transporter-1 exposes a calcium entry pathway. PLoS Biol. 2, E50 (2004).

    Article  Google Scholar 

  40. Iolascon, A. & De Falco, L. Mutations in the gene encoding DMT1: clinical presentation and treatment. Semin. Hematol. 46, 358–370 (2009).

    Article  CAS  Google Scholar 

  41. Courville, P., Chaloupka, R. & Cellier, M.F. Recent progress in structure-function analyses of Nramp proton-dependent metal-ion transporters. Biochem. Cell Biol. 84, 960–978 (2006).

    Article  CAS  Google Scholar 

  42. Forrest, L.R. et al. Mechanism for alternating access in neurotransmitter transporters. Proc. Natl. Acad. Sci. USA 105, 10338–10343 (2008).

    Article  CAS  Google Scholar 

  43. Casadaban, M.J. & Cohen, S.N. Analysis of gene control signals by DNA fusion and cloning in Escherichia coli. J. Mol. Biol. 138, 179–207 (1980).

    Article  CAS  Google Scholar 

  44. Kawate, T. & Gouaux, E. Fluorescence-detection size-exclusion chromatography for precrystallization screening of integral membrane proteins. Structure 14, 673–681 2006).

    Article  CAS  Google Scholar 

  45. Van Duyne, G.D., Standaert, R.F., Karplus, P.A., Schreiber, S.L. & Clardy, J. Atomic structures of the human immunophilin FKBP-12 complexes with FK506 and rapamycin. J. Mol. Biol. 229, 105–124 (1993).

    Article  CAS  Google Scholar 

  46. Kabsch, W. Automatic processing of rotation diffraction data from crystals of initially unknown symmetry and cell constants. J. Appl. Crystallogr. 26, 795–800 (1993).

    Article  CAS  Google Scholar 

  47. Collaborative Computational Project, Number 4. The CCP4 Suite: programs for X-ray crystallography. Acta Crystallogr. D Biol. Crystallogr. 50, 760–763 (1994).

  48. Schneider, T.R. & Sheldrick, G.M. Substructure solution with SHELXD. Acta Crystallogr. D Biol. Crystallogr. 58, 1772–1779 (2002).

    Article  Google Scholar 

  49. Pape, T. & Schneider, T.R. HKL2MAP: a graphical user interface for phasing with SHELX programs. J. Appl. Crystallogr. 37, 843–844 (2004).

    Article  CAS  Google Scholar 

  50. De La Fortelle, E. & Bricogne, G. Maximum-likelihood heavy-atom parameter refinement for multiple isomorphous replacement and multiwavelength anomalous diffraction methods. Methods Enzymol. 276, 472–494 (1997).

    Article  CAS  Google Scholar 

  51. Cowtan, K. dm: an automated procedure for phase improvement by density modification. Joint CCP4 and ESF-EACBM Newslett. Protein Crystallogr. 31, 34–38 (1994).

    Google Scholar 

  52. Jones, T.A., Zou, J.Y., Cowan, S.W. & Kjeldgaard, M. Improved methods for building protein models in electron density maps and the location of errors in these models. Acta Crystallogr. A 47, 110–119 (1991).

    Article  Google Scholar 

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

    Article  Google Scholar 

  54. Brünger, A.T. et al. Crystallography & NMR system: a new software suite for macromolecular structure determination. Acta Crystallogr. D Biol. Crystallogr. 54, 905–921 (1998).

    Article  Google Scholar 

  55. Adams, P.D. et al. PHENIX: building new software for automated crystallographic structure determination. Acta Crystallogr. D Biol. Crystallogr. 58, 1948–1954 (2002).

    Article  Google Scholar 

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

    Article  CAS  Google Scholar 

  57. Geertsma, E.R., Nik Mahmood, N.A., Schuurman-Wolters, G.K. & Poolman, B. Membrane reconstitution of ABC transporters and assays of translocator function. Nat. Protoc. 3, 256–266 (2008).

    Article  CAS  Google Scholar 

  58. Keller, S. et al. High-precision isothermal titration calorimetry with automated peak-shape analysis. Anal. Chem. 84, 5066–5073 (2012).

    Article  CAS  Google Scholar 

  59. Lorenz, C., Pusch, M. & Jentsch, T.J. Heteromultimeric CLC chloride channels with novel properties. Proc. Natl. Acad. Sci. USA 93, 13362–13366 (1996).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

This research was supported by the Swiss National Science Foundation through the National Centre of Competence in Research TransCure. We thank the staff of the X06SA beamline for support during data collection, B. Blattman and C. Stutz-Ducommun of the Protein Crystallization Center at the University of Zurich for support with crystallization, B. Dreier for help with MALS experiments, the Center for Microscopy and Image Analysis at the University of Zurich for help with freeze-fracture EM, M. Hediger (University of Bern) for providing the cDNA of human DMT1 and E. Beke (Vrije Universiteit Brussel) for help with nanobody selection. All members of the Dutzler laboratory are acknowledged for help in all stages of the project. E.R.G. acknowledges a long-term postdoctoral fellowship from the Human Frontier Science Program (LT-00899/2008). I.A.E. is affiliated with the Biomolecular Structure and Mechanism PhD program of the University of Zurich (UZH) and the Swiss Federal Institute of Technology (ETH) Zurich. Data collection was performed at the X06SA beamline at the Swiss Light Source of the Paul Scherrer Institute.

Author information

Authors and Affiliations

Authors

Contributions

I.A.E. carried out all experiments except for the initial nanobody selection. E.R.G. supported the high-throughput expression screening and transport assays and initiated nanobody selection by phage display. E.P. performed immunization, cloned and expressed nanobodies and performed the initial selections. J.S. supervised nanobody production. R.D. assisted I.A.E. in structure determination. I.A.E. and R.D. jointly planned the experiments, analyzed the data and wrote the manuscript.

Corresponding author

Correspondence to Raimund Dutzler.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Sequence alignment.

Sequences of ScaDMT and human DMT1 (isoform 3) were aligned with ClustalW. Identical residues are highlighted in green and homologous residues in yellow. Numbering corresponds to ScaDMT. Secondary structure elements are shown below. Labels indicate: (black triangle) Start of the truncated construct ScaDMTtru, (red circle) residues of the transition metal binding site, (blue circle) residues involved in pH dependence and H+ transport of DMT1, (green circle) DMT1 disease mutation in the Belgrade rat and the mk/mk mouse and (brown circle) DMT1 disease mutations in human.

Supplementary Figure 2 Multiangle light-scattering and freeze-fracture EM.

Gel filtration and light scattering results for different protein constructs in the detergent DM. Continuous black traces correspond to the UV280 elution profiles. The respective molecular weight of the protein-detergent complex obtained from light scattering is shown at its corresponding position on the chromatogram in green, the molecular weight of the protein component alone in red. Panels show (a) ScaDMT, (b) ScaDMTtru, (c) nanobody, (d) ScaDMT–nanobody complex, (e) ScaDMTtru–nanobody complex. The first peak in panels (c) and (d) corresponds to a homo-dimer of the respective transporter-nanobody complex. (f) Freeze-fracture electron micrograph of proteoliposomes containing ScaDMT. The transporter was reconstituted at a protein to lipid ratio of 1:40 (w/w).

Supplementary Figure 3 Electron density.

(a) Stereo view of the ion-binding region. Experimental electron density calculated at 3.5 Å with Se-Met SAD phases that were improved by solvent flattening and cyclic 2-fold NCS symmetry averaging (blue mesh, contoured at 1 σ) is superimposed on the refined model of the ScaDMTtru–NB complex. (b) The same region of the protein is shown with 2Fo-Fc electron density superimposed (cyan mesh, contoured at 1 σ). The density at 3.1 Å was calculated with phases from the refined model. (c) Anomalous difference electron density of Se atoms (calculated at 3.5 Å and contoured at 5 σ) superimposed on the ScaDMT structure. Protein is shown as Cα-trace, methionine side-chains as sticks. (d) Anomalous difference density of data collected at a wavelength of 1.95 Å. Electron density calculated at 4.5 Å and contoured at 4 σ is superimposed on the model. The density shows peaks for sulfur atoms of methionine and cysteine residues but not for bound Ca2+. (e) Anomalous difference density of Cs+ (calculated at 4.5 Å and contoured at 5 σ, red) is shown superimposed on the ScaDMT structure. ScaDMT is represented as Cα-trace, the position of Mn2+ as black sphere.

Supplementary Figure 4 Nanobody complex and comparison with related proteins.

Stereo views of ScaDMT–nanobody interactions. (a) Structure of the ScaDMT–nanobody complex. Proteins are displayed as Cα-trace. The coloring of the protein is as in Fig. 2, the nanobody is colored in green. (b) Close-up of the interaction interface. Interacting residues are shown as sticks. (c) Stereo view of the superposition of ScaDMT on the inward-facing conformation of LeuT. The superposition was based on the Cα positions of equivalent residues in transmembrane helices (RMSD 3.2 Å). ScaDMT is colored in orange, LeuT in red. (d) Stereo view of the superposition of ScaDMT on the inward-facing conformation of vSGLT. The superposition was based on the Cα positions of equivalent residues in transmembrane helices (r.m.s.d. 3.9 Å). ScaDMT is colored in orange, vSGLT in green.

Supplementary Figure 5 Structure of full-length ScaDMT.

Stereo view of the full-length ScaDMT structure at 6.5 Å resolution. The structure shows one of the two transporters present in the asymmetric unit. The electron density calculated from phases obtained by molecular replacement with the program Phaser is colored in blue, electron density after rigid-body refinement in PHENIX is shown in cyan. Both electron densities are contoured at 1 σ. ScaDMT is colored in orange, α-helix 1a of LeuT from the superimposed model is shown in red. (a) View from within the membrane, (b) view from the cytoplasm.

Supplementary Figure 6 Transport properties of ScaDMT binding-site mutants.

Transport of Mn2+ into proteoliposomes containing the ScaDMT binding site mutants D49A (a), N52A (b) and M226A (c). Left: Time-dependent quenching of the fluorophore calcein, that is trapped inside the vesicle, upon addition of 300 μM Mn2+ to the external medium. Addition of Mn2+ and the ionophore calcimycin (Cal.), which acts as Mn2+/H+ exchanger, are indicated. A control trace from liposomes devoid of protein upon addition of 300 μM Mn2+ is shown in black, transport into proteoliposomes containing WT is shown in grey. Compared to WT, the proteoliposomes containing the mutants D49A, N52A and M226A show decreased activity. Right: ScaDMT-mediated transport of Mn2+ in the presence of Cd2+. Time-dependent quenching of the fluorophore calcein, was monitored upon addition of 100 μM of Mn2+ and 100 μM Cd2+ to the external medium. Transport by WT under the same conditions is shown in grey for comparison, traces from liposomes devoid of protein are shown in black. In all cases the presence of Cd2+ decreases the transport of Mn2+ into liposomes.

Supplementary Figure 7 Location of disease mutations.

The position of disease causing mutations in DMT1 in rodents (green) and human (magenta) are mapped on equivalent positions on the ScaDMT structure. The numbering corresponds to human DMT1. The protein is shown as Cα-trace, the mutated residues are indicated as spheres, the approximate membrane boundaries are indicated in grey.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–7 and Supplementary Table 1 (PDF 5088 kb)

Supplementary Data Set 1

Western blot of human DMT1 mutants (PDF 178 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ehrnstorfer, I., Geertsma, E., Pardon, E. et al. Crystal structure of a SLC11 (NRAMP) transporter reveals the basis for transition-metal ion transport. Nat Struct Mol Biol 21, 990–996 (2014). https://doi.org/10.1038/nsmb.2904

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nsmb.2904

This article is cited by

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