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:

X-ray structure of a calcium-activated TMEM16 lipid scramblase

Abstract

The TMEM16 family of proteins, also known as anoctamins, features a remarkable functional diversity. This family contains the long sought-after Ca2+-activated chloride channels as well as lipid scramblases and cation channels. Here we present the crystal structure of a TMEM16 family member from the fungus Nectria haematococca that operates as a Ca2+-activated lipid scramblase. Each subunit of the homodimeric protein contains ten transmembrane helices and a hydrophilic membrane-traversing cavity that is exposed to the lipid bilayer as a potential site of catalysis. This cavity harbours a conserved Ca2+-binding site located within the hydrophobic core of the membrane. Mutations of residues involved in Ca2+ coordination affect both lipid scrambling in N. haematococca TMEM16 and ion conduction in the Cl− channel TMEM16A. The structure reveals the general architecture of the family and its mode of Ca2+ activation. It also provides insight into potential scrambling mechanisms and serves as a framework to unravel the conduction of ions in certain TMEM16 proteins.

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

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

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

Figure 1: Phospholipid scrambling by nhTMEM16.
Figure 2: nhTMEM16 structure.
Figure 3: Dimer interface and subunit cavity.
Figure 4: Ca2+-binding site.
Figure 5: Functional properties of Ca2+-binding site mutants.
Figure 6: Mechanism.

Similar content being viewed by others

Accession codes

Primary accessions

Protein Data Bank

Data deposits

Coordinates and structure factors have been deposited in the Protein Data Bank under accession codes 4WIS (nhTMEM16 CF1) and 4WIT (nhTMEM16 CF2).

References

  1. Milenkovic, V. M., Brockmann, M., Stohr, H., Weber, B. H. & Strauss, O. Evolution and functional divergence of the anoctamin family of membrane proteins. BMC Evol. Biol. 10, 319 (2010)

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  2. Pedemonte, N. & Galietta, L. J. Structure and function of TMEM16 proteins (anoctamins). Physiol. Rev. 94, 419–459 (2014)

    Article  CAS  PubMed  Google Scholar 

  3. Yang, Y. D. et al. TMEM16A confers receptor-activated calcium-dependent chloride conductance. Nature 455, 1210–1215 (2008)

    Article  ADS  CAS  PubMed  Google Scholar 

  4. Caputo, A. et al. TMEM16A, a membrane protein associated with calcium-dependent chloride channel activity. Science 322, 590–594 (2008)

    Article  ADS  CAS  PubMed  Google Scholar 

  5. Schroeder, B. C., Cheng, T., Jan, Y. N. & Jan, L. Y. Expression cloning of TMEM16A as a calcium-activated chloride channel subunit. Cell 134, 1019–1029 (2008)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Pifferi, S., Cenedese, V. & Menini, A. Anoctamin 2/TMEM16B: a calcium-activated chloride channel in olfactory transduction. Exp. Physiol. 97, 193–199 (2012)

    Article  CAS  PubMed  Google Scholar 

  7. Ferrera, L., Zegarra-Moran, O. & Galietta, L. J. Ca2+-activated Cl– channels. Compr. Physiol. 1, 2155–2174 (2011)

    Article  PubMed  Google Scholar 

  8. Huang, F., Wong, X. & Jan, L. Y. International Union of Basic and Clinical Pharmacology. LXXXV: calcium-activated chloride channels. Pharmacol. Rev. 64, 1–15 (2012)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Billig, G. M., Pal, B., Fidzinski, P. & Jentsch, T. J. Ca2+-activated Cl– currents are dispensable for olfaction. Nature Neurosci. 14, 763–769 (2011)

    Article  CAS  PubMed  Google Scholar 

  10. Yang, H. et al. TMEM16F forms a Ca2+-activated cation channel required for lipid scrambling in platelets during blood coagulation. Cell 151, 111–122 (2012)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Martins, J. R. et al. Anoctamin 6 is an essential component of the outwardly rectifying chloride channel. Proc. Natl Acad. Sci. USA 108, 18168–18172 (2011)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  12. Suzuki, J., Umeda, M., Sims, P. J. & Nagata, S. Calcium-dependent phospholipid scrambling by TMEM16F. Nature 468, 834–838 (2010)

    Article  ADS  CAS  PubMed  Google Scholar 

  13. Suzuki, J. et al. Calcium-dependent phospholipid scramblase activity of TMEM16 protein family members. J. Biol. Chem. 288, 13305–13316 (2013)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Malvezzi, M. et al. Ca2+-dependent phospholipid scrambling by a reconstituted TMEM16 ion channel. Nature Commun. 4, 2367 (2013)

    Article  ADS  Google Scholar 

  15. Kunzelmann, K. et al. Molecular functions of anoctamin 6 (TMEM16F): a chloride channel, cation channel, or phospholipid scramblase? Pflugers Arch. 466, 407–414 (2014)

    Article  CAS  PubMed  Google Scholar 

  16. Kuruma, A. & Hartzell, H. C. Bimodal control of a Ca2+-activated Cl− channel by different Ca2+ signals. J. Gen. Physiol. 115, 59–80 (2000)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Ni, Y. L., Kuan, A. S. & Chen, T. Y. Activation and inhibition of TMEM16A calcium-activated chloride channels. PLoS ONE 9, e86734 (2014)

    Article  ADS  PubMed  PubMed Central  CAS  Google Scholar 

  18. Yu, K., Duran, C., Qu, Z., Cui, Y. Y. & Hartzell, H. C. Explaining calcium-dependent gating of anoctamin-1 chloride channels requires a revised topology. Circ. Res. 110, 990–999 (2012)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Terashima, H., Picollo, A. & Accardi, A. Purified TMEM16A is sufficient to form Ca2+-activated Cl– channels. Proc. Natl Acad. Sci. USA 110, 19354–19359 (2013)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  20. Menon, I. et al. Opsin is a phospholipid flippase. Curr. Biol. 21, 149–153 (2011)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Tien, J. et al. A comprehensive search for calcium binding sites critical for TMEM16A calcium-activated chloride channel activity. eLife 3, e02772 (2014)

    Article  PubMed Central  CAS  Google Scholar 

  22. Fallah, G. et al. TMEM16A(a)/anoctamin-1 shares a homodimeric architecture with CLC chloride channels. Mol. Cell. Proteomics 10, M110.004697 (2011)

    Article  PubMed  CAS  Google Scholar 

  23. Sheridan, J. T. et al. Characterization of the oligomeric structure of the Ca2+-activated Cl– channel Ano1/TMEM16A. J. Biol. Chem. 286, 1381–1388 (2011)

    Article  CAS  PubMed  Google Scholar 

  24. Tien, J., Lee, H. Y., Minor, D. L., Jr, Jan, Y. N. & Jan, L. Y. Identification of a dimerization domain in the TMEM16A calcium-activated chloride channel (CaCC). Proc. Natl Acad. Sci. USA 110, 6352–6357 (2013)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  25. Hartzell, H. C., Yu, K., Xiao, Q., Chien, L. T. & Qu, Z. Anoctamin/TMEM16 family members are Ca2+-activated Cl− channels. J. Physiol. (Lond.) 587, 2127–2139 (2009)

    Article  CAS  Google Scholar 

  26. Sanyal, S. & Menon, A. K. Flipping lipids: why an’ what’s the reason for? ACS Chem. Biol. 4, 895–909 (2009)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Suzuki, J., Denning, D. P., Imanishi, E., Horvitz, H. R. & Nagata, S. Xk-related protein 8 and CED-8 promote phosphatidylserine exposure in apoptotic cells. Science 341, 403–406 (2013)

    Article  ADS  CAS  PubMed  Google Scholar 

  28. Sanyal, S. & Menon, A. K. Stereoselective transbilayer translocation of mannosyl phosphoryl dolichol by an endoplasmic reticulum flippase. Proc. Natl Acad. Sci. USA 107, 11289–11294 (2010)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  29. Mohammadi, T. et al. Identification of FtsW as a transporter of lipid-linked cell wall precursors across the membrane. EMBO J. 30, 1425–1432 (2011)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Zwaal, R. F., Comfurius, P. & Bevers, E. M. Scott syndrome, a bleeding disorder caused by defective scrambling of membrane phospholipids. Biochim. Biophys. Acta 1636, 119–128 (2004)

    Article  CAS  PubMed  Google Scholar 

  31. Dekkers, D. W., Comfurius, P., Bevers, E. M. & Zwaal, R. F. Comparison between Ca2+-induced scrambling of various fluorescently labelled lipid analogues in red blood cells. Biochem. J. 362, 741–747 (2002)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Sham, L. T. et al. Bacterial cell wall. MurJ is the flippase of lipid-linked precursors for peptidoglycan biogenesis. Science 345, 220–222 (2014)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  33. Suzuki, T., Suzuki, J. & Nagata, S. Functional swapping between transmembrane proteins TMEM16A and TMEM16F. J. Biol. Chem. 289, 7438–7447 (2014)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Miller, C. Open-state substructure of single chloride channels from Torpedo electroplax. Phil. Trans. R. Soc. Lond. B 299, 401–411 (1982)

    Article  ADS  CAS  Google Scholar 

  35. Kmit, A. et al. Calcium-activated and apoptotic phospholipid scrambling induced by Ano6 can occur independently of Ano6 ion currents. Cell Death Dis. 4, e611 (2013)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Tian, Y., Schreiber, R. & Kunzelmann, K. Anoctamins are a family of Ca2+-activated Cl− channels. J. Cell Sci. 125, 4991–4998 (2012)

    CAS  PubMed  Google Scholar 

  37. Baldridge, R. D. & Graham, T. R. Identification of residues defining phospholipid flippase substrate specificity of type IV P-type ATPases. Proc. Natl Acad. Sci. USA 109, E290–E298 (2012)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  38. Vestergaard, A. L. et al. Critical roles of isoleucine-364 and adjacent residues in a hydrophobic gate control of phospholipid transport by the mammalian P4-ATPase ATP8A2. Proc. Natl Acad. Sci. USA 111, E1334–E1343 (2014)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Pifferi, S., Dibattista, M. & Menini, A. TMEM16B induces chloride currents activated by calcium in mammalian cells. Pflugers Arch. 458, 1023–1038 (2009)

    Article  CAS  PubMed  Google Scholar 

  40. Woodhull, A. M. Ionic blockage of sodium channels in nerve. J. Gen. Physiol. 61, 687–708 (1973)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Arreola, J., Melvin, J. E. & Begenisich, T. Activation of calcium-dependent chloride channels in rat parotid acinar cells. J. Gen. Physiol. 108, 35–47 (1996)

    Article  CAS  PubMed  Google Scholar 

  42. Xiao, Q. et al. Voltage- and calcium-dependent gating of TMEM16A/Ano1 chloride channels are physically coupled by the first intracellular loop. Proc. Natl Acad. Sci. USA 108, 8891–8896 (2011)

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  43. Vocke, K. et al. Calmodulin-dependent activation and inactivation of anoctamin calcium-gated chloride channels. J. Gen. Physiol. 142, 381–404 (2013)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Yu, K., Zhu, J., Qu, Z., Cui, Y. Y. & Hartzell, H. C. Activation of the Ano1 (TMEM16A) chloride channel by calcium is not mediated by calmodulin. J. Gen. Physiol. 143, 253–267 (2014)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Tian, Y. et al. Calmodulin-dependent activation of the epithelial calcium-dependent chloride channel TMEM16A. FASEB J. 25, 3. 1058–1068 (2011)

    Article  CAS  PubMed  Google Scholar 

  46. Jung, J. et al. Dynamic modulation of ANO1/TMEM16A HCO3− permeability by Ca2+/calmodulin. Proc. Natl Acad. Sci. USA 110, 360–365 (2013)

    Article  ADS  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  48. Keefe, A. D., Wilson, D. S., Seelig, B. & Szostak, J. W. One-step purification of recombinant proteins using a nanomolar-affinity streptavidin-binding peptide, the SBP-tag. Protein Expr. Purif. 23, 440–446 (2001)

    Article  CAS  PubMed  Google Scholar 

  49. Nagai, T. et al. A variant of yellow fluorescent protein with fast and efficient maturation for cell-biological applications. Nature Biotechnol. 20, 87–90 (2002)

    Article  CAS  Google Scholar 

  50. Schenck, S., Wojcik, S. M., Brose, N. & Takamori, S. A chloride conductance in VGLUT1 underlies maximal glutamate loading into synaptic vesicles. Nature Neurosci. 12, 156–162 (2009)

    Article  CAS  PubMed  Google Scholar 

  51. Drew, D. et al. GFP-based optimization scheme for the overexpression and purification of eukaryotic membrane proteins in Saccharomyces cerevisiae. Nature Protocols 3, 784–798 (2008)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. 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 

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

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

    Article  CAS  PubMed  Google Scholar 

  55. 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 

  56. de La Fortelle, E. & Bricogne, G. in Methods in Enzymology Vol. 276 (eds Carter, C. W. & Sweet, R. M. ) 492–494 (Academic, 1997)

    Google Scholar 

  57. Cowtan, K. ‘dm’: An automated procedure for phase improvement by density modification. Joint CCP4 and ESF-EACBM Newsletter on Protein Crystallography 31, 34–38 (1994)

    Google Scholar 

  58. 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  PubMed  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

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

    Article  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  64. Garcia-Celma, J., Szydelko, A. & Dutzler, R. Functional characterization of a ClC transporter by solid-supported membrane electrophysiology. J. Gen. Physiol. 141, 479–491 (2013)

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  65. Accardi, A., Kolmakova-Partensky, L., Williams, C. & Miller, C. Ionic currents mediated by a prokaryotic homologue of CLC Cl− channels. J. Gen. Physiol. 123, 109–119 (2004)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Zimmermann, I. & Dutzler, R. Ligand activation of the prokaryotic pentameric ligand-gated ion channel ELIC. PLoS Biol. 9, e1001101 (2011)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Sievers, F. et al. Fast, scalable generation of high-quality protein multiple sequence alignments using Clustal Omega. Mol. Syst. Biol. 7, 539 (2011)

    Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

This research was supported by a grant from the European Research Council (no. 339116, AnoBest) and 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 UZH, for their support with crystallization, B. Dreier for help with MALS experiments, A. Szydelko for providing ecClC as negative control in the scramblase assay and D. Drew for the FGY217 yeast strain. All members of the Dutzler laboratory are acknowledged for help in all stages of the project.

Author information

Authors and Affiliations

Authors

Contributions

J.D.B. screened homologues, purified and crystallized nhTMEM16 and performed scrambling experiments. N.K.L. screened and crystallized homologues, performed electrophysiological recordings and did the MALS measurement. S.S. started the project, made expression vectors and aided in cell culture. A.D. screened homologues. R.D. assisted J.D.B. and N.K.L. during structure determination. J.D.B., N.K.L., S.S. and R.D. jointly planned experiments, analysed data and wrote the manuscript.

Corresponding author

Correspondence to Raimund Dutzler.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Extended data figures and tables

Extended Data Figure 1 Structure-based sequence alignment.

Sequences were aligned with Clustal Omega67 and edited manually. Identical residues are highlighted in green, homologous residues in yellow and residues of the Ca2+-binding site in red. Secondary structure elements are shown below. a, Comparison of nhTMEM16 and afTMEM16. The numbering corresponds to nhTMEM16. b, Comparison of the membrane domains of selected TMEM16 proteins, m refers to murine, hs to human proteins. Long insertions in loop regions of mammalian family members (indicated by -xxx-) are not shown in the alignment. The positions of residues in α-helix 10 involved in an inter-subunit salt bridge at the dimer interface are highlighted in cyan. c, Comparison of the observed and predicted topology of TMEM16 proteins. Sequence alignment of the membrane spanning regions of mTMEM16A and nhTMEM16 with the observed (green) and predicted topology25 (red) indicated. Identical residues are highlighted in green, homologous residues in yellow, residues of the Ca2+-binding site in red and the inter-subunit salt bridge at the dimer interface in cyan. The difference between the predicted and observed transmembrane segments is due to the failure of sequence-based approaches to identify the correct boundaries of several helices and to detect α-helix 6 at all and helices 7 and 8 as separate entities.

Extended Data Figure 2 Multi-angle light scattering and lipid scrambling.

a, Gel filtration and light scattering results for nhTMEM16 in the detergent DDM. The continuous black trace corresponds to the absorption at 280 nm. Molecular weights of the protein and the protein-detergent complex are shown in red and green, respectively. b, Inaccessibility of NBD groups trapped within liposomes. Dithionite is incapable of reducing the soluble NBD-dextran trapped in the interior of proteoliposomes containing nhTMEM16. Traces of proteoliposomes containing nhTMEM16 and empty liposomes are shown in red and black respectively. Traces from proteoliposomes of nhTMEM16 containing NBD-PE at equivalent dithionite levels are shown for comparison (blue). Asterisk marks addition of 2.5 mM dithionite. c, SDS–PAGE gel of the Ca2+-binding site triple-mutant E452Q/E535Q/D539N (M) and nhTMEM16 (WT) used for reconstitution illustrating the purity of the sample. The molecular weight marker (MW) is shown on the left with selected bands labelled. d, Analysis of phospholipid scrambling. Time dependent fluorescence decrease of NBD-PE upon reduction by 30 mM dithionite (t = 0). The traces are as in Figs 1b–d and 5a. A fit to a single exponential decay is shown as dotted lines for protein-free/Ca2+ and nhTMEM16/Ca2+ with time constants of 15 s and 22 s, respectively. A fit to a sum of two exponential functions is shown for nhTMEM16/no Ca2+ and E452Q/E535Q/D539N/Ca2+ with time constants of 25 and 21 s for the fast component and 175 and 803 s for the slow component, respectively.

Extended Data Figure 3 Search for ion channel activity in nhTMEM16.

a, Freeze-fracture electron microscopy image of a proteoliposome containing nhTMEM16 formed from a 3:1 mixture of E. coli polar lipids/egg PC. Reconstituted proteins are labelled with red asterisks. b, Planar lipid bilayer experiments. Currents recorded after fusion of proteoliposomes containing nhTMEM16 expressed and purified from S. cerevisiae (Sc) in the absence of Ca2+ (top, left), with 300 μM Ca2+ added on both sides of the bilayer (top right) and of proteoliposomes containing nhTMEM16 expressed and purified from HEK tsA201 cells in the presence of 300 μM Ca2+ added on both sides of the bilayer (bottom left). Currents recorded after fusion of liposomes of the same lipid composition not containing any protein are shown for comparison (bottom right). Displayed traces were recorded at a holding potential of 100 mV in symmetric solutions containing 150 mM NaCl and 10 mM HEPES pH 7.4. Selected current levels (in pA) are indicated on the left. c, Fluorescence confocal microscopy images of HEK tsA201 cells expressing a mTMEM16A–YFP fusion construct (left) or a nhTMEM16–YFP fusion construct (right). d, Recordings from excised inside-out patches. Representative current response in a membrane patch excised from cells expressing a mTMEM16A–YFP fusion construct upon rapid exchange into solutions containing the indicated amount of Ca2+ (left) and equivalent recordings from patches that were excised from cells expressing a nhTMEM16–YFP fusion protein (right). The voltage was clamped at 80 mV. The fluorescence of transfected cells expressing mTMEM16A–YFP used for recording is shown below. No activity of nhTMEM16–YFP was observed in any of more than 30 patches. e, Patch-clamp recording in the whole-cell configuration. Representative currents from a HEK293T cell expressing a mTMEM16A–YFP construct recorded from a solution containing either 0.1 µM (left) or 20 µM (right) free Ca2+ in the patch pipette. f, Representative currents from a cell expressing a nhTMEM16–YFP fusion protein recorded from a solution containing 20 µM free Ca2+ in the patch pipette (left). Current response from mock-transfected cells recorded under the same conditions is shown for comparison (right). Insets show part of the traces with magnified current scale.

Extended Data Figure 4 Crystallography.

a, Table describing data collection and refinement statistics of five data sets presented in this study. nhTMEM16/CF1 and nhTMEM16/CF2 are data sets used for the building and refinement of the crystal structures of CF1 and CF2 respectively that have been deposited in the PDB. nhTMEM16/SeMet, a data set of a selenomethionine derivative collected at the Se anomalous absorption edge, was used for obtaining initial phases of CF2. nhTMEM16/anom is a data set used for the identification of the Ca2+-binding site by anomalous scattering and nhTMEM16/no Ca2+ is from a protein purified in the presence of EDTA and crystallized without addition of Ca2+. b, Stereo view of the Ca2+-binding region in CF1. The model of the protein displayed as sticks is shown with experimental electron density superimposed. The map was calculated at 3.3 Å with Se-Met SAD phases that were improved by solvent flattening, cyclic twofold NCS and cross-crystal averaging (blue mesh, contoured at 1σ). Ca2+ ions are shown as blue spheres.

Extended Data Figure 5 Electron density.

a, Stereo view of the Ca2+-binding region in CF1. The model of the protein displayed as sticks is shown with 2Fo – Fc electron density superimposed (cyan mesh, contoured at 1σ after sharpening with b = 50). The density at 3.3 Å was calculated with phases from the refined model. Ca2+ ions are shown as blue spheres. b, 2Fo – Fc electron density of the Ca2+-binding region in CF2 (calculated at 3.4 Å and contoured at 1σ after sharpening with b = 50, orange) superimposed on the refined model. c, Stereo view of the Ca2+-binding region of a structure obtained from protein purified in the presence of EDTA and crystallized in CF2 without addition of Ca2+. 2Fo – Fc electron density (cyan mesh, calculated at 4.2 Å and contoured at 1σ after sharpening with a b = 50) and Fo – Fc density (contoured at 3σ, green) is superimposed on the refined model. No ions were included in the refinement. d, Close-up of the Ca2+-binding site. Anomalous difference density (left, calculated at 6 Å and contoured at 4σ, magenta) and Fo – Fc density (right, contoured at 3σ, green) indicates the presence of bound Ca2+ ions.

Extended Data Figure 6 nhTMEM16 dimer.

Stereo views of a ribbon representation of the dimeric protein. Bound Ca2+ ions are shown as blue spheres. a, View from within the membrane; b, view from the extracellular side; c, view from the cytoplasm.

Extended Data Figure 7 Model of lipid interactions and dimer cavity.

a, Model of nhTMEM16 embedded in a lipid membrane (left). The protein was positioned within the model of a PC bilayer (obtained from http://www.lobos.nih.gov/mbs/coords.shtml). A ribbon representation of the protein and the molecular surface are shown. Lipids are displayed as CPK models. Same view of the protein with regions on the surface presumably in contact with the membrane coloured in orange (right). b, Putative location of α-helices 0a and 0b relative to the lipid bilayer. c, Inter-subunit interactions between residues of α-helix 10. The protein is shown as sticks with 2Fo – Fc density (CF1, calculated at 3.3 Å and contoured at 1σ after sharpening with b = 50, cyan mesh) superimposed (left). A sequence alignment of the corresponding region underlines the conservation of interacting residues. Amino acids of the salt bridge in nhTMEM16 are highlighted in cyan, the numbering corresponds to nhTMEM16. d, View on the dimer cavity from the dimer interface. The molecular surface is coloured according to the properties of contacting residues (red, acidic; blue, basic; green, polar). A modelled lipid indicates the boundary of the inner leaflet of the bilayer. e, Stereo view of the cleft between α-helices 3 and 10. The protein is shown as stick model. The molecular surface is coloured according to the properties of contacting residues (yellow, hydrophobic; orange, aromatic). Lipids indicate the membrane boundary. f, Residual density in the dimer cavity. The molecular surface is coloured in white. 2Fo – Fc density (CF2, contoured at 1σ after sharpening with b = 50, orange) and Fo – Fc density (contoured at 3σ, green) are shown. The view is as in d.

Extended Data Figure 8 Subunit cavity and Ca2+-binding site.

a, Stereo view of the subunit cavity viewed from within the membrane. Protein residues and the molecular surface are shown. b, Residual density in the subunit cavity. The molecular surface of the protein is shown. 2Fo – Fc density (CF2, contoured at 1σ after sharpening with b = 50, orange) and Fo – Fc density (contoured at 3σ, green) are displayed. c, Model of the subunit cavity in different TMEM16 proteins. The molecular surface is coloured according to the properties of contacting residues (red, acidic; blue, basic; green, polar). Putative surface-exposed residues were obtained from a sequence alignment with nhTMEM16. d, Location of the Ca2+-binding site in relation to the lipid bilayer. Modelled lipids of the inner leaflet of the bilayer are shown as sticks.

Extended Data Figure 9 Electrophysiology.

Current response in HEK293T cells overexpressing mTMEM16A–YFP and point mutants of the Ca2+-binding site. All recordings were measured from single excised patches in the inside-out configuration after changing to intracellular solutions containing the indicated Ca2+ concentrations. a, WT, with voltage protocol shown as inset. b, Mutant N650A, with the voltage protocol shown as inset. c–g, Recordings of mutants E654A, c, E702Q, d, E705Q, e, E634Q, f, and mutant D738N, g.

Extended Data Figure 10 Ca2+ activation of mTMEM16A.

Representative current traces of mTMEM16A and mutants of the Ca2+-binding site. Currents were measured from excised inside-out patches of HEK293T cells expressing the respective protein at 80 mV. The Ca2+ concentration is indicated, selected traces are shown in colour. a, WT, b, N650A, c, E734Q, d, E702Q, e, D738N f, E705Q and g, E654A. h, Fluorescence microscopy image of mTMEM16A mutants expressed in HEK293T cells. Expression of fluorescently labelled protein is shown for mutants E654Q (left) and E654A (right). i, Analysis of the EC50 of Ca2+ activation for different binding site mutants. The data show averages of fits to 3–4 independent recordings. Errors are s.d.

Supplementary information

Supplementary Information

This file contains a Supplementary Discussion and additional references. (PDF 249 kb)

PowerPoint slides

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Brunner, J., Lim, N., Schenck, S. et al. X-ray structure of a calcium-activated TMEM16 lipid scramblase. Nature 516, 207–212 (2014). https://doi.org/10.1038/nature13984

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nature13984

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