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:

Architecture of the mammalian mechanosensitive Piezo1 channel

This article has been updated

Abstract

Piezo proteins are evolutionarily conserved and functionally diverse mechanosensitive cation channels. However, the overall structural architecture and gating mechanisms of Piezo channels have remained unknown. Here we determine the cryo-electron microscopy structure of the full-length (2,547 amino acids) mouse Piezo1 (Piezo1) at a resolution of 4.8 Å. Piezo1 forms a trimeric propeller-like structure (about 900 kilodalton), with the extracellular domains resembling three distal blades and a central cap. The transmembrane region has 14 apparently resolved segments per subunit. These segments form three peripheral wings and a central pore module that encloses a potential ion-conducting pore. The rather flexible extracellular blade domains are connected to the central intracellular domain by three long beam-like structures. This trimeric architecture suggests that Piezo1 may use its peripheral regions as force sensors to gate the central ion-conducting pore.

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: Piezo1 forms a homotrimer.
Figure 2: Overall structure of Piezo1.
Figure 3: Organization of the transmembrane skeleton.
Figure 4: Putative ion-conducting pore.
Figure 5: Conformational heterogeneity of the ‘blade’ and a proposed model of force-induced gating of Piezo channels.

Similar content being viewed by others

Accession codes

Primary accessions

Electron Microscopy Data Bank

Protein Data Bank

Data deposits

The 3D cryo-electron microscopy density map has been deposited in the Electron Microscopy Data Bank (EMDB), with accession code EMD-6343. The coordinates of atomic models have been deposited in the Protein Data Bank (PDB) under the accession codes 4RAX for the CED and 3JAC for the full length.

Change history

  • 04 November 2015

    A grant number in the Acknowledgements was corrected.

References

  1. Chalfie, M. Neurosensory mechanotransduction. Nature Rev. Mol. Cell Biol. 10, 44–52 (2009)

    Article  CAS  Google Scholar 

  2. Coste, B. et al. Piezo1 and Piezo2 are essential components of distinct mechanically activated cation channels. Science 330, 55–60 (2010)

    Article  ADS  CAS  Google Scholar 

  3. Coste, B. et al. Piezo proteins are pore-forming subunits of mechanically activated channels. Nature 483, 176–181 (2012)

    Article  ADS  CAS  Google Scholar 

  4. Nilius, B. & Honoré, E. Sensing pressure with ion channels. Trends Neurosci. 35, 477–486 (2012)

    Article  CAS  Google Scholar 

  5. Volkers, L., Mechioukhi, Y. & Coste, B. Piezo channels: from structure to function. Pflugers Arch. 467, 95–99 (2015)

    Article  CAS  Google Scholar 

  6. Bae, C., Gottlieb, P. A. & Sachs, F. Human PIEZO1: removing inactivation. Biophys. J. 105, 880–886 (2013)

    Article  ADS  CAS  Google Scholar 

  7. Gottlieb, P. A. & Sachs, F. Piezo1: properties of a cation selective mechanical channel. Channels 6, 214–219 (2012)

    Article  CAS  Google Scholar 

  8. Kim, S. E., Coste, B., Chadha, A., Cook, B. & Patapoutian, A. The role of Drosophila Piezo in mechanical nociception. Nature 483, 209–212 (2012)

    Article  ADS  CAS  Google Scholar 

  9. Faucherre, A., Nargeot, J., Mangoni, M. E. & Jopling, C. piezo2b regulates vertebrate light touch response. J. Neurosci. 33, 17089–17094 (2013)

    Article  CAS  Google Scholar 

  10. Schneider, E. R. et al. Neuronal mechanism for acute mechanosensitivity in tactile-foraging waterfowl. Proc. Natl Acad. Sci. USA 111, 14941–14946 (2014)

    Article  ADS  CAS  Google Scholar 

  11. Maksimovic, S. et al. Epidermal Merkel cells are mechanosensory cells that tune mammalian touch receptors. Nature 509, 617–621 (2014)

    Article  ADS  CAS  Google Scholar 

  12. Woo, S. H. et al. Piezo2 is required for Merkel-cell mechanotransduction. Nature 509, 622–626 (2014)

    Article  ADS  CAS  Google Scholar 

  13. Ranade, S. S. et al. Piezo2 is the major transducer of mechanical forces for touch sensation in mice. Nature 516, 121–125 (2014)

    Article  ADS  CAS  Google Scholar 

  14. Ikeda, R. et al. Merkel cells transduce and encode tactile stimuli to drive Aβ-afferent impulses. Cell 157, 664–675 (2014)

    Article  CAS  Google Scholar 

  15. Schrenk-Siemens, K. et al. PIEZO2 is required for mechanotransduction in human stem cell-derived touch receptors. Nature Neurosci. 18, 10–16 (2015)

    Article  CAS  Google Scholar 

  16. Ranade, S. S. et al. Piezo1, a mechanically activated ion channel, is required for vascular development in mice. Proc. Natl Acad. Sci. USA 111, 10347–10352 (2014)

    Article  ADS  CAS  Google Scholar 

  17. Li, J. et al. Piezo1 integration of vascular architecture with physiological force. Nature 515, 279–282 (2014)

    Article  ADS  CAS  Google Scholar 

  18. Faucherre, A., Kissa, K., Nargeot, J., Mangoni, M. E. & Jopling, C. Piezo1 plays a role in erythrocyte volume homeostasis. Haematologica 99, 70–75 (2014)

    Article  CAS  Google Scholar 

  19. Cahalan, S. M. et al. Piezo1 links mechanical forces to red blood cell volume. eLife 4, 07370 (2015)

    Article  Google Scholar 

  20. McHugh, B. J. et al. Integrin activation by Fam38A uses a novel mechanism of R-Ras targeting to the endoplasmic reticulum. J. Cell Sci. 123, 51–61 (2010)

    Article  Google Scholar 

  21. Pathak, M. M. et al. Stretch-activated ion channel Piezo1 directs lineage choice in human neural stem cells. Proc. Natl Acad. Sci. USA 111, 16148–16153 (2014)

    Article  ADS  CAS  Google Scholar 

  22. Albuisson, J. et al. Dehydrated hereditary stomatocytosis linked to gain-of-function mutations in mechanically activated PIEZO1 ion channels. Nature Commun. 4, 1884 (2013)

    Article  ADS  Google Scholar 

  23. Andolfo, I. et al. Multiple clinical forms of dehydrated hereditary stomatocytosis arise from mutations in PIEZO1 . Blood 121, 3925–3935 (2013)

    Article  CAS  Google Scholar 

  24. Bae, C., Gnanasambandam, R., Nicolai, C., Sachs, F. & Gottlieb, P. A. Xerocytosis is caused by mutations that alter the kinetics of the mechanosensitive channel PIEZO1. Proc. Natl Acad. Sci. USA 110, E1162–E1168 (2013)

    Article  ADS  CAS  Google Scholar 

  25. Beneteau, C. et al. Recurrent mutation in the PIEZO1 gene in two families of hereditary xerocytosis with fetal hydrops. Clin. Genet. 85, 293–295 (2014)

    Article  CAS  Google Scholar 

  26. Shmukler, B. E. et al. Dehydrated stomatocytic anemia due to the heterozygous mutation R2456H in the mechanosensitive cation channel PIEZO1: a case report. Blood Cells Mol. Dis. 52, 53–54 (2014)

    Article  CAS  Google Scholar 

  27. Zarychanski, R. et al. Mutations in the mechanotransduction protein PIEZO1 are associated with hereditary xerocytosis. Blood 120, 1908–1915 (2012)

    Article  CAS  Google Scholar 

  28. Coste, B. et al. Gain-of-function mutations in the mechanically activated ion channel PIEZO2 cause a subtype of distal arthrogryposis. Proc. Natl Acad. Sci. USA 110, 4667–4672 (2013)

    Article  ADS  CAS  Google Scholar 

  29. McMillin, M. J. et al. Mutations in PIEZO2 cause Gordon syndrome, Marden–Walker syndrome, and distal arthrogryposis type 5. Am. J. Hum. Genet. 94, 734–744 (2014)

    Article  CAS  Google Scholar 

  30. Zhang, X. et al. Crystal structure of an orthologue of the NaChBac voltage-gated sodium channel. Nature 486, 130–134 (2012)

    Article  ADS  CAS  Google Scholar 

  31. Payandeh, J., Scheuer, T., Zheng, N. & Catterall, W. A. The crystal structure of a voltage-gated sodium channel. Nature 475, 353–358 (2011)

    Article  CAS  Google Scholar 

  32. Kawate, T., Michel, J. C., Birdsong, W. T. & Gouaux, E. Crystal structure of the ATP-gated P2X(4) ion channel in the closed state. Nature 460, 592–598 (2009)

    Article  ADS  CAS  Google Scholar 

  33. Paulsen, C. E., Armache, J. P., Gao, Y., Cheng, Y. & Julius, D. Structure of the TRPA1 ion channel suggests regulatory mechanisms. Nature 520, 511–517 (2015)

    Article  ADS  CAS  Google Scholar 

  34. Liao, M., Cao, E., Julius, D. & Cheng, Y. Structure of the TRPV1 ion channel determined by electron cryo-microscopy. Nature 504, 107–112 (2013)

    Article  ADS  CAS  Google Scholar 

  35. Liu, Z., Gandhi, C. S. & Rees, D. C. Structure of a tetrameric MscL in an expanded intermediate state. Nature 461, 120–124 (2009)

    Article  ADS  CAS  Google Scholar 

  36. Bass, R. B., Strop, P., Barclay, M. & Rees, D. C. Crystal structure of Escherichia coli MscS, a voltage-modulated and mechanosensitive channel. Science 298, 1582–1587 (2002)

    Article  ADS  CAS  Google Scholar 

  37. Chang, G., Spencer, R. H., Lee, A. T., Barclay, M. T. & Rees, D. C. Structure of the MscL homolog from Mycobacterium tuberculosis: a gated mechanosensitive ion channel. Science 282, 2220–2226 (1998)

    Article  ADS  CAS  Google Scholar 

  38. Kung, C., Martinac, B. & Sukharev, S. Mechanosensitive channels in microbes. Annu. Rev. Microbiol. 64, 313–329 (2010)

    Article  CAS  Google Scholar 

  39. Brohawn, S. G., del Mármol, J. & MacKinnon, R. Crystal structure of the human K2P TRAAK, a lipid- and mechano-sensitive K+ ion channel. Science 335, 436–441 (2012)

    Article  ADS  CAS  Google Scholar 

  40. Coste, B. et al. Piezo1 ion channel pore properties are dictated by C-terminal region. Nature Commun. 6, 7223 (2015)

    Article  ADS  Google Scholar 

  41. Hou, X., Pedi, L., Diver, M. M. & Long, S. B. Crystal structure of the calcium release-activated calcium channel Orai. Science 338, 1308–1313 (2012)

    Article  ADS  CAS  Google Scholar 

  42. Penna, A. et al. The CRAC channel consists of a tetramer formed by Stim-induced dimerization of Orai dimers. Nature 456, 116–120 (2008)

    Article  ADS  CAS  Google Scholar 

  43. Kamajaya, A., Kaiser, J. T., Lee, J., Reid, M. & Rees, D. C. The structure of a conserved Piezo channel domain reveals a topologically distinct β sandwich fold. Structure 22, 1520–1527 (2014)

    Article  CAS  Google Scholar 

  44. Prole, D. L. & Taylor, C. W. Identification and analysis of putative homologues of mechanosensitive channels in pathogenic protozoa. PLoS One 8, e66068 (2013)

    Article  ADS  CAS  Google Scholar 

  45. Gonzales, E. B., Kawate, T. & Gouaux, E. Pore architecture and ion sites in acid-sensing ion channels and P2X receptors. Nature 460, 599–604 (2009)

    Article  ADS  CAS  Google Scholar 

  46. Vilceanu, D. & Stucky, C. L. TRPA1 mediates mechanical currents in the plasma membrane of mouse sensory neurons. PLoS One 5, e12177 (2010)

    Article  ADS  Google Scholar 

  47. Kwan, K. Y., Glazer, J. M., Corey, D. P., Rice, F. L. & Stucky, C. L. TRPA1 modulates mechanotransduction in cutaneous sensory neurons. J. Neurosci. 29, 4808–4819 (2009)

    Article  CAS  Google Scholar 

  48. Smart, O. S., Goodfellow, J. M. & Wallace, B. A. The pore dimensions of gramicidin A. Biophys. J. 65, 2455–2460 (1993)

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  50. Winn, M. D. et al. Overview of the CCP4 suite and current developments. Acta Crystallogr. D 67, 235–242 (2011)

    Article  CAS  Google Scholar 

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

    Article  Google Scholar 

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

    Article  CAS  Google Scholar 

  53. Cowtan, K. D. & Main, P. Improvement of macromolecular electron-density maps by the simultaneous application of real and reciprocal space constraints. Acta Crystallogr. D 49, 148–157 (1993)

    Article  CAS  Google Scholar 

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

    Article  Google Scholar 

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

    Article  Google Scholar 

  56. Tang, G. et al. EMAN2: an extensible image processing suite for electron microscopy. J. Struct. Biol. 157, 38–46 (2007)

    Article  CAS  Google Scholar 

  57. Scheres, S. H. RELION: implementation of a Bayesian approach to cryo-EM structure determination. J. Struct. Biol. 180, 519–530 (2012)

    Article  CAS  Google Scholar 

  58. Radermacher, M., Wagenknecht, T., Verschoor, A. & Frank, J. Three-dimensional reconstruction from a single-exposure, random conical tilt series applied to the 50S ribosomal subunit of Escherichia coli . J. Microsc. 146, 113–136 (1987)

    Article  CAS  Google Scholar 

  59. Shaikh, T. R. et al. SPIDER image processing for single-particle reconstruction of biological macromolecules from electron micrographs. Nature Protocols 3, 1941–1974 (2008)

    Article  CAS  Google Scholar 

  60. Li, X. et al. Electron counting and beam-induced motion correction enable near-atomic-resolution single-particle cryo-EM. Nature Methods 10, 584–590 (2013)

    Article  CAS  Google Scholar 

  61. Mindell, J. A. & Grigorieff, N. Accurate determination of local defocus and specimen tilt in electron microscopy. J. Struct. Biol. 142, 334–347 (2003)

    Article  Google Scholar 

  62. Scheres, S. H. & Chen, S. Prevention of overfitting in cryo-EM structure determination. Nature Methods 9, 853–854 (2012)

    Article  CAS  Google Scholar 

  63. Chen, S. et al. High-resolution noise substitution to measure overfitting and validate resolution in 3D structure determination by single particle electron cryomicroscopy. Ultramicroscopy 135, 24–35 (2013)

    Article  CAS  Google Scholar 

  64. Voorhees, R. M., Fernández, I. S., Scheres, S. H. & Hegde, R. S. Structure of the mammalian ribosome-Sec61 complex to 3.4 Å resolution. Cell 157, 1632–1643 (2014)

    Article  CAS  Google Scholar 

  65. Greber, B. J. et al. The complete structure of the large subunit of the mammalian mitochondrial ribosome. Nature 515, 283–286 10.1038/nature13895 (2014)

    Article  ADS  CAS  Google Scholar 

  66. Brown, A. et al. Structure of the large ribosomal subunit from human mitochondria. Science 346, 718–722 (2014)

    Article  ADS  CAS  Google Scholar 

  67. Fernández, I. S., Bai, X. C., Murshudov, G., Scheres, S. H. & Ramakrishnan, V. Initiation of translation by cricket paralysis virus IRES requires its translocation in the ribosome. Cell 157, 823–831 (2014)

    Article  Google Scholar 

  68. Kucukelbir, A., Sigworth, F. J. & Tagare, H. D. Quantifying the local resolution of cryo-EM density maps. Nature Methods 11, 63–65 (2014)

    Article  CAS  Google Scholar 

  69. Pettersen, E. F. et al. UCSF Chimera – a visualization system for exploratory research and analysis. J. Comput. Chem. 25, 1605–1612 (2004)

    Article  CAS  Google Scholar 

  70. Emsley, P., Lohkamp, B., Scott, W. G. & Cowtan, K. Features and development of Coot. Acta Crystallogr. D 66, 486–501 (2010)

    Article  CAS  Google Scholar 

  71. Larkin, M. A. et al. Clustal W and Clustal X version 2.0. Bioinformatics 23, 2947–2948 (2007)

    Article  CAS  Google Scholar 

  72. Yachdav, G. et al. PredictProtein – an open resource for online prediction of protein structural and functional features. Nucleic Acids Res. 42, W337–W343 (2014)

    Article  CAS  Google Scholar 

  73. Bernsel, A., Viklund, H., Hennerdal, A. & Elofsson, A. TOPCONS: consensus prediction of membrane protein topology. Nucleic Acids Res. 37, W465–W468 (2009)

    Article  CAS  Google Scholar 

  74. Sonnhammer, E. L., von Heijne, G. & Krogh, A. A hidden Markov model for predicting transmembrane helices in protein sequences. Proc. Int. Conf. Intell. Syst. Mol. Biol. 6, 175–182 (1998)

    CAS  PubMed  Google Scholar 

  75. Tusnády, G. E. & Simon, I. The HMMTOP transmembrane topology prediction server. Bioinformatics 17, 849–850 (2001)

    Article  Google Scholar 

  76. Käll, L., Krogh, A. & Sonnhammer, E. L. Advantages of combined transmembrane topology and signal peptide prediction – the Phobius web server. Nucleic Acids Res. 35, W429–W432 (2007)

    Article  Google Scholar 

Download references

Acknowledgements

We thank H. Yu and J. Chai for discussion and proofreading of the manuscript. We thank the staff at beamline BL17U of the Shanghai Synchrotron Radiation Facility (SSRF) and beamline 3W1A of the Beijing Synchrotron Radiation Facility (BSRF) for their assistance in data collection. K. Wu and H. Wang are acknowledged for technique help. We also thank the National Center for Protein Sciences (Beijing, China) for technical support with cryo-electron microscopy data collection and for computation resources. This work was supported by grants from the Ministry of Science and Technology (2012CB911101 and 2011CB910502 to M.Y., 2015CB910102 to B.X. and 2013CB910404 to N.G.), the National Natural Science Foundation of China (21532004, 31570733, 31030020 and 31170679 to M.Y., 31422016 to N.G. and 31422027 to B.X.) and the Ministry of Education (the Young Thousand Talent program to B.X.).

Author information

Authors and Affiliations

Authors

Contributions

M.Y. directed the study. J.G., M.C. and R.L. performed protein purification, detergent screening and crystallization. W.L. performed electron microscopy sample preparation, data collection and structural determination with N.L; Q.Z. was responsible for molecular cloning (with P.Z.), protein purification, detergent screening and biochemical and confocal imaging studies. N.G. directed electron microscopy studies and wrote part of the manuscript. B.X. initiated the project and directed molecular cloning, protein expression and purification and wrote most of the manuscript. All authors contributed to discussion of the data and editing of the manuscript.

Corresponding authors

Correspondence to Ning Gao, Bailong Xiao or Maojun Yang.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Extended data figures and tables

Extended Data Figure 1 Biochemical characterization of the recombinant protein of Piezo1–pp–GST.

a, A representative trace of gel filtration chromatography of the Piezo1–pp–GST protein. b, Protein samples of the indicated fractions were subjected to SDS–PAGE and Coomassie blue staining. Fractions of 8.0 ml and 8.5 ml (elution volume) were used for the negative-staining electron microscopy and native gel analyses, respectively.

Extended Data Figure 2 Negative-staining electron microscopy examination of Piezo1 in different detergents.

a, A representative micrograph of negatively stained Piezo1 purified with C12E10. b, 2D class averages of Piezo1 particles (C12E10). c, A representative micrograph of negatively stained Piezo1 purified with C12E8. d, 2D class averages of Piezo1 particles (C12E8). e, A representative micrograph of negatively stained Piezo1, with amphipol A8-35 as detergent. f, 2D class averages of Piezo1 particles (amphipol A8-35).

Extended Data Figure 3 Initial model of Piezo1 generated from the random conical tilt method and validation of the model using cryo-electron microscopy data from a TF20 microscope.

a, b, Representative micrographs of negatively stained Piezo1 in C12E10 collected in random conical tilt (RCT) pairs (a, untilted and b, 50° tilted). c, Top view of an RCT reconstruction, showing an overall threefold symmetry for the Piezo1 complex, is shown on the left. The right-hand side shows the top view of the refined model, obtained by a structural refinement of all particles from untilted micrographs. d, e, Model validation was performed by refinement of cryo-electron microscopy particles collected with TF20, with a Gaussian ball (d) or the RCT model (e) as initial reference. The 3D volumes are shown in top, side and bottom views. During the refinement, both the symmetry-free (C1) and symmetry-imposed (C3) reconstructions were tested. Note that some of these reconstructions have incorrect handedness.

Extended Data Figure 4 Representative raw particles of Piezo1 collected with the Titan Krios electron microscope fitted with a K2 electron detector.

A collection of raw particles of Piezo1 (eluted with C12E10), collected with Titan Krios (300 kV).

Extended Data Figure 5 Workflow of 3D classification of Piezo1 particles.

a, The schematic diagram of a series of 3D classification procedures with RELION is shown (also see Methods). After several rounds of 2D classification, the remaining 120,000 particles were subjected to three rounds of 3D classification without imposing any symmetry. A final set of particles (class 4 after the second round of 3D classification), with its reconstruction best matching threefold symmetry, was subjected to 3D refinement (C3 imposed). Notably, further 3D classification of this class resulted in generally similar structures (vertically arranged panels) without detectable improvement of conformational homogeneity. A top view of the soft mask used in structural refinement is also shown (yellow). b, Distribution of particle orientations in the last iteration of the refinement. c, Gold-standard Fourier shell correlation (FSC) curves of the final density map. The FSC curves were calculated with (red) or without (blue) the application of a soft mask to the two half-set maps. The final FSC curve (red) was corrected for the soft-mask-induced effect. Reported resolutions were based on FSC = 0.143 criteria.

Extended Data Figure 6 The trimeric CEDs form the cap domain of Piezo1.

a, A representative micrograph of negatively stained Piezo1(ΔCED) in C12E10. b, 2D class averages of negatively stained Piezo1(ΔCED) particles. It is evident that the central cap domain is absent from these average images. c, Sequence alignment of the CED region of Piezo1 from Mus musculus and Caenorhabditis elegans. Identical residues are highlighted in blue. Secondary structures are indicated by cartoons above the primary sequence. Sequence alignment was performed using Clustal W2 (ref. 71). d, Structure alignment of the trimeric CED of Piezo1 from M. musculus and C. elegans. The three CEDs are coloured in purple, cyan and green, respectively. The CED of C. elegans is coloured in orange. e, A representative trace of gel filtration of the CED of Piezo1. The molecular weights are labelled. Protein samples of the indicated fractions were subjected to SDS–PAGE and Coomassie blue staining (bottom). f, Transparent surface representation of the segmented density map of the cap, superimposed with the trimeric CED crystal structure. The trimeric CEDs are coloured as in d.

Extended Data Figure 7 Local resolution map of the final density map.

a, The final 3D density map of Piezo1 is coloured according to the local resolutions estimated by the software of ResMap. The density map is shown in three different views (top, bottom and side, respectively). b, The final 3D density map (transparent) is superimposed with a poly-alanine model and the crystal structure of the trimeric CED. Three protomers are coloured cyan, purple and green, respectively.

Extended Data Figure 8 Density connections between the transmembrane helices and between the helices in the compact CTD.

a, Alanine models of five representative pairs of transmembrane helices are displayed with their densities (mesh) superimposed. The transmembrane region is highlighted by a light purple shade with the intracellular and extracellular sides indicated. b, An alanine model of the anchor motif with its density superimposed (mesh). Four helices (α1anchor–α4anchor) connecting PH1 and OH are labelled. The transmembrane region is highlighted by a light purple shade with the intracellular and extracellular sides indicated. c, An alanine model of the last four helices (α1CTD–α4CTD) of the trimeric CTD, superimposed with the density of the CTD (mesh).

Extended Data Figure 9 Secondary structure analyses of the C-terminal segments of Piezo1 proteins from different species.

Sequence alignment of the C-terminal regions of Piezo1 from different species. The alignment was performed using Clustal W2 (ref. 71). The anchor motif and the CTD are highlighted in green and pink, respectively. For clarification, the sequences of CEDs were omitted and are indicated by red dashed lines. Secondary structures (α-helices) predicted with PredictProtein72 are shown as black lines. Transmembrane segments were predicted using multiple web servers including Topcons73 (green lines), TMHMM2 (ref. 74) (blue lines), HMMTOP75 (orange lines) and Phobius76 (pink lines).

Extended Data Table 1 Statistics of data collection and structure refinement.

PowerPoint slides

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ge, J., Li, W., Zhao, Q. et al. Architecture of the mammalian mechanosensitive Piezo1 channel. Nature 527, 64–69 (2015). https://doi.org/10.1038/nature15247

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nature15247

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