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:

The role of lipids in mechanosensation

Abstract

The ability of proteins to sense membrane tension is pervasive in biology. A higher-resolution structure of the Escherichia coli small-conductance mechanosensitive channel MscS identifies alkyl chains inside pockets formed by the transmembrane helices (TMs). Purified MscS contains E. coli lipids, and fluorescence quenching demonstrates that phospholipid acyl chains exchange between bilayer and TM pockets. Molecular dynamics and biophysical analyses show that the volume of the pockets and thus the number of lipid acyl chains within them decreases upon channel opening. Phospholipids with one acyl chain per head group (lysolipids) displace normal phospholipids (with two acyl chains) from MscS pockets and trigger channel opening. We propose that the extent of acyl-chain interdigitation in these pockets determines the conformation of MscS. When interdigitation is perturbed by increased membrane tension or by lysolipids, the closed state becomes unstable, and the channel gates.

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: Pockets formed between the TM helices; both pockets and helices change upon gating.
Figure 2: Lipids pack in the pockets created by the TM helices.
Figure 3: Lipid exchange between the pockets and the bilayer.
Figure 4: The conformational state of MscS can be altered by perturbing the interactions between the phospholipid and the protein.
Figure 5: A model for mechanosensation.

Similar content being viewed by others

Accession codes

Primary accessions

Protein Data Bank

Referenced accessions

Protein Data Bank

References

  1. Pivetti, C.D. et al. Two families of mechanosensitive channel proteins. Microbiol. Mol. Biol. Rev. 67, 66–85 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Kung, C. A possible unifying principle for mechanosensation. Nature 436, 647–654 (2005).

    Article  CAS  PubMed  Google Scholar 

  3. Arnadóttir, J. & Chalfie, M. Eukaryotic mechanosensitive channels. Annu. Rev. Biophys. 39, 111–137 (2010).

    Article  PubMed  CAS  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  5. Naismith, J.H. & Booth, I.R. Bacterial mechanosensitive channels–MscS: evolution’s solution to creating sensitivity in function. Annu. Rev. Biophys. 41, 157–177 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Haswell, E.S., Phillips, R. & Rees, D.C. Mechanosensitive channels: what can they do and how do they do it? Structure 19, 1356–1369 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Martinac, B. Bacterial mechanosensitive channels as a paradigm for mechanosensory transduction. Cell. Physiol. Biochem. 28, 1051–1060 (2011).

    Article  CAS  PubMed  Google Scholar 

  8. Levina, N. et al. Protection of Escherichia coli cells against extreme turgor by activation of MscS and MscL mechanosensitive channels: identification of genes required for MscS activity. EMBO J. 18, 1730–1737 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Reuter, M. et al. Mechanosensitive channels and bacterial cell wall integrity: does life end with a bang or a whimper? J. R. Soc. Interface 11, 20130850 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  10. Booth, I.R. Bacterial mechanosensitive channels: progress towards an understanding of their roles in cell physiology. Curr. Opin. Microbiol. 18, 16–22 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Shapovalov, G. & Lester, H.A. Gating transitions in bacterial ion channels measured at 3 microns resolution. J. Gen. Physiol. 124, 151–161 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Edwards, M.D. et al. Pivotal role of the glycine-rich TM3 helix in gating the MscS mechanosensitive channel. Nat. Struct. Mol. Biol. 12, 113–119 (2005).

    Article  CAS  PubMed  Google Scholar 

  13. Anishkin, A. & Sukharev, S. State-stabilizing interactions in bacterial mechanosensitive channel gating and adaptation. J. Biol. Chem. 284, 19153–19157 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Cox, C.D. et al. Selectivity mechanism of the mechanosensitive channel MscS revealed by probing channel subconducting states. Nat. Commun. 4, 2137 (2013).

    Article  CAS  PubMed  Google Scholar 

  15. Perozo, E., Cortes, D.M., Sompornpisut, P., Kloda, A. & Martinac, B. Open channel structure of MscL and the gating mechanism of mechanosensitive channels. Nature 418, 942–948 (2002).

    Article  CAS  PubMed  Google Scholar 

  16. Vásquez, V., Sotomayor, M., Cordero-Morales, J., Schulten, K. & Perozo, E. A structural mechanism for MscS gating in lipid bilayers. Science 321, 1210–1214 (2008).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  17. Sukharev, S.I., Martinac, B., Arshavsky, V.Y. & Kung, C. Two types of mechanosensitive channels in the Escherichia coli cell envelope: solubilization and functional reconstitution. Biophys. J. 65, 177–183 (1993).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  19. Wang, W. et al. The structure of an open form of an E. coli mechanosensitive channel at 3.45 A resolution. Science 321, 1179–1183 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. 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  CAS  PubMed  Google Scholar 

  21. Anishkin, A. & Sukharev, S. Water dynamics and dewetting transitions in the small mechanosensitive channel MscS. Biophys. J. 86, 2883–2895 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Akitake, B., Anishkin, A., Liu, N. & Sukharev, S. Straightening and sequential buckling of the pore-lining helices define the gating cycle of MscS. Nat. Struct. Mol. Biol. 14, 1141–1149 (2007).

    Article  CAS  PubMed  Google Scholar 

  23. Pliotas, C. et al. Conformational state of the MscS mechanosensitive channel in solution revealed by pulsed electron-electron double resonance (PELDOR) spectroscopy. Proc. Natl. Acad. Sci. USA 109, E2675–E2682 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Ward, R. et al. Probing the structure of the mechanosensitive channel of small conductance in lipid bilayers with pulsed electron-electron double resonance. Biophys. J. 106, 834–842 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Lai, J.Y., Poon, Y.S., Kaiser, J.T. & Rees, D.C. Open and shut: crystal structures of the dodecylmaltoside solubilized mechanosensitive channel of small conductance from Escherichia coli and Helicobacter pylori at 4.4 A and 4.1 A resolutions. Protein Sci. 22, 502–509 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Zhang, X. et al. Structure and molecular mechanism of an anion-selective mechanosensitive channel of small conductance. Proc. Natl. Acad. Sci. USA 109, 18180–18185 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Miller, A.N. & Long, S.B. Crystal structure of the human two-pore domain potassium channel K2P1. Science 335, 432–436 (2012).

    Article  CAS  PubMed  Google Scholar 

  29. Perozo, E. & Rees, D.C. Structure and mechanism in prokaryotic mechanosensitive channels. Curr. Opin. Struct. Biol. 13, 432–442 (2003).

    Article  CAS  PubMed  Google Scholar 

  30. Moe, P. & Blount, P. Assessment of potential stimuli for mechano-dependent gating of MscL: effects of pressure, tension, and lipid headgroups. Biochemistry 44, 12239–12244 (2005).

    Article  CAS  PubMed  Google Scholar 

  31. Schmidt, D., del Marmol, J. & MacKinnon, R. Mechanistic basis for low threshold mechanosensitivity in voltage-dependent K+ channels. Proc. Natl. Acad. Sci. USA 109, 10352–10357 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Brohawn, S.G., Su, Z. & Mackinnon, R. Mechanosensitivity is mediated directly by the lipid membrane in TRAAK and TREK1 K+ channels. Proc. Natl. Acad. Sci. USA 111, 3614–3619 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Anishkin, A. & Kung, C. Stiffened lipid platforms at molecular force foci. Proc. Natl. Acad. Sci. USA 110, 4886–4892 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Teng, J., Loukin, S., Anishkin, A. & Kung, C. The force-from-lipid (FFL) principle of mechanosensitivity, at large and in elements. Pflugers Arch. 467, 27–37 (2015).

    Article  CAS  PubMed  Google Scholar 

  35. Sukharev, S.I., Sigurdson, W.J., Kung, C. & Sachs, F. Energetic and spatial parameters for gating of the bacterial large conductance mechanosensitive channel, MscL. J. Gen. Physiol. 113, 525–540 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Sukharev, S. Purification of the small mechanosensitive channel of Escherichia coli (MscS): the subunit structure, conduction, and gating characteristics in liposomes. Biophys. J. 83, 290–298 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Brohawn, S.G., Campbell, E.B. & MacKinnon, R. Physical mechanism for gating and mechanosensitivity of the human TRAAK K+ channel. Nature 516, 126–130 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Branigan, E., Pliotas, C., Hagelueken, G. & Naismith, J.H. Quantification of free cysteines in membrane and soluble proteins using a fluorescent dye and thermal unfolding. Nat. Protoc. 8, 2090–2097 (2013).

    Article  CAS  PubMed  Google Scholar 

  39. Palsdottir, H. & Hunte, C. Lipids in membrane protein structures. Biochim. Biophys. Acta 1666, 2–18 (2004).

    Article  CAS  PubMed  Google Scholar 

  40. Laganowsky, A. et al. Membrane proteins bind lipids selectively to modulate their structure and function. Nature 510, 172–175 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Oursel, D. et al. Lipid composition of membranes of Escherichia coli by liquid chromatography/tandem mass spectrometry using negative electrospray ionization. Rapid Commun. Mass Spectrom. 21, 1721–1728 (2007).

    Article  CAS  PubMed  Google Scholar 

  42. Rasmussen, T. et al. Tryptophan in the pore of the mechanosensitive channel MscS: assessment of pore conformations by fluorescence spectroscopy. J. Biol. Chem. 285, 5377–5384 (2010).

    Article  CAS  PubMed  Google Scholar 

  43. Phillips, R., Ursell, T., Wiggins, P. & Sens, P. Emerging roles for lipids in shaping membrane-protein function. Nature 459, 379–385 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Sotomayor, M. & Schulten, K. Molecular dynamics study of gating in the mechanosensitive channel of small conductance MscS. Biophys. J. 87, 3050–3065 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Rasmussen, A. et al. The role of tryptophan residues in the function and stability of the mechanosensitive channel MscS from Escherichia coli. Biochemistry 46, 10899–10908 (2007).

    Article  CAS  PubMed  Google Scholar 

  46. Carney, J. et al. Fluorescence quenching methods to study lipid-protein interactions. Curr. Protoc. Protein Sci. 45, 19.12 (2006).

    Google Scholar 

  47. Racker, E., Chien, T.F. & Kandrach, A. A cholate-dilution procedure for the reconstitution of the Ca. pump, 32Pi–ATP exchange, and oxidative phosphorylation. FEBS Lett. 57, 14–18 (1975).

    Article  CAS  PubMed  Google Scholar 

  48. Wiener, M.C. & White, S.H. Transbilayer distribution of bromine in fluid bilayers containing a specifically brominated analogue of dioleoylphosphatidylcholine. Biochemistry 30, 6997–7008 (1991).

    Article  CAS  PubMed  Google Scholar 

  49. Knol, J., Sjollema, K. & Poolman, B. Detergent-mediated reconstitution of membrane proteins. Biochemistry 37, 16410–16415 (1998).

    Article  CAS  PubMed  Google Scholar 

  50. Nomura, T. et al. Differential effects of lipids and lyso-lipids on the mechanosensitivity of the mechanosensitive channels MscL and MscS. Proc. Natl. Acad. Sci. USA 109, 8770–8775 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Anishkin, A., Kamaraju, K. & Sukharev, S. Mechanosensitive channel MscS in the open state: modeling of the transition, explicit simulations, and experimental measurements of conductance. J. Gen. Physiol. 132, 67–83 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Martinac, B., Buechner, M., Delcour, A.H., Adler, J. & Kung, C. Pressure-sensitive ion channel in Escherichia coli. Proc. Natl. Acad. Sci. USA 84, 2297–2301 (1987).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Batiza, A.F., Kuo, M.M., Yoshimura, K. & Kung, C. Gating the bacterial mechanosensitive channel MscL in vivo. Proc. Natl. Acad. Sci. USA 99, 5643–5648 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Marsh, D. & Horvath, L.I. Structure, dynamics and composition of the lipid-protein interface: perspectives from spin-labelling. Biochim. Biophys. Acta 1376, 267–296 (1998).

    Article  CAS  PubMed  Google Scholar 

  55. Gullingsrud, J. & Schulten, K. Lipid bilayer pressure profiles and mechanosensitive channel gating. Biophys. J. 86, 3496–3509 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Lee, A.G. Biological membranes: the importance of molecular detail. Trends Biochem. Sci. 36, 493–500 (2011).

    Article  CAS  PubMed  Google Scholar 

  57. Iscla, I. & Blount, P. Sensing and responding to membrane tension: the bacterial MscL channel as a model system. Biophys. J. 103, 169–174 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Perozo, E., Kloda, A., Cortes, D.M. & Martinac, B. Physical principles underlying the transduction of bilayer deformation forces during mechanosensitive channel gating. Nat. Struct. Biol. 9, 696–703 (2002).

    Article  CAS  PubMed  Google Scholar 

  59. Tanaka, K., Caaveiro, J.M., Morante, K., Gonzalez-Manas, J.M. & Tsumoto, K. Structural basis for self-assembly of a cytolytic pore lined by protein and lipid. Nat. Commun. 6, 6337 (2015).

    Article  PubMed  CAS  Google Scholar 

  60. Anishkin, A., Loukin, S.H., Teng, J. & Kung, C. Feeling the hidden mechanical forces in lipid bilayer is an original sense. Proc. Natl. Acad. Sci. USA 111, 7898–7905 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Leslie, A.G.W. Recent changes to the MOSFLM package for processing film and image plate data. Joint CCP4 ESF-EACBM Newsl. Protein Crystallogr. 26 (1992).

  62. CCP4. The CCP4 suite: programs for protein crystallography. Acta Crystallogr. D Biol. Crystallogr. 50, 760–763 (1994).

  63. Diederichs, K. & Karplus, P.A. Better models by discarding data? Acta Crystallogr. D Biol. Crystallogr. 69, 1215–1222 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Karplus, P.A. & Diederichs, K. Linking crystallographic model and data quality. Science 336, 1030–1033 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Joosten, R.P., Long, F., Murshudov, G.N. & Perrakis, A. The PDB_REDO server for macromolecular structure model optimization. IUCrJ 1, 213–220 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Binkowski, T.A., Naghibzadeh, S. & Liang, J. CASTp: computed atlas of surface topography of proteins. Nucleic Acids Res. 31, 3352–3355 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Pellegrini-Calace, M., Maiwald, T. & Thornton, J.M. PoreWalker: a novel tool for the identification and characterization of channels in transmembrane proteins from their three-dimensional structure. PLOS Comput. Biol. 5, e1000440 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  68. Humphrey, W., Dalke, A. & Schulten, K. VMD: visual molecular dynamics. J. Mol. Graph. 14, 33–38 (1996).

    Article  CAS  PubMed  Google Scholar 

  69. Booth, I.R. et al. Physiological analysis of bacterial mechanosensitive channels. Methods Enzymol. 428, 47–61 (2007).

    Article  PubMed  Google Scholar 

  70. Ladokhin, A.S., Jayasinghe, S. & White, S.H. How to measure and analyze tryptophan fluorescence in membranes properly, and why bother? Anal. Biochem. 285, 235–245 (2000).

    Article  CAS  PubMed  Google Scholar 

  71. Schumann, U. et al. YbdG in Escherichia coli is a threshold-setting mechanosensitive channel with MscM activity. Proc. Natl. Acad. Sci. USA 107, 12664–12669 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  72. Sobott, F., Hernandez, H., McCammon, M.G., Tito, M.A. & Robinson, C.V. A tandem mass spectrometer for improved transmission and analysis of large macromolecular assemblies. Anal. Chem. 74, 1402–1407 (2002).

    Article  CAS  PubMed  Google Scholar 

  73. Hernández, H. & Robinson, C.V. Determining the stoichiometry and interactions of macromolecular assemblies from mass spectrometry. Nat. Protoc. 2, 715–726 (2007).

    Article  PubMed  CAS  Google Scholar 

  74. Fyffe, S.A. et al. Reevaluation of the PPAR-beta/delta ligand binding domain model reveals why it exhibits the activated form. Mol. Cell 21, 1–2 (2006).

    Article  CAS  PubMed  Google Scholar 

  75. Bligh, E.G. & Dyer, W.J. A rapid method of total lipid extraction and purification. Can. J. Biochem. Physiol. 37, 911–917 (1959).

    Article  CAS  PubMed  Google Scholar 

  76. Richmond, G.S. et al. Lipidomic analysis of bloodstream and procyclic form Trypanosoma brucei. Parasitology 137, 1357–1392 (2010).

    Article  CAS  PubMed  Google Scholar 

  77. Monticelli, L. et al. The MARTINI coarse-grained force field: extension to proteins. J. Chem. Theory Comput. 4, 819–834 (2008).

    Article  CAS  PubMed  Google Scholar 

  78. Scott, K.A. et al. Coarse-grained MD simulations of membrane protein-bilayer self-assembly. Structure 16, 621–630 (2008).

    Article  CAS  PubMed  Google Scholar 

  79. Hess, B., Kutzner, C., van der Spoel, D. & Lindahl, E. Gromacs 4, load-balanced, and scalable molecular simulation: algorithms for highly efficient. J. Chem. Theory Comput. 4, 435–447 (2008).

    Article  CAS  PubMed  Google Scholar 

  80. Stansfeld, P.J. & Sansom, M.S.P. From coarse grained to atomistic: a serial multiscale approach to membrane protein simulations. J. Chem. Theory Comput. 7, 1157–1166 (2011).

    Article  CAS  PubMed  Google Scholar 

  81. Scott, W.R.P. et al. The GROMOS biomolecular simulation program package. J. Phys. Chem. A 103, 3596–3607 (1999).

    Article  CAS  Google Scholar 

  82. Hermans, J., Berendsen, H.J.C., Van Gunsteren, W.F. & Postma, J.P.M. A consistent empirical potential for water–protein interactions. Biopolymers 23, 1513–1518 (1984).

    Article  CAS  Google Scholar 

  83. Parrinello, M. & Rahman, A. Polymorphic transitions in single crystals: a new molecular dynamics method. J. Appl. Phys. 52, 7182–7190 (1981).

    Article  CAS  Google Scholar 

  84. Berendsen, H.J.C., Postma, J.P.M., van Gunsteren, W.F., DiNola, A.R.H.J. & Haak, J.R. Molecular dynamics with coupling to an external bath. J. Chem. Phys. 81, 3684–3690 (1984).

    Article  CAS  Google Scholar 

  85. Hess, B., Bekker, H., Berendsen, H.J.C. & Fraaije, J.G.E.M. LINCS: a linear constraint solver for molecular simulations. J. Comput. Chem. 18, 1463–1472 (1997).

    Article  CAS  Google Scholar 

  86. Darden, T., York, D. & Pedersen, L. Particle mesh Ewald: an N· log (N) method for Ewald sums in large systems. J. Chem. Phys. 98, 10089–10092 (1993).

    Article  CAS  Google Scholar 

  87. Montal, M. & Mueller, P. Formation of bimolecular membranes from lipid monolayers and a study of their electrical properties. Proc. Natl. Acad. Sci. USA 69, 3561–3566 (1972).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Kramp, W., Pieroni, G., Pinckard, R.N. & Hanahan, D.J. Observations on the critical micellar concentration of 1-O-alkyl-2-acetyl-sn-glycero-3-phosphocholine and a series of its homologs and analogs. Chem. Phys. Lipids 35, 49–62 (1984).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

This work was supported by Wellcome Trust grants WT092552MA (J.H.N. and I.R.B.), Senior Investigator Award WT100209MA (J.H.N.), 093228 (T.K.S.) and 092970 (M.S.P.S.), and Biotechnology and Biological Sciences Research Council grants BB/I019855/1 (M.S.P.S.), BB/H017917/1 (J.H.N. and I.R.B.) and BB/J009784/1 (H.B.). We acknowledge the Diamond Light Source for beam time. I.R.B. is supported as a Leverhulme Emeritus Fellow. J.H.N. is supported as a Royal Society Wolfson Merit Award holder and as a 1000 Talent Scholar at Sichuan University. A.C.E.D. was supported by an Engineering and Physical Sciences Research Council Systems Biology Doctoral Training Centre student fellowship. We thank R. Phillips, A. Lee and S. Conway for helpful discussions.

Author information

Authors and Affiliations

Authors

Contributions

C.P. purified and spin-labeled MscS for lipid analysis, single-molecule analysis and crystallization; obtained and analyzed the new crystal structure; and participated in the single-molecule and lipid-analysis experiments. T.R. purified MscS and carried out and analyzed the fluorescence studies including synthesis of brominated lipids. A.C.E.D. wrote the MD-analysis software and performed and analyzed the simulations. K.R.M. performed and analyzed single-molecule experiments. A.R. made mutants of MscS and performed osmotic downshock assays. T.B. assisted in fluorescence experiments. T.K.S. performed lipidomic mass spectrometry. C.V.R. and J.G. carried out native mass spectrometry. P.M. performed TLC experiments. S.M., H.B., M.S.P.S., I.R.B. and J.H.N. conceived and supervised the study. All authors wrote, reviewed and approved the paper.

Corresponding author

Correspondence to James H Naismith.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Additional structural data.

A: The spin label on D67R1 is clearly visible.

B: Mean spin-to-spin x-ray (measured at O of the NO group) and PELDOR distances, these are taken from the original paper1. The vectors are from subunit A to B (D1-2), A to C, (D1-3) and A to D (D1-4). The D1-3 spin-to-spin distance (46.7 ± 0.2) Å differs only by 1Å from the PELDOR value (Fig S1B), whilst the D1-4 distance (58.3 ± 0.1) Å differs by 1.5 Å1. The increased deviation in the higher order vectors may be measurement error or be a manifestation of multispin effects2.

C: Electron density 2Fo-Fc at 1σ for the acyl chains.

D: The internal pore diameter was measured using Porewalker software3 in 3 Å steps starting from the resolved Y27 of the 2OAU structure (closed) and L23 of the open structure in this study with D67R1 mutated in silico to native. With this approach the structures have essentially the same length of 105 Å (Z-axis) using residues 23-278 (open) and 27-278 (closed). This is because the change in the tilt of the helices, means one residue in one structure is not in the same lateral plane as the same residue in the other structure. The Z-axis can be split into 35 steps along and diameter of the pore measured at each point, the pore volume can be integrated along the Z-axis. The cumulative pore volume difference of the structures for 45 Å starting at the periplasmic top of the structure along pore Z-axis (includes the whole of the TM domain) is 10,389 Å3. To assess the change in pocket volumes the structures (with D67R1 again mutated in silico to native) were analyzed with the CASTp server4 with probes of radii 1.5, 2.0, 2.5 and 3.0 Å. At all probe radii, the pockets in the closed structure are identified as a single volume, thus the volume of each pocket is estimated at 1/7 of this volume. The pocket volumes are 4500 Å3 (3 Å probe), 5161 Å3 (2.5 Å probe), 5629 Å3 (2 Å probe) and 6262 Å3 (1.5 Å probe); the corresponding central cavity volumes are 60,895, 62, 727, 64,965 and 68,344 Å3 respectively (for the two larger probes the central cavity is split by the hydrophobic seal at L105). In the open structure analyzed with a probe of 3.0 Å the central cavity is separated from the pockets with a volume of 65,276 Å3 but the pockets are split into sub pockets so were not analyzed at this radius. With a probe of 2.5 Å radius, three pockets and the central cavity are considered a single volume of 78,386 Å3, the remaining four pockets are identified individually, the largest and most complete was 3,892 Å3. Assuming this is correct pocket would give a central cavity volume of 78,386 – 3 x 3,892 = 66,710 Å3 in close agreement with the 3.0 Å probe. Thus with a 2.5 Å probe we estimate the pocket volume is reduced by 1,269 Å3. With a radius of 2.5 Å probe, all seven pockets and the central cavity are considered as one volume of 98,140 Å3. To estimate the reduction in the pocket volume, we combine the volumes from the closed structure for the central cavity and pockets (64,965 + 39,403 = 104,368 Å3) then substrate the volume of the seven pockets and central cavity from the open structure (98,140 Å3), dividing the result by seven to yield reduction of 890 Å3 in volume per pocket. This approach estimates the central cavity in the open structure as 64,967 Å3 similar to the other probes. Repeating this approach with the 1.5 Å probe reveals a single volume of 107,457 Å3 which using the approach detailed for the 2.0 Å probe yields a reduction in each pocket volume of 674 Å3 upon opening and a central cavity of 68,343 Å3. Discussion of the cross sectional area is given below.

Supplementary Figure 2 Additional lipid analysis.

A: Mass spectrum of the total lipid from MJF612 E. coli strain expressing MscS.

B: DM-solubilized MscS shows the same shift in lipid profile relative to the total lipid profile (Fig S2A) as DDM, suggesting it is not a simple property of the detergent.

C: Mass spectrum of the total lipid from MJF612 E. coli strain which are not expressing MscS. The lipid profile is not altered by presence or absence of MscS.

D: DDM-solubilized membrane protein (uniprot G1FG65) shows the same profile as the E. coli cell.

Supplementary Figure 3 Additional molecular dynamics analysis.

A: Representative view of lipid interaction with a single subunit of the closed MscS structure. Full heptameric structure of closed MscS is shown. A section of the TM domain and the lipids within the pocket is enlarged to show lipid interactions with the polypeptide chain of a single monomer (salmon). The TM3ab helices of the adjacent subunit are shown in yellow, with the hydrophobic gate residues in orange (in space-filling representation). Phospholipids are shown as space-filling representations, with the colours set to distinguish the acyl tails (turquoise) from the headgroups. Note the orientation of the lipids: headgroups pointing downwards with respect to the tails signify a lipid in the cytosolic bilayer leaflet. The opposite orientation signifies a lipid in the periplasmic leaflet (c.f. Fig 3. A, B).

B: Representative view of lipid interaction with a single subunit of the open MscS structure. The heptameric structure of open MscS is shown, with a section of the TM domain and the lipids within the pocket enlarged to show lipid interactions with the polypeptide chain of a single monomer (salmon). The hydrophobic gate residues are shown in orange space-filling representations. Phospholipid representations and significance of orientation as in Fig S3A.

C: Protein proximity to lipid during AT-MD is correlated with proximity to proposed lipid in the crystal structure. Average separation between a single subunit from the new open structure and persistent lipid in the lower region of the pocket during 50-100 ns AT-MD (orange, error bars show one standard deviation) compared to the separation between the new crystal structure and electron density, modelled as carbon (blue).

D: Lipid contacts to the closed and open structures differ significantly. Data were collected from 50 to 100 ns during each AT-MD simulation, and averaged across 5 simulations per state. Statistically different numbers of lipids within 6 Å between closed and open states are shown per residue as heat-mapped residues on the polypeptide chain of the TM domain from a single subunit in the closed structure (p < 0.01). A dark blue residue indicates more contacts to the closed state (maximum 1.7 lipids more) whereas red indicates that the open state has more lipid contacts to the residue (maximum 1.4 lipids more). A white residue indicates no significant difference in contact number between the two states.

E: ‘Persistent’ (those that remain with 6 Å contact of residues in the pockets during the length simulation) and total (including those that are exchanged with the bilayer during the simulation). Dark bars are those in the closed, white in the open structure.

F: (i) The headgroups of the persisting POPE lipids shown in (iii) were analysed for contact preferences to MscS residues. Headgroups make most contacts to charges along the TM1/TM2 helices. ‘Persistent’ lipid configurations in the lower pocket region in the closed (ii) and open (iii) state of MscS. Lipids that make contact (6 Å cutoff) to the lower pocket region throughout 50 to 100 ns AT-MD are displayed in snapshots from the end of the respective simulations. Lipid molecules in the open structure reach the hydrophobic seal (iii), whereas the lipids in the closed structure are prevented from reaching the seal (ii).

Supplementary Figure 4 Additional data on MscS mutants and the effect of LPC on MscS.

A: MscS tryptophan mutants transformed into MJF612 were exposed to an osmotic downshock of 0.5 M NaCl. Samples were uninduced (blue) or induced by 0.3 mM IPTG (orange). Survival of samples are shown relative to samples diluted into control medium. All survival experiments were performed using transformants of either MJF641 (ΔyggB, ΔmscL, ΔmscK ΔybdG, ΔybiO, ΔynaI, ΔyjeP) or MJF612 (ΔyggB, ΔmscL, ΔmscK, ΔybdG). First, cells were grown at 37 °C in Luria−Bertani (LB) medium, and both induced (0.3 mM IPTG added when OD650nm 0.2) and uninduced cultures were studied. The culture was adapted to high osmolarity by growth to an OD650nm of 0.3 in the presence of 0.5 M NaCl, and an osmotic downshock was then applied by a 1:20 dilution into LB medium (shock) or the medium containing 0.5 M NaCl (control). After 10 min incubation at 37 °C, 5 μL serial dilutions of these cultures were spread onto LB-agar plates in the presence (control) or absence (shock) of 0.5 M NaCl. The survival rates were then assessed by counting the number of colonies after incubation overnight at 37 °C. Data are reported as means ± standard deviation.

B: A western blot (anti His tag), 3 second exposure of blue native gels. The lanes correspond to:

1: A119W: reconstituted in DOPC, +BrLPC, then concentrated

2: A119W: after purification in DDM, +BrLPC

3: A119W: after purification in DDM

4: M47W: reconstituted in DOPC, +BrLPC, then concentrated

5: M47W: after purification in DDM, +BrLPC

6: M47W: after purification in DDM

The data show that brominated LPC 18:1 does not dissociate the heptamer in lipid bilayers but LPC (either 18:1 or 14:0) does dissociate MscS partially in detergent (consistent with gel filtration S4B).

C: When LPC is added to detergent (not reconstituted in bilayer) solubilized MscS (mole ratio 0.3 LPC), some dissociation of the MscS heptamer is observed, we only analysed the heptameric fraction.

D: Fragmentation of the lipid extract from the heptamer confirms the identity of the lipid in Fig 1E to be LPC 14:0 to be present by comparison to LPC 14:0 standard (shown below).

E: The same control protein as used in Supplementary Fig 2D incubated with LPC 14:0 using the same protocol as MscS does not show (unlike MscS) retention of LPC 14:0 in ESI-MS/MS.

F: MscS mutants transformed into MJF641 were induced by 0.3 mM IPTG and exposed to an osmotic downshock of 0.3 M NaCl. Samples with (orange) or without (blue) addition of 0.5 mM MTSSL during the shock phase are shown.

Supplementary Figure 5 Additional single-molecule data on MscS.

A: Typical current recordings of MscS channels after spontaneous opening at +20 mV, +50 mV, -20 mV and -50 mV. All-points amplitude histograms (right panel). Electrolyte: 200 mM KCl, 90 mM MgCl2, 10 mM CaCl2, 10 mM HEPES, pH 7.5. 70 % spontaneously opened channel were stable with higher open probability without any closures recorded at +50 mV, -50 mV, +20 mV and -20 mV.

B: Typical current recordings at +100 mV of spontaneously opened MscS channels. Upper: gating in a single channel manifested as fast flickering. Lower: gating events in three channels. Electrolyte: 200 mM KCl, 90 mM MgCl2, 10 mM CaCl2, 10 mM HEPES, pH 7.5. 30 % of spontaneously opened channel showed fast flickering activity only at higher voltages +100 mV and we have not observed this fast flickering spontaneous channel opening at lower voltages (+50 mV and +20 mV). The observation of multiple channels is an important control. If the conductance arises from a single molecule, then multiple insertions should give rise to conductance that is multiples thereof. It also eliminates concerns about protein heterogeneity, that is, it is MscS, not another protein, that gives rise to the conductance data in the presence of LPC. We did not explore further the voltage dependence of spontaneous opening as this was outside the scope of the study.

C: I-V curves obtained with single MscS channels in planar lipid bilayer measured at applied potential -100 mV to 100 mV. Electrolyte: 200 mM KCl, 90 mM MgCl2, 10 mM CaCl2, 10 mM HEPES, pH 7.5.

D: Selected ion current recordings at + 20 mV and + 50 mV of MscS channels reconstituted into a planar lipid membrane and activated by 10 µM LPC 14:0 (cis). In both cases, the activation of a second channel with identical conductance to the first can be observed, shown as an important control. All-points amplitude histograms (right panel). Electrolyte: 200 mM KCl, 90 mM MgCl2, 10 mM CaCl2, 10 mM HEPES, pH 7.5. The single channel is mostly open 100% of the time with a second channel that appears to exhibit a much smaller open probability was a very rare event but importantly it demonstrated a similar conductance (thus providing a control). We do not have enough observations to determine the open probability of the ‘second’ channel and this was outwith the scope of our study.

Supplementary Figure 6 Bromolipids.

1H NMR spectrum of 1,2-di-(9,10-dibromo)stearoyl-sn-glycero-3-phosphocholine. The sample was solved in CDCl3 and the spectrum recorded at room temperature on a 400 MHz Bruker Avance III spectrometer: CDCl3): δH 0.91 (6H, m, ω-CH3), 1.30 (40H, m, CH2), 1.61 (7H, m, β-CH2 and H2O), 1.87 (5H, m, trans CH2CHBr), 2.06 (3H, m, gauche CH2CHBr), 2.32 (4H, m, α-CH2), 3.47 (9H, s, N(CH3)3), 3.54 (4H, m, CHBr), 4.06 (2H, m, PO3CH2 (glycerol)), 4.16 (1H, m, CH2O (glycerol)), 4.22 (2H, m, CH2N), 4.39 (1H, m, CH2O (glycerol)), 4.48 (2H, m, CH2O (glycerol)), 5.24 (1H, m, CHO)5,6. The sample was also examined with positive ion electrospray mass spectroscopy in the presence of formic acid with an Agilent Technologies 6120 ESI-MS. Expected highest peak for (M+H)+ = 1106.3 m/z (observed 1106.2 m/z). The peak pattern was as expected for an isotope cluster with four bromine atoms. The highest intensity peaks cluster in the experimental spectrum can be explained with the (M-Br) fragment. The peak pattern was as expected for an isotope cluster with three bromine atoms. Isotope clusters were simulated with the program IsoPro 3.1 (Mike Senko; https://sites.google.com/site/isoproms/).

Supplementary information

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Pliotas, C., Dahl, A., Rasmussen, T. et al. The role of lipids in mechanosensation. Nat Struct Mol Biol 22, 991–998 (2015). https://doi.org/10.1038/nsmb.3120

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nsmb.3120

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