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:

A molecular mechanism for calcium-mediated synaptotagmin-triggered exocytosis

Abstract

The regulated exocytotic release of neurotransmitter and hormones is accomplished by a complex protein machinery whose core consists of SNARE proteins and the calcium sensor synaptotagmin-1. We propose a mechanism in which the lipid membrane is intimately involved in coupling calcium sensing to release. We found that fusion of dense core vesicles, derived from rat PC12 cells, was strongly linked to the angle between the cytoplasmic domain of the SNARE complex and the plane of the target membrane. We propose that, as this tilt angle increases, force is exerted on the SNARE transmembrane domains to drive the merger of the two bilayers. The tilt angle markedly increased following calcium-mediated binding of synaptotagmin to membranes, strongly depended on the surface electrostatics of the membrane, and was strictly coupled to the lipid order of the target membrane.

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

Fig. 1: Correlating conformational changes in SNARE proteins with fusion efficiency during Ca2+-triggered vesicle fusion.
Fig. 2: C2AB induces trans–cis conformational transition of neuronal SNARE complex.
Fig. 3: SNARE complex conformation and DCV fusion depend on the ratio of saturated and unsaturated fatty acids in the target membrane.
Fig. 4: Mutations in C2AB and SNAREs impair trans–cis conformational transition of the SNARE complex and DCV fusion.
Fig. 5: Model for Ca2+-triggered fusion mechanism.

Similar content being viewed by others

Data availability

Source data for Figs. 2, 3a,b,d,e, and 4 are available with the paper online. The custom code used to acquire and analyze the data, as well other data in this study, is available from the corresponding author upon reasonable request.

References

  1. Sabatini, B. L. & Regehr, W. G. Timing of neurotransmission at fast synapses in the mammalian brain. Nature 384, 170–172 (1996).

    Article  CAS  Google Scholar 

  2. Schneggenburger, R. & Neher, E. Intracellular calcium dependence of transmitter release rates at a fast central synapse. Nature 406, 889–893 (2000).

    Article  CAS  Google Scholar 

  3. Südhof, T. C. Neurotransmitter release: the last millisecond in the life of a synaptic vesicle. Neuron 80, 675–690 (2013).

    Article  Google Scholar 

  4. Südhof, T. C. A molecular machine for neurotransmitter release: synaptotagmin and beyond. Nat. Med. 19, 1227–1231 (2013).

    Article  Google Scholar 

  5. Jahn, R. & Fasshauer, D. Molecular machines governing exocytosis of synaptic vesicles. Nature 490, 201–207 (2012).

    Article  CAS  Google Scholar 

  6. Jahn, R. & Scheller, R. H. SNAREs—engines for membrane fusion. Nat. Rev. Mol. Cell Biol. 7, 631–643 (2006).

    Article  CAS  Google Scholar 

  7. Weber, T. et al. SNAREpins: minimal machinery for membrane fusion. Cell 92, 759–772 (1998).

    Article  CAS  Google Scholar 

  8. Geppert, M. et al. Synaptotagmin I: a major Ca2+ sensor for transmitter release at a central synapse. Cell 79, 717–727 (1994).

    Article  CAS  Google Scholar 

  9. Stein, A., Radhakrishnan, A., Riedel, D., Fasshauer, D. & Jahn, R. Synaptotagmin activates membrane fusion through a Ca2+-dependent trans interaction with phospholipids. Nat. Struct. Mol. Biol. 14, 904–911 (2007).

    Article  CAS  Google Scholar 

  10. Kiessling, V. et al. Rapid fusion of synaptic vesicles with reconstituted target SNARE membranes. Biophys. J. 104, 1950–1958 (2013).

    Article  CAS  Google Scholar 

  11. Kyoung, M., Zhang, Y., Diao, J., Chu, S. & Brunger, A. T. Studying calcium-triggered vesicle fusion in a single vesicle-vesicle content and lipid-mixing system. Nat. Protoc. 8, 1–16 (2013).

    Article  CAS  Google Scholar 

  12. Lee, H. K. et al. Dynamic Ca2+-dependent stimulation of vesicle fusion by membrane-anchored synaptotagmin 1. Science 328, 760–763 (2010).

    Article  CAS  Google Scholar 

  13. Tucker, W. C., Weber, T. & Chapman, E. R. Reconstitution of Ca2+-regulated membrane fusion by synaptotagmin and SNAREs. Science 304, 435–438 (2004).

    Article  CAS  Google Scholar 

  14. Malsam, J. et al. Complexin arrests a pool of docked vesicles for fast Ca2+-dependent release. EMBO J. 31, 3270–3281 (2012).

    Article  CAS  Google Scholar 

  15. Südhof, T. C. & Rothman, J. E. Membrane fusion: grappling with SNARE and SM proteins. Science 323, 474–477 (2009).

    Article  Google Scholar 

  16. Rizo, J. & Xu, J. The synaptic vesicle release machinery. Annu. Rev. Biophys. 44, 339–367 (2015).

    Article  CAS  Google Scholar 

  17. Ledesma, M. D., Martin, M. G. & Dotti, C. G. Lipid changes in the aged brain: effect on synaptic function and neuronal survival. Prog. Lipid Res. 51, 23–35 (2012).

    Article  CAS  Google Scholar 

  18. Ma, C., Su, L., Seven, A. B., Xu, Y. & Rizo, J. Reconstitution of the vital functions of Munc18 and Munc13 in neurotransmitter release. Science 339, 421–425 (2013).

    Article  CAS  Google Scholar 

  19. Kreutzberger, A. J. B. et al. Reconstitution of calcium-mediated exocytosis of dense-core vesicles. Sci. Adv. 3, e1603208 (2017).

    Article  Google Scholar 

  20. Liang, B., Dawidowski, D., Ellena, J. F., Tamm, L. K. & Cafiso, D. S. The SNARE motif of synaptobrevin exhibits an aqueous-interfacial partitioning that is modulated by membrane curvature. Biochemistry 53, 1485–1494 (2014).

    Article  CAS  Google Scholar 

  21. Liang, B., Kiessling, V. & Tamm, L. K. Prefusion structure of syntaxin-1A suggests pathway for folding into neuronal trans-SNARE complex fusion intermediate. Proc. Natl. Acad. Sci. USA 110, 19384–19389 (2013).

    Article  CAS  Google Scholar 

  22. Zdanowicz, R. et al. Complexin binding to membranes and acceptor t-SNAREs explains its clamping effect on fusion. Biophys. J. 113, 1235–1250 (2017).

    Article  CAS  Google Scholar 

  23. Gong, J. et al. C-terminal domain of mammalian complexin-1 localizes to highly curved membranes. Proc. Natl. Acad. Sci. USA 113, E7590–E7599 (2016).

    Article  CAS  Google Scholar 

  24. Dawidowski, D. & Cafiso, D. S. Allosteric control of syntaxin 1a by Munc18-1: characterization of the open and closed conformations of syntaxin. Biophys. J. 104, 1585–1594 (2013).

    Article  CAS  Google Scholar 

  25. Cremona, O. & De Camilli, P. Phosphoinositides in membrane traffic at the synapse. J. Cell. Sci. 114, 1041–1052 (2001).

    CAS  PubMed  Google Scholar 

  26. Holz, R. W. & Axelrod, D. Localization of phosphatidylinositol 4,5-P2 important in exocytosis and a quantitative analysis of chromaffin granule motion adjacent to the plasma membrane. Ann. NY Acad. Sci. 971, 232–243 (2002).

    Article  CAS  Google Scholar 

  27. Milosevic, I. et al. Plasmalemmal phosphatidylinositol-4,5-bisphosphate level regulates the releasable vesicle pool size in chromaffin cells. J. Neurosci. 25, 2557–2565 (2005).

    Article  CAS  Google Scholar 

  28. Martin, T. F. Role of PI(4,5)P2 in vesicle exocytosis and membrane fusion. Subcell. Biochem. 59, 111–130 (2012).

    Article  CAS  Google Scholar 

  29. Kiessling, V. & Tamm, L. K. Measuring distances in supported bilayers by fluorescence interference-contrast microscopy: polymer supports and SNARE proteins. Biophys. J. 84, 408–418 (2003).

    Article  CAS  Google Scholar 

  30. Lambacher, A. & Fromherz, P. Luminescence of dye moleculars on oxidized silicon and fluorescence interference contrast (FLIC) microscopy of biomembranes. J. Opt. Soc. Am. B 19, 1435–1453 (2002).

    Article  CAS  Google Scholar 

  31. Kreutzberger, A. J., Liang, B., Kiessling, V. & Tamm, L. K. Assembly and comparison of plasma membrane SNARE acceptor complexes. Biophys. J. 110, 2147–2150 (2016).

    Article  CAS  Google Scholar 

  32. Stein, A., Weber, G., Wahl, M. C. & Jahn, R. Helical extension of the neuronal SNARE complex into the membrane. Nature 460, 525–528 (2009).

    Article  CAS  Google Scholar 

  33. Lai, A. L., Tamm, L. K., Ellena, J. F. & Cafiso, D. S. Synaptotagmin 1 modulates lipid acyl chain order in lipid bilayers by demixing phosphatidylserine. J. Biol. Chem. 286, 25291–25300 (2011).

    Article  CAS  Google Scholar 

  34. Gaffaney, J. D., Dunning, F. M., Wang, Z., Hui, E. & Chapman, E. R. Synaptotagmin C2B domain regulates Ca2+-triggered fusion in vitro: critical residues revealed by scanning alanine mutagenesis. J. Biol. Chem. 283, 31763–31775 (2008).

    Article  CAS  Google Scholar 

  35. Rizo, J. & Rosenmund, C. Synaptic vesicle fusion. Nat. Struct. Mol. Biol. 15, 665–674 (2008).

    Article  CAS  Google Scholar 

  36. Pérez-Lara, Á. et al. PtdInsP2 and PtdSer cooperate to trap synaptotagmin-1 to the plasma membrane in the presence of calcium. eLife 5, e15886 (2016).

    Article  Google Scholar 

  37. Park, Y. et al. Synaptotagmin-1 binds to PIP2-containing membrane but not to SNAREs at physiological ionic strength. Nat. Struct. Mol. Biol. 22, 815–823 (2015).

    Article  CAS  Google Scholar 

  38. Zhou, Q. et al. Architecture of the synaptotagmin–SNARE machinery for neuronal exocytosis. Nature 525, 62–67 (2015).

    Article  CAS  Google Scholar 

  39. Wang, S., Li, Y. & Ma, C. Synaptotagmin-1 C2B domain interacts simultaneously with SNAREs and membranes to promote membrane fusion. eLife 5, e14211 (2016).

    Article  Google Scholar 

  40. McMahon, H. T., Kozlov, M. M. & Martens, S. Membrane curvature in synaptic vesicle fusion and beyond. Cell 140, 601–605 (2010).

    Article  CAS  Google Scholar 

  41. Hui, E., Johnson, C. P., Yao, J., Dunning, F. M. & Chapman, E. R. Synaptotagmin-mediated bending of the target membrane is a critical step in Ca2+-regulated fusion. Cell 138, 709–721 (2009).

    Article  CAS  Google Scholar 

  42. Wagner, M. L. & Tamm, L. K. Reconstituted syntaxin1a/SNAP25 interacts with negatively charged lipids as measured by lateral diffusion in planar supported bilayers. Biophys. J. 81, 266–275 (2001).

    Article  CAS  Google Scholar 

  43. van den Bogaart, G. et al. Membrane protein sequestering by ionic protein-lipid interactions. Nature 479, 552–555 (2011).

    Article  CAS  Google Scholar 

  44. Honigmann, A. et al. Phosphatidylinositol 4,5-bisphosphate clusters act as molecular beacons for vesicle recruitment. Nat. Struct. Mol. Biol. 20, 679–686 (2013).

    Article  CAS  Google Scholar 

  45. Simons, K. & Ikonen, E. Functional rafts in cell membranes. Nature 387, 569–572 (1997).

    Article  CAS  Google Scholar 

  46. Veatch, S. L. et al. Critical fluctuations in plasma membrane vesicles. ACS Chem. Biol. 3, 287–293 (2008).

    Article  CAS  Google Scholar 

  47. Wang, J. et al. Calcium sensitive ring-like oligomers formed by synaptotagmin. Proc. Natl. Acad. Sci. USA 111, 13966–13971 (2014).

    Article  CAS  Google Scholar 

  48. Shin, O. H. et al. Munc13 C2B domain is an activity-dependent Ca2+ regulator of synaptic exocytosis. Nat. Struct. Mol. Biol. 17, 280–288 (2010).

    Article  CAS  Google Scholar 

  49. Snead, D., Wragg, R. T., Dittman, J. S. & Eliezer, D. Membrane curvature sensing by the C-terminal domain of complexin. Nat. Commun. 5, 4955 (2014).

    Article  CAS  Google Scholar 

  50. Braun, D. & Fromherz, P. Fluorescence interferometry of neuronal cell adhesion on microstructured silicon. Phys. Rev. Lett. 81, 5241–5244 (1998).

    Article  CAS  Google Scholar 

  51. Braun, D. & Fromherz, P. Fluorescence interference-contrast microscopy of cell adhesion on oxidized silicon. Appl. Phys., A Mater. Sci. Process 65, 341–348 (1997).

    Article  CAS  Google Scholar 

  52. Crane, J. M., Kiessling, V. & Tamm, L. K. Measuring lipid asymmetry in planar supported bilayers by fluorescence interference contrast microscopy. Langmuir 21, 1377–1388 (2005).

    Article  CAS  Google Scholar 

  53. Pabst, S. et al. Selective interaction of complexin with the neuronal SNARE complex. Determination of the binding regions. J. Biol. Chem. 275, 19808–19818 (2000).

    Article  CAS  Google Scholar 

  54. Katti, S. et al. Non-native metal ion reveals the role of electrostatics in synaptotagmin 1–membrane interactions. Biochemistry 56, 3283–3295 (2017).

    Article  CAS  Google Scholar 

  55. Domanska, M. K., Kiessling, V., Stein, A., Fasshauer, D. & Tamm, L. K. Single vesicle millisecond fusion kinetics reveals number of SNARE complexes optimal for fast SNARE-mediated membrane fusion. J. Biol. Chem. 284, 32158–32166 (2009).

    Article  CAS  Google Scholar 

  56. Wagner, M. L. & Tamm, L. K. Tethered polymer-supported planar lipid bilayers for reconstitution of integral membrane proteins: silane-polyethyleneglycol-lipid as a cushion and covalent linker. Biophys. J. 79, 1400–1414 (2000).

    Article  CAS  Google Scholar 

  57. Kalb, E., Frey, S. & Tamm, L. K. Formation of supported planar bilayers by fusion of vesicles to supported phospholipid monolayers. Biochim. Biophys. Acta 1103, 307–316 (1992).

    Article  CAS  Google Scholar 

  58. van den Ent, F. & Löwe, J. RF cloning: a restriction-free method for inserting target genes into plasmids. J. Biochem. Biophys. Methods 67, 67–74 (2006).

    Article  Google Scholar 

  59. Kiessling, V., Crane, J. M. & Tamm, L. K. Transbilayer effects of raft-like lipid domains in asymmetric planar bilayers measured by single molecule tracking. Biophys. J. 91, 3313–3326 (2006).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank A. Lambacher and P. Fromherz for providing FLIC analysis software and for their help with the production of FLIC substrates. This work was supported by US NIH grant P01 GM72694 to L.K.T., D.S.C., B.L., and V.K.

Author information

Authors and Affiliations

Authors

Contributions

V.K. performed and analyzed all of the sdFLIC experiments. A.J.B.K. performed and analyzed all of the DCV fusion experiments. B.L. and S.B.N. prepared protein samples. P.S. prepared shRNA for SytKD DCVs. V.K. designed the work. V.K., D.S.C., J.D.C., and L.K.T. wrote the paper.

Corresponding author

Correspondence to Volker Kiessling.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Integrated supplementary information

Supplementary Figure 1 SdFLIC example images and evaluation of different SNARE complexes in bPC/bPE/bPS/bPIP2/cholesterol.

ac, Example sdFLIC images (second column) with a yellow grid of 12 × 12 oxides that were evaluated, extracted intensity data from one set of 16 oxides and FLIC fit curve (third column), and example histogram of fit results (fourth column) obtained from one sample under one condition for Syx*192/SNAP-25 (a), Syx*192/SNAP-25/Syb(1-96) (b), and Syx*192(183–288)/SNAP-25/Syb (c). Images were rotated and scaled for printing and show an area of 201 × 201 μm2.

Supplementary Figure 2 SdFLIC example images and evaluation for Syx*192/SNAP-25/Syb(1-96) in different lipid environments.

al, Example sdFLIC images with a yellow grid of 12 × 12 oxides that were evaluated, extracted intensity data from one set of 16 oxides and FLIC fit curve, and example histogram of fit results obtained from one sample under one condition for Syx*192/SNAP-25/Syb(1-96). Images in af were taken before and after the addition of Ca2+/C2AB in different lipid headgroup environments, as described in Fig. 1d and indicated in each panel. Images in gl were taken in different lipid environments, as described in Fig. 2a and indicated in each panel. Images were rotated and scaled for printing and show an area of 201 × 201 μm2.

Supplementary Figure 3 SdFLIC example images and evaluation for Syx*192/SNAP-25/Syb(1-96) in different lipid environments (continued).

ag, Example sdFLIC images with a yellow grid of 12 × 12 oxides that were evaluated, extracted intensity data from one set of 16 oxides and FLIC fit curve, and example histogram of fit results obtained from one sample under one condition for Syx*192/SNAP-25/Syb(1-96). Images were taken in different lipid environments, as described in Fig. 2a (ad) and Fig. 2c (eg), and as indicated in each panel. Images were rotated and scaled for printing and show an area of 201 × 201 μm2.

Supplementary Figure 4 SdFLIC and DCV fusion results obtained with different SNARE complex constructs and supported membrane control.

a, sdFLIC results from Syx*/SNAP-25/Syb(1-96) in bPC/bPE/bPS/bPIP2/cholesterol, labeled at residues 192 and 249 for comparison in EDTA (solid bars) and after addition of Ca2+/C2AB (open bars). b, sdFLIC results from Syb*28(1-96) after it was added to reconstituted Syx1a/SNAP-25 acceptor complex30 in different lipid environments as indicated in EDTA (solid bars) and after addition of Ca2+/C2AB (open bars). c, sdFLIC results from Syb*28(1-96) after binding to reconstituted Syx1a/SNAP-25 acceptor complex at different bPS concentrations with and without 1 mol% PIP2 as indicated. d, Distance of the lipid-bilayer surface from substrate as determined by FLIC in EDTA and after addition of Ca2+/C2AB. e, sdFLIC results from Syx*192/SNAP-25/Syb(1-96) in bPC/bPE/bPS/bPIP2/cholesterol in EDTA (solid bar) and after addition of C2AB/EDTA, Ca2+ and EDTA in successive order. f, sdFLIC and g, fusion probabilities when complexin, Munc18, or both have been added to form the ‘trigger-ready’ primed prefusion state19. When complexin is added to Syx*/SNAP-25/Syb(1-96), the complex moves slightly further away from the membrane. Adding Ca2+/C2AB straightens the complex in the same way as without complexin. When complexin is added to acceptor complex in the supported membrane, it inhibits Syb binding and DCV docking in the absence of Ca2+ (ref. 22). In the presence of Ca2+/C2AB, Syt-deficient DCVs dock and proceed to fusion with the supported membrane, similar to the WT DCV fusion in the presence of Ca2+ (ref. 19). When Munc18 and soluble Syb2 (Syb1-96) is added to reconstituted Syx*192/SNAP-25 (Fig. 1c), the measured distance of Syx*192 within this complex increases by more than 6 nm. When Ca2+/C2AB is added to this in the membrane assembled complex, its structure becomes more upright although the distance does not increase to the same height as in the isolated preassembled SNARE complex. Fusion of SytKD-DCVs in the absence of Ca2+ is increased by Munc18 and is further stimulated by Ca2+/C2AB. The height of Syx*192 within a complex that has been assembled by adding Munc18, complexin, and Syb1-96 to Syx*192/SNAP-25, is raised to 9.5 nm. When Ca2+/C2AB is added, the distance of Syx*192 from the membrane increases further to 11.6 nm, about the same as that of the cis-SNARE complex. Adding Munc18 and complexin at the same time to Syx1a/SNAP-25 in the supported membrane results in a ‘trigger competent’ prefusion state that allows DCV docking in the absence of Ca2+ without significant fusion19. Adding Ca2+/C2AB to the docked DCVs after a 15-min incubation time increases the fusion probability of the SytKD-DCVs from almost zero to 58%. Note that when multiprotein complexes are assembled on the membrane, the measured distance of Syx*192 most likely represents an average originating from different multiprotein assemblies. All data represent means from at least five independent experiments. Error bars represent s.e.m.

Supplementary Figure 5 SNARE and Ca2+/C2AB-dependent fusion in proteo-liposome/DCV fusion assay and Syx*192-membrane distance changes induced by C2AB-WT or the C2AB-KAKA mutant with and without PIP2.

ad, Example (NBD) fluorescence dequenching curves after WT-DCVs have been added to Syx/SNAP-25-containing proteoliposomes in EDTA (black curves) or in the presence of Ca2+/C2AB. Syx/SNAP-25 is reconstituted into DP (a), PO (b), DO (c), or brain lipids (d). e, Average fluorescence increase due to NBD dequenching after proteoliposome/DCV fusion in EDTA (solid bars) and in the presence of Ca2+/C2AB. The lipid headgroup composition is PC/PE/PS/PIP2/cholesterol (34/30/15/1/20) in all cases, and the acyl chain composition is as indicated. Averages represent means from three repeats, and error bars represent s.d. f, Syx*192-membrane distance changes in response to C2AB-WT with (red) and without (orange) 1 mol% PIP2 and C2AB-KAKA with (light blue) and without (dark blue) 1 mol% PIP2. Data points represent means from at least five independent experiments. Error bars represent s.e.m.

Supplementary information

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kiessling, V., Kreutzberger, A.J.B., Liang, B. et al. A molecular mechanism for calcium-mediated synaptotagmin-triggered exocytosis. Nat Struct Mol Biol 25, 911–917 (2018). https://doi.org/10.1038/s41594-018-0130-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41594-018-0130-9

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