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:

In situ architecture of the ER–mitochondria encounter structure

Abstract

The endoplasmic reticulum and mitochondria are main hubs of eukaryotic membrane biogenesis that rely on lipid exchange via membrane contact sites1,2,3, but the underpinning mechanisms remain poorly understood. In yeast, tethering and lipid transfer between the two organelles is mediated by the endoplasmic reticulum–mitochondria encounter structure (ERMES), a four-subunit complex of unresolved stoichiometry and architecture4,5,6. Here we determined the molecular organization of ERMES within Saccharomyces cerevisiae cells using integrative structural biology by combining quantitative live imaging, cryo-correlative microscopy, subtomogram averaging and molecular modelling. We found that ERMES assembles into approximately 25 discrete bridge-like complexes distributed irregularly across a contact site. Each bridge consists of three synaptotagmin-like mitochondrial lipid binding protein domains oriented in a zig-zag arrangement. Our molecular model of ERMES reveals a pathway for lipids. These findings resolve the in situ supramolecular architecture of a major inter-organelle lipid transfer machinery and provide a basis for the mechanistic understanding of lipid fluxes in eukaryotic cells.

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: The number of molecules of ERMES components per MCS.
Fig. 2: ERMES-mediated MCS consist of bridge structures connecting the two membranes.
Fig. 3: Supramolecular organization of ERMES within MCS.
Fig. 4: Integrative modelling of the ERMES complex.

Similar content being viewed by others

Data availability

Data generated in this study have been deposited at the Electron Microscopy Data Bank76 (www.ebi.ac.uk/emdb) under accessions EMD-15355, EMD-16871, EMD-16872 and EMD-16873. Raw cryo-ET data has been deposited at EMPIAR77 (https://www.ebi.ac.uk/empiar) under accession EMPIAR-11462. Live fluorescence microscopy data and integrative molecular models have been deposited at www.zenodo.org at https://doi.org/10.5281/zenodo.6784812, https://doi.org/10.5281/zenodo.7392153, https://doi.org/10.5281/zenodo.7753491 and https://doi.org/10.5281/zenodo.7736245. Source data are provided with this paper.

Code availability

The SpotQuant package20 used for quantification of fluorescence intensities is available at https://github.com/apicco/spotquant and https://github.com/apicco/spotquant/tree/b10. The STA package subTOM50 is available at https://github.com/DustinMorado/subTOM. Other code available is for conversion of Dynamo dipole model coordinates to a MOTL file https://github.com/mwozn/DYNAMO_dipoles_to_MOTL, calculation and classification of bridge lengths from dipole ER- and OMM anchor points https://github.com/mwozn/getDipoleLength_binByLength, nearest-neighbour analysis https://github.com/mwozn/MOTL_neighbourAnalysis, and fluorescent protein levels on mitochondria https://github.com/mwozn/totalFluorPerCell.

References

  1. Vance, J. E. Phospholipid synthesis in a membrane fraction associated with mitochondria. J. Biol. Chem. 265, 7248–7256 (1990).

    Article  CAS  PubMed  Google Scholar 

  2. Vance, J. E. Newly made phosphatidylserine and phosphatidylethanolamine are preferentially translocated between rat liver mitochondria and endoplasmic reticulum. J. Biol. Chem. 266, 89–97 (1991).

    Article  CAS  PubMed  Google Scholar 

  3. Achleitner, G. et al. Association between the endoplasmic reticulum and mitochondria of yeast facilitates interorganelle transport of phospholipids through membrane contact. Eur. J. Biochem. 264, 545–553 (1999).

    Article  CAS  PubMed  Google Scholar 

  4. Kornmann, B. et al. An ER-mitochondria tethering complex revealed by a synthetic biology screen. Science 325, 477–481 (2009).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  5. Kawano, S. et al. Structure-function insights into direct lipid transfer between membranes by Mmm1-Mdm12 of ERMES. J. Cell Biol. 217, 959–974 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. John Peter, A. T., Petrungaro, C., Peter, M. & Kornmann, B. METALIC reveals interorganelle lipid flux in live cells by enzymatic mass tagging. Nat. Cell Biol. 24, 996–1004 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Reinisch, K. M. & Prinz, W. A. Mechanisms of nonvesicular lipid transport. J. Cell Biol. 220, e202012058 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Scorrano, L. et al. Coming together to define membrane contact sites. Nat. Commun. 10, 1287 (2019).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  9. Murley, A. et al. ER-associated mitochondrial division links the distribution of mitochondria and mitochondrial DNA in yeast. eLife 2, e00422 (2013).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  10. Kopec, K. O., Alva, V. & Lupas, A. N. Homology of SMP domains to the TULIP superfamily of lipid-binding proteins provides a structural basis for lipid exchange between ER and mitochondria. Bioinformatics 26, 1927–1931 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Toulmay, A. & Prinz, W. A. A conserved membrane-binding domain targets proteins to organelle contact sites. J. Cell Sci. 125, 49–58 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Wong, L. H. & Levine, T. P. Tubular lipid binding proteins (TULIPs) growing everywhere. Biochim. Biophys. Acta 1864, 1439–1449 (2017).

    Article  CAS  PubMed Central  Google Scholar 

  13. Schauder, C. M. et al. Structure of a lipid-bound extended synaptotagmin indicates a role in lipid transfer. Nature 510, 552–555 (2014).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  14. AhYoung, A. P. et al. Conserved SMP domains of the ERMES complex bind phospholipids and mediate tether assembly. Proc. Natl Acad. Sci. USA 112, E3179–E3188 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Lees, J. A. et al. Lipid transport by TMEM24 at ER–plasma membrane contacts regulates pulsatile insulin secretion. Science 355, eaah6171 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  16. Jeong, H., Park, J., Jun, Y. & Lee, C. Crystal structures of Mmm1 and Mdm12–Mmm1 reveal mechanistic insight into phospholipid trafficking at ER–mitochondria contact sites. Proc. Natl Acad. Sci. USA 114, E9502–E9511 (2017).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  17. Hoffmann, P. C. et al. Tricalbins contribute to cellular lipid flux and form curved ER–PM contacts that are bridged by rod-shaped structures. Dev. Cell 51, 488–502.e488 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Jeong, H., Park, J. & Lee, C. Crystal structure of Mdm12 reveals the architecture and dynamic organization of the ERMES complex. EMBO Rep. 17, 1857–1871 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Picco, A., Mund, M., Ries, J., Nedelec, F. & Kaksonen, M. Visualizing the functional architecture of the endocytic machinery. eLife 4, e04535 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  20. Picco, A. & Kaksonen, M. Precise tracking of the dynamics of multiple proteins in endocytic events. Methods Cell. Biol. 139, 51–68 (2017).

    Article  CAS  PubMed  Google Scholar 

  21. Ellenrieder, L. et al. Separating mitochondrial protein assembly and endoplasmic reticulum tethering by selective coupling of Mdm10. Nat. Commun. 7, 13021 (2016).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  22. Stroud, D. A. et al. Composition and topology of the endoplasmic reticulum-mitochondria encounter structure. J. Mol. Biol. 413, 743–750 (2011).

    Article  CAS  PubMed  Google Scholar 

  23. Wagner, F. R. et al. Preparing samples from whole cells using focused-ion-beam milling for cryo-electron tomography. Nat. Protoc. 15, 2041–2070 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Cohen, Y. et al. Peroxisomes are juxtaposed to strategic sites on mitochondria. Mol. Biosyst. 10, 1742–1748 (2014).

    Article  CAS  PubMed  Google Scholar 

  25. Mattiazzi Usaj, M. et al. Genome-wide localization study of yeast Pex11 identifies peroxisome-mitochondria interactions through the ERMES complex. J. Mol. Biol. 427, 2072–2087 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Bryant, P., Pozzati, G. & Elofsson, A. Improved prediction of protein-protein interactions using AlphaFold2. Nat. Commun. 13, 1265 (2022).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  27. Jumper, J. et al. Highly accurate protein structure prediction with AlphaFold. Nature 596, 583–589 (2021).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  28. Kornmann, B., Osman, C. & Walter, P. The conserved GTPase Gem1 regulates endoplasmic reticulum–mitochondria connections. Proc. Natl Acad. Sci. USA 108, 14151–14156 (2011).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  29. Rasul, F. et al. Emr1 regulates the number of foci of the endoplasmic reticulum–mitochondria encounter structure complex. Nat. Commun. 12, 521 (2021).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  30. Li, P., Lees, J. A., Lusk, C. P. & Reinisch, K. M. Cryo-EM reconstruction of a VPS13 fragment reveals a long groove to channel lipids between membranes. J. Cell Biol. 219, e202001161 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Rogers, J. R., Espinoza Garcia, G. & Geissler, P. L. Membrane hydrophobicity determines the activation free energy of passive lipid transport. Biophys. J. 120, 3718–3731 (2021).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  32. Moser von Filseck, J., Vanni, S., Mesmin, B., Antonny, B. & Drin, G. A phosphatidylinositol-4-phosphate powered exchange mechanism to create a lipid gradient between membranes. Nat. Commun. 6, 6671 (2015).

    Article  ADS  PubMed  Google Scholar 

  33. Cai, S. et al. In situ architecture of the lipid transport protein VPS13C at ER–lysosome membrane contacts. Proc. Natl Acad. Sci. USA 119, e2203769119 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. de la Mora, E. et al. Nanoscale architecture of a VAP-A–OSBP tethering complex at membrane contact sites. Nat. Commun. 12, 3459 (2021).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  35. Bian, X., Zhang, Z., Xiong, Q., De Camilli, P. & Lin, C. A programmable DNA-origami platform for studying lipid transfer between bilayers. Nat. Chem. Biol. 15, 830–837 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Lawrimore, J., Bloom, K. S. & Salmon, E. D. Point centromeres contain more than a single centromere-specific Cse4 (CENP-A) nucleosome. J. Cell Biol. 195, 573–582 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Trabuco, L. G., Villa, E., Mitra, K., Frank, J. & Schulten, K. Flexible fitting of atomic structures into electron microscopy maps using molecular dynamics. Structure 16, 673–683 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Janke, C. et al. A versatile toolbox for PCR-based tagging of yeast genes: new fluorescent proteins, more markers and promoter substitution cassettes. Yeast 21, 947–962 (2004).

    Article  CAS  PubMed  Google Scholar 

  39. Shaner, N. C. et al. A bright monomeric green fluorescent protein derived from Branchiostoma lanceolatum. Nat. Methods 10, 407–409 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Laughery, M. F. et al. New vectors for simple and streamlined CRISPR–Cas9 genome editing in Saccharomyces cerevisiae. Yeast 32, 711–720 (2015).

    Article  CAS  PubMed  Google Scholar 

  41. Schindelin, J. et al. Fiji: an open-source platform for biological-image analysis. Nat. Methods 9, 676–682 (2012).

    Article  CAS  PubMed  Google Scholar 

  42. Joglekar, A. P., Bouck, D. C., Molk, J. N., Bloom, K. S. & Salmon, E. D. Molecular architecture of a kinetochore-microtubule attachment site. Nat. Cell Biol. 8, 581–585 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Russo, C. J., Scotcher, S. & Kyte, M. A precision cryostat design for manual and semi-automated cryo-plunge instruments. Rev. Sci. Instrum. 87, 114302 (2016).

    Article  ADS  PubMed  Google Scholar 

  44. Schaffer, M. et al. Cryo-focused ion beam sample preparation for imaging vitreous cells by cryo-electron tomography. Bio Protoc. 5, e1575 (2015).

    Article  PubMed  Google Scholar 

  45. Wolff, G. et al. Mind the gap: micro-expansion joints drastically decrease the bending of FIB-milled cryo-lamellae. J. Struct. Biol. 208, 107389 (2019).

    Article  PubMed  Google Scholar 

  46. Mastronarde, D. N. Automated electron microscope tomography using robust prediction of specimen movements. J. Struct. Biol. 152, 36–51 (2005).

    Article  PubMed  Google Scholar 

  47. Paul-Gilloteaux, P. et al. eC-CLEM: flexible multidimensional registration software for correlative microscopies. Nat. Methods 14, 102–103 (2017).

    Article  CAS  PubMed  Google Scholar 

  48. Hagen, W. J., Wan, W. & Briggs, J. A. Implementation of a cryo-electron tomography tilt-scheme optimized for high resolution subtomogram averaging. J. Struct. Biol. 197, 191–198 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  49. Bharat, T. A. M., Hoffmann, P. C. & Kukulski, W. Correlative microscopy of vitreous sections provides insights into BAR-domain organization in situ. Structure 26, 879–886 e873 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Tremel, S. et al. Structural basis for VPS34 kinase activation by Rab1 and Rab5 on membranes. Nat. Commun. 12, 1564 (2021).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  51. Kremer, J. R., Mastronarde, D. N. & McIntosh, J. R. Computer visualization of three-dimensional image data using IMOD. J. Struct. Biol. 116, 71–76 (1996).

    Article  CAS  PubMed  Google Scholar 

  52. Xiong, Q., Morphew, M. K., Schwartz, C. L., Hoenger,A. H. & Mastronarde, D. N. CTF determination and correction for low dose tomographic tilt series. J. Struct. Biol. 168, 378–387 (2009).

  53. Castano-Diez, D., Kudryashev, M., Arheit, M. & Stahlberg, H. Dynamo: a flexible, user-friendly development tool for subtomogram averaging of cryo-EM data in high-performance computing environments. J. Struct. Biol. 178, 139–151 (2012).

    Article  PubMed  Google Scholar 

  54. Nickell, S. et al. TOM software toolbox: acquisition and analysis for electron tomography. J. Struct. Biol. 149, 227–234 (2005).

    Article  PubMed  Google Scholar 

  55. Forster, F., Medalia, O., Zauberman, N., Baumeister, W. & Fass, D. Retrovirus envelope protein complex structure in situ studied by cryo-electron tomography. Proc. Natl Acad. Sci. USA 102, 4729–4734 (2005).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  56. Forster, F. & Hegerl, R. Structure determination in situ by averaging of tomograms. Methods Cell. Biol. 79, 741–767 (2007).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  58. Qu, K. et al. Structure and architecture of immature and mature murine leukemia virus capsids. Proc. Natl Acad. Sci. USA 115, E11751–E11760 (2018).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  59. 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  PubMed  PubMed Central  Google Scholar 

  60. Pettersen, E. F. et al. UCSF ChimeraX: structure visualization for researchers, educators, and developers. Protein Sci. 30, 70–82 (2021).

    Article  CAS  PubMed  Google Scholar 

  61. Wickham, H. ggplot2: Elegant Graphics for Data Analysis (Springer, 2016).

  62. Allen, M., Poggiali, D., Whitaker, K., Rhys Marshall, T. & Kievit, R. A. in Wellcome Open Research Vol. 4 https://doi.org/10.12688/wellcomeopenres.15191.2 (2019).

  63. Flinner, N. et al. Mdm10 is an ancient eukaryotic porin co-occurring with the ERMES complex. Biochim. Biophys. Acta 1833, 3314–3325 (2013).

    Article  CAS  PubMed  Google Scholar 

  64. Huang, J. et al. CHARMM36m: an improved force field for folded and intrinsically disordered proteins. Nat. Methods 14, 71–73 (2017).

  65. Evans, R. et al. Protein complex prediction with AlphaFold-Multimer. Preprint at bioRxiv https://doi.org/10.1101/2021.10.04.463034 (2022).

  66. Sperka-Gottlieb, C. D., Hermetter, A., Paltauf, F. & Daum, G. Lipid topology and physical properties of the outer mitochondrial membrane of the yeast, Saccharomyces cerevisiae. Biochim. Biophys. Acta 946, 227–234 (1988).

    Article  CAS  PubMed  Google Scholar 

  67. Jo, S., Kim, T., Iyer, V. G. & Im, W. CHARMM-GUI: a web-based graphical user interface for CHARMM. J. Comput. Chem. 29, 1859–1865 (2008).

    Article  CAS  PubMed  Google Scholar 

  68. Essmann, U. et al. A smooth particle mesh Ewald method. J. Chem. Phys. 103, 8577–8593 (1995).

    Article  ADS  CAS  Google Scholar 

  69. McGreevy, R., Teo, I., Singharoy, A. & Schulten, K. Advances in the molecular dynamics flexible fitting method for cryo-EM modeling. Methods 100, 50–60 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Monroe, L., Terashi, G. & Kihara, D. Variability of protein structure models from electron microscopy. Structure 25, 592–602 e592 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Olechnovic, K. & Venclovas, C. VoroMQA: assessment of protein structure quality using interatomic contact areas. Proteins 85, 1131–1145 (2017).

    Article  CAS  PubMed  Google Scholar 

  72. Phillips, J. C. et al. Scalable molecular dynamics on CPU and GPU architectures with NAMD. J. Chem. Phys. 153, 044130 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  74. Goddard, T. D. et al. UCSF ChimeraX: meeting modern challenges in visualization and analysis. Protein Sci. 27, 14–25 (2018).

    Article  CAS  PubMed  Google Scholar 

  75. Chovancova, E. et al. CAVER 3.0: a tool for the analysis of transport pathways in dynamic protein structures. PLoS Comput. Biol. 8, e1002708 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Lawson, C. L. et al. EMDataBank unified data resource for 3DEM. Nucleic Acids Res. 44, D396–403 (2016).

    Article  CAS  PubMed  Google Scholar 

  77. Iudin, A. et al. EMPIAR: the Electron Microscopy Public Image Archive. Nucleic Acids Res. 51, D1503–D1511 (2023).

    Article  PubMed  Google Scholar 

  78. Dahan, N. et al. Peroxisome function relies on organelle-associated mRNA translation. Sci. Adv. 8, eabk2141 (2022).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  79. Wu, H. et al. Peroxisome development in yeast is associated with the formation of Pex3-dependent peroxisome–vacuole contact sites. Biochim. Biophys. Acta Mol. Cell. Res. 1866, 349–359 (2019).

    Article  CAS  PubMed  Google Scholar 

  80. Takeda, H. et al. Mitochondrial sorting and assembly machinery operates by beta-barrel switching. Nature 590, 163–169 (2021).

    Article  ADS  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

The authors thank the MRC LMB facilities for light microscopy, electron microscopy and scientific computing for support with data collection and processing; M. Kaksonen for encouragement and helpful discussions; and the Microscopy Imaging Center (MIC) of the University of Bern for support. M.R.W. was supported by the Natural Sciences and Engineering Research Council (NSERC) of Canada (PGSD). This work was supported by the Medical Research Council, as part of United Kingdom Research and Innovation (also known as UK Research and Innovation) under awards MC_UP_1201/8 to W.K. and MC_UP_1201/10 to E.A.M. For the purpose of open access, the authors have applied a CC BY public copyright licence to any Author Accepted Manuscript version arising. Work in the group of W.K. was also supported by the NCCR TransCure, a National Center of Competence in Research of the Swiss National Science Foundation (SNSF) (185544) and the SNSF project funding (201158). Work in the group of S.V. was supported by the SNSF (194807) and by the European Research Council under the European Union’s Horizon 2020 research and innovation programme (grant agreement no. 803952). This work was supported by grants from the Swiss National Supercomputing Centre under project ID s1132.

Author information

Authors and Affiliations

Authors

Contributions

M.R.W. generated yeast strains, collected data for quantification of numbers of molecules, prepared cellular lamellae and collected cryo-CLEM and cryo-ET data. A.D.L. and P.C. performed the computational integrative modelling, supervised by S.V. A.D.L., P.C. and S.V. analysed the integrative modelling data. D.R.M. and M.R.W. developed the STA alignment procedure. M.R.W. performed STA with help from D.R.M. A.P. and M.R.W. implemented the quantitative live-cell imaging pipeline and analysed the resulting data. M.R.W., W.K. and E.A.M. analysed cryo-ET data and conceived the experimental strategy of the project. R.K. and L.I. participated in yeast strain generation, collected and analysed fluorescence imaging data and assessed yeast mutants. P.C.H. participated in yeast strain generation, cryo-FIB milling and cryo-ET data acquisition. W.K. conceived and supervised the project. M.R.W. and W.K. wrote the paper, with direct help by E.A.M., A.D.L. and S.V., as well as input from all authors.

Corresponding authors

Correspondence to Stefano Vanni or Wanda Kukulski.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature thanks the anonymous reviewer(s) for their contribution to the peer review of this work. Peer review reports are available.

Additional information

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

Extended data figures and tables

Extended Data Fig. 1 Correlation between simultaneously labelled ERMES components.

a: Fluorescence live cell microscopy of cells expressing Mdm34-eGFP and Mdm12-mCherry. Individual channels and merge are shown. b: Fluorescence live cell microscopy of cells expressing Mdm34-eGFP and Mmm1-mCherry. Individual channels and merge are shown. c: The fluorescence intensity of Mdm34-eGFP and Mdm12-mCherry, labelled in the same cells, correlates in puncta. n = 128 puncta, Pearson’s correlation R = 0.566, p = 3.38*10−12 (two-sided). d: The fluorescence intensity in a.u. of Mdm34-EGFP and Mmm1-mCherry, labelled in the same cells, correlates in puncta. n = 134 puncta, Pearson’s correlation R = 0.529, p = 5.18*10−11 (two-sided). Scale bars are 3 µm.

Source Data

Extended Data Fig. 2 MCS imaged by cryo-ET at Mdm34-mNeonGreen signals.

a—f: Diversity of membrane morphologies of ER-mitochondria MCS imaged by cryo-ET at locations of Mdm34-mNeonGreen. Six representative examples from a data set of 51 tomograms are shown. Left panels are virtual slices through electron cryo-tomograms. ER-mitochondria MCS are indicated by white arrows. Right panels are segmentation models of the ER (blue), the OMM (yellow), the IMM (mustard) and peroxisomes (pink). The segmentation models are rotated relative to the virtual slices to better visualize the MCS. The example in e is the same as shown in Fig. 3f. g, h: ER-peroxisome contacts imaged by cryo-ET at locations of Mdm34-mNeonGreen. Approximately 15% of Mdm34-mNeonGreen puncta contained ER-peroxisome MCS (indicated by white arrows) rather than ER-mitochondria MCS. Peroxisomes are identified as spheroid single-membrane vesicles more than 100 nm in diameter with dense interior78,79. Scale bars are 100 nm. Note that due to the perspective view, scale bars in the panels showing segmentation models apply only to the front plane of the scenes.

Extended Data Fig. 3 Subtomogram averaging pipeline.

The strategy using sequential alignment steps is outlined in steps a—h. The data set consisted of manually picked coordinates of the ER and OMM anchor points of 1098 bridge structures from 51 electron cryo-tomograms at positions of Mdm34-mNeonGreen signals. The resolution of the final STA map is potentially affected by a minor fraction of bridge-like particles of different identity, which could contribute noise. Images in a are the same as shown in Fig. 2b and c, with modifications to the overlay. Scale bars in a are 100 nm (left image) and 20 nm (images in pink dashed boxes).

Extended Data Fig. 4 Resolution estimate of STA map.

a: Fourier shell correlation (FSC), calculated according to59. The final resolution corresponds to the value at FSC = 0.143. b: Left panel: Three virtual slices through the unmasked STA map. Right panel: Three virtual slices through the STA map masked for FSC calculation.

Extended Data Fig. 5 STA maps of short and long bridges.

a, b: STA map of short class (pink). c, d: STA map of long class (teal). The two classes were obtained by halving the data set of 1098 bridge structures according to bridge length and subjecting both classes separately to the procedure shown in Extended Data Fig. 3. In a and c the maps are shown at lower contour levels than in b and d. Scale bars are 5 nm. Note that due to the perspective view, scale bars apply only to the front plane of the scenes.

Extended Data Fig. 6 Orientation of the bridges relative to the membranes.

a—d: STA alignment strategy to determine the angle between the bridges and the OMM. e: This graph is the same as shown in Fig. 3d. Large point indicates median, vertical lines MAD. n = 1098 bridges from 51 tomograms. f—i: STA alignment strategy to determine the angle between the bridges and the ER. j: This graph is the same as shown in Fig. 3e. Large point indicates median, vertical lines MAD. n = 1098 bridges from 51 tomograms. k: Comparison of the STA maps obtained from different alignments. Grey: Full STA map obtained from the alignment strategy depicted in Extended Data Fig. 3. The STA map is the same as shown in Fig. 2f and g, displayed at different contour level. Green: STA map obtained from the alignment strategy used to determine the angle between bridges and OMM, as depicted in panels a-d. Red: STA map obtained from the alignment strategy used to determine the angle between bridges and ER, as depicted in panels f—i. Scale bars are 5 nm. Note that due to the perspective view, scale bars apply only to the front plane of the scenes.

Extended Data Fig. 7 The distribution of bridges within the MCS ultrastructure.

a—c: Segmentation models of MCS in three different tomograms, with STA maps of short (pink) and long (teal) classes placed back to the tomographic positions of individual bridges. OMM is in yellow, IMM in mustard, and ER in transparent light blue. d: Distribution of bridges of the long and short classes, across MCS. Each dot represents one MCS (n = 64), plotted according to the percentage of long (left y-axis) and short bridges (right y-axis) it contains. e: Distribution of bridges of the long and short classes, within MCS. The number of short (left bar) and long (right bar) bridges that have a nearest neighbour belonging to the long (light grey bar fraction) and short (dark grey bar fraction) class. f: Dot plot and half-violin plot of the surface area of the ER membrane serviced by one ERMES bridge, determined per MCS. Large point indicates median, vertical lines indicate MAD. One outlier is not shown but included in median and MAD determination. n = 63 MCS from 49 tomograms.

Source Data

Extended Data Fig. 8 Integrative modelling of the ERMES complex.

a: Differences between FD (in colour) and AF multimer (grey) predicted structures. While initial predictions of the complex structure using AF multimer27,65 and FD26 yield a nearly identical Mmm1-Mdm12 interface, the Mdm12-Mdm34 FD-interface shows a larger aperture (black arrowheads) which tends to close upon fitting to the cryo-ET STA map (black arrow). This difference between AF and FD highlights a lack of dynamic information as a current limitation of structural prediction. b—d: Quality estimation of the dimer predictions (Mmm1/Mdm12, Mdm12/Mdm34, Mdm34/Mdm10) using the FoldDock (FD) protocol. For each protein, the local pLDDT score is shown, together with the overall DockQ score of the dimer. For all protein-protein interfaces, the DockQ score is above the value generally considered to successfully predict heteromeric interfaces (DockQ > 0.23)26,65. e: Final conformations of the trimeric Mmm1-Mdm12-Mdm34 complex, obtained from the MDFF simulations including 4 POPE lipids after fitting with different scaling factors (g_scale, numerical values given below models). The scaling factor determines the weight of the experimental STA map on the total molecular potential. f: Assessment of MDFF-derived models obtained using different scaling factors, shown in panel e. ccc refers to the cross-correlation coefficient between map and model, voroMQAtotal and voroMQAinter refer to the global voroMQA score and the component including only inter-subunit contacts, respectively. Based on the gscore which combines the assessment parameters (see Methods), g_scale = 0.3 was considered best and is thus shown in other figures. g: Models obtained when the conformation of the Mmm1-Mdm12-Mdm34 complex was biased by MDFF into the short (left) and long (right) STA maps. The interaction between Mmm1 and Mdm12 (indicated by black arrows) appears to be diminished in the long conformation. In both cases, the model contained 4 lipids and the scaling factor was g_scale = 0.1. See also Supplementary Table 1.

Extended Data Fig. 9 Effect of bound lipids on the ERMES complex model.

a: Model of the Mmm1-Mdm12-Mdm34 complex obtained by MDFF when the starting complex contained 4 POPE lipid molecules (left) as compared to 12 POPE lipid molecules (right). Note that the MDFF-derived model with 4 POPE molecules is also shown in Fig. 4a,b and c as well as in Extended Data Fig. 8e (g_scale = 0.3), without visualization of the lipids. b: Distributions of detected cavity radii at the interfaces between subunits Mmm1-Mdm12 (magenta) and Mdm12-Mdm34 (green) in the presence of 12 lipid molecules. The radii were computed from frames taken over the last 100 ns of the MDFF simulation and are shown as probability density distributions. c: Structural comparison between the heterotrimeric complex obtained by MDFF with the STA map (blue), and the same complex after an additional unbiased MD simulation of 120 ns. The left model is with 4 lipids bound, the right model with 12 lipids bound. d: Root mean square deviations (RMSD) of individual subunits and of the heterotrimeric complex, measured over the additional 120 ns unbiased MD simulations of the MDFF-obtained heterotrimeric complex with either 12 or 4 lipids bound (red and orange, respectively). The end points (120 ns) of the red and orange plots correspond to red and orange conformations of the complex shown in d, respectively. The differences in the RMSD of Mmm1 with 12 vs. 4 lipids can be attributed to its N-terminal region, see also Supplementary Fig. 2.

Extended Data Fig. 10 Mdm10 mutation in the predicted Mdm10-Mdm34 interface.

a: MDFF-derived predicted interface between Mdm10 (blue) and Mdm34 (green). Residues in wild type Mdm10 that were mutated to Mdm10W238A/G240L/L274A/F275A to disrupt the interface are indicated in yellow. The mutated residues are not involved in known interactions within the SAM complex21,80. b: Spot growth assay on YPD, comparing wild type, Mdm10Y296A/F298A/Y301A (shown before to disrupt the interface between Mdm10 and Mdm3421) and Mdm10W238A/G240L/L274A/F275A. All three strains express Tom20-Cherry. c: Light microscopy of strains expressing wild type and mutant Mdm10. Tom20-mCherry is used as an indicator of mitochondrial morphology. Mdm10Y296A/F298A/Y301A and Mdm10W238A/G240L/L274A/F275A display a similar mitochondrial phenotype, indicative of ERMES complex disruption4,21. Scale bars are 3 µm. See also Supplementary Fig. 3.

Extended Data Table 1 Cryo-ET data collection and processing

Supplementary information

Supplementary Information

This file contains Supplementary materials and methods, Supplementary Figs. 1–3 and Supplementary Table 1.

Reporting Summary

Peer Review File

Supplementary Video 1

The STA map calculated from bridge structures. In the first half of the movie the STA map is shown at low contour level to visualize parts of the ER membrane (top) and the OMM (bottom). In the second half of the movie, the contour level is increased. The scale bar is 5 nm. Note that due to the perspective view, the scale bar applies only to the front of the scene.

Supplementary Video 2

Segmentation model of ER–mitochondria contact site, imaged by cryo-ET at location of Mdm34–mNeonGreen. The membrane segmentation is the same as shown in Extended Data Fig. 2a. In addition, the STA map was placed at the positions of individual bridge structures (shown in green). The ER is shown in cyan (semi-transparent), OMM in yellow, IMM in mustard. Scale bar is 100 nm. Note that due to the perspective view, the scale bar applies only to the front of the scene.

Supplementary Video 3

Segmentation model of ER–mitochondria contact site, imaged by cryo-ET at location of Mdm34–mNeonGreen. The membrane segmentation is the same as shown in Extended Data Fig. 2b. In addition, the STA map was placed at the positions of individual bridge structures (shown in green). The ER is shown in cyan (semi-transparent), OMM in yellow, IMM in mustard, peroxisome in pink. Scale bar is 100 nm. Note that due to the perspective view, the scale bar applies only to the front of the scene.

Supplementary Video 4

Segmentation model of ER–mitochondria contact site, imaged by cryo-ET at location of Mdm34–mNeonGreen. The membrane segmentation is the same as shown in Extended Data Fig. 2c. In addition, the STA map was placed at the positions of individual bridge structures (shown in green). The ER is shown in cyan (semi-transparent), OMM in yellow, IMM in mustard. Scale bar is 100 nm. Note that due to the perspective view, the scale bar applies only to the front of the scene.

Supplementary Video 5

Segmentation model of ER–mitochondria contact site, imaged by cryo-ET at location of Mdm34–mNeonGreen. The membrane segmentation is the same as shown in Extended Data Fig. 2d. In addition, the STA map was placed at the positions of individual bridge structures (shown in green). The ER is shown in cyan (semi-transparent), OMM in yellow, IMM in mustard. Scale bar is 100 nm. Note that due to the perspective view, the scale bar applies only to the front of the scene.

Supplementary Video 6

Segmentation model of ER–mitochondria contact site, imaged by cryo-ET at location of Mdm34–mNeonGreen. The membrane segmentation is the same as shown in Extended Data Fig. 2e. In addition, the STA map was placed at the positions of individual bridge structures (shown in green). The ER is shown in cyan (semi-transparent), OMM in yellow, IMM in mustard. Scale bar is 100 nm. Note that due to the perspective view, the scale bar applies only to the front of the scene.

Supplementary Video 7

Segmentation model of ER–mitochondria contact site, imaged by cryo-ET at location of Mdm34–mNeonGreen. The membrane segmentation is the same as shown in Extended Data Fig. 2f. In addition, the STA map was placed at the positions of individual bridge structures (shown in green). The ER is shown in cyan (semi-transparent), OMM in yellow, IMM in mustard, peroxisomes in pink. Scale bar is 100 nm. Note that due to the perspective view, the scale bar applies only to the front of the scene.

Supplementary Video 8

Model of the ERMES heterotrimeric Mmm1–Mdm12–Mdm34 complex, obtained by integrative modelling. FoldDock prediction was followed by MDFF into the STA map. First, the STA map (grey) and a space-filling representation of the heterotrimeric complex are shown, before they sequentially fade and a ribbon representation of the heterotrimeric complex remains. Mmm1 is shown in orange, Mdm12 in purple and Mdm34 in green.

Source data

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wozny, M.R., Di Luca, A., Morado, D.R. et al. In situ architecture of the ER–mitochondria encounter structure. Nature 618, 188–192 (2023). https://doi.org/10.1038/s41586-023-06050-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-023-06050-3

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