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:

CD63 sorts cholesterol into endosomes for storage and distribution via exosomes

Abstract

Extracellular vesicles such as exosomes are now recognized as key players in intercellular communication. Their role is influenced by the specific repertoires of proteins and lipids, which are enriched when they are generated as intraluminal vesicles (ILVs) in multivesicular endosomes. Here we report that a key component of small extracellular vesicles, the tetraspanin CD63, sorts cholesterol to ILVs, generating a pool that can be mobilized by the NPC1/2 complex, and exported via exosomes to recipient cells. In the absence of CD63, cholesterol is retrieved from the endosomes by actin-dependent vesicular transport, placing CD63 and cholesterol at the centre of a balance between inward and outward budding of endomembranes. These results establish CD63 as a lipid-sorting mechanism within endosomes, and show that ILVs and exosomes are alternative providers of cholesterol.

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: Characterization of WT and CD63 KO HeLa and MNT-1 cells and sEVs.
Fig. 2: Proteome analysis of WT and CD63 KO HeLa and MNT-1 cells and sEVs.
Fig. 3: Lipidome analysis of WT and CD63 KO HeLa and MNT-1 cells and sEVs.
Fig. 4: Cholesterol localization in WT and CD63 KO HeLa cells.
Fig. 5: Absence of CD63 induces accumulation of cholesterol at the Golgi apparatus and actin-dependent endosomal tubulations.
Fig. 6: CD63 presents an intramembrane cavity able to accommodate cholesterol.
Fig. 7: CD63 modulates cholesterol trafficking to sEVs and exosomes, affecting their biophysical properties.
Fig. 8: CD63 sEVs can transfer cholesterol to recipient cells.

Similar content being viewed by others

Data availability

The mass spectrometry proteomics raw data have been deposited to the ProteomeXchange Consortium via the PRIDE partner repository. The datasets are publicly available at the following links. ProteomeXchange title: Human – Cells and Extracellular vesicles of Hela and MNT-1 WT vs CD63 KO Proteomes 1: ProteomeXchange accession PXD037146, project webpage http://www.ebi.ac.uk/pride/archive/projects/PXD037146, FTP download https://ftp.pride.ebi.ac.uk/pride/data/archive/2024/04/PXD037146. ProteomeXchange title: Human – Cells and Extracellular vesicles of Hela and MNT-1 WT vs CD63 KO Proteomes 2: ProteomeXchange accession PXD037147, project webpage http://www.ebi.ac.uk/pride/archive/projects/PXD037147, FTP download https://ftp.pride.ebi.ac.uk/pride/data/archive/2024/04/PXD037147. ProteomeXchange title: Human – Cells and Extracellular vesicles of Hela and MNT-1 WT vs CD63 KO Exosomes: ProteomeXchange accession PXD037149, project webpage http://www.ebi.ac.uk/pride/archive/projects/PXD037149, FTP download https://ftp.pride.ebi.ac.uk/pride/data/archive/2024/04/PXD037149. UniProt accession codes: Homo Sapiens (UP000005640), P08962 for CD63, P60033 for CD81 and P21926 for CD9. Lipidomics datasets have been included in full in the Supplementary tables, provided as source data. All relevant data are included in the Article. Data supporting the findings of this study are available from the corresponding author on reasonable request. Source data are provided with this paper.

References

  1. Naslavsky, N. & Caplan, S. The enigmatic endosome—sorting the ins and outs of endocytic trafficking. J. Cell Sci. 131, jcs216499 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  2. Cullen, P. J. & Steinberg, F. To degrade or not to degrade: mechanisms and significance of endocytic recycling. Nat. Rev. Mol. Cell Biol. 19, 679–696 (2018).

    Article  CAS  PubMed  Google Scholar 

  3. Simonetti, B. & Cullen, P. J. Actin-dependent endosomal receptor recycling. Curr. Opin. Cell Biol. 56, 22–33 (2019).

    Article  CAS  PubMed  Google Scholar 

  4. Hurley, J. H. ESCRTs are everywhere. EMBO J. 34, 2398–2407 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Baietti, M. F. et al. Syndecan–syntenin–ALIX regulates the biogenesis of exosomes. Nat. Cell Biol. 14, 677–685 (2012).

    Article  CAS  PubMed  Google Scholar 

  6. van Niel, G., D’Angelo, G. & Raposo, G. Shedding light on the cell biology of extracellular vesicles. Nat. Rev. Mol. Cell Biol. 19, 213–228 (2018).

    Article  PubMed  Google Scholar 

  7. van Niel, G. et al. Challenges and directions in studying cell–cell communication by extracellular vesicles. Nat. Rev. Mol. Cell Biol. 23, 369–382 (2022).

    Article  PubMed  Google Scholar 

  8. Tekirdag, K. & Cuervo, A. M. Chaperone-mediated autophagy and endosomal microautophagy: joint by a chaperone. J. Biol. Chem. 293, 5414–5424 (2018).

    Article  CAS  PubMed  Google Scholar 

  9. Charrin, S. et al. A physical and functional link between cholesterol and tetraspanins. Eur. J. Immunol. 33, 2479–2489 (2003).

    Article  CAS  PubMed  Google Scholar 

  10. Odintsova, E. et al. Gangliosides play an important role in the organization of CD82-enriched microdomains. Biochem. J. 400, 315–325 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Charrin, S. et al. Lateral organization of membrane proteins: tetraspanins spin their web. Biochem. J. 420, 133–154 (2009).

    Article  CAS  PubMed  Google Scholar 

  12. Escola, J. M. et al. Selective enrichment of tetraspan proteins on the internal vesicles of multivesicular endosomes and on exosomes secreted by human B-lymphocytes. J. Biol. Chem. 273, 20121–20127 (1998).

    Article  CAS  PubMed  Google Scholar 

  13. Corso, G. et al. Systematic characterization of extracellular vesicle sorting domains and quantification at the single molecule – single vesicle level by fluorescence correlation spectroscopy and single particle imaging. J. Extracell. Vesicles 8, 1663043 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Gupta, D. et al. Amelioration of systemic inflammation via the display of two different decoy protein receptors on extracellular vesicles. Nat. Biomed. Eng. 5, 1084–1098 (2021).

    Article  CAS  PubMed  Google Scholar 

  15. Pols, M. S. & Klumperman, J. Trafficking and function of the tetraspanin CD63. Exp. Cell. Res. 315, 1584–1592 (2009).

    Article  CAS  PubMed  Google Scholar 

  16. Yoshida, T., Ebina, H. & Koyanagi, Y. N-linked glycan-dependent interaction of CD63 with CXCR4 at the Golgi apparatus induces downregulation of CXCR4. Microbiol. Immunol. 53, 629–635 (2009).

    Article  CAS  PubMed  Google Scholar 

  17. Doyle, E. L. et al. CD63 is an essential cofactor to leukocyte recruitment by endothelial P-selectin. Blood 118, 4265–4273 (2011).

    Article  CAS  PubMed  Google Scholar 

  18. van Niel, G. et al. The tetraspanin CD63 regulates ESCRT-independent and -dependent endosomal sorting during melanogenesis. Dev. Cell 21, 708–721 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  19. Colombo, M. et al. Analysis of ESCRT functions in exosome biogenesis, composition and secretion highlights the heterogeneity of extracellular vesicles. J. Cell Sci. 126, 5553–5565 (2013).

    CAS  PubMed  Google Scholar 

  20. van Niel, G. et al. Apolipoprotein E regulates amyloid formation within endosomes of pigment cells. Cell Rep. 13, 43–51 (2015).

    Article  PubMed  Google Scholar 

  21. Hurwitz, S. N. et al. CD63 regulates Epstein-Barr virus LMP1 exosomal packaging, enhancement of vesicle production, and noncanonical NF-κB signaling. J. Virol. 91, e02251-16 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Verweij, F. J. et al. LMP1 association with CD63 in endosomes and secretion via exosomes limits constitutive NF‐κB activation. EMBO J. 30, 2115–2129 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Mathieu, M. et al. Specificities of exosome versus small ectosome secretion revealed by live intracellular tracking of CD63 and CD9. Nat. Commun. 12, 4389 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Hemler, M. E. Targeting of tetraspanin proteins—potential benefits and strategies. Nat. Rev. Drug Discov. 7, 747–758 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. De Silva, N. S. et al. Nuclear envelope disruption triggers hallmarks of aging in lung alveolar macrophages. Nat. Aging 3, 1251–1268 (2023).

    Article  PubMed  Google Scholar 

  26. Dooley, K. et al. A versatile platform for generating engineered extracellular vesicles with defined therapeutic properties. Mol. Ther. 29, 1729–1743 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Filippini, F. et al. Secretion of VGF relies on the interplay between LRRK2 and post-Golgi v-SNAREs. Cell Rep. 42, 112221 (2023).

    Article  CAS  PubMed  Google Scholar 

  28. Fordjour, F. K., Guo, C., Ai, Y., Daaboul, G. G. & Gould, S. J. A shared, stochastic pathway mediates exosome protein budding along plasma and endosome membranes. J. Biol. Chem. 298, 102394 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Fan, Y. et al. Differential proteomics argues against a general role for CD9, CD81 or CD63 in the sorting of proteins into extracellular vesicles. J. Extracell. Vesicles 12, 12352 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Skotland, T., Sagini, K., Sandvig, K. & Llorente, A. An emerging focus on lipids in extracellular vesicles. Adv. Drug Deliv. Rev. 159, 308–321 (2020).

    Article  CAS  PubMed  Google Scholar 

  31. Subra, C., Laulagnier, K., Perret, B. & Record, M. Exosome lipidomics unravels lipid sorting at the level of multivesicular bodies. Biochimie 89, 205–212 (2007).

    Article  CAS  PubMed  Google Scholar 

  32. Bissig, C. & Gruenberg, J. Lipid sorting and multivesicular endosome biogenesis. Cold Spring Harb. Perspect. Biol. 5, a016816 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  33. Trajkovic, K. et al. Ceramide triggers budding of exosome vesicles into multivesicular endosomes. Science 319, 1244–1247 (2008).

    Article  CAS  PubMed  Google Scholar 

  34. Zimmerman, B. et al. Crystal structure of a full-length human tetraspanin reveals a cholesterol-binding pocket. Cell 167, 1041–1051 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Huang, C., Hays, F. A., Tomasek, J. J., Benyajati, S. & Zhang, X. A. Tetraspanin CD82 interaction with cholesterol promotes extracellular vesicle-mediated release of ezrin to inhibit tumour cell movement. J. Extracell. Vesicles 9, 1692417 (2020).

    Article  CAS  PubMed  Google Scholar 

  36. Möbius, W. et al. Recycling compartments and the internal vesicles of multivesicular bodies harbor most of the cholesterol found in the endocytic pathway. Traffic 4, 222–231 (2003).

    Article  PubMed  Google Scholar 

  37. Meng, Y., Heybrock, S., Neculai, D. & Saftig, P. Cholesterol handling in lysosomes and beyond. Trends Cell Biol. 30, 452–466 (2020).

    Article  CAS  PubMed  Google Scholar 

  38. Iaea, D. B. & Maxfield, F. R. Cholesterol trafficking and distribution. Essays Biochem. 57, 43–55 (2015).

    Article  PubMed  Google Scholar 

  39. Shimada, Y., Maruya, M., Iwashita, S. & Ohno‐Iwashita, Y. The C-terminal domain of perfringolysin O is an essential cholesterol-binding unit targeting to cholesterol-rich microdomains. Eur. J. Biochem. 269, 6195–6203 (2002).

    Article  CAS  PubMed  Google Scholar 

  40. Wilhelm, L. P. et al. STARD3 mediates endoplasmic reticulum-to-endosome cholesterol transport at membrane contact sites. EMBO J. 36, 1412–1433 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Edgar, J. R., Willén, K., Gouras, G. K. & Futter, C. E. ESCRTs regulate amyloid precursor protein sorting in multivesicular bodies and intracellular amyloid-β accumulation. J. Cell Sci. 128, 2520–2528 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Hölttä-Vuori, M. et al. BODIPY-cholesterol: a new tool to visualize sterol trafficking in living cells and organisms. Traffic 9, 1839–1849 (2008).

    Article  PubMed  Google Scholar 

  43. Hölttä-Vuori, M., Sezgin, E., Eggeling, C. & Ikonen, E. Use of BODIPY-Cholesterol (TF-Chol) for visualizing lysosomal cholesterol accumulation. Traffic 17, 1054–1057 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  44. Ying, M., Grimmer, S., Iversen, T.-G., Van Deurs, B. & Sandvig, K. Cholesterol loading induces a block in the exit of VSVG from the TGN. Traffic 4, 772–784 (2003).

    Article  CAS  PubMed  Google Scholar 

  45. Miwako, I., Yamamoto, A., Kitamura, T., Nagayama, K. & Ohashi, M. Cholesterol requirement for cation-independent mannose 6-phosphate receptor exit from multivesicular late endosomes to the Golgi. J. Cell Sci. 114, 1765–1776 (2001).

    Article  CAS  PubMed  Google Scholar 

  46. Boncompain, G. et al. Synchronization of secretory protein traffic in populations of cells. Nat. Methods 9, 493–498 (2012).

    Article  CAS  PubMed  Google Scholar 

  47. Derivery, E. et al. The Arp2/3 activator WASH controls the fission of endosomes through a large multiprotein complex. Dev. Cell 17, 712–723 (2009).

    Article  CAS  PubMed  Google Scholar 

  48. Dubois, L., Ronquist, K. K. G., Ek, B., Ronquist, G. & Larsson, A. Proteomic profiling of detergent resistant membranes (lipid rafts) of prostasomes. Mol. Cell. Proteom. 14, 3015–3022 (2015).

    Article  CAS  Google Scholar 

  49. McNamara, R. P. et al. Imaging of surface microdomains on individual extracellular vesicles in 3-D. J. Extracell. Vesicles 11, e12191 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Savinov, S. N. & Heuck, A. P. Interaction of cholesterol with Perfringolysin O: what have we learned from functional analysis? Toxins 9, 381 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  51. Sezgin, E., Levental, I., Mayor, S. & Eggeling, C. The mystery of membrane organization: composition, regulation and roles of lipid rafts. Nat. Rev. Mol. Cell Biol. 18, 361–374 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Maekawa, M. Domain 4 (D4) of perfringolysin O to visualize cholesterol in cellular membranes—the update. Sensors 17, 504 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  53. Arthur, J. R., Heinecke, K. A. & Seyfried, T. N. Filipin recognizes both GM1 and cholesterol in GM1 gangliosidosis mouse brain. J. Lipid Res. 52, 1345–1351 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Edgar, J. R., Manna, P. T., Nishimura, S., Banting, G. & Robinson, M. S. Tetherin is an exosomal tether. eLife 5, e17180 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  55. Collot, M. et al. MemBright: a family of fluorescent membrane probes for advanced cellular imaging and neuroscience. Cell Chem. Biol. 26, 600–614 (2019).

    Article  CAS  PubMed  Google Scholar 

  56. Matsuo, H. et al. Role of LBPA and Alix in multivesicular liposome formation and endosome organization. Science 303, 531–534 (2004).

    Article  CAS  PubMed  Google Scholar 

  57. Möbius, W. et al. Immunoelectron microscopic localization of cholesterol using biotinylated and non-cytolytic perfringolysin O. J. Histochem. Cytochem. 50, 43–55 (2002).

    Article  PubMed  Google Scholar 

  58. Palor, M. et al. Cholesterol sensing by CD81 is important for hepatitis C virus entry. J. Biol. Chem. 295, 16931–16948 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Molostvov, G. et al. Tspan6 stimulates the chemoattractive potential of breast cancer cells for B cells in an EV- and LXR-dependent manner. Cell Rep. 42, 112207 (2023).

    Article  CAS  PubMed  Google Scholar 

  60. Edgar, J. R., Eden, E. R. & Futter, C. E. Hrs- and CD63-dependent competing mechanisms make different sized endosomal intraluminal vesicles. Traffic 15, 197–211 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  61. Latysheva, N. et al. Syntenin-1 is a new component of tetraspanin-enriched microdomains: mechanisms and consequences of the interaction of syntenin-1 with CD63. Mol. Cell. Biol. 26, 7707–7718 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Verweij, F. J. et al. Live tracking of inter-organ communication by endogenous exosomes in vivo. Dev. Cell 48, 573–589 (2019).

    Article  CAS  PubMed  Google Scholar 

  63. Colombo, A. et al. Loss of NPC1 enhances phagocytic uptake and impairs lipid trafficking in microglia. Nat. Commun. 12, 1158 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Strauss, K. et al. Exosome secretion ameliorates lysosomal storage of cholesterol in Niemann-Pick type C disease. J. Biol. Chem. 285, 26279–26288 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Hurwitz, S. N., Cheerathodi, M. R., Nkosi, D., York, S. B. & Meckes, D. G. Tetraspanin CD63 bridges autophagic and endosomal processes to regulate exosomal secretion and intracellular signaling of Epstein-Barr virus LMP1. J. Virol. 92, e01969-17 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  66. Castellano, B. M. et al. Lysosomal cholesterol activates mTORC1 via an SLC38A9–Niemann-Pick C1 signaling complex. Science 355, 1306–1311 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Lippincott-Schwartz, J. & Phair, R. D. Lipids and cholesterol as regulators of traffic in the endomembrane system. Annu. Rev. Biophys. 39, 559–578 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Dominko, K. et al. Impaired retromer function in Niemann-Pick Type C disease is dependent on intracellular cholesterol accumulation. Int. J. Mol. Sci. 22, 13256 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Marquer, C. et al. Arf6 controls retromer traffic and intracellular cholesterol distribution via a phosphoinositide-based mechanism. Nat. Commun. 7, 11919 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Bucki, R. et al. Lateral distribution of phosphatidylinositol 4,5-bisphosphate in membranes regulates formin- and ARP2/3-mediated actin nucleation. J. Biol. Chem. 294, 4704–4722 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Petersen, S. H. et al. The role of tetraspanin CD63 in antigen presentation via MHC class II. Eur. J. Immunol. 41, 2556–2561 (2011).

    Article  CAS  PubMed  Google Scholar 

  72. Roy, K., Ghosh, M., Pal, T. K., Chakrabarti, S. & Roy, S. Cholesterol lowering drug may influence cellular immune response by altering MHC II function. J. Lipid Res. 54, 3106–3115 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Horiuchi, Y. et al. Characterization of the cholesterol efflux of apolipoprotein E-containing high-density lipoprotein in THP-1 cells. Biol. Chem. 400, 209–218 (2019).

    Article  CAS  PubMed  Google Scholar 

  74. Setiadi, H. & McEver, R. P. Clustering endothelial E-selectin in clathrin-coated pits and lipid rafts enhances leukocyte adhesion under flow. Blood 111, 1989–1998 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Verweij, F. J. et al. Exosomal sorting of the viral oncoprotein LMP1 is restrained by TRAF2 association at signalling endosomes. J. Extracell. Vesicles 4, 26334 (2015).

    Article  PubMed  Google Scholar 

  76. Flannery, A. R., Czibener, C. & Andrews, N. W. Palmitoylation-dependent association with CD63 targets the Ca2+ sensor synaptotagmin VII to lysosomes. J. Cell Biol. 191, 599–613 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Liao, G. et al. Enhanced expression of matrix metalloproteinase-12 contributes to Npc1 deficiency-induced axonal degeneration. Exp. Neurol. 269, 67–74 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Loats, A. E. et al. Cholesterol is required for transcriptional repression by BASP1. Proc. Natl Acad. Sci. USA 118, e2101671118 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Yoshida, T. et al. A CD63 mutant inhibits T-cell tropic human immunodeficiency virus type 1 entry by disrupting CXCR4 trafficking to the plasma membrane. Traffic 9, 540–558 (2008).

    Article  CAS  PubMed  Google Scholar 

  80. Gordón-Alonso, M. et al. The PDZ-adaptor protein syntenin-1 regulates HIV-1 entry. Mol. Biol. Cell 23, 2253–2263 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  81. D’Acunzo, P. et al. Enhanced generation of intraluminal vesicles in neuronal late endosomes in the brain of a Down syndrome mouse model with endosomal dysfunction. Dev. Neurobiol. 79, 656–663 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  82. Williams, E. T., Chen, X., Otero, P. A. & Moore, D. J. Understanding the contributions of VPS35 and the retromer in neurodegenerative disease. Neurobiol. Dis. 170, 105768 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Kaprio, T., Hagström, J., Andersson, L. C. & Haglund, C. Tetraspanin CD63 independently predicts poor prognosis in colorectal cancer. Histol. Histopathol. 35, 887–892 (2020).

    CAS  PubMed  Google Scholar 

  84. Gupta, D. et al. Quantification of extracellular vesicles in vitro and in vivo using sensitive bioluminescence imaging. J. Extracell. Vesicles 9, 1800222 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Ran, F. A. et al. Genome engineering using the CRISPR-Cas9 system. Nat. Protoc. 8, 2281–2308 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Wilhelm, L. P., Voilquin, L., Kobayashi, T., Tomasetto, C. & Alpy, F. Intracellular and plasma membrane cholesterol labeling and quantification using filipin and GFP-D4. Methods Mol. Biol. Clifton NJ 1949, 137–152 (2019).

    Article  CAS  Google Scholar 

  87. Ripoll, L. et al. Myosin VI and branched actin filaments mediate membrane constriction and fission of melanosomal tubule carriers. J. Cell Biol. 217, 2709–2726 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Raposo, G., Tenza, D., Murphy, D. M., Berson, J. F. & Marks, M. S. Distinct protein sorting and localization to premelanosomes, melanosomes and lysosomes in pigmented melanocytic cells. J. Cell Biol. 152, 809–824 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Poullet, P., Carpentier, S. & Barillot, E. myProMS, a web server for management and validation of mass spectrometry-based proteomic data. Proteomics 7, 2553–2556 (2007).

    Article  CAS  PubMed  Google Scholar 

  90. The, M., MacCoss, M. J., Noble, W. S. & Käll, L. Fast and accurate protein false discovery rates on large-scale proteomics data sets with Percolator 3.0. J. Am. Soc. Mass Spectrom. 27, 1719–1727 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Valot, B., Langella, O., Nano, E. & Zivy, M. MassChroQ: a versatile tool for mass spectrometry quantification. Proteomics 11, 3572–3577 (2011).

    Article  CAS  PubMed  Google Scholar 

  92. Kowal, J. et al. Proteomic comparison defines novel markers to characterize heterogeneous populations of extracellular vesicle subtypes. Proc. Natl Acad. Sci. USA 113, E968–E977 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Perez-Riverol, Y. et al. The PRIDE database and related tools and resources in 2019: improving support for quantification data. Nucleic Acids Res. 47, D442–D450 (2019).

    Article  CAS  PubMed  Google Scholar 

  94. Ejsing, C. S. et al. Automated identification and quantification of glycerophospholipid molecular species by multiple precursor ion scanning. Anal. Chem. 78, 6202–6214 (2006).

    Article  CAS  PubMed  Google Scholar 

  95. Saeed, A. I. et al. TM4: a free, open-source system for microarray data management and analysis. BioTechniques 34, 374–378 (2003).

    Article  CAS  PubMed  Google Scholar 

  96. Benjamini, Y. & Hochberg, Y. Controlling the false discovery rate: a practical and powerful approach to multiple testing. J. R. Stat. Soc. Ser. B Methodol. 57, 289–300 (1995).

    Article  Google Scholar 

  97. Zhang, Y. I-TASSER server for protein 3D structure prediction. BMC Bioinformatics 9, 40 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  98. van Zundert, G. C. P. et al. The HADDOCK2.2 Web Server: user-friendly integrative modeling of biomolecular complexes. J. Mol. Biol. 428, 720–725 (2016).

    Article  PubMed  Google Scholar 

  99. Vorselen, D., MacKintosh, F. C., Roos, W. H. & Wuite, G. J. L. Competition between bending and internal pressure governs the mechanics of fluid nanovesicles. ACS Nano 11, 2628–2636 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Vorselen, D., Piontek, M. C., Roos, W. H. & Wuite, G. J. L. Mechanical characterization of liposomes and extracellular vesicles, a protocol. Front. Mol. Biosci. 7, 139 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Vorselen, D. et al. The fluid membrane determines mechanics of erythrocyte extracellular vesicles and is softened in hereditary spherocytosis. Nat. Commun. 9, 4960 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  102. Sorkin, R. et al. Nanomechanics of extracellular vesicles reveals vesiculation pathways. Small 14, 1801650 (2018).

    Article  Google Scholar 

  103. Sorkin, R. et al. Synaptotagmin-1 and Doc2b exhibit distinct membrane-remodeling mechanisms. Biophys. J. 118, 643–656 (2020).

    Article  CAS  PubMed  Google Scholar 

  104. Mary, B., Ghoroghi, S., Hyenne, V. & Goetz, J. G. in Methods in Enzymology (eds Spada, S. & Galluzzi, L.) vol. 645, 243–275 (Academic, 2020).

Download references

Acknowledgements

This work was funded by the Institut Curie International PhD Program (to R.P.), the Fondation ARC pour la Recherche sur le Cancer (DOC20180507506 to R.P. and PGA1 RF20190208474 to M.C.), the Fondation pour la Recherche Médicale (AJE20160635884 to G.v.N.), the Institut National du Cancer (grant no. 2019-125 PLBIO19°059), ANR (ANR-20-CE18-0026-01 to G.v.N. and ANR-18-CE13-0017-02 to E.R.), Région Ile-de-France and Fondation pour la Recherche Médicale grants (to D.L.) and the STW Cancer-ID program (project no. 14192 to W.H.R.). We thank the Cell and Tissue Imaging core facility (PICT IBiSA) and Nikon Imaging Centre at Institut Curie–CNRS, member of the French National Research Infrastructure France–BioImaging (ANR-10-INBS-04) and the NeurImag core facility team for their technical and scientific support. NeurImag is part of IPNP, Inserm U1266 and Université Paris Cité and a member of the national infrastructure France-BioImaging supported by the French National Research Agency (ANR-10-INBS-04). We also thank the Leducq establishment for funding the Leica SP8 confocal/STED 3DX system, the Bettencourt Foundation for funding the Leica/Yokogawa spinning disc system, A. Canette and M. Trichet at the Service de Microscopie Électronique (SME) de l’Institut de Biologie Paris-Seine, and C. Durieu at ImagoSeine core facility of Institut Jacques Monod, member of France–BioImaging (ANR-10-INBS-04) and IBiSA, with the support of Labex ‘Who Am I’, Inserm Plan Cancer, Region Ile-de-France and Fondation Bettencourt Schueller. We thank the Structure and Membrane Compartment laboratory and the Endosomal Dynamic in Neuropathies laboratory for insightful discussions. We thank F. Alpy for discussions and providing essential reagents.

Author information

Authors and Affiliations

Authors

Contributions

R.P., M.C. and G.v.N. designed, performed and analysed most of the experiments. R.P. and G.v.N. wrote the paper with help from M.C., M.C.P. and W.H.R. M.C.P., S.S. and W.H.R. performed and analysed atomic force microscopy. F.D. and D.L. performed and analysed proteome mass spectrometry. M.P., M.L. and A.K. performed lipidomic analysis and/or analysed lipidomic data. F.J.V., S.C., M.T. and E.R. performed experiments and/or generated tools and cell lines. G.R. and G.v.N. acquired funding. G.v.N. conceived and supervised the study. All authors read and approved the paper.

Corresponding author

Correspondence to Guillaume van Niel.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Cell Biology thanks the anonymous reviewers for their contribution to the peer review of this work.

Additional information

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

Extended data

Extended Data Fig. 1 Characterization of WT, CD63 KO, siRNA treated, cells and sEVs.

a. NTA analysis of sEVs derived from WT or CD63 KO HeLa cells and isolated by SEC. Average number of particles/ml (n = 7 independent experiments, Ordinary one-way ANOVA, ns P = 0.6449 and ns P = 0.9359). b. DLS analysis of sEVs derived from WT or CD63 KO HeLa cells grown in FBS supplemented media. Average number of particles/ml (n = 5 independent experiments, two-tailed unpaired t test, ns P = 0.9135). c. DLS analysis of sEVs derived from WT or CD63 KO HeLa or MNT-1 cells grown in LPDS. Average number of particles/ml (n = 5 independent experiments, two-tailed Mann-Whitney test for Hela cells, ns P = 0.5952; two-tailed unpaired t test for MNT-1 cells, ns P = 0.3579). d. Western blot analysis of cell lysates and sEVs derived from Ctrl or CD63 siRNA-treated MNT-1 cells and isolated by ultracentrifugation. Quantification of protein content normalized to Ctrl siRNA-treated MNT-1 for cell lysates (n= independent experiments, n = 5 for CD9, n = 6 for alix, syntenin, n = 7 for CD63, ApoE, unpaired multiple t test, **** P = 0.000027) and for sEVs (n = independent experiments, n = 3 for CD9, n = 5 for other markers, unpaired multiple t test, ** CD63 P = 0.01129, ** ApoE P = 0.0065). e. NTA analysis of sEVs derived from Ctrl or CD63 siRNA-treated MNT-1 cells and isolated by ultracentrifugation. Average number of particles/ml (mean of 2 independent experiments) or size distribution (data are presented as the mean of 2 independent experiments). Data are presented as mean values +/− SEM. Source numerical data and unprocessed blots are available in source data.

Source data

Extended Data Fig. 2 Proteomic analysis of WT and CD63 KO cells and sEVs.

a. Volcano plot of proteins identified by at least 2 peptides in at least 3 replicates. Shown are the fold changes of peptide abundance between WT or CD63 KO HeLa cells and the p-value of this quantification. b. List of proteins enriched in WT HeLa cell proteome. List of Cellular components GO terms in WT HeLa cells. c. Volcano plot of proteins identified by at least 2 peptides in at least 3 replicates. Shown are the fold changes of peptide abundance between WT or CD63 KO HeLa sEVs and the p-value of this quantification. d. List of proteins enriched in WT HeLa sEVs proteome. List of Cellular components GO terms in WT HeLa sEVs. e. Volcano plot of proteins identified by at least 2 peptides in at least 3 replicates. Shown are the fold changes of peptide abundance between WT or CD63 KO MNT-1 cells and the p-value of this quantification. f. List of proteins enriched in WT MNT-1 cell proteome. List of Cellular components GO terms in WT MNT-1 cells. g. Volcano plot of proteins identified by at least 2 peptides in at least 3 replicates. Shown are the fold changes of peptide abundance between WT or CD63 KO MNT-1 sEVs and the p-value of this quantification. h. List of proteins enriched in WT MNT-1 sEVs. List of Cellular components GO terms in WT MNT-1 sEVs. A linear model (adjusted on peptides and biological replicates) was performed, and a two sided T-test was applied on the fold change estimated by the model. The p-values were then adjusted using the Benjamini–Hochberg FDR procedure. Source numerical data are available in source data.

Source data

Extended Data Fig. 3 Lipidomic analysis of WT, CD63 KO cells and sEVs.

a. Clustered heatmap of significantly altered lipid features (following FDR correction) identified in WT or CD63 KO HeLa cells (n = independent experiments, 3 for WT and 7 for KO, two-tailed Mann–Whitney test). b. Clustered heatmap of significantly altered lipid features (following FDR correction) identified in WT or CD63 KO MNT-1 cells (n = 3 independent experiments, two-tailed Mann-Whitney test). P values: *P = < 0.05, **P = < 0.01, ***P = < 0.001. Source numerical data are available in source data.

Source data

Extended Data Fig. 4 Cholesterol trafficking in WT, CD63 KO, and siRNA treated cells.

a. Clustered heatmap of significantly altered lipid features (following FDR correction) identified in WT or CD63 KO HeLa sEVs (n = 8 independent experiments, two-tailed Student’s paired t test). b. Clustered heatmap of significantly altered lipid features (following FDR correction) identified in WT or CD63 KO MNT-1 sEVs (n = 5 independent experiments, two-tailed Student’s paired t test). c. PC/PE molar ratio in HeLa cells and sEVs (n = 3 independent experiments for WT cells (min, Q1-min, median-Q1, Q3-median, max-Q3; (2.75, 0.3, 0.3, 0.13, 0.13)), n = 7 independent experiments for KO cells (min, Q1-min, median-Q1, Q3-median, max-Q3; (3.03, 0.21, 0.29, 1.06, 0.59)), n = 4 independent experiments for WT sEVs (min, Q1-min, median-Q1, Q3-median, max-Q3; (1.47, 0.06, 0.05, 0.14, 0.32)), n = 8 independent experiments for KO sEVs (min, Q1-min, median-Q1, Q3-median, max-Q3; (1.63, 0.36, 0.08, 0.15, 0.28)), two-tailed student t test). d. PC/PE molar ratio in MNT-1 cells and sEVs (n = 3 independent experiments for WT cells (min, Q1-min, median-Q1, Q3-median, max-Q3; (5.48, 0.1, 0.11, 0.47, 0.47)), n = 3 independent experiments for KO cells (min, Q1-min, median-Q1, Q3-median, max-Q3; (0.13, 2.59, 0.99, 0.2, 0.22)), (n = 5 independent experiments for WT sEVs (min, Q1-min, median-Q1, Q3-median, max-Q3; (2.72, 0.12, 0.32, 0.55, 1.04)), n = 5 independent experiments for KO sEVs, (min, Q1-min, median-Q1, Q3-median, max-Q3; (3.14, 0.05, 0.05, 0.51, 0.47)) two-tailed student t test). P values: *P = < 0.05, **P = < 0.01, ***P = < 0.001. Source numerical data are available in source data.

Source data

Extended Data Fig. 5 Effect of CD63 depletion on the anterograde and retrograde trafficking of cargoes.

a. HeLa cells WT or CD63 KO were grown in FBS, processed for ultrathin cryo-sectioning and immunogold labeled with D4-GFP (PAG 10 nm) and CD63 (PAG 5 nm). Bars = 200 nm. Representative of 2 independent experiments. b. Cholesterol content of WT and CD63 KO HeLa or MNT-1 cells measured by Amplex Red assay. (n = 3 independent experiments, Mann-Whitney test for HeLa, ns P > 0.9999, unpaired t test for MNT-1, ns P = 0.9256). c. Western blot analysis of WT or CD63 KO HeLa cell lysates. Quantification of protein content, shown as the ratio between nuclear SREBP2 and total SREBP2, normalized to HeLa WT (n = 3 independent experiments, Ordinary one-way ANOVA, ns P = 0.4827 and 0.5698). d. Localization of Nile Red (staining lipid droplets) in WT and CD63 KO HeLa cells. Bars=10 µm. Quantification of Nile Red fluorescence intensity (n= number of cells, H WT n = 27, H CD63 KO#1 n = 25, H CD63 KO#2 n = 35, from 1 independent experiment, Ordinary one-way ANOVA, ns P > 0.05). e. WT or CD63 KO MNT-1 cells were grown in LPDS, processed for ultrathin cryo-sectioning and immunogold labeled with D4-GFP (PAG 10 nm). White arrows indicate MVEs containing cholesterol. Bars=200 nm. Representative of 2 independent experiments. f. Ctrl or CD63 siRNA-treated MNT-1 cells were grown in LPDS, processed for ultrathin cryo-sectioning and immunogold labeled with D4-GFP (PAG 10 nm) and CD63 (PAG 5 nm). White arrows indicate MVEs and black arrows Golgi stacks enriched in cholesterol. Bars=500 nm. Representative of 2 independent experiments. g. Localization of cholesterol stained with D4-GFP in WT or CD63 KO non-permeabilized Hela cells, Bars=10 µm. Quantification of D4-GFP intensity of fluorescence (n = 42 WT cells and, n = 89 KO cells pooled across 2 independent experiments, two-tailed Mann-Whitney test, **** P < 0.0001). Data are presented as mean values +/− SEM. Source numerical data and unprocessed blots are available in source data.

Source data

Extended Data Fig. 6 Endosomal actin-dependent tubulation is required for cholesterol trafficking in absence of CD63.

a. Localization of VSVG-EGFP-RUSH in WT or CD63 KO HeLa cells at different time (T) points after biotin addition. Biotin was added at T = 0. Bars = 10 µm. Representative of 2 independent experiments. b, c. Pulse - chase of anti CI-M6PR antibody in WT or CD63 KO HeLa cells (b) or MNT-1 cells treated with control or CD63 siRNA (c), and IFM with secondary antibody anti-mouse for CI-M6PR. CI-M6PR staining is shown as pseudo-color. Bars = 10 µm. CI-M6PR fluorescence intensity in the whole cells or in the Golgi area was measured. Quantifications show the ratio of vesicular /Golgi fluorescence intensity associated with CI-M6PR. (H WT n = 51 cells, H CD63 KO#1 n = 55 cells, H CD63 KO#2 n = 52 cells, pooled across 3 independent experiments, Ordinary one-way ANOVA, ** P = 0.0028, **** P < 0.0001; M siCtrl n = 18 cells, M siCD63 n = 18 cells, pooled across 2 independent experiments, two-tailed unpaired Student’s t test with Welch’s correction, ns P = 0.1292). d. IFM of WT or CD63 KO MNT-1 cells and of Ctrl or CD63 siRNA treated MNT-1 cells grown in LPDS and co-stained for endogenous ApoE and TGN46. Pearson’s correlation coefficient (n = 33 cells pooled across 3 independent experiments, two-tailed unpaired Student’s t test, **** P < 0.0001). e. IFM of Ctrl or CD63 siRNA treated MNT-1 cells grown in LPDS and co-stained for endogenous ApoE and GM130. Bars=10 µm. Pearson’s correlation coefficient (M siCtrl n = 22 cells and, M siCD63 n = 15 cells pooled across 3 independent experiments, two-tailed unpaired t-test, ** P = 0.0018). f. IFM of Ctrl or CD63 siRNA treated MNT-1 cells grown in LPDS and co-stained for endogenous ApoE and LAMP-1. Bars=10 µm. Pearson’s correlation coefficient (M siCtrl n = 27 cells and, M siCD63 n = 31 cells pooled across 3 independent experiments, two-tailed unpaired Student’s t test, *** P = 0.0002). Data are presented as mean values +/− SEM. Source numerical data are available in source data.

Source data

Extended Data Fig. 7 Rescue of CD63 KO cells with WT or E217Q CD63.

a. Electron micrograph of high-pressure frozen WT or CD63 KO MNT-1 cells. Bars= 500 nm. Representative of 2 independent experiments. b. SR-IFM of WT or CD63 KO HeLa cells treated with DMSO or CK666 and co-stained for EEA1 and actin. Bars=10 µm, bars= 5 µm for magnifications. Normalized phalloidin fluorescence on EEA1-positive endosomes (n = number of endosomes, H WT DMSO n = 174, H CD63 KO#1 DMSO n = 229, H CD63 KO#2 DMSO n = 214, H WT CK666 n = 102, H CD63 KO#1 CK666 n = 104, H CD63 KO#2 CK666 n = 90; pooled across 3 independent experiments, Ordinary one-way ANOVA, ns P > 0.05, **** P < 0.001). c. SR-IFM of MNT-1 cells WT or CD63 KO and co-stained for EEA1 and actin. Bars=10 µm, bars= 5 µm for magnifications. Normalized phalloidin fluorescence on EEA1-positive endosomes (n=number of endosomes, M siCtrl n = 348, M siCD63 n = 352; pooled across 3 independent experiments, two-tailed unpaired t-test with Welch’s correction, **** P > 0.001). d. Electron micrograph of HeLa cells WT or CD63 KO treated with CK666 and analyzed by 2D EM. Bars= 200 nm. Quantification of the number of tubules/buds per endosome (n= number of MVEs, DMSO treated cells are as shown in Fig. 5f, H WT CK666 n = 15, H CD63 KO#1 CK666 n = 49, H CD63 KO#2 n = 39; pooled across 2 independent experiments, Ordinary one-way ANOVA, ns P > 0.05, * P = 0.0164, ** P = 0.0066). e. Localization of TF-chol in WT and CD63 KO cells treated with CK666. Bars = 10 µm. Representative of 2 independent experiments. f. CD63 KO HeLa cells were grown in LPDS, treated with CK666, processed for ultrathin cryo-sectioning, and immunogold labeled with D4-GFP (PAG 10 nm). Bars = 200 nm. Quantification of number of gold particles on tubules per MVE (n = number of MVEs, H WT n = 12, H CD63 KO n = 82, pooled across 2 independent experiments, two-tailed unpaired t test with Welch’s correction, **** P < 0.0001). Data are presented as mean values +/− SEM. Source numerical data are available in source data.

Source data

Extended Data Fig. 8 Uptake of TF-cholesterol and MemBright labeled sEVs by recipient cells.

a. Localization of cholesterol stained with D4-mCherry in CD63 KO Hela cells rescued with CD63 WT or E217Q. Bars=10um. Quantification of D4-mCherry integrated density of fluorescence (n = 14 H CD63 KO + CD63 WT cells, 15 H CD63 KO + CD63 E217Q cells, pooled across 2 independent experiments, two-tailed Mann-Whitney test, ns P = 0.0848). b. Cholesterol content (Amplex Red assay) of WT or CD63 KO sEVs isolated from HeLa or MNT-1 cells grown in LPDS. Cholesterol content is normalized to the number of vesicles (g per sEV) (n = 4 independent experiments, Kruskal-Wallis test for Hela, ns P = 0.8593, ns P > 0.9999; two-tailed paired Wilcoxon test for MNT-1, ns P = 0.5000). c. Relative frequency of sEVs derived from HeLa cells (grown in LPDS) WT or CD63 KO or CD63 KO rescued with CD63 WT or E217Q in function of the number of D4 associated PAG per vesicle (2-way ANOVA, only significant results are shown, 2 independent experiments for H CD63 KO#1, H CD63 KO#2 + CD63 WT, H CD63 KO#2 + CD63 E217Q, 4 independent experiments for H WT and H CD63 KO#2, ** P = 0.0016, ** P = 0.0055,**** P < 0.0001, * P = 0.0277, *** P = 0.0010, * P = 0.0167). d. Relative frequency of sEVs derived from WT or CD63 KO MNT-1 cells (grown in LPDS) in function of the number of D4 associated PAG per vesicle (data are presented as the mean of 2 independent experiments). e. DLS analysis of sEVs derived from WT or CD63 KO HeLa cells or rescued with CD63 WT or E217Q (grown in LPDS). Average number of particles/ml (n= independent experiments, n = 5 for H WT, n = 6 for H CD63 KO, n = 3 for H CD63 KO + CD63 WT and H CD63 KO + CD63 E217Q, Kruskal-Wallis test, ns P = 0.8010, ns P > 0.9999). f. Quantification of D4 associated PAG on ILVs or MVE limiting membrane in CD63 KO Hela cells or rescued with CD63 WT or E217Q (n = n° of endosomes, CD63 KO n = 43, CD63 WT n = 37, CD63 E217Q n = 59, pooled across 2 independent experiments, Kruskal-Wallis test, ns P > 0.9999, *** P = 0.0003, **** P < 0.0001). g. DLS analysis of sEVs derived from Bafilomycin-A1 treated WT or CD63 KO HeLa (grown in LPDS). Average number of particles/ml (n = 7 independent experiments, two-tailed Mann Whitney test, ns P = 0.7104). Data are presented as mean values +/− SEM. Source numerical data are available in source data.

Source data

Extended Data Fig. 9

a. Workflow used to load sEVs with TF-chol and label them with the membrane dye MemBright-640. b. CD63 KO HeLa cells were labeled with TF-chol, sEVs containing TF-chol were isolated, labeled with MemBright-640, and incubated overnight with WT or CD63 KO HeLa cells treated with U18666A. Cells were observed using a confocal microscope. Bars = 10 µm. Representative of 2 independent experiments. c. WT or CD63 KO Hela cells were labeled with TF-chol, sEVs containing TF-chol were isolated, and incubated overnight with WT Hela cells treated with vehicle or U18666A. Cells were observed using a confocal microscope. Bars = 10 µm. Representative of 2 independent experiments.

Supplementary information

Source data

Source Data Fig. 1

Statistical source data.

Source Data Fig. 2

Statistical source data.

Source Data Fig. 3

Statistical source data.

Source Data Fig. 4

Statistical source data.

Source Data Fig. 5

Statistical source data.

Source Data Fig. 6

Statistical source data.

Source Data Fig. 7

Statistical source data.

Source Data Fig. 8

Statistical source data.

Source Data Extended Data Figs. 1–8

Statistical source data Extended Data Fig. 1–8.

UnprocessedBlots_Palmullietal

Unprocessed western blots.

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

Palmulli, R., Couty, M., Piontek, M.C. et al. CD63 sorts cholesterol into endosomes for storage and distribution via exosomes. Nat Cell Biol 26, 1093–1109 (2024). https://doi.org/10.1038/s41556-024-01432-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41556-024-01432-9

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