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:

Cancer-cell-secreted miR-122 suppresses O-GlcNAcylation to promote skeletal muscle proteolysis

Abstract

A decline in skeletal muscle mass and low muscular strength are prognostic factors in advanced human cancers. Here we found that breast cancer suppressed O-linked N-acetylglucosamine (O-GlcNAc) protein modification in muscle through extracellular-vesicle-encapsulated miR-122, which targets O-GlcNAc transferase (OGT). Mechanistically, O-GlcNAcylation of ryanodine receptor 1 (RYR1) competed with NEK10-mediated phosphorylation and increased K48-linked ubiquitination and proteasomal degradation; the miR-122-mediated decrease in OGT resulted in increased RYR1 abundance. We further found that muscular protein O-GlcNAcylation was regulated by hypoxia and lactate through HIF1A-dependent OGT promoter activation and was elevated after exercise. Suppressed O-GlcNAcylation in the setting of cancer, through increasing RYR1, led to higher cytosolic Ca2+ and calpain protease activation, which triggered cleavage of desmin filaments and myofibrillar destruction. This was associated with reduced skeletal muscle mass and contractility in tumour-bearing mice. Our findings link O-GlcNAcylation to muscular protein homoeostasis and contractility and reveal a mechanism of cancer-associated muscle dysregulation.

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: Cancer-secreted EVs suppress O-GlcNAcylation in skeletal muscle through miR-122.
Fig. 2: O-GlcNAcylation increases RYR1 ubiquitination and proteasomal degradation.
Fig. 3: miR-122 regulates RYR1 and SERCA3 by directly targeting OGT.
Fig. 4: OGT expression is induced by lactate and hypoxia through HIF1A activation.
Fig. 5: Cytosolic Ca2+ level is regulated by the miR-122–OGT–RYR1 axis.
Fig. 6: BC-secreted miR-122 induces calpain activity and protein cleavage in myotubes through regulating RYR1.
Fig. 7: Regulation of calpain activity and protein cleavage in skeletal muscle by tumour-derived EVs.
Fig. 8: Exogenous OGT expression alleviates BC-associated proteolysis and skeletal muscle loss.

Similar content being viewed by others

Data availability

The RNA-seq data that support the findings of this study have been deposited in the Gene Expression Omnibus (GEO) under accession code GSE156909. MS data have been deposited in ProteomeXchange Consortium via the PRIDE partner repository with the primary accession code PXD021232. The previously published GEO dataset GSE50429 was re-analysed for miRNA levels in the cells and EVs of MDA-MB-231 and MCF-10A (Supplementary Table 1). The GSEA data in Fig. 1b involved re-analyses of the Reactome pathway datasets R-HSA-1630316 and R-HSA-445355 as well as the Kyoto Encyclopedia of Genes and Genomes (KEGG) pathway dataset hsa04020. All other data supporting the findings of this study are available from the corresponding authors on reasonable request. Source data are provided with this paper.

References

  1. Tkach, M. & Thery, C. Communication by extracellular vesicles: where we are and where we need to go. Cell 164, 1226–1232 (2016).

    Article  CAS  PubMed  Google Scholar 

  2. Becker, A. et al. Extracellular vesicles in cancer: cell-to-cell mediators of metastasis. Cancer Cell 30, 836–848 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Redzic, J. S., Balaj, L., van der Vos, K. E. & Breakefield, X. O. Extracellular RNA mediates and marks cancer progression. Semin. Cancer Biol. 28, 14–23 (2014).

    Article  CAS  PubMed  Google Scholar 

  4. Wang, S. E. Extracellular vesicles and metastasis. Cold Spring Harbor Perspect. Med. 10, a037275 (2020).

    Article  CAS  Google Scholar 

  5. Sato, S. & Weaver, A. M. Extracellular vesicles: important collaborators in cancer progression. Essays Biochem. 62, 149–163 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  6. He, W. A. et al. Microvesicles containing miRNAs promote muscle cell death in cancer cachexia via TLR7. Proc. Natl Acad. Sci. USA 111, 4525–4529 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Zhang, G. et al. Tumor induces muscle wasting in mice through releasing extracellular Hsp70 and Hsp90. Nat. Commun. 8, 589 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  8. Fearon, K., Arends, J. & Baracos, V. Understanding the mechanisms and treatment options in cancer cachexia. Nat. Rev. Clin. Oncol. 10, 90–99 (2013).

    Article  CAS  PubMed  Google Scholar 

  9. Argiles, J. M., Busquets, S., Stemmler, B. & Lopez-Soriano, F. J. Cancer cachexia: understanding the molecular basis. Nat. Rev. Cancer 14, 754–762 (2014).

    Article  CAS  PubMed  Google Scholar 

  10. Siegel, R. L., Miller, K. D. & Jemal, A. Cancer statistics, 2018. CA Cancer J. Clin. 68, 7–30 (2018).

    Article  PubMed  Google Scholar 

  11. Fox, K. M., Brooks, J. M., Gandra, S. R., Markus, R. & Chiou, C. F. Estimation of cachexia among cancer patients based on four definitions. J. Oncol. 2009, 693458 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  12. Fearon, K. C., Glass, D. J. & Guttridge, D. C. Cancer cachexia: mediators, signaling, and metabolic pathways. Cell Metab. 16, 153–166 (2012).

    Article  CAS  PubMed  Google Scholar 

  13. Caan, B. J. et al. Association of muscle and adiposity measured by computed tomography with survival in patients with nonmetastatic breast cancer. JAMA Oncol. 4, 798–804 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  14. Rier, H. N. et al. Low muscle attenuation is a prognostic factor for survival in metastatic breast cancer patients treated with first line palliative chemotherapy. Breast 31, 9–15 (2017).

    Article  PubMed  Google Scholar 

  15. Shachar, S. S. et al. Skeletal muscle measures as predictors of toxicity, hospitalization, and survival in patients with metastatic breast cancer receiving taxane-based chemotherapy. Clin. Cancer Res. 23, 658–665 (2017).

    Article  CAS  PubMed  Google Scholar 

  16. Prado, C. M. et al. Sarcopenia as a determinant of chemotherapy toxicity and time to tumor progression in metastatic breast cancer patients receiving capecitabine treatment. Clin. Cancer Res. 15, 2920–2926 (2009).

    Article  CAS  PubMed  Google Scholar 

  17. Villasenor, A. et al. Prevalence and prognostic effect of sarcopenia in breast cancer survivors: the HEAL Study. J. Cancer Surviv. 6, 398–406 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  18. Mueller, T. C., Bachmann, J., Prokopchuk, O., Friess, H. & Martignoni, M. E. Molecular pathways leading to loss of skeletal muscle mass in cancer cachexia—can findings from animal models be translated to humans? BMC Cancer 16, 75 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  19. Sandri, M. Autophagy in skeletal muscle. FEBS Lett. 584, 1411–1416 (2010).

    Article  CAS  PubMed  Google Scholar 

  20. Aweida, D., Rudesky, I., Volodin, A., Shimko, E. & Cohen, S. GSK3-β promotes calpain-1-mediated desmin filament depolymerization and myofibril loss in atrophy. J. Cell Biol. 217, 3698–3714 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Cohen, S. Role of calpains in promoting desmin filaments depolymerization and muscle atrophy. Biochim. Biophys. Acta Mol. Cell. Res. 1867, 118788 (2020).

    Article  CAS  PubMed  Google Scholar 

  22. Alderton, J. M. & Steinhardt, R. A. Calcium influx through calcium leak channels is responsible for the elevated levels of calcium-dependent proteolysis in dystrophic myotubes. J. Biol. Chem. 275, 9452–9460 (2000).

    Article  CAS  PubMed  Google Scholar 

  23. Lanner, J. T., Georgiou, D. K., Joshi, A. D. & Hamilton, S. L. Ryanodine receptors: structure, expression, molecular details, and function in calcium release. Cold Spring Harb. Perspect. Biol. 2, a003996 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Zalk, R., Lehnart, S. E. & Marks, A. R. Modulation of the ryanodine receptor and intracellular calcium. Annu. Rev. Biochem. 76, 367–385 (2007).

    Article  CAS  PubMed  Google Scholar 

  25. Andersson, D. C. et al. Ryanodine receptor oxidation causes intracellular calcium leak and muscle weakness in aging. Cell Metab. 14, 196–207 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Waning, D. L. et al. Excess TGF-β mediates muscle weakness associated with bone metastases in mice. Nat. Med. 21, 1262–1271 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Bellinger, A. M. et al. Hypernitrosylated ryanodine receptor calcium release channels are leaky in dystrophic muscle. Nat. Med. 15, 325–330 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Zalk, R. et al. Structure of a mammalian ryanodine receptor. Nature 517, 44–49 (2015).

    Article  CAS  PubMed  Google Scholar 

  29. Robinson, R., Carpenter, D., Shaw, M. A., Halsall, J. & Hopkins, P. Mutations in RYR1 in malignant hyperthermia and central core disease. Hum. Mutat. 27, 977–989 (2006).

    Article  CAS  PubMed  Google Scholar 

  30. Yang, X. & Qian, K. Protein O-GlcNAcylation: emerging mechanisms and functions. Nat. Rev. Mol. Cell Biol. 18, 452–465 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Hardiville, S. & Hart, G. W. Nutrient regulation of signaling, transcription, and cell physiology by O-GlcNAcylation. Cell Metab. 20, 208–213 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Martinez, M. R., Dias, T. B., Natov, P. S. & Zachara, N. E. Stress-induced O-GlcNAcylation: an adaptive process of injured cells. Biochem. Soc. Trans. 45, 237–249 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Hart, G. W., Slawson, C., Ramirez-Correa, G. & Lagerlof, O. Cross talk between O-GlcNAcylation and phosphorylation: roles in signaling, transcription, and chronic disease. Annu. Rev. Biochem. 80, 825–858 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Benediktsson, A. M., Schachtele, S. J., Green, S. H. & Dailey, M. E. Ballistic labeling and dynamic imaging of astrocytes in organotypic hippocampal slice cultures. J. Neurosci. Methods 141, 41–53 (2005).

    Article  PubMed  Google Scholar 

  35. Fong, M. Y. et al. Breast-cancer-secreted miR-122 reprograms glucose metabolism in premetastatic niche to promote metastasis. Nat. Cell Biol. 17, 183–194 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Wu, X. et al. De novo sequencing of circulating miRNAs identifies novel markers predicting clinical outcome of locally advanced breast cancer. J. Transl. Med 10, 42 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Zhou, W. et al. Cancer-secreted miR-105 destroys vascular endothelial barriers to promote metastasis. Cancer Cell 25, 501–515 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Shen, M. et al. Chemotherapy-induced extracellular vesicle miRNAs promote breast cancer stemness by targeting ONECUT2. Cancer Res. 79, 3608–3621 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Kao, H. J. et al. A two-layered machine learning method to identify protein O-GlcNAcylation sites with O-GlcNAc transferase substrate motifs. BMC Bioinformatics 16, S10 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  40. Chen, Q., Chen, Y., Bian, C., Fujiki, R. & Yu, X. TET2 promotes histone O-GlcNAcylation during gene transcription. Nature 493, 561–564 (2013).

    Article  CAS  PubMed  Google Scholar 

  41. Hornbeck, P. V., Chabra, I., Kornhauser, J. M., Skrzypek, E. & Zhang, B. PhosphoSite: a bioinformatics resource dedicated to physiological protein phosphorylation. Proteomics 4, 1551–1561 (2004).

    Article  CAS  PubMed  Google Scholar 

  42. Povlsen, L. K. et al. Systems-wide analysis of ubiquitylation dynamics reveals a key role for PAF15 ubiquitylation in DNA-damage bypass. Nat. Cell Biol. 14, 1089–1098 (2012).

    Article  CAS  PubMed  Google Scholar 

  43. Ameln, H. et al. Physiological activation of hypoxia inducible factor-1 in human skeletal muscle. FASEB J. 19, 1009–1011 (2005).

    Article  CAS  PubMed  Google Scholar 

  44. Gaudelli, N. M. et al. Programmable base editing of A•T to G•C in genomic DNA without DNA cleavage. Nature 551, 464–471 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Hoshino, A. et al. Tumour exosome integrins determine organotropic metastasis. Nature 527, 329–335 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Chow, A. et al. Macrophage immunomodulation by breast cancer-derived exosomes requires Toll-like receptor 2-mediated activation of NF-κB. Sci. Rep. 4, 5750 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Buren, S. et al. Regulation of OGT by URI in response to glucose confers c-MYC-dependent survival mechanisms. Cancer Cell 30, 290–307 (2016).

    Article  CAS  PubMed  Google Scholar 

  48. Yi, W. et al. Phosphofructokinase 1 glycosylation regulates cell growth and metabolism. Science 337, 975–980 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Slawson, C. & Hart, G. W. O-GlcNAc signalling: implications for cancer cell biology. Nat. Rev. Cancer 11, 678–684 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Yan, Z. et al. Structure of the rabbit ryanodine receptor RyR1 at near-atomic resolution. Nature 517, 50–55 (2015).

    Article  CAS  PubMed  Google Scholar 

  51. Ruan, H. B. et al. Calcium-dependent O-GlcNAc signaling drives liver autophagy in adaptation to starvation. Genes Dev. 31, 1655–1665 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Zou, L. et al. The identification of a novel calcium-dependent link between NAD+ and glucose deprivation-induced increases in protein O-GlcNAcylation and ER stress. Front Mol. Biosci. 8, 780865 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Favier, F. B., Britto, F. A., Freyssenet, D. G., Bigard, X. A. & Benoit, H. HIF-1-driven skeletal muscle adaptations to chronic hypoxia: molecular insights into muscle physiology. Cell. Mol. Life Sci. 72, 4681–4696 (2015).

    Article  CAS  PubMed  Google Scholar 

  54. Cai, D. et al. IKKβ/NF-κB activation causes severe muscle wasting in mice. Cell 119, 285–298 (2004).

    Article  CAS  PubMed  Google Scholar 

  55. Wani, W. Y., Chatham, J. C., Darley-Usmar, V., McMahon, L. L. & Zhang, J. O-GlcNAcylation and neurodegeneration. Brain Res. Bull. 133, 80–87 (2017).

    Article  CAS  PubMed  Google Scholar 

  56. Lon, H. K. et al. Pharmacokinetics, safety, tolerability, and pharmacodynamics of alicapistat, a selective inhibitor of human calpains 1 and 2 for the treatment of Alzheimer disease: an overview of phase 1 studies. Clin. Pharm. Drug Dev. 8, 290–303 (2019).

    Article  CAS  Google Scholar 

  57. Tsuyada, A. et al. CCL2 mediates cross-talk between cancer cells and stromal fibroblasts that regulates breast cancer stem cells. Cancer Res. 72, 2768–2779 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. 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 

  59. Shin, H. R. et al. Small-molecule inhibitors of histone deacetylase improve CRISPR-based adenine base editing. Nucleic Acids Res. 49, 2390–2399 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Kim, S., Bae, T., Hwang, J. & Kim, J. S. Rescue of high-specificity Cas9 variants using sgRNAs with matched 5′ nucleotides. Genome Biol. 18, 218 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  61. Gray, J. T. & Zolotukhin, S. Design and construction of functional AAV vectors. Methods Mol. Biol. 807, 25–46 (2011).

    Article  CAS  PubMed  Google Scholar 

  62. Sato, K., Pollock, N. & Stowell, K. M. Functional studies of RYR1 mutations in the skeletal muscle ryanodine receptor using human RYR1 complementary DNA. Anesthesiology 112, 1350–1354 (2010).

    Article  CAS  PubMed  Google Scholar 

  63. Yan, W. et al. Cancer-cell-secreted exosomal miR-105 promotes tumour growth through the MYC-dependent metabolic reprogramming of stromal cells. Nat. Cell Biol. 20, 597–609 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Stoner, S. A. et al. High sensitivity flow cytometry of membrane vesicles. Cytometry A 89, 196–206 (2016).

    Article  CAS  PubMed  Google Scholar 

  65. Thery, C. et al. Minimal information for studies of extracellular vesicles 2018 (MISEV2018): a position statement of the International Society for Extracellular Vesicles and update of the MISEV2014 guidelines. J. Extracell. Vesicles 7, 1535750 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  66. Lim, K. L. et al. Parkin mediates nonclassical, proteasomal-independent ubiquitination of synphilin-1: implications for Lewy body formation. J. Neurosci. 25, 2002–2009 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Stephens, J. et al. Functional analysis of RYR1 variants linked to malignant hyperthermia. Temperature 3, 328–339 (2016).

    Article  Google Scholar 

  68. Shevchenko, A., Wilm, M., Vorm, O. & Mann, M. Mass spectrometric sequencing of proteins silver-stained polyacrylamide gels. Anal. Chem. 68, 850–858 (1996).

    Article  CAS  PubMed  Google Scholar 

  69. DeBalsi, K. L. et al. Targeted metabolomics connects thioredoxin-interacting protein (TXNIP) to mitochondrial fuel selection and regulation of specific oxidoreductase enzymes in skeletal muscle. J. Biol. Chem. 289, 8106–8120 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Deacon, R. M. Measuring the strength of mice. J. Vis. Exp. 76, e2610 (2013).

    Google Scholar 

  71. Svensson, K. et al. Combined overexpression of SIRT1 and knockout of GCN5 in adult skeletal muscle does not affect glucose homeostasis or exercise performance in mice. Am. J. Physiol. Endocrinol. Metab. 318, E145–E151 (2020).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

This work was supported by the National Institutes of Health (NIH) National Cancer Institute (NCI) grants R01CA218140 (S.E.W.) and R01CA206911 (S.E.W.), National Institute of Arthritis and Musculoskeletal and Skin Diseases (NIAMS) grant R21AR072882 (S.S.) and National Institute of General Medical Sciences (NIGMS) grant R01GM102362 (D.W.). Research reported in this publication includes work performed in core facilities supported by the NIH NCI under grant numbers P30CA23100 (UC San Diego Cancer Center) and P30CA030199 (Sanford Burnham Prebys Cancer Center Flow Cytometry Core). We thank M. C. Hogan for constructive suggestions,A. de Maio and his group for assistance with EV characterization by NTA, X. Yu for kindly providing the expression plasmids of WT and enzymatic dead (G482S) human OGT, K. Stowell for kindly providing the pcDNA3.0-RYR1 expression plasmid and Q. Zhang for kindly providing the expression plasmids of WT and dpA mutant of HIF1A.

Author information

Authors and Affiliations

Authors

Contributions

S.E.W., S.S. and W.Y. conceived ideas. S.E.W., S.S., W.Y. and J.D.E. contributed to project planning and manuscript writing. W.Y. and S.E.W. designed and performed most of the experiments. S.L., A.L. and S.S. performed muscle mechanics tests and analyses. M.G. assisted with MS analysis. M.C. and A.R.C. assisted with cell line construction. M.C., X.R., L.J. and Y.W. assisted with mouse experiments. Y.Q. assisted with data analysis. E.D. and J.P.N. performed EV characterization by flow cytometry. D.P.P. assisted with tissue processing and histological analyses. D.W. contributed to characterization of RYR1 PTMs and functional domains.

Corresponding authors

Correspondence to Wei Yan, Simon Schenk or Shizhen Emily Wang.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Cell Biology thanks Shenhav Cohen, Xiaoyong Yang and the other, anonymous, reviewer(s) 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 EVs used in this study.

(a) Flow cytometry showing size distribution and expression of GFP and selected tetraspanins in EVs collected from MDA-MB-231/Lck-GFP and MCF-10A/Lck-GFP cells. GFP-positive EVs are displayed in green on dot plots of tetraspanin expression. (b) GFP fluorescence (green) in GA sections from NSG mice receiving tail vein injections of Lck-GFP-labelled EVs or bearing MDA-MB-231/Lck-GFP tumours. Scale bar=100 µm. (c) Western blots of nuclear, cytosolic, and mitochondrial fractions prepared from GA of tumour-free or MDA-MB-231/Lck-GFP-tumour-bearing mice showing the subcellular localization of GFP in the cytosol. Histone H3, GAPDH, and COXIV served as the markers for the nuclear, cytosolic, and mitochondrial fractions, respectively. (d) Western blots of indicated whole cell lysates and EVs showing EV markers and a Golgi marker (GM130, as a negative control for EV-specific proteins). (e,g) NTA of indicated EVs showing size distribution (n=3 biological replicates). Western blots showing Rab27a knockdown in 4T1 cells were included next to the NTA in (g). (f,h) Levels of indicated EV proteins and RT-qPCR-determined miR-122 levels (normalized to a synthetic cel-miR-39-3p spike-in control added into each fraction) in OptiPrep gradient fractions (n=3 biological replicates). In bar graphs, values are shown as mean ± SEM. Unprocessed original scans of blots are shown in source data file.

Source data

Extended Data Fig. 2 MS-identified O‑GlcNAcylation of RYR1 and SERCA3 in skeletal muscle.

(a) Western blots showing O‑GlcNAcylation in lung, liver, adipose tissue, kidney, and brain of tumour-free and tumour-bearing mice described in Fig. 1e. (b) Western blots of GA from mice that have received repeated injections of indicated EVs as in Fig. 1e. (c) The MS/MS spectra of O-GlcNAcylated peptides of mouse RYR1 (residues 4324–4353) and SERCA3 (residues 534–550) with the triply charged precursor ion m/z 638.9881 (M + 3H)3+. The c- and z-type product ions were assigned. The O-GlcNAc oxonium ion (m/z, 204.09) and a series of its fragments (m/z, 186.08, 168.06, 138.05, and 126.05) were also assigned. (d) OGTSite analysis predicting the herein identified O-GlcNAcylation sites in RYR1 and SERCA3 of human and mouse.

Source data

Extended Data Fig. 3 MS-identified RYR1 phosphorylation and the predicted kinases and nearby ubiquitylation sites.

(a) IF of C2C12 myotubes showing colocalization of exogenously expressed HA-tagged RYR1 and an ER marker PDI. DAPI stains the nuclei. Bar=100 μm. (b) MS-identified phosphorylation in immunoprecipitated, endogenous RYR1 from C2C12 with OGT knockdown but not from control C2C12. (c) Scansite 4.0 analysis predicting kinases that can potentially phosphorylate the identified threonine. (d) PhosphoSitePlus analysis showing published phosphorylation and ubiquitylation sites in the surrounding region in human and mouse RYR1.

Extended Data Fig. 4 O-GlcNAcylation increases the abundance of SERCA3.

(a,b) C2C12 expressing FLAG-tagged wild-type SERCA3 or T534A mutant and with OGT overexpression (a) or knockdown (b) were analysed by western blots. Unprocessed original scans of blots are shown in source data file.

Source data

Extended Data Fig. 5 miR-122 regulates O-GlcNAcylation in myotubes in the presence of LA or under hypoxia in an HIF1A-dependent manner.

(a) GFP signals in C2C12 myotubes treated with indicated EVs for 24 h indicating EV uptake. Bar=100 μm. (b,c) Western blots showing indicated protein levels in treated C2C12 (b) and L6 myotubes (c) grown in the absence of LA and under 21% O2. (d,e) Protein levels in L6 treated as indicated and grown in 2 mM LA (d) or under 1% O2 (e). (f) Western blots showing indicated protein levels in L6 myotubes grown under indicated concentrations of LA for 24 h. (g,h) Protein levels (g) and ROS levels (h) in C2C12 myotubes treated with HCl to match the pH of indicated concentrations of LA (2mM: 2 mM LA). Values are shown as mean ± SEM in h (one-way ANOVA with Tukey’s multiple comparisons test, n=3 biologically independent wells of cells). ns: not significant. (i,j) Protein levels in L6 treated with 2 mM LA (for 2 h, for 2 h followed by 22 h of incubation in LA-free medium, or for 24 h) and cultured under 21% or 1% O2, and transfected with control siRNA or siRNA targeting HIF1A (i) or OGT (j), as indicated. Unprocessed original scans of blots are shown in source data file.

Source data

Extended Data Fig. 6 IHC of skeletal muscle and calpain activity assessments after aerobic exercise.

(a,b) IHC staining of GA from EV-treated (a) or tumour-bearing (b) mice. For tumour-bearing mice, muscle was analysed when tumours reached ~300 mm3 in all groups. Bar=100 μm. (c,d) GA were collected from female C57BL/6J mice immediately after 1 h of high-intensity treadmill running (aerobic exercise/AEX) and from mice under sedentary lifestyle (SED). Tissue lysates were assessed for calpain protease activity (c; unpaired two-tailed t-test, n=6 mice per group) and by western blots for indicated protein cleavage (d). The boxes in the box-and-whiskers plot show the median (centre line) and the quartile range (25-75%), and the whiskers extend from the quartile to the minimum and maximum values. ***P<0.0001. Unprocessed original scans of blots are shown in source data file.

Source data

Extended Data Fig. 7 AAV-delivered gene expression in various organs.

(a) Fluorescent microscopy showing the presence or absence of GFP in various organs. Bar=100 μm. (b) Western blots of indicated tumour lysates from AAV-treated mice showing lack of GFP expression and lack of changes in OGT protein levels. Unprocessed original scans of blots are shown in source data file.

Source data

Extended Data Fig. 8 Structure of RYR1 monomer showing the O-GlcNAcylated threonine possibly residing in an EF-hand domain.

(a) Previously resolved structure of rabbit RYR1 monomer (PDB: 3J8H). Structure is displayed by PyMOL. (b) Partial structure showing the position of A4330 in rabbit RYR1 (equivalent to the O-GlcNAcylated threonine in human/mouse RYR1) within the linker region between central domain and channel domain. (c) JPred-predicted secondary structure of the peptide region with poor EM quality showing the O-GlcNAcylated threonine residing in the E alpha-helix of an EF hand.

Extended Data Fig. 9 Method of vesicle flow cytometry (VFC).

(a) Schematic of VFC workflow. Vesicle flow cytometry (vFCTM) is a homogeneous assay in which a cell-free sample, prepared by centrifugation, is stained with a fluorogenic membrane stain and one or more additional fluorescence probes then analysed by flow cytometry with detection triggered by membrane fluorescence. (b) VFC gating. Flow cytometry data is gated on time (left, to eliminate a fluidics-related background that occurs at the beginning of each sample measured in the CytoFlex), Membrane Stain fluorescence pulse height vs area (middle, vFRed-H v- A, which can help resolve EV-associated events from some sources of background), and by a gate (right, Vesicles) on Membrane fluorescence vs VSSC, which selects events with characteristic fluorescence and light scatter characteristics. (c) Vesicle size estimation. Vesicle size was estimated from the relationship between fluorescence intensity and vesicle surface area, as determined by the staining and analysis of a well-characterized synthetic vesicle size standard (Lipo100, Cellarcus Biosciences). The Lipo100 diameter distribution measured by NTA (upper left) was used to calculate the surface area distribution (upper middle), which, when compared to the fluorescence intensity distribution after staining with vFRed (upper right), showed a linear relationship (lower left) that was used to convert the arbitrary vFRed fluorescence intensity into units of estimated equivalent surface area (lower middle) and diameter (lower right). (d) vFC Buffer +reagent and dilution controls. The single EV specificity of the measured marker positive events (left) was demonstrated by comparison to a Buffer +vFRed (no vesicle) sample and by serial dilution (middle) which demonstrated a ~2 log dynamic range (right) with no detectable coincidence/swarm. (e) Negligible Fc Receptor-mediated antibody binding as measured using an isotype-matched irrelevant control antibody. A representative EV sample immunostained with the indicated PE-labelled antibodies. MESF: mean equivalent soluble fluorochromes.

Supplementary information

Supplementary Information

Supplementary Methods: vesicle flow cytometry.

Reporting Summary

Supplementary Tables

Supplementary Tables 1–5.

Source data

Source Data Fig. 1

Statistical source data.

Source Data Fig. 1

Unprocessed western blots.

Source Data Fig. 2

Statistical source data.

Source Data Fig. 2

Unprocessed western blots.

Source Data Fig. 3

Statistical source data.

Source Data Fig. 3

Unprocessed western blots.

Source Data Fig. 4

Statistical source data.

Source Data Fig. 4

Unprocessed western blots.

Source Data Fig. 5

Statistical source data.

Source Data Fig. 5

Unprocessed western blots.

Source Data Fig. 6

Statistical source data.

Source Data Fig. 6

Unprocessed western blots.

Source Data Fig. 7

Statistical source data.

Source Data Fig. 7

Unprocessed western blots.

Source Data Fig. 8

Statistical source data.

Source Data Fig. 8

Unprocessed western blots.

Source Data Extended Data Fig. 1

Statistical source data.

Source Data Extended Data Fig. 1

Unprocessed western blots.

Source Data Extended Data Fig. 2

Unprocessed western blots.

Source Data Extended Data Fig. 4

Statistical source data.

Source Data Extended Data Fig. 4

Unprocessed western blots.

Source Data Extended Data Fig. 5

Statistical source data.

Source Data Extended Data Fig. 5

Unprocessed western blots.

Source Data Extended Data Fig. 6

Statistical source data.

Source Data Extended Data Fig. 6

Unprocessed western blots.

Source Data Extended Data Fig. 7

Unprocessed western blots.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Yan, W., Cao, M., Ruan, X. et al. Cancer-cell-secreted miR-122 suppresses O-GlcNAcylation to promote skeletal muscle proteolysis. Nat Cell Biol 24, 793–804 (2022). https://doi.org/10.1038/s41556-022-00893-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41556-022-00893-0

This article is cited by

Search

Quick links

Nature Briefing: Cancer

Sign up for the Nature Briefing: Cancer newsletter — what matters in cancer research, free to your inbox weekly.

Get what matters in cancer research, free to your inbox weekly. Sign up for Nature Briefing: Cancer