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:

O-GlcNAcylation determines the translational regulation and phase separation of YTHDF proteins

Abstract

N6-methyladenosine (m6A) is the most abundant internal mRNA nucleotide modification in mammals, regulating critical aspects of cell physiology and differentiation. The YTHDF proteins are the primary readers of m6A modifications and exert physiological functions of m6A in the cytosol. Elucidating the regulatory mechanisms of YTHDF proteins is critical to understanding m6A biology. Here we report a mechanism that protein post-translational modifications control the biological functions of the YTHDF proteins. We find that YTHDF1 and YTHDF3, but not YTHDF2, carry high levels of nutrient-sensing O-GlcNAc modifications. O-GlcNAcylation attenuates the translation-promoting function of YTHDF1 and YTHDF3 by blocking their interactions with proteins associated with mRNA translation. We further demonstrate that O-GlcNAc modifications on YTHDF1 and YTHDF3 regulate the assembly, stability and disassembly of stress granules to enable better recovery from stress. Therefore, our results discover an important regulatory pathway of YTHDF functions, adding an additional layer of complexity to the post-transcriptional regulation function of mRNA m6A.

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: YTHDF1/3 are modified by O-GlcNAc.
Fig. 2: Reversible O-GlcNAcylation sites of YTHDF1/3.
Fig. 3: O-GlcNAcylation of YTHDF1/3 represses their protein interaction.
Fig. 4: O-GlcNAcylation of YTHDF1 represses target mRNA translation efficiency.
Fig. 5: Dynamic O-GlcNAcylation of YTHDF1/3 regulates m6A mRNA translation efficiency in a cell cycle-dependent manner.
Fig. 6: O-GlcNAcylation of YTHDF1/3 increases their dynamic nature in SGs.
Fig. 7: Validation of the effect of O-GlcNAcylation on YTHDF1/3 condensation in vitro.

Similar content being viewed by others

Data availability

The MS proteomics data have been deposited to the ProteomeXchange Consortium (https://proteomecentral.proteomexchange.org) via the iProX partner repository104 with the dataset identifier PXD037017. High-throughput sequencing data of CLIP and RiboLace can be accessed in the Gene Expression Omnibus under the accession number GSE216570. Previously published datasets that were used for reanalysis33,35,72. Swiss-Prot human database (release 10 December 2018) containing 26,448 entries from UniProt was used for proteomics analysis. Previously published data used for design of constructs and plasmids are available from RefSeq under the accession numbers NM_017798.4, NM_016258.3 and NM_152758.6. 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. Dominissini, D. et al. Topology of the human and mouse m6A RNA methylomes revealed by m6A-seq. Nature 485, 201–206 (2012).

    Article  CAS  PubMed  Google Scholar 

  2. Lin, Z. et al. Mettl3-/Mettl14-mediated mRNA N6-methyladenosine modulates murine spermatogenesis. Cell Res. 27, 1216–1230 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Du, Y. et al. SUMOylation of the m6A-RNA methyltransferase METTL3 modulates its function. Nucleic Acids Res. 46, 5195–5208 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Wei, J. et al. Differential m6A, m6Am, and m1A demethylation mediated by FTO in the cell nucleus and cytoplasm. Mol. Cell 71, 973–985 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Liu, J. et al. Landscape and regulation of m6A and m6Am methylome across human and mouse tissues. Mol. Cell 77, 426–440 (2020).

    Article  CAS  PubMed  Google Scholar 

  6. Liu, J. et al. A METTL3–METTL14 complex mediates mammalian nuclear RNA N6-adenosine methylation. Nat. Chem. Biol. 10, 93–95 (2014).

    Article  CAS  PubMed  Google Scholar 

  7. Ping, X. L. et al. Mammalian WTAP is a regulatory subunit of the RNA N6-methyladenosine methyltransferase. Cell Res. 24, 177–189 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Wang, Y. et al. N6-methyladenosine modification destabilizes developmental regulators in embryonic stem cells. Nat. Cell Biol. 16, 191–198 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Sledz, P. & Jinek, M. Structural insights into the molecular mechanism of the m6A writer complex. eLife 5, e18434 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  10. Wang, P., Doxtader, K. A. & Nam, Y. Structural basis for cooperative function of Mettl3 and Mettl14 methyltransferases. Mol. Cell 63, 306–317 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Wang, X. et al. Structural basis of N6-adenosine methylation by the METTL3–METTL14 complex. Nature 534, 575–578 (2016).

    Article  CAS  PubMed  Google Scholar 

  12. Jia, G. et al. N6-methyladenosine in nuclear RNA is a major substrate of the obesity-associated FTO. Nat. Chem. Biol. 7, 885–887 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Zheng, G. et al. ALKBH5 is a mammalian RNA demethylase that impacts RNA metabolism and mouse fertility. Mol. Cell 49, 18–29 (2013).

    Article  CAS  PubMed  Google Scholar 

  14. Batista, P. J. et al. m6A RNA modification controls cell fate transition in mammalian embryonic stem cells. Cell Stem Cell 15, 707–719 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Geula, S. et al. Stem cells. m6A mRNA methylation facilitates resolution of naive pluripotency toward differentiation. Science 347, 1002–1006 (2015).

    Article  CAS  PubMed  Google Scholar 

  16. Batista, P. J. The RNA modification N6-methyladenosine and its implications in human disease. Genomics Proteom. Bioinform. 15, 154–163 (2017).

    Article  CAS  Google Scholar 

  17. De Jesus, D. F. et al. m6A mRNA methylation regulates human β-cell biology in physiological states and in type 2 diabetes. Nat. Metab. 1, 765–774 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  18. Ianniello, Z., Paiardini, A. & Fatica, A. N6-methyladenosine (m6A): a promising new molecular target in acute myeloid leukemia. Front. Oncol. 9, 251 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  19. Yang, Y. et al. Glucose is involved in the dynamic regulation of m6A in patients with type 2 diabetes. J. Clin. Endocrinol. Metab. 104, 665–673 (2019).

    Article  PubMed  Google Scholar 

  20. Huang, H. L., Weng, H. Y., Deng, X. L. & Chen, J. J. RNA modifications in cancer: functions, mechanisms, and therapeutic implications. Annu. Rev. Cancer Biol. 4, 221–240 (2020).

    Article  Google Scholar 

  21. He, P. C. & He, C. m6A RNA methylation: from mechanisms to therapeutic potential. EMBO J. 40, e105977 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Meyer, K. D. et al. 5′ UTR m6A promotes cap-independent translation. Cell 163, 999–1010 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Choi, J. et al. N6-methyladenosine in mRNA disrupts tRNA selection and translation–elongation dynamics. Nat. Struct. Mol. Biol. 23, 110–115 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Barbieri, I. et al. Promoter-bound METTL3 maintains myeloid leukaemia by m6A-dependent translation control. Nature 552, 126–131 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Slobodin, B. et al. Transcription impacts the efficiency of mRNA translation via co-transcriptional N6-adenosine methylation. Cell 169, 326–337 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Huang, H. et al. Recognition of RNA N6-methyladenosine by IGF2BP proteins enhances mRNA stability and translation. Nat. Cell Biol. 20, 285–295 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Zhou, J. et al. N6-methyladenosine guides mRNA alternative translation during integrated stress response. Mol. Cell 69, 636–647 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Mao, Y. et al. m6A in mRNA coding regions promotes translation via the RNA helicase-containing YTHDC2. Nat. Commun. 10, 5332 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  29. Xu, C. et al. Structural basis for selective binding of m6A RNA by the YTHDC1 YTH domain. Nat. Chem. Biol. 10, 927–929 (2014).

    Article  CAS  PubMed  Google Scholar 

  30. Shi, H., Wei, J. & He, C. Where, when, and how: context-dependent functions of RNA methylation writers, readers, and erasers. Mol. Cell 74, 640–650 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Li, Y., Bedi, R. K., Moroz-Omori, E. V. & Caflisch, A. Structural and dynamic insights into redundant function of YTHDF proteins. J. Chem. Inf. Model. 60, 5932–5935 (2020).

    Article  CAS  PubMed  Google Scholar 

  32. Wang, X. et al. N6-methyladenosine-dependent regulation of messenger RNA stability. Nature 505, 117–120 (2014).

    Article  PubMed  Google Scholar 

  33. Wang, X. et al. N6-methyladenosine modulates messenger RNA translation efficiency. Cell 161, 1388–1399 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Shi, H. et al. YTHDF3 facilitates translation and decay of N6-methyladenosine-modified RNA. Cell Res. 27, 315–328 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Zaccara, S. & Jaffrey, S. R. A unified model for the function of YTHDF proteins in regulating m6A-modified mRNA. Cell 181, 1582–1595 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Liu, J. et al. m6A mRNA methylation regulates AKT activity to promote the proliferation and tumorigenicity of endometrial cancer. Nat. Cell Biol. 20, 1074–1083 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Zou, Z., Sepich-Poore, C., Zhou, X., Wei, J. & He, C. The mechanism underlying redundant functions of the YTHDF proteins. Genome Biol. 24, 17 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Kontur, C., Jeong, M., Cifuentes, D. & Giraldez, A. J. Ythdf m6A readers function redundantly during zebrafish development. Cell Rep. 33, 108598 (2020).

    Article  CAS  PubMed  Google Scholar 

  39. Heckel, D. et al. Novel immunogenic antigen homologous to hyaluronidase in meningioma. Hum. Mol. Genet. 7, 1859–1872 (1998).

    Article  CAS  PubMed  Google Scholar 

  40. Lubas, W. A. & Hanover, J. A. Functional expression of O-linked GlcNAc transferase. Domain structure and substrate specificity. J. Biol. Chem. 275, 10983–10988 (2000).

    Article  CAS  PubMed  Google Scholar 

  41. Comtesse, N., Maldener, E. & Meese, E. Identification of a nuclear variant of MGEA5, a cytoplasmic hyaluronidase and a β-N-acetylglucosaminidase. Biochem. Biophys. Res. Commun. 283, 634–640 (2001).

    Article  CAS  PubMed  Google Scholar 

  42. Hanover, J. A. et al. Mitochondrial and nucleocytoplasmic isoforms of O-linked GlcNAc transferase encoded by a single mammalian gene. Arch. Biochem. Biophys. 409, 287–297 (2003).

    Article  CAS  PubMed  Google Scholar 

  43. Pathak, S. et al. The active site of O-GlcNAc transferase imposes constraints on substrate sequence. Nat. Struct. Mol. Biol. 22, 744–750 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Levine, Z. G. & Walker, S. The biochemistry of O-GlcNAc transferase: which functions make it essential in mammalian cells? Annu. Rev. Biochem. 85, 631–657 (2016).

    Article  CAS  PubMed  Google Scholar 

  45. Roth, C. et al. Structural and functional insight into human O-GlcNAcase. Nat. Chem. Biol. 13, 610–612 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. 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 

  47. Gorelik, A. et al. Genetic recoding to dissect the roles of site-specific protein O-GIcNAcylation. Nat. Struct. Mol. Biol. 26, 1071–1090 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Khoury, G. A., Baliban, R. C. & Floudas, C. A. Proteome-wide post-translational modification statistics: frequency analysis and curation of the swiss-prot database. Sci. Rep. 1, 90 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Patwardhan, A., Cheng, N. & Trejo, J. Post-translational modifications of G protein-coupled receptors control cellular signaling dynamics in space and time. Pharm. Rev. 73, 120–151 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Khidekel, N. et al. A chemoenzymatic approach toward the rapid and sensitive detection of O-GlcNAc posttranslational modifications. J. Am. Chem. Soc. 125, 16162–16163 (2003).

    Article  CAS  PubMed  Google Scholar 

  51. Agard, N. J., Prescher, J. A. & Bertozzi, C. R. A strain-promoted [3 + 2] azide-alkyne cycloaddition for covalent modification of biomolecules in living systems. J. Am. Chem. Soc. 126, 15046–15047 (2004).

    Article  CAS  PubMed  Google Scholar 

  52. Clark, P. M. et al. Direct in-gel fluorescence detection and cellular imaging of O-GlcNAc-modified proteins. J. Am. Chem. Soc. 130, 11576–11577 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Wang, Z. et al. Enrichment and site mapping of O-linked N-acetylglucosamine by a combination of chemical/enzymatic tagging, photochemical cleavage, and electron transfer dissociation mass spectrometry. Mol. Cell Proteom. 9, 153–160 (2010).

    Article  CAS  Google Scholar 

  54. Boyce, M. et al. Metabolic cross-talk allows labeling of O-linked β-N-acetylglucosamine-modified proteins via the N-acetylgalactosamine salvage pathway. Proc. Natl Acad. Sci. USA 108, 3141–3146 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Darabedian, N., Thompson, J. W., Chuh, K. N., Hsieh-Wilson, L. C. & Pratt, M. R. Optimization of chemoenzymatic mass tagging by strain-promoted cycloaddition (SPAAC) for the determination of O-GlcNAc stoichiometry by western blotting. Biochemistry 57, 5769–5774 (2018).

    Article  CAS  PubMed  Google Scholar 

  56. Thompson, J. W., Griffin, M. E. & Hsieh-Wilson, L. C. Methods for the detection, study, and dynamic profiling of O-GlcNAc glycosylation. Methods Enzymol. 598, 101–135 (2018).

    Article  CAS  PubMed  Google Scholar 

  57. Ma, J. F. et al. O-GlcNAc site mapping by using a combination of chemoenzymatic labeling, copper-free click chemistry, reductive cleavage, and electron-transfer dissociation mass spectrometry. Anal. Chem. 91, 2620–2625 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Rexach, J. E. et al. Dynamic O-GlcNAc modification regulates CREB-mediated gene expression and memory formation. Nat. Chem. Biol. 8, 253–261 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Liu, C. et al. O-GlcNAcylation of myosin phosphatase targeting subunit 1 (MYPT1) dictates timely disjunction of centrosomes. J. Biol. Chem. 295, 7341–7349 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Schwein, P. A. & Woo, C. M. The O-GlcNAc modification on kinases. ACS Chem. Biol. 15, 602–617 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Pedowitz, N. J., Batt, A. R., Darabedian, N. & Pratt, M. R. MYPT1 O-GlcNAc modification regulates sphingosine-1-phosphate mediated contraction. Nat. Chem. Biol. 17, 169–177 (2021).

    Article  CAS  PubMed  Google Scholar 

  62. Gross, B. J., Kraybill, B. C. & Walker, S. Discovery of O-GlcNAc transferase inhibitors. J. Am. Chem. Soc. 127, 14588–14589 (2005).

    Article  CAS  PubMed  Google Scholar 

  63. Yuzwa, S. A. et al. A potent mechanism-inspired O-GlcNAcase inhibitor that blocks phosphorylation of tau in vivo. Nat. Chem. Biol. 4, 483–490 (2008).

    Article  CAS  PubMed  Google Scholar 

  64. Gloster, T. M. et al. Hijacking a biosynthetic pathway yields a glycosyltransferase inhibitor within cells. Nat. Chem. Biol. 7, 174–181 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Ortiz-Meoz, R. F. et al. A small molecule that inhibits OGT activity in cells. ACS Chem. Biol. 10, 1392–1397 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Trapannone, R., Rafie, K. & van Aalten, D. M. O-GlcNAc transferase inhibitors: current tools and future challenges. Biochem. Soc. Trans. 44, 88–93 (2016).

    Article  CAS  PubMed  Google Scholar 

  67. Du, H. et al. YTHDF2 destabilizes m6A-containing RNA through direct recruitment of the CCR4–NOT deadenylase complex. Nat. Commun. 7, 12626 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Park, O. H. et al. Endoribonucleolytic cleavage of m6A-containing RNAs by RNase P/MRP complex. Mol. Cell 74, 494–507 (2019).

    Article  CAS  PubMed  Google Scholar 

  69. Gao, Y. et al. Multivalent m6A motifs promote phase separation of YTHDF proteins. Cell Res. 29, 767–769 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Ries, R. J. et al. m6A enhances the phase separation potential of mRNA. Nature 571, 424–428 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Fu, Y. & Zhuang, X. m6A-binding YTHDF proteins promote stress granule formation. Nat. Chem. Biol. 16, 955–963 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Park, J. E., Yi, H., Kim, Y., Chang, H. & Kim, V. N. Regulation of poly(A) tail and translation during the somatic cell cycle. Mol. Cell 62, 462–471 (2016).

    Article  CAS  PubMed  Google Scholar 

  73. Slawson, C., Lakshmanan, T., Knapp, S. & Hart, G. W. A mitotic GlcNAcylation/phosphorylation signaling complex alters the posttranslational state of the cytoskeletal protein vimentin. Mol. Biol. Cell 19, 4130–4140 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Tian, Y. et al. One-step enzymatic labeling reveals a critical role of O-GlcNAcylation in cell-cycle progression and DNA damage response. Angew. Chem. Int. Ed. 60, 26128–26135 (2021).

    Article  CAS  Google Scholar 

  75. Yuzwa, S. A. et al. Increasing O-GlcNAc slows neurodegeneration and stabilizes tau against aggregation. Nat. Chem. Biol. 8, 393–399 (2012).

    Article  CAS  PubMed  Google Scholar 

  76. Gambetta, M. C. & Muller, J. O-GlcNAcylation prevents aggregation of the Polycomb group repressor polyhomeotic. Dev. Cell 31, 629–639 (2014).

    Article  CAS  PubMed  Google Scholar 

  77. Marotta, N. P. et al. O-GlcNAc modification blocks the aggregation and toxicity of the protein alpha-synuclein associated with Parkinson’s disease. Nat. Chem. 7, 913–920 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Nosella, M. L. et al. O-linked-N-acetylglucosaminylation of the RNA-binding protein EWS N-terminal low complexity region reduces phase separation and enhances condensate dynamics. J. Am. Chem. Soc. 143, 11520–11534 (2021).

    Article  CAS  PubMed  Google Scholar 

  79. Wang, J. et al. Binding to m6A RNA promotes YTHDF2-mediated phase separation. Protein Cell 11, 304–307 (2020).

    Article  PubMed  Google Scholar 

  80. Aulas, A. et al. Stress-specific differences in assembly and composition of stress granules and related foci. J. Cell Sci. 130, 927–937 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  81. Hans, F., Glasebach, H. & Kahle, P. J. Multiple distinct pathways lead to hyperubiquitylated insoluble TDP-43 protein independent of its translocation into stress granules. J. Biol. Chem. 295, 673–689 (2020).

    Article  PubMed  Google Scholar 

  82. Anderson, P. & Kedersha, N. RNA granules. J. Cell Biol. 172, 803–808 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Jain, S. et al. ATPase-modulated stress granules contain a diverse proteome and substructure. Cell 164, 487–498 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Ohn, T., Kedersha, N., Hickman, T., Tisdale, S. & Anderson, P. A functional RNAi screen links O-GlcNAc modification of ribosomal proteins to stress granule and processing body assembly. Nat. Cell Biol. 10, 1224–1231 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Lv, P. et al. O-GlcNAcylation modulates liquid–liquid phase separation of SynGAP/PSD-95. Nat. Chem. 14, 831–840 (2022).

    Article  CAS  PubMed  Google Scholar 

  86. Hart, G. W., Housley, M. P. & Slawson, C. Cycling of O-linked β-N-acetylglucosamine on nucleocytoplasmic proteins. Nature 446, 1017–1022 (2007).

    Article  CAS  PubMed  Google Scholar 

  87. Shu, X. E., Mao, Y., Jia, L. & Qian, S. B. Dynamic eIF3a O-GlcNAcylation controls translation reinitiation during nutrient stress. Nat. Chem. Biol. 18, 134–141 (2022).

    Article  CAS  PubMed  Google Scholar 

  88. Wheeler, J. R., Jain, S., Khong, A. & Parker, R. Isolation of yeast and mammalian stress granule cores. Methods 126, 12–17 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Cox, J. & Mann, M. MaxQuant enables high peptide identification rates, individualized p.p.b.-range mass accuracies and proteome-wide protein quantification. Nat. Biotechnol. 26, 1367–1372 (2008).

    Article  CAS  PubMed  Google Scholar 

  90. Cox, J. et al. Andromeda: a peptide search engine integrated into the MaxQuant environment. J. Proteome Res. 10, 1794–1805 (2011).

    Article  CAS  PubMed  Google Scholar 

  91. Oates, M. E. et al. D²P²: database of disordered protein predictions. Nucleic Acids Res. 41, 508–516 (2013).

    Article  Google Scholar 

  92. Wang, L. H. et al. pFind 2.0: a software package for peptide and protein identification via tandem mass spectrometry. Rapid Commun. Mass Spectrom. 21, 2985–2991 (2007).

    Article  CAS  PubMed  Google Scholar 

  93. Yu, G., Wang, L. G., Han, Y. & He, Q. Y. clusterProfiler: an R package for comparing biological themes among gene clusters. OMICS 16, 284–287 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Szklarczyk, D. et al. STRING v11: protein-protein association networks with increased coverage, supporting functional discovery in genome-wide experimental datasets. Nucleic Acids Res. 47, 607–613 (2019).

    Article  Google Scholar 

  95. Shannon, P. et al. Cytoscape: a software environment for integrated models of biomolecular interaction networks. Genome Res. 13, 2498–2504 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Martin, M. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet.journal 17, 3 (2011).

  97. Church, D. M. et al. Modernizing reference genome assemblies. PLoS Biol. 9, e1001091 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Dobin, A. et al. STAR: ultrafast universal RNA-seq aligner. Bioinformatics 29, 15–21 (2013).

    Article  CAS  PubMed  Google Scholar 

  99. Li, H. et al. The Sequence Alignment/Map format and SAMtools. Bioinformatics 25, 2078–2079 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  100. Drewe-Boss, P., Wessels, H. H. & Ohler, U. omniCLIP: probabilistic identification of protein-RNA interactions from CLIP-seq data. Genome Biol. 19, 183 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  101. Schneider, V. A. et al. Evaluation of GRCh38 and de novo haploid genome assemblies demonstrates the enduring quality of the reference assembly. Genome Res. 27, 849–864 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Li, B. & Dewey, C. N. RSEM: accurate transcript quantification from RNA-Seq data with or without a reference genome. BMC Bioinform. 12, 323 (2011).

    Article  CAS  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  104. Ma, J. et al. iProX: an integrated proteome resource. Nucleic Acids Res. 47, D1211–D1217 (2019).

    Article  PubMed  Google Scholar 

Download references

Acknowledgements

This study was supported by the National Key R&D Program of China (grant no. 2019YFA09006600 to S.L.), the National Natural Science Foundation of China (grant nos 22222705, 91953113 and 92253302 to S.L. and 22277080 to Y.G.) and the Startup Fund from Shenzhen Bay Laboratory (grant no. 21230102 to Y.G.). C.H. is a Howard Hughes Medical Institute Investigator. We thank the core facility of the Life Sciences Institute Zhejiang University and the Bioimaging Core of Shenzhen Bay Laboratory. We also thank X. He for editing and helpful discussions.

Author information

Authors and Affiliations

Authors

Contributions

S.L. and C.H. conceived the idea and supervised the study. Y.G. supervised the in vitro LLPS experiments. Y.C., R.W. and Z.Z. designed and conducted all experiments and also analysed the data together. L.L., Y.Z. and L.T. assisted with experiments and provided valuable discussion. G.S. performed the in vitro LLPS experiments. Y.Y. provided technical advice. S.L., C.H., Y.C., R.W. and Z.Z. wrote the paper. All authors commented on the final draft of the paper.

Corresponding authors

Correspondence to Yun Ge, Chuan He or Shixian Lin.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Cell Biology thanks Zhi Qi, and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Peer reviewer 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

Extended Data Fig. 1 Post-translational modifications on YTHDFs identified by LC–MS/MS.

a) Filtered identified modifications in HEK293T were annotated on the sites (score > 100 for all modifications and diagnostic peak for O-GlcNAcylation). b, O-GlcNAcylation ratio of each site of YTHDF1 and YTHDF3 by LC–MS/MS analysis in HEK293T calculated with intensities (in blue) and MSMS counts (in orange) of the respective modified and unmodified peptide. (c) Estimated ratio of unmodified, single modified and double modified of YTHDF1 and YTHDF3 by the intensity ratio in (b). The estimated modification level is highly similar to the O-GlcNAcylation stoichiometry analysis in Fig. 1e. (d) O-GlcNAcylation stoichiometry analysis of Flag-tagged MYPT1, CSNK2A1 and CREB1 in HEK293T and HeLa. Arrowheads indicates the O-GlcNAcylation modified proteins. Flag* denoted samples were performed O-GlcNAcylation stoichiometry analysis and immunoblotting by anti-Flag. (e) Quantification result of O-GlcNAcylation level of MYPT1, CSNK2A1 and CREB1 in (d). Data are presented as mean ± s.d. (n = 3 biologically independent repeats). Significance was calculated using two-sided t-test.

Source data

Extended Data Fig. 2 Validation O-GlcNAcylation sites on YTHDF1 and Reversible O-GlcNAcylation sites of YTHDF3.

(a) O-GlcNAcylation stoichiometry analysis of WT-YTHDF1 and S157A mutant. The experiment was repeated twice with similar results. (b) O-GlcNAcylation stoichiometry analysis of WT-YTHDF1 and S196A mutant. The experiment was repeated twice with similar results. (c) O-GlcNAcylation stoichiometry analysis shows O-GlcNAcylation can be transferred in S196, S197 and S198, while S196 is the major O-GlcNAcylation site of YTHDF1. The experiment was repeated twice with similar results. (d) O-GlcNAcylation stoichiometry analysis of WT-YTHDF1 and S196/197/198 A mutant. The experiment was repeated twice with similar results. (e) O-GlcNAcylation stoichiometry analysis of WT-YTHDF3 and YTHDF3 with mutations on modification sites. The experiment was repeated twice with similar results. (f) Tandem mass spectrum of YTHDF3 peptide with T205 modified by O-GlcNAcylation. The spectrum of unmodified peptide was used for side-by-side comparison. The arrows pointed to the signature b–y ions in the modification spectrum. (g) Tandem mass spectrum of YTHDF3 peptide with S229 modified by O-GlcNAcylation. The spectrum of unmodified peptide was used for side-by-side comparison. The arrows pointed to the signature b–y ions in the modification spectrum. (h) O-GlcNAcylation stoichiometry analysis of YTHDF3 treated without and with 50 µM OSMI-1 or 10 µM Thiamet G for 5 h in HEK293T. The experiment was repeated twice with similar results. Arrowheads indicates the O-GlcNAcylation modified proteins. Flag* denotes samples were performed O-GlcNAcylation stoichiometry analysis and immunoblotting with anti-Flag.

Source data

Extended Data Fig. 3 The region of YTHDF1/3 modified by O-GlcNAcylation is essential for translation related protein binding but irrelevant for m6A binding.

(a, b) Pulldown assay of purified Flag-EGFP-tagged YTHDF1/3 by purified GST tagged EIF2S3, EIF3M and EIF4E. YTHDF1/3 can directly bind to these translation related proteins. The experiment was repeated twice with similar results. (c, d) Pulldown assay of purified his tagged EIF2S3 by purified Flag-EGFP-tagged YTHDF1/3 truncations (bottom) and the truncation scheme (up). Only the truncated YTHDF1/3 containing O-GlcNAcylation sites can directly bind to EIF2S3. (e, f) EMSA assay using vehicle, OGT or OGT-K852A treated EGFP–YTHDF1/3 and Cy3-m6A-RNA probe, (GGm6ACUC)10. Vehicle treated YTHDF1/3 contained no O-GlcNAcylation modification, OGT can modify YTHDF1/3 by O-GlcNAcylation and OGT-K852M mutant contains no catalytic activity. (g, h) Quantification analysis of shift RNA probe ratio in (e-f). Data are presented as mean ± s.d. (n = 3 biologically independent repeats). O-GlcNAcylation on the YTHDF1/3 contains no effect for the m6A RNA binding.

Source data

Extended Data Fig. 4 O-GlcNAcylation of YTHDF3 repressed target mRNA translation efficiency.

(a, b) Expression analysis of N-WT-YTHDF1/3-λ and N-YTHDF1/3-Mut-λ in of Fig. 4b,c and Extended Data Fig. 4f-g. (c) Translation efficiency analysis of N-WT-YTHDF1-λ containing no, either or both mutation on O-GlcNAcylation sites. Error bars, mean ± s.d., n = 3 biologically independent repeats. Significance was calculated using two-sided t-test. (d) Translation efficiency analysis of N-WT-YTHDF1-λ and λ treated with different concentration of OGT inhibitor OSMI-1 in HEK293T. Error bars, mean ± s.d., n = 3 biologically independent repeats. Significance was calculated using two-sided t-test. (e) Translation efficiency analysis of N-WT-YTHDF1-λ and λ treated with different concentration of OGA inhibitor Thiamet G in HeLa. Error bars, mean ± s.d., n = 3 biologically independent repeats. Significance was calculated using two-sided t-test. (f) Translation efficiency analysis of N-WT-YTHDF3-λ, N-YTHDF3-Mut-λ and λ in HEK293T. Error bars, mean ± s.d., n = 3 biologically independent repeats. Significance was calculated using two-sided t-test. (g) Translation efficiency analysis of N-WT-YTHDF3-λ, N-YTHDF3-Mut-λ and λ in HeLa. Error bars, mean ± s.d., n = 3 biologically independent repeats. Significance was calculated using two-sided t-test. (h) RNA abundance analysis of N-WT-YTHDF3-λ, N-YTHDF3-Mut-λ and λ in HEK293T by qPCR. Error bars, mean ± s.d., n = 3 biologically independent repeats. Significance was calculated using two-sided t-test. (i) Translation efficiency analysis of N-WT-YTHDF3-λ containing no, either or both mutation on O-GlcNAcylation sites. Error bars, mean ± s.d., n = 3 biologically independent repeats. Significance was calculated using two-sided t-test. (j) Comparison of fold change of label-free quantification of immunoprecipitated translation related proteins interacting with WT-YTHDF1/3 and YTHDF1/3-Mut in HEK293T and HeLa by LC–MS/MS. The red lines denoted the mean fold change of the proteins. Mean fold change in HEK293T is lower than in HeLa, which might be due to the low O-GlcNAcylation level of YTHDF1/3 in HeLa.

Source data

Extended Data Fig. 5 Co-localization of YTHDF1/3 with P-body.

Co-localization of mCherry-YTHDF1/EGFP–YTHDF3, m6A RNA and EDC4 (P-body marker) in HeLa shows YTHDF1/3 localized in P-body with m6A RNA. OE, overexpress. Scale bars: 10 μm. The experiment was repeated twice with similar results.

Extended Data Fig. 6 Cellular localization and expression level of YTHDF1/3.

(a, b) Immunoblotting of endogenous YTHDF1/3 in HEK293T, HeLa, OSMI-1 treated HEK293T and Thiamet G treated HeLa. OSMI-1: 50 µM, 5 hr; Thiamet G: 10 µM, 5 h. c, d) Quantification analysis of YTHDF1/3 expression level in (a-b). Error bars, mean ± s.d., n = 3 biologically independent repeats. Significance was calculated using two-sided t-test. The expression level of YTHDF1/3 in HeLa was lower than in HEK293T, but the inhibitors contained no effect on the expression level of YTHDF1/3. (e) Immunostaining of endogenous YTHDF1/3 in HEK293T, HeLa and Thiamet G treated HeLa. The localization of YTHDF1/3 shows not significantly different among cell lines. Scale bars: 10 μm. The experiment was repeated twice with similar results. (f) Immunostaining of Flag-tagged WT-YTHDF1/3 and YTHDF1/3-Mut in HEK293T and HeLa. The localization of YTHDF1/3 shows not significantly different among cell lines or between WT and Mut. Scale bars: 10 μm. The experiment was repeated twice with similar results.

Source data

Extended Data Fig. 7 Features of YTHDF1/3 PAR-CLIP data and mRNA-seq.

(a) Immunoblotting of Ythdf1/3 knockdown HEK293T and HeLa rescued by YTHDF1 or YTHDF3. Labelled transfection conditions were used for YTHDF1/3 PAR-CLIP and RNA-seq. The experiment was performed once. (b) FPKM of YTHDFs in HEK293T (b) and HeLa (c). Ctrl, Ythdf1/3 knockdown. WT, Ythdf1/3 knockdown rescued by WT-YTHDF1/3. Mut, Ythdf1/3 knockdown rescued by YTHDF1/3-Mut. Cumulative distribution log2-fold changes of mRNA input between Ythdf1/3 knockdown and Ythdf1/3 knockdown rescued by WT-YTHDF1/3 (d, f) or YTHDF1/3-Mut (e, g) in HEK293T (d, e) and HeLa (f, g) for non-target (grey) and YTHDF1/3 target (red). Distribution with boxplot and p values were calculated using two-sided t-test. Boxplot bounds depict quartile 1, median and quartile 3, with whiskers at 1.5× interquartile range. n = 58813 genes in total. The experiment was performed with 3 independent replicates. (h) Translation efficiency analysis of N-YTHDF1-λ and N-YTHDF3-λ treated with 1 or 9 g/L glucose in HeLa and the F-luc/R-luc was normalized by respective λ group. Error bars, mean ± s.d., n = 3 biologically independent repeats. Significance was calculated using two-sided t-test. (i) Distribution of translation efficiency of mRNA in HeLa cultured with no glucose or high glucose medium. n = 4655 genes and the experiment was performed with 3 independent replicates. (j) Distribution of log2-fold changes of translation efficiency of YTHDF1 target mRNAs in HeLa cultured with no glucose or high glucose medium. n = 1095 and 2268 genes. The experiment was performed with 3 independent replicates. (k) Distribution of log2-fold changes of translation efficiency of mRNAs containing different m6A numbers in HeLa cultured with no glucose or high glucose medium. n = 1694, 600, 1054 and 1306 genes. The experiment was performed with 3 independent replicates. Distribution with boxplot and p values were calculated using two-sided t-test. Boxplot bounds depict quartile 1, median and quartile 3, with whiskers at 1.5× interquartile range and outlier points.

Source data

Extended Data Fig. 8 Dynamic O-GlcNAcylation of YTHDF1/3 regulated m6A mRNA translation efficiency in a cell cycle-dependent manner.

(a) Detailed description of Fig. 6a. mRNAs in YTHDF1 PAR-CLIP and IP dataset were enriched by YTHDF1 PAR-CLIP or YTHDF1 IP in HeLa. mRNAs in YTHDFs iCLIP dataset were enriched by YTHDF1/2/3 iCLIP in HeLa. Translation efficiency in dataset was calculated by bulk RNA-seq and ribosome profiling seq results. Three datasets were combined to analyse the translation efficiency of m6A containing mRNAs. (b) Distribution of log2-fold changes of translation efficiency of mRNAs containing different m6A numbers between M and S phase. The dataset of m6A-modified mRNA in HeLa cells is generated by iCLIP. n = 4,384 genes in total. Boxplot bounds depict quartile 1, median and quartile 3, with whiskers at 1.5× interquartile range and outlier points. The analysis was performed using previously reported dataset (X. Wang et al.33, S. Zaccara et al.35 and J. Park et al.72) with 3 independent replicates. (c) O-GlcNAcylation stoichiometry analysis for YTHDF1 in HeLa at different cell cycle stages by double thymidine release. YTHDF1* denotes samples were performed O-GlcNAcylation stoichiometry analysis and immunoblotting by YTHDF1 antibodies. Arrowheads indicates the O-GlcNAcylation modified proteins. Quantification analysis was shown in Fig. 5g. (d) Immunoblotting of HeLa synchronized to M phase by nocodazole and released for 14 h. The experiment was repeated twice with similar results. (e) Immunoblotting for the cell cycle marker proteins in (f). H3pS10, M phase. Cyclin A2, late G1-S-G2. p-Cyclin E1-T395, G1. Cyclin B1, G2-M. The experiment was repeated twice with similar results. (f) Immunostaining of endogenous YTHDF1 in HeLa released from nocodazole for 4 (M), 7 (early G1) or 11 h (late G1). The localization of YTHDF1 shows not significantly different though different cell cycles. Scale bars: 10 μm. The experiment was repeated twice with similar results.

Source data

Extended Data Fig. 9 O-GlcNAcylation of YTHDF1 increased its dynamic nature in stress granules.

Immunostaining of EGFP–WT-YTHDF1 and EGFP–YTHDF1-Mut treated with or without OSMI-1 (a) or Thiamet G (b) in HEK293T (a) and HeLa (b). Scale bars: 10 μm. (c, d) Quantification result of SGs per cell of (a-b), n = 30 cells per condition from 3 independent experiments. Boxplot bounds depict quartile 1, median and quartile 3, with whiskers at 1.5× interquartile range. Significance was calculated using two-sided t-test. (e) Immunoblotting of expression control of (a). (f) Immunoblotting of expression control of (b). (g) Validation of Ythdf1/3 knockdown efficiency in HeLa by immunoblotting. The experiment was repeated twice with similar results. (h) Immunostaining of G3BP1 and YTHDF1 in parental and Ythdf1/3 knockdown HeLa treated with 0.6 M sorbitol stress. Scale bars: 10 μm. (i) Quantification result of SGs per cell of (h), n = 60 cells per condition from 3 independent experiments. Error bar, mean ± s.d. Significance was calculated using two-sided t-test. (j) Quantification result of fraction of YTHDF1 out of SGs of Fig. 6e, n = 69, 55 cells for WT and Mut group from 3 independent experiments. Error bar, mean ± s.d. Significance is calculated using two-sided t-test. (k) Quantification result of fraction of G3BP1 in SGs of Fig. 6e, n = 42, 37 cells for WT and Mut group from 3 independent experiments. Error bar, mean ± s.d. Significance was calculated using two-sided t-test. (l-n) FRAP analysis of EGFP–WT-YTHDF1 and EGFP–YTHDF1-Mut under 0.2 M NaCl stress in HEK293T. Scale bars: 1 μm. Error bar, mean ± s.d., n = 3 SGs per condition from 3 independent experiments. Significance was calculated using two-sided t-test. The white circles represent the photobleaching region. (o, p) O-GlcNAcylation stoichiometry analysis for YTHDF1 in HEK293T under sorbitol stress treatment and recovery. Arrowheads indicates the O-GlcNAcylation modified proteins. The experiment was performed once. YTHDF1* denotes samples were performed O-GlcNAcylation stoichiometry analysis and immunoblotting by YTHDF1 antibodies.

Source data

Extended Data Fig. 10 O-GlcNAcylation of YTHDF1 increased the dynamic nature of stress granules.

a–e, FRAP analysis of mCherry-WT-YTHDF1/YTHDF1-Mut and EGFP-G3BP1 treated with 0.6 M sorbitol stress in Ythdf1/3 knockdown HeLa. YTHDF1-Mut shows lower and slower recovery in fluorescence than WT-YTHDF1 both in YTHDF1 itself and SGs (G3BP1 as marker). Scale bars: 1 μm. Error bar, mean ± s.d., n = 3 SGs per condition from 3 independent experiments. Significance was calculated using two-sided t-test. The white circles represent the photobleaching region. (f, g) Time lapse imaging of mCherry-WT-YTHDF1/YTHDF1-Mut and EGFP-G3BP1 released from 0.6 M sorbitol stress in living Ythdf1/3 knockdown HeLa cells. Scale bars: 10 μm. Error bar, mean ± s.d., n = 7, 6 cells for WT and Mut group from 3 independent experiments. Significance was calculated using two-sided t-test. (hj) FRAP analysis of Cy3-m6A-RNA probe, BFP-G3BP1 and EGFP–YTHDF1 or O-GlcNAcylated EGFP–YTHDF1 in the droplets in vitro. Scale bars: 5 μm. Error bar, mean ± s.d., n = 3 droplets per condition from 3 independent experiments. Significance was calculated using a two-sided t-test. OG denoted O-GlcNAcylated form by OGT reaction. The white circles represent the photobleaching region.

Source data

Supplementary information

Reporting Summary

Peer Review File

Supplementary Data 1

Post-translational modifications identified on YTHDFs.

Supplementary Data 2

Interaction proteins of YTHDF1 and YTHDF3 by IP-MS.

Supplementary Data 3

CLIP peaks of the wild type and mutant.

Supplementary Data 4

RiboLace reads.

Supplementary Data 5

Translation efficiency analysis of m6A mRNAs in different glucose concentrations.

Supplementary Data 6

Translation efficiency analysis of cell cycle by reported RNA-seq datasets in cell cycles.

Source data

Source Data Fig. 1

Unprocessed western blots.

Source Data Fig. 1

Statistical source data.

Source Data Fig. 2

Unprocessed western blots.

Source Data Fig. 3

Unprocessed western blots.

Source Data Fig. 4

Statistical source data.

Source Data Fig. 5

Unprocessed western blots.

Source Data Fig. 5

Statistical source data.

Source Data Fig. 6

Unprocessed western blots.

Source Data Fig. 6

Statistical source data.

Source Data Fig. 7

Unprocessed western blots.

Source Data Fig. 7

Statistical source data.

Source Data Extended Data Fig. 1

Unprocessed western blots.

Source Data Extended Data Fig. 1

Statistical source data.

Source Data Extended Data Fig. 2

Unprocessed western blots.

Source Data Extended Data Fig. 3

Unprocessed western blots.

Source Data Extended Data Fig. 3

Statistical source data.

Source Data Extended Data Fig. 4

Unprocessed western blots.

Source Data Extended Data Fig. 4

Statistical source data.

Source Data Extended Data Fig. 6

Unprocessed western blots.

Source Data Extended Data Fig. 6

Statistical source data.

Source Data Extended Data Fig. 7

Unprocessed western blots.

Source Data Extended Data Fig. 7

Statistical source data.

Source Data Extended Data Fig. 8

Unprocessed western blots.

Source Data Extended Data Fig. 9

Unprocessed western blots.

Source Data Extended Data Fig. 9

Statistical source data.

Source Data Extended Data Fig. 10

Statistical 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

Chen, Y., Wan, R., Zou, Z. et al. O-GlcNAcylation determines the translational regulation and phase separation of YTHDF proteins. Nat Cell Biol 25, 1676–1690 (2023). https://doi.org/10.1038/s41556-023-01258-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41556-023-01258-x

This article is cited by

Search

Quick links

Nature Briefing: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research