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:

Dynamic eIF3a O-GlcNAcylation controls translation reinitiation during nutrient stress

Abstract

In eukaryotic cells, many messenger RNAs (mRNAs) possess upstream open reading frames (uORFs) in addition to the main coding region. After uORF translation, the ribosome could either recycle at the stop codon or resume scanning for downstream start codons in a process known as reinitiation. Accumulating evidence suggests that some initiation factors, including eukaryotic initiation factor 3 (eIF3), linger on the early elongating ribosome, forming an eIF3–80S complex. Very little is known about how eIF3 is carried along with the 80S during elongation and whether the eIF3–80S association is subject to regulation. Here, we report that eIF3a undergoes dynamic O-linked N-acetylglucosamine (O-GlcNAc) modification in response to nutrient starvation. Stress-induced de-O-GlcNAcylation promotes eIF3 retention on the elongating ribosome and facilitates activating transcription factor 4 (ATF4) reinitiation. Eliminating the modification site from eIF3a via CRISPR genome editing induces ATF4 reinitiation even under the nutrient-rich condition. Our findings illustrate a mechanism in balancing ribosome recycling and reinitiation, thereby linking the nutrient stress response and translational reprogramming.

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: Prolonged eIF3–80S association in response to amino acid starvation.
Fig. 2: Amino acid starvation triggers de-O-GlcNAcylation of eIF3a.
Fig. 3: Dynamic O-GlcNAcylation of eIF3a regulates translation reinitiation.
Fig. 4: Deficient O-GlcNAcylation of eIF3a promotes eIF3–80S association.
Fig. 5: Deficient O-GlcNAcylation of eIF3a affects general translation reinitiation.
Fig. 6: A model of eIF3–80S association orchestrated by dynamic O-GlcNAcylation of eIF3a.

Similar content being viewed by others

Data availability

All sequencing data have been deposited in the Gene Expression Omnibus under accession number GSE181040. Source data are provided with this paper.

Code availability

All the codes used in this study can be found in https://github.com/QianLab-Cornell/Count_Ribo_Reads.git

References

  1. Hinnebusch, A. G. The scanning mechanism of eukaryotic translation initiation. Annu. Rev. Biochem. 83, 779–812 (2014).

    Article  CAS  PubMed  Google Scholar 

  2. Jackson, R. J., Hellen, C. U. & Pestova, T. V. The mechanism of eukaryotic translation initiation and principles of its regulation. Nat. Rev. Mol. Cell Biol. 11, 113–127 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Hershey, J. W. B., Sonenberg, N. & Mathews, M. B. Principles of translational control. Cold Spring Harb. Perspect. Biol. 11, a032607 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Gunisova, S., Hronova, V., Mohammad, M. P., Hinnebusch, A. G. & Valasek, L. S. Please do not recycle! Translation reinitiation in microbes and higher eukaryotes. FEMS Microbiol. Rev. 42, 165–192 (2018).

    Article  CAS  PubMed  Google Scholar 

  5. Hinnebusch, A. G. Translational regulation of GCN4 and the general amino acid control of yeast. Annu. Rev. Microbiol. 59, 407–450 (2005).

    Article  CAS  PubMed  Google Scholar 

  6. Lu, P. D., Harding, H. P. & Ron, D. Translation reinitiation at alternative open reading frames regulates gene expression in an integrated stress response. J. Cell Biol. 167, 27–33 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Vattem, K. M. & Wek, R. C. Reinitiation involving upstream ORFs regulates ATF4 mRNA translation in mammalian cells. Proc. Natl Acad. Sci. USA 101, 11269–11274 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Kozak, M. Constraints on reinitiation of translation in mammals. Nucleic Acids Res. 29, 5226–5232 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Luukkonen, B. G., Tan, W. & Schwartz, S. Efficiency of reinitiation of translation on human immunodeficiency virus type 1 mRNAs is determined by the length of the upstream open reading frame and by intercistronic distance. J. Virol. 69, 4086–4094 (1995).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Kozak, M. Effects of intercistronic length on the efficiency of reinitiation by eucaryotic ribosomes. Mol. Cell. Biol. 7, 3438–3445 (1987).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Wek, R. C., Jiang, H. Y. & Anthony, T. G. Coping with stress: eIF2 kinases and translational control. Biochem. Soc. Trans. 34, 7–11 (2006).

    Article  CAS  PubMed  Google Scholar 

  12. Skabkin, M. A., Skabkina, O. V., Hellen, C. U. & Pestova, T. V. Reinitiation and other unconventional posttermination events during eukaryotic translation. Mol. Cell. 51, 249–264 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Mohammad, M. P., Munzarova Pondelickova, V., Zeman, J., Gunisova, S. & Valasek, L. S. In vivo evidence that eIF3 stays bound to ribosomes elongating and terminating on short upstream ORFs to promote reinitiation. Nucleic Acids Res. 45, 2658–2674 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Park, H. S., Himmelbach, A., Browning, K. S., Hohn, T. & Ryabova, L. A. A plant viral “reinitiation” factor interacts with the host translational machinery. Cell 106, 723–733 (2001).

    Article  CAS  PubMed  Google Scholar 

  15. Hronova, V. et al. Does eIF3 promote reinitiation after translation of short upstream ORFs also in mammalian cells? RNA Biol. 14, 1660–1667 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  16. Hershey, J. W. The role of eIF3 and its individual subunits in cancer. Biochim. Biophys. Acta 1849, 792–800 (2015).

    Article  CAS  PubMed  Google Scholar 

  17. Valasek, L. S. et al. Embraced by eIF3: structural and functional insights into the roles of eIF3 across the translation cycle. Nucleic Acids Res. 45, 10948–10968 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Lee, A. S., Kranzusch, P. J. & Cate, J. H. eIF3 targets cell-proliferation messenger RNAs for translational activation or repression. Nature 522, 111–114 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. des Georges, A. et al. Structure of mammalian eIF3 in the context of the 43S preinitiation complex. Nature 525, 491–495 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Siridechadilok, B., Fraser, C. S., Hall, R. J., Doudna, J. A. & Nogales, E. Structural roles for human translation factor eIF3 in initiation of protein synthesis. Science 310, 1513–1515 (2005).

    Article  CAS  PubMed  Google Scholar 

  21. Wagner, S. et al. Selective translation complex profiling reveals staged initiation and co-translational assembly of initiation factor complexes. Mol. Cell 79, 546–560 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Bohlen, J., Fenzl, K., Kramer, G., Bukau, B. & Teleman, A. A. Selective 40S footprinting reveals cap-tethered ribosome scanning in human cells. Mol. Cell 79, 561–574 (2020).

    Article  CAS  PubMed  Google Scholar 

  23. Lin, Y. et al. eIF3 associates with 80S ribosomes to promote translation elongation, mitochondrial homeostasis, and muscle health. Mol. Cell 79, 575–587 (2020).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  25. Bond, M. R. & Hanover, J. A. A little sugar goes a long way: the cell biology of O-GlcNAc. J. Cell Biol. 208, 869–880 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. 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 

  27. 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 

  28. Li, X. et al. O-GlcNAcylation of core components of the translation initiation machinery regulates protein synthesis. Proc. Natl Acad. Sci. USA 116, 7857–7866 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Zhang, X., Shu, X. E. & Qian, S. B. O-GlcNAc modification of eIF4GI acts as a translational switch in heat shock response. Nat. Chem. Biol. 14, 909–916 (2018).

    Article  CAS  PubMed  Google Scholar 

  30. Archer, S. K., Shirokikh, N. E., Beilharz, T. H. & Preiss, T. Dynamics of ribosome scanning and recycling revealed by translation complex profiling. Nature 535, 570–574 (2016).

    Article  CAS  PubMed  Google Scholar 

  31. Xing, L., Li, J., Xu, Y., Xu, Z. & Chong, K. Phosphorylation modification of wheat lectin VER2 is associated with vernalization-induced O-GlcNAc signaling and intracellular motility. PLoS ONE 4, e4854 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  32. Kazemi, Z., Chang, H., Haserodt, S., McKen, C. & Zachara, N. E. O-linked beta-N-acetylglucosamine (O-GlcNAc) regulates stress-induced heat shock protein expression in a GSK-3beta-dependent manner. J. Biol. Chem. 285, 39096–39107 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Zhou, J. et al. N6-Methyladenosine guides mRNA alternative translation during integrated stress response. Mol. Cell 69, 636–647 e7 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Bohlen, J. et al. DENR promotes translation reinitiation via ribosome recycling to drive expression of oncogenes including ATF4. Nat. Commun. 11, 4676 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Calvo, S. E., Pagliarini, D. J. & Mootha, V. K. Upstream open reading frames cause widespread reduction of protein expression and are polymorphic among humans. Proc. Natl Acad. Sci. USA 106, 7507–7512 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Orr, M. W., Mao, Y., Storz, G. & Qian, S. B. Alternative ORFs and small ORFs: shedding light on the dark proteome. Nucleic Acids Res. 48, 1029–1042 (2020).

    Article  CAS  PubMed  Google Scholar 

  37. Pisarev, A. V., Hellen, C. U. & Pestova, T. V. Recycling of eukaryotic posttermination ribosomal complexes. Cell 131, 286–299 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Zhang, F. et al. O-GlcNAc modification is an endogenous inhibitor of the proteasome. Cell 115, 715–725 (2003).

    Article  CAS  PubMed  Google Scholar 

  39. Thiebeauld, O. et al. A new plant protein interacts with eIF3 and 60S to enhance virus-activated translation re-initiation. EMBO J. 28, 3171–3184 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Sanjana, N. E., Shalem, O. & Zhang, F. Improved vectors and genome-wide libraries for CRISPR screening. Nat. Methods 11, 783–784 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Shalem, O. et al. Genome-scale CRISPR–Cas9 knockout screening in human cells. Science 343, 84–87 (2014).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank N. E. Zachara for providing inducible MEF Ogt KO cell lines and R. C. Wek for ATF4 uORF mutants. We also thank Qian lab members for helpful discussions. We are grateful to Cornell University Life Sciences Core Laboratory Center for sequencing support. We thank the Proteomic and MS Facility of Cornell University for help with the mass spectrometry. This work was supported by US National Institutes of Health (R01GM1222814 and DP1GM142101) and HHMI Faculty Scholar (55108556) to S.-B.Q.

Author information

Authors and Affiliations

Authors

Contributions

S.-B.Q. conceived the project and designed the experiments. X.E.S. designed and performed the majority of experiments. Y.M. conducted the majority of sequencing data analysis. L.J. assisted HiBiT reporter assay and Ribo-seq experiments. All authors discussed the results and edited the manuscript.

Corresponding author

Correspondence to Shu-Bing Qian.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature Chemical Biology thanks Wen Yi and other, anonymous, reviewer(s) for their contribution to the peer review of this work.

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 eIF3-80S association in response to amino acid starvation.

a. MEFs with or without amino acid starvation for 2 hr were subjected polysome profiling using sucrose gradient. Ribosome fractions were collected for immunoblotted using antibodies indicated. The bottom panel shows the eIF3a levels relative to the RpL4. Error bars, mean ± s.e.m.; t-test, two-tailed; n = 2 independent experiments. b. The top panel shows a scatter plot of read density on individual transcripts between 80 S (Ribo-seq, x-axis) and eIF3-80S (eIF3-seq, y-axis) footprints from MEF cells. The bottom panel shows the GO analysis of eIF3-enriched (red) or eIF3-depleted (blue) transcripts. c. The top panel shows a scatter plot of read density on individual transcripts between 80 S (Ribo-seq, x-axis) and eIF3-80S (eIF3-seq, y-axis) footprints from MEF cells subjected to amino acid starvation for 2 hr. The bottom panel shows the GO analysis of eIF3-enriched (red) or eIF3-depleted (blue) transcripts.

Source data

Extended Data Fig. 2 Amino acid starvation induces de-O-GlcNAcylation of eIF3a.

a. Schematic of quantitative mass spectrometry using iTRAQ to compare phosphoproteomics. b. A venn diagram shows the number of proteins identified in control (198), 2 hr amino acid starvation (291), or both (467) from the phosphoproteomics data in (A). c. A scatter plot shows the differential phosphoproteins in MEF cells with and without amino acid starvation. Phosphoproteins with >2 fold increase or decrease are colored in red and blue, respectively. d. MEF/Ogt F/Y(mER-Cre) cells were pre-treated with (+) or without (–) 0.5 μM 4-OHT for 24 hr followed by normal media for another 24 hr. Cells were then subjected to amino acid starvation for 2 hr prior to immunoblotting. Representative results of 3 independent experiments were shown. e. MEF cells as in (D) were subjected to immunoprecipitation using WGA antibody followed by immunoblotting. Representative results of 3 independent experiments were shown. f. HEK293 cells were transfected with scramble or siRNA targeting OGT for 24 hr followed by RL2 (O-GlcNAc) immunoprecipitation and immunoblotting. Representative results of 3 independent experiments were shown. g. MEF cells were treated with increasing doses of Thiamet-G for 16 hr followed by immunoblotting. Representative results of 3 independent experiments were shown.

Source data

Extended Data Fig. 3 Characterization of eIF3-80S association in the presence of Thiamet-G.

a. MEF cells were treated with Thiamet-G for 16 hr with or without amino acid starvation for 2 hr followed by RNA extraction and RT-qPCR. Relative Atf4 mRNA levels are normalized to the Actb mRNA level. Error bars, mean ± s.d.; n = 3 independent experiments. b. MEF cells were treated with 0.4 μM Thiamet-G for 1 hr followed by amino acid starvation for 2 hr. Whole-cell lysates were immunoprecipitated using the eIF3a antibody followed by immunoblotting using indicated antibodies. Representative results of 3 independent experiments were shown. c. Same as (b), total RNAs and eIF3-associated RNAs were extracted followed by RT-PCR measuring Actb and Aft4 mRNA levels. Relative Atf4/Actb ratios were normalized to the corresponding input. Error bars, mean ± s.d.; *p = 0.019; ***p = 0.00078; two-way ANOVA; n = 3 independent experiments. d. MEF cells as in (b) were subjected to immunoblotting using antibodies indicated. The right panel shows the relative ATF4 protein levels normalized to α-tubulin. Error bars, mean ± s.d.; ***p = 0.00014; two-way ANOVA; n = 3 independent experiments. e. MEF cells were treated with Thiamet-G for 16 hr were subjected to polysome profiling using sucrose gradient. Ribosome fractions were collected for immunoblotted using antibodies indicated. The bottom panel shows the densitometry of eIF3a and RpL4 protein levels. Representative results of 3 independent experiments were shown. f. MEF cells as (e) were subjected to amino acid starvation for 2 h followed by polysome profiling using sucrose gradient. Ribosome fractions were collected for immunoblotted using antibodies indicated. The bottom panel shows the densitometry of eIF3a and RpL4 protein levels. Representative results of 3 independent experiments were shown.

Source data

Extended Data Fig. 4 Further characterization of eIF3-80S association in the presence of Thiamet-G.

a. MEF cells were treated with Thiamet-G for 16 hr followed by amino acid starvation for 2 hr. Polysome profiling was conducted using sucrose gradient. Ribosome fractions were collected for immunoblotted using antibodies indicated. Representative results of 3 independent experiments were shown. b. The top panel shows a scatter plot of read density on individual transcripts between 80 S (Ribo-seq, x-axis) and eIF3-80S (eIF3-seq, y-axis) footprints from MEF cells treated with Thiamet-G. The bottom panel shows the GO analysis of eIF3-enriched (red) or eIF3-depleted (blue) transcripts. c. The top panel shows a scatter plot of read density on individual transcripts between 80 S (Ribo-seq, x-axis) and eIF3-80S (eIF3-seq, y-axis) footprints from MEF cells treated with Thiamet-G and 2 hr amino acid starvation. The bottom panel shows the GO analysis of eIF3-enriched (red) or eIF3-depleted (blue) transcripts. d. The distribution of 80 S footprints in the Atf4 transcript from MEF cells with or without Thiamet-G treatment, before and after amino acid starvation.

Source data

Extended Data Fig. 5 Deficient O-GlcNAcylation attenuates global protein synthesis.

a. MEF/Ogt F/Y(mER-Cre) cells were pre-treated with (+) or without (–) 0.5 μM 4-OHT for 24 hr followed by normal media for another 24 hr. Cells were then subjected to amino acid starvation for 2 hr followed by immunoprecipitation using the eIF3a antibody. Total RNAs and eIF3-associated RNAs were extracted followed by RT-PCR measuring Actb and Aft4 mRNA levels. Relative Atf4/Actb ratios were normalized to the corresponding input. Error bars, mean ± s.e.m.; *P = 0.0036; two-way ANOVA; n = 3 independent experiments. b. MEF cells as in (a) were subjected to immunoblotting using antibodies indicated. The right panel shows the relative ATF4 protein levels normalized to α-tubulin. Error bars, mean ± s.e.m.; *P = 0.015, **P = 0.0067; two-way ANOVA; n = 3 independent experiments. c. MEF cells as in (a) were subjected to polysome profiling using sucrose gradient. Ribosome fractions were collected for immunoblotted using antibodies indicated. The bottom panel shows the eIF3a levels relative to the RpL4. Error bars, mean ± s.e.m.; t-test, two-tailed; n = 2 independent experiments. The bottom panel shows the densitometry of eIF3a and RpL4 protein levels. d. MEF cells as in (a) were incubated with 200 ng/µl of puromycin for 15 min. Whole cell lysates were used for immunoblotting. Representative results of 3 independent experiments were shown. e. MEF cells as in (a) were subjected to polysome profiling using sucrose gradient. Ribosome fractions were collected for immunoblotted using antibodies indicated. Representative results of 3 independent experiments were shown.

Source data

Extended Data Fig. 6 Deficient O-GlcNAcylation of eIF3a promotes eIF3-80S association.

a. The top panel shows a scatter plot of read density on individual transcripts between 80 S (Ribo-seq, x-axis) and eIF3-80S (eIF3-seq, y-axis) footprints from MEF cells with 4-OHT treatment. The bottom panel shows the GO analysis of eIF3-enriched (red) or eIF3-depleted (blue) transcripts. b. The top panel shows a scatter plot of read density on individual transcripts between 80 S (Ribo-seq, x-axis) and eIF3-80S (eIF3-seq, y-axis) footprints from MEF cells with 4-OHT treatment and 2 hr amino acid starvation. The bottom panel shows the GO analysis of eIF3-enriched (red) or eIF3-depleted (blue) transcripts. c. A Scatter plot shows the correlation of fold changes (log2) of CDS ribosome density in MEF/Ogt F/Y(mER-Cre) cells between Ogt knockout and amino acid starvation. d. Cumulative distribution of fold changes of read density in 5’UTR (blue) and 3’UTR (red) relative to the CDS in MEF cells with OGT knockout (top panel) or together with amino acid starvation (bottom panel). P values are based on Wilcoxon signed-rank test.

Extended Data Fig. 7 Characterization of eIF3a mutants with deficient O-GlcNAcylation.

a. HEK 293 cells were transfected with plasmids expressing EGFP-eIF3a for 24 hr. Cells lysates were immunoprecipitated using an anti-GFP antibody followed by immunoblotting. Bottom panel, HEK cells transfected with GFP-eIF3a were subjected to a 15%-45% sucrose gradient, followed by western blotting using indicated antibodies. Representative results of 3 independent experiments were shown. b. HEK 293 cells were transfected with plasmids expressing wild type eIF3a (WT) or T336A mutant fused to EGFP for 24 hr. Cells lysates were immunoprecipitated using an anti-GFP antibody followed by immunoblotting. Representative results of 3 independent experiments were shown. c. HEK 293 cells bearing eIF3a wildtype (WT) and S225G mutation were incubated with 200 ng/µl of puromycin for 15 min. Whole cell lysates were used for immunoblotting. d. Growth rates of HEK 293 cells bearing eIF3a wildtype (WT) and S225G mutation were measured using CCK-8 assay. Error bars, mean ± s.d.; t-test, two-tailed; n = 3 biological replicates. e. HEK293 cells as in (C) were subjected to amino acid starvation for 2 hr followed by total RNA extraction and RT-qPCR. Relative Atf4 mRNA levels are normalized to the Actb mRNA level. Error bars, mean ± s.d.; n = 3 independent experiments.

Source data

Extended Data Fig. 8 Deficient O-GlcNAcylation of eIF3a affects general translation reinitiation.

a. The left panel shows the schematic of ATF4-Fluc reporters with uORF deletion. The right panel shows the Fluc activities in transfected MEF cells with and without amino acid starvation. Error bars, mean ± s.d.; *p = 0.037; two-way ANOVA; n = 3 independent experiments. b. MEF cells transfected with ATF4-Fluc reporter mutants were subjected to amino acid starvation for 2 hr. Whole cell lysates were immunoprecipitated using the eIF3a antibody. Total RNAs and eIF3-associated RNAs were extracted followed by RT-PCR measuring Actb and Aft4 mRNA levels. Relative Atf4/Actb ratios were normalized to the corresponding input. Error bars, mean ± s.d.; ***p = 0.00035; two-way ANOVA; n = 3 independent experiments. c. HEK293 cells bearing S225G mutation of eIF3a were transfected with ATF4-Fluc reporter mutants followed by amino acid starvation for 2 hr. Whole cell lysates were immunoprecipitated using the eIF3a antibody. Total RNAs and eIF3-associated RNAs were extracted followed by RT-PCR measuring Actb and Aft4 mRNA levels. Relative Atf4/Actb ratios were normalized to the corresponding input. Error bars, mean ± s.d.; n = 3 independent experiments. d. EK293 cells with (+) or without (–) amino acid starvation for 2 hr were subjected to immuno-precipitation using eIF3a antibody. Total RNAs and eIF3-associated RNAs were extracted followed by RT-PCR measuring Actb and Map2k6 mRNA levels. Relative Map2k6/Actb ratios were normalized to the corresponding input. Error bars, mean ± s.d.; *p = 0.018; two-way ANOVA; n = 3 independent experiments. e. Steady-state mRNA levels of Map2k6 in HEK293 cells as in (d) were normalized with Actb, Error bars, mean ± s.d.; n = 3 independent experiments.

Source data

Extended Data Fig. 9 A model summarizing eIF3-80S complex association orchestrated by dynamic O-GlcNAcylation of eIF3a.

Schematic Cryo-EM structure of mammalian eIF3 (Adopted from des Georges et al. 2015). eIF3a is shown in red and the putative O-GlcNAc modification site shown as the white star.

Supplementary information

Reporting Summary

Supplementary Table 1

Phosphoproteomics of MEF cells with or without amino acid starvation.

Supplementary Table 2

Quantitative proteomics of O-GlcNAc modification in MEF cells with or without amino acid starvation.

Supplementary Table 3

Quantitative proteomics of O-GlcNAc modification in MEF cells with or without OGT knockdown.

Supplementary Table 4

Tandem mass spectrometry of purified eIF3a for putative O-GlcNAc modification sites.

Source data

Source Data Fig. 1

Unprocessed western blots in Fig. 1.

Source Data Fig. 2

Unprocessed western blots in Fig. 2.

Source Data Fig. 3

Unprocessed western blots in Fig. 3.

Source Data Fig. 4

Unprocessed western blots in Fig. 4.

Source Data Fig. 5

Unprocessed western blots in Fig. 5.

Source Data Extended Data Fig. 1

Unprocessed western blots.

Source Data Extended Data Fig. 2

Unprocessed western blots.

Source Data Extended Data Fig. 3

Unprocessed western blots.

Source Data Extended Data Fig. 4

Unprocessed western blots.

Source Data Extended Data Fig. 5

Unprocessed western blots.

Source Data Extended Data Fig. 7

Unprocessed western blots.

Source Data Extended Data Fig. 8

Unprocessed western blots.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Shu, X.E., Mao, Y., Jia, L. et al. Dynamic eIF3a O-GlcNAcylation controls translation reinitiation during nutrient stress. Nat Chem Biol 18, 134–141 (2022). https://doi.org/10.1038/s41589-021-00913-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41589-021-00913-4

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing