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:

Site-specific proteasome phosphorylation controls cell proliferation and tumorigenesis

Abstract

Despite the fundamental importance of proteasomal degradation in cells, little is known about whether and how the 26S proteasome itself is regulated in coordination with various physiological processes. Here we show that the proteasome is dynamically phosphorylated during the cell cycle at Thr 25 of the 19S subunit Rpt3. CRISPR/Cas9-mediated genome editing, RNA interference and biochemical studies demonstrate that blocking Rpt3-Thr25 phosphorylation markedly impairs proteasome activity and impedes cell proliferation. Through a kinome-wide screen, we have identified dual-specificity tyrosine-regulated kinase 2 (DYRK2) as the primary kinase that phosphorylates Rpt3-Thr25, leading to enhanced substrate translocation and degradation. Importantly, loss of the single phosphorylation of Rpt3-Thr25 or knockout of DYRK2 significantly inhibits tumour formation by proteasome-addicted human breast cancer cells in mice. These findings define an important mechanism for proteasome regulation and demonstrate the biological significance of proteasome phosphorylation in regulating cell proliferation and tumorigenesis.

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

Figure 1: Rpt3-Thr25 is dynamically phosphorylated during the cell cycle.
Figure 2: Blockade of Rpt3-Thr25 phosphorylation impedes cell proliferation.
Figure 3: Loss of Rpt3-Thr25 phosphorylation downregulates 26S proteasome activity.
Figure 4: DYRK2 is the primary Rpt3-Thr25 kinase.
Figure 5: DYRK2 is a positive regulator of proteasome activity.
Figure 6: Mechanisms by which DYRK2 regulates the proteasome.
Figure 7: DYRK2 positively regulates cell growth.
Figure 8: Rpt3-Thr25 phosphorylation is required for tumour growth in vivo.

Similar content being viewed by others

References

  1. Coux, O., Tanaka, K. & Goldberg, A. L. Structure and functions of the 20S and 26S proteasomes. Annu. Rev. Biochem. 65, 801–847 (1996).

    CAS  PubMed  Google Scholar 

  2. Tai, H. C. & Schuman, E. M. Ubiquitin, the proteasome and protein degradation in neuronal function and dysfunction. Nat. Rev. Neurosci. 9, 826–838 (2008).

    CAS  PubMed  Google Scholar 

  3. López-Otín, C., Blasco, M. A., Partridge, L., Serrano, M. & Kroemer, G. The hallmarks of aging. Cell 153, 1194–1217 (2013).

    PubMed  PubMed Central  Google Scholar 

  4. Hoeller, D. & Dikic, I. Targeting the ubiquitin system in cancer therapy. Nature 458, 438–444 (2009).

    CAS  PubMed  Google Scholar 

  5. Orlowski, R. Z. & Kuhn, D. J. Proteasome inhibitors in cancer therapy: lessons from the first decade. Clin. Cancer Res. 14, 1649–1657 (2008).

    CAS  PubMed  Google Scholar 

  6. Finley, D. Recognition and processing of ubiquitin-protein conjugates by the proteasome. Annu. Rev. Biochem. 78, 477–513 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Bhattacharyya, S., Yu, H., Mim, C. & Matouschek, A. Regulated protein turnover: snapshots of the proteasome in action. Nat. Rev. Mol. Cell Biol. 15, 122–133 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Ehlinger, A. & Walters, K. J. Structural insights into proteasome activation by the 19S regulatory particle. Biochemistry 52, 3618–3628 (2013).

    CAS  PubMed  Google Scholar 

  9. Murata, S., Yashiroda, H. & Tanaka, K. Molecular mechanisms of proteasome assembly. Nat. Rev. Mol. Cell Biol. 10, 104–115 (2009).

    CAS  PubMed  Google Scholar 

  10. Schmidt, M. & Finley, D. Regulation of proteasome activity in health and disease. Biochim. Biophys. Acta 1843, 13–25 (2014).

    CAS  PubMed  Google Scholar 

  11. Radhakrishnan, S. K. et al. Transcription factor Nrf1 mediates the proteasome recovery pathway after proteasome inhibition in mammalian cells. Mol. Cell 38, 17–28 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Tomko, R. J. & Hochstrasser, M. Molecular architecture and assembly of the eukaryotic proteasome. Annu. Rev. Biochem. 82, 415–445 (2013).

    CAS  PubMed  Google Scholar 

  13. Wang, X. et al. Mass spectrometric characterization of the affinity-purified human 26S proteasome complex. Biochemistry 46, 3553–3565 (2007).

    CAS  PubMed  Google Scholar 

  14. Wang, X. & Huang, L. Identifying dynamic interactors of protein complexes by quantitative mass spectrometry. Mol. Cell Proteomics 7, 46–57 (2008).

    PubMed  Google Scholar 

  15. Hershko, A. Roles of ubiquitin-mediated proteolysis in cell cycle control. Curr. Opin. Cell Biol. 9, 788–799 (1997).

    CAS  PubMed  Google Scholar 

  16. Teixeira, L. K. & Reed, S. I. Ubiquitin ligases and cell cycle control. Annu. Rev. Biochem. 82, 387–414 (2013).

    CAS  PubMed  Google Scholar 

  17. Bence, N. F., Sampat, R. M. & Kopito, R. R. Impairment of the ubiquitin-proteasome system by protein aggregation. Science 292, 1552–1555 (2001).

    CAS  PubMed  Google Scholar 

  18. Min, M. & Lindon, C. Substrate targeting by the ubiquitin–proteasome system in mitosis. Semin. Cell Dev. Biol. 23, 482–491 (2012).

    CAS  PubMed  Google Scholar 

  19. Dephoure, N. et al. A quantitative atlas of mitotic phosphorylation. Proc. Natl Acad. Sci. USA 105, 10762–10767 (2008).

    CAS  PubMed  Google Scholar 

  20. Olsen, J. V. et al. Quantitative phosphoproteomics reveals widespread full phosphorylation site occupancy during mitosis. Sci. Signal. 3, ra3 (2010).

    PubMed  Google Scholar 

  21. Nagano, K. et al. Phosphoproteomic analysis of distinct tumor cell lines in response to nocodazole treatment. Proteomics 9, 2861–2874 (2009).

    CAS  PubMed  Google Scholar 

  22. Kettenbach, A. N. et al. Quantitative phosphoproteomics identifies substrates and functional modules of aurora and polo-like kinase activities in mitotic cells. Sci. Signal. 4, rs5 (2011).

    CAS  PubMed  Google Scholar 

  23. Bian, Y. et al. An enzyme assisted RP-RPLC approach for in-depth analysis of human liver phosphoproteome. J. Proteomics 96, 253–262 (2014).

    CAS  PubMed  Google Scholar 

  24. Mayya, V. et al. Quantitative phosphoproteomic analysis of T cell receptor signaling reveals system-wide modulation of protein–protein interactions. Sci. Signal. 2, ra46 (2009).

    PubMed  Google Scholar 

  25. Kisselev, A. F. & Goldberg, A. L. Monitoring activity and inhibition of 26S proteasomes with fluorogenic peptide substrates. Methods Enzymol. 398, 364–378 (2005).

    CAS  PubMed  Google Scholar 

  26. Besche, H. C. et al. Autoubiquitination of the 26S proteasome on Rpn13 regulates breakdown of ubiquitin conjugates. EMBO J. 33, 1159–1176 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Taipale, M. et al. Quantitative analysis of HSP90-client interactions reveals principles of substrate recognition. Cell 150, 987–1001 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Manning, G., Whyte, D. B., Martinez, R., Hunter, T. & Sudarsanam, S. The protein kinase complement of the human genome. Science 298, 1912–1934 (2002).

    CAS  PubMed  Google Scholar 

  29. Djuranovic, S. et al. Structure and activity of the N-terminal substrate recognition domains in proteasomal ATPases. Mol. Cell 34, 580–590 (2009).

    CAS  PubMed  Google Scholar 

  30. Beckwith, R., Estrin, E., Worden, E. J. & Martin, A. Reconstitution of the 26S proteasome reveals functional asymmetries in its AAA + unfoldase. Nat. Struct. Mol. Biol. 20, 1164–1172 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Liu, C. W. et al. ATP binding and ATP hydrolysis play distinct roles in the function of 26S proteasome. Mol. Cell 24, 39–50 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Smith, D. M. et al. ATP binding to PAN or the 26S ATPases causes association with the 20S proteasome, gate opening, and translocation of unfolded proteins. Mol. Cell 20, 687–698 (2005).

    CAS  PubMed  Google Scholar 

  33. Peth, A., Kukushkin, N., Bosse, M. & Goldberg, A. L. Ubiquitinated proteins activate the proteasomal ATPases by binding to Usp14 or Uch37 homologs. J. Biol. Chem. 288, 7781–7790 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Śledź, P. et al. Structure of the 26S proteasome with ATP-γ S bound provides insights into the mechanism of nucleotide-dependent substrate translocation. Proc. Natl Acad. Sci. USA 110, 7264–7269 (2013).

    PubMed  Google Scholar 

  35. Santarius, T., Shipley, J., Brewer, D., Stratton, M. R. & Cooper, C. S. A census of amplified and overexpressed human cancer genes. Nat. Rev. Cancer 10, 59–64 (2010).

    CAS  PubMed  Google Scholar 

  36. Györffy, B. et al. An online survival analysis tool to rapidly assess the effect of 22,277 genes on breast cancer prognosis using microarray data of 1,809 patients. Breast Cancer Res. Treat. 123, 725–731 (2010).

    PubMed  Google Scholar 

  37. Petrocca, F. et al. A genome-wide siRNA screen identifies proteasome addiction as a vulnerability of basal-like triple-negative breast cancer cells. Cancer Cell 24, 182–196 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Vilchez, D. et al. Increased proteasome activity in human embryonic stem cells is regulated by PSMD11. Nature 489, 304–308 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Shabaneh, T. B. et al. Molecular basis of differential sensitivity of myeloma cells to clinically relevant bolus treatment with bortezomib. PLoS ONE 8, e56132 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Mason, G. G., Murray, R. Z., Pappin, D. & Rivett, A. J. Phosphorylation of ATPase subunits of the 26S proteasome. FEBS Lett. 430, 269–274 (1998).

    CAS  PubMed  Google Scholar 

  41. Bose, S., Stratford, F. L., Broadfoot, K. I., Mason, G. G. & Rivett, A. J. Phosphorylation of 20S proteasome alpha subunit C8 (alpha7) stabilizes the 26S proteasome and plays a role in the regulation of proteasome complexes by gamma-interferon. Biochem. J. 378, 177–184 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Satoh, K., Sasajima, H., Nyoumura, K.-i., Yokosawa, H. & Sawada, H. Assembly of the 26S proteasome is regulated by phosphorylation of the p45/Rpt6 ATPase subunit. Biochemistry 40, 314–319 (2000).

    Google Scholar 

  43. Feng, Y., Longo, D. L. & Ferris, D. K. Polo-like kinase interacts with proteasomes and regulates their activity. Cell Growth Differ. 12, 29–37 (2001).

    CAS  PubMed  Google Scholar 

  44. Zhang, F. et al. Proteasome function is regulated by cyclic AMP-dependent protein kinase through phosphorylation of Rpt6. J. Biol. Chem. 282, 22460–22471 (2007).

    CAS  PubMed  Google Scholar 

  45. Djakovic, S. N., Schwarz, L. A., Barylko, B., DeMartino, G. N. & Patrick, G. N. Regulation of the proteasome by neuronal activity and calcium/calmodulin-dependent protein kinase II. J. Biol. Chem. 284, 26655–26665 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Bingol, B. et al. Autophosphorylated CaMKIIα acts as a scaffold to recruit proteasomes to dendritic spines. Cell 140, 567–578 (2010).

    CAS  PubMed  Google Scholar 

  47. Djakovic, S. N. et al. Phosphorylation of Rpt6 regulates synaptic strength in hippocampal neurons. J. Neurosci. 32, 5126–5131 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Hamilton, A. M. et al. Activity-dependent growth of new dendritic spines is regulated by the proteasome. Neuron 74, 1023–1030 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  49. Ranek, M. J., Terpstra, E. J., Li, J., Kass, D. A. & Wang, X. Protein kinase g positively regulates proteasome-mediated degradation of misfolded proteins. Circulation 128, 365–376 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Aranda, S., Laguna, A. & de la Luna, S. DYRK family of protein kinases: evolutionary relationships, biochemical properties, and functional roles. FASEB J. 25, 449–462 (2011).

    CAS  PubMed  Google Scholar 

  51. Becker, W. Emerging role of DYRK family protein kinases as regulators of protein stability in cell cycle control. Cell Cycle 11, 3389–3394 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Lander, G. C. et al. Complete subunit architecture of the proteasome regulatory particle. Nature 482, 186–191 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  53. da Fonseca, P. C. A., He, J. & Morris, E. P. Molecular model of the human 26S proteasome. Mol. Cell 46, 54–66 (2012).

    PubMed  Google Scholar 

  54. Matyskiela, M. E., Lander, G. C. & Martin, A. Conformational switching of the 26S proteasome enables substrate degradation. Nat. Struct. Mol. Biol. 20, 781–788 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Miller, C. T. et al. Amplification and overexpression of the dual-specificity tyrosine-(Y)-phosphorylation regulated kinase 2 (DYRK2) gene in esophageal and lung adenocarcinomas. Cancer Res. 63, 4136–4143 (2003).

    CAS  PubMed  Google Scholar 

  56. Bonifaci, N. et al. Exploring the link between germline and somatic genetic alterations in breast carcinogenesis. PLoS ONE 5, e14078 (2010).

    PubMed  PubMed Central  Google Scholar 

  57. Gorringe, K. L., Boussioutas, A., Bowtell, D. D. & Melbourne Gastric Cancer Group, P. M. M. A. F. Novel regions of chromosomal amplification at 6p21, 5p13, and 12q14 in gastric cancer identified by array comparative genomic hybridization. Genes Chromosomes Cancer 42, 247–259 (2005).

    CAS  PubMed  Google Scholar 

  58. Taira, N. et al. DYRK2 priming phosphorylation of c-Jun and c-Myc modulates cell cycle progression in human cancer cells. J. Clin. Invest. 122, 859–872 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Guo, X. et al. Axin and GSK3-β control Smad3 protein stability and modulate TGF-β signaling. Genes Dev. 22, 106–120 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  60. Inobe, T., Fishbain, S., Prakash, S. & Matouschek, A. Defining the geometry of the two-component proteasome degron. Nat. Chem. Biol. 7, 161167 (2011).

    Google Scholar 

  61. Fishbain, S. et al. Sequence composition of disordered regions fine-tunes protein half-life. Nat. Struct. Mol. Biol. 22, 214221 (2015).

    Google Scholar 

  62. Soundararajan, M. et al. Structures of Down syndrome kinases, DYRKs, reveal mechanisms of kinase activation and substrate recognition. Structure 21, 986–996 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Kaake, R. M., Milenković, T., Pržulj, N., Kaiser, P. & Huang, L. Characterization of cell cycle specific protein interaction networks of the yeast 26S proteasome complex by the QTAX strategy. J. Proteome Res. 9, 2016–2029 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  64. Ong, S. E. & Mann, M. A practical recipe for stable isotope labeling by amino acids in cell culture (SILAC). Nat. Protoc. 1, 2650–2660 (2006).

    CAS  PubMed  Google Scholar 

  65. Guo, X. et al. UBLCP1 is a 26S proteasome phosphatase that regulates nuclear proteasome activity. Proc. Natl Acad. Sci. USA 108, 18649–18654 (2011).

    CAS  PubMed  Google Scholar 

  66. Smith, D. M., Fraga, H., Reis, C., Kafri, G. & Goldberg, A. L. ATP binds to proteasomal ATPases in pairs with distinct functional effects, implying an ordered reaction cycle. Cell 144, 526–538 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  67. Jacobson, A. D., MacFadden, A., Wu, Z., Peng, J. & Liu, C.-W. Autoregulation of the 26S proteasome by in situ ubiquitination. Mol. Biol. Cell 25, 1824–1835 (2014).

    PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  69. Rape, M. & Kirschner, M. W. Autonomous regulation of the anaphase-promoting complex couples mitosis to S-phase entry. Nature 432, 588–595 (2004).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

The authors would like to thank members of the Dixon laboratory and K.-L. Guan, E. Bennett, B. Yang, M. Kaulich and G. Lander for insightful suggestions and discussions of the manuscript. We acknowledge E. Durrant and A. Ly for excellent technical assistance. We thank S. Murata, A. Matouschek, S. Lindquist and W. Becker for providing crucial cDNAs. We owe a debt of gratitude to C. Bashore and A. Martin for the polyubiquitylated GFP substrate. This work was supported by the NIH grants RO1DK018849 to J.E.D., RO1GM074830 to L.H., and 1RO1CA168689 and 1R01CA174869 to J.Y. X.G. was partly supported by the Susan G. Komen postdoctoral fellowship for breast cancer research (KG111280).

Author information

Authors and Affiliations

Authors

Contributions

X.G. and J.E.D. conceived the study. X.G., X.W. and L.H. made the original observation of Rpt3-Thr25 phosphorylation. X.G. designed most experiments. X.G., X.W., Z.W. and S.B. executed the experiments. X.G., X.W., S.B. and L.H. performed data analyses. X.W. and L.H. generated and analysed all mass spectrometry data. Tumour xenograft study was performed by X.G. with J.Y.’s support and guidance. X.G., Z.W. and J.E.D. wrote the paper.

Corresponding authors

Correspondence to Xing Guo or Jack E. Dixon.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 3 Phosphorylation of Rpt3 at Thr25.

(a) Endogenous Rpt3-T25 phosphorylation in multiple cell lines. Rpn11-TBHA was stably engineered in each of the cell lines shown. 26S proteasomes were purified from these cells using streptavidin pulldown and treated with or without λ-phosphatase. Phospho-T25 and total Rpt3 were blotted. MEF, mouse embryonic fibroblast. (b) Differential T25 phosphorylation at G1/S and G2/M phases. Four different breast cancer cell lines were treated with either aphidicolin (Aph, 10 μM, 12 hrs) or nocodazole (Ndz, 100 ng/ml, 16 hrs), and pT25 was probed from whole cell extracts. (c) HaCaT cells infected with pLL3.7-HA-Rpt3-WT or T25V (see Supplementary Figure 3) were treated as in (b), and pT25 was probed from whole cell extracts. The complete lack of T25 phosphorylation in the T25V cells indicates efficient replacement of endogenous Rpt3 with the T25V mutant. (d) An estimation of the fraction of Rpt3 that is phosphorylated at T25. HaCaT cells (parental and T25A knock-in, upper panel) and 293A cells (lower panel) were enriched in G2/M phase by sequential treatments with hydroxyurea (0.4 mM, 24 hrs) and nocodazole (100 ng/ml, 8 hrs). Cells were lysed in buffer containing phosphatase inhibitors, and 100 μg of each cell lysate was mixed with 10 μg of anti-pT25 antibody pre-bound to 100 μl of sheep-anti-rabbit IgG-conjugated Dynabeads (Life Technology). T25-phosphorylated Rpt3 was immunoprecipitated for 1 hr at 4 °C. After extensive washes, the immunocaptured phospho-Rpt3 was run side by side with increasing amounts of whole cell lysates (WCL, serving as standards) on SDS-PAGE. Total Rpt3 was blotted using a rabbit polyclonal primary antibody (Bethyl Laboratory) and a mouse-anti-rabbit, light chain-specific secondary antibody (Jackson ImmunoResearch). For HaCaT cells, 5, 10 and 20 μg of WCL were loaded. For 293A cells, 2.5, 5 and 10 μg of WCL were loaded. Intensity of signals were quantified using the ChemiDoc MP imaging system (Bio-Rad) and representative gels are shown. The amount of immunoprecipitated Rpt3 (i.e. T25-phosphorylated) from 100 μg of lysate equals that of total Rpt3 from approximately 15 μg of WCL, in both cell types. Note that the anti-pT25 antibody is very specific for the phosphorylated form and does not pull down Rpt3-T25A (upper panel). Considering that the majority (but not all) of phospho-Rpt3 was immunodepleted from the input sample (data not shown), we conclude that at least 15% of total Rpt3 is phosphorylated at T25 in these cells under this condition.

Supplementary Figure 4 Rpt3-T25A knock-in impedes cell proliferation.

(a) Rpt3-T25A is properly incorporated into the 26S proteasome. MDA-MB-468 parental cells (P) and the two T25A knock-in clones (#4 and #5) were subjected to anti-Rpt3 IP (Bethyl Laboratories). Rpt3, Rpn2 and 20S subunits from the immunoprecipitates and from the whole cell lysates were blotted. (b) Growth curves of the indicated cells stably expressing Rpt3 shRNA and exogenous Rpt3 (WT or T25V). Results are mean ± s.e.m. from n = 3 independent experiments. p < 0.01, p < 0.05, two-tailed paired Student’s T-test. Source data can be found in Supplementary Table 3. (c) MDA-MB-468 parental and T25A knock-in cells were synchronized in early S phase by HU and treated with CHX for the indicated periods of time. Cell lysates were probed for p21Cip1 and p27Kip1. (d) MDA-MB-468 parental and T25A knock-in cells were synchronized by aphidicolin and released into regular medium. Cell lysates collected at each time point were used for western blotting. (e) MDA-MB-468 parental and T25A knock-in cells were synchronized by aphidicolin and released into regular medium containing nocodazole. Cell cycle progression was analyzed by FACS using BrdU and PI double labeling. (f) Mild treatment with HU (0.4 mM, 24 hours) or Aph (10 μM, 12 hrs) used in this study does not cause DNA damage response in HaCaT cells. Total cell lysates were probed for γ-H2AX. High-concentration HU treatment (2 mM) was shown as a positive control. (g) T25A knock-in does not cause ubiquitin depletion. HaCaT cells were released from aphidicolin block into S phase. Cells were lysed in SDS buffer with 10 mM N-ethylmaleimide (NEM). Lysates were sonicated and probed for free ubiquitin and ubiquitinated histone H2B.

Supplementary Figure 5 Loss of Rpt-T25 phosphorylation inhibits proteasome activity.

(a) A schematic of the pLL3.7 lentiviral construct used for generating stable lines with simultaneous Rpt3 knockdown and ectopic HA-Rpt3 expression. (b) Replacement of endogenous Rpt3 with HA-Rpt3-T25V reduces proteasome activity in multiple cell lines. Cells transduced with pLL3.7-HA-Rpt3 (WT or T25V) were partly synchronized with nocodazole treatment for 12 hours. After anti-HA IP, 26S proteasome activity was measured using Suc-LLVY-AMC (top), and the immunoprecipitates were then boiled and probed for pT25 and HA-Rpt3 (bottom). Results are mean ± s.e.m. from n = 3 independent experiments. p < 0.01,p < 0.05, two-tailed paired Student’s T-test. Source data can be found in Supplementary Table 3.

Supplementary Figure 6 DYRK2 as the Rpt3-T25 kinase.

(a) Overexpression of DYRK kinases increases T25 phosphorylation. 293T cells were transfected with the indicated GFP-tagged DYRK kinases. Total cell extracts were probed with the anti-pT25 antibody. GFP alone was used as a negative control. Note that all five DYRK family members could phosphorylate Rpt3-T25 when overexpressed, with DYRK2 and the closely related DYRK3 being the most active against this site. (b) Knockdown of other DYRKs has little effect on T25 phosphorylation. 293T cells stably expressing control or DYRK shRNAs were transfected with HA-Rpt3 (WT). Phospho-T25 was blotted after anti-HA IP. The remaining mRNA levels of each DYRK kinase in the corresponding knockdown cells were determined by qPCR and shown as percentage of that in control cells. (c) DYRK2 expression is regulated by serum starvation and during cell cycle. HaCaT cells were serum-starved (left) or synchronized (right) as described in Fig. 1. DYRK2 protein and mRNA levels were determined by western blot and qPCR, respectively. p < 0.001,p < 0.05 (two-tailed Student’s T-test, mean ± s.e.m. from n = 3 independent experiments). Source data can be found in Supplementary Table 3. (d) Purification of His-DYRK2-WT and D275N (aa. 74-479) shown by Coomassie staining. The inactive D275N mutant lacks co-translational autophosphorylation and therefore migrates faster on gel than the WT kinase. (e) Tandem mass spectrometry identification of Rpt3-T25 phosphorylation (indicated by arrow) from purified 26S proteasomes treated in vitro with DYRK2.

Supplementary Figure 7 DYRK2 positively regulates proteasome activity.

(a) Guide RNA design for DYRK2 targeting and sequencing verification of DYRK2 disruption in MDA-MB-468 cells. Both guide RNAs target regions close to the start codon, and both clones (#1 and #2) have homozygous single nucleotide insertions. The resulting frame shift leads to a premature stop after 83 amino acids in Clone #1 and after 21 amino acids in Clone #2. (b) Proteasome activity (top) and levels of the indicated proteins (bottom) from parental (P) MDA-MB-468 cells and two independent DYRK2 KO clones were determined as in Fig. 3a. p < 0.01,p < 0.05 (two-tailed Student’s T-test, compared to parental cells, mean ± s.e.m. from n = 3 independent experiments). Source data can be found in Supplementary Table 3. (c) In vitro degradation of casein. DYRK2-treated 26S proteasomes as in Fig. 5f was incubated with BODIPY-labeled casein at 37 °C in the presence of 1 mM ATP. The rate of fluorescence increase is shown as mean ± s.e.m. from n = 3 independent experiments. p < 0.05, two-tailed paired Student’s T-test. Source data can be found in Supplementary Table 3.

Supplementary Figure 8 DYRK2 knockout also sensitizes HaCaT cells to Bortezomib.

(a) Sequencing verification of DYRK2 KO caused by a single nucleotide insertion (highlighted) at the Cas9 cleavage site. (b) Parental and DYRK2 KO HaCaT cells were seeded in triplicates and treated with Bortezomib for 48 hrs, and the MTS assay was performed as in Fig. 7c. p < 0.01, Student’s T-test. n.s., non-significant. Source data can be found in Supplementary Table 3.

Supplementary Figure 9 DYRK2 alterations in cancer.

Overview of genetic changes of the DYRK2 gene across a panel of human cancers (http://www.cbioportal.org).

Supplementary Table 1 SILAC/MS of affinity purified 26S proteasome.
Supplementary Table 2 Sequences of primers, shRNAs and gRNAs.

Supplementary information

Supplementary Information

Supplementary Information (PDF 17767 kb)

Supplementary Table 3

Supplementary Information (XLSX 40 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Guo, X., Wang, X., Wang, Z. et al. Site-specific proteasome phosphorylation controls cell proliferation and tumorigenesis. Nat Cell Biol 18, 202–212 (2016). https://doi.org/10.1038/ncb3289

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/ncb3289

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