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:

Decoding the selectivity of eIF2α holophosphatases and PPP1R15A inhibitors

Abstract

The reversible phosphorylation of proteins controls most cellular functions. Protein kinases have been popular drug targets, unlike phosphatases, which remain a drug discovery challenge. Guanabenz and Sephin1 are selective inhibitors of the phosphatase regulatory subunit PPP1R15A (R15A) that prolong the benefit of eIF2α phosphorylation, thereby protecting cells from proteostatic defects. In mice, Sephin1 prevents two neurodegenerative diseases, Charcot–Marie–Tooth 1B (CMT-1B) and SOD1-mediated amyotrophic lateral sclerosis (ALS). However, the molecular basis for R15A inhibition is unknown. Here we reconstituted human recombinant eIF2α holophosphatases, R15A–PP1 and R15B–PP1, whose activity depends on both the catalytic subunit PP1 (protein phosphatase 1) and either R15A or R15B. This system enabled the functional characterization of these holophosphatases and revealed that Guanabenz and Sephin1 induced a selective conformational change in R15A, detected by resistance to limited proteolysis. This altered the recruitment of eIF2α, preventing its dephosphorylation. This work demonstrates that regulatory subunits of phosphatases are valid drug targets and provides the molecular rationale to expand this concept to other phosphatases.

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: Reconstitution of functional eIF2α holophosphatases with recombinant proteins.
Figure 2: Defining the R15 domains required for PP1 binding and eIF2α holophosphatase activity.
Figure 3: Functional R15 holoenzymes have a higher affinity for their substrate than PP1.
Figure 4: R15A inhibitors alter the protease sensitivity of R15A and selectively decrease substrate binding to R15A.
Figure 5: An activity assay with functional recombinant R15 holoenzymes recapitulates selective inhibition of R15A by Guanabenz and Sephin1.

Similar content being viewed by others

References

  1. Rask-Andersen, M., Zhang, J., Fabbro, D. & Schiöth, H.B. Advances in kinase targeting: current clinical use and clinical trials. Trends Pharmacol. Sci. 35, 604–620 (2014).

    Article  CAS  Google Scholar 

  2. Tonks, N.K. Protein tyrosine phosphatases—from housekeeping enzymes to master regulators of signal transduction. FEBS J. 280, 346–378 (2013).

    Article  CAS  Google Scholar 

  3. Tsaytler, P. & Bertolotti, A. Exploiting the selectivity of protein phosphatase 1 for pharmacological intervention. FEBS J. 280, 766–770 (2013).

    Article  CAS  Google Scholar 

  4. Gilmartin, A.G. et al. Allosteric Wip1 phosphatase inhibition through flap-subdomain interaction. Nat. Chem. Biol. 10, 181–187 (2014).

    Article  CAS  Google Scholar 

  5. Chen, Y.-N.P. Allosteric inhibition of SHP2 phosphatase inhibits cancers driven by receptor tyrosine kinases. Nature 535, 148–152 (2016).

    Article  CAS  Google Scholar 

  6. Hubbard, M.J. & Cohen, P. On target with a new mechanism for the regulation of protein phosphorylation. Trends Biochem. Sci. 18, 172–177 (1993).

    Article  CAS  Google Scholar 

  7. Cohen, P.T. Protein phosphatase 1–targeted in many directions. J. Cell Sci. 115, 241–256 (2002).

    CAS  PubMed  Google Scholar 

  8. Heroes, E. et al. The PP1 binding code: a molecular-lego strategy that governs specificity. FEBS J. 280, 584–595 (2013).

    Article  CAS  Google Scholar 

  9. Brautigan, D.L. Protein Ser/Thr phosphatases—the ugly ducklings of cell signalling. FEBS J. 280, 324–345 (2013).

    Article  CAS  Google Scholar 

  10. Roy, J. & Cyert, M.S. Cracking the phosphatase code: docking interactions determine substrate specificity. Sci. Signal. 2, re9 (2009).

    Article  Google Scholar 

  11. Virshup, D.M. & Shenolikar, S. From promiscuity to precision: protein phosphatases get a makeover. Mol. Cell 33, 537–545 (2009).

    Article  CAS  Google Scholar 

  12. De Munter, S., Köhn, M. & Bollen, M. Challenges and opportunities in the development of protein phosphatase-directed therapeutics. ACS Chem. Biol. 8, 36–45 (2013).

    Article  CAS  Google Scholar 

  13. Ron, D. & Harding, H.P. in Translational Control in Biology and Medicine Vol. 48 345–368 (Cold Spring Harbor Laboratory Press, 2007).

  14. Schneider, K. & Bertolotti, A. Surviving protein quality control catastrophes–from cells to organisms. J. Cell Sci. 128, 3861–3869 (2015).

    Article  CAS  Google Scholar 

  15. Novoa, I., Zeng, H., Harding, H.P. & Ron, D. Feedback inhibition of the unfolded protein response by GADD34-mediated dephosphorylation of eIF2alpha. J. Cell Biol. 153, 1011–1022 (2001).

    Article  CAS  Google Scholar 

  16. Jousse, C. et al. Inhibition of a constitutive translation initiation factor 2alpha phosphatase, CReP, promotes survival of stressed cells. J. Cell Biol. 163, 767–775 (2003).

    Article  CAS  Google Scholar 

  17. Tsaytler, P., Harding, H.P., Ron, D. & Bertolotti, A. Selective inhibition of a regulatory subunit of protein phosphatase 1 restores proteostasis. Science 332, 91–94 (2011).

    Article  CAS  Google Scholar 

  18. Harding, H.P. et al. Ppp1r15 gene knockout reveals an essential role for translation initiation factor 2 alpha (eIF2alpha) dephosphorylation in mammalian development. Proc. Natl. Acad. Sci. USA 106, 1832–1837 (2009).

    Article  CAS  Google Scholar 

  19. Scheuner, D. et al. Double-stranded RNA-dependent protein kinase phosphorylation of the alpha-subunit of eukaryotic translation initiation factor 2 mediates apoptosis. J. Biol. Chem. 281, 21458–21468 (2006).

    Article  CAS  Google Scholar 

  20. Boyce, M. et al. A selective inhibitor of eIF2alpha dephosphorylation protects cells from ER stress. Science 307, 935–939 (2005).

    Article  CAS  Google Scholar 

  21. Balch, W.E., Morimoto, R.I., Dillin, A. & Kelly, J.W. Adapting proteostasis for disease intervention. Science 319, 916–919 (2008).

    Article  CAS  Google Scholar 

  22. Holmes, B., Brogden, R.N., Heel, R.C., Speight, T.M. & Avery, G.S. Guanabenz. A review of its pharmacodynamic properties and therapeutic efficacy in hypertension. Drugs 26, 212–229 (1983).

    Article  CAS  Google Scholar 

  23. Hall, A.H., Smolinske, S.C., Kulig, K.W. & Rumack, B.H. Guanabenz overdose. Ann. Intern. Med. 102, 787–788 (1985).

    Article  CAS  Google Scholar 

  24. Tribouillard-Tanvier, D. et al. Antihypertensive drug guanabenz is active in vivo against both yeast and mammalian prions. PLoS One 3, e1981 (2008).

    Article  Google Scholar 

  25. Das, I. et al. Preventing proteostasis diseases by selective inhibition of a phosphatase regulatory subunit. Science 348, 239–242 (2015).

    Article  CAS  Google Scholar 

  26. Pennuto, M. et al. Ablation of the UPR-mediator CHOP restores motor function and reduces demyelination in Charcot-Marie-Tooth 1B mice. Neuron 57, 393–405 (2008).

    Article  CAS  Google Scholar 

  27. D'Antonio, M. et al. Resetting translational homeostasis restores myelination in Charcot-Marie-Tooth disease type 1B mice. J. Exp. Med. 210, 821–838 (2013).

    Article  CAS  Google Scholar 

  28. Wang, L., Popko, B. & Roos, R.P. An enhanced integrated stress response ameliorates mutant SOD1-induced ALS. Hum. Mol. Genet. 23, 2629–2638 (2014).

    Article  CAS  Google Scholar 

  29. Choy, M.S. et al. Structural and functional analysis of the GADD34:PP1 eIF2α phosphatase. Cell Rep. 11, 1885–1891 (2015).

    Article  CAS  Google Scholar 

  30. Crespillo-Casado, A., Chambers, J.E., Fischer, P.M., Marciniak, S.J. & Ron, D. PPP1R15A-mediated dephosphorylation of eIF2α is unaffected by Sephin1 or Guanabenz. eLife 6, e26109 (2017).

    Article  Google Scholar 

  31. Cohen, P.T.W., Browne, G.J., Delibegovic, M. & Munro, S. Assay of protein phosphatase 1 complexes. Methods Enzymol. 366, 135–144 (2003).

    CAS  PubMed  Google Scholar 

  32. Hendrickx, A. et al. Docking motif-guided mapping of the interactome of protein phosphatase-1. Chem. Biol. 16, 365–371 (2009).

    Article  CAS  Google Scholar 

  33. Connor, J.H., Weiser, D.C., Li, S., Hallenbeck, J.M. & Shenolikar, S. Growth arrest and DNA damage-inducible protein GADD34 assembles a novel signaling complex containing protein phosphatase 1 and inhibitor 1. Mol. Cell. Biol. 21, 6841–6850 (2001).

    Article  CAS  Google Scholar 

  34. Boens, S., Szekér, K., Van Eynde, A. & Bollen, M. Interactor-guided dephosphorylation by protein phosphatase-1. Methods Mol. Biol. 1053, 271–281 (2013).

    Article  CAS  Google Scholar 

  35. Peti, W., Nairn, A.C. & Page, R. Structural basis for protein phosphatase 1 regulation and specificity. FEBS J. 280, 596–611 (2013).

    Article  CAS  Google Scholar 

  36. Verbinnen, I., Ferreira, M. & Bollen, M. Biogenesis and activity regulation of protein phosphatase 1. Biochem. Soc. Trans. 45, 89–99 (2017).

    Article  CAS  Google Scholar 

  37. Zhang, J., Zhang, Z., Brew, K. & Lee, E.Y. Mutational analysis of the catalytic subunit of muscle protein phosphatase-1. Biochemistry 35, 6276–6282 (1996).

    Article  CAS  Google Scholar 

  38. Huang, H.B., Horiuchi, A., Goldberg, J., Greengard, P. & Nairn, A.C. Site-directed mutagenesis of amino acid residues of protein phosphatase 1 involved in catalysis and inhibitor binding. Proc. Natl. Acad. Sci. USA 94, 3530–3535 (1997).

    Article  CAS  Google Scholar 

  39. Rojas, M., Vasconcelos, G. & Dever, T.E. An eIF2α-binding motif in protein phosphatase 1 subunit GADD34 and its viral orthologs is required to promote dephosphorylation of eIF2α. Proc. Natl. Acad. Sci. USA 112, E3466–E3475 (2015).

    Article  CAS  Google Scholar 

  40. Krishnan, N. et al. Targeting the disordered C terminus of PTP1B with an allosteric inhibitor. Nat. Chem. Biol. 10, 558–566 (2014).

    Article  CAS  Google Scholar 

  41. Beullens, M., Stalmans, W. & Bollen, M. The biochemical identification and characterization of new species of protein phosphatase 1. Methods Mol. Biol. 93, 145–155 (1998).

    CAS  PubMed  Google Scholar 

  42. Ito, T., Marintchev, A. & Wagner, G. Solution structure of human initiation factor eIF2alpha reveals homology to the elongation factor eEF1B. Structure 12, 1693–1704 (2004).

    Article  CAS  Google Scholar 

  43. McAvoy, T. & Nairn, A.C. Serine/threonine protein phosphatase assays. Curr. Protoc. Mol. Biol. 92, 18.18.1–18.18.11 (2010).

    Google Scholar 

  44. Wilcken, R., Wang, G.Z. & Boeckler, F.M. Kinetic mechanism of p53 oncogenic mutant aggregation and its inhibition. Proc. Natl. Acad. Sci. USA 109, 13584–13589 (2012).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank members of the Bertolotti laboratory for advice and discussions, S. McLaughlin for help with light scattering and M. Goedert and R. Taylor for comments on the manuscript. This work was supported by the Medical Research Council (UK) grant MC_U105185860 and the European Research Council (ERC) under the European Union's Seventh Framework Programme (FP7/2007-2013)/ERC grant 309516. A.B. is an honorary fellow of the Clinical Neurosciences Department of Cambridge University.

Author information

Authors and Affiliations

Authors

Contributions

M.C. designed, performed and analyzed all experiments, prepared the figures and helped with the manuscript. A.S. discovered compound C3 and performed cytotoxicity experiments. A.B. designed and guided the study and wrote the manuscript.

Corresponding author

Correspondence to Anne Bertolotti.

Ethics declarations

Competing interests

M.C. and A.B. are co-inventors on Great Britain patent application 1709927.6 on the activity assays and methods described in this manuscript.

Integrated supplementary information

Supplementary Figure 1 Coomassie-stained gel of purified recombinant proteins

Coomassie-stained gel of ~ 3 μg of purified R15A, R15B, R3A, eIF2α and PP1 proteins. Molecular weight markers, in kilodalton (kDa), are indicated.

Supplementary Figure 2 Reconstitution of functional eIF2α holophosphatases with recombinant proteins

(a,c,d,e,f) Phos-tag gels of experiments shown in Fig 1. Supplementary Figure 2a,c,d,e,f correspond to samples analyzed by immunoblotting shown in Fig 1a,d,e,f,g respectively. Samples were run on 15% Phos-tag gels and visualized by Coomassie staining. (b) Representative immunoblot of P-eIF2α and eIF2α following a dephosphorylation reaction, of 1 μM P-eIF2α, by increasing amounts of PP1 used for the titration curve in Fig. 1d. The concentration of PP1 used is indicated. Dephosphorylation reactions were carried out at 30 °C for 16 h. Representative results of three independent experiments are shown (n = 3; biological replicates).

Supplementary Figure 3 Functional R15-PP1 do not dephosphorylate 33P Phosphorylase a

Phosphorimaging and Coomassie gel of 33P Phosphorylase a following a dephosphorylation reaction, of 1 μM 33P Phosphorylase a, by PP1 (1 μM and 10 nM) in the presence or absence of 1 μM R15A or R15B. All dephosphorylation reactions were carried out at 30 °C for 16 h. Representative results of three independent experiments are shown (n = 3; biological replicates).

Supplementary Figure 4 PP1D95A is catalytically dead

Immunoblot of P-eIF2α and eIF2α following a dephosphorylation reaction, of 1 μM P-eIF2α, by PP1 (10 nM) or PP1D95A (10 nM) in the presence of 1 μM R15A, or R15B. All dephosphorylation reactions were carried out at 30 °C for 16 h. Representative results of three independent experiments are shown (n = 3; biological replicates).

Supplementary Figure 5 Cytoprotection of Tunicamycin-treated cells by Guanabenz, but not by C3

Dose dependent cytoprotection of HeLa cells by Guanabenz, but not the inactive derivative C3, in response to Tunicamycin stress. Vehicle concentrations represent the corresponding amounts of DMSO used in compound treated samples (from 0 to 0.04% highest concentration). Cell viability after 72 h of treatment was monitored using the IncuCyte ZOOM system. Representative results of three independent experiments are shown (n = 3; biological replicates).

Supplementary Figure 6 Limited trypsin digestion of MBP

Coomassie-stained gel showing limited trypsin digestion of MBP in the presence or absence of Guanabenz or Sephin1. Trypsin digestions were carried out using 2.5 nM of trypsin. Reactions were allowed to proceed for 0 h (first lane in each gel), 30 min, 1 h, 2 h or 3 h at 22 °C. Reactions were terminated by addition of 4% SDS Laemmli sample buffer. Representative results of three independent experiments are shown (n = 3; biological replicates).

Supplementary Figure 7 Aggregation of R15A with Salubrinal but not with saturating concentrations of Guanabenz, Sephin1 or C3

Light scattering measurements of 5 μM R15A in the presence of 1 mM Guanabenz, 1 mM Sephin1, 1 mM C3, 50 μM Salubrinal or DMSO vehicle. Absorbance at 380 nm was monitored over 10 min at 20 °C with constant stirring. 100 data points are plotted for each sample.

Supplementary Figure 8 Binding of recombinant PP1 to R15s in the presence or absence of Guanabenz or Sephin1

Binding of recombinant PP1 to MBP-tagged R15s immobilized on magnetic amylose beads (see methods) in the presence or absence of Guanabenz or Sephin1. Immunoblots of input and bound samples, probed with α-MBP (to reveal R15s) or α-PPP1A (to reveal PP1) antibodies are shown. Representative results of three independent experiments are shown (n = 3; biological replicates).

Supplementary Figure 9 Salubrinal inhibits PP1 and causes aggregation of P-eIF2α, in contrast to Sephin1 and Guanabenz that have no effect in these assays

(a-d) Immunoblots of P-eIF2α and eIF2α following a dephosphorylation reactions, of 1 μM P-eIF2α, by increasing amounts of PP1, in the presence of (a) DMSO vehicle control, (b) Guanabenz, (c) Sephin1, or (d) Salubrinal. The concentration of PP1 used is indicated. All dephosphorylation reactions were carried out at 30 °C for 16 h. (e,f) Images of the effect of adding increasing amounts (from 30 μM up to 2 mM) Salubrinal to (e) 1μM P-eIF2α and (f) 1μM PP1. (g,h) Images of the effect of adding 2 mM Guanabenz or Sephin1 to (g) 1 μM P-eIF2α and (h) 1 μM PP1.

Supplementary Figure 10 Time course of P-eIF2α dephosphorylation by PP1 and R15A-PP1 holophosphatase

Phos-tag gel showing P-eIF2α and eIF2α following a dephosphorylation reaction, of 1μM P-eIF2α, using free PP1 (10 nM) or PP1 (10 nM) plus 50 nM R15A. Dephosphorylation reactions were carried out at 30 °C for the time indicated (min (‘) or hours (h)). Note that similar to what has been reported (Crespillo-Casado, A. et al., eLife. 6, e26109, 2017), no dephosphorylation was seen after 20 min in either condition. Representative results of three independent experiments are shown (n = 3; biological replicates).

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–10 and Supplementary Table 1. (PDF 1011 kb)

Life Sciences Reporting Summary (PDF 140 kb)

Supplementary Table 2

Raw values of Figs. 1, 2c, 2d, 3, 4f, 4h, 5, Supplementary Fig 5, Supplementary Fig 7. (XLSX 145 kb)

Supplementary Data Set 1

Uncropped images of all gels. (PDF 8028 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Carrara, M., Sigurdardottir, A. & Bertolotti, A. Decoding the selectivity of eIF2α holophosphatases and PPP1R15A inhibitors. Nat Struct Mol Biol 24, 708–716 (2017). https://doi.org/10.1038/nsmb.3443

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nsmb.3443

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