Desumoylation of RNA polymerase III lies at the core of the Sumo stress response in yeast

Aurélie Nguéa P, Joseph Robertson, Maria Carmen Herrera, Pierre Chymkowitch, Jorrit M Enserink, Aurélie Nguéa P, Joseph Robertson, Maria Carmen Herrera, Pierre Chymkowitch, Jorrit M Enserink

Abstract

Post-translational modification by small ubiquitin-like modifier (Sumo) regulates many cellular processes, including the adaptive response to various types of stress, referred to as the Sumo stress response (SSR). However, it remains unclear whether the SSR involves a common set of core proteins regardless of the type of stress or whether each particular type of stress induces a stress-specific SSR that targets a unique, largely nonoverlapping set of Sumo substrates. In this study, we used MS and a Gene Ontology approach to identify differentially sumoylated proteins during heat stress, hyperosmotic stress, oxidative stress, nitrogen starvation, and DNA alkylation in Saccharomyces cerevisiae cells. Our results indicate that each stress triggers a specific SSR signature centered on proteins involved in transcription, translation, and chromatin regulation. Strikingly, whereas the various stress-specific SSRs were largely nonoverlapping, all types of stress tested here resulted in desumoylation of subunits of RNA polymerase III, which correlated with a decrease in tRNA synthesis. We conclude that desumoylation and subsequent inhibition of RNA polymerase III constitutes the core of all stress-specific SSRs in yeast.

Keywords: Nutrient starvation; RNA polymerase III; Saccharomyces cerevisiae; gene regulation; mass spectrometry (MS); post-translational modification (PTM); small ubiquitin-like modifier (SUMO); stress response; transcription regulation; transfer RNA (tRNA).

Conflict of interest statement

The authors declare that they have no conflicts of interest with the contents of this article

© 2019 Nguéa P et al.

Figures

Figure 1.
Figure 1.
Characterization of Sumo proteomes in five stress conditions by MS.A, experimental workflow for Ni–NTA enrichment of sumoylated proteins in diverse stress conditions. Exponentially growing cells expressing HIS6-FLAG-SMT3 were exposed to different cellular stressors as indicated on the scheme. The cells were lysed under denaturing conditions, and sumoylated proteins were enriched with Ni–NTA affinity matrix. Sumo substrates in each condition were identified by MS. B and C, heat maps showing specific signatures of differentially sumoylated proteins under different stress conditions. All identified Sumo targets are shown in B, whereas C shows only transcription machineries subunits. Relative changes in the sumoylation status of identified targets after each stress are displayed, where green denotes an increase in sumoylation, and red denotes a decrease. Values from two biological replicates per stress are depicted. The asterisk to the right of each heat map highlights proteins displaying a SSR that is consistent across all five stresses. D and E, Circos plot representation of overlaps between the different stresses for proteins enriched in Sumo pulldowns. The outer arcs (multiple colors) on each plot represent a given stress, whereas the inner arcs (blue) represent proteins depleted (D) or enriched (E) from Sumo pulldowns in the given stress. Lines connect the same proteins shared by multiple stresses. F and G, heat map of differentially enriched biological processes (F) or cellular components (G) based on GO identifiers generated using Metascape (52). The GO terms were assigned from the list of proteins identified in Sumo pulldowns in stressed versus unstressed samples. The respective −log10 (p values) are visualized with a color scale ranging from 0 for no representation to +20 for overrepresentation. HS, heat stress; Os, osmotic stress; Ox, oxidative stress; DA, DNA alkylation; -N, nitrogen starvation.
Figure 2.
Figure 2.
Identification of sumo-modified complexes: heat stress and osmotic stress. Networks generated for differentially sumoylated substrates following heat stress (A) and osmotic stress (B). Nodes represent proteins, and edges represent protein–protein associations. Desumoylated proteins are circled in red, whereas hypersumoylated proteins are circled in blue. Enriched complexes are depicted with different colors (see legend). Black nodes represent proteins not associated with enriched complexes. Networks were generated using the STRING database (55).
Figure 3.
Figure 3.
Identification of sumo-modified complexes: oxidative stress, DNA alkylation, and nitrogen starvation. Networks were generated for differentially sumoylated substrates following oxidative stress (A), DNA alkylation (B), and nitrogen starvation (C). Nodes represent proteins and edges represent protein–protein associations. Desumoylated proteins are circled in red, whereas more sumoylated proteins are circled in blue. Enriched complexes are depicted with different colors (see legend). Black nodes represent proteins not associated with enriched complexes. Networks were generated using the STRING database (55).
Figure 4.
Figure 4.
Desumoylation of Ret1 and Rpc82 upon stress rewire their binding to class III genes and affect RNAPIII transcriptional activity.A and B, validation of MS data. Sumo pulldowns were performed under denaturing conditions using cell lysates of 6HF-SMT3 cells grown in the indicated conditions. The levels of copurifying Rpc82-HA and Ret1-MYC were analyzed by Western blotting (WB) using anti-HA (A) and anti-MYC (B) antibodies, respectively. The lower graphs display the quantification of Rpc82-HA (A) and Ret1-MYC (B) signals. Error bars indicate the deviation from the average of three different experiments. *, p < 0.05; **, p < 0.01; ***, p < 0.001. n.s., nonsignificant (Student's t test); UT, untagged; NT, no treatment; HS, heat stress; Os, osmotic stress; Ox, oxidative stress; DA, DNA alkylation; -N, nitrogen starvation. C–E, stress reduces the recruitment of RNAPIII holoenzyme and general transcription factor TFIIIB at class III genes. The occupancy of RNAPIII and TFIIIB subunits were analyzed by ChIP-qPCR. Yeasts were stressed as indicated in Fig. 1, and the relative levels of Rpc82-HA (C), Ret1-MYC (D), and Brf1-GFP (E) at tDNALeu and SCR1 were determined by ChIP-qPCR using anti-HA, -MYC, and -GFP antibodies, respectively. F, schematic representation of the regions where primers used for ChIP anneal. G, schematic representation or Norther blotting probes. H, precursor tRNAs (pre-tRNA) do not accumulate under stress. Pre-tRNA levels were analyzed by Northern blotting. Northern blots are representative of three independent biological replicates of experiments. WT cells were grown at 30 °C to exponential phase and then stressed as specified in Fig. 1. Total RNA was isolated and analyzed by Northern hybridization with oligonucleotide probes specific to precursor and mature tRNALeu(CAA) (top panel) and tRNATrp(CCA) (middle panel). Loading control is provided by ethidium bromide–stained 5.8S rRNA (bottom panel). I, ratios of precursor over mature tRNAs. Errors bars, S.E. of three different experiments. *, p < 0.05; **, p < 0.01; ***, p < 0.001.

References

    1. Wilkinson K. A., and Henley J. M. (2010) Mechanisms, regulation and consequences of protein SUMOylation. Biochem. J. 428, 133–145 10.1042/BJ20100158
    1. Giaever G., Chu A. M., Ni L., Connelly C., Riles L., Véronneau S., Dow S., Lucau-Danila A., Anderson K., André B., Arkin A. P., Astromoff A., El-Bakkoury M., Bangham R., Benito R., et al. (2002) Functional profiling of the Saccharomyces cerevisiae genome. Nature 418, 387–391 10.1038/nature00935
    1. Li S. J., and Hochstrasser M. (1999) A new protease required for cell-cycle progression in yeast. Nature 398, 246–251 10.1038/18457
    1. Li S. J., and Hochstrasser M. (2000) The yeast ULP2 (SMT4) gene encodes a novel protease specific for the ubiquitin-like Smt3 protein. Mol. Cell. Biol. 20, 2367–2377 10.1128/MCB.20.7.2367-2377.2000
    1. Desterro J. M., Thomson J., and Hay R. T. (1997) Ubch9 conjugates SUMO but not ubiquitin. FEBS Lett. 417, 297–300 10.1016/S0014-5793(97)01305-7
    1. Johnson E. S., and Blobel G. (1997) Ubc9p is the conjugating enzyme for the ubiquitin-like protein Smt3p. J. Biol. Chem. 272, 26799–26802 10.1074/jbc.272.43.26799
    1. Gareau J. R., and Lima C. D. (2010) The SUMO pathway: emerging mechanisms that shape specificity, conjugation and recognition. Nat. Rev. Mol. Cell Biol. 11, 861–871 10.1038/nrm3011
    1. Takahashi Y., Toh-e A., and Kikuchi Y. (2001) A novel factor required for the SUMO1/Smt3 conjugation of yeast septins. Gene 275, 223–231 10.1016/S0378-1119(01)00662-X
    1. Smet-Nocca C., Wieruszeski J. M., Léger H., Eilebrecht S., and Benecke A. (2011) SUMO-1 regulates the conformational dynamics of thymine–DNA glycosylase regulatory domain and competes with its DNA binding activity. BMC Biochem. 12, 4 10.1186/1471-2091-12-4
    1. Ulrich H. D. (2005) SUMO modification: wrestling with protein conformation. Curr. Biol. 15, R257–R259 10.1016/j.cub.2005.03.021
    1. Kolesar P., Altmannova V., Silva S., Lisby M., and Krejci L. (2016) Pro-recombination role of Srs2 protein requires SUMO (small ubiquitin-like modifier) but is independent of PCNA (proliferating cell nuclear antigen) interaction. J. Biol. Chem. 291, 7594–7607 10.1074/jbc.M115.685891
    1. Li Y. J., Stark J. M., Chen D. J., Ann D. K., and Chen Y. (2010) Role of SUMO:SIM-mediated protein–protein interaction in non-homologous end joining. Oncogene 29, 3509–3518 10.1038/onc.2010.108
    1. Rojas-Fernandez A., Plechanovová A., Hattersley N., Jaffray E., Tatham M. H., and Hay R. T. (2014) SUMO chain-induced dimerization activates RNF4. Mol. Cell 53, 880–892 10.1016/j.molcel.2014.02.031
    1. Wu C. S., Ouyang J., Mori E., Nguyen H. D., Maréchal A., Hallet A., Chen D. J., and Zou L. (2014) SUMOylation of ATRIP potentiates DNA damage signaling by boosting multiple protein interactions in the ATR pathway. Genes Dev. 28, 1472–1484 10.1101/gad.238535.114
    1. Churikov D., Charifi F., Eckert-Boulet N., Silva S., Simon M. N., Lisby M., and Géli V. (2016) SUMO-dependent relocalization of eroded telomeres to nuclear pore complexes controls telomere recombination. Cell Reports 15, 1242–1253 10.1016/j.celrep.2016.04.008
    1. Escobar-Ramirez A., Vercoutter-Edouart A. S., Mortuaire M., Huvent I., Hardivillé S., Hoedt E., Lefebvre T., and Pierce A. (2015) Modification by SUMOylation controls both the transcriptional activity and the stability of delta-lactoferrin. PLoS One 10, e0129965 10.1371/journal.pone.0129965
    1. Van Rechem C., Boulay G., Pinte S., Stankovic-Valentin N., Guérardel C., and Leprince D. (2010) Differential regulation of HIC1 target genes by CtBP and NuRD, via an acetylation/SUMOylation switch, in quiescent versus proliferating cells. Mol. Cell. Biol. 30, 4045–4059 10.1128/MCB.00582-09
    1. Desterro J. M., Keegan L. P., Jaffray E., Hay R. T., O'Connell M. A., and Carmo-Fonseca M. (2005) SUMO-1 modification alters ADAR1 editing activity. Mol. Biol. Cell 16, 5115–5126 10.1091/mbc.e05-06-0536
    1. El McHichi B., Regad T., Maroui M. A., Rodriguez M. S., Aminev A., Gerbaud S., Escriou N., Dianoux L., and Chelbi-Alix M. K. (2010) SUMOylation promotes PML degradation during encephalomyocarditis virus infection. J. Virol. 84, 11634–11645 10.1128/JVI.01321-10
    1. Psakhye I., and Jentsch S. (2012) Protein group modification and synergy in the SUMO pathway as exemplified in DNA repair. Cell 151, 807–820 10.1016/j.cell.2012.10.021
    1. Chymkowitch P., Nguéa P. A., and Enserink J. M. (2015) SUMO-regulated transcription: challenging the dogma. Bioessays 37, 1095–1105 10.1002/bies.201500065
    1. Chymkowitch P., Nguéa A. P., Aanes H., Koehler C. J., Thiede B., Lorenz S., Meza-Zepeda L. A., Klungland A., and Enserink J. M. (2015) Sumoylation of Rap1 mediates the recruitment of TFIID to promote transcription of ribosomal protein genes. Genome Res. 25, 897–906 10.1101/gr.185793.114
    1. Chymkowitch P., Nguéa P. A., Aanes H., Robertson J., Klungland A., and Enserink J. M. (2017) TORC1-dependent sumoylation of Rpc82 promotes RNA polymerase III assembly and activity. Proc. Natl. Acad. Sci. U.S.A. 114, 1039–1044 10.1073/pnas.1615093114
    1. Enserink J. M. (2015) Sumo and the cellular stress response. Cell Div. 10, 4 10.1186/s13008-015-0010-1
    1. Lewicki M. C., Srikumar T., Johnson E., and Raught B. (2015) The S. cerevisiae SUMO stress response is a conjugation-deconjugation cycle that targets the transcription machinery. J. Proteomics 118, 39–48 10.1016/j.jprot.2014.11.012
    1. Johnson E. S., and Blobel G. (1999) Cell cycle–regulated attachment of the ubiquitin-related protein SUMO to the yeast septins. J. Cell Biol. 147, 981–994 10.1083/jcb.147.5.981
    1. Miller M. J., and Vierstra R. D. (2011) Mass spectrometric identification of SUMO substrates provides insights into heat stress-induced SUMOylation in plants. Plant Signal. Behav. 6, 130–133 10.4161/psb.6.1.14256
    1. Akache B., and Turcotte B. (2002) New regulators of drug sensitivity in the family of yeast zinc cluster proteins. J. Biol. Chem. 277, 21254–21260 10.1074/jbc.M202566200
    1. Mendizabal I., Rios G., Mulet J. M., Serrano R., and de Larrinoa I. F. (1998) Yeast putative transcription factors involved in salt tolerance. FEBS Lett. 425, 323–328 10.1016/S0014-5793(98)00249-X
    1. Zhao X. J., Raitt D., Burke P. V., Clewell A. S, Kwast K. E., and Poyton R. O. (1996) Function and expression of flavohemoglobin in Saccharomyces cerevisiae: evidence for a role in the oxidative stress response. J. Biol. Chem. 271, 25131–25138 10.1074/jbc.271.41.25131
    1. Sri Theivakadadcham V. S., Bergey B. G., and Rosonina E. (2019) Sumoylation of DNA-bound transcription factor Sko1 prevents its association with nontarget promoters. PLoS Genet. 15, e1007991 10.1371/journal.pgen.1007991
    1. Hinnebusch A. G., and Natarajan K. (2002) Gcn4p, a master regulator of gene expression, is controlled at multiple levels by diverse signals of starvation and stress. Eukaryot. Cell 1, 22–32 10.1128/EC.01.1.22-32.2002
    1. Esteras M., Liu I. C., Snijders A. P., Jarmuz A., and Aragon L. (2017) Identification of SUMO conjugation sites in the budding yeast proteome. Microb. Cell 4, 331–341 10.15698/mic2017.10.593
    1. Motorin Y., Muller S., Behm-Ansmant I., and Branlant C. (2007) Identification of modified residues in RNAs by reverse transcription-based methods. Methods Enzymol. 425, 21–53 10.1016/S0076-6879(07)25002-5
    1. Wilusz J. E. (2015) Removing roadblocks to deep sequencing of modified RNAs. Nat. Methods 12, 821–822 10.1038/nmeth.3516
    1. Manza L. L., Codreanu S. G., Stamer S. L., Smith D. L., Wells K. S., Roberts R. L., and Liebler D. C. (2004) Global shifts in protein sumoylation in response to electrophile and oxidative stress. Chem. Res. Toxicol. 17, 1706–1715 10.1021/tx049767l
    1. Niskanen E. A., Malinen M., Sutinen P., Toropainen S., Paakinaho V., Vihervaara A., Joutsen J., Kaikkonen M. U., Sistonen L., and Palvimo J. J. (2015) Global SUMOylation on active chromatin is an acute heat stress response restricting transcription. Genome Biol. 16, 153 10.1186/s13059-015-0717-y
    1. Xiao Z., Chang J. G., Hendriks I. A., Sigurethsson J. O., Olsen J. V., and Vertegaal A. C. (2015) System-wide analysis of SUMOylation dynamics in response to replication stress reveals novel small ubiquitin-like modified target proteins and acceptor lysines relevant for genome stability. Mol. Cell. Proteomics 14, 1419–1434 10.1074/mcp.O114.044792
    1. Loewith R., and Hall M. N. (2011) Target of rapamycin (TOR) in nutrient signaling and growth control. Genetics 189, 1177–1201 10.1534/genetics.111.133363
    1. Chymkowitch P., and Enserink J. M. (2018) Regulation of tRNA synthesis by posttranslational modifications of RNA polymerase III subunits. Biochim. Biophys. Acta Gene Regul. Mech. 1861, 310–319 10.1016/j.bbagrm.2017.11.001
    1. Khoo S. K., Wu C. C., Lin Y. C., and Chen H. T. (2018) The TFIIE-related Rpc82 subunit of RNA polymerase III interacts with the TFIIB-related transcription factor Brf1 and the polymerase cleft for transcription initiation. Nucleic Acids Res. 46, 1157–1166 10.1093/nar/gkx1179
    1. Wu C. C., Lin Y. C., and Chen H. T. (2011) The TFIIF-like Rpc37/53 dimer lies at the center of a protein network to connect TFIIIC, Bdp1, and the RNA polymerase III active center. Mol. Cell. Biol. 31, 2715–2728 10.1128/MCB.05151-11
    1. Hendriks I. A., D'Souza R. C., Chang J. G., Mann M., and Vertegaal A. C. (2015) System-wide identification of wild-type SUMO-2 conjugation sites. Nat. Commun. 6, 7289 10.1038/ncomms8289
    1. Hendriks I. A., Treffers L. W., Verlaan-de Vries M., Olsen J. V., and Vertegaal A. C. (2015) SUMO-2 orchestrates chromatin modifiers in response to DNA damage. Cell Reports 10, 1778–1791 10.1016/j.celrep.2015.02.033
    1. Cossec J. C., Theurillat I., Chica C., Búa Aguín S., Gaume X., Andrieux A., Iturbide A., Jouvion G., Li H., Bossis G., Seeler J. S., Torres-Padilla M. E., and Dejean A. (2018) SUMO safeguards somatic and pluripotent cell identities by enforcing distinct chromatin states. Cell Stem Cell 23, 742–757.e8 10.1016/j.stem.2018.10.001
    1. Neyret-Kahn H., Benhamed M., Ye T., Le Gras S., Cossec J. C., Lapaquette P., Bischof O., Ouspenskaia M., Dasso M., Seeler J., Davidson I., and Dejean A. (2013) Sumoylation at chromatin governs coordinated repression of a transcriptional program essential for cell growth and proliferation. Genome Res. 23, 1563–1579 10.1101/gr.154872.113
    1. Cieśla M., and Boguta M. (2008) Regulation of RNA polymerase III transcription by Maf1 protein. Acta Biochim. Pol. 55, 215–225
    1. Willis I. M., and Moir R. D. (2007) Integration of nutritional and stress signaling pathways by Maf1. Trends Biochem. Sci. 32, 51–53 10.1016/j.tibs.2006.12.001
    1. Chen C., and Kolodner R. D. (1999) Gross chromosomal rearrangements in Saccharomyces cerevisiae replication and recombination defective mutants. Nat. Genet. 23, 81–85 10.1038/12687
    1. Cox J., and Mann M. (2008) MaxQuant enables high peptide identification rates, individualized p.p.b.-range mass accuracies and proteome-wide protein quantification. Nat. Biotechnol. 26, 1367–1372 10.1038/nbt.1511
    1. Tyanova S., Temu T., Sinitcyn P., Carlson A., Hein M. Y., Geiger T., Mann M., and Cox J. (2016) The Perseus computational platform for comprehensive analysis of (prote)omics data. Nat. Methods 13, 731–740 10.1038/nmeth.3901
    1. Zhou Y., Zhou B., Pache L., Chang M., Khodabakhshi A. H., Tanaseichuk O., Benner C., and Chanda S. K. (2019) Metascape provides a biologist-oriented resource for the analysis of systems-level datasets. Nat. Commun. 10, 1523 10.1038/s41467-019-09234-6
    1. Kruitwagen T., Chymkowitch P., Denoth-Lippuner A., Enserink J., and Barral Y. (2018) Centromeres license the mitotic condensation of yeast chromosome arms. Cell 175, 780–795.e15 10.1016/j.cell.2018.09.012
    1. Perez-Riverol Y., Csordas A., Bai J., Bernal-Llinares M., Hewapathirana S., Kundu D. J., Inuganti A., Griss J., Mayer G., Eisenacher M., Pérez E., Uszkoreit J., Pfeuffer J., Sachsenberg T., Yilmaz S., et al. (2019) The PRIDE database and related tools and resources in 2019: improving support for quantification data. Nucleic Acids Res. 47, D442–D450 10.1093/nar/gky1106
    1. Jensen L. J., Kuhn M., Stark M., Chaffron S., Creevey C., Muller J., Doerks T., Julien P., Roth A., Simonovic M., Bork P., and von Mering C. (2009) STRING 8: a global view on proteins and their functional interactions in 630 organisms. Nucleic Acids Res. 37, D412–D416 10.1093/nar/gkn760

Source: PubMed

3
Abonnere