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:

Nascent-protein ubiquitination is required for heat shock–induced gene downregulation in human cells

Abstract

Proteotoxic stress such as heat shock causes heat-shock factor (HSF)-dependent transcriptional upregulation of chaperones. Heat shock also leads to a rapid and reversible downregulation of many genes, a process we term stress-induced transcriptional attenuation (SITA). The mechanism underlying this conserved phenomenon is unknown. Here we report that enhanced recruitment of negative transcription elongation factors to gene promoters in human cell lines induces SITA. A chemical inhibitor screen showed that active translation and protein ubiquitination are required for the response. We further find that proteins translated during heat shock are subjected to ubiquitination and that p38 kinase signaling connects cytosolic translation with gene downregulation. Notably, brain samples of subjects with Huntington’s disease also show transcriptional attenuation, which is recapitulated in cellular models of protein aggregation similar to heat shock. Thus our work identifies an HSF-independent mechanism that links nascent-protein ubiquitination to transcriptional downregulation during heat shock, with potential ramifications in neurodegenerative diseases.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: N-TEFs are enriched at chromatin upon heat shock.
Fig. 2: Active translation is required for stress-induced increase of N-TEFs at chromatin and SITA.
Fig. 3: Ubiquitination of nascent proteins is required for SITA.
Fig. 4: p38α kinase (MAPK14)-mediated signaling is required for SITA.
Fig. 5: Acute and chronic proteotoxic stress cause gene downregulation.

Similar content being viewed by others

Data availability

All the deep-sequencing data reported in this study are deposited in Gene Expression Omnibus and are available under accession number GSE112379. The mass spectrometry proteomics data have been deposited with the ProteomeXchange Consortium via the PRIDE partner repository with the data set identifier PXD012077. The source data for Figs. 2d, 3c, 4g and 5c,f are available online. Uncropped images of Figs. 1c,d, 2ac, 3a,b,fj, 4b,c,f and 5a,b,e are available with the paper online. All other data will be made available upon request.

References

  1. Morimoto, R. I. Regulation of the heat shock transcriptional response: cross talk between a family of heat shock factors, molecular chaperones, and negative regulators. Genes Dev. 12, 3788–3796 (1998).

    Article  CAS  PubMed  Google Scholar 

  2. Akerfelt, M., Morimoto, R. I. & Sistonen, L. Heat shock factors: integrators of cell stress, development and lifespan. Nat. Rev. Mol. Cell Biol. 11, 545–555 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Gomez-Pastor, R., Burchfiel, E. T. & Thiele, D. J. Regulation of heat shock transcription factors and their roles in physiology and disease. Nat. Rev. Mol. Cell Biol. 19, 4–19 (2018).

    Article  CAS  PubMed  Google Scholar 

  4. Li, J., Labbadia, J. & Morimoto, R. I. Rethinking HSF1 in stress, development, and organismal health. Trends. Cell Biol. 27, 895–905 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Spradling, A., Penman, S. & Pardue, M. L. Analysis of drosophila mRNA by in situ hybridization: sequences transcribed in normal and heat shocked cultured cells. Cell 4, 395–404 (1975).

    Article  CAS  PubMed  Google Scholar 

  6. Jamrich, M., Greenleaf, A. L. & Bautz, E. K. Localization of RNA polymerase in polytene chromosomes of Drosophila melanogaster. Proc. Natl Acad. Sci. USA 74, 2079–2083 (1977).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Mahat, D. B., Salamanca, H. H., Duarte, F. M., Danko, C. G. & Lis, J. T. Mammalian heat shock response and mechanisms underlying its genome-wide transcriptional regulation. Mol. Cell 62, 63–78 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Gasch, A. P. et al. Genomic expression programs in the response of yeast cells to environmental changes. Mol. Biol. Cell. 11, 4241–4257 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Duarte, F. M. et al. Transcription factors GAF and HSF act at distinct regulatory steps to modulate stress-induced gene activation. Genes Dev. 30, 1731–1746 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Vihervaara, A. et al. Transcriptional response to stress is pre-wired by promoter and enhancer architecture. Nat. Commun. 8, 255 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  11. Core, L. J. & Lis, J. T. Transcription regulation through promoter-proximal pausing of RNA polymerase II. Science 319, 1791–1792 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Adelman, K. & Lis, J. T. Promoter-proximal pausing of RNA polymerase II: emerging roles in metazoans. Nat. Rev. Genet. 13, 720–731 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Kwak, H. & Lis, J. T. Control of transcriptional elongation. Annu. Rev. Genet. 47, 483–508 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Bunch, H. et al. TRIM28 regulates RNA polymerase II promoter-proximal pausing and pause release. Nat. Struct. Mol. Biol. 21, 876–883 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Stadelmayer, B. et al. Integrator complex regulates NELF-mediated RNA polymerase II pause/release and processivity at coding genes. Nat. Commun. 5, 5531 (2014).

    Article  CAS  PubMed  Google Scholar 

  16. Chen, F. X. et al. PAF1, a molecular regulator of promoter-proximal pausing by rna polymerase II. Cell 162, 1003–1015 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Paparidis, N. F., Durvale, M. C. & Canduri, F. The emerging picture of CDK9/P-TEFb: more than 20 years of advances since PITALRE. Mol. Biosyst. 13, 246–276 (2017).

    Article  PubMed  CAS  Google Scholar 

  18. Kwak, H., Fuda, N. J., Core, L. J. & Lis, J. T. Precise maps of RNA polymerase reveal how promoters direct initiation and pausing. Science 339, 950–953 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Palozola, K. C. et al. Mitotic transcription and waves of gene reactivation during mitotic exit. Science 358, 119–122 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Ong, S. E. et al. Stable isotope labeling by amino acids in cell culture, SILAC, as a simple and accurate approach to expression proteomics. Mol. Cell. Proteomics 1, 376–386 (2002).

    Article  CAS  PubMed  Google Scholar 

  21. Gilchrist, D. A. et al. NELF-mediated stalling of Pol II can enhance gene expression by blocking promoter-proximal nucleosome assembly. Genes Dev. 22, 1921–1933 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Chen, F. X. et al. PAF1 regulation of promoter-proximal pause release via enhancer activation. Science 357, 1294–1298 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Court, S. J., Waclaw, B. & Allen, R. J. Lower glycolysis carries a higher flux than any biochemically possible alternative. Nat. Commun. 6, 8427 (2015).

    Article  CAS  PubMed  Google Scholar 

  24. Bersuker, K., Brandeis, M. & Kopito, R. R. Protein misfolding specifies recruitment to cytoplasmic inclusion bodies. J. Cell Biol. 213, 229–241 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Santagata, S. et al. Tight coordination of protein translation and HSF1 activation supports the anabolic malignant state. Science 341, 1238303 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  26. Salomons, F. A. et al. Selective accumulation of aggregation-prone proteasome substrates in response to proteotoxic stress. Mol. Cell. Biol. 29, 1774–1785 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Xu, G. et al. Vulnerability of newly synthesized proteins to proteostasis stress. J. Cell Sci. 129, 1892–1901 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Hanna, J., Leggett, D. S. & Finley, D. Ubiquitin depletion as a key mediator of toxicity by translational inhibitors. Mol. Cell. Biol. 23, 9251–9261 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Abravaya, K., Phillips, B. & Morimoto, R. I. Attenuation of the heat shock response in HeLa cells is mediated by the release of bound heat shock transcription factor and is modulated by changes in growth and in heat shock temperatures. Genes Dev. 5, 2117–2127 (1991).

    Article  CAS  PubMed  Google Scholar 

  30. Duttler, S., Pechmann, S. & Frydman, J. Principles of cotranslational ubiquitination and quality control at the ribosome. Mol. Cell 50, 379–393 (2013).

    Article  CAS  PubMed  Google Scholar 

  31. Wang, F., Durfee, L. A. & Huibregtse, J. M. A cotranslational ubiquitination pathway for quality control of misfolded proteins. Mol. Cell 50, 368–378 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Cuadrado, A. & Nebreda, A. R. Mechanisms and functions of p38 MAPK signalling. Biochem. J. 429, 403–417 (2010).

    Article  CAS  PubMed  Google Scholar 

  33. Davis, R. J. Signal transduction by the JNK group of MAP kinases. Cell 103, 239–252 (2000).

    Article  CAS  PubMed  Google Scholar 

  34. Hotamisligil, G. S. & Davis, R. J. Cell signaling and stress responses. Cold Spring Harb. Perspect. Biol. 8, a006072 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  35. Wood, C. D., Thornton, T. M., Sabio, G., Davis, R. A. & Rincon, M. Nuclear localization of p38 MAPK in response to DNA damage. Int. J. Biol. Sci. 5, 428–437 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. de Nadal, E., Ammerer, G. & Posas, F. Controlling gene expression in response to stress. Nat. Rev. Genet. 12, 833–845 (2011).

    Article  PubMed  CAS  Google Scholar 

  37. Young, P. R. et al. Pyridinyl imidazole inhibitors of p38 mitogen-activated protein kinase bind in the ATP site. J. Biol. Chem. 272, 12116–12121 (1997).

    Article  CAS  PubMed  Google Scholar 

  38. Tatham, M. H., Matic, I., Mann, M. & Hay, R. T. Comparative proteomic analysis identifies a role for SUMO in protein quality control. Sci. Signal. 4, rs4 (2011).

    Article  CAS  PubMed  Google Scholar 

  39. Hageman, J. et al. A DNAJB chaperone subfamily with HDAC-dependent activities suppresses toxic protein aggregation. Mol. Cell 37, 355–369 (2010).

    Article  CAS  PubMed  Google Scholar 

  40. Brehme, M. et al. A chaperome subnetwork safeguards proteostasis in aging and neurodegenerative disease. Cell Rep. 9, 1135–1150 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Hodges, A. et al. Regional and cellular gene expression changes in human Huntington’s disease brain. Hum. Mol. Genet. 15, 965–977 (2006).

    Article  CAS  PubMed  Google Scholar 

  42. Hadizadeh Esfahani, A., Sverchkova, A., Saez-Rodriguez, J., Schuppert, A. A. & Brehme, M. A systematic atlas of chaperome deregulation topologies across the human cancer landscape. PLoS Comput. Biol. 14, e1005890 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  43. Niskanen, E. A. et al. Global SUMOylation on active chromatin is an acute heat stress response restricting transcription. Genome Biol. 16, 1–19 (2015).

    Article  CAS  Google Scholar 

  44. Matunis, M. J., Zhang, X. D. & Ellis, N. A. SUMO: the glue that binds. Dev. Cell 11, 596–597 (2006).

    Article  CAS  PubMed  Google Scholar 

  45. Niskanen, E. A. & Palvimo, J. J. Chromatin SUMOylation in heat stress: to protect, pause and organise?: SUMO stress response on chromatin. Bioessays. https://doi.org/10.1002/bies.201600263 (2017).

    Article  PubMed  Google Scholar 

  46. Borisova, M. E. et al. p38-MK2 signaling axis regulates RNA metabolism after UV-light-induced DNA damage. Nat. Commun. 9, 1017 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  47. Gianfriddo, M., Melani, A., Turchi, D., Giovannini, M. G. & Pedata, F. Adenosine and glutamate extracellular concentrations and mitogen-activated protein kinases in the striatum of Huntington transgenic mice. Selective antagonism of adenosine A2A receptors reduces transmitter outflow. Neurobiol. Dis. 17, 77–88 (2004).

    Article  CAS  PubMed  Google Scholar 

  48. Taylor, D. M. et al. MAP kinase phosphatase 1 (MKP-1/DUSP1) is neuroprotective in Huntington’s disease via additive effects of JNK and p38 inhibition. J. Neurosci. 33, 2313–2325 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Shalgi, R. et al. Widespread regulation of translation by elongation pausing in heat shock. Mol. Cell 49, 439–452 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Pavitt, G. D. & Ron, D. New insights into translational regulation in the endoplasmic reticulum unfolded protein response. Cold Spring Harb. Perspect. Biol. 4, a012278 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  51. Mahat, D. B. & Lis, J. T. Use of conditioned media is critical for studies of regulation in response to rapid heat shock. Cell Stress Chaperones 22, 155–162 (2017).

    Article  PubMed  Google Scholar 

  52. Margueron, R. et al. Ezh1 and Ezh2 maintain repressive chromatin through different mechanisms. Mol. Cell 32, 503–518 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Henikoff, S., Henikoff, J. G., Sakai, A., Loeb, G. B. & Ahmad, K. Genome-wide profiling of salt fractions maps physical properties of chromatin. Genome Res. 19, 460–469 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Kustatscher, G., Wills, K. L., Furlan, C. & Rappsilber, J. Chromatin enrichment for proteomics. Nat. Protoc. 9, 2090–2099 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Tresini, M. et al. The core spliceosome as target and effector of non-canonical ATM signalling. Nature 523, 53–58 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Vinayagam, A. et al. Protein complex-based analysis framework for high-throughput data sets. Sci. Signal. 6, rs5 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  57. Arrigoni, L. et al. Standardizing chromatin research: a simple and universal method for ChIP-seq. Nucleic Acids Res. 44, e67 (2016).

    Article  PubMed  CAS  Google Scholar 

  58. Lee, T. I., Johnstone, S. E. & Young, R. A. Chromatin immunoprecipitation and microarray-based analysis of protein location. Nat. Protoc. 1, 729–748 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Bonhoure, N. et al. Quantifying ChIP-seq data: a spiking method providing an internal reference for sample-to-sample normalization. Genome Res. 24, 1157–1168 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Langmead, B. & Salzberg, S. L. Fast gapped-read alignment with Bowtie 2. Nat. Methods 9, 357–359 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Zhang, Y. et al. Model-based analysis of ChIP-Seq (MACS). Genome Biol. 9, R137 (2008).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  62. Quinlan, A. R. & Hall, I. M. BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics 26, 841–842 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Love, M. I., Huber, W. & Anders, S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 15, 550 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  64. Ramirez, F. et al. deepTools2: a next generation web server for deep-sequencing data analysis. Nucleic Acids Res. 44, W160–W165 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. H. Wickham. ggplot2: Elegant Graphics for Data Analysis. (Springer-Verlag New York, 2009).

  66. Warnes, G. W. et al. gplots: Various R programming tools for plotting data. https://cran.r-project.org/web/packages/gplots/index.html (2016).

  67. Ritchie, M. E. et al. limma powers differential expression analyses for RNA-sequencing and microarray studies. Nucleic Acids Res. 43, e47 (2015).

  68. Adler, D. vioplot: Violin plot. R package version 0.2. https://cran.r-project.org/web/packages/vioplot/index.html (2005).

  69. Mahat, D. B. et al. Base-pair-resolution genome-wide mapping of active RNA polymerases using precision nuclear run-on (PRO-seq). Nat. Protoc. 11, 1455–1476 (2016).

  70. Dieterich, D. C. et al. Labeling, detection and identification of newly synthesized proteomes with bioorthogonal non-canonical amino-acid tagging. Nat. Protoc. 2, 532–540 (2007).

  71. Efron, B. & Tibshirani, R. On testing the significance of sets of genes. Ann. Appl. Stat. 1, 23 (2007).

Download references

Acknowledgements

We thank Y.L. Dréan (University of Rennes), H. Kampinga (University of Groningen), R. Kopito (Stanford University), J.T. Lis (Cornell University), R. Paro (ETH), A. Shilatifard (Northwestern University), L. Sistonen (University of Turku), E. Trompouki (Max Planck Institute of Immunobiology and Epigenetics) and U. Wölfle (University of Freiburg) for sharing cell lines, protocols and plasmids; and J. Büscher, G. Mittler and sequencing and bioinformatics facilities for data acquisition and help with the analyses. We acknowledge critical discussions with P. Beli, U. Hartl, J. Palvimo, L. Sistonen and colleagues at the Max Planck Institute of Immunobiology and Epigenetics. Excellent technical assistance from S. Bares and A. Antonova is acknowledged. R.S. acknowledges financial support by the Max Planck Society, the German Research Foundation through the collaborative research center Medical Epigenetics, through the research grant SA 3190 (R.S.) and through Germany’s Excellence Strategy (CIBSS, EXC-2189, project ID 390939984).

Author information

Authors and Affiliations

Authors

Contributions

R.S. conceived the project; R.S., F.A.G. and P.T. designed the study; F.A.G., P.T. and A.K. performed experiments and interpreted the results; F.A.G. conceived and performed experiments on nascent-protein ubiquitination during the revision process; B.H. performed all the computational analyses; F.A.G. and R.S. wrote the manuscript with input from all the authors.

Corresponding author

Correspondence to Ritwick Sawarkar.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Integrated supplementary information

Supplementary Figure 1 Heat-shock-induced SITA in human cells quantified by RNA Pol II ChIP-seq and PRO-seq.

(a) Heatmap showing normalized RNA Pol II ChIP-seq signal in gene body of all significantly downregulated genes upon heat-shock in HEK293 cells. Each row represents a gene. Individual replicates from Fig. 1a are shown to demonstrate reproducibility. Gene subset with high basal expression levels and highest transcriptional downregulation upon HS, listed in Supplementary Data Set 1, are indicated by blue vertical bar adjacent to the heat-map (‘Top Downregulated genes’). These genes are used for all the genomic analyses in this study unless otherwise stated. The high basal expression of these genes allows a better statistical assessment of SITA, as detailed in Methods. The rest of the downregulated genes are indicated with a red bar, and are also listed in Supplementary Data Set 1. (b) Top left panel: metaplot of RNA Pol II binding in promoter vicinity around transcriptional start sites (TSS) in NHS and HS HEK293 cells. Bottom left panel (in gray): metaplot showing RNA Pol II occupancy in gene body regions (TSS + 1.5 kb to TES – 0.5kb). y-axes indicate normalized ChIP-seq occupancy of RNA Pol II in reads per million (rpm). Right panels: Genome browser tracks showing ChIP-seq occupancy of RNA Pol II in NHS and HS HEK293 cells. Two representative genes - RABGGTB and RPS20 - are shown. In gray: RNA Pol II occupancy in gene body regions shown at a scale different than promoter region. The vertical scale indicates normalized read density in reads per million (rpm). Histograms on the right: quantification of gene body RNA Pol II occupancy for the depicted genes across experimental replicates (n = 5 independent experiments, error bars depict standard error of the mean). (c) Heatmap showing normalized RNA Pol II ChIP-seq signal in gene body of all significantly downregulated genes upon heat-shock in K562 cells. Each row represents a gene. Individual replicates from Supplementary Fig. 1d are shown to demonstrate reproducibility. Gene subset with high basal expression levels and highest transcriptional downregulation upon HS, listed in Supplementary Data Set 2, are indicated by blue vertical bar adjacent to the heat-map (‘Top Downregulated genes’). The rest of the downregulated genes are indicated with a red bar, and are also listed in Supplementary Data Set 2. (d) MA plot of RNA Pol II ChIP-seq occupancy in gene body in K562 cells with or without heat-shock. Genes with significant (red dots) or nonsignificant changes (gray dots) in gene-body RNA Pol II occupancy (p-value <0.01 as calculated by DESeq2 analysis, n = 2 independent experiments) are indicated. Blue dots: gene subset with high basal expression levels and highest transcriptional downregulation upon HS, listed in Supplementary Data Set 2 and used for the genomic analyses in Fig. 1e. (e) Heatmap of normalized gene body counts from Precision Run-On sequencing (PRO-seq) signal in HEK293 cells with or without heat-shock. Each row represents a gene. The set of genes and their vertical order in the heatmap is the same as in Supplementary Fig. 1a. Individual replicates are shown to demonstrate reproducibility. (f) Effect of heat-shock on downregulation of transcription in HEK293 (right panels) and K562 cells (left panels) as assessed in individual replicates of RNA Pol II ChIP-seq. The y-axis represents the ratio of ChIP-seq signal in gene body in HS over NHS HEK293 cells. Violin plots show population distribution of the data. Upper panels: all downregulated genes; lower panels: most downregulated genes. (g) The empirical cumulative distribution function (ECDF) plot of the distribution of gene body ratios (HS over NHS) of RNA Pol II ChIP-seq values in K562 and HEK cells. Dotted lines: examples showing that 50% of HEK293 and K562 downregulated genes have more than 40% and 47% reduction in their expression levels respectively.

Supplementary Figure 2 Nascent-transcript quantification at individual loci by RT-qPCR as an assay for SITA.

(a) Primer validation for detecting nascent transcripts in HEK293 cells. 5,6-Dichlorobenzimidazole 1-β-D-ribofuranoside (DRB) is an inhibitor of Cdk9, and is used to rapidly block transcriptional elongation. DRB treatment reduces nascent transcripts, as shown recently (Palozola et al., 2017 (ref. 19)) and can be recapitulated in the shown quantification. Primer pairs were designed on exon-intron junction for indicated genes to capture the nascent unspliced transcript, similar to the recent study (Palozola et al., 2017 (ref. 19)). (b) Genome browser track view of HNRNPA2B1 gene showing no change in RNA Pol II occupancy during heat shock. This transcript is used as a normalizing control for all RT-qPCR quantification of nascent transcripts in this work, unless otherwise stated. Right: quantification of gene body RNA Pol II occupancy assessed by ChIP-seq (n = 5 independent experiments, error bars depict standard error of the mean). (c) Effect of HS on the nascent transcript levels of indicated genes in HEK293, K562 and SH-SY5Y cells. Value of 1.0 on y-axis indicates expression level of genes in NHS cells, shown by a dashed line. Error bars represent standard error of the mean (n = 3 independent experiments). (d) Effect of HS on the nascent transcript levels of indicated genes in T cell blasts derived from human primary PBMCs. Error bars represent standard error of the mean (n = 2 independent experiments).

Supplementary Figure 3 N-TEF enrichment at chromatin upon heat shock.

Western blot analyses of N-TEFs in total cell extracts and chromatin fractions of NHS and HS in different cell lines or by various methodologies as indicated. Histone H3 was used as loading control. (a) K562 and HCT116 cells. (b) HEK293 cells were either exposed to standard HS treatment as in the rest of this work (HS) or an acute HS treatment as previously reported, referred to as HS (acute) (Mahat et al., 2016 (ref. 7)). Detailed procedures can be found in the Methods. (c) Chromatin fractions from HEK293 cells were either obtained using a high salt-based protocol as in the rest of this work or using a low salt-based protocol, as in Henikoff, 2009 (ref. 53). (d and e) Chromatin extraction was performed using protocol described by Kustatscher et al., 2014 (ref. 54) (d), or with the protocol described in Tresini et al., 2015 (ref. 55) (e). (f) NELFA and NELFE enrichment at promoters of downregulated genes as detected by ChIP-seq. Histogram showing the ratio of HS to NHS occupancy at promoters (TSS +/- 500bp) for NELFA (left) and NELFE (right). Majority of genes have a higher peak upon HS (threshold of HS/NHS>1.3 showed by the vertical dotted line). (g) Metaplots of NELFA, NELFE and RNA Pol II binding around transcriptional start sites (TSS) of control genes in K562 cells with or without heat shock. y-axes indicate normalized ChIP-seq occupancy in reads per million (rpm). Lower right panel (in gray): metaplot showing RNA Pol II occupancy in gene body regions (TSS+1.5kb to TES-0.5kb). Control genes are defined as genes with no change in NELFE ChIP-seq signal upon heat shock in the region TSS +/-250bp. A list of these genes is available in Supplementary Data Set 6. (h) Effect of NELFE depletion on control genes as assessed by Pol II ChIP-seq occupancy in gene body. The y-axis represents the ratio of ChIP-seq signal in gene body in HS over NHS HEK293 cells treated with siRNA against NELFE (SiNELFE) or scrambled (SiScr). Violin plots show population distribution of the data. List of genes used for this analysis is available in Supplementary Data Set 7. (i) Effect of NELFE and PAF1 knockdown on nascent transcription as measured by RT-qPCR-based nascent transcript quantification in NHS and HS HEK293 cells. Value of 1.0 on y-axis indicates expression level of genes in NHS cells. Error bars represent standard error of the mean (n = 3 independent experiments).

Supplementary Figure 4 SITA of metabolic genes.

(a) Schematic diagram of the lower pathway of glycolysis from glyceraldehyde 3-phosphate to pyruvate and further to lactate formation. Enzymes that are downregulated upon HS are highlighted in red. GAPDH: glyceraldehyde-3-phosphate dehydrogenase; PGK: phosphoglycerate kinase; PGAM: phosphoglycerate mutase; ENO: enolase; PKM: pyruvate kinase; LDHA: L-lactate dehydrogenase. (b) Quantification of gene body RNA Pol II occupancy for genes encoding glycolytic enzymes in scramble and NELFE-depleted HEK293 cells with or without heat shock. Value of 1.0 on y-axis indicates expression level of genes in NHS cells. Error bars represent standard error of the mean (n = 2 independent experiments). (c) Genome browser tracks showing RNA Pol II ChIP-seq occupancy in gene-body regions of metabolic genes in NHS and HS HEK293 cells treated with siRNA against NELFE or scramble controls. The vertical scale indicates normalized read density in reads per million (rpm). (d) Mass spectrometry-based quantification of three metabolites shown in Supplementary Fig. 4a, which are reaction products of enzymes that are transcriptionally downregulated in heat-shock. Quantification was performed in scramble and NELFE-depleted HEK293 cells with or without heat-shock. Results are expressed as % decrease in each condition relative to NHS metabolite levels. Error bars represent standard error of the mean (n = 5 independent experiments).

Supplementary Figure 5 Heat-shock factor (HSF) is not required for stress-induced increase of NELFE at chromatin and consequent transcriptional downregulation.

(a) Nascent transcript quantification for indicated genes in HS and NHS HEK293 cells treated with siRNA against HSF1 and HSF2. Value of 1.0 on y-axis indicates expression level of genes in NHS cells, shown by a dashed line. Error bars represent standard error of the mean (n=2 independent experiments). (b) HSP70 transcription quantification to functionally validate HSF knockdown in the same samples as in Supplementary Fig. 5a. HS-induction of HSP70 transcription requires HSF. These data accompany Western blot analyses in Fig. 2a. (c) Western blot analysis of NELFE in chromatin fractions of HEK293 cells treated with specific inhibitors of HSF, namely KRIBB11 and Quercetin. Histone H3 was used as loading control. (d) RT-qPCR quantification of nascent transcripts of indicated genes in NHS and HS HEK293 cells treated with HSF inhibitors. Error bars represent standard error of the mean (n=3 independent experiments). (e) HSP70 transcription quantification to functionally validate HSF inhibition. (f) Western blot analysis of NELFE in total extracts and chromatin fractions of wild-type or HSF1and HSF2 double knockout mouse embryonic fibroblasts with or without heat shock. Histone H3 was used as loading control. (g) Representative blots of the small-molecule screen to identify regulators of stress-induced N-TEF recruitment to chromatin. Western blot analysis of NELFE in chromatin fractions of HEK293 cells treated with either translation inhibitor cycloheximide (left panel) or inhibitor of ubiquitin-activating enzyme E1 (right panel) NSC 624206 (Ub-E1 inhibitor). Histone H3 was used as loading control. (h) Western blot analysis of newly translated protein labeling. Cells were incubated for the indicated times with the amino acid analog AHA, an azide to label newly translated proteins. After click-chemistry reactions in total cell extracts to biotinylate azide-labeled new proteins, the biotinylated proteins were detected with a streptavidin-HRP conjugated reagent on western blot. Vimentin was used as a loading control. Cycloheximide incubation during AHA labeling completely abrogated its incorporation, confirming translation inhibition. Labeling strategy depicted in Fig. 3f. (i) Western blot analysis of HSP70 in total fractions of HEK293 cells treated with cycloheximide or vehicle. Ponceau staining was used as loading control. (j) RT-qPCR quantification of heat-shock genes HSP70, HSP60 and HSP90 in NHS and HS HEK293 cells treated with cycloheximide or vehicle. Error bars represent standard error of the mean (n = 2 independent experiments). (k) RT-qPCR-based nascent transcription quantification of indicated genes in NHS and HS cells treated with cycloheximide or vehicle. Top panel: HEK293 cells, lower panel: K562 cells. Error bars represent standard error of the mean (n = 3 independent experiments). (l) Effect of heat shock or cycloheximide on gene expression as assessed by RNA Pol II ChIP-seq occupancy in gene body. The y-axis represents the ratio of ChIP-seq signal in gene body in each condition over the NHS vehicle condition. Violin plots show population distribution of the data.

Supplementary Figure 6 Ubiquitination and translation during heat shock.

(a) Western blot analyses of K48-linked protein ubiquitination in several human cell lines and T cell blasts derived from human primary PBMCs with or without heat-shock. POLR2C was used as loading control. (b) Western blot analysis of K48-linked protein ubiquitination in total extracts of HEK293 cells with or without heat shock treated with vehicle or translation inhibitors cycloheximide, puromycin or anisomycin. POLR2C was used as loading control. (c) Effect of heat shock or Ub-E1 inhibitor on gene expression as assessed by RNA Pol II ChIP-seq occupancy in gene body. The y-axis represents the ratio of ChIP-seq signal in gene body in each condition over the NHS vehicle condition. Violin plots show population distribution of the data. (d) Western blot analysis of total free ubiquitin levels in HEK293 cells. Cells were treated with cycloheximide or vehicle and subjected to heat-shock. Free ubiquitin is shown. POLR2C was used as loading control. (e) Western blot analysis of total free ubiquitin levels in HEK293 cells. Cells were treated with Ub-E1 inhibitor or vehicle and subjected to heat-shock. Free ubiquitin is shown. POLR2C was used as loading control. (f) Western blot analysis of newly translated proteins from HEK293 cells treated with Ub-E1 inhibitor or vehicle and subjected to heat-shock. The labeling of nascent proteins is depicted in Fig. 3f. Biotinylated proteins were detected with a Streptavidin-HRP conjugated reagent. Vimentin was used as a loading control. Heat shock causes a decrease in total translation amount.

Supplementary Figure 7 SITA in an acute model of heat shock.

(a and c) Western blot analyses of K48-linked protein ubiquitination in extracts from HeLa cells exposed to acute HS at 42 °C for the indicated times. Vimentin was used as loading control. (b, d and e) Nascent transcript quantification of the indicated genes. Value of 1.0 on y-axis indicates expression level of NHS cells. Error bars represent standard error of the mean (at least n = 3 independent experiments). (f) Western blot analysis of NELFE in total extracts and chromatin fractions of HeLa cells subjected to acute heat-shock at 42 °C for the indicated times. Histone H3 was used as loading control.

Supplementary Figure 8 p38α kinase (MAPK14)-mediated signaling is required for SITA.

(a) Western blot analyses of MAPK14 in total extracts and chromatin fractions of HEK293 cells with or without heat shock treated with p38 inhibitor. Histone H3 was used as loading control. (b) Top: Schematic representation of how the Pausing Index was calculated. Bottom: The empirical cumulative distribution function (ECDF) plot of the pausing index of RNA Pol II ChIP-seq values in HEK293 cells treated with p38 inhibitor or vehicle subjected or not to heat-shock. (c) Genome browser tracks showing RNA Pol II ChIP-seq occupancy in a representative gene RASL10B in NHS and HS HEK293 cells treated with p38 inhibitor or vehicle. The vertical scale indicates normalized read density in reads per million.

Supplementary Figure 9 Proteasome inhibition does not lead to SITA.

(a) Western blot analyses of total free ubiquitin levels. HEK293 cells were pretreated with Ub-E1 inhibitor or vehicle and incubated for 2 h with L-arginine or its analog L-canavanine. Free ubiquitin is shown. POLR2C was used as loading control. (b) Western blot analysis of newly translated proteins from HEK293 cells treated with Ub-E1 inhibitor or vehicle and incubated for 2 h with L-arginine or its analog L-canavanine. The labeling of nascent proteins is depicted in Fig. 3f. Biotinylated proteins were detected with a Streptavidin-HRP conjugated reagent. Vimentin was used as a loading control. (c) Western blot analyses of NELFE in total extracts and chromatin fractions of HEK293 cells with or without proteasome inhibitor Bortezomib (Bort). Histone H3 was used as loading control. (d) ChIP-qPCR measuring NELFE occupancy at indicated gene promoters in HEK293 cells incubated with Bortezomib or vehicle. ‘NO PEAK’ primer set amplifies genomic region not expected to bind NELFE and acts as a negative control. The y-axis indicates the amount of immunoprecipitated DNA relative to starting input material. Error bars represent standard error of the mean (n = 2 independent experiments). (e) Effect of Bortezomib on the nascent transcript levels of indicated genes in HEK293 cells. Value of 1.0 on y-axis indicates expression level of genes in vehicle-incubated cells, shown by a dashed line. Error bars represent standard error of the mean (n = 2 independent experiments).

Supplementary Figure 10 SITA in cells with Huntingtin misfolding.

(a and b) Effect of transient overexpression of HTT in neuronal SH-SY5Y cells (a) and HEK293 cells (b). Top: Fluorescence microscopy images of cells overexpressing GFP-tagged variants of either HTT25Q or HTT103Q. White arrow indicates HTT aggregate. Bottom: RT-qPCR-based nascent transcript quantification in cells expressing mutant HTT103Q. Value 1.0 on y-axis represents normalized expression levels of genes in cells expressing wild-type HTT25Q. Error bars represent standard error of the mean (n = 2 independent experiments for a and 3 replicates for b). (c) HTT 103Q or HTT 25Q transient overexpression in neuronal SH-SY5Y cell line. Left: Western blotting with antibody against K48-linked protein ubiquitination in total extracts and NELFE in chromatin fractions. POLR2C was used as loading control. Right: Corresponding densitometry immunoblot quantification. Enrichment of NELFE at chromatin in HTT 103Q-expressing cells relative to HTT 25Q-expressing cells. Error bars represent standard error of the mean (n = 2 independent experiments). (d) Western blot analysis of K48-linked protein ubiquitination in proteins translated during HTT expression or total proteins of HEK293 cells induced for 48 h to express either soluble HTT25Q or aggregating HTT119Q and labeled with AHA for the last 1h before lysis. Vimentin and biotinylated proteins were used as a loading control. (e) Densitometry immunoblot quantification of experiment depicted in Fig 5e. Enrichment of NELFE at chromatin after induction of cells harboring inducible variants of HTT relative to uninduced cells. Error bars represent standard error of the mean (n = 3 independent experiments). (f and g) Normalized expression levels in brain samples of healthy controls in comparison with Huntington’s disease subjects for genes encoding metabolic enzymes indicated in Supplementary Fig. 4a (f) and genes encoding chaperones (g). Statistical significance determined by Student’s t-test against Healthy controls is indicated by asterisks. *: p < 0.5; **: p < 0.05 and ***: p < 0.005.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–10 and Supplementary Tables 1–4

Reporting Summary

Supplementary Dataset 1

Differentially expressed genes assessed from HEK293 RNA Pol II ChIP-seq.

Supplementary Dataset 2

Differentially expressed genes assessed from K562 RNA Pol II ChIP-seq.

Supplementary Dataset 3

Comparison of differentially expressed genes between HEK293 and K562.

Supplementary Dataset 4

HS-induced changes in the chromatin proteome.

Supplementary Dataset 5

Blots in main figures shown as uncropped images.

Supplementary Dataset 6

Control genes used for analyzing NELF ChIP-seq and Pol II ChIP-seq in Supplementary Fig. 3g.

Supplementary Dataset 7

Control genes used for analyzing RNA Pol II ChIP-seq in Supplementary Fig. 3h.

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Aprile-Garcia, F., Tomar, P., Hummel, B. et al. Nascent-protein ubiquitination is required for heat shock–induced gene downregulation in human cells. Nat Struct Mol Biol 26, 137–146 (2019). https://doi.org/10.1038/s41594-018-0182-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41594-018-0182-x

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