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:

PRC1 drives Polycomb-mediated gene repression by controlling transcription initiation and burst frequency

Abstract

The Polycomb repressive system plays a fundamental role in controlling gene expression during mammalian development. To achieve this, Polycomb repressive complexes 1 and 2 (PRC1 and PRC2) bind target genes and use histone modification-dependent feedback mechanisms to form Polycomb chromatin domains and repress transcription. The inter-relatedness of PRC1 and PRC2 activity at these sites has made it difficult to discover the specific components of Polycomb chromatin domains that drive gene repression and to understand mechanistically how this is achieved. Here, by exploiting rapid degron-based approaches and time-resolved genomics, we kinetically dissect Polycomb-mediated repression and discover that PRC1 functions independently of PRC2 to counteract RNA polymerase II binding and transcription initiation. Using single-cell gene expression analysis, we reveal that PRC1 acts uniformly within the cell population and that repression is achieved by controlling transcriptional burst frequency. These important new discoveries provide a mechanistic and conceptual framework for Polycomb-dependent transcriptional control.

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

Access options

Buy this article

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

Fig. 1: Acute depletion of PRC1 reveals a rapid turnover of H2AK119ub1.
Fig. 2: Perturbing PRC1 has immediate effects on gene expression.
Fig. 3: PRC1-mediated gene repression does not rely directly on PRC2 or H3K27me3.
Fig. 4: Removal of PRC1 causes rapid new binding of Pol II and accumulation of H3K4me3.
Fig. 5: New initiation is required for Polycomb target gene derepression.
Fig. 6: Single-cell analysis of Polycomb target gene expression.
Fig. 7: The Polycomb system represses gene expression by limiting transcription burst frequency.

Similar content being viewed by others

Data availability

High-throughput sequencing datasets generated for this study are available in the GEO database under the accession number GSE159400. Published data used in this study include BioCAP-seq data (GEO, GSE43512) from ref. 118, cnRNA-seq of Ring1a−/−;Ring1bfl/fl (PRC1CKO) (GEO, GSE119619) samples from ref. 29 and cChIP–seq data for JARID2, AEBP2, PCL2 and EPOP in untreated (wild-type) Ring1aI50A/D53K;Ring1b(WT->I53A/D56K)fl/fl ES cells (GEO, GSE132754) from ref. 30. For cnRNA-seq processing, we used mm10 (GenBank, BK000964.3) and dm6 (GenBank, M21017.1) rDNA genomic datasets. Source data are provided with this paper.

Code availability

All R and Perl scripts used for genomic data analysis in this study are available at https://github.com/pauladobrinic/PRC1degron-2021. A custom-made ImageJ script for preprocessing 3D images (ThunderFISH) is publicly available with a detailed manual for sample preparation and script use at https://github.com/aleks-szczure/ThunderFISH.

References

  1. Kouzarides, T. Chromatin modifications and their function. Cell 128, 693–705 (2007).

    Article  CAS  PubMed  Google Scholar 

  2. Bannister, A. J. & Kouzarides, T. Regulation of chromatin by histone modifications. Cell Res. 21, 381–395 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Blackledge, N. P., Rose, N. R. & Klose, R. J. Targeting Polycomb systems to regulate gene expression: modifications to a complex story. Nat. Rev. Mol. Cell Biol. 16, 643–649 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Schuettengruber, B., Bourbon, H.-M., Di Croce, L. & Cavalli, G. Genome regulation by Polycomb and Trithorax: 70 years and counting. Cell 171, 34–57 (2017).

    Article  CAS  PubMed  Google Scholar 

  5. Voncken, J. W. et al. Rnf2 (Ring1b) deficiency causes gastrulation arrest and cell cycle inhibition. Proc. Natl Acad. Sci. USA 100, 2468–2473 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Faust, C., Schumacher, A., Holdener, B. & Magnuson, T. The Eed mutation disrupts anterior mesoderm production in mice. Development 121, 273–285 (1995).

    Article  CAS  PubMed  Google Scholar 

  7. Pasini, D., Bracken, A. P., Jensen, M. R., Lazzerini Denchi, E. & Helin, K. Suz12 is essential for mouse development and for EZH2 histone methyltransferase activity. EMBO J. 23, 4061–4071 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. O’Carroll, D. et al. The Polycomb-group gene Ezh2 is required for early mouse development. Mol. Cell. Biol. 21, 4330–4336 (2001).

    Article  PubMed  PubMed Central  Google Scholar 

  9. Wang, H. et al. Role of histone H2A ubiquitination in Polycomb silencing. Nature 431, 873–878 (2004).

    Article  CAS  PubMed  Google Scholar 

  10. Li, Z. et al. Structure of a Bmi-1–Ring1B Polycomb group ubiquitin ligase complex. J. Biol. Chem. 281, 20643–20649 (2006).

    Article  CAS  PubMed  Google Scholar 

  11. Buchwald, G. et al. Structure and E3-ligase activity of the Ring–Ring complex of Polycomb proteins Bmi1 and Ring1b. EMBO J. 25, 2465–2474 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Cao, R. et al. Role of histone H3 lysine 27 methylation in Polycomb-group silencing. Science 298, 1039–1043 (2002).

    Article  CAS  PubMed  Google Scholar 

  13. Kuzmichev, A., Nishioka, K., Erdjument-Bromage, H., Tempst, P. & Reinberg, D. Histone methyltransferase activity associated with a human multiprotein complex containing the enhancer of Zeste protein. Genes Dev. 16, 2893–2905 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Müller, J. et al. Histone methyltransferase activity of a Drosophila Polycomb group repressor complex. Cell 111, 197–208 (2002).

    Article  PubMed  Google Scholar 

  15. Czermin, B. et al. Drosophila enhancer of Zeste/ESC complexes have a histone H3 methyltransferase activity that marks chromosomal Polycomb sites. Cell 111, 185–196 (2002).

    Article  CAS  PubMed  Google Scholar 

  16. Farcas, A. M. et al. KDM2B links the Polycomb repressive complex 1 (PRC1) to recognition of CpG islands. eLife 1, e00205 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  17. He, J. et al. Kdm2b maintains murine embryonic stem cell status by recruiting PRC1 complex to CpG islands of developmental genes. Nat. Cell Biol. 15, 373–384 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Li, H. et al. Polycomb-like proteins link the PRC2 complex to CpG islands. Nature 549, 287–291 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Perino, M. et al. MTF2 recruits Polycomb repressive complex 2 by helical-shape-selective DNA binding. Nat. Genet. 50, 1002–1010 (2018).

    Article  CAS  PubMed  Google Scholar 

  20. Wu, X., Johansen, J. V. & Helin, K. Fbxl10/Kdm2b recruits Polycomb repressive complex 1 to CpG islands and regulates H2A ubiquitylation. Mol. Cell 49, 1134–1146 (2013).

    Article  CAS  PubMed  Google Scholar 

  21. Mikkelsen, T. S. et al. Genome-wide maps of chromatin state in pluripotent and lineage-committed cells. Nature 448, 553–560 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Ku, M. et al. Genomewide analysis of PRC1 and PRC2 occupancy identifies two classes of bivalent domains. PLoS Genet. 4, e1000242 (2008).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  23. Blackledge, N. P. et al. Variant PRC1 complex-dependent H2A ubiquitylation drives PRC2 recruitment and Polycomb domain formation. Cell 157, 1445–1459 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Cooper, S. et al. Targeting Polycomb to pericentric heterochromatin in embryonic stem cells reveals a role for H2AK119u1 in PRC2 recruitment. Cell Rep. 7, 1456–1470 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Kalb, R. et al. Histone H2A monoubiquitination promotes histone H3 methylation in Polycomb repression. Nat. Struct. Mol. Biol. 21, 569–571 (2014).

    Article  CAS  PubMed  Google Scholar 

  26. Wang, L. et al. Hierarchical recruitment of Polycomb group silencing complexes. Mol. Cell 14, 637–646 (2004).

    Article  CAS  PubMed  Google Scholar 

  27. Min, J., Zhang, Y. & Xu, R.-M. Structural basis for specific binding of Polycomb chromodomain to histone H3 methylated at Lys 27. Genes Dev. 17, 1823–1828 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Klose, R. J., Cooper, S., Farcas, A. M., Blackledge, N. P. & Brockdorff, N. Chromatin sampling—an emerging perspective on targeting Polycomb repressor proteins. PLoS Genet. 9, e1003717 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Fursova, N. A. et al. Synergy between variant PRC1 complexes defines Polycomb-mediated gene repression. Mol. Cell 74, 1020–1036 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Blackledge, N. P. et al. PRC1 catalytic activity is central to Polycomb system function. Mol. Cell 77, 857–874 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Endoh, M. et al. Polycomb group proteins Ring1A/B are functionally linked to the core transcriptional regulatory circuitry to maintain ES cell identity. Development 135, 1513–1524 (2008).

    Article  CAS  PubMed  Google Scholar 

  32. Natsume, T., Kiyomitsu, T., Saga, Y. & Kanemaki, M. T. Rapid protein depletion in human cells by auxin-inducible degron tagging with short homology donors. Cell Rep. 15, 210–218 (2016).

    Article  CAS  PubMed  Google Scholar 

  33. Nishimura, K., Fukagawa, T., Takisawa, H., Kakimoto, T. & Kanemaki, M. An auxin-based degron system for the rapid depletion of proteins in nonplant cells. Nat. Methods 6, 917–922 (2009).

    Article  CAS  PubMed  Google Scholar 

  34. Rhodes, J. D. P. et al. Cohesin disrupts Polycomb-dependent chromosome interactions in embryonic stem cells. Cell Rep. 30, 820–835 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Fujimura, Y. et al. Distinct roles of Polycomb group gene products in transcriptionally repressed and active domains of Hoxb8. Development 133, 2371–2381 (2006).

    Article  CAS  PubMed  Google Scholar 

  36. Leeb, M. & Wutz, A. Ring1B is crucial for the regulation of developmental control genes and PRC1 proteins but not X inactivation in embryonic cells. J. Cell Biol. 178, 219–229 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Cooper, S. et al. Jarid2 binds mono-ubiquitylated H2A lysine 119 to mediate crosstalk between Polycomb complexes PRC1 and PRC2. Nat. Commun. 7, 13661 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Beltran, M. et al. G-tract RNA removes Polycomb repressive complex 2 from genes. Nat. Struct. Mol. Biol. 26, 899–909 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Davidovich, C., Zheng, L., Goodrich, K. J. & Cech, T. R. Promiscuous RNA binding by Polycomb repressive complex 2. Nat. Struct. Mol. Biol. 20, 1250–1257 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Wang, X. et al. Molecular analysis of PRC2 recruitment to DNA in chromatin and its inhibition by RNA. Nat. Struct. Mol. Biol. 24, 1028–1038 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Margueron, R. et al. Role of the Polycomb protein EED in the propagation of repressive histone marks. Nature 461, 762–767 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Xu, C. et al. Binding of different histone marks differentially regulates the activity and specificity of Polycomb repressive complex 2 (PRC2). Proc. Natl Acad. Sci. USA 107, 19266–19271 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Hansen, K. H. et al. A model for transmission of the H3K27me3 epigenetic mark. Nat. Cell Biol. 10, 1291–1300 (2008).

    Article  CAS  PubMed  Google Scholar 

  44. Laprell, F., Finkl, K. & Müller, J. Propagation of Polycomb-repressed chromatin requires sequence-specific recruitment to DNA. Science 356, 85–88 (2017).

    Article  CAS  PubMed  Google Scholar 

  45. Nabet, B. et al. The dTAG system for immediate and target-specific protein degradation. Nat. Chem. Biol. 14, 431–441 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Huseyin, M. K. & Klose, R. J. Live-cell single particle tracking of PRC1 reveals a highly dynamic system with low target site occupancy. Nat. Commun. 12, 887 (2021).

  47. Lehmann, L. et al. Polycomb repressive complex 1 (PRC1) disassembles RNA polymerase II preinitiation complexes. J. Biol. Chem. 287, 35784–35794 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Nakagawa, T. et al. Deubiquitylation of histone H2A activates transcriptional initiation via trans-histone cross-talk with H3K4 di- and trimethylation. Genes Dev. 22, 37–49 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Stock, J. K. et al. Ring1-mediated ubiquitination of H2A restrains poised RNA polymerase II at bivalent genes in mouse ES cells. Nat. Cell Biol. 9, 1428–1435 (2007).

    Article  CAS  PubMed  Google Scholar 

  50. Zhou, W. et al. Histone H2A monoubiquitination represses transcription by inhibiting RNA polymerase II transcriptional elongation. Mol. Cell 29, 69–80 (2008).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  51. Chopra, V. S. et al. The Polycomb group mutant esc leads to augmented levels of paused Pol II in the Drosophila embryo. Mol. Cell 42, 837–844 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Dellino, G. I. et al. Polycomb silencing blocks transcription initiation. Mol. Cell 13, 887–893 (2004).

    Article  CAS  PubMed  Google Scholar 

  53. King, I. F. G., Francis, N. J. & Kingston, R. E. Native and recombinant Polycomb group complexes establish a selective block to template accessibility to repress transcription in vitro. Mol. Cell. Biol. 22, 7919–7928 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Hughes, A. L., Kelley, J. R. & Klose, R. J. Understanding the interplay between CpG island-associated gene promoters and H3K4 methylation. Biochim. Biophys. Acta Gene Regul. Mech. 1863, 194567 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Harlen, K. M. & Churchman, L. S. The code and beyond: transcription regulation by the RNA polymerase II carboxy-terminal domain. Nat. Rev. Mol. Cell Biol. 18, 263–273 (2017).

    Article  CAS  PubMed  Google Scholar 

  56. Brookes, E. et al. Polycomb associates genome-wide with a specific RNA polymerase II variant, and regulates metabolic genes in ESCs. Cell Stem Cell 10, 157–170 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Min, I. M. et al. Regulating RNA polymerase pausing and transcription elongation in embryonic stem cells. Genes Dev. 25, 742–754 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Williams, L. H. et al. Pausing of RNA polymerase II regulates mammalian developmental potential through control of signaling networks. Mol. Cell 58, 311–322 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Kar, G. et al. Flipping between Polycomb repressed and active transcriptional states introduces noise in gene expression. Nat. Commun. 8, 36 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  60. Torre, E. et al. Rare cell detection by single-cell RNA sequencing as guided by single-molecule RNA FISH. Cell Syst. 6, 171–179 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Raj, A., Peskin, C. S., Tranchina, D., Vargas, D. Y. & Tyagi, S. Stochastic mRNA synthesis in mammalian cells. PLoS Biol. 4, e309 (2006).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  62. Rodriguez, J. & Larson, D. R. Transcription in living cells: molecular mechanisms of bursting. Annu. Rev. Biochem. 89, 189–212 (2020).

    Article  CAS  PubMed  Google Scholar 

  63. Peccoud, J. & Ycart, B. Markovian modeling of gene-product synthesis. Theor. Popul. Biol. 48, 222–234 (1995).

    Article  Google Scholar 

  64. Zoller, B., Little, S. C. & Gregor, T. Diverse spatial expression patterns emerge from unified kinetics of transcriptional bursting. Cell 175, 835–847 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Fritzsch, C. et al. Estrogen-dependent control and cell-to-cell variability of transcriptional bursting. Mol. Syst. Biol. 14, e7678 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  66. Levesque, M. J. & Raj, A. Single-chromosome transcriptional profiling reveals chromosomal gene expression regulation. Nat. Methods 10, 246–248 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Tamburri, S. et al. Histone H2AK119 mono-ubiquitination is essential for Polycomb-mediated transcriptional repression. Mol. Cell 77, 840–856 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Gao, Z. et al. PCGF homologs, CBX proteins, and RYBP define functionally distinct PRC1 family complexes. Mol. Cell 45, 344–356 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Taherbhoy, A. M., Huang, O. W. & Cochran, A. G. BMI1–RING1B is an autoinhibited RING E3 ubiquitin ligase. Nat. Commun. 6, 7621 (2015).

    Article  PubMed  Google Scholar 

  70. Rose, N. R. et al. RYBP stimulates PRC1 to shape chromatin-based communication between Polycomb repressive complexes. eLife 5, 18591 (2016).

    Article  CAS  Google Scholar 

  71. Morey, L. et al. Polycomb regulates mesoderm cell fate-specification in embryonic stem cells through activation and repression mechanisms. Cell Stem Cell 17, 300–315 (2015).

    Article  CAS  PubMed  Google Scholar 

  72. Morey, L. et al. Nonoverlapping functions of the Polycomb group Cbx family of proteins in embryonic stem cells. Cell Stem Cell 10, 47–62 (2012).

    Article  CAS  PubMed  Google Scholar 

  73. Kundu, S. et al. Polycomb repressive complex 1 generates discrete compacted domains that change during differentiation. Mol. Cell 65, 432–446 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Zepeda-Martinez, J. A. et al. Parallel PRC2/cPRC1 and vPRC1 pathways silence lineage-specific genes and maintain self-renewal in mouse embryonic stem cells. Sci. Adv. 6, eaax5692 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Cohen, I., Bar, C. & Ezhkova, E. Activity of PRC1 and histone H2AK119 monoubiquitination: revising popular misconceptions. BioEssays 42, e1900192 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  76. Aihara, H. et al. Histone H2A T120 phosphorylation promotes oncogenic transformation via upregulation of cyclin D1. Mol. Cell 64, 176–188 (2016).

    Article  CAS  PubMed  Google Scholar 

  77. Tsuboi, M. et al. Ubiquitination-independent repression of PRC1 targets during neuronal fate restriction in the developing mouse neocortex. Dev. Cell 47, 758–772 (2018).

    Article  CAS  PubMed  Google Scholar 

  78. Højfeldt, J. W. et al. Accurate H3K27 methylation can be established de novo by SUZ12-directed PRC2. Nat. Struct. Mol. Biol. 25, 225–232 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  79. Alabert, C. et al. Two distinct modes for propagation of histone PTMs across the cell cycle. Genes Dev. 29, 585–590 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Oksuz, O. et al. Capturing the onset of PRC2-mediated repressive domain formation. Mol. Cell 70, 1149–1162 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Reverón-Gómez, N. et al. Accurate recycling of parental histones reproduces the histone modification landscape during DNA replication. Mol. Cell 72, 239–249 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  82. Alabert, C. et al. Domain model explains propagation dynamics and stability of histone H3K27 and H3K36 methylation landscapes. Cell Rep. 30, 1223–1234 (2020).

    Article  CAS  PubMed  Google Scholar 

  83. Miller, S. A., Damle, M., Kim, J. & Kingston, R. E. Full methylation of H3K27 by PRC2 is dispensable for initial embryoid body formation but required to maintain differentiated cell identity. Development 148, dev196329 (2021).

  84. Francis, N. J., Saurin, A. J., Shao, Z. & Kingston, R. E. Reconstitution of a functional core Polycomb repressive complex. Mol. Cell 8, 545–556 (2001).

    Article  CAS  PubMed  Google Scholar 

  85. Francis, N. J., Kingston, R. E. & Woodcock, C. L. Chromatin compaction by a Polycomb group protein complex. Science 306, 1574–1577 (2004).

    Article  CAS  PubMed  Google Scholar 

  86. Shao, Z. et al. Stabilization of chromatin structure by PRC1, a Polycomb complex. Cell 98, 37–46 (1999).

    Article  CAS  PubMed  Google Scholar 

  87. Isono, K. et al. SAM domain polymerization links subnuclear clustering of PRC1 to gene silencing. Dev. Cell 26, 565–577 (2013).

    Article  CAS  PubMed  Google Scholar 

  88. Lavigne, M., Francis, N. J., King, I. F. G. & Kingston, R. E. Propagation of silencing. Mol. Cell 13, 415–425 (2004).

    Article  CAS  PubMed  Google Scholar 

  89. King, H. W., Fursova, N. A., Blackledge, N. P. & Klose, R. J. Polycomb repressive complex 1 shapes the nucleosome landscape but not accessibility at target genes. Genome Res. 28, 1494–1507 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Lavarone, E., Barbieri, C. M. & Pasini, D. Dissecting the role of H3K27 acetylation and methylation in PRC2 mediated control of cellular identity. Nat. Commun. 10, 1679 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  91. Hodges, H. C. et al. Dominant-negative SMARCA4 mutants alter the accessibility landscape of tissue-unrestricted enhancers. Nat. Struct. Mol. Biol. 25, 61–72 (2018).

    Article  CAS  PubMed  Google Scholar 

  92. Bartman, C. R. et al. Transcriptional burst initiation and polymerase pause release are key control points of transcriptional regulation. Mol. Cell 73, 519–532 (2019).

    Article  CAS  PubMed  Google Scholar 

  93. Ochiai, H. et al. Genome-wide kinetic properties of transcriptional bursting in mouse embryonic stem cells. Sci. Adv. 6, eaaz6699 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Larsson, A. J. M. et al. Genomic encoding of transcriptional burst kinetics. Nature 565, 251–254 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Bartman, C. R., Hsu, S. C., Hsiung, C. C.-S., Raj, A. & Blobel, G. A. Enhancer regulation of transcriptional bursting parameters revealed by forced chromatin looping. Mol. Cell 62, 237–247 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Mermet, J. et al. Clock-dependent chromatin topology modulates circadian transcription and behavior. Genes Dev. 32, 347–358 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Fukaya, T., Lim, B. & Levine, M. Enhancer control of transcriptional bursting. Cell 166, 358–368 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Chen, L.-F. et al. Enhancer histone acetylation modulates transcriptional bursting dynamics of neuronal activity-inducible genes. Cell Rep. 26, 1174–1188 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Larson, D. R. et al. Direct observation of frequency modulated transcription in single cells using light activation. eLife 2, e00750 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  100. Senecal, A. et al. Transcription factors modulate c-Fos transcriptional bursts. Cell Rep. 8, 75–83 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Nicolas, D., Zoller, B., Suter, D. M. & Naef, F. Modulation of transcriptional burst frequency by histone acetylation. Proc. Natl Acad. Sci. USA 115, 7153–7158 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  102. Dimitrova, E. et al. FBXL19 recruits CDK-mediator to CpG islands of developmental genes priming them for activation during lineage commitment. eLife 7, e37084 (2018).

  103. Atsuta, T. et al. Production of monoclonal antibodies against mammalian Ring1B proteins. Hybridoma 20, 43–46 (2001).

    Article  CAS  PubMed  Google Scholar 

  104. 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 

  105. Hu, B. et al. Biological chromodynamics: a general method for measuring protein occupancy across the genome by calibrating ChIP–seq. Nucleic Acids Res. 43, e132 (2015).

    PubMed  PubMed Central  Google Scholar 

  106. Orlando, D. A. et al. Quantitative ChIP–seq normalization reveals global modulation of the epigenome. Cell Rep. 9, 1163–1170 (2014).

    Article  CAS  PubMed  Google Scholar 

  107. Turberfield, A. H. et al. KDM2 proteins constrain transcription from CpG island gene promoters independently of their histone demethylase activity. Nucleic Acids Res. 47, 9005–9023 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. 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 

  109. Dobin, A. et al. STAR: ultrafast universal RNA-seq aligner. Bioinformatics 29, 15–21 (2013).

    Article  CAS  PubMed  Google Scholar 

  110. Tarasov, A., Vilella, A. J., Cuppen, E., Nijman, I. J. & Prins, P. Sambamba: fast processing of NGS alignment formats. Bioinformatics 31, 2032–2034 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  PubMed Central  CAS  Google Scholar 

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

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  113. Kent, W. J. et al. The human genome browser at UCSC. Genome Res. 12, 996–1006 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. 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 

  115. Zhu, A., Ibrahim, J. G. & Love, M. I. Heavy-tailed prior distributions for sequence count data: removing the noise and preserving large differences. Bioinformatics 35, 2084–2092 (2019).

    Article  CAS  PubMed  Google Scholar 

  116. Taruttis, F. et al. External calibration with Drosophila whole-cell spike-ins delivers absolute mRNA fold changes from human RNA-seq and qPCR data. BioTechniques 62, 53–61 (2017).

    Article  CAS  PubMed  Google Scholar 

  117. Gu, Z., Eils, R. & Schlesner, M. Complex heatmaps reveal patterns and correlations in multidimensional genomic data. Bioinformatics 32, 2847–2849 (2016).

    Article  CAS  PubMed  Google Scholar 

  118. Long, H. K. et al. Epigenetic conservation at gene regulatory elements revealed by non-methylated DNA profiling in seven vertebrates. eLife 2, e00348 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  119. Ovesný, M., Křížek, P., Borkovec, J., Svindrych, Z. & Hagen, G. M. ThunderSTORM: a comprehensive ImageJ plug-in for PALM and STORM data analysis and super-resolution imaging. Bioinformatics 30, 2389–2390 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  120. Bolte, S. & Cordelières, F. P. A guided tour into subcellular colocalization analysis in light microscopy. J. Microsc. 224, 213–232 (2006).

    Article  CAS  PubMed  Google Scholar 

  121. Singer, Z. S. et al. Dynamic heterogeneity and DNA methylation in embryonic stem cells. Mol. Cell 55, 319–331 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Singh, A., Razooky, B., Cox, C. D., Simpson, M. L. & Weinberger, L. S. Transcriptional bursting from the HIV-1 promoter is a significant source of stochastic noise in HIV-1 gene expression. Biophys. J. 98, L32–L34 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Foreman, R. & Wollman, R. Mammalian gene expression variability is explained by underlying cell state. Mol. Syst. Biol. 16, e9146 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Padovan-Merhar, O. et al. Single mammalian cells compensate for differences in cellular volume and DNA copy number through independent global transcriptional mechanisms. Mol. Cell 58, 339–352 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Mueller, F. et al. FISH-quant: automatic counting of transcripts in 3D FISH images. Nat. Methods 10, 277–278 (2013).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

Work in the Klose laboratory is supported by the Wellcome Trust (209400/Z/17/Z) (R.J.K.), the European Research Council (681440) (R.J.K.) and the Lister Institute of Preventive Medicine (R.J.K.). We are grateful to A. Williams at the Department of Zoology, University of Oxford, for sequencing support on the NextSeq 500. We express our gratitude to D. Ennis and I. Davis for their advice on RNA-FISH probe design. We thank N. Fursova for her valuable advice on computational analysis. We are grateful to N. Fursova, N. Blackledge and E. Dimitrova for critical reading of the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

Conceptualization, P.D., A.T.S. and R.J.K.; methodology, investigation and formal analysis, P.D. and A.T.S.; writing (original draft), P.D., A.T.S. and R.J.K.; writing (review and editing), P.D., A.T.S. and R.J.K.; funding acquisition and supervision, R.J.K.

Corresponding author

Correspondence to Robert J. Klose.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature Structural and Molecular Biology thanks the anonymous reviewers for their contribution to the peer review of this work. Editor recognition statement Anke Sparmann and Carolina Perdigoto were the primary editors on this article and managed its editorial process and peer review in collaboration with the rest of the editorial team.

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

Extended data

Extended Data Fig. 1 Detailed characterisation of the PRC1deg.

a, Schematic of the PRC1deg system. Both endogenous Ring1B alleles are N-terminally fused to the full-length AID tag in a Ring1A-/- background, while OsTIR1 is inserted in Rosa26 locus. b, Western blot analysis of RING1A and RING1B in the control cell line (TIR1) and PRC1deg cells (left panel) with a quantification of RING1B levels (right panel). Shown are values and mean from n = 2 independent experiments. c, Genomic snapshots of typical Polycomb target genes showing cChIP-seq signal for RING1B in wild-type (TIR1) and PRC1deg cells before auxin treatment. d, Metaplot analysis of RING1B cChIP-seq at RING1B-bound sites in wild-type (TIR1) and PRC1deg cells before auxin treatment. Maximal read density in PRC1deg cells was set to e, A scatterplot showing the relationship between RING1B cChIP-seq signal at RING1B-bound sites in wild-type (TIR1) and PRC1deg cells before auxin treatment. R2 is the coefficient of determination for linear regression and cor is Pearson correlation coefficient. f, Immunoprecipitation of RING1B from TIR1 and PRC1deg nuclear extracts followed by western blot analysis of PRC1 components, and SUZ12 (PRC2 component) as a negative control. Shown are representative results from the two independent experiments. vPRC1 – variant PRC1 components, cPRC1 – canonical PRC1 components. g, Chromatin occupancy of PCGF2 (a canonical PRC1 complex component) at Polycomb target gene promoters and control loci in wild-type (TIR1) and PRC1deg cells before auxin treatment assessed by ChIP-qPCR. Individual biological replicates (n = 3) are shown as dots along with mean and SEM. h, As in (G), but for RYBP (a variant PRC1 complex component). (i) Western blot analysis of PRC1 components and SUZ12 in the nuclear extracts of the PRC1deg cells following IAA treatment. CFP1 serves as a loading control. Shown are representative results from the two independent experiments. j, Boxplots comparing H2AK119ub1 cChIP-seq signal before and after IAA treatment at RING1B-bound sites (n = 7240) and over 100 kb genomic windows (n = 27282). Read coverages represent merged spike-in normalised values from n = 3 biologically independent experiments. All signal is normalised to the median RPKM value of RING1B-bound sites in untreated cells. Boxes show interquartile range, center line represents median, whiskers extend by 1.5x IQR, while notches extend by 1.58x IQR/sqrt(n), giving a roughly 95% confidence interval for comparing medians.

Source data

Extended Data Fig. 2 Expression and chromatin features of genes derepressed after PRC1 removal.

a, Scatterplots comparing the log2 fold changes in gene expression (cnRNA-seq) at different time points after auxin treatment in PRC1deg cells with gene expression changes in PRC1CKO (72 h OHT) cells. R2 is the coefficient of determination for linear regression and cor is Pearson correlation coefficient. b, An UpSet plot for genes with significantly increased expression (p-adj < 0.05, fold change > 1.5) following auxin treatment of PRC1deg cells at the indicated time points, or a 72 h OHT treatment of PRC1CKO cells. The total number of genes in each set is shown in the bar chart on the left. Non-empty intersections were sorted by size, while excluding any intersection with less than 10 members for clarity. Intersections not included in the PRC1CKO (72 h OHT) set are highlighted in green and may imply compensatory effects on expression, while the blue column denotes gene changes exclusive to the PRC1CKO and could correspond to secondary effects on expression. c, Boxplots comparing the log2 fold changes in expression of genes split into three groups based on the earliest time point when they became derepressed. Log2 fold changes are derived from n = 3 biologically independent experiments. Dotted line represents 1.5-fold change. Boxes show interquartile range, center line represents median, whiskers extend by 1.5x IQR, while notches extend by 1.58x IQR/sqrt(n), giving a roughly 95% confidence interval for comparing medians. d, Boxplots comparing the gene expression levels in the untreated PRC1deg cells for the three groups of derepressed genes defined in (B). RPKM values represent mean of n = 3 biologically independent experiments. Boxplots are defined as in (C). e, Boxplots comparing the promoter (TSS ± 2.5 kb) cChIP-seq signal for RING1B, H2AK119ub1, SUZ12 and H3K27me3 in the untreated PRC1deg cells for the three groups of derepressed genes defined in (B). Read coverages represent merged spike-in normalised values from n = 3 biologically independent experiments. Boxplots are defined as in (C). f, Boxplots comparing the promoter (TSS ± 2.5 kb) cChIP-seq signal for total Pol II and H3K4me3 in the untreated PRC1deg cells for the three groups of derepressed genes defined in (B). Read coverages represent merged spike-in normalised values from n = 3 biologically independent experiments. Boxplots are defined as in (C).

Extended Data Fig. 3 Characterisation of PRC2 and H3K27me3 after PRC1 removal.

a, Heatmap analysis of SUZ12 cChIP-seq at TSSs for the three groups of genes defined by the earliest time of derepression in PRC1deg cells treated with IAA for the indicated times. Heatmaps are sorted by RING1B cChIP-seq signal in untreated cells. b, Boxplots illustrating log2 fold changes in expression (cnRNA-seq, left panel), SUZ12 cChIP-seq signal (middle) or H3K27me3 cChIP-seq signal (right) at promoters (TSS ± 2.5 kb) of Polycomb target genes showing a significant reduction in SUZ12 levels following 2 hours of IAA treatment. Log2 fold changes are derived from n = 3 biologically independent experiments. Genes are divided into Polycomb target genes that become derepressed (UP, n = 955) and those that do not change in expression (n.s., n = 3739) by 2 hours. SUZ12 binding was reduced in both groups, while H3K27me3 levels were only modestly affected. Boxes show interquartile range, center line represents median, whiskers extend by 1.5x IQR, while notches extend by 1.58x IQR/sqrt(n), giving a roughly 95% confidence interval for comparing medians. c, Boxplots comparing the promoter (TSS ± 2.5 kb) cChIP-seq signal of PRC2 components for the two groups of genes defined in (B), in the wild-type state. Polycomb target genes derepressed after 2 hours of PRC1 depletion are characterised by high levels of PRC2 components, in agreement with this group of genes being highly enriched in Polycomb chromatin domain features. SUZ12 cChIP-seq data is from this study (PRC1deg cells), while JARID2, AEBP2, PCL2 and EPOP are from30. Read coverages represent merged spike-in normalised values from n = 3 biologically independent experiments. Boxplots are defined as in (B). d, As in (A) but for H3K27me3 cChIP-seq. e,The dynamics of reduction in H3K27me3 cChIP-seq signal at RING1B-bound sites which show a significant reduction in H3K27me3 levels by 24 hours of IAA treatment (n = 5926). Central line represents median, with interquartile range shown as shaded area. A theoretical exponential decay function is shown as dashed line, assuming that H3K27me3 levels are halved with every cell cycle if maintenance is completely disrupted. The doubling time of mouse ES cells is approximately 12 hours. f, Genomic snapshots of typical Polycomb target genes showing cChIP-seq signal for H3K27me3 in wild-type (E14) and PRC2deg cells before dTAG-13 treatment. g, Metaplot analysis of H3K27me3 cChIP-seq at RING1B-bound sites in wild-type (E14) and PRC2deg cells before dTAG-13 treatment. Maximal read density in E14 cells was set to 1. h, A scatterplot showing the relationship between H3K27me3 cChIP-seq signal at RING1B-bound sites in wild-type (E14) and PRC2deg cells before dTAG-13 treatment. R2 is the coefficient of determination for linear regression and cor is Pearson correlation coefficient. i, Metaplot analysis of H3K27me3 cChIP-seq at RING1B-bound sites in PRC2deg cells before and after 2 h dTAG-13 treatment. Maximal read density in untreated cells was set to 1.

Extended Data Fig. 4 Effects on total Pol II and H3K4me3 at genes after PRC1 removal.

a, Heatmaps illustrating RING1B binding in untreated cells, and total Pol II and H3K4me3 cChIP-seq signal at TSSs of the three groups of genes defined by the earliest time of derepression in PRC1deg cells treated with IAA for the indicated times. Heatmaps are sorted by RING1B signal in untreated cells. b, Boxplots comparing the log2 fold changes in total Pol II cChIP-seq signal following IAA treatment for genes split into three groups defined by the earliest time of derepression (2 hours, n = 1044; 4 hours, n = 1822; 8 hours, n = 2017). Log2 fold changes are derived from n = 3 biologically independent experiments. The analysis was done at promoters (TSS ± 2.5 kb) and over gene bodies (TSS to TES). Boxes show interquartile range, center line represents median, whiskers extend by 1.5x IQR, while notches extend by 1.58x IQR/sqrt(n), giving a roughly 95% confidence interval for comparing medians. c, Boxplots comparing the log2 fold changes in H3K4me3 cChIP-seq signal following IAA treatment at promoter regions (TSS ± 2.5 kb) for genes split into three groups defined by the earliest time of derepression (2 hours, n = 1044; 4 hours, n = 1822; 8 hours, n = 2017). Log2 fold changes are derived from n = 3 biologically independent experiments. Boxplots are defined as in (B).

Extended Data Fig. 5 Effects on Pol II phosphorylation state after PRC1 removal.

a, Genomic snapshot of Hoxd locus showing cChIP-seq signal for RING1B, total Pol II, Ser5P-Pol II and Ser2P-Pol II in untreated PRC1deg cells and 2 hours after IAA addition. b, Heatmaps illustrating RING1B, total Pol II, Ser5P-Pol II and Ser2P-Pol II levels and changes in polymerase occupancy following 2 h of IAA treatment for three groups of genes defined by the earliest time of derepression. Heatmaps are sorted by RING1B signal in untreated cells. c, Metaplot analysis of Ser5P-Pol II cChIP-seq at promoters of the three groups of genes defined by the earliest time of derepression in untreated PRC1deg cells and 2 hours after IAA addition. The profiles represent the median signal over the shown genomic region. d, As in (C) but for Ser2P-Pol II cChIP-seq over gene bodies.

Extended Data Fig. 6 Detailed characterisation of the smRNA-FISH transcript counting approach.

a, A schematic illustrating our automated approach to analyse smRNA-FISH in a high throughput manner. This approach enables effective single-cell segmentation, conversion of the field of view into single-cell smRNA-FISH images, in which individual transcripts can then be counted using ThunderSTORM119. b, An illustration of the effectiveness of ThunderSTORM threshold factor (TF) value identification for spot (transcript) detection in cells. An Hspg2 dataset is shown as an example. The TF unit is set at a standard deviation of the image background. Spots per cell (mean ± SD; top) and heterogeneity of detected transcript levels expressed as coefficient of variation (middle) are shown for a range of TF values. Vertical red lines indicate the range of TF values yielding similar spot-counting outcomes and which were ultimately employed in this study. Density plots (bottom) demonstrate transcript per cell distributions depending on the TF used. TFs from 6 to 10 yield comparable spot-counting values. Very large or very low TF values lead either to overcounting or undercounting of transcript signals. c, An illustration of the reproducibility between technical and biological replicates for our smRNA-FISH transcript counting approach. Mean transcripts per cell (top), coefficient of variation (middle), and mean transcription burst size inferred from the 2-state model (bottom) are shown. Error bars in the top panel represent 10% of standard deviation of transcripts per cell distributions around the mean. d, To ensure the robustness of our transcript counting approach, we compared it to other spot (transcript) counting methods and manual counting of transcripts in 50 cells. The methods compared are: 3D-FISH QUANT125 (left), 3D Objects Counter120 (middle), and technique used in this study (right). The right panel indicates that our technique can be prone to a slight undercounting when number of transcripts per cell exceeds 60-70, but otherwise performs comparably to other approaches. Pearson correlation coefficient (r) and slope derived from linear regression are presented. e, Transcript counting using 3D-FISH Quant and 3D Objects Counter correlate well with the transcripts counting using our approach.

Source data

Extended Data Fig. 7 Effects on absolute transcript numbers in single cells after PRC1 removal.

a, An MA plot depicting gene expression (cnRNA-seq) changes following 4 h IAA treatment in PRC1deg cells with candidate genes for smRNA-FISH highlighted. The genes chosen span a wide range of initial transcript levels and transcript increase upon RING1B removal. The control genes are highlighted in blue and Polycomb target genes in black. b, Correlation of log2 fold changes between cnRNA-seq and smRNA-FISH after 4 hours of IAA treatment. Hollow dots represent genes excluded from the linear fit. Dots represent mean values while error bars represent standard error of 3 biological replicates. c, Normalised histograms illustrating the distribution of transcripts per cell over the time course of RING1B removal for all genes studied with smRNA-FISH. Shown is a representative biological replicate of 3 independent experiments. n indicates minimum number of cells measured at any given time-point per gene dataset. d, A heatmap of mean fold changes in transcripts per cell over the time course of RING1B removal. Numbers represent the mean of 3 biological replicates. e, As in (D) but representing the mean number of transcripts per cell.

Source data

Extended Data Fig. 8 Analysis of transcription burst size and frequency after PRC1 removal.

a, Examples of model fits to cellular transcripts distributions for all the genes examined in untreated state. The negative binomial fit is in red and density in blue. The Poisson distribution fit is indicated for comparison as a dashed purple line. b, A heatmap of the goodness of fit p-value (one-sided Chi-square test) throughout the time course of RING1B depletion. High p-value (light yellow) represents a good negative binomial fit to the data. c, Barplots representing mean inferred transcription burst sizes for Polycomb target genes and control genes. Error bars correspond to standard error of 3 biological replicates. d, The relationship between nascent spot intensity (active transcription site, measured using nascent smRNA-FISH targeting intronic sequences) and transcription burst size inferred from mRNA-FISH (targeting exonic sequences) reveals that genes with higher predicted burst size values have greater nascent spot intensity. Dots represent mean value while error bars correspond to standard deviation of 3 biological replicates. e, Heatmaps illustrating the fold change in inferred transcription burst size (left) and burst frequency (right) over the time course of RING1B depletion for the panel of genes studied. f, Examples of the relationship between square coefficient of variation (CV) and mean number of transcripts per cell demonstrate that Polycomb target genes derepressed upon PRC1 removal experience increase in transcript number while simultaneously retaining constant transcription burst size. Dashed lines represent theoretical relation between CV2 and the mean number of transcripts at constant burst size values (CV2 = b/mean#mRNA) with changing burst frequencies. Dots represent mean values while error bars correspond to the standard deviation of 3 biological replicates.

Source data

Extended Data Fig. 9 Active allele distribution in the cell population and transcription noise decomposition.

(a) Frequency of nascent spots representing active alleles in cell population (mean percentage of cells with 0, 1, or >1 nascent spot per cell obtained in 3 biological replicates). For comparison a simulated negative binomial distribution of active alleles per cell is shown assuming their expression is independent and random. The simulated distribution assumes the same frequency of spots as in the experimental data and 3 alleles per cell because a large proportion of ES cells will exist in S-phase at any given point. Importantly, the independence of allele expression is maintained 4 h after removal of PRC1 by the addition of IAA. A minimum of 887 cells was measured per biological replicate. (b) Pearson’s correlation coefficient between number of transcripts per cell and either DAPI intensity per cell (a proxy for the cell cycle stage as G2 cells have generally 2x the genetic material of the cells in G1), or cell area (representing cell volume) – the two primary sources of extrinsic noise. (c) i – iii) Examples of area or DAPI signal per cell plotted against transcripts per cell for a control (Hspg2) or Polycomb target genes (Zic2, E2f6). Each data-point represents a single cell measurement. Shading around regression line represents 95% confidence interval. iv) Correlation between cell area and DAPI signal intensity (Pearson’s r = 0.74) suggests that volume of ES cells is strongly related to their cell cycle phase. We note that the cell nucleus in ES cells occupies a significant portion of the total cell volume. (d) Specific heterogeneity (noise) in transcripts per cell expressed as % of the total heterogeneity measured as square coefficient of variation of the transcripts per cell distributions. Cell area and DAPI signal intensity contribute very little to the overall variability in number of transcripts per cell in a population for Polycomb target genes. We note, that this measure of extrinsic variability likely suffers from overestimation as DAPI intensity and cell area are moderately correlated, and hence to some extent represent the same source of variability (see C).

Source data

Supplementary information

Source data

Source Data Fig. 1

Unprocessed Western Blots

Source Data Fig. 1

Statistical source data.

Source Data Fig. 3

Unprocessed western blots.

Source Data Fig. 5

Statistical source data.

Source Data Fig. 6

Statistical source data.

Source Data Fig. 7

Statistical source data.

Source Data Extended Data Fig. 1

Unprocessed western blots.

Source Data Extended Data Fig. 1

Statistical source data.

Source Data Extended Data Fig. 6

Statistical source data.

Source Data Extended Data Fig. 7

Statistical source data.

Source Data Extended Data Fig. 8

Statistical source data.

Source Data Extended Data Fig. 9

Statistical source data.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Dobrinić, P., Szczurek, A.T. & Klose, R.J. PRC1 drives Polycomb-mediated gene repression by controlling transcription initiation and burst frequency. Nat Struct Mol Biol 28, 811–824 (2021). https://doi.org/10.1038/s41594-021-00661-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41594-021-00661-y

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