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:

RNA is essential for PRC2 chromatin occupancy and function in human pluripotent stem cells

Abstract

Many chromatin-binding proteins and protein complexes that regulate transcription also bind RNA. One of these, Polycomb repressive complex 2 (PRC2), deposits the H3K27me3 mark of facultative heterochromatin and is required for stem cell differentiation. PRC2 binds RNAs broadly in vivo and in vitro. Yet, the biological importance of this RNA binding remains unsettled. Here, we tackle this question in human induced pluripotent stem cells by using multiple complementary approaches. Perturbation of RNA–PRC2 interaction by RNase A, by a chemical inhibitor of transcription or by an RNA-binding-defective mutant all disrupted PRC2 chromatin occupancy and localization genome wide. The physiological relevance of PRC2–RNA interactions is further underscored by a cardiomyocyte differentiation defect upon genetic disruption. We conclude that PRC2 requires RNA binding for chromatin localization in human pluripotent stem cells and in turn for defining cellular state.

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: RNA is essential for PRC2 chromatin occupancy.
Fig. 2: Disruption of PRC2–RNA interaction in iPSCs leads to loss of EZH2 occupancy and H3K27me3 level on PRC2-repressed genes.
Fig. 3: PRC2–RNA interaction is essential for cardiomyocyte differentiation.
Fig. 4: A molecular scheme of RNA regulation of PRC2 activity.

Similar content being viewed by others

Data availability

Genomic data have been uploaded to GEO (accession GSE128135). All data, plasmids and cell lines used in the analysis are available upon request. Plasmids and cell lines are available upon request subject to a material transfer agreement (UBMTA) with the University of Colorado Boulder. Custom computational codes are available upon request as well as in GitHub (https://github.com/taeyoungh/PRC2). Gene set information for GSEA is available at https://www.gsea-msigdb.org/gsea/msigdb/index.jsp. Source data are provided with this paper.

References

  1. Hudson, W. H. & Ortlund, E. A. The structure, function and evolution of proteins that bind DNA and RNA. Nat. Rev. Mol. Cell Biol. 15, 749–760 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Hendrickson, D. G., Kelley, D. R., Tenen, D., Bernstein, B. & Rinn, J. L. Widespread RNA binding by chromatin-associated proteins. Genome Biol. 17, 28 (2016).

    Google Scholar 

  3. Long, Y., Wang, X., Youmans, D. T. & Cech, T. R. How do lncRNAs regulate transcription? Sci. Adv. 3, eaao2110 (2017).

    PubMed  PubMed Central  Google Scholar 

  4. Khalil, A. M. et al. Many human large intergenic noncoding RNAs associate with chromatin-modifying complexes and affect gene expression. Proc. Natl Acad. Sci. USA 106, 11667–11672 (2009).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  6. Muller, J. & Verrijzer, P. Biochemical mechanisms of gene regulation by polycomb group protein complexes. Curr. Opin. Genet. Dev. 19, 150–158 (2009).

    PubMed  Google Scholar 

  7. Schuettengruber, B. & Cavalli, G. Recruitment of Polycomb group complexes and their role in the dynamic regulation of cell fate choice. Development 136, 3531–3542 (2009).

    CAS  PubMed  Google Scholar 

  8. Margueron, R. & Reinberg, D. The Polycomb complex PRC2 and its mark in life. Nature 469, 343–349 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Helin, K. & Dhanak, D. Chromatin proteins and modifications as drug targets. Nature 502, 480–488 (2013).

    CAS  PubMed  Google Scholar 

  10. Pasini, D., Bracken, A. P., Hansen, J. B., Capillo, M. & Helin, K. The polycomb group protein Suz12 is required for embryonic stem cell differentiation. Mol. Cell. Biol. 27, 3769–3779 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Laugesen, A., Hojfeldt, J. W. & Helin, K. Molecular mechanisms directing PRC2 recruitment and H3K27 methylation. Mol. Cell 74, 8–18 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Davidovich, C. & Cech, T. R. The recruitment of chromatin modifiers by long noncoding RNAs: lessons from PRC2. RNA 21, 2007–2022 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. 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).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Beltran, M. et al. The interaction of PRC2 with RNA or chromatin is mutually antagonistic. Genome Res. 26, 896–907 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Kaneko, S., Son, J., Bonasio, R., Shen, S. S. & Reinberg, D. Nascent RNA interaction keeps PRC2 activity poised and in check. Genes Dev. 28, 1983–1988 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Cifuentes-Rojas, C., Hernandez, A. J., Sarma, K. & Lee, J. T. Regulatory interactions between RNA and Polycomb repressive complex 2. Mol. Cell 55, 171–185 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  17. Wei, C. et al. RBFox2 binds nascent RNA to globally regulate Polycomb complex 2 targeting in mammalian genomes. Mol. Cell 62, 875–889 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Guil, S. et al. Intronic RNAs mediate EZH2 regulation of epigenetic targets. Nat. Struct. Mol. Biol. 19, 664–670 (2012).

    CAS  PubMed  Google Scholar 

  19. Piunti, A. & Shilatifard, A. Epigenetic balance of gene expression by Polycomb and COMPASS families. Science 352, aad9780 (2016).

    PubMed  Google Scholar 

  20. Riising, E. M. et al. Gene silencing triggers Polycomb repressive complex 2 recruitment to CpG islands genome wide. Mol. Cell 55, 347–360 (2014).

    CAS  PubMed  Google Scholar 

  21. Youmans, D. T., Schmidt, J. C. & Cech, T. R. Live-cell imaging reveals the dynamics of PRC2 and recruitment to chromatin by SUZ12-associated subunits. Genes Dev. 32, 794–805 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Davidovich, C. et al. Toward a consensus on the binding specificity and promiscuity of PRC2 for RNA. Mol. Cell 57, 552–558 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Davidovich, C., Zheng, L., Goodrich, K. J. & Cech, T. R. Promiscuous RNA binding by Polycomb repressive complex 2. Nat. Struct. Mol. Biol. 20, 1250–1273 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Wang, X. et al. Targeting of Polycomb repressive complex 2 to RNA by short repeats of consecutive guanines. Mol. Cell 65, 1056–1067 e5 (2017).

    CAS  PubMed  Google Scholar 

  26. Long, Y. et al. Conserved RNA-binding specificity of polycomb repressive complex 2 is achieved by dispersed amino acid patches in EZH2. eLife 6, e31558 (2017).

    PubMed  PubMed Central  Google Scholar 

  27. Abruzzi, K. C., Lacadie, S. & Rosbash, M. Biochemical analysis of TREX complex recruitment to intronless and intron-containing yeast genes. EMBO J. 23, 2620–2631 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Thompson, P. J. et al. hnRNP K coordinates transcriptional silencing by SETDB1 in embryonic stem cells. PLoS Genet. 11, e1004933 (2015).

    PubMed  PubMed Central  Google Scholar 

  29. Casale, A. M., Cappucci, U., Fanti, L. & Piacentini, L. Heterochromatin protein 1 (HP1) is intrinsically required for post-transcriptional regulation of Drosophila Germline Stem Cell (GSC) maintenance. Sci. Rep. 9, 4372 (2019).

    PubMed  PubMed Central  Google Scholar 

  30. Bernstein, E. et al. Mouse polycomb proteins bind differentially to methylated histone H3 and RNA and are enriched in facultative heterochromatin. Mol. Cell Biol. 26, 2560–2569 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Barutcu, A. R., Blencowe, B. J. & Rinn, J. L. Differential contribution of steady-state RNA and active transcription in chromatin organization. EMBO Rep. 20, e48068 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Lee, C. H. et al. Automethylation of PRC2 promotes H3K27 methylation and is impaired in H3K27M pediatric glioma. Genes Dev. 33, 1428–1440 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Wang, X. et al. Regulation of histone methylation by automethylation of PRC2. Genes Dev. 33, 1416–1427 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Miyaoka, Y. et al. Isolation of single-base genome-edited human iPS cells without antibiotic selection. Nat. Methods 11, 291–U345 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Mandegar, M. A. et al. CRISPR interference efficiently induces specific and reversible gene silencing in human iPSCs. Cell Stem Cell 18, 541–553 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Wang, J. et al. Identification of LBX2 as a novel causal gene of atrial septal defect. Int. J. Cardiol. 265, 188–194 (2018).

    PubMed  Google Scholar 

  37. Ochi, H. & Westerfield, M. Lbx2 regulates formation of myofibrils. BMC Dev. Biol. 9, 13 (2009).

    PubMed  PubMed Central  Google Scholar 

  38. Hoffmann, A. D. et al. Foxf genes integrate tbx5 and hedgehog pathways in the second heart field for cardiac septation. PLoS Genet. 10, e1004604 (2014).

    PubMed  PubMed Central  Google Scholar 

  39. Lian, X. et al. Directed cardiomyocyte differentiation from human pluripotent stem cells by modulating Wnt/β-catenin signaling under fully defined conditions. Nat. Protoc. 8, 162–175 (2013).

    CAS  PubMed  Google Scholar 

  40. Delgado-Olguin, P. et al. Epigenetic repression of cardiac progenitor gene expression by Ezh2 is required for postnatal cardiac homeostasis. Nat. Genet. 44, 343–347 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Amendt, B. A., Sutherland, L. B. & Russo, A. F. Transcriptional antagonism between Hmx1 and Nkx2.5 for a shared DNA-binding site. J. Biol. Chem. 274, 11635–11642 (1999).

    CAS  PubMed  Google Scholar 

  42. Rinn, J. L. et al. Functional demarcation of active and silent chromatin domains in human HOX loci by noncoding RNAs. Cell 129, 1311–1323 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Zhao, J. et al. Genome-wide identification of Polycomb-associated RNAs by RIP-seq. Mol. Cell 40, 939–953 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Tsai, M. C. et al. Long noncoding RNA as modular scaffold of histone modification complexes. Science 329, 689–693 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Yu, J. R., Lee, C. H., Oksuz, O., Stafford, J. M. & Reinberg, D. PRC2 is high maintenance. Genes Dev. 33, 903–935 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Lee, C. H. et al. Allosteric activation dictates PRC2 activity independent of its recruitment to chromatin. Mol. Cell 70, 422–434 e6 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Oksuz, O. et al. Capturing the onset of PRC2-mediated repressive domain formation. Mol. Cell 70, 1149–1162 e5 (2018).

    CAS  PubMed  Google Scholar 

  48. Pirrotta, V. Polycomb Bodies. in The Functional Nucleus (eds Bazett-Jones, D. P., & Dellaire, G.) 157–173 (Springer, 2016).

  49. Cheutin, T. & Cavalli, G. Polycomb silencing: from linear chromatin domains to 3D chromosome folding. Curr. Opin. Genet. Dev. 25, 30–37 (2014).

    CAS  PubMed  Google Scholar 

  50. Schertzer, M. D. et al. lncRNA-induced spread of Polycomb controlled by genome architecture, RNA abundance, and CpG island DNA. Mol. Cell 75, 523–537 e10 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Fang, H. et al. Trans- and cis-acting effects of the lncRNA Firre on epigenetic and structural features of the inactive X chromosome. Preprint at https://www.biorxiv.org/content/10.1101/687236v1 (2019).

  52. Hacisuleyman, E. et al. Topological organization of multichromosomal regions by the long intergenic noncoding RNA Firre. Nat. Struct. Mol. Biol. 21, 198–206 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Martin, M. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet. J. 17, 10–12 (2011).

    Google Scholar 

  54. Li, H. & Durbin, R. Fast and accurate short read alignment with Burrows–Wheeler transform. Bioinformatics 25, 1754–1760 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    PubMed  PubMed Central  Google Scholar 

  56. Kharchenko, P. V., Tolstorukov, M. Y. & Park, P. J. Design and analysis of ChIP-seq experiments for DNA-binding proteins. Nat. Biotechnol. 26, 1351–1359 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  57. Wickham, H. Ggplot2: Elegant Graphics for Data Analysis (Springer, 2009).

  58. O’Leary, N. A. et al. Reference sequence (RefSeq) database at NCBI: current status, taxonomic expansion, and functional annotation. Nucleic Acids Res. 44, D733–D745 (2016).

    PubMed  Google Scholar 

  59. 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).

    PubMed  PubMed Central  Google Scholar 

  60. Strimmer, K. fdrtool: a versatile R package for estimating local and tail area-based false discovery rates. Bioinformatics 24, 1461–1462 (2008).

    CAS  PubMed  Google Scholar 

  61. Cong, L. et al. Multiplex genome engineering using CRISPR/Cas systems. Science 339, 819–823 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  62. Xi, L. H., Schmidt, J. C., Zaug, A. J., Ascarrunz, D. R. & Cech, T. R. A novel two-step genome editing strategy with CRISPR–Cas9 provides new insights into telomerase action and TERT gene expression. Genome Biol. 16, 231 (2015).

    PubMed  PubMed Central  Google Scholar 

  63. Liao, Y., Smyth, G. K. & Shi, W. featureCounts: an efficient general purpose program for assigning sequence reads to genomic features. Bioinformatics 30, 923–930 (2014).

    CAS  PubMed  Google Scholar 

  64. The Gene Ontology Consortium. The Gene Ontology Resource: 20 years and still GOing strong. Nucleic Acids Res. 47, D330–D338 (2019).

    Google Scholar 

  65. Ashburner, M. et al. Gene ontology: tool for the unification of biology. Nat. Genet. 25, 25–29 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  66. Liberzon, A. et al. Molecular signatures database (MSigDB) 3.0. Bioinformatics 27, 1739–1740 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  67. Wu, D. & Smyth, G. K. Camera: a competitive gene set test accounting for inter-gene correlation. Nucleic Acids Res. 40, e133 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  68. Huebsch, N. et al. Automated video-based analysis of contractility and calcium flux in human-induced pluripotent stem cell-derived cardiomyocytes cultured over different spatial scales. Tissue Eng. Part C Methods 21, 467–479 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank L. Leinwand, O. Uhlenbeck, D. Youmans, T. Rowland, and Cech and Rinn laboratory members for useful discussions and technical assistance. We thank core facility directors J. Dragavon (BioFrontiers Advanced Light Microscopy Core), T. Nahreini (Biochemistry Cell Culture Facility and Flow Cytometry Shared Core; grant no. S10ODO21601), A. Scott (BioFrontiers Next Generation Sequencing Facility) and D. Timmons (BioFrontiers IT). We also thank B. Conklin (Gladstone Institute) for providing the WTC-11 iPSC cell line, J. Wheeler and J. Silver (CU-Boulder) for protocols of iPSC culture, M. Regnier (University of Washington) for advice on cardiomyocyte differentiation, N. Huebsch (Washington University in St. Louis) for providing the motion-tracking MATLAB platform/algorithm, and R. Parker and C. Decker (CU-Boulder) for access to the DeltaVision microscope. Y.L. is supported by NIH K99 award no. K99GM132546. T.R.C. is an investigator of the Howard Hughes Medical Institute (HHMI). J.L.R. is an HHMI Faculty Scholar and holds a Marvin H. Caruthers Endowed Chair for Early Career Faculty. J.L.R. and T.H. are supported by NIH P01 award no. P01GM099117.

Author information

Authors and Affiliations

Authors

Contributions

Y.L., T.H., J.L.R. and T.R.C. conceived the study and designed experiments. Y.L., A.R.G. and K.J.G. performed the biochemical characterization and data analysis of purified protein complexes. Y.L. performed and analyzed genome editing of iPSCs, cardiomyocyte differentiation, imaging, flow cytometry, ChIP and fRIP experiments, and deep sequencing library preparation. T.H. developed computational pipelines and performed genomic and statistical analyses. Y.L., T.H., J.L.R. and T.R.C. wrote the manuscript.

Corresponding authors

Correspondence to John L. Rinn or Thomas R. Cech.

Ethics declarations

Competing interests

T.R.C. is on the Merck board and is a consultant for Storm Therapeutics.

Additional information

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

Extended data

Extended Data Fig. 1 rChIP-seq of TBP, Rpb1, EZH2 and SUZ12 in human iPSCs.

a, Venn diagrams with number of identified peaks in EZH2 and SUZ12 ChIP-seq experiments in -RNase A and +RNase A samples. The numbers of peaks identified in either of two independent clones were merged and counted. b, Metaplot of fold change (pulldown/input) for TBP rChIP-seq in two WT clones (black: WT-A, blue: WT-B) at transcription start site (TSS) + /− 5 kb. c, Motif analysis of TBP ChIP-seq identified the A/T rich sequence of TBP binding sites. This motif (statistical significance quantified by E-value: 7.5*10−16) was found by using MEME (http://meme-suite.org/) with top 1000 peaks. d, Metaplot of Rpb1 ChIP-seq in two WT clones (black: WT-A, blue: WT-B) at transcription start site (TSS) + /− 5 kb. e, Metaplot of Rpb1 ChIP-seq in two WT clones (black: WT-A, blue: WT-B) around gene bodies [from TSS – 5 kb to transcription termination site (TTS) + 5 kb] of the genes with peaks. Every gene is length-normalized to have the same number of tiles. f–i, RNase A treatment (y-axis) was compared to control (x-axis) in terms of fold change (log scale) of genes in pulldown relative to input. Two independent clones in each condition (N = 4) were used for empirical Wald test (two-sided, see methods). Each dot represents a gene. The red denotes the genes with significant fold changes [multiple test-corrected (FDR) p-value < 0.1] between the treatment and the control. Dashed lines, expectation for RNA-independent ChIP. j–m, heatmap comparison of minus and plus RNase A treatment in terms of fold changes in pulldown relative to input. These are the same analyses as in Fig. 1b–e but using an independent clone.

Extended Data Fig. 2 EZH2 ChIP with or without triptolide treatment of human iPSCs.

a, Heatmaps of EZH2 ChIP fold change (pulldown/input) without and with (columns) triptolide in two different iPSC clones (rows). Each row describes a region of 5 kb upstream and downstream from the center of a single ChIP-seq peak. b, Triptolide treatment (y-axis) was compared to wild-type/untreated control (x-axis) in terms of fold change (log scale) of genes in pulldown relative to input. Two independent clones in each condition (N = 4) were used for empirical Wald tests (two-sided, see methods). Each dot represents a gene. The red denotes the genes with significant fold changes [FDR-adjusted p-value < 0.1] between the treatment and the wild-type control. Diagonal dashed line is the expectation for triptolide-independent EZH2 binding to chromatin. c, EZH2 ChIP-qPCR results of three PRC2 target genes upon triptolide treatment. WT and MT (RNA binding mutant) are the two CRISPR-engineered iPSC strains presented in Fig. 2. The bar represents mean value and error bar represents the standard deviation of four biological replicates (two clones, each in two independent experiments). d, Distribution of EZH2 ChIP enrichment (log2(IP/input)) at four groups of genes. These four groups are divided based on their EZH2 ChIP enrichment in control samples (WT) as evaluated by empirical Wald tests (two-sided, see methods): Non-target group consists of the depleted genes in WT EZH2 ChIP, based on FDR-adjusted p-value<0.05 and fold change of pulldown/input<1 (N = 1139); Weak target group consists of the enriched genes in WT EZH2 ChIP, based on multiple-test corrected (FDR) p-value<0.05 and 1 < = fold change of pulldown/input <2 (N = 3183); Strong target group consists of the enriched genes in IP relative to Input based on FDR-adjusted p-value<0.05 and fold change of pulldown/input > =2 (N = 285); all the other genes comprise background group (N = 31504). e, Fold change of EZH2 occupancy (triptolide divided by WT/untreated control) at individual genes. The distribution is described by a typical box plot: the box represents the first and third quartiles of percentage distribution with the horizontal line showing the median, and whiskers indicate 10th and 90th percentiles. All comparisons against background are statistically significant by Wilcoxon rank sum test (*** indicates p < 10−15 in two-sided test). The number of genes in statistical tests is same as in d.

Extended Data Fig. 3 Characterization of a separation-of-function human PRC2 mutant that is specifically defective in RNA interaction.

a, (Left) The mutant contains mutations at two sites of the EZH2 subunit, indicated with stars on the cryo-EM structure of PRC2 (PDB 6c23). Dashed red line, flexible loop. (Right) Table of all tested functions of PRC2. Green check mark, no substantial change compared to wild type (WT); red cross, substantial defect for the mutant. b, Coomassie-stained SDS-PAGE gel of recombinant 5-mer PRC2 WT and MT as used for following biochemical experiments. (cf) WT in red and MT (mutant) in blue for all curves. For all binding curves, center of the dot represents mean value and the error bar represents the standard deviation of three independent experiments. c, Equilibrium binding of PRC2 to a G-quadruplex 40-mer RNA (GGAA)10 was measured using Electrophoretic Mobility Shift Assay (EMSA). Kdapp values of WT and MT: 2 and 12 nM, p-value = 0.0007, calculated by Student’s T-test with two-tailed distribution and two-sample equal variance). First lane of gel is RNA only, no protein. Subsequent lanes have successive three-fold dilutions of PRC2 from 5 μM. d, (left) Kinetic dissociation of PRC2-(GGAA)10 RNA complex was measured using fluorescence anisotropy, with unlabeled (GGAA)10 RNA added at time 0 to trap dissociated PRC2. Dissociation rate constants (koff) for 4 m and 5 m complexes are indicated and the dissociation curves for 4 m are shown. (right) Kinetic dissociation of PRC2-(CG)30 dsDNA. e, PRC2-DNA binding was tested using EMSA on a TERT promoter DNA (48 bp) with either unmethylated (left) or methylated (right) CpG dinucleotides. f, Binding of PRC2 to mono-nucleosomes (left) or tri-nucleosomes (right) was measured using EMSA. g, Methyltransferase activities of 4 m PRC2 complexes were measured on either histone H3, the entire histone octamer, reconstituted mononucleosomes or native polynucleosomes (Amsbio 52015). Automethylation of EZH2 also quantified. For all assays, center of the dot represents mean value and error bars give standard derivation of three independent experiments.

Extended Data Fig. 4 In vivo binding defect of the MT EZH2 exhibited by fRIP-seq.

fRIP experiment was used to examine the RNAs associated with PRC2. EZH2 MT (y-axis) was compared to EZH2 WT control (x-axis) for RNAs in terms of enrichment score defined as fold change (log scale) of pulldown relative to input divided by standard deviation. Two independent clones were used in each genotype to calculate enrichment scores for individual genes. Each dot represents the RNA product of a gene. The significantly enriched genes in WT, determined by empirical Wald tests with two independent clones (N = 2, multiple test-corrected (FDR) p-value<0.05, see methods) and denoted by red, are located off the dashed 45-degree line that indicates the expectation if the MT EZH2 had no effect on RNA pulldown.

Extended Data Fig. 5 Introduction of the separation-of-function RNA mutant did not change the undifferentiated state of the iPSCs.

a, (Left) Schematic representation of the CRISPR genome editing to construct the WT and MT iPSC strains. Locations of the primer binding sites are shown for the four primers that are used in the PCR validation. (Right) Validation of the correct integration of the donor DNA at the CRISPR guide RNA cleavage site by PCR on genomic DNA. Two clones of each strain were validated, and the expected PCR amplicon sizes of edited and unedited allele are shown in the table below the gels. Data indicate absence of unedited alleles for all WT and MT clones. DNA ladders are labeled in base pairs on the side of the image. b, Appearance of the stem cell colonies under bright field microscope. White scale bar represents 1000 µm. c, Karyotyping of two clones of each edited strain; no abnormality of the 23 pairs was observed. d, Immunofluorescence results of the CRISPR-edited clones to evaluate expression of pluripotency markers (SOX2, OCT4 and SSEA4) and EZH2. e, Zoomed-in image of d comparing subcellular localization of EZH2 in WT and MT clones. f, EZH2 co-immunoprecipitation (co-IP) experiment in each of the two clones of the WT and MT iPSCs. Western blot for EZH2, SUZ12, EED, RBBP4, JARID2 and MTF2 are shown. Note that the IgG control pulldown doesn’t precipitate any of the PRC2 subunits, and the IgG bands in these blots are lighted by the secondary antibody (anti-mouse IgG HRP).

Extended Data Fig. 6 ChIP-seq of EZH2 and H3K27me3 in WT iPSCs.

a, Comparison of the EZH2 ChIP (x-axis) and H3K27me3 ChIP (y-axis) in the WT iPSCs. Log2 fold changes of IP relative to input are compared for every gene with two independent clones in each ChIP-seq. Colors indicate the statistical significance determined by empirical Wald tests [multiple test-corrected (FDR) p-value<0.05 in two-sided] of log2 fold change. The number of these genes are 3231 (blue: EZH2 only), 17 (orange: H3K27me3 only), 237 (red: both) and 33920 (gray: no significance). Venn diagram is shown on the top of scatter plot. Significant association of enriched genes between EZH2 ChIP and H3K27me3 ChIP was found (Fisher-exact test, p-value < 10−15). b, An example of a red dot gene in a (note that the EZH2 and H3K27me3 peaks correlate very well). c, Two examples of blue dot genes in a (note that there are no significant H3K27me3 peaks in the region).

Extended Data Fig. 7 Comparison of ChIP-seq between EZH2 WT and MT, and validation of G4 RNA-PRC2 binding using TMPyP4.

a, Fold changes of EZH2 ChIP-seq in two WT clones (red) and two MT clones (blue) at transcription start site (TSS) + /− 5 kb. b, TMPyP4 treatment phenocopies the RNA-binding defective EZH2 mutant. EZH2 ChIP-qPCR results of WT, WT with TMPyP4 treatment, and MT. Enrichment percentage (pulldown/input) plotted as the Y-axis. Five genes that had significant change of EZH2 occupancy with the mutant EZH2 are shown in the top panels, and the insignificantly changed genes are shown in the lower panels. Each point represents a replicate (N = 3 or 4 biological independent samples). The bar represents mean value and the error bars show standard deviation. c, Fold change of H3K27me3 ChIP-seq. In a heatmap, each row describes a region of 5 kb upstream and downstream from the center of a single peak. d, Comparison of the differential enrichment between WT and MT between EZH2 ChIP (x-axis) and H3K27me3 ChIP (y-axis). The differential enrichment was quantified by subtraction of log2 fold change in WT from log2 fold change in MT. Colors indicate the statistical significance [multiple test corrected (FDR) p-value<0.1 in two-sided empirical Wald test, see methods] of differential enrichment. The number of these genes are 146 (blue: EHZ2 only), 9 (orange: H3K27me3 only), 96 (red: both) and 36946 (gray: no significance). Venn diagram is shown on the top of scatter plot. Significant overlap of differential genes was found between EZH2 ChIP and H3K27me3 ChIP (Fisher-exact test, p-value < 10−15).

Extended Data Fig. 8 Genome-wide gene expression changes in iPSCs caused by disruption of PRC2-RNA interaction.

a, Global gene expression change in MT relative to WT by RNA-seq. Two independent clones generating two sequencing libraries each in both WT and MT (N = 8) were used for statistical significance (two-sided Wald tests, see methods). Each dot represents a gene where x-value is Transcripts Per Million (TPM) mean value of WT and y-value is TPM mean value of MT. Color indicates three groups of genes in comparison of mean TPM between WT and MT: gray, insignificant [multiple test-corrected (FDR) p-value > =0.05]; orange, significant (FDR-adjusted p-value<0.05) and log2(FC) < 1; red, significant (FDR-adjusted p-value<0.05) and log2(FC) > = 1, where FC (fold change) is ratio of MT to WT in normalized read counts. b, Comparison of cumulative distribution of gene expression between the two groups of genes (N is the number of genes): red, significantly perturbed genes in both EZH2 and H3K27me3 ChIP-seq (N = 51); grey, all other genes (N = 21842). The curve shifted to right (red relative to gray) indicates higher overall x-values (expression increase in MT compared to WT): p-value=0.0003714 by two-sample Kolmogorov-Smirnov test (two-sided, D = 0.29056). c, GSEA top 10 significant gene sets for differential gene expression. All of these gene sets had higher expression in MT relative to WT (Supplementary Table 2). Blue texts indicate heart-related terms. d, An example of significant term from c related to heart development: in the top panel, vertical bar with gene name indicates the location of genes in this term on the ranked score of differential expression from down-regulation to up-regulation in MT. The two regions are marked by colors for visualization indicating lower (blue) and upper (red) intervals of differential gene expression statistic/score for gene set enrichment analysis. Default values in R package limma are used for the two cut-offs. The bottom panel shows the updates of enrichment score (y-axis) along the ranked differential score (horizontal bar) in GSEA method. Differential score for a gene is a statistic calculated by DESeq2 for differential gene expression (see method).

Extended Data Fig. 9 Immunofluorescence staining of neuronal progenitor cell (NPC) markers (Nestin, SOX1 and PAX6) in iPSC-derived NPCs.

a, day 11 NPCs were stained with Nestin (green) and SOX1 (red). b, day 8 NPCs were stained with PAX6 (green) and OCT4 (red). OCT4 is a pluripotency marker while PAX6, Nestin and SOX1 are all NPC markers. Scale bars represent 15 µm.

Extended Data Fig. 10 Molecular details of the cardiomyocyte differentiation defect.

a, Western blot (cropped) shows expression levels of the transfected WT or mutant FLAG-tagged EZH2 in the rescue experiment. b, Molecular response in the rescue experiment. qRT-PCR experiments were used to compare expression of two cardiac genes (TNNT2 and MYH6) and one control gene (SOX2) in each clone rescued with each WT or MT EZH2. Three biological replicates are represented with dots. The bar represents mean value and error bar represents standard deviation. Expressions were normalized to GAPDH level. P-values calculated in Student’s T-test with two-tailed distribution and two-sample equal variance are shown on top of each bar plot. c, RNA-seq results of WT and MT strains at day 8 of differentiation. Sample size (N = 12), statistical test and color legends are exactly the same as described in Fig. 3g. d, GSEA analysis of significantly differentially-expressed genes between WT and MT in RNA-seq. e, Mean differential expression (MT/WT, expression quantified by RNA-seq) was compared across three time points [day 0 (undifferentiated), day 8 and day 12 of differentiation] for the 18 genes among the 96 genes that have significant changes in both EZH2 and H3K27me3 ChIP-seqs between WT and MT (genes denoted by red in the Extended Data Fig. 7d). These 18 genes were selected as they have significant changes in gene expression between WT and MT at any time points. Time points at which a gene is significantly differentially expressed are shown as larger triangles. f, The distribution of the 96 genes that have significant changes in both EZH2 and H3K27me3 occupancies (measured by ChIP-seq) between WT and MT (the genes denoted by red in the Extended Data Fig. 8e) in terms of mean differential expression (MT/WT, expression quantified by RNA-seq) across three differentiation time points. The distribution is described by a typical box plot: the box represents the second and third quartiles with the horizontal line showing the median, and whiskers indicate 10th and 90th percentiles (the number of detected genes, N = 51, 65, and 63 for day 0, day 8, and day 12, respectively). g, Expression levels (Mean TPM with standard error, N = 3 biological replicates for each bar) of transcription factors HMX1 and NKX2–5 at day 0, 8 and 12 of differentiation determined by RNA-seq analysis.

Supplementary information

Supplementary Information

Supplementary Figs. 1–5

Reporting Summary

41588_2020_662_MOESM3_ESM.xlsx

Supplementary Tables 1–4. Supplementary Table 1. List of genes that have significant change in EZH2 occupancy or H3K27me3 between wild-type and mutant strain. Supplementary Table 2. List of significant gene ontology terms from gene set enrichment analysis (GSEA) of day 0 RNA-seq data. Supplementary Table 3. List of genes that have significant change in expression level between wild-type and mutant strain on day 0, 8 or 12 of cardiomyocyte differentiation. Supplementary Table 4. Nucleotide sequences used in this study.

Supplementary Video 1

30-s movies of day 11 cardiomyocytes for WT clone A.

Supplementary Video 2

30-s movies of day 11 cardiomyocytes for WT clone B.

Supplementary Video 3

30-s movies of day 11 cardiomyocytes for MT clone A

Supplementary Video 4

30-s movies of day 11 cardiomyocytes for MT clone B.

Source data

Source Data Fig. 1

Uncropped immunoblot from Fig. 2.

Source Data Fig. 2

Uncropped immunoblot from Extended Data Fig. 5f.

Source Data Fig. 3

Uncropped immunoblot from Extended Data Fig. 10a.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Long, Y., Hwang, T., Gooding, A.R. et al. RNA is essential for PRC2 chromatin occupancy and function in human pluripotent stem cells. Nat Genet 52, 931–938 (2020). https://doi.org/10.1038/s41588-020-0662-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41588-020-0662-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