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:

Whsc1 links pluripotency exit with mesendoderm specification

Abstract

How pluripotent stem cells differentiate into the main germ layers is a key question of developmental biology. Here, we show that the chromatin-related factor Whsc1 (also known as Nsd2 and MMSET) has a dual role in pluripotency exit and germ layer specification of embryonic stem cells. On induction of differentiation, a proportion of Whsc1-depleted embryonic stem cells remain entrapped in a pluripotent state and fail to form mesendoderm, although they are still capable of generating neuroectoderm. These functions of Whsc1 are independent of its methyltransferase activity. Whsc1 binds to enhancers of the mesendodermal regulators Gata4, T (Brachyury), Gata6 and Foxa2, together with Brd4, and activates the expression of these genes. Depleting each of these regulators also delays pluripotency exit, suggesting that they mediate the effects observed with Whsc1. Our data indicate that Whsc1 links silencing of the pluripotency regulatory network with activation of mesendoderm lineages.

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: Whsc1 is involved in the exit from pluripotency of mouse ESCs.
Fig. 2: Whsc1 is required for mesendoderm differentiation.
Fig. 3: Ablation of Whsc1 results in delayed downregulation of pluripotency genes and impairment of mesendoderm formation.
Fig. 4: The N terminus of Whsc1 is sufficient to rescue pluripotency exit and mesendoderm differentiation in Whsc1−/− cells.
Fig. 5: Whsc1 controls the enhancer activity of mesendoderm transcription factors.
Fig. 6: Whsc1 interacts with Brd4 and Gata6 on enhancers of mesendoderm transcription factors.
Fig. 7: Mesendoderm transcription factors are required for efficient pluripotency exit.

Similar content being viewed by others

Data availability

The RNA-Seq and UMI-4C data generated have been deposited in the Gene Expression Omnibus under accession number GSE126618. The accession numbers of the published datasets used are as follows: HiC in ESCs: GSE96611 (ref. 61); H3K27ac ChIP-Seq in Flk1+ mesendodermal progenitors: GSE47082 (ref. 62); H3K27ac ChIP-Seq in Eomes + mesendodermal progenitors: GSE103262 (ref. 63); H3K27ac ChIP-Seq in activin A-induced mesendodermal precursors: GSE38596 (ref. 64); and H3K27ac ChIP-Seq in neural progenitor cells: GSE35496 (ref. 65). Source data for all figures have been provided as Supplementary Table 4. All other data supporting the findings of this study are available from the corresponding authors on reasonable request.

References

  1. Nichols, J. & Smith, A. The origin and identity of embryonic stem cells. Development 138, 3–8 (2011).

    Article  CAS  PubMed  Google Scholar 

  2. Jaenisch, R. & Young, R. Stem cells, the molecular circuitry of pluripotency and nuclear reprogramming. Cell 132, 567–582 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Young, R. A. Control of the embryonic stem cell state. Cell 144, 940–954 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Martello, G. & Smith, A. The nature of embryonic stem cells. Annu. Rev. Cell Dev. Biol. 30, 647–675 (2014).

    Article  CAS  PubMed  Google Scholar 

  5. Clevers, H., Loh, K. M. & Nusse, R. Stem cell signaling. An integral program for tissue renewal and regeneration: Wnt signaling and stem cell control. Science 346, 1248012 (2014).

    Article  PubMed  Google Scholar 

  6. Loh, K. M., Lim, B. & Ang, L. T. Ex uno plures: molecular designs for embryonic pluripotency. Physiol. Rev. 95, 245–295 (2015).

    Article  PubMed  Google Scholar 

  7. Loh, K. M. & Lim, B. A precarious balance: pluripotency factors as lineage specifiers. Cell Stem Cell 8, 363–369 (2011).

    Article  CAS  PubMed  Google Scholar 

  8. Wray, J. et al. Inhibition of glycogen synthase kinase-3 alleviates Tcf3 repression of the pluripotency network and increases embryonic stem cell resistance to differentiation. Nat. Cell Biol. 13, 838–845 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Respuela, P. et al. Foxd3 promotes exit from naive pluripotency through enhancer decommissioning and inhibits germline specification. Cell Stem Cell 18, 118–133 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Martello, G. et al. Esrrb is a pivotal target of the Gsk3/Tcf3 axis regulating embryonic stem cell self-renewal. Cell Stem Cell 11, 491–504 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Leeb, M., Dietmann, S., Paramor, M., Niwa, H. & Smith, A. Genetic exploration of the exit from self-renewal using haploid embryonic stem cells. Cell Stem Cell 14, 385–393 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Geula, S. et al. m6A mRNA methylation facilitates resolution of naive pluripotency toward differentiation. Science 347, 1002–1006 (2015).

    Article  CAS  PubMed  Google Scholar 

  13. Betschinger, J. et al. Exit from pluripotency is gated by intracellular redistribution of the bHLH transcription factor Tfe3. Cell 153, 335–347 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Ho, L. & Crabtree, G. R. Chromatin remodelling during development. Nature 463, 474–484 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Di Croce, L. & Helin, K. Transcriptional regulation by Polycomb group proteins. Nat. Struct. Mol. Biol. 20, 1147–1155 (2013).

    Article  CAS  PubMed  Google Scholar 

  16. Zhang, Y. et al. H3K36 histone methyltransferase Setd2 is required for murine embryonic stem cell differentiation toward endoderm. Cell Rep. 8, 1989–2002 (2014).

    Article  CAS  PubMed  Google Scholar 

  17. 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 

  18. Aloia, L. et al. Zrf1 is required to establish and maintain neural progenitor identity. Genes Dev. 28, 182–197 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. 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 

  20. Arai, S. & Miyazaki, T. Impaired maturation of myeloid progenitors in mice lacking novel Polycomb group protein MBT-1. EMBO J. 24, 1863–1873 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Bergemann, A. D., Cole, F. & Hirschhorn, K. The etiology of Wolf–Hirschhorn syndrome. Trends Genet. 21, 188–195 (2005).

    Article  CAS  PubMed  Google Scholar 

  22. Fei, T. et al. Smad2 mediates Activin/Nodal signaling in mesendoderm differentiation of mouse embryonic stem cells. Cell Res. 20, 1306–1318 (2010).

    Article  CAS  PubMed  Google Scholar 

  23. Wang, Q. et al. The p53 family coordinates wnt and nodal inputs in mesendodermal differentiation of embryonic stem cells. Cell Stem Cell 20, 70–86 (2017).

    Article  CAS  PubMed  Google Scholar 

  24. Toyooka, Y., Shimosato, D., Murakami, K., Takahashi, K. & Niwa, H. Identification and characterization of subpopulations in undifferentiated ES cell culture. Development 135, 909–918 (2008).

    Article  CAS  PubMed  Google Scholar 

  25. Hayashi, K., Ohta, H., Kurimoto, K., Aramaki, S. & Saitou, M. Reconstitution of the mouse germ cell specification pathway in culture by pluripotent stem cells. Cell 146, 519–532 (2011).

    Article  CAS  PubMed  Google Scholar 

  26. Costello, I. et al. The T-box transcription factor Eomesodermin acts upstream of Mesp1 to specify cardiac mesoderm during mouse gastrulation. Nat. Cell Biol. 13, 1084–1091 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Morrison, G., Scognamiglio, R., Trumpp, A. & Smith, A. Convergence of cMyc and β-catenin on Tcf7l1 enables endoderm specification. EMBO J. 35, 356–368 (2016).

    Article  CAS  PubMed  Google Scholar 

  28. Aloia, L., Di Stefano, B. & Di Croce, L. Polycomb complexes in stem cells and embryonic development. Development 140, 2525–2534 (2013).

    Article  CAS  PubMed  Google Scholar 

  29. Kuo, A. J. et al. NSD2 links dimethylation of histone H3 at lysine 36 to oncogenic programming. Mol. Cell 44, 609–620 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Popovic, R. et al. Histone methyltransferase MMSET/NSD2 alters EZH2 binding and reprograms the myeloma epigenome through global and focal changes in H3K36 and H3K27 methylation. PLoS Genet. 10, e1004566 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  31. Nimura, K. et al. A histone H3 lysine 36 trimethyltransferase links Nkx2-5 to Wolf–Hirschhorn syndrome. Nature 460, 287–291 (2009).

    Article  CAS  PubMed  Google Scholar 

  32. Batista, P. J. et al. m6A RNA modification controls cell fate transition in mammalian embryonic stem cells. Cell Stem Cell 15, 707–719 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Le Dily, F. et al. Distinct structural transitions of chromatin topological domains correlate with coordinated hormone-induced gene regulation. Genes Dev. 28, 2151–2162 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Rowley, M. J. & Corces, V. G. Organizational principles of 3D genome architecture. Nat. Rev. Genet. 19, 789–800 (2018).

    Article  CAS  PubMed  Google Scholar 

  35. Devaiah, B. N. et al. BRD4 is a histone acetyltransferase that evicts nucleosomes from chromatin. Nat. Struct. Mol. Biol. 23, 540–548 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Sarai, N. et al. WHSC1 links transcription elongation to HIRA-mediated histone H3.3 deposition. EMBO J. 32, 2392–2406 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Filippakopoulos, P. et al. Selective inhibition of BET bromodomains. Nature 468, 1067–1073 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Di Micco, R. et al. Control of embryonic stem cell identity by BRD4-dependent transcriptional elongation of super-enhancer-associated pluripotency genes. Cell Rep. 9, 234–247 (2014).

    Article  CAS  PubMed  Google Scholar 

  39. Hoffman, J. A., Wu, C. I. & Merrill, B. J. Tcf7l1 prepares epiblast cells in the gastrulating mouse embryo for lineage specification. Development 140, 1665–1675 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Wen, J. et al. Single-cell analysis reveals lineage segregation in early post-implantation mouse embryos. J. Biol. Chem. 292, 9840–9854 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Morgani, S. M., Metzger, J. J., Nichols, J., Siggia, E. D. & Hadjantonakis, A. K. Micropattern differentiation of mouse pluripotent stem cells recapitulates embryo regionalized cell fate patterning. eLife 7, e32839 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  42. Teo, A. K. et al. Pluripotency factors regulate definitive endoderm specification through eomesodermin. Genes Dev. 25, 238–250 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Loh, K. M. & Lim, B. Epigenetics: actors in the cell reprogramming drama. Nature 488, 599–600 (2012).

    Article  CAS  PubMed  Google Scholar 

  44. Radzisheuskaya, A. et al. A defined Oct4 level governs cell state transitions of pluripotency entry and differentiation into all embryonic lineages. Nat Cell Biol. 15, 579–590 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Wamaitha, S. E. et al. Gata6 potently initiates reprograming of pluripotent and differentiated cells to extraembryonic endoderm stem cells. Genes Dev. 29, 1239–1255 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Jaffe, J. D. et al. Global chromatin profiling reveals NSD2 mutations in pediatric acute lymphoblastic leukemia. Nat. Genet. 45, 1386–1391 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Shen, C. et al. NSD3-short is an adaptor protein that couples BRD4 to the CHD8 chromatin remodeler. Mol. Cell 60, 847–859 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Bennett, R. L., Swaroop, A., Troche, C. & Licht, J. D. The role of nuclear receptor-binding SET domain family histone lysine methyltransferases in cancer. Cold Spring Harb. Perspect. Med. 7, a026708 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  49. Rutherford, E. L. & Lowery, L. A. Exploring the developmental mechanisms underlying Wolf–Hirschhorn syndrome: evidence for defects in neural crest cell migration. Dev. Biol. 420, 1–10 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Lee, Y. F., Nimura, K., Lo, W. N., Saga, K. & Kaneda, Y. Histone H3 lysine 36 methyltransferase Whsc1 promotes the association of Runx2 and p300 in the activation of bone-related genes. PLoS ONE 9, e106661 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  51. Gillich, A. et al. Epiblast stem cell-based system reveals reprogramming synergy of germline factors. Cell Stem Cell 10, 425–439 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Ran, F. A. et al. Genome engineering using the CRISPR–Cas9 system. Nat. Protoc. 8, 2281–2308 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Aparicio-Prat, E. et al. DECKO: single-oligo, dual-CRISPR deletion of genomic elements including long non-coding RNAs. BMC Genomics 16, 846 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  54. Di Stefano, B. et al. C/EBPα poises B cells for rapid reprogramming into induced pluripotent stem cells. Nature 506, 235–239 (2014).

    Article  CAS  PubMed  Google Scholar 

  55. Schneider, C. A., Rasband, W. S. & Eliceiri, K. W. NIH Image to ImageJ: 25 years of image analysis. Nat. Methods 9, 671–675 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  57. Schwartzman, O. et al. UMI-4C for quantitative and targeted chromosomal contact profiling. Nat. Methods 13, 685–691 (2016).

    Article  CAS  PubMed  Google Scholar 

  58. LI, H. Aligning sequence reads, clone sequences and assembly contigs with BWA-MEM. Preprint at https://arxiv.org/abs/1303.3997 (2013).

  59. Bates, D. Mächler, M., Bolker, B. & Walker, S. Fitting linear mixed-effects models using lme4. Preprint at https://arxiv.org/abs/1406.5823 (2014).

  60. R Development Core Team R: A Language and Environment for Statistical Computing (R Foundation for Statistical Computing, 2018).

  61. Stadhouders, R. et al. Transcription factors orchestrate dynamic interplay between genome topology and gene regulation during cell reprogramming. Nat. Genet. 50, 238–249 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Org, T. et al. Scl binds to primed enhancers in mesoderm to regulate hematopoietic and cardiac fate divergence. EMBO J. 34, 759–777 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Alexanian, M. et al. A transcribed enhancer dictates mesendoderm specification in pluripotency. Nat. Commun. 8, 1806 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  64. Yu, P. et al. Spatiotemporal clustering of the epigenome reveals rules of dynamic gene regulation. Genome Res. 23, 352–364 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Lodato, M. A. et al. SOX2 co-occupies distal enhancer elements with distinct POU factors in ESCs and NPCs to specify cell state. PLoS Genet. 9, e1003288 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank A. Smith for providing the Rex1GFPd2 cells, K. Nimura for the Whsc1 ΔSET cells, E. J. Robertson for the Eomes-GFP cells, A. Surani for the XGFP EpiSCs, J. E. Bradner for the JQ1 compound, L. Batlle and C. Berenguer for technical help, the CRG core facilities of flow cytometry, advanced light microscopy and tissue engineering, and L. Di Croce, B. Payer, M. Behringer, J. Licht, K. M. Loh, K. Kaji and R. Stadhouders for advice and discussions. T.V.T. and J.L.S. were supported by Juan de la Cierva postdoctoral fellowships (MINECO; FJCI-2014-22946 and IJCI-2014-21872), B.D.S. by an EMBO long-term fellowship (number ALTF 1143-2015), G.S. by a Marie Sklodowska-Curie fellowship (H2020-MSCA-IF-2016, miRStem), A.D. by a Severo Ochoa fellowship and J.G. by a Boehringer Ingelheim Graduate Student Fellowship. R.J. was supported by NIH grants R01 NS088538-01 and 2R01MH104610-15. This work was supported by the EU-FP7 project BLUEPRINT, the Spanish Ministry of Economy, Industry and Competitiveness to the EMBL partnership, Centro de Excelencia Severo Ochoa 2013–2017 and the CERCA Program Generalitat de Catalunya. T.V.T., J.L.S. and B.D.S. were supported by a CRG award for junior collaborative projects.

Author information

Authors and Affiliations

Authors

Contributions

T.V.T. and T.G. conceived the project, designed the experimental work and wrote the manuscript. T.V.T., B.D.S., G.S., A.D., J.L.S., C.S.M., L.D.A.A. and J.G. performed the experiments. R.J. provided the reagents. E.V., M.V.-C. and A.G. conducted the bioinformatics analysis.

Corresponding authors

Correspondence to Tian V. Tian or Thomas Graf.

Ethics declarations

Competing interests

R.J. is an advisor/co-founder of Fate Therapeutics, Fulcrum Therapeutics, Omega Therapeutics and Dewpoint Therapeutics.

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 An in silico screen reveals a role of Whsc1 in the exit from pluripotency of ESCs.

a, Heatmap showing gene expression dynamics of Chromatin-related Factors (CRFs) during embryoid body (EB) differentiation of R1 (GSE2972) (left panel) and J1 ESCs (GSE3749) (right panel) (BMC Genomics 8, 85, 2007). Selected candidate genes, Cbx4, L3mbtl3 and Whsc1, are indicated. b, Expression of pluripotency, mesendoderm and neuroectoderm markers during exit from pluripotency induced by N2B27, Activin A and FBS. Relative expression (fold changes relative to 0 hr) is indicated as numbers which represent mean from n=3 independent experiments. c, Left panel: Evaluation of knockdown efficiency by RT-qPCR after shRNA transductions in Rex1GFPd2 cells. Data represent mean±s.d. from n=3 independent experiments and p-values were calculated by two-tailed unpaired t-test. Right panel: Flow cytometry analysis of GFP (Rex1) expression in Rex1GFPd2 cells depleted of Cbx4 or L3mbtl3 at 0 hr and 72 hrs after induction. Three independent experiments were performed with similar results. d, Evaluation of serial induction of exit from pluripotency. Left-top panel: schematic representation of experimental procedure; left-bottom panel: quantification of AP+ colonies. Right panel: representative images of control (shScr) and Whsc1 depleted Rex1GFPd2 cells subjected to two rounds of exit from pluripotency. Differentiated cells (cells lost GFP/Rex1 expression) were eliminated by Blasticidin and pluripotent cells were stained by AP. Scale bar: 500 μm. Two independent experiments were performed with similar results. e, Cell proliferation assay of control (shScr) and Whsc1 depleted (shWhsc1.2 and shWhsc1.4) Rex1GFPd2 cells. Data represent mean±s.d. from n=3 independent experiments and p-values were calculated by two-tailed unpaired t-test. f, Evaluation of pluripotency gene expression in control (shScr) and Whsc1 depleted (shWhsc1.2 and shWhsc1.4) Rex1GFPd2 cells by RT-qPCR. Data represent mean±s.d. from n=3 independent experiments. g, Western blot analysis of pluripotency factors, Nanog, Sox2 and Oct4 in control (shScr) and Whsc1-depleted (shWhsc1.2 and shWhsc1.4) Rex1GFPd2 cells. Tubulin was used as loading control. Two independent experiments were performed with similar results. Scanned images of unprocessed blots are shown in Supplementary Fig. 8.

Supplementary Figure 2 Whsc1 depletion results in reduced mesendoderm differentiation in EpiSCs and in teratomas in vivo.

a, Expression of Whsc1, mesendoderm and ectoderm markers and Pou5f1 was quantified by RT-qPCR in control (shScr) and Whsc1 depleted (shWhsc1.2 and shWhsc1.4) EBs generated from EpiSCs (XGFP EpiSCs). Data represent mean±s.d. from n=3 independent experiments and p-values were calculated by two-tailed unpaired t-test. b, HE staining of teratomas from control (shScr) and Whsc1 depleted (Whsc1.2 and Whsc1.4) cells. Endodermal (gut), mesodermal (cartilage) and ectodermal (neuroepithelia) were indicated on the figures. Scale Bar: 200 μm. Ten teratomas from each group were examined, and representative images are shown. c, Expression of Whsc1, mesendoderm, ectoderm and pluripotency makers was quantified by RT-qPCR in control (shScr) and Whsc1 depleted teratomas (shWhsc1.2 and shWhsc1.4). Data represent mean±s.d. from n=3 independent experiments and p-values were calculated by two-tailed unpaired t-test.

Supplementary Figure 3 Whsc1 is required for efficient generation of cardiac progenitors and definitive endoderm but not of neural progenitor cells.

a, Percentage of EBs with beating cells from control and Whsc1-depleted ESCs counted from Day 6 to 10 after induction of differentiation. Data represent mean from n=2 independent experiments with similar results. b-c, RT-qPCR quantification of the expression of cardiac progenitor and pluripotency markers 10 days after induction of differentiation. Data represent mean±s.d. from n=3 independent experiments and p-values were calculated by two-tailed unpaired t-test. d-e, Definitive endoderm differentiation using Eomes-GFP ESCs. (d) FACS profiles of ESCs 0 and 7 days after induction of differentiation and (e) quantification of Eomes-GFP positive cells in control and Whsc1-depleted cells 7 days after induction of differentiation. Data represent mean±s.d. from n=3 independent experiments and p-values were calculated by two-tailed unpaired t-test. f-g, Expression levels by RT-qPCR of definitive endoderm (f) and pluripotency markers (g) during induction of definitive endoderm in shScr and shWhsc1.4 cells. Data represent mean±s.d. from n=3 independent experiments and p-values were calculated by two-tailed unpaired t-test. h, Expression levels by RT-qPCR of neural markers at several time points after neural progenitor induction and quantification of pluripotency gene expression by RT-qPCR 8 days after induction of differentiation. Data represent mean±s.d. from n=3 independent experiments and p-values were calculated by two-tailed unpaired t-test.

Supplementary Figure 4 The SET-domain is dispensable for the functions of Whsc1 in pluripotency exit and mesendodermal differentiation.

a, Schematics of wild type (WT) and the SET-domain deleted (ΔSET) Whsc1 proteins (ΔSET). b, Expression levels by RT-qPCR of Whsc1 and pluripotency genes in WT and ΔSET cells, as well as in ΔSET cells expressing shScr or shWhsc1.4 constructs at 48 and 72hrs after induction of exit from pluripotency by transferring cells into N2B27 medium containing Activin A and FBS. Data represent mean±s.d. from n=3 independent experiments. c, Expression levels by RT-qPCR of Whsc1, mesendoderm, ectoderm markers and pluripotency markers in Day 6 EBs derived from WT and ΔSET ESCs, as well as in ΔSET cells with shScr or shWhsc1.4 constructs. Data represent mean±s.d. from n=6 independent experiments. d, Western blots of Whsc1 expression in WT and ΔSET cells, as well as ΔSET cells with control or Whsc1 shRNA (shWhsc1.4). Three independent experiments were performed with similar results. Scanned images of unprocessed blots are shown in Supplementary Fig. 8. e, Western blots of H3K36me2, H3K36me3 and H3K27me3 levels in Whsc1 +/+ and Whsc1 −/− ESCs. Total H3 was used as loading control. Three independent experiments were performed with similar results. Scanned images of unprocessed blots are shown in Supplementary Fig. 8.

Supplementary Figure 5 Whsc1 binds specifically to the enhancers of mesendoderm TFs.

a, Top panel: HiC contact maps at 10kb resolution from ESCs around Foxa2, Gata4, Sox1 and Pax6 loci (GSE96611). TADs are marked by solid black lines and subTAD are indicated by dashed black lines. The regions analysed using H3K27ac ChIP-seq data are indicated as black boxes; Bottom panel: H3K27ac ChIP-seq profiles on the loci of Foxa2, Gata4, Sox1 and Pax6 loci in neural progenitor cells (NPC) (GSE35496), Eomes+ mesendodermal progenitors (GSE103262), Activin A-induced mesendodermal precursors (MP) (GSE38596) and Flk1+ mesendodermal progenitors (GSE47082). The black arrows below each panel correspond to putative regulatory regions up or downstream of the respective gene. b, ChIP-qPCR quantification of Whsc1 occupancy on the same enhancers as shown in Fig. 5a (above) in D6 EBs from Whsc1 +/+ and Whsc1 −/− cells with numbers indicating distance to the TSS in kb. Data represent mean±s.d. from n=3 independent experiments and p-values were calculated by two-tailed unpaired t-test. c, Replicated UMI-4C profiles for baits located on the Foxa2 (left) and Gata4 (right) promoters assayed in Whsc1 +/+ and Whsc1 −/− ESCs and D6 EBs. Top panel, contact profiles generated from the average of two independent biological replicates; bottom panel: average contact fold change D6 EBs versus D0 ESCs from two independent biological replicates. d, H3K27ac enrichments on putative regulatory regions of Foxa2, Gata4, Sox1 and Pax6 were quantified by ChIP-qPCR in Day 6 Whsc1 +/+ and Whsc1 −/− EBs. Data represent mean±s.d. from n=3 independent experiments and p-values were calculated by two-tailed unpaired t-test. e, ChIP-qPCR quantification of H3K4me2 occupancies on putative regulatory regions of T, Gata6, Foxa2, Gata4, Sox1 and Pax6 in D6 Whsc1 +/+ and Whsc1 −/− EBs. Data represent mean±s.d. from n=3 independent experiments and p-values were calculated by two-tailed unpaired t-test. f, ChIP-qPCR quantification of Whsc1 (left) and H3K27ac (right) occupancies on putative regulatory regions of T, Gata6, Gata4 and Foxa2 in Whsc1 +/+ and Whsc1 −/− ESCs. These regulatory regions were bound by Whsc1 in D6 with increase H3K27ac enrichments (Fig. 5b, d, Supplementary Fig. 5b and d). Data represent mean±s.d. from n=3 independent experiments.

Supplementary Figure 6 Mesendoderm TFs are required for efficient exit from pluripotency.

a, siRNA knockdown efficiency was evaluated by RT-qPCR in Rex1GFPd2 cells 72 hrs after induction of exit from pluripotency. Data represent mean±s.d. from n=3 independent experiments and p-values were calculated by two-tailed unpaired t-test. b, Kinetics of GFP expression determined by FACS in Rex1GFPd2 ESCs transfected with siRNAs against T, Gata6, Gata4, Foxa2 at different times after differentiation induction. Data represent mean±s.d. from n=3 independent experiments.

Supplementary Figure 7

Gating strategy for FACS analysis.

Supplementary Figure 8

Scans of unprocessed Western Blots.

Supplementary information

Supplementary Information

Supplementary Figs. 1–8, and titles and footnotes of Supplementary Tables 1–4.

Reporting Summary

Supplementary Table 1

List of CRFs

Supplementary Table 2

Differentially expressed genes in Whsc1−/− ESCs

Supplementary Table 3

Table of primers, antibodies, single guide RNAs and siRNAs

Supplementary Table 4

Statistical source data

Supplementary Video 1

Representative day 10 beating EB derived from shScrambled ESCs under cardiac differentiation conditions

Supplementary Video 2

Representative day 10 EB derived from shWhsc1.4 ESCs under cardiac differentiation conditions

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Tian, T.V., Di Stefano, B., Stik, G. et al. Whsc1 links pluripotency exit with mesendoderm specification. Nat Cell Biol 21, 824–834 (2019). https://doi.org/10.1038/s41556-019-0342-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41556-019-0342-1

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