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:

Transcriptional repression upon S phase entry protects genome integrity in pluripotent cells

Abstract

Coincident transcription and DNA replication causes replication stress and genome instability. Rapidly dividing mouse pluripotent stem cells are highly transcriptionally active and experience elevated replication stress, yet paradoxically maintain genome integrity. Here, we study FOXD3, a transcriptional repressor enriched in pluripotent stem cells, and show that its repression of transcription upon S phase entry is critical to minimizing replication stress and preserving genome integrity. Acutely deleting Foxd3 leads to immediate replication stress, G2/M phase arrest, genome instability and p53-dependent apoptosis. FOXD3 binds near highly transcribed genes during S phase entry, and its loss increases the expression of these genes. Transient inhibition of RNA polymerase II in S phase reduces observed replication stress and cell cycle defects. Loss of FOXD3-interacting histone deacetylases induces replication stress, while transient inhibition of histone acetylation opposes it. These results show how a transcriptional repressor can play a central role in maintaining genome integrity through the transient inhibition of transcription during S phase, enabling faithful DNA replication.

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: FOXD3 loss results in G2/M arrest before broad transcriptional changes and independent of p53.
Fig. 2: FOXD3 limits replication stress and 2C gene activation and protects genome integrity.
Fig. 3: FOXD3 preferentially binds chromatin near the highest transcribed genes of the OSN network at G1/S.
Fig. 4: FOXD3 phasically limits transcriptional activity of nearby genes.
Fig. 5: Transcriptional repression by FOXD3 precedes and is functionally linked to replication stress markers.
Fig. 6: Loss of FOXD3-interacting HDAC1/HDAC2 results in replication stress, which is opposed by inhibition of histone acetylation after FOXD3 loss.

Similar content being viewed by others

Data availability

The 4SU-seq, streptavidin pulldown sequencing, CUT&Tag and MULTI-seq data have been deposited in the Gene Expression Omnibus (GEO) database under accession number GSE183420. Mass spectrometry data have been deposited to the ProteomeXchange Consortium via the PRIDE repository under accession number PXD035284. The following GEO datasets were analyzed in this study: ATAC-seq (GSE93147) and ChIP–seq for OCT4, SOX2 and NANOG (GSE56312), H3K27ac (GSE93147), Pol II (GSE23943) and γH2AX (GSE69143). The University of Alabama at Birmingham Cancer data analysis portal is accessible through https://ualcan.path.uab.edu. The STRING database is accessible through https://string-db.org. Source data are provided with this paper.

Code availability

No custom code was developed in this study.

References

  1. Liu, L., Michowski, W., Kolodziejczyk, A. & Sicinski, P. The cell cycle in stem cell proliferation, pluripotency and differentiation. Nat. Cell Biol. 21, 1060–1067 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Ciemerych, M. A. & Sicinski, P. Cell cycle in mouse development. Oncogene 24, 2877–2898 (2005).

    Article  CAS  PubMed  Google Scholar 

  3. Lawson, K. A., Meneses, J. J. & Pedersen, R. A. Clonal analysis of epiblast fate during germ layer formation in the mouse embryo. Development 113, 891–911 (1991).

    Article  CAS  PubMed  Google Scholar 

  4. Snow, M. H. L. Gastrulation in the mouse: growth and regionalization of the epiblast. Development 42, 293–303 (1977).

    Article  Google Scholar 

  5. Efroni, S. et al. Global transcription in pluripotent embryonic stem cells. Cell Stem Cell 2, 437–447 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Percharde, M., Bulut-Karslioglu, A. & Ramalho-Santos, M. Hypertranscription in development, stem cells, and regeneration. Dev. Cell 40, 9–21 (2017).

    Article  CAS  PubMed  Google Scholar 

  7. Solter, D., Škreb, N. & Damjanov, I. Cell cycle analysis in the mouse egg-cylinder. Exp. Cell. Res. 64, 331–334 (1971).

    Article  CAS  PubMed  Google Scholar 

  8. Ahuja, A. K. et al. A short G1 phase imposes constitutive replication stress and fork remodelling in mouse embryonic stem cells. Nat. Commun. 7, 10660 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Liu, L. et al. G1 cyclins link proliferation, pluripotency and differentiation of embryonic stem cells. Nat. Cell Biol. 19, 177–188 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Hamperl, S., Bocek, M. J., Saldivar, J. C., Swigut, T. & Cimprich, K. A. Transcription–replication conflict orientation modulates R-loop levels and activates distinct DNA damage responses. Cell 170, 774–786 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. García-Muse, T. & Aguilera, A. Transcription–replication conflicts: how they occur and how they are resolved. Nat. Rev. Mol. Cell Biol. 17, 553–563 (2016).

    Article  PubMed  Google Scholar 

  12. Saldivar, J. C. et al. An intrinsic S/G2 checkpoint enforced by ATR. Science 361, 806–810 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Saldivar, J. C., Cortez, D. & Cimprich, K. A. The essential kinase ATR: ensuring faithful duplication of a challenging genome. Nat. Rev. Mol. Cell Biol. 18, 622–636 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Zeman, M. K. & Cimprich, K. A. Causes and consequences of replication stress. Nat. Cell Biol. 16, 2–9 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Banáth, J. P. et al. Explanation for excessive DNA single-strand breaks and endogenous repair foci in pluripotent mouse embryonic stem cells. Exp. Cell. Res. 315, 1505–1520 (2009).

    Article  PubMed  Google Scholar 

  16. Chuykin, I. A., Lianguzova, M. S., Pospelova, T. V. & Pospelov, V. A. Activation of DNA damage response signaling in mouse embryonic stem cells. Cell Cycle 7, 2922–2928 (2008).

    Article  CAS  PubMed  Google Scholar 

  17. Turinetto, V. et al. High basal γH2AX levels sustain self-renewal of mouse embryonic and induced pluripotent stem cells. Stem Cells 30, 1414–1423 (2012).

    Article  CAS  PubMed  Google Scholar 

  18. Ziegler-Birling, C., Helmrich, A., Tora, L. & Torres-Padilla, M.-E. Distribution of p53 binding protein 1 (53BP1) and phosphorylated H2A.X during mouse preimplantation development in the absence of DNA damage. Int. J. Dev. Biol. 53, 1003–1011 (2009).

    Article  CAS  PubMed  Google Scholar 

  19. Choi, E.-H., Yoon, S. & Kim, K. P. Combined ectopic expression of homologous recombination factors promotes embryonic stem cell differentiation. Mol. Ther. 26, 1154–1165 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Cervantes, R. B., Stringer, J. R., Shao, C., Tischfield, J. A. & Stambrook, P. J. Embryonic stem cells and somatic cells differ in mutation frequency and type. Proc. Natl Acad. Sci. USA 99, 3586–3590 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Sutton, J. et al. Genesis, a winged helix transcriptional repressor with expression restricted to embryonic stem cells. J. Biol. Chem. 271, 23126–23133 (1996).

    Article  CAS  PubMed  Google Scholar 

  22. Labosky, P. A. & Kaestner, K. H. The winged helix transcription factor Hfh2 is expressed in neural crest and spinal cord during mouse development. Mech. Dev. 76, 185–190 (1998).

    Article  CAS  PubMed  Google Scholar 

  23. Hromas, R. et al. Genesis, a winged helix transcriptional repressor, has embryonic expression limited to the neural crest, and stimulates proliferation in vitro in a neural development model. Cell Tissue Res. 297, 371–382 (1999).

    Article  CAS  PubMed  Google Scholar 

  24. Dottori, M., Gross, M. K., Labosky, P. & Goulding, M. The winged-helix transcription factor Foxd3 suppresses interneuron differentiation and promotes neural crest cell fate. Development 128, 4127–4138 (2001).

    Article  CAS  PubMed  Google Scholar 

  25. Pohl, B. S. & Knöchel, W. Overexpression of the transcriptional repressor FoxD3 prevents neural crest formation in Xenopus embryos. Mech. Dev. 103, 93–106 (2001).

    Article  CAS  PubMed  Google Scholar 

  26. Steiner, A. B. et al. FoxD3 regulation of Nodal in the Spemann organizer is essential for Xenopus dorsal mesoderm development. Development 133, 4827–4838 (2006).

    Article  CAS  PubMed  Google Scholar 

  27. Krishnakumar, R. et al. FOXD3 regulates pluripotent stem cell potential by simultaneously initiating and repressing enhancer activity. Cell Stem Cell 18, 104–117 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. 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 

  29. Hanna, L. A., Foreman, R. K., Tarasenko, I. A., Kessler, D. S. & Labosky, P. A. Requirement for Foxd3 in maintaining pluripotent cells of the early mouse embryo. Genes Dev. 16, 2650–2661 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Liu, Y. & Labosky, P. A. Regulation of embryonic stem cell self-renewal and pluripotency by Foxd3. Stem Cells 26, 2475–2484 (2008).

    Article  CAS  PubMed  Google Scholar 

  31. Pijuan-Sala, B. et al. A single-cell molecular map of mouse gastrulation and early organogenesis. Nature 566, 490–495 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Tabula Sapiens Consortium et al. The Tabula Sapiens: a multiple-organ, single-cell transcriptomic atlas of humans. Science 376, eabl4896 (2022).

    Article  Google Scholar 

  33. Paglini, M. G. & Rovasio, R. A. Cell cycle of neural crest cells in the early migratory stage in vivo. Cell Prolif. 27, 571–578 (1994).

    Article  Google Scholar 

  34. Katsuta, E. et al. H2AX mRNA expression reflects DNA repair, cell proliferation, metastasis, and worse survival in breast cancer. Am. J. Cancer Res. 12, 793–804 (2022).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Bonner, W. M. et al. γH2AX and cancer. Nat. Rev. Cancer 8, 957–967 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Niwa, H., Miyazaki, J. & Smith, A. G. Quantitative expression of Oct-3/4 defines differentiation, dedifferentiation or self-renewal of ES cells. Nat. Genet. 24, 372–376 (2000).

    Article  CAS  PubMed  Google Scholar 

  37. Masamsetti, V. P. et al. Replication stress induces mitotic death through parallel pathways regulated by WAPL and telomere deprotection. Nat. Commun. 10, 4224 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  38. Mankouri, H. W., Huttner, D. & Hickson, I. D. How unfinished business from S-phase affects mitosis and beyond. EMBO J. 32, 2661–2671 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Lukas, C. et al. 53BP1 nuclear bodies form around DNA lesions generated by mitotic transmission of chromosomes under replication stress. Nat. Cell Biol. 13, 243–253 (2011).

    Article  CAS  PubMed  Google Scholar 

  40. van der Heijden, T. et al. Real-time assembly and disassembly of human RAD51 filaments on individual DNA molecules. Nucleic Acids Res. 35, 5646–5657 (2007).

    Article  PubMed  PubMed Central  Google Scholar 

  41. Ma, C. J., Gibb, B., Kwon, Y., Sung, P. & Greene, E. C. Protein dynamics of human RPA and RAD51 on ssDNA during assembly and disassembly of the RAD51 filament. Nucleic Acids Res. 45, 749–761 (2017).

    Article  CAS  PubMed  Google Scholar 

  42. Despras, E., Daboussi, F., Hyrien, O., Marheineke, K. & Kannouche, P. L. ATR/Chk1 pathway is essential for resumption of DNA synthesis and cell survival in UV-irradiated XP variant cells. Hum. Mol. Genet. 19, 1690–1701 (2010).

    Article  CAS  PubMed  Google Scholar 

  43. Atashpaz, S. et al. ATR expands embryonic stem cell fate potential in response to replication stress. eLife 9, e54756 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Nakatani, T. et al. DNA replication fork speed underlies cell fate changes and promotes reprogramming. Nat. Genet. 54, 318–327 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Macfarlan, T. S. et al. Embryonic stem cell potency fluctuates with endogenous retrovirus activity. Nature 487, 57–63 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Choi, Y. J. et al. Deficiency of microRNA miR-34a expands cell fate potential in pluripotent stem cells. Science 355, eaag1927 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  47. Buecker, C. et al. Reorganization of enhancer patterns in transition from naive to primed pluripotency. Cell Stem Cell 14, 838–853 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Galonska, C., Ziller, M. J., Karnik, R. & Meissner, A. Ground state conditions induce rapid reorganization of core pluripotency factor binding before global epigenetic reprogramming. Cell Stem Cell 17, 462–470 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Flynn, R. A. et al. 7SK–BAF axis controls pervasive transcription at enhancers. Nat. Struct. Mol. Biol. 23, 231–238 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. McGinnis, C. S. et al. MULTI-seq: sample multiplexing for single-cell RNA sequencing using lipid-tagged indices. Nat. Methods 16, 619–626 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Clouaire, T. et al. Comprehensive mapping of histone modifications at DNA double-strand breaks deciphers repair pathway chromatin signatures. Mol. Cell 72, 250–262 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Kim, D. I. et al. An improved smaller biotin ligase for BioID proximity labeling. Mol. Biol. Cell 27, 1188–1196 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Cole, M. F., Johnstone, S. E., Newman, J. J., Kagey, M. H. & Young, R. A. Tcf3 is an integral component of the core regulatory circuitry of embryonic stem cells. Genes Dev. 22, 746–755 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. van den Berg, D. L. C. et al. An Oct4-centered protein interaction network in embryonic stem cells. Cell Stem Cell 6, 369–381 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  55. Guo, Y. et al. The embryonic stem cell transcription factors Oct-4 and FoxD3 interact to regulate endodermal-specific promoter expression. Proc. Natl Acad. Sci. USA 99, 3663–3667 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Chen, Y.-H. et al. Transcription shapes DNA replication initiation and termination in human cells. Nat. Struct. Mol. Biol. 26, 67–77 (2019).

    Article  CAS  PubMed  Google Scholar 

  57. Ma, Y., Kanakousaki, K. & Buttitta, L. How the cell cycle impacts chromatin architecture and influences cell fate. Front. Genet. 6, 19 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  58. Ginno, P. A., Burger, L., Seebacher, J., Iesmantavicius, V. & Schübeler, D. Cell cycle-resolved chromatin proteomics reveals the extent of mitotic preservation of the genomic regulatory landscape. Nat. Commun. 9, 4048 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  59. Tichy, E. D. et al. Mouse embryonic stem cells, but not somatic cells, predominantly use homologous recombination to repair double-strand DNA breaks. Stem Cells Dev. 19, 1699–1711 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Teng, L., Mundell, N. A., Frist, A. Y., Wang, Q. & Labosky, P. A. Requirement for Foxd3 in the maintenance of neural crest progenitors. Development 135, 1615–1624 (2008).

    Article  CAS  PubMed  Google Scholar 

  61. Montero-Balaguer, M. et al. The mother superior mutation ablates foxd3 activity in neural crest progenitor cells and depletes neural crest derivatives in zebrafish. Dev. Dyn. 235, 3199–3212 (2006).

    Article  CAS  PubMed  Google Scholar 

  62. Stewart, R. A. et al. Zebrafish foxd3 is selectively required for neural crest specification, migration and survival. Dev. Biol. 292, 174–188 (2006).

    Article  CAS  PubMed  Google Scholar 

  63. Lister, J. A. et al. Zebrafish Foxd3 is required for development of a subset of neural crest derivatives. Dev. Biol. 290, 92–104 (2006).

    Article  CAS  PubMed  Google Scholar 

  64. Xu, M. et al. FOXD3, frequently methylated in colorectal cancer, acts as a tumor suppressor and induces tumor cell apoptosis under ER stress via p53. Carcinogenesis 41, 1253–1262 (2020).

    Article  CAS  PubMed  Google Scholar 

  65. Li, D. et al. FOXD3 is a novel tumor suppressor that affects growth, invasion, metastasis and angiogenesis of neuroblastoma. Oncotarget 4, 2021–2044 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  66. Yan, J.-H., Zhao, C.-L., Ding, L.-B. & Zhou, X. FOXD3 suppresses tumor growth and angiogenesis in non-small cell lung cancer. Biochem. Biophys. Res. Commun. 466, 111–116 (2015).

    Article  CAS  PubMed  Google Scholar 

  67. Chen, A. F. et al. GRHL2-dependent enhancer switching maintains a pluripotent stem cell transcriptional subnetwork after exit from naive pluripotency. Cell Stem Cell 23, 226–238 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Marks, H. et al. The transcriptional and epigenomic foundations of ground state pluripotency. Cell 149, 590–604 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Dovey, O. M., Foster, C. T. & Cowley, S. M. Histone deacetylase 1 (HDAC1), but not HDAC2, controls embryonic stem cell differentiation. Proc. Natl Acad. Sci. USA 107, 8242–8247 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Beard, C., Hochedlinger, K., Plath, K., Wutz, A. & Jaenisch, R. Efficient method to generate single-copy transgenic mice by site-specific integration in embryonic stem cells. Genesis 44, 23–28 (2006).

    Article  CAS  PubMed  Google Scholar 

  71. Freimer, J. W., Hu, T. J. & Blelloch, R. Decoupling the impact of microRNAs on translational repression versus RNA degradation in embryonic stem cells. eLife 7, e38014 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  72. Kaya-Okur, H. S., Janssens, D. H., Henikoff, J. G., Ahmad, K. & Henikoff, S. Efficient low-cost chromatin profiling with CUT&Tag. Nat. Protoc. 15, 3264–3283 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Kaya-Okur, H. S. et al. CUT&Tag for efficient epigenomic profiling of small samples and single cells. Nat. Commun. 10, 1930 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  75. 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).

    Article  CAS  PubMed  Google Scholar 

  76. Robinson, M. D., McCarthy, D. J. & Smyth, G. K. edgeR: a Bioconductor package for differential expression analysis of digital gene expression data. Bioinformatics 26, 139–140 (2010).

    Article  CAS  PubMed  Google Scholar 

  77. 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  Google Scholar 

  78. Jin, Y., Tam, O. H., Paniagua, E. & Hammell, M. TEtranscripts: a package for including transposable elements in differential expression analysis of RNA-seq datasets. Bioinformatics 31, 3593–3599 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. 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  Google Scholar 

  80. Ewels, P. A. et al. The nf-core framework for community-curated bioinformatics pipelines. Nat. Biotechnol. 38, 276–278 (2020).

    Article  CAS  PubMed  Google Scholar 

  81. Liu, X. et al. In situ capture of chromatin interactions by biotinylated dCas9. Cell 170, 1028–1043 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. 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 

  83. Danecek, P. et al. Twelve years of SAMtools and BCFtools. Gigascience 10, giab008 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  PubMed Central  Google Scholar 

  85. Li, Q., Brown, J. B., Huang, H. & Bickel, P. J. Measuring reproducibility of high-throughput experiments. Ann. Appl. Stat. 5, 1752–1779 (2011).

    Article  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Heinz, S. et al. Simple combinations of lineage-determining transcription factors prime cis-regulatory elements required for macrophage and B cell identities. Mol. Cell 38, 576–589 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Pohl, A. & Beato, M. bwtool: a tool for bigWig files. Bioinformatics 30, 1618–1619 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Hao, Y. et al. Integrated analysis of multimodal single-cell data. Cell 184, 3573–3587 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Lun, A. T. L., McCarthy, D. J. & Marioni, J. C. A step-by-step workflow for low-level analysis of single-cell RNA-seq data with Bioconductor. F1000Res. 5, 2122 (2016).

    PubMed  PubMed Central  Google Scholar 

  91. Ge, S. X., Jung, D. & Yao, R. ShinyGO: a graphical gene-set enrichment tool for animals and plants. Bioinformatics 36, 2628–2629 (2020).

    Article  CAS  PubMed  Google Scholar 

  92. Chandrashekar, D. S. et al. UALCAN: a portal for facilitating tumor subgroup gene expression and survival analyses. Neoplasia 19, 649–658 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. 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 

  94. Mootha, V. K. et al. PGC-1α-responsive genes involved in oxidative phosphorylation are coordinately downregulated in human diabetes. Nat. Genet. 34, 267–273 (2003).

    Article  CAS  PubMed  Google Scholar 

  95. Subramanian, A. et al. Gene set enrichment analysis: a knowledge-based approach for interpreting genome-wide expression profiles. Proc. Natl Acad. Sci. USA 102, 15545–15550 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Szklarczyk, D. et al. STRING v11: protein–protein association networks with increased coverage, supporting functional discovery in genome-wide experimental datasets. Nucleic Acids Res. 47, D607–D613 (2019).

    Article  CAS  PubMed  Google Scholar 

  97. Huang, H. H. et al. Proteasome inhibitor-induced modulation reveals the spliceosome as a specific therapeutic vulnerability in multiple myeloma. Nat. Commun. 11, 1931 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Cox, J. & Mann, M. MaxQuant enables high peptide identification rates, individualized p.p.b.-range mass accuracies and proteome-wide protein quantification. Nat. Biotechnol. 26, 1367–1372 (2008).

    Article  CAS  PubMed  Google Scholar 

  99. Gierlinski, M., Gastaldello, F., Cole, C. & Barton, G. J. Proteus: an R package for downstream analysis of MaxQuant output. Preprint at bioRxiv https://doi.org/10.1101/416511 (2018).

  100. Bai, G. et al. HLTF promotes fork reversal, limiting replication stress resistance and preventing multiple mechanisms of unrestrained DNA synthesis. Mol. Cell 78, 1237–1251 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Fischer, M. Census and evaluation of p53 target genes. Oncogene 36, 3943–3956 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank K. Cimprich for helpful discussions and consultation. We thank M. T. McManus for providing access to equipment. We thank D. Lim, B. Schwer and B. Marsh for critical reading of the manuscript. Technical support by the University of California, San Francisco (UCSF), Parnassus Flow Cytometry Core, The PCAT at UCSF and the UCSF Center for Advanced Technology is greatly appreciated. This publication includes data generated at the University of California, San Diego, IGM Genomics Center using an Illumina NovaSeq 6000 that was purchased with funding from a National Institutes of Health SIG grant (S10 OD026929). This work was supported by the National Institute of General Medical Sciences of the National Institutes of Health grant R01GM125089 (to R.B.) and the German Research Foundation Research Fellowship 363120021 (to D.G.).

Author information

Authors and Affiliations

Authors

Contributions

Conceptualization: D.G. and R.B. Methodology: D.G. Investigation: D.G., K.L., R.M.B., J.B. and H.H. Visualization: D.G. Software: D.G. Formal analysis: D.G. Resources: A.P.W., D.J.L. and R.B. Data curation: D.G. Funding acquisition: D.G. and R.B. Project administration: R.B. Supervision: D.G. and R.B. Writing, original draft: D.G. Writing, review and editing: D.G. and R.B.

Corresponding author

Correspondence to Robert Blelloch.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Structural & Molecular Biology thanks David Gilbert, Duanqing Pei and Jianlong Wang for their contribution to the peer review of this work. Primary Handling Editors: Sara Osman, Tiago Faial and Carolina Perdigoto, in collaboration with the Nature Structural & Molecular Biology team. Peer reviewer reports are available.

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 Single cell sequencing data analysis of FOXD3 expressing cells in early mouse embryogenesis and adult human tissues.

a,b, UMAP plots of single cell transcriptomes at different days during prenatal mouse development31 showing identified cell populations (a) and Foxd3 mRNA level (b). Populations with high Foxd3 mRNA levels are highlighted in legend (cross). c,d, UMAP plots of single cell transcriptomes32 in adult human stromal (c) and immune tissues (d) showing identified cell populations. e,f, Overlay of FOXD3 mRNA expression on UMAP plots from adult human stromal (e) and immune tissues (f) shown in c and d, respectively.

Source data

Extended Data Fig. 2 Coexpression analysis of FOXD3 expressing cells in early mouse embryogenesis.

a, GO biological term analysis of transcripts co-expressed with Foxd3 (see Methods). b, Protein interaction network analysis using STRING database with transcripts co-expressed with Foxd3. c, Violin plots of H2afx mRNA levels in cells with >1 normalized count of Foxd3 mRNA (Foxd3+) or <1 normalized count of Foxd3 mRNA (Foxd3). Statistical test used: Wilcoxon rank sum test. ****P < 0.0001.

Source data

Extended Data Fig. 3 Loss of FOXD3 results in increased apoptosis preceded by cell cycle arrest in G2/M.

a, Model and kinetics of FOXD3 loss following treatment of Foxd3iKO ESCs with 4HT (T) versus EtOH (E) as measured by immunoblots. See quantifications of immunoblots in Fig. 1b. EtOH serves as solvent control. Triangles in model represent loxP sites. Approximate location of molecular weight marker shown. b, Flow cytometry measurement of fraction of cells positive for the apoptotic marker Annexin V after 14-, 24- and 48-hour treatment of Foxd3iKO cells with 4HT versus EtOH (n = 3 independent cultures per treatment and timepoint). c,d, Time course of DNA content histogram comparisons (c) and quantification of cell cycle distributions (d) in Foxd3iKO cells treated with 4HT or EtOH for 8, 14, 24 and 32 hours (n = 3 independent cultures per treatment and timepoint). e, Model (left) and kinetics of OCT4 loss following 16- and 24-hour treatment of Oct4tetOff ESCs with doxycycline (D) as measured by immunoblots (middle). Quantification of independent repeats relative to H2O (H) solvent control (right) (n = 2 independent cultures per treatment and timepoint). Approximate location of molecular weight marker shown. Mean and s.e.m. are shown (b,d,e). Statistical test used: Two-way ANOVA with Šidák correction (b,d,e). *P < 0.05,**P < 0.01,***P < 0.001,****P < 0.0001 (b,d,e).

Source data

Extended Data Fig. 4 Differential transcription measurements using 4SU-seq data at different timepoints following treatments of Foxd3iKO and Oct4tetOff cells.

a, b, MA plots of Foxd3iKO (a) and Oct4tetOff (b) at shown hours of treatment. Genes with FDR < 0.05 shown in blue. p53 target genes101 with FDR < 0.05 are shown in red. Names of consistently differentially expressed genes in a are shown. c, Number of differentially expressed (FDR < 0.05) genes shown for each treatment time and genotype. d, Top10 significant KEGG pathways for upregulated genes in Foxd3iKO at 24 hours of 4HT treatment. Statistical test used: Fisher’s exact test (d).

Source data

Extended Data Fig. 5 Analysis of role of p53 and ID2 in apoptosis and cell cycle phenotypes associated with FOXD3 loss.

a, Western analysis of p53 activation in 24-hour EtOH (E) and 4HT (T) treated Foxd3iKO cells. Blot was probed for p53, p53 (S15P) and GAPDH loading control. Two replicates from two independent cultures are shown. Approximate location of molecular weight marker shown. b, Western analysis of p53 loss in multiple candidate Trp53 CRISPR-Cas9 deletion clones in Foxd3iKO background (Foxd3iKO-Trp53KO). Blots were probed for p53 and GAPDH loading control. Clones 17 and 48 were chosen for further experiments (highlighted with triangle). Approximate location of molecular weight marker shown. Results were independently confirmed once. c, Annexin V and SYTOX staining flow cytometry quantification of two independent 48-hour 4HT versus EtOH treated Foxd3iKO-Trp53KO clones (n = 3 independent cultures per treatment of clone 48 and 4HT-treated clone 17, and n = 2 for EtOH-treated clone 17). d, Percentage of Foxd3iKO-Trp53KO cells in different cell cycle phases following 24 (n = 2 independent cultures per treatment) and 48 hours (n = 3 independent cultures per treatment) of 4HT versus EtOH treatment. Data for two independent clones shown. e, qRT-PCR analysis of two independent Id2 CRISPR-Cas9 induced deletion clones in a Foxd3iKO background (Foxd3iKO-Id2KO). Id2 mRNA was normalized on Gapdh (n = 2 independent cultures per condition). f, Annexin V and SYTOX staining flow cytometry quantifications of two independent Foxd3iKO-Id2KO clones following 24 and 48 hours of 4HT versus EtOH treatment (n = 3 independent cultures per treatment of clone 18 and n = 1 for clone 46 treatments). g, Quantification fraction of cells of two independent Foxd3iKO-Id2KO clones in each cell cycle phase as measured by DNA content following 48 hours of 4HT versus EtOH treatment (n = 3 independent cultures per treatment of clone 18 and n = 1 for clone 46 treatments). h, Quantifications of 53BP1 foci in high-content confocal microscopy images of Foxd3iKO-Trp53KO cells following 24 hours 4HT versus EtOH treatment (n = 4 independent cultures per treatment). Foxd3iKO data from Fig. 2a are shown in comparison. i, Quantifications of 53BP1 foci in high-content confocal microscopy images of Foxd3iKO cells following 24 hours 4HT versus EtOH treatment using different indicated culture conditions (n = 8 independent cultures per treatment and culture condition). Note Y axis is represented on log scale which visually results in increased resolution at low values and decreased resolution at high values (c,f). Data points represent independent cultures (c-i). Mean and s.e.m. are shown (c-i). Statistical tests used: Two-way ANOVA with Šidák correction (c,d,f), two-tailed Student’s t-test (d,g,i), one-way ANOVA with Tukey correction (e,h). *P < 0.05,**P < 0.01,***P < 0.001,****P < 0.0001 (c-i).

Source data

Extended Data Fig. 6 Loss of FOXD3 during embryogenesis results in increased DNA damage markers.

a, Experimental design of embryo aggregations. b, Whole mount stainings of non-aggregated embryos for trophectoderm marker CDX2 and inner cell mass marker OCT4. Results were independently confirmed once. c, Representative confocal microscopy images of whole mount embryo stainings for DAPI, 53BP1 and GFP. d, Quantifications of fraction of GFP+ cells with more than two large 53BP1 foci using high-content microscopy of whole mount embryo stainings previously with prior treatment with 4HT (n = 90 GFP+ cells from 20 embryos) versus EtOH (n = 62 GFP+ cells from 20 embryos). e, Representative confocal microscopy images of whole mount embryo stainings for DAPI, γH2AX and GFP. f, Quantifications of fraction of GFP+ cells with more than two large γH2AX foci using high-content microscopy of whole mount embryo stainings with prior treatment with 4HT (n = 271 GFP+ cells from 20 embryos) versus EtOH (n = 51 GFP+ cells from 20 embryos). Scalebar equals 25 µm (b), 10 µm (c,e). Means are shown (d,f). Statistical test used: Two-tailed Mann-Whitney test (d,f). ****P < 0.0001 (d,f).

Source data

Extended Data Fig. 7 Gene set enrichment analysis of 2C gene sets in differential nascent expression data upon FOXD3 loss.

a,b, Gene set enrichment plots for 2C gene sets45,46 (a) and (b) using differential nascent expression data from 24- and 32-hour 4HT versus EtOH treated Foxd3iKO cells. c, Significantly differential expressed repeat RNAs in 32-hour 4HT versus EtOH treated Foxd3iKO cells. Mean is shown (c). *FDR < 0.05 (c).

Source data

Extended Data Fig. 8 Phasic FOXD3 SAseq quality control and peak analysis.

a, Scheme of endogenous and exogenous Avi-HA tag modified FOXD3 (see Methods). b, Immunoblot validation of exogenous (exo) and endogenous (endo) FOXD3-Avi ESC lines probed with anti-FOXD3, anti-HA and anti-GAPDH antibodies. Doxycycline (Dox) concentrations (ng/ml) shown. Approximate location of molecular weight marker shown. Results were independently confirmed at least twice. c,d, Experimental design (c) and flow cytometry cell percentages per cell cycle phase (d) of SA-seq experiment in endogenously FOXD3 Avi-Tag expressing ESCs synchronized and released from G2/M using CDK1i. Triangles and squares indicate samples used for sequencing and are shown relative to WT ESCs (circles). e, HOMER motif enrichment analysis ±50 bp from the center of FOXD3 peaks. Resulting P values for top 10 motifs shown. f,g, Genome tracks at Fhit (f) and Ddit4 (g) loci of endogenous (FOXD3endo) and exogenous (FOXD3exo) FOXD3-AviTag SA-seq data from synchronized and asynchronous FOXD3 Avi-Tag expressing ESCs, and ChIP-seq data for OCT4, SOX2, NANOG and H3K27ac plotted as CPM. h, HOMER motif enrichment analysis ±200 bp from the center of FOXD3 peaks. Resulting P values for top 10 motifs shown. i, Box plots of 4SU-seq read counts (CPM) at genes bodies nearby all FOXD3 sites called with IDR stringency thresholds of 0.1 (N = 1092 sites), 0.05 (N = 869 sites), 0.01 (N = 559 sites), 0.001 (N = 328 sites) and 0.0001 (N = 328 sites) (see Methods). Box plots show median center, 25th and 75th percentile box bounds, and 1.5× interquartile range whisker limits.

Source data

Extended Data Fig. 9 Expanded FOXD3 binding and transcriptional regulation analysis.

a, Box plots of 4SU-seq read counts (CPM) at all OSN (N = 10397 sites), FOXD3 (N = 1110 sites) or all other ATAC-accessible sites (N = 116976 sites). Box plots show median center, 25th and 75th percentile box bounds, and 1.5× interquartile range whisker limits. b, qRT-PCR for major satellite repeats and Foxd3 mRNA in Foxd3iKO at 24 hours of 4HT versus EtOH treatment (n = 3 independent cultures per treatment and target). Data normalized on Gapdh mRNA. c, Metagene plots (top) and heatmap plots (bottom) of SA-seq data at all ATAC-accessible sites (N = 127719) for endogenous (FOXD3endo) and exogenous (FOXD3exo) FOXD3-AviTag SA-seq data from synchronized and asynchronous FOXD3 Avi-Tag expressing ESCs, no biotin FOXD3 control and ChIP-seq data for OCT4, SOX2 and NANOG plotted as CPM. Mean and s.e.m. shown (b). Statistical tests used: Wilcoxon rank sum test with Benjamini-Hochberg correction (a), two-tailed Student’s t-test (b). *P < 0.05, ****P < 0.0001 (a,b).

Source data

Extended Data Fig. 10 Phasic sequencing data analysis and FOXD3 cancer meta-analysis.

a,b, Schematic (a) and results (b) of DNA content FACS analysis of Foxd3iKO cells used for 4SU-seq at different release times after double thymidine block and 24-hour 4HT vs EtOH treatment. c, Immunoblot for experiment shown in a and b probed for FOXD3 at 0 and 1-hour release from double thymidine block. See below quantification of FOXD3 protein level normalized on loading control (LC) – a non-specific lower kDa band consistently shown by FOXD3 antibody. Approximate location of molecular weight marker shown. Comparable results were independently confirmed once. d,e, Experimental design (d) and box plots (e) of log2 fold changes of CUT&Tag sequencing data for H3K4me3 at TSS nearby FOXD3 sites obtained from 24-hour 4HT versus EtOH treated Foxd3iKO cells sorted into G1/early S, mid S and late S/G2/M. Wedges represent quartiles with increasing nascent transcription rate based on 4SU-seq data (for FOXD3 nearby TSS: Nq1 = 359, Nq2 = 360, Nq3 = 352, Nq4 = 337 and for other TSS: Nq1 = 26822, Nq2 = 26820, Nq3 = 26828, Nq4 = 26844). f,g, Schematic (f) and results (g) of DNA content FACS analysis of double thymidine blocked cells treated for 14 hours with 4HT versus EtOH before being released for shown hours. h, Quantification of γH2AX foci from high content confocal microscopy data of 24-hour 4HT versus EtOH treatment with or without triptolide addition (n = 8 independent cultures for each treatment). i, Heatmap analysis of stage specific Foxd3 mRNA levels in cancer types available via UALCAN. Clustered by K-means. j, Box plot representation of stage dependent FOXD3 expression in uveal melanoma available via UALCAN. See full statistical test results in Supplementary Table 3 (e). Box plots show median center, 25th and 75th percentile box bounds, and 1.5× interquartile range whisker limits (e). Mean with s.e.m. are shown (h). Statistical tests used: Wilcoxon rank sum test with Benjamini-Hochberg correction (e), two-tailed Student’s t-test (h). *P < 0.05 (e).

Source data

Supplementary information

Reporting Summary

Peer Review File

Supplementary Tables 1–4

Supplementary Table 1. Oligonucleotides and synthesized sequences. Supplementary Table 2. Antibody list. Supplementary Table 3. Statistical test results for Fig. 4 and Extended Data Fig. 10c comparisons. Supplementary Table 4. Statistical test results for Fig. 5j comparisons.

Source data

Source Data Fig. 1

Source data of graphs.

Source Data Fig. 2

Source data of graphs.

Source Data Fig. 3

Source data of graphs.

Source Data Fig. 4

Source data of graphs.

Source Data Fig. 5

Source data of graphs.

Source Data Fig. 6

Source data of graphs.

Source Data Extended Data Fig. 1

Source data of graphs.

Source Data Extended Data Fig. 2

Source data of graphs.

Source Data Extended Data Fig. 3

Source data of graphs and uncropped immunoblots.

Source Data Extended Data Fig. 4

Source data of graphs.

Source Data Extended Data Fig. 5

Source data of graphs and uncropped immunoblots.

Source Data Extended Data Fig. 6

Source data of graphs.

Source Data Extended Data Fig. 7

Source data of graphs.

Source Data Extended Data Fig. 8

Source data of graphs and uncropped immunoblots.

Source Data Extended Data Fig. 9

Source data of graphs.

Source Data Extended Data Fig. 10

Source data of graphs and uncropped immunoblots.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Gökbuget, D., Lenshoek, K., Boileau, R.M. et al. Transcriptional repression upon S phase entry protects genome integrity in pluripotent cells. Nat Struct Mol Biol 30, 1561–1570 (2023). https://doi.org/10.1038/s41594-023-01092-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41594-023-01092-7

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