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:

The histone chaperone CAF-1 safeguards somatic cell identity

Abstract

Cellular differentiation involves profound remodelling of chromatic landscapes, yet the mechanisms by which somatic cell identity is subsequently maintained remain incompletely understood. To further elucidate regulatory pathways that safeguard the somatic state, we performed two comprehensive RNA interference (RNAi) screens targeting chromatin factors during transcription-factor-mediated reprogramming of mouse fibroblasts to induced pluripotent stem cells (iPS cells). Subunits of the chromatin assembly factor-1 (CAF-1) complex, including Chaf1a and Chaf1b, emerged as the most prominent hits from both screens, followed by modulators of lysine sumoylation and heterochromatin maintenance. Optimal modulation of both CAF-1 and transcription factor levels increased reprogramming efficiency by several orders of magnitude and facilitated iPS cell formation in as little as 4 days. Mechanistically, CAF-1 suppression led to a more accessible chromatin structure at enhancer elements early during reprogramming. These changes were accompanied by a decrease in somatic heterochromatin domains, increased binding of Sox2 to pluripotency-specific targets and activation of associated genes. Notably, suppression of CAF-1 also enhanced the direct conversion of B cells into macrophages and fibroblasts into neurons. Together, our findings reveal the histone chaperone CAF-1 to be a novel regulator of somatic cell identity during transcription-factor-induced cell-fate transitions and provide a potential strategy to modulate cellular plasticity in a regenerative setting.

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

Figure 1: Arrayed and multiplexed shRNAmiR screening strategies to identify suppressors of reprogramming.
Figure 2: CAF-1 suppression accelerates reprogramming and yields developmentally competent iPS cells.
Figure 3: Enhanced reprogramming depends on optimal CAF-1 and OKSM dosage.
Figure 4: CAF-1 suppression enhances reprogramming in different cell conversion systems.
Figure 5: CAF-1 suppression affects chromatin dynamics and facilitates activation of pluripotency genes during iPS cell generation.

Similar content being viewed by others

Accession codes

Primary accessions

Gene Expression Omnibus

Data deposits

All SONO-seq, ATAC-seq, ChIP-seq, RNA-seq and microarray data have been deposited in the Gene Expression Omnibus database under accession number GSE66534.

References

  1. Lee, T. I. & Young, R. A. Transcriptional regulation and its misregulation in disease. Cell 152, 1237–1251 (2013)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Takahashi, K. & Yamanaka, S. Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell 126, 663–676 (2006)

    Article  CAS  PubMed  Google Scholar 

  3. Vierbuchen, T. & Wernig, M. Molecular roadblocks for cellular reprogramming. Mol. Cell 47, 827–838 (2012)

    Article  CAS  PubMed  Google Scholar 

  4. Yang, C. S., Chang, K. Y. & Rana, T. M. Genome-wide functional analysis reveals factors needed at the transition steps of induced reprogramming. Cell Rep. 8, 327–337 (2014)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Mikkelsen, T. S. et al. Dissecting direct reprogramming through integrative genomic analysis. Nature 454, 49–55 10.1038/nature07056 (2008)

    Article  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  6. Onder, T. T. et al. Chromatin-modifying enzymes as modulators of reprogramming. Nature 483, 598–602 (2012)

    Article  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  7. Rais, Y. et al. Deterministic direct reprogramming of somatic cells to pluripotency. Nature 502, 65–70 (2013)

    Article  CAS  PubMed  ADS  Google Scholar 

  8. Dejosez, M., Ura, H., Brandt, V. L. & Zwaka, T. P. Safeguards for cell cooperation in mouse embryogenesis shown by genome-wide cheater screen. Science 341, 1511–1514 (2013)

    Article  CAS  PubMed  ADS  Google Scholar 

  9. Qin, H. et al. Systematic identification of barriers to human iPSC generation. Cell 158, 449–461 (2014)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Stadtfeld, M., Maherali, N., Borkent, M. & Hochedlinger, K. A reprogrammable mouse strain from gene-targeted embryonic stem cells. Nature Methods 7, 53–55 (2010)

    Article  CAS  PubMed  Google Scholar 

  11. Zuber, J. et al. RNAi screen identifies Brd4 as a therapeutic target in acute myeloid leukaemia. Nature 478, 524–528 (2011)

    Article  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  12. Fellmann, C. et al. An optimized microRNA backbone for effective single-copy RNAi. Cell Rep. 5, 1704–1713 (2013)

    Article  CAS  PubMed  Google Scholar 

  13. Smith, S. & Stillman, B. Purification and characterization of CAF-I, a human cell factor required for chromatin assembly during DNA replication in vitro. Cell 58, 15–25 (1989)

    Article  CAS  PubMed  Google Scholar 

  14. Chen, J. et al. H3K9 methylation is a barrier during somatic cell reprogramming into iPSCs. Nature Genet. 45, 34–42 (2013)

    Article  CAS  PubMed  Google Scholar 

  15. Soufi, A., Donahue, G. & Zaret, K. S. Facilitators and impediments of the pluripotency reprogramming factors’ initial engagement with the genome. Cell 151, 994–1004 (2012)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Sridharan, R. et al. Proteomic and genomic approaches reveal critical functions of H3K9 methylation and heterochromatin protein-1γ in reprogramming to pluripotency. Nature Cell Biol. 15, 872–882 (2013)

    Article  CAS  PubMed  Google Scholar 

  17. Krizhanovsky, V. & Lowe, S. W. Stem cells: the promises and perils of p53. Nature 460, 1085–1086 (2009)

    Article  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  18. Polo, J. M. et al. A molecular roadmap of reprogramming somatic cells into iPS cells. Cell 151, 1617–1632 (2012)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Quivy, J. P., Gerard, A., Cook, A. J., Roche, D. & Almouzni, G. The HP1-p150/CAF-1 interaction is required for pericentric heterochromatin replication and S-phase progression in mouse cells. Nature Struct. Mol. Biol. 15, 972–979 (2008)

    Article  CAS  Google Scholar 

  20. Hoek, M. & Stillman, B. Chromatin assembly factor 1 is essential and couples chromatin assembly to DNA replication in vivo. Proc. Natl Acad. Sci. USA 100, 12183–12188 (2003)

    Article  CAS  PubMed  ADS  PubMed Central  Google Scholar 

  21. Houlard, M. et al. CAF-1 is essential for heterochromatin organization in pluripotent embryonic cells. PLoS Genet. 2, e181 (2006)

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  22. Ye, X. et al. Defective S phase chromatin assembly causes DNA damage, activation of the S phase checkpoint, and S phase arrest. Mol. Cell 11, 341–351 (2003)

    Article  CAS  PubMed  Google Scholar 

  23. Rolef Ben-Shahar, T. et al. Two fundamentally distinct PCNA interaction peptides contribute to chromatin assembly factor 1 function. Mol. Cell. Biol. 29, 6353–6365 (2009)

    Article  PubMed  Google Scholar 

  24. Chanda, S. et al. Generation of induced neuronal cells by the single reprogramming factor ASCL1. Stem Cell Rep. 3, 282–296 (2014)

    Article  CAS  Google Scholar 

  25. Bussmann, L. H. et al. A robust and highly efficient immune cell reprogramming system. Cell Stem Cell 5, 554–566 (2009)

    Article  CAS  PubMed  Google Scholar 

  26. Auerbach, R. K. et al. Mapping accessible chromatin regions using Sono-seq. Proc. Natl Acad. Sci. USA 106, 14926–14931 (2009)

    Article  CAS  PubMed  ADS  PubMed Central  Google Scholar 

  27. Buenrostro, J. D., Giresi, P. G., Zaba, L. C., Chang, H. Y. & Greenleaf, W. J. Transposition of native chromatin for fast and sensitive epigenomic profiling of open chromatin, DNA-binding proteins and nucleosome position. Nature Methods 10, 1213–1218 (2013)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Huang, H. et al. Drosophila CAF-1 regulates HP1-mediated epigenetic silencing and pericentric heterochromatin stability. J. Cell Sci. 123, 2853–2861 (2010)

    Article  CAS  PubMed  Google Scholar 

  29. Matoba, S. et al. Embryonic development following somatic cell nuclear transfer impeded by persisting histone methylation. Cell 159, 884–895 (2014)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Hathaway, N. A. et al. Dynamics and memory of heterochromatin in living cells. Cell 149, 1447–1460 (2012)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Ishiuchi, T. et al. Early embryonic-like cells are induced by downregulating replication-dependent chromatin assembly. Nature Struct. Mol. Biol. (2015)

  32. Ray-Gallet, D. et al. Dynamics of histone H3 deposition in vivo reveal a nucleosome gap-filling mechanism for H3.3 to maintain chromatin integrity. Mol. Cell 44, 928–941 (2011)

    Article  CAS  PubMed  Google Scholar 

  33. Jullien, J. et al. HIRA dependent H3.3 deposition is required for transcriptional reprogramming following nuclear transfer to Xenopus oocytes. Epigenetics Chromatin 5, 17 (2012)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Wen, D., Banaszynski, L. A., Rosenwaks, Z., Allis, C. D. & Rafii, S. H3.3 replacement facilitates epigenetic reprogramming of donor nuclei in somatic cell nuclear transfer embryos. Nucleus 5, 369–375 (2014)

    Article  PubMed  Google Scholar 

  35. Premsrirut, P. K. et al. A rapid and scalable system for studying gene function in mice using conditional RNA interference. Cell 145, 145–158 (2011)

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Zuber, J. et al. Toolkit for evaluating genes required for proliferation and survival using tetracycline-regulated RNAi. Nature Biotechnol. 29, 79–83 (2011)

    Article  CAS  Google Scholar 

  37. Lengner, C. J. et al. Oct4 expression is not required for mouse somatic stem cell self-renewal. Cell Stem Cell 1, 403–415 (2007)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Bernstein, B. E. et al. Genomic maps and comparative analysis of histone modifications in human and mouse. Cell 120, 169–181 (2005)

    Article  CAS  PubMed  Google Scholar 

  39. Langmead, B., Trapnell, C., Pop, M. & Salzberg, S. L. Ultrafast and memory-efficient alignment of short DNA sequences to the human genome. Genome Biol. 10, R25 (2009)

    PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  Google Scholar 

  41. Shen, Y. et al. A map of the cis-regulatory sequences in the mouse genome. Nature 488, 116–120 (2012)

    CAS  PubMed  PubMed Central  ADS  Google Scholar 

  42. 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 

  43. Whyte, W. A. et al. Master transcription factors and mediator establish super-enhancers at key cell identity genes. Cell 153, 307–319 (2013)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Sabo, P. J. et al. Discovery of functional noncoding elements by digital analysis of chromatin structure. Proc. Natl Acad. Sci. USA 101, 16837–16842 (2004)

    Article  CAS  PubMed  ADS  PubMed Central  Google Scholar 

  45. Ross-Innes, C. S. et al. Differential oestrogen receptor binding is associated with clinical outcome in breast cancer. Nature 481, 389–393 10.1038/nature10730 (2012)

    Article  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  46. 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 

  47. Marson, A. et al. Connecting microRNA genes to the core transcriptional regulatory circuitry of embryonic stem cells. Cell 134, 521–533 (2008)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. 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 

  49. Pezic, D., Manakov, S. A., Sachidanandam, R. & Aravin, A. A. piRNA pathway targets active LINE1 elements to establish the repressive H3K9me3 mark in germ cells. Genes Dev. 28, 1410–1428 (2014)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Smyth, G. K. Linear models and empirical Bayes methods for assessing differential expression in microarray experiments. Stat. Appl. Genet. Mol. Biol. 3, 1–25 (2004)

    Article  MathSciNet  MATH  ADS  Google Scholar 

  51. 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  ADS  Google Scholar 

  52. Dowen, J. M. et al. Control of cell identity genes occurs in insulated neighborhoods in mammalian chromosomes. Cell 159, 374–387 (2014)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Davis, M. P., van Dongen, S., Abreu-Goodger, C., Bartonicek, N. & Enright, A. J. Kraken: a set of tools for quality control and analysis of high-throughput sequence data. Methods 63, 41–49 (2013)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Anders, S. & Huber, W. Differential expression analysis for sequence count data. Genome Biol. 11, R106 (2010)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank B. Kingston, C. Vakoc, M. Tolstorukov and G. Hannon for guidance and discussions, B. Bernstein, K. Plath, K. Chronis, Y. Shen and O. Tam for advice on the ATAC-seq analysis, P. Brown for providing the Dot1l inhibitor, B. Stillman for sharing the Chaf1b antibody and T. Graf for sharing the C10 cell line. We thank C. Nakada and Y. Kiyota (Nikon) for providing software to quantify iPS cell formation and A. Huebner for help with transdifferentiation experiments. We are grateful to H. Hock and the HSCI-CRM flow cytometry core for help with flow data analysis and to W. Mallard for initial RNA-sequencing analysis. We further thank B. Ma, S. Muller, M. Weissenboeck and the IMP/IMBA Biooptics and Transgenic core facility as well the CSF NGS laboratory for technical assistance and all members of the Hochedlinger, Zuber, Penninger, Elling, Shi and Kingston laboratories for their feedback on various aspects of this project. We thank A. Stark, A. Deaton and L. Barrett for critical reading of the manuscript. S.C. was supported by the PRCRP at the Department of Defense (CA 120212). H.Y.C. by was supported by NIH P50-HG007735. U.E. was funded by grants from IMBA and the Austrian National Foundation. S.W.L. was supported by a cancer center support grant and program project grant from the NCI and is an HHMI investigator. J.M.P. was supported by IMBA, ERC GA (number 341036) and the Innovator Award/Era of Hope Award Number W81XWH-12-1-0093. J.Z. was funded by an ERC starting grant (number 336860) and generous institutional funding from Boehringer Ingelheim. K.H. was supported by funds from the MGH, HHMI, NIH (R01 HD058013-06) and the Gerald and Darlene Jordan Chair in Regenerative Medicine.

Author information

Authors and Affiliations

Authors

Contributions

Y. L. Jung and B. Hopfgartner contributed equally to this work. S.C., K.H., U.E. and J.Z. designed primary screens, analysed and interpreted data. S.C., J.M. and N.A. performed the arrayed screen and S.C. conducted follow-up cell biology and chromatin studies. U.E. and B.H. performed the multiplexed screen. U.E. performed validation experiments, genetic interaction assays and cell biology experiments with support from B.H., M.H. and D.W. Human reprogramming experiments were performed by S.C. and J.B.; N.T. and S.W.L. assisted in the generation of inducible Col1a1::tetOP-Chaf1a shRNA cell lines. S.C., A.I.B., A.B. and Y.S. performed B-cell to macrophage conversion experiments. C.E.A. and M.W. conducted MEF to induced neuron transdifferentiation experiments. Y.L.J., M.N., A.A., F.F. and P.J.P. performed bioinformatics analyses. M.H. and U.E. conducted the CiA assay with support from O.B. D.J.W. assisted with the SONO-seq experiments and H.Y.C. helped with the ATAC-seq assay. J.M., M.H. and M.Z. assisted with western blot and chromatin studies. D.T. and J.R. conducted ChIP experiments and library construction. M.S. and S.E.V. provided secondary Oct4–tdTomato MEFs. J.Z. and S.W.L. provided the arrayed library. J.Z. and P.R. designed the extended chromatin library. M.F., J.J. and B.H. generated lentiviral vectors and RNAi reagents. J.M.P. and G.A. provided intellectual support and mentoring. K.H., S.C., J.Z. and U.E. wrote the paper with input from all co-authors.

Corresponding authors

Correspondence to Johannes Zuber or Konrad Hochedlinger.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Extended data figures and tables

Extended Data Figure 1 Validation of hits from chromatin-focused shRNA screens.

a, Quantitative RT–PCR analysis to confirm suppression of Chaf1a and Chaf1b expression with miR-30-based vectors from arrayed screen. Sh Chaf1a pool, sh Chaf1b pool and sh CAF-1 pool denote pools of shRNAs targeting either Chaf1a, Chaf1b or both. b, Western blot analysis to confirm knockdown of CAF-1 components using the top-scoring miR-30-based shRNAs from arrayed screen (see Supplementary Fig. 1 for full scans). c, Quantification of data shown in Fig. 1f. d, Quantitative RT–PCR analysis confirming knockdown with top-scoring miR-E-based shRNAmiRs targeting Chaf1a, Chaf1b or Ube2i from the multiplexed screen. Error bars show s.d. from biological triplicates. RNA and protein were extracted from reprogrammable MEFs 72 h after doxycycline induction in panels ad. e, Suppression of CAF-1 components, Ube2i and Setdb2 enhances reprogramming in the presence or absence of ascorbic acid (AA) as well as in serum replacement media containing LIF (SR-LIF). Oct4–GFP+ cells were scored by flow cytometry on day 11 after 7 days of OKSM induction and 4 days of transgene-independent growth. Error bars show s.d. from biological triplicates. f, Number of doxycycline-independent, alkaline phosphatase (AP)-positive colonies emerging two weeks after plating 10,000 reprogrammable MEFs carrying shRNA vectors against indicated targets and cultured in serum replacement media containing 2i (SR-2i), n = 1 experiment. g, Effect of suppressing SUMO E2 ligase Ube2i, E1 ligases Sae1 and Uba2 on iPS cell formation. Shown is fraction of Oct4–GFP+ cells at day 11 (7 days of OKSM induction, 4 days of transgene-independent growth). Error bars depict s.d. from biological triplicates.

Extended Data Figure 2 Germline transmission of iPS cells, genetic interaction of shRNA hits and effect of CAF-1 or Ube2i suppression on reprogramming dynamics.

a, Germline transmission of agouti chimaeras generated from iPS cells using doxycycline-inducible shRNA vectors targeting Chaf1a, Chaf1b or Ube2i. Germline transmission was determined by scoring for agouti coat colour offspring upon breeding chimaeras with albino females. Germline transmission was observed in 8/8, 4/4 and 6/8 cases for Chaf1a iPS-cell-derived chimaeras, in 7/7, 4/4, 7/7 and 9/9 cases for Chaf1b iPS-cell-derived chimaeras, and in 5/5, 7/7 and 5/5 cases for Ube2i iPS-cell-derived chimaeras. b, Table summarizing effects of co-suppressing pairs of targets on emergence of Oct4–GFP+ cells, shown as the ratio of Oct4–GFP+ to Oct4–GFP cells relative to an empty vector control. Experiment equivalent to Fig. 2b except that second shRNAs were transduced two days after induction of reprogramming. c, Representative FACS plots showing effects of Chaf1a/b or Ube2i suppression on emergence of Oct4–GFP+ cells at days 7, 9 and 11 of OKSM expression. Histogram plots show fraction of Nanog+ cells within Oct4–GFP+ cells.

Extended Data Figure 3 Effect of CAF-1 suppression on OKSM levels and cellular growth, and shRNA rescue experiment.

a, Quantitative RT–PCR for transgenic OKSM expression using reprogrammable MEFs transduced with indicated shRNA vectors. Error bars show s.d. from biological triplicates. b, RNA-seq analysis of OKSM transgene expression in reprogrammable MEFs transduced with Renilla and Chaf1a shRNAs and exposed to doxycycline for 0, 3 or 6 days. Error bars indicate s.d. from biological triplicates. c, Western blot analysis for Sox2 and Tbp (loading control) in reprogrammable MEFs transduced with shRNA vectors targeting Renilla (Ren.713) or different CAF-1 components and exposed to doxycycline for 3 days (see Supplementary Fig. 1 for full scans). The same membrane was probed with anti-CAF-1 p150 and anti-CAF-1 p60 antibody to confirm knockdown (data not shown). d, Rescue experiment to demonstrate specificity of Chaf1b.367 shRNA vector. Reprogrammable MEFs carrying Oct4–tomato knock-in reporter were infected with lentiviral vectors expressing either EGFP or human CAF-1 p60 (CHAF1B) before transducing cells with Renilla or Cha1fb.367 shRNAs and applying doxycycline for 6 days. Colonies were counted at day 11. Note that CAF-1 p60 overexpression attenuates enhanced reprogramming elicited by Chaf1b suppression. e, f, Competitive proliferation assay between shRNA vector-infected and non-infected reprogrammable cells using indicated shRNAs in the presence or absence of doxycycline (OKSM expression). Note that CAF-1 suppression does not substantially affect the proliferation potential of reprogrammable MEFs after 1–3 days of doxycycline (OKSM) induction while it impairs the long-term growth potential of uninduced MEFs. Data were normalized to cell counts in ‘no OKSM’ condition for e and ‘day 2’ time point for f. Error bars show s.d. from biological triplicates.

Extended Data Figure 4 Confirmation of CAF-1 reprogramming phenotype with alternative transgenic and non-transgenic vector systems.

a, Alkaline phosphatase (AP)-positive, transgene-independent iPS cell colonies at day 14 following transduction of R26-M2rtTA MEFs with tetO-STEMCCA lentiviral OKSM expression vector and either Chaf1a.164 or Ren.713 shRNA vectors and treatment with high (2 μg ml−1) or low (0.2 μg ml−1) doses of doxycycline for 10 days. b, Quantification of data shown in a. Experiment was performed at 3 different plating densities (n = 1 experiment per density), representative data are shown. c, Comparison of reprogramming efficiencies between Col1a1::tetOP-OKSM; R26-M2rtTA reprogrammable MEFs and wild-type MEFs infected directly with OKSM-expressing lentiviral vectors containing either a strong Ef1a full-length promoter (Ef1a-OKSM long) or a weaker truncated promoter (Ef1a-OKSM short). TRE3G-OKSM is a lentiviral vector with a strong promoter, whose activity is downregulated over time upon infection of CAGS-rtTA3 transgenic MEFs (see below). Error bars show s.d. from biological triplicates. d, Quantitative RT–PCR data showing variability in OKSM expression levels over time using different vector systems. Cells were analysed after 3 and 6 days of infection (lentiviral vectors) or doxycycline exposure (reprogrammable MEFs). Error bars show s.d. from biological triplicates. OGR MEF, transgenic MEFs carrying Oct4–GFP and CAGS-rtTA3 alleles. e, Quantification of Oct4 protein levels by intracellular flow cytometry (top) and cellular granularity/complexity by side scatter (SSC) analysis of indicated samples (bottom). Error bars show s.d. from biological triplicates.

Extended Data Figure 5 Effects of CAF-1 dose on NIH3T3 growth and reprogramming potential.

a, Competitive proliferation assay to determine effect of indicated Chaf1a and Chaf1b shRNA vectors on long-term growth potential of immortalized NIH3T3 cell line. Cells were infected with indicated constructs and the fraction of shRNA vector-positive cells was measured by flow cytometry at different time points. Data were normalized to cell counts at day 2 post-transduction. Rpa3.455, validated control shRNA targeting the broadly essential replication protein A3. Error bars show s.d. from biological triplicates. b, Histogram plots of MEFs harbouring R26-M2rtTA allele and either Col1a1::tetOP-miR30-tRFP-Ren.713 or Col1a1::tetOP-miR30-tRFP-Chaf1a.164 shRNA knock-in allele after transduction with pHAGE (Ef1a-OKSM) lentiviral vector and exposure of cells to different doses of doxycycline for 2, 4 and 6 days. Low doses of doxycycline (0.2 µg ml−1) result in lower expression of the shRNA miR cassettes than high doses of doxycycline (2 µg ml−1). c, Quantification of data shown in b using the geometric mean (n = 1 experiment for 3 indicated time points). d, Reprogramming efficiency of Col1a1::tetOP-miR30-tRFP-Chaf1a.164; R26-M2rtTA MEFs infected with pHAGE (Ef1a-OKSM) vector and induced with high (2 μg ml−1) or low (0.2 μg ml−1) doses of doxycycline for indicated number of days before scoring for Nanog+ iPS cells by immunocytochemistry on day 9. e, Classification of CRISPR/Cas9-induced mutations by sequence analysis of representative iPS cell clones (wt, wild type; indel, insertion/deletion; fs, frameshift; *, point mutation). f, Western blot analysis for CAF-1 subunits p150 and p60 in 6 representative iPS cell clones after CRISPR/Cas9-induced modifications of the Chaf1a locus (see Supplementary Fig. 1 for full scans). Wt/wt samples show unmodified wild-type control samples.

Extended Data Figure 6 Effect of CAF-1 suppression on HSP cell reprogramming and transdifferentiation.

a, Gating strategy for determining Pecam+ fraction (shaded area) in panel b; data identical to Fig. 4c. b, Quantification of the fraction of Pecam+ cells at day 4 and day 6 of reprogramming. Data obtained from one experiment using two different Chaf1 shRNAs. c, Transgene dependence assay during the reprogramming of haematopoietic stem and progenitor cells (HSP cells) into iPS cells in the presence of Chaf1a or Renilla shRNAs. Doxycycline pulses were given for 3 or 6 days and alkaline phosphatase (AP)-positive colonies were scored at day 10. d, Quantitative RT–PCR analysis of Chaf1a expression to confirm knockdown after 3 days of doxycycline induction, that is, coexpression of shRNAmiR and Ascl1 (n = 4 independent infections of the same Col1a1::tetOP-Chaf1a.164 shRNA MEF line; mean value ± s.d.). e, Gating strategy for determining Cd14+ and Mac1+ fractions (shaded area) shown in f; data identical to Fig. 4g. Positive gates were based on untreated (0 h) control cells. f, Quantification of the fraction of Cd14+ and Mac1+ cells at 0, 24 and 48 h of transdifferentiation using indicated CAF-1 shRNA or empty control vector (n = 2 independent infections; rep, replicate). g, Quantitative RT–PCR analysis of Chaf1a and Chaf1b expression to confirm knockdown in transduced pre-B cell line before induction of transdifferentiation (kd/ctrl, knockdown/empty vector control; n = 1 experiment, representative of 2 independent infections).

Extended Data Figure 7 CAF-1 suppression promotes chromatin accessibility at enhancer elements.

a, Experimental outline and assays (SONO-seq, ATAC-seq, Sox2 ChIP-seq, H3K9me3 ChIP-seq, microarrays and RNA-seq) to dissect effect of CAF-1 suppression on chromatin accessibility, transcription factor binding, heterochromatin patterns and gene expression. Assays were performed either in early reprogramming intermediates (day 3) or throughout the reprogramming time course (ATAC-seq and gene expression). b, SONO-seq analysis of CAF-1 knockdown and control cells at day 3 of reprogramming to determine accessible chromatin regions across promoters (n = 5,513) and ES-cell-specific enhancers (n = 14,265). CAF-1 shRNA vectors Chaf1a.164, Chaf1a.2120, Chaf1b.365 and Chaf1b.1221 were pooled for this experiment. c, ATAC-seq peak distribution across different genomic features. Shown is classification of peaks that are gained in CAF-1 knockdown cells compared to Renilla. d, ATAC-seq analysis of Chaf1a and Renilla control cells at day 3 of reprogramming to measure global chromatin accessibility over pluripotency-specific super-enhancer elements. ATAC-seq data from Chaf1a.164 shRNA- and Chaf1a.2120 shRNA-transduced cells were merged for this analysis. e, ATAC-seq accessibility maps at super-enhancer elements associated with the Sall4 locus. Shaded grey bars highlight more accessible sites in Chaf1a knockdown samples at days 3 and 6 of reprogramming compared to Renilla shRNA controls. f, ATAC-seq analysis of Chaf1a and control cells at day 3 of reprogramming to measure global chromatin accessibility over lineage-specific super-enhancer elements43 (C2C12, myoblast cell line; proB, progenitor B cells; Th, T helper cells). The n denotes number of examined enhancer elements for each cell type. ATAC-seq data from Chaf1a.164 shRNA- and Chaf1a.2120 shRNA-transduced cells were merged for this analysis.

Extended Data Figure 8 CAF-1 suppression facilitates Sox2 binding to chromatin.

a, Sox2 ChIP-seq enrichment across pluripotency-specific super-enhancer elements at day 3 of reprogramming in the presence of indicated shRNA vectors. b, Venn diagram depicting shared and unique Sox2 targets in Chaf1a and Renilla knockdown cells. c, Bar graph shows the number and fraction of ES-cell-specific Sox2 targets (blue colour) among Sox2-bound sites that are unique to Chaf1a or Renilla knockdown cells at day 3 of OKSM expression. d, Sox2 ChIP-seq analysis of Chaf1a and control shRNA-infected cells at day 3 of reprogramming to determine enrichment of Sox2 binding across lineage-specific super-enhancer elements43 (C2C12, myoblast cell line; proB, progenitor B cells; Th, T helper cells; P value <10−15 for all comparisons between Chaf1a knockdown cells and control).

Extended Data Figure 9 CAF-1 suppression induces specific depletion of H3K9me3 at somatic heterochromatin domains.

a, Scatter plots comparing H3K9me3 enrichment nearby ATAC-seq sensitive and super-enhancer regions between control (Ren.713) and Chaf1a knockdown cells (Chaf1a.164 and Chaf1a.2122) at day 3 of reprogramming. Values reflect normalized H3K9me3 ChIP signal (IP/input) for 5-kb genomic regions overlapping ATAC-seq sensitive regions (red), super-enhancer regions (orange) and regions within 50-kb upstream and downstream of super-enhancers (black). b, Scatter plots comparing H3K9me3 enrichment over transposable element (TE) families in control and Chaf1a knockdown cells at day 3 of reprogramming. Values reflect normalized H3K9me3 ChIP-seq signal (IP/input) over families of TEs in the mouse genome. c, Heatmap shows the relative changes (z-normalized) of TE family expression as estimated by RNA sequencing in control and Chaf1a knockdown cells at day 0, 3 and 6 of reprogramming. Data are clustered using the k-means algorithm. d, Cumulative histogram showing the relative fraction of reprogramming-resistant regions (RRRs)29 (x axis) that display negative or positive enrichment (fold change) of average H3K9me3 signal at day 3 of reprogramming in control and Chaf1a knockdown cells. Note that more RRRs exhibit depletion of H3K9me3 in Chaf1a knockdown samples. e, H3K9me3 ChIP-seq analysis of RRRs after 0 and 3 days of reprogramming. Box plots depict representative RRRs on chromosome 7 (P < 0.05 for both shRNAs). See also Fig. 5d. f, Histogram plot showing activation of UAS–Oct4–GFP transgene upon suppression of Chaf1b (shRNA+ line) in the presence of Gal4–VP16 fusion protein. See Fig. 5f for quantification.

Supplementary information

Supplementary Figure 1

This file contains full scans of Western Blot data for Figures 1f and Extended Data Figures 1b, 3c and 5f. (PDF 1421 kb)

Supplementary Table 1

This file contains Primary data from the arrayed shRNAmiR screen including all 1,075 tested shRNAmiRs, their sequences, corresponding scores for each biological replicate (FC replicate A/B) and their average (FC avg), which are plotted. (XLSX 130 kb)

Supplementary Table 2

This file contains primary data from the multiplexed shRNAmiR screen. Sheet 1 includes all 5,049 tested shRNAmiRs, their sequences and corresponding average scores for each time point. Sheet 2 shows raw data as normalized reads for all biological replicates (3-A01-D12, 48 replicates with OKSM induction after 3 days of G418 selection; 6-A01-D12, 48 replicates with OKSM induction after 3 days of G418 selection). Fold-change (FC) ratios for each shRNA were calculated by dividing the normalized reads of each shRNA by the normalized reads in MEFs 3 days after viral transduction. The mean of all FC values for the day 3 and day 6 time point were used to calculate the average FC for both time points (AvgFC_D3/6). In addition, FC ratios were used to calculate scores for each shRNA in each replicate (default score=0; score=1 if FC>1, score=3 if FC>10), which were summed up over all replicates separately for the day 3 and day 6 time point (Score_D3/6). (XLSX 2875 kb)

Supplementary Table 3

This file shows accessible super-enhancers in reprogramming intermediates upon CAF-1 suppression. Genomic coordinates of significantly more accessible super-enhancer (SE) sites in CAF-1 knockdown samples compared to Renilla control including log2 fold-enrichment scores (log2FC) and p-values for ATAC-seq data with indication of neighboring genes. (XLSX 12 kb)

Supplementary Table 4

This file shows correlation analysis between Sox2 binding sites and chromatin accessibility at super-enhancer sites in control and CAF-1 knockdown cells expressing OKSM. Sheet 1 shows Sox2 binding site counts in reprogramming intermediates (day 3) containing Chaf1a or Renilla shRNAs (unique and shared sites), ESC-specific sites, overlap with super-enhancers and positive correlation with ATAC-seq data. Sheet 2 depicts unique Sox2-bound super-enhancer sites in Chaf1a knockdown cells that are also more accessible by ATAC-Seq analysis, including genomic coordinates, neighboring genes and log2 fold-enrichment for both ChIP-seq and ATAC-seq data. (XLSX 10 kb)

Supplementary Table 5

This file shows significantly up-regulated genes in CAF-1 shRNA transduced MEFs relative to Renilla control by day 6 of reprogramming. Sheet1 contains genes that are either (1) significantly up-regulated at day 6 compared to day 3 in CAF-1 knockdown cells relative to Renilla control as determined by expression array data, or (2) significantly up-regulated at day 6 in CAF-1 shRNA transduced MEFs compared to Renilla control by RNA-seq analysis. (XLSX 9 kb)

PowerPoint slides

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Cheloufi, S., Elling, U., Hopfgartner, B. et al. The histone chaperone CAF-1 safeguards somatic cell identity. Nature 528, 218–224 (2015). https://doi.org/10.1038/nature15749

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nature15749

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

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