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.

  • Review Article
  • Published:

Open chromatin in pluripotency and reprogramming

A Corrigendum to this article was published on 23 March 2011

This article has been updated

Key Points

  • Pluripotent stem cells, such as embryonic stem cells, maintain the capacity to differentiate into all cell types of the body through a complex regulatory mechanism that involves a particular chromatin landscape.

  • Pluripotent stem cells have been shown, by a variety of approaches, to have an open chromatin state with reduced levels of heterochromatin, both in vitro and in vivo. This open chromatin state is thought to be important for the maintenance of pluripotency.

  • Open chromatin may be regulated by several chromatin regulators that are abundant in embryonic stem cells. These factors seem to actively prevent heterochromatin from expanding in the undifferentiated state.

  • In the context of a globally open chromatin, other chromatin regulators contribute locally to the silencing of lineage-specific genes until differentiation is triggered, keeping pluripotent stem cells in a poised undifferentiated state.

  • Reprogramming of somatic cells to pluripotent stem cells requires re-opening of chromatin in a process that probably involves some of the same factors that maintain open chromatin. Chromatin re-opening during reprogramming may not always be complete and thus leaves an epigenetic memory of the original cell type.

  • The overcoming of epigenetic barriers during somatic cell reprogramming to pluripotency appears to have molecular parallels with cellular transformation in cancer.

Abstract

Pluripotent stem cells can be derived from embryos or induced from adult cells by reprogramming. They are unique among stem cells in that they can give rise to all cell types of the body. Recent findings indicate that a particularly 'open' chromatin state contributes to maintenance of pluripotency. Two principles are emerging: specific factors maintain a globally open chromatin state that is accessible for transcriptional activation; and other chromatin regulators contribute locally to the silencing of lineage-specific genes until differentiation is triggered. These same principles may apply during reacquisition of an open chromatin state upon reprogramming to pluripotency, and during de-differentiation in cancer.

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: Chromatin in pluripotent stem cells versus differentiated cells.
Figure 2: The balance between euchromatin and heterochromatin in ES cells.

Similar content being viewed by others

Change history

  • 23 March 2011

    In the reference list for this article, reference 39 was incorrect. The reference should have been "Azuara V. et al. Chromatin signatures of pluripotent cell lines. Nature Cell Biol. 8, 532–538 (2006)". The authors apologize for this error.

References

  1. Evans, M. J. & Kaufman, M. H. Establishment in culture of pluripotential cells from mouse embryos. Nature 292, 154–156 (1981).

    CAS  PubMed  Google Scholar 

  2. Martin, G. Isolation of a pluripotent cell line from early mouse embryos cultured in medium conditioned by teratocarcinoma stem cells. Proc. Natl Acad. Sci. USA 78, 7634–7638 (1981).

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Thomson, J. A. et al. Embryonic stem cell lines derived from human blastocysts. Science 282, 1145–1147 (1998).

    CAS  PubMed  Google Scholar 

  4. Bradley, A., Evans, M., Kaufman, M. H. & Robertson, E. Formation of germ-line chimaeras from embryo-derived teratocarcinoma cell lines. Nature 309, 255–266 (1984).

    CAS  PubMed  Google Scholar 

  5. Kriegstein, A. & Alvarez-Buylla, A. The glial nature of embryonic and adult neural stem cells. Annu. Rev. Neurosci. 32, 149–184 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Dzierzak, E. The emergence of definitive hematopoietic stem cells in the mammal. Curr. Opin. Hematol. 12, 197–202 (2005).

    PubMed  Google Scholar 

  7. Lee, G. et al. Modelling pathogenesis and treatment of familial dysautonomia using patient-specific iPSCs. Nature 461, 402–406 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Carvajal-Vergara, X. et al. Patient-specific induced pluripotent stem-cell-derived models of LEOPARD syndrome. Nature 465, 808–812 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Matsui, Y., Zsebo, K. & Hogan, B. L. Derivation of pluripotential embryonic stem cells from murine primordial germ cells in culture. Cell 70, 841–847 (1992).

    CAS  PubMed  Google Scholar 

  10. Resnick, J. L., Bixler, L. S., Cheng, L. & Donovan, P. J. Long-term proliferation of mouse primordial germ cells in culture. Nature 359, 550–551 (1992).

    CAS  PubMed  Google Scholar 

  11. Ko, K. et al. Induction of pluripotency in adult unipotent germline stem cells. Cell Stem Cell 5, 87–96 (2009).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  14. Takahashi, K. et al. Induction of pluripotent stem cells from adult human fibroblasts by defined factors. Cell 131, 861–872 (2007).

    CAS  PubMed  Google Scholar 

  15. Lowry, W. E. et al. Generation of human induced pluripotent stem cells from dermal fibroblasts. Proc. Natl Acad. Sci. USA 105, 2883–2888 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Park, I.-H. et al. Reprogramming of human somatic cells to pluripotency with defined factors. Nature 451, 141–146 (2008).

    CAS  PubMed  Google Scholar 

  17. Yu, J. et al. Induced pluripotent stem cell lines derived from human somatic cells. Science 318, 1917–1920 (2007).

    CAS  PubMed  Google Scholar 

  18. Meshorer, E. & Misteli, T. Chromatin in pluripotent embryonic stem cells and differentiation. Nature Rev. Mol. Cell Biol. 7, 540–546 (2006).

    CAS  Google Scholar 

  19. Paweletz, N. Walther Flemming: pioneer of mitosis research. Nature Rev. Mol. Cell Biol. 2, 72–75 (2001).

    CAS  Google Scholar 

  20. Flemming, W. Zellsubstanz, Kern und Zelltheilung (ed. Vogel, F. C. W.; Leipzig, Germany, 1882).

    Google Scholar 

  21. Heitz, E. Das heterochromatin der moose. I. Jahrb Wiss Botanik 69, 762–818 (1928).

    Google Scholar 

  22. Reddien, P. W. & Sánchez-Alvarado, A. Fundamentals of planarian regenaration. Annu. Rev. Cell Dev. Biol. 20, 725–757 (2004).

    CAS  PubMed  Google Scholar 

  23. Spangrude, G. J., Heimfeld, S. & Weissman, I. L. Purification and characterization of mouse hematopoietic stem cells. Science 241, 58–62 (1988).

    CAS  PubMed  Google Scholar 

  24. Mattout, A. & Meshorer, E. Chromatin plasticity and genome organization in pluripotent embryonic stem cells. Curr. Opin. Cell Biol. 22, 334–341 (2010).

    CAS  PubMed  Google Scholar 

  25. Park, S.-H. et al. Ultrastructure of human embryonic stem cells and spontaneous and retinoic acid-induced differentiating cells. Ultrastruct. Pathol. 28, 229–238 (2004).

    PubMed  Google Scholar 

  26. Efroni, S. et al. Global transcription in pluripotent embryonic stem cells. Cell Stem Cell 2, 437–447 (2008). This study identifies the ES cell genome as transcriptionally hyperactive, and demonstrates — using ESI — increased heterochromatin in differentiating cells, as well as an abundance of chromatin remodellers in the undifferentiated state.

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Ahmed, K. et al. Global chromatin architecture reflects pluripotency and lineage commitment in the early mouse embryo. PLoS ONE 5, e10531 (2010). Using ESI, the authors show that there are changes in chromatin structure in the early embryo, demonstrating that pluripotent epiblast cells in the ICM of the blastocyst have a less condensed chromatin than lineage-committed cells.

    PubMed  PubMed Central  Google Scholar 

  28. Schaniel, C. et al. Smarcc1/Baf155 couples self-renewal gene repression with changes in chromatin structure in mouse embryonic stem cells. Stem Cells 27, 2979–2991 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Meshorer, E. et al. Hyperdynamic plasticity of chromatin proteins in pluripotent embryonic stem cells. Dev. Cell 10, 105–116 (2006). This paper shows increased chromatin plasticity in ES cells and reduced levels of heterochromatin-associated histone modifications compared with differentiating cells.

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Wen, B., Wu, H., Shinkai, Y., Irizarry, R. & Feinberg, A. Large histone H3 lysine 9 dimethylated chromatin blocks distinguish differentiated from embryonic stem cells. Nature Genet. 41, 246–250 (2009).

    CAS  PubMed  Google Scholar 

  31. Hawkins, R. D. et al. Distinct epigenomic landscapes of pluripotent and lineage-committed human cells. Cell Stem Cell 6, 479–491 (2010). This article describes changes in the epigenomic landscapes of ES cells compared with differentiated cells and shows the expansion of H3K9me3 and H3K27me3 marks during differentiation.

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Krejcí, J. et al. Genome-wide reduction in H3K9 acetylation during human embryonic stem cell differentiation. J. Cell Physiol. 219, 677–687 (2009).

    PubMed  Google Scholar 

  33. Bhattacharya, D., Talwar, S., Mazumder, A. & Shivashankar, G. V. Spatio-temporal plasticity in chromatin organization in mouse cell differentiation and during Drosophila embryogenesis. Biophys. J. 96, 3832–3839 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Szutorisz, H., Georgiou, A., Tora, L. & Dillon, N. The proteasome restricts permissive transcription at tissue-specific gene loci in embryonic stem cells. Cell 127, 1375–1388 (2006).

    CAS  PubMed  Google Scholar 

  35. Bernstein, B. E. et al. A bivalent chromatin structure marks key developmental genes in embryonic stem cells. Cell 125, 315–326 (2006).

    CAS  PubMed  Google Scholar 

  36. Pan, G. et al. Whole-genome analysis of histone H3 lysine 4 and lysine 27 methylation in human embryonic stem cells. Cell Stem Cell 1, 299–312 (2007).

    CAS  PubMed  Google Scholar 

  37. Boyer, L. A., Mathur, D. & Jaenisch, R. Molecular control of pluripotency. Curr. Opin. Genet. Dev. 16, 455–462 (2006).

    CAS  PubMed  Google Scholar 

  38. Kashyap, V. et al. Regulation of stem cell pluripotency and differentiation involves a mutual regulatory circuit of the NANOG, OCT4, and SOX2 pluripotency transcription factors with polycomb repressive complexes and stem cell microRNAs. Stem Cells Dev. 18, 1093–1108 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Azuara, V. et al. Chromatin signatures of pluripotent cell lines. Nature Cell Biol. 8, 532–538 (2006).

    CAS  PubMed  Google Scholar 

  40. Boyer, L. A. et al. Polycomb complexes repress developmental regulators in murine embryonic stem cells. Nature 441, 349–353 (2006).

    CAS  PubMed  Google Scholar 

  41. Lee, T. I. et al. Control of developmental regulators by Polycomb in human embryonic stem cells. Cell 125, 301–313 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Chamberlain, S. J., Yee, D. & Magnuson, T. Polycomb repressive complex 2 is dispensable for maintenance of embryonic stem cell pluripotency. Stem Cells 26, 1496–1505 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Shen, X. et al. EZH1 mediates methylation on histone H3 lysine 27 and complements EZH2 in maintaining stem cell identity and executing pluripotency. Mol. Cell 32, 491–502 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Pasini, D. et al. JARID2 regulates binding of the Polycomb repressive complex 2 to target genes in ES cells. Nature 464, 306–310 (2010).

    CAS  PubMed  Google Scholar 

  45. Li, G. et al. Jarid2 and PRC2, partners in regulating gene expression. Genes Dev. 24, 368–380 (2010).

    PubMed  PubMed Central  Google Scholar 

  46. Peng, J. C. et al. Jarid2/Jumonji coordinates control of PRC2 enzymatic activity and target gene occupancy in pluripotent cells. Cell 139, 1290–1302 (2009).

    PubMed  PubMed Central  Google Scholar 

  47. Shen, X. et al. Jumonji modulates polycomb activity and self-renewal versus differentiation of stem cells. Cell 139, 1303–1314 (2009).

    PubMed  PubMed Central  Google Scholar 

  48. Feldman, N. et al. G9a-mediated irreversible epigenetic inactivation of Oct-3/4 during early embryogenesis. Nature Cell Biol. 8, 188–194 (2006).

    CAS  PubMed  Google Scholar 

  49. Epsztejn-Litman, S. et al. De novo DNA methylation promoted by G9a prevents reprogramming of embryonically silenced genes. Nature Struct. Mol. Biol. 15, 1176–1183 (2008).

    CAS  Google Scholar 

  50. Loh, Y.-H., Zhang, W., Chen, X., George, J. & Ng, H.-H. Jmjd1a and Jmjd2c histone H3 Lys 9 demethylases regulate self-renewal in embryonic stem cells. Genes Dev. 21, 2545–2557 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Meissner, A. et al. Genome-scale DNA methylation maps of pluripotent and differentiated cells. Nature 454, 766–770 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Fouse, S. D. et al. Promoter CpG methylation contributes to ES cell gene regulation in parallel with Oct4/Nanog, PcG complex, and histone H3 K4/K27 trimethylation. Cell Stem Cell 2, 160–169 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Lister, R. et al. Human DNA methylomes at base resolution show widespread epigenomic differences. Nature 462, 315–322 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. de la Serna, I. L., Ohkawa, Y. & Imbalzano, A. N. Chromatin remodelling in mammalian differentiation: lessons from ATP-dependent remodellers. Nature Rev. Genet. 7, 461–473 (2006).

    CAS  PubMed  Google Scholar 

  55. Cairns, B. R. The logic of chromatin architecture and remodelling at promoters. Nature 461, 193–198 (2009).

    CAS  PubMed  Google Scholar 

  56. Clapier, C. R. & Cairns, B. R. The biology of chromatin remodeling complexes. Annu. Rev. Biochem. 78, 273–304 (2009).

    CAS  PubMed  Google Scholar 

  57. Moshkin, Y. M., Mohrmann, L., van Ijcken, W. F. J. & Verrijzer, C. P. Functional differentiation of SWI/SNF remodelers in transcription and cell cycle control. Mol. Cell Biol. 27, 651–661 (2007).

    CAS  PubMed  Google Scholar 

  58. Lessard, J. A. & Crabtree, G. R. Chromatin regulatory mechanisms in pluripotency. Annu. Rev. Cell Dev. Biol. 26, 503–532 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Ho, L. et al. An embryonic stem cell chromatin remodeling complex, esBAF, is essential for embryonic stem cell self-renewal and pluripotency. Proc. Natl Acad. Sci. USA 106, 5181–5186 (2009). The authors characterize a special composition of the SWI/SNF chromatin-remodelling complex in ES cells, esBAF, and its role in ES cell maintenance and pluripotency.

    CAS  PubMed  PubMed Central  Google Scholar 

  60. Kaeser, M. D., Aslanian, A., Dong, M.-Q., Yates, J. R. & Emerson, B. M. BRD7, a novel PBAF-specific SWI/SNF subunit, is required for target gene activation and repression in embryonic stem cells. J. Biol. Chem. 283, 32254–32263 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Bultman, S. et al. A Brg1 null mutation in the mouse reveals functional differences among mammalian SWI/SNF complexes. Mol. Cell 6, 1287–1295 (2000).

    CAS  PubMed  Google Scholar 

  62. Fazzio, T. G., Huff, J. T. & Panning, B. An RNAi screen of chromatin proteins identifies Tip60–p400 as a regulator of embryonic stem cell identity. Cell 134, 162–174 (2008). This paper reports an RNA interference screen of chromatin proteins in ES cells and a characterization of the TIP60–p400 complex, which is necessary to maintain ES cell identity.

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Kidder, B. L., Palmer, S. & Knott, J. G. SWI/SNF–Brg1 regulates self-renewal and occupies core pluripotency-related genes in embryonic stem cells. Stem Cells 27, 317–328 (2009).

    CAS  PubMed  Google Scholar 

  64. Ho, L. et al. An embryonic stem cell chromatin remodeling complex, esBAF, is an essential component of the core pluripotency transcriptional network. Proc. Natl Acad. Sci. USA 106, 5187–5191 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  65. Gao, X. et al. ES cell pluripotency and germ-layer formation require the SWI/SNF chromatin remodeling component BAF250a. Proc. Natl Acad. Sci. USA 105, 6656–6661 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  66. Yan, Z. et al. BAF250B-associated SWI/SNF chromatin-remodeling complex is required to maintain undifferentiated mouse embryonic stem cells. Stem Cells 26, 1155–1165 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  67. Bajpai, R. et al. CHD7 cooperates with PBAF to control multipotent neural crest formation. Nature 463, 958–962 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  68. Schnetz, M. P. et al. Genomic distribution of CHD7 on chromatin tracks H3K4 methylation patterns. Genome Res. 19, 590–601 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  69. Gaspar-Maia, A. et al. Chd1 regulates open chromatin and pluripotency of embryonic stem cells. Nature 460, 863–868 (2009). This study identified a chromatin-remodelling protein that binds to active genes and is required to maintain an open chromatin state and pluripotency in ES cells.

    CAS  PubMed  PubMed Central  Google Scholar 

  70. Kaji, K. et al. The NuRD component Mbd3 is required for pluripotency of embryonic stem cells. Nature Cell Biol. 8, 285–292 (2006).

    CAS  PubMed  Google Scholar 

  71. Zhu, D., Fang, J., Li, Y. & Zhang, J. Mbd3, a component of NuRD/Mi-2 complex, helps maintain pluripotency of mouse embryonic stem cells by repressing trophectoderm differentiation. PLoS ONE 4, e7684 (2009).

    PubMed  PubMed Central  Google Scholar 

  72. 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).

    CAS  PubMed  PubMed Central  Google Scholar 

  73. Guan, J.-S. et al. HDAC2 negatively regulates memory formation and synaptic plasticity. Nature 459, 55–60 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  74. Montgomery, R. L. et al. Histone deacetylases 1 and 2 redundantly regulate cardiac morphogenesis, growth, and contractility. Genes Dev. 21, 1790–1802 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  75. Zimmermann, S. et al. Reduced body size and decreased intestinal tumor rates in HDAC2-mutant mice. Cancer Res. 67, 9047–9054 (2007).

    CAS  PubMed  Google Scholar 

  76. Trivedi, C. M. et al. Hdac2 regulates the cardiac hypertrophic response by modulating Gsk3β activity. Nature Med. 13, 324–331 (2007).

    CAS  PubMed  Google Scholar 

  77. Landry, J. et al. Essential role of chromatin remodeling protein Bptf in early mouse embryos and embryonic stem cells. PLoS Genet. 4, e1000241 (2008).

    PubMed  PubMed Central  Google Scholar 

  78. Fazzio, T. G. & Panning, B. Condensin complexes regulate mitotic progression and interphase chromatin structure in embryonic stem cells. J. Cell Biol. 188, 491–503 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  79. Venkatasubrahmanyam, S., Hwang, W. W., Meneghini, M. D., Tong, A. H. Y. & Madhani, H. D. Genome-wide, as opposed to local, antisilencing is mediated redundantly by the euchromatic factors Set1 and H2AZ. Proc. Natl Acad. Sci. USA 104, 16609–16614 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  80. Kimura, A., Umehara, T. & Horikoshi, M. Chromosomal gradient of histone acetylation established by Sas2p and Sir2p functions as a shield against gene silencing. Nature Genet. 32, 370–377 (2002).

    PubMed  Google Scholar 

  81. Osborne, E. A., Dudoit, S. & Rine, J. The establishment of gene silencing at single-cell resolution. Nature Genet. 41, 800–806 (2009).

    CAS  PubMed  Google Scholar 

  82. McKittrick, E., Gafken, P. R., Ahmad, K. & Henikoff, S. Histone H3.3 is enriched in covalent modifications associated with active chromatin. Proc. Natl Acad. Sci. USA, 101, 1525–1530 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  83. Hake, S. B. et al. Expression patterns and post-translational modifications associated with mammalian histone H3 variants. J. Biol. Chem. 281, 559–568 (2006).

    CAS  PubMed  Google Scholar 

  84. Tagami, H., Ray-Gallet, D., Almouzni, G. & Nakatani, Y. Histone H3.1 and H3.3 complexes mediate nucleosome assembly pathways dependent or independent of DNA synthesis. Cell 116, 51–61 (2004).

    CAS  PubMed  Google Scholar 

  85. Goldberg, A. D. et al. Distinct factors control histone variant H3.3 localization at specific genomic regions. Cell 140, 678–691 (2010). This article describes the genome-wide incorporation of the histone variant H3.3 in ES cells, and shows that incorporation in gene promoters requires HIRA, but incorporation at enhancer elements and telomeres is HIRA-independent and requires ATRX and DAXX.

    CAS  PubMed  PubMed Central  Google Scholar 

  86. Mito, Y., Henikoff, J. G. & Henikoff, S. Genome-scale profiling of histone H3.3 replacement patterns. Nature Genet. 37, 1090–1097 (2005).

    CAS  PubMed  Google Scholar 

  87. Ng, R. K. & Gurdon, J. B. Epigenetic memory of an active gene state depends on histone H3.3 incorporation into chromatin in the absence of transcription. Nature Cell Biol. 10, 102–109 (2008).

    CAS  PubMed  Google Scholar 

  88. Konev, A. Y. et al. CHD1 motor protein is required for deposition of histone variant H3.3 into chromatin in vivo. Science 317, 1087–1090 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  89. Sims, R. J. et al. Human but not yeast CHD1 binds directly and selectively to histone H3 methylated at lysine 4 via its tandem chromodomains. J. Biol. Chem. 280, 41789–41792 (2005).

    CAS  PubMed  Google Scholar 

  90. Sims, R. J. et al. Recognition of trimethylated histone H3 lysine 4 facilitates the recruitment of transcription postinitiation factors and pre-mRNA splicing. Mol. Cell 28, 665–676 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  91. Nishioka, K. et al. Set9, a novel histone H3 methyltransferase that facilitates transcription by precluding histone tail modifications required for heterochromatin formation. Genes Dev. 16, 479–489 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  92. Zegerman, P., Canas, B., Pappin, D. & Kouzarides, T. Histone H3 lysine 4 methylation disrupts binding of nucleosome remodeling and deacetylase (NuRD) repressor complex. J. Biol. Chem. 277, 11621–11624 (2002).

    CAS  PubMed  Google Scholar 

  93. Ooi, S. K. T. et al. DNMT3L connects unmethylated lysine 4 of histone H3 to de novo methylation of DNA. Nature 448, 714–717 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  94. Ramalho-Santos, M. iPS cells: insights into basic biology. Cell 138, 616–618 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  95. Brambrink, T. et al. Sequential expression of pluripotency markers during direct reprogramming of mouse somatic cells. Cell Stem Cell 2, 151–159 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  96. Boyer, L. A. et al. Core transcriptional regulatory circuitry in human embryonic stem cells. Cell 122, 947–956 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  97. Chen, X. et al. Integration of external signaling pathways with the core transcriptional network in embryonic stem cells. Cell 133, 1106–1117 (2008).

    CAS  PubMed  Google Scholar 

  98. Nakagawa, M. et al. Generation of induced pluripotent stem cells without Myc from mouse and human fibroblasts. Nature Biotechnol. 26, 101–106 (2008).

    CAS  Google Scholar 

  99. Wernig, M. et al. A drug-inducible transgenic system for direct reprogramming of multiple somatic cell types. Nature Biotechnol. 26, 916–924 (2008).

    CAS  Google Scholar 

  100. Knoepfler, P. S. Why myc? An unexpected ingredient in the stem cell cocktail. Cell Stem Cell 2, 18–21 (2008).

    CAS  PubMed  Google Scholar 

  101. Sridharan, R. et al. Role of the murine reprogramming factors in the induction of pluripotency. Cell 136, 364–377 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  102. Meissner, A., Wernig, M. & Jaenisch, R. Direct reprogramming of genetically unmodified fibroblasts into pluripotent stem cells. Nature Biotechnol. 25, 1177–1181 (2007).

    CAS  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  104. Hanna, J. et al. Direct cell reprogramming is a stochastic process amenable to acceleration. Nature 462, 595–601 (2009). The authors show that all somatic cells are potentially reprogrammable to iPS cells through ectopic expression of the required factors, that the process is stochastic and that inhibition of the p53/p21 pathway or increased expression of LIN28 or NANOG increases its efficiency.

    CAS  PubMed  PubMed Central  Google Scholar 

  105. Huangfu, D. et al. Induction of pluripotent stem cells by defined factors is greatly improved by small-molecule compounds. Nature Biotechnol. 26, 795–797 (2008).

    CAS  Google Scholar 

  106. Huangfu, D. et al. Induction of pluripotent stem cells from primary human fibroblasts with only Oct4 and Sox2. Nature Biotechnol. 26, 1269–1275 (2008).

    CAS  Google Scholar 

  107. Shi, Y. et al. A combined chemical and genetic approach for the generation of induced pluripotent stem cells. Cell Stem Cell 2, 525–528 (2008).

    CAS  PubMed  Google Scholar 

  108. Singhal, N. et al. Chromatin-remodeling components of the BAF complex facilitate reprogramming. Cell 141, 943–955 (2010). This paper describes a proteomic analysis of the reprogramming factors contained in nuclear fractions of pluripotent stem cells and identifies BAF complex elements, which are shown to increase efficiency of reprogramming.

    CAS  PubMed  Google Scholar 

  109. Yamanaka, S. & Blau, H. M. Nuclear reprogramming to a pluripotent state by three approaches. Nature 465, 704–712 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  110. Egli, D. & Eggan, K. Recipient cell nuclear factors are required for reprogramming by nuclear transfer. Development 137, 1953–1963 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  111. Kishigami, S. et al. Significant improvement of mouse cloning technique by treatment with trichostatin A after somatic nuclear transfer. Biochem. Biophys. Res. Commun. 340, 183–189 (2006).

    CAS  PubMed  Google Scholar 

  112. Chin, M. H. et al. Induced pluripotent stem cells and embryonic stem cells are distinguished by gene expression signatures. Cell Stem Cell 5, 111–123 (2009). This article describes some differences in gene and microRNA expression between ES and iPS cells, and analyses changes in gene expression and chromatin structure that occur in early- versus late-passage iPS cells.

    CAS  PubMed  PubMed Central  Google Scholar 

  113. Chin, M. H., Pellegrini, M., Plath, K. & Lowry, W. E. Molecular analyses of human induced pluripotent stem cells and embryonic stem cells. Cell Stem Cell 7, 263–269 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  114. Polo, J. M. et al. Cell type of origin influences the molecular and functional properties of mouse induced pluripotent stem cells. Nature Biotech. 28, 848–855 (2010)

    CAS  Google Scholar 

  115. Kim, K. et al. Epigenetic memory in induced pluripotent stem cells. Nature 467, 285–290 (2010). The authors define DNA methylation differences in iPS cells depending on their somatic cell of origin, which can be reset by serial reprogramming or through the use of chromatin-modifying drugs.

    CAS  PubMed  PubMed Central  Google Scholar 

  116. Filion, G. J. et al. Systematic protein location mapping reveals five principal chromatin types in Drosophila cells. Cell 143, 212–224 (2010). This study identifies five major chromatin domains in D. melanogaster cells, each defined by the binding of different combinations of chromatin proteins, and also reports a repressive domain that covers about half of the genome and lacks classic heterochromatin markers.

    CAS  PubMed  PubMed Central  Google Scholar 

  117. Bjerkvig, R., Tysnes, B. B., Aboody, K. S., Najbauer, J. & Terzis, A. J. A. The origin of the cancer stem cell: current controversies and new insights. Nature Rev. Cancer 5, 899–904 (2005).

    CAS  Google Scholar 

  118. Ruiz i Altaba, A., Sánchez, P. & Dahmane, N. Gli and hedgehog in cancer: tumours, embryos and stem cells. Nature Rev. Cancer 2, 361–372 (2002).

    CAS  Google Scholar 

  119. Reya, T., Morrison, S. J., Clarke, M. F. & Weissman, I. L. Stem cells, cancer, and cancer stem cells. Nature 414, 105–111 (2001).

    CAS  PubMed  Google Scholar 

  120. Ben-Porath, I. et al. An embryonic stem cell-like gene expression signature in poorly differentiated aggressive human tumors. Nature Genet. 40, 499–507 (2008).

    CAS  PubMed  Google Scholar 

  121. Komashko, V. M. et al. Using ChIP–chip technology to reveal common principles of transcriptional repression in normal and cancer cells. Genome Res. 18, 521–532 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  122. Widschwendter, M. et al. Epigenetic stem cell signature in cancer. Nature Genet. 39, 157–158 (2007).

    CAS  PubMed  Google Scholar 

  123. Ohm, J. E. et al. A stem cell-like chromatin pattern may predispose tumor suppressor genes to DNA hypermethylation and heritable silencing. Nature Genet. 39, 237–242 (2007).

    CAS  PubMed  Google Scholar 

  124. Wong, D. J. et al. Module map of stem cell genes guides creation of epithelial cancer stem cells. Cell Stem Cell 2, 333–344 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  125. Knoepfler, P. S. et al. Myc influences global chromatin structure. EMBO J. 25, 2723–2734 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  126. Cotterman, R. et al. N-Myc regulates a widespread euchromatic program in the human genome partially independent of its role as a classical transcription factor. Cancer Res. 68, 9654–9662 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  127. Frye, M., Fisher, A. G. & Watt, F. M. Epidermal stem cells are defined by global histone modifications that are altered by Myc-induced differentiation. PLoS ONE 2, e763 (2007).

    PubMed  PubMed Central  Google Scholar 

  128. Kim, J. et al. A Myc Network Accounts for Similarities between Embryonic Stem and Cancer Cell Transcription Programs. Cell 143, 313–324 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  129. Dialynas, G. K., Vitalini, M. W. & Wallrath, L. L. Linking Heterochromatin Protein 1 (HP1) to cancer progression. Mutat. Res. 647, 13–20 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  130. Kirschmann, D. A. et al. Down-regulation of HP1Hsα expression is associated with the metastatic phenotype in breast cancer. Cancer Res. 60, 3359–3363 (2000).

    CAS  PubMed  Google Scholar 

  131. Tiwari, V. K. et al. PcG proteins, DNA methylation, and gene repression by chromatin looping. PLoS Biol. 6, 2911–2927 (2008).

    CAS  PubMed  Google Scholar 

  132. Daley, G. Q. Common themes of dedifferentiation in somatic cell reprogramming and cancer. Cold Spring Harb. Symp. Quant. Biol. 73, 171–174 (2008).

    CAS  PubMed  Google Scholar 

  133. Kim, J. K. et al. Srg3, a mouse homolog of yeast SWI3, is essential for early embryogenesis and involved in brain development. Mol. Cell Biol. 21, 7787–7795 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  134. Klochendler-Yeivin, A. et al. The murine SNF5/INI1 chromatin remodeling factor is essential for embryonic development and tumor suppression. EMBO Rep. 1, 500–506 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  135. Shawlot, W., Deng, J. M., Fohn, L. E. & Behringer, R. R. Restricted β-galactosidase expression of a hygromycin-lacZ gene targeted to the β-actin locus and embryonic lethality of β-actin mutant mice. Transgenic Res. 7, 95–103 (1998).

    CAS  PubMed  Google Scholar 

  136. Marfella, C. G. A. et al. Mutation of the SNF2 family member Chd2 affects mouse development and survival. J. Cell Physiol. 209, 162–171 (2006).

    CAS  PubMed  Google Scholar 

  137. Hurd, E. A. et al. Loss of Chd7 function in gene-trapped reporter mice is embryonic lethal and associated with severe defects in multiple developing tissues. Mamm. Genome 18, 94–104 (2007).

    CAS  PubMed  Google Scholar 

  138. Nishiyama, M. et al. CHD8 suppresses p53-mediated apoptosis through histone H1 recruitment during early embryogenesis. Nature Cell Biol. 11, 172–182 (2009).

    CAS  PubMed  Google Scholar 

  139. Marino, S. & Nusse, R. Mutants in the mouse NuRD/Mi2 component P66α are embryonic lethal. PLoS ONE 2, e519 (2007).

    PubMed  PubMed Central  Google Scholar 

  140. Hu, Y. et al. Homozygous disruption of the Tip60 gene causes early embryonic lethality. Dev. Dyn. 238, 2912–2921 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  141. Tominaga, K. et al. MRG15 regulates embryonic development and cell proliferation. Mol. Cell Biol. 25, 2924–2937 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  142. Herceg, Z. et al. Disruption of Trrap causes early embryonic lethality and defects in cell cycle progression. Nature Genet. 29, 206–211 (2001).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank E. Bernstein, F. M. Koh, M. Sachs and three anonymous reviewers for constructive comments. E.M. is a Joseph H. and Belle R. Braun senior lecturer in life sciences and is supported by the Israel Science Foundation (ISF 215/07 and 943/09), the Israel Cancer Research Foundation, the Israel Ministry of Health (6007), the European Union (IRG-206872 and 238176) and an Alon Fellowship. A.A. is a Safra fellow. Work in the M.R.-S. laboratory is supported by a US National Institutes of Health Director's New Innovator Award, the California Institute for Regenerative Medicine and the Helmsley Trust.

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to Eran Meshorer or Miguel Ramalho-Santos.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Related links

Related links

FURTHER INFORMATION

Miguel Ramalho-Santos's homepage

Eran Meshorer's homepage

Glossary

Endoderm

The innermost of the three germ layers that are formed during embryonic development. Prominent examples of endodermal tissues include the epithelia of the gastrointestinal and respiratory tracts, thyroid, liver and pancreas, as well as of the auditory and urinary systems.

Mesoderm

The middle of the three germ layers that are formed during embryonic development. Prominent examples of mesodermal tissues include bone, cartilage, blood, muscle, heart, connective tissue and kidney.

Ectoderm

The outermost of the three germ layers that are formed during embryonic development. Prominent examples of ectodermal tissues include the nervous system, hair, skin, nails and eyes, as well as the various derivatives of the neural crest, including bones of the head and peripheral nerves.

Heterochromatin

Highly compacted chromatin that is transcriptionally inactive. Includes structural regions of the chromosome, such as centromeres, that lack genes ('constitutive' heterochromatin) and regions in which genes are silenced in a given cell type ('facultative' heterochromatin).

Euchromatin

A form of chromatin that is relatively decondensed and often transcriptionally active during interphase.

Electron spectroscopic imaging

(ESI). Energy-filtered transmission electron microscopy, in which the image is formed only by electrons transmitted within a certain energy window. It allows direct quantitative imaging of elements within the specimen.

Embryoid body

(EB). A cellular aggregate that is produced when ES cells are induced to differentiate in non-adherent conditions that mimic the early stages of embryogenesis.

ChIP–chip

Chromatin immunoprecipitation (ChIP) followed by microarray. ChIP is a method that allows isolation of DNA sequences that are bound to a protein of interest using specific antibodies. DNA isolated by ChIP is denatured and hybridized to a tiling array, which typically includes probes covering the entire genome. Paired probes indicate that the protein of interest was bound to that particular region of DNA.

ChIP–seq

Chromatin immunoprecipitation (ChIP) followed by sequencing. Refers to high-throughput sequencing of ChIP-isolated DNA, and provides genome-wide information of the DNA binding sites of the protein of interest.

Heterochromatin protein 1

(HP1). A heterochromatin-binding protein that recognizes and binds to histone H3 tri-methylated on Lys9. It includes three isoforms (α, β and γ), which are encoded by three different genes (CBX5, CBX1 and CBX3, respectively).

Proteasome

A large multisubunit protein complex that degrades proteins. Undesired proteins are labelled for degradation by the addition of a chain of the small protein ubiquitin; a process that is mediated by a family of enzymes called ubiquitin ligases.

CpG island

A genomic region which contains a high content of cytosine (C) and guanine (G) dinucleotides (the 'p' refers to the phosphodiester bond linking the two bases). CpG islands are found in many mammalian promoters, and unlike scattered CpGs throughout the genome, which are usually hypermethylated, promoter CpG islands are normally hypomethylated.

Helicase

A protein that can unwind DNA or RNA.

Teratoma

A confined tumour, originating from pluripotent cells, that includes tissues of the three germ layers, endoderm, mesoderm and ectoderm.

Telomeric region

A region of repetitive DNA at the ends of chromosomes that protects the chromosomes from premature deterioration, rearrangements and chromosome fusion.

Histone hyperacetylation

A state in which many Lys residues are acetylated on many of the histones present in a given region of chromatin.

Genetic epistasis

The relationship or order in which two genes act in a pathway (that is, upstream or downstream, synergistic or antagonistic), which can be studied by analysing single and double mutants.

DamID

A method that is used to analyse binding of proteins to DNA. Genetically modified Drosophila melanogaster culture cell lines express a protein of interest fused with a bacterial DNA adenine methyltransferase. Local DNA methyltransferase activity indicates protein binding.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Gaspar-Maia, A., Alajem, A., Meshorer, E. et al. Open chromatin in pluripotency and reprogramming. Nat Rev Mol Cell Biol 12, 36–47 (2011). https://doi.org/10.1038/nrm3036

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrm3036

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