Skip to main content
The Journal of Experimental Medicine logoLink to The Journal of Experimental Medicine
. 2010 Oct 25;207(11):2287–2295. doi: 10.1084/jem.20101438

Epigenetic control of embryonic stem cell fate

Nicolaj Strøyer Christophersen 1,2, Kristian Helin 1,2,
PMCID: PMC2964577  PMID: 20975044

In embryonic stem cells, alterations in chromatin modifications influence the balance between self-renewal and differentiation.

Abstract

Embryonic stem (ES) cells are derived from the inner cell mass of the preimplantation embryo and are pluripotent, as they are able to differentiate into all cell types of the adult organism. Once established, the pluripotent ES cells can be maintained under defined culture conditions, but can also be induced rapidly to differentiate. Maintaining this balance of stability versus plasticity is a challenge, and extensive studies in recent years have focused on understanding the contributions of transcription factors and epigenetic enzymes to the “stemness” properties of these cells. Identifying the molecular switches that regulate ES cell self-renewal versus differentiation can provide insights into the nature of the pluripotent state and enhance the potential use of these cells in therapeutic applications. Here, we review the latest models for how changes in chromatin methylation can modulate ES cell fate, focusing on two major repressive pathways, Polycomb group (PcG) repressive complexes and promoter DNA methylation.


Embryonic stem (ES) cells have the ability to self-renew, producing daughter cells with equivalent developmental potential, or to differentiate into more specialized cells. ES cells are derived from the inner cell mass of the preimplantation embryo and are pluripotent, as they are able to differentiate into cells of the three germ layers, both in vitro and in vivo (Evans and Kaufman, 1981; Thomson et al., 1998). Three transcription factors, OCT4, SOX2, and NANOG, cooperate to ensure the self-renewal and pluripotency of ES cells (Boyer et al., 2005; Loh et al., 2006). These factors are highly expressed in undifferentiated ES cells and physically interact with each other in large protein complexes (Wang et al., 2006; van den Berg et al., 2010). OCT4, SOX2, and NANOG are transcriptionally interconnected and co-occupy promoters of actively transcribed genes that promote ES cell self-renewal such as KLF4 (Boyer et al., 2005; Loh et al., 2006; Kim et al., 2008). They also occupy genes encoding a large set of developmental regulators that are silent in ES cells, but whose expression is associated with lineage commitment and cellular differentiation (Fig. 1; Boyer et al., 2005; Loh et al., 2006).

Figure 1.

Figure 1.

ES cell self-renewal and differentiation. ES cells have the ability to self-renew or differentiate into cells of all three germ layers (endoderm, ectoderm, and mesoderm). In ES cells, OCT4, SOX2, and NANOG form a core transcriptional network influencing the stem cell self-renewal machinery. Several hundred target genes co-occupied by OCT4, SOX2, and NANOG can be classified into two groups of downstream genes exerting opposing functions. One group includes actively transcribed genes associated with proliferation and transcription factors necessary to maintain the ES cell state. The other group includes transcriptionally silent genes encoding developmental regulators that are only activated as cells differentiate and commit to particular lineages.

The fact that the three key regulators can activate some genes and repress others is thought to be caused by the chromatin packaging in ES cells that are regulated by epigenetic factors. Epigenetics refers to heritable changes in gene expression that are independent of nucleotide sequence. This is achieved by regulating gene activity through alterations of chromatin structure, such as posttranslational modifications of the histones and DNA methylation, which can be either permissive or restrictive for transcription. These changes are catalyzed by histone and DNA modification enzymes that work in coordination to coregulate the balance between pluripotency and lineage-specific differentiation. The question is how these multiple regulatory mechanisms are coordinated to control the transcriptional state of pluripotent versus developmental genes in ES cells and during in vitro differentiation of ES cells. In this Review, we describe recent advances in the understanding of the role of the repressive epigenetic marks deposited by Polycomb group (PcG) repressive complexes and DNA methyl transferases (DNMTs) on ES cell self-renewal and differentiation.

PcG proteins in ES cells

Genome-wide approaches revealed that the transcriptionally silent developmental genes targeted by OCT4, SOX2, and NANOG in ES cells are also occupied by PcG repressive complexes (Boyer et al., 2006; Lee et al., 2006). PcG proteins facilitate maintenance of cell states through gene silencing and were first identified in Drosophila as a result of their essential roles as repressors of body patterning genes, including homeobox (HOX) genes during fruit fly development (Lewis, 1978; Struhl, 1981). HOX genes are expressed in distinct domains along the body axis and act to give cells of diverse tissues their unified regional cell identities.

PcG proteins act as multimers in two main protein complexes, Polycomb repressive complex 1 (PRC1) and PRC2, which are discriminated according to their biochemical functions and compositions. The three PcG proteins EED, SUZ12, and EZH2 are part of PRC2, which catalyzes di- and trimethylation of histone H3 lysine 27 (H3K27me2/me3; Czermin et al., 2002; Kuzmichev et al., 2002; Müller et al., 2002). Although the direct function of H3K27me3 is not fully understood, the modification correlates with transcriptional repression, and it can function as a docking site for PRC1, which catalyzes the monoubiquitinylation of histone H2A at lysine 119 (H2AK119Ub1; Cao et al., 2002; Min et al., 2003; de Napoles et al., 2004). In vitro studies suggested that H2AK119Ub1 blocks RNA polymerase II activity, thereby leading to transcriptional repression (Stock et al., 2007). Interestingly, the mechanism by which PRC1 represses transcription does not involve a block of RNA polymerase II recruitment to target genes, but rather attenuation of RNA polymerase II elongation (Stock et al., 2007). The H3K27me3 mark can also attract PRC2 itself, and the H3K27me3 mark is heritably transmitted to daughter cells as a self-perpetuating mark to maintain specific gene expression programs (Hansen et al., 2008; Margueron et al., 2009). Additionally, both PRC1 and PRC2 can mediate repression by direct chromatin compaction as a result of alternative subunit compositions (Francis et al., 2004; Margueron et al., 2008; Eskeland et al., 2010), and PRC1-mediated H2AK119Ub1 can occur in the absence of H3K27me3 (Leeb et al., 2010), suggesting several alternative mechanisms leading to PcG-mediated repression.

The PcG proteins are required for early mammalian embryo development (O’Carroll et al., 2001; Wang et al., 2002; Voncken et al., 2003; Pasini et al., 2004), but not for maintaining ES cell pluripotency. PcG mutant ES cells can still self-renew, maintain normal morphology, and express OCT4, SOX2, and NANOG (Pasini et al., 2007; Chamberlain et al., 2008a; Leeb et al., 2010). Moreover, although the PcG knockout ES cells do not differentiate efficiently into the three germ layers, they can still contribute to their formation, in vivo and in vitro (Pasini et al., 2007; Chamberlain et al., 2008b; Leeb et al., 2010). However, loss of individual PRC components in ES cells does lead to marginally increased expression of various lineage-affiliated genes and unscheduled differentiation (Morin-Kensicki et al., 2001; Pasini et al., 2007; Chamberlain et al., 2008b), an effect that is even more pronounced in ES cells carrying targeted deletions of both PRC1 and PRC2 genes (Leeb et al., 2010).

Genome-wide studies of PRC1 and PRC2 in ES cells have shown that they target promoters of >2,000 genes, of which a large subset overlaps with target genes of OCT4, NANOG, and SOX2 (Boyer et al., 2006; Lee et al., 2006). Further studies on the chromatin landscape in ES cells revealed that virtually all of these sites of PcG activity contain large regions of the repressive H3K27me3 modification and are strongly enriched in the activation-associated H3 lysine 4 trimethylation (H3K4me3) mark around the transcriptional start site (Bernstein et al., 2006; Guenther et al., 2007; Mikkelsen et al., 2007; Zhao et al., 2007). This mark is mediated through the SETD1 and/or MLL methyltransferase complexes (Miller et al., 2001; Roguev et al., 2001; Milne et al., 2002; Nagy et al., 2002), which are chromatin-activating factors that generally antagonize PcG silencing.

These genomic regions with opposing modifications have been termed “bivalent domains” and are proposed to silence developmental regulators while keeping them “poised.” Bivalent domains are not, however, exclusive to pluripotent cells; they can also be found in cells of restricted potency, including T cells and fibroblasts, where genes are unlikely to be poised in preparation for subsequent activation (Roh et al., 2006; Barski et al., 2007; Mikkelsen et al., 2007; Pan et al., 2007; Zhao et al., 2007; Cui et al., 2009). However, whether bivalent domains are functionally important is thus far not known. Genes carrying H3K4me3 modifications are highly common among ES cells and T cells, whereas H3K27me3 modification patterns do not overlap between ES cells and T cells (Zhao et al., 2007), indicating that PcG-mediated histone methylations show a higher level of tissue specificity than H3K4me3 marks. Interestingly, recent data indicate that the presence of H3K4me3 mark might not itself be indicative or predictive of transcriptional activity (Guenther et al., 2007), and that the H3K4me3 mark is found on genes in the absence of sequence-specific activators in ES cells and without the stable association of RNA polymerase II (Thomson et al., 2010; Vastenhouw et al., 2010); thus, the significance of genome-wide marking of promoters by H3K4me3 is not clear at this moment. This means that the ability of the bivalent mark to silence lineage-specific gene expression is likely caused by the dominant effect of the H3K27me3 over the H3K4me3 mark, yet preserves the potential for rapid gene activation upon differentiation of stimuli-induced removal of the H3K27me3 mark.

Cell fate specifications in mammals permit the formation of ∼200 different cell types consistent with a multitude of different combinations of the epigenome in mammalian development. ES cellular differentiation entails loss of pluripotency and parallel gain of first lineage-specific and, ultimately, cell-type specific characteristics. The process of tissue fate specification is initiated by signaling molecules that drive the dynamic equilibrium of ES cells toward a particular lineage.

As some results suggest that the repressive H3K27me3 mark can be heritably transmitted to daughter cells to maintain specific gene expression programs (Hansen et al., 2008; Margueron et al., 2009), the expression of developmentally regulated genes would require the removal of the H3K27me3 mark. The UTX and JMJD3 genes encode H3K27me3/me2 demethylases, suggesting a mechanism by which PcG-repressed promoters are activated (Agger et al., 2007; De Santa et al., 2007; Lan et al., 2007; Lee et al., 2007b). For example, UTX is absent from the transcriptionally silent HOXA9 gene promoter in ES cells, but is present in this region in primary fibroblasts, in which HOXA9 gene is expressed (Lan et al., 2007). In embryonic carcinoma cells undergoing differentiation, the removal of the H3K27me3 mark to permit the activation of gene expression is associated with recruitment of the UTX histone demethylase to promoters during the transcriptional activation (Agger et al., 2007). Interestingly, UTX is associated with the H3K4 methyltransferases MLL2–4, whereas JMJD3 has not been reported to bind to any of the MLL proteins or the H3K4 methyltransferase SETD1 (Agger et al., 2007; Cho et al., 2007; Issaeva et al., 2007; Lee et al., 2007a). In a similar fashion, PRC2 can interact with and recruit a histone H3K4 demethylase (Pasini et al., 2008), thereby coordinating the removal of an “active” mark with the deposition of a repressive mark.

Importantly, PcG proteins are also recruited to promoters of non–ES cell–specified PcG target genes in response to differentiation signals, and this recruitment is required for the silencing of these genes during differentiation (Pasini et al., 2007; Mohn et al., 2008; Oktaba et al., 2008; Ezhkova et al., 2009). Moreover, some ES cell–specific genes such as NANOG are marked by H3K27me3 during differentiation (Pan et al., 2007; Hawkins et al., 2010). This means that when cell lineage commitment occurs, pluripotency transcription factors are silenced, whereas the appropriate regulators of development lose PcG-mediated repression and are activated (Fig. 2). By this mechanism, which reduces the likelihood of inappropriate reactivation of stem cell–specific or lineage-unrelated genes, the PcG repressive complexes are believed to contribute to the robustness of differentiation. In support of this view, PcG-deficient ES cells can enter differentiation, but fail to maintain the differentiated phenotype (Pasini et al., 2007; Leeb et al., 2010).

Figure 2.

Figure 2.

Dynamic recruitment of PcG proteins to chromatin during lineage specification. In ES cells, differentiation and development-promoting genes (Dev. A, B, and C) are repressed by bivalent domains, whereas late differentiation genes are not marked by H3K27me3, but not expressed. Pluripotency genes such as OCT4 are methylated at H3K4 and expressed. Differentiation signals generate cells committed to various somatic lineages, and activate lineage-specific genes that lose the repressive H3K27me3 mark (Dev. A). However, many genes preserve the bivalent domains and are not expressed (Dev. B and C); a few of these genes (e.g., those that are selectively expressed in other somatic cell lineages) also gain promoter DNA methylation during lineage commitment to ensure silencing (Dev. C). Late differentiation genes become marked by H3K27 in a manner dependent on the particular committed cell type, resulting in the formation of new bivalent domains that may be resolved in more mature differentiated cells. Examples of the aforementioned dynamics during neuronal differentiation of ES cells are NEUROG1, encoding for a neurogenic transcription factor (Dev. A); GATA4, encoding for an endodermal marker (Dev. B); TPARP, encoding for a germline-specific polyadenylate polymerase (Dev. C); and SCN1B, encoding for a neuronal voltage-gated sodium channel (late diff. gene; Mohn et al., 2008). The population of de novo DNA-methylated genes is also enriched in pluripotency-specific genes such as OCT4, ensuring the stable repression of transcripts that are required for ES cell maintenance.

A key question of importance for ES cell self-renewal and differentiation is how PcG binding to chromatin is regulated to ensure proper PcG recruitment to and dissociation from chromatin. Because the PcG proteins themselves do not have the ability to bind DNA-specific sites, recruitment is believed to require the interaction with sequence-specific transcription factors. The availability of these transcription factors or competing transcription factors may be involved in regulating the sustained binding of PcG proteins to their target genes. In Drosophila, the PcG proteins are recruited to Polycomb repressive elements (PREs). These elements are stretches of DNA of >1,000 bp containing DNA-binding sites for different transcription factors. Several Drosophila transcription factors have been shown to bind to the PREs and to be required for the recruitment of the Drosophila PcG proteins to target genes. Importantly, however, the results obtained so far suggest that no single transcription factor is sufficient to recruit the PcG proteins, but that a combination of transcription factors is required. Moreover, only one of the PRE-associated Drosophila transcription factors is conserved in mammalian cells (YY1 in mammals, PHO in Drosophila), and YY1 binding in mammals does not overlap with PcG target genes (Squazzo et al., 2006).

PREs have not been mapped in mammalian cells, despite genome-wide PRC-binding maps (Boyer et al., 2006; Bracken et al., 2006; Lee et al., 2006), though a region between two HOX genes might target PcG proteins to a reporter gene in human cells (Woo et al., 2010). In addition to YY1, it has been speculated that OCT4 could be involved in PcG recruitment (Wang et al., 2006; van den Berg et al., 2010), though these studies have not been independently confirmed.

Recent results have, however, shed new light on PcG protein recruitment in ES cells (Fig. 3). Five independent studies identified the Jumonji C (JmjC) protein JARID2 as a novel component of PRC2, displaying a significant overlap with PRC2 target sites in ES cells (Peng et al., 2009; Shen et al., 2009; Landeira et al., 2010; Li et al., 2010; Pasini et al., 2010). JmjC domains are characteristic of lysine demethylases, but JARID2 is unusual among the JmjC proteins in that it lacks crucial residues for cofactor binding and is catalytically inactive (Cloos et al., 2008). JARID2, which can bind directly to DNA, cofractions with PRC2 in high molecular weight complexes, and localization of PRC2 to its respective target sites is dependent on JARID2. It remains unclear whether JARID2 regulates the intrinsic histone methyltransferase activity of the PRC2 complex because the five different studies reached different conclusions on this point and major global reductions in H3K27me3 levels were not detected in ES cells lacking JARID2. Importantly, however, JARID2 was shown to be required for the execution of differentiation pathways in ES cells (Shen et al., 2009; Landeira et al., 2010; Li et al., 2010; Pasini et al., 2010). In addition to lack of recruitment of PRC2, this effect might be caused by a role of JARID2 in recruiting PRC1 and poised RNA polymerase II to PcG target genes (Landeira et al., 2010).

Figure 3.

Figure 3.

Potential mechanisms of PRC2 recruitment to target genes. PRC2 is recruited to target genes by a combination of transcription factors and ncRNAs. A fraction of PRC2 associates with JARID2, which is required for PRC2 binding to its target genes in ES cells. JARID2 might therefore represent a core component of PRC2 in ES cells, although other PRC2 complexes exist; these contain MTF2, PHF1, and other uncharacterized factors that could represent alternative targeting mechanisms operative both during ES cell self-renewal and differentiation. Sequence-specific transcription factors (TF) and ncRNA might also recruit the PRC2 core complex to target genes during differentiation. Finally, PcG target genes are CpG-rich, and proteins binding to CpG elements such as TET1 or the histone demethylase FBXL10 might have a role in recruiting Polycomb to target genes. For additional information on the components of PRC1 and PRC2 complexes, see Morey and Helin (2010).

PRC2 also interacts with the protein MTF2 (also named PCL2; Shen et al., 2009; Li et al., 2010). MTF2 is one of three mammalian homologues of Drosophila Polycomb-like, which can stimulate PRC2 activity in Drosophila (Nekrasov et al., 2007). Interestingly, MTF2 was recently found in a screen for novel regulators of ES cell self-renewal and shown to be involved in recruiting PRC2 to a subset of PcG target genes in ES cells (Walker et al., 2010). Whether JARID2 and MTF2 are both found in a subset of PRC2 complexes is unclear, but coimmunoprecipitation studies indicate that JARID2 and MTF2 can reside in distinct PCR2 subcomplexes. These results highlight the importance of dissecting the composition and activities of different PRC2 complexes and their associated “recruiters” to fully understand how PcG proteins control thousands of genes in many different cellular and developmental contexts. For example, JARID2 expression declines rapidly as ES cells differentiate (Boyer et al., 2005), and other proteins must therefore regulate recruitment of PRC2 to target genes during differentiation. Strong candidates for this include PHF1/PCL1, PHF19/PCL3, AEBP2, and NIPP1, which have all been previously associated with PRC2 (Cao et al., 2008; Nuytten et al., 2008; Sarma et al., 2008; Kim et al., 2009).

Recent advances suggest that noncoding RNAs (ncRNAs) also play a role in the recruitment of PcG complexes in ES cells. DNA microarray analysis showed that short ncRNAs (<200 nt) were transcribed from the 5′ end of several hundred PcG target genes in ES cells (Kanhere et al., 2010). Interestingly, these ncRNAs interact with PRC2 and are involved in stabilizing PRC2 association with chromatin. Moreover, the ncRNAs were depleted from PcG target genes that are derepressed during cell differentiation (Kanhere et al., 2010). This indicates that short ncRNAs might function as the interface between DNA and specific chromatin remodeling activities, though the importance of direct base pairing at specific sequence motifs is still unknown. Long ncRNAs (lincRNAs; transcripts longer than 200 nt) have also been reported to recruit PRC2 to specific targets, including the HOXD cluster and the inactive X chromosome (Rinn et al., 2007; Pandey et al., 2008; Zhao et al., 2008). In addition, lincRNAs are now emerging as a new class of noncoding transcripts, with >3,000 members of which 20% might associate with PRC2 (Guttman et al., 2009; Khalil et al., 2009). Importantly, lincRNAs show dynamic patterns of expression with suggested roles in cell fate choices, but the mechanisms by which these trans-acting ncRNAs recruit PcG complexes to specific sites is not fully understood. Interestingly, the lincRNA HOTAIR was recently shown to target both PRC2 and the H3K4me2 demethylase LSD1 to hundreds of genes thereby acting as a modular scaffold (Tsai et al., 2010).

Collectively, these studies demonstrate that both DNA-binding factors and ncRNAs can guide de novo histone modification by PcG proteins and establish repressive chromatin domains.

DNA methylation in ES cells

A second major repressive epigenetic pathway is mediated by DNMTs. Methylation of cytosine residues in CpG dinucleotides in promoter regions catalyzed by DNMTs is an important epigenetic mark that maintains long-term repression by controlling DNA accessibility (Bird, 2002). DNMTs are essential for embryonic development (Okano et al., 1999), whereas they are not required for self-renewal or genomic integrity of ES cells (Tsumura et al., 2006). This probably reflects the fact that blastocysts undergo global demethylation immediately before the derivation of ES cells from blastocysts (Mayer et al., 2000).

Promoters can be classified according to whether they contain high or low CpG content within a certain region around the transcriptional start sites. These distinct promoter classes show differences in their methylation levels and in the mechanisms through which they are regulated. In ES cells, high CpG promoters have low DNA methylation levels, whereas low CpG promoters have relatively high DNA methylation levels (Fouse et al., 2008; Meissner et al., 2008; Mohn et al., 2008). CpG-rich promoters—almost by default—appear to associate with nucleosomes carrying the H3K4me3 mark. Some of these regulate constitutively expressed housekeeping genes, but others corresponding to developmental regulators also contain the H3K27me3 mark. This observation raises the question of why these poised, yet inactive, CpG-rich promoters are protected from DNA methylation. Recent data showed that the CPF1 protein has affinity for nonmethylated CpG islands, and because CPF1 associates with the H3K4 methyltransferase SETD1 it leads to H3K4 trimethylation of CpG-rich promoters (Thomson et al., 2010). As some DNMT domain–containing proteins bind to unmethylated, but not methylated, H3K4 and recruit DNMTs (Ooi et al., 2007), this SETD1-induced H3K4 methylation may protect both active and inactive CpG-rich promoters from DNA methylation. Methylated low CpG promoters are marked neither by H3K4me3 nor by H3K27me3 and are mostly repressed in ES cells (Fouse et al., 2008; Meissner et al., 2008; Mohn et al., 2008).

The conversion of ES cells into somatic cells only leads to modest changes in DNA methylation at promoter regions and the majority of promoters maintain their methylation levels during differentiation (Lagarkova et al., 2006; Fouse et al., 2008; Meissner et al., 2008; Mohn et al., 2008; Hawkins et al., 2010). These observations question the importance of DNA methylation as a controlling element of cellular differentiation. However, a small number of genes do become methylated during development, an event that is accompanied by silencing of the associated promoters (Lagarkova et al., 2006; Fouse et al., 2008; Meissner et al., 2008; Mohn et al., 2008; Hawkins et al., 2010). Many of the identified targets of differentiation-coupled de novo methylation are promoters of stem cell–specific genes, including those encoding pluripotency transcription factors (Lagarkova et al., 2006; Fouse et al., 2008; Meissner et al., 2008; Mohn et al., 2008; Hawkins et al., 2010). DNA methylation is thought to keep such genes silent in differentiated cells and prevent their aberrant reactivation and the risk of de-differentiation (Fig. 2). In line with this hypothesis, in vitro differentiation of DNMT mutant ES cells leads to rapid apoptotic cell death (Panning and Jaenisch, 1996), reflecting the importance of DNA methylation in differentiated cells, which is also described for other cell types (Jackson-Grusby et al., 2001).

As a consequence it is a prerequisite that some promoters of stem cell genes are demethylated during reprogramming to enable reacquisition of pluripotency (Simonsson and Gurdon, 2004; Takahashi and Yamanaka, 2006). This demethylation may occur through a DNA repair process (Bhutani et al., 2010; Hajkova et al., 2010), or involve the recently isolated enzymes of the TET family that can convert methylated cytosine to hydroxymethylated cytosine (Tahiliani et al., 2009).

Concluding remarks

Progress in the past few years has greatly enhanced our understanding of epigenetic control of ES cells. Epigenetic factors appear to be essential for regulating cell fate decisions and maintaining the cellular state of differentiated cells. Surprisingly, however, the results so far have suggested that epigenetic regulation may be dispensable for maintaining ES cell identity. Epigenetic mechanisms of gene silencing contribute to the overall stability of pluripotency, but are downstream in this setting. Thus, ES cell identity might primarily be regulated by transcription factors, whereas epigenetic chromatin-based repressive and activating modifiers may serve transcriptional corepressor and co-activating functions in this process. Nevertheless, the stability of a given cell state relies on the silencing of genes encoding inducers of other cell states. For example, expression of only a few transcription factors that produce cell type–specific gene expression programs can cause cells to adopt new states (Vierbuchen et al., 2010). By setting thresholds, PcG proteins may help prevent inappropriate gene expression induced by weak activating signals and thereby limit noise-induced errors and neutralize extrinsic perturbations. DNA methylation is used to lower the chance of spurious activation, which is more likely to occur in a bivalent chromatin environment.

Our understanding of how different regulatory networks are activated to guide lineage commitment is still limited. For example, the identity of differentiation signals that lead to selective activation of genes encoding specific developmental regulators by turning off PcG repression and recruiting UTX or JMD3 remains unknown. Moreover, mechanisms of recruiting PcG proteins to PcG targets in a developmental stage-specific context by transcription factors remain to be dissected. Further characterization of epigenetic regulation will help to provide a more comprehensive view of the molecular mechanisms that govern the balance between self-renewal and lineage commitment. This advance may lead to improved protocols for directing the differentiation of ES cells toward particular lineages and for the generation of ES-like cells from somatic cells.

Acknowledgments

We apologize to those researchers whose work could not be cited due to space constraints.

The work in the Helin laboratory is supported by grants from the Danish National Research Foundation, the Danish Cancer Society, the Lundbeck Foundation, the Novo Nordisk Foundation, the Villum Kann Rasmussen foundation, the Danish Medical Research Council, and the Excellence Program of the University of Copenhagen.

Footnotes

Abbreviations used:

DNMT
DNA methyl transferase
ES
embryonic stem
JmjC
Jumonji C
lincRNA
long ncRNA
ncRNA
noncoding RNA
PcG
Polycomb group
PRC
Polycomb repressive complex
PRE
Polycomb repressive element

References

  1. Agger K., Cloos P.A., Christensen J., Pasini D., Rose S., Rappsilber J., Issaeva I., Canaani E., Salcini A.E., Helin K. 2007. UTX and JMJD3 are histone H3K27 demethylases involved in HOX gene regulation and development. Nature. 449:731–734 10.1038/nature06145 [DOI] [PubMed] [Google Scholar]
  2. Barski A., Cuddapah S., Cui K., Roh T.Y., Schones D.E., Wang Z., Wei G., Chepelev I., Zhao K. 2007. High-resolution profiling of histone methylations in the human genome. Cell. 129:823–837 10.1016/j.cell.2007.05.009 [DOI] [PubMed] [Google Scholar]
  3. Bernstein B.E., Mikkelsen T.S., Xie X., Kamal M., Huebert D.J., Cuff J., Fry B., Meissner A., Wernig M., Plath K., et al. 2006. A bivalent chromatin structure marks key developmental genes in embryonic stem cells. Cell. 125:315–326 10.1016/j.cell.2006.02.041 [DOI] [PubMed] [Google Scholar]
  4. Bhutani N., Brady J.J., Damian M., Sacco A., Corbel S.Y., Blau H.M. 2010. Reprogramming towards pluripotency requires AID-dependent DNA demethylation. Nature. 463:1042–1047 10.1038/nature08752 [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Bird A. 2002. DNA methylation patterns and epigenetic memory. Genes Dev. 16:6–21 10.1101/gad.947102 [DOI] [PubMed] [Google Scholar]
  6. Boyer L.A., Lee T.I., Cole M.F., Johnstone S.E., Levine S.S., Zucker J.P., Guenther M.G., Kumar R.M., Murray H.L., Jenner R.G., et al. 2005. Core transcriptional regulatory circuitry in human embryonic stem cells. Cell. 122:947–956 10.1016/j.cell.2005.08.020 [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Boyer L.A., Plath K., Zeitlinger J., Brambrink T., Medeiros L.A., Lee T.I., Levine S.S., Wernig M., Tajonar A., Ray M.K., et al. 2006. Polycomb complexes repress developmental regulators in murine embryonic stem cells. Nature. 441:349–353 10.1038/nature04733 [DOI] [PubMed] [Google Scholar]
  8. Bracken A.P., Dietrich N., Pasini D., Hansen K.H., Helin K. 2006. Genome-wide mapping of Polycomb target genes unravels their roles in cell fate transitions. Genes Dev. 20:1123–1136 10.1101/gad.381706 [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Cao R., Wang L., Wang H., Xia L., Erdjument-Bromage H., Tempst P., Jones R.S., Zhang Y. 2002. Role of histone H3 lysine 27 methylation in Polycomb-group silencing. Science. 298:1039–1043 10.1126/science.1076997 [DOI] [PubMed] [Google Scholar]
  10. Cao R., Wang H., He J., Erdjument-Bromage H., Tempst P., Zhang Y. 2008. Role of hPHF1 in H3K27 methylation and Hox gene silencing. Mol. Cell. Biol. 28:1862–1872 10.1128/MCB.01589-07 [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Chamberlain S.J., Yee D., Magnuson T. 2008a. Polycomb repressive complex 2 is dispensable for maintenance of embryonic stem cell pluripotency. Stem Cells. 26:1496–1505 10.1634/stemcells.2008-0102 [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Chamberlain S.J., Yee D., Magnuson T. 2008b. Polycomb repressive complex 2 is dispensable for maintenance of embryonic stem cell Pluripotency. Stem Cells. 26:1495–1505 [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Cho Y.W., Hong T., Hong S., Guo H., Yu H., Kim D., Guszczynski T., Dressler G.R., Copeland T.D., Kalkum M., Ge K. 2007. PTIP associates with MLL3- and MLL4-containing histone H3 lysine 4 methyltransferase complex. J. Biol. Chem. 282:20395–20406 10.1074/jbc.M701574200 [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Cloos P.A., Christensen J., Agger K., Helin K. 2008. Erasing the methyl mark: histone demethylases at the center of cellular differentiation and disease. Genes Dev. 22:1115–1140 10.1101/gad.1652908 [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Cui K., Zang C., Roh T.Y., Schones D.E., Childs R.W., Peng W., Zhao K. 2009. Chromatin signatures in multipotent human hematopoietic stem cells indicate the fate of bivalent genes during differentiation. Cell Stem Cell. 4:80–93 10.1016/j.stem.2008.11.011 [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Czermin B., Melfi R., McCabe D., Seitz V., Imhof A., Pirrotta V. 2002. Drosophila enhancer of Zeste/ESC complexes have a histone H3 methyltransferase activity that marks chromosomal Polycomb sites. Cell. 111:185–196 10.1016/S0092-8674(02)00975-3 [DOI] [PubMed] [Google Scholar]
  17. de Napoles M., Mermoud J.E., Wakao R., Tang Y.A., Endoh M., Appanah R., Nesterova T.B., Silva J., Otte A.P., Vidal M., et al. 2004. Polycomb group proteins Ring1A/B link ubiquitylation of histone H2A to heritable gene silencing and X inactivation. Dev. Cell. 7:663–676 10.1016/j.devcel.2004.10.005 [DOI] [PubMed] [Google Scholar]
  18. De Santa F., Totaro M.G., Prosperini E., Notarbartolo S., Testa G., Natoli G. 2007. The histone H3 lysine-27 demethylase Jmjd3 links inflammation to inhibition of polycomb-mediated gene silencing. Cell. 130:1083–1094 10.1016/j.cell.2007.08.019 [DOI] [PubMed] [Google Scholar]
  19. Eskeland R., Leeb M., Grimes G.R., Kress C., Boyle S., Sproul D., Gilbert N., Fan Y., Skoultchi A.I., Wutz A., Bickmore W.A. 2010. Ring1B compacts chromatin structure and represses gene expression independent of histone ubiquitination. Mol. Cell. 38:452–464 10.1016/j.molcel.2010.02.032 [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Evans M.J., Kaufman M.H. 1981. Establishment in culture of pluripotential cells from mouse embryos. Nature. 292:154–156 10.1038/292154a0 [DOI] [PubMed] [Google Scholar]
  21. Ezhkova E., Pasolli H.A., Parker J.S., Stokes N., Su I.H., Hannon G., Tarakhovsky A., Fuchs E. 2009. Ezh2 orchestrates gene expression for the stepwise differentiation of tissue-specific stem cells. Cell. 136:1122–1135 10.1016/j.cell.2008.12.043 [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Fouse S.D., Shen Y., Pellegrini M., Cole S., Meissner A., Van Neste L., Jaenisch R., Fan G. 2008. 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 10.1016/j.stem.2007.12.011 [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Francis N.J., Kingston R.E., Woodcock C.L. 2004. Chromatin compaction by a polycomb group protein complex. Science. 306:1574–1577 10.1126/science.1100576 [DOI] [PubMed] [Google Scholar]
  24. Guenther M.G., Levine S.S., Boyer L.A., Jaenisch R., Young R.A. 2007. A chromatin landmark and transcription initiation at most promoters in human cells. Cell. 130:77–88 10.1016/j.cell.2007.05.042 [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Guttman M., Amit I., Garber M., French C., Lin M.F., Feldser D., Huarte M., Zuk O., Carey B.W., Cassady J.P., et al. 2009. Chromatin signature reveals over a thousand highly conserved large non-coding RNAs in mammals. Nature. 458:223–227 10.1038/nature07672 [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Hajkova P., Jeffries S.J., Lee C., Miller N., Jackson S.P., Surani M.A. 2010. Genome-wide reprogramming in the mouse germ line entails the base excision repair pathway. Science. 329:78–82 10.1126/science.1187945 [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Hansen K.H., Bracken A.P., Pasini D., Dietrich N., Gehani S.S., Monrad A., Rappsilber J., Lerdrup M., Helin K. 2008. A model for transmission of the H3K27me3 epigenetic mark. Nat. Cell Biol. 10:1291–1300 10.1038/ncb1787 [DOI] [PubMed] [Google Scholar]
  28. Hawkins R.D., Hon G.C., Lee L.K., Ngo Q., Lister R., Pelizzola M., Edsall L.E., Kuan S., Luu Y., Klugman S., et al. 2010. Distinct epigenomic landscapes of pluripotent and lineage-committed human cells. Cell Stem Cell. 6:479–491 10.1016/j.stem.2010.03.018 [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Issaeva I., Zonis Y., Rozovskaia T., Orlovsky K., Croce C.M., Nakamura T., Mazo A., Eisenbach L., Canaani E. 2007. Knockdown of ALR (MLL2) reveals ALR target genes and leads to alterations in cell adhesion and growth. Mol. Cell. Biol. 27:1889–1903 10.1128/MCB.01506-06 [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Jackson-Grusby L., Beard C., Possemato R., Tudor M., Fambrough D., Csankovszki G., Dausman J., Lee P., Wilson C., Lander E., Jaenisch R. 2001. Loss of genomic methylation causes p53-dependent apoptosis and epigenetic deregulation. Nat. Genet. 27:31–39 10.1038/83730 [DOI] [PubMed] [Google Scholar]
  31. Kanhere A., Viiri K., Araújo C.C., Rasaiyaah J., Bouwman R.D., Whyte W.A., Pereira C.F., Brookes E., Walker K., Bell G.W., et al. 2010. Short RNAs are transcribed from repressed polycomb target genes and interact with polycomb repressive complex-2. Mol. Cell. 38:675–688 10.1016/j.molcel.2010.03.019 [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Khalil A.M., Guttman M., Huarte M., Garber M., Raj A., Rivea Morales D., Thomas K., Presser A., Bernstein B.E., van Oudenaarden A., et al. 2009. Many human large intergenic noncoding RNAs associate with chromatin-modifying complexes and affect gene expression. Proc. Natl. Acad. Sci. USA. 106:11667–11672 10.1073/pnas.0904715106 [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Kim J., Chu J., Shen X., Wang J., Orkin S.H. 2008. An extended transcriptional network for pluripotency of embryonic stem cells. Cell. 132:1049–1061 10.1016/j.cell.2008.02.039 [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Kim H., Kang K., Kim J. 2009. AEBP2 as a potential targeting protein for Polycomb Repression Complex PRC2. Nucleic Acids Res. 37:2940–2950 10.1093/nar/gkp149 [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Kuzmichev A., Nishioka K., Erdjument-Bromage H., Tempst P., Reinberg D. 2002. Histone methyltransferase activity associated with a human multiprotein complex containing the Enhancer of Zeste protein. Genes Dev. 16:2893–2905 10.1101/gad.1035902 [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Lagarkova M.A., Volchkov P.Y., Lyakisheva A.V., Philonenko E.S., Kiselev S.L. 2006. Diverse epigenetic profile of novel human embryonic stem cell lines. Cell Cycle. 5:416–420 [DOI] [PubMed] [Google Scholar]
  37. Lan F., Bayliss P.E., Rinn J.L., Whetstine J.R., Wang J.K., Chen S., Iwase S., Alpatov R., Issaeva I., Canaani E., et al. 2007. A histone H3 lysine 27 demethylase regulates animal posterior development. Nature. 449:689–694 10.1038/nature06192 [DOI] [PubMed] [Google Scholar]
  38. Landeira D., Sauer S., Poot R., Dvorkina M., Mazzarella L., Jørgensen H.F., Pereira C.F., Leleu M., Piccolo F.M., Spivakov M., et al. 2010. Jarid2 is a PRC2 component in embryonic stem cells required for multi-lineage differentiation and recruitment of PRC1 and RNA Polymerase II to developmental regulators. Nat. Cell Biol. 12:618–624 10.1038/ncb2065 [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Lee T.I., Jenner R.G., Boyer L.A., Guenther M.G., Levine S.S., Kumar R.M., Chevalier B., Johnstone S.E., Cole M.F., Isono K., et al. 2006. Control of developmental regulators by Polycomb in human embryonic stem cells. Cell. 125:301–313 10.1016/j.cell.2006.02.043 [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Lee M.G., Villa R., Trojer P., Norman J., Yan K.P., Reinberg D., Di Croce L., Shiekhattar R. 2007a. Demethylation of H3K27 regulates polycomb recruitment and H2A ubiquitination. Science. 318:447–450 10.1126/science.1149042 [DOI] [PubMed] [Google Scholar]
  41. Lee M.G., Villa R., Trojer P., Norman J., Yan K.P., Reinberg D., Di Croce L., Shiekhattar R. 2007b. Demethylation of H3K27 regulates polycomb recruitment and H2A ubiquitination. Science. 318:447–450 10.1126/science.1149042 [DOI] [PubMed] [Google Scholar]
  42. Leeb M., Pasini D., Novatchkova M., Jaritz M., Helin K., Wutz A. 2010. Polycomb complexes act redundantly to repress genomic repeats and genes. Genes Dev. 24:265–276 10.1101/gad.544410 [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Lewis E.B. 1978. A gene complex controlling segmentation in Drosophila. Nature. 276:565–570 10.1038/276565a0 [DOI] [PubMed] [Google Scholar]
  44. Li G., Margueron R., Ku M., Chambon P., Bernstein B.E., Reinberg D. 2010. Jarid2 and PRC2, partners in regulating gene expression. Genes Dev. 24:368–380 10.1101/gad.1886410 [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Loh Y.H., Wu Q., Chew J.L., Vega V.B., Zhang W., Chen X., Bourque G., George J., Leong B., Liu J., et al. 2006. The Oct4 and Nanog transcription network regulates pluripotency in mouse embryonic stem cells. Nat. Genet. 38:431–440 10.1038/ng1760 [DOI] [PubMed] [Google Scholar]
  46. Margueron R., Li G., Sarma K., Blais A., Zavadil J., Woodcock C.L., Dynlacht B.D., Reinberg D. 2008. Ezh1 and Ezh2 maintain repressive chromatin through different mechanisms. Mol. Cell. 32:503–518 10.1016/j.molcel.2008.11.004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Margueron R., Justin N., Ohno K., Sharpe M.L., Son J., Drury W.J., III, Voigt P., Martin S.R., Taylor W.R., De Marco V., et al. 2009. Role of the polycomb protein EED in the propagation of repressive histone marks. Nature. 461:762–767 10.1038/nature08398 [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Mayer W., Niveleau A., Walter J., Fundele R., Haaf T. 2000. Demethylation of the zygotic paternal genome. Nature. 403:501–502 10.1038/35000656 [DOI] [PubMed] [Google Scholar]
  49. Meissner A., Mikkelsen T.S., Gu H., Wernig M., Hanna J., Sivachenko A., Zhang X., Bernstein B.E., Nusbaum C., Jaffe D.B., et al. 2008. Genome-scale DNA methylation maps of pluripotent and differentiated cells. Nature. 454:766–770 [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Mikkelsen T.S., Ku M., Jaffe D.B., Issac B., Lieberman E., Giannoukos G., Alvarez P., Brockman W., Kim T.K., Koche R.P., et al. 2007. Genome-wide maps of chromatin state in pluripotent and lineage-committed cells. Nature. 448:553–560 10.1038/nature06008 [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Miller T., Krogan N.J., Dover J., Erdjument-Bromage H., Tempst P., Johnston M., Greenblatt J.F., Shilatifard A. 2001. COMPASS: a complex of proteins associated with a trithorax-related SET domain protein. Proc. Natl. Acad. Sci. USA. 98:12902–12907 10.1073/pnas.231473398 [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Milne T.A., Briggs S.D., Brock H.W., Martin M.E., Gibbs D., Allis C.D., Hess J.L. 2002. MLL targets SET domain methyltransferase activity to Hox gene promoters. Mol. Cell. 10:1107–1117 10.1016/S1097-2765(02)00741-4 [DOI] [PubMed] [Google Scholar]
  53. Min J., Zhang Y., Xu R.M. 2003. Structural basis for specific binding of Polycomb chromodomain to histone H3 methylated at Lys 27. Genes Dev. 17:1823–1828 10.1101/gad.269603 [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Mohn F., Weber M., Rebhan M., Roloff T.C., Richter J., Stadler M.B., Bibel M., Schübeler D. 2008. Lineage-specific polycomb targets and de novo DNA methylation define restriction and potential of neuronal progenitors. Mol. Cell. 30:755–766 10.1016/j.molcel.2008.05.007 [DOI] [PubMed] [Google Scholar]
  55. Morey L., Helin K. 2010. Polycomb group protein-mediated repression of transcription. Trends Biochem. Sci. 35:323–332 10.1016/j.tibs.2010.02.009 [DOI] [PubMed] [Google Scholar]
  56. Morin-Kensicki E.M., Faust C., LaMantia C., Magnuson T. 2001. Cell and tissue requirements for the gene eed during mouse gastrulation and organogenesis. Genesis. 31:142–146 10.1002/gene.10017 [DOI] [PubMed] [Google Scholar]
  57. Müller J., Hart C.M., Francis N.J., Vargas M.L., Sengupta A., Wild B., Miller E.L., O’Connor M.B., Kingston R.E., Simon J.A. 2002. Histone methyltransferase activity of a Drosophila Polycomb group repressor complex. Cell. 111:197–208 10.1016/S0092-8674(02)00976-5 [DOI] [PubMed] [Google Scholar]
  58. Nagy P.L., Griesenbeck J., Kornberg R.D., Cleary M.L. 2002. A trithorax-group complex purified from Saccharomyces cerevisiae is required for methylation of histone H3. Proc. Natl. Acad. Sci. USA. 99:90–94 10.1073/pnas.221596698 [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Nekrasov M., Klymenko T., Fraterman S., Papp B., Oktaba K., Köcher T., Cohen A., Stunnenberg H.G., Wilm M., Müller J. 2007. Pcl-PRC2 is needed to generate high levels of H3-K27 trimethylation at Polycomb target genes. EMBO J. 26:4078–4088 10.1038/sj.emboj.7601837 [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Nuytten M., Beke L., Van Eynde A., Ceulemans H., Beullens M., Van Hummelen P., Fuks F., Bollen M. 2008. The transcriptional repressor NIPP1 is an essential player in EZH2-mediated gene silencing. Oncogene. 27:1449–1460 10.1038/sj.onc.1210774 [DOI] [PubMed] [Google Scholar]
  61. O’Carroll D., Erhardt S., Pagani M., Barton S.C., Surani M.A., Jenuwein T. 2001. The polycomb-group gene Ezh2 is required for early mouse development. Mol. Cell. Biol. 21:4330–4336 10.1128/MCB.21.13.4330-4336.2001 [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Okano M., Bell D.W., Haber D.A., Li E. 1999. DNA methyltransferases Dnmt3a and Dnmt3b are essential for de novo methylation and mammalian development. Cell. 99:247–257 10.1016/S0092-8674(00)81656-6 [DOI] [PubMed] [Google Scholar]
  63. Oktaba K., Gutiérrez L., Gagneur J., Girardot C., Sengupta A.K., Furlong E.E., Müller J. 2008. Dynamic regulation by polycomb group protein complexes controls pattern formation and the cell cycle in Drosophila. Dev. Cell. 15:877–889 10.1016/j.devcel.2008.10.005 [DOI] [PubMed] [Google Scholar]
  64. Ooi S.K., Qiu C., Bernstein E., Li K., Jia D., Yang Z., Erdjument-Bromage H., Tempst P., Lin S.P., Allis C.D., et al. 2007. DNMT3L connects unmethylated lysine 4 of histone H3 to de novo methylation of DNA. Nature. 448:714–717 10.1038/nature05987 [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Pan G., Tian S., Nie J., Yang C., Ruotti V., Wei H., Jonsdottir G.A., Stewart R., Thomson J.A. 2007. Whole-genome analysis of histone H3 lysine 4 and lysine 27 methylation in human embryonic stem cells. Cell Stem Cell. 1:299–312 10.1016/j.stem.2007.08.003 [DOI] [PubMed] [Google Scholar]
  66. Pandey R.R., Mondal T., Mohammad F., Enroth S., Redrup L., Komorowski J., Nagano T., Mancini-Dinardo D., Kanduri C. 2008. Kcnq1ot1 antisense noncoding RNA mediates lineage-specific transcriptional silencing through chromatin-level regulation. Mol. Cell. 32:232–246 10.1016/j.molcel.2008.08.022 [DOI] [PubMed] [Google Scholar]
  67. Panning B., Jaenisch R. 1996. DNA hypomethylation can activate Xist expression and silence X-linked genes. Genes Dev. 10:1991–2002 10.1101/gad.10.16.1991 [DOI] [PubMed] [Google Scholar]
  68. Pasini D., Bracken A.P., Jensen M.R., Lazzerini Denchi E., Helin K. 2004. Suz12 is essential for mouse development and for EZH2 histone methyltransferase activity. EMBO J. 23:4061–4071 10.1038/sj.emboj.7600402 [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Pasini D., Bracken A.P., Hansen J.B., Capillo M., Helin K. 2007. The polycomb group protein Suz12 is required for embryonic stem cell differentiation. Mol. Cell. Biol. 27:3769–3779 10.1128/MCB.01432-06 [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Pasini D., Hansen K.H., Christensen J., Agger K., Cloos P.A., Helin K. 2008. Coordinated regulation of transcriptional repression by the RBP2 H3K4 demethylase and Polycomb-Repressive Complex 2. Genes Dev. 22:1345–1355 10.1101/gad.470008 [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Pasini D., Cloos P.A., Walfridsson J., Olsson L., Bukowski J.P., Johansen J.V., Bak M., Tommerup N., Rappsilber J., Helin K. 2010. JARID2 regulates binding of the Polycomb repressive complex 2 to target genes in ES cells. Nature. 464:306–310 10.1038/nature08788 [DOI] [PubMed] [Google Scholar]
  72. Peng J.C., Valouev A., Swigut T., Zhang J., Zhao Y., Sidow A., Wysocka J. 2009. Jarid2/Jumonji coordinates control of PRC2 enzymatic activity and target gene occupancy in pluripotent cells. Cell. 139:1290–1302 10.1016/j.cell.2009.12.002 [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Rinn J.L., Kertesz M., Wang J.K., Squazzo S.L., Xu X., Brugmann S.A., Goodnough L.H., Helms J.A., Farnham P.J., Segal E., Chang H.Y. 2007. Functional demarcation of active and silent chromatin domains in human HOX loci by noncoding RNAs. Cell. 129:1311–1323 10.1016/j.cell.2007.05.022 [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Roguev A., Schaft D., Shevchenko A., Pijnappel W.W., Wilm M., Aasland R., Stewart A.F. 2001. The Saccharomyces cerevisiae Set1 complex includes an Ash2 homologue and methylates histone 3 lysine 4. EMBO J. 20:7137–7148 10.1093/emboj/20.24.7137 [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Roh T.Y., Cuddapah S., Cui K., Zhao K. 2006. The genomic landscape of histone modifications in human T cells. Proc. Natl. Acad. Sci. USA. 103:15782–15787 10.1073/pnas.0607617103 [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Sarma K., Margueron R., Ivanov A., Pirrotta V., Reinberg D. 2008. Ezh2 requires PHF1 to efficiently catalyze H3 lysine 27 trimethylation in vivo. Mol. Cell. Biol. 28:2718–2731 10.1128/MCB.02017-07 [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Shen X., Kim W., Fujiwara Y., Simon M.D., Liu Y., Mysliwiec M.R., Yuan G.C., Lee Y., Orkin S.H. 2009. Jumonji modulates polycomb activity and self-renewal versus differentiation of stem cells. Cell. 139:1303–1314 10.1016/j.cell.2009.12.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Simonsson S., Gurdon J. 2004. DNA demethylation is necessary for the epigenetic reprogramming of somatic cell nuclei. Nat. Cell Biol. 6:984–990 10.1038/ncb1176 [DOI] [PubMed] [Google Scholar]
  79. Squazzo S.L., O’Geen H., Komashko V.M., Krig S.R., Jin V.X., Jang S.W., Margueron R., Reinberg D., Green R., Farnham P.J. 2006. Suz12 binds to silenced regions of the genome in a cell-type-specific manner. Genome Res. 16:890–900 10.1101/gr.5306606 [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Stock J.K., Giadrossi S., Casanova M., Brookes E., Vidal M., Koseki H., Brockdorff N., Fisher A.G., Pombo A. 2007. Ring1-mediated ubiquitination of H2A restrains poised RNA polymerase II at bivalent genes in mouse ES cells. Nat. Cell Biol. 9:1428–1435 10.1038/ncb1663 [DOI] [PubMed] [Google Scholar]
  81. Struhl G. 1981. A gene product required for correct initiation of segmental determination in Drosophila. Nature. 293:36–41 10.1038/293036a0 [DOI] [PubMed] [Google Scholar]
  82. Tahiliani M., Koh K.P., Shen Y., Pastor W.A., Bandukwala H., Brudno Y., Agarwal S., Iyer L.M., Liu D.R., Aravind L., Rao A. 2009. Conversion of 5-methylcytosine to 5-hydroxymethylcytosine in mammalian DNA by MLL partner TET1. Science. 324:930–935 10.1126/science.1170116 [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Takahashi K., Yamanaka S. 2006. Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell. 126:663–676 10.1016/j.cell.2006.07.024 [DOI] [PubMed] [Google Scholar]
  84. Thomson J.A., Itskovitz-Eldor J., Shapiro S.S., Waknitz M.A., Swiergiel J.J., Marshall V.S., Jones J.M. 1998. Embryonic stem cell lines derived from human blastocysts. Science. 282:1145–1147 10.1126/science.282.5391.1145 [DOI] [PubMed] [Google Scholar]
  85. Thomson J.P., Skene P.J., Selfridge J., Clouaire T., Guy J., Webb S., Kerr A.R., Deaton A., Andrews R., James K.D., et al. 2010. CpG islands influence chromatin structure via the CpG-binding protein Cfp1. Nature. 464:1082–1086 10.1038/nature08924 [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Tsai M.C., Manor O., Wan Y., Mosammaparast N., Wang J.K., Lan F., Shi Y., Segal E., Chang H.Y. 2010. Long noncoding RNA as modular scaffold of histone modification complexes. Science. 329:689–693 10.1126/science.1192002 [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Tsumura A., Hayakawa T., Kumaki Y., Takebayashi S., Sakaue M., Matsuoka C., Shimotohno K., Ishikawa F., Li E., Ueda H.R., et al. 2006. Maintenance of self-renewal ability of mouse embryonic stem cells in the absence of DNA methyltransferases Dnmt1, Dnmt3a and Dnmt3b. Genes Cells. 11:805–814 10.1111/j.1365-2443.2006.00984.x [DOI] [PubMed] [Google Scholar]
  88. van den Berg D.L., Snoek T., Mullin N.P., Yates A., Bezstarosti K., Demmers J., Chambers I., Poot R.A. 2010. An Oct4-centered protein interaction network in embryonic stem cells. Cell Stem Cell. 6:369–381 10.1016/j.stem.2010.02.014 [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Vastenhouw N.L., Zhang Y., Woods I.G., Imam F., Regev A., Liu X.S., Rinn J., Schier A.F. 2010. Chromatin signature of embryonic pluripotency is established during genome activation. Nature. 464:922–926 10.1038/nature08866 [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Vierbuchen T., Ostermeier A., Pang Z.P., Kokubu Y., Südhof T.C., Wernig M. 2010. Direct conversion of fibroblasts to functional neurons by defined factors. Nature. 463:1035–1041 10.1038/nature08797 [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Voncken J.W., Roelen B.A., Roefs M., de Vries S., Verhoeven E., Marino S., Deschamps J., van Lohuizen M. 2003. Rnf2 (Ring1b) deficiency causes gastrulation arrest and cell cycle inhibition. Proc. Natl. Acad. Sci. USA. 100:2468–2473 10.1073/pnas.0434312100 [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Walker E., Chang W.Y., Hunkapiller J., Cagney G., Garcha K., Torchia J., Krogan N.J., Reiter J.F., Stanford W.L. 2010. Polycomb-like 2 associates with PRC2 and regulates transcriptional networks during mouse embryonic stem cell self-renewal and differentiation. Cell Stem Cell. 6:153–166 10.1016/j.stem.2009.12.014 [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Wang J., Mager J., Schnedier E., Magnuson T. 2002. The mouse PcG gene eed is required for Hox gene repression and extraembryonic development. Mamm. Genome. 13:493–503 10.1007/s00335-002-2182-7 [DOI] [PubMed] [Google Scholar]
  94. Wang J., Rao S., Chu J., Shen X., Levasseur D.N., Theunissen T.W., Orkin S.H. 2006. A protein interaction network for pluripotency of embryonic stem cells. Nature. 444:364–368 10.1038/nature05284 [DOI] [PubMed] [Google Scholar]
  95. Woo C.J., Kharchenko P.V., Daheron L., Park P.J., Kingston R.E. 2010. A region of the human HOXD cluster that confers polycomb-group responsiveness. Cell. 140:99–110 10.1016/j.cell.2009.12.022 [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Zhao X.D., Han X., Chew J.L., Liu J., Chiu K.P., Choo A., Orlov Y.L., Sung W.K., Shahab A., Kuznetsov V.A., et al. 2007. Whole-genome mapping of histone H3 Lys4 and 27 trimethylations reveals distinct genomic compartments in human embryonic stem cells. Cell Stem Cell. 1:286–298 10.1016/j.stem.2007.08.004 [DOI] [PubMed] [Google Scholar]
  97. Zhao J., Sun B.K., Erwin J.A., Song J.J., Lee J.T. 2008. Polycomb proteins targeted by a short repeat RNA to the mouse X chromosome. Science. 322:750–756 10.1126/science.1163045 [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from The Journal of Experimental Medicine are provided here courtesy of The Rockefeller University Press

RESOURCES