The Polycomb group (PcG) of proteins is a major mechanism of epigenetic regulation that has been broadly linked to cancer. This system can repress gene expression by chromatin modification and is essential for establishing cell identity. PcG proteins are important for stem cell function and differentiation and have a profound impact during hematopoiesis. In recent years, several published studies have deepened our knowledge of the biology of the PcG in health and disease. In this article, we review the current understanding of the mechanisms of PcG-mediated repression and their relation to DNA methylation, and we discuss the role of the PcG system in hematopoiesis and hematologic malignancies. We suggest that alteration of different PcG members is a frequent event in leukemia and lymphomas that confers the stem cell properties on tumor cells. Thus, drugs targeting Polycomb complexes could be useful for treating patients with these diseases.

In the past few years, evidence has built up concerning the critical function of the Polycomb group of proteins (PcG) and other epigenetic regulators for the establishment of cell identity, stem cell function, and differentiation. The PcG is a major mechanism of epigenetic regulation that has been broadly linked to cancer. Taking into account the increasingly extensive literature reflecting recent advances in this field and the abundant data on PcG protein alterations and functions, we review PcG functions in lymphopoiesis and hematopoietic stem cell function as well as their role in the initiation and maintenance of hematologic malignancies. Specifically, we also discuss the use of PcG genes as prognostic or therapeutic markers.

Chromatin structure and modifications and, therefore, chromatin regulatory factors are involved in the control of gene expression affecting stemness, differentiation, and proliferation processes and, consequently, tumorigenesis. Among these factors, the PcG and trithorax group (trxG), together with histone demethylases, histone acetylases, and deacetylases, among others, are important sets of proteins that regulate gene activity at the chromatin level.

PcG and trxG proteins were first described in Drosophila,1  where they are responsible for maintaining homeotic gene activity in the appropriate segments during fly development. These systems act by silencing (PcG) or activating (TrxG) gene expression when they bind to specific regions of DNA. In Drosophila, most of the genes regulated by PcG contain consensus sequences called Polycomb-repressed elements (PREs).2  However, although some specific regions in the human genome have been shown to have a similar function than PREs,3  a general DNA sequence specific for Polycomb binding in mammals has not been identified.

PcG proteins form DNA-binding protein complexes, and although the composition of these is variable and dynamic,4,5  2 biochemical and major functional PcG complexes, termed Polycomb-repressive complexes (PRCs), have been described in mammals. One complex is composed of enhancer of zeste homolog 2 (EZH2), suppressor of zeste 12 homolog (SUZ12), and embryonic ectoderm development (EED), which form the nucleus of the PRC2 complex. EZH2 has histone methyltransferase activity and di- and trimethylates lysine 27 of histone H3 (H3K27me2 and H3K27me3). However, PRC2 composition is variable. In fact, some investigators have postulated the existence of PRC3 and PRC4, depending on the partners present in the complex.6,7  These include different EED isoforms, RbAp48, RbAp46, PHD finger protein 1 (PHF1), adipocyte enhances binding protein 2 (AEBP2), sirtuin 1 (SIRT1), histone deacetylase 1 (HDAC1), and HDAC2, among others. The patterns of expression of these various complexes depend on the differentiated status (eg, PRC4 is detected in undifferentiated, but not in differentiated, cells) or the normal/tumoral status of the cell (PRC4 is found in tumoral, but not in normal, differentiated cells).6 

The second complex, PRC1, is even more heterogeneous than PRC2, but the core of the complex is formed by RING finger protein 1 (RING1), RNF2 (RING finger protein 2) BMI1 (B lym-phoma Moloney murine leukemia virus integration site 1), MEL18/PCGF2 (melanoma nuclear protein 18/polycomb group ring finger 2), polyhomeotic homolog 1 (PH), nervous system polycomb 1 (NSPC1), MEL18- and BMI1-like ring-finger protein (MBLR), and chromobox homolog (CBX) proteins (Table 1). PRC1 is believed to recognize the PRC2-methyl mark, H3K27me3, that is recruited to the DNA at the appropriate genomic locations. RNF2 in PRC1 complex has histone ubiquitination E3 ligase activity (H2K119ub),8  and this modification is also associated with gene repression.8,9  These complexes could have different compositions with distinct targets, depending on the cellular context, cell type, cell cycle, or differentiation stage, and on the association with other proteins (reviewed in Kerppola10 ). For example, although BMI1 and MEL18 are both homologous to Posterior sex combs (Psc) in Drosophila, their expression seems to be mutually exclusive and confers different specificities on the PRC1 complex.11-13 

Table 1

Common members of PRCs and their alterations in cancer

HumanMouse homologAlteration in cancerTumor type
PRC1 complex    
BMI1 Bmi1 Overexpression/amplification Acute myeloid leukemia,155  Hodgkin lymphoma,5,102,103,113  B cell non-Hodgkin lymphoma,5,54,112,126  gastrointestinal tumors,5  pituitary adenoma,5  parathyroid adenoma,6  breast cancer,156  glioma,157  neuroblastoma,158  medulloblastoma,159  non-small-cell lung cancer,160  skin cancer161  
MEL18/RNF110/PCGF2 Mel18 Loss Cutaneous squamous-cell carcinoma,6  gynecological tumors,5  prostate cancer158  
RING1A Ring1a Overexpression Large B-cell lymphoma126  
  Loss Clear-cell renal-cell carcinoma and testicular germ-cell tumors5  
RNF2 Ring1b Overexpression Gastric and colonic tumors,5  melanoma,5  lung cancer,5  gynecological tumors,5  diffuse large B-cell lymphoma,5  follicular lymphoma,5  Burkitt lymphoma,5  Hodgkin lymphoma5,102  
HPH1, HPH2, HPH3 Rae28/Mph1, Phc2, Phc3 Lack of expression Acute lymphocytic leukemia162  
Overexpression Large B-cell lymphoma,156  central nervous system tumors6  
CBX2, CBX4, CBX6, CBX7, CBX8 M33/Cbx2, Cbx4, Cbx6, Cbx7, Cbx8 Loss Pancreatic cancer,159  aggressive thyroid carcinomas160  
Overexpression Follicular lymphoma163  
PCGF6 Pcgf6 Loss Mantle-cell lymphoma164  
RYBP Rybp Overexpression Oligodendroglioma,5  pituitary adenoma,5  T cell lymphoma,5  Hodgkin Lymphoma5,102  
PRC2 complex    
EZH2/EZH1 Ezh2/Ezh1 Overexpression/amplification Prostate carcinoma,109,165  breast carcinoma,127,165,166  melanoma,160  bladder cancer,127,167  glioma,168  endometrial carcinoma,165  lymphomas,111,127  colon carcinoma,127  glioblastoma,127  Ewing tumor169  
Deletion/mutation Lymphomas,114  myeloid malignancies117,118  
Altered localization Acute myeloid leukemia104  
SUZ12 Suz12 Overexpression/amplification/translocation Gastrointestinal tumors,108,170  lung tumors,108  thyroid follicular carcinoma,108  pituitary adenoma,108  parathyroid adenoma,108  skin tumors,108  breast and gynecological tumors,108,170  mantle-cell lymphoma,108  chronic myeloid leukemia,119  endometrial stromal tumor119  
EED Eed Overexpression Salivary gland adenoid cystic carcinoma,171  Hodgkin lymphoma,103  prostate cancer6  
Lack of expression Prostate cancer172  
RBBP7 Rbbp7 Overexpression Non-small-cell lung cancer,173  breast cancer174  
RBBP4 Rbbp4 Down-regulation Cervical cancer175  
PCL3 Phf19 Overexpression Cancers of the colon, skin, lung, rectum, cervix, uterus, and liver 176  
HumanMouse homologAlteration in cancerTumor type
PRC1 complex    
BMI1 Bmi1 Overexpression/amplification Acute myeloid leukemia,155  Hodgkin lymphoma,5,102,103,113  B cell non-Hodgkin lymphoma,5,54,112,126  gastrointestinal tumors,5  pituitary adenoma,5  parathyroid adenoma,6  breast cancer,156  glioma,157  neuroblastoma,158  medulloblastoma,159  non-small-cell lung cancer,160  skin cancer161  
MEL18/RNF110/PCGF2 Mel18 Loss Cutaneous squamous-cell carcinoma,6  gynecological tumors,5  prostate cancer158  
RING1A Ring1a Overexpression Large B-cell lymphoma126  
  Loss Clear-cell renal-cell carcinoma and testicular germ-cell tumors5  
RNF2 Ring1b Overexpression Gastric and colonic tumors,5  melanoma,5  lung cancer,5  gynecological tumors,5  diffuse large B-cell lymphoma,5  follicular lymphoma,5  Burkitt lymphoma,5  Hodgkin lymphoma5,102  
HPH1, HPH2, HPH3 Rae28/Mph1, Phc2, Phc3 Lack of expression Acute lymphocytic leukemia162  
Overexpression Large B-cell lymphoma,156  central nervous system tumors6  
CBX2, CBX4, CBX6, CBX7, CBX8 M33/Cbx2, Cbx4, Cbx6, Cbx7, Cbx8 Loss Pancreatic cancer,159  aggressive thyroid carcinomas160  
Overexpression Follicular lymphoma163  
PCGF6 Pcgf6 Loss Mantle-cell lymphoma164  
RYBP Rybp Overexpression Oligodendroglioma,5  pituitary adenoma,5  T cell lymphoma,5  Hodgkin Lymphoma5,102  
PRC2 complex    
EZH2/EZH1 Ezh2/Ezh1 Overexpression/amplification Prostate carcinoma,109,165  breast carcinoma,127,165,166  melanoma,160  bladder cancer,127,167  glioma,168  endometrial carcinoma,165  lymphomas,111,127  colon carcinoma,127  glioblastoma,127  Ewing tumor169  
Deletion/mutation Lymphomas,114  myeloid malignancies117,118  
Altered localization Acute myeloid leukemia104  
SUZ12 Suz12 Overexpression/amplification/translocation Gastrointestinal tumors,108,170  lung tumors,108  thyroid follicular carcinoma,108  pituitary adenoma,108  parathyroid adenoma,108  skin tumors,108  breast and gynecological tumors,108,170  mantle-cell lymphoma,108  chronic myeloid leukemia,119  endometrial stromal tumor119  
EED Eed Overexpression Salivary gland adenoid cystic carcinoma,171  Hodgkin lymphoma,103  prostate cancer6  
Lack of expression Prostate cancer172  
RBBP7 Rbbp7 Overexpression Non-small-cell lung cancer,173  breast cancer174  
RBBP4 Rbbp4 Down-regulation Cervical cancer175  
PCL3 Phf19 Overexpression Cancers of the colon, skin, lung, rectum, cervix, uterus, and liver 176  

Bold text indicates members known to be frequently altered in cancer.

In embryonic stem (ES) cells, which have been the subject of most of the studies, a high level of cooccupancy among different PcG proteins (eg, EED, SUZ12, Ring1B, and Ph1) has been described in H3K27me3 promoters. The presence of PRC1 and PRC2 is interrelated, and, as previously mentioned, H3K27me3 has been described as the epigenetic mark that recruits PRC1 to specific promoters.14-16  PRC1 and PRC2 then provoke changes in chromatin structure,17,18  and/or the ubiquitin ligase activity of Ring1B contributes to gene silencing. However, independence from this mark has also been proposed for PRC1 recruitment.19,20  Therefore, the mechanisms for gene repression mediated by PcG proteins depend on PRC1 and PRC2, alone or in association (for a recent review, see Simon and Kingston21 ). These epigenetic modifications of histone proteins and chromatin conformational changes seem to be the main way by which PcG mediates gene silencing. An interrelationship between PcG silencing and DNA hypermethylation has also been suggested for EZH222,23  and has been proposed as being critical to transcriptional silencing. PRC2 can recruit DNMTs to PcG target genes, leading to de novo methylation in these loci. BMI1 is associated with DNMT1 and DMAP1 (DNMT1-associated protein).24  Other studies showing association between PcG targets and methylated cytosine guanine dinucleotide (CpG) islands corroborate the relationship between PcG and DNA methylation25  (see below). However, other authors report the presence of PcG marks (H3K27me3) outside CpG islands in tumoral cells,26  suggesting that there is a mechanism of gene silencing that is independent of promoter DNA methylation in cancer.

Once bound to chromatin, PcG complexes can interfere with other systems. They can inhibit ATP-dependent chromatin remodeling by SWI/SNF,27  interact with components of the transcriptional machinery, as is suggested by the presence of TFIID in the PRC1 complex,28  or block RNA polymerase II association with the target promoter, leading to the inhibition of transcriptional initiation and elongation.29 

As mentioned above, the DNA sequences specific for Polycomb binding have not been identified in mammals. Therefore, the question remains as to what mechanisms direct PcG proteins to target promoters in vertebrates. One of those proposed is mediated by PcG-interacting proteins that provide the sequence specificity for PcG binding to target promoters, such as transcription regulators with DNA-binding capacity. This has been described for E2F6, a member of the E2F transcription factor protein family, but unlike other E2F members, this behaves as a transcriptional repressor by interacting with Ring1a and Ring1b, together with YAF2 and DP1.30  BCOR (BCL-6 interacting corepressor), a POZ that is required for germinal center formation,31  is also associated with RING1, RNF2, and Ring1 and YY1 binding protein (RYBP) PcG proteins, among others.32,33  This formation of the BCOR complex with PcG proteins could explain some of the enzymatic activities that can be recruited by BCL6 and, also, how BCL6 may determine some of the targets of PcG proteins. Other transcription regulators, such as OCT4, NANOG, and SOX, interact with PcG proteins from both PRC2 (SUZ12)34  and PRC1 complexes (RNF2)35  in ES cells. These factors could repress differentiation-promoting genes by recruiting PcG proteins to their promoters (see below).34,36 

Gene-specific recruitment by ncRNA14,37-39  has been implicated in PcG target repression and specificity. Recently, Kanhere et al40  showed that short RNAs transcribed from Polycomb target genes by RNApolII interact with PRC2, forming a stem-loop structure stabilizing the association of PRC2 with chromatin through SUZ12 and, consequently, repressing the gene expression in cis.

Long ncRNA, such as HOTAIR (Hox transcript antisense intergenic RNA, encoded in the HOXC cluster) interacts with PRC2 and is required for SUZ12 recruitment and trimethylation of H3K27 in HOXD loci.37  Something similar happens with Kcnq1ot1, an ncRNA situated in the intronic region of the Kcnq1 gene. Kcnq1ot1 interacts with chromatin and PRC2 complex in a lineage-specific manner (it seems to function in placenta, but not in fetal liver),38  and this correlates with H3K27me3 enrichment at their target promoters (for review, see Hekimoglu and Ringrose41 ). This mechanism is similar to that producing X-chromosome inactivation and genomic imprinting, mediated by Xist, which targets PRC2 to the inactive X-chromosome (for review, see Chandrasekhar et al42 ).

RNAi machinery can play a major role in transcriptional gene silencing by targeting promoter sequences, among other mechanisms, and this has been associated with PcG silencing. In Drosophila, Argonaute-1 (AGO1; the effector protein in the complexes involved in the RNAi mechanism) and other RNAi components colocalize with PcG bodies, suggesting a role for the RNAi machinery in the chromatin organization of PcG targets.39  In humans, AGO1 is colocalized with EZH2 and TRBP2 proteins in some siRNA-targeted promoters, such as RASFF1A.43 

A growing number of studies reflect the relevance of PcG in stem cell function, differentiation, development, and cell-cycle control, and consequently in cancer. It has also been implicated in cellular senescence, X-inactivation, genomic imprinting, and actin polymerization (for review, see Rajasekhar and Begemann44 ).

The role of BMI1 in self-renewal of hematopoietic stem cells (HSCs)45  and the fact that CDKN2, a cell-cycle regulator, is the paradigm of polycomb targets, both indicate the involvement of PcG in cell proliferation. There are several lines of evidence for the role of PcG proteins in controlling cell proliferation. Several cell-cycle regulators (both activators and repressors) are found among PcG targets, such as inhibitor of kinase 4/alternate reading frame (INK4a/ARF), MYC, JUN, FOS, CDC25, human telomerase reverse transcriptase, cyclins, cyclin-dependent kinases (CDKs), checkpoint kinase 1 homolog (CHEK1), mitotic arrest deficient–like 1 (MAD2L1), and budding uninhibited by benzimidazoles 3 (BUB3).46-48  Recent studies in Drosophila show that PcG can also regulate cell growth by signaling pathways, such as JAK/STAT.49,50  Changes occur in the expression levels of PcG proteins during hematopoietic differentiation, at least, in part, through the regulation of Hox genes, whose dysregulation leads to alterations in lymphocyte proliferation.51-53  These alterations of PcG members occur in a wide variety of cancers (described below), highlighting the role in controlling cellular proliferation, whereby some PcG genes act as oncogenes (BMI15,54 ), and others function as tumor suppressor genes (MEL18)5,55  (reviewed in Martinez and Cavalli56 ). A special case is Rnf2, which has been shown to restrict proliferation of early myeloid progenitors through the inhibition of cyclin D2 and Cdc6 and, at the same time, to promote the expansion of maturing cell lineage–committed precursors of the myeloid and lymphoid compartments. Moreover, Rnf2 deficiency accelerates lymphomagenesis in the absence of Ink4a.57 

Studies of various species have demonstrated the importance of the PcG system not only in embryonic development, but also in adult differentiation and homeostasis.58  PcG regulates genes involved in development and differentiation pathways and are critical regulators during embryogenesis, as is demonstrated by the fact that disruption of several PcG genes is embryonic lethal or leads to postnatal lethality.51-53,59,60 

In ES cells, PRC2, and PRC1 are both present in the promoter region of key development regulators. These development regulator genes in mammalian stem cells have 2 epigenetic marks: H3K27me3, a signal of gene silencing, and H3K4me, a mark of gene activation associated with TrxG.61  These 2 marks, with their opposite functions in the same region, result in a bivalent promoter with both negative and positive possibilities and are thought to be important for the repressive or activating function in stem cells, to switch gene expression upon changes in cell status. These bivalent marks have also been found in nonstem cells, such as T cells and human lung fibroblasts,62-65  indicating that in more highly differentiated cells, the dual PcG/TrxG system also enabled fine-scale control of the expression of certain genes with bivalent domains in response to external or internal signals.66,67 

It is worth mentioning that ES cells can be generated in the absence of H3K27me3.68  Thus, H3K27me3 seems to mark genes that need to be activated during differentiation, rather than having a direct role on pluripotency at the bivalent loci. As proposed by Bernstein and colleagues,69  during ES cell differentiation, the abundance of bivalent promoters might be correlated with differentiation. However, the possibility that other histone modifications may have a bigger impact in these domains cannot be ruled out.

The localization and expression patterns of PcG genes change during the stages of differentiation.6,12  PcG proteins are displaced from one set of target genes while being directed to another during lineage specification (see below), but the mechanisms are not well understood. Genome-wide mapping by chromatin immunoprecipitation (ChIP) experiments of PcG members in human and murine ES cells has shown that PcG proteins bind and repress several genes involved in the regulation of development: transcription factors involved in regulating early stages in neurogenesis or hematopoiesis (eg, Pax, Lhx) and members of the Sox, Tbx, and Gata families.34,36  Therefore, there is a prevalence of developmental regulators and genes involved in cell-fate decisions among PcG target genes.34,36,70 

A large proportion of PcG targets in ES cells are also occupied by the transcription factors, OCT4, SOX2, and NANOG, which are known to be essential for stemness and also indicate that PcG is involved in this process.71  These factors could repress differentiation-promoting genes by recruiting PcG genes to their promoters, thereby maintaining the stem cell capacity of these cells.34,71  PcG complexes are displaced from their target promoters when cells are committed to differentiation.

The PcG system has been implicated not only in ES cells, but also in adult stem cell maintenance, and the role of Bmi1 and Ring1b in neural control,72-74  hematopoietic stemness,45  and other processes75  has been demonstrated (see below).

All blood cells derive from a common undifferentiated progenitor, which undergoes subsequent differentiation through a series of binary decisions; the cell must be able both to self-renew and to differentiate into any of the hematopoietic cell lineages. The function of HSCs is regulated by both extrinsic and intrinsic signals. Several systems that are active during embryonic development and act extrinsically on HSCs, such as the Sonic Hedgehog, Notch, and Wnt pathways,76,77  have been shown to induce self-renewal in adults. However, very few effectors are known that are downstream of these external signals. Intrinsic effectors important for HSCs include the HOX genes, Hoxa5, Hoxa9, Hoxa10, Hoxb3, Hoxb4, and Hoxb6.78  For instance, Hoxb4 is a transcription factor that, when overexpressed in HSCs, leads to the expansion of this cellular subset.79  It also seems to protect HSCs from multiple extrinsic signals, such as tumor necrosis factor-α (TNF-α).80  Strikingly, the development of knockout mice for Hoxb4 has shown this gene not to be required for the generation of HSCs or the maintenance of steady-state hematopoiesis. Hoxb4−/− HSCs show only mild defects in proliferation. The conclusion is that either Hoxb4 promotes proliferation of HSC, but not self-renewal, or that Hoxb4 deficiency can be compensated for by neighboring or paralogous Hox genes.81,82  In any case, there must be other intrinsic effectors that are essential to the function and survival of HSCs.

PcG genes play a major role in regulating hematopoietic function.11,45,83-85  It has recently been demonstrated that the levels of BMI1 and MEL18 determine, to a great extent, the capability of HSCs to function as progenitors,13  whereby BMI1 is associated with the enhancement of HSC properties and MEL18 is more closely related to the differentiated phenotypes. These data are consistent with the findings of a previous report showing that BMI1 is expressed in primitive human bone marrow cells, while many other components, such as MEL18, RAE28, and EZH2,12  are increasingly expressed upon differentiation. Moreover, Bmi1-deficient mice have severely impaired HSC self-renewal, and bone marrow progenitors lacking Bmi1 have a restricted proliferative potential,45,52,86  while forced expression of Bmi1 leads to enhanced lymphoproliferation, promotion of HSC self-renewal, and a higher probability of inheriting stemness through cell division.87 

Until now, the main mechanism known to regulate HSC capacities by Bmi1 is that acting by regulating the Ink4a/Arf locus. Bmi1−/− HSCs have markedly higher levels of Ink4a and Arf,46  and forced expression of the latter causes a decrease in HSC compartment due to the induction of cell-cycle arrest and p53-mediated apoptosis.45  Because the deletion of both genes partially restores the proliferative capacity of HSCs in a Bmi1−/− background, it seems that Bmi1 protects HSCs from premature loss through the inhibition of Ink4a and Arf,88  although other downstream pathways must mediate the effect of Bmi1 on proliferation.

Alterations in other PcG components also affect hematopoietic function. Overexpression of Ezh2 confers long-term repopulating potential on HSCs, preventing its exhaustion after replicative stress.89  Expression of one null eed allele or 2 hypomorphic eed alleles results in greater lymphoproliferation, developmental blockade during thymocyte differentiation, and a greater risk of developing hematologic tumors.90  On the other hand, targeted deletion of Mel-18 causes severe defects in lymphoid organs, including hypoplasia, while, in bone marrow, most hematopoietic cells are replaced by adipocytes.51  However, the defects in Mel-18−/− mice seem to be due to the impaired response to cytokines, because down-regulation of Mel-18 in bone marrow–derived HSCs promotes their self-renewal, while its forced expression reduces self-renewal capacity of HSCs.11  Similar to Bmi1 and Mel-18, M33-deficient mice also show hypoplasia in the spleen and thymus, and defects in T and B cells.53  Mph1/Rae-28 is another important component of the PcG system that affects proper hematopoietic function. Mph1/Rae-28 mutant embryos die because HSC activity in these animals cannot maintain the hematopoietic system during embryo development and show a progressive impairment in the numbers and proliferative capacity of colony-forming cells, compared with their wild-type littermates. Moreover, HSCs lacking the Mph1/Rae-28 gene cannot reconstitute the bone marrow after the irradiation of mice.91  Likewise, deletion of Rnf2 in the HSC compartment affects HSC function. Rnf2 is able to restrict the proliferation of progenitor cells while promoting the expansion of their maturing progeny, having a dual role in this system.57  All these data suggest a complex regulation of HSC function, based on a delicate equilibrium between enhancing and repressing PcG complexes that, when altered, induces mainly a differentiation blockade and/or impairment of the HSC function.

Human tumors are characterized by a broad spectrum of genetic alterations. It is well documented that progression from normal to tumoral cells also involves epigenetic changes, including extensive DNA methylation at promoter-associated CpG islands, and an aberrant pattern of histone modifications. Chromatin structure is crucial for the regulation of DNA accessibility and thus for the regulation of gene expression. Furthermore, an altered chromatin structure provokes altered gene patterns and genomic instability that can be propagated to daughter cells, causing cellular transformation to a malignant status.92 

There is increasing evidence that many tumors depend on the presence of a subpopulation of cancer stem cells (CSCs) with a genetic program that partially resembles that of normal stem cells. In fact, several signaling pathways that are important for tumor development, such as Sonic Hedgehog, Notch, and Wnt, also regulate self-renewal in stem cells (reviewed in Taipale and Beachy93 ), and some embryonic genes are reexpressed in human cancers.94  Several studies have demonstrated the presence of CSCs in tumors, such as acute myeloid leukemia (AML), breast cancer, glioblastoma, and others (for review, see Gupta et al95 ). This is accompanied by a stem cell–like signature at several levels, including gene expression96  and chromatin structure.97 

There are several ways in which CSCs can be generated. Tumor-initiating cells with stem cell properties could arise from a stem cell that loses the capacity to regulate its mitotic potential, or a downstream progenitor (or committed progenitor) cell could transiently acquire the ability to self-renew through some molecular alterations, possibly induced by the microenvironment (Figure 1). Both possibilities can be integrated in a single hypothesis through the demonstrated cell plasticity that cancer cells harbor, as suggested by a recent report on melanoma.98 

Figure 1

Models of tumorigenesis. Cancer cells can originate from a cell with stem cell properties, such as an HSC, that acquire an altered phenotype through chromosomal aberrations and/or mutations. This leads to an abnormal PcG activity that represses several tumor-suppressor genes promoting cancer development. Another possibility is that the first chromosomal aberration and/or mutation occurs in a committed progenitor cell. In this case, the abnormal PcG activity confers stem cell properties and self-renewal capacity as well as tumorigenic potential.

Figure 1

Models of tumorigenesis. Cancer cells can originate from a cell with stem cell properties, such as an HSC, that acquire an altered phenotype through chromosomal aberrations and/or mutations. This leads to an abnormal PcG activity that represses several tumor-suppressor genes promoting cancer development. Another possibility is that the first chromosomal aberration and/or mutation occurs in a committed progenitor cell. In this case, the abnormal PcG activity confers stem cell properties and self-renewal capacity as well as tumorigenic potential.

Close modal

PcG proteins have been identified as being important proteins in tumorigenesis due to their potential to repress tumor suppressor genes and regulate genes related to stemness and differentiation.19,34,36,45,70,72,91  Thus, genetic phenomena that alter the expression of PcG members such as BMI1 could be one of the key events that allow the malignant cell to acquire a stem cell phenotype.

Among the most important genes that confer transforming capacity to the cell, Cdkn2a is one of the main targets of Bmi1. In fact, Bmi1 has been identified as an important factor cooperating with Myc-induced lymphomagenesis,99,100  and Bmi1 knockdown has been shown to promote cancer-specific cell death in neuroblastoma.101  However, in some tumor types, such as Hodgkin lymphoma (HL), the regulation of CDKN2A by BMI1 might be disrupted by other mechanisms.102,103  A recent report has found the presence of EZH2 and the H3K27me3 polycomb repressive mark not only at the CDKN2A locus but also at the CDKN2B locus in AML cell lines and patient samples.104  The regulation of the Cdkn2a locus is a mechanism shared by another PcG protein, Cbx7, which is critical for expanding the cellular lifespan and for bypassing senescence by inhibiting the expression of Ink4a/Arf. In contrast to Bmi1, however, Cbx7 is unable to induce telomerase activity, although it still can initiate T-cell lymphomagenesis in mouse models.105,106  Other oncogenic mechanisms have been ascribed to BMI1, such as the inhibition of phosphatase and tensin homolog.107  Furthermore, the effects of many other PcG members are independent of Cdkn2a and are much more diverse.

Several studies have shown an extensive relationship between alterations in different PcG members and cancers originating from various tissues (Table 1). Our group and others have found overexpression/down-regulation of PcG members in many tumors, including several lymphomas5,102,108  (Figure 2). Many solid tumors also present alterations in members of the Polycomb system. Prostate cancer is an important example of this, in which EZH2 levels are correlated with aggressiveness of the disease in a mechanism independent of DNA methylation.109  The cancer types studied are diverse, but particular attention has been paid to hematologic malignancies. Several observations support the role of PcG genes in abnormal hematopoiesis through the regulation of HSC self-renewal/proliferation. BMI1 up-regulation is associated with leukemia and mantle cell lymphoma (MCL) and is linked to bad prognosis, and its locus is found to be amplified in MCL.5,54  Furthermore, hematopoietic progenitors of Bmi1-deficient mice are resistant to transformation by the E2a-Pbx1 fusion gene that is frequently present in human acute pre-B lymphoblastic leukemias.110  Bmi1 also seems to be important for the maintenance of the leukemic phenotype (or for the maintenance of leukemia) initiated by the Hoxa9 and Meis1a collaborative oncogenes in a model of AML.86  EZH2 is also up-regulated in MCL111  and anomalously coexpressed with BMI1 in several non-HLs, including small lymphocytic lymphoma, follicular lymphoma (FL), diffuse large B-cell lymphoma (DLBCL), and Burkitt lymphoma (BL),112  as well as in HL itself103,113  (Figure 2). EZH2 is also a target of other genetic abnormalities. For instance, mutations have been described in the catalytic domain of EZH2 in FL and DLBCL of germinal center origin.114  This mutation leads to a reduction of the in vitro enzymatic activity of EZH2, suggesting that, at least in DLBCL and FL of germinal center origin, the mechanism by which EZH2 collaborates to produce malignant transformation is different from those of other types of cancer, such as prostate cancer. Although we do not know for certain, it is possible that the mutations in EZH2 alter not only its enzymatic activity, but also the product or target specificity of the complex. This alteration might be especially important for germinal center–derived B cells, because EZH2 is necessary in these cells for early B-cell development, including rearrangement of the immunoglobulin heavy-chain (IGH) locus.115  Furthermore, in AML, EZH2 seems to be aberrantly localized in the genome,104  and the chromosomal location of EZH2 is a hotspot of genomic aberrations present in myeloid disorders.116  Strikingly, 2 recent studies reported genetic aberrations affecting EZH2, including deletions, mutations, and uniparental disomy, in myeloid malignancies.117,118  All the studied alterations result in the inactivation of EZH2 and loss of H3K27 trimethylation, suggesting that EZH2 may have a dual role as an oncogene or tumor-suppressor gene, depending on cell context. SUZ12 is also the target of genomic aberrations and has been found to be translocated in endometrial stromal tumors.119  The fusion protein is able to restore H3K27me3 after SUZ12 knockdown, inhibits apoptosis, and promotes cell proliferation.120  Our laboratory has shown that SUZ12 is overexpressed in several germinal center (GC)–derived lymphomas, although normal GCs also express SUZ12.108  On the other hand, we found that SUZ12 is also overexpressed in MCL, while the normal mantle zone cells are negative for SUZ12 expression.108  In some cases, this overexpression was accompanied by genetic gains and/or amplifications. The overexpression of SUZ12 seems to be relevant for the tumoral cells, because silencing of SUZ12 in MCL-derived cell lines induces the reexpression of several SUZ12 target genes and increases apoptosis.108  Likewise, the knockdown of SUZ12 in acute promyelocytic leukemia (APL)–derived cells is able to induce granulocytic differentiation. In this leukemia, the fusion gene PML-RARα interacts with the PRC2 complex and directs it to specific target genes121 ; this is a good example of how targeting Polycomb complexes by an oncogene can be of great relevance in tumorigenesis.122  Another example is the WNT pathway, which is altered in several malignancies and constitutively active in chronic myeloid leukemia (CML). In this tumor, during the blastic crisis of the disease, SUZ12 is overexpressed due to the activation of its promoter by some members of the WNT pathway, such as WNT5A and WNT11.123 

Figure 2

Immunohistochemical staining for PcG members reveals frequent alterations of these proteins in hematologic malignancies. RLT: reactive lymphoid tissue; MCL: mantle cell lymphoma; DLBCL: diffuse large B-cell lymphoma; HL: Hodgkin lymphoma.

Figure 2

Immunohistochemical staining for PcG members reveals frequent alterations of these proteins in hematologic malignancies. RLT: reactive lymphoid tissue; MCL: mantle cell lymphoma; DLBCL: diffuse large B-cell lymphoma; HL: Hodgkin lymphoma.

Close modal

EED, another component of the PRC2, exerts a protective influence upon several lymphomagenesis inducers, such as carcinogens, Moloney murine leukemia virus, and irradiation.90,124,125  Intriguingly, mice with a hypomorphic Eed allele show a blockade during thymocyte differentiation at the β-selection checkpoint. These results support the notion that many hematologic malignancies are derived from not fully differentiated cells or cells with either stem cell or progenitor cell properties.

In some lymphomas, such as HL and DLBCL, the expression pattern of PcG members has been studied extensively. In fact, HL is characterized by a unique pattern of expression of PcG proteins, including coexpression of BMI1, MEL18, RING1, human PH1, HPC1, EED, and EZH2.102,103  Moreover, overexpression of RYBP, a PcG member found in Ring1A/Ring1B-containing complexes, is associated with poorer prognosis in HL patients.102  DLBCL samples also overexpress several members of the PRC1 complex, including BMI1, RING1, and HPH1.126 

Regulation of PcG in cancer

Direct alteration of PcG members is not the only phenomenon in cancer that affects PcG function. Other PcG-related systems may be altered in cancer. The E2F/Rb pathway, which is frequently deregulated in cancer, may account for the frequent up-regulation of several PcG members, because E2F transcription factors bind and activate many PcG genes.127  This relationship implies that some PcG members are regulated during cell cycle, but there is also evidence that PcG regulates cell cycle.46,90,128-130 

MicroRNAs are another type of PcG regulator with aberrations in cancer, and the down-regulation of several of these is a general hallmark of cancer. MiR-101 and miR-26a can down-regulate EZH2 and are known to be down-regulated in bladder cancer and BL, respectively,131,132  linking other epigenetic systems to PcG and cancer. In the case of miR-26a, it has been shown that MYC, which is frequently translocated in BL, can repress miR-26a, leading to the up-regulation of EZH2. Other microRNAs, such as miR-137 and miR-214, regulate differentiation through the control of Ezh2 protein levels.133,134  Bmi1 has also been shown to be regulated by microRNAs. In fact, the down-regulation of miR-200c, miR-203, and miR-183 seems to be important to ensure the expression of Bmi1 in cancer cells and mouse embryonic stem cells.135,136  However, the relationship between Polycomb and microRNAs is even more profound. We and others have found that Polycomb complexes also target several microRNA promoters in cancer and during embryogenesis.108,137,138  Moreover, in recent years, it has been suggested that some microRNAs, such as miR-320, can direct gene silencing through the interaction between AGO1 and Polycomb complexes.139 

There are some clues that highlight the importance of PcG in cancer even in the absence of any obvious alteration of the PcG components. In recent years, it has been demonstrated that among the genes that suffer anomalous repression in cancer, there is significant enrichment in those targeted by PcG in ES cells.23,140-146  In fact, some investigators have proposed the existence of a Polycomb repression signature in cancer. This signature is constructed on the basis of the identification of PRC2 cancer-occupied genes by ChIP-on-chip, many of which coincide with those targets found in ES cells, and of down-regulated targets in the tumor by gene-expression profiling. This Polycomb fingerprint can predict clinical outcome in several tumors.147  However, it seems that this repression is not brought about uniquely by the direct action of PcG members; the H3K27me3 mark is able to generate the necessary machinery to induce de novo methylation in cancer (Figure 3), because the 2 systems might be connected. EZH2 can directly control DNA methylation, in some models, by serving as a recruitment platform for DNA methyltransferases.22  For instance, many PcG target genes, such as WT1, RARβ, KLF4, inhibitor of DNA binding 4 (ID4), GATA3, CHD5, and SPI1, accumulate DNA methylation at their promoters in cancer.148  This phenomenon occurs in several hematologic malignancies, including primary lymphomas of the central nervous system and DLBCL,141,142,146  FL,143  BL,141,146  acute lymphoblastic leukemia, and CML.144  However, methylation of PcG target genes appears to be an early event in tumorigenesis, because it may be present in early as well as advanced stages of the disease. The fact that EZH2 and the H3K27me3 modification are present at the p15INK4b gene in AML cell lines and human samples, and that this repression can be accompanied by either H3K4me3 activation mark or DNA methylation in a mutually exclusive pattern, probably indicates that DNA methylation induces a more permanent repression, because activation marks also disappear when DNA is methylated.104 

Figure 3

Cell identity is reflected by changes in chromatin status. In HSCs, PRC1 and PRC2 complexes help to repress genes involved in differentiation to allow the maintenance of stem cell properties. Differentiation induces displacement of PcG members and relocation to promoters of stemness genes. In cancer, the aberrant PcG activity induces a repression of differentiation promoting genes as well as tumor suppressor genes. In this case, the recruitment of DNMTs induces permanent repression of the genes.

Figure 3

Cell identity is reflected by changes in chromatin status. In HSCs, PRC1 and PRC2 complexes help to repress genes involved in differentiation to allow the maintenance of stem cell properties. Differentiation induces displacement of PcG members and relocation to promoters of stemness genes. In cancer, the aberrant PcG activity induces a repression of differentiation promoting genes as well as tumor suppressor genes. In this case, the recruitment of DNMTs induces permanent repression of the genes.

Close modal

Other histone modifiers and complexes

The recent arrival on the scene of new players, the histone demethylases (HDMs), in particular, H3K27me3 demethylases UTX and JMJD3, makes the PcG regulation of normal and tumoral cellular processes even more dynamic, complex, and exciting. These HDMs remove the methyl mark, making it possible to derepress genes marked for silencing by PcG complexes. This is evidence of the dynamic regulation of activation/repression of genes involved in the processes described in this review that are controlled by PcG/TrxG complexes.

UTX and JMJD3 have been found in complex with MLL proteins (TrxG members responsible for H3K4 methylation) and proteins associated with them, such as RbBP5. This represents new evidence of the cooperation between H3K4 methylation and H3K27 demethylation.66,149 

The first evidence of the role of these enzymes in cancer in general and lymphomas in particular, in the form of mutations of UTX and other HDMs in several types of cancer, has recently been published.150 

Implications for therapy

Taken together, the data reviewed here and the new findings from HDMs suggest that alterations in the chromatin modification machinery are a frequent event in cancer and sometimes have prognostic relevance, especially in hematologic malignancies, and that dysregulation of PcG members induces the development of lymphomas and leukemias by blocking the normal differentiation pattern of HSCs and/or conferring stem cell properties on progenitor or differentiated cells.

This has important implications for therapy. Because most chemotherapeutic agents target actively dividing cells, cancer stem cells may be relatively resistant to these kinds of drugs, leading to treatment failure and patient death. Targeting pathways that are especially important for stem cells may offer a better therapeutic window for patient care.

In this regard, there continue to be few drugs developed for use against Polycomb members, despite the importance of this system in cancer cells. Indeed, Tan et al151  reported how DNZep, a drug that disrupts the function of the PRC2 complex, can induce apoptosis in cancer, but not in normal, cells. This is accompanied by the reexpression of genes involved in development and differentiation,152  probably meaning that this kind of treatment induces loss of stemness properties in cancer cells. Some histone deacetylase inhibitors (HDACi), such as LBH589 and LAQ824, can also deplete the PRC2 complex in tumoral cells in AML.153  The authors of this study showed that the inhibition of EZH2 by small interfering RNAs has a synergistically negative effect on the survival of AML cells, favoring the combined use of anti-EZH2 drugs with HDACi for the treatment of AML. In fact, in 2009, the same group published a study in which they combined DNZep with panobinostat, a pan-histone deacetylase inhibitor in AML. The treatment was synergistic and induced apoptosis in AML cells but not in normal CD34+ bone marrow progenitor cells.154  APL patients could also benefit from a PRC2-directed therapy. In this leukemia the PRC2 collaborates with the oncogenic fusion gene PML-RAR-α to repress genes involved in differentiation and to promote tumor development.122  These patients are typically treated with retinoic acid, but depletion of SUZ12 in APL patients reverses the epigenetic marks induced by PML-RAR-α, suggesting that PRC2-targeted therapy could also be used to treat this type of leukemia.

We express our gratitude to Dr Miguel A. Vidal for his critical reading of the manuscript.

This work was supported by grants from Ministerio de Educación y Ciencia (SAF2007-65 957-C02-02); Ministerio de Ciencia e Innovacion (SAF2008-03 871); Fondo de Investigaciones Sanitarias (RTICC RD06/0020/0107); and Fundación Científica de la Asociación Española contra el Cáncer, Spain. D.M.-P. was supported by a fellowship from the Ministerio de Educación y Ciencia (AP2005-3972).

Contribution: D.M.-P. wrote the manuscript; M.A.P. revised the manuscript; and M.S.-B. wrote and revised the manuscript.

Conflict-of-interest disclosure: The authors declare no competing financial interests.

Correspondence: Margarita Sanchez-Beato, Lymphoma Group, Molecular Pathology Program, Centro Nacional de Investigaciones Oncológicas (CNIO), c/ Melchor Fernández Almagro 3, E-28029 Madrid, Spain; e-mail: msbeato@cnio.es.

1
Lewis
 
EB
A gene complex controlling segmentation in Drosophila.
Nature
1978
, vol. 
276
 
5688
(pg. 
565
-
570
)
2
Muller
 
J
Kassis
 
JA
Polycomb response elements and targeting of Polycomb group proteins in Drosophila.
Curr Opin Genet Dev
2006
, vol. 
16
 
5
(pg. 
476
-
484
)
3
Woo
 
CJ
Kharchenko
 
PV
Daheron
 
L
Park
 
PJ
Kingston
 
RE
A region of the human HOXD cluster that confers polycomb-group responsiveness.
Cell
2010
, vol. 
140
 
1
(pg. 
99
-
110
)
4
Gunster
 
MJ
Raaphorst
 
FM
Hamer
 
KM
, et al. 
Differential expression of human Polycomb group proteins in various tissues and cell types.
J Cell Biochem Suppl
2001
suppl 36
(pg. 
129
-
143
)
5
Sanchez-Beato
 
M
Sanchez
 
E
Gonzalez-Carrero
 
J
, et al. 
Variability in the expression of polycomb proteins in different normal and tumoral tissues. A pilot study using tissue microarrays.
Mod Pathol
2006
, vol. 
19
 
5
(pg. 
684
-
694
)
6
Kuzmichev
 
A
Margueron
 
R
Vaquero
 
A
, et al. 
Composition and histone substrates of polycomb repressive group complexes change during cellular differentiation.
Proc Natl Acad Sci U S A
2005
, vol. 
102
 
6
(pg. 
1859
-
1864
)
7
Kuzmichev
 
A
Jenuwein
 
T
Tempst
 
P
Reinberg
 
D
Different EZH2-containing complexes target methylation of histone H1 or nucleosomal histone H3.
Mol Cell
2004
, vol. 
14
 
2
(pg. 
183
-
193
)
8
Wang
 
H
Wang
 
L
Erdjument-Bromage
 
H
, et al. 
Role of histone H2A ubiquitination in Polycomb silencing.
Nature
2004
, vol. 
431
 
7010
(pg. 
873
-
878
)
9
Cao
 
R
Tsukada
 
Y
Zhang
 
Y
Role of Bmi-1 and Ring1A in H2A ubiquitylation and Hox gene silencing.
Mol Cell
2005
, vol. 
20
 
6
(pg. 
845
-
854
)
10
Kerppola
 
TK
Polycomb group complexes—many combinations, many functions.
Trends Cell Biol
2009
, vol. 
19
 
12
(pg. 
692
-
704
)
11
Kajiume
 
T
Ninomiya
 
Y
Ishihara
 
H
Kanno
 
R
Kanno
 
M
Polycomb group gene mel-18 modulates the self-renewal activity and cell cycle status of hematopoietic stem cells.
Exp Hematol
2004
, vol. 
32
 
6
(pg. 
571
-
578
)
12
Lessard
 
J
Baban
 
S
Sauvageau
 
G
Stage-specific expression of polycomb group genes in human bone marrow cells.
Blood
1998
, vol. 
91
 
4
(pg. 
1216
-
1224
)
13
Kajiume
 
T
Ohno
 
N
Sera
 
Y
Kawahara
 
Y
Yuge
 
L
Kobayashi
 
M
Reciprocal expression of Bmi1 and Mel-18 is associated with functioning of primitive hematopoietic cells.
Exp Hematol
2009
, vol. 
37
 
7
(pg. 
857
-
866
)pg. 
e852
 
14
Bernstein
 
E
Duncan
 
EM
Masui
 
O
Gil
 
J
Heard
 
E
Allis
 
CD
Mouse polycomb proteins bind differentially to methylated histone H3 and RNA and are enriched in facultative heterochromatin.
Mol Cell Biol
2006
, vol. 
26
 
7
(pg. 
2560
-
2569
)
15
Lee
 
MG
Villa
 
R
Trojer
 
P
, et al. 
Demethylation of H3K27 regulates polycomb recruitment and H2A ubiquitination.
Science
2007
, vol. 
318
 
5849
(pg. 
447
-
450
)
16
Terranova
 
R
Yokobayashi
 
S
Stadler
 
MB
, et al. 
Polycomb group proteins Ezh2 and Rnf2 direct genomic contraction and imprinted repression in early mouse embryos.
Dev Cell
2008
, vol. 
15
 
5
(pg. 
668
-
679
)
17
Francis
 
NJ
Kingston
 
RE
Woodcock
 
CL
Chromatin compaction by a polycomb group protein complex.
Science
2004
, vol. 
306
 
5701
(pg. 
1574
-
1577
)
18
Margueron
 
R
Li
 
G
Sarma
 
K
, et al. 
Ezh1 and Ezh2 maintain repressive chromatin through different mechanisms.
Mol Cell
2008
, vol. 
32
 
4
(pg. 
503
-
518
)
19
Pasini
 
D
Bracken
 
AP
Hansen
 
JB
Capillo
 
M
Helin
 
K
The polycomb group protein, Suz12, is required for embryonic stem cell differentiation.
Mol Cell Biol
2007
, vol. 
27
 
10
(pg. 
3769
-
3779
)
20
Schoeftner
 
S
Sengupta
 
AK
Kubicek
 
S
, et al. 
Recruitment of PRC1 function at the initiation of X inactivation independent of PRC2 and silencing.
EMBO J
2006
, vol. 
25
 
13
(pg. 
3110
-
3122
)
21
Simon
 
JA
Kingston
 
RE
Mechanisms of polycomb gene silencing: knowns and unknowns.
Nat Rev Mol Cell Biol
2009
, vol. 
10
 
10
(pg. 
697
-
708
)
22
Vire
 
E
Brenner
 
C
Deplus
 
R
, et al. 
The Polycomb group protein EZH2 directly controls DNA methylation.
Nature
2006
, vol. 
439
 
7078
(pg. 
871
-
874
)
23
Schlesinger
 
Y
Straussman
 
R
Keshet
 
I
, et al. 
Polycomb-mediated methylation on Lys27 of histone H3 pre-marks genes for de novo methylation in cancer.
Nat Genet
2007
, vol. 
39
 
2
(pg. 
232
-
236
)
24
Negishi
 
M
Saraya
 
A
Miyagi
 
S
, et al. 
Bmi1 cooperates with Dnmt1-associated protein 1 in gene silencing.
Biochem Biophys Res Commun
2007
, vol. 
353
 
4
(pg. 
992
-
998
)
25
Eden
 
E
Lipson
 
D
Yogev
 
S
Yakhini
 
Z
Discovering motifs in ranked lists of DNA sequences.
PLoS Comput Biol
2007
, vol. 
3
 
3
pg. 
e39
 
26
Kondo
 
Y
Shen
 
L
Cheng
 
AS
, et al. 
Gene silencing in cancer by histone H3 lysine 27 trimethylation independent of promoter DNA methylation.
Nat Genet
2008
, vol. 
40
 
6
(pg. 
741
-
750
)
27
Shao
 
Z
Raible
 
F
Mollaaghababa
 
R
, et al. 
Stabilization of chromatin structure by PRC1, a Polycomb complex.
Cell
1999
, vol. 
98
 
1
(pg. 
37
-
46
)
28
Breiling
 
A
Turner
 
BM
Bianchi
 
ME
Orlando
 
V
General transcription factors bind promoters repressed by Polycomb group proteins.
Nature
2001
, vol. 
412
 
6847
(pg. 
651
-
655
)
29
Stock
 
JK
Giadrossi
 
S
Casanova
 
M
, et al. 
Ring1-mediated ubiquitination of H2A restrains poised RNA polymerase II at bivalent genes in mouse ES cells.
Nat Cell Biol
2007
, vol. 
9
 
12
(pg. 
1428
-
1435
)
30
Ogawa
 
H
Ishiguro
 
K
Gaubatz
 
S
Livingston
 
DM
Nakatani
 
Y
A complex with chromatin modifiers that occupies E2F- and Myc-responsive genes in G0 cells.
Science
2002
, vol. 
296
 
5570
(pg. 
1132
-
1136
)
31
Huynh
 
KD
Fischle
 
W
Verdin
 
E
Bardwell
 
VJ
BCoR, a novel corepressor involved in BCL-6 repression.
Genes Dev
2000
, vol. 
14
 
14
(pg. 
1810
-
1823
)
32
Gearhart
 
MD
Corcoran
 
CM
Wamstad
 
JA
Bardwell
 
VJ
Polycomb group and SCF ubiquitin ligases are found in a novel BCOR complex that is recruited to BCL6 targets.
Mol Cell Biol
2006
, vol. 
26
 
18
(pg. 
6880
-
6889
)
33
Sanchez
 
C
Sanchez
 
I
Demmers
 
JA
Rodriguez
 
P
Strouboulis
 
J
Vidal
 
M
Proteomics analysis of Ring1B/Rnf2 interactors identifies a novel complex with the Fbxl10/Jhdm1B histone demethylase and the Bcl6 interacting corepressor.
Mol Cell Proteomics
2007
, vol. 
6
 
5
(pg. 
820
-
834
)
34
Lee
 
TI
Jenner
 
RG
Boyer
 
LA
, et al. 
Control of developmental regulators by Polycomb in human embryonic stem cells.
Cell
2006
, vol. 
125
 
2
(pg. 
301
-
313
)
35
Endoh
 
M
Endo
 
TA
Endoh
 
T
, et al. 
Polycomb group proteins Ring1A/B are functionally linked to the core transcriptional regulatory circuitry to maintain ES cell identity.
Development
2008
, vol. 
135
 
8
(pg. 
1513
-
1524
)
36
Boyer
 
LA
Plath
 
K
Zeitlinger
 
J
, et al. 
Polycomb complexes repress developmental regulators in murine embryonic stem cells.
Nature
2006
, vol. 
441
 
7091
(pg. 
349
-
353
)
37
Rinn
 
JL
Kertesz
 
M
Wang
 
JK
, et al. 
Functional demarcation of active and silent chromatin domains in human HOX loci by noncoding RNAs.
Cell
2007
, vol. 
129
 
7
(pg. 
1311
-
1323
)
38
Pandey
 
RR
Mondal
 
T
Mohammad
 
F
, et al. 
Kcnq1ot1 antisense noncoding RNA mediates lineage-specific transcriptional silencing through chromatin-level regulation.
Mol Cell
2008
, vol. 
32
 
2
(pg. 
232
-
246
)
39
Grimaud
 
C
Bantignies
 
F
Pal-Bhadra
 
M
Ghana
 
P
Bhadra
 
U
Cavalli
 
G
RNAi components are required for nuclear clustering of Polycomb group response elements.
Cell
2006
, vol. 
124
 
5
(pg. 
957
-
971
)
40
Kanhere
 
A
Viiri
 
K
Araújo
 
CC
, et al. 
Short RNAs are transcribed from repressed polycomb target genes and interact with polycomb repressive complex-2.
Mol Cell
2010
, vol. 
38
 
5
(pg. 
675
-
688
)
41
Hekimoglu
 
B
Ringrose
 
L
Non-coding RNAs in polycomb/trithorax regulation.
RNA Biol
2009
, vol. 
6
 
2
(pg. 
129
-
137
)
42
Chandrasekhar
 
K
Joanne
 
W
Faizaan
 
M
The long and the short of it: RNA-directed chromatin asymmetry in mammalian X-chromosome inactivation.
FEBS Lett
2009
, vol. 
583
 
5
(pg. 
857
-
864
)
43
Kim
 
DH
Villeneuve
 
LM
Morris
 
KV
Rossi
 
JJ
Argonaute-1 directs siRNA-mediated transcriptional gene silencing in human cells.
Nat Struct Mol Biol
2006
, vol. 
13
 
9
(pg. 
793
-
797
)
44
Rajasekhar
 
VK
Begemann
 
M
Concise review: roles of polycomb group proteins in development and disease: a stem cell perspective.
Stem Cells
2007
, vol. 
25
 
10
(pg. 
2498
-
2510
)
45
Park
 
IK
Qian
 
D
Kiel
 
M
, et al. 
Bmi-1 is required for maintenance of adult self-renewing haematopoietic stem cells.
Nature
2003
, vol. 
423
 
6937
(pg. 
302
-
305
)
46
Jacobs
 
JJ
Scheijen
 
B
Voncken
 
JW
Kieboom
 
K
Berns
 
A
van Lohuizen
 
M
Bmi-1 collaborates with c-Myc in tumorigenesis by inhibiting c-Myc-induced apoptosis via INK4a/ARF.
Genes Dev
1999
, vol. 
13
 
20
(pg. 
2678
-
2690
)
47
Bruggeman
 
SW
Valk-Lingbeek
 
ME
van der Stoop
 
PP
, et al. 
Ink4a and Arf differentially affect cell proliferation and neural stem cell self-renewal in Bmi1-deficient mice.
Genes Dev
2005
, vol. 
19
 
12
(pg. 
1438
-
1443
)
48
Martin-Perez
 
D
Sanchez
 
E
Maestre
 
L
, et al. 
Deregulated expression of the polycomb-group protein SUZ12 target genes characterizes mantle cell lymphoma.
Am J Pathol
2010
, vol. 
177
 
2
(pg. 
930
-
942
)
49
Classen
 
AK
Bunker
 
BD
Harvey
 
KF
Vaccari
 
T
Bilder
 
D
A tumor suppressor activity of Drosophila Polycomb genes mediated by JAK-STAT signaling.
Nat Genet
2009
, vol. 
41
 
10
(pg. 
1150
-
1155
)
50
Gonzalez
 
I
Simon
 
R
Busturia
 
A
The Polyhomeotic protein induces hyperplastic tissue overgrowth through the activation of the JAK/STAT pathway.
Cell Cycle
2009
, vol. 
8
 
24
(pg. 
4103
-
4111
)
51
Akasaka
 
T
Tsuji
 
K
Kawahira
 
H
, et al. 
The role of mel-18, a mammalian Polycomb group gene, during IL-7-dependent proliferation of lymphocyte precursors.
Immunity
1997
, vol. 
7
 
1
(pg. 
135
-
146
)
52
van der Lugt
 
NM
Domen
 
J
Linders
 
K
, et al. 
Posterior transformation, neurological abnormalities, and severe hematopoietic defects in mice with a targeted deletion of the bmi-1 proto-oncogene.
Genes Dev
1994
, vol. 
8
 
7
(pg. 
757
-
769
)
53
Core
 
N
Bel
 
S
Gaunt
 
SJ
, et al. 
Altered cellular proliferation and mesoderm patterning in Polycomb-M33–deficient mice.
Development
1997
, vol. 
124
 
3
(pg. 
721
-
729
)
54
Bea
 
S
Tort
 
F
Pinyol
 
M
, et al. 
BMI-1 gene amplification and overexpression in hematological malignancies occur mainly in mantle cell lymphomas.
Cancer Res
2001
, vol. 
61
 
6
(pg. 
2409
-
2412
)
55
Kanno
 
M
Hasegawa
 
M
Ishida
 
A
Isono
 
K
Taniguchi
 
M
mel-18, a Polycomb group-related mammalian gene, encodes a transcriptional negative regulator with tumor suppressive activity.
EMBO J
1995
, vol. 
14
 
22
(pg. 
5672
-
5678
)
56
Martinez
 
AM
Cavalli
 
G
The role of polycomb group proteins in cell cycle regulation during development.
Cell Cycle
2006
, vol. 
5
 
11
(pg. 
1189
-
1197
)
57
Cales
 
C
Roman-Trufero
 
M
Pavon
 
L
, et al. 
Inactivation of the polycomb group protein Ring1B unveils an antiproliferative role in hematopoietic cell expansion and cooperation with tumorigenesis associated with Ink4a deletion.
Mol Cell Biol
2008
, vol. 
28
 
3
(pg. 
1018
-
1028
)
58
Sparmann
 
A
van Lohuizen
 
M
Polycomb silencers control cell fate, development, and cancer.
Nat Rev Cancer
2006
, vol. 
6
 
11
(pg. 
846
-
856
)
59
Pasini
 
D
Bracken
 
AP
Jensen
 
MR
Lazzerini Denchi
 
E
Helin
 
K
Suz12 is essential for mouse development and for EZH2 histone methyltransferase activity.
EMBO J
2004
, vol. 
23
 
20
(pg. 
4061
-
4071
)
60
Niswander
 
L
Yee
 
D
Rinchik
 
EM
Russell
 
LB
Magnuson
 
T
The albino deletion complex and early postimplantation survival in the mouse.
Development
1988
, vol. 
102
 
1
(pg. 
45
-
53
)
61
Bernstein
 
BE
Mikkelsen
 
TS
Xie
 
X
, et al. 
A bivalent chromatin structure marks key developmental genes in embryonic stem cells.
Cell
2006
, vol. 
125
 
2
(pg. 
315
-
326
)
62
Roh
 
TY
Cuddapah
 
S
Cui
 
K
Zhao
 
K
The genomic landscape of histone modifications in human T cells.
Proc Natl Acad Sci U S A
2006
, vol. 
103
 
43
(pg. 
15782
-
15787
)
63
Araki
 
Y
Wang
 
Z
Zang
 
C
, et al. 
Genome-wide analysis of histone methylation reveals chromatin state-based regulation of gene transcription and function of memory CD8+ T cells.
Immunity
2009
, vol. 
30
 
6
(pg. 
912
-
925
)
64
Wei
 
G
Wei
 
L
Zhu
 
J
, et al. 
Global mapping of H3K4me3 and H3K27me3 reveals specificity and plasticity in lineage fate determination of differentiating CD4+ T cells.
Immunity
2009
, vol. 
30
 
1
(pg. 
155
-
167
)
65
Barski
 
A
Cuddapah
 
S
Cui
 
K
, et al. 
High-resolution profiling of histone methylations in the human genome.
Cell
2007
, vol. 
129
 
4
(pg. 
823
-
837
)
66
De Santa
 
F
Totaro
 
MG
Prosperini
 
E
Notarbartolo
 
S
Testa
 
G
Natoli
 
G
The histone H3 lysine-27 demethylase Jmjd3 links inflammation to inhibition of polycomb-mediated gene silencing.
Cell
2007
, vol. 
130
 
6
(pg. 
1083
-
1094
)
67
Bracken
 
AP
Kleine-Kohlbrecher
 
D
Dietrich
 
N
, et al. 
The Polycomb group proteins bind throughout the INK4A-ARF locus and are disassociated in senescent cells.
Genes Dev
2007
, vol. 
21
 
5
(pg. 
525
-
530
)
68
Chamberlain
 
SJ
Yee
 
D
Magnuson
 
T
Polycomb repressive complex 2 is dispensable for maintenance of embryonic stem cell pluripotency.
Stem Cells
2008
, vol. 
26
 
6
(pg. 
1496
-
1505
)
69
Mikkelsen
 
TS
Ku
 
M
Jaffe
 
DB
, et al. 
Genome-wide maps of chromatin state in pluripotent and lineage-committed cells.
Nature
2007
, vol. 
448
 
7153
(pg. 
553
-
560
)
70
Bracken
 
AP
Dietrich
 
N
Pasini
 
D
Hansen
 
KH
Helin
 
K
Genome-wide mapping of Polycomb target genes unravels their roles in cell fate transitions.
Genes Dev
2006
, vol. 
20
 
9
(pg. 
1123
-
1136
)
71
Boyer
 
LA
Lee
 
TI
Cole
 
MF
, et al. 
Core transcriptional regulatory circuitry in human embryonic stem cells.
Cell
2005
, vol. 
122
 
6
(pg. 
947
-
956
)
72
Molofsky
 
AV
Pardal
 
R
Iwashita
 
T
Park
 
I-K
Clarke
 
MF
Morrison
 
SJ
Bmi-1 dependence distinguishes neural stem cell self-renewal from progenitor proliferation.
Nature
2003
, vol. 
425
 
6961
(pg. 
962
-
967
)
73
Molofsky
 
AV
He
 
S
Bydon
 
M
Morrison
 
SJ
Pardal
 
R
Bmi-1 promotes neural stem cell self-renewal and neural development but not mouse growth and survival by repressing the p16Ink4a and p19Arf senescence pathways.
Genes Dev
2005
, vol. 
19
 
12
(pg. 
1432
-
1437
)
74
Roman-Trufero
 
M
Mendez-Gomez
 
HR
Perez
 
C
, et al. 
Maintenance of undifferentiated state and self-renewal of embryonic neural stem cells by Polycomb protein Ring1B.
Stem Cells
2009
, vol. 
27
 
7
(pg. 
1559
-
1570
)
75
Liu
 
S
Dontu
 
G
Mantle
 
ID
, et al. 
Hedgehog signaling and Bmi-1 regulate self-renewal of normal and malignant human mammary stem cells.
Cancer Res
2006
, vol. 
66
 
12
(pg. 
6063
-
6071
)
76
Reya
 
T
Duncan
 
AW
Ailles
 
L
, et al. 
A role for Wnt signalling in self-renewal of haematopoietic stem cells.
Nature
2003
, vol. 
423
 
6938
(pg. 
409
-
414
)
77
Bhardwaj
 
G
Murdoch
 
B
Wu
 
D
, et al. 
Sonic hedgehog induces the proliferation of primitive human hematopoietic cells via BMP regulation.
Nat Immunol
2001
, vol. 
2
 
2
(pg. 
172
-
180
)
78
Argiropoulos
 
B
Humphries
 
RK
Hox genes in hematopoiesis and leukemogenesis.
Oncogene
2007
, vol. 
26
 
47
(pg. 
6766
-
6776
)
79
Antonchuk
 
J
Sauvageau
 
G
Humphries
 
RK
HOXB4-induced expansion of adult hematopoietic stem cells ex vivo.
Cell
2002
, vol. 
109
 
1
(pg. 
39
-
45
)
80
Schiedlmeier
 
B
Santos
 
AC
Ribeiro
 
A
, et al. 
HOXB4's road map to stem cell expansion.
Proc Natl Acad Sci U S A
2007
, vol. 
104
 
43
(pg. 
16952
-
16957
)
81
Bjornsson
 
JM
Larsson
 
N
Brun
 
AC
, et al. 
Reduced proliferative capacity of hematopoietic stem cells deficient in Hoxb3 and Hoxb4.
Mol Cell Biol
2003
, vol. 
23
 
11
(pg. 
3872
-
3883
)
82
Brun
 
AC
Bjornsson
 
JM
Magnusson
 
M
, et al. 
Hoxb4-deficient mice undergo normal hematopoietic development but exhibit a mild proliferation defect in hematopoietic stem cells.
Blood
2004
, vol. 
103
 
11
(pg. 
4126
-
4133
)
83
Ohta
 
H
Sawada
 
A
Kim
 
JY
, et al. 
Polycomb group gene rae28 is required for sustaining activity of hematopoietic stem cells.
J Exp Med
2002
, vol. 
195
 
6
(pg. 
759
-
770
)
84
Jacobs
 
JJ
van Lohuizen
 
M
Polycomb repression: from cellular memory to cellular proliferation and cancer.
Biochim Biophys Acta
2002
, vol. 
1602
 
2
(pg. 
151
-
161
)
85
Orlando
 
V
Polycomb, epigenomes, and control of cell identity.
Cell
2003
, vol. 
112
 
5
(pg. 
599
-
606
)
86
Lessard
 
J
Sauvageau
 
G
Bmi-1 determines the proliferative capacity of normal and leukaemic stem cells.
Nature
2003
, vol. 
423
 
6937
(pg. 
255
-
260
)
87
Iwama
 
A
Oguro
 
H
Negishi
 
M
, et al. 
Enhanced self-renewal of hematopoietic stem cells mediated by the polycomb gene product, Bmi-1.
Immunity
2004
, vol. 
21
 
6
(pg. 
843
-
851
)
88
Oguro
 
H
Iwama
 
A
Morita
 
Y
Kamijo
 
T
van Lohuizen
 
M
Nakauchi
 
H
Differential impact of Ink4a and Arf on hematopoietic stem cells and their bone marrow microenvironment in Bmi1-deficient mice.
J Exp Med
2006
, vol. 
203
 
10
(pg. 
2247
-
2253
)
89
Kamminga
 
LM
Bystrykh
 
LV
de Boer
 
A
, et al. 
The Polycomb group gene, Ezh2, prevents hematopoietic stem cell exhaustion.
Blood
2006
, vol. 
107
 
5
(pg. 
2170
-
2179
)
90
Lessard
 
J
Schumacher
 
A
Thorsteinsdottir
 
U
van Lohuizen
 
M
Magnuson
 
T
Sauvageau
 
G
Functional antagonism of the Polycomb-group genes, eed and Bmi1, in hemopoietic cell proliferation.
Genes Dev
1999
, vol. 
13
 
20
(pg. 
2691
-
2703
)
91
Kim
 
JY
Sawada
 
A
Tokimasa
 
S
, et al. 
Defective long-term repopulating ability in hematopoietic stem cells lacking the Polycomb-group gene, rae28.
Eur J Haematol
2004
, vol. 
73
 
2
(pg. 
75
-
84
)
92
Chen
 
J
Odenike
 
O
Rowley
 
JD
Leukaemogenesis: more than mutant genes.
Nat Rev Cancer
2010
, vol. 
10
 
1
(pg. 
23
-
36
)
93
Taipale
 
J
Beachy
 
PA
The Hedgehog and Wnt signalling pathways in cancer.
Nature
2001
, vol. 
411
 
6835
(pg. 
349
-
354
)
94
Monk
 
M
Holding
 
C
Human embryonic genes re-expressed in cancer cells.
Oncogene
2001
, vol. 
20
 
56
(pg. 
8085
-
8091
)
95
Gupta
 
PB
Chaffer
 
CL
Weinberg
 
RA
Cancer stem cells: mirage or reality?
Nat Med
2009
, vol. 
15
 
9
(pg. 
1010
-
1012
)
96
Ben-Porath
 
I
Thomson
 
MW
Carey
 
VJ
, et al. 
An embryonic stem cell-like gene expression signature in poorly differentiated aggressive human tumors.
Nat Genet
2008
, vol. 
40
 
5
(pg. 
499
-
507
)
97
Widschwendter
 
M
Fiegl
 
H
Egle
 
D
, et al. 
Epigenetic stem cell signature in cancer.
Nat Genet
2007
, vol. 
39
 
2
(pg. 
157
-
158
)
98
Roesch
 
A
Fukunaga-Kalabis
 
M
Schmidt
 
EC
, et al. 
A temporarily distinct subpopulation of slow-cycling melanoma cells is required for continuous tumor growth.
Cell
2010
, vol. 
141
 
4
(pg. 
583
-
594
)
99
van Lohuizen
 
M
Verbeek
 
S
Scheijen
 
B
Wientjens
 
E
van der Gulden
 
H
Berns
 
A
Identification of cooperating oncogenes in E mu-myc transgenic mice by provirus tagging.
Cell
1991
, vol. 
65
 
5
(pg. 
737
-
752
)
100
Haupt
 
Y
Alexander
 
WS
Barri
 
G
Klinken
 
SP
Adams
 
JM
Novel zinc finger gene implicated as myc collaborator by retrovirally accelerated lymphomagenesis in E mu-myc transgenic mice.
Cell
1991
, vol. 
65
 
5
(pg. 
753
-
763
)
101
Liu
 
L
Andrews
 
LG
Tollefsbol
 
TO
Loss of the human polycomb group protein BMI1 promotes cancer-specific cell death.
Oncogene
2006
, vol. 
25
 
31
(pg. 
4370
-
4375
)
102
Sanchez-Beato
 
M
Sanchez
 
E
Garcia
 
JF
, et al. 
Abnormal PcG protein expression in Hodgkin's lymphoma. Relation with E2F6 and NFkappaB transcription factors.
J Pathol
2004
, vol. 
204
 
5
(pg. 
528
-
537
)
103
Dukers
 
DF
van Galen
 
JC
Giroth
 
C
, et al. 
Unique polycomb gene expression pattern in Hodgkin's lymphoma and Hodgkin's lymphoma-derived cell lines.
Am J Pathol
2004
, vol. 
164
 
3
(pg. 
873
-
881
)
104
Paul
 
TA
Bies
 
J
Small
 
D
Wolff
 
L
Signatures of polycomb repression and reduced H3K4 trimethylation are associated with p15INK4b DNA methylation in AML.
Blood
2010
, vol. 
115
 
15
(pg. 
3098
-
3108
)
105
Gil
 
J
Bernard
 
D
Martinez
 
D
Beach
 
D
Polycomb CBX7 has a unifying role in cellular lifespan.
Nat Cell Biol
2004
, vol. 
6
 
1
(pg. 
67
-
72
)
106
Dimri
 
GP
Martinez
 
JL
Jacobs
 
JJ
, et al. 
The Bmi-1 oncogene induces telomerase activity and immortalizes human mammary epithelial cells.
Cancer Res
2002
, vol. 
62
 
16
(pg. 
4736
-
4745
)
107
Song
 
LB
Li
 
J
Liao
 
WT
, et al. 
The polycomb group protein Bmi-1 represses the tumor suppressor, PTEN, and induces epithelial-mesenchymal transition in human nasopharyngeal epithelial cells.
J Clin Invest
2009
, vol. 
119
 
12
(pg. 
3626
-
3636
)
108
Martin-Perez
 
D
Sanchez
 
E
Maestre
 
L
, et al. 
Deregulated expression of the Polycomb-group protein, SUZ12 target genes characterizes mantle cell lymphoma.
Am J Pathol
2010
, vol. 
177
 
2
(pg. 
933
-
942
)
109
Varambally
 
S
Dhanasekaran
 
SM
Zhou
 
M
, et al. 
The polycomb group protein, EZH2, is involved in progression of prostate cancer.
Nature
2002
, vol. 
419
 
6907
(pg. 
624
-
629
)
110
Smith
 
KS
Chanda
 
SK
Lingbeek
 
M
, et al. 
Bmi-1 regulation of INK4A-ARF is a downstream requirement for transformation of hematopoietic progenitors by E2a-Pbx1.
Mol Cell
2003
, vol. 
12
 
2
(pg. 
393
-
400
)
111
Visser
 
HP
Gunster
 
MJ
Kluin-Nelemans
 
HC
, et al. 
The Polycomb group protein, EZH2, is upregulated in proliferating, cultured human mantle cell lymphoma.
Br J Haematol
2001
, vol. 
112
 
4
(pg. 
950
-
958
)
112
Raaphorst
 
FM
Otte
 
AP
van Kemenade
 
FJ
, et al. 
Distinct BMI-1 and EZH2 expression patterns in thymocytes and mature T cells suggest a role for Polycomb genes in human T cell differentiation.
J Immunol
2001
, vol. 
166
 
10
(pg. 
5925
-
5934
)
113
Raaphorst
 
FM
van Kemenade
 
FJ
Blokzijl
 
T
, et al. 
Coexpression of BMI-1 and EZH2 polycomb group genes in Reed-Sternberg cells of Hodgkin's disease.
Am J Pathol
2000
, vol. 
157
 
3
(pg. 
709
-
715
)
114
Morin
 
RD
Johnson
 
NA
Severson
 
TM
, et al. 
Somatic mutations altering EZH2 (Tyr641) in follicular and diffuse large B-cell lymphomas of germinal-center origin.
Nat Genet
2010
, vol. 
42
 
2
(pg. 
181
-
185
)
115
Su
 
IH
Basavaraj
 
A
Krutchinsky
 
AN
, et al. 
Ezh2 controls B cell development through histone H3 methylation and Igh rearrangement.
Nat Immunol
2003
, vol. 
4
 
2
(pg. 
124
-
131
)
116
Cardoso
 
C
Mignon
 
C
Hetet
 
G
Grandchamps
 
B
Fontes
 
M
Colleaux
 
L
The human EZH2 gene: genomic organisation and revised mapping in 7q35 within the critical region for malignant myeloid disorders.
Eur J Hum Genet
2000
, vol. 
8
 
3
(pg. 
174
-
180
)
117
Nikoloski
 
G
Langemeijer
 
SMC
Kuiper
 
RP
, et al. 
Somatic mutations of the histone methyltransferase gene EZH2 in myelodysplastic syndromes
Nat Genet
2010
, vol. 
42
 
8
(pg. 
665
-
667
)
118
Ernst
 
T
Chase
 
AJ
Score
 
J
, et al. 
Inactivating mutations of the histone methyltransferase gene, EZH2, in myeloid disorders
Nat Genet
2010
, vol. 
42
 
8
(pg. 
722
-
726
)
119
Koontz
 
JI
Soreng
 
AL
Nucci
 
M
, et al. 
Frequent fusion of the JAZF1 and JJAZ1 genes in endometrial stromal tumors.
Proc Natl Acad Sci U S A
2001
, vol. 
98
 
11
(pg. 
6348
-
6353
)
120
Li
 
H
Ma
 
X
Wang
 
J
Koontz
 
J
Nucci
 
M
Sklar
 
J
Effects of rearrangement and allelic exclusion of JJAZ1/SUZ12 on cell proliferation and survival.
Proc Natl Acad Sci U S A
2007
, vol. 
104
 
50
(pg. 
20001
-
20006
)
121
Boukarabila
 
H
Saurin
 
AJ
Batsche
 
E
, et al. 
The PRC1 Polycomb group complex interacts with PLZF/RARA to mediate leukemic transformation.
Genes Dev
2009
, vol. 
23
 
10
(pg. 
1195
-
1206
)
122
Villa
 
R
Pasini
 
D
Gutierrez
 
A
, et al. 
Role of the polycomb repressive complex 2 in acute promyelocytic leukemia.
Cancer Cell
2007
, vol. 
11
 
6
(pg. 
513
-
525
)
123
Pizzatti
 
L
Binato
 
R
Cofre
 
J
, et al. 
SUZ12 is a candidate target of the non-canonical WNT pathway in the progression of chronic myeloid leukemia.
Genes Chrom Cancer
2010
, vol. 
49
 
2
(pg. 
107
-
118
)
124
Sauvageau
 
M
Miller
 
M
Lemieux
 
S
Lessard
 
J
Hebert
 
J
Sauvageau
 
G
Quantitative expression profiling guided by common retroviral insertion sites reveals novel and cell-type–specific cancer genes in leukemia.
Blood
2008
, vol. 
111
 
2
(pg. 
790
-
799
)
125
Richie
 
ER
Schumacher
 
A
Angel
 
JM
Holloway
 
M
Rinchik
 
EM
Magnuson
 
T
The Polycomb-group gene, eed, regulates thymocyte differentiation and suppresses the development of carcinogen-induced T-cell lymphomas.
Oncogene
2002
, vol. 
21
 
2
(pg. 
299
-
306
)
126
Raaphorst
 
FM
Vermeer
 
M
Fieret
 
E
, et al. 
Site-specific expression of polycomb-group genes encoding the HPC-HPH/PRC1 complex in clinically defined primary nodal and cutaneous large B-cell lymphomas.
Am J Pathol
2004
, vol. 
164
 
2
(pg. 
533
-
542
)
127
Bracken
 
AP
Pasini
 
D
Capra
 
M
Prosperini
 
E
Colli
 
E
Helin
 
K
EZH2 is downstream of the pRB-E2F pathway, essential for proliferation, and amplified in cancer.
EMBO J
2003
, vol. 
22
 
20
(pg. 
5323
-
5335
)
128
Lee
 
JY
Jang
 
KS
Shin
 
DH
, et al. 
Mel-18 negatively regulates INK4a/ARF-independent cell cycle progression via Akt inactivation in breast cancer.
Cancer Res
2008
, vol. 
68
 
11
(pg. 
4201
-
4209
)
129
Martinez
 
AM
Colomb
 
S
Dejardin
 
J
Bantignies
 
F
Cavalli
 
G
Polycomb group–dependent Cyclin A repression in Drosophila.
Genes Dev
2006
, vol. 
20
 
4
(pg. 
501
-
513
)
130
Chun
 
T
Rho
 
SB
Byun
 
HJ
Lee
 
JY
Kong
 
G
The polycomb group gene product, Mel-18, interacts with cyclin D2 and modulates its activity.
FEBS Lett
2005
, vol. 
579
 
24
(pg. 
5275
-
5280
)
131
Sander
 
S
Bullinger
 
L
Klapproth
 
K
, et al. 
MYC stimulates EZH2 expression by repression of its negative regulator, miR-26a.
Blood
2008
, vol. 
112
 
10
(pg. 
4202
-
4212
)
132
Friedman
 
JM
Liang
 
G
Liu
 
CC
, et al. 
The putative tumor suppressor, microRNA-101, modulates the cancer epigenome by repressing the polycomb group protein, EZH2.
Cancer Res
2009
, vol. 
69
 
6
(pg. 
2623
-
2629
)
133
Szulwach
 
KE
Li
 
X
Smrt
 
RD
, et al. 
Cross talk between microRNA and epigenetic regulation in adult neurogenesis.
J Cell Biol
2010
, vol. 
189
 
1
(pg. 
127
-
141
)
134
Juan
 
AH
Kumar
 
RM
Marx
 
JG
Young
 
RA
Sartorelli
 
V
Mir-214–dependent regulation of the polycomb protein, Ezh2, in skeletal muscle and embryonic stem cells.
Mol Cell
2009
, vol. 
36
 
1
(pg. 
61
-
74
)
135
Wellner
 
U
Schubert
 
J
Burk
 
UC
, et al. 
The EMT-activator ZEB1 promotes tumorigenicity by repressing stemness-inhibiting microRNAs.
Nat Cell Biol
2009
, vol. 
11
 
12
(pg. 
1487
-
1495
)
136
Shimono
 
Y
Zabala
 
M
Cho
 
RW
, et al. 
Downregulation of miRNA-200c links breast cancer stem cells with normal stem cells.
Cell
2009
, vol. 
138
 
3
(pg. 
592
-
603
)
137
Marson
 
A
Levine
 
SS
Cole
 
MF
, et al. 
Connecting microRNA genes to the core transcriptional regulatory circuitry of embryonic stem cells.
Cell
2008
, vol. 
134
 
3
(pg. 
521
-
533
)
138
Wang
 
H
Garzon
 
R
Sun
 
H
, et al. 
NF-kappaB-YY1-miR-29 regulatory circuitry in skeletal myogenesis and rhabdomyosarcoma.
Cancer Cell
2008
, vol. 
14
 
5
(pg. 
369
-
381
)
139
Kim
 
DH
Saetrom
 
P
Snove
 
O
Rossi
 
JJ
MicroRNA-directed transcriptional gene silencing in mammalian cells.
Proc Natl Acad Sci U S A
2008
, vol. 
105
 
42
(pg. 
16230
-
16235
)
140
Wu
 
X
Rauch
 
TA
Zhong
 
X
, et al. 
CpG island hypermethylation in human astrocytomas.
Cancer Res
2010
, vol. 
70
 
7
(pg. 
2718
-
2727
)
141
Martin-Subero
 
JI
Kreuz
 
M
Bibikova
 
M
, et al. 
New insights into the biology and origin of mature aggressive B-cell lymphomas by combined epigenomic, genomic, and transcriptional profiling.
Blood
2009
, vol. 
113
 
11
(pg. 
2488
-
2497
)
142
Richter
 
J
Ammerpohl
 
O
Martin-Subero
 
JI
, et al. 
Array-based DNA methylation profiling of primary lymphomas of the central nervous system.
BMC Cancer
2009
, vol. 
9
 
1
pg. 
455
 
143
O'Riain
 
C
O'Shea
 
DM
Yang
 
Y
, et al. 
Array-based DNA methylation profiling in follicular lymphoma.
Leukemia
2009
, vol. 
23
 
10
(pg. 
1858
-
1866
)
144
Dunwell
 
T
Hesson
 
L
Rauch
 
TA
, et al. 
A genome-wide screen identifies frequently methylated genes in haematological and epithelial cancers.
Mol Cancer
2010
, vol. 
9
 pg. 
44
 
145
Martin-Subero
 
JI
Ammerpohl
 
O
Bibikova
 
M
, et al. 
A comprehensive microarray-based DNA methylation study of 367 hematological neoplasms.
PLoS One
2009
, vol. 
4
 
9
pg. 
e6986
 
146
Wang
 
XM
Greiner
 
TC
Bibikova
 
M
, et al. 
Identification and functional relevance of de novo DNA methylation in cancerous B-cell populations.
J Cell Biochem
2010
, vol. 
109
 
4
(pg. 
818
-
827
)
147
Yu
 
J
Rhodes
 
DR
Tomlins
 
SA
, et al. 
A polycomb repression signature in metastatic prostate cancer predicts cancer outcome.
Cancer Res
2007
, vol. 
67
 
22
(pg. 
10657
-
10663
)
148
Bracken
 
AP
Helin
 
K
Polycomb group proteins: navigators of lineage pathways led astray in cancer.
Nat Rev Cancer
2009
, vol. 
9
 
11
(pg. 
773
-
784
)
149
Issaeva
 
I
Zonis
 
Y
Rozovskaia
 
T
, et al. 
Knockdown of ALR (MLL2) reveals ALR target genes and leads to alterations in cell adhesion and growth.
Mol Cell Biol
2007
, vol. 
27
 
5
(pg. 
1889
-
1903
)
150
van Haaften
 
G
Dalgliesh
 
GL
Davies
 
H
, et al. 
Somatic mutations of the histone H3K27 demethylase gene UTX in human cancer.
Nat Genet
2009
, vol. 
41
 
5
(pg. 
521
-
523
)
151
Tan
 
J
Yang
 
X
Zhuang
 
L
, et al. 
Pharmacologic disruption of Polycomb-repressive complex 2–mediated gene repression selectively induces apoptosis in cancer cells.
Genes Dev
2007
, vol. 
21
 
9
(pg. 
1050
-
1063
)
152
Miranda
 
TB
Cortez
 
CC
Yoo
 
CB
, et al. 
DZNep is a global histone methylation inhibitor that reactivates developmental genes not silenced by DNA methylation.
Mol Cancer Ther
2009
, vol. 
8
 
6
(pg. 
1579
-
1588
)
153
Fiskus
 
W
Pranpat
 
M
Balasis
 
M
, et al. 
Histone deacetylase inhibitors deplete enhancer of zeste 2 and associated polycomb repressive complex 2 proteins in human acute leukemia cells.
Mol Cancer Ther
2006
, vol. 
5
 
12
(pg. 
3096
-
3104
)
154
Fiskus
 
W
Wang
 
Y
Sreekumar
 
A
, et al. 
Combined epigenetic therapy with the histone methyltransferase EZH2 inhibitor, 3-deazaneplanocin A, and the histone deacetylase inhibitor, panobinostat, against human AML cells.
Blood
2009
, vol. 
114
 
13
(pg. 
2733
-
2743
)
155
Sawa
 
M
Yamamoto
 
K
Yokozawa
 
T
, et al. 
BMI-1 is highly expressed in M0-subtype acute myeloid leukemia.
Int J Hematol
2005
, vol. 
82
 
1
(pg. 
42
-
47
)
156
Hoenerhoff
 
MJ
Chu
 
I
Barkan
 
D
, et al. 
BMI1 cooperates with H-RAS to induce an aggressive breast cancer phenotype with brain metastases.
Oncogene
2009
, vol. 
28
 
34
(pg. 
3022
-
3032
)
157
Bruggeman
 
SW
Hulsman
 
D
Tanger
 
E
, et al. 
Bmi1 controls tumor development in an Ink4a/Arf-independent manner in a mouse model for glioma.
Cancer Cell
2007
, vol. 
12
 
4
(pg. 
328
-
341
)
158
Nowak
 
K
Kerl
 
K
Fehr
 
D
, et al. 
BMI1 is a target gene of E2F-1 and is strongly expressed in primary neuroblastomas.
Nucleic Acids Res
2006
, vol. 
34
 
6
(pg. 
1745
-
1754
)
159
Leung
 
C
Lingbeek
 
M
Shakhova
 
O
, et al. 
Bmi1 is essential for cerebellar development and is overexpressed in human medulloblastomas.
Nature
2004
, vol. 
428
 
6980
(pg. 
337
-
341
)
160
Vonlanthen
 
S
Heighway
 
J
Altermatt
 
HJ
, et al. 
The bmi-1 oncoprotein is differentially expressed in non-small cell lung cancer and correlates with INK4A-ARF locus expression.
Br J Cancer
2001
, vol. 
84
 
10
(pg. 
1372
-
1376
)
161
Balasubramanian
 
S
Adhikary
 
G
Eckert
 
RL
The Bmi-1 polycomb protein antagonizes the (-)-epigallocatechin-3-gallate-dependent suppression of skin cancer cell survival.
Carcinogenesis
2010
, vol. 
31
 
3
(pg. 
496
-
503
)
162
Tokimasa
 
S
Ohta
 
H
Sawada
 
A
, et al. 
Lack of the Polycomb-group gene, rae28, causes maturation arrest at the early B-cell developmental stage.
Exp Hematol
2001
, vol. 
29
 
1
(pg. 
93
-
103
)
163
Scott
 
CL
Gil
 
J
Hernando
 
E
, et al. 
Role of the chromobox protein CBX7 in lymphomagenesis.
Proc Natl Acad Sci U S A
2007
, vol. 
104
 
13
(pg. 
5389
-
5394
)
164
Ferreira
 
BI
Garcia
 
JF
Suela
 
J
, et al. 
Comparative genome profiling across subtypes of low-grade B-cell lymphoma identifies type-specific and common aberrations that target genes with a role in B-cell neoplasia.
Haematologica
2008
, vol. 
93
 
5
(pg. 
670
-
679
)
165
Bachmann
 
IM
Halvorsen
 
OJ
Collett
 
K
, et al. 
EZH2 expression is associated with high proliferation rate and aggressive tumor subgroups in cutaneous melanoma and cancers of the endometrium, prostate, and breast.
J Clin Oncol
2006
, vol. 
24
 
2
(pg. 
268
-
273
)
166
Puppe
 
J
Drost
 
R
Liu
 
X
, et al. 
BRCA1-deficient mammary tumor cells are dependent on EZH2 expression and sensitive to Polycomb repressive complex 2-inhibitor, 3-deazaneplanocin A.
Breast Cancer Res
2009
, vol. 
11
 
4
pg. 
R63
 
167
Raman
 
JD
Mongan
 
NP
Tickoo
 
SK
Boorjian
 
SA
Scherr
 
DS
Gudas
 
LJ
Increased expression of the polycomb group gene, EZH2, in transitional cell carcinoma of the bladder.
Clin Cancer Res
2005
, vol. 
11
 
24 Pt 1
(pg. 
8570
-
8576
)
168
Suva
 
ML
Riggi
 
N
Janiszewska
 
M
, et al. 
EZH2 is essential for glioblastoma cancer stem cell maintenance.
Cancer Res
2009
, vol. 
69
 
24
(pg. 
9211
-
9218
)
169
Burdach
 
S
Plehm
 
S
Unland
 
R
, et al. 
Epigenetic maintenance of stemness and malignancy in peripheral neuroectodermal tumors by EZH2.
Cell Cycle
2009
, vol. 
8
 
13
(pg. 
1991
-
1996
)
170
Kirmizis
 
A
Bartley
 
SM
Farnham
 
PJ
Identification of the polycomb group protein SU(Z)12 as a potential molecular target for human cancer therapy.
Mol Cancer Ther
2003
, vol. 
2
 
1
(pg. 
113
-
121
)
171
Vekony
 
H
Raaphorst
 
FM
Otte
 
AP
, et al. 
High expression of Polycomb group protein EZH2 predicts poor survival in salivary gland adenoid cystic carcinoma.
J Clin Pathol
2008
, vol. 
61
 
6
(pg. 
744
-
749
)
172
van Leenders
 
GJ
Dukers
 
D
Hessels
 
D
, et al. 
Polycomb-group oncogenes EZH2, BMI1, and RING1 are overexpressed in prostate cancer with adverse pathologic and clinical features.
Eur Urol
2007
, vol. 
52
 
2
(pg. 
455
-
463
)
173
Wang
 
CL
Wang
 
CI
Liao
 
PC
, et al. 
Discovery of retinoblastoma-associated binding protein 46 as a novel prognostic marker for distant metastasis in nonsmall cell lung cancer by combined analysis of cancer cell secretome and pleural effusion proteome.
J Proteome Res
2009
, vol. 
8
 
10
(pg. 
4428
-
4440
)
174
Thakur
 
A
Rahman
 
KW
Wu
 
J
, et al. 
Aberrant expression of X-linked genes RbAp46, Rsk4, and Cldn2 in breast cancer.
Mol Cancer Res
2007
, vol. 
5
 
2
(pg. 
171
-
181
)
175
Kong
 
L
Yu
 
XP
Bai
 
XH
, et al. 
RbAp48 is a critical mediator controlling the transforming activity of human papillomavirus type 16 in cervical cancer.
J Biol Chem
2007
, vol. 
282
 
36
(pg. 
26381
-
26391
)
176
Wang
 
S
Robertson
 
GP
Zhu
 
J
A novel human homologue of Drosophila polycomb-like gene is up-regulated in multiple cancers.
Gene
2004
, vol. 
343
 
1
(pg. 
69
-
78
)
Sign in via your Institution