Abstract
Malignant gliomas are central nervous system tumors and remain amongst the most treatment-resistant cancers. Exome sequencing has revealed significant heterogeneity and important insights into the molecular pathogenesis of gliomas. Mutations in chromatin modifiers – proteins that shape the epigenomic landscape through remodeling and regulation of post-translational modifications on chromatin - are very frequent and often define specific glioma subtypes. This suggests epigenomic reprogramming may be a fundamental driver of glioma. Here, we describe the key chromatin regulatory pathways disrupted in gliomas, delineating their physiological function and our current understanding of how their dysregulation may contribute to gliomagenesis.
Keywords: Glioma, epigenomics, chromatin, transcription
Introduction
Chromatin is the physiological state in which genetic material is packaged in the nucleus and consists of genomic DNA and its interacting proteins. The basic repeating subunit of chromatin is comprised of two copies of the four core histone proteins H2A, H2B, H3, and H4, that together form an octamer around which DNA is wrapped to form the nucleosome (Luger et al., 1997). Unlike the DNA sequence, chromatin structure is highly dynamic allowing information encoded within the DNA sequence to be transmitted in different ways in a cell type specific fashion. Cellular signals reach the nucleus to affect gene expression through the activity of DNA binding transcription factors – this engagement occurs in the context of chromatin, which can be either permissive or restrictive thereby providing an additional layer of regulation to tailor gene expression and cellular phenotypes. Phenotypic variations that cannot be explained by alterations in the DNA template are often referred to as being ‘epigenomic’ in nature, although mechanistic details into how epigenomic states are inherited during differentiation and development remain poorly understood.
The chromatin landscape can be modulated through several mechanisms including: 1) post-translational modifications (PTMs) on histones and DNA methylation, 2) remodeling of histone:DNA and histone:histone contacts within nucleosomes and their genomic positions and 3) the incorporation of minor histone ‘variants’ with distinct function into nucleosomes (Allis and Jenuwein, 2016). Modifications on histones and DNA are established by ‘writer’ enzymes (Brownell et al., 1996), removed by ‘eraser’ enzymes (Taunton et al., 1996) and exert function in some cases when recognized by ‘reader’ proteins (Jones et al., 1998) which mediate specific effects on chromatin (i.e. opening or closing of the chromatin fiber). On DNA, methylation of cytosines at position 5 on the pyrimidine ring (5mC) is the more abundant modification and is often associated with gene repression when present at transcription factor binding sites at gene promoters (Razin and Riggs, 1980). In contrast, an impressive array of PTMs have been described on histone proteins (Huang et al., 2014), perhaps the most well characterized being acetyl or methyl modifications on specific lysine residues of the unstructured amino-terminal ‘tail’ of histone H3. PTMs have characteristic genomic locations and can either mediate, or are correlated with, specific transcriptional states. For example, methylation of H3-lysine 27 (H3K27me3) mediates gene repression (Cao et al., 2002; Müller et al., 2002; Pengelly et al., 2013), while acetylation of multiple residues on H3 are associated with gene activation (e.g. H3K9ac, H3K27ac) (Allfrey et al., 1964; Brownell et al., 1996). The second broad class of proteins to shape the chromatin landscape are ‘remodelers’ (Hirschhorn et al., 1992) which alter access for chromatin binding factors by changing nucleosome position or via direct exclusion. Some remodelers can also replace the histones within nucleosomes with variant isoforms which alter chromatin structure locally for specialized functions (Goldberg et al., 2010). For example, canonical H3 can be replaced by the variant H3.3, whose deposition is replication-independent and important for transcriptional regulation in post-mitotic cells in the brain (Maze et al., 2015).
The discovery of mutations in chromatin regulators at a high frequency across human cancer was an unexpected finding from pivotal sequencing efforts (The Cancer Genome Atlas Reseach Network, 2008; Mardis, 2018). This is exemplified in malignant gliomas, where remarkably, genes involved in most of the chromatin regulatory mechanisms described above are either directly mutated or indirectly dysregulated, making glioma an instructive model of epigenomic reprogramming in cancer (The Cancer Genome Atlas Reseach Network, 2008; Verhaak et al., 2010; Wu et al., 2012; Schwartzentruber et al., 2012; Brennan et al., 2013). Gliomas arise in the central nervous system and were historically classified based on their morphological resemblance to specific glial cells (e.g. astrocytes or oligodendrocytes) and the degree of anaplasia they exhibit, otherwise known as ‘grade’. Malignant gliomas can broadly be considered those tumors ≥ grade III (and ≥ grade II in adults) due to their locally invasive behavior, meaning surgery cannot safely remove the entire tumor from the brain. Even with additional treatments such as radiation and/or chemotherapy they remain inevitably fatal. The histopathological designation of gliomas has been augmented, and in many respects superseded, by molecular characteristics which identify biological subtypes - often defined by chromatin regulators - with specific distributions across age and anatomical location in the brain (Fig. 1).
Figure 1. Simplified molecular subgrouping of malignant gliomas defined by common somatic mutations.
Subtypes broadly defined by mutational status of H3 (in children), and IDH (in adults) with common accompanying molecular alterations for each subtype shown. Alterations in genes which regulate chromatin function shown in red. Subtypes were initially identified in either adults or children, though are distributed across age-groups at different frequencies e.g. H3 mutations were defined in children but also occurs in young adults. The designation ‘H3K27M-midline glioma’ is replacing ‘diffuse intrinsic pontine glioma or DIPG’ to reflect a variety of midline anatomical locations (e.g. pons, spinal cord) in which these mutations occur. IDH mutant subtype further defined by either ATRX loss (‘astrocytomas’) or 1p/19q co-deletion (‘oligodendrogliomas’). IDH wild-type group in adults and H3 wild-type group in children are highly heterogeneous and can be further subdivided based on DNA methylation analysis (reviewed in Sturm et al., 2014). Grade denotes typical grade at presentation, though tumors can present clinically at different points in their anaplastic evolution.
Here, we review the main chromatin pathways which are dysregulated in malignant glioma using their function in normal physiology as a primer to illustrate how they may promote cancer. We speculate as to why aberrant chromatin drives cancer in specific contexts and highlight some of the fundamental questions, which need addressing to translate mechanistic insights into therapies for what remains largely incurable disease.
Misregulation of H3K27 methylation: H3K27M mutation and EZHIP
H3 lysine 27 methylation is a PTM which functions as an important effector mechanism in gene repression mediated by a group of evolutionarily conserved multi-protein complexes collectively called the Polycomb repressive complexes (Schuettengruber et al., 2017). These complexes were discovered in Drosophila as critical regulators of cell fate during development, as evidenced by abnormal body patterning induced by mutations in Polycomb proteins (Lewis, 1978). Two Polycomb repressive complexes (PRC1 and PRC2) act cooperatively to repress genes. PRC2 is responsible for establishing H3K27 methylation, predominantly through EZH2 (Enhancer of zeste homologue 2), and to a lesser extent, EZH1 enzymatic (“writer”) subunits. H3K27 can be mono-, di-, or tri-methylated (H3K27me1, H3K27me2, H3K27me3, respectively) with H3K27me3 the best characterized PTM. The functional importance of H3K27 as a substrate for methylation was inferred by genetic experiments in which H3 lysine-27 was mutated to arginine, which cannot undergo methylation at that site – leading to aberrant cell fate decisions, phenocopying polycomb mutants (Pengelly et al., 2013). H3K27me3 interacts directly with a second complex, PRC1, through its ‘reader’ component, a chromobox (CBX) protein. CBX proteins also have nucleosome compacting (Francis et al., 2004; Lau et al., 2017) and phase-separating activity (Plys et al., 2019; Tatavosian et al., 2018), which may be essential for PRC1-mediated repression, potentially by restricting access of transcription factors to genes (Fig. 2a). Polycomb repression is both heritable through cell division (‘epigenetic memory’) (Angel et al., 2011; Berry et al., 2015), while capable of being dynamically reversed when necessary, making it a versatile repressive mechanism, particularly in development.
Figure 2. Reprogramming of the H3K27 methylation landscape by H3K27M mutation and EZHIP.
(A) PRC2 establishes H3K27me3 through its enzymatic component EZH2. HK27me3 is recognized by a CBX ‘reader’ protein in the PRC1 complex, which compacts chromatin to repress genes. (B) PRC2 is unable to methylate methionine residue in H3K27M, but also has reduced methylation activity in ‘trans’ i.e. at nucleosomes which contain wildtype H3. Amino acid sequence adjacent to K-to-M mutation aligned with EZHIP sequence showing similarity: an example of ‘oncohistone mimicry’. (C) Schematic of genome-wide H3K27me3 landscape contrasting WT, H3K27M and PRC2 knockout cells. In H3K27M cells small areas of H3K27me3 are retained at PRC2 nucleation sites, in contrast to PRC2 knockout cells where H3K27me3 is totally ablated.
There has been a surge in interest in the function of H3K27 methylation in glioma following landmark studies identifying mutations in histone H3 at a high-frequency in pediatric glioma, implicating mutant histones as potential oncogenes in cancer for the first time (Schwartzentruber et al., 2012; Wu et al., 2012). One class of the mutations described occurred recurrently at H3K27, leading to a lysine to methionine substitution (H3K27M), the exact site at which PTMs such as H3K27me3 is established. Interestingly, these H3K27M tumors occur in midline brain structures (Aihara et al., 2013; Schwartzentruber et al., 2012; Sturm et al., 2012; Wu et al., 2012), possibly reflecting a propensity for this mutation to transform a cell-of-origin in a specific anatomic location. The strict occurrence of these mutations at lysine 27, in addition to their heterozygous nature – especially given there are a total of 16 genes (and 32 alleles) from which H3 protein can be produced - suggests that the H3K27M mutation has dominant or neomorphic function. Indeed, biochemical data is consistent with this - in vitro, H3K27M inhibits the enzyme EZH2 and furthermore, H3K27M glioma tissues or cells in which H3K27M has been experimentally introduced, exhibit dramatically reduced H3K27me3 (Bender et al., 2013; Chan et al., 2013; Lewis et al., 2013). The H3K27M mutation in histone H3, as well as other recently-identified histone mutations in H3 as well as in H2A, H2B and H4 have been collectively described as ‘oncohistones’ (Nacev et al., 2019). The function of these mutations in tumorigenesis, if any, is not well understood and is under active investigation.
The molecular details of how loss of H3K27 methylation occurs in H3K27M cells are still being unraveled (Fig. 2b). The increased affinity of H3K27M for EZH2 in vitro (Lewis et al., 2013) and the crystal structure showing tight H3K27M/EZH2 binding (Justin et al., 2016) supports a model in which the mutant histone sequesters EZH2 (and PRC2) on chromatin (Lewis et al., 2013). In contrast, studies mapping occupancy of H3K27M and EZH2 on chromatin show localization at predominantly distinct sites, arguing against sequestration (Fang et al., 2018a; Mohammad et al., 2017; Piunti et al., 2017; Stafford et al., 2018). There is evidence of a persistent effect on PRC2 function following disengagement from H3K27M: PRC2 complex purified from H3K27M expressing cells exhibits reduced methyltransferase activity in vitro (Stafford et al., 2018). How PRC2 may be ‘poisoned’ following its engagement with H3K27M remains to be seen, though one possibility would be through a conformational allosteric change which diminishes enzyme activity. EZH2 was recently shown to auto-methylate at specific lysine residues, and this activity is also inhibited by H3K27M (Lee et al., 2019; Wang et al., 2019). It will be important to understand the structural/conformational effect of this auto-methylation on EZH2 and the entire PRC2 complex.
It is important to note the inhibition of PRC2 is not total, a minority of loci in the genome retain H3K27me3 in H3K27M cells (Chan et al., 2013; Mohammad et al., 2017). PRC2-mediated methylation is nucleated on chromatin at particular sites, typically large CpG islands (regions of the genome that contain a large number of CG dinucleotide repeats) and then spreads outward establishing broad domains of H3K27me3 (Oksuz et al., 2018). In H3K27M cells, PRC2 is still recruited to its nucleation sites (which are typically distinct from H3K27M incorporation sites in the genome), but broad H3K27me3 domains do not form suggesting the spreading mechanism is defective (Harutyunyan et al., 2019; Stafford et al., 2018). The generalized loss of H3K27me3 with selective retention at PRC2 nucleation sites is a distinct epigenomic landscape not observed in cancers with loss-of-function mutations in PRC2 components (Fig. 2c) (Wojcik et al., 2019). This may explain why other PRC2 mutations are not found in midline gliomas – the retained H3K27me3 sites may silence tumor suppressor genes, such as CDKN2A, which are essential for cell survival (Mohammad et al., 2017). This also explains why inhibiting this residual PRC2 function may be a promising strategy in H3K27M midline gliomas through de-repression of tumor-suppressor genes (Mohammad et al., 2017; Piunti et al., 2017).
How might loss of H3K27me3 drive tumorigenesis in neural cells? Given the role of H3K27me3 and the Polycomb system in cell fate regulation, the effect of H3K27M on differentiation has been explored (Funato et al., 2014; Silveira et al., 2019). To faithfully reflect molecular alterations that co-occur in midline gliomas, human neural stem cells were engineered to express H3K27M, a TP53 shRNA, and constitutively active PDGFRA (Funato et al., 2014). This led to an impairment in differentiation and acquisition of a more primitive gene expression program and was sufficient to drive gliomagenesis when the transformed cells were injected in vivo (Funato et al., 2014). Though there is broad reduction of H3K27me3, a restricted number of genes are inappropriately activated following loss of H3K27me3, with an enrichment of ‘bivalent’ genes (bearing both H3K27me3 and H3K4me3), which ordinarily are activated only in a highly-orchestrated fashion during differentiation (Funato et al., 2014; Larson et al., 2019). Depletion of H3K27M in midline glioma cells derived from patients also promotes expression of neural differentiation genes (Harutyunyan et al., 2019; Silveira et al., 2019). Thus, disruption of normal neural differentiation is the prevailing hypothesis of how H3K27M contributes to oncogenesis. It is notable that many of the differentiation phenotypes described above required additional molecular perturbations alongside H3K27M - it remains poorly understood whether and how H3K27M alone may confer an oncogenic advantage.
Interestingly, a striking loss of H3K27me3 is also a characteristic feature of posterior fossa ependymomas (PFAs) (Pajtler et al., 2015), a rare glioma variant which develops in the posterior part of the brain and histologically resemble ependymocytes, the cells lining the fluid-filled brain ventricles (Wu et al., 2016). However, only 4% of PFAs express H3K27M, and recent studies have implicated mutation or overexpression of a different gene, CXorf67, as a driver of this altered H3K27 methylation landscape in PFAs (Pajtler et al., 2018). Notably, the protein sequence of CXorf67 resembles the region surrounding lysine-to-methionine (K-to-M) sequence within the H3 tail in H3K27M mutants (Fig. 2b), and remarkably, expression of protein derived from this gene, or a synthetic peptide incorporating the K-to-M sequence, inhibits EZH2 via a similar mechanism (Hübner et al., 2019; Jain et al., 2019; Ragazzini et al., 2019; Piunti et al., 2019). Hence, this gene has now been re-named EZH2 inhibitory protein (EZHIP) as an example of ‘oncohistone mimicry’. These data lend further support to the idea that hindbrain derived cells may have enhanced sensitivity to transformation by proteins which antagonize EZH2 function during development.
H3K27me3 domains can block the placement of other PTMs and the loss of H3K27me3 in H3K27M tumors is also accompanied by downstream epigenomic changes. For example, several studies have shown expression of H3K27M leads to an increase in H3K27 acetylation on wild-type H3 (Lewis et al., 2013; Piunti et al., 2017). This PTM is frequently present at active regulatory elements (enhancers and promoters), and it has been proposed that H3K27M may induce enhancer reorganization (Piunti et al., 2017) though analysis in isogenic models suggests the enhancer landscape is determined largely by the cell of origin rather than the presence of the H3K27M mutation (Krug et al., 2019). Increased H3K27ac was also observed in intergenic regions, leading to higher basal endogenous retroviral element (ERV) expression (Krug et al., 2019). These acetylated ERVs may prime H3K27M gliomas for epigenetic therapies aiming to augment anti-tumor immunity by increasing ERV expression to activate the innate interferon response (Krug et al., 2019). H3K36me2 has also been shown to be increased in the presence of H3K27M) (Stafford et al., 2018), forming new domains which invade into regions where H3K27me3 is lost - further study is required to elucidate whether these epigenomic changes are causal or consequential to oncogenesis induced by H3K27M.
Misregulation of H3K36 methylation: H3G34R/V and SETD2 mutations
Misregulation of histone lysine 36 methylation (H3K36) has been implicated as an important event in gliomagenesis by two key observations. Firstly, there is a subtype of malignant gliomas in children and young adults characterized by recurrent mutations in H3 at a residue in close proximity to H3K36, leading to substitution of glycine to either arginine or valine (H3G34R/V) (Schwartzentruber et al., 2012; Wu et al., 2012) (Fig. 1). Biochemically, substituting G34 with bulky sidechain such as arginine or valine causes steric hindrance and restricts access to H3K36 by the ‘writer’ enzyme SETD2 (SET Domain Containing 2, Histone Lysine Methyltransferase), leading to loss of H3K36 methylation (Yang et al., 2016) (Fig. 3b). Secondly, loss-of-function mutations in SETD2 have been identified at a high frequency in malignant gliomas (Brennan et al., 2013; Fontebasso et al., 2013). Though they are not subtype defining, they represent the second most frequent mutation in a chromatin modifier in adults (Brennan et al., 2013; Zehir et al., 2017). H3G34 mutations differ from SETD2 mutations in that the loss of H3K36me3 occurs in cis (i.e. only at nucleosomes where this mutant histone is present), as opposed to more globally in SETD2 mutant cells (Fig. 3c).
Figure 3. Reprogramming of the H3K36me3 landscape by H3G34 and SETD2 mutations.
(A) H3K36me3 is coupled with transcription elongation and resides at gene bodies to recruit chromatin effectors to suppress cryptic transcripts (DNMT3B), and regulate splicing (MRG15), and DNA repair (MSH6, PHF1) (see text for details). (B) H3G34R/V sterically restricts access of SETD2 to H3K36 leading to loss of H3K36me3 in cis i.e. in nucleosomes where the mutant histone is incorporated, while SETD2 remains functional in trans on nucleosomes containing wild-type H3. (C) Loss-of-function mutations of SETD2 lead to global reduction in H3K36me3.
Insights into how reduced H3K36me3 levels promote glioma may be deduced from understanding its function in normal cells (Fig. 3a). Physiologically, H3K36me3 occurs on chromatin at the gene bodies of actively expressed loci and its establishment is coupled with transcription elongation (Krogan et al., 2003; Li et al., 2002). Functional studies suggest H3K36me3 plays an important role in regulating the ‘quality’ of transcription by repressing generation of aberrant or unwanted transcripts (McDaniel and Strahl, 2017). For example, loss of H3K36 methylation through mutation of SETD2 (or its homologues) (Kaplan et al., 2003; Neil et al., 2009; Xu et al., 2009) leads to the emergence of cryptic transcripts – RNA species whose production is initiated within the gene body or at other sites outside the canonical promoter region at the 5’ end of a gene. H3K36me3 may prevent cryptic transcript production by promoting DNA methylation in the gene body through direct recruitment of the methyltransferase, DNMT3B (Neri et al., 2017). If cryptic transcripts occur in the antisense direction they may suppress production of the physiological transcript (and thus the encoded protein) derived from the 5’ promoter in the ‘sense’ direction (Huber et al., 2016; Kim et al., 2016). It has been shown that some cryptic transcripts are translated (Cheung et al., 2008); whether cryptic proteins provide a selective advantage (or neomorphic function) for cancer cells remains to be determined.
H3K36me3 also plays an important role in recruiting RNA splicing factors thereby regulating exon choice (Luco et al., 2010) and SETD2 mutant cells show genome-wide disruption of splicing as evidenced by the presence of alternative and intron-retaining transcripts (Simon et al., 2014). Spliceosome disruption is an established driver of cancer (Dvinge et al., 2016), broadly by either generating oncogenic splice variants, or by abrogating expression of tumor-suppressive isoforms (Simon et al., 2014). Thus, splicing defects secondary to loss of H3K36me3 have plausible potential to promote gliomagenesis, though the specific mis-splicing events and/or protein isoforms which are oncogenic in this context remain to be characterized.
In addition to its described role in regulating transcription, H3K36me3 plays an important role in DNA damage repair, which has a well-established role in tumor suppression (Jeggo et al., 2016). For example, deletion of SETD2 is sufficient to disrupt mismatch repair (MMR) and increase mutation frequency, as MSH6 (MutS homolog 6), a component of the MutSαmismatch recognition complex, interacts with the H3K36me3 mark through its PWWP ‘reader’ domain (Li et al., 2013). This phenotype has also been demonstrated in cells bearing H3G34R mutations (Fang et al., 2018b). Consistent with this observation, unbiased analyses of >1000 cancer genomes demonstrated mutation frequency is reduced in H3K36me3-enriched regions of the genome (Clendening et al., 2010). H3K36me3 also plays a role in the early stages of the DNA damage response, by recruiting the protein PHF1 to sites of DNA double strand breaks through its H3K36me3 interacting Tudor domain (Musselman et al., 2012), where it may repair damage through non-homologous end joining (Musselman et al., 2012).
While the loss of H3K36me3 and resultant failure to recruit H3K36me3 ‘reader’ proteins may be important for oncogenesis, other epigenomic changes that are gained may be worthy of further attention. For example, reduction in H3K36me3 can facilitate H3K27me3 placement, as EZH2 ‘senses’ methylation status at H3K36 and is active when the residue is unmethylated (Jani et al., 2019). It will thus be interesting to define the function of any increases in H3K27me3 in H3G34R/V induced phenotypes. Additionally, a recent study profiling the chromatin state in embryonic stem cells genome-edited to express H3G34R demonstrated relative increases in both H3K36me3 and H3K9me3 in ‘trans’ (Voon et al., 2018). This was attributed to an ability of H3G34R to inhibit the function of lysine demethylase 4B (KDM4B), a dual H3K9/H3K36 demethylase, in a dominant fashion (Voon et al., 2018). These findings are intriguing because this demethylase is a target of the neometabolite, 2-HG produced by mutant IDH (see below). Interestingly, H3G34 and IDH mutant gliomas have several shared features, such as their age distribution, location in the brain, and the co-occurrence of ATRX loss and p53 mutations (Fig. 1) (Sturm et al., 2012). One hypothesis is that overlapping effects of IDH and H3G34 mutants on the chromatin landscape promote gliomagenesis and influence the co-mutations that are acquired in the process.
It remains unexplained why H3G34 mutations only occur in the gene encoding the H3 variant, H3.3 (H3F3A gene), in contrast to the H3K27M mutation which can occur in genes encoding either variant H3.3 or canonical H3.1/3.2 (HIST1H3B and HIST2H3C genes) (Fig. 1). H3.3 only differs from canonical H3 by a few amino acids yet this results in its recognition and deposition into chromatin by distinct chaperones (see below), with a characteristic distribution (Elsässer et al., 2012; Goldberg et al., 2010; Lewis et al., 2010) and in a replication-independent fashion (Ahmad and Henikoff, 2002). How these and other functional differences (Armache et al., 2019; Martire et al., 2019) from canonical H3 interact with mutation at H3G34 are important areas for future investigation.
Disruption of chromatin through oncometabolites: IDH mutations
Mutations in isocitrate dehydrogenase (IDH) have been identified as the earliest somatic event (Johnson et al., 2014; Parsons et al., 2008; Suzuki et al., 2015; Watanabe et al., 2009) in a subtype of gliomas which most frequently present with: 1) lower histological grade (often grade II) (The Cancer Genome Atlas Reseach Network, 2015), 2) arise in the frontal lobes of the brain (Lai et al., 2011), 3) a peak age of presentation between 20 and 35 (Watanabe et al., 2009) (Fig. 1). IDHs are important metabolic enzymes which convert the TCA metabolite isocitrate into alpha-ketoglutarate (alpha-KG) (Pavlova and Thompson, 2016). This can occur in the cytosol (IDH1) and mitochondria (IDH2), generating NADPH for metabolic biosynthesis and redox homeostasis (Pavlova and Thompson, 2016). IDH1 is most frequently mutated in gliomas leading to a heterozygous arginine 132 to histidine mutation (R132H) (Parsons et al., 2008; Verhaak et al., 2010). Less commonly, arginine 172 in IDH2 is heterozygously mutated to histidine in gliomas. These gain-of-function mutations confer neomorphic function upon IDH which then favors alpha-KG as a substrate, converting it into 2-hydroxyglutarate (2-HG) (Dang et al., 2009; Ward et al., 2010) (Fig. 4a). 2-HG may act as a competitive inhibitor of over 60 different enzymes which use alpha-KG as a cofactor (Rose et al., 2011). These alpha-KG dependent enzymes span a variety of functions including metabolism, chromatin regulation and DNA repair (Rose et al., 2011), thus mutant IDH has broad potential to directly perturb normal cell physiology.
Figure 4. Pleiotropic effects of IDH mutation on the epigenome and other alpha-ketoglutarate dependent pathways.
(A) Gain-of-function mutations in IDH leads enzyme to favor alpha-ketoglutarate as a substrate leading to production of oncometabolite 2-HG. 2-HG inhibits many alpha-ketoglutarate dependent enzymes including histone (KDMs) and DNA demethylases (TETs), leading to aberrant accumulation of methylation on chromatin. Some of these changes such as H3K9 methylation prevent appropriate gene activation required for normal differentiation (B) Mutant IDH induced DNA methylation at CTCF binding sites disrupts chromatin architecture allowing oncogenes such as PDGFRA to hijack enhancers in different TADs.
Which are the key enzymes 2-HG inhibits to drive tumorigenesis? Studies indicate alpha-KG dependent demethylases (of histones and DNA) may be critical targets of 2-HG in driving tumor development by reshaping the epigenomic landscape (Lu et al., 2012; Schvartzman et al., 2019; Turcan et al., 2012) (Fig. 4a). For example, 2-HG inhibits demethylation of several histone marks – H3K9, H3K27, H3K36, H3K79 - leading to their accumulation on chromatin (Lu et al., 2012). The composite effects of these changes are poorly understood – though H3K9 accumulation is the earliest change induced by mutant IDH and may be mechanistically important (Lu et al., 2012). In normal cells, H3K9 methylation is enriched at heterochromatin where it recruits HP1 ‘reader’ proteins (Bannister et al., 2001; James and Elgin, 1986; Lachner et al., 2001; Nakayama et al., 2001), which compact chromatin and act as scaffolds for other chromatin modifiers (H3K9 methylases and histone deacetylases) to establish a transcription refractory state. During normal differentiation, H3K9 methylation is dynamically reorganized (Hawkins et al., 2010; Nicetto et al., 2019) such that lineage-specific genes are activated where loss of H3K9 methylation occurs, while stem cell genes and genes pertaining to other lineages gain H3K9 methylation (Hawkins et al., 2010; Nicetto et al., 2019). In the context of mutant IDH, this process becomes disorganized and there is a block in cellular differentiation as lineage specific transcription factor access is restricted by aberrant retention of H3K9 methylation (Lu et al., 2012; Modrek et al., 2017; Schvartzman et al., 2019). Interestingly, 2-HG driven increases in H3K9 methylation was recently demonstrated to impair recruitment of DNA damage repair factors to sites of double-stranded DNA breaks, resulting in potential vulnerability to therapies targeting DNA damage pathways (Sulkowski et al., 2017, 2020)
Through inhibition of TET (ten-eleven translocation) family of DNA demethylases, 2-HG is sufficient to induce accumulation of DNA methylation at CpG islands leading to a characteristic epigenomic phenotype called G-CIMP (CpG Island methylator phenotype) (Turcan et al., 2012). DNA methylation at promoter CpG islands has long been associated with gene-repression (Comb and Goodman, 1990; Prendergast and Ziff, 1991); the mechanism underlying this is incompletely understood but is in part by interfering with transcription factor binding (Stadler et al., 2011). By increasing promoter DNA methylation, mutant IDH can broadly silence multiple putative tumor suppressor genes. Notably, in acute myeloid leukemia IDH and TET mutations occur in a mutually exclusive manner (Figueroa et al., 2010) - one interpretation of this is that these two distinct mutations are alternative molecular routes to produce a similar downstream effect, i.e. dysregulation of DNA methylation. TET mutations are observed in glioma though they are extremely rare and not mutually exclusive with IDH1 mutations (Zehir et al., 2017) hence there are likely distinct, lineage specific differences in the impact of these mutations on DNA methylation and oncogenesis.
Recent work has highlighted the presence of CpG island DNA methylation at non-promoter sites in IDH mutant cells (Bledea et al., 2019; Flavahan et al., 2016; Modrek et al., 2017), which may have distinct effects on chromatin structure and gene expression (Fig. 4b). The transcription factor CTCF (CCCTC-binding factor) binds to ‘insulator sites’ regulating the structure of the genome into domains called TADs (topologically associated domains) which insulate genes in one domain from activation by enhancers in different domains (Dixon et al., 2012; Nakahashi et al., 2013). DNA methylation can in certain circumstances prevent CTCF binding and indeed loss of these insulation sites in an IDH mutant glioma model increased expression of the PDGFRA oncogene, potentially by abrogating the TAD boundary and permitting the PDGFRA gene access to an active enhancer in a distinct domain (Flavahan et al., 2016). Another study highlighted extensive methylation of putative enhancer CpG islands in IDH mutant tumors (Bledea et al., 2019), the transcriptional impact of which will require further characterization (Bledea et al., 2019). In summary, the data showing mutant IDH disrupts DNA methylation genome-wide is robust, though further work is needed to decipher the transcriptional impact of these changes at different genomic locations are (promoter vs gene-body vs enhancer), and determine which changes are oncogenic or simply incidental.
Chromatin remodeler dysfunction: ATRX mutation
Loss of function mutations in ATRX (Alpha-Thalassemia/Mental Retardation Syndrome, X-Linked) frequently co-occur with IDH mutations and define a specific IDH mutant subgroup which have an astrocytic appearance histopathologically (IDH mutant astrocytoma, Fig. 1) (Jiao et al., 2012; Liu et al., 2012). In addition, similar ATRX mutations occur in pediatric gliomas where they are uniformly present in the H3G34-mutant gliomas (Fig. 1), frequently present in the rare entity anaplastic astrocytoma with piloid features (Reinhardt et al., 2018) and occasionally in H3K27M-mutant gliomas (Khuong-Quang et al., 2012; Schwartzentruber et al., 2012; Wu et al., 2012). Genetic mouse modeling has shown ATRX loss accelerates glioma formation (Koschmann et al., 2016), supporting a role for this event as a driver, however ATRX has complex function physiologically and how its loss promotes tumorigenesis is poorly understood.
ATRX protein contains two highly conserved domains, which are sites of increased mutation frequency in glioma: the ADD (ATRX-DNMT3-DNMT3L) domain and the ATP-dependent helicase domain. The ADD region of ATRX is a dual ‘reader’ domain which recognizes the combination of H3K9me3 and unmethylated H3K4 (Eustermann et al., 2011; Iwase et al., 2011), found in certain areas of constitutive heterochromatin to which ATRX is recruited. H3K4 methylation is largely correlated with active chromatin (promoter and enhancer elements) and may block ATRX recruitment through the ADD domain. This illustrates an example of an elaborate ‘code’, where PTM status at different positions on the H3 amino-tail are ‘sensed’ to regulate effector:chromatin interactions (Strahl and Allis, 2000). What is the function ATRX serves in heterochromatic regions (Fig. 5)? Several studies show that ATRX forms a chaperone complex with another protein called DAXX (Death-Domain associated protein) to deposit the histone variant H3.3 to telomeres and pericentromeric chromatin (Drané et al., 2010; Goldberg et al., 2010; Lewis et al., 2010). How deposition of H3.3 regulates heterochromatin function is still being elucidated: studies suggest it is essential for maintaining genomic integrity and normal replication in heterochromatin, and its loss can induce chromosome instability (Clynes et al., 2014; De La Fuente et al., 2015; Huh et al., 2012).
Figure 5: Consequences of ATRX loss in different chromatin regions.
(A) ATRX acts in concert with DAXX as a chaperone to deposit the histone variant H3.3 into heterochromatic regions bearing H3K9me3 at telomeres and peri-centromeres, endogenous retroviral elements and some euchromatic sites. Loss of ATRX and H3.3. deposition at telomeres leads to ALT and supports cell immortalization. Consequences of ATRX loss in euchromatin are poorly defined but may affect gene expression.
Loss off ATRX function at telomeric heterochromatin, promotes a phenomenon called alternative lengthening of telomeres (ALT) (Fig. 5). This is a recombination-mediated replication mechanism where telomeres become lengthened via use of telomeric DNA as a copy template for replication (Cesare and Reddel, 2010). Induction of ALT has been proposed as the major evolutionary advantage conferred to cancer cells by ATRX loss (Clynes et al., 2015; Heaphy et al., 2011), as it enables maintenance of telomere length which is important for cellular immortality by preventing senescence associated with telomere shortening (Cesare and Reddel, 2010). Repetitive DNA at telomeres contain a high content of guanine which is prone to form secondary DNA structures, called G4 quadruplexes which can impede both transcription and DNA replication (Varshney et al., 2020). ATRX co-localizes with regions predicted to adopt these structures and is thought to either suppress or reverse their formation through deposition of H3.3 – in the absence of ATRX, homologous recombination of telomeres is promoted. One suggestion is that repeat DNA at heterochromatin can destabilize nucleosomes, and H3.3 deposition can help maintain nucleosome density optimal for heterochromatin formation (Lewis et al., 2010).
Though there is a compelling correlation between loss of ATRX function and ALT, it is notable that a proportion of cells in which ATRX is not expressed, do not develop ALT (Clynes et al., 2014; Fishbein et al., 2015; Brosnan-Cashman et al., 2018; Danussi et al., 2018). This suggests a specific cellular context and/or additional changes are needed to induce ALT. Moreover, it is likely ATRX loss confers selective advantages to cancer cells through ALT-independent mechanisms. A recent study showed that deletion of ATRX in neural stem cells induced transcriptional changes associated with astrocytic differentiation and increase migratory capacity (Danussi et al., 2018), both characteristic features of the of the IDH mutant astrocytomas in which ATRX loss invariably co-occurs (Danussi et al., 2018). Notably, ATRX was preferentially enriched at gene promoters, enhancers and at gene bodies in neural stem cells, raising the possibility of a direct role in active chromatin regulation (Danussi et al., 2018). Further work will be necessary to explore this hypothesis and the molecular function of ATRX, DAXX and H3.3 in these regions.
Overrepresentation of chromatin mutations in pediatric glioma
It is noteworthy that chromatin regulators are most frequently mutated in the glioma subtypes which occur in children and young adults (Fig. 1) this reflects a pattern which is observed across cancer types (Schwartzentruber et al., 2012; Sturm et al., 2012; Verhaak et al., 2010; Wu et al., 2012). The biological basis of this is unclear though a number of non-mutually exclusive hypotheses can emerge (Fig. 6a). Chromatin regulators have highly cell-type specific effects, likely due to: 1) cell-type specific patterns of genomic occupancy for chromatin factors and 2) cell-type specific activity of genes regulated by chromatin. Hence, the high-frequency of chromatin regulator mutations in younger individuals may reflect a propensity to transform a particular cell-of-origin, which may only accessible or abundant during an age-restricted window, or in a microenvironment unique during a specific developmental stage. Indeed, these mechanisms may contribute to the distinct age distributions of H3 and IDH mutations even amongst young individuals (Fig. 1).
Figure 6: Putative hallmarks of chromatin factor mutations which drive gliomagenesis in young patients.
(A) Context dependent effects of chromatin may be oncogenic in a developmentally restricted cell of origin. (B) Developmentally active period may be more sensitive to perturbation in chromatin factors given their function in establishing and maintaining cell identity. (C) ‘Single-hit’ may disrupt multiple cancer hallmarks as chromatin factors reside at hundreds of different genes regulating distinct pathways.
Another related factor may be the initial mechanism of transformation mediated by mutated chromatin factors, which may be distinct from mutations in non-chromatin genes. Mutations in chromatin regulators often co-occur alongside other classes of mutation in more traditional oncogenic pathways such as growth factor signaling, genomic stability, and cell cycle control (Schwartzentruber et al., 2012; Sturm et al., 2012; Verhaak et al., 2010; Wu et al., 2012). One interpretation of these data is that mutations of chromatin regulators occupy a distinct, and specific ‘niche’ in oncogenesis. But what might this niche be? Perturbation of normal differentiation appears to be a common feature and a potential candidate as a more general phenotype induced by chromatin mutations in glioma and other cancers (Funato et al., 2014; Lu et al., 2012). Evidence from single cell transcriptomics has nominated specific glial precursors in the developing brain as the cells from which different pediatric gliomas arise (Filbin et al., 2018; Jessa et al., 2019). We propose that dysregulation of the chromatin landscape may be particularly impactful during early development, when a larger number of cells are undergoing both self-renewal and fate transitions, which require highly coordinated chromatin remodeling. Moreover, if it is correct to consider epigenetic regulation broadly as a ‘lock down’ mechanism after traditional transcription factor-mediated gene regulatory programs are initiated, this window of ‘plasticity’ may be more prone to error earlier in development when epigenetic machinery is being established de novo (Fig. 6b).
Finally, cancers such as glioma in younger individuals in general bear a lower total number of mutations than gliomas in the elderly (Gröbner et al., 2018). We postulate that some chromatin regulators, through their ability to regulate hundreds of genomic loci, have the potential to simultaneously dysregulate multiple cancer hallmarks with a single hit, meaning fewer additional mutations are required to enter a fully transformed state (Fig. 6c).
Prospects for therapeutic translation
Not a single targeted therapy has shown efficacy in malignant glioma in a randomized control trial, highlighting the difficulty in identifying tractable therapeutic targets, and developing brain penetrant small molecules to engage those targets in the CNS (Fine, 2015). The recognition that aberrant chromatin is central to glioma development has generated much interest in mechanism-based therapies targeting chromatin regulators. The most direct strategy is to directly inhibit gain-of function proteins e.g. mutant IDH and H3. Mutant IDH is particularly amenable to small molecule inhibition given its defined enzymatic domain, which can be engaged with a small molecule. However, while mutant IDH may be an early driving event, more established tumors may not rely on the effects of the mutation to drive tumor growth once additional molecular events have been acquired. This is evidenced by variable responses to IDH inhibition in pre-clinical glioma models (Johannessen et al., 2016; Rohle et al., 2013; Tateishi et al., 2015) and other studies which suggest deletion of mutant IDH may be selected for as tumors evolve in patients or in vitro cell culture (Luchman et al., 2013; Mazor et al., 2017). In the context of leukemia however, IDH inhibition has shown success in human clinical trials (DiNardo et al., 2018) and the ultimate test of IDH as a therapeutic target in glioma will be determined through similar human clinical studies that are now underway.
For other gain-of-function mutants such H3K27M and H3G34R which are not enzymes, more creative therapeutic approaches may be necessary such as inhibition of protein-protein interactions or targeted degradation (Bondeson and Crews, 2017). For H3K27M, such approaches are supported by data suggesting the mutant histone is required by tumors to maintain the oncogenic state (Harutyunyan et al., 2019; Silveira et al., 2019).
Targeting chromatin through non-oncogene dependencies is another promising anti-cancer strategy which may have utility in glioma. HDAC inhibitors were the first class of approved ‘epigenetic’ therapy (Marks, 2010), followed by DNA methylation inhibitors (Kaminskas et al., 2005) and most recently by EZH2 inhibitors in sarcoma (Rothbart and Baylin, 2020; Stacchiotti et al., 2019). EZH2 inhibition in this context built on basic research delineating a synthetic lethal relationship between SWI/SNF complex (mutated in ~30% of cancers) and PRC2 (Wilson and Roberts, 2011). The loss of SWI/SNF complex on chromatin permits recruitment of PRC2 in ‘cis’ and placement of H3K27me3 (Kadoch et al., 2017; Wilson et al., 2010). This compensatory change is essential for cell survival and blocking PRC2 activity exhibits anti-tumor activity in this context (Knutson et al., 2013). This is an example of synthetic lethality in chromatin factors that oppose each other to maintain transcriptional control, and other similar therapeutic opportunities will likely arise as our understanding of chromatin cross-talk in glioma deepens, given the dynamic and highly interconnected manner in which chromatin factors regulate the epigenome. Already it has been shown that the inhibitory effect of H3K27M on EZH2 in vitro can be partially reversed by the presence of histone acetylation on nearby lysines (Brown et al., 2014). Building on this finding, histone deacetylase inhibitors, which increase steady state histone acetylation, are in clinical trials for H3K27M glioma, based on encouraging activity in pre-clinical studies (Grasso et al., 2015).
Finally, a growing number of studies show that chromatin regulation is an important mechanism of immune-evasion in cancer (Pan et al., 2018), with clinical trials combining chromatin targeting agents with checkpoint blockade now underway (Daver et al., 2019; Gandhi et al., 2018). Such mechanisms are relatively unexplored in malignant glioma and will be important areas for future investigation.
Conclusion
Epigenomic reprogramming is now recognized as a key feature in the molecular etiology of malignant glioma. Important initial progress has been made to define the proximal biochemical effects of mutated chromatin regulators in glioma though key questions remain. A fundamental challenge in studying chromatin regulators lies in their inherent ability to simultaneously impact hundreds of different genomic loci and hence biological pathways, even in some cases acting on non-chromatin substrates. Deciphering which events are essential for tumorigenesis and which are passenger events will require use of systems-based technologies which can induce simultaneous perturbations in a controlled fashion, and an ability to causally link these changes to oncogenic phenotypes. Such techniques are emerging and have the potential to reveal important insights when employed in the context of genetically-engineered glioma models which have been developed (Funato et al., 2014; Larson et al., 2019; Mohammad et al., 2017; Pathania et al., 2017). These models will also permit detailed interrogation of the functional interplay between changes in the epigenomic landscape caused by chromatin mutations, cell-of-origin, and aberrations in signaling pathways all of which co-occur in a characteristic manner in different glioma subtypes (e.g. ACVR1 with H3.1K27M, and PDGFRA/FGFR1 mutations with H3.3K27M, Fig. 1). Finally, a wealth of evidence has shown the chromatin landscape is modulated by non-genetic stimuli including ageing, inflammation and metabolic state within the local microenvironment all of which can signal to chromatin. The role of these factors in glioma initiation, maintenance and progression are relatively understudied and will be an important area for exploration in the future. Exciting progress has been made unraveling fundamental mechanisms that utilize a plethora of epigenetic regulators drawing upon biochemical and cell-based studies. Extending these studies in faithful animal models of glioma with insights gleaned from human genetics will undoubtedly lead to much-needed advances in the treatment of these cancers.
Acknowledgements
We apologize to colleagues whose work could not be cited because of space limitations. We thank members of the Allis lab for their feedback regarding this work. C.D.A. and A.A.S. are supported by funding from The Rockefeller University and the National Institutes of Health (NIH) Grants P01CA196539 and R01CA234561. R.E.P is supported by the MSKCC IDEA Award, The Center for Experimental Therapeutics, and Team Jack Foundation.
Footnotes
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Declaration of Interests
The authors declare no conflict of interest.
REFERENCES
- Ahmad K, and Henikoff S (2002). The Histone Variant H3.3 Marks Active Chromatin by Replication-Independent Nucleosome Assembly. Molecular Cell 9, 1191–1200. [DOI] [PubMed] [Google Scholar]
- Aihara K, Mukasa A, Gotoh K, Saito K, Nagae G, Tsuji S, Tatuno K, Yamamoto S, Takayanagi S, Narita Y, et al. (2013). H3F3A K27M mutations in thalamic gliomas from young adult patients. Neuro Oncol not144. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Allfrey VG, Faulkner R, and Mirsky AE (1964). ACETYLATION AND METHYLATION OF HISTONES AND THEIR POSSIBLE ROLE IN THE REGULATION OF RNA SYNTHESIS*. Proc Natl Acad SciU SA 51, 786–794. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Allis CD, and Jenuwein T (2016). The molecular hallmarks of epigenetic control. Nat Rev Genet 17, 487–500. [DOI] [PubMed] [Google Scholar]
- Angel A, Song J, Dean C, and Howard M (2011). A Polycomb-based switch underlying quantitative epigenetic memory. Nature 476, 105–108. [DOI] [PubMed] [Google Scholar]
- Armache A, Yang S, Robbins LE, Durmaz C, Daman AW, Jeong JQ, Paz A.M. de, Ravishankar A, Arslan T, Lin S, et al. (2019). Phosphorylation of the ancestral histone variant H3.3 amplifies stimulation-induced transcription. BioRxiv 808048. [Google Scholar]
- Bannister AJ, Zegerman P, Partridge JF, Miska EA, Thomas JO, Allshire RC, and Kouzarides T (2001). Selective recognition of methylated lysine 9 on histone H3 by the HP1 chromo domain. Nature 410, 120–124. [DOI] [PubMed] [Google Scholar]
- Bender S, Tang Y, Lindroth AM, Hovestadt V, Jones DTW, Kool M, Zapatka M, Northcott PA, Sturm D, Wang W, et al. (2013). Reduced H3K27me3 and DNA Hypomethylation Are Major Drivers of Gene Expression in K27M Mutant Pediatric High-Grade Gliomas. Cancer Cell 24, 660–672. [DOI] [PubMed] [Google Scholar]
- Berry S, Hartley M, Olsson TSG, Dean C, and Howard M (2015). Local chromatin environment of a Polycomb target gene instructs its own epigenetic inheritance. ELife 4, e07205. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bledea R, Vasudevaraja V, Patel S, Stafford J, Serrano J, Esposito G, Tredwin LM, Goodman N, Kloetgen A, Golfinos JG, et al. (2019). Functional and topographic effects on DNA methylation in IDH1/2 mutant cancers. Sci Rep 9, 1–12. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bondeson DP, and Crews CM (2017). Targeted Protein Degradation by Small Molecules. Annual Review of Pharmacology and Toxicology 57, 107–123. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brennan CW, Verhaak RGW, McKenna A, Campos B, Noushmehr H, Salama SR, Zheng S, Chakravarty D, Sanborn JZ, Berman SH, et al. (2013). The somatic genomic landscape of glioblastoma. Cell 155, 462–477. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brosnan-Cashman JA, Yuan M, Graham MK, Rizzo AJ, Myers KM, Davis C, Zhang R, Esopi DM, Raabe EH, Eberhart CG, et al. (2018). ATRX loss induces multiple hallmarks of the alternative lengthening of telomeres (ALT) phenotype in human glioma cell lines in a cell line-specific manner. PLOS ONE 13, e0204159. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brown ZZ, Müller MM, Jain SU, Allis CD, Lewis PW, and Muir TW (2014). Strategy for “Detoxification” of a Cancer-Derived Histone Mutant Based on Mapping Its Interaction with the Methyltransferase PRC2. J Am Chem Soc 136, 13498–13501. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brownell JE, Zhou J, Ranalli T, Kobayashi R, Edmondson DG, Roth SY, and Allis CD (1996). Tetrahymena Histone Acetyltransferase A: A Homolog to Yeast Gcn5p Linking Histone Acetylation to Gene Activation. Cell 84, 843–851. [DOI] [PubMed] [Google Scholar]
- Cao R, Wang L, Wang H, Xia L, Erdjument-Bromage H, Tempst P, Jones RS, and Zhang Y (2002). Role of histone H3 lysine 27 methylation in Polycomb-group silencing. Science 298, 1039–1043. [DOI] [PubMed] [Google Scholar]
- Cesare AJ, and Reddel RR (2010). Alternative lengthening of telomeres: models, mechanisms and implications. Nat Rev Genet 11, 319–330. [DOI] [PubMed] [Google Scholar]
- Chan K-M, Fang D, Gan H, Hashizume R, Yu C, Schroeder M, Gupta N, Mueller S, James CD, Jenkins R, et al. (2013). The histone H3.3K27M mutation in pediatric glioma reprograms H3K27 methylation and gene expression. Genes & Development 27, 985–990. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cheung V, Chua G, Batada NN, Landry CR, Michnick SW, Hughes TR, and Winston F (2008). Chromatin-and transcription-related factors repress transcription from within coding regions throughout the Saccharomyces cerevisiae genome. PLoS Biol. 6, e277. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Clendening JW, Pandyra A, Li Z, Boutros PC, Martirosyan A, Lehner R, Jurisica I, Trudel S, and Penn LZ (2010). Exploiting the mevalonate pathway to distinguish statin-sensitive multiple myeloma. Blood 115, 4787–4797. [DOI] [PubMed] [Google Scholar]
- Clynes D, Jelinska C, Xella B, Ayyub H, Taylor S, Mitson M, Bachrati CZ, Higgs DR, and Gibbons RJ (2014). ATRX Dysfunction Induces Replication Defects in Primary Mouse Cells. PLoS One 9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Clynes D, Jelinska C, Xella B, Ayyub H, Scott C, Mitson M, Taylor S, Higgs DR, and Gibbons RJ (2015). Suppression of the alternative lengthening of telomere pathway by the chromatin remodelling factor ATRX. Nat Commun 6, 7538. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Comb M, and Goodman HM (1990). CpG methylation inhibits proenkephalin gene expression and binding of the transcription factor AP-2. Nucleic Acids Res. 18, 3975–3982. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dang L, White DW, Gross S, Bennett BD, Bittinger MA, Driggers EM, Fantin VR, Jang HG, Jin S, Keenan MC, et al. (2009). Cancer-associated IDH1 mutations produce 2-hydroxyglutarate. Nature 462, 739. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Danussi C, Bose P, Parthasarathy PT, Silberman PC, Van Arnam JS, Vitucci M, Tang OY, Heguy A, Wang Y, Chan TA, et al. (2018). Atrx inactivation drives disease-defining phenotypes in glioma cells of origin through global epigenomic remodeling. Nat Commun 9, 1057. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Daver NG, Garcia-Manero G, Konopleva MY, Alfayez M, Pemmaraju N, Kadia TM, DiNardo CD, Cortes JE, Ravandi F, Abbas H, et al. (2019). Azacitidine (AZA) with Nivolumab (Nivo), and AZA with Nivo + Ipilimumab (Ipi) in Relapsed/Refractory Acute Myeloid Leukemia: A Non-Randomized, Prospective, Phase 2 Study. Blood 134, 830–830. [Google Scholar]
- De La Fuente R, Baumann C, and Viveiros MM (2015). ATRX contributes to epigenetic asymmetry and silencing of major satellite transcripts in the maternal genome of the mouse embryo. Development 142, 1806–1817. [DOI] [PMC free article] [PubMed] [Google Scholar]
- DiNardo CD, Stein EM, de Botton S, Roboz GJ, Altman JK, Mims AS, Swords R, Collins RH, Mannis GN, Pollyea DA, et al. (2018). Durable Remissions with Ivosidenib in IDH1-Mutated Relapsed or Refractory AML. N. Engl. J. Med 378, 2386–2398. [DOI] [PubMed] [Google Scholar]
- Dixon JR, Selvaraj S, Yue F, Kim A, Li Y, Shen Y, Hu M, Liu JS, and Ren B (2012). Topological domains in mammalian genomes identified by analysis of chromatin interactions. Nature 485, 376–380. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Drané P, Ouararhni K, Depaux A, Shuaib M, and Hamiche A (2010). The death-associated protein DAXX is a novel histone chaperone involved in the replication-independent deposition of H3.3. Genes Dev 24, 1253–1265. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dvinge H, Kim E, Abdel-Wahab O, and Bradley RK (2016). RNA splicing factors as oncoproteins and tumor suppressors. Nat Rev Cancer 16, 413–430. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Elsässer SJ, Huang H, Lewis PW, Chin JW, Allis CD, and Patel DJ (2012). DAXX envelops a histone H3.3–H4 dimer for H3.3-specific recognition. Nature 491, 560–565. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Eustermann S, Yang J-C, Law MJ, Amos R, Chapman LM, Jelinska C, Garrick D, Clynes D, Gibbons RJ, Rhodes D, et al. (2011). Combinatorial readout of histone H3 modifications specifies localization of ATRX to heterochromatin. Nat. Struct. Mol. Biol 18, 777–782. [DOI] [PubMed] [Google Scholar]
- Fang D, Gan H, Cheng L, Lee J-H, Zhou H, Sarkaria JN, Daniels DJ, and Zhang Z (2018a). H3.3K27M mutant proteins reprogram epigenome by sequestering the PRC2 complex to poised enhancers. ELife 7, e36696. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fang J, Huang Y, Mao G, Yang S, Rennert G, Gu L, Li H, and Li G-M (2018b). Cancer-driving H3G34V/R/D mutations block H3K36 methylation and H3K36me3-MutSαinteraction. Proc. Natl. Acad. Sci. U.S.A 115, 9598–9603. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Figueroa ME, Abdel-Wahab O, Lu C, Ward PS, Patel J, Shih A, Li Y, Bhagwat N, Vasanthakumar A, Fernandez HF, et al. (2010). Leukemic IDH1 and IDH2 mutations result in a hypermethylation phenotype, disrupt TET2 function, and impair hematopoietic differentiation. Cancer Cell 18, 553–567. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Filbin MG, Tirosh I, Hovestadt V, Shaw ML, Escalante LE, Mathewson ND, Neftel C, Frank N, Pelton K, Hebert CM, et al. (2018). Developmental and oncogenic programs in H3K27M gliomas dissected by single-cell RNA-seq. Science 360, 331–335. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fine HA (2015). New Strategies in Glioblastoma: Exploiting the New Biology. Clinical Cancer Research 21, 1984–1988. [DOI] [PubMed] [Google Scholar]
- Fishbein L, Khare S, Wubbenhorst B, DeSloover D, D’Andrea K, Merrill S, Cho NW, Greenberg RA, Else T, Montone K, et al. (2015). Whole-exome sequencing identifies somatic ATRX mutations in pheochromocytomas and paragangliomas. Nature Communications 6, 6140. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Flavahan WA, Drier Y, Liau BB, Gillespie SM, Venteicher AS, Stemmer-Rachamimov AO, Suvà ML, and Bernstein BE (2016). Insulator dysfunction and oncogene activation in IDH mutant gliomas. Nature 529, 110–114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fontebasso AM, Schwartzentruber J, Khuong-Quang D-A, Liu X-Y, Sturm D, Korshunov A, Jones DTW, Witt H, Kool M, Albrecht S, et al. (2013). Mutations in SETD2 and genes affecting histone H3K36 methylation target hemispheric high-grade gliomas. Acta Neuropathol. 125, 659–669. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Francis NJ, Kingston RE, and Woodcock CL (2004). Chromatin compaction by a polycomb group protein complex. Science 306, 1574–1577. [DOI] [PubMed] [Google Scholar]
- Funato K, Major T, Lewis PW, Allis CD, and Tabar V (2014). Use of human embryonic stem cells to model pediatric gliomas with H3.3K27M histone mutation. Science 346, 1529–1533. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gandhi L, Janne PA, Opyrchal M, Ramalingam SS, Rybkin II, Hafez N, Raez LE, Gabrilovich D, Wang F, Ordentlich P, et al. (2018). Efficacy and safety of entinostat (ENT) and pembrolizumab (PEMBRO) in patients with non-small cell lung cancer (NSCLC) previously treated with anti-PD-(L)1 therapy. JCO 36, 9036–9036. [Google Scholar]
- Goldberg AD, Banaszynski LA, Noh K-M, Lewis PW, Elsaesser SJ, Stadler S, Dewell S, Law M, Guo X, Li X, et al. (2010). Distinct factors control histone variant H3.3 localization at specific genomic regions. Cell 140, 678–691. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Grasso CS, Tang Y, Truffaux N, Berlow NE, Liu L, Debily M-A, Quist MJ, Davis LE, Huang EC, Woo PJ, et al. (2015). Functionally defined therapeutic targets in diffuse intrinsic pontine glioma. Nat. Med 21, 555–559. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gröbner SN, Worst BC, Weischenfeldt J, Buchhalter I, Kleinheinz K, Rudneva VA, Johann PD, Balasubramanian GP, Segura-Wang M, Brabetz S, et al. (2018). The landscape of genomic alterations across childhood cancers. Nature 555, 321–327. [DOI] [PubMed] [Google Scholar]
- Harutyunyan AS, Krug B, Chen H, Papillon-Cavanagh S, Zeinieh M, Jay ND, Deshmukh S, Chen CCL, Belle J, Mikael LG, et al. (2019). H3K27M induces defective chromatin spread of PRC2-mediated repressive H3K27me2/me3 and is essential for glioma tumorigenesis. Nat Commun 10, 1–13. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hawkins RD, Hon GC, Lee LK, Ngo Q, Lister R, Pelizzola M, Edsall LE, 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. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Heaphy CM, de Wilde RF, Jiao Y, Klein AP, Edil BH, Shi C, Bettegowda C, Rodriguez FJ, Eberhart CG, Hebbar S, et al. (2011). Altered Telomeres in Tumors with ATRX and DAXX Mutations. Science 333, 425. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hirschhorn JN, Brown SA, Clark CD, and Winston F (1992). Evidence that SNF2/SWI2 and SNF5 activate transcription in yeast by altering chromatin structure. Genes Dev. 6, 2288–2298. [DOI] [PubMed] [Google Scholar]
- Huang H, Sabari BR, Garcia BA, Allis CD, and Zhao Y (2014). SnapShot: Histone Modifications. Cell 159, 458–458.e1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huber F, Bunina D, Gupta I, Khmelinskii A, Meurer M, Theer P, Steinmetz LM, and Knop M (2016). Protein Abundance Control by Non-coding Antisense Transcription. Cell Rep 15, 2625–2636. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hübner J-M, Müller T, Papageorgiou DN, Mauermann M, Krijgsveld J, Russell RB, Ellison DW, Pfister SM, Pajtler KW, and Kool M (2019). EZHIP/CXorf67 mimics K27M mutated oncohistones and functions as an intrinsic inhibitor of PRC2 function in aggressive posterior fossa ependymoma. Neuro-Oncology 21, 878–889. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huh MS, Price O’Dea T, Ouazia D, McKay BC, Parise G, Parks RJ, Rudnicki MA, and Picketts DJ (2012). Compromised genomic integrity impedes muscle growth after Atrx inactivation. J Clin Invest 122, 4412–4423. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Iwase S, Xiang B, Ghosh S, Ren T, Lewis PW, Cochrane JC, Allis CD, Picketts DJ, Patel DJ, Li H, et al. (2011). ATRX ADD domain links an atypical histone methylation recognition mechanism to human mental-retardation syndrome. Nat. Struct. Mol. Biol 18, 769–776. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jain SU, Do TJ, Lund PJ, Rashoff AQ, Diehl KL, Cieslik M, Bajic A, Juretic N, Deshmukh S, Venneti S, et al. (2019). PFA ependymoma-associated protein EZHIP inhibits PRC2 activity through a H3 K27M-like mechanism. Nat Commun 10, 1–14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- James TC, and Elgin SC (1986). Identification of a nonhistone chromosomal protein associated with heterochromatin in Drosophila melanogaster and its gene. Mol. Cell. Biol 6, 3862–3872. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jani KS, Jain SU, Ge EJ, Diehl KL, Lundgren SM, Müller MM, Lewis PW, and Muir TW (2019). Histone H3 tail binds a unique sensing pocket in EZH2 to activate the PRC2 methyltransferase. PNAS 116, 8295–8300. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jeggo PA, Pearl LH, and Carr AM (2016). DNA repair, genome stability and cancer: a historical perspective. Nat Rev Cancer 16, 35–42. [DOI] [PubMed] [Google Scholar]
- Jessa S, Blanchet-Cohen A, Krug B, Vladoiu M, Coutelier M, Faury D, Poreau B, De Jay N, Hébert S, Monlong J, et al. (2019). Stalled developmental programs at the root of pediatric brain tumors. Nature Genetics 51, 1702–1713. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jiao Y, Killela PJ, Reitman ZJ, Rasheed AB, Heaphy CM, de Wilde RF, Rodriguez FJ, Rosemberg S, Oba-Shinjo SM, Nagahashi Marie SK, et al. (2012). Frequent ATRX, CIC, FUBP1 and IDH1 mutations refine the classification of malignant gliomas. Oncotarget 3, 709–722. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Johannessen T-CA, Mukherjee J, Viswanath P, Ohba S, Ronen SM, Bjerkvig R, and Pieper RO (2016). Rapid Conversion of Mutant IDH1 from Driver to Passenger in a Model of Human Gliomagenesis. Mol. Cancer Res 14, 976–983. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Johnson BE, Mazor T, Hong C, Barnes M, Aihara K, McLean CY, Fouse SD, Yamamoto S, Ueda H, Tatsuno K, et al. (2014). Mutational Analysis Reveals the Origin and Therapy-Driven Evolution of Recurrent Glioma. Science 343, 189–193. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jones PL, Veenstra GJ, Wade PA, Vermaak D, Kass SU, Landsberger N, Strouboulis J, and Wolffe AP (1998). Methylated DNA and MeCP2 recruit histone deacetylase to repress transcription. Nat. Genet 19, 187–191. [DOI] [PubMed] [Google Scholar]
- Justin N, Zhang Y, Tarricone C, Martin SR, Chen S, Underwood E, Marco VD, Haire LF, Walker PA, Reinberg D, et al. (2016). Structural basis of oncogenic histone H3K27M inhibition of human polycomb repressive complex 2. Nat Commun 7, 1–11. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kadoch C, Williams RT, Calarco JP, Miller EL, Weber CM, Braun SG, Pulice JL, Chory EJ, and Crabtree GR (2017). Dynamics of BAF-Polycomb Complex Opposition on Heterochromatin in Normal and Oncogenic States. Nat Genet 49, 213–222. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kaminskas E, Farrell A, Abraham S, Baird A, Hsieh L-S, Lee S-L, Leighton JK, Patel H, Rahman A, Sridhara R, et al. (2005). Approval summary: azacitidine for treatment of myelodysplastic syndrome subtypes. Clin. Cancer Res 11, 3604–3608. [DOI] [PubMed] [Google Scholar]
- Kaplan CD, Laprade L, and Winston F (2003). Transcription elongation factors repress transcription initiation from cryptic sites. Science 301, 1096–1099. [DOI] [PubMed] [Google Scholar]
- Khuong-Quang D-A, Buczkowicz P, Rakopoulos P, Liu X-Y, Fontebasso AM, Bouffet E, Bartels U, Albrecht S, Schwartzentruber J, Letourneau L, et al. (2012). K27M mutation in histone H3.3 defines clinically and biologically distinct subgroups of pediatric diffuse intrinsic pontine gliomas. Acta Neuropathol 124, 439–447. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim JH, Lee BB, Oh YM, Zhu C, Steinmetz LM, Lee Y, Kim WK, Lee SB, Buratowski S, and Kim T (2016). Modulation of mRNA and lncRNA expression dynamics by the Set2-Rpd3S pathway. Nat Commun 7, 13534. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Knutson SK, Warholic NM, Wigle TJ, Klaus CR, Allain CJ, Raimondi A, Scott MP, Chesworth R, Moyer MP, Copeland RA, et al. (2013). Durable tumor regression in genetically altered malignant rhabdoid tumors by inhibition of methyltransferase EZH2. PNAS 110, 7922–7927. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Koschmann C, Calinescu A-A, Nunez FJ, Mackay A, Fazal-Salom J, Thomas D, Mendez F, Kamran N, Dzaman M, Mulpuri L, et al. (2016). ATRX loss promotes tumor growth and impairs nonhomologous end joining DNA repair in glioma. Sci Transl Med 8, 328ra28. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Krogan NJ, Kim M, Tong A, Golshani A, Cagney G, Canadien V, Richards DP, K. Beattie B, Emili A, Boone C, et al. (2003). Methylation of Histone H3 by Set2 in Saccharomyces cerevisiae Is Linked to Transcriptional Elongation by RNA Polymerase II. Mol Cell Biol 23, 4207–4218. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Krug B, De Jay N, Harutyunyan AS, Deshmukh S, Marchione DM, Guilhamon P, Bertrand KC, Mikael LG, McConechy MK, Chen CCL, et al. (2019). Pervasive H3K27 Acetylation Leads to ERV Expression and a Therapeutic Vulnerability in H3K27M Gliomas. Cancer Cell 35, 782–797.e8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lachner M, O’Carroll D, Rea S, Mechtler K, and Jenuwein T (2001). Methylation of histone H3 lysine 9 creates a binding site for HP1 proteins. Nature 410, 116–120. [DOI] [PubMed] [Google Scholar]
- Lai A, Kharbanda S, Pope WB, Tran A, Solis OE, Peale F, Forrest WF, Pujara K, Carrillo JA, Pandita A, et al. (2011). Evidence for sequenced molecular evolution of IDH1 mutant glioblastoma from a distinct cell of origin. J. Clin. Oncol 29, 4482–4490. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Larson JD, Kasper LH, Paugh BS, Jin H, Wu G, Kwon C-H, Fan Y, Shaw TI, Silveira AB, Qu C, et al. (2019). Histone H3.3 K27M Accelerates Spontaneous Brainstem Glioma and Drives Restricted Changes in Bivalent Gene Expression. Cancer Cell 35, 140–155.e7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lau MS, Schwartz MG, Kundu S, Savol AJ, Wang PI, Marr SK, Grau DJ, Schorderet P, Sadreyev RI, Tabin CJ, et al. (2017). Mutation of a nucleosome compaction region disrupts Polycomb-mediated axial patterning. Science 355, 1081–1084. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee C-H, Yu J-R, Granat J, Saldaña-Meyer R, Andrade J, LeRoy G, Jin Y, Lund P, Stafford JM, Garcia BA, et al. (2019). Automethylation of PRC2 promotes H3K27 methylation and is impaired in H3K27M pediatric glioma. Genes Dev. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lewis EB (1978). A gene complex controlling segmentation in Drosophila. Nature 276, 565–570. [DOI] [PubMed] [Google Scholar]
- Lewis PW, Elsaesser SJ, Noh K-M, Stadler SC, and Allis CD (2010). Daxx is an H3.3-specific histone chaperone and cooperates with ATRX in replication-independent chromatin assembly at telomeres. Proc Natl Acad Sci U S A 107, 14075–14080. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lewis PW, Muller MM, Koletsky MS, Cordero F, Lin S, Banaszynski LA, Garcia BA, Muir TW, Becher OJ, and Allis CD (2013). Inhibition of PRC2 Activity by a Gain-of-Function H3 Mutation Found in Pediatric Glioblastoma. Science 340, 857–861. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li F, Mao G, Tong D, Huang J, Gu L, Yang W, and Li G-M (2013). The histone mark H3K36me3 regulates human DNA mismatch repair through its interaction with MutSα. Cell 153, 590–600. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li J, Moazed D, and Gygi SP (2002). Association of the histone methyltransferase Set2 with RNA polymerase II plays a role in transcription elongation. J. Biol. Chem 277, 49383–49388. [DOI] [PubMed] [Google Scholar]
- Liu X-Y, Gerges N, Korshunov A, Sabha N, Khuong-Quang D-A, Fontebasso AM, Fleming A, Hadjadj D, Schwartzentruber J, Majewski J, et al. (2012). Frequent ATRX mutations and loss of expression in adult diffuse astrocytic tumors carrying IDH1/IDH2 and TP53 mutations. Acta Neuropathol. 124, 615–625. [DOI] [PubMed] [Google Scholar]
- Lu C, Ward PS, Kapoor GS, Rohle D, Turcan S, Abdel-Wahab O, Edwards CR, Khanin R, Figueroa ME, Melnick A, et al. (2012). IDH mutation impairs histone demethylation and results in a block to cell differentiation. Nature 483, 474–478. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Luchman HA, Chesnelong C, Cairncross JG, and Weiss S (2013). Spontaneous loss of heterozygosity leading to homozygous R132H in a patient-derived IDH1 mutant cell line. Neuro Oncol 15, 979–980. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Luco RF, Pan Q, Tominaga K, Blencowe BJ, Pereira-Smith OM, and Misteli T (2010). Regulation of alternative splicing by histone modifications. Science 327, 996–1000. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Luger K, Mäder AW, Richmond RK, Sargent DF, and Richmond TJ (1997). Crystal structure of the nucleosome core particle at 2.8 A resolution. Nature 389, 251–260. [DOI] [PubMed] [Google Scholar]
- Mardis ER (2018). Insights from Large-Scale Cancer Genome Sequencing. Annual Review of Cancer Biology 2, 429–444. [Google Scholar]
- Marks PA (2010). The clinical development of histone deacetylase inhibitors as targeted anticancer drugs. Expert Opin Investig Drugs 19, 1049–1066. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Martire S, Gogate AA, Whitmill A, Tafessu A, Nguyen J, Teng Y-C, Tastemel M, and Banaszynski LA (2019). Phosphorylation of histone H3.3 at serine 31 promotes p300 activity and enhancer acetylation. Nat Genet 51, 941–946. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Maze I, Wenderski W, Noh K-M, Bagot RC, Tzavaras N, Purushothaman I, Elsässer SJ, Guo Y, Ionete C, Hurd YL, et al. (2015). Critical Role of Histone Turnover in Neuronal Transcription and Plasticity. Neuron 87, 77–94. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mazor T, Chesnelong C, Pankov A, Jalbert LE, Hong C, Hayes J, Smirnov IV, Marshall R, Souza CF, Shen Y, et al. (2017). Clonal expansion and epigenetic reprogramming following deletion or amplification of mutant IDH1. PNAS 201708914. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McDaniel SL, and Strahl BD (2017). Shaping the Cellular Landscape with Set2/SETD2 Methylation. Cell Mol Life Sci 74, 3317–3334. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Modrek AS, Golub D, Khan T, Bready D, Prado J, Bowman C, Deng J, Zhang G, Rocha PP, Raviram R, et al. (2017). Low-Grade Astrocytoma Mutations in IDH1, P53, and ATRX Cooperate to Block Differentiation of Human Neural Stem Cells via Repression of SOX2. Cell Rep 21, 1267–1280. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mohammad F, Weissmann S, Leblanc B, Pandey DP, Højfeldt JW, Comet I, Zheng C, Johansen JV, Rapin N, Porse BT, et al. (2017). EZH2 is a potential therapeutic target for H3K27M-mutant pediatric gliomas. Nat Med advance online publication. [DOI] [PubMed] [Google Scholar]
- Müller J, Hart CM, Francis NJ, Vargas ML, Sengupta A, Wild B, Miller EL, O’Connor MB, Kingston RE, and Simon JA (2002). Histone methyltransferase activity of a Drosophila Polycomb group repressor complex. Cell 111, 197–208. [DOI] [PubMed] [Google Scholar]
- Musselman CA, Avvakumov N, Watanabe R, Abraham CG, Lalonde M-E, Hong Z, Allen C, Roy S, Nuñez JK, Nickoloff J, et al. (2012). Molecular basis for H3K36me3 recognition by the Tudor domain of PHF1. Nat Struct Mol Biol 19, 1266–1272. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nacev BA, Feng L, Bagert JD, Lemiesz AE, Gao J, Soshnev AA, Kundra R, Schultz N, Muir TW, and Allis CD (2019). The expanding landscape of ‘oncohistone’ mutations in human cancers. Nature 567, 473–478. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nakahashi H, Kieffer Kwon K-R, Resch W, Vian L, Dose M, Stavreva D, Hakim O, Pruett N, Nelson S, Yamane A, et al. (2013). A genome-wide map of CTCF multivalency redefines the CTCF code. Cell Rep 3, 1678–1689. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nakayama J, Rice JC, Strahl BD, Allis CD, and Grewal SIS (2001). Role of Histone H3 Lysine 9 Methylation in Epigenetic Control of Heterochromatin Assembly. Science 292, 110–113. [DOI] [PubMed] [Google Scholar]
- Neil H, Malabat C, d’Aubenton-Carafa Y, Xu Z, Steinmetz LM, and Jacquier A (2009). Widespread bidirectional promoters are the major source of cryptic transcripts in yeast. Nature 457, 1038–1042. [DOI] [PubMed] [Google Scholar]
- Neri F, Rapelli S, Krepelova A, Incarnato D, Parlato C, Basile G, Maldotti M, Anselmi F, and Oliviero S (2017). Intragenic DNA methylation prevents spurious transcription initiation. Nature 543, 72–77. [DOI] [PubMed] [Google Scholar]
- Nicetto D, Donahue G, Jain T, Peng T, Sidoli S, Sheng L, Montavon T, Becker JS, Grindheim JM, Blahnik K, et al. (2019). H3K9me3-heterochromatin loss at protein-coding genes enables developmental lineage specification. Science 363, 294–297. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Oksuz O, Narendra V, Lee C-H, Descostes N, LeRoy G, Raviram R, Blumenberg L, Karch K, Rocha PP, Garcia BA, et al. (2018). Capturing the Onset of PRC2-Mediated Repressive Domain Formation. Molecular Cell 70, 1149–1162.e5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pajtler KW, Witt H, Sill M, Jones DTW, Hovestadt V, Kratochwil F, Wani K, Tatevossian R, Punchihewa C, Johann P, et al. (2015). Molecular Classification of Ependymal Tumors across All CNS Compartments, Histopathological Grades, and Age Groups. Cancer Cell 27, 728–743. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pajtler KW, Wen J, Sill M, Lin T, Orisme W, Tang B, Hübner J-M, Ramaswamy V, Jia S, Dalton JD, et al. (2018). Molecular heterogeneity and CXorf67 alterations in posterior fossa group A (PFA) ependymomas. Acta Neuropathol. 136, 211–226. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pan D, Kobayashi A, Jiang P, de Andrade LF, Tay RE, Luoma AM, Tsoucas D, Qiu X, Lim K, Rao P, et al. (2018). A major chromatin regulator determines resistance of tumor cells to T cell–mediated killing. Science 359, 770–775. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Parsons DW, Jones S, Zhang X, Lin JC-H, Leary RJ, Angenendt P, Mankoo P, Carter H, Siu I-M, Gallia GL, et al. (2008). An Integrated Genomic Analysis of Human Glioblastoma Multiforme. Science 321, 1807–1812. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pathania M, De Jay N, Maestro N, Harutyunyan AS, Nitarska J, Pahlavan P, Henderson S, Mikael LG, Richard-Londt A, Zhang Y, et al. (2017). H3.3K27M Cooperates with Trp53 Loss and PDGFRA Gain in Mouse Embryonic Neural Progenitor Cells to Induce Invasive High-Grade Gliomas. Cancer Cell 32, 684–700.e9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pavlova NN, and Thompson CB (2016). THE EMERGING HALLMARKS OF CANCER METABOLISM. Cell Metab 23, 27–47. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pengelly AR, Copur Ö, Jäckle H, Herzig A, and Muller J (2013). A Histone Mutant Reproduces the Phenotype Caused by Loss of Histone-Modifying Factor Polycomb. Science 339, 698–699. [DOI] [PubMed] [Google Scholar]
- Piunti A, Hashizume R, Morgan MA, Bartom ET, Horbinski CM, Marshall SA, Rendleman EJ, Ma Q, Takahashi Y-H, Woodfin AR, et al. (2017). Therapeutic targeting of polycomb and BET bromodomain proteins in diffuse intrinsic pontine gliomas. Nat. Med [DOI] [PMC free article] [PubMed] [Google Scholar]
- Piunti A, Smith ER, Morgan MAJ, Ugarenko M, Khaltyan N, Helmin KA, Ryan CA, Murray DC, Rickels RA, Yilmaz BD, et al. (2019). CATACOMB: An endogenous inducible gene that antagonizes H3K27 methylation activity of Polycomb repressive complex 2 via an H3K27M-like mechanism. Sci Adv 5, eaax2887. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Plys AJ, Davis CP, Kim J, Rizki G, Keenen MM, Marr SK, and Kingston RE (2019). Phase separation of Polycomb-repressive complex 1 is governed by a charged disordered region of CBX2. Genes Dev. 33, 799–813. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Prendergast GC, and Ziff EB (1991). Methylation-sensitive sequence-specific DNA binding by the c-Myc basic region. Science 251, 186–189. [DOI] [PubMed] [Google Scholar]
- Ragazzini R, Pérez-Palacios R, Baymaz IH, Diop S, Ancelin K, Zielinski D, Michaud A, Givelet M, Borsos M, Aflaki S, et al. (2019). EZHIP constrains Polycomb Repressive Complex 2 activity in germ cells. Nat Commun 10, 1–18. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Razin A, and Riggs AD (1980). DNA methylation and gene function. Science 210, 604–610. [DOI] [PubMed] [Google Scholar]
- Reinhardt A, Stichel D, Schrimpf D, Sahm F, Korshunov A, Reuss DE, Koelsche C, Huang K, Wefers AK, Hovestadt V, et al. (2018). Anaplastic astrocytoma with piloid features, a novel molecular class of IDH wildtype glioma with recurrent MAPK pathway, CDKN2A/B and ATRX alterations. Acta Neuropathol. 136, 273–291. [DOI] [PubMed] [Google Scholar]
- Rohle D, Popovici-Muller J, Palaskas N, Turcan S, Grommes C, Campos C, Tsoi J, Clark O, Oldrini B, Komisopoulou E, et al. (2013). An Inhibitor of Mutant IDH1 Delays Growth and Promotes Differentiation of Glioma Cells. Science 340, 626–630. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rose NR, McDonough MA, King ONF, Kawamura A, and Schofield CJ (2011). Inhibition of 2-oxoglutarate dependent oxygenases. Chem Soc Rev 40, 4364–4397. [DOI] [PubMed] [Google Scholar]
- Rothbart SB, and Baylin SB (2020). Epigenetic Therapy for Epithelioid Sarcoma. Cell 181, 211. [DOI] [PubMed] [Google Scholar]
- Schuettengruber B, Bourbon H-M, Di Croce L, and Cavalli G (2017). Genome Regulation by Polycomb and Trithorax: 70 Years and Counting. Cell 171, 34–57. [DOI] [PubMed] [Google Scholar]
- Schvartzman J-M, Reuter VP, Koche RP, and Thompson CB (2019). 2-hydroxyglutarate inhibits MyoD-mediated differentiation by preventing H3K9 demethylation. Proc. Natl. Acad. Sci. U.S.A 116, 12851–12856. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schwartzentruber J, Korshunov A, Liu X-Y, Jones DTW, Pfaff E, Jacob K, Sturm D, Fontebasso AM, Quang D-AK, Tönjes M, et al. (2012). Driver mutations in histone H3.3 and chromatin remodelling genes in paediatric glioblastoma. Nature 482, 226–231. [DOI] [PubMed] [Google Scholar]
- Silveira AB, Kasper LH, Fan Y, Jin H, Wu G, Shaw TI, Zhu X, Larson JD, Easton J, Shao Y, et al. (2019). H3.3 K27M depletion increases differentiation and extends latency of diffuse intrinsic pontine glioma growth in vivo. Acta Neuropathol. 137, 637–655. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Simon JM, Hacker KE, Singh D, Brannon AR, Parker JS, Weiser M, Ho TH, Kuan P-F, Jonasch E, Furey TS, et al. (2014). Variation in chromatin accessibility in human kidney cancer links H3K36 methyltransferase loss with widespread RNA processing defects. Genome Res. 24, 241–250. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stacchiotti S, Schoffski P, Jones R, Agulnik M, Villalobos VM, Jahan TM, Chen TW-W, Italiano A, Demetri GD, Cote GM, et al. (2019). Safety and efficacy of tazemetostat, a first-in-class EZH2 inhibitor, in patients (pts) with epithelioid sarcoma (ES) (NCT02601950). JCO 37, 11003–11003. [Google Scholar]
- Stadler MB, Murr R, Burger L, Ivanek R, Lienert F, Schöler A, van Nimwegen E, Wirbelauer C, Oakeley EJ, Gaidatzis D, et al. (2011). DNA-binding factors shape the mouse methylome at distal regulatory regions. Nature 480, 490–495. [DOI] [PubMed] [Google Scholar]
- Stafford JM, Lee C-H, Voigt P, Descostes N, Saldaña-Meyer R, Yu J-R, Leroy G, Oksuz O, Chapman JR, Suarez F, et al. (2018). Multiple modes of PRC2 inhibition elicit global chromatin alterations in H3K27M pediatric glioma. Science Advances 4, eaau5935. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Strahl BD, and Allis CD (2000). The language of covalent histone modifications. Nature 403, 41–45. [DOI] [PubMed] [Google Scholar]
- Sturm D, Witt H, Hovestadt V, Khuong-Quang D-A, Jones DTW, Konermann C, Pfaff E, Tönjes M, Sill M, Bender S, et al. (2012). Hotspot Mutations in H3F3A and IDH1 Define Distinct Epigenetic and Biological Subgroups of Glioblastoma. Cancer Cell 22, 425–437. [DOI] [PubMed] [Google Scholar]
- Sturm D, Bender S, Jones DTW, Lichter P, Grill J, Becher O, Hawkins C, Majewski J, Jones C, Costello JF, et al. (2014). Paediatric and adult glioblastoma: multiform (epi)genomic culprits emerge. Nat Rev Cancer 14, 92–107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sulkowski PL, Corso CD, Robinson ND, Scanlon SE, Purshouse KR, Bai H, Liu Y, Sundaram RK, Hegan DC, Fons NR, et al. (2017). 2-Hydroxyglutarate produced by neomorphic IDH mutations suppresses homologous recombination and induces PARP inhibitor sensitivity. Sci Transl Med 9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sulkowski PL, Oeck S, Dow J, Economos NG, Mirfakhraie L, Liu Y, Noronha K, Bao X, Li J, Shuch BM, et al. (2020). Oncometabolites suppress DNA repair by disrupting local chromatin signalling. Nature 582, 586–591. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Suzuki H, Aoki K, Chiba K, Sato Y, Shiozawa Y, Shiraishi Y, Shimamura T, Niida A, Motomura K, Ohka F, et al. (2015). Mutational landscape and clonal architecture in grade II and III gliomas. Nat Genet 47, 458–468. [DOI] [PubMed] [Google Scholar]
- Tatavosian R, Kent S, Brown K, Yao T, Duc HN, Huynh TN, Zhen CY, Ma B, Wang H, and Ren X (2018). Nuclear condensates of the Polycomb protein chromobox 2 (CBX2) assemble through phase separation. J. Biol. Chem. jbc.RA118.006620 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tateishi K, Wakimoto H, Iafrate AJ, Tanaka S, Loebel F, Lelic N, Wiederschain D, Bedel O, Deng G, Zhang B, et al. (2015). Extreme vulnerability of IDH1 mutant cancers to NAD+ depletion. Cancer Cell 28, 773–784. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Taunton J, Hassig CA, and Schreiber SL (1996). A mammalian histone deacetylase related to the yeast transcriptional regulator Rpd3p. Science 272, 408–411. [DOI] [PubMed] [Google Scholar]
- The Cancer Genome Atlas Reseach Network (2008). Comprehensive genomic characterization defines human glioblastoma genes and core pathways. Nature 455, 1061–1068. [DOI] [PMC free article] [PubMed] [Google Scholar]
- The Cancer Genome Atlas Reseach Network (2015). Comprehensive, Integrative Genomic Analysis of Diffuse Lower-Grade Gliomas. N Engl J Med 372, 2481–2498. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Turcan S, Rohle D, Goenka A, Walsh LA, Fang F, Yilmaz E, Campos C, Fabius AWM, Lu C, Ward PS, et al. (2012). IDH1 mutation is sufficient to establish the glioma hypermethylator phenotype. Nature 483, 479–483. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Varshney D, Spiegel J, Zyner K, Tannahill D, and Balasubramanian S (2020). The regulation and functions of DNA and RNA G-quadruplexes. Nature Reviews Molecular Cell Biology 1–16. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Verhaak RGW, Hoadley KA, Purdom E, Wang V, Qi Y, Wilkerson MD, Miller CR, Ding L, Golub T, Mesirov JP, et al. (2010). Integrated genomic analysis identifies clinically relevant subtypes of glioblastoma characterized by abnormalities in PDGFRA, IDH1, EGFR, and NF1. Cancer Cell 17, 98–110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Voon HPJ, Udugama M, Lin W, Hii L, Law RHP, Steer DL, Das PP, Mann JR, and Wong LH (2018). Inhibition of a K9/K36 demethylase by an H3.3 point mutation found in paediatric glioblastoma. Nat Commun 9, 1–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang X, Long Y, Paucek RD, Gooding AR, Lee T, Burdorf RM, and Cech TR (2019). Regulation of histone methylation by automethylation of PRC2. Genes Dev. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ward PS, Patel J, Wise DR, Abdel-Wahab O, Bennett BD, Coller HA, Cross JR, Fantin VR, Hedvat CV, Perl AE, et al. (2010). The common feature of leukemia-associated IDH1 and IDH2 mutations is a neomorphic enzyme activity converting alpha-ketoglutarate to 2-hydroxyglutarate. Cancer Cell 17, 225–234. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Watanabe T, Nobusawa S, Kleihues P, and Ohgaki H (2009). IDH1 Mutations Are Early Events in the Development of Astrocytomas and Oligodendrogliomas. Am J Pathol 174, 1149–1153. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wilson BG, and Roberts CWM (2011). SWI/SNF nucleosome remodellers and cancer. Nat Rev Cancer 11, 481–492. [DOI] [PubMed] [Google Scholar]
- Wilson BG, Wang X, Shen X, McKenna ES, Lemieux ME, Cho Y-J, Koellhoffer EC, Pomeroy SL, Orkin SH, and Roberts CWM (2010). Epigenetic antagonism between Polycomb and SWI/SNF complexes during oncogenic transformation. Cancer Cell 18, 316–328. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wojcik JB, Marchione DM, Sidoli S, Djedid A, Lisby A, Majewski J, and Garcia BA (2019). Epigenomic Reordering Induced by Polycomb Loss Drives Oncogenesis but Leads to Therapeutic Vulnerabilities in Malignant Peripheral Nerve Sheath Tumors. Cancer Res. 79, 3205–3219. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu G, Broniscer A, McEachron TA, Lu C, Paugh BS, Becksfort J, Qu C, Ding L, Huether R, Parker M, et al. (2012). Somatic histone H3 alterations in pediatric diffuse intrinsic pontine gliomas and non-brainstem glioblastomas. Nat. Genet 44, 251–253. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu J, Armstrong TS, and Gilbert MR (2016). Biology and management of ependymomas. Neuro Oncol 18, 902–913. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xu Z, Wei W, Gagneur J, Perocchi F, Clauder-Münster S, Camblong J, Guffanti E, Stutz F, Huber W, and Steinmetz LM (2009). Bidirectional promoters generate pervasive transcription in yeast. Nature 457, 1033–1037. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yang S, Zheng X, Lu C, Li G-M, Allis CD, and Li H (2016). Molecular basis for oncohistone H3 recognition by SETD2 methyltransferase. Genes Dev. 30, 1611–1616. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zehir A, Benayed R, Shah RH, Syed A, Middha S, Kim HR, Srinivasan P, Gao J, Chakravarty D, Devlin SM, et al. (2017). Mutational landscape of metastatic cancer revealed from prospective clinical sequencing of 10,000 patients. Nature Medicine 23, 703–713. [DOI] [PMC free article] [PubMed] [Google Scholar]