Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2021 Aug 22.
Published in final edited form as: Nat Rev Cancer. 2020 Aug 11;20(10):608–623. doi: 10.1038/s41568-020-0288-4

The epigenomics of sarcoma

Benjamin A Nacev 1,2, Kevin B Jones 3, Andrew M Intlekofer 4, Jamie SE Yu 5, C David Allis 2, William D Tap 1, Marc Ladanyi 6, Torsten O Nielsen 5,*
PMCID: PMC8380451  NIHMSID: NIHMS1720406  PMID: 32782366

Abstract

Epigenetic regulation is critical to physiologic control of development, cell fate, cell proliferation, genomic integrity, and fundamentally, transcriptional regulation. This epigenetic control occurs at multiple levels including through DNA methylation, histone modification, nucleosome remodeling, and modulation of three-dimensional chromatin structure. Alterations in genes that encode chromatin regulators are common among mesenchymal neoplasms, a collection of more than 160 tumor types including over 60 malignant variants (sarcomas) that have unique and varied genetic, biologic, and clinical characteristics. Here, we review sarcomas in which chromatin pathway alterations drive disease biology. Specifically, we emphasize examples of dysregulation of each level of epigenetic control though mechanisms that include metabolic effects on enzymes that regulate DNA methylation and histone posttranslational modification, mutations in histone genes, subunit loss or fusions in chromatin remodeling and modifying complexes, and disruption of higher-order chromatin structure. Epigenetic mechanisms of tumorigenesis have been implicated in mesenchymal tumors ranging from chondroblastoma and giant cell tumor of bone to chondrosarcoma, malignant peripheral nerve sheath tumor, synovial sarcoma, epithelioid sarcoma and Ewing sarcoma: aggressive diseases which present in a younger patient population than most cancers. Finally, we review current and potential future approaches for the development of sarcoma therapies based on this emerging understanding of chromatin dysregulation.

INTRODUCTION

Much of the focus of cancer biology has rightly emphasized cancer genetics owing to the observation that most cancers are driven by somatic genetic changes, ranging from viral introduction of an oncogene, to mutations that inappropriately activate or silence a gene, to copy number changes that amplify or remove a gene, or to rearrangements that aberrantly connect disparate sequences into a fusion transcript. Notably, a small but important fraction of these genetic changes physiologically dysregulate transcription across the whole genome. Herein enters the study of the epigenome, which is relevant in a number of cancer types, but appears particularly so in sarcomas, leukemias1,2, and lymphomas3,4 for reasons that remain unclear, but may be related to the mesenchymal and hematopoietic origins of these cancers, in contrast to carcinomas. The broad application of next generation sequencing provides additional granularity to our understanding of the alterations in chromatin regulators that can occur across myriad sarcoma histotypes.

Sarcoma is not a single entity, but rather a collection of more than 60 malignancies from within a broader set of over 160 different bone & soft tissue neoplasms, with diverse biologic and clinical characteristics. While the precise cell of origin is unknown for most sarcomas, sarcomas arise in tissues of mesenchymal lineage such as muscle, adipose tissue, and bone. Despite mesenchymal tissues making up the bulk of body mass, sarcomas are uncommon, comprising approximately 1% of human malignancies5. Clinically, sarcomas display a range of behavior from low grade tumors with minimal metastatic potential, to highly aggressive cancers with a tendency for widespread metastasis. For localized disease, surgery is typically the preferred treatment modality whereas chemotherapy or targeted therapy is used to treat metastatic disease6. The use of immunotherapy in sarcoma is largely still investigational, although immune checkpoint blockade is supported by expert guidelines in one common histotype (undifferentiated pleomorphic sarcoma)6.

The genetics of sarcoma are highly variable. Some histotypes have relatively simple genetics and are driven by chromosomal translocations leading to fusion oncogenes, classically Ewing sarcoma (EWSR1-FLI1) and synovial sarcoma (SSX-SS18). Other sarcomas have complex genomic alterations (osteosarcoma), are driven by copy number alterations (well-differentiated/dedifferentiated liposarcoma), or by mutations in canonical drivers (gastrointestinal stromal tumors)7.

Increasingly, many types of sarcoma have come to be considered as predominantly epigenetic diseases, with widespread epigenetic dysregulation initiated by a small number of, or even a single, genetic change7,8. The many translocation-associated sarcomas achieve substantial changes in the transcriptome of transformed cells from very few alterations in genomic DNA coding or copy number917. This fact in and of itself suggests that each of the fusion oncoproteins that associate with specific sarcoma subtypes taps into fundamental epigenomic mechanisms of transcriptome control.

For the purpose of this review, we will work from a specific definition of epigenetics. The meaning of the term has evolved over the decades and has been used in settings ranging from population genetics, to polygenic traits, to macromolecular structure, to the molecular genetics of DNA methylation. Here, we will define epigenetics as non-sequence alterations (or at least non-local sequence changes) in the genome that impact transcriptional programs of downstream genes. Epigenetic regulation exists in multiple levels, for many of which there exist instructive examples of altered function in sarcomas (Figure 1, Table 1). Working outward from the DNA sequence itself, there are some sarcomas driven by alteration of gene and promoter DNA methylation, such as a subset of chondrosarcomas that are driven by IDH mutations18. As DNA is packaged by wrapping around nucleosomes, some secondary modifications of the histone proteins that comprise those nucleosomes represent a second layer of epigenetic control of transcription. Two recently elucidated examples in mesenchymal tumors that appear to be driven by somatic mutations in genes encoding histones (‘oncohistones’) are giant cell tumor of bone and chondroblastoma19, although these and other histone mutations are observed in other sarcomas a well. The enzymatic complexes that act on nucleosomes provide a third layer of epigenetic complexity. These include enzymes that perform the ‘writing’, ‘reading’, or ‘erasing’ of posttranslational modifications (PTMs such as acetylation, methylation, phosphorylation, etc.) to histone proteins20. Malignant peripheral nerve sheath tumors are examples of sarcomas with alterations in a chromatin writing complex, whereas synovial sarcoma, malignant rhabdoid tumors, and epithelioid sarcoma are examples of sarcomas that are characterized by genetic alterations disrupting chromatin remodeling complexes involved in nucleosome repositioning or ejection (a fourth layer of complexity). Finally, the three-dimensional structure of chromatin (a fifth layer), with the creation of enhancer regions and their regulatory association with often distant genes, is a burgeoning area of investigation: FET (FUS, EWSR1, TAF15) domain proteins, the most common fusion partners in translocation-associated sarcoma fusion oncoproteins, are major drivers of phase-separation events that have been associated with these subnuclear organizational structures21. In this review, we will discuss how each of these five levels of epigenetic control can be lost in different sarcomas as a result of genetic alterations with an emphasis on how a better understanding of this dysregulation may provide opportunities for therapeutic intervention.

Figure 1 |. Schematic of five layers of epigenomics that drive transcriptional programs in sarcomas.

For Figure 1, please refer to Figure 1 of the manuscript Nacev BA, Jones KB, Intlekofer AM, Yu JSE, Allis CD, Tap WD, Ladanyi M, Nielsen TO. The epigenomics of sarcoma. Nat Rev Cancer. 2020 Oct;20(10):608–623. doi: 10.1038/s41568-020-0288-4. Epub 2020 Aug 11. PMID: 32782366.

Table 1 |.

Chromatin pathway mutations in sarcoma

Histotype Genetic alteration Epigenetic role of altered pathway Investigational targeted therapy
DNA methylation
Chondrosarcoma IDH1/IDH2 neomorphs Oncometabolite → alteration of DNA methylation and Histone modification ivosidenib (NCT02073994)180, enasidenib (NCT02273739)179
Histones
Giant cell tumor of bone (> 90%), chondroblastoma (> 90%), chondrosarcoma, osteosarcoma, UPS Histone H3 K36 or G34 mutations Common substrate for chromatin regulation Pressing need
Histones PTM ‘writers’
MPNST SUZ12 or EED loss of function Histone PTM ‘Writer’ Pressing need
PRC2
BCOR-Rearranged BCOR-CCNB3 Histone PTM ‘Writer’ Pressing need
small blue round cell tumors BCOR-MAML3 non-canonical PRC1
ZC3H7B-BCOR
KMT2D-BCOR
Endometrial stroma sarcoma JAZF1-SUZ12 Histone PTM ‘Writer’ Pressing need
JAZF1-PHF1 PRC2; non-canonical PRC1
MBTD1-EZHIP
ZC3H7B-BCOR
Chromatin remodeling complexes
Epithelioid sarcoma SMARCB1 loss Remodeling tazemetostat (FDA approved, 2020)
Malignant rhabdoid tumor SMARCB1 loss Remodeling EZH2 inhibitors (NCT02601937; NCT03213665; NCT02875548)211213
Synovial sarcoma SSX-SS18 fusion Remodeling Brd9 inhibitors
Higher order chromatin organization
FUS/EWSR1/TAF15 translocation-associated sarcomas FUS/EWSR1/TAF15-associated fusions Three-dimensional chromatin structure Pressing need

DNA METHYLATION

Using technologies based on microarrays and more recently reduced representation bisulfite sequencing, DNA methylation status, particularly 5-Me cytosine deposition at CpG islands, is currently one of the most heavily studied epigenetic changes in cancer. Research into DNA methylation is facilitated by its stability – patterns are retained biologically across cell states, and even after formalin-fixation and paraffin-embedding – as well as by its binary nature which lends itself to relatively straightforward bioinformatics. Ewing sarcoma, for example, has been shown to have a very distinct pattern of DNA methylation (including hypomethylation at enhancers targeted by the EWSR1-FLI1 oncoprotein) that clearly distinguishes it from other cancer types22,23. Indeed, even within the class of small blue round cell sarcomas, methylomes are subtype-specific and may have value in the differential diagnosis of histologically indistinguishable entities24,25. Methylation patterns also serve to distinguish among subtypes of rhabdomyosarcoma26, nerve sheath tumors27 and even within otherwise genetically homogeneous entities such as malignant rhabdoid tumors28. In dedifferentiated liposarcoma, specific methylation profiles correlate with clinical outcomes11. In many cases, promoter elements are hypomethylated, while enhancers and coding sequences are hypermethylated, although the net consequences on transcription in vivo are not entirely predictable. Moreover, drugs that alter DNA methylation status with clear benefit in hematopoietic diseases have never found a role in sarcoma treatment, despite a long history of attempts29. This may be because of crosstalk between at least some layers of epigenetic regulation30, that put DNA methylation changes downstream of the proximal or driving genetic alteration in these other layers (Figure 1), such as Polycomb deletions in malignant nerve sheath tumors, SMARCB1 deletions in malignant rhabdoid tumors, and recently proposed phase transition effects in EWSR1 fusion-driven sarcomas31,32. How such genetic alterations mechanistically alter the DNA methylome is an area of active research, in which there have been recent advances in at least one major class of sarcomas, the central chondrosarcomas.

Central Chondrosarcoma

The discovery of oncogenic mutations in the isocitrate dehydrogenase (IDH) enzymes provided fundamental new insights into the mechanisms by which metabolites regulate epigenetic marks and how dysregulation of cellular metabolism can drive oncogenesis3335. Somatic mutations in cytosolic IDH1 and mitochondrial IDH2 are prevalent in acute myeloid leukemia (AML), glioma, cholangiocarcinoma, T cell lymphoma and several other tumor types36. Within sarcomas, IDH mutations are almost exclusively found in cartilaginous tumors, including 50–80% of central chondrosarcomas and in both their benign precursor (enchondromas) and in their advanced form (dedifferentiated chondrosarcomas)18,37. Somatic mosaic mutations in IDH1 or IDH2 underlie the pathogenesis of Ollier disease and Maffucci syndrome – non-hereditary diseases characterized by multiple enchondromas with increased risk of malignant transformation to chondrosarcoma38,392.

IDH enzymes normally function to catalyze the NADP(H)-dependent interconversion of isocitrate and alpha-ketoglutarate (aKG)36. Cancer-associated IDH mutations occur at specific arginine residues within the catalytic sites of IDH1 (R132) and IDH2 (R172 and R140). These mutations disrupt normal enzyme function, while simultaneously conferring a neomorphic enzymatic activity that enables efficient reduction of aKG to the ‘oncometabolite’ 2-hydroxyglutarate (2HG)40,41. Given its structural similarity to aKG, 2HG acts as a competitive inhibitor of a large family of aKG-dependent dioxygenases with diverse biologic functions, including regulation of DNA hydroxymethylation, RNA demethylation, histone demethylation, DNA repair, and prolyl hydroxylation of collagen and hypoxia-inducible factors36.

While the relative importance of different 2HG targets to oncogenesis remains an active area of investigation, 2HG-mediated inhibition of DNA and histone demethylation has been shown to play an important role in AML and glioma42. In particular, 2HG inhibits the TET DNA cytosine hydroxymethyltransferases and Jumonji histone demethylases resulting in increased methylation of DNA and histones, respectively4345. Disruption of this metabolic–epigenetic axis leads to a repressive chromatin state that impairs expression of genetic programs required for normal differentiation and locks IDH-mutant cancer cells into an undifferentiated state. In AML and glioma, small-molecule inhibitors of IDH-mutant enzymes can block 2HG production, reverse repressive chromatin states, and restore cellular differentiation4648. The mutant IDH1 inhibitor ivosidenib and the mutant IDH2 inhibitor enasidenib were recently approved for the treatment of AML49,50.

There have been more limited investigations into the mechanisms by which IDH mutations drive oncogenesis in chondrosarcoma. In mesenchymal stem/progenitor cells, mutant IDH-derived 2HG promotes hypermethylation of DNA CpG islands and increases repressive histone marks, resulting in impaired adipocyte and osteocyte differentiation in vitro5153. The precise effects of mutant IDH and 2HG on the chondrocyte lineage are less clear, with impaired chondrocyte differentiation observed in mouse models but enhanced chondrocyte differentiation observed in human mesenchymal stem cells in vitro5154. Introduction of mutant IDH2 R172K into a mouse mesenchymal progenitor cell line resulted in loss of contact inhibition in vitro and was sufficient to promote growth of dedifferentiated tumor xenografts in vivo53. In contrast, a mouse model with mutant IDH1 R132Q expression restricted to Col2a1-expressing chondrocytes exhibited multiple enchondroma-like lesions but no malignant progression to chondrosarcoma54. Notably, 2HG directly inhibits the aKG-dependent collagen prolyl 4-hydroxylase, resulting in impaired collagen maturation and defective extracellular matrix formation55,56; however, the importance of this 2HG-target in chondrosarcoma pathogenesis remains unknown.

In patients with chondrosarcoma, there has been no clear association between IDH-mutation status and clinical outcome, though a recent analysis of 89 chondrosarcoma cases, of which 41 were IDH1 or IDH2 mutated, showed an improved relapse-free and metastasis-free (though not overall) survival in subset of high grade chondrosarcoma bearing these mutations (22/47)57,58. The presence of IDH mutations in both primary and locally recurrent/metastatic lesions suggests that the IDH mutation may be an early event in chondrosarcoma pathogenesis59. The clinical phenotypes of Ollier disease and Maffucci syndrome indicate that IDH mutations may be sufficient for development of enchondromas but that additional genetic alterations may be required for progression to chondrosarcoma38,39. In clinical chondrosarcoma specimens, IDH-mutant tumors exhibit increased methylation of DNA CpG islands, but no clear association with histone methylation or DNA cytosine hydroxymethylation53,57. Unfortunately, treatment of IDH-mutant patient-derived chondrosarcoma cell lines with mutant-specific IDH inhibitors resulted in potent suppression of 2HG levels but did not reverse methylation of DNA/histones nor inhibit cell growth51.

Notably, a different enzyme in the same tricarboxylic acid cycle pathway as IDH, succinate dehydrogenase (SDH), is implicated in the pathogenesis another sarcoma histotype. SDH deficiency, either as part of the Carney-Stratakis syndrome or in sporadic cases, is responsible for a subset of gastrointestinal stromal tumors (GIST) that lack more common KIT or PDGFRA mutations60,61. Loss of SDH activity results in an accumulation of succinate, which, similarly to 2HG, inhibits the activity of histone demethylases and the TET enzymes62. The latter leads to marked differences in DNA methylation profiles between SDH-deficient and KIT-mutated GIST63. These changes in DNA-methylation in SDH-deficient GIST have recently been linked to downstream alterations in 3D chromatin structure through disruption of binding sites for the CTCF insulator64. Thus, there is an intriguing parallel between altered metabolism and epigenetic dysregulation in multiple rare sarcoma histotypes.

HISTONE MUTATIONS

Despite the relatively high frequency of driver mutations in chromatin modifying-enzymes in sarcomas and other cancers, it was only recently recognized that the fundamental substrates for this machinery, histone proteins, are also mutated in certain cancers (Figure 1). The first report of a histone driver mutation in cancer (i.e. an ‘oncohistone’) was in the setting of pediatric gliomas where they occur in two histone H3 variants, H3.3 or H3.1. Lysine to methionine mutations occur at lysine 27 in H3.1 or H3.3, and in H3.3 at glycine 34 with substitutions to arginine/valine65,66 (Figure 2).

Figure 2 |. Classical oncohistone mutations alter histone PTM ‘writer’ complex activity:

For Figure 2, please refer to Figure 2 of the manuscript Nacev BA, Jones KB, Intlekofer AM, Yu JSE, Allis CD, Tap WD, Ladanyi M, Nielsen TO. The epigenomics of sarcoma. Nat Rev Cancer. 2020 Oct;20(10):608–623. doi: 10.1038/s41568-020-0288-4. Epub 2020 Aug 11. PMID: 32782366.

Shortly after the discovery of oncohistones in pediatric gliomas, methionine substitutions at lysine 36 in H3.3 (H3F3B) were reported in > 90% of chondroblastomas, and missense mutations (nearly all resulting in tryptophan substitution) were observed at glycine 34 in the H3F3A gene in > 90% of giant cell tumors of bone67. In the same study, lower frequency mutations also occurred in osteosarcoma (H3.3 G34R) and in chondrosarcoma (H3.3 K36M). Thus, oncohistones function as possible drivers in multiple mesenchymal lineages. Subsequent studies confirmed the presence of H3.3 G34 missense mutations in osteosarcoma and identified K36M/I mutations in H3.1, an H3 variant closely related to H3.3, in pediatric undifferentiated sarcoma6870.

The initial discovery of oncohistones has led to several important questions including why two different H3.3 genes (which encode identical proteins) are differentially mutated depending on the affected amino acid and cancer type, why specific oncohistone mutations occur at high frequency in tumors of mesenchymal lineage, whether these or other mutations occur in other cancers, and how oncohistones affect the local and long-distance chromatin landscape. While some of these questions remain unanswered, significant progress has been made in addressing the last two.

Through the work of several labs, the ‘K-to-M’ class of histone mutations has been shown to function as inhibitors of the cognate methyltransferases that normally act on lysine as a substrate leading to reorganization of the chromatin landscape69,71,72. In the case of H3.3 K36M, inhibition of the H3K36 methyltransferases NSD2 and SETD2 leads to the loss of H3K36 di- and tri-methylation and induces a gain in H3K27 methylation, particularly in intergenic regions69. As a result, the repressive PRC1 complex, which engages H3K27 trimethylation (H3K27me3), is titrated away from genes that are normally repressed, and redistributed to intergenic regions that contain aberrantly localized H3K27me3. The ultimate consequence of this reorganization is the expression of genes that are normally silenced during cell fate commitment thereby promoting the persistence of an undifferentiated progenitor population despite appropriate in vitro differentiation signals69. Accordingly, expression of H3.3 or H3.1 K36M in mesenchymal progenitor cells is sufficient to generate sarcoma-like tumors in a murine allograft model69.

Interestingly, the ‘K-to-M’ paradigm commonly observed in oncohistone mutations has now been extended to the non-histone protein EZHIP. EZHIP contains a peptide sequence similar to the mutated region of the H3K27M oncohistone and is overexpressed in posterior fossa type A ependymomas where it inhibits PRC27377. Interestingly, EZHIP is also a fusion partner in a subset of endometrial stromal sarcomas, as discussed in the Fusions with Polycomb Complex Components section. Thus, the concept of ‘onco-histone mimicry’ broadens the mutational landscape underlying a variety of cancers.

In contrast, H3G34 mutations are less well understood. One notable difference is that mutations at H3G34 block H3K36 methylation exclusively in cis (i.e. on the same histone tail that harbors the mutations) while H3 K36M/I mutations also reduce H3K36 methylation in trans (i.e. on wildtype nucleosome tails)69,71. Inhibition of H3K36 methylation is hypothesized to result from impaired binding of the H3K36 substrate by SETD2, an H3K36 methyltransferase, due to H3G34 substitutions that introduce a bulky sidechain into a G33-G34 binding pocket near the active site of SETD278.

Despite differences in the cis/trans effects on oncohistone mutations at H3G34 and H3K36, both classes of oncohistone mutations have characteristics of oncogenic drivers. In which contexts effects in cis, which likely only affect local chromatin structure, versus effects in trans, which induce a global change in the chromatin landscape, are necessary to promote oncogenesis is an intriguing question that remains to be answered69,72,79.

Since the identification of these first (i.e. ‘classical’) oncohistones, several groups have recognized other non-classical histone mutations in rare cases involving multiple cancer types70,8083. These include AML, bladder cancer, uterine carcinosarcoma and mesenchymal tumors such as Ewing sarcoma, gastrointestinal stromal tumor, undifferentiated pleomorphic sarcoma, angiosarcoma, and desmoplastic small round cell tumor2. These non-classical histone mutations affect all four core histone families with frequent mutations in both the N-terminal tails, where the classical oncohistones mutants are located, and in the histone fold domains, which form the core of the nucleosome structure. The location of these novel oncohistone mutations is hypothesized to affect regulatory histone posttranslational modifications and/or nucleosome structure70. This may lead to downstream effects on chromatin-regulated processes including transcription and DNA repair. Additional work is ongoing to understand the potential function of these novel oncohistones in sarcomas and other tumor types.

HISTONE PTM ‘WRITER’ COMPLEXES

Malignant Peripheral Nerve Sheath Tumors

As discussed in the preceding section, the fundamental mechanism of two of the classical oncohistone mutations, H3.3 K27M and H3.3 K36M, is inhibition of histone PTM writer complexes. This paradigm extends from the histone proteins themselves to genetic alterations leading to loss of function in the enzyme complexes depositing the histone marks (Figure 3). Within sarcomas, malignant peripheral nerve sheath tumors (MPNST) are a classical example. The development of MPNST is associated with the germline syndrome of neurofibromatosis, which accounts for roughly half of cases84. The remainder of cases occur either sporadically or more rarely as a result of prior radiotherapy, but in all settings commonly harbor precise somatic mutations in NF1 (82% (18/22) in one series)85. Efforts to better define the molecular events that lead to the development of MPNST identified mutually exclusive loss of function alterations in two core components of Polycomb Repressor Complex 2 (PRC2): SUZ12 (which can be codeleted with NF1 by virtue of its adjacent location on 17q11.2) and EED85,86.

Figure 3 |. Genetic alterations in histone PTM ‘writer’ complex components are found in various sarcoma subtypes.

For Figure 3, please refer to Figure 3 of the manuscript Nacev BA, Jones KB, Intlekofer AM, Yu JSE, Allis CD, Tap WD, Ladanyi M, Nielsen TO. The epigenomics of sarcoma. Nat Rev Cancer. 2020 Oct;20(10):608–623. doi: 10.1038/s41568-020-0288-4. Epub 2020 Aug 11. PMID: 32782366.

PRC2 is responsible for depositing the repressive histone mark, H3K27me3, and has an important role in development and cell fate87. These genetic alterations in PRC2 components occur in the majority of NF1 associated, sporadic, and radiation-associated MPNST85. Notably, genetic alterations in a catalytic subunit of PRC2, EZH2, are observed in other cancers but not in MPNST88. As one would predict, SUZ12 or EED loss leads to suppressed PRC2 activity and decreased abundance of the PRC2-catalyzed histone mark, H3K27me3. Lack of H3K27me3 staining by immunohistochemistry is a biomarker that has been shown to be useful clinically in the pathologic evaluation of tumors suspected to be MPNST89. More recently, loss of H3K27 dimethylation has been suggested to be even more specific than H3K27me3 loss90. Because PRC2 is an important mediator of chromatin-regulated transcriptional repression, upregulation of gene expression is the most common transcriptional change observed in EED- and SUZ12-deficient MPNST compared to MPNST that are wildtype with respect to these genes85. Interestingly, the H3.3 K27M oncohistone mutation found in gliomas directly inhibits the EZH2 subunit of the PRC2 complex, suggesting that PRC2 loss of function may be a common mechanism for malignant transformation, but with important contextual or cell type-specific differences that are yet to be fully understood71.

Fusions with Polycomb Complex Components

In addition to the loss of function alterations in PRC2 components seen in MPNST, polycomb complexes can also be perturbed in mesenchymal tumors through fusion events involving polycomb genes (Table 1). For instance, a member of the PRC1.1 complex, BCOR, is fused in a subset of small blue round cell tumors with a variety of different partners (e.g. BCOR-CCNB3) with considerable overlap in transcriptional programs induced by the different fusions7,9193. The BCOR-PRC1.1 complex plays an important role in maintaining pluripotency in stem cell populations and suppresses mesodermal transcriptional programs; how sarcoma-associated BCOR fusions alter this function and contribute to sarcomagenesis is an ongoing area of research94,95.

BCOR fusions are also identified in a subset of high-grade endometrial stromal sarcoma (ESS), and the PRC2 component, SUZ12, is fused to the transcriptional repressor JAZF1 in low-grade ESS9698. The JAZF1-SUZ12 fusion reduces PRC2 activity and disrupts transcriptional repression99. In other ESS cases, JAZF1 is fused to PHF1, which targets the PRC2 complex to chromatin100,101. Finally, in yet another subset of ESS, the ‘oncohistone mimic’ EZHIP, which inhibits the catalytic activity of PRC2, is fused to a member of the NuA4 histone acetyltransferase complex, MBTD1102,103. This MBTD1-EZHIP fusion protein interacts with EZH2, the catalytic subunit of PRC2, and reduces the methyltransferase activity of the complex76. Taking the above together, we suggest that impaired polycomb function may represent a common mechanism for tumorigenesis in a large fraction of ESSs.

CHROMATIN REMODELING COMPLEXES

Synovial Sarcoma

Synovial sarcoma, characterized by the SS18-SSX fusion, represents a model disease for the involvement of chromatin remodeling complexes in fusion oncoprotein-mediated sarcomagenesis. In mammals, the SWI/SNF family of multiprotein complexes are critical effectors of chromatin remodelling104, existing in normal cells in at least three forms (Figure 4A): canonical BAF (BRG1/BRM Associated Factor, named after the ATPase components of the complex encoded respectively by the SMARCA4 and SMARCA2 genes), PBAF (Polybromo-associated BAF complex), and GBAF (GLTSCR1/1L-containing BAF, also known as non-canonical BAF or ncBAF). The amino terminal fusion partner, SS18, almost all of which is included in the chimeric oncoprotein, was identified as a frequent interactor with canonical BAF complexes shortly after its discovery105107. Middeljans et al. were the first to report (in 2012) that SS18 was a stable member of canonical BAF (but not PBAF) complexes, and indeed that the t(X;18)-derived chimeric fusion oncoprotein, SS18-SSX, associates with the members of the BAF complexes in which native SS18 participates108. The mechanism by which SS18-SSX alters BAF complex function has since been interrogated through a variety of methods, revealing altered and disrupted associations. SS18-SSX can serve a bridging function connecting ATF2 to the PRC2 complex member TLE1 and in doing so represses the expression of important tumor suppressor genes including CDKN2A and EGR1109111 (Figure 4B). Another group identified a relationship with β-catenin resulting in the activations of specific target genes112. Incorporation of SS18-SSX into canonical BAF complexes was noted by Kadoch and Crabtree to result in the ejection of a member of the complex known to be a tumor suppressor, SMARCB1 (a.k.a. BAF47, hSNF5, INI1)113 (Figure 4C). Although subsequent work has shown that synovial sarcoma-associated BAF complexes are quite distinct in their function from SMARCB1-absent BAF complexes114, this altered membership of canonical BAF complexes is nonetheless likely important to disease biology not only in synovial sarcoma, but also in other mesenchymal tumors where SMARCB1 loss is a defining diagnostic feature115, including the epithelioid variant of MPNST, poorly differentiated chordoma, malignant rhabdoid tumors and epithelioid sarcoma (discussed below).

Figure 4 |. Mechanisms of action of SS18-SSX in synovial sarcoma.

For Figure 4, please refer to Figure 4 of the manuscript Nacev BA, Jones KB, Intlekofer AM, Yu JSE, Allis CD, Tap WD, Ladanyi M, Nielsen TO. The epigenomics of sarcoma. Nat Rev Cancer. 2020 Oct;20(10):608–623. doi: 10.1038/s41568-020-0288-4. Epub 2020 Aug 11. PMID: 32782366.

Another subset of BAF complexes (that constitutively lack SMARCB1), termed GBAFs116 (Figure 4A), comprise a functional dependency for both synovial sarcoma cells and cells of other cancer types such as acute myelogenous leukemia and malignant rhabdoid tumors. The functional dependency on BRD9, a stable member of GBAFs, was first reported in the target drive database117, then followed up by multiple groups118120 (Figure 4C, bottom) (see Targeting Epigenetic Drivers in Sarcoma).

The distribution of BAFs across the epigenome has been reported to derive in part from an association of the SSX C-terminus with KDM2B and other members of the PRC1.1 complex121. As this complex normally functions to prime chromatin for subsequent transcriptional repression by PRC2, which BAF generally opposes, it is intriguing that the fusion repurposes PRC1.1 as a localizing mechanism for BAF-mediated PRC2 antagonism at many target genes, leading to their activation (Figure 4D).

Epithelioid Sarcoma and Malignant Rhabdoid Tumor

Deficiencies in SMARCB1 are characteristic of epithelioid sarcoma and malignant rhabdoid tumor, although there are notable differences between these entities. The age distribution for epithelioid sarcoma peaks in the adolescent-young adult population – younger than for most cancer patients but considerably older than for malignant rhabdoid tumor122. In terms of its primary site of disease, epithelioid sarcoma has an unusual predilection for superficial tissues in distal extremities; a less common “proximal” variant is based in the deep soft tissues of the limb girdle123. Rhabdoid cytomorphology is seen occasionally, particularly in proximal cases, but is not a diagnostic requirement124. Instead, epithelioid sarcoma typically shows a blend of spindled and epithelioid cells, all possessing enlarged nuclei with open chromatin. While the most reliable diagnostic marker in epithelioid sarcoma is loss of SMARCB1 protein expression in tumor nuclei125, both epithelial (keratin, epithelial membrane antigen) and mesenchymal (vimentin, CD34) biomarkers are concurrently expressed, consistent with a polyphenotypic state of differentiation126.

Whereas malignant rhabdoid tumor is characterized by bi-allelic deletion of SMARCB1 on a particularly quiet genomic background127, genome-wide studies of epithelioid sarcoma show a more prominent landscape of copy number alterations and a higher mutational burden (on par with e.g. glioblastoma multiforme), with SMARCB1 expression lost by several alternative mechanisms128. These include biallelic or monoallelic deletion of SMARCB1 (which may be heterogeneous in the cell population) and/or overexpression of inhibitory miRNAs128131. In transfected cell line models, SMARCB1 rescue leads to decreased proliferation and a reversal of polycomb-mediated repression, albeit not to as great an extent in epithelioid sarcoma as in malignant rhabdoid tumor cells132. In malignant rhabdoid tumors, recent work has shown the underlying mechanism of SMARCB1 loss to be associated with DNA methylation status and microenvironmental changes: homozygous large regional deletions of SMARCB1 correlate with global hypomethylation and prominent cytotoxic T cell infiltration and immune checkpoint expression133.

SWI/SNF complexes have roles not only in transcription, but also in DNA repair134, with cBAF involved in the DNA end resection step of homologous recombination repair135,136 and PBAF silencing transcription adjacent to double strand breaks137. Impaired SWI/SNF function can compromise DNA repair138, which may explain the higher rate of genomic, abnormalities in epithelioid sarcoma in the context of a clinically more protracted tumor evolution than is seen in malignant rhabdoid tumors. Further supporting this concept are initial results from correlative sequencing studies on an epithelioid sarcoma clinical trial (NCT02601950)139 which, while largely confirming prior genomic findings128, also identified mutations in genes mediating DNA double strand break repair in a subset of epithelioid sarcomas (unpublished conference abstract)140.

Rare sarcomas retain SMARCB1 but instead lose other members of the cBAF complex: SMARCA4, SMARCC1 or SMARCC2141. While SMARCA4 deficiency may be an infrequent alternative to SMARCB1 in epithelioid sarcoma, its loss via inactivating mutation is the defining feature of a recently-discovered entity termed SMARCA4-deficient thoracic sarcomatoid tumors, considered by some to be sarcomas142, while others have recently proposed that they represent a subset of smoking-associated, sarcomatoid lung carcinomas143. Usually presenting as mediastinal tumors in middle-aged males, these very aggressive cancers (median survival 6 months) can also show rhabdoid morphology, and typically lose expression not only of SMARCA4, but also of SMARCA2, while retaining SMARCB1143. A similar aggressive clinical course as well as molecular and histologic phenotype is also characteristic of small cell carcinoma of the ovary, hypercalcemic type144, and the even more recently-recognized entity of SMARCA4-deficient undifferentiated uterine sarcomas145,146. Perhaps surprisingly, although these BAF-deficient sarcomas (including synovial sarcoma) share some apparent commonalities – such as a propensity for epithelial-mesenchymal biphasic differentiation, and evidence for a dependency on GBAF complexes localizing to CTCF sites proximal to promoters – they have very distinct transcriptional profiles114. While evidence suggests these cancers share induced dependencies on GBAF117120, the involved loci and consequent patterns of induced gene expression appear to be highly cell context-dependent.

HIGHER ORDER CHROMATIN ORGANIZATION

Ewing Sarcoma

There has recently been substantial progress in our understanding of the small family of proteins that become the amino terminus partners in more than half of known sarcoma-associated fusion oncoproteins. These FET domain proteins, namely FUS (previously TLS), EWSR1, and TAF15, all of which can interact with the SWI/SNF chromatin remodeling complex and demonstrate some capacity for interchangeable roles in sarcoma fusions147, were initially poorly understood. By peptide sequence similarity, the three FET family members have an RNA binding domain, partly homologous to splicing factors, as well as a “transactivation” domain148. However, the target of the transactivating domain was unclear. Notably, from the native protein sequence only the transactivation domain is obligatorily included in the fusion oncoproteins149. An early observation was that this domain contains intrinsically disordered peptide sequences. When serine residues were experimentally substituted for the many tyrosines in these domains in EWSR1-FLI1, the transcriptional activation of target genes (including NKX2–2, PRKCB and EZH2) was abrogated31. Furthermore, both biochemical experiments with recombinant proteins or domains and experiments in cell lines showed FET proteins to be mediators of liquid-liquid phase separation (Figure 1E, panel 5; Box 1)150156. These phase-separated droplets have also been demonstrated to interact with RNA Polymerase II and to organize transcription hubs, with much higher transcriptional outputs than traditional transcription factors binding to singular binding sites in promoters32,157,158.

BOX 1: Liquid-liquid phase separation.

Phase separation is a physical process for liquid-liquid demixing by which a supersaturated solution of components spontaneously separates into two distinct but stable phases, a high-concentration phase and a low-concentration phase192194. An everyday example of would be the demixing, or phase separation, of immiscible fluids such as oil and water.

While phase separation is well-known in polymer science195, the concept of phase separation as a possible mechanism for membraneless compartmentalization and spatiotemporal regulation of biological reactions is a recent development in biology192194,196. These membraneless compartments, recently termed biomolecular condensates193,194, are highly diverse in molecular composition, subcellular localization and functions, and include subnuclear bodies such as nucleoli, Cajal bodies, PML nuclear bodies and nuclear speckles197, as well as cytoplasmic structures such as stress granules and P-bodies198.

Formation of biomolecular condensates is driven by multivalent protein-protein or protein-RNA interactions199,200 involving two major classes of proteins that can phase separate under physiological conditions. The first class of proteins contain multiple folded domains or modules that frequently interact with linear motifs of other proteins, with increasing number of modules conferring higher propensity to phase separate, and is exemplified by the clustering of signaling molecules to facilitate signal transduction199,201,202.

The second class of proteins contain intrinsically disordered regions which lack a defined 3-dimensional structure and are typically enriched in low complexity domains – repeat sequences of a limited number of amino acid residues that drive phase separation, such as asparagine, glycine, glutamine, serine, arginine, lysine, aspartic acid, glutamic acid, phenylalanine and tyrosine203205. For example, the ~30 members of the FUS family of proteins (including the sarcoma-associated FET proteins FUS, EWSR1 and TAF15) share similar domain structures206, and undergo phase separation primarily by interactions between arginine and tyrosine residues200.

There is increasing evidence that transcription factors mechanistically activate genes through phase separation207, and that aberrant formation of biomolecular condensates is associated with neurodegenerative diseases and cancer194,208. Nevertheless, in the rapidly emerging field of liquid-liquid phase separation biology, it should be noted that while many proteins have been shown in elegant studies to phase separate in vitro, extrapolating or experimentally determining their in vivo functional consequences has proven far more challenging209,210.

The ETS family of transcription factors that provide the C-terminal partners in Ewing sarcoma fusion oncoproteins contribute to the phase separation of chromatin by their binding (more significantly than to traditional singular ETS consensus binding sequences in promoters) to multimeric sites present in microsatellite regions with many GGAA repeats159. The multimeric associations between multiple contiguous binding sites on chromatin with the low-energy but multiplied prion-like domain interactions of EWSR1-ETS fusions generate novel superenhancers in these microsatellite regions, often regulating the transcription of genes located at a great linear distance via higher order chromatin looping31,32,160. When the EWSR1-FLI fusion oncoprotein is depleted in Ewing sarcoma cells, changes in the epigenome at promoters, enhancers, and super-enhancers imply an important role for the fusion in regulating the epigenetic landscape161.

These transcription hubs and target genes associated by three-dimensional chromatin structure have implications for nucleosome distribution and specific histone marks, as well as with the chromatin remodeling complexes that enzymatically drive these changes in epigenomic structure and function. SWI/SNF or BAF complexes are involved in the ejection of nucleosomes from enhancer elements and have been found to associate with both these native FET protein transcription hubs as well as with sarcomagenic fusion oncoproteins31,147.

Interestingly, the binding pattern of BAF complexes that incorporate SS18-SSX in synovial sarcoma suggests that these may also participate in some types of phase-separated transcription hubs, because the BAF enrichment patterns on ChIPseq broaden from the typical promoters and enhancers114, to include more of the gene bodies of target loci.

CLINICAL TRANSLATION: TARGETING EPIGENETIC DRIVERS IN SARCOMA

Although not the focus of this review, it is worth noting that insights into sarcoma biology from epigenetics-driven research have rapidly impacted clinical diagnostics. Histone mutation-specific antibodies exist: H3FA G34W immunohistochemistry has value in the diagnosis of difficult or malignant cases of giant cell tumor of bone162. As mentioned above, nuclei demonstrating loss of di- or tri- methylated H3K27 help define MPNST89,90, while loss of SMARCB1 or SMARCA4 can seal the diagnosis of the epithelioid variant of MPNST, epithelioid sarcoma, chordoma, malignant rhabdoid tumor and related thoracic and gynecologic malignancies (all entities that were especially problematic for pathologists before these tools became available)115. Mutation-specific PCR has value to identify IDH1/2 mutations in dedifferentiated chondrosarcoma163, and karyotyping, FISH or RT-PCR have historically aided the diagnosis of fusion oncogene sarcomas (including not only Ewing sarcoma, but a growing list of additional entities bearing EWSR1, FUS or other translocations)164. Such methods are now being replaced by more modern multiplexed hybrid capture sequencing, anchored PCR or color coded probe pair technologies to identify most fusion events with a single assay165167. Finally, DNA methylation arrays are emerging as a pan sarcoma diagnostic tool based on the recognition of sarcoma subtype-specific methylomes23,24. “Liquid biopsy” approaches are in advanced states of translational research, but have yet to enter routine clinical practice for epigenetically-driven sarcomas168.

The identification of the genetic events that lead to downstream epigenetic alterations in some sarcomas has opened the door for targeted therapeutic interventions. This has been made possible by 1) an understanding of the detailed nature of the chromatin alterations that occur in the setting of the genetic defect, and 2) the development of second-generation chromatin targeting drugs. The latter has moved the field beyond early HDAC and DNA methyltransferase inhibitors; emerging classes of small molecule epigenetic drugs hold promise for manipulating the activity of chromatin-modifying enzymes with a level of specificity not previously feasible169,170.

Despite this progress, the majority of chromatin pathway alterations in sarcoma are not sufficiently understood to develop targeted therapeutic approaches, although there are several notable exceptions (Table 1). For example, epithelioid sarcoma and malignant rhabdoid tumor, as discussed above, are characterized by loss of the SWI/SNF (BAF) remodeling complex subunit SMARCB1, which interacts with a regulatory interface of the nucleosome (the acidic patch) through its C-terminal domain171173. Loss of SMARCB1 induces dependency on the PRC2 methyltransferase complex which has subsequently been targeted via inhibition of its catalytic subunit EZH2174,175. Similarly, SMARCA4-deficient neoplasms show evidence for unopposed Polycomb activity with resulting sensitivity to EZH2 inhibitors176. A phase 2 study (NCT02601950)139 of the EZH2 inhibitor tazemetostat included advanced/metastatic epithelioid sarcoma patients; initial results (disease control in 10 / 31 patients) exceeded prespecified criteria for success and led to a doubling of the cohort size. As existing sarcoma drugs (cytotoxic chemotherapy and tyrosine kinase inhibitors) confer minimal benefit in epithelioid sarcoma, these phase 2 findings led to approval by the FDA in 2020. Updated results (unpublished conference abstract)177 on the final cohort of 62 patients showed disease control in 16 (26%) and objective responses in 9 (15%), which is better than what had been achieved in epithelioid sarcoma with existing chemotherapy regimens. Interestingly, EZH2 inhibition has not been effective clinically in synovial sarcoma (unpublished conference abstract)178, perhaps owing to the fact that the SSX-SS18 fusion protein forms a BAF complex with altered composition and aberrant activity instead of loss of function113,121.

Two additional examples of potential mechanism-based targeting of chromatin pathway alterations in sarcoma include IDH-mutated chondrosarcoma, and targetable dependencies of the (as yet untargetable) SS18-SSX fusion oncoprotein in synovial sarcoma. Inhibitors of IDH1 (ivosidenib) and IDH2 (enasidenib) have been developed in the setting of IDH-mutant acute myeloid leukemia and are now being evaluated in clinical trials (e.g. NCT02273739, NCT02073994)179,180 for the treatment of the subset of chondrosarcomas that harbor IDH1 or 2 mutations181. A recent a Phase I study of IDH1 inhibition in IDH1-mutant chondrosarcomas suggest a potential but modest benefit in the conventional, but not dedifferentiated, subtype182. Two hypotheses for this subtype-specific difference are that the dedifferentiation leads to IDH1-independent growth or that the changes to chromatin environment are ‘locked-in’ and cannot be reversed by a decrease in the oncometabolite. The final results of this and other clinical trials183 will be informative in evaluating the efficacy of this approach.

In synovial sarcoma and malignant rhabdoid tumors, recent preclinical data demonstrate a possible role for targeting the bromodomain containing protein, BRD9, a GBAF subunit as described above (Figure 4)119,184. Loss of SMARCB1 leads to compensatory use of alternative SWI/SNF (BAF) complexes, such as those incorporating SMARC paralogs and BRD9, with associated widespread changes in gene expression (including those controlling stem cell differentiation) and CTCF-regulated higher-order chromatin structure120,132. BRD9 can be targeted with emerging agents119, although those most effective in vitro (chemical degraders) have delivery and toxicity issues that are necessitating further pharmacological development before they can be used clinically. Additional work is needed to validate BRD9 as a target, and to develop compounds that can safely target this protein clinically.

Future successful therapeutic strategies in sarcomas with chromatin driver mutations will likely result from targeting either induced functional dependencies or through a reversal of one or more elements of the pathologically altered chromatin landscape. A hypothetical example of the latter is to leverage the reciprocal nature of multiple posttranslational modifications (PTMs), which have different functions but occur at the same residue. The aberrant gain in one PTM because of a loss of the reciprocal mark due to a genetic lesion in the ‘writer’ complex could be targeted by inhibiting the ‘writer’ or ‘reader’ of the inappropriately gained mark. One setting for exploring this approach is in those subsets of MPNST and endometrial stromal sarcoma which have mutations in the H3K27me3 writer complex, PRC2, since H3K27me3 and H3K27ac are reciprocal and have transcriptionally repressive and activating functions, respectively113,119,120,185. For instance, inhibition of the H3K27 acetyltransferase, p300, or H3K27ac reader domain containing proteins such as BRD4 could be potential strategies in PRC2 deficient sarcomas. Combining inhibition of the BET family of histone PTM readers and targeted degradation of one BET family member, BRD4, has also been proposed as a synthetic lethal strategy in MPNST186.

One challenge in designing targeted therapies for sarcomas with chromatin pathway drivers is that chromatin regulators have diverse and incompletely understood functions, which complicates efforts to understand their pathological dysregulation let alone the activity of small molecules targeting them. Pursuing therapies in this setting may be more complex than studying inhibitors of signaling pathway drivers where at least the proximal downstream events are better characterized. This challenge is particularly pronounced in the setting of sarcomas with complex genomics such as undifferentiated pleomorphic sarcoma, where a subset harbor heterogeneous chromatin pathway alterations without a defined driver11. The first step to developing epigenetic therapeutics in these diseases is to identify which chromatin pathway genetic alterations affect the epigenome. Further investigations into which of these alterations affect chromatin regulated processes such as transcription, control of antitumor immunity, differentiation, and DNA damage repair will be needed. Having thus defined ‘functional’ alterations, selection and pre-clinical testing of targeted therapies can follow paradigms similar to sarcomas with clearly defined epigenetic drivers.

Another potential challenge is that chromatin changes resulting from mutations in chromatin regulators may be relatively difficult to reverse, which could explain the modest response rates of EZH2 inhibition in epithelioid sarcoma and IDH1 inhibition in chondrosarcoma. Combinations of chromatin targeting drugs may be needed to ultimately reverse the effects of chromatin dysregulation in these sarcomas.

That said, we are optimistic that in depth and rigorous investigations of both normal and aberrant chromatin-modifying enzyme complexes will continue to inform efforts to therapeutically intervene in epigenetically-driven sarcomas. We also suggest that because of this complexity, therapeutically targeting tumors driven by derangements of chromatin regulation may represent an ideal opportunity to revisit the strategy of phenotypic drug discovery, which has fallen out of favor in the era of great success in targeted therapies but may be one valuable approach to in addition to the target-based strategies described above.

CONCLUSIONS

As this review demonstrates, there has been a remarkable evolution in our appreciation of the key role of epigenetic dysregulation in many if not most sarcomas. It is humbling to recall that, historically, targeted sequencing approaches focusing on proliferation and apoptosis-related “cancer genes” in sarcomas (and gliomas) failed to include many genes involved in epigenetic control9,187 and thus, for instance, IDH1 mutations were therefore instead first identified by a whole exome approach in gliomas33. Even to this day, the conventional precision oncology approach of matching kinase inhibitors to genetic alterations activating signaling through the MAPK pathway has been largely disappointing in sarcomas (with a few remarkable exceptions such as gastrointestinal stromal tumor188, and NTRK fusion-driven sarcomas189191). Thus, further clinical progress in targeting epigenetic dysregulation in sarcomas will depend on expanded clinical genomic testing that includes genes involved in epigenetic pathways as well as robust profiling of DNA methylation and histone modifications carefully paired with new agents that can specifically target these aberrant epigenetic states.

ACKNOWLEDGEMENT

The authors would like to thank Drs. Peter Stirling, T. Michael Underhill, and Martin Hirst, Brendan Dickson and Elizabeth Demicco for helpful discussions about some of the topics covered in this review, and Drs. Alexey Soshnev, Martin Hirst and Ms. Linda Wang for assistance in figure preparation.

GLOSSARY

Chondrosarcomas

A malignant cartilaginous matrix-producing tumor often driven by IDH1/2 mutations. Typically arising in the axial skeleton of middle-aged patients, these sarcomas can be slow growing but resistant to existing systemic therapy and radiotherapy.

Nucleosomes

Basic repeating structural units of the chromosome consisting of eight histone proteins (2 each of four core histones, H3, H4, H2A, and H2B) and 147 base pairs of DNA wrapping the structure.

Giant cell tumor of bone

A benign but often locally aggressive neoplasm of bone in young adults driven by missense histone mutations at H3.3G34. These tumors have a propensity for local recurrence and present as destructive, radiolytic lesions that destroy bone underneath articular surfaces.

Chondroblastoma

A benign cartilaginous neoplasm that characteristically arises at the ends of the body’s long bones, close to the joints, and occurs predominantly in adolescents. Chondroblastoma is driven by the H3.3K36M histone mutation.

Malignant peripheral nerve sheath tumors

Sarcomas arising within peripheral nerves, about half of which are sporadic and half which occur in the context of congenital NF1 inactivating mutations (neurofibromatosis type 1). Development of malignant peripheral nerve sheath tumors is additionally driven by (mutually exclusive) loss of function alterations in SUZ12 or EED.

Synovial sarcoma

A malignant translocation-associated sarcoma driven by the SS18-SSX gene fusion. Although frequently arising in extremities near joints, the term is a misnomer as the cell of origin is unknown and the tumor is not derived from synovium, nor does it differentiate into synovial-type tissue. Exists in monophasic spindle cell forms or as a biphasic type with areas of epithelial differentiation.

Malignant rhabdoid tumors

Highly aggressive, malignant tumors that occur in infants and young children. Three presentations exist: kidney, extrarenal, and brain (termed atypical teratoid/rhabdoid tumor); all are characterized by bi-allelic deletion of SMARCB1.

Epithelioid sarcoma

Malignant soft tissue sarcoma in distal extremities, with mixed features of mesenchymal and epithelial differentiation. Typically affects adolescents and young adults, metastasizes aggressively and is resistant to conventional chemotherapies. Characterized by inactivating mutations in SMARCB1.

Chromatin remodeling complexes

Multiple families of protein complexes that alter chromatin structure to regulate gene expression. Their functions include alteration of nucleosome assembly (maturation and spacing), chromatin access (nucleosome repositioning or ejection), and nucleosome editing (histone exchange or eviction).

Phase separation

A physical process in which a single homogenous liquid phase spontaneously separates into two distinct phases due to changes in environment such as pH, temperature, salt and protein concentration.

Bisulfite sequencing

Current gold standard sequencing strategy for detecting DNA methylation based on the conversion of unmethylated cytosine to uracil after treatment with sodium bisulfite (without modification of methylated cytosine).

CpG islands

Segments of genomic DNA, several hundred base pairs in length, that contain a large number of CpG dinucleotide repeats. When occurring near promoters of expressed genes, CpG islands are usually unmethylated; in contrast, CpG dinucleotides occurring in other contexts tend to be methylated.

Ewing sarcoma

A malignant bone or soft tissue tumor comprised of uniform small blue round cells, typically affecting children and adolescents. Driven by chromosomal translocations resulting in transcripts fusing FET (FUS/EWSR1/TAF15) genes with ETS family transcription factors, EWSR1-FLI1 is the most common variant.

Enhancers

Gene regulatory elements that bind transcription factors and cofactors to activate transcription of target genes that may be located a relatively great linear distance away, and independently of their orientation on DNA.

Polycomb

A group of proteins originally discovered in Drosophila involved in establishment and maintenance of developmental gene expression programs through formation of PRC complexes that repress gene expression by methylation of histone H3K27 (PRC2) and ubiquitination of H2A119 (PRC1).

Gastrointestinal stromal tumor

Mesenchymal neoplasm of the gastrointestinal tract, derived from the interstitial cells of Cajal. Activating mutations in the KIT (or PDGFRA) receptor tyrosine kinases are the key initiating oncogenic events in the majority of cases, making imatinib and related tyrosine kinase inhibitors an effective targeted therapy for this disease.

Undifferentiated pleomorphic sarcoma

A malignant mesenchymal tumor of undefined histogenesis, histologically characterized high grade spindle cells producing a nonspecific collagenous matrix. Previously termed malignant fibrous histiocytoma and considered a diagnosis of exclusion.

Angiosarcoma

An aggressive, malignant endothelial cell tumor of vascular or lymphatic origin that can arise anywhere in the body, sporadically or sometimes in association with radiation exposure or lymphedema. Angiosarcomas are especially infiltrative and prone to metastatic spread.

Desmoplastic small round cell tumor

An aggressive, malignant neoplasm that typically presents as a large mass in the abdomen of adolescent and young males. Characterized by a translocation resulting in EWSR1-WT1 fusion transcripts, this sarcoma does not respond well to any currently available systemic therapies.

Endometrial stromal sarcoma

A type of uterine malignancy with low and high grade forms that are associated with distinct genetic rearrangements and fusion oncogenes. Typically presenting in middle age, the disease is relatively slower to progress than most other types of sarcoma.

Sarcomatoid tumors

A descriptive term for neoplasms of non-mesenchymal origin that develop a sarcoma-like histologic phenotype (characterized by spindle cell cytomorphology, matrix production and cell-matrix interactions) – for example, carcinomas that have undergone epithelial-mesenchymal transition.

Mediastinal tumors

A term for primary neoplasms of the thoracic cavity, other than lung cancers.

Superenhancers

Clusters of enhancers in close genomic proximity with high concentrations of bound transcriptional co-activators that control expression programs to regulate cell identity.

Chromatin looping

A model for long-range control of gene expression to allow for direct contact of promoters and enhancers over long linear distances by looping out the intervening chromatin. Loops are mediated and stabilized by proteins and complexes including CCCTC-binding factor (CTCF), Mediator, and Cohesin.

Chemical degrader

A class of compounds that bind a target protein through one chemical domain and through a second domain recruit the cereblon E3 ubiquitin ligase complex leading to degradation of the target protein. Degradation can have a distinct biologic effect from small molecule inhibition of the target in cases where the target protein has non-enzymatic functions.

REFERENCES

  • 1.Cancer Genome Atlas Research Network et al. Genomic and epigenomic landscapes of adult de novo acute myeloid leukemia. N. Engl. J. Med 368, 2059–2074 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Papaemmanuil E et al. Genomic Classification and Prognosis in Acute Myeloid Leukemia. N. Engl. J. Med 374, 2209–2221 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Lohr JG et al. Discovery and prioritization of somatic mutations in diffuse large B-cell lymphoma (DLBCL) by whole-exome sequencing. Proc. Natl. Acad. Sci. U.S.A 109, 3879–3884 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Okosun J et al. Integrated genomic analysis identifies recurrent mutations and evolution patterns driving the initiation and progression of follicular lymphoma. Nat. Genet 46, 176–181 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Siegel RL, Miller KD & Jemal A Cancer statistics, 2018. CA Cancer J Clin 68, 7–30 (2018). [DOI] [PubMed] [Google Scholar]
  • 6.Mehren, von M et al. Soft Tissue Sarcoma, Version 2.2016, NCCN Clinical Practice Guidelines in Oncology. Journal of the National Comprehensive Cancer Network : JNCCN 14, 758–786 (2016). [DOI] [PubMed] [Google Scholar]
  • 7.Schaefer I-M, Cote GM & Hornick JL Contemporary Sarcoma Diagnosis, Genetics, and Genomics. J Clin Oncol 36, 101–110 (2018). [DOI] [PubMed] [Google Scholar]
  • 8.Jain S, Xu R, Prieto VG & Lee P Molecular classification of soft tissue sarcomas and its clinical applications. Int J Clin Exp Pathol 3, 416–428 (2010). [PMC free article] [PubMed] [Google Scholar]
  • 9.Barretina J et al. Subtype-specific genomic alterations define new targets for soft-tissue sarcoma therapy. Nat. Genet 42, 715–721 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Vlenterie M et al. Next generation sequencing in synovial sarcoma reveals novel gene mutations. Oncotarget 6, 34680–34690 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Cancer Genome Atlas Research Network. Comprehensive and Integrated Genomic Characterization of Adult Soft Tissue Sarcomas. 171, 950–965.e28 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Brohl AS et al. The genomic landscape of the Ewing Sarcoma family of tumors reveals recurrent STAG2 mutation. PLoS Genet. 10, e1004475 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Crompton BD et al. The genomic landscape of pediatric Ewing sarcoma. Cancer Discov 4, 1326–1341 (2014). [DOI] [PubMed] [Google Scholar]
  • 14.Tirode F et al. Genomic landscape of Ewing sarcoma defines an aggressive subtype with co-association of STAG2 and TP53 mutations. Cancer Discov 4, 1342–1353 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Shern JF et al. Comprehensive genomic analysis of rhabdomyosarcoma reveals a landscape of alterations affecting a common genetic axis in fusion-positive and fusion-negative tumors. Cancer Discov 4, 216–231 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Chalmers ZR et al. Analysis of 100,000 human cancer genomes reveals the landscape of tumor mutational burden. Genome Med 9, 34 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Joseph CG et al. Exomic analysis of myxoid liposarcomas, synovial sarcomas, and osteosarcomas. Genes Chromosomes Cancer 53, 15–24 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Amary MF et al. IDH1 and IDH2 mutations are frequent events in central chondrosarcoma and central and periosteal chondromas but not in other mesenchymal tumours. J. Pathol 224, 334–343 (2011). [DOI] [PubMed] [Google Scholar]
  • 19.Venneker S, Szuhai K, Hogendoorn PCW & Bovée JVMG Mutation-driven epigenetic alterations as a defining hallmark of central cartilaginous tumours, giant cell tumour of bone and chondroblastoma. Virchows Arch. path Anat 476, 135–146 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Ho PJ, Lloyd SM & Bao X Unwinding chromatin at the right places: how BAF is targeted to specific genomic locations during development. Development 146, dev178780 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Sawyer IA, Bartek J & Dundr M Phase separated microenvironments inside the cell nucleus are linked to disease and regulate epigenetic state, transcription and RNA processing. Semin. Cell Dev. Biol 90, 94–103 (2019). [DOI] [PubMed] [Google Scholar]
  • 22.Sheffield NC et al. DNA methylation heterogeneity defines a disease spectrum in Ewing sarcoma. Nat. Med 23, 386–395 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Wu SP et al. DNA Methylation-Based Classifier for Accurate Molecular Diagnosis of Bone Sarcomas. JCO Precis Oncol 2017, 1–11 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Koelsche C et al. Array-based DNA-methylation profiling in sarcomas with small blue round cell histology provides valuable diagnostic information. Mod. Pathol 31, 1246–1256 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Koelsche C et al. DNA methylation profiling distinguishes Ewing-like sarcoma with EWSR1-NFATc2 fusion from Ewing sarcoma. J. Cancer Res. Clin. Oncol 145, 1273–1281 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Seki M et al. Integrated genetic and epigenetic analysis defines novel molecular subgroups in rhabdomyosarcoma. Nat Commun 6, 7557 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Röhrich M et al. Methylation-based classification of benign and malignant peripheral nerve sheath tumors. Acta Neuropathol 131, 877–887 (2016). [DOI] [PubMed] [Google Scholar]
  • 28.Johann PD et al. Atypical Teratoid/Rhabdoid Tumors Are Comprised of Three Epigenetic Subgroups with Distinct Enhancer Landscapes. Cancer Cell 29, 379–393 (2016). [DOI] [PubMed] [Google Scholar]
  • 29.Srinivasan U, Reaman GH, Poplack DG, Glaubiger DL & LeVine AS Phase II study of 5-azacytidine in sarcomas of bone. Am. J. Clin. Oncol 5, 411–415 (1982). [DOI] [PubMed] [Google Scholar]
  • 30.Viré E et al. The Polycomb group protein EZH2 directly controls DNA methylation. Nature 439, 871–874 (2006). [DOI] [PubMed] [Google Scholar]
  • 31.Boulay G et al. Cancer-Specific Retargeting of BAF Complexes by a Prion-like Domain. Cell 171, 163–178.e19 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Chong S et al. Imaging dynamic and selective low-complexity domain interactions that control gene transcription. Science 361, eaar2555 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Parsons DW et al. An integrated genomic analysis of human glioblastoma multiforme. Science 321, 1807–1812 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Mardis ER et al. Recurring mutations found by sequencing an acute myeloid leukemia genome. N. Engl. J. Med 361, 1058–1066 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Lu C & Thompson CB Metabolic regulation of epigenetics. Cell Metab. 16, 9–17 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Losman J-A & Kaelin WG What a difference a hydroxyl makes: mutant IDH, (R)-2-hydroxyglutarate, and cancer. Genes Dev. 27, 836–852 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Tarpey PS et al. Frequent mutation of the major cartilage collagen gene COL2A1 in chondrosarcoma. Nat. Genet 45, 923–926 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Pansuriya TC et al. Somatic mosaic IDH1 and IDH2 mutations are associated with enchondroma and spindle cell hemangioma in Ollier disease and Maffucci syndrome. Nat. Genet 43, 1256–1261 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Amary MF et al. Ollier disease and Maffucci syndrome are caused by somatic mosaic mutations of IDH1 and IDH2. Nat. Genet 43, 1262–1265 (2011). [DOI] [PubMed] [Google Scholar]
  • 40.Dang L et al. Cancer-associated IDH1 mutations produce 2-hydroxyglutarate. Nature 462, 739–744 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Ward PS et al. 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 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.McCarthy N Metabolism: unmasking an oncometabolite. Nature Reviews Cancer 12, 229–229 (2012). [DOI] [PubMed] [Google Scholar]
  • 43.Figueroa ME et al. Leukemic IDH1 and IDH2 mutations result in a hypermethylation phenotype, disrupt TET2 function, and impair hematopoietic differentiation. Cancer Cell 18, 553–567 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Lu C et al. IDH mutation impairs histone demethylation and results in a block to cell differentiation. Nature 483, 474–478 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Turcan S et al. IDH1 mutation is sufficient to establish the glioma hypermethylator phenotype. Nature 483, 479–483 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Yen K et al. AG-221, a First-in-Class Therapy Targeting Acute Myeloid Leukemia Harboring Oncogenic IDH2 Mutations. Cancer Discov 7, 478–493 (2017). [DOI] [PubMed] [Google Scholar]
  • 47.Wang F et al. Targeted inhibition of mutant IDH2 in leukemia cells induces cellular differentiation. Science 340, 622–626 (2013). [DOI] [PubMed] [Google Scholar]
  • 48.Rohle D et al. An inhibitor of mutant IDH1 delays growth and promotes differentiation of glioma cells. Science 340, 626–630 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.DiNardo CD et al. Durable Remissions with Ivosidenib in IDH1-Mutated Relapsed or Refractory AML. N. Engl. J. Med 378, 2386–2398 (2018). [DOI] [PubMed] [Google Scholar]
  • 50.Stein EM et al. Enasidenib in mutant IDH2 relapsed or refractory acute myeloid leukemia. Blood 130, 722–731 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Suijker J et al. Inhibition of mutant IDH1 decreases D-2-HG levels without affecting tumorigenic properties of chondrosarcoma cell lines. Oncotarget 6, 12505–12519 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Jin Y et al. Mutant IDH1 Dysregulates the Differentiation of Mesenchymal Stem Cells in Association with Gene-Specific Histone Modifications to Cartilage- and Bone-Related Genes. PLoS ONE 10, e0131998 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Lu C et al. Induction of sarcomas by mutant IDH2. Genes Dev. 27, 1986–1998 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Hirata M et al. Mutant IDH is sufficient to initiate enchondromatosis in mice. Proc. Natl. Acad. Sci. U.S.A 112, 2829–2834 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Sasaki M et al. D-2-hydroxyglutarate produced by mutant IDH1 perturbs collagen maturation and basement membrane function. Genes Dev. 26, 2038–2049 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Koivunen P et al. Transformation by the (R)-enantiomer of 2-hydroxyglutarate linked to EGLN activation. Nature 483, 484–488 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Cleven AHG et al. IDH1 or −2 mutations do not predict outcome and do not cause loss of 5-hydroxymethylcytosine or altered histone modifications in central chondrosarcomas. Clin Sarcoma Res 7, 8–10 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Zhu GG et al. Genomic Profiling Identifies Association of IDH1/IDH2 Mutation with Longer Relapse-Free and Metastasis-Free Survival in High-Grade Chondrosarcoma. Clin. Cancer Res 26, 419–427 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Amary MF et al. Isocitrate dehydrogenase 1 mutations (IDH1) and p16/CDKN2A copy number change in conventional chondrosarcomas. Virchows Arch. path Anat 466, 217–222 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Janeway KA et al. Defects in succinate dehydrogenase in gastrointestinal stromal tumors lacking KIT and PDGFRA mutations. Proc. Natl. Acad. Sci. U.S.A 108, 314–318 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Pantaleo MA et al. SDHA loss-of-function mutations in KIT-PDGFRA wild-type gastrointestinal stromal tumors identified by massively parallel sequencing. J. Natl. Cancer Inst 103, 983–987 (2011). [DOI] [PubMed] [Google Scholar]
  • 62.Xiao M et al. Inhibition of α-KG-dependent histone and DNA demethylases by fumarate and succinate that are accumulated in mutations of FH and SDH tumor suppressors. Genes Dev. 26, 1326–1338 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Killian JK et al. Succinate dehydrogenase mutation underlies global epigenomic divergence in gastrointestinal stromal tumor. Cancer Discov 3, 648–657 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Flavahan WA et al. Altered chromosomal topology drives oncogenic programs in SDH-deficient GISTs. Nature 575, 229–233 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Schwartzentruber J et al. Driver mutations in histone H3.3 and chromatin remodelling genes in paediatric glioblastoma. Nature 482, 226–231 (2012). [DOI] [PubMed] [Google Scholar]
  • 66.Wu G et al. Somatic histone H3 alterations in pediatric diffuse intrinsic pontine gliomas and non-brainstem glioblastomas. Nat. Genet 44, 251–253 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Behjati S et al. Distinct H3F3A and H3F3B driver mutations define chondroblastoma and giant cell tumor of bone. Nat. Genet 45, 1479–1482 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Koelsche C et al. Histone 3.3 hotspot mutations in conventional osteosarcomas: a comprehensive clinical and molecular characterization of six H3F3A mutated cases. Clin Sarcoma Res 7, 9–11 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Lu C et al. Histone H3K36 mutations promote sarcomagenesis through altered histone methylation landscape. Science 352, 844–849 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Nacev BA et al. The expanding landscape of ‘oncohistone’ mutations in human cancers. Nature 567, 473–478 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Lewis PW et al. Inhibition of PRC2 activity by a gain-of-function H3 mutation found in pediatric glioblastoma. Science 340, 857–861 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Fang D et al. The histone H3.3K36M mutation reprograms the epigenome of chondroblastomas. Science 352, 1344–1348 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Hübner J-M et al. 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 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Jain SU et al. PFA ependymoma-associated protein EZHIP inhibits PRC2 activity through a H3 K27M-like mechanism. Nat Commun 10, 2146–14 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Pajtler KW et al. Molecular heterogeneity and CXorf67 alterations in posterior fossa group A (PFA) ependymomas. Acta Neuropathol 136, 211–226 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Piunti A et al. CATACOMB: An endogenous inducible gene that antagonizes H3K27 methylation activity of Polycomb repressive complex 2 via an H3K27M-like mechanism. Sci Adv 5, eaax2887 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Ragazzini R et al. EZHIP constrains Polycomb Repressive Complex 2 activity in germ cells. Nat Commun 10, 3858–18 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Fang J et al. Cancer-driving H3G34V/R/D mutations block H3K36 methylation and H3K36me3-MutSα interaction. Proc. Natl. Acad. Sci. U.S.A 115, 9598–9603 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Shi L, Shi J, Shi X, Li W & Wen H Histone H3.3 G34 Mutations Alter Histone H3K36 and H3K27 Methylation In Cis. J. Mol. Biol 430, 1562–1565 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Zhao S et al. Mutational landscape of uterine and ovarian carcinosarcomas implicates histone genes in epithelial-mesenchymal transition. Proc. Natl. Acad. Sci. U.S.A 113, 12238–12243 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Arimura Y et al. Cancer-associated mutations of histones H2B, H3.1 and H2A.Z.1 affect the structure and stability of the nucleosome. Nucleic Acids Res. 46, 10007–10018 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Boileau M et al. Mutant H3 histones drive human pre-leukemic hematopoietic stem cell expansion and promote leukemic aggressiveness. Nat Commun 10, 2891 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Bennett RL et al. A Mutation in Histone H2B Represents a New Class of Oncogenic Driver. Cancer Discov 9, 1438–1451 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Widemann BC & Italiano A Biology and Management of Undifferentiated Pleomorphic Sarcoma, Myxofibrosarcoma, and Malignant Peripheral Nerve Sheath Tumors: State of the Art and Perspectives. J Clin Oncol 36, 160–167 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Lee W et al. PRC2 is recurrently inactivated through EED or SUZ12 loss in malignant peripheral nerve sheath tumors. Nat. Genet 46, 1227–1232 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.De Raedt T et al. PRC2 loss amplifies Ras-driven transcription and confers sensitivity to BRD4-based therapies. Nature 514, 247–251 (2014). [DOI] [PubMed] [Google Scholar]
  • 87.Schuettengruber B, Bourbon H-M, Di Croce L & Cavalli G Genome Regulation by Polycomb and Trithorax: 70 Years and Counting. Cell 171, 34–57 (2017). [DOI] [PubMed] [Google Scholar]
  • 88.Morin RD et al. Somatic mutations altering EZH2 (Tyr641) in follicular and diffuse large B-cell lymphomas of germinal-center origin. Nat. Genet 42, 181–185 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Cleven AHG et al. Loss of H3K27 tri-methylation is a diagnostic marker for malignant peripheral nerve sheath tumors and an indicator for an inferior survival. Mod. Pathol 29, 582–590 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Marchione DM et al. Histone H3K27 dimethyl loss is highly specific for malignant peripheral nerve sheath tumor and distinguishes true PRC2 loss from isolated H3K27 trimethyl loss. Mod. Pathol 32, 1434–1446 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Pierron G et al. A new subtype of bone sarcoma defined by BCOR-CCNB3 gene fusion. Nat. Genet 44, 461–466 (2012). [DOI] [PubMed] [Google Scholar]
  • 92.Specht K et al. Novel BCOR-MAML3 and ZC3H7B-BCOR Gene Fusions in Undifferentiated Small Blue Round Cell Sarcomas. Am. J. Surg. Pathol 40, 433–442 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Kao Y-C et al. BCOR-CCNB3 Fusion Positive Sarcomas: A Clinicopathologic and Molecular Analysis of 36 Cases With Comparison to Morphologic Spectrum and Clinical Behavior of Other Round Cell Sarcomas. Am. J. Surg. Pathol 42, 604–615 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Wang Z et al. A Non-canonical BCOR-PRC1.1 Complex Represses Differentiation Programs in Human ESCs 22, 235–251.e9 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Astolfi A et al. BCOR involvement in cancer. Epigenomics 11, 835–855 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Koontz JI et al. Frequent fusion of the JAZF1 and JJAZ1 genes in endometrial stromal tumors. Proc. Natl. Acad. Sci. U.S.A 98, 6348–6353 (2001). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Lewis N et al. ZC3H7B-BCOR high-grade endometrial stromal sarcomas: a report of 17 cases of a newly defined entity. Mod. Pathol 31, 674–684 (2018). [DOI] [PubMed] [Google Scholar]
  • 98.Panagopoulos I et al. Fusion of the ZC3H7B and BCOR genes in endometrial stromal sarcomas carrying an X;22-translocation. Genes Chromosomes Cancer 52, 610–618 (2013). [DOI] [PubMed] [Google Scholar]
  • 99.Ma X et al. The JAZF1-SUZ12 fusion protein disrupts PRC2 complexes and impairs chromatin repression during human endometrial stromal tumorogenesis. Oncotarget 8, 4062–4078 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Cao R et al. Role of hPHF1 in H3K27 methylation and Hox gene silencing. Mol. Cell. Biol 28, 1862–1872 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Micci F, Panagopoulos I, Bjerkehagen B & Heim S Consistent rearrangement of chromosomal band 6p21 with generation of fusion genes JAZF1/PHF1 and EPC1/PHF1 in endometrial stromal sarcoma. Cancer Res. 66, 107–112 (2006). [DOI] [PubMed] [Google Scholar]
  • 102.Dewaele B et al. Identification of a novel, recurrent MBTD1-CXorf67 fusion in low-grade endometrial stromal sarcoma. Int J Cancer 134, 1112–1122 (2014). [DOI] [PubMed] [Google Scholar]
  • 103.Jacquet K et al. The TIP60 Complex Regulates Bivalent Chromatin Recognition by 53BP1 through Direct H4K20me Binding and H2AK15 Acetylation. Mol. Cell 62, 409–421 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Alfert A, Moreno N & Kerl K The BAF complex in development and disease. Epigenetics Chromatin 12, 19–15 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Perani M, Ingram CJE, Cooper CS, Garrett MD & Goodwin GH Conserved SNH domain of the proto-oncoprotein SYT interacts with components of the human chromatin remodelling complexes, while the QPGY repeat domain forms homo-oligomers. Oncogene 22, 8156–8167 (2003). [DOI] [PubMed] [Google Scholar]
  • 106.Ishida M, Tanaka S, Ohki M & Ohta T Transcriptional co-activator activity of SYT is negatively regulated by BRM and Brg1. Genes Cells 9, 419–428 (2004). [DOI] [PubMed] [Google Scholar]
  • 107.de Bruijn DRH et al. The synovial-sarcoma-associated SS18-SSX2 fusion protein induces epigenetic gene (de)regulation. Cancer Res. 66, 9474–9482 (2006). [DOI] [PubMed] [Google Scholar]
  • 108.Middeljans E et al. SS18 together with animal-specific factors defines human BAF-type SWI/SNF complexes. PLoS ONE 7, e33834 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Su L et al. Deconstruction of the SS18-SSX fusion oncoprotein complex: insights into disease etiology and therapeutics. Cancer Cell 21, 333–347 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Jones KB et al. SS18-SSX2 and the mitochondrial apoptosis pathway in mouse and human synovial sarcomas. Oncogene 32, 2365–71– 2375.e1–5 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Garcia CB, Shaffer CM & Eid JE Genome-wide recruitment to Polycomb-modified chromatin and activity regulation of the synovial sarcoma oncogene SYT-SSX2. BMC Genomics 13, 189 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Pretto D et al. The synovial sarcoma translocation protein SYT-SSX2 recruits beta-catenin to the nucleus and associates with it in an active complex. Oncogene 25, 3661–3669 (2006). [DOI] [PubMed] [Google Scholar]
  • 113.Kadoch C & Crabtree GR Reversible Disruption of mSWI/SNF (BAF) Complexes by the SS18-SSX Oncogenic Fusion in Synovial Sarcoma. Cell 153, 71–85 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.McBride MJ et al. The SS18-SSX Fusion Oncoprotein Hijacks BAF Complex Targeting and Function to Drive Synovial Sarcoma. Cancer Cell 33, 1128–1141.e7 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Thway K & Folpe AL Update on selected advances in the immunohistochemical and molecular genetic analysis of soft tissue tumors. Virchows Arch. path Anat 476, 3–15 (2020). [DOI] [PubMed] [Google Scholar]
  • 116.Alpsoy A & Dykhuizen EC Glioma tumor suppressor candidate region gene 1 (GLTSCR1) and its paralog GLTSCR1-like form SWI/SNF chromatin remodeling subcomplexes. J. Biol. Chem 293, 3892–3903 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.McDonald ER et al. Project DRIVE: A Compendium of Cancer Dependencies and Synthetic Lethal Relationships Uncovered by Large-Scale, Deep RNAi Screening. Cell 170, 577–592.e10 (2017). [DOI] [PubMed] [Google Scholar]
  • 118.Krämer KF, Moreno N, Frühwald MC & Kerl K BRD9 Inhibition, Alone or in Combination with Cytostatic Compounds as a Therapeutic Approach in Rhabdoid Tumors. Int J Mol Sci 18, 1537 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Brien GL et al. Targeted degradation of BRD9 reverses oncogenic gene expression in synovial sarcoma. Elife 7, 3892 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Michel BC et al. A non-canonical SWI/SNF complex is a synthetic lethal target in cancers driven by BAF complex perturbation. Nature Cell Biology 20, 1410–1420 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Banito A et al. The SS18-SSX Oncoprotein Hijacks KDM2B-PRC1.1 to Drive Synovial Sarcoma. Cancer Cell 33, 527–541.e8 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Pawel BR SMARCB1-deficient Tumors of Childhood: A Practical Guide. Pediatr. Dev. Pathol 21, 6–28 (2018). [DOI] [PubMed] [Google Scholar]
  • 123.Thway K, Jones RL, Noujaim J & Fisher C Epithelioid Sarcoma: Diagnostic Features and Genetics. Advances In Anatomic Pathology 23, 41–49 (2016). [DOI] [PubMed] [Google Scholar]
  • 124.Guillou L, Wadden C, Coindre JM, Krausz T & Fletcher CD ‘Proximal-type’ epithelioid sarcoma, a distinctive aggressive neoplasm showing rhabdoid features. Clinicopathologic, immunohistochemical, and ultrastructural study of a series. Am. J. Surg. Pathol 21, 130–146 (1997). [DOI] [PubMed] [Google Scholar]
  • 125.Hornick JL, Dal Cin P & Fletcher CDM Loss of INI1 expression is characteristic of both conventional and proximal-type epithelioid sarcoma. Am. J. Surg. Pathol 33, 542–550 (2009). [DOI] [PubMed] [Google Scholar]
  • 126.Laskin WB & Miettinen M Epithelioid sarcoma: new insights based on an extended immunohistochemical analysis. Arch. Pathol. Lab. Med 127, 1161–1168 (2003). [DOI] [PubMed] [Google Scholar]
  • 127.Lee RS et al. A remarkably simple genome underlies highly malignant pediatric rhabdoid cancers. J Clin Invest 122, 2983–2988 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Jamshidi F et al. The genomic landscape of epithelioid sarcoma cell lines and tumours. J. Pathol 238, 63–73 (2016). [DOI] [PubMed] [Google Scholar]
  • 129.Sullivan LM, Folpe AL, Pawel BR, Judkins AR & Biegel JA Epithelioid sarcoma is associated with a high percentage of SMARCB1 deletions. Mod. Pathol 26, 385–392 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Papp G, Krausz T, Stricker TP, Szendrői M & Sápi Z SMARCB1 expression in epithelioid sarcoma is regulated by miR-206, miR-381, and miR-671–5p on Both mRNA and protein levels. Genes Chromosomes Cancer 53, 168–176 (2014). [DOI] [PubMed] [Google Scholar]
  • 131.Sápi Z et al. Epigenetic regulation of SMARCB1 By miR-206, −381 and −671–5p is evident in a variety of SMARCB1 immunonegative soft tissue sarcomas, while miR-765 appears specific for epithelioid sarcoma. A miRNA study of 223 soft tissue sarcomas. Genes Chromosomes Cancer 55, 786–802 (2016). [DOI] [PubMed] [Google Scholar]
  • 132.Nakayama RT et al. SMARCB1 is required for widespread BAF complex-mediated activation of enhancers and bivalent promoters. Nat. Genet 49, 1613–1623 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Chun H-JE et al. Identification and Analyses of Extra-Cranial and Cranial Rhabdoid Tumor Molecular Subgroups Reveal Tumors with Cytotoxic T Cell Infiltration. Cell Rep 29, 2338–2354.e7 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Brownlee PM, Meisenberg C & Downs JA The SWI/SNF chromatin remodelling complex: Its role in maintaining genome stability and preventing tumourigenesis. DNA Repair 32, 127–133 (2015). [DOI] [PubMed] [Google Scholar]
  • 135.Peng G et al. BRIT1/MCPH1 links chromatin remodelling to DNA damage response. Nature Cell Biology 11, 865–872 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Qi W et al. BRG1 promotes the repair of DNA double-strand breaks by facilitating the replacement of RPA with RAD51. J. Cell. Sci 128, 317–330 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Kakarougkas A et al. Requirement for PBAF in transcriptional repression and repair at DNA breaks in actively transcribed regions of chromatin. Mol. Cell 55, 723–732 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Park J-H et al. Mammalian SWI/SNF complexes facilitate DNA double-strand break repair by promoting gamma-H2AX induction. EMBO J 25, 3986–3997 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.US National Library of Medicine. ClinicalTrials.gov Available at: https://clinicaltrials.gov/ct2/show/NCT02601950. (Accessed: 25 February 2020)
  • 140.Daigle S et al. Molecular characterization of epithelioid sarcoma (ES) tumors derived from patients enrolled in a phase II study of tazemetostat (NCT02601950). Annals of Oncology 29, viii670–viii682 (2018). [Google Scholar]
  • 141.Kohashi K et al. SWI/SNF Chromatin-remodeling Complex Status in SMARCB1/INI1-preserved Epithelioid Sarcoma. Am. J. Surg. Pathol 42, 312–318 (2018). [DOI] [PubMed] [Google Scholar]
  • 142.Perret R et al. SMARCA4-deficient Thoracic Sarcomas: Clinicopathologic Study of 30 Cases With an Emphasis on Their Nosology and Differential Diagnoses. Am. J. Surg. Pathol 43, 455–465 (2019). [DOI] [PubMed] [Google Scholar]
  • 143.Rekhtman N et al. Thoracic SMARCA4-deficient sarcomatoid tumors represent primarily smoking-related undifferentiated carcinomas rather than primary thoracic sarcomas. Journal of Thoracic Oncology 1–30 (2019). doi: 10.1016/j.jtho.2019.10.023 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Karnezis AN et al. Dual loss of the SWI/SNF complex ATPases SMARCA4/BRG1 and SMARCA2/BRM is highly sensitive and specific for small cell carcinoma of the ovary, hypercalcaemic type. J. Pathol 238, 389–400 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Kolin DL et al. SMARCA4-deficient undifferentiated uterine sarcoma (malignant rhabdoid tumor of the uterus): a clinicopathologic entity distinct from undifferentiated carcinoma. Mod. Pathol 31, 1442–1456 (2018). [DOI] [PubMed] [Google Scholar]
  • 146.Lin DI et al. SMARCA4 inactivation defines a subset of undifferentiated uterine sarcomas with rhabdoid and small cell features and germline mutation association. Mod. Pathol 32, 1675–1687 (2019). [DOI] [PubMed] [Google Scholar]
  • 147.Lindén M et al. FET family fusion oncoproteins target the SWI/SNF chromatin remodeling complex. EMBO Rep. 20, 1741 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Schwartz JC, Cech TR & Parker RR Biochemical Properties and Biological Functions of FET Proteins. Annu. Rev. Biochem 84, 141210135300003 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Sankar S & Lessnick SL Promiscuous partnerships in Ewing’s sarcoma. Cancer Genet 204, 351–365 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Couthouis J et al. A yeast functional screen predicts new candidate ALS disease genes. Proc. Natl. Acad. Sci. U.S.A 108, 20881–20890 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Couthouis J et al. Evaluating the role of the FUS/TLS-related gene EWSR1 in amyotrophic lateral sclerosis. Human Molecular Genetics 21, 2899–2911 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Kato M et al. Cell-free Formation of RNA Granules: Low Complexity Sequence Domains Form Dynamic Fibers within Hydrogels. 149, 753–767 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Kwon I et al. Phosphorylation-regulated binding of RNA polymerase II to fibrous polymers of low-complexity domains. Cell 155, 1049–1060 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Schwartz JC, Wang X, Podell ER & Cech TR RNA seeds higher-order assembly of FUS protein. Cell Rep 5, 918–925 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Burke KA, Janke AM, Rhine CL & Fawzi NL Residue-by-Residue View of In Vitro FUS Granules that Bind the C-Terminal Domain of RNA Polymerase II. Mol. Cell 60, 231–241 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Patel A et al. A Liquid-to-Solid Phase Transition of the ALS Protein FUS Accelerated by Disease Mutation. Cell 162, 1066–1077 (2015). [DOI] [PubMed] [Google Scholar]
  • 157.Lu H et al. Phase-separation mechanism for C-terminal hyperphosphorylation of RNA polymerase II. Nature 558, 318–323 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Sabari BR et al. Coactivator condensation at super-enhancers links phase separation and gene control. Science 361, eaar3958 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Riggi N et al. EWS-FLI1 Utilizes Divergent Chromatin Remodeling Mechanisms to Directly Activate or Repress Enhancer Elements in Ewing Sarcoma. Cancer Cell 26, 668–681 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Boulay G et al. Epigenome editing of microsatellite repeats defines tumor-specific enhancer functions and dependencies. Genes Dev. 32, 1008–1019 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Tomazou EM et al. Epigenome mapping reveals distinct modes of gene regulation and widespread enhancer reprogramming by the oncogenic fusion protein EWS-FLI1. Cell Rep 10, 1082–1095 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Yoshida K-I et al. Absence of H3F3A mutation in a subset of malignant giant cell tumor of bone. Mod. Pathol 32, 1751–1761 (2019). [DOI] [PubMed] [Google Scholar]
  • 163.Chen S et al. Diagnostic utility of IDH1/2 mutations to distinguish dedifferentiated chondrosarcoma from undifferentiated pleomorphic sarcoma of bone. Hum. Pathol 65, 239–246 (2017). [DOI] [PubMed] [Google Scholar]
  • 164.Noujaim J et al. The spectrum of EWSR1-rearranged neoplasms at a tertiary sarcoma centre; assessing 772 tumour specimens and the value of current ancillary molecular diagnostic modalities. Br J Cancer 116, 669–678 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165.Qadir MA et al. ChildSeq-RNA: A next-generation sequencing-based diagnostic assay to identify known fusion transcripts in childhood sarcomas. J Mol Diagn 16, 361–370 (2014). [DOI] [PubMed] [Google Scholar]
  • 166.Zhu G et al. Diagnosis of known sarcoma fusions and novel fusion partners by targeted RNA sequencing with identification of a recurrent ACTB-FOSB fusion in pseudomyogenic hemangioendothelioma. Mod. Pathol 32, 609–620 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Chang KTE et al. Development and Evaluation of a Pan-Sarcoma Fusion Gene Detection Assay Using the NanoString nCounter Platform. J Mol Diagn 20, 63–77 (2018). [DOI] [PubMed] [Google Scholar]
  • 168.Salguero-Aranda C, Amaral AT, Olmedo-Pelayo J, Diaz-Martin J & de. Álava E Breakthrough Technologies Reshape the Ewing Sarcoma Molecular Landscape. Cells 9, 804 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Cermakova K & Hodges HC Next-Generation Drugs and Probes for Chromatin Biology: From Targeted Protein Degradation to Phase Separation. Molecules 23, 1958 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Pfister SX & Ashworth A Marked for death: targeting epigenetic changes in cancer. Nat Rev Drug Discov 16, 241–263 (2017). [DOI] [PubMed] [Google Scholar]
  • 171.Versteege I et al. Truncating mutations of hSNF5/INI1 in aggressive paediatric cancer. Nature 394, 203–206 (1998). [DOI] [PubMed] [Google Scholar]
  • 172.Modena P et al. SMARCB1/INI1 tumor suppressor gene is frequently inactivated in epithelioid sarcomas. Cancer Res. 65, 4012–4019 (2005). [DOI] [PubMed] [Google Scholar]
  • 173.Valencia AM et al. Recurrent SMARCB1 Mutations Reveal a Nucleosome Acidic Patch Interaction Site That Potentiates mSWI/SNF Complex Chromatin Remodeling. Cell 179, 1342–1356.e23 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174.Wilson BG et al. Epigenetic antagonism between polycomb and SWI/SNF complexes during oncogenic transformation. Cancer Cell 18, 316–328 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175.Knutson SK et al. Durable tumor regression in genetically altered malignant rhabdoid tumors by inhibition of methyltransferase EZH2. Proc. Natl. Acad. Sci. U.S.A 110, 7922–7927 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176.Wang Y et al. The histone methyltransferase EZH2 is a therapeutic target in small cell carcinoma of the ovary, hypercalcaemic type. J. Pathol 242, 371–383 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Stacchiotti S et al. Safety and efficacy of tazemetostat, a first-in-class EZH2 inhibitor, in patients (pts) with epithelioid sarcoma (ES) (NCT02601950). JCO 37, 11003–11003 (2019). [Google Scholar]
  • 178.Schöffski P et al. Phase 2 multicenter study of the EZH2 inhibitor tazemetostat in adults with synovial sarcoma (NCT02601950). JCO 35, 11057–11057 (2017). [Google Scholar]
  • 179.US National Library of Medicine. ClinicalTrials.gov Available at: https://clinicaltrials.gov/ct2/show/NCT02273739. (Accessed: 25 February 2020)
  • 180.US National Library of Medicine. ClinicalTrials.gov Available at: https://clinicaltrials.gov/ct2/show/NCT02073994. (Accessed: 25 February 2020)
  • 181.Fan B et al. Clinical pharmacokinetics and pharmacodynamics of ivosidenib, an oral, targeted inhibitor of mutant IDH1, in patients with advanced solid tumors. Invest New Drugs 462, 739–12 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182.Tap W et al. Phase I Study of the Mutant IDH1 Inhibitor Ivosidenib: Safety and Clinical Activity in Patients With Advanced Chondrosarcoma. J Clin Oncol JCO1902492 (2020). doi: 10.1200/JCO.19.02492 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Tap W et al. A phase 1 study of AG-120, an IDH1 mutant enzyme inhibitor: results from the chondrosarcoma dose escalation and expansion cohorts. Connective Tissue Oncology Society Annual Meeting (2016). [Google Scholar]
  • 184.Wang X et al. BRD9 defines a SWI/SNF sub-complex and constitutes a specific vulnerability in malignant rhabdoid tumors. Nat Commun 10, 1881–11 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.McBride MJ & Kadoch C Disruption of mammalian SWI/SNF and polycomb complexes in human sarcomas: mechanisms and therapeutic opportunities. J. Pathol 244, 638–649 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Cooper JM et al. Overcoming BET Inhibitor Resistance in Malignant Peripheral Nerve Sheath Tumors. Clin. Cancer Res 25, 3404–3416 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187.Cancer Genome Atlas Research Network. Comprehensive genomic characterization defines human glioblastoma genes and core pathways. Nature 455, 1061–1068 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Demetri GD et al. Efficacy and safety of imatinib mesylate in advanced gastrointestinal stromal tumors. N. Engl. J. Med 347, 472–480 (2002). [DOI] [PubMed] [Google Scholar]
  • 189.Doebele RC et al. An Oncogenic NTRK Fusion in a Patient with Soft-Tissue Sarcoma with Response to the Tropomyosin-Related Kinase Inhibitor LOXO-101. Cancer Discov 5, 1049–1057 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190.Drilon A et al. Efficacy of Larotrectinib in TRK Fusion-Positive Cancers in Adults and Children. N. Engl. J. Med 378, 731–739 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.Laetsch TW et al. Larotrectinib for paediatric solid tumours harbouring NTRK gene fusions: phase 1 results from a multicentre, open-label, phase 1/2 study. Lancet Oncol. 19, 705–714 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.Hyman AA, Weber CA & Jülicher F Liquid-liquid phase separation in biology. Annu. Rev. Cell Dev. Biol 30, 39–58 (2014). [DOI] [PubMed] [Google Scholar]
  • 193.Banani SF, Lee HO, Hyman AA & Rosen MK Biomolecular condensates: organizers of cellular biochemistry. Nat. Rev. Mol. Cell Biol 18, 285–298 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 194.Shin Y & Brangwynne CP Liquid phase condensation in cell physiology and disease. Science 357, eaaf4382 (2017). [DOI] [PubMed] [Google Scholar]
  • 195.Overbeek JT & Voorn MJ Phase separation in polyelectrolyte solutions; theory of complex coacervation. J Cell Physiol Suppl 49, 7–22–discussion–22–6 (1957). [PubMed] [Google Scholar]
  • 196.Boeynaems S et al. Protein Phase Separation: A New Phase in Cell Biology. Trends Cell Biol. 28, 420–435 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197.Mao YS, Zhang B & Spector DL Biogenesis and function of nuclear bodies. Trends Genet. 27, 295–306 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 198.Decker CJ & Parker R P-bodies and stress granules: possible roles in the control of translation and mRNA degradation. Cold Spring Harb Perspect Biol 4, a012286–a012286 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 199.Li P et al. Phase transitions in the assembly of multivalent signalling proteins. Nature 483, 336–340 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 200.Wang J et al. A Molecular Grammar Governing the Driving Forces for Phase Separation of Prion-like RNA Binding Proteins. Cell 174, 688–699.e16 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 201.Banjade S & Rosen MK Phase transitions of multivalent proteins can promote clustering of membrane receptors. Elife 3, 641 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 202.Su X et al. Phase separation of signaling molecules promotes T cell receptor signal transduction. Science 352, 595–599 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 203.Romero P et al. Sequence complexity of disordered protein. Proteins 42, 38–48 (2001). [DOI] [PubMed] [Google Scholar]
  • 204.Vucetic S, Brown CJ, Dunker AK & Obradovic Z Flavors of protein disorder. Proteins 52, 573–584 (2003). [DOI] [PubMed] [Google Scholar]
  • 205.Brangwynne CP, Tompa P & Pappu RV Polymer physics of intracellular phase transitions. Nature Phys 11, 899–904 (2015). [Google Scholar]
  • 206.King OD, Gitler AD & Shorter J The tip of the iceberg: RNA-binding proteins with prion-like domains in neurodegenerative disease. Brain Res. 1462, 61–80 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 207.Boija A et al. Transcription Factors Activate Genes through the Phase-Separation Capacity of Their Activation Domains. Cell 175, 1842–1855.e16 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 208.Alberti S & Dormann D Liquid-Liquid Phase Separation in Disease. Annu. Rev. Genet 53, 171–194 (2019). [DOI] [PubMed] [Google Scholar]
  • 209.Alberti S, Gladfelter A & Mittag T Considerations and Challenges in Studying Liquid-Liquid Phase Separation and Biomolecular Condensates. Cell 176, 419–434 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 210.McSwiggen DT, Mir M, Darzacq X & Tjian R Evaluating phase separation in live cells: diagnosis, caveats, and functional consequences. Genes Dev. (2019). doi: 10.1101/gad.331520.119 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 211.US National Library of Medicine. ClinicalTrials.gov Available at: https://clinicaltrials.gov/ct2/show/NCT02601937. (Accessed: 25 February 2020)
  • 212.US National Library of Medicine. ClinicalTrials.gov Available at: https://clinicaltrials.gov/ct2/show/NCT03213665. (Accessed: 25 February 2020)
  • 213.US National Library of Medicine. ClinicalTrials.gov Available at: https://clinicaltrials.gov/ct2/show/NCT02875548. (Accessed: 25 February 2020)

RESOURCES