Abstract
The enhancer of zeste homolog 2 (EZH2) is a catalytic component of Polycomb repressive complex 2 (PRC2) mediating the methylation of histone 3 lysine 27 (H3K27me3) and hence the epigenetic repression of target genes, known as canonical function. Growing evidence indicates that EZH2 has non-canonical roles that are exerted as PRC2-dependent and PRC2-independent methylation of non-histone proteins, and methyltransferase-independent interactions of EZH2 with various proteins contributing to gene expression regulation and alterations in the protein stability. EZH2 is frequently mutated and/or its expression is deregulated in various cancer types. The cancer sensitivity to inhibitors of EZH2 enzymatic activity and state-of-the-art approaches to deplete EZH2 with chemical degraders are discussed. This review also presents the clinical trials in various phases that evaluate the use of EZH2 inhibitors, both as monotherapy and in combination with other agents for the treatment of patients with diverse types of cancers.
Keywords: cancer, epigenetic regulation, EZH2, methyltransferase activity, targeted protein degradation
Background
Enhancer of zeste homolog 2 (EZH2) is one of the epigenetic modifiers that contribute to the global and focal transcriptional output, thus defining normal organismal development and developmental disorders, as well as cancer and mental health diseases. 1 It is the catalytic subunit of Polycomb repressive complex 2 (PRC2), a protein complex that contains also embryonic ectoderm development (EED), suppressor of zeste 12 (SUZ12), and retinoblastoma cell tumor-associated proteins 46 and 48 (RbAp46/48).1,2 In addition to these core components, several facultative protein subunits such as AE-binding protein 2 (AEBP2), jumonji and AT-rich interaction domain containing 2 (JARID2), Polycomb-like (PCL) proteins, and various RNA molecules can interact with PRC2 defining the specificity of its binding to chromatin and activity.1,3–5 PRC2 mediates mono-, di-, and trimethylation of the lysine 27 on the tail of histone 3 (H3K27me1/2/3) (Figure 1). While H3K27me1 is found in actively transcribed loci, H3K27me3 is associated with chromatin compaction and epigenetic silencing. 6 The transfer of a methyl group from S-adenosyl methionine (SAM) to H3K27 is catalyzed by an enzymatic domain of EZH2 named SET (Su(var)3-9, enhancer of zeste, and Trithorax). SET domain is not active unless EZH2 interacts with SUZ12 and EED.7,8
Figure 1.
Schematic presentation of canonical and non-canonical function of EZH2. Canonically, EZH2 as an enzymatic subunit of PRC2 silences target genes by trimethylation of H3K27. Several facultative protein subunits, such as PCL, JARID2, and AEBP2 can bind to PRC2 and redirect the repressive complex to specific targets. DNMT-mediated methylation of DNA complements gene silencing. Non-canonically, EZH2 can methylate non-histone proteins leading to either gene silencing or activation and protein stability regulation. The non-canonical activity of EZH2 can be also methyltransferase independent and result from EZH2 direct interactions with several transcription regulatory proteins. See Figures 2–5 for details.
AEBP2, AE-binding protein 2; DNMT, DNA methyltransferase; EED, embryonic ectoderm development; EZH2, enhancer of zeste homolog 2; H3K27, histone 3 lysine 27; JARID2, jumonji and AT-rich interaction domain containing 2; Me, methyl group; P, phosphate; PCL, Polycomb-like proteins; PRC2, Polycomb repressive complex 2; RbAp, retinoblastoma associated protein; SUZ12, suppressor of zeste 12.
H3K27me3 is a marker of transcriptional repression, and EZH2-mediated epigenetic silencing of cancer suppressors contributes to the development of various malignancies, including hepatocellular carcinoma (HCC), 9 prostate cancer (PCa),10,11 breast cancer, 12 ovarian cancer, 13 lung cancer, 14 B-cell lymphomas, 15 glioblastoma, 16 colorectal cancer, 17 and melanoma. 18 Augmented expression of EZH2 in cancer often correlates with poor prognosis.19,20 Gain-of-function (GoF) mutations of the EZH2 SET domain, including Y646 (Y641 in mouse), have been associated with enhanced histone methyltransferase (HMT) activity and downregulation of cancer suppressor genes. 21 In hematological malignancies, changes in the EZH2 level are mainly caused by somatic mutations, either GoF mutations predominantly detected in lymphomas22,23 or loss-of-function (LoF) mutations mostly found in myeloid disorders.24,25 Analysis of LoF mutations in EZH2 suggests that EZH2 can act as a tumor suppressor in selected hematologic malignancies.24–26 Abnormal levels and activity of EZH2 in cancer cells can also result from various transcriptional and post-transcriptional modifications of EZH2, as well as cancer-type-specific alterations in signaling pathways (see below).
Several recent reviews summarize the canonical function of EZH2.8,27–29 In this review, we focus on the non-canonical role of EZH2 related to the methylation of non-histone proteins and methyltransferase-independent activity resulting from the direct binding of EZH2 to various nuclear and cytoplasmic proteins (Figure 1). These non-canonical roles of EZH2 are considered as a part of its oncogenic activity. We discuss the experimental approaches undertaken to inhibit EZH2 activity, and dual inhibitors and degraders of EZH2 protein are reviewed in more detail. We also summarize current data from clinical trials investigating EZH2 inhibitors as monotherapy or in combination with other agents for the treatment of various cancer types.
Non-canonical methyltransferase-dependent EZH2 activity
EZH2 activity contributes to gene repression through H3K27 methylation, and this activity is referred to as canonical. Methylation of various non-histone proteins is a part of a non-canonical EZH2 activity. It includes methylation of PRC2 components to enhance HMT activity and non-histone protein methylation participating in the positive and negative regulation of gene expression and the control of protein turnover.
Methylation of PRC2 components and cofactors to support canonical activity
Non-canonical methylation of the components of the PRC2 complex enhances its HMT activity (Figure 2(a)). Automethylation of EZH2 occurs in cis (intramolecularly) at three lysine residues K510, K514, and K515 that are close to each other on the conserved flexible loop containing many positively charged amino acids. 30 Utomethylation of EZH2 in response to cellular SAM concentration and the status of H3K27 methylation has been shown in vitro and in vivo as promoting HMT activity by increased accessibility of PRC2 to H3K27.30,31 In addition, EZH2 automethylation is supported by AEBP2, a PRC2 cofactor, and diminished by another PRC2 cofactor, JARID2. 31 EZH2 automethylation is important during stem cell reprogramming and differentiation, the processes that require de novo methylation of H3K27. 30 A triple mutation of EZH2 at K510, K514, and K515 preventing automethylation reduces the methylating activity of PRC2, specifically conversion of H3K27me2 to H3K27me3, which is a rate-limiting step.30,31
Figure 2.
PRC2-dependent methylation of non-histone proteins. (a) The autoregulatory feedback loop, in which EZH2 methylates the components of PRC2 complex to enhance H3K27 trimethylation and repression of PRC2 target genes. (b) EZH2 methylates EloA, the RNA polymerase II transcription elongation factor, to attenuate transcription. Methylated EloA may either recruit its repressive reader (red circle) or evict a positive transcription factor (green circle). (c) PRC2-dependent gene repression can be enhanced by direct methylation of transcription factors such as GATA4.
Ac, acetyl group; EED, embryonic ectoderm development; EloA, Elongin A; EZH2, enhancer of zeste homolog 2; GATA4, GATA binding protein 4; H3K27, histone 3 lysine 27; H3K27me3, H3K27 trimethylation; JARID2, jumonji and AT-rich interaction domain containing 2; Me, methyl group; Pol II, RNA polymerase II; PRC2, Polycomb repressive complex 2; RbpAp46/48, retinoblastoma-associated protein 46/48; SAM, S-adenosyl methionine; SUZ12, suppressor of zeste 12.
Apart from automethylation, EZH2 can mediate SUZ12 methylation, which is diminished by AEBP2.30,31 Methylation of JARID2 at K116 is another PRC2-dependent non-canonical activity of EZH2. JARID2me3 by binding to an aromatic cage of EED promotes allosteric activation of PRC2 catalytic activity. 3
PRC2-dependent non-canonical activity of EZH2 in the repression of transcription
EZH2 in the PRC2 complex can repress transcription not only by trimethylation of H3K27 but also by methylation of non-histone targets (Figure 2, Table 1). Elongin A (EloA), a subunit of ternary complex EloA-EloB-EloC, which interacts with RNA polymerase II and stimulates transcription elongation, is monomethylated by EZH2 at K754. 32 This methylation complements PRC2-dependent gene silencing and is important for normal embryonic stem cell differentiation. 32 In the proposed model, methylated EloA may either recruit its repressive reader (red circle) or evict a positive transcription factor (green circle), in both cases downregulating the expression of target genes (Figure 2(b)).
Table 1.
EZH2-mediated methylation of non-histone targets.
Substrate | Site and type of methylation | Cell type | Mode of action | Molecular effect | Reference |
---|---|---|---|---|---|
EZH2 | K510me1/2/3 K514me2/3 K515me1 |
Mouse embryonic stem cell lines; HEK293T cells |
PRC2 dependent | Increasing HMT activity of PRC2 | Wang et al. (2019), 30 Lee et al. (2019) 31 |
SUZ12 | N/A | Mouse embryonic stem cell lines; HEK293T cells |
PRC2 dependent | Increasing HMT activity of PRC2 | Wang et al. (2019), 30 Lee et al. (2019) 31 |
JARID2 | K116me2/3 | HEK293T cells and embryonic stem cell lines | PRC2 dependent | EED binding; increasing HMT activity of PRC2 | Sanulli et al. (2015) 3 |
EloA | K754me1 | Mouse embryonic stem cell lines | PRC2 dependent | Gene repression in ESCs | Ardehali et al. (2017) 32 |
GATA4 | K299me1 | Cardiac muscle cell lines and embryonic stem cell lines | PRC2 dependent | Repression of GATA4 target genes; diminution of p300 binding | He et al. (2012) 33 |
AR | N/A | CRPC cell lines | PRC2 independent | Activation of AR target gene expression | Xu et al. (2012) 11 |
STAT3 | K180me1 K709me2 |
Glioblastoma stem cell lines | PRC2 dependent | Activation of STAT3 target gene expression | Kim et al. (2013) 34 |
K180me1 | Breast cancer cell lines | N/A | Zhao et al. (2021) 35 | ||
K49me2 | Colon cancer cell lines | N/A | Dasgupta et al. (2015) 36 | ||
N/A | Melanoma mouse model | N/A | Hypermethylation of STAT3 by EZH2Y641F and activation of interferon-regulated gene expression | Zimmerman et al. (2022) 37 | |
β-catenin | K49me3 | Colorectal cell lines | N/A | Induction of β-catenin interaction with TCF1 and RNA polymerase II | Ghobasi et al. (2023) 38 |
RORα | K38me1 | Breast cancer cell lines | N/A | Promoting RORα ubiquitination and degradation | Lee et al. (2012) 39 |
PLZF | K430me1 | NKT cell lines | N/A | Promoting PLZF ubiquitination and degradation | Vasanthakumar et al. (2017) 40 |
DLC1 | K678me1 | NSCLC cell lines | N/A | Promoting DCT1 ubiquitination and degradation | Tripathi et al. (2021) 41 |
FOXA1 | K295me1 | Prostate cancer cell lines | PRC2 dependent | Preventing FOXA1 degradation | Park et al. (2021) 42 |
p38 | K139me1/3 K165me1 |
TNBC cell lines | PRC2 dependent | Increasing p38 stability | Gonzalez et al. (2022) 43 |
talin1 | K2454me3 | Neutrophils and dendritic cells | PRC2 dependent | Modulation of talin1-F-actin interaction | Gunawan et al. (2015) 44 Venkatesan et al. (2018) 45 |
AR, androgen receptor; CRPC, castration-resistant prostate cancer; DLC1, deleted in liver cancer 1; EED, embryonic ectoderm development; EloA, elongin A; ESC, embryonic stem cell; EZH2, enhancer of zeste homolog 2; FOXA1, forkhead box A1; GATA4, GATA-binding protein 4; HMT, histone methyltransferase; JARID2, jumonji and AT-rich interaction domain containing 2; me1,2,3, mono-, di-, trimethylation; N/A, not available; NKT, natural killer T cell; NSCLC, non-small-cell lung cancer; PLZF, promyelocytic leukemia zinc finger; PRC2, Polycomb repressive complex 2; RORα, retinoic acid-related orphan nuclear receptor alpha; STAT3, signal transducer and activator of transcription 3; SUZ12, suppressor of zeste 12; TCF1, T-cell factor 1; TNBC, triple-negative breast cancer.
EZH2 as a part of PRC2 can also directly methylate transcription factors to diminish their activity in normal lineage-specific differentiation and during cancerogenesis (Figure 1, Table 1). It has been demonstrated that PRC2 can interact with GATA-binding protein 4 (GATA4), a transcription factor crucial for heart development. Methylation of GATA4 at K299 attenuates its acetylation by transcriptional coactivator p300, thus repressing the transcriptional potency of GATA433 (Figure 2(c)). An interaction of EZH2 with GATA4 has been also detected in cardiac rhabdomyosarcoma. 46
Non-canonical methyltransferase activity of EZH2 in gene activation
Growing evidence indicates that EZH2 can also act as a transcriptional co-activator (Figure 3, Table 1). This activity can be mediated by its trans-activating domain (TAD) that spans amino acids 135–254. 47 In normal conditions, TAD is structurally blocked but can be unlocked by cancer-specific phosphorylation of EZH2. 47 EZH2 can be phosphorylated at serine 21 (S21) by pAKT as shown in breast cancer, 48 castration-resistant prostate cancer (CRPC), 11 glioblastoma stem cell subpopulation, 34 and colorectal cancer. 38 EZH2 phosphorylation at S21 reduces the EZH2 affinity to H3K27 leading to de-repression of PRC2 target genes. 48 It has been shown in PCa cells that the chromatin loci enriched in phosphorylated EZH2 are devoid of SUZ12, EED, and other components of PRC2, but contain RNA polymerase II and di- and tri-methylated H3K4, which is a mark of transcriptionally active chromatin. 11 Phosphorylation of EZH2 at S21 turns EZH2 into a transcriptional co-activator promoting proliferation, invasion, and stem-like properties of cancer cells by methylating specific transcription factors, such as androgen receptor (AR), 11 signal transducer and activator of transcription 3 (STAT3),34,36 and β-catenin 38 (Figure 3, Table 1). It has been shown in colorectal cancer that S21-phosphorylated EZH2 trimethylates β-catenin at K49, which induces β-catenin interaction with RNA polymerase II and T-cell factor 1 and modulates the expression of genes involved in cell migration and metabolic processes. 38 While the pEZH2-STAT3 interaction is common for various cancer types, its molecular effects are context dependent. For instance, EZH2-mediated STAT3 monomethylation at K180 has been identified in the glioblastoma stem-like cell subpopulation characterized by highly active phosphoinositide 3 kinase (PI3K)/AKT pathway and increased level of S21-phosphorylated EZH2. 34 EZH2-mediated methylation of STAT3 at K180 in breast cancer cells reinforces its nuclear retention and expression of STAT3 targets such as matrix metalloproteinase 2 and cyclin D1. 35 STAT3 dimethylation at K49 in colon cancer cells favors the expression of interleukin 6 (IL-6)-dependent genes. 36 In murine melanoma cells, the GoF EZH2Y641F mutant cooperates with STAT3 to drive the upregulation of a subset of interferon-regulated genes. 37
Figure 3.
Non-canonical EZH2 methyltransferase activity promoting gene expression. Phosphorylation of EZH2 at S21 by pAKT has been recognized as a molecular switch that turns EZH2 into a transcriptional activator. pS21-EZH2 can methylate AR, STAT3, and β-catenin, which leads to an activation of gene sets supporting the progression of various cancers.
AKT, serine/threonine kinase; AR, androgen receptor; EZH2, enhancer of zeste homolog 2; Me, methyl group; P, phosphate; S21, serine 21; STAT3, signal transducer and activator of transcription 3.
Non-canonical EZH2 methyltransferase activity involved in protein stability
Another non-canonical function of EZH2 is its contribution to the regulation of protein stability, particularly in methylation-dependent ubiquitination. In this mechanism, EZH2 methylates the protein substrate at lysine, which is subsequently selectively recognized by a specific adaptor linking monomethylated substrate to E3 ubiquitin ligase complex composed of damage-specific DNA-binding protein 1 (DDB1) and cullin 4 (CUL4). Such a mechanism of protein degradation has been shown for retinoic acid-related orphan nuclear receptor α (RORα) with DDB1 and CUL4-associated factor 1 (DCAF1) as an adaptor. 39 Monomethylation of RORα at K38 by EZH2 leading to RORα ubiquitination and degradation has been found in breast cancer cell lines 39 (Figure 4(a), Table 1). An elevated level of EZH2 correlates with a reduced level of RORα in breast tumor specimens further suggesting that EZH2 may facilitate RORα degradation, thus contributing to the repression of RORα target genes. 39 Interaction of EZH2 with a ubiquitin ligase complex has been also detected in natural killer T (NKT) cells. 40 In these cells, EZH2 monomethylates promyelocytic leukemia zinc finger transcription factor at K430 promoting its ubiquitination and degradation 40 (Figure 4(a), Table 1). EZH2 playing a role in protein stability has been also found in liver cancer cells. 41 Deleted in liver cancer 1 (DLC1) can form a complex with EZH2 in the cytoplasm, which results in DLC1 methylation at K678 followed by the reduction of its steady-state level (Figure 4(a), Table 1).
Figure 4.
EZH2 methyltransferase activity is involved in protein turnover. (a) EZH2 mediates monomethylation of RORα, PLZF, and DLC1. Methylated proteins are then recognized by DCAF1 associated with the CUL3/CUL4 E3 ligase complex and directed to proteasomal degradation. (b) EZH2-mediated methylation of transcription factors such as FOXA1 stimulates their deubiquitylation and prevents their degradation, which in the case of FOXA1 contributes to prostate cancer development. (c) EZH2-p38α cooperation is needed to enhance the activity of the PI3K/AKT pathway in breast cancer. p38α phosphorylates EZH2 to trigger its cytoplasmic localization, which, in turn, causes methylation of p38α, thereby enhancing its activity and stability.
AEBP2, AE binding protein 2; AKT, serine/threonine kinase; BUB3, BUB3 mitotic checkpoint protein; CUL3/4, cullin 3/4; DCAF1, DDB1 and CUL4-associated factor 1; DLC1, deleted in liver cancer 1; DDB1, damage-specific DNA binding protein 1; EED, embryonic ectoderm development; EZH2, enhancer of zeste homolog 2; FOXA1, forkhead box A1; JARID2 jumonji and AT-rich interaction domain containing 2; Me, methyl group; P, phosphate; PI3K, phosphoinositide 3 kinase; PLZF, promyelocytic leukemia zinc finger; RORα, retinoic acid-related orphan nuclear receptor alpha; SAM, S-adenosyl methionine; SUZ12, suppressor of zeste 12; Ub, ubiquitin; USP7, ubiquitin-specific peptidase 7.
Ubiquitination is balanced by deubiquitylating enzymes that play distinct roles in tumorigenesis. 49 It has been reported that EZH2-mediated methylation of forkhead box A1 (FOXA1) at K295 is recognized by protein BUB3, which subsequently stimulates the recruitment of ubiquitin-specific peptidase 7 to induce deubiquitylation, thereby preventing FOXA1 degradation along the proteasome proteolytic pathway 42 (Figure 4(b), Table 1). EZH2-enhanced stabilization of transcription factor FOXA1 has been found to promote cell cycle progression and growth of PCa. 42 In HCC, EZH2 associated with a long noncoding RNA called lnc-β-Catm and β-catenin induces β-catenin methylation at K49, which, in turn, suppresses the ubiquitination of β-catenin sustaining a high activity of the WNT/β-catenin pathway. 50
Other non-canonical EZH2 methyltransferase activities in cytoplasmic localization
The enhanced activity of the PI3K/AKT pathway has been reported to be associated with increased EZH2 levels in triple-negative breast cancer (TNBC) progression. 43 The molecular mechanisms are complex and involve EZH2-p38α cooperation in the AKT activation (Figure 4(c), Table 1). EZH2 phosphorylation at T367 by p38α in the nucleus is needed for the cytoplasmic localization of EZH2. Cytoplasmic pEZH2, in turn, induces p38α methylation at K165 and trimethylation at K139 enhancing p38α stability and activity leading to AKT activation. Thus, the study suggests that EZH2 exerts a dual function in promoting breast cancer through the H3K27me3 silencing mechanism and by methylation of p38α. 43
The methylation of cytoplasmic integrin adaptor talin1 is another example of the non-canonical EZH2 methyltransferase activity in cytoplasmic localization. 44 This extranuclear function of EZH2 affects the binding of talin1 to F-actin, which disturbs the migration of neutrophils and dendritic cells, and immune responses. Further study has shown that this cytoplasmic activity of EZH2 can promote tumorigenesis through the interaction with the cytoskeleton remodeling effector, VAV. 45
Non-canonical methyltransferase-independent activity of EZH2
A growing number of methyltransferase-independent EZH2 activities have been reported recently. This chapter will focus only on the most prominent EZH2 interactions leading to stabilization of target proteins, and fine-tuning the activity of various oncogenic pathways. The interactions of EZH2 with non-coding RNAs that additionally modify the mode of action of EZH2 are beyond the scope of this review and are summarized elsewhere.4,5
Protein stabilization by EZH2 binding
EZH2 can control protein turnover either in a methyltransferase-dependent (Table 1, Figure 4) or non-catalytic manner (Table 2). EZH2 can bind directly to specific target proteins to impede their ubiquitination and proteasomal degradation.51,52
Table 2.
Methyltransferase-independent EZH2 interactions.
Interacting molecules | Cell type | Molecular function | Reference |
---|---|---|---|
MYCN | Neuroblastoma and SCLC cell lines | Preventing FBXW7-mediated MYC and MYCN polyubiquitination and proteasomal degradation | Wang et al. (2022) 51 |
DDB2 | SCLC cell lines | Stabilization of DDB1/DDB2 complex; impeding ubiquitination | Koyen et al. (2020) 52 |
p300/MYC | AML and MLL cell lines | Activation of MYC-dependent genes | Wang et al. (2022) 53 |
MYCN | PTCL cell lines | Promoting cancer cell proliferation | Vanden Bemt et al. (2022) 54 |
RNA polymerase II | Breast cancer cell lines | TGFβ activation/promoting bone metastases | Zhang et al. (2022) 55 |
NKTL cell lines | Activation of genes involved in DNA replication, cell cycle, stemness, and invasion | Yan et al. (2016) 56 | |
Melanoma brain metastases | Upregulation of pro-tumorigenic inflammatory cytokines | Zhang et al. (2020) 57 | |
TRIM28 | ER+ breast cancer cell lines | Enhancement of expression of genes related to stemness and metastasis | Li et al. (2017) 58 |
ERα/β-catenin | Breast cancer cell lines | Activation of ER and Wnt/β-catenin pathways | Shi et al. (2007) 59 |
PAF/β-catenin | Colon cancer cell lines | Activation of Wnt/β-catenin pathway | Jung et al. (2013) 60 |
β-catenin | Hepatoblastoma cell lines and human tumor samples | Activation of Wnt/β-catenin pathway | Glaser et al. (2024) 61 |
HP1BP3 | Glioma stem cells | Upregulation of WNT7B expression | Yu et al. (2023) 62 |
RelA/RelB | ER-negative breast cancer cell lines | Activation of NF-κB target genes, e.g., IL-6 and TNFα | Dardis et al. (2023), 63 Lee et al. (2011) 64 |
FOXM1 | TNBC cell lines | Activation of EMT genes | Mahara et al. (2016) 65 |
AR | Prostate cancer cell lines | Activation of AR target gene expression | Wang et al. (2022) 66 |
E2F1 | Prostate cancer cell lines | Activation of E2F1 and its target genes | Yi et al. (2022) 67 |
FBL | Prostate cancer cell lines | Enhancing rRNA 2′-O methylation and translation | Yi et al. (2021) 68 |
AML, acute myeloid leukemia; AR, androgen receptor; DDB1/2, damage-specific DNA-binding protein 1/2; E2F1, E2F transcription factor 1; EMT, epithelial–mesenchymal transition; ER, estrogen receptor; EZH2, enhancer of zeste homolog 2; FBL, fibrillarin; FBXW7, F-box and WD repeat domain containing 7; FOXM1, forkhead box M1; HP1BP3, heterochromatin protein 1 binding protein 3; IL6, interleukin 6; MLL, mixed lineage leukemia; MYC(N), proto-oncogene, bHLH transcription factor; NF-κB, nuclear factor kappa B; NKTL, natural killer T-cell lymphoma; PAF, PCNA-associated factor; PTCL, peripheral T-cell lymphoma; SCLC, small-cell lung cancer; TGFβ, transforming growth factor beta; TNBC, triple-negative breast cancer; TNFα, tumor necrosis factor alpha; TRIM28, tripartite motif containing 28; WNT7B, Wnt family member 7B.
In neuroblastoma and small-cell lung carcinoma (SCLC), EZH2 interacts with two oncogenic proteins, MYCN and MYC, respectively. EZH2 prevents F-box and WD repeat domain-containing 7 (FBXW7)-mediated MYCN polyubiquitination and proteasomal degradation, and EZH2 depletion decreases MYCN transcriptional activity. 51
Methyltransferase-independent interference of EZH2 with ubiquitin-proteasome pathway leads to increased stability of oncogenic proteins such as DDB2. DDB2 recognizes DNA lesions and recruits the nucleotide excision repair factors. 52 DDB2 is then ubiquitylated and targeted for degradation by the CUL4-RING E3 ubiquitin ligase complex. EZH2 localizes at DNA damage sites in SCLC cells and directly binds to DDB2. This interaction promotes the stabilization of the DDB1–DDB2 complex and enhances DNA repair, contributing to cisplatin resistance of SCLC. 52
EZH2 non-canonical role in the activation of gene expression
EZH2 methyltransferase-independent activity can be also executed after binding general transcriptional regulators or specific transcription factors (Table 2). These interactions are highly context dependent and mediated by distinct mechanisms. In these interactions, different domains of EZH2 can participate, including TAD,59,63,66 central region with C-X-C cysteine-rich domain, 60 and carboxyl end with pre-SET and SET domains (Figure 5(a)). 58
Figure 5.
The non-canonical role of EZH2 in the co-activation of transcription. (a) Schematic representation of EZH2 functional domains. Gray ovals represent various kinases responsible for specific EZH2 phosphorylation. (b–e) To enhance gene expression, EZH2 can bind general transcriptional regulators or specific transcription factors. (b) Interaction of EZH2 with MYCN stimulated by CDK1-dependent EZH2 phosphorylation at T487 in PTCL leads to MYCN-dependent gene upregulation, including EZH2. (c) The binding of EZH2 to RNA polymerase II leads to overexpression of ITGB1, which enhances TGFβ signaling in breast cancer. JAK3-mediated EZH2 phosphorylation at Y244 supports EZH2 binding to RNA polymerase II in NKTL, whereas SRC-dependent phosphorylation of EZH2 at Y696 contributes to c-JUN upregulation. (d) Context-dependent interplay between EZH2 and NF-κB. In ER-positive breast cancer, EZH2 represses NF-κB target genes through the canonical pathway, whereas in the absence of ER EZH2 stimulates NF-κB target genes such as IL-6 and TNF, which through a positive feedback loop causes constitutive activation of NF-κB signaling. (e) Interaction of EZH2 with E2F1 facilitates the self-activation of E2F1 expression, which, in turn, upregulates CDCA8 in prostate cancer.
AKT, serine/threonine kinase; CCND1, gene encoding cyclin D1; CDCA8, cell division cycle-associated 8; CDK1, cyclin-dependent kinase 1; CXC, cysteine-rich domain; E2F1, E2F transcription factor 1; EBD, EED binding domain; ER, estrogen receptor; EZH2, enhancer of zeste homolog 2; IL6, interleukin 6; IL8, interleukin-8; ITGB1, integrin β; JAK3, Janus kinase 3; KRAS, KRAS proto-oncogene; GTPase; MYCN, MYCN proto-oncogene, bHLH transcription factor; NF-κB, nuclear factor kappa B; NKTL, natural killer/T-cell lymphoma; NLS, nuclear localization sequence; P, phosphate; Pol II, RNA polymerase II; PTCL, peripheral T-cell lymphoma; RAD51C, DNA repair protein RAD51 homolog C; S, serine; SRC, SRC non-receptor tyrosine kinase; T, threonine; TAD, trans-activating domain; TGFβ, transforming growth factor; TNFα, tumor necrosis factor alpha; Y, tyrosine.
EZH2 can interact with general transcriptional regulators and specific transcription factors to enhance gene expression (Figure 5(b)–(e)). It has been demonstrated that phosphorylated EZH2 with active TAD can directly interact with transcriptional co-activator p300 to stimulate gene expression. 47 Active TAD is used to recruit a co-activator to enhance MYC-dependent genes in acute myeloid leukemia (AML) and mixed lineage leukemia (MLL) cell lines. 53 EZH2 has been also identified as a transcriptional co-activator for MYCN-dependent gene expression in peripheral T-cell lymphoma, including expression of EZH2 itself 54 (Figure 5(b)). A direct interaction between EZH2 and MYCN is stimulated by cyclin-dependent kinase 1-dependent EZH2 phosphorylation at T487. 54
EZH2 has been identified to function as a binding partner of RNA polymerase II (Figure 5(c)) to upregulate transcription of selected genes in breast cancer 55 and natural killer/T-cell lymphoma (NKTL) cell lines, 56 and to trigger melanoma brain metastases. 57 In breast cancer cells, the EZH2-polymerase II cooperation enhances the expression of ITGB1 encoding integrin β1, which leads to activation of the focal adhesion kinase and the transforming growth factor beta pathway promoting bone metastases. 55 The interaction of EZH2 with polymerase II in NKTL cells, supported by Janus kinase 3-mediated EZH2 phosphorylation at tyrosine 244 (Y244), upregulates genes involved in DNA replication, cell cycle, stemness, and invasion. 56 EZH2 phosphorylated at tyrosine 696 (Y696) by SRC non-receptor tyrosine kinase binds RNA polymerase II and induces c-JUN expression that promotes the expression of protumorigenic inflammatory cytokines that recruit immunosuppressive neutrophils to support melanoma brain metastases. 57 A complex of EZH2 with the subunits of chromatin remodeler, tripartite motif containing 28, and SWI/SNF positively regulates a set of genes associated with carcinogenesis, metastasis, and stemness as shown in ER-positive breast cancer cell lines. 58
The activation of the Wnt pathway can be achieved by direct interaction of EZH2 with β-catenin and estrogen receptor α (ERα) in breast cancer cells. 59 PCNA-associated factor (PAF) forms a complex with EZH2 and β-catenin to support Wnt signaling in colon cancer cells. 60 EZH2 directly interacts with β-catenin contributing to hepatoblastoma pathogenesis. 61 Interaction of EZH2 with heterochromatin protein 1 binding protein 3 (HP1BP3) in glioma stem cells epigenetically activates Wnt/β-catenin signaling through upregulation of the Wnt family member 7B (WNT7B). 62
In cancer, EZH2 can modulate the NF-κB pathway in a context-dependent manner. 69 In ER-negative basal-like breast cancer cells, EZH2 plays a transactivation non-canonical function by forming the ternary complex with a dimer RelA/RelB to support the constitutive expression of NF-κB target genes such as tumor necrosis factor and IL-6 (Figure 5(d)). 64 In ER-positive luminal-like breast cancer cells, ER recruits PRC2 to NF-κB target gene promoters, and expression of RelB and NF-κB-dependent genes is repressed by canonical EZH2 activity. 64 Thus, mutually exclusive occupancy of NF-κB target gene promoters associated with either activating or repressive function of EZH2 might require a context-specific therapeutic strategy for targeting EZH2 in breast cancer. In addition, the migratory and stem-like properties of breast cancer cells can be enhanced by EZH2 interactions with forkhead box M1 (FOXM1) transcription factor, which is supported by high levels of hypoxia-inducible factor subunit alpha in TNBC cells. 65
The molecular context determining the function of EZH2 has been also found in PCa. 70 EZH2 can occupy and induce gene expression from the promoters marked with H3K27 acetylation (H3K27ac), in contrast to promoters bound by PRC2 and marked with H3K27me3. This indicates that the local chromatin structure can determine the EZH2 activity. 70 EZH2 can activate the transcription of AR, and inhibitors of EZH2 combined with AR antagonists exert a strong synergy in suppressing PCa in vitro and in vivo. 70 EZH2 with active TAD can bind AR but also its constitutively active splice variant AR-V7. 66 EZH2 can enhance transcription of E2F1 by recruiting E2F1 to its promoter. An elevated level of E2F1, in turn, upregulates the cell division cycle-associated 8, a subunit of the chromosomal passenger complex, which plays an important role in mitosis and positively correlates with poor clinical outcomes of PCa patients (Figure 5(e)). 67 In addition to the non-canonical activities of EZH2 in the co-activation of transcription, its role in the regulation of translation in PCa cells has been demonstrated as a result of the direct functional interaction between EZH2 and fibrillarin, a SAM-dependent methyltransferase involved in pre-rRNA processing. 68
EZH2 inhibitors and degraders
EZH2 is considered an attractive therapeutic target for pharmacological interventions in cancer. Based on their mechanisms of action, EZH2 inhibitors can be categorized as (1) inhibiting EZH2 methyltransferase activity, (2) disrupting the interaction between PRC2 components, and (3) causing EZH2 degradation.
Inhibitors of EZH2 methyltransferase activity
The EZH2 methyltransferase activity can be inhibited by S-adenosylhomocysteine (SAH) accumulation or by competing with the methyl donor SAM for the binding pocket. S-adenosylhomocysteine hydrolase inhibitors such as 3-deazaneplanocin A (DZNep), first investigated in breast and colon cancer cell lines, 71 belong to inhibitors of EZH2 methyltransferase activity that cause SAH accumulation, which, in turn, inhibits SAM-mediated methyl transfer. It is not specific for EZH2 as other methyltransferases are also dependent on SAM as a methyl donor. Interestingly, a combination of DZNep and AC1Q3QWB, a selective disruptor of HOTAIR-EZH2 interaction, has been found more efficient than either agent alone in killing tumor cells in breast cancer and glioblastoma patient-derived xenograft models. 72
A number of highly selective EZH2 inhibitors, including GSK126, 73 EPZ-6438 (tazemetostat), 74 CPI-1205 (lirametostat), 75 PF-06821497, 76 and SHR2554, 77 directly inhibiting methyltransferase activity of EZH2 by occupying its SAM-binding site yielded promising results in preclinical studies and some of them entered clinical trials for the treatment of diverse cancer types (Table 3). While numerous clinical trials have been initiated or completed, only tazemetostat 74 has been approved by the Food and Drug Administration (FDA) for cancer treatment, specifically for epithelioid sarcoma and follicular lymphoma.78,79 Tazemetostat competitively binds to the SET domain of EZH2 to prevent binding of SAM. The above-mentioned AC1Q3QWB, an agent targeting HOTAIR-EZH2 interaction, has been recently shown to increase the efficacy of tazemetostat in inhibiting proliferation and migration of endometrial cancer cells, 80 whereas in combination with GSK-LSD1 lysine-specific demethylase 1 inhibitor, AC1Q3QWB has shown the enhanced anticancer activity in glioblastoma patient-derived xenograft models. 81 Results of the study performed on CRPC cells, but also referred to public data on diverse types of solid tumors with wild-type EZH2, have revealed that expression of DNA damage repair (DDR) genes, especially those encoding elements of the base excision repair pathway, is significantly correlated with tumor sensitivity to EZH2 inhibitors. 82 Mechanistically, EZH2 inhibitors block methylation of FOXA1, and this suppression of the non-canonical transactivation function of EZH2 contributes to reduced expression of DDR genes. In this way, EZH2 inhibitors may enhance the sensitivity of solid tumors to genotoxic stress, such as chemotherapy or radiation. Another example of preferential sensitivity to EZH2 inhibition has been reported recently in the context of ARID1A LoF mutations in bladder, ovarian, and endometrial tumor models. 83 Tulmimetostat (CPI-0209), an orally available, clinical-stage inhibitor of EZH1 and EZH2, which induces selected changes in gene expression in addition to an extensive reduction in H3K27me3 levels, has been found efficacious alone or in combination with cisplatin. 83 Inhibition of EZH2 acting in a synthetic lethal manner in ARID1A-mutated tumors has been previously shown as causing downregulation of PI3K/AKT signaling in ovarian clear cell carcinoma. 84 Mutation rates in the gene encoding a chromatin remodeler ARID1A is one of the highest across several tumor types. Therefore, EZH2 inhibition might be a potential therapeutic strategy for solid tumors with ARID1A LoF mutations.
Table 3.
Clinical trials of inhibitors of EZH2 histone methyltransferase activity alone or in combination with other anticancer agents.
Drug | Clinical trial number | Company | Drug action | Condition/disease | Phase | Enrollment | Status | References |
---|---|---|---|---|---|---|---|---|
Tazemetostat | NCT02601950 | Epizyme, Inc. | EZH2 HMT activity inhibitor | ES; solid tumors with EZH2 GoF mutation | II | 267 participants | Completed, no results posted | N/A |
Tazemetostat | NCT02860286 | Epizyme, Inc. | EZH2 HMT activity inhibitor | Malignant mesothelioma | II | 74 participants | Completed, results posted | Zauderer et al. (2022) 85 |
Tazemetostat | NCT03155620 | National Cancer Institute | EZH2 HMT activity inhibitor | Pediatric relapsed/refractory advanced solid tumors and (N)HL | II | 2316 participants | Recruiting, no results posted | Parsons et al. (2022) 86 |
Tazemetostat | NCT02601937 | Epizyme, Inc. | EZH2 HMT activity inhibitor | Relapsed/refractory INI1-negative tumors or synovial sarcoma |
I | 109 participants | Completed, results submitted | N/A |
Tazemetostat | NCT04917042 | University of Florida | EZH2 HMT activity inhibitor | Peripheral nerve sheath tumor | II | 24 participants | Recruiting, no results posted | N/A |
Tazemetostat | NCT03456726 | Eisai Co., Ltd | EZH2 HMT activity inhibitor | Relapsed/refractory (N)HL with EZH2 mutation | II | 20 participants | Completed, results posted | N/A |
Tazemetostat | NCT03213665 | National Cancer Institute | EZH2 HMT activity inhibitor | Relapsed/refractory advanced solid tumors and (N)HL with EZH2 mutation | II | 20 participants | Active, not recruiting, results posted | N/A |
Tazemetostat | NCT05983965 | University of Alabama at Birmingham | EZH2 HMT activity inhibitor | TCL | I | 30 participants | Not yet recruiting, no results posted | N/A |
Tazemetostat | NCT06068881 | Epizyme, Inc. | EZH2 HMT activity inhibitor | Relapsed/refractory FL | II | 55 participants | Not yet recruiting, no results posted | N/A |
Tazemetostat | NCT05467943 | Hutchison Medipharma Limited | EZH2 HMT activity inhibitor | Relapsed/refractory FL | II | 39 participants | Recruiting, no results posted | N/A |
Tazemetostat | NCT03028103 | Epizyme, Inc. | EZH2 HMT activity inhibitor | Advanced solid tumors and B-cell lymphomas | I | 32 participants | Completed, results posted | N/A |
Tazemetostat | NCT05023655 | Prisma Health-Upstate | EZH2 HMT activity inhibitor | Solid tumors with ARID1A mutation | II | 40 participants | Recruiting, no results posted | N/A |
Tazemetostat + mosunetuzumab | NCT05994235 | Weill Medical College of Cornell University | EZH2 HMT activity inhibitor + anti-CD20/CD3 monoclonal antibody | Untreated FL | II | 50 participants | Recruiting, no results posted | N/A |
Tazemetostat + talazoparib | NCT04846478 | Dana-Farber Cancer Institute | EZH2 HMT activity inhibitor + PARP inhibitor | CRPC | I | 35 participants | Active, not recruiting, no results posted | N/A |
Tazemetostat + prednisolone | NCT01897571 | Epizyme, Inc. | EZH2 HMT activity inhibitor + corticosteroid | Relapsed/refractory FL, DLBCL, advanced solid tumors | I/II | 400 participants | Completed, results posted | Italiano et al. (2018)
87
Morschhauser et al. (2020) 79 |
Tazemetostat + rituximab + bendamustine | NCT05551936 | Vaishalee Kenkre; Epizyme, Inc.; University of Wisconsin, Madison |
EZH2 HMT activity inhibitor + anti-CD20 monoclonal antibody + DNA alkylating agent | FL | I/II | 42 participants | Recruiting, no results posted | N/A |
Tazemetostat + rituximab + lenalidomide | NCT04224493 | Epizyme, Inc. | EZH2 HMT activity inhibitor + anti-CD20 monoclonal antibody + proteasomal degrader | Relapsed/refractory FL | III | 612 participants | Recruiting, no results posted | N/A |
Tazemetostat + tafasitamab + lenalidomide | NCT05890352 | SWOG Cancer Research Network | EZH2 HMT activity inhibitor + anti-CD19 monoclonal antibody + proteasomal degrader | Relapsed/non-responsive large BCL | II | 227 participants | Recruiting, no results posted | N/A |
Tazemetostat + CPX-351 | NCT05627232 | Thomas Jefferson University | EZH2 HMT activity inhibitor + liposome-encapsulated daunorubicin and cytarabine | Relapsed/refractory AML | I | 24 participants | Recruiting, no results posted | N/A |
Tazemetostat + enzalutamide | NCT04179864 | Epizyme, Inc. | EZH2 HMT activity inhibitor + AR antagonist | CRPC | Ib/II | 102 participants | Active, not recruiting | N/A |
Tazemetostat + doxorubicin | NCT04204941 | Epizyme, Inc. | EZH2 HMT activity inhibitor + topoisomerase poison | Advanced ES | III | 164 participants | Active, recruiting, no results posted | Gounder, et al. (2020) 78 |
Tazemetostat + pembrolizumab | NCT03854474 | National Cancer Institute | EZH2 HMT activity inhibitor + ICI | Advanced UC | I/II | 30 participants | Active, not recruiting, no results posted | N/A |
Tazemetostat + pembrolizumab | NCT05467748 | VA Office of Research and Development | EZH2 HMT activity inhibitor + ICI | Advanced NSCLC | I/II | 66 participants (estimated) | Not yet recruiting, no results posted | N/A |
Tazemetostat + belinostat | NCT05627245 | National Cancer Institute | EZH2 HMT activity inhibitor + HDAC inhibitor | Drug-resistant lymphomas | I | 64 participants | Recruiting no results posted | N/A |
Tazemetostat + dabrafenib + trametinib | NCT04557956 | National Cancer Institute | EZH2 HMT activity inhibitor + BRAFV600 and MEK inhibitors | Metastatic melanoma | I/II | 58 participants | Recruiting, no results posted | N/A |
CPI-1205 | NCT02395601 | Constellation Pharmaceuticals | EZH2 HMT activity inhibitor | BCL | I | 41 participants | Completed, no results posted | N/A |
CPI-1205 + ipilimumab | NCT03525795 | Constellation Pharmaceuticals | EZH2 HMT activity inhibitor + ICI | Advanced solid tumors | I | 24 participants | Completed, no results posted | N/A |
CPI-1205 + enzalutamide or abiraterone/prednisone | NCT03480646 | Constellation Pharmaceuticals | EZH2 HMT activity inhibitor + AR inhibitors and immunosuppressants | Metastatic CRPC | I | 242 participants | Unknown, results submitted | N/A |
CPI-0209 | NCT04104776 | Constellation Pharmaceuticals | Dual EZH2/EZH1 small-molecule inhibitor | Advanced solid tumors DLBCL, T-Cell Mesothelioma, CRPC, EC, OCCC |
I/II | 210 participants | Recruiting, no results posted | N/A |
CPI-0209 + carboplatin | NCT05942300 | Lan Coffman, University of Pittsburgh | Dual EZH2/EZH1 small-molecule inhibitor + DNA-damaging agent | Recurrent ovarian cancer | I | 30 participants | Recruiting, no results posted | N/A |
PF-06821497 | NCT03460977 | Pfizer | EZH2 HMT activity inhibitor | Relapsed/refractory SCLC, CRPC, and FL | I | 267 participants | Recruiting, no results posted | N/A |
PF-06821497 | NCT06392230 | Pfizer | EZH2 HMT activity inhibitor | Healthy participants | I | 6 participants | Not yet recruiting, no results posted | N/A |
MAK683 | NCT02900651 | Novartis Pharmaceuticals | EED inhibitor | DLBCL | I/II | 139 participants | Active, not recruiting, no results posted | N/A |
Valemetostat | NCT06294548 | University of Alabama at Birmingham | Dual EZH2/EZH1 small-molecule inhibitor | Hepatocellular carcinoma | I/II | 45 participants (estimated) | Not yet recruiting, no results posted | N/A |
Valemetostat | NCT04842877 | The Lymphoma Academic Research Organisation | Dual EZH2/EZH1 small-molecule inhibitor | BCL | II | 141 participants | Active, not recruiting, no results posted | N/A |
SHR2554 | NCT03603951 | Jiangsu HengRui Medicine Co., Ltd | EZH2 HMT activity inhibitor | Relapsed/refractory mature lymphoid neoplasms | I | 231 participants | Unknown, no results posted | Song et al. (2022) 77 |
SHR2554 + SHR3162 | NCT04355858 | Fudan University | EZH2 HMT activity inhibitor + PARP inhibitor | Metastatic breast cancer | II | 319 participants | Recruiting, no results posted | Chen (2021) 88 |
SHR2554 + SHR1701 | NCT05896046 | Chinese PLA General Hospital | EZH2 HMT activity inhibitor + bifunctional fusion protein targeting PD-L1 and TGFβ | Relapsed/refractory HL | I/II | 100 participants | Recruiting, no results posted | N/A |
SHR2554 + SHR1701 | NCT04407741 | Chinese PLA General Hospital | EZH2 HMT activity inhibitor + bifunctional fusion protein targeting PD-L1 and TGFβ | Advanced solid tumors and BCL | I/II | 100 participants | Recruiting, no results posted | N/A |
AXT-1003 | NCT05965505 | Axter Therapeutics (Beijing) Co., Ltd | EZH2 HMT activity inhibitor | Relapsed/refractory (N)HL | I | 79 participants | Recruiting, no results posted | N/A |
XNW5004 + pembrolizumab | NCT06022757 | Evopoint Biosciences Inc. | EZH2 HMT activity inhibitor + ICI | Advanced solid tumors that failed standard treatments | I/II | 204 participants | Recruiting, no results posted | N/A |
Source: Data obtained from ClinicalTrials.org site (as of June 14, 2024).
Clinical trials labeled as terminated, withdrawn, or with unknown status or no longer available have been ruled out.
AML, acute myeloid leukemia; AR, androgen receptor; BCL, B-cell lymphoma; CRPC, castration-resistant prostate cancer; DLBCL, diffuse large B-cell lymphoma; EC, endometrial cancer; EED, embryonic ectoderm development; ES, epithelioid sarcoma; EZH2, enhancer of zeste homolog 2; FL, follicular lymphoma; GoF, gain of function; HC, hepatocellular carcinoma; HDAC, histone deacetylase; HMT, enhanced histone methyltransferase; ICI, immune checkpoint inhibitor; MM, multiple myeloma; N/A, not available; (N)HL, (non-)Hodgkin lymphoma; NSCLC, non-small-cell lung carcinoma; OC, ovarian cancer; OCCC, ovarian clear cell carcinoma; PARP, poly-ADP ribose polymerase; SCLC, small cell lung cancer; TCL, T-cell lymphoma; TCM, T-cell mesothelioma; TGFβ, transforming growth factor beta; UC, urothelial carcinoma.
Comparative analysis of the impact of EZH2 inhibitors on the pattern of posttranslational histone modifications between EZH2 sensitive and insensitive solid tumor cell lines has revealed that while H3K27me3 levels were similarly reduced by EZH2 inhibitors, H3K27 acetylation (H3K27ac) levels were markedly increased only in the insensitive cells. 89 Mechanistically, the MLL1 forms a complex with p300/CBP to facilitate histone acetylation (H3K27ac) in response to EZH2 inhibitors, and H3K27ac-associated transcriptional output contributes to the resistance to EZH2 inhibitors in solid tumors. The efficacy of EZH2 inhibitors can be improved by H3K27ac co-inhibition but it leads to oncogenic reprogramming such as MAPK activation in some cancers. This, in turn, highlights the need for a combination therapy after the patient stratification based on tumor-intrinsic expression of MLL1. The authors suggest that the combination therapy might expand the potential of EZH2 inhibitors from hematological cancers to solid tumors. 89 These and several other reports indicate that the efficacy of EZH2 inhibitors in solid tumors is highly context dependent, and overexpression of EZH2 is only one of the determining factors.
Tazemetostat has been evaluated in clinical trials in combination with drugs already approved for cancer treatment (Table 3). For example, a combination of tazemetostat with dabrafenib (BRAFV600 inhibitor) and trametinib (MEK inhibitor) has been evaluated in a phase I/II study in BRAF-mutant melanoma patients who progressed on BRAF/MEK inhibitor therapy (NCT04557956). Tazemetostat in combination with talazoparib, a poly-ADP ribose polymerase (PARP) inhibitor, has been evaluated in metastatic CRPC (NCT04846478), and the use of this combination is considered as one of the potential synergies between different molecular pathways. 90
EZH2 plays a critical role not only in cancer cells but also in cells of the tumor microenvironment. 91 It has been demonstrated that the modulation of EZH2 expression reprograms intratumoral regulatory T cells 92 and improves the efficacy of anti-CTLA-4 therapy. 93 Therefore, the EZH2 inhibition is considered a promising modality to enhance current immunotherapies. 94 Tazemetostat and other inhibitors of EZH2 methyltransferase activity have been combined with various drugs modulating immune response such as pembrolizumab (e.g., NCT03854474, NCT05467748, and NCT06022757) or ipilimumab (NCT03525795) in different types of cancers (Table 3).
Dual-acting EZH2 inhibitors
Several research groups have applied the concept of synthetic lethality to expand the EZH2 inhibitory activity by adding other anticancer pharmacophores to existing EZH2 inhibitory templates, thus creating dual-acting inhibitors. 95 Dual HDAC and EZH2 inhibitors have displayed a capacity to decrease cancer cell viability at micromolar concentrations in several cancer cell lines. 96 ZLD-2 designed as an inhibitor of EZH2 and bromodomain-containing protein 4 (BRD4) has shown potent antiproliferative activity against cells of solid tumors, including bladder, breast, lung, and pancreatic tumors. 97 The dual-acting inhibitor that can diminish the EZH2 methylation activity and affect the AR level has been found effective against PCa cells, including those resistant to enzalutamide. 98 The dual inhibitor of PARP and EZH2 activities has shown higher inhibitory potential against TNBC cells than the combination of olaparib and tazemetostat. 99 One of the recently synthesized dual-target PARP/EZH2 inhibitors has demonstrated a capacity to induce excessive autophagy and inhibit TNBC growth. 100 The dual inhibitor of HMTs G9A and EZH2 has demonstrated the capacity to induce chromatin changes followed by the transcriptional activation of the immunostimulatory gene network modulating the immune microenvironment and extending the survival of mice with ovarian cancer. 101 Most recently, a novel dual inhibitor of EZH2 and heat shock protein 90 has exerted promising activity against temozolomide-resistant glioblastoma. 102
Agents disrupting PRC2 composition
Knowing the dependence of EZH2 activity on the scaffolding provided by EED, various inhibitors of the EZH2-EED interaction have been developed, recently reviewed in the paper of Bao et al. 103 Astemizole, a drug approved by the FDA, has been identified as an agent destabilizing the PRC2 complex in cancer cells by the disruption of the EZH2-EED interaction. 104 An agent EED226 can form with EED-EZH2, a ternary complex capable of inhibiting a PRC2 activity, and also PRC2 containing a mutant EZH2 protein resistant to SAM-competitive inhibitors. 105 MAK683, another selective inhibitor of EED, has been found active in cancer cells with EZH2-mutation-driven acquired resistance to EZH2 methyltransferase inhibitors. 106 Due to its balanced pharmacokinetics/pharmacodynamics profile and efficacy, it has been selected for clinical development and it is currently in phase I/II study for the treatment of patients with diffuse large B-cell lymphoma (DLBCL) and advanced solid tumors (NCT02900651).
EZH2 degraders
The studies showing the ineffectiveness of inhibitors of EZH2 catalytic activity in several cancer types but cell sensitivity to a complete depletion of EZH2 protein provide the rationale for a pharmacological degradation of EZH2 as a therapeutic strategy for treating EZH2-dependent cancers. Inhibitors that covalently bind cysteine residues in the SET domain of EZH2 can cause an inability to form disulfide bonds resulting in inappropriate folding that triggers EZH2 degradation by carboxyl terminus of Hsp70-interacting protein-mediated polyubiquitination. Two such compounds, gambogenic acid derivative GNA002 that binds C668 of EZH2 and IHMT-337 that binds C663, have shown specific cytotoxic effects in head and neck cancer cell lines and mouse xenografts, 107 and TNBC and DLBCL cell lines and xenografts, 108 respectively. Most recently, a novel EZH2 degrader with the capacity to form the covalent bonding with C663 in the SET domain of EZH2 and potent antiproliferation activity in B-cell lymphoma and TNBC cell lines has been reported. 109
Using targeted protein degradation (TPD) technology that has gradually become widespread and potentially beneficial in terms of therapy,110,111 specific EZH2 degraders have been developed to suppress all types of EZH2 activities in cancer. A first-in-class EZH2 selective degrader, hydrophobic tag-based MS1943, can induce TNBC cell death associated with the activation of the unfolded protein response pathway and prolonged endoplasmic reticulum stress. 112 Bifunctional degraders, also known as proteolysis-targeting chimeras (PROTACs), contain a specific ligand for a distinct E3 ubiquitin ligase, and another ligand specific for a target protein, and an optimized chemical linker connecting the two.113,114 Compared to small inhibitors of enzymatic activity that need to steadily occupy an active site of the enzyme, PROTACs by binding to a protein of interest trigger its irreversible degradation. Although approximately 600 different E3 ligases have been identified in human cells, only a few of them, including cereblon (CRBN), von Hippel-Lindau (VHL), and murine double minute 2, have been used to design PROTACs against oncogenes.115–118 VHL-recruiting compounds named YM181 and YM281 designed to degrade EZH2 have shown potent anticancer activity in various subtypes of lymphomas. 119 Other EZH2 degraders recruiting VHL, MS8815, and MS8847 have effectively inhibited the growth of TNBC and AML cells.120,121 Three other PROTACs such as E7, 122 U3i, 99 and MS17753 have been designed to cooperate with CRBN E3 ligase. U3i exerts a high specificity toward EZH2 degradation in TNBC cells while sparing normal cells. 99 E7 and MS177 have shown wider specificity. E7 is also capable of degrading EED and SUZ12 in various cancer cell lines. 122 MS177, in turn, can effectively deplete both canonical EZH2-PRC2 and noncanonical EZH2-cMYC complexes leading to suppression of cancer cell proliferation. 53 In multiple myeloma cells, MS177 has shown the capacity to reactivate PRC2-repressed genes and genes associated with the immune response with the simultaneous suppression of c-MYC-associated oncogenes. 123
The above reports suggest that EZH2 can be targeted for degradation using different approaches (Figure 6). Several EZH2 degraders, their mode of action, and tumor specificity, are summarized in Table 4.
Table 4.
EZH2 degraders investigated in various cancers.
Compound | Mode of action | Tumor samples | Reference |
---|---|---|---|
GNA002 | CHIP-mediated ubiquitination | Head and neck cancer, breast cancer, and hepatocarcinoma cell lines; head and neck mouse xenografts | Wang et al. (2017) 107 |
IHMT-337 | CHIP-mediated ubiquitination | TNBC and lymphoma cell lines; lymphoma mouse model | Mei et al. (2023) 108 |
MS1943 | Hydrophobic tag-mediated degradation | Breast cancer cell lines; breast cancer mouse xenografts | Ma et al. (2020) 112 |
YM181 YM281 |
VHL-mediated ubiquitination | Human prostate, colon, and DLBCL cancer cell lines, and mouse xenografts | Tu et al. (2021) 119 |
MS8815 MS8847 |
VHL-mediated ubiquitination | TNBC cell lines and mouse xenografts leukemia, AML, and TNBC cell lines |
Dale et al. (2022), 120 Velez et al. (2024) 121 |
E7 | CRBN-mediated ubiquitination | Prostate, ovarian, NSCLC, TNBC, and DLBCL cell lines | Liu et al. (2021) 122 |
U3i | CRBN-mediated ubiquitination | TNBC and leukemia cell lines | Wang et al. (2022) 99 |
MS177 | CRBN-mediated ubiquitination | Leukemia, lymphoma, neuroblastoma, cervical cancer, and hematopoietic stem cell lines; leukemia and lymphoma mouse models |
Wang et al. (2022), 53 Yu et al. (2023) 123 |
AML, acute myeloid leukemia; CHIP, co-chaperone C terminus of Hsp70-interacting protein; CRBN, cereblon; DLBCL, diffuse large B-cell lymphoma; EZH2, enhancer of zeste homolog 2; NSCLC, non-small-cell lung cancer; TNBC, triple-negative breast cancer; VHL, von Hippel-Lindau.
Figure 6.
Mechanism of action of EZH2 degraders. (a) Agents that bind the sulfhydryl groups in the SET domain of EZH2 trigger (CHIP)-mediated ubiquitination and degradation of EZH2. (b) PROTAC which consists of ligand-binding EZH2 and ligand for E3 ligase, both connected with a linker, mediates polyubiquitylation and proteasomal degradation of EZH2. (c) Hydrophobic tagging of EZH2 triggers destabilization of the protein followed by degradation. After each cycle of reaction, EZH2 degraders are being regenerated.
CHIP, co-chaperone C terminus of Hsc70-interacting protein; E2/E3, E2/E3 ligases; EZH2, enhancer of zeste homolog 2; PROTACs, proteolysis-targeting chimeras; SET, Su(var)3-9, Enhancer-of-zeste, and Trithorax; SH, sulfhydryl group; Ub, ubiquitin.
The challenges of targeting EZH2 in cancer
Numerous small-molecule agents that inhibit EZH2 methyltransferase activity or block the formation of active PRC2 complex have been examined in various types of cancer in preclinical settings and clinical trials. Unfortunately, the inhibition of histone methylating activity of EZH2 fails to diminish the progression of several types of cancer. The target residence time of EZH2 inhibitors might be one of the factors influencing their cellular efficacy. 124 In addition, the emergence of resistant clones with acquired mutations at the interface of PRC2-inhibitor or in crucial epigenome factors with cooperative function in gene silencing125–128 is an inevitable limitation of durable clinical response. A detailed analysis of resistance mechanisms developed in the patients with adult T-cell leukemia/lymphoma enrolled for the valemetostat clinical trials has been published most recently. 129
As EZH2 inhibitors that only inhibit the enzymatic activity of EZH2 may fail to suppress its non-canonical methylation-independent functions, selective EZH2 degraders are being developed. Several drawbacks and challenges must be addressed to increase the effectiveness of TPD technology. Some cancer subtypes might be insensitive to EZH2 degradation. For instance, while MDA-MB-231 breast cancer cell line is resistant to EZH2 degradation, BT549 and MDA-MB-468 cell lines are sensitive. 112 The final effects of this treatment are determined by the basal EZH2 levels and cell dependencies on EZH2 activity. The context-dependent interactions between EZH2 and non-coding RNAs may also affect EZH2 activities.5,130 Other aspects that need to be considered include the limited degradation scope, low water solubility of degraders, and overall toxicity of EZH2 degrading agents. It also becomes clear that the identification of E3 ligases with higher cancer cell specificity is crucial for the effectiveness of TPD technology, and this issue is largely unexplored considering the large number of E3 ligases. Prolonged depletion of EZH2 can alter the epigenetic status leading to tumor progression as shown for glioblastoma. 131 As these effects have not been observed after the short-term depletion of EZH2, precise dosing regimens of anti-EZH2 agents in the clinic might be crucial for the effectiveness of treatment. Therefore, while TPD technology is recognized as a therapeutic approach with great clinical potential, several unresolved problems require further studies.
Conclusion
This review gathers recent knowledge regarding non-canonical multifaceted functions of EZH2 in cancer development and progression. EZH2-mediated PRC2-dependent and PRC2-independent methylation of non-histone proteins, and methyltransferase-independent interactions of EZH2 with various non-histone proteins leading to gene activation or repression, and influencing protein turnover, are highlighted. More research is necessary to identify non-canonical functionalities of EZH2 in a particular cancer type, and possible synthetic lethality between inhibition of EZH2 and other epigenomic abnormalities or cancer type-specific alterations in signaling pathways. It is also important to elucidate the still not well-recognized complexity of disturbances caused by the inhibition of aberrant EZH2 activity since EZH2 inhibition can also promote cancer. All this knowledge is crucial for the development of treatment strategies more specific for each malignancy. Considering the limited efficacy of epigenetic therapies and the risk of drug resistance, EZH2 degrading agents used alone or in combination with therapy targeting cancer-type specific protein(s) and/or immunotherapies have become more and more attractive.
Acknowledgments
None.
Footnotes
ORCID iDs: Michal Wozniak
https://orcid.org/0000-0003-2251-8647
Malgorzata Czyz
https://orcid.org/0000-0002-8279-4742
Contributor Information
Michal Wozniak, Department of Molecular Biology of Cancer, Medical University of Lodz, Lodz, Poland.
Malgorzata Czyz, Department of Molecular Biology of Cancer, Medical University of Lodz, Mazowiecka 6/8, Lodz 92-215, Poland.
Declarations
Ethics approval and consent to participate: Not applicable.
Consent for publication: Not applicable.
Author contributions: Michal Wozniak: Writing – original draft.
Malgorzata Czyz: Writing – review & editing.
Funding: The authors received no financial support for the research, authorship, and/or publication of this article.
The authors declare that there is no conflict of interest.
Availability of data and materials: Not applicable.
References
- 1. Margueron R, Reinberg D. The Polycomb complex PRC2 and its mark in life. Nature 2011; 469: 343–349. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2. Chammas P, Mocavini I, Di Croce L. Engaging chromatin: PRC2 structure meets function. Br J Cancer 2020; 122: 315–328. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3. Sanulli S, Justin N, Teissandier A, et al. Jarid2 methylation via the PRC2 complex regulates H3K27me3 deposition during cell differentiation. Mol Cell 2015; 57: 769–783. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4. Mirzaei S, Gholami MH, Hushmandi K, et al. The long and short non-coding RNAs modulating EZH2 signaling in cancer. J Hematol Oncol 2022; 15: 18. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5. Wozniak M, Czyz M. lncRNAs-EZH2 interaction as promising therapeutic target in cutaneous melanoma. Front Mol Biosci 2023; 10: 1170026. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6. Wiles ET, Selker EU. H3K27 methylation: a promiscuous repressive chromatin mark. Curr Opin Genet Dev 2017; 43: 31–37. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7. Jiao L, Liu X. Structural basis of histone H3K27 trimethylation by an active polycomb repressive complex 2. Science 2015; 350: aac4383. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8. Laugesen A, Højfeldt JW, Helin K. Molecular mechanisms directing PRC2 recruitment and H3K27 methylation. Mol Cell 2019; 74: 8–18. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9. Wang B, Liu Y, Liao Z, et al. EZH2 in hepatocellular carcinoma: progression, immunity, and potential targeting therapies. Exp Hematol Oncol 2023; 12: 52. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10. Varambally S, Dhanasekaran SM, Zhou M, et al. The polycomb group protein EZH2 is involved in progression of prostate cancer. Nature 2002; 419: 624–629. [DOI] [PubMed] [Google Scholar]
- 11. Xu K, Wu ZJ, Groner AC, et al. EZH2 oncogenic activity in castration-resistant prostate cancer cells is Polycomb-independent. Science 2012; 338: 1465–1469. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12. Knudsen ES, Dervishaj O, Kleer CG, et al. EZH2 and ALDH1 expression in ductal carcinoma in situ: complex association with recurrence and progression to invasive breast cancer. Cell Cycle 2013; 12: 2042–2050. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13. Li T, Cai J, Ding H, et al. EZH2 participates in malignant biological behavior of epithelial ovarian cancer through regulating the expression of BRCA1. Cancer Biol Ther 2014; 15: 271–278. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14. Zhang H, Qi J, Reyes JM, et al. Oncogenic deregulation of EZH2 as an opportunity for targeted therapy in lung cancer. Cancer Discov 2016; 6:1006–1021. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15. Béguelin W, Popovic R, Teater M, et al. EZH2 is required for germinal center formation and somatic EZH2 mutations promote lymphoid transformation. Cancer Cell 2013; 23: 677–692. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16. Suvà ML, Riggi N, Janiszewska M, et al. EZH2 is essential for glioblastoma cancer stem cell maintenance. Cancer Res 2009; 69: 9211–9218. [DOI] [PubMed] [Google Scholar]
- 17. Luo M, Yang X, Chen HN, et al. Drug resistance in colorectal cancer: an epigenetic overview. Biochim Biophys Acta Rev Cancer 2021; 1876: 188623. [DOI] [PubMed] [Google Scholar]
- 18. Zingg D, Debbache J, Schaefer SM, et al. The epigenetic modifier EZH2 controls melanoma growth and metastasis through silencing of distinct tumour suppressors. Nat Commun 2015; 6: 6051. [DOI] [PubMed] [Google Scholar]
- 19. Li F, Wang P, Ye J, et al. Serum EZH2 is a novel biomarker for bladder cancer diagnosis and prognosis. Front Oncol 2024; 14: 1303918. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20. Wang X, Zhao H, Lv L, et al. Prognostic significance of EZH2 expression in non-small cell lung cancer: a meta-analysis. Sci Rep 2016; 6: 19239. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21. Tiffen JC, Gunatilake D, Gallagher SJ, et al. Targeting activating mutations of EZH2 leads to potent cell growth inhibition in human melanoma by derepression of tumor suppressor genes. Oncotarget 2015; 6: 27023–27036. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22. Morin RD, Johnson NA, Severson TM, et al. Somatic mutations altering EZH2 (Tyr641) in follicular and diffuse large B-cell lymphomas of germinal-center origin. Nat Genet 2010; 42: 181–185. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23. Sneeringer CJ, Scott MP, Kuntz KW, et al. Coordinated activities of wild-type plus mutant EZH2 drive tumor-associated hypertrimethylation of lysine 27 on histone H3 (H3K27) in human B-cell lymphomas. Proc Natl Acad Sci U S A 2010; 107: 20980–20985. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24. Ernst T, Chase AJ, Score J, et al. Inactivating mutations of the histone methyltransferase gene EZH2 in myeloid disorders. Nat Genet 2010; 42: 722–726. [DOI] [PubMed] [Google Scholar]
- 25. Nikoloski G, Langemeijer SM, Kuiper RP, et al. Somatic mutations of the histone methyltransferase gene EZH2 in myelodysplastic syndromes. Nat Genet 2010; 42: 665–667. [DOI] [PubMed] [Google Scholar]
- 26. Ntziachristos P, Tsirigos A, Van Vlierberghe P, et al. Genetic inactivation of the polycomb repressive complex 2 in T cell acute lymphoblastic leukemia. Nat Med 2012; 18: 298–301. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27. Piunti A, Shilatifard A. The roles of Polycomb repressive complexes in mammalian development and cancer. Nat Rev Mol Cell Biol 2021; 22: 326–345. [DOI] [PubMed] [Google Scholar]
- 28. Kim JJ, Kingston RE. Context-specific Polycomb mechanisms in development. Nat Rev Genet 2022; 23: 680–695. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29. Parreno V, Martinez AM, Cavalli G. Mechanisms of Polycomb group protein function in cancer. Cell Res 2022; 32: 231–253. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30. Wang X, Long Y, Paucek RD, et al. Regulation of histone methylation by automethylation of PRC2. Genes Dev 2019; 33: 1416–1427. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31. Lee CH, Yu JR, Granat J, et al. Automethylation of PRC2 promotes H3K27 methylation and is impaired in H3K27M pediatric glioma. Genes Dev 2019; 33: 1428–1440. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32. Ardehali MB, Anselmo A, Cochrane JC, et al. Polycomb repressive complex 2 methylates elongin A to regulate transcription. Mol Cell 2017; 68: 872–884.e6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33. He A, Shen X, Ma Q, et al. PRC2 directly methylates GATA4 and represses its transcriptional activity. Genes Dev 2012; 26: 37–42. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34. Kim E, Kim M, Woo DH, et al. Phosphorylation of EZH2 activates STAT3 signaling via STAT3 methylation and promotes tumorigenicity of glioblastoma stem-like cells. Cancer Cell 2013; 23: 839–852. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35. Zhao Y, Hu Z, Li J, et al. EZH2 exacerbates breast cancer by methylating and activating STAT3 directly. J Cancer 2021; 12: 5220–5230. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36. Dasgupta M, Dermawan JK, Willard B, et al. STAT3-driven transcription depends upon the dimethylation of K49 by EZH2. Proc Natl Acad Sci U S A 2015; 112: 3985–3990. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37. Zimmerman SM, Nixon SJ, Chen PY, et al. Ezh2 Y641F mutations co-operate with Stat3 to regulate MHC class I antigen processing and alter the tumor immune response in melanoma. Oncogene 2022; 41: 4983–4993. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38. Ghobashi AH, Vuong TT, Kimani JW, et al. Activation of AKT induces EZH2-mediated β-catenin trimethylation in colorectal cancer. iScience 2023; 26: 107630. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39. Lee JM, Lee JS, Kim H, et al. EZH2 generates a methyl degron that is recognized by the DCAF1/DDB1/CUL4 E3 ubiquitin ligase complex. Mol Cell 2012; 48: 572–586. [DOI] [PubMed] [Google Scholar]
- 40. Vasanthakumar A, Xu D, Lun AT, et al. A non-canonical function of Ezh2 preserves immune homeostasis. EMBO Rep 2017; 18: 619–631. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41. Tripathi BK, Anderman MF, Bhargava D, et al. Inhibition of cytoplasmic EZH2 induces antitumor activity through stabilization of the DLC1 tumor suppressor protein. Nat Commun 2021; 12: 6941. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42. Park SH, Fong KW, Kim J, et al. Posttranslational regulation of FOXA1 by Polycomb and BUB3/USP7 deubiquitin complex in prostate cancer. Sci Adv 2021; 7: eabe2261. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43. Gonzalez ME, Naimo GD, Anwar T, et al. EZH2 T367 phosphorylation activates p38 signaling through lysine methylation to promote breast cancer progression. iScience 2022; 25: 104827. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44. Gunawan M, Venkatesan N, Loh JT, et al. The methyltransferase Ezh2 controls cell adhesion and migration through direct methylation of the extranuclear regulatory protein talin. Nat Immunol 2015; 16: 505–516. [DOI] [PubMed] [Google Scholar]
- 45. Venkatesan N, Wong JF, Tan KP, et al. EZH2 promotes neoplastic transformation through VAV interaction-dependent extranuclear mechanisms. Oncogene 2018; 37: 461–477. [DOI] [PubMed] [Google Scholar]
- 46. Song YL, Yang MH, Zhang S, et al. A GRIP-1-EZH2 switch binding to GATA-4 is linked to the genesis of rhabdomyosarcoma through miR-29a. Oncogene 2022; 41: 5223–5237. [DOI] [PubMed] [Google Scholar]
- 47. Jiao L, Shubbar M, Yang X, et al. A partially disordered region connects gene repression and activation functions of EZH2. Proc Natl Acad Sci U S A 2020; 117: 16992–17002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48. Cha TL, Zhou BP, Xia W, et al. Akt-mediated phosphorylation of EZH2 suppresses methylation of lysine 27 in histone H3. Science 2005; 310: 306–310. [DOI] [PubMed] [Google Scholar]
- 49. Cheng J, Guo J, North BJ, et al. Functional analysis of deubiquitylating enzymes in tumorigenesis and development. Biochim Biophys Acta Rev Cancer 2019; 1872: 188312. [DOI] [PubMed] [Google Scholar]
- 50. Zhu P, Wang Y, Huang G, et al. lnc-β-Catm elicits EZH2-dependent β-catenin stabilization and sustains liver CSC self-renewal. Nat Struct Mol Biol 2016; 23: 631–639. [DOI] [PubMed] [Google Scholar]
- 51. Wang L, Chen C, Song Z, et al. EZH2 depletion potentiates MYC degradation inhibiting neuroblastoma and small cell carcinoma tumor formation. Nat Commun 2022; 13: 12. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52. Koyen AE, Madden MZ, Park D, et al. EZH2 has a non-catalytic and PRC2-independent role in stabilizing DDB2 to promote nucleotide excision repair. Oncogene 2020; 39: 4798–4813. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53. Wang J, Yu X, Gong W, et al. EZH2 noncanonically binds cMyc and p300 through a cryptic transactivation domain to mediate gene activation and promote oncogenesis. Nat Cell Biol 2022; 24: 384–399. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54. Vanden Bempt M, Debackere K, Demeyer S, et al. Aberrant MYCN expression drives oncogenic hijacking of EZH2 as a transcriptional activator in peripheral T-cell lymphoma. Blood 2022; 140: 2463–2476. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55. Zhang L, Qu J, Qi Y, et al. EZH2 engages TGFβ signaling to promote breast cancer bone metastasis via integrin β1-FAK activation. Nat Commun 2022; 13: 2543. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56. Yan J, Li B, Lin B, et al. EZH2 phosphorylation by JAK3 mediates a switch to noncanonical function in natural killer/T-cell lymphoma. Blood 2016; 128: 948–958. [DOI] [PubMed] [Google Scholar]
- 57. Zhang L, Yao J, Wei Y, et al. Blocking immunosuppressive neutrophils deters pY696-EZH2-driven brain metastases. Sci Transl Med 2020; 12(545): eaaz5387. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58. Li J, Xi Y, Li W, et al. TRIM28 interacts with EZH2 and SWI/SNF to activate genes that promote mammosphere formation. Oncogene 2017; 36: 2991–3001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59. Shi B, Liang J, Yang X, et al. Integration of estrogen and Wnt signaling circuits by the polycomb group protein EZH2 in breast cancer cells. Mol Cell Biol 2007; 27: 5105–5119. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60. Jung HY, Jun S, Lee M, et al. PAF and EZH2 induce Wnt/β-catenin signaling hyperactivation. Mol Cell 2013; 52: 193–205. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61. Glaser K, Schepers EJ, Zwolshen HM, et al. EZH2 is a key component of hepatoblastoma tumor cell growth. Pediatr Blood Cancer 2024; 71: e30774. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62. Yu T, Zhou F, Tian W, et al. EZH2 interacts with HP1BP3 to epigenetically activate WNT7B that promotes temozolomide resistance in glioblastoma. Oncogene 2023; 42: 461–470. [DOI] [PubMed] [Google Scholar]
- 63. Dardis GJ, Wang J, Simon JM, et al. An EZH2-NF-κB regulatory axis drives expression of pro-oncogenic gene signatures in triple negative breast cancer. iScience 2023; 26: 107115. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64. Lee ST, Li Z, Wu Z, et al. Context-specific regulation of NF-κB target gene expression by EZH2 in breast cancers. Mol Cell 2011; 43: 798–810. [DOI] [PubMed] [Google Scholar]
- 65. Mahara S, Lee PL, Feng M, et al. HIFI-α activation underlies a functional switch in the paradoxical role of Ezh2/PRC2 in breast cancer. Proc Natl Acad Sci U S A 2016; 113: E3735–E3744. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66. Wang J, Park KS, Yu X, et al. A cryptic transactivation domain of EZH2 binds AR and AR’s splice variant, promoting oncogene activation and tumorous transformation. Nucleic Acids Res 2022; 50: 10929–10946. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67. Yi Y, Li Y, Li C, et al. Methylation-dependent and -independent roles of EZH2 synergize in CDCA8 activation in prostate cancer. Oncogene 2022; 41: 1610–1621. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68. Yi Y, Li Y, Meng Q, et al. A PRC2-independent function for EZH2 in regulating rRNA 2′-O methylation and IRES-dependent translation. Nat Cell Biol 2021; 23: 341–354. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69. Kaur P, Verma S, Kushwaha PP, et al. EZH2 and NF-κB: a context-dependent crosstalk and transcriptional regulation in cancer. Cancer Lett 2023; 560: 216143. [DOI] [PubMed] [Google Scholar]
- 70. Kim J, Lee Y, Lu X, et al. Polycomb- and methylation-independent roles of EZH2 as a transcription activator. Cell Rep 2018; 25: 2808–2820.e4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71. Tan J, Yang X, Zhuang L, et al. Pharmacologic disruption of Polycomb-repressive complex 2-mediated gene repression selectively induces apoptosis in cancer cells. Genes Dev 2007; 21: 1050–1063. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72. Li Y, Ren Y, Wang Y, et al. A compound AC1Q3QWB selectively disrupts HOTAIR-mediated recruitment of PRC2 and enhances cancer therapy of DZNep. Theranostics 2019; 9: 4608–4623. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73. McCabe MT, Ott HM, Ganji G, et al. EZH2 inhibition as a therapeutic strategy for lymphoma with EZH2-activating mutations. Nature 2012; 492: 108–112. [DOI] [PubMed] [Google Scholar]
- 74. Knutson SK, Warholic NM, Wigle TJ, et al. Durable tumor regression in genetically altered malignant rhabdoid tumors by inhibition of methyltransferase EZH2. Proc Natl Acad Sci U S A 2013; 110: 7922–7927. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75. Vaswani RG, Gehling VS, Dakin LA, et al. Identification of (R)-N-((4-Methoxy-6-methyl-2-oxo-1,2-dihydropyridin-3-yl)methyl)-2-methyl-1-(1-(1-(2,2,2-trifluoroethyl)piperidin-4-yl)ethyl)-1H-indole-3-carboxamide (CPI-1205), a potent and selective inhibitor of histone methyltransferase EZH2, suitable for phase I clinical trials for B-cell lymphomas. J Med Chem 2016; 59: 9928–9941. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76. Kung PP, Bingham P, Brooun A, et al. Optimization of orally bioavailable enhancer of zeste homolog 2 (EZH2) inhibitors using ligand and property-based design strategies: identification of development candidate (R)-5,8-dichloro-7-(methoxy(oxetan-3-yl)methyl)-2-((4-methoxy-6-methyl-2-oxo-1,2-dihydropyridin-3-yl)methyl)-3,4-dihydroisoquinolin-1(2H)-one (PF-06821497). J Med Chem 2018; 61: 650–665. [DOI] [PubMed] [Google Scholar]
- 77. Song Y, Jin Z, Li ZM, et al. Enhancer of zeste homolog 2 inhibitor SHR2554 in relapsed or refractory peripheral T-cell lymphoma: data from the first-in-human phase I study. Clin Cancer Res 2024; 30:1248–1255. [DOI] [PubMed] [Google Scholar]
- 78. Gounder M, Schöffski P, Jones RL, et al. Tazemetostat in advanced epithelioid sarcoma with loss of INI1/SMARCB1: an international, open-label, phase 2 basket study. Lancet Oncol 2020; 21: 1423–1432. [DOI] [PubMed] [Google Scholar]
- 79. Morschhauser F, Tilly H, Chaidos A, et al. Tazemetostat for patients with relapsed or refractory follicular lymphoma: an open-label, single-arm, multicentre, phase 2 trial. Lancet Oncol 2020; 21: 1433–1442. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80. Chen L, Zheng X, Liu W, et al. Compound AC1Q3QWB upregulates CDKN1A and SOX17 by interrupting the HOTAIR-EZH2 interaction and enhances the efficacy of tazemetostat in endometrial cancer. Cancer Lett 2023; 578: 216445. [DOI] [PubMed] [Google Scholar]
- 81. Zhao J, Jin W, Yi K, et al. Combination LSD1 and HOTAIR-EZH2 inhibition disrupts cell cycle processes and induces apoptosis in glioblastoma cells. Pharmacol Res 2021; 171: 105764. [DOI] [PubMed] [Google Scholar]
- 82. Liao Y, Chen CH, Xiao T, et al. Inhibition of EZH2 transactivation function sensitizes solid tumors to genotoxic stress. Proc Natl Acad Sci U S A 2022; 119: e2105898119. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83. Keller PJ, Adams EJ, Wu R, et al. Comprehensive target engagement by the EZH2 inhibitor tulmimetostat allows for targeting of ARID1A mutant cancers. Cancer Res 2024; 84: 2501–2517. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84. Bitler BG, Aird KM, Garipov A, et al. Synthetic lethality by targeting EZH2 methyltransferase activity in ARID1A-mutated cancers. Nat Med 2015; 21: 231–238. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85. Zauderer MG, Szlosarek PW, Le Moulec S, et al. EZH2 inhibitor tazemetostat in patients with relapsed or refractory, BAP1-inactivated malignant pleural mesothelioma: a multicentre, open-label, phase 2 study. Lancet Oncol 2022; 23: 758–767. [DOI] [PubMed] [Google Scholar]
- 86. Parsons DW, Janeway KA, Patton DR, et al.; NCI-COG Pediatric MATCH Team. Actionable tumor alterations and treatment protocol enrollment of pediatric and young adult patients with refractory cancers in the National Cancer Institute-Children’s Oncology Group Pediatric MATCH Trial. J Clin Oncol 2022; 40: 2224–2234. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87. Italiano A, Soria JC, Toulmonde M, et al. Tazemetostat, an EZH2 inhibitor, in relapsed or refractory B-cell non-Hodgkin lymphoma and advanced solid tumours: a first-in-human, open-label, phase 1 study. Lancet Oncol 2018; 19: 649–659. [DOI] [PubMed] [Google Scholar]
- 88. Chen MK. Efficacy of PARP inhibition combined with EZH2 inhibition depends on BRCA mutation status and microenvironment in breast cancer. FEBS J 2021; 288: 2884–2887. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89. Huang X, Yan J, Zhang M, et al. Targeting epigenetic crosstalk as a therapeutic strategy for EZH2-aberrant solid tumors. Cell 2018; 175: 186–199.e19. [DOI] [PubMed] [Google Scholar]
- 90. Garofoli M, Maiorano BA, Bruno G, et al. Androgen receptor, PARP signaling, and tumor microenvironment: the ‘perfect triad’ in prostate cancer? Ther Adv Med Oncol 2024; 16: 17588359241258443. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91. Shao FF, Chen BJ, Wu GQ. The functions of EZH2 in immune cells: principles for novel immunotherapies. J Leukoc Biol 2021; 110: 77–87. [DOI] [PubMed] [Google Scholar]
- 92. Wang D, Quiros J, Mahuron K, et al. Targeting EZH2 reprograms intratumoral regulatory T cells to enhance cancer immunity. Cell Rep 2018; 23: 3262–3274. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93. Goswami S, Apostolou I, Zhang J, et al. Modulation of EZH2 expression in T cells improves efficacy of anti-CTLA-4 therapy. J Clin Invest 2018; 128: 3813–3818. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94. Kim HJ, Cantor H, Cosmopoulos K. Overcoming immune checkpoint blockade resistance via EZH2 inhibition. Trends Immunol 2020; 41: 948–963. [DOI] [PubMed] [Google Scholar]
- 95. Yang X, Xu L, Yang L. Recent advances in EZH2-based dual inhibitors in the treatment of cancers. Eur J Med Chem 2023; 256: 115461. [DOI] [PubMed] [Google Scholar]
- 96. Romanelli A, Stazi G, Fioravanti R, et al. Design of first-in-class dual EZH2/HDAC inhibitor: biochemical activity and biological evaluation in cancer cells. ACS Med Chem Lett 2020; 11: 977–983. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97. Huang N, Liao P, Zuo Y, et al. Design, synthesis, and biological evaluation of a potent dual EZH2-BRD4 inhibitor for the treatment of some solid tumors. J Med Chem 2023; 66: 2646–2662. [DOI] [PubMed] [Google Scholar]
- 98. Han Z, Rimal U, Khatiwada P, et al. Dual-acting peptides target EZH2 and AR: a new paradigm for effective treatment of castration-resistant prostate cancer. Endocrinology 2023; 164: 1–11. [DOI] [PubMed] [Google Scholar]
- 99. Wang C, Chen X, Liu X, et al. Discovery of precision targeting EZH2 degraders for triple-negative breast cancer. Eur J Med Chem 2022; 238: 114462. [DOI] [PubMed] [Google Scholar]
- 100. Li X, Wang C, Li S, et al. Dual target PARP1/EZH2 inhibitors inducing excessive autophagy and producing synthetic lethality for triple-negative breast cancer therapy. Eur J Med Chem 2024; 265: 116054. [DOI] [PubMed] [Google Scholar]
- 101. Spiliopoulou P, Spear S, Mirza H, et al. Dual G9A/EZH2 inhibition stimulates antitumor immune response in ovarian high-grade serous carcinoma. Mol Cancer Ther 2022; 21: 522–534. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102. Sharma S, Wang SA, Yang WB, et al. First-in-class dual EZH2-HSP90 inhibitor eliciting striking antiglioblastoma activity. J Med Chem 2024; 67: 2963–2985. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103. Bao Q, Kumar A, Wu D, et al. Targeting EED as a key PRC2 complex mediator toward novel epigenetic therapeutics. Drug Discov Today 2024; 29: 103986. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104. Kong X, Chen L, Jiao L, et al. Astemizole arrests the proliferation of cancer cells by disrupting the EZH2-EED interaction of polycomb repressive complex 2. J Med Chem 2014; 57: 9512–9521. [DOI] [PubMed] [Google Scholar]
- 105. Qi W, Zhao K, Gu J, et al. An allosteric PRC2 inhibitor targeting the H3K27me3 binding pocket of EED. Nat Chem Biol 2017; 13: 381–388. [DOI] [PubMed] [Google Scholar]
- 106. Huang Y, Sendzik M, Zhang J, et al. Discovery of the clinical candidate MAK683: an EED-directed, allosteric, and selective PRC2 inhibitor for the treatment of advanced malignancies. J Med Chem 2022; 65: 5317–5333. [DOI] [PubMed] [Google Scholar]
- 107. Wang X, Cao W, Zhang J, et al. A covalently bound inhibitor triggers EZH2 degradation through CHIP-mediated ubiquitination. EMBO J 2017; 36: 1243–1260. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108. Mei H, Wu H, Yang J, et al. Discovery of IHMT-337 as a potent irreversible EZH2 inhibitor targeting CDK4 transcription for malignancies. Signal Transduct Target Ther 2023; 8: 18. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109. Zhou B, Wang B, Zou F, et al. Discovery of dihydropyridinone derivative as a covalent EZH2 degrader. Eur J Med Chem 2023; 261: 115825. [DOI] [PubMed] [Google Scholar]
- 110. Békés M, Langley DR, Crews CM. PROTAC targeted protein degraders: the past is prologue. Nat Rev Drug Discov 2022; 21: 181–200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111. Wang C, Zhang Y, Chen W, et al. Epidermal growth factor receptor PROTACs as an effective strategy for cancer therapy: a review. Biochim Biophys Acta Rev Cancer 2023; 1878: 188927. [DOI] [PubMed] [Google Scholar]
- 112. Ma A, Stratikopoulos E, Park KS, et al. Discovery of a first-in-class EZH2 selective degrader. Nat Chem Biol 2020; 16: 214–222. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113. Sun X, Gao H, Yang Y, et al. PROTACs: great opportunities for academia and industry. Signal Transduct Target Ther 2019; 4: 64. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114. Schreiber SL. The rise of molecular glues. Cell 2021; 184: 3–9. [DOI] [PubMed] [Google Scholar]
- 115. Li X, Pu W, Zheng Q, et al. Proteolysis-targeting chimeras (PROTACs) in cancer therapy. Mol Cancer 2022; 21: 99. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116. Marei H, Tsai WK, Kee YS, et al. Antibody targeting of E3 ubiquitin ligases for receptor degradation. Nature 2022; 610: 182–189. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117. Wang C, Zhang Y, Wang J, et al. VHL-based PROTACs as potential therapeutic agents: recent progress and perspectives. Eur J Med Chem 2022; 227: 113906. [DOI] [PubMed] [Google Scholar]
- 118. He Q, Zhao X, Wu D, et al. Hydrophobic tag-based protein degradation: development, opportunity and challenge. Eur J Med Chem 2023; 260: 115741. [DOI] [PubMed] [Google Scholar]
- 119. Tu Y, Sun Y, Qiao S, et al. Design, synthesis, and evaluation of VHL-based EZH2 degraders to enhance therapeutic activity against lymphoma. J Med Chem 2021; 64: 10167–10184. [DOI] [PubMed] [Google Scholar]
- 120. Dale B, Anderson C, Park KS, et al. Targeting triple-negative breast cancer by a novel proteolysis targeting chimera degrader of enhancer of zeste homolog 2. ACS Pharmacol Transl Sci 2022; 5: 491–507. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121. Velez J, Dale B, Park KS, et al. Discovery of a novel, highly potent EZH2 PROTAC degrader for targeting non-canonical oncogenic functions of EZH2. Eur J Med Chem 2024; 267: 116154. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122. Liu Z, Hu X, Wang Q, et al. Design and synthesis of EZH2-based PROTACs to degrade the PRC2 complex for targeting the noncatalytic activity of EZH2. J Med Chem 2021; 64: 2829–2848. [DOI] [PubMed] [Google Scholar]
- 123. Yu X, Wang J, Gong W, et al. Dissecting and targeting noncanonical functions of EZH2 in multiple myeloma via an EZH2 degrader. Oncogene 2023; 42: 994–1009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124. Falkenstern L, Georgi V, Bunse S, et al. A miniaturized mode-of-action profiling platform enables high throughput characterization of the molecular and cellular dynamics of EZH2 inhibition. Sci Rep 2024; 14: 1739. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125. Bisserier M, Wajapeyee N. Mechanisms of resistance to EZH2 inhibitors in diffuse large B-cell lymphomas. Blood 2018; 131: 2125–2137. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126. Brooun A, Gajiwala KS, Deng YL, et al. Polycomb repressive complex 2 structure with inhibitor reveals a mechanism of activation and drug resistance. Nat Commun 2016; 7: 11384. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127. Gibaja V, Shen F, Harari J, et al. Development of secondary mutations in wild-type and mutant EZH2 alleles cooperates to confer resistance to EZH2 inhibitors. Oncogene 2016; 35: 558–566. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128. Chu L, Tan D, Zhu M, et al. EZH2 W113C is a gain-of-function mutation in B-cell lymphoma enabling both PRC2 methyltransferase activation and tazemetostat resistance. J Biol Chem 2023; 299: 103073. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129. Yamagishi M, Kuze Y, Kobayashi S, et al. Mechanisms of action and resistance in histone methylation-targeted therapy. Nature 2024; 627: 221–228. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130. Yan KS, Lin CY, Liao TW, et al. EZH2 in cancer progression and potential application in cancer therapy: a friend or foe? Int J Mol Sci 2017; 18: 1172. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131. de Vries NA, Hulsman D, Akhtar W, et al. Prolonged Ezh2 depletion in glioblastoma causes a robust switch in cell fate resulting in tumor progression. Cell Rep 2015; 10: 383–397. [DOI] [PubMed] [Google Scholar]