Next Article in Journal
Grapefruit-Derived Micro and Nanovesicles Show Distinct Metabolome Profiles and Anticancer Activities in the A375 Human Melanoma Cell Line
Next Article in Special Issue
Chronic Myeloid Leukemia: A Model Disease of the Past, Present and Future
Previous Article in Journal
Implication of the Association of Fibrinogen Citrullination and Osteoclastogenesis in Bone Destruction in Rheumatoid Arthritis
Previous Article in Special Issue
t(8;21) Acute Myeloid Leukemia as a Paradigm for the Understanding of Leukemogenesis at the Level of Gene Regulation and Chromatin Programming
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Targeting Chromatin Complexes in Myeloid Malignancies and Beyond: From Basic Mechanisms to Clinical Innovation

by
Florian Perner
1,2,* and
Scott A. Armstrong
1
1
Department of Pediatric Oncology, Dana-Farber Cancer Institute, Harvard Medical School, Boston, MA 02215, USA
2
Internal Medicine II, Hematology and Oncology, Friedrich Schiller University Medical Center, 07747 Jena, Germany
*
Author to whom correspondence should be addressed.
Submission received: 19 November 2020 / Revised: 13 December 2020 / Accepted: 20 December 2020 / Published: 21 December 2020
(This article belongs to the Special Issue Pathophysiology and Molecular Targets in Myeloid Neoplasia)

Abstract

:
The aberrant function of chromatin regulatory networks (epigenetics) is a hallmark of cancer promoting oncogenic gene expression. A growing body of evidence suggests that the disruption of specific chromatin-associated protein complexes has therapeutic potential in malignant conditions, particularly those that are driven by aberrant chromatin modifiers. Of note, a number of enzymatic inhibitors that block the catalytic function of histone modifying enzymes have been established and entered clinical trials. Unfortunately, many of these molecules do not have potent single-agent activity. One potential explanation for this phenomenon is the fact that those drugs do not profoundly disrupt the integrity of the aberrant network of multiprotein complexes on chromatin. Recent advances in drug development have led to the establishment of novel inhibitors of protein–protein interactions as well as targeted protein degraders that may provide inroads to longstanding effort to physically disrupt oncogenic multiprotein complexes on chromatin. In this review, we summarize some of the current concepts on the role epigenetic modifiers in malignant chromatin states with a specific focus on myeloid malignancies and recent advances in early-phase clinical trials.

1. Histone-Code and Epigenetic Networks: Implications for Myeloid Malignancies and Other Cancer Types

The genomic DNA in eukaryotic cells is wrapped around a core histone octamer composed of Histone H2A (H2A), Histone H2B (H2B), Histone H3 (H3), and Histone H4 (H4) [1,2,3]. This complex of DNA and histones is compacted at different densities to chromatin, ranging from a simple chain of DNA and histones to highly condensed metaphase chromosomes [4,5,6,7,8,9]. The tight regulation of chromatin condensation and de-condensation is vital for a healthy organism since it is crucial for the equal distribution of genetic information to the daughter cells during cell division [10,11]. Furthermore, the state of chromatin condensation and therefore the degree of DNA accessibility, restricts or allows transcription factors and the RNA-polymerase machinery to physically interact with the DNA. Therefore, the stringent regulation of chromatin accessibility is a critical determinant of proper spatial and temporal regulation of gene transcription [12,13,14]. In cancer cells, the healthy chromatin homeostasis is disrupted by a variety of mechanisms promoting aberrant transcription and potentially also cytogenetic alterations [15,16,17].
The dynamics of chromatin biology are regulated by a variety of post-translational modifications at the unstructured core-histone tails [18]. These modifications involve methylation, acetylation, and phosphorylation, as well as many other less well characterized modifications like sumoylation, ADP-ribosylation, deimination, proline-isomerization, and crotonylation [18,19,20,21,22,23]. Early studies in the field have provided evidence that histone acetylation, a mark that is broadly associated with active chromatin, diminishes the positive charge of histones, thereby loosening the interaction between the histones and the negatively charged phosphate groups on the DNA, resulting in increased DNA accessibility [23,24,25,26,27,28,29]. In 2000, Brian Strahl and David Allis suggested a groundbreaking concept by proposing the model of the “histone code” [22]. Several observations from their and other groups prompted the idea that the complex array of post-translational modifications on histone tails serve as a binding platform for reader proteins and associated multiprotein complexes and thereby orchestrate a complex network of regulatory proteins on chromatin [22,30,31,32,33,34]. This concept is the basis for our current understanding of the regulatory dynamics on chromatin involving writers, erasers, and readers of histone modifications (Figure 1). An example is H3K27 acetylation, which is associated with active enhancers and promotors [18]. The “writers” of this mark are the acetyltransferases p300 and CBP that deposit H3K27ac to activate specific enhancers and promotors [35,36]. Histone-deacetylases (like HDAC1 and 2) serve in an antagonistic fashion as “erasers” to remove the mark and repress these regulatory elements. BRD4, that can associate with H3K27ac nucleosomes via its bromodomain, is a “reader” that gets recruited to active promotors and enhancers through binding to H3K27ac [37,38,39,40]. Importantly, BRD4 associates with the multiprotein-complex of the transcription machinery and is essential for the activation of RNA-polymerase II (RNA-PolII) by the superelongation complex [41]. Therefore, altering the function of readers, writers, or erasers in this system will have dramatic consequences for the homeostasis of gene transcription [40,42,43,44,45,46].
In actual physiological systems, the complexity of the protein-network generated by the histone code is complicated, since the regulatory elements of different marks are highly interconnected. Writers and erasers of histone modifications get recruited by reader proteins of other histone modifications, leading to a topologically organized array of co-existing modifications and associated multi-protein complexes that define distinct regulatory microenvironments at specific regions on chromatin impacting transcription and 3D-chromosome architecture [47,48,49,50,51,52]. These regulatory networks are central mediators of embryonic development, lineage determination during differentiation and cellular homeostasis [53,54,55,56,57,58,59,60,61,62,63,64,65]. Consequently, aberrant regulation of this network is a hallmark of cancer development and progression [17,66,67,68,69,70]. Several of the recurrently detected genetic abnormalities in patients with myeloid malignancies involve epigenetic modifiers like DNMT3A, TET2, ASXL1, EZH2, IDH1/2, KMT2A, KAT6A, KDM5A, KDM6A, or NSD1 [71,72,73,74]. In the recent years, it has become clear that mutations in some of these genes arise in hematopoietic stem cells of healthy individuals during the process of aging [75,76,77,78]. The occurrence of these mutations was found to be associated with an increased risk to develop myeloid malignancies as well as cardiovascular disease [75,79,80]. The application of novel single-cell sequencing approaches in primary material from patients with acute myeloid leukemia (AML) and myeloproliferative neoplasia (MPN) revealed that these clonal mutations in epigenetic modifiers tend to co-occur in dominant clones of patients that ultimately developed leukemia, implicating a synergistic pathogenic function if occurring in the same cell. Mutations in signaling molecules, on the other hand, appeared to originate later during disease evolution and showed a tendency not to co-occur within the same clone [81]. These recent findings strongly suggest, that epigenetic rewiring in specific hematopoietic clones is the foundation for the development of myeloid malignancies and may explain the increasing incidence of myelodysplastic syndrome (MDS), MPN, and AML in elderly individuals. Nevertheless, no pharmacological targeting approaches that show promise in blocking expansion and molecular progression of these pre-leukemic clones have been established yet clinically. Gaining a better understanding of the chromatin biology during this preleukemic disease state and developing valid predictors of progression, as well as clinically active epigenetic mono- or combination therapies, will be a central goal to prevent leukemia development in the future. However, much progress has been made in the preclinical and clinical development of compounds interfering with the function of chromatin modifiers for use in patients with fully developed cancer. In both myeloid malignancies and solid cancers, several of these molecules have entered clinical trials and show activity in a variety of diseases [74,82,83,84,85]. In this review, we will discuss several clinical and pre-clinical epigenetic targeting approaches for cancer therapy in the context of the molecular mechanisms of action of different drug-classes and their potential to disrupt oncogenic epigenetic signatures. We will particularly focus on therapeutic tools that target specific chromatin complexes and will not discuss the clinical use of inhibitors of DNA methylation. The demethylating agents 5-azacitidine and decitabine were the first “epigenetic” drugs that were approved for the treatment of MDS and AML by the US Food and Drug Administration (FDA) in 2004 and are successfully used in different clinical settings to date [84]. Nevertheless, their exact mechanism of action is not fully understood and they have been previously reviewed [82,83,84]. Therefore, we will not discuss these compounds in further detail in this review.

2. Inhibition of Histone Methylation on Chromatin: Recent Clinical Advances and Challenges

Targeting histone-methylation pattern on chromatin is a scientifically and clinically exiting opportunity to interrogate with disease-related epigenetic programs in vitro and in vivo. Given the fact, that a very specific set of methyltransferases and de-methylases is responsible for the deposition and removal of specific methylation-marks, the interrogation with distinct de-regulated pattern within the histone-code is possible. In contrast, histone acetyltransferases and deacetylases act in a more promiscuous way. Those enzymes have a very broad spectrum of targets among histones as well as non-histone proteins. Therefore, the molecular consequences of inhibiting those enzymes are more complex and much less specific for certain epigenetic programs. Multiple inhibitory drugs for methyltransferases, acetyltransferases, demethylases, and deacetylases have been developed and are currently under preclinical or clinical investigation [83]. In the history of drug-discovery, much experience has been gathered regarding the development and evaluation of enzymatic inhibitors [86,87,88]. Those drugs act as competitive or allosteric inhibitors of catalytically active proteins, preventing the binding of an enzyme to its target substrate and/or disturbing the structural integrity of its catalytic center. In the epigenetic space, proteins with enzymatic activity play a crucial role as “writers” and “erasers” of histone modifications, as discussed above. Therefore, molecules that inhibit the activity of those proteins have the potential to specifically interfere with the global deposition of certain histone-marks. A number of these molecules have been developed for the treatment of hematologic malignancies and solid cancers [83]. The first inhibitors of histone modifiers that have been approved for the treatment of cancer were inhibitors of histone-deacetylases. The first-in class molecule Vorinostat was initially approved for cutaneous T cell lymphoma [89]. Drug development efforts to improve the pharmacologic properties led to the approval of Belinostat in 2014 for T-cell lymphoma and Panobinostat in 2015 for refractory multiple myeloma. In myeloid malignancies, HDAC inhibitors show promise in clinical trials as monotherapy or in combination with other drugs [83,90,91,92]. Using this class of drugs for the precise interrogation of specific epigenetic programs/mechanisms has however remained difficult. As mentioned above, HDACs act in a promiscuous way erasing histone acetylation broadly across the genome. Furthermore, clinically active compounds are mostly pan-HDAC-inhibitors without particular selectivity for a specific molecule. Therefore, distinct oncogenic signatures or somatic mutations that confer sensitivity to HDAC-inhibition have not been established. More recently, the evolving knowledge about epigenetic mechanisms and subsequent drug development efforts have led to the establishment of a number of inhibitors of histone-methyltransferases. Histone methyltransferases appear to act in a more selective fashion than acetyltransferases. This class of compounds may enable scientists and clinicians for the first time to develop mechanistically driven and relatively well tolerated treatment regimens for cancer therapy. In this review, we will focus on specific inhibitors of histone-modifying complexes that have recently entered clinical trials (Table 1).

2.1. Inhibition of EZH2—The Catalytic Subunit of the PRC2-Complex

The tight regulation of the interplay between activating and repressive mediators on chromatin is crucial for the determination of cell identity during embryonic development and differentiation and also a critical element of aberrant chromatin biology in cancer cells. The competition between activating Trithorax-group proteins and repressive Polycomb complexes on central developmental genes (e.g., Hox-cluster genes) and the importance of this balance for segmental identity has been defined in Drosophila melanogaster [55,57,93,94] (Figure 2). The importance of the interplay between proteins associated with homologues of these chromatin-regulator families is highly conserved from yeast to mammalian cells and de-regulation of this balance is frequently observed in cancer cells [55,57,95,96,97,98,99,100,101].
One of the histone modifications that is associated with polycomb (PRC)-mediated repression is H3K27me3. This modification is catalyzed by the histone methyltransferase “Enhancer of zeste 2” (EZH2), which is the catalytic subunit of the PRC2-complex [102]. H3K27me3 is strongly associated with repressed chromatin and is crucial for the repression of developmental programs in a timely manner [103,104,105]. Mutations and aberrant expression of EZH2 are frequently found in cancer and targeting this methyltransferase may have therapeutic potential in different disease entities [106,107,108,109,110,111,112]. Based on pre-clinical data and the early experiences from clinical trials, sensitivity to EZH2-inhibition seems to be determined by molecular settings that result in a specific dependency on PRC2 function. Activating mutations in EZH2 are frequently found in lymphoma and are considered disease drivers that display a target for pharmacological inhibition [110,113,114,115]. Overexpression of EZH2 is observed in a variety of different cancers, suggesting potential therapeutic opportunities [111,112,115,116]. Furthermore, specific mutations in other chromatin modifying complexes have been shown to create synthetic-lethality with EZH2-inhibition. As mentioned above, the interplay between activating trithorax-group proteins and PCR-complexes is crucial for the regulation of developmental genes (Figure 2, left). Inactivating mutations within members of the SWI/SNF chromatin remodeling complexes, as members of the trithorax-group, lead to a shift of this balance towards the dominance of PRC-mediated repression (Figure 2, right). SWI/SNF complex alterations are among the most frequently found genetic alterations across all cancer types and are disease drivers in a variety of entities, most intensively characterized in rhabdoid sarcoma [117,118,119,120,121]. In rhabdoid tumors with SMARCB1 loss-of-function mutations, EZH2 displays a specific dependency which can be exploited using pharmacologic inhibition of EZH2′s enzymatic function [122,123]. In myeloid malignancies, the relevance of altered EZH2 function largely remains elusive. In MPN, MDS and AML, mutations of EZH2 and other members of the PRC2 complex are as recurrent events. Of note, the vast majority of these mutations confer a loss of PRC2 function [124,125,126]. In contrast to the observations in sarcoma, EZH2 seems to act as a tumor suppressor in these myeloid conditions, since its inactivation by somatic mutations is associated with adverse outcomes [127,128,129,130]. Conversely, homologues to drosophila’s trithorax-family like the histone methyltransferase MLL1 (KMT2A) and certain SWI/SNF complex members appear to act as oncogenes/dependencies in myeloid malignancies, suggesting fundamentally different consequences of the trithorax–polycomb balance in myeloid neoplasms compared to rhabdoid- and other tumors [131,132,133]. Therefore, targeting of EZH2 using pharmacologic inhibitors does not appear to be intuitive in these conditions. Nevertheless, specific molecular constellations and/or the exploitation of so far unknown synthetic lethalities might allow the use of these drugs in certain settings in the future. For example, it has been demonstrated that chronic myeloid leukemia (CML) stem cells rely on EZH2 for self-renewal and disease maintenance and that EZH2 inhibition in combination with targeting of BCR-ABL can be utilized to eradicate these disease-initiating cell populations in a pre-clinical setting [134,135]. Furthermore, the PRC2 complex appears to be required for proliferation and BET-inhibitor resistance in MLL-rearranged AML models [136,137]. More detailed molecular insights into the chromatin biology of these diseases, particularly in the disease-driving cell populations, are required to evaluate the potential of pharmacologic inhibition of the PRC2-repressive complex in myeloid conditions.
Tazemetostat, a selective inhibitor of EZH2, was the first compound of this class that was approved by the FDA for the treatment of locally advanced epitheloid sarcoma and follicular lymphoma in 2020 [138,139] (Figure 3). The response rate in SMARCB1-negative epitheloid sarcoma was modest with an overall response rate of 15%, while only 1.6% of patients reached a complete remission. Most responses lasted 6 months or longer and the drug was well tolerated [139]. Due to the lack of therapeutic alternatives, Tazemetostat was approved based on this phase 2 clinical trial in January 2020. In relapsed/refractory follicular lymphoma harboring activating mutations of EZH2, the clinical responses were much more encouraging, with an overall response rate of 69% and 12% complete responses [138]. The results of this clinical trial led to approval of Tazemetostat for relapsed/refractory follicular lymphoma with EZH2 mutations in June 2020. A number of clinical trials evaluating Tazemetostat in different clinical settings in lymphoma and solid tumors are currently ongoing [83,111,140]. GSK2816126, another EZH2-inhibitor, was recently investigated in a phase 1 clinical trial in lymphoma and solid cancers. The drug was well tolerated, but failed to potently repress H3K27me3, as measured in the peripheral blood of patients potentially due to unfavorable pharmacologic properties [141]. The second-generation compound CPI-1205 that has improved potency for inhibition of EZH2 is currently investigated in phase 1–2 clinical trials in lymphoma and solid cancers [142,143]. Data from these trials have not been released.

2.2. Inhibition of the Histone Methyltransferase DOT1L in MLL-Rearranged Leukemia and Beyond

DOT1L is the only histone-methyltransferase that catalyzes H3K79me1, -me2, and -me3 (Figure 3). H3K79me is deposited across the gene body and is associated with transcriptionally active genes [18,144,145,146]. In adult mammalian organisms, DOT1L is dispensable for homeostasis of most cell types except hematopoietic stem cells [147,148]. MLL1 (KMT2A), which is the mammalian homologue of Drosophila trithorax, is recurrently affected by chromosomal translocations during the development of acute leukemia, leading to the generation of oncogenic fusion proteins [149,150,151,152]. These MLL1-fusion oncogenes recruit DOT1L directly via the fusion partner (e.g., AF9, AF10, AF17, or ENL) or indirectly to their target genes to establish an active chromatin environment and prevent repression [153,154,155,156,157,158,159,160,161,162]. Consequently, genetic inactivation or pharmacologic inhibition of DOT1L has been established as a tool to target the MLL-fusion driven transcriptional program [154,163,164,165,166,167]. Of note, other subtypes of leukemia that show a high dependency on wild-type MLL1, like NPM1-, DNMT3A-, and IDH1/2-mutated as well as NUP98-rearranged AML, have been shown to be likewise sensitive to inhibition of DOT1L [153,168,169,170,171]. Given the fact that DOT1L is not required in normal tissues, pharmacological inhibition of DOT1L could be a highly selective treatment strategy in MLL-rearranged and MLL-dependent leukemia. EPZ004777, the first selective DOT1L inhibitor reported in 2011, led to a potent reduction in H3K79me and cell proliferation in the micromolar range [167]. In an initial in vivo assessment using a cell-line xenograft system, the drug showed only modest activity due to poor pharmacologic properties. The second-generation DOT1L-inhibitor EPZ5676 showed a higher potency and selectivity to inhibit DOT1L with activity in the low nanomolar range in vitro [166]. Nevertheless, also EPZ5676 showed relatively poor pharmacologic properties in vivo and had to be administrated via continuous infusion to achieve relevant plasma concentrations and clinical responses. In a rat cell-line xenograft model, EPZ5676 led to complete regression of tumors after 3 weeks of administration while no relevant signs of toxicity were reported from this pre-clinical trial [166]. Recent efforts in structure-guided probe optimization have led to the establishment of a set of DOT1L-inhibitors with improved pharmacologic properties including attenuated plasma concentrations and oral bioavailability that were recently reported to have promising pre-clinical activity in patient-derived xenograft models of AML [172,173]. The availability of these novel compounds enables pre-clinical investigators for the first time to probe combination therapies of DOT1L-inhibitors with other epigenetic compounds using clinically applicable dosing regimens in animal model systems. Collectively, these pre-clinical reports provided proof that pharmacological targeting of DOT1L could be an effective and selective treatment approach in MLL-rearranged leukemia. However, they also pointed out difficulties regarding DOT1L as a drug target: Both the in vitro and in vivo experiences showed that only complete and prolonged inhibition of DOT1L and consequently H3K79me was sufficient to induce relevant therapeutic responses. The sub-optimal pharmacological properties of both EPZ004777 and EPZ5676 may therefore be problematic.
Despite these pharmacologic limitations, EPZ5676 was investigated as a single-agent in a phase 1 clinical trial in relapsed/refractory MLL-rearranged AML (NCT01684150) [165]. In this study, patients with AML and ALL, 39 of which had MLL-rearranged leukemia, were enrolled in 6 dose-escalation and 2 expansion cohorts. Similar to the pre-clinical settings, EPZ5676 (Pinometostat) was administrated via continuous infusion over 28 days. Steady-state plasma concentration was reached after 4–8 h after start of the infusion but decreased rapidly upon discontinuation. Overall, the treatment was well tolerated and a maximum tolerated dose was not reached in part due to physical limitations in solubility and infusion rate of the drug. Out of the 51 patients that received the study medication, only 3 patients completed the study. The main reason for discontinuation was insufficient response and disease progression (31 patients, 60%). Two patients experienced a complete remission on the study medication but ultimately relapsed shortly after the last cycle of EPZ5676. In summary, the trial provided proof of concept for the tolerability and clinical activity of DOT1L-inhibition in MLL-rearranged leukemia. Nevertheless, efficacy of the treatment as a monotherapy was limited, inducing modest responses that did not last [165]. Currently, EPZ5676 is under active investigation in two clinical trials, testing its efficacy in relapsed and therapy refractory AML patients with MLL-rearrangements in combination with 5-azacitidine (NCT03701295) and in patients with newly diagnosed MLL-rearranged AML in combination with standard induction chemotherapy (NCT03724084). Both trials are still recruiting participants. In summary, DOT1L-inhibition may be a useful pharmacological opportunity in myeloid malignancies and potentially also in other cancers. However, to achieve sufficient clinical responses, the establishment of compounds with more favorable pharmacological properties and synergistic combination therapy regimens are likely necessary.

2.3. Inhibition of LSD1 in Myeloid Leukemia and Beyond

The lysine-specific demethylase 1 (LSD1) has been identified as a critical eraser for H3K9me1/2 and most importantly for H3K4me1/2 [174,175] on chromatin. H3K4me1 marks enhancers and its tight regulation during cellular differentiation is critical to execute lineage switches and other cell fate decisions [18]. K3K4me1 is catalyzed by MLL3 and MLL4 (KMT2C, KMT2D), a group of trithorax-group proteins that are crucial for shaping enhancers. LSD1 antagonizes this process by demethylation of H3K4 and is therefore able to suppress previously primed enhancers [175,176]. Of note, the demethylase function of LSD1 is specific for mono- and dimethylated H3K4 and H3K9 and has therefore no direct impact on active promotors (H3K4me3) or the heterochromatin (H3K9me3). In embryonic stem cells, it has been shown that loss of LSD1 prevents terminal differentiation, since LSD1 mediated decommissioning of stem-cell enhancers is essential to license lineage-specific rearrangement of the enhancer landscape and associated transcriptional programs [176,177]. In cancer cells on the other hand, LSD1 seems to be a gatekeeper of cancer stemness in different malignant entities, and therefore, a therapeutic target for differentiation therapy [178,179]. A possible explanation for these contrary roles in embryonic stem- and cancer cells might be the fact that cancer cells already underwent lineage priming but actively repress enhancers that drive essential genes for terminal differentiation via LSD1. This epigenetic block of cellular differentiation programs can be released by pharmacologic inhibition of LSD1. Evidence suggests that LSD1-inhibition in AML drives differentiation by disrupting stemness-signatures and at the same time, recommissions PU.1 and C/EBPa-dependent enhancers to facilitate myeloid differentiation [180]. The ability of LSD1-inhibitors to drive differentiation in AML has been well described in several different molecular subtypes to date [181,182,183,184,185,186,187]. In solid cancers, LSD1 activity has been reported to be linked to metastasis, cell proliferation, and inhibition of p53 [188,189,190,191].
Tranylcypromine (TCP) was initially approved by the FDA in 1961 as a MAO-inhibitor for mood and anxiety disorders but recently identified as the first inhibitor of LSD1 demethylase function [192]. TCP and other more selective and potent LSD1-inhibitors like ORY-1001, ORY-2001, or IMG-7289 are currently investigated in a number of early phase clinical trials in different conditions, including AML, MPN, MDS, solid cancers, Alzheimer’s, and multiple sclerosis [83,186] (Figure 3). As a single agent, LSD1-inhibitors were able to induce differentiation of AML blasts and control disease burden in a phase 1 clinical trial [193]. Even though monotherapy with highly potent LSD1-inhibitors shows modest activity, synergistic combination therapies seem to be a more promising strategy to address aggressive malignancies like AML. The potential of LSD1-inhibition to sensitize AML-cells to other differentiation-inducing agents, like ATRA, has been recognized in pre-clinical settings as well as in an initial clinical trial [194,195]. Several other early-phase clinical trials probing combinations of LSD1-inhibitors with ATRA or conventional chemotherapy in AML are currently ongoing and have not been published yet [186]. In pre-leukemic conditions, including MPN and MDS, preclinical and emerging clinical data suggests that LSD1-inhibition can control clonal burden and potentially impact disease progression [196,197]. Clinical trials probing the activity of LSD1-inhibition in MDS and MPN are ongoing [186].

3. Disrupting Protein-Protein Interactions to Target Epigenetic Complexes

Despite the recent advances in the development and clinical establishment of enzymatic inhibitors of chromatin modifiers, major challenges remain on the way to development of more clinically efficacious drugs that target chromatin complexes. As described above, most enzymatic inhibitors of histone-modifying enzymes that have been assessed to date show only modest clinical activity, particularly when administrated as single agents. Several of these compounds showed sufficient on-target activity in vivo (as measured by global reduction of histone methylation) and were relatively well tolerated. Still, the clinical responses observed were not as extensive as might have been predicted based on the phenotypes described in preclinical model systems using genetic-inactivation and small molecule inhibitors. In recent years, a growing body of evidence suggests that inhibition of any single enzymatic activity/function in a chromatin complex may be insufficient to perturb epigenetic homeostasis and control gene expression in cancers cells. This has recently been demonstrated for complexes containing EZH2 and LSD1 [198,199,200]. Similarly, pharmacologic inhibition of DOT1L’s methyltransferase function shows modest and delayed responses in vitro and in vivo when compared to genetic inactivation that demonstrated DOT1L’s essential function in MLL-rearranged leukemia [147,154,164,172,201,202,203,204]. Therefore, we expect approaches that disrupt chromatin associated complexes more extensively may be more therapeutically efficacious.
Another major issue is that only a relatively small fraction of essential proteins on chromatin have enzymatic activity, so the spectrum of potential applications for catalytic inhibitors is naturally limited. Recent advances in the development of inhibitors of protein–protein interactions open up a novel field of potential applications, particularly in the epigenetic space, were a large proportion of the dense network of multiprotein-complexes cannot be pharmacologically addressed using enzymatic inhibitors. In the past few years, a variety of compounds disrupting epigenetic protein–protein interactions have been synthesized and established mostly in vitro [205]. In this section, we will focus on Menin-MLL inhibitors and bromodomain-inhibitors since these molecules have already been intensively investigated in preclinical studies and entered early-phase clinical trials.

3.1. Menin-MLL1 Interaction Inhibitors: Novel Compounds with Disruptive Potential

As described above, gain-of-function of the mammalian trithorax-homologue MLL1 (KMT2A) is a recurring pattern in cancer, particularly in acute leukemia. The increased activation of MLL1-target genes, which are part of an evolutionary conserved stem-cell program, is driven by MLL-fusion oncogenes or aberrantly activated WT-MLL1-complexes [149,169,206,207,208]. Menin is an adaptor protein that binds to the N-terminal proportion of MLL1 and is important for MLL1 function [209,210,211]. Therefore, genetic inactivation of Menin or disruption of the Menin-MLL1-interaction has the potential to disturb the integrity of oncogenic MLL1-chromatin complexes [168,169,206,209,212,213,214,215,216,217]. In 2012, the first potent and selective menin-MLL1 interaction inhibitor, MI2-2, was reported to reduce cellular proliferation and oncogenic gene expression in MLL-rearranged leukemia [217,218]. Further structural optimization of this lead compound led to the development of MI-503 and MI-463, two highly potent and orally bioavailable compounds with pre-clinical activity against cell-line xenografts of MLL-rearranged leukemia [216]. Moreover, it has been demonstrated that the MLL1-menin-interaction is a central vulnerability in NPM1c-mutated leukemia, a much more frequent subtype of AML in adult patients [168]. In this type of leukemia, an intact MLL1-Menin interaction was shown to be required for the expression of critical stem cell-associated genes and could be successfully targeted using MI-503 [168]. Recently, a further improved but structurally related molecule, MI-3454, was reported to have potent effects on proliferation and gene expression in human leukemia cell lines and primary patient leukemia cells. Importantly, MI-3454 induced potent and long-lasting responses in patient-derived xenografts of MLL-rearranged and NPM1c mutated leukemia [213]. Recent work has also demonstrated that a different structurally unrelated highly specific inhibitor of the MLL1-menin interaction, VTP50469, is active in the low nanomolar range and suppressed MLL-target gene expression and cell proliferation in models of MLL-rearranged leukemia. In a large panel of patient derived xenografts, VTP50469 induced long-lasting complete responses and eradicated disease in a number of MLL-rearranged AML- and ALL-grafts [206]. In a model of NPM1c-mutated AML, the same molecule could be utilized intercept AML development and ultimately prevent manifestation of the disease in mice. Additionally, VTP50469 could also be used to target leukemia after disease onset [169]. In both studies, the expression of HOXA-cluster genes was not- or only slightly affected by treatment with the inhibitor. The expression of MEIS1 and PBX3, on the other hand, was rapidly and dramatically affected by VTP50469 treatment serving as molecular markers of response, similar to the observations made under treatment of MI-3454 [169,206,213]. ChIP-sequencing studies investigating chromatin occupancy of MLL-fusion complex members at MLL-target genes have shed light on the mechanisms behind this differential response at different loci. Treatment with the Menin-inhibitor VTP50469 leads to a global loss of Menin and DOT1L from chromatin [206], thereby disrupting the integrity of the oncogenic MLL-complex. Importantly, at the highly responsive target genes (MEIS1, PBX3, etc.) VTP50469 treatment also leads to a loss of the MLL-fusion protein itself [169,206]. Therefore, the ability of a Menin-inhibitor to displace MLL1 from its target genes seems to be the best predictor of the impact of the drug on gene expression.
These observations point out, that the potency of pharmacologic inhibition of the menin-MLL1 interaction at certain target genes is critically determined by its ability to physically disrupt the oncogenic multiprotein complex. Along these lines, it is tempting to speculate if this might also be the reason why this class of drugs appears to be more potent than inhibitors of DOT1Ls methyltransferase function. Besides the superior pharmacological features, particularly of VTP50469, MI-503, and MI-3454, their ability to structurally disrupt the oncogenic MLL-complex distinguishes their mechanism of action from DOT1L-inhibition (Figure 4). Using VTP50469, it was demonstrated that the cellular effects induced by this drug are not only more dramatic, but also more rapid compared to EPZ5676 in vitro [206]. Presumably, these features of Menin-inhibitors help to prevent early adaptation of leukemia cells to the drug by not giving them the time they need to adapt their programs. Therefore, Menin-inhibitors are currently a particularly promising group of compounds for selective, potent, and minimally toxic targeted therapy in MLL-rearranged and NPM1c-mutated leukemia and other MLL1-dependent dependent cancer types.
In November 2019, the AUGMENT-101 trial started to enroll patients on the Menin-inhibitor SDNX-5613 (NCT04065399), which is a close analogue of VTP-50469 [214]. This phase 1/2 clinical trial will enroll 156 participants with relapsed or refractory MLL-rearranged and NPM1c-mutated leukemia on the study medication. In an early report at the virtual AACR meeting 2020, Syndax pharmaceuticals reported on the first responses observed in patients treated with SDNX-5613. Importantly, the study medication was well tolerated at the first steps of dose-escalation and no dose-limiting toxicities were observed. Two out of three of the patients with MLL-rearranged leukemia that were enrolled in the trial showed a response after 28 days of treatment. Remarkably, one of these patients achieved a complete remission irrespective of a low daily drug dose, presumably due to co-medication with CYP3A4 inhibitor. The second patient achieved a partial response after the first cycle of treatment and was continued on the study medication. Treatment of the third patient with MLL-rearranged leukemia was discontinued as a result of disease progression at insufficient plasma concentrations of the drug [214]. Another phase 1 clinical trial (KOMET) investigating Kura Oncology’s compound KO-539 (an analogue of MI-3454) is also currently recruiting patients with relapsed/refractory AML (NCT04067336) and recently reported results from the first 6 patients treated, including 2 complete responses [219].
In summary, Menin-MLL interaction inhibitors are currently among the most promising epigenetic compounds under clinical investigation for the treatment of AML with impressive single-agent activity in pre-clinical trials. Their potential to disrupt oncogenic MLL-complexes in contrast to the isolated inhibition of the catalytic activity of DOT1L is presumably one important reason for the increased potency of this therapeutic approach.

3.2. Bromodomain-Inhibition: Breaking Apart Acetyl-Reader-Complexes

As discussed above, pharmacologic targeting of specific acetylation patterns is challenging due to the fact that acetyltransferases and deacteylases are generally promiscuous, so the interrogation with the deposition of distinct marks is difficult. Bromodomain-containing proteins are readers of acetylated histones, which are part of several different multiprotein complexes that are guided to certain sites on chromatin by these proteins [220,221]. Distinct structural differences within the bromodomains of different proteins of this family enabled the design and synthesis of a large set of fairly specific small molecule inhibitors of protein–protein interactions between acetylated histones and bromodomain proteins [44,220,221,222]. The first bromodomain-inhibitors that proceeded into clinical development were compounds targeting the bromodomains of bromodomain-containing protein 4 (BRD4) and BRD2. BRD4 binds to acetylated H3K27 at active enhancers and super-enhancers and is crucial for transcriptional co-activation, particularly in developmental and oncogenic contexts [37,38,39,107]. In more detail, BRD4 aids in recruitment and activation P-TEFb (CDK9/cyclinT1), thereby controlling transcriptional elongation [41,222] (Figure 5). Efforts to target transcriptional co-activation by BRD4 led to the discovery of the first effective bromodomain-inhibitors, I-BET and JQ1, in 2010 [44,223]. Despite a certain degree of general toxicity, at low doses, BET-inhibitors are fairly selective in inhibiting oncogenic programs, in part by disrupting super-enhancer function [40]. Furthermore, BRD4 was identified as a genetic vulnerability in AML using unbiased screening approaches [43,224]. In these studies, genetic- (RNAi) and pharmacologic inhibition of Brd4 induced differentiation and reduced disease burden in mouse models of MLL-rearranged leukemia. Both of these studies found reduced gene expression of MYC as a consequence of enhancer silencing by BET-inhibition as a major factor contributing to the therapeutic efficacy of JQ1 and I-BET151, respectively [43,224]. This phenomenon might be an explanation for the relatively broad, yet fairly selective activity of BET-inhibitors in cancer. Several clinical trials have investigated BET-inhibitors for the treatment of advanced cancers [225]. OTX015 was investigated in a number of clinical trials in hematologic and non-hematologic cancers (NCT02698176, NCT02698189, NCT02259114, NCT01713582, NCT02296476). A common issue in these early phase clinical trials was the fact that dose-limiting toxicities were observed at doses that were not sufficient to induce remission or stable disease and required an intermittent dosing regimen to manage adverse effects [226]. Nevertheless, oral bioavailability and reasonable pharmacokinetics of OTX015 were demonstrated in these trials. The structurally related yet improved compound TEN-101 showed superior pharmacokinetics and first clinical efficacy in NUT-midline carcinoma [227]. A number of clinical trials investigating TEN-101 in AML, myelodysplastic syndromes, multiple myeloma, high-grade B-cell lymphoma, and solid tumors have recently been completed (NCT02308761, NCT03068351, NCT01987362, NCT03292172, NCT03255096). Detailed results from most of these trials are not published yet. Several other BET-inhibitors are currently in clinical- and preclinical development for different malignant conditions [225,228]. Nevertheless, the apparently narrow therapeutic window along with the modest single-agent clinical efficacy of BRD4/2 (and pan-BRD) inhibitors have moderated clinical expectations.

4. Hijacking the Proteasome for Targeted Protein Degradation

In the recent years, targeted protein degradation has evolved as a novel and emerging strategy in drug development [229,230,231]. Two major concepts to achieve targeted protein degradation have been established and are currently exploited in drug discovery:
“Immunomodulatory drugs” like the FDA-approved compounds Thalidomide, Lenalidomide, and Pomalidomide exert their therapeutic function by selective degradation of a variety of zinc-finger transcription factors, including IKZF1 and IKZF3. Those molecules act as a “molecular glue” that binds to its target and subsequently recruits a cereblon E3-ligase complex, leading to polyubiquitination and proteasomal degradation of the target [232,233]. Recent advances in drug discovery strategies have led to screening approaches that allow to exploit this molecular glue concept to design degraders that specifically target other target proteins, opening up the opportunity to address a variety of drug targets including epigenetic modifiers in the future [234,235].
PROteolysis TArgeting Chimeras (PROTACs) are bifunctional molecules high affinity ligands of ubiquitin ligases like cereblon (CRBN), Von Hippel-Lindau (VHL), or DCAF16 are linked to a highly specific binder to the protein of interest [230,231]. The design and application of PROTACs is rapidly evolving in part due to the fact that highly selective binders of target protein are available for a variety of target proteins in the form of established small molecules that bind to enzymatic pockets or other protein domains. PROTACs successfully targeting kinases, nuclear receptors, and epigenetic modulators have recently been established [230].
In the space of bromodomain-targeting, the differences between bromodomain-inhibition and targeted bromodomain-protein-degradation have been demonstrated and exemplify the potential of this novel targeting approach. In contrast to BET-competitive inhibition, targeted degradation of BRD4 using the optimized degrader molecule dBET6 led to a global breakdown of transcription, therefore having more rapid and dramatic consequences in cancer cells [236]. The partial selectivity of BRD4-inhibtion for cancer specific enhancer and super-enhancer programs was not observed with this BRD4 degrader. This data demonstrated, that indeed BRD4 is a common essential gene with a non-redundant function in transcriptional elongation in malignant as well as normal cells. Therefore, the degradation of BRD4 has detrimental consequences, indicating a clear mechanistic difference to bromodomain-inhibition. Presumably, these differences are due to a more rapid and profound disruption of the transcription machinery at actively transcribed genes. Even though it might be challenging to manage toxicity of this type of treatment in more advanced pre-clinical stages of development, BET-degradation by dBET6 has been shown to be an efficient and tolerable treatment strategy in a mouse model of glioblastoma, indicating a potential therapeutic window in some transcriptionally driven cancers [237]. Nevertheless, based on the current data, it is questionable if BRD4 degradation is a viable therapeutic approach for targeted therapy rather than a cytotoxic treatment with a therapeutic window in aggressive cancers, similar to chemotherapy. However, these data clearly demonstrate different functional consequences when one uses small molecules to inhibit a protein’s function as compared to complete eradication of the protein.
A clearer example for the potential therapeutic advantages of bromodomain-protein degradation are the pre-clinical observations made in synovial sarcoma regarding targeting of BRD9. As opposed to BRD4, BRD9 is a selective dependency in certain cancer types, including AML and synovial sarcoma [238,239,240]. BRD9 is a part of the mammalian SWI/SNF chromatin remodeling complex, more specifically the ncBAF complex [241,242]. Dysfunctional BAF-complexes have a specific relevance in synovial sarcoma, since the disease is driven by SS18-SSX fusion oncogenes. SS18 itself is a part of mammalian BAF-complexes, therefore, its involvement in oncogenic fusions in synovial sarcoma drives aberrant SWI/SNF function. Importantly, genetic targeting of BRD9 using different single-guide RNAs was detrimental for synovial sarcoma cells. A selective inhibitor of the BRD9-bromodomain on the other hand showed only modest cellular activity at reasonable drug concentrations. ChIP-sequencing demonstrated that BRD9-bromodomain inhibition only led to a sub-complete dissociation of BRD9 from chromatin (Figure 6). When utilizing a PROTAC of BRD9 (dBRD9A), the therapeutic efficacy in synovial sarcoma could be substantially improved due to elimination of BRD9 and a suspected higher degree of ncBAF complex disruption [238]. These findings have prompted the development of orally bioavailable degraders of BRD9. The first clinical trial investigating one of these compounds in synovial sarcoma are projected to start by the end of 2021.

5. Conclusions

Targeting epigenetic modulators for cancer therapy is a highly dynamic field that has led to the pre-clinical and clinical establishment of several molecules during the last decade. A major limitation in this process is the fact that a number of compounds, including most enzymatic inhibitors of epigenetic writers and erasers, have only modest single-agent activity. One opportunity to overcome this limitation is the establishment of efficacious combination therapy regimens. Another important consideration is that most enzymatic inhibitors have only limited potential to disrupt the integrity of chromatin-bound multiprotein complexes, potentially allowing for adaptation and resistance development by recruitment of alternative effectors. The evolving development of inhibitors of protein–protein interactions and targeted protein degraders offers an exciting opportunity to overcome these issues. It has been shown that the degree of physical disruption of certain oncogenic programs is significantly improved using these targeting approaches. Furthermore, proteins without a tractable enzymatic pocket can now be targeted using small molecules. With Menin-MLL1-interaction inhibitors the first class of compounds disrupting epigenetic protein–protein interactions recently entered phase 1 clinical trials. Beyond this, a number of compounds that aim to target previously “undruggable” proteins are now in preclinical and clinical pipelines and will significantly expand our toolbox for pharmacologic interrogation of aberrant epigenetic programs in myeloid malignancies and solid cancers in the years to come.

Author Contributions

F.P. and S.A.A. performed literature research, wrote and edited the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by a grant of the German Research Association (DFG) to FP (PE 3217/1-1). S.A.A is supported by NIH grants CA176745, CA206963, CA204639 and CA066996.

Acknowledgments

The results shown in Section 2.3 are in part based upon data generated by the TCGA Research Network: https://www.cancer.gov/tcga.

Conflicts of Interest

F.P. has no conflict of interests to disclose. S.A.A. has been a consultant and/or shareholder for Epizyme Inc, Vitae/Allergan Pharmaceuticals, Imago Biosciences, Cyteir Therapeutics, C4 Therapeutics, OxStem Oncology, Accent Therapeutics, Neomorph Inc, and Mana Therapeutics. S.A.A. has received research support from Janssen, Novartis, and AstraZeneca.

References

  1. Richmond, T.J.; Davey, C.A. The structure of DNA in the nucleosome core. Nature 2003, 423, 145–150. [Google Scholar] [CrossRef] [PubMed]
  2. Luger, K.; Mader, A.W.; Richmond, R.K.; Sargent, D.F.; Richmond, T.J. Crystal structure of the nucleosome core particle at 2.8 A resolution. Nature 1997, 389, 251–260. [Google Scholar] [CrossRef] [PubMed]
  3. Richmond, T.J.; Finch, J.T.; Rushton, B.; Rhodes, D.; Klug, A. Structure of the nucleosome core particle at 7 A resolution. Nature 1984, 311, 532–537. [Google Scholar] [CrossRef]
  4. Nicodemi, M.; Pombo, A. Models of chromosome structure. Curr. Opin. Cell Biol. 2014, 28, 90–95. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Kornberg, R.D. Chromatin structure: A repeating unit of histones and DNA. Science 1974, 184, 868–871. [Google Scholar] [CrossRef] [PubMed]
  6. Olins, A.L.; Olins, D.E. Spheroid chromatin units (v bodies). Science 1974, 183, 330–332. [Google Scholar] [CrossRef]
  7. Widom, J.; Klug, A. Structure of the 300A chromatin filament: X-ray diffraction from oriented samples. Cell 1985, 43, 207–213. [Google Scholar] [CrossRef]
  8. Belmont, A.S.; Bruce, K. Visualization of G1 chromosomes: A folded, twisted, supercoiled chromonema model of interphase chromatid structure. J. Cell Biol. 1994, 127, 287–302. [Google Scholar] [CrossRef]
  9. Gerchman, S.E.; Ramakrishnan, V. Chromatin higher-order structure studied by neutron scattering and scanning transmission electron microscopy. Proc. Natl. Acad. Sci. USA 1987, 84, 7802–7806. [Google Scholar] [CrossRef] [Green Version]
  10. Vagnarelli, P. Chapter Six—Chromatin Reorganization Through Mitosis. In Advances in Protein Chemistry and Structural Biology; Donev, R., Ed.; Academic Press: Boston, MA, USA, 2013; Volume 90, pp. 179–224. [Google Scholar]
  11. Antonin, W.; Neumann, H. Chromosome condensation and decondensation during mitosis. Curr. Opin. Cell Biol. 2016, 40, 15–22. [Google Scholar] [CrossRef] [Green Version]
  12. Li, B.; Carey, M.; Workman, J.L. The role of chromatin during transcription. Cell 2007, 128, 707–719. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Voss, T.C.; Hager, G.L. Dynamic regulation of transcriptional states by chromatin and transcription factors. Nat. Rev. Genet. 2014, 15, 69–81. [Google Scholar] [CrossRef] [PubMed]
  14. Rando, O.J.; Winston, F. Chromatin and transcription in yeast. Genetics 2012, 190, 351–387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Valencia, A.M.; Kadoch, C. Chromatin regulatory mechanisms and therapeutic opportunities in cancer. Nat. Cell Biol. 2019, 21, 152–161. [Google Scholar] [CrossRef]
  16. Morgan, M.A.; Shilatifard, A. Chromatin signatures of cancer. Genes Dev. 2015, 29, 238–249. [Google Scholar] [CrossRef] [Green Version]
  17. Chi, P.; Allis, C.D.; Wang, G.G. Covalent histone modifications–miswritten, misinterpreted and mis-erased in human cancers. Nat. Rev. Cancer 2010, 10, 457–469. [Google Scholar] [CrossRef] [Green Version]
  18. Kouzarides, T. Chromatin modifications and their function. Cell 2007, 128, 693–705. [Google Scholar] [CrossRef] [Green Version]
  19. Hyun, K.; Jeon, J.; Park, K.; Kim, J. Writing, erasing and reading histone lysine methylations. Exp. Mol. Med. 2017, 49, e324. [Google Scholar] [CrossRef] [Green Version]
  20. Wang, Z.; Zang, C.; Rosenfeld, J.A.; Schones, D.E.; Barski, A.; Cuddapah, S.; Cui, K.; Roh, T.Y.; Peng, W.; Zhang, M.Q.; et al. Combinatorial patterns of histone acetylations and methylations in the human genome. Nat. Genet. 2008, 40, 897–903. [Google Scholar] [CrossRef] [Green Version]
  21. Tan, M.; Luo, H.; Lee, S.; Jin, F.; Yang, J.S.; Montellier, E.; Buchou, T.; Cheng, Z.; Rousseaux, S.; Rajagopal, N.; et al. Identification of 67 histone marks and histone lysine crotonylation as a new type of histone modification. Cell 2011, 146, 1016–1028. [Google Scholar] [CrossRef] [Green Version]
  22. Strahl, B.D.; Allis, C.D. The language of covalent histone modifications. Nature 2000, 403, 41–45. [Google Scholar] [CrossRef] [PubMed]
  23. Wolffe, A.P.; Hayes, J.J. Chromatin disruption and modification. Nucleic Acids Res. 1999, 27, 711–720. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Bauer, W.R.; Hayes, J.J.; White, J.H.; Wolffe, A.P. Nucleosome structural changes due to acetylation. J. Mol. Biol. 1994, 236, 685–690. [Google Scholar] [CrossRef] [PubMed]
  25. Krajewski, W.A.; Becker, P.B. Reconstitution of hyperacetylated, DNase I-sensitive chromatin characterized by high conformational flexibility of nucleosomal DNA. Proc. Natl. Acad. Sci. USA 1998, 95, 1540–1545. [Google Scholar] [CrossRef] [Green Version]
  26. Kuo, M.H.; Allis, C.D. Roles of histone acetyltransferases and deacetylases in gene regulation. Bioessays 1998, 20, 615–626. [Google Scholar] [CrossRef]
  27. Hong, L.; Schroth, G.P.; Matthews, H.R.; Yau, P.; Bradbury, E.M. Studies of the DNA binding properties of histone H4 amino terminus. Thermal denaturation studies reveal that acetylation markedly reduces the binding constant of the H4 “tail” to DNA. J. Biol. Chem. 1993, 268, 305–314. [Google Scholar]
  28. Brownell, J.E.; Zhou, J.; Ranalli, T.; Kobayashi, R.; Edmondson, D.G.; Roth, S.Y.; Allis, C.D. Tetrahymena histone acetyltransferase A: A homolog to yeast Gcn5p linking histone acetylation to gene activation. Cell 1996, 84, 843–851. [Google Scholar] [CrossRef] [Green Version]
  29. Durrin, L.K.; Mann, R.K.; Kayne, P.S.; Grunstein, M. Yeast histone H4 N-terminal sequence is required for promoter activation in vivo. Cell 1991, 65, 1023–1031. [Google Scholar] [CrossRef]
  30. Turner, B.M. Decoding the nucleosome. Cell 1993, 75, 5–8. [Google Scholar] [CrossRef]
  31. Grunstein, M. Histone acetylation in chromatin structure and transcription. Nature 1997, 389, 349–352. [Google Scholar] [CrossRef]
  32. Mizzen, C.; Kuo, M.H.; Smith, E.; Brownell, J.; Zhou, J.; Ohba, R.; Wei, Y.; Monaco, L.; Sassone-Corsi, P.; Allis, C.D. Signaling to chromatin through histone modifications: How clear is the signal? Cold Spring Harb. Symp. Quant. Biol. 1998, 63, 469–481. [Google Scholar] [CrossRef] [PubMed]
  33. Lopez-Rodas, G.; Brosch, G.; Georgieva, E.I.; Sendra, R.; Franco, L.; Loidl, P. Histone deacetylase. A key enzyme for the binding of regulatory proteins to chromatin. FEBS Lett. 1993, 317, 175–180. [Google Scholar] [CrossRef] [Green Version]
  34. Tordera, V.; Sendra, R.; Perez-Ortin, J.E. The role of histones and their modifications in the informative content of chromatin. Experientia 1993, 49, 780–788. [Google Scholar] [CrossRef] [PubMed]
  35. Dancy, B.M.; Cole, P.A. Protein lysine acetylation by p300/CBP. Chem. Rev. 2015, 115, 2419–2452. [Google Scholar] [CrossRef]
  36. Chan, H.M.; La Thangue, N.B. p300/CBP proteins: HATs for transcriptional bridges and scaffolds. J. Cell Sci. 2001, 114, 2363–2373. [Google Scholar]
  37. Brown, J.D.; Feldman, Z.B.; Doherty, S.P.; Reyes, J.M.; Rahl, P.B.; Lin, C.Y.; Sheng, Q.; Duan, Q.; Federation, A.J.; Kung, A.L.; et al. BET bromodomain proteins regulate enhancer function during adipogenesis. Proc. Natl. Acad. Sci. USA 2018, 115, 2144–2149. [Google Scholar] [CrossRef] [Green Version]
  38. Lee, J.E.; Park, Y.K.; Park, S.; Jang, Y.; Waring, N.; Dey, A.; Ozato, K.; Lai, B.; Peng, W.; Ge, K. Brd4 binds to active enhancers to control cell identity gene induction in adipogenesis and myogenesis. Nat. Commun. 2017, 8, 2217. [Google Scholar] [CrossRef]
  39. Chapuy, B.; McKeown, M.R.; Lin, C.Y.; Monti, S.; Roemer, M.G.; Qi, J.; Rahl, P.B.; Sun, H.H.; Yeda, K.T.; Doench, J.G.; et al. Discovery and characterization of super-enhancer-associated dependencies in diffuse large B cell lymphoma. Cancer Cell 2013, 24, 777–790. [Google Scholar] [CrossRef] [Green Version]
  40. Loven, J.; Hoke, H.A.; Lin, C.Y.; Lau, A.; Orlando, D.A.; Vakoc, C.R.; Bradner, J.E.; Lee, T.I.; Young, R.A. Selective inhibition of tumor oncogenes by disruption of super-enhancers. Cell 2013, 153, 320–334. [Google Scholar] [CrossRef] [Green Version]
  41. Yang, Z.; Yik, J.H.; Chen, R.; He, N.; Jang, M.K.; Ozato, K.; Zhou, Q. Recruitment of P-TEFb for stimulation of transcriptional elongation by the bromodomain protein Brd4. Mol. Cell 2005, 19, 535–545. [Google Scholar] [CrossRef]
  42. Bolden, J.E.; Tasdemir, N.; Dow, L.E.; van Es, J.H.; Wilkinson, J.E.; Zhao, Z.; Clevers, H.; Lowe, S.W. Inducible in vivo silencing of Brd4 identifies potential toxicities of sustained BET protein inhibition. Cell Rep. 2014, 8, 1919–1929. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Zuber, J.; Shi, J.; Wang, E.; Rappaport, A.R.; Herrmann, H.; Sison, E.A.; Magoon, D.; Qi, J.; Blatt, K.; Wunderlich, M.; et al. RNAi screen identifies Brd4 as a therapeutic target in acute myeloid leukaemia. Nature 2011, 478, 524–528. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Filippakopoulos, P.; Qi, J.; Picaud, S.; Shen, Y.; Smith, W.B.; Fedorov, O.; Morse, E.M.; Keates, T.; Hickman, T.T.; Felletar, I.; et al. Selective inhibition of BET bromodomains. Nature 2010, 468, 1067–1073. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Raisner, R.; Kharbanda, S.; Jin, L.; Jeng, E.; Chan, E.; Merchant, M.; Haverty, P.M.; Bainer, R.; Cheung, T.; Arnott, D.; et al. Enhancer Activity Requires CBP/P300 Bromodomain-Dependent Histone H3K27 Acetylation. Cell Rep. 2018, 24, 1722–1729. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Lasko, L.M.; Jakob, C.G.; Edalji, R.P.; Qiu, W.; Montgomery, D.; Digiammarino, E.L.; Hansen, T.M.; Risi, R.M.; Frey, R.; Manaves, V.; et al. Discovery of a selective catalytic p300/CBP inhibitor that targets lineage-specific tumours. Nature 2017, 550, 128–132. [Google Scholar] [CrossRef] [PubMed]
  47. Wilson, S.; Filipp, F.V. A network of epigenomic and transcriptional cooperation encompassing an epigenomic master regulator in cancer. NPJ Syst. Biol. Appl. 2018, 4, 24. [Google Scholar] [CrossRef] [PubMed]
  48. Wang, Y.; Jiang, R.; Wong, W.H. Modeling the causal regulatory network by integrating chromatin accessibility and transcriptome data. Natl. Sci. Rev. 2016, 3, 240–251. [Google Scholar] [CrossRef] [Green Version]
  49. Tu, X.; Mejia-Guerra, M.K.; Valdes Franco, J.A.; Tzeng, D.; Chu, P.Y.; Shen, W.; Wei, Y.; Dai, X.; Li, P.; Buckler, E.S.; et al. Reconstructing the maize leaf regulatory network using ChIP-seq data of 104 transcription factors. Nat. Commun. 2020, 11, 5089. [Google Scholar] [CrossRef]
  50. Malysheva, V.; Mendoza-Parra, M.A.; Blum, M.; Spivakov, M.; Gronemeyer, H. Gene regulatory network reconstruction incorporating 3D chromosomal architecture reveals key transcription factors and DNA elements driving neural lineage commitment. bioRxiv 2019, 303842. [Google Scholar] [CrossRef] [Green Version]
  51. Ashoor, H.; Chen, X.; Rosikiewicz, W.; Wang, J.; Cheng, A.; Wang, P.; Ruan, Y.; Li, S. Graph embedding and unsupervised learning predict genomic sub-compartments from HiC chromatin interaction data. Nat. Commun. 2020, 11, 1173. [Google Scholar] [CrossRef] [Green Version]
  52. Hnisz, D.; Day, D.S.; Young, R.A. Insulated Neighborhoods: Structural and Functional Units of Mammalian Gene Control. Cell 2016, 167, 1188–1200. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Atlasi, Y.; Stunnenberg, H.G. The interplay of epigenetic marks during stem cell differentiation and development. Nat. Rev. Genet. 2017, 18, 643–658. [Google Scholar] [CrossRef] [PubMed]
  54. Amit, I.; Winter, D.R.; Jung, S. The role of the local environment and epigenetics in shaping macrophage identity and their effect on tissue homeostasis. Nat. Immunol. 2016, 17, 18–25. [Google Scholar] [CrossRef]
  55. Piunti, A.; Shilatifard, A. Epigenetic balance of gene expression by Polycomb and COMPASS families. Science 2016, 352, aad9780. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Ho, L.; Crabtree, G.R. Chromatin remodelling during development. Nature 2010, 463, 474–484. [Google Scholar] [CrossRef] [PubMed]
  57. Schuettengruber, B.; Bourbon, H.M.; Di Croce, L.; Cavalli, G. Genome Regulation by Polycomb and Trithorax: 70 Years and Counting. Cell 2017, 171, 34–57. [Google Scholar] [CrossRef] [Green Version]
  58. Long, H.K.; Prescott, S.L.; Wysocka, J. Ever-Changing Landscapes: Transcriptional Enhancers in Development and Evolution. Cell 2016, 167, 1170–1187. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Papp, B.; Plath, K. Epigenetics of reprogramming to induced pluripotency. Cell 2013, 152, 1324–1343. [Google Scholar] [CrossRef] [Green Version]
  60. Surani, M.A.; Hayashi, K.; Hajkova, P. Genetic and epigenetic regulators of pluripotency. Cell 2007, 128, 747–762. [Google Scholar] [CrossRef] [Green Version]
  61. Yadav, T.; Quivy, J.P.; Almouzni, G. Chromatin plasticity: A versatile landscape that underlies cell fate and identity. Science 2018, 361, 1332–1336. [Google Scholar] [CrossRef] [Green Version]
  62. Apostolou, E.; Hochedlinger, K. Chromatin dynamics during cellular reprogramming. Nature 2013, 502, 462–471. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Johnson, D.G.; Dent, S.Y. Chromatin: Receiver and quarterback for cellular signals. Cell 2013, 152, 685–689. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Sen, P.; Shah, P.P.; Nativio, R.; Berger, S.L. Epigenetic Mechanisms of Longevity and Aging. Cell 2016, 166, 822–839. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Dulac, C. Brain function and chromatin plasticity. Nature 2010, 465, 728–735. [Google Scholar] [CrossRef] [Green Version]
  66. Feinberg, A.P.; Koldobskiy, M.A.; Gondor, A. Epigenetic modulators, modifiers and mediators in cancer aetiology and progression. Nat. Rev. Genet. 2016, 17, 284–299. [Google Scholar] [CrossRef]
  67. Kinnaird, A.; Zhao, S.; Wellen, K.E.; Michelakis, E.D. Metabolic control of epigenetics in cancer. Nat. Rev. Cancer 2016, 16, 694–707. [Google Scholar] [CrossRef]
  68. Flavahan, W.A.; Gaskell, E.; Bernstein, B.E. Epigenetic plasticity and the hallmarks of cancer. Science 2017, 357. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Dawson, M.A. The cancer epigenome: Concepts, challenges, and therapeutic opportunities. Science 2017, 355, 1147–1152. [Google Scholar] [CrossRef]
  70. Suva, M.L.; Riggi, N.; Bernstein, B.E. Epigenetic reprogramming in cancer. Science 2013, 339, 1567–1570. [Google Scholar] [CrossRef] [Green Version]
  71. Fathi, A.T.; Abdel-Wahab, O. Mutations in epigenetic modifiers in myeloid malignancies and the prospect of novel epigenetic-targeted therapy. Adv. Hematol. 2012, 2012, 469592. [Google Scholar] [CrossRef] [Green Version]
  72. Shih, A.H.; Abdel-Wahab, O.; Patel, J.P.; Levine, R.L. The role of mutations in epigenetic regulators in myeloid malignancies. Nat. Rev. Cancer 2012, 12, 599–612. [Google Scholar] [CrossRef] [PubMed]
  73. Woods, B.A.; Levine, R.L. The role of mutations in epigenetic regulators in myeloid malignancies. Immunol. Rev. 2015, 263, 22–35. [Google Scholar] [CrossRef] [PubMed]
  74. Wouters, B.J.; Delwel, R. Epigenetics and approaches to targeted epigenetic therapy in acute myeloid leukemia. Blood 2016, 127, 42–52. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Jaiswal, S.; Fontanillas, P.; Flannick, J.; Manning, A.; Grauman, P.V.; Mar, B.G.; Lindsley, R.C.; Mermel, C.H.; Burtt, N.; Chavez, A.; et al. Age-related clonal hematopoiesis associated with adverse outcomes. N. Engl. J. Med. 2014, 371, 2488–2498. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Shlush, L.I.; Zandi, S.; Mitchell, A.; Chen, W.C.; Brandwein, J.M.; Gupta, V.; Kennedy, J.A.; Schimmer, A.D.; Schuh, A.C.; Yee, K.W.; et al. Identification of pre-leukaemic haematopoietic stem cells in acute leukaemia. Nature 2014, 506, 328–333. [Google Scholar] [CrossRef] [PubMed]
  77. Chan, S.M.; Majeti, R. Role of DNMT3A, TET2, and IDH1/2 mutations in pre-leukemic stem cells in acute myeloid leukemia. Int. J. Hematol. 2013, 98, 648–657. [Google Scholar] [CrossRef]
  78. Corces-Zimmerman, M.R.; Majeti, R. Pre-leukemic evolution of hematopoietic stem cells: The importance of early mutations in leukemogenesis. Leukemia 2014, 28, 2276–2282. [Google Scholar] [CrossRef] [PubMed]
  79. Jaiswal, S.; Ebert, B.L. Clonal hematopoiesis in human aging and disease. Science 2019, 366. [Google Scholar] [CrossRef]
  80. Jaiswal, S.; Natarajan, P.; Silver, A.J.; Gibson, C.J.; Bick, A.G.; Shvartz, E.; McConkey, M.; Gupta, N.; Gabriel, S.; Ardissino, D.; et al. Clonal Hematopoiesis and Risk of Atherosclerotic Cardiovascular Disease. N. Engl. J. Med. 2017, 377, 111–121. [Google Scholar] [CrossRef]
  81. Miles, L.A.; Bowman, R.L.; Merlinsky, T.R.; Csete, I.S.; Ooi, A.T.; Durruthy-Durruthy, R.; Bowman, M.; Famulare, C.; Patel, M.A.; Mendez, P.; et al. Single-cell mutation analysis of clonal evolution in myeloid malignancies. Nature 2020. [Google Scholar] [CrossRef]
  82. Bennett, R.L.; Licht, J.D. Targeting Epigenetics in Cancer. Annu. Rev. Pharmacol. Toxicol. 2018, 58, 187–207. [Google Scholar] [CrossRef] [PubMed]
  83. Cheng, Y.; He, C.; Wang, M.; Ma, X.; Mo, F.; Yang, S.; Han, J.; Wei, X. Targeting epigenetic regulators for cancer therapy: Mechanisms and advances in clinical trials. Signal Transduct. Target. Ther. 2019, 4, 62. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Ganesan, A.; Arimondo, P.B.; Rots, M.G.; Jeronimo, C.; Berdasco, M. The timeline of epigenetic drug discovery: From reality to dreams. Clin. Epigenet. 2019, 11, 174. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Park, J.W.; Han, J.W. Targeting epigenetics for cancer therapy. Arch. Pharm. Res. 2019, 42, 159–170. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Holdgate, G.A.; Meek, T.D.; Grimley, R.L. Mechanistic enzymology in drug discovery: A fresh perspective. Nat. Rev. Drug Discov. 2018, 17, 115–132. [Google Scholar] [CrossRef] [PubMed]
  87. Copeland, R.A.; Harpel, M.R.; Tummino, P.J. Targeting enzyme inhibitors in drug discovery. Expert Opin. Ther. Targets 2007, 11, 967–978. [Google Scholar] [CrossRef]
  88. Silverman, R.B.; Holladay, M.W. Chapter 5—Enzyme Inhibition and Inactivation. In The Organic Chemistry of Drug Design and Drug Action, 3rd ed.; Silverman, R.B., Holladay, M.W., Eds.; Academic Press: Boston, MA, USA, 2014; pp. 207–274. [Google Scholar]
  89. Marks, P.A.; Breslow, R. Dimethyl sulfoxide to vorinostat: Development of this histone deacetylase inhibitor as an anticancer drug. Nat. Biotechnol. 2007, 25, 84–90. [Google Scholar] [CrossRef]
  90. Bose, P.; Verstovsek, S. Investigational histone deacetylase inhibitors (HDACi) in myeloproliferative neoplasms. Expert Opin. Investig. Drugs 2016, 25, 1393–1403. [Google Scholar] [CrossRef]
  91. San Jose-Eneriz, E.; Gimenez-Camino, N.; Agirre, X.; Prosper, F. HDAC Inhibitors in Acute Myeloid Leukemia. Cancers 2019, 11, 1794. [Google Scholar] [CrossRef] [Green Version]
  92. Quintas-Cardama, A.; Santos, F.P.; Garcia-Manero, G. Histone deacetylase inhibitors for the treatment of myelodysplastic syndrome and acute myeloid leukemia. Leukemia 2011, 25, 226–235. [Google Scholar] [CrossRef] [Green Version]
  93. Kuroda, M.I.; Kang, H.; De, S.; Kassis, J.A. Dynamic Competition of Polycomb and Trithorax in Transcriptional Programming. Annu. Rev. Biochem. 2020. [Google Scholar] [CrossRef] [PubMed]
  94. Kingston, R.E.; Tamkun, J.W. Transcriptional regulation by trithorax-group proteins. Cold Spring Harb. Perspect. Biol. 2014, 6, a019349. [Google Scholar] [CrossRef] [PubMed]
  95. Yokoyama, A.; Wang, Z.; Wysocka, J.; Sanyal, M.; Aufiero, D.J.; Kitabayashi, I.; Herr, W.; Cleary, M.L. Leukemia proto-oncoprotein MLL forms a SET1-like histone methyltransferase complex with menin to regulate Hox gene expression. Mol. Cell. Biol. 2004, 24, 5639–5649. [Google Scholar] [CrossRef] [Green Version]
  96. Milne, T.A.; Briggs, S.D.; Brock, H.W.; Martin, M.E.; Gibbs, D.; Allis, C.D.; Hess, J.L. MLL targets SET domain methyltransferase activity to Hox gene promoters. Mol. Cell 2002, 10, 1107–1117. [Google Scholar] [CrossRef]
  97. Yu, B.D.; Hess, J.L.; Horning, S.E.; Brown, G.A.; Korsmeyer, S.J. Altered Hox expression and segmental identity in Mll-mutant mice. Nature 1995, 378, 505–508. [Google Scholar] [CrossRef] [PubMed]
  98. Djabali, M.; Selleri, L.; Parry, P.; Bower, M.; Young, B.D.; Evans, G.A. A trithorax-like gene is interrupted by chromosome 11q23 translocations in acute leukaemias. Nat. Genet. 1992, 2, 113–118. [Google Scholar] [CrossRef] [PubMed]
  99. Tkachuk, D.C.; Kohler, S.; Cleary, M.L. Involvement of a homolog of Drosophila trithorax by 11q23 chromosomal translocations in acute leukemias. Cell 1992, 71, 691–700. [Google Scholar] [CrossRef]
  100. Vidal, M.; Starowicz, K. Polycomb complexes PRC1 and their function in hematopoiesis. Exp. Hematol. 2017, 48, 12–31. [Google Scholar] [CrossRef] [Green Version]
  101. Rossi, A.; Ferrari, K.J.; Piunti, A.; Jammula, S.; Chiacchiera, F.; Mazzarella, L.; Scelfo, A.; Pelicci, P.G.; Pasini, D. Maintenance of leukemic cell identity by the activity of the Polycomb complex PRC1 in mice. Sci. Adv. 2016, 2, e1600972. [Google Scholar] [CrossRef] [Green Version]
  102. Jiao, L.; Liu, X. Structural basis of histone H3K27 trimethylation by an active polycomb repressive complex 2. Science 2015, 350, aac4383. [Google Scholar] [CrossRef] [Green Version]
  103. Cao, R.; Wang, L.; Wang, H.; Xia, L.; Erdjument-Bromage, H.; Tempst, P.; Jones, R.S.; Zhang, Y. Role of histone H3 lysine 27 methylation in Polycomb-group silencing. Science 2002, 298, 1039–1043. [Google Scholar] [CrossRef] [Green Version]
  104. Onder, T.T.; Kara, N.; Cherry, A.; Sinha, A.U.; Zhu, N.; Bernt, K.M.; Cahan, P.; Marcarci, B.O.; Unternaehrer, J.; Gupta, P.B.; et al. Chromatin-modifying enzymes as modulators of reprogramming. Nature 2012, 483, 598–602. [Google Scholar] [CrossRef]
  105. Cyrus, S.; Burkardt, D.; Weaver, D.D.; Gibson, W.T. PRC2-complex related dysfunction in overgrowth syndromes: A review of EZH2, EED, and SUZ12 and their syndromic phenotypes. Am. J. Med. Genet. C Semin Med. Genet. 2019, 181, 519–531. [Google Scholar] [CrossRef] [PubMed]
  106. Fillmore, C.M.; Xu, C.; Desai, P.T.; Berry, J.M.; Rowbotham, S.P.; Lin, Y.J.; Zhang, H.; Marquez, V.E.; Hammerman, P.S.; Wong, K.K.; et al. EZH2 inhibition sensitizes BRG1 and EGFR mutant lung tumours to TopoII inhibitors. Nature 2015, 520, 239–242. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Roberts, T.C.; Etxaniz, U.; Dall’Agnese, A.; Wu, S.Y.; Chiang, C.M.; Brennan, P.E.; Wood, M.J.A.; Puri, P.L. BRD3 and BRD4 BET Bromodomain Proteins Differentially Regulate Skeletal Myogenesis. Sci. Rep. 2017, 7, 6153. [Google Scholar] [CrossRef] [PubMed]
  108. Lewis, P.W.; Muller, M.M.; Koletsky, M.S.; Cordero, F.; Lin, S.; Banaszynski, L.A.; Garcia, B.A.; Muir, T.W.; Becher, O.J.; Allis, C.D. Inhibition of PRC2 activity by a gain-of-function H3 mutation found in pediatric glioblastoma. Science 2013, 340, 857–861. [Google Scholar] [CrossRef] [Green Version]
  109. Rothbart, S.B.; Baylin, S.B. Epigenetic Therapy for Epithelioid Sarcoma. Cell 2020, 181, 211. [Google Scholar] [CrossRef]
  110. McCabe, M.T.; Ott, H.M.; Ganji, G.; Korenchuk, S.; Thompson, C.; Van Aller, G.S.; Liu, Y.; Graves, A.P.; Della Pietra, A., 3rd; Diaz, E.; et al. EZH2 inhibition as a therapeutic strategy for lymphoma with EZH2-activating mutations. Nature 2012, 492, 108–112. [Google Scholar] [CrossRef]
  111. Eich, M.-L.; Athar, M.; Ferguson, J.E.; Varambally, S. EZH2-Targeted Therapies in Cancer: Hype or a Reality. Cancer Res. 2020. [Google Scholar] [CrossRef]
  112. Varambally, S.; Dhanasekaran, S.M.; Zhou, M.; Barrette, T.R.; Kumar-Sinha, C.; Sanda, M.G.; Ghosh, D.; Pienta, K.J.; Sewalt, R.G.; Otte, A.P.; et al. The polycomb group protein EZH2 is involved in progression of prostate cancer. Nature 2002, 419, 624–629. [Google Scholar] [CrossRef]
  113. Bodor, C.; Grossmann, V.; Popov, N.; Okosun, J.; O’Riain, C.; Tan, K.; Marzec, J.; Araf, S.; Wang, J.; Lee, A.M.; et al. EZH2 mutations are frequent and represent an early event in follicular lymphoma. Blood 2013, 122, 3165–3168. [Google Scholar] [CrossRef] [PubMed]
  114. Kruger, R.G.; Graves, A.P.; McCabe, M.T. Chapter 11—Activating Mutations of the EZH2 Histone Methyltransferase in Cancer. In Polycomb Group Proteins; Pirrotta, V., Ed.; Academic Press: Boston, MA, USA, 2017; pp. 259–288. [Google Scholar]
  115. Kim, K.H.; Roberts, C.W. Targeting EZH2 in cancer. Nat. Med. 2016, 22, 128–134. [Google Scholar] [CrossRef] [PubMed]
  116. Bachmann, I.M.; Halvorsen, O.J.; Collett, K.; Stefansson, I.M.; Straume, O.; Haukaas, S.A.; Salvesen, H.B.; Otte, A.P.; Akslen, L.A. EZH2 expression is associated with high proliferation rate and aggressive tumor subgroups in cutaneous melanoma and cancers of the endometrium, prostate, and breast. J. Clin. Oncol. 2006, 24, 268–273. [Google Scholar] [CrossRef]
  117. Lu, C.; Allis, C.D. SWI/SNF complex in cancer. Nat. Genet. 2017, 49, 178–179. [Google Scholar] [CrossRef] [PubMed]
  118. Orlando, K.A.; Nguyen, V.; Raab, J.R.; Walhart, T.; Weissman, B.E. Remodeling the cancer epigenome: Mutations in the SWI/SNF complex offer new therapeutic opportunities. Expert Rev. Anticancer Ther. 2019, 19, 375–391. [Google Scholar] [CrossRef]
  119. Wang, X.; Lee, R.S.; Alver, B.H.; Haswell, J.R.; Wang, S.; Mieczkowski, J.; Drier, Y.; Gillespie, S.M.; Archer, T.C.; Wu, J.N.; et al. SMARCB1-mediated SWI/SNF complex function is essential for enhancer regulation. Nat. Genet. 2017, 49, 289–295. [Google Scholar] [CrossRef]
  120. St Pierre, R.; Kadoch, C. Mammalian SWI/SNF complexes in cancer: Emerging therapeutic opportunities. Curr. Opin. Genet. Dev. 2017, 42, 56–67. [Google Scholar] [CrossRef] [Green Version]
  121. Wilson, B.G.; Roberts, C.W. SWI/SNF nucleosome remodellers and cancer. Nat. Rev. Cancer 2011, 11, 481–492. [Google Scholar] [CrossRef]
  122. Wilson, B.G.; Wang, X.; Shen, X.; McKenna, E.S.; Lemieux, M.E.; Cho, Y.J.; Koellhoffer, E.C.; Pomeroy, S.L.; Orkin, S.H.; Roberts, C.W. Epigenetic antagonism between polycomb and SWI/SNF complexes during oncogenic transformation. Cancer Cell 2010, 18, 316–328. [Google Scholar] [CrossRef] [Green Version]
  123. Knutson, S.K.; Warholic, N.M.; Wigle, T.J.; Klaus, C.R.; Allain, C.J.; Raimondi, A.; Porter Scott, M.; Chesworth, R.; Moyer, M.P.; Copeland, R.A.; et al. Durable tumor regression in genetically altered malignant rhabdoid tumors by inhibition of methyltransferase EZH2. Proc. Natl. Acad. Sci. USA 2013, 110, 7922–7927. [Google Scholar] [CrossRef] [Green Version]
  124. Ernst, T.; Chase, A.J.; Score, J.; Hidalgo-Curtis, C.E.; Bryant, C.; Jones, A.V.; Waghorn, K.; Zoi, K.; Ross, F.M.; Reiter, A.; et al. Inactivating mutations of the histone methyltransferase gene EZH2 in myeloid disorders. Nat. Genet. 2010, 42, 722–726. [Google Scholar] [CrossRef] [PubMed]
  125. Nikoloski, G.; Langemeijer, S.M.; Kuiper, R.P.; Knops, R.; Massop, M.; Tonnissen, E.R.; van der Heijden, A.; Scheele, T.N.; Vandenberghe, P.; de Witte, T.; et al. Somatic mutations of the histone methyltransferase gene EZH2 in myelodysplastic syndromes. Nat. Genet. 2010, 42, 665–667. [Google Scholar] [CrossRef] [PubMed]
  126. Score, J.; Hidalgo-Curtis, C.; Jones, A.V.; Winkelmann, N.; Skinner, A.; Ward, D.; Zoi, K.; Ernst, T.; Stegelmann, F.; Dohner, K.; et al. Inactivation of polycomb repressive complex 2 components in myeloproliferative and myelodysplastic/myeloproliferative neoplasms. Blood 2012, 119, 1208–1213. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Bejar, R.; Stevenson, K.; Abdel-Wahab, O.; Galili, N.; Nilsson, B.; Garcia-Manero, G.; Kantarjian, H.; Raza, A.; Levine, R.L.; Neuberg, D.; et al. Clinical effect of point mutations in myelodysplastic syndromes. N. Engl. J. Med. 2011, 364, 2496–2506. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Zhang, Q.; Han, Q.; Zi, J.; Ma, J.; Song, H.; Tian, Y.; McGrath, M.; Song, C.; Ge, Z. Mutations in EZH2 are associated with poor prognosis for patients with myeloid neoplasms. Genes Dis. 2019, 6, 276–281. [Google Scholar] [CrossRef]
  129. Guglielmelli, P.; Biamonte, F.; Score, J.; Hidalgo-Curtis, C.; Cervantes, F.; Maffioli, M.; Fanelli, T.; Ernst, T.; Winkelman, N.; Jones, A.V.; et al. EZH2 mutational status predicts poor survival in myelofibrosis. Blood 2011, 118, 5227–5234. [Google Scholar] [CrossRef]
  130. Mechaal, A.; Menif, S.; Abbes, S.; Safra, I. EZH2, new diagnosis and prognosis marker in acute myeloid leukemia patients. Adv. Med. Sci. 2019, 64, 395–401. [Google Scholar] [CrossRef]
  131. Hu, D.; Shilatifard, A. Epigenetics of hematopoiesis and hematological malignancies. Genes Dev. 2016, 30, 2021–2041. [Google Scholar] [CrossRef]
  132. Shilatifard, A. The COMPASS family of histone H3K4 methylases: Mechanisms of regulation in development and disease pathogenesis. Annu. Rev. Biochem. 2012, 81, 65–95. [Google Scholar] [CrossRef] [Green Version]
  133. Wang, P.; Lin, C.; Smith, E.R.; Guo, H.; Sanderson, B.W.; Wu, M.; Gogol, M.; Alexander, T.; Seidel, C.; Wiedemann, L.M.; et al. Global analysis of H3K4 methylation defines MLL family member targets and points to a role for MLL1-mediated H3K4 methylation in the regulation of transcriptional initiation by RNA polymerase II. Mol. Cell Biol. 2009, 29, 6074–6085. [Google Scholar] [CrossRef] [Green Version]
  134. Scott, M.T.; Korfi, K.; Saffrey, P.; Hopcroft, L.E.; Kinstrie, R.; Pellicano, F.; Guenther, C.; Gallipoli, P.; Cruz, M.; Dunn, K.; et al. Epigenetic Reprogramming Sensitizes CML Stem Cells to Combined EZH2 and Tyrosine Kinase Inhibition. Cancer Discov. 2016, 6, 1248–1257. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Xie, H.; Peng, C.; Huang, J.; Li, B.E.; Kim, W.; Smith, E.C.; Fujiwara, Y.; Qi, J.; Cheloni, G.; Das, P.P.; et al. Chronic Myelogenous Leukemia- Initiating Cells Require Polycomb Group Protein EZH2. Cancer Discov. 2016, 6, 1237–1247. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Neff, T.; Sinha, A.U.; Kluk, M.J.; Zhu, N.; Khattab, M.H.; Stein, L.; Xie, H.; Orkin, S.H.; Armstrong, S.A. Polycomb repressive complex 2 is required for MLL-AF9 leukemia. Proc. Natl. Acad. Sci. USA 2012, 109, 5028–5033. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Rathert, P.; Roth, M.; Neumann, T.; Muerdter, F.; Roe, J.S.; Muhar, M.; Deswal, S.; Cerny-Reiterer, S.; Peter, B.; Jude, J.; et al. Transcriptional plasticity promotes primary and acquired resistance to BET inhibition. Nature 2015, 525, 543–547. [Google Scholar] [CrossRef] [PubMed]
  138. Italiano, A.; Soria, J.C.; Toulmonde, M.; Michot, J.M.; Lucchesi, C.; Varga, A.; Coindre, J.M.; Blakemore, S.J.; Clawson, A.; Suttle, B.; 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. [Google Scholar] [CrossRef]
  139. Stacchiotti, S.; Blay, J.Y.; Jones, R.L.; Demetri, G.D.D.; Mir, O.; Italiano, A.; Thomas, D.; Chen, T.W.W.; Schöffski, P.; Gil, T.; et al. A phase II, multicenter study of the EZH2 inhibitor tazemetostat in adults (INI1-negative tumors cohort) (NCT02601950). Ann. Oncol. 2018, 29, viii580. [Google Scholar] [CrossRef]
  140. Gulati, N.; Beguelin, W.; Giulino-Roth, L. Enhancer of zeste homolog 2 (EZH2) inhibitors. Leuk. Lymphoma 2018, 59, 1574–1585. [Google Scholar] [CrossRef]
  141. Yap, T.A.; Winter, J.N.; Giulino-Roth, L.; Longley, J.; Lopez, J.; Michot, J.M.; Leonard, J.P.; Ribrag, V.; McCabe, M.T.; Creasy, C.L.; et al. Phase I Study of the Novel Enhancer of Zeste Homolog 2 (EZH2) Inhibitor GSK2816126 in Patients with Advanced Hematologic and Solid Tumors. Clin. Cancer Res. 2019, 25, 7331–7339. [Google Scholar] [CrossRef] [Green Version]
  142. Frankel, A.E.; Liu, X.; Minna, J.D. Developing EZH2-Targeted Therapy for Lung Cancer. Cancer Discov. 2016, 6, 949–952. [Google Scholar] [CrossRef] [Green Version]
  143. Vaswani, R.G.; Gehling, V.S.; Dakin, L.A.; Cook, A.S.; Nasveschuk, C.G.; Duplessis, M.; Iyer, P.; Balasubramanian, S.; Zhao, F.; Good, A.C.; 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. [Google Scholar] [CrossRef] [Green Version]
  144. Steger, D.J.; Lefterova, M.I.; Ying, L.; Stonestrom, A.J.; Schupp, M.; Zhuo, D.; Vakoc, A.L.; Kim, J.E.; Chen, J.; Lazar, M.A.; et al. DOT1L/KMT4 recruitment and H3K79 methylation are ubiquitously coupled with gene transcription in mammalian cells. Mol. Cell Biol. 2008, 28, 2825–2839. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Kouskouti, A.; Talianidis, I. Histone modifications defining active genes persist after transcriptional and mitotic inactivation. EMBO J. 2005, 24, 347–357. [Google Scholar] [CrossRef] [Green Version]
  146. Barski, A.; Cuddapah, S.; Cui, K.; Roh, T.Y.; Schones, D.E.; Wang, Z.; Wei, G.; Chepelev, I.; Zhao, K. High-resolution profiling of histone methylations in the human genome. Cell 2007, 129, 823–837. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Nguyen, A.T.; He, J.; Taranova, O.; Zhang, Y. Essential role of DOT1L in maintaining normal adult hematopoiesis. Cell Res. 2011, 21, 1370–1373. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Nguyen, A.T.; Zhang, Y. The diverse functions of Dot1 and H3K79 methylation. Genes Dev. 2011, 25, 1345–1358. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Armstrong, S.A.; Staunton, J.E.; Silverman, L.B.; Pieters, R.; den Boer, M.L.; Minden, M.D.; Sallan, S.E.; Lander, E.S.; Golub, T.R.; Korsmeyer, S.J. MLL translocations specify a distinct gene expression profile that distinguishes a unique leukemia. Nat. Genet. 2002, 30, 41–47. [Google Scholar] [CrossRef] [PubMed]
  150. Broeker, P.L.; Super, H.G.; Thirman, M.J.; Pomykala, H.; Yonebayashi, Y.; Tanabe, S.; Zeleznik-Le, N.; Rowley, J.D. Distribution of 11q23 breakpoints within the MLL breakpoint cluster region in de novo acute leukemia and in treatment-related acute myeloid leukemia: Correlation with scaffold attachment regions and topoisomerase II consensus binding sites. Blood 1996, 87, 1912–1922. [Google Scholar] [CrossRef]
  151. Corral, J.; Lavenir, I.; Impey, H.; Warren, A.J.; Forster, A.; Larson, T.A.; Bell, S.; McKenzie, A.N.; King, G.; Rabbitts, T.H. An Mll-AF9 fusion gene made by homologous recombination causes acute leukemia in chimeric mice: A method to create fusion oncogenes. Cell 1996, 85, 853–861. [Google Scholar] [CrossRef] [Green Version]
  152. Ziemin-van der Poel, S.; McCabe, N.R.; Gill, H.J.; Espinosa, R., 3rd; Patel, Y.; Harden, A.; Rubinelli, P.; Smith, S.D.; LeBeau, M.M.; Rowley, J.D.; et al. Identification of a gene, MLL, that spans the breakpoint in 11q23 translocations associated with human leukemias. Proc. Natl. Acad. Sci. USA 1991, 88, 10735–10739. [Google Scholar] [CrossRef] [Green Version]
  153. Deshpande, A.J.; Deshpande, A.; Sinha, A.U.; Chen, L.; Chang, J.; Cihan, A.; Fazio, M.; Chen, C.W.; Zhu, N.; Koche, R.; et al. AF10 regulates progressive H3K79 methylation and HOX gene expression in diverse AML subtypes. Cancer Cell 2014, 26, 896–908. [Google Scholar] [CrossRef] [Green Version]
  154. Bernt, K.M.; Armstrong, S.A. A role for DOT1L in MLL-rearranged leukemias. Epigenomics 2011, 3, 667–670. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Song, X.; Yang, L.; Wang, M.; Gu, Y.; Ye, B.; Fan, Z.; Xu, R.M.; Yang, N. A higher-order configuration of the heterodimeric DOT1L-AF10 coiled-coil domains potentiates their leukemogenenic activity. Proc. Natl. Acad. Sci. USA 2019, 116, 19917–19923. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Mueller, D.; Bach, C.; Zeisig, D.; Garcia-Cuellar, M.P.; Monroe, S.; Sreekumar, A.; Zhou, R.; Nesvizhskii, A.; Chinnaiyan, A.; Hess, J.L.; et al. A role for the MLL fusion partner ENL in transcriptional elongation and chromatin modification. Blood 2007, 110, 4445–4454. [Google Scholar] [CrossRef] [PubMed]
  157. Kuntimaddi, A.; Achille, N.J.; Thorpe, J.; Lokken, A.A.; Singh, R.; Hemenway, C.S.; Adli, M.; Zeleznik-Le, N.J.; Bushweller, J.H. Degree of recruitment of DOT1L to MLL-AF9 defines level of H3K79 Di- and tri-methylation on target genes and transformation potential. Cell Rep. 2015, 11, 808–820. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Zhang, H.; Zhou, B.; Qin, S.; Xu, J.; Harding, R.; Tempel, W.; Nayak, V.; Li, Y.; Loppnau, P.; Dou, Y.; et al. Structural and functional analysis of the DOT1L-AF10 complex reveals mechanistic insights into MLL-AF10-associated leukemogenesis. Genes Dev. 2018, 32, 341–346. [Google Scholar] [CrossRef] [PubMed]
  159. Sarno, F.; Nebbioso, A.; Altucci, L. DOT1L: A key target in normal chromatin remodelling and in mixed-lineage leukaemia treatment. Epigenetics 2020, 15, 439–453. [Google Scholar] [CrossRef] [Green Version]
  160. Godfrey, L. The Role of DOT1L in MLL-AF4 Leukaemia; University of Oxford: Oxford, UK, 2018. [Google Scholar]
  161. Krivtsov, A.V.; Feng, Z.; Lemieux, M.E.; Faber, J.; Vempati, S.; Sinha, A.U.; Xia, X.; Jesneck, J.; Bracken, A.P.; Silverman, L.B.; et al. H3K79 methylation profiles define murine and human MLL-AF4 leukemias. Cancer Cell 2008, 14, 355–368. [Google Scholar] [CrossRef] [Green Version]
  162. Krivtsov, A.V.; Hoshii, T.; Armstrong, S.A. Mixed-Lineage Leukemia Fusions and Chromatin in Leukemia. Cold Spring Harb. Perspect. Med. 2017, 7. [Google Scholar] [CrossRef]
  163. Bernt, K.M.; Zhu, N.; Sinha, A.U.; Vempati, S.; Faber, J.; Krivtsov, A.V.; Feng, Z.; Punt, N.; Daigle, A.; Bullinger, L.; et al. MLL-rearranged leukemia is dependent on aberrant H3K79 methylation by DOT1L. Cancer Cell 2011, 20, 66–78. [Google Scholar] [CrossRef] [Green Version]
  164. Chen, C.W.; Koche, R.P.; Sinha, A.U.; Deshpande, A.J.; Zhu, N.; Eng, R.; Doench, J.G.; Xu, H.; Chu, S.H.; Qi, J.; et al. DOT1L inhibits SIRT1-mediated epigenetic silencing to maintain leukemic gene expression in MLL-rearranged leukemia. Nat. Med. 2015, 21, 335–343. [Google Scholar] [CrossRef] [Green Version]
  165. Stein, E.M.; Garcia-Manero, G.; Rizzieri, D.A.; Tibes, R.; Berdeja, J.G.; Savona, M.R.; Jongen-Lavrenic, M.; Altman, J.K.; Thomson, B.; Blakemore, S.J.; et al. The DOT1L inhibitor pinometostat reduces H3K79 methylation and has modest clinical activity in adult acute leukemia. Blood 2018, 131, 2661–2669. [Google Scholar] [CrossRef] [PubMed]
  166. Daigle, S.R.; Olhava, E.J.; Therkelsen, C.A.; Basavapathruni, A.; Jin, L.; Boriack-Sjodin, P.A.; Allain, C.J.; Klaus, C.R.; Raimondi, A.; Scott, M.P.; et al. Potent inhibition of DOT1L as treatment of MLL-fusion leukemia. Blood 2013, 122, 1017–1025. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Daigle, S.R.; Olhava, E.J.; Therkelsen, C.A.; Majer, C.R.; Sneeringer, C.J.; Song, J.; Johnston, L.D.; Scott, M.P.; Smith, J.J.; Xiao, Y.; et al. Selective killing of mixed lineage leukemia cells by a potent small-molecule DOT1L inhibitor. Cancer Cell 2011, 20, 53–65. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Kuhn, M.W.; Song, E.; Feng, Z.; Sinha, A.; Chen, C.W.; Deshpande, A.J.; Cusan, M.; Farnoud, N.; Mupo, A.; Grove, C.; et al. Targeting Chromatin Regulators Inhibits Leukemogenic Gene Expression in NPM1 Mutant Leukemia. Cancer Discov. 2016, 6, 1166–1181. [Google Scholar] [CrossRef] [Green Version]
  169. Uckelmann, H.J.; Kim, S.M.; Wong, E.M.; Hatton, C.; Giovinazzo, H.; Gadrey, J.Y.; Krivtsov, A.V.; Rucker, F.G.; Dohner, K.; McGeehan, G.M.; et al. Therapeutic targeting of preleukemia cells in a mouse model of NPM1 mutant acute myeloid leukemia. Science 2020, 367, 586–590. [Google Scholar] [CrossRef]
  170. Rau, R.E.; Rodriguez, B.A.; Luo, M.; Jeong, M.; Rosen, A.; Rogers, J.H.; Campbell, C.T.; Daigle, S.R.; Deng, L.; Song, Y.; et al. DOT1L as a therapeutic target for the treatment of DNMT3A-mutant acute myeloid leukemia. Blood 2016, 128, 971–981. [Google Scholar] [CrossRef]
  171. Sarkaria, S.M.; Christopher, M.J.; Klco, J.M.; Ley, T.J. Primary acute myeloid leukemia cells with IDH1 or IDH2 mutations respond to a DOT1L inhibitor in vitro. Leukemia 2014, 28, 2403–2406. [Google Scholar] [CrossRef] [Green Version]
  172. Perner, F.; Gadrey, J.Y.; Xiong, Y.; Hatton, C.; Eschle, B.K.; Weiss, A.; Stauffer, F.; Gaul, C.; Tiedt, R.; Perry, J.A.; et al. Novel Inhibitors of the Histone-Methyltransferase DOT1L Show Potent Antileukemic Activity in Patient-derived Xenografts. Blood 2020. [Google Scholar] [CrossRef]
  173. Stauffer, F.; Weiss, A.; Scheufler, C.; Mobitz, H.; Ragot, C.; Beyer, K.S.; Calkins, K.; Guthy, D.; Kiffe, M.; Van Eerdenbrugh, B.; et al. New Potent DOT1L Inhibitors for in Vivo Evaluation in Mouse. ACS Med. Chem. Lett. 2019, 10, 1655–1660. [Google Scholar] [CrossRef]
  174. Shi, Y.; Lan, F.; Matson, C.; Mulligan, P.; Whetstine, J.R.; Cole, P.A.; Casero, R.A.; Shi, Y. Histone demethylation mediated by the nuclear amine oxidase homolog LSD1. Cell 2004, 119, 941–953. [Google Scholar] [CrossRef] [Green Version]
  175. Amente, S.; Lania, L.; Majello, B. The histone LSD1 demethylase in stemness and cancer transcription programs. Biochim. Biophys. Acta (BBA) Gene Regul. Mech. 2013, 1829, 981–986. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Whyte, W.A.; Bilodeau, S.; Orlando, D.A.; Hoke, H.A.; Frampton, G.M.; Foster, C.T.; Cowley, S.M.; Young, R.A. Enhancer decommissioning by LSD1 during embryonic stem cell differentiation. Nature 2012, 482, 221–225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Wang, J.; Scully, K.; Zhu, X.; Cai, L.; Zhang, J.; Prefontaine, G.G.; Krones, A.; Ohgi, K.A.; Zhu, P.; Garcia-Bassets, I.; et al. Opposing LSD1 complexes function in developmental gene activation and repression programmes. Nature 2007, 446, 882–887. [Google Scholar] [CrossRef] [PubMed]
  178. Karakaidos, P.; Verigos, J.; Magklara, A. LSD1/KDM1A, a Gate-Keeper of Cancer Stemness and a Promising Therapeutic Target. Cancers 2019, 11, 1821. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Rai, K.; Sarkar, S.; Broadbent, T.J.; Voas, M.; Grossmann, K.F.; Nadauld, L.D.; Dehghanizadeh, S.; Hagos, F.T.; Li, Y.; Toth, R.K.; et al. DNA demethylase activity maintains intestinal cells in an undifferentiated state following loss of APC. Cell 2010, 142, 930–942. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Cusan, M.; Cai, S.F.; Mohammad, H.P.; Krivtsov, A.; Chramiec, A.; Loizou, E.; Witkin, M.D.; Smitheman, K.N.; Tenen, D.G.; Ye, M.; et al. LSD1 inhibition exerts its antileukemic effect by recommissioning PU.1- and C/EBPalpha-dependent enhancers in AML. Blood 2018, 131, 1730–1742. [Google Scholar] [CrossRef] [PubMed]
  181. Barth, J.; Abou-El-Ardat, K.; Dalic, D.; Kurrle, N.; Maier, A.M.; Mohr, S.; Schutte, J.; Vassen, L.; Greve, G.; Schulz-Fincke, J.; et al. LSD1 inhibition by tranylcypromine derivatives interferes with GFI1-mediated repression of PU.1 target genes and induces differentiation in AML. Leukemia 2019, 33, 1411–1426. [Google Scholar] [CrossRef]
  182. Bell, C.C.; Fennell, K.A.; Chan, Y.C.; Rambow, F.; Yeung, M.M.; Vassiliadis, D.; Lara, L.; Yeh, P.; Martelotto, L.G.; Rogiers, A.; et al. Targeting enhancer switching overcomes non-genetic drug resistance in acute myeloid leukaemia. Nat. Commun. 2019, 10, 2723. [Google Scholar] [CrossRef] [Green Version]
  183. Maes, T.; Mascaro, C.; Tirapu, I.; Estiarte, A.; Ciceri, F.; Lunardi, S.; Guibourt, N.; Perdones, A.; Lufino, M.M.P.; Somervaille, T.C.P.; et al. ORY-1001, a Potent and Selective Covalent KDM1A Inhibitor, for the Treatment of Acute Leukemia. Cancer Cell 2018, 33, 495–511.e12. [Google Scholar] [CrossRef] [Green Version]
  184. Bose, P.; Konopleva, M.Y. ORY-1001: Overcoming the Differentiation Block in AML. Cancer Cell 2018, 33, 342–343. [Google Scholar] [CrossRef] [Green Version]
  185. Romussi, A.; Cappa, A.; Vianello, P.; Brambillasca, S.; Cera, M.R.; Dal Zuffo, R.; Faga, G.; Fattori, R.; Moretti, L.; Trifiro, P.; et al. Discovery of Reversible Inhibitors of KDM1A Efficacious in Acute Myeloid Leukemia Models. ACS Med. Chem. Lett. 2020, 11, 754–759. [Google Scholar] [CrossRef] [PubMed]
  186. Fang, Y.; Liao, G.; Yu, B. LSD1/KDM1A inhibitors in clinical trials: Advances and prospects. J. Hematol. Oncol. 2019, 12, 129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Braun, T.P.; Coblentz, C.; Smith, B.M.; Coleman, D.J.; Schonrock, Z.; Carratt, S.A.; Callahan, R.L.; Maniaci, B.; Druker, B.J.; Maxson, J.E. Combined inhibition of JAK/STAT pathway and lysine-specific demethylase 1 as a therapeutic strategy in CSF3R/CEBPA mutant acute myeloid leukemia. Proc. Natl. Acad. Sci. USA 2020, 117, 13670–13679. [Google Scholar] [CrossRef] [PubMed]
  188. Liu, J.; Feng, J.; Li, L.; Lin, L.; Ji, J.; Lin, C.; Liu, L.; Zhang, N.; Duan, D.; Li, Z.; et al. Arginine methylation-dependent LSD1 stability promotes invasion and metastasis of breast cancer. EMBO Rep. 2020, 21, e48597. [Google Scholar] [CrossRef]
  189. Wang, Y.; Zhang, H.; Chen, Y.; Sun, Y.; Yang, F.; Yu, W.; Liang, J.; Sun, L.; Yang, X.; Shi, L.; et al. LSD1 is a subunit of the NuRD complex and targets the metastasis programs in breast cancer. Cell 2009, 138, 660–672. [Google Scholar] [CrossRef] [Green Version]
  190. Metzger, E.; Wissmann, M.; Yin, N.; Muller, J.M.; Schneider, R.; Peters, A.H.; Gunther, T.; Buettner, R.; Schule, R. LSD1 demethylates repressive histone marks to promote androgen-receptor-dependent transcription. Nature 2005, 437, 436–439. [Google Scholar] [CrossRef]
  191. Huang, J.; Sengupta, R.; Espejo, A.B.; Lee, M.G.; Dorsey, J.A.; Richter, M.; Opravil, S.; Shiekhattar, R.; Bedford, M.T.; Jenuwein, T.; et al. p53 is regulated by the lysine demethylase LSD1. Nature 2007, 449, 105–108. [Google Scholar] [CrossRef]
  192. Binda, C.; Valente, S.; Romanenghi, M.; Pilotto, S.; Cirilli, R.; Karytinos, A.; Ciossani, G.; Botrugno, O.A.; Forneris, F.; Tardugno, M.; et al. Biochemical, structural, and biological evaluation of tranylcypromine derivatives as inhibitors of histone demethylases LSD1 and LSD2. J. Am. Chem. Soc. 2010, 132, 6827–6833. [Google Scholar] [CrossRef]
  193. Anonymous. The LSD1 Inhibitor Iadademstat Is Active in Acute Myeloid Leukemia. Cancer Discov. 2020. [Google Scholar] [CrossRef]
  194. Schenk, T.; Chen, W.C.; Gollner, S.; Howell, L.; Jin, L.; Hebestreit, K.; Klein, H.U.; Popescu, A.C.; Burnett, A.; Mills, K.; et al. Inhibition of the LSD1 (KDM1A) demethylase reactivates the all-trans-retinoic acid differentiation pathway in acute myeloid leukemia. Nat. Med. 2012, 18, 605–611. [Google Scholar] [CrossRef]
  195. Wass, M.; Gollner, S.; Besenbeck, B.; Schlenk, R.F.; Mundmann, P.; Gothert, J.R.; Noppeney, R.; Schliemann, C.; Mikesch, J.H.; Lenz, G.; et al. A proof of concept phase I/II pilot trial of LSD1 inhibition by tranylcypromine combined with ATRA in refractory/relapsed AML patients not eligible for intensive therapy. Leukemia 2020. [Google Scholar] [CrossRef] [PubMed]
  196. Srivastava, P.; Tzetzo, S.L.; Gomez, E.C.; Eng, K.H.; Jani Sait, S.N.; Kuechle, J.B.; Singh, P.K.; De Jong, K.; Wiatrowski, K.R.; Peresie, J.; et al. Inhibition of LSD1 in MDS progenitors restores differentiation of CD141(Hi) conventional dendritic cells. Leukemia 2020, 34, 2460–2472. [Google Scholar] [CrossRef] [PubMed]
  197. Jutzi, J.S.; Kleppe, M.; Dias, J.; Staehle, H.F.; Shank, K.; Teruya-Feldstein, J.; Gambheer, S.M.M.; Dierks, C.; Rienhoff, H.Y., Jr.; Levine, R.L.; et al. LSD1 Inhibition Prolongs Survival in Mouse Models of MPN by Selectively Targeting the Disease Clone. Hemasphere 2018, 2, e54. [Google Scholar] [CrossRef] [PubMed]
  198. Kim, J.; Lee, Y.; Lu, X.; Song, B.; Fong, K.W.; Cao, Q.; Licht, J.D.; Zhao, J.C.; Yu, J. Polycomb- and Methylation-Independent Roles of EZH2 as a Transcription Activator. Cell Rep. 2018, 25, 2808–2820.e4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Kim, K.H.; Kim, W.; Howard, T.P.; Vazquez, F.; Tsherniak, A.; Wu, J.N.; Wang, W.; Haswell, J.R.; Walensky, L.D.; Hahn, W.C.; et al. SWI/SNF-mutant cancers depend on catalytic and non-catalytic activity of EZH2. Nat. Med. 2015, 21, 1491–1496. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  200. Ravasio, R.; Ceccacci, E.; Nicosia, L.; Hosseini, A.; Rossi, P.L.; Barozzi, I.; Fornasari, L.; Zuffo, R.D.; Valente, S.; Fioravanti, R.; et al. Targeting the scaffolding role of LSD1 (KDM1A) poises acute myeloid leukemia cells for retinoic acid-induced differentiation. Sci. Adv. 2020, 6, eaax2746. [Google Scholar] [CrossRef] [Green Version]
  201. Chan, A.K.N.; Chen, C.W. Rewiring the Epigenetic Networks in MLL-Rearranged Leukemias: Epigenetic Dysregulation and Pharmacological Interventions. Front. Cell Dev. Biol. 2019, 7, 81. [Google Scholar] [CrossRef]
  202. Okuda, H.; Stanojevic, B.; Kanai, A.; Kawamura, T.; Takahashi, S.; Matsui, H.; Takaori-Kondo, A.; Yokoyama, A. Cooperative gene activation by AF4 and DOT1L drives MLL-rearranged leukemia. J. Clin. Investig. 2017, 127, 1918–1931. [Google Scholar] [CrossRef]
  203. Chen, C.W.; Armstrong, S.A. Targeting DOT1L and HOX gene expression in MLL-rearranged leukemia and beyond. Exp. Hematol. 2015, 43, 673–684. [Google Scholar] [CrossRef] [Green Version]
  204. Deshpande, A.J.; Chen, L.; Fazio, M.; Sinha, A.U.; Bernt, K.M.; Banka, D.; Dias, S.; Chang, J.; Olhava, E.J.; Daigle, S.R.; et al. Leukemic transformation by the MLL-AF6 fusion oncogene requires the H3K79 methyltransferase Dot1l. Blood 2013, 121, 2533–2541. [Google Scholar] [CrossRef] [Green Version]
  205. Linhares, B.M.; Grembecka, J.; Cierpicki, T. Targeting epigenetic protein-protein interactions with small-molecule inhibitors. Future Med. Chem. 2020, 12, 1305–1326. [Google Scholar] [CrossRef] [PubMed]
  206. Krivtsov, A.V.; Evans, K.; Gadrey, J.Y.; Eschle, B.K.; Hatton, C.; Uckelmann, H.J.; Ross, K.N.; Perner, F.; Olsen, S.N.; Pritchard, T.; et al. A Menin-MLL Inhibitor Induces Specific Chromatin Changes and Eradicates Disease in Models of MLL-Rearranged Leukemia. Cancer Cell 2019, 36, 660–673.e11. [Google Scholar] [CrossRef] [PubMed]
  207. Krivtsov, A.V.; Armstrong, S.A. MLL translocations, histone modifications and leukaemia stem-cell development. Nat. Rev. Cancer 2007, 7, 823–833. [Google Scholar] [CrossRef] [PubMed]
  208. Krivtsov, A.V.; Twomey, D.; Feng, Z.; Stubbs, M.C.; Wang, Y.; Faber, J.; Levine, J.E.; Wang, J.; Hahn, W.C.; Gilliland, D.G.; et al. Transformation from committed progenitor to leukaemia stem cell initiated by MLL-AF9. Nature 2006, 442, 818–822. [Google Scholar] [CrossRef] [PubMed]
  209. Kuhn, M.W.; Armstrong, S.A. Designed to kill: Novel menin-MLL inhibitors target MLL-rearranged leukemia. Cancer Cell 2015, 27, 431–433. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Yokoyama, A.; Cleary, M.L. Menin critically links MLL proteins with LEDGF on cancer-associated target genes. Cancer Cell 2008, 14, 36–46. [Google Scholar] [CrossRef] [Green Version]
  211. Yokoyama, A.; Somervaille, T.C.; Smith, K.S.; Rozenblatt-Rosen, O.; Meyerson, M.; Cleary, M.L. The menin tumor suppressor protein is an essential oncogenic cofactor for MLL-associated leukemogenesis. Cell 2005, 123, 207–218. [Google Scholar] [CrossRef]
  212. Gundry, M.C.; Goodell, M.A.; Brunetti, L. It’s All About MEis: Menin-MLL Inhibition Eradicates NPM1-Mutated and MLL-Rearranged Acute Leukemias in Mice. Cancer Cell 2020, 37, 267–269. [Google Scholar] [CrossRef]
  213. Klossowski, S.; Miao, H.; Kempinska, K.; Wu, T.; Purohit, T.; Kim, E.; Linhares, B.M.; Chen, D.; Jih, G.; Perkey, E.; et al. Menin inhibitor MI-3454 induces remission in MLL1-rearranged and NPM1-mutated models of leukemia. J. Clin. Investig. 2020, 130, 981–997. [Google Scholar] [CrossRef] [Green Version]
  214. McGeehan, J. A first-in-class Menin-MLL1 antagonist for the treatment of MLL-r and NPM1 mutant leukemias. In Proceedings of the AACR Virtual Annual Meeting; AACR: Philadelphia, PA, USA, 2020. [Google Scholar]
  215. Uckelmann, H.J.; Kim, S.M.; Antonissen, N.J.C.; Krivtsov, A.V.; Hatton, C.; McGeehan, G.M.; Levine, R.L.; Vassiliou, G.S.; Armstrong, S.A. MLL-Menin Inhibition Reverses Pre-Leukemic Progenitor Self-Renewal Induced By NPM1 Mutations and Prevents AML Development. Blood 2018, 132, 546. [Google Scholar] [CrossRef]
  216. Borkin, D.; He, S.; Miao, H.; Kempinska, K.; Pollock, J.; Chase, J.; Purohit, T.; Malik, B.; Zhao, T.; Wang, J.; et al. Pharmacologic inhibition of the Menin-MLL interaction blocks progression of MLL leukemia in vivo. Cancer Cell 2015, 27, 589–602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Grembecka, J.; He, S.; Shi, A.; Purohit, T.; Muntean, A.G.; Sorenson, R.J.; Showalter, H.D.; Murai, M.J.; Belcher, A.M.; Hartley, T.; et al. Menin-MLL inhibitors reverse oncogenic activity of MLL fusion proteins in leukemia. Nat. Chem. Biol. 2012, 8, 277–284. [Google Scholar] [CrossRef] [PubMed]
  218. Shi, A.; Murai, M.J.; He, S.; Lund, G.; Hartley, T.; Purohit, T.; Reddy, G.; Chruszcz, M.; Grembecka, J.; Cierpicki, T. Structural insights into inhibition of the bivalent menin-MLL interaction by small molecules in leukemia. Blood 2012, 120, 4461–4469. [Google Scholar] [CrossRef]
  219. Wang, E.S.; Altman, J.K.; Pettit, K.; De Botton, S.; Walter, R.P.; Fenaux, P.; Burrows, F.; Tomkinson, B.E.; Martell, B.; Fathi, A.T. Preliminary Data on a Phase 1/2A First in Human Study of the Menin-KMT2A (MLL) Inhibitor KO-539 in Patients with Relapsed or Refractory Acute Myeloid Leukemia. Blood 2020, 136, 7–8. [Google Scholar] [CrossRef]
  220. Filippakopoulos, P.; Knapp, S. Targeting bromodomains: Epigenetic readers of lysine acetylation. Nat. Rev. Drug Discov. 2014, 13, 337–356. [Google Scholar] [CrossRef]
  221. Filippakopoulos, P.; Picaud, S.; Mangos, M.; Keates, T.; Lambert, J.P.; Barsyte-Lovejoy, D.; Felletar, I.; Volkmer, R.; Muller, S.; Pawson, T.; et al. Histone recognition and large-scale structural analysis of the human bromodomain family. Cell 2012, 149, 214–231. [Google Scholar] [CrossRef] [Green Version]
  222. Xu, Y.; Vakoc, C.R. Targeting Cancer Cells with BET Bromodomain Inhibitors. Cold Spring Harb. Perspect. Med. 2017, 7. [Google Scholar] [CrossRef] [Green Version]
  223. Nicodeme, E.; Jeffrey, K.L.; Schaefer, U.; Beinke, S.; Dewell, S.; Chung, C.W.; Chandwani, R.; Marazzi, I.; Wilson, P.; Coste, H.; et al. Suppression of inflammation by a synthetic histone mimic. Nature 2010, 468, 1119–1123. [Google Scholar] [CrossRef]
  224. Dawson, M.A.; Prinjha, R.K.; Dittmann, A.; Giotopoulos, G.; Bantscheff, M.; Chan, W.I.; Robson, S.C.; Chung, C.W.; Hopf, C.; Savitski, M.M.; et al. Inhibition of BET recruitment to chromatin as an effective treatment for MLL-fusion leukaemia. Nature 2011, 478, 529–533. [Google Scholar] [CrossRef] [Green Version]
  225. Alqahtani, A.; Choucair, K.; Ashraf, M.; Hammouda, D.M.; Alloghbi, A.; Khan, T.; Senzer, N.; Nemunaitis, J. Bromodomain and extra-terminal motif inhibitors: A review of preclinical and clinical advances in cancer therapy. Future Sci. OA 2019, 5, FSO372. [Google Scholar] [CrossRef] [Green Version]
  226. Amorim, S.; Stathis, A.; Gleeson, M.; Iyengar, S.; Magarotto, V.; Leleu, X.; Morschhauser, F.; Karlin, L.; Broussais, F.; Rezai, K.; et al. Bromodomain inhibitor OTX015 in patients with lymphoma or multiple myeloma: A dose-escalation, open-label, pharmacokinetic, phase 1 study. Lancet Haematol. 2016, 3, e196–e204. [Google Scholar] [CrossRef]
  227. Shapiro, G.I.; Dowlati, A.; LoRusso, P.M.; Eder, J.P.; Anderson, A.; Do, K.T.; Kagey, M.H.; Sirard, C.; Bradner, J.E.; Landau, S.B. Abstract A49: Clinically efficacy of the BET bromodomain inhibitor TEN-010 in an open-label substudy with patients with documented NUT-midline carcinoma (NMC). Mol. Cancer Ther. 2015, 14, A49. [Google Scholar] [CrossRef]
  228. Stathis, A.; Bertoni, F. BET Proteins as Targets for Anticancer Treatment. Cancer Discov. 2018, 8, 24–36. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  229. Chamberlain, P.P.; Hamann, L.G. Development of targeted protein degradation therapeutics. Nat. Chem. Biol. 2019, 15, 937–944. [Google Scholar] [CrossRef] [PubMed]
  230. Wu, T.; Yoon, H.; Xiong, Y.; Dixon-Clarke, S.E.; Nowak, R.P.; Fischer, E.S. Targeted protein degradation as a powerful research tool in basic biology and drug target discovery. Nat. Struct. Mol. Biol. 2020, 27, 605–614. [Google Scholar] [CrossRef]
  231. Schapira, M.; Calabrese, M.F.; Bullock, A.N.; Crews, C.M. Targeted protein degradation: Expanding the toolbox. Nat. Rev. Drug Discov. 2019, 18, 949–963. [Google Scholar] [CrossRef]
  232. Kronke, J.; Udeshi, N.D.; Narla, A.; Grauman, P.; Hurst, S.N.; McConkey, M.; Svinkina, T.; Heckl, D.; Comer, E.; Li, X.; et al. Lenalidomide causes selective degradation of IKZF1 and IKZF3 in multiple myeloma cells. Science 2014, 343, 301–305. [Google Scholar] [CrossRef] [Green Version]
  233. Lu, G.; Middleton, R.E.; Sun, H.; Naniong, M.; Ott, C.J.; Mitsiades, C.S.; Wong, K.K.; Bradner, J.E.; Kaelin, W.G., Jr. The myeloma drug lenalidomide promotes the cereblon-dependent destruction of Ikaros proteins. Science 2014, 343, 305–309. [Google Scholar] [CrossRef] [Green Version]
  234. Mayor-Ruiz, C.; Bauer, S.; Brand, M.; Kozicka, Z.; Siklos, M.; Imrichova, H.; Kaltheuner, I.H.; Hahn, E.; Seiler, K.; Koren, A.; et al. Rational discovery of molecular glue degraders via scalable chemical profiling. Nat. Chem. Biol. 2020, 16, 1199–1207. [Google Scholar] [CrossRef]
  235. Slabicki, M.; Kozicka, Z.; Petzold, G.; Li, Y.D.; Manojkumar, M.; Bunker, R.D.; Donovan, K.A.; Sievers, Q.L.; Koeppel, J.; Suchyta, D.; et al. The CDK inhibitor CR8 acts as a molecular glue degrader that depletes cyclin K. Nature 2020, 585, 293–297. [Google Scholar] [CrossRef]
  236. Winter, G.E.; Mayer, A.; Buckley, D.L.; Erb, M.A.; Roderick, J.E.; Vittori, S.; Reyes, J.M.; di Iulio, J.; Souza, A.; Ott, C.J.; et al. BET Bromodomain Proteins Function as Master Transcription Elongation Factors Independent of CDK9 Recruitment. Mol. Cell 2017, 67, 5–18.e19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  237. Xu, L.; Chen, Y.; Mayakonda, A.; Koh, L.; Chong, Y.K.; Buckley, D.L.; Sandanaraj, E.; Lim, S.W.; Lin, R.Y.; Ke, X.Y.; et al. Targetable BET proteins- and E2F1-dependent transcriptional program maintains the malignancy of glioblastoma. Proc. Natl. Acad. Sci. USA 2018, 115, E5086–E5095. [Google Scholar] [CrossRef] [Green Version]
  238. Brien, G.L.; Remillard, D.; Shi, J.; Hemming, M.L.; Chabon, J.; Wynne, K.; Dillon, E.T.; Cagney, G.; Van Mierlo, G.; Baltissen, M.P.; et al. Targeted degradation of BRD9 reverses oncogenic gene expression in synovial sarcoma. Elife 2018, 7. [Google Scholar] [CrossRef] [PubMed]
  239. Hohmann, A.F.; Martin, L.J.; Minder, J.L.; Roe, J.S.; Shi, J.; Steurer, S.; Bader, G.; McConnell, D.; Pearson, M.; Gerstberger, T.; et al. Sensitivity and engineered resistance of myeloid leukemia cells to BRD9 inhibition. Nat. Chem. Biol. 2016, 12, 672–679. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  240. Michel, B.C.; D’Avino, A.R.; Cassel, S.H.; Mashtalir, N.; McKenzie, Z.M.; McBride, M.J.; Valencia, A.M.; Zhou, Q.; Bocker, M.; Soares, L.M.M.; et al. A non-canonical SWI/SNF complex is a synthetic lethal target in cancers driven by BAF complex perturbation. Nat. Cell Biol. 2018, 20, 1410–1420. [Google Scholar] [CrossRef] [PubMed]
  241. Mashtalir, N.; Suzuki, H.; Farrell, D.P.; Sankar, A.; Luo, J.; Filipovski, M.; D’Avino, A.R.; St Pierre, R.; Valencia, A.M.; Onikubo, T.; et al. A Structural Model of the Endogenous Human BAF Complex Informs Disease Mechanisms. Cell 2020, 183, 802–817.e24. [Google Scholar] [CrossRef] [PubMed]
  242. Mashtalir, N.; D’Avino, A.R.; Michel, B.C.; Luo, J.; Pan, J.; Otto, J.E.; Zullow, H.J.; McKenzie, Z.M.; Kubiak, R.L.; St Pierre, R.; et al. Modular Organization and Assembly of SWI/SNF Family Chromatin Remodeling Complexes. Cell 2018, 175, 1272–1288.e20. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Schematic illustrating the concept of writers, erasers, and readers on the example of the H3K27ac-mark. Multiprotein-complexes are recruited and/or associate with reader proteins for locus specific chromatin localization. In this example, the superelongation-complex, which is required for RNAPolII activation, associates with BRD4.
Figure 1. Schematic illustrating the concept of writers, erasers, and readers on the example of the H3K27ac-mark. Multiprotein-complexes are recruited and/or associate with reader proteins for locus specific chromatin localization. In this example, the superelongation-complex, which is required for RNAPolII activation, associates with BRD4.
Cells 09 02721 g001
Figure 2. Illustration of the balance between polycomb-mediated repression and trithorax-mediated activation of developmental gene expression programs. Embryonic development and cellular differentiation are regulated by a fine balance between activating and repressive mediators of gene expression and chromatin state (left). In rhabdoid tumors and certain lymphomas, this balance is shifted towards a dominance of the polycomb-repressive complex (right).
Figure 2. Illustration of the balance between polycomb-mediated repression and trithorax-mediated activation of developmental gene expression programs. Embryonic development and cellular differentiation are regulated by a fine balance between activating and repressive mediators of gene expression and chromatin state (left). In rhabdoid tumors and certain lymphomas, this balance is shifted towards a dominance of the polycomb-repressive complex (right).
Cells 09 02721 g002
Figure 3. Enzymatic inhibitors of epigenetic writers and readers in clinical trials. Illustration of the targets of EZH2, DOT1L, and LSD1-inhibition with annotation of compounds that are approved by the FDA or entered clinical trials.
Figure 3. Enzymatic inhibitors of epigenetic writers and readers in clinical trials. Illustration of the targets of EZH2, DOT1L, and LSD1-inhibition with annotation of compounds that are approved by the FDA or entered clinical trials.
Cells 09 02721 g003
Figure 4. Schematic comparison between the consequences of DOT1L- and Menin-inhibition for the integrity of the MLL-AF9 complex on chromatin. The intact MLL-AF9 complex stably binds to its target genes via Menin and recruits the DOT1L-complex to drive H3K79me2 and gene expression of canonical targets (e.g., MEIS1) (left). DOT1L-inhibition removes H3K79me2 by blocking the enzymatic activity of the methyltransferase, while leaving the complex largely intact (middle). Inhibition of the menin-MLL1 interaction disrupts the integrity of the MLL-AF9 complex on chromatin, leading to a more rapid and profound response (right).
Figure 4. Schematic comparison between the consequences of DOT1L- and Menin-inhibition for the integrity of the MLL-AF9 complex on chromatin. The intact MLL-AF9 complex stably binds to its target genes via Menin and recruits the DOT1L-complex to drive H3K79me2 and gene expression of canonical targets (e.g., MEIS1) (left). DOT1L-inhibition removes H3K79me2 by blocking the enzymatic activity of the methyltransferase, while leaving the complex largely intact (middle). Inhibition of the menin-MLL1 interaction disrupts the integrity of the MLL-AF9 complex on chromatin, leading to a more rapid and profound response (right).
Cells 09 02721 g004
Figure 5. Illustration of the disruption of transcriptional elongation by bromodomain-inhibitors of BRD4. BRD4 binds K3K27ac and associates with the superelongation-complex (SEC) to license phosphorylation and activation of RNA-polymerase II to drive productive transcriptional elongation (left). Bromodomain-inhibition of BRD4 prevents binding of BRD4 to H3K27ac preventing association with the SEC and activation of transcription (right).
Figure 5. Illustration of the disruption of transcriptional elongation by bromodomain-inhibitors of BRD4. BRD4 binds K3K27ac and associates with the superelongation-complex (SEC) to license phosphorylation and activation of RNA-polymerase II to drive productive transcriptional elongation (left). Bromodomain-inhibition of BRD4 prevents binding of BRD4 to H3K27ac preventing association with the SEC and activation of transcription (right).
Cells 09 02721 g005
Figure 6. Schematic of the ncBAF-complex and the molecular consequences of BRD9-inhibition and BRD9-degradation. The chromatin remodeling function of the ncBAF complex is essential in maintenance of SS18-fusion driven synovial sarcoma (left). BRD9-bromodomain-inhibition leads to only a partial dissociation of the complex limiting cellular efficacy (middle). Targeted degradation of BRD9, on the other hand, de-stabilizes the complex to a greater extent, enabling increased cellular efficacy (right).
Figure 6. Schematic of the ncBAF-complex and the molecular consequences of BRD9-inhibition and BRD9-degradation. The chromatin remodeling function of the ncBAF complex is essential in maintenance of SS18-fusion driven synovial sarcoma (left). BRD9-bromodomain-inhibition leads to only a partial dissociation of the complex limiting cellular efficacy (middle). Targeted degradation of BRD9, on the other hand, de-stabilizes the complex to a greater extent, enabling increased cellular efficacy (right).
Cells 09 02721 g006
Table 1. Clinical development stage of histone methyltransferase enzymatic inhibitors.
Table 1. Clinical development stage of histone methyltransferase enzymatic inhibitors.
DrugTargetClinical PhaseIndication
Tazemetostat
(EPZ-6438)
EZH2approvedRhabdoid tumors, Follicular lymphoma
Tazemetostat
(EPZ-6438)
EZH2phase 2Diffuse Large B-Cell Lymphoma, Prostate Cancer, Synovial Sarcoma, Epitheloid Sarcoma, Mesothelioma, Squamous-cell Carcinoma, Urothelial Carcinoma, Ovarian and Endometrial Carcinoma, Melanoma,
GSK2816126EZH2phase 1Diffuse Large B Cell Lymphoma, Follicular Lymphoma, Other Non-Hodgkin’s Lymphomas, Solid Tumors, Multiple Myeloma
CPI-1202EZH2phase 2Diffuse Large B-Cell Lymphoma, Prostate Cancer, Advanced Solid Tumors
Pinometostat
(EPZ-5676)
DOT1Lphase 2Acute Myeloid Leukemia, Acute Lymphatic Leukemia
TranylcypromineLSD1phase 2Acute Myeloid Leukemia, Myelodysplastic Syndrome
Iadademstat
(ORY-2001)
LSD1phase 2Alzheimer’s Disease
Bomedemstat
(IMG-7289)
LSD1phase 2Essential Thrombocythemia, Polycythemia vera, Myelofibrosis, Acute Myeloid Leukemia, Myelodysplastic Syndrome
Summary of drugs discussed in this manuscript with their target, clinical stage, and indications.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Perner, F.; Armstrong, S.A. Targeting Chromatin Complexes in Myeloid Malignancies and Beyond: From Basic Mechanisms to Clinical Innovation. Cells 2020, 9, 2721. https://0-doi-org.brum.beds.ac.uk/10.3390/cells9122721

AMA Style

Perner F, Armstrong SA. Targeting Chromatin Complexes in Myeloid Malignancies and Beyond: From Basic Mechanisms to Clinical Innovation. Cells. 2020; 9(12):2721. https://0-doi-org.brum.beds.ac.uk/10.3390/cells9122721

Chicago/Turabian Style

Perner, Florian, and Scott A. Armstrong. 2020. "Targeting Chromatin Complexes in Myeloid Malignancies and Beyond: From Basic Mechanisms to Clinical Innovation" Cells 9, no. 12: 2721. https://0-doi-org.brum.beds.ac.uk/10.3390/cells9122721

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop