Epigenetic deregulation is now a well-recognized although not yet fully understood mechanism that contributes to the development and progression of myeloid malignancies. In the past 15 years, next-generation sequencing studies have revealed patterns of aberrant DNA methylation, altered chromatin states, and mutations in chromatin modifiers across the spectrum of myeloid malignancies. Studies into the mechanisms that drive these diseases through mouse modeling have helped identify new avenues for therapeutic interventions, from initial treatment to resistant or relapsed disease. This is particularly significant when chemotherapy with cytotoxic agents remains the general standard of care. In this review, we will discuss some of the recent findings of epigenetic mechanisms and how these are informing the development of more targeted strategies for therapeutic intervention in myeloid malignancies.

Spatiotemporal regulation of gene expression is controlled by chromatin status and regulatory components of the epigenome, which primarily consists of DNA methylation and histone modifications at the nucleosomal level. The combination of DNA methylation and histone modifications, which are re versible, promotes or inhibits transcriptional activity without changing the underlying DNA sequence, and disruption of these processes can lead to cancer and other disorders (Figure 1). Compared with cancers in other tissue types, myeloid malignancies have a relatively low frequency of somatic mutations. By contrast, epigenetic deregulation is now understood to be an important contributing mechanism to the development of myeloid malignancies. Within this landscape of an overall paucity of mutations, genes that regulate DNA methylation, histone modifications, and chromatin conformation are found to be among the most frequently mutated in hematologic malignancies.1  In normal hematopoiesis, these different layers of epigenetic information have been shown to be important for regulating transcriptional programs that instruct hematopoietic stem cell (HSC) self-renewal and lineage differentiation by regulating hematopoietic transcription factor access to regulatory elements such as enhancers and promoters.2  Any kind of imbalance can lead to abnormal hematopoiesis and the development of hematologic malignancies.

Figure 1.

Schematic of regulatory components of the epigenome. The epigenome primarily consists of DNA methylation and histone modifications at the nucleosomal level. Most cytosine guanine dinucleotides (CpGs) in the mammalian genome are methylated in the form of 5-methylcytosine (5mC), except for those in CpG islands (CGIs), commonly found near promoter regions. The methylation is catalyzed by DNA methyltransferases (DNMTs). CpG shores, located within 2 kb of CGIs, as well as CpG canyons, which are large conserved regions of hypomethylation, can also be differentially methylated. Active removal of 5mC is catalyzed by the ten-eleven translocation (TET) family of dioxygenases through oxidation of 5mC to 5-hydroxymethylcytosine (5hmC). Isocitrate dehydrogenase 1 (IDH1) and IDH2 catalyze the oxidative decarboxylation of isocitrate to α-ketoglutarate (α-KG), a metabolite required for TET catalytic function. Enhancer of zeste homolog 2 (EZH2) is the catalytic subunit of polycomb group (PcG) repressive complex 2 (PRC2), responsible for trimethylation of lysine 27 of histone H3 (H3K27me3), a repressive histone mark that is also regulated by additional sex combs–like 1 (ASXL1). The methylation of H3K4 by mixed-lineage leukemia (MLL), H3K79 by DOT1L, H3R2/R8 by protein arginine methyltransferase 5 (PRMT5), and histone acetylation by histone acetyltransferases (HATs) all promote active transcription. Lysine methylation and acetylation can be removed by demethylases such as lysine-specific demethylase 1 (LSD1) and histone deacetylases (HDACs). DOT1L, disruptor of telomeric silencing 1–like; me1, monomethylation; me2, dimethylation. Professional illustration by Somersault18:24.

Figure 1.

Schematic of regulatory components of the epigenome. The epigenome primarily consists of DNA methylation and histone modifications at the nucleosomal level. Most cytosine guanine dinucleotides (CpGs) in the mammalian genome are methylated in the form of 5-methylcytosine (5mC), except for those in CpG islands (CGIs), commonly found near promoter regions. The methylation is catalyzed by DNA methyltransferases (DNMTs). CpG shores, located within 2 kb of CGIs, as well as CpG canyons, which are large conserved regions of hypomethylation, can also be differentially methylated. Active removal of 5mC is catalyzed by the ten-eleven translocation (TET) family of dioxygenases through oxidation of 5mC to 5-hydroxymethylcytosine (5hmC). Isocitrate dehydrogenase 1 (IDH1) and IDH2 catalyze the oxidative decarboxylation of isocitrate to α-ketoglutarate (α-KG), a metabolite required for TET catalytic function. Enhancer of zeste homolog 2 (EZH2) is the catalytic subunit of polycomb group (PcG) repressive complex 2 (PRC2), responsible for trimethylation of lysine 27 of histone H3 (H3K27me3), a repressive histone mark that is also regulated by additional sex combs–like 1 (ASXL1). The methylation of H3K4 by mixed-lineage leukemia (MLL), H3K79 by DOT1L, H3R2/R8 by protein arginine methyltransferase 5 (PRMT5), and histone acetylation by histone acetyltransferases (HATs) all promote active transcription. Lysine methylation and acetylation can be removed by demethylases such as lysine-specific demethylase 1 (LSD1) and histone deacetylases (HDACs). DOT1L, disruptor of telomeric silencing 1–like; me1, monomethylation; me2, dimethylation. Professional illustration by Somersault18:24.

Close modal

Myeloid malignancies comprise a group of biologically heterogeneous diseases, encompassing acute myeloid leukemia (AML), myelodysplastic syndromes (MDSs), myeloproliferative neoplasms (MPNs), and MDS/MPN overlap syndromes, all of which arise more frequently in the elderly. Despite their heterogeneity in clinical presentation and response to therapy, they have in common the presence of profound epigenetic abnormalities of varying types. This epigenetic deregulation remains a key factor that can be potentially exploited for the development of novel targeted therapies that can prevent relapse and improve long-term survival. In this review, we will discuss the current understanding of how deregulation of the different layers of the epigenome contributes to myeloid malignancies and how this is being exploited for the development of novel therapies.

Aberrant DNA methylation is a hallmark of many cancers, and it has been extensively studied in different types of myeloid malignancies.3  DNA methylation in mammals occurs primarily at CpGs in the form of 5mC and is catalyzed by the family of DNMTs, including de novo methyltransferases DNMT3A and DNMT3B and maintenance methyltransferase DNMT1 (Figure 2). Under normal circumstances, most CpGs in the genome are methylated, except for those in CGIs, which are short clusters of CpG-rich sites ∼1000 bases long commonly found near promoter regions. CpGs located within 2 kb of CGIs, CpG shores, and large conserved regions of hypomethylation known as CpG canyons have also been identified as regions that can be differentially methylated.4,5  Although DNA methylation is most frequently associated with gene silencing, in reality, the precise impact on transcription depends on its genomic context. Promoter CGI methylation is indeed linked to gene silencing; however, 5mC across gene bodies is associated with active transcription.6  Seminal work by Tahiliani et al7  demonstrated for the first time in 2009 that 5mC can be actively removed through oxidation of 5mC to 5hmC, catalyzed by the TET family of dioxygenases. TET enzymes are oxygen-, α-KG–, and vitamin C–dependent dioxygenases that play a role in active DNA demethylation. Demethylation proceeds through the iterative oxidation of 5mC to 5hmC, 5-formylcytosine, and 5-carboxylcytosine, which are finally replaced by unmethylated cytosines through base excision repair (Figure 2).8,9  In addition to active demethylation, passive demethylation results from inadequate methylation maintenance by DNMT1, a situation which can arise through the increased number and rate of cell divisions of cancerous cells, because DNA maintenance is linked to DNA replication.10-12  The implication of this phenomenon in the context of cancer development is that aberrant methylation patterns may thus exist even in the absence of mutations in the DNA methylation machinery.

Figure 2.

DNA methylation pathway. DNA methylation occurs on the C5 position of cytosine to form 5mC and is catalyzed by the family of DNA methyltransferases that includes DNMT3A. Active demethylation is catalyzed by TET enzymes, which are oxygen-, α-KG–, and vitamin C (VitC)–dependent dioxygenases. Demethylation proceeds through iterative oxidation of 5mC to 5hmC, 5-formylcytosine, and 5-carboxylcytosine. Final replacement with unmethylated cytosines is mediated by thymine DNA glycosylase (TDG) and the base excision repair (BER) pathway. Passive DNA demethylation occurs through inadequate methylation maintenance after DNA replication. Professional illustration by Somersault18:24.

Figure 2.

DNA methylation pathway. DNA methylation occurs on the C5 position of cytosine to form 5mC and is catalyzed by the family of DNA methyltransferases that includes DNMT3A. Active demethylation is catalyzed by TET enzymes, which are oxygen-, α-KG–, and vitamin C (VitC)–dependent dioxygenases. Demethylation proceeds through iterative oxidation of 5mC to 5hmC, 5-formylcytosine, and 5-carboxylcytosine. Final replacement with unmethylated cytosines is mediated by thymine DNA glycosylase (TDG) and the base excision repair (BER) pathway. Passive DNA demethylation occurs through inadequate methylation maintenance after DNA replication. Professional illustration by Somersault18:24.

Close modal

In cancer, there is a redistribution of DNA methylation, usually described as aberrant promoter methylation in the context of global hypomethylation. However, specific cancers and cancer subtypes may differ in the pattern of aberrant methylation observed.13  Aberrant CGI methylation has been linked to the silencing of tumor suppressor genes, whereas aberrant loss of methylation has been associated with the overexpression of oncogenes.14,15  However, changes in DNA methylation cannot always be easily linked to changes in gene expression, indicating that its role in transcription regulation is more complex than previously thought. In recent years, studies of different myeloid malignancies using deep sequencing have identified recurrent mutations in genes that regulate DNA methylation: DNMT3A, TET2, and IDH1/2.1,16-18 

Mutations in DNMT3A

DNMT3A is mutated in ∼23% of AMLs, ∼10% of MDSs, ∼5% of MPNs, and ∼7% of chronic myelomonocytic leukemias (CMMLs); it is also the most frequently mutated gene in clonal hematopoiesis of indeterminate potential. The most common mutation in myeloid malignancies occurs at the R882 residue located within the catalytic domain.1,16-26  The heterozygous R882H mutation acts in a dominant-negative manner whereby the mutant DNMT3A prevents the wild-type enzyme from forming functional tetramers, thus reducing the methyltransferase activity by ∼80%. This leads to focal hypomethylation of CpGs at different types of genomic locations (eg, promoter, gene body, intergenic, CGI, or CpG shore or canyon regions).27  However, no clear correlation was found between differentially methylated regions (DMRs) and gene expression or other epigenetic modifications at nearby genes in the context of AML; therefore, how these DMRs contribute to leukemogenesis remains to be understood, although it is possible they may play a role in determining chromatin accessibility and chromosomal conformation.28  Other non-R882 mutations can increase or decrease the enzymatic activity of DNMT3A in vitro, but their impact on the methylome in the cellular context may not always be significant.29  In 1 study, mean methylation levels were not significantly different between non-R882 mutant and wild-type DNMT3A AMLs.27  How the different types of DNMT3A mutations contribute to pathogenesis, whether through catalytic or noncatalytic mechanisms, remains an area of active investigation.

Physiologically, loss of Dnmt3a in mice, where it is highly expressed in long-term HSCs, has been shown to bias HSCs toward self-renewal.30  Even though once again there was no clear correlation between DMRs and gene expression in mutant HSCs, there were multiple genes associated with HSC and leukemia function near hypomethylated DMRs, with some of the most significant regions found at enhancer regions and canyon edges, making them accessible to transcription factors such as FLI1, LMO2, RUNX1, and PU.1 that regulate hematopoietic cell fate.31,32  Mice transplanted with Dnmt3a knockout HSCs ultimately develop a range of hematologic malignancies as expansion of the HSC compartment coupled with aberrant DNA methylation increases the opportunity for secondary hits and subsequent leukemic transformation.33  Knockin of the murine mutant homolog Dnmt3aR878H, alone or in combination with cooperating oncogenes such as FMS-related tyrosine kinase 3–internal tandem duplication (Flt3ITD) and nucleophosmin 1 (Npm1c), also drives clonal expansion and leukemia development.34-36  These studies underscore the critical role that DNMT3A plays in normal hematopoiesis.

Mutations in TET2

TET2 and 5hmC have been found to have important biologic functions in clonal myeloid diseases. In their recent work, Rasmussen et al37  found that Tet2 binding in mouse embryonic stem cells is enriched at regions of open chromatin that overlap with enhancer regions marked by the presence of acetylation of H3K27 and monomethylation of H3K4, whereas it is depleted in promoters and CGIs. Notably, profiling of 5hmC in human AML identified better correlation between 5hmC and gene expression, regardless of genomic location, than 5mC at promoters. Nearly 90% of differential 5hmC regions are outside of CGIs, with approximately half of those located in enhancer regions. Indeed, binding sites for FLI1, RUNX1, and CEBPA/B were enriched in enhancer regions that are enriched for 5hmC.37,38  Tet2 deficiency in murine hematopoietic cells results in global loss of 5hmC, as expected, and changes in chromatin accessibility, leading to changes in transcription factor binding activity.37,39,40  Similar to Dnmt3a mutants, Tet2-deficient murine models exhibited increased self-renewal of hematopoietic stem and progenitor cells (HSPCs) while impairing differentiation, with hypermethylation of DMRs observed at distal enhancers.39-42  This resulted in a CMML-like phenotype found in human patients that includes splenomegaly, neutrophilia, and monocytosis.39,41 

Loss-of-function TET2 mutations occur in ∼12% of AMLs, ∼22% of MDSs, ∼13% of MPNs, and ∼50% of CMMLs.1,16-18,21-23,43-45  Importantly, most TET2 mutations in leukemia patients are monoallelic, and recent work by Cimmino et al46  showed that restoring Tet2 activity in Tet2-deficient cells through the administration of vitamin C, a cofactor required for TET enzymatic activity, inhibits leukemogenesis. Treatment with vitamin C impaired self-renewal, promoted myeloid differentiation, and resulted in the reversal of DNA hypermethylation in murine hematopoietic cells and human AML cell lines.46 

Mutations of IDH1 and IDH2

TET2 deficiency is also found in the context of IDH1 and IDH2 mutations. IDH1 and IDH2 belong to the family of IDH enzymes that catalyze the oxidative decarboxylation of isocitrate to α-KG. Mutations confer on these enzymes neomorphic activity to catalyze the reduction of α-KG into the oncometabolite 2-hydroxyglutarate (2-HG; Figure 3).47  High levels of 2-HG disrupt multiple cellular processes that include inhibition of α-KG–dependent dioxygenases like TET2, resulting in global hypermethylation and a block in differentiation.48,49  Likewise, histone demethylases like jumonji C domain–containing proteins are also affected by 2-HG, leading to an increase in histone methylation.50  Conditional knockin of IDH1R132H into mouse hematopoietic cells resulted in the expansion of HSPCs, with age, and DNA hypermethylation as observed in human AML.51  Mutations in IDH1 and IDH2 are found in ∼17% of AMLs (7% for IDH1 and 10% for IDH2), but they are rare in MDSs, MPNs, and MDS/MPN syndromes.1,16,18,21-23 

Figure 3.

Wild-type and mutant activities of IDH1/2. Mitochondrial metabolite α-KG is produced by reduced NAD phosphate (NADP; NAPDH)-dependent isocitrate dehydrogenases IDH1 and IDH2 in the mitochondrial matrix and in the cytosol by oxidative decarboxylation of isocitrate. Mutations (IDH1/2-mut) confer these enzymes with neomorphic activity to catalyze the reduction of α-KG into oncometabolite 2-HG. High levels of 2-HG disrupt multiple cellular processes that include inhibition of α-KG–dependent dioxygenases like TET2 and histone demethylases. Professional illustration by Somersault18:24.

Figure 3.

Wild-type and mutant activities of IDH1/2. Mitochondrial metabolite α-KG is produced by reduced NAD phosphate (NADP; NAPDH)-dependent isocitrate dehydrogenases IDH1 and IDH2 in the mitochondrial matrix and in the cytosol by oxidative decarboxylation of isocitrate. Mutations (IDH1/2-mut) confer these enzymes with neomorphic activity to catalyze the reduction of α-KG into oncometabolite 2-HG. High levels of 2-HG disrupt multiple cellular processes that include inhibition of α-KG–dependent dioxygenases like TET2 and histone demethylases. Professional illustration by Somersault18:24.

Close modal

Other mutations affecting DNA methylation

Much remains to be understood about how either global or focal DNA methylation or hydroxymethylation changes in hematopoietic cells drive malignant disease. However, it is clear that in addition to mutations of the epigenetic modifiers themselves, mutations in transcription factors may also affect the epigenome through the impaired recruitment of epigenetic modifiers to their normal target sites. Mutations in Wilms tumor 1, a zinc finger transcription factor that interacts with TET2, can impair either its binding to DNA or its recruitment of TET2 to the transcriptional complex. Both types of mutations lead to a phenocopy of TET2 deficiency.38,52 

Another layer of complexity is added when combinations of different mutations are considered. On the basis of genomic classification studies in AML, a driver mutation can be identified in up to 99% of samples, with 86% of samples containing ≥2 driver mutations. Mutations of genes within the same molecular subgroup are usually mutually exclusive, such as transcription factor fusions, cohesin complex mutations, and kinase signaling. For cooccurring mutations, DNA methylation pathway mutations often cooccur with FLT3ITD mutations, which are found in ∼30% of AMLs.1,21  Modeling in mice has shown that Flt3ITD combined with Tet2 loss leads to leukemic transformation, and the number of DMRs is greatly increased in double mutants compared with single mutants, showing significant hypermethylation at promoters and CpG shores.53  Similarly, Dnmt3a deficiency in combination with Flt3ITD also leads to leukemic transformation, with leukemic cells showing significant hypomethylation at promoters and various CpG regions.32,54  The addition of an Npm1 mutation is sufficient to transform Dnmt3aR878H mutant clones into AML.36 Tet2−/−Flt3ITD AML cells were found to be more refractory to chemotherapy and FLT3 inhibition (quizartinib), and patients with the triple combination of DNMT3A, NPM1, and FLT3ITD mutations have a significantly worse prognosis than those with FLT3ITD plus either a DNMT3A or NPM1 mutation.21,53  Consequently, understanding how various combinations of mutations affect disease phenotype is equally important for identifying new strategies to improve therapeutic treatments of myeloid malignancies.

Malignant methylome

Aberrant DNA methylation profiles correlate strongly with molecular and cytogenetic subtypes as well as clinical outcomes in AML, irrespective of whether the leukemia carries mutations in the DNA methylation machinery.55  Each AML subtype is characterized by a distinct epigenetic profile, with different degrees of aberrant hyper- and hypomethylation when compared with normal CD34+ cells, indicating that there is no such thing as a uniquely characteristic malignant methylation profile. MLL-rearranged and DNMT3a mutant AMLs are characterized by predominant hypomethylation, whereas AMLs harboring silenced CEBPA, del(5q)/del(7q), or IDH1/2 mutations have a predominantly hypermethylated profile.13  Furthermore, specific DNA methylation profiles were sufficient to identify the existence of previously unrecognized AML subtypes, with distinct clinical outcomes, which could not be captured by gene expression profiling.55  Two of these 5 novel subtypes were shortly afterward identified as corresponding to IDH1/2 mutant cases, demonstrating that unlike gene expression profiling, DNA methylation could robustly segregate these distinct cases from other AML subtypes.48  Notably, although the cancer field focused on the role of aberrant promoter hypermethylation for many years, next-generation bisulfite sequencing studies of AML in recent years have revealed that enhancer methylation profiles play a much stronger role in determining the epigenetic identity of AMLs than promoter CGI profiles, stressing the important role that these regulatory elements play in malignant epigenomic regulation.13 

Targeting DNA methylation

The first 2 epigenetic drugs approved for clinical treatment of myeloid malignancies were cytosine analogs 5-azacytidine (AZA) and 5-aza-2′-deoxycytidine (DAC), both of which function as DNA hypomethylating agents (HMAs; Figure 4). Both are currently used in the standard care of higher-risk MDS patients and as alternatives for AML patients ineligible for conventional chemotherapy. DAC is incorporated directly into the DNA, whereas AZA is predominantly incorporated into RNA, with a small percentage going into DNA.56  Incorporation of these analogs inhibits DNMT activity, resulting in DNA hypomethylation and upregulation of some silenced genes, although genome-wide studies have failed to report significant transcriptional changes.57,58  Seminal work by Chiappinelli et al59  and Roulois et al60  in 2015 demonstrated that HMAs lead to reexpression of endogenous retroviral elements capable of inducing an immune response through interferon signaling. However, ∼40% to ∼50% of MDS patients do not respond to HMA therapy even after 6 months of treatment, and there are still no clear clinical biomarkers that can predict response.61,62  Evidence for cell-cycle quiescence as a mechanism of resistance in mutant HSPCs to AZA has emerged, consistent with the need for DNA replication to occur for drug incorporation.63,64  Thus far, although some mutational profiling studies in MDS and CMML have shown a correlation between TET2 mutation status and response to HMAs, this did not correlate with overall survival.65-67  In mice, AZA treatment reduced the competitive advantage of Tet2-null hematopoietic cells.65  Given the strong correlation between DNA methylation profiles and clinical subtypes, it is possible that DNA methylation–based biomarkers may offer an alternative for outcome prediction in HMA-treated patients. In CMML, a 21-DMR biomarker predicts response to DAC with 87% accuracy.68 

Figure 4.

Targets of epigenetic therapies. Examples of epigenetic drugs developed to treat aberrant DNA methylation and chromatin in myeloid malignancies. Nucleoside analogs like AZA are incorporated into the DNA and inhibit DNMT activity, resulting in DNA hypomethylation. Mutant IDH1 and IDH2 (IDH1/2-mut) convert α-KG normally produced by wild-type enzymes into oncometabolite 2-HG, and the oncogenic effects caused by 2-HG can be reversed by IDH1 and IDH2 inhibitors ivosidenib and enasidenib, respectively. Chromatin complexes that drive leukemic gene expression or inhibit differentiation programs can also be targeted. Disruption of MLL fusion complexes includes inhibition of DOT1L methyltransferase by pinometostat and menin by menin inhibitors (MIs). Bromodomain and extra terminal (BET) inhibitors (BETi) target the bromodomains of BET proteins like bromodomain-containing protein 4 (BRD4) to prevent their binding to acetylated histones and interaction with transcription factors that promote leukemic gene expression programs. Conversely, repressed differentiation programs can be reversed with the use of LSD1 inhibitors (LSD1i). Disrupting the interaction between LSD1, GFI, and the CoREST repressive complex allows for upregulation of critical myeloid transcription factors and differentiation. me1, monomethylation; me2, dimethylation; NADP, NAD phosphate; NADPH, reduced NAD phosphate. Professional illustration by Somersault18:24.

Figure 4.

Targets of epigenetic therapies. Examples of epigenetic drugs developed to treat aberrant DNA methylation and chromatin in myeloid malignancies. Nucleoside analogs like AZA are incorporated into the DNA and inhibit DNMT activity, resulting in DNA hypomethylation. Mutant IDH1 and IDH2 (IDH1/2-mut) convert α-KG normally produced by wild-type enzymes into oncometabolite 2-HG, and the oncogenic effects caused by 2-HG can be reversed by IDH1 and IDH2 inhibitors ivosidenib and enasidenib, respectively. Chromatin complexes that drive leukemic gene expression or inhibit differentiation programs can also be targeted. Disruption of MLL fusion complexes includes inhibition of DOT1L methyltransferase by pinometostat and menin by menin inhibitors (MIs). Bromodomain and extra terminal (BET) inhibitors (BETi) target the bromodomains of BET proteins like bromodomain-containing protein 4 (BRD4) to prevent their binding to acetylated histones and interaction with transcription factors that promote leukemic gene expression programs. Conversely, repressed differentiation programs can be reversed with the use of LSD1 inhibitors (LSD1i). Disrupting the interaction between LSD1, GFI, and the CoREST repressive complex allows for upregulation of critical myeloid transcription factors and differentiation. me1, monomethylation; me2, dimethylation; NADP, NAD phosphate; NADPH, reduced NAD phosphate. Professional illustration by Somersault18:24.

Close modal

Clinical progress on the use of HMAs will depend on next-generation compounds or synergistic drug combinations. Next-generation HMAs like guadecitabine, which has an increased half-life, showed improved responses with upfront treatment, but whether they are superior remains to be determined.69  Various drug combinations with HMAs are being tested in clinical trials, from HDAC inhibitors to IDH inhibitors.70  One of the new promising combinations is with vitamin C; a recent study found it synergizes with DAC to significantly improve the complete remission (CR) rate and overall survival in AML patients.71 

The oncogenic effects of 2-HG caused by IDH mutations can be reversed with the use of allosteric inhibitors ivosidenib and enasidenib, which target mutant IDH1 and IDH2, respectively (Figure 4). A recent phase 1 trial showed similar results for both inhibitors, achieving an overall response rate of ∼40% in patients with relapsed/refractory IDH mutant AML, reducing plasma levels of 2-HG in nearly all patients, and inducing blast maturation in responders.72,73  They are now being tested in frontline therapies, with initial results of ivosidenib treatment for newly diagnosed AML in patients age >75 years showing a CR rate of 30%.74  Moreover, combining the inhibitors with induction and consolidation therapy produced even more impressive results. Patients with newly diagnosed AML achieved CR in ∼55% of cases for ivosidenib and ∼47% for enasidenib.75  Despite this efficacy, primary and secondary mechanisms of resistance remain a problem. Recent results from studying patients with mutant IDH1 AML treated with ivosidenib showed resistance was associated with mutations in receptor tyrosine kinase pathway genes (NRAS, PTPN11), secondary mutations in IDH1 that reduce drug binding, and/or emergence of IDH2 mutant clones that restore 2-HG production.76  Thus, combinatorial treatment with kinase inhibitors or combined IDH1/2 treatment may improve responses and reduce relapse. The combination of ivosidenib and AZA also produced durable responses, with ∼61% of patients achieving CR, and of the responders, ∼71% showed clearance of mutant IDH1 clones.77 

Another compartment of the epigenome that becomes deregulated in myeloid malignancies involves posttranslational modifications of nucleosomal histone proteins at their N-terminal tails. Histones can be modified by acetylation, methylation, phosphorylation, ubiquitylation, and sumoylation, among others, and each type of modification alone or in combination is associated with particular patterns of gene activity and chromatin conformation, regulating accessibility of other factors to the DNA.78  Each type of modification is catalyzed and recognized by different classes of chromatin factors, which are frequently mutated or rearranged in myeloid malignancies.

MLL rearrangements

Rearrangements of MLL genes result in a class of aggressive leukemias. MLL genes are the mammalian homologs for the trithorax genes that catalyze the methylation of H3K4, a mark of active transcription, and are important for developmental homeotic (Hox) gene expression (Figure 1).79  Rearrangements involving the MLL1 gene on 11q23 create a chimeric gene that encodes the N terminus of MLL1 fused to the C terminus of at least 1 of the 70 known fusion partners.80  Although the methyltransferase domain of MLL is lost, some of the more common MLL fusion partners include members of the super elongation complex and DOT1L complex, both of which promote active transcription.79,81  The MLL fusion complexes induce leukemic transformation through transcriptional deregulation of target genes that activate or maintain a stem cell self-renewal gene expression program in HSPCs.82-84  Overexpression of HOXA cluster genes and MEIS1 is characteristic of MLL fusion leukemias, and recent work has shown that Hoxa9 overexpression actively drives leukemic transformation by increasing chromatin accessibility of enhancers and upregulating gene expression, including Msi1 and Bcl2, which have been previously implicated in myeloid leukemias.85 

Given that MLL fusion complexes drive leukemic expression programs, they can be disrupted using small-molecule inhibitors. DOT1L is the only known H3K79 methyltransferase, and inhibition of its activity has been shown to be effective at targeting MLL-transformed leukemic cells (Figure 4). DOT1L inhibitors were the first histone methyltransferase inhibitors to enter human clinical trials. DOT1L inhibition reduces aberrant H3K79 dimethylation and gene expression driven by MLL fusions, consequently inhibiting proliferation and inducing apoptosis of MLL-transformed leukemic cells.86,87  A recent phase 1 study of the inhibitor pinometostat (EPZ-5676), however, showed moderate efficacy, with only 2 of 51 patients achieving CR.88 

Menin, another cofactor of the MLL complex, is also required for MLL fusion–mediated transformation, and Borkin et al89  were the first to show in 2015 that menin inhibitors can also reverse the leukemic gene signature and block leukemia progression in preclinical models (Figure 4).90  Since then, 2 menin inhibitors have entered phase 1 clinical trials.91,92  Intriguing new findings suggest that the same targeting strategies can be applied to other leukemias that share an MLL dependency. NPM1c AMLs show similar HOXA/MEIS1 overexpression, and the use of menin inhibitors could eradicate preleukemic clones even in these non–MLL-rearranged leukemias.93,94  Other approaches include stabilizing wild-type MLL through IRAK inhibition.95 

Mutations in members of the PcG complexes

The PcG proteins play an important role in gene silencing during embryonic development and tissue differentiation through H3K27me3 (Figure 1). Increased expression of PcG proteins and that of loss-of-function mutations have both been described for myeloid malignancies, indicating that balanced PRC2 activity is important for normal function.96  PcG proteins exert their activity through 2 main complexes: PRC1 and PRC2. Mutations in EZH2, the catalytic subunit of PRC2 responsible for H3K27me3, leads to loss of H3K27me3 and upregulation of HOX oncogenes as well as genes that confer chemoresistance, such as FHL1 and UBE2E1.97,98 ASXL1 is also frequently mutated in myeloid leukemias.1,21-23  Loss of ASXL1 leads to global loss of H3K27me3 and PRC2 activity and the development of an MDS-like phenotype.99-101  Mutations in EZH2 and other PRC2 components are relatively infrequent in myeloid malignancies.1,16-18,21-23  When present, they are loss-of-function mutations, but EZH2 deficiency is also found in SRSF2 splicing mutants.102  Patients with EZH2 or ASXL1 mutations consistently show worse prognosis in various malignancies.

Current approaches for EZH2 mutant malignancies are targeted at immediate downstream dependencies: targeting oncogenes derepressed by the loss of H3K27me3 such as HOX genes, proteasome inhibition (bortezomib) to restore EZH2 levels, inhibition of H3K27me3 demethylases like UTX/KDM6A, and dual inhibition with EZH1, which can compensate partially for EZH2.97,103,104  Recent work has shown that leukemia stem cells overexpress EZH1 and EZH2, and the reduction in H3K27me3 with dual inhibition allows for expression of cell-cycle genes and exit from quiescence.105 

Deregulation of expression of epigenetic modifiers

In addition to mutations, other factors involved in regulating histone modifications have been shown to be abnormally expressed in myeloid malignancies and required for establishing or maintaining the leukemic state. Lysine methylation is targeted for removal by demethylases such as LSD1, which in complex with other proteins can demethylate mono- and dimethylation of H3K4 (Figure 1).106  LSD1 is overexpressed in myeloid malignancies, and its inhibition can induce differentiation of blasts in mouse models of MLL-rearranged leukemia.107,108  LSD1 inhibitors iadademstat and GSK-LSD1, even though they target the catalytic site, did not affect global histone methylation levels but rather disrupted LSD1's interaction with GFI and the CoREST repressive complex, leading to activation of enhancers and upregulation of critical myeloid transcription factors SPI1 and CEBPA to induce differentiation (Figure 4).109,110  Tranylcypromine and GSK2879552 were shown to have synergistic effects with all-trans retinoic acid to inhibit proliferation and induce differentiation and apoptosis in AML cell lines and in patients from a phase 1 study with relapsed/refractory AML and MDS.111,112  Multiple clinical trials studying the use of LSD1 inhibitors as mono- or combination therapy are ongoing. In vitro data in primary AML specimens have shown synergism between enhancer reprogramming through LSD1 inhibition and promoter reprogramming through HMAs, which was most effective in TET2 mutant AML.113 

The family of PRMTs catalyzes histone arginine mono- and dimethylation, among many other cellular substrates, and regulates a wide variety of cellular processes, from gene transcription to RNA splicing and protein translation. Multiple family members have been found to be important for normal and malignant hematopoiesis.114  PRMT5-mediated methylation of histones is generally repressive, but recent work in AML cells has shown that PRMT5-mediated dimethylation of histone H3 on R2 and R8 antagonizes the deposition of H3K27me3 by PRC2 (Figure 1).115  PRMT5 is required for HSC self-renewal and is overexpressed in MPNs and lymphocytic leukemias.116-118  Although the mechanism of drug action remains to be elucidated, PRMT5 inhibitors activate the p53 pathway through alternative splicing of its target genes.118,119  Clinical trials are underway for AML, MDS, and CMML. Unlike PRMT5, coactivator-associated methyltransferase 1 is essential for leukemogenesis but not for normal hematopoiesis, providing an attractive therapeutic target for AML.114 

Acetylation of histones is found in euchromatic regions and is dynamically regulated by the activity of histone lysine acetyltransferases and HDACs (Figure 1). Although they are not known to be mutated in myeloid malignancies, histone lysine acetyltransferases have been identified as leukemic fusion partners with other transcriptional activators such as the monocytic leukemia zinc finger and translation initiation factor MOZ-TIF2 fusion. Not only does expression of the fusion increase self-renewal and drive leukemic transformation of HSPCs in mice, its recruitment of another acetyltransferase, CREB binding protein, is essential for leukemogenesis.120,121 HDAC mutations in myeloid malignancies are likewise rare, but aberrant expression has been observed, as has aberrant recruitment by leukemia fusion proteins.1  The therapeutic effects of HDAC inhibitors, which target the catalytic site, range from restoration of gene expression programs leading to differentiation to induction of DNA damage and apoptosis.122,123  However, HDAC inihbitors (vorinostat, panobinostat) have not been effective clinically either as monotherapy or in combination with HMAs for the treatment of AML or MDS, even with newer agents like pracinostat.124-127  The broad effects of HDAC inhibitors on histone and nonhistone proteins make it challenging to identify signatures or somatic mutations that would confer sensitivity to treatment.

Acetylated lysine residues are recognized by BET domain–containing proteins. These proteins act as transcriptional coactivators and serve as scaffolds for the recruitment of other transcriptional regulators.128  Bromodomain-containing protein 4 (BRD4) colocalizes with hematopoietic transcription factors at gene promoters and enhancers such that inhibition by BET inhibitors, which target the conserved bromodomains, exerts strong antileukemic effects through bromodomain-containing protein 4 loss from enhancers of oncogenic drivers like MYC in AML (Figure 4).129,130  Given their efficacy, multiple BET inhibitors have been developed that are already in preclinical and clinical trials for several types of cancers, including advanced or refractory myeloid malignancies. However, reports from phase 1 trials completed in recent years have shown dose-limiting toxicities at doses insufficient to achieve clinical response.131-134  Furthermore, response to BET inhibitors is not uniform across subtypes. Notably, BET inhibitor–resistant AML cells have compensatory mechanisms to restore MYC expression after treatment through activation of a focal MYC enhancer located within PVT1, a long noncoding RNA.135,136  As a result, combinations of BET inhibitors with other therapies are being tested to overcome resistance and improve clinical response. In addition, although MPNs are mainly driven by mutations that activate tyrosine kinase signaling, recent work has shown that combination with BET inhibitors attenuates disease burden in a manner not seen with the single inhibitors because of the cooperation between BET proteins and NF-κB.137 

Unlike genetic abnormalities, the cancer epigenome is amenable to reprogramming and therefore remains a target of interest for the development of novel therapies. Both somatic mutations and oncogenic fusions can hijack different chromatin regulators to enforce epigenomic reprogramming. This new oncogenic landscape results in the deregulation of critical regulatory elements such as promoters and enhancers and can promote oncogene expression and/or repress tumor suppressor programs. Despite recent advances in our understanding of the mechanisms involved in the development and progression of myeloid malignancies, epigenetic therapies thus far have not been curative in most cases, and resistance to these therapeutic agents involves a variety of mechanisms. The quiescent nature of HSPCs, tumor heterogeneity, and secondary mutations allows for the outgrowth of resistant clones and has proven to be a difficult hurdle to overcome. Although it remains to be seen how effective epigenetic therapies will be as a broad class, rational combinations are expected to provide the greatest benefit given the diverse landscape of myeloid malignancies.

This work was supported by Leukemia and Lymphoma Society Scholar Award 1357-19 (M.E.F.).

Contribution: M.E.F. and H.-T.H. wrote the manuscript.

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

Correspondence: Maria E. Figueroa, 1501 NW 10th Ave, BRB 709A, Miami, FL 33136; e-mail: mefigueroa@miami.edu.

1.
Ley
TJ
,
Miller
C
,
Ding
L
, et al
;
Cancer Genome Atlas Research Network
.
Genomic and epigenomic landscapes of adult de novo acute myeloid leukemia
.
N Engl J Med.
2013
;
368
(
22
):
2059
-
2074
.
2.
Hu
D
,
Shilatifard
A
.
Epigenetics of hematopoiesis and hematological malignancies
.
Genes Dev
.
2016
;
30
(
18
):
2021
2041
.
3.
Baylin
SB
,
Jones
PA.
Epigenetic determinants of cancer
.
Cold Spring Harb Perspect Biol.
2016
;
8
(
9
):
a019505
.
4.
Irizarry
RA
,
Ladd-Acosta
C
,
Wen
B
, et al
.
The human colon cancer methylome shows similar hypo- and hypermethylation at conserved tissue-specific CpG island shores
.
Nat Genet.
2009
;
41
(
2
):
178
-
186
.
5.
Jeong
M
,
Sun
D
,
Luo
M
, et al
.
Large conserved domains of low DNA methylation maintained by Dnmt3a
.
Nat Genet.
2014
;
46
(
1
):
17
-
23
.
6.
Jones
PA.
Functions of DNA methylation: islands, start sites, gene bodies and beyond
.
Nat Rev Genet.
2012
;
13
(
7
):
484
-
492
.
7.
Tahiliani
M
,
Koh
KP
,
Shen
Y
, et al
.
Conversion of 5-methylcytosine to 5-hydroxymethylcytosine in mammalian DNA by MLL partner TET1
.
Science.
2009
;
324
(
5929
):
930
-
935
.
8.
He
YF
,
Li
BZ
,
Li
Z
, et al
.
Tet-mediated formation of 5-carboxylcytosine and its excision by TDG in mammalian DNA
.
Science.
2011
;
333
(
6047
):
1303
-
1307
.
9.
Ito
S
,
Shen
L
,
Dai
Q
, et al
.
Tet proteins can convert 5-methylcytosine to 5-formylcytosine and 5-carboxylcytosine
.
Science.
2011
;
333
(
6047
):
1300
-
1303
.
10.
Wild
L
,
Flanagan
JM.
Genome-wide hypomethylation in cancer may be a passive consequence of transformation
.
Biochim Biophys Acta.
2010
;
1806
(
1
):
50
-
57
.
11.
Zhang
W
,
Klinkebiel
D
,
Barger
CJ
, et al
.
Global DNA hypomethylation in epithelial ovarian cancer: passive demethylation and association with genomic instability
.
Cancers (Basel).
2020
;
12
(
3
):
764
.
12.
Zhou
W
,
Dinh
HQ
,
Ramjan
Z
, et al
.
DNA methylation loss in late-replicating domains is linked to mitotic cell division
.
Nat Genet.
2018
;
50
(
4
):
591
-
602
.
13.
Glass
JL
,
Hassane
D
,
Wouters
BJ
, et al
.
Epigenetic identity in AML depends on disruption of nonpromoter regulatory elements and is affected by antagonistic effects of mutations in epigenetic modifiers
.
Cancer Discov.
2017
;
7
(
8
):
868
-
883
.
14.
Toyota
M
,
Ahuja
N
,
Ohe-Toyota
M
,
Herman
JG
,
Baylin
SB
,
Issa
JP.
CpG island methylator phenotype in colorectal cancer
.
Proc Natl Acad Sci USA.
1999
;
96
(
15
):
8681
-
8686
.
15.
Feinberg
AP
,
Vogelstein
B.
Hypomethylation of ras oncogenes in primary human cancers
.
Biochem Biophys Res Commun.
1983
;
111
(
1
):
47
-
54
.
16.
Papaemmanuil
E
,
Gerstung
M
,
Malcovati
L
, et al
;
Chronic Myeloid Disorders Working Group of the International Cancer Genome Consortium
.
Clinical and biological implications of driver mutations in myelodysplastic syndromes
.
Blood.
2013
;
122
(
22
):
3616
-
3627, quiz 3699
.
17.
Patel
BJ
,
Przychodzen
B
,
Thota
S
, et al
.
Genomic determinants of chronic myelomonocytic leukemia
.
Leukemia.
2017
;
31
(
12
):
2815
-
2823
.
18.
Grinfeld
J
,
Nangalia
J
,
Baxter
EJ
, et al
.
Classification and personalized prognosis in myeloproliferative neoplasms
.
N Engl J Med.
2018
;
379
(
15
):
1416
-
1430
.
19.
Ley
TJ
,
Ding
L
,
Walter
MJ
, et al
.
DNMT3A mutations in acute myeloid leukemia
.
N Engl J Med.
2010
;
363
(
25
):
2424
-
2433
.
20.
Yan
XJ
,
Xu
J
,
Gu
ZH
, et al
.
Exome sequencing identifies somatic mutations of DNA methyltransferase gene DNMT3A in acute monocytic leukemia
.
Nat Genet.
2011
;
43
(
4
):
309
-
315
.
21.
Papaemmanuil
E
,
Gerstung
M
,
Bullinger
L
, et al
.
Genomic classification and prognosis in acute myeloid leukemia
.
N Engl J Med.
2016
;
374
(
23
):
2209
-
2221
.
22.
Metzeler
KH
,
Herold
T
,
Rothenberg-Thurley
M
, et al
;
AMLCG Study Group
.
Spectrum and prognostic relevance of driver gene mutations in acute myeloid leukemia
.
Blood.
2016
;
128
(
5
):
686
-
698
.
23.
Silva
P
,
Neumann
M
,
Schroeder
MP
, et al
.
Acute myeloid leukemia in the elderly is characterized by a distinct genetic and epigenetic landscape
.
Leukemia.
2017
;
31
(
7
):
1640
-
1644
.
24.
Genovese
G
,
Kähler
AK
,
Handsaker
RE
, et al
.
Clonal hematopoiesis and blood-cancer risk inferred from blood DNA sequence
.
N Engl J Med.
2014
;
371
(
26
):
2477
-
2487
.
25.
Jaiswal
S
,
Fontanillas
P
,
Flannick
J
, et al
.
Age-related clonal hematopoiesis associated with adverse outcomes
.
N Engl J Med.
2014
;
371
(
26
):
2488
-
2498
.
26.
Yamashita
Y
,
Yuan
J
,
Suetake
I
, et al
.
Array-based genomic resequencing of human leukemia
.
Oncogene.
2010
;
29
(
25
):
3723
-
3731
.
27.
Russler-Germain
DA
,
Spencer
DH
,
Young
MA
, et al
.
The R882H DNMT3A mutation associated with AML dominantly inhibits wild-type DNMT3A by blocking its ability to form active tetramers
.
Cancer Cell.
2014
;
25
(
4
):
442
-
454
.
28.
Spencer
DH
,
Russler-Germain
DA
,
Ketkar
S
, et al
.
CpG island hypermethylation mediated by DNMT3A is a consequence of AML progression
.
Cell.
2017
;
168
(
5
):
801
-
816.e13
.
29.
Sandoval
JE
,
Huang
YH
,
Muise
A
,
Goodell
MA
,
Reich
NO.
Mutations in the DNMT3A DNA methyltransferase in acute myeloid leukemia patients cause both loss and gain of function and differential regulation by protein partners
.
J Biol Chem.
2019
;
294
(
13
):
4898
-
4910
.
30.
Challen
GA
,
Sun
D
,
Jeong
M
, et al
.
Dnmt3a is essential for hematopoietic stem cell differentiation
.
Nat Genet.
2011
;
44
(
1
):
23
-
31
.
31.
Jeong
M
,
Park
HJ
,
Celik
H
, et al
.
Loss of Dnmt3a immortalizes hematopoietic stem cells in vivo
.
Cell Rep.
2018
;
23
(
1
):
1
-
10
.
32.
Yang
L
,
Rodriguez
B
,
Mayle
A
, et al
.
DNMT3A loss drives enhancer hypomethylation in FLT3-ITD-associated leukemias [published correction appears in Cancer Cell. 2016;30(2):363-365]
.
Cancer Cell.
2016
;
29
(
6
):
922
-
934
.
33.
Mayle
A
,
Yang
L
,
Rodriguez
B
, et al
.
Dnmt3a loss predisposes murine hematopoietic stem cells to malignant transformation
.
Blood.
2015
;
125
(
4
):
629
-
638
.
34.
Guryanova
OA
,
Shank
K
,
Spitzer
B
, et al
.
DNMT3A mutations promote anthracycline resistance in acute myeloid leukemia via impaired nucleosome remodeling
.
Nat Med.
2016
;
22
(
12
):
1488
-
1495
.
35.
Dai
YJ
,
Wang
YY
,
Huang
JY
, et al
.
Conditional knockin of Dnmt3a R878H initiates acute myeloid leukemia with mTOR pathway involvement
.
Proc Natl Acad Sci USA.
2017
;
114
(
20
):
5237
-
5242
.
36.
Loberg
MA
,
Bell
RK
,
Goodwin
LO
, et al
.
Sequentially inducible mouse models reveal that Npm1 mutation causes malignant transformation of Dnmt3a-mutant clonal hematopoiesis
.
Leukemia.
2019
;
33
(
7
):
1635
-
1649
.
37.
Rasmussen
KD
,
Berest
I
,
Keβler
S
, et al
.
TET2 binding to enhancers facilitates transcription factor recruitment in hematopoietic cells
.
Genome Res.
2019
;
29
(
4
):
564
-
575
.
38.
Rampal
R
,
Alkalin
A
,
Madzo
J
, et al
.
DNA hydroxymethylation profiling reveals that WT1 mutations result in loss of TET2 function in acute myeloid leukemia
.
Cell Rep.
2014
;
9
(
5
):
1841
-
1855
.
39.
Li
Z
,
Cai
X
,
Cai
CL
, et al
.
Deletion of Tet2 in mice leads to dysregulated hematopoietic stem cells and subsequent development of myeloid malignancies
.
Blood.
2011
;
118
(
17
):
4509
-
4518
.
40.
Rasmussen
KD
,
Jia
G
,
Johansen
JV
, et al
.
Loss of TET2 in hematopoietic cells leads to DNA hypermethylation of active enhancers and induction of leukemogenesis
.
Genes Dev.
2015
;
29
(
9
):
910
-
922
.
41.
Moran-Crusio
K
,
Reavie
L
,
Shih
A
, et al
.
Tet2 loss leads to increased hematopoietic stem cell self-renewal and myeloid transformation
.
Cancer Cell.
2011
;
20
(
1
):
11
-
24
.
42.
Quivoron
C
,
Couronné
L
,
Della Valle
V
, et al
.
TET2 inactivation results in pleiotropic hematopoietic abnormalities in mouse and is a recurrent event during human lymphomagenesis
.
Cancer Cell.
2011
;
20
(
1
):
25
-
38
.
43.
Delhommeau
F
,
Dupont
S
,
Della Valle
V
, et al
.
Mutation in TET2 in myeloid cancers
.
N Engl J Med.
2009
;
360
(
22
):
2289
-
2301
.
44.
Kosmider
O
,
Gelsi-Boyer
V
,
Ciudad
M
, et al
;
Groupe Francophone des Myélodysplasies
.
TET2 gene mutation is a frequent and adverse event in chronic myelomonocytic leukemia
.
Haematologica.
2009
;
94
(
12
):
1676
-
1681
.
45.
Abdel-Wahab
O
,
Mullally
A
,
Hedvat
C
, et al
.
Genetic characterization of TET1, TET2, and TET3 alterations in myeloid malignancies
.
Blood.
2009
;
114
(
1
):
144
-
147
.
46.
Cimmino
L
,
Dolgalev
I
,
Wang
Y
, et al
.
Restoration of TET2 function blocks aberrant self-renewal and leukemia progression
.
Cell.
2017
;
170
(
6
):
1079
-
1095.e20
.
47.
Gross
S
,
Cairns
RA
,
Minden
MD
, et al
.
Cancer-associated metabolite 2-hydroxyglutarate accumulates in acute myelogenous leukemia with isocitrate dehydrogenase 1 and 2 mutations
.
J Exp Med.
2010
;
207
(
2
):
339
-
344
.
48.
Figueroa
ME
,
Abdel-Wahab
O
,
Lu
C
, et al
.
Leukemic IDH1 and IDH2 mutations result in a hypermethylation phenotype, disrupt TET2 function, and impair hematopoietic differentiation
.
Cancer Cell.
2010
;
18
(
6
):
553
-
567
.
49.
Losman
JA
,
Looper
RE
,
Koivunen
P
, et al
.
(R)-2-hydroxyglutarate is sufficient to promote leukemogenesis and its effects are reversible
.
Science.
2013
;
339
(
6127
):
1621
-
1625
.
50.
Lu
C
,
Ward
PS
,
Kapoor
GS
, et al
.
IDH mutation impairs histone demethylation and results in a block to cell differentiation
.
Nature.
2012
;
483
(
7390
):
474
-
478
.
51.
Sasaki
M
,
Knobbe
CB
,
Munger
JC
, et al
.
IDH1(R132H) mutation increases murine haematopoietic progenitors and alters epigenetics
.
Nature.
2012
;
488
(
7413
):
656
-
659
.
52.
Wang
Y
,
Xiao
M
,
Chen
X
, et al
.
WT1 recruits TET2 to regulate its target gene expression and suppress leukemia cell proliferation
.
Mol Cell.
2015
;
57
(
4
):
662
-
673
.
53.
Shih
AH
,
Jiang
Y
,
Meydan
C
, et al
.
Mutational cooperativity linked to combinatorial epigenetic gain of function in acute myeloid leukemia
.
Cancer Cell.
2015
;
27
(
4
):
502
-
515
.
54.
Meyer
SE
,
Qin
T
,
Muench
DE
, et al
.
DNMT3A haploinsufficiency transforms FLT3ITD myeloproliferative disease into a rapid, spontaneous, and fully penetrant acute myeloid leukemia
.
Cancer Discov.
2016
;
6
(
5
):
501
-
515
.
55.
Figueroa
ME
,
Lugthart
S
,
Li
Y
, et al
.
DNA methylation signatures identify biologically distinct subtypes in acute myeloid leukemia
.
Cancer Cell.
2010
;
17
(
1
):
13
-
27
.
56.
Christman
JK.
5-Azacytidine and 5-aza-2'-deoxycytidine as inhibitors of DNA methylation: mechanistic studies and their implications for cancer therapy
.
Oncogene.
2002
;
21
(
35
):
5483
-
5495
.
57.
Figueroa
ME
,
Skrabanek
L
,
Li
Y
, et al
.
MDS and secondary AML display unique patterns and abundance of aberrant DNA methylation
.
Blood.
2009
;
114
(
16
):
3448
-
3458
.
58.
Flotho
C
,
Claus
R
,
Batz
C
, et al
.
The DNA methyltransferase inhibitors azacitidine, decitabine and zebularine exert differential effects on cancer gene expression in acute myeloid leukemia cells
.
Leukemia.
2009
;
23
(
6
):
1019
-
1028
.
59.
Chiappinelli
KB
,
Strissel
PL
,
Desrichard
A
, et al
.
Inhibiting DNA methylation causes an interferon response in cancer via dsRNA including endogenous retroviruses [published corrections appear in Cell. 2016;164(5):1073 and Cell. 2017;169(2):361]
.
Cell.
2015
;
162
(
5
):
974
-
986
.
60.
Roulois
D
,
Loo Yau
H
,
Singhania
R
, et al
.
DNA-demethylating agents target colorectal cancer cells by inducing viral mimicry by endogenous transcripts
.
Cell.
2015
;
162
(
5
):
961
-
973
.
61.
Fenaux
P
,
Ades
L.
Review of azacitidine trials in intermediate-2-and high-risk myelodysplastic syndromes
.
Leuk Res.
2009
;
33
(
suppl 2
):
S7
-
S11
.
62.
Prébet
T
,
Gore
SD
,
Esterni
B
, et al
.
Outcome of high-risk myelodysplastic syndrome after azacitidine treatment failure
.
J Clin Oncol.
2011
;
29
(
24
):
3322
-
3327
.
63.
Unnikrishnan
A
,
Papaemmanuil
E
,
Beck
D
, et al
.
Integrative genomics identifies the molecular basis of resistance to azacitidine therapy in myelodysplastic syndromes
.
Cell Rep.
2017
;
20
(
3
):
572
-
585
.
64.
Ali
A
,
Penneroux
J
,
Dal Bello
R
Jr
, et al
.
Granulomonocytic progenitors are key target cells of azacytidine in higher risk myelodysplastic syndromes and acute myeloid leukemia
.
Leukemia.
2018
;
32
(
8
):
1856
-
1860
.
65.
Bejar
R
,
Lord
A
,
Stevenson
K
, et al
.
TET2 mutations predict response to hypomethylating agents in myelodysplastic syndrome patients
.
Blood.
2014
;
124
(
17
):
2705
-
2712
.
66.
Coltro
G
,
Mangaonkar
AA
,
Lasho
TL
, et al
.
Clinical, molecular, and prognostic correlates of number, type, and functional localization of TET2 mutations in chronic myelomonocytic leukemia (CMML)-a study of 1084 patients
.
Leukemia.
2020
;
34
(
5
):
1407
-
1421
.
67.
Hunter
AM
,
Komrokji
RS
,
Yun
S
, et al
.
Baseline and serial molecular profiling predicts outcomes with hypomethylating agents in myelodysplastic syndromes
.
Blood Adv.
2021
;
5
(
4
):
1017
-
1028
.
68.
Meldi
K
,
Qin
T
,
Buchi
F
, et al
.
Specific molecular signatures predict decitabine response in chronic myelomonocytic leukemia
.
J Clin Invest.
2015
;
125
(
5
):
1857
-
1872
.
69.
Garcia-Manero
G
,
Roboz
G
,
Walsh
K
, et al
.
Guadecitabine (SGI-110) in patients with intermediate or high-risk myelodysplastic syndromes: phase 2 results from a multicentre, open-label, randomised, phase 1/2 trial
.
Lancet Haematol
.
2019
;
6
(
6
):
e317
-
e327
.
70.
Duchmann
M
,
Itzykson
R.
Clinical update on hypomethylating agents
.
Int J Hematol.
2019
;
110
(
2
):
161
-
169
.
71.
Zhao
H
,
Zhu
H
,
Huang
J
, et al
.
The synergy of vitamin C with decitabine activates TET2 in leukemic cells and significantly improves overall survival in elderly patients with acute myeloid leukemia
.
Leuk Res.
2018
;
66
:
1
-
7
.
72.
Stein
EM
,
DiNardo
CD
,
Pollyea
DA
, et al
.
Enasidenib in mutant IDH2 relapsed or refractory acute myeloid leukemia
.
Blood.
2017
;
130
(
6
):
722
-
731
.
73.
DiNardo
CD
,
Stein
EM
,
de Botton
S
, et al
.
Durable remissions with ivosidenib in IDH1-mutated relapsed or refractory AML
.
N Engl J Med.
2018
;
378
(
25
):
2386
-
2398
.
74.
Roboz
GJ
,
DiNardo
CD
,
Stein
EM
, et al
.
Ivosidenib induces deep durable remissions in patients with newly diagnosed IDH1-mutant acute myeloid leukemia
.
Blood
.
2020
;
135
(
7
):
463
-
471
.
75.
Stein
EM
,
DiNardo
CD
,
Fathi
AT
, et al
.
Ivosidenib or enasidenib combined with intensive chemotherapy in patients with newly diagnosed AML: a phase 1 study
.
Blood.
2021
;
137
(
13
):
1792
-
1803
.
76.
Choe
S
,
Wang
H
,
DiNardo
CD
, et al
.
Molecular mechanisms mediating relapse following ivosidenib monotherapy in IDH1-mutant relapsed or refractory AML
.
Blood Adv.
2020
;
4
(
9
):
1894
-
1905
.
77.
DiNardo
CD
,
Stein
AS
,
Stein
EM
, et al
.
Mutant isocitrate dehydrogenase 1 inhibitor ivosidenib in combination with azacitidine for newly diagnosed acute myeloid leukemia [published correction appears in J Clin Oncol. 2021;39(4):341]
.
J Clin Oncol.
2021
;
39
(
1
):
57
-
65
.
78.
Bannister
AJ
,
Kouzarides
T.
Regulation of chromatin by histone modifications
.
Cell Res.
2011
;
21
(
3
):
381
-
395
.
79.
Smith
E
,
Lin
C
,
Shilatifard
A.
The super elongation complex (SEC) and MLL in development and disease
.
Genes Dev.
2011
;
25
(
7
):
661
-
672
.
80.
Meyer
C
,
Burmeister
T
,
Gröger
D
, et al
.
The MLL recombinome of acute leukemias in 2017
.
Leukemia.
2018
;
32
(
2
):
273
-
284
.
81.
Uckelmann
HJ
,
Armstrong
SA.
Chromatin complexes maintain self-renewal of myeloid progenitors in AML: opportunities for therapeutic intervention
.
Stem Cell Reports.
2020
;
15
(
1
):
6
-
12
.
82.
So
CW
,
Karsunky
H
,
Passegué
E
,
Cozzio
A
,
Weissman
IL
,
Cleary
ML.
MLL-GAS7 transforms multipotent hematopoietic progenitors and induces mixed lineage leukemias in mice
.
Cancer Cell.
2003
;
3
(
2
):
161
-
171
.
83.
Cozzio
A
,
Passegué
E
,
Ayton
PM
,
Karsunky
H
,
Cleary
ML
,
Weissman
IL.
Similar MLL-associated leukemias arising from self-renewing stem cells and short-lived myeloid progenitors
.
Genes Dev.
2003
;
17
(
24
):
3029
-
3035
.
84.
Krivtsov
AV
,
Twomey
D
,
Feng
Z
, et al
.
Transformation from committed progenitor to leukaemia stem cell initiated by MLL-AF9
.
Nature.
2006
;
442
(
7104
):
818
-
822
.
85.
Sun
Y
,
Zhou
B
,
Mao
F
, et al
.
HOXA9 reprograms the enhancer landscape to promote leukemogenesis
.
Cancer Cell.
2018
;
34
(
4
):
643
-
658.e5
.
86.
Bernt
KM
,
Zhu
N
,
Sinha
AU
, et al
.
MLL-rearranged leukemia is dependent on aberrant H3K79 methylation by DOT1L
.
Cancer Cell.
2011
;
20
(
1
):
66
-
78
.
87.
Daigle
SR
,
Olhava
EJ
,
Therkelsen
CA
, et al
.
Potent inhibition of DOT1L as treatment of MLL-fusion leukemia
.
Blood.
2013
;
122
(
6
):
1017
-
1025
.
88.
Stein
EM
,
Garcia-Manero
G
,
Rizzieri
DA
, et al
.
The DOT1L inhibitor pinometostat reduces H3K79 methylation and has modest clinical activity in adult acute leukemia
.
Blood.
2018
;
131
(
24
):
2661
-
2669
.
89.
Borkin
D
,
He
S
,
Miao
H
, et al
.
Pharmacologic inhibition of the menin-MLL interaction blocks progression of MLL leukemia in vivo
.
Cancer Cell.
2015
;
27
(
4
):
589
-
602
.
90.
Yokoyama
A
,
Somervaille
TC
,
Smith
KS
,
Rozenblatt-Rosen
O
,
Meyerson
M
,
Cleary
ML.
The menin tumor suppressor protein is an essential oncogenic cofactor for MLL-associated leukemogenesis
.
Cell.
2005
;
123
(
2
):
207
-
218
.
91.
Klossowski
S
,
Miao
H
,
Kempinska
K
, et al
.
Menin inhibitor MI-3454 induces remission in MLL1-rearranged and NPM1-mutated models of leukemia
.
J Clin Invest.
2020
;
130
(
2
):
981
-
997
.
92.
Krivtsov
AV
,
Evans
K
,
Gadrey
JY
, et al
.
A menin-MLL inhibitor induces specific chromatin changes and eradicates disease in models of MLL-rearranged leukemia
.
Cancer Cell.
2019
;
36
(
6
):
660
-
673.e11
.
93.
Uckelmann
HJ
,
Kim
SM
,
Wong
EM
, et al
.
Therapeutic targeting of preleukemia cells in a mouse model of NPM1 mutant acute myeloid leukemia
.
Science.
2020
;
367
(
6477
):
586
-
590
.
94.
Kühn
MW
,
Song
E
,
Feng
Z
, et al
.
Targeting chromatin regulators inhibits leukemogenic gene expression in NPM1 mutant leukemia
.
Cancer Discov.
2016
;
6
(
10
):
1166
-
1181
.
95.
Liang
K
,
Volk
AG
,
Haug
JS
, et al
.
Therapeutic targeting of MLL degradation pathways in MLL-rearranged leukemia
.
Cell.
2017
;
168
(
1-2
):
59
-
72.e13
.
96.
Di Carlo
V
,
Mocavini
I
,
Di Croce
L.
Polycomb complexes in normal and malignant hematopoiesis
.
J Cell Biol.
2019
;
218
(
1
):
55
-
69
.
97.
Göllner
S
,
Oellerich
T
,
Agrawal-Singh
S
, et al
.
Loss of the histone methyltransferase EZH2 induces resistance to multiple drugs in acute myeloid leukemia
.
Nat Med.
2017
;
23
(
1
):
69
-
78
.
98.
Kempf
JM
,
Weser
S
,
Bartoschek
MD
, et al
.
Loss-of-function mutations in the histone methyltransferase EZH2 promote chemotherapy resistance in AML
.
Sci Rep.
2021
;
11
(
1
):
5838
.
99.
Abdel-Wahab
O
,
Adli
M
,
LaFave
LM
, et al
.
ASXL1 mutations promote myeloid transformation through loss of PRC2-mediated gene repression
.
Cancer Cell.
2012
;
22
(
2
):
180
-
193
.
100.
Inoue
D
,
Kitaura
J
,
Togami
K
, et al
.
Myelodysplastic syndromes are induced by histone methylation–altering ASXL1 mutations
.
J Clin Invest.
2013
;
123
(
11
):
4627
-
4640
.
101.
Wang
J
,
Li
Z
,
He
Y
, et al
.
Loss of Asxl1 leads to myelodysplastic syndrome-like disease in mice
.
Blood.
2014
;
123
(
4
):
541
-
553
.
102.
Kim
E
,
Ilagan
JO
,
Liang
Y
, et al
.
SRSF2 mutations contribute to myelodysplasia by mutant-specific effects on exon recognition
.
Cancer Cell.
2015
;
27
(
5
):
617
-
630
.
103.
Ezponda
T
,
Dupéré-Richer
D
,
Will
CM
, et al
.
UTX/KDM6A loss enhances the malignant phenotype of multiple myeloma and sensitizes cells to EZH2 inhibition
.
Cell Rep.
2017
;
21
(
3
):
628
-
640
.
104.
Honma
D
,
Kanno
O
,
Watanabe
J
, et al
.
Novel orally bioavailable EZH1/2 dual inhibitors with greater antitumor efficacy than an EZH2 selective inhibitor
.
Cancer Sci.
2017
;
108
(
10
):
2069
-
2078
.
105.
Fujita
S
,
Honma
D
,
Adachi
N
, et al
.
Dual inhibition of EZH1/2 breaks the quiescence of leukemia stem cells in acute myeloid leukemia
.
Leukemia.
2018
;
32
(
4
):
855
-
864
.
106.
Shi
Y
,
Lan
F
,
Matson
C
, et al
.
Histone demethylation mediated by the nuclear amine oxidase homolog LSD1
.
Cell.
2004
;
119
(
7
):
941
-
953
.
107.
Niebel
D
,
Kirfel
J
,
Janzen
V
,
Höller
T
,
Majores
M
,
Gütgemann
I.
Lysine-specific demethylase 1 (LSD1) in hematopoietic and lymphoid neoplasms
.
Blood.
2014
;
124
(
1
):
151
-
152
.
108.
Harris
WJ
,
Huang
X
,
Lynch
JT
, et al
.
The histone demethylase KDM1A sustains the oncogenic potential of MLL-AF9 leukemia stem cells [published correction appears in Cancer Cell. 2012;21(6):856]
.
Cancer Cell.
2012
;
21
(
4
):
473
-
487
.
109.
Maiques-Diaz
A
,
Spencer
GJ
,
Lynch
JT
, et al
.
Enhancer activation by pharmacologic displacement of LSD1 from GFI1 induces differentiation in acute myeloid leukemia
.
Cell Rep.
2018
;
22
(
13
):
3641
-
3659
.
110.
Cusan
M
,
Cai
SF
,
Mohammad
HP
, et al
.
LSD1 inhibition exerts its antileukemic effect by recommissioning PU.1- and C/EBPα-dependent enhancers in AML
.
Blood.
2018
;
131
(
15
):
1730
-
1742
.
111.
Smitheman
KN
,
Severson
TM
,
Rajapurkar
SR
, et al
.
Lysine specific demethylase 1 inactivation enhances differentiation and promotes cytotoxic response when combined with all-trans retinoic acid in acute myeloid leukemia across subtypes
.
Haematologica.
2019
;
104
(
6
):
1156
-
1167
.
112.
Tayari
MM
,
Santos
HGD
,
Kwon
D
, et al
.
Clinical responsiveness to all-trans retinoic acid is potentiated by LSD1 inhibition and associated with a quiescent transcriptome in myeloid malignancies
.
Clin Cancer Res.
2021
;
27
(
7
):
1893
-
1903
.
113.
Duy
C
,
Teater
M
,
Garrett-Bakelman
FE
, et al
.
Rational targeting of cooperating layers of the epigenome yields enhanced therapeutic efficacy against AML
.
Cancer Discov.
2019
;
9
(
7
):
872
-
889
.
114.
Greenblatt
SM
,
Liu
F
,
Nimer
SD.
Arginine methyltransferases in normal and malignant hematopoiesis
.
Exp Hematol.
2016
;
44
(
6
):
435
-
441
.
115.
Liu
F
,
Xu
Y
,
Lu
X
, et al
.
PRMT5-mediated histone arginine methylation antagonizes transcriptional repression by polycomb complex PRC2
.
Nucleic Acids Res.
2020
;
48
(
6
):
2956
-
2968
.
116.
Liu
F
,
Cheng
G
,
Hamard
PJ
, et al
.
Arginine methyltransferase PRMT5 is essential for sustaining normal adult hematopoiesis
.
J Clin Invest.
2015
;
125
(
9
):
3532
-
3544
.
117.
Wang
L
,
Pal
S
,
Sif
S.
Protein arginine methyltransferase 5 suppresses the transcription of the RB family of tumor suppressors in leukemia and lymphoma cells
.
Mol Cell Biol.
2008
;
28
(
20
):
6262
-
6277
.
118.
Pastore
F
,
Bhagwat
N
,
Pastore
A
, et al
.
PRMT5 inhibition modulates E2F1 methylation and gene-regulatory networks leading to therapeutic efficacy in JAK2V617F-mutant MPN
.
Cancer Discov.
2020
;
10
(
11
):
1742
-
1757
.
119.
Hamard
PJ
,
Santiago
GE
,
Liu
F
, et al
.
PRMT5 regulates DNA repair by controlling the alternative splicing of histone-modifying enzymes
.
Cell Rep.
2018
;
24
(
10
):
2643
-
2657
.
120.
Deguchi
K
,
Ayton
PM
,
Carapeti
M
, et al
.
MOZ-TIF2-induced acute myeloid leukemia requires the MOZ nucleosome binding motif and TIF2-mediated recruitment of CBP
.
Cancer Cell.
2003
;
3
(
3
):
259
-
271
.
121.
Huntly
BJ
,
Shigematsu
H
,
Deguchi
K
, et al
.
MOZ-TIF2, but not BCR-ABL, confers properties of leukemic stem cells to committed murine hematopoietic progenitors
.
Cancer Cell.
2004
;
6
(
6
):
587
-
596
.
122.
Côté
S
,
Rosenauer
A
,
Bianchini
A
, et al
.
Response to histone deacetylase inhibition of novel PML/RARalpha mutants detected in retinoic acid-resistant APL cells
.
Blood.
2002
;
100
(
7
):
2586
-
2596
.
123.
Petruccelli
LA
,
Dupéré-Richer
D
,
Pettersson
F
,
Retrouvey
H
,
Skoulikas
S
,
Miller
WH
Jr
.
Vorinostat induces reactive oxygen species and DNA damage in acute myeloid leukemia cells
.
PLoS One.
2011
;
6
(
6
):
e20987
.
124.
Schaefer
EW
,
Loaiza-Bonilla
A
,
Juckett
M
, et al
;
Mayo P2C Phase II Consortium
.
A phase 2 study of vorinostat in acute myeloid leukemia
.
Haematologica.
2009
;
94
(
10
):
1375
-
1382
.
125.
Garcia-Manero
G
,
Sekeres
MA
,
Egyed
M
, et al
.
A phase 1b/2b multicenter study of oral panobinostat plus azacitidine in adults with MDS, CMML or AML with ≤30% blasts
.
Leukemia.
2017
;
31
(
12
):
2799
-
2806
.
126.
Abaza
YM
,
Kadia
TM
,
Jabbour
EJ
, et al
.
Phase 1 dose escalation multicenter trial of pracinostat alone and in combination with azacitidine in patients with advanced hematologic malignancies
.
Cancer.
2017
;
123
(
24
):
4851
-
4859
.
127.
Garcia-Manero
G
,
Montalban-Bravo
G
,
Berdeja
JG
, et al
.
Phase 2, randomized, double-blind study of pracinostat in combination with azacitidine in patients with untreated, higher-risk myelodysplastic syndromes
.
Cancer.
2017
;
123
(
6
):
994
-
1002
.
128.
Belkina
AC
,
Denis
GV.
BET domain co-regulators in obesity, inflammation and cancer
.
Nat Rev Cancer.
2012
;
12
(
7
):
465
-
477
.
129.
Delmore
JE
,
Issa
GC
,
Lemieux
ME
, et al
.
BET bromodomain inhibition as a therapeutic strategy to target c-Myc
.
Cell.
2011
;
146
(
6
):
904
-
917
.
130.
Roe
JS
,
Mercan
F
,
Rivera
K
,
Pappin
DJ
,
Vakoc
CR.
BET bromodomain inhibition suppresses the function of hematopoietic transcription factors in acute myeloid leukemia
.
Mol Cell.
2015
;
58
(
6
):
1028
-
1039
.
131.
Amorim
S
,
Stathis
A
,
Gleeson
M
, 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
(
4
):
e196
-
e204
.
132.
Ameratunga
M
,
Braña
I
,
Bono
P
, et al
.
First-in-human Phase 1 open label study of the BET inhibitor ODM-207 in patients with selected solid tumours
.
Br J Cancer.
2020
;
123
(
12
):
1730
-
1736
.
133.
Falchook
G
,
Rosen
S
,
LoRusso
P
, et al
.
Development of 2 bromodomain and extraterminal inhibitors with distinct pharmacokinetic and pharmacodynamic profiles for the treatment of advanced malignancies
.
Clin Cancer Res.
2020
;
26
(
6
):
1247
-
1257
.
134.
Shapiro
GI
,
LoRusso
P
,
Dowlati
A
, et al
.
A phase 1 study of RO6870810, a novel bromodomain and extra-terminal protein inhibitor, in patients with NUT carcinoma, other solid tumours, or diffuse large B-cell lymphoma
.
Br J Cancer.
2021
;
124
(
4
):
744
-
753
.
135.
Rathert
P
,
Roth
M
,
Neumann
T
, et al
.
Transcriptional plasticity promotes primary and acquired resistance to BET inhibition
.
Nature.
2015
;
525
(
7570
):
543
-
547
.
136.
Guo
L
,
Li
J
,
Zeng
H
, et al
.
A combination strategy targeting enhancer plasticity exerts synergistic lethality against BETi-resistant leukemia cells
.
Nat Commun.
2020
;
11
(
1
):
740
.
137.
Kleppe
M
,
Koche
R
,
Zou
L
, et al
.
Dual targeting of oncogenic activation and inflammatory signaling increases therapeutic efficacy in myeloproliferative neoplasms [published correction appears in Cancer Cell. 2018;33(4):785–787]
.
Cancer Cell.
2018
;
33
(
1
):
29
-
43.e7
.
Sign in via your Institution