Skip to main content

REVIEW article

Front. Genet., 01 March 2019
Sec. Epigenomics and Epigenetics
This article is part of the Research Topic Epigenetic Modulating Agents: Complementary and Alternative Treatment Approaches View all 11 articles

Epigenetic Modifications in Acute Myeloid Leukemia: Prognosis, Treatment, and Heterogeneity

  • 1Department of Physiology and Biophysics, Weill Cornell Medicine, New York, NY, United States
  • 2The HRH Prince Alwaleed Bin Talal Bin Abdulaziz Alsaud Institute for Computational Biomedicine, Weill Cornell Medicine, New York, NY, United States
  • 3University of Maryland, College Park, MD, United States
  • 4Yale College, New Haven, CT, United States
  • 5The WorldQuant Initiative for Quantitative Prediction, Weill Cornell Medicine, New York, NY, United States
  • 6The Feil Family Brain and Mind Research Institute, Weill Cornell Medicine, New York, NY, United States

Leukemia, specifically acute myeloid leukemia (AML), is a common malignancy that can be differentiated into multiple subtypes based on leukemogenic history and etiology. Although genetic aberrations, particularly cytogenetic abnormalities and mutations in known oncogenes, play an integral role in AML development, epigenetic processes have been shown as a significant and sometimes independent dynamic in AML pathophysiology. Here, we summarize how tumors evolve and describe AML through an epigenetic lens, including discussions on recent discoveries that include prognostics from epialleles, changes in RNA function for hematopoietic stem cells and the epitranscriptome, and novel epigenetic treatment options. We further describe the limitations of treatment in the context of the high degree of heterogeneity that characterizes acute myeloid leukemia.

Introduction

Acute myeloid leukemia is a common, heterogeneous, and aggressive hematopoietic malignancy (Hasserjian, 2013; Saultz and Garzon, 2016; De Kouchkovsky and Abdul-Hay, 2016), which can be differentiated into subtypes based on the nature of the (often cytogenetic) initiating leukemogenic event (Lazarus and Litzow, 2012; Hou et al., 2013; Hasserjian, 2013; Saultz and Garzon, 2016). There are two common classification systems for the characterization of acute myeloid leukemia subtypes: the French-American-British (FAB) system, which morphologically distinguishes leukemias by progenitor cell type and cell maturation; and the World Health Organization (WHO) system, which distinguishes leukemias by their characteristic cytogenetic or genetic abnormalities (Saultz and Garzon, 2016). Though there are many acute myeloid leukemia subtypes driven by epigenetic dysregulation, a brief survey of some common AML subtypes and relevant epigenetic context is given below.

t(15;17)

The t(15;17) chromosomal abnormality is a translocation that fuses the promyelocytic locus (PML) gene, which encodes an eponymous tumor-suppressor protein (Salomoni and Pandolfi, 2002) on chromosome 15, with the retinoic acid receptor-α (RARA) gene on chromosome 17, resulting in the PML-RARA fusion gene (Fu et al., 1995). Both PML and RARA genes are known to be involved in regular cellular functions: PML is involved in the regulation of cellular proliferation (Salomoni and Pandolfi, 2002), and RARa is integral in myeloid differentiation and regular hematopoietic development (Melnick and Licht, 1999). Thus, the disruption of both can contribute to a hematopoietic malignancy characterized by enhanced cellular proliferation, such as what is seen in acute myeloid leukemia. Indeed, it has been shown that the introduction of PML-RARa into myeloid cells (a TF-1 leukemia cell line) inhibits regularly induced cellular apoptosis (Fu et al., 1995) and blocks myeloid differentiation (Grignani et al., 1993).

Though the t(15;17) translocation is reciprocal – it generates both PML-RARa and RARa-PML gene fusions (Gallagher et al., 1995), with both being implicated in leukemogenesis (Lafage-Pochitaloff et al., 1995; Richter et al., 2016) – the PML-RARa product seems to be of primary importance (Gallagher et al., 1995). The PML-RARa fusion is pathognomonic of acute promyelocytic leukemia (Fu et al., 1995), a subtype of acute myeloid leukemia characterized by a high response rate to all-trans retinoic acid (ATRA)/arsenic trioxide (ATO) therapy and a comparatively favorable prognosis (de Thé and Chen, 2010; Coombs et al., 2015).

The PML-RARa product works in tandem with DNA methyltransferases (DNMTs) to induce hypermethylation in PML-RARa targets, particularly RARβ (Di Croce, 2002). Indeed, the PML-RARa product seems to be a necessary component for the development of a hypermethylated phenotype. For instance, mice possessing DNMT3a1 but lacking the PML-RARa product did not display a hypermethylated phenotype, while leukemic mice possessing both the PML-RARa product and DNMT3a1 presented with the hypermethylated phenotype (Subramanyam et al., 2010). Notably, retinoic acid reverses this hypermethylated phenotype (Di Croce, 2002), indicating that utility of ATRA treatment for acute promyelocytic leukemia functions in part due to the epigenetic nature of its mechanism of action.

t(8;21)

The t(8;21) chromosomal translocation fuses the RUNX1 and RUNX1T1 genes (also known as AML1-ETO) (Nishii et al., 2003; Krauth et al., 2014). The RUNX1 gene encodes runt-related transcription factor 1 (or acute myeloid leukemia factor 1/AML1) and plays a regulatory role in hematopoietic development (Okuda et al., 2001; Wichmann et al., 2015), including in the generation and differentiation of hematopoietic stem cells (Asou, 2003). Mutational aberrations in the RUNX1 gene have been shown to reduce the expression of CCAAT enhancer-binding protein alpha (CEBPA), a transcription factor involved in cell cycle regulation and myeloid differentiation (Grossmann et al., 2012). The AML1-ETO protein generated by the t(8;21) fusion gene downregulates CEBPA expression (Pabst, 2001), playing a role in the development and pathology of acute myeloid leukemia (Yan et al., 2004; Pabst and Mueller, 2009).

In the AML1-ETO fusion protein, the ETO domain aids in the recruitment of histone deacetylases (Yan et al., 2004; Liu et al., 2007), epigenetically driving the arrest of myeloid differentiation in t(8;21)-positive leukemia (Liu et al., 2007; Wichmann et al., 2015; Loke et al., 2017) and contributing to leukemogenesis (Liu et al., 2007; Loke et al., 2017). Indeed, upon the selective removal of AML1-ETO, previously blocked myeloid differentiation is induced and leukemic proliferation halts (Loke et al., 2017).

The leukemogenic ability of the AML1-ETO product may be partly dependent on post-translational lysine acetylation of the fusion protein. Wang et al. found median survival in leukemic mouse models was increased via inhibition of the lysine acetyltransferase p300 (Wang et al., 2011), which reduced Lys43 acetylation levels in AML1-ETO9a, a splice isoform of AML1-ETO (Zhang et al., 2007; Link et al., 2016). Though p300 knockdown leads to decreased acetylation, the therapeutic effects of p300 knockdown could be due to effects largely unrelated to AML1-ETO9a acetylation, indicating p300 may be a broader therapeutic target (Wang et al., 2011).

Similarly, post-translational arginine methylation of the AML1-ETO9a protein via protein arginine methyltransferase 1 (PRMT1) may affect leukemic potential (Shia et al., 2012). Though PRMT1 methylates AML1-ETO9a and PRMT1 knockdown reduces leukemic proliferation, it is important to note that PRMT1 weakly methylates the AML1-ETO9a arginine (Shia et al., 2012). Thus, similar to the case of p300 knockdown, it is unclear whether leukemic proliferation is reduced by virtue of lesser arginine methylation or by virtue of inhibiting additional PRMT1-mediated interactions – for example, the recruitment of PRMT1 by AML1-ETO9a to methylate histone H4 to upregulate transcription (Shia et al., 2012).

The bulk of the therapeutic potential of PRMT1 inhibition, then, may be derived less from the consequent reduction of arginine methylation and more from the inhibition of interaction of PRMT1 with additional substrates, indicating the role of PRMT1 as a broader therapeutic target (Shia et al., 2012). Indeed, many PRMTs have been indicated as potential therapeutic targets for AML. For instance, PRMT4 inhibition enhances myeloid differentiation and inhibits myeloid proliferation (Vu et al., 2013), while PRMT5 inhibition promotes myeloid differentiation (Kaushik et al., 2017).

More recently, it has been shown that the AML1-ETO induced silencing of brain acid-soluble protein 1 (BASP1) contributes to the leukemic phenotype: AML1-ETO product-mediated recruitment of a DNA methyltransferase to the BASP1-promoter region leads to promoter methylation and ultimate silencing of BASP1, the expression of which regulates the cell cycle and arrests the cellular proliferation (Zhou et al., 2018).

Though acute myeloid leukemias characterized by the t(8;21) cytogenetic abnormality show generally favorable response to conventional chemotherapy and have a relatively favorable prognosis (Nishii et al., 2003; Krauth et al., 2014), accompanying aberrations in other respects can impact prognosis. For instance, mutations related to tyrosine-protein kinase kit, encoded by the KIT gene, can confer chemoresistance (Wichmann et al., 2015) and negatively impact the overall survival prognosis in t(8;21)-positive acute myeloid leukemias (Krauth et al., 2014).

inv(16)/t(16;16)

The inv(16) chromosomal inversion is an inversion that fuses the core-binding factor subunit beta (CBFB) gene to the myosin heavy chain 11 (MYH11) gene, producing a CBFB-MYH11 fusion gene and resulting in the expression of a CBFB-SMMHC fusion protein (Eghtedar et al., 2012; Richter et al., 2016). While the CBFB-SMMHC fusion protein plays a critical role in leukemogenesis, it does not independently initiate the disease, since CBFB-SMMHC initiated leukemogenesis has been shown in mouse models to require the cooperation of additional gene partners (Castilla et al., 1999; Castilla et al., 2004).

Unlike the t(15;17) reciprocal translocation, in which both the PML-RARa product and the reciprocal RARa-PML product are implicated in leukemogenesis (Lafage-Pochitaloff et al., 1995), the reciprocal product MYH11-CBFB for inv(16) is thought to be largely irrelevant to leukemogenesis (Marlton et al., 1995; Richter et al., 2016), with the lion’s share of leukemogenic activity coming from the CBFB-MYH11 transcript (Marlton et al., 1995). Functionally, the inv(16) cytogenetic inversion produces the same fusion product as the t(16;16) cytogenetic translocation (Eghtedar et al., 2012), and it has been shown that there is no statistically significant survival difference between t(16;16) cases and inv(16) cases of acute myeloid leukemia (Larson et al., 1986). Furthermore, AMLs characterized by inv(16) and the related t(16;16) show favorable prognosis (Larson et al., 1986; Delaunay, 2003; Eghtedar et al., 2012; Richter et al., 2016).

A driver of the leukemic phenotype in inv(16) leukemia is the histone deacetylase 8 (HDAC8)-mediated deacetylation of the p53 protein (Qi et al., 2015). HDAC8 expression is significantly enriched in inv(16) + AML cells compared to hematopoietic controls, and HDAC8 inhibition reduced inv(16) + AML cell viability, reducing cellular proliferation and inducing apoptosis in the AML cells while leaving control hematopoietic cells unaffected (Qi et al., 2015).

It seems that comparatively little work has been done to elucidate the mechanisms of epigenetic dysregulation in the inv(16) subtype. However, it has been shown that inv(16) AMLs tend to feature a hypomethylated phenotype relative to non-inv(16) AMLs, with the bulk of this hypomethylation occurring in the meningioma 1 (MN1) oncogene (Larmonie et al., 2017).

Epigenetic Discriminants within Acute Myeloid Leukemia

Epigenetic modifications comprise a class of heritable, non-genetic changes in gene expression, which commonly include DNA methylation, histone modification, and chromatin remodeling (Allis and Jenuwein, 2016). In healthy hematopoietic stem cells, epigenetic processes play a critical role in cell differentiation and hematopoiesis (Cullen et al., 2014; Kramer and Challen, 2017). Moreover, aberrations in epigenetic processes that disrupt normal hematopoiesis are implicated as contributory to the development of healthy hematopoietic cells into their leukemogenic counterparts (Hu and Shilatifard, 2016).

Indeed, epigenetics has been implicated as relevant to acute myeloid leukemia pathophysiology, as well as in colon cancer and in the CpG Island Methylator Phenotype (CIM) in glioblastoma (Toyota et al., 1999; Plass et al., 2008; Turcan et al., 2012). Further, Welch et al. have demonstrated that the mutational distribution of hematopoietic stem cells collected from healthy subjects is largely indifferent from the mutational distribution of the genomes of cells collected from AML samples. Some of the mutational burden observed in AML genomes is seemingly a result of random mutations that necessarily accumulate as a healthy progenitor cell ages, a process referred to as clonal hematopoiesis of indeterminate potential (CHIP) (Busque et al., 2012; Steensma et al., 2015). Upon the initiation of a leukemogenic event, the pre-existing mutations naturally accumulated by age in the progenitor cell are propagated into the now-cancerous genome (Welch et al., 2012).

AML genomes additionally tend to be less mutated relative to the genomes of other cancers (The Cancer Genome Atlas Research Network, 2013), with similar mutational distributions pre-relapse and post-relapse (Li et al., 2016a,b); furthermore, of these mutations, many tend to occur in genes associated with DNA methylation and epigenetic regulation, including DNMT3A, TET1/2, and IDH1/2 (Guillamot et al., 2016). These results suggest a relative independence between the progression of acute myeloid leukemia and the strictly genetic landscape of the disease, further indicating an increased importance for understanding AML through a non-genetic lens (Hassan et al., 2017), with a particular focus on DNA methylation and other epigenetic modalities.

A recent five-stage, relapsed AML study revealed several aspects of this genetic and epigenetic independence, specifically using an analysis of cytosine methylation at a series of consecutive CpG dinucleotides, with each of 16 possible methylation patterns at these loci termed an “epiallele” (Li et al., 2016a,b). The overall shift in these methylation patterns (the degree of “epigenetic shift”) across the genome of an AML specimen was quantified (as an entropy shift – delta Boltzmann entropy), with the magnitude of the overall methylation shift across a genome exhibiting prognostic relevance (Li et al., 2016a,b). An analysis of a cohort of AML patients split by median epigenetic shift demonstrated that patients with a high magnitude of epigenetic shift at diagnosis relapsed more quickly than did patients with a low magnitude of epigenetic shift (p = 0.0396, Mantel-Cox log rank test). The prognostic significance of epigenetic shift was preserved (p = 0.024, Cox proportional hazards regression model) in a subsequent multivariate analysis controlling for additional clinical variables including age, white blood cell count at diagnosis, and gender. None of these additional clinical variables was singularly significantly associated with time-to-relapse in the studied AML cohort, and Li et al. additionally found no significant association between the degree of somatic mutations and time-to-relapse (Li et al., 2016a,b).

Methylation signatures are further discriminative of acute myeloid leukemia subtypes. Figueroa et al. compiled a set of methylation profiles (50,000 CpG dinucleotides across ~14,000 unique gene loci) of blast cells collected from 344 AML subjects, clustering them via an unsupervised hierarchical method into 16 distinct clusters (Figueroa et al., 2010), each characterized by a unique epigenetic signature. Three of the generated epigenetic signature-differentiated clusters featured strong overlap with three established cytogenetic classes of acute myeloid leukemia commonly designated as having a relatively favorable prognosis: t(8;21) (Nishii et al., 2003); inv(16)/t(16;16) (Delaunay, 2003); and t(15;17) (Lavau and Dejean, 1994; Figueroa et al., 2010; Iland and Seymour, 2013). These subclasses have already shown some significance for stratifying treatment options in combination therapy, based on TET and IDH mutations (Shih et al., 2017).

Interestingly, cluster 3, an epigenetically distinct cluster that features strong overlap with the established t(8;21) cytogenetic class, also contains individual leukemia cases that do not feature the cytogenetic abnormality characteristic of their cluster’s class. Though cluster 3 is defined as a t(8;21) cluster due to being significantly enriched (p < 1.85E-25, Fisher’s exact test) with t(8;21) positive patients, the epigenetic cluster contains individual cases that are not positive for the cluster’s characteristic t(8;21) abnormality. Despite the cytogenetic difference, there is no significant difference (p = 0.83, log-rank test) in survival between cluster 3 patients with t(8;21) and cluster 3 patients without t(8;21), indicating that the underlying methylation signature is the primarily discriminative feature that predicts survival (Figueroa et al., 2010). An additional five of the generated clusters are characterized only by their unique epigenetic signatures – these clusters are not significantly enriched with or differentiated by other molecular, cytogenetic, or clinical factors known to be associated with acute myeloid leukemia. Figueroa et al. further report an epigenetic signature consisting of 45 genes, many hypermethylated, that feature consistent dysregulation across multiple AML subtypes and a supermajority of the AML cases studied. Additionally, the authors report a 16-gene methylation-based classifier predictive of AML-patient survival (p < 0.001) (Figueroa et al., 2010).

More recently, Singh et al. have identified DNA methylation and histone acetylation signatures discriminative of high-risk acute promyelocytic leukemia (APL), which, as described above, is often induced by a PML-RARa fusion oncogene (Lavau and Dejean, 1994; Figueroa et al., 2010; Iland and Seymour, 2013 Singh et al., 2018). Specifically, Singh et al. analyzed both the genomes and the epigenomes of two “high-risk” APL patients, characterized by a resistance to the all-trans-retinoic acid (ATRA) therapy commonly used to successfully treat APL (de Thé and Chen, 2010; Coombs et al., 2015; Singh et al., 2018) in comparison with ATRA-sensitive “lower-risk” APL patients. While there were no significantly discriminative genetic features separating the two risk groups in comparison with lower-risk APL cases, high-risk APL was characterized by a set of distinct epigenetic signatures. Through supervised clustering, Singh et al. identified a set of differentially methylated CpG sites between the high-risk APL and the low-risk APL groups, with a 23 CpG signature capable of differentiating high-risk APL cases (Singh et al., 2018).

A separate unsupervised hierarchical clustering analysis by Schoofs et al. found that, compared to healthy controls, AML genomes tend to be characterized by increased methylation variability (à la the “epigenetic shift” described by Li et al.). (Li et al., 2016a,b) and tend to feature DNA hypermethylation (with hypomethylation occurring at chromosomal ends) (Schoofs et al., 2012). Interestingly, Schoofs et al. noted that PML-RARa-binding sites were generally lacking of aberrant methylation signatures, despite the PML-RARa fusion being the hallmark initiating cytogenetic event for acute promyelocytic leukemia (Schoofs et al., 2012; De Braekeleer et al., 2014; Singh et al., 2018). Singh et al., however, found a distinct hyperacetylation of PML-RARa-binding sites in high-risk APL patients compared to low-risk APL samples, both untreated and post-treatment with all-trans-retinoic acid (Singh et al., 2018).

Finally, Gebhard et al. have recently differentiated CG-rich classes of AML through methylation profiling, which revealed that AML could be divided into two subgroups based on the likely mechanism of methylation. The first group was associated with repression of polycomb group (PcG) proteins, a class of proteins which contributes to epigenetic silencing of genes via chromatin remodeling. Methylation of these genes was found in all of the AML samples analyzed. The second group was comprised of highly heterogeneous hypo- and hyper-methylated samples. These methylation patterns demonstrate the true complexity of methylation in AML (Gebhard et al., 2018).

Approaches to Epigenetic Therapy

As the role of DNA methylation, histone acetylation, and related epigenetic events is, as described above, relevant to AML pathophysiology, the use of epigenetically focused therapies for AML is an area of active research. Several treatments targeting the epigenetic mechanisms and heterogeneity of AML are currently being explored.

Histone Methyltransferase Inhibitors

Histone methyltransferases can regulate gene transcription through the addition of a methyl group to a lysine (Murray, 1964) or arginine residue (Byvoet et al., 1972). Depending on the location of histone methylation and the number of methyl groups added to the residue, this process can lead to either gene transcription (Bernstein et al., 2002) or gene repression (Snowden et al., 2002). Many histone methyltransferases belong to the SET domain methyltransferase protein superfamily (Dillon et al., 2005), which transfer a methyl group from S-adenosyl-L-methionine (SAM) to the lysine residue, leaving a methylated lysine residue on the histone protein and S-adeno-L-homocysteine (SAH) as a byproduct of the reaction (Dillon et al., 2005). A similar process is also implicated in RNA methylation (Saletore et al., 2012), which has also recently been implicated in controlling stem cell differentiation in AML (Vu et al., 2017) and which can now be detected with single-molecule methods to create epialleles of phased RNA modifications on RNA (Novoa et al., 2017), just as we have described here for DNA.

Enhancer of Zest Homologue 2 (EZH2) is a histone lysine methyltransferase that catalyzes the Polycomb Repressive Complex 2 (PRC2) (Cao et al., 2002). PRC2 is involved in regular cellular division, and upregulation of its catalytic subunit, EZH2, has been shown to be a marker for prostate (Xu et al., 2012) and breast cancer (Kleer et al., 2003). Mutations in EZH2 can also change its substrate specificity. While the wild-type version of EZH2 normally leads to the conversion of H3K27 from mono-methylated to di-methylated, mutations in EZH2 further converts the H3K27 from di-methylated to tri-methylated. These mutations have been implicated in various types of lymphoma (Wigle et al., 2011).

Singh et al. explored the inhibition of EZH2 with a novel EZH2-inhibitor, MC2884. In an ex vivo analysis of high-risk acute promyelocytic leukemia (APL), Singh et al. found that the administration of MC2884 directly reduced the expression of EZH2 and induced apoptosis with implications that MC2884 reduced both methylation and acetylation at H3K27. The potential of this inhibitor to shape the leukemic epigenome and upregulate apoptosis has important implications for APL treatment (Singh et al., 2018).

Disruptor of telomeric silencing 1 (DOT1L) is a histone methyltransferase that targets the histone H3 lysine 79 position, or H3K79 (Feng et al., 2002). Uniquely, DOT1L does not bind in the same way as SET domain methyltransferase proteins tend to (Feng et al., 2002), and the unique AdoMet-binding method of DOT1L makes it an important target for therapeutic treatments (Sawada et al., 2004). Notably, differential degrees of methylation of the histone H3 lysine 79 position lead to varying cellular responses. H3K79 mono-methylation (H3K79me1) and di-methylation (H3K79me2) have been correlated with transcriptional activation (Zhang et al., 2004; Morillon et al., 2005), and analyses in a Saccharomyces cerevisiae yeast model have shown differential levels of H3K79me2 at various stages of the cell cycle (Schulze et al., 2009), indicating a potential regulatory role of H3K79 methylation in the cell cycle. Though there are some genes for which tri-methylation of H3K79 (H3K79me3) has an activating effect, H3K79me3 has been generally correlated with gene repression in human T cells.

DOT1L has additionally been implicated in oncogenesis for multiple leukemias, including mixed lineage leukemia (MLL), acute lymphoblastic leukemia (ALL), and acute myeloid leukemia (AML) (McLean et al., 2014). In particular, high levels of H3K79 methylation, indicative of high levels of DOT1L activity, were found in an analysis of MLL (Deshpande et al., 2013). Human DOT1L further interacts with AF10, an MLL-fusion partner, resulting in the upregulation of multiple leukemia-relevant genes, including HOXA9 and HOXA7 (Okada et al., 2005).

Due to the suspected role of DOT1L in oncogenesis, inhibition of this histone methyltransferase has been of interest. While multiple DOT1L inhibitors such as EPZ04777 (Daigle et al., 2011) and SGC0946 (Yu et al., 2012) have shown ability to inhibit the histone methyltransferase, they have non-optimal pharmacokinetic properties (Daigle et al., 2011). There has been some evidence that a DOT1L inhibitor such as pinometostat may have clinical efficacy if paired with other inhibitors or drugs (Stein et al., 2018), but further research needs to be conducted to design a robust DOT1L inhibitor that can be used safely for treatment in humans.

There exist additional epigenetic modalities of treatment, including histone demethylase inhibitors (D’Oto et al., 2016; Li et al., 2018), bromodomain-containing protein inhibitors (Zuber et al., 2011; Pérez-Salvia and Esteller, 2017), and mutant isocitrate dehydrogenase inhibitors (DiNardo et al., 2017; Nassereddine et al., 2017). These therapies offer promising further avenues for the epigenetic modulation of acute myeloid leukemia.

Histone Deacetylase Inhibitors

Histone deacetylase (HDAC) inhibitors are posited to treat cancers by preventing excessive deacetylation of histones and transcription factors, particularly those that regulate tumor-suppressor genes (Garcia-Manero et al., 2008). Alternatively, HDAC inhibitors could function through hyperacetylation that leads either to transcription factor overexpression or to gene acetylation (Glaser et al., 2003). One benefit of using HDAC inhibitors as clinical treatments is that they primarily mediate cancerous cell death and exhibit less cytotoxicity in healthy cells (Xu et al., 2007), thereby making them a more attractive therapeutic agent.

Generally, HDACs deacetylate lysine residues on proteins, including those that regulate cell death and cell proliferation, and are classified into one of four classes based on their similarity to yeast HDACs (Xu et al., 2007). The class I HDACs are HDACs 1, 2, 3, 8, and 11 (Glaser et al., 2003). This class of HDAC is categorized based on its homology to the yeast Rpd3 HDAC (Haberland et al., 2009). Structurally, the most common HDAC inhibitors include a hydroxamic acid or a benzamide zinc-binding group; however, current research has identified potential HDAC inhibitors that can bind zinc without the “pharmacokinetic liability” of the hydroxamic acid or toxicity of benzamide groups (Li and Woster, 2015). HDAC inhibitors can function by inhibiting cells from exiting the G1 stage of the cell cycle, by inducing apoptosis through upregulating death receptors and ligands or inducing the mitochondria driven apoptotic pathway, by causing DNA defects leading to cell death during mitosis, and by causing autophagy, among many possible pathways (Xu et al., 2007).

The drug panobinostat, for example, is an HDAC inhibitor used to treat AML. It functions to inhibit class I HDACs and degrade the oncoprotein AML1/ETO9a, which drives AML disease (Bots et al., 2014). It shows promise for treating leukemia that has an acquired resistance to apoptosis, including in p53 knockout cells and in cells treated with anti-apoptotic agents (Bots et al., 2014) and which can play a role in controlling the cell cycle (Chiron et al., 2013). Vorinostat is another HDAC inhibitor that shows promise in the treatment of AML. Phase 1 trials of vorinostat combined with cytarabine chemotherapy for patients with relapsed or refractory AML demonstrated positive clinical outcomes: 6 of 17 (35%) patients achieved complete remission, although 5 of those patients later relapsed and died (Mims et al., 2018). Interestingly, concurrent treatment of vorinostat with azacitidine did not improve clinical outcomes in comparison with patients receiving only azacitidine, which could be an indication that vorinostat interferes with cellular uptake of azacitidine (Craddock et al., 2017). In contrast, a JAK2-HDAC1 dual inhibitor showed significant in vivo anti-proliferative potential while also working to prevent opportunistic fungal infections, which are a consistent threat to immunocompromised AML patients (Huang et al., 2018). Overall, HDAC inhibitors show strong promise to treat some of the epigenetic bases of AML, although it is necessary to further examine combination therapies to ensure that the drugs do not interfere with each other’s ability to function.

DNA Methyltransferase Inhibitors

Decitabine (2′-deoxy-5-azacytidine) and azacytidine (5-azacytosine) are cytosine analogues (azanucleosides) that inhibit the function of DNA methyltransferase, an enzyme responsible for catalyzing the methylation of DNA (Jones and Taylor, 1980). As the inhibition of methylation can revert the hypermethylation-induced silencing of tumor-suppressor genes (Mund et al., 2006), the use of DNA methyltransferase inhibitors as therapeutic agents offers promising implications for acute myeloid leukemia and related malignancies.

Decitabine inhibits DNA methylation and suppresses growth in multiple human tumor cell lines by reactivating methylation-silenced tumor suppressors and regulators (Bender et al., 1999). Decitabine has been shown, for example, to demethylate and reactivate hypermethylated versions of the p51 tumor-suppressor gene, aiding in the restoration of normal cell cycle and function (Daskalakis et al., 2002). Decitabine has also been shown to upregulate BASP1, which is epigenetically silenced (via methylation of the BASP1-promoter region) in acute myeloid leukemias characterized by the t(8;21) translocation (Zhou et al., 2018). In a clinical analysis, continuous administration of decitabine resulted in genome-wide hypomethylation in patients with refractory solid tumors, though this hypomethylation was impermanent, with genomic methylation levels returning to pre-treatment baseline levels within 4–5 weeks (Samlowski et al., 2005).

The incorporation of azanucleosides into DNA and the associated inhibition of DNA methylation involve a complex set of biochemical mechanisms involving cellular uptake, intracellular metabolism, and azanucleoside incorporation. Though the precise mechanism of azanucleoside cellular uptake is uncertain, nucleoside transporters (Hubeek et al., 2005) and the equilibrative transporter ENT-1 (Huang et al., 2004) have been implicated in the process. Following cellular uptake, azanucleosides are metabolically activated and modified into 5-aza-2′-deoxycytidine-5′-triphosphate. The nucleoside is phosphorylated (ATP-dependent) into the monophosporylated nucleotide. The metabolic activation of both azacytidine and decitabine is catalyzed by enzymes specific to each compound (Momparler and Derse, 1979); deoxycytidine kinase has, for example, been implicated in the metabolic activation of decitabine (Stegmann et al., 1995).

Following their metabolic activation, azanucleosides are incorporated into DNA during DNA replication as replacements for cytosine. This leads to the formation of azacytosine-guanine dinucleotides, which are recognized by DNA methyltransferase. After recognition, methylation is initiated via a nucleophilic attack, resulting in the formation of a covalent bond between the carbon-6 atom on the azacytosine ring and the DNA methyltransferase enzyme. Unlike in the case of the interaction between regular cytosine and DNA methyltransferase, where a β-elimination occurs through the carbon-5 atom on the cytosine ring, azacytosine (where the carbon-5 is replaced by a nitrogen) and DNA methyltransferase remain covalently bound to one another until the methyltransferase enzyme is eventually degraded or removed (Santi et al., 1984). This mechanism serves to inhibit the function of the DNA methyltransferase, preventing methylation.

While some enzymes aid the process of azanucleoside activation and incorporation, there exist enzymes (e.g., cytidine deaminase), which have been implicated in the inhibition of this metabolic process (Chabot et al., 1983; Chuang et al., 2010). Thus, it is important to consider ways of countering these inhibitors and enhancing azanucleoside stability in order to optimize azanucleoside therapy. Indeed, the chemical stability of azanucleosides is an important consideration for determining the clinical viability of these compounds. In particular, the plasma stability of azanucleoside-class compounds is an important consideration for clinical contexts. S110, a dinucleotide that contains 5-aza-2′-deoxycytidine and a deoxyguanosine, is comparatively resistant to deamination by cytidine deaminase (Chuang et al., 2010) while still retaining a comparable ability to inhibit methylation (Yoo et al., 2007). Promisingly, azanucleosides further have considerable stability at both room and physiological temperatures, with this class of compounds still retaining efficiency in inhibiting genomic cytosine methylation subsequent to storage at room temperature (Stresemann and Lyko, 2008).

Though demethylation-based approaches are promising avenues of therapy for acute myeloid leukemia and similar malignancies, it is important to note that there are concerns associated with excessive hypomethylation. Hypomethylation has been shown in mouse models to lead to tumors via the genetic amplification (Gaudet et al., 2003) and insertional activation of oncogenic loci (Howard et al., 2008). Overall, while azanucleoside therapy is currently one of the most advanced epigenetic cancer therapies with a significant therapeutic potential in acute myeloid leukemia, further study is needed to better understand the molecular mechanisms and safety considerations of such therapies. Moreover, new combined CRISPR systems that use specific guide RNAs to selectively target specific loci for methylation or de-methylation can also turn the genome-wide therapy into one with more focused impact (Afshinnekoo and Mason, 2016).

Menin-MLL Inhibitors

Patients with de novo AML are found to have a chromosomal translocation of the mixed lineage leukemia gene located at chromosome band 11q23 (Cierpicki and Grembecka, 2014), a chromosomal translocation that results in the fusion of the MLL gene with one of over 60 different protein partners (Prange et al., 2017), with the AF9 protein being a common fusion partner in mixed lineage leukemia (Prange et al., 2017).

The disruption of the MLL gene caused by the gene fusion leads to the upregulation of HOXA9 and MEIS1 gene expression (Zeisig et al., 2003), the upregulation of which ultimately results in the inhibition of hematopoietic differentiation and the enhancement of cellular proliferation, resulting in acute leukemia (Zeisig et al., 2003). While chemotherapy is the standard treatment for this type of leukemia, the event-free survival is a comparatively low 50% for both pediatric and adult patients (Neff and Armstrong, 2013), necessitating novel treatment approaches.

MLL fusion proteins are dependent on interactions with menin, a protein encoded by the MEN1 gene, for their oncogenic function (Yokoyama et al., 2005). Menin binds to N-terminus of MLL in the MLL fusion protein (Caslini et al., 2007) and recruits the MLL fusion protein to activate genes, including HOXA9 (Chen et al., 2006).

Inhibiting the ability of menin to bind to MLL fusion proteins has been shown to inhibit cellular proliferation and induce cellular differentiation (Grembecka et al., 2012). Both the direct genetic disruption of the menin-MLL fusion protein interaction (Yokoyama et al., 2005) and the use of small-molecule inhibitors to block the menin-MLL protein interaction (Grembecka et al., 2012; He et al., 2014) induced cellular differentiation and reduced cellular proliferation in vitro (Yokoyama et al., 2005; Grembecka et al., 2012; He et al., 2014).

Borkin et al. characterized two additional small-molecule inhibitors – MI-463 and MI-503 – for in vivo use (Borkin et al., 2015). Introduction of MI-503 into a leukemic mouse model resulted in more than an 80% reduction in MV4;11 tumor volume in two mice, and treatment with MI-463 or MI-503 resulted in the median survival of MLL-AF9 leukemic mice increasing by ~70% and ~45%, respectively (Borkin et al., 2015). Further, MI-503 was noted to be metabolically stable, orally bioavailable, and lacking any noticeable hepatotoxic or nephrotoxic effects after prolonged, 38-day treatment in mice (Borkin et al., 2015), indicating therapeutic promise for the in vivo applications of this menin-MLL inhibitor.

Bromodomain-Containing Protein Inhibitors

Bromodomains (BRDs) are protein domains involved in the regulation of gene expression via recruitment of various molecular partners. BRDs recognize acetylated lysine residues, such as those found on histone tails, and recruit chromatin-modifying enzymes such as histone acetyltransferases (HATs) and histone deacetylases (HDACs) along with other transcriptional machinery to specific sites in the chromatin (Pérez-Salvia and Esteller, 2017).

The bromodomain and extra terminal (BET) protein family includes the proteins BRD2, BRD3, BRD4, BRD9, and BRDT (Pérez-Salvia and Esteller, 2017). BET proteins have been implicated in the regulation of some cancer-related genes (Delmore et al., 2011): for example, BRD4 recruits the positive transcription elongation factor complex (P-TEFb) to acetylated chromatin, leading to the transcriptional initiation and elongation of genes controlling cell proliferation (Yang et al., 2005); and BRD9 facilitates maintenance of the leukemic phenotype (Bakshi et al., 2010; Shi et al., 2013) as a component of the SWItch/Sucrose Non-Fermentable (SWI-SNF) chromatin-remodeling complex (Hohmann et al., 2016).

Delmore et al. demonstrated the use of a small-molecule inhibitor, JQ1, to inhibit BET proteins as regulatory factors for c-Myc (Delmore et al., 2011), which has been implicated in the pathogenesis of many cancers (Miller et al., 2012), including acute myeloid leukemia, in which c-Myc tends to be upregulated (Luo, 2005). The inhibition of BET by JQ1 has been shown to downregulate Myc transcription and, ultimately, Myc-dependent target genes, resulting in an overall reduction of cellular proliferation (Delmore et al., 2011).

Additional BET inhibitors have also been characterized. OTX015 (MK-8628) – a BRD2, BRD3, and BRD4 inhibitor – was demonstrated to have a stronger anti-proliferative effect than JQ1 (Berenguer-Daizé et al., 2016). In a similar manner to JQ1, OTX015 results in cell growth inhibition and apoptosis in acute leukemia cell lines (Coudé et al., 2015), and this inhibitor is currently undergoing clinical trials for AML treatment (Dombret et al., 2014).

Additionally, Hohmann et al. demonstrated that the BRD9 domain of SWI-SNF facilitates the leukemogenic ability of this chromatin-remodeling complex by facilitating the Myc transcription, the induction of cellular proliferation, and the inhibition of hematopoietic differentiation (Hohmann et al., 2016) that characterizes acute myeloid leukemia. The authors of this work further demonstrated the use of small-molecule BRD9 inhibitors – particularly, BI-7273 – that were effective in curbing cellular proliferation in AML cell lines (Hohmann et al., 2016). Martin et al. described another BRD9 inhibitor – BI-9564 – with an improved pharmacokinetic profile relative to that of BI-7273 (Martin et al., 2016).

Mutant Isocitrate Dehydrogenase (IDH1) Inhibitors

Approximately 6–10% of patients with acute myeloid leukemia (AML) have mutations in the isocitrate dehydrogenase 1 (IDH1) gene (DiNardo et al., 2018). IDH1 is involved in the regulation of cellular metabolism, specifically lipid metabolism given the ability of IDH1 to produce cytoplasmic NADPH (Koh et al., 2004) and in glucose sensing (Joseph et al., 2006). Wild-type isocitrate dehydrogenases (IDH) are enzymes that catalyze the oxidative decarboxylation of isocitrate to α-ketoglutarate (Cairns and Mak, 2013), and it has been suggested that the IDH1 Arg132 mutation causes a change in enzyme function, leading to the reduction of α-ketoglutarate to R(−)-2-hydroxyglutarate (2HG) (Dang et al., 2009). This excess of R(−)-2-hydroxyglutarate leads to an increase in cellular proliferation and impairs cellular differentiation (Losman et al., 2013).

Furthermore, 2HG has been shown to inhibit histone demethylation (Xu et al., 2011), a problematic effect given that histone methylation can result in the silencing of tumor-suppressor genes (Agrawal et al., 2007). Thus, researchers are exploring 2HG inhibition to reverse this hypomethylation, for instance, Li et al. demonstrated the use of AGI-5198, an IDH1 inhibitor, to inhibit 2HG production (Li et al., 2015). Treatment with AGI-5198 of cells containing mutated IDH1 indicated anti-tumoral activity and induced apoptosis (Li et al., 2015), suggesting potential clinical promise.

In 2018, the Food and Drug Administration (FDA) approved ivosidenib (Tibsovo®), an inhibitor of IDH1/2HG production, for adult AML patients with the IDH1 mutation (Dhillon, 2018), following clinical trials, which indicated that ivosidenib induced remissions and improved patient outcomes (DiNardo et al., 2018).

Histone Demethylase Inhibitors

Lysine-specific demethylase 1 (LSD1) is an enzyme responsible for the demethylation of histone H3 lysine 4, playing an important role in epigenetic regulation (Shi et al., 2004). LSD1 expression is enriched in multiple cancer cell types (Hayami et al., 2010), and LSD1 inhibition has been shown to induce cellular differentiation and inhibit cellular proliferation (Schulte et al., 2009).

In a mixed lineage leukemia mouse model, mice treated with LSD1 inhibitor showed substantial decreases in leukemic cell proliferation, but the treated mice presented with greater levels of anemia and thrombocytopenia compared to vehicle-treated control mice (Harris et al., 2012). Thus, many of the mice treated with LSD1 inhibitor died of anemia rather than leukemia, though this problem could be addressed in humans via adjustments to dosage and via blood transfusions (Harris et al., 2012). Though LSD1 presents as a potential therapeutic target in mixed lineage leukemias, additional considerations are necessary before usage of LSD1 inhibitors in human treatment.

It has been suggested that the pairing of an LSD1 inhibitor with other therapeutic agents may be more effective than the use of the LSD1 inhibitor alone. Schenk et al. demonstrated in mouse models of AML that combination therapy – all-trans retinoic acid (ATRA), which enhances cellular differentiation, in conjunction with tranylcypromine (an LSD1-inhibitor) – decreased tumor burden more substantially than the use of either agent alone (Schenk et al., 2012). Furthermore, their work suggests that LSD1 inhibits the pro-differentiative function of ATRA, and thus the inhibition of LSD1 via tranylcypromine allowed for ATRA to induce cellular differentiation to a more prominent extent (Schenk et al., 2012). This, along with other results from other studies (Ishikawa et al., 2017), points to the therapeutic potential in AML of combination therapy utilizing LSD1 inhibitors.

The Impact of Heterogeneity

Cellular heterogeneity is increasingly implicated in oncogenesis. Tumors evolve clonally through a series of stepwise, somatic mutations (Nowell, 1976). Like most cancers, acute myeloid leukemia often presents with intratumoral genetic and epigenetic heterogeneity, and it has been suggested that this heterogeneity is what leads to the high rate of relapse among patients post-therapy. However, relative to other cancers, AML has a comparatively low level of genetic heterogeneity (The Cancer Genome Atlas Research Network, 2013), suggesting that epigenetic heterogeneity is of primary importance (Li et al., 2016a,b), as well as other regulators such as non-coding RNAs and miRNAs (Hu et al., 2015). Here, we discuss epigenetic heterogeneity in acute myeloid leukemia.

Heterogeneity and Cancer Cell Population Fitness

Like in other cancers, heterogeneity in acute myeloid leukemia is likely maintained due to the increase in fitness associated with a heterogeneous population. In an environment with fluctuating selective pressures (such as one exposed to therapeutic drugs and radiation), it is beneficial for the population to employ a bet-hedging strategy (Brutovsky and Horváth, 2012). Although therapeutic agents (chemotherapy, radiation, etc.) are designed to eradicate cancerous cells, they are also a source of selective pressure in which therapy-resistant subclones are selected for (Turner and Reis-Filho, 2012; Greaves and Maley, 2012). In this way, cancer evolution can be thought of as a traditional Darwinian process in which less fit subclones are eliminated from the population by therapeutic agents (i.e., bet-hedging). More recently, there has been evidence that bet-hedging is not the only reason why heterogeneity is maintained in a population: increased epigenetic heritability for mutations that hasten mean cell division time and increased heterogeneity have been found to be positively correlated with mean population fitness in yeast (Cerulus et al., 2016).

Next Generation Sequencing and AML Heterogeneity

Recent research coupled with the development of next generation single-cell sequencing has advanced the characterization of AML and elucidated both the true complexity of and the genetic heterogeneity associated with the disease. Many genes, such as FLT3 and NPM1, display both homogeneity and heterogeneity in AML populations and appear to be undergoing convergent evolution (Paguirigan et al., 2015). More difficult to ascertain is the level of epigenetic heterogeneity in a sample, as epigenetic modifications are often dynamic. The study of epigenetic heterogeneity in AML is further complicated by the detection limit of sequencing (Chhangawala et al., 2015) and the high level of temporal and environmental phenotypic plasticity of AML epigenetic changes (Li et al., 2016a,b).

Epigenetic heterogeneity in acute myeloid leukemia is also present in differential chromatin landscapes and gene expression profiles. Through the comparison of two AML subtypes with two different translocations [(t8;21), in which the RUNX1-binding domain is fused to the ETO regulator, and t(3;21), in which the RUNX1-binding domain is fused to the EVI1 regulator], Loke et al. determined that these two subtypes each displayed a unique epigenetic landscape and gene expression profile. Each type had a different transcriptional network, which likely explains the differing clinical outcomes for the two types (Loke et al., 2017; Agirre et al., 2018). Clearly, it is necessary to not only consider genetic heterogeneity but also consider epigenetic heterogeneity to comprehensively characterize acute myeloid leukemia pathophysiology.

Genetic Heterogeneity

Interestingly, epiallele diversity may develop before AML and is enhanced by cooperation between somatic mutations (The Cancer Genome Atlas Research Network, 2013; Li et al., 2017). For example, mutational cooperativity between FLT3 mutations and TET2 mutations has been shown to confer a gain-of-function DNA methylation, leading to differentiated gene expression (Shih et al., 2015). In acute myeloid leukemia, somatic mutations are most often classified as nonsynonymous mutations in genes relevant for pathogenesis. Some of these changes influence epigenetic modifications, such as mutations in DNA-methylation genes and chromatin-modifying genes (The Cancer Genome Atlas Research Network, 2013). However, there is debate as to whether there is an association between the degree of epigenetic heterogeneity and the somatic mutational burden in epigenetic modifier genes (Li et al., 2016b) or not (Li et al., 2016a,b).

Epiallele diversity is inversely correlated with clinical outcome in acute myeloid leukemia, and more complex epialleles are associated with higher-risk AMLs (Li et al., 2016a,b; Li et al., 2017). More heterogeneous epialleles and higher epigenetic burden have been correlated with increased rate of relapse (Garrett-Bakelman et al., 2015). For example, children with the common t(8;21)(q22;q22)/RUNX1-RUNX1T1 mutation are more likely to relapse if they display heterogeneous DNA methylation at the time of diagnosis. Patients who relapsed had a unique epigenetic signature that led to aberrant activation of several cell-to-cell adhesion and cell-motility pathways, which likely contributed to their relapse (Zampini et al., 2018). This relationship is unsurprising, as epigenetic modification is a core component of plasticity, and increased plasticity usually confers increased fitness. A significant increase in the level of epigenetic heterogeneity precedes a significant change in somatic mutational burden, which also likely contributes to increased tumoral fitness and to therapeutic resistance upon relapse (Garrett-Bakelman et al., 2015).

Heterogeneity as a Driving Force

Acute myeloid leukemia can be generally characterized as “driven” by one of two driving factors in disease progression: higher epigenetic heterogeneity and lower mutational burden (epigenetically driven), and lesser epigenetic heterogeneity and higher mutational burden (genetically driven). The genetically driven cluster increases in epigenetic heterogeneity as the disease progresses (Li et al., 2016a,b). As relapse is often followed by mortality in acute myeloid leukemia, relapse states are important considerations for AML prognosis and progression. While the bulk of the research regarding the effect of heterogeneity on treatment efficacy, relapse occurrence, and disease prognosis is focused on genetic heterogeneity (Ding et al., 2012; Fisher et al., 2013; Hackl et al., 2017), several models have been developed that can predict prognosis and clinical outcome from methylation levels (Bullinger et al., 2009).

Targeted Epigenetic Therapies and the Heterogeneity Problem

Mutations in the isocitrate dehydrogenase 2 (IDH2) gene, which commonly lead to hematopoietic differentiation arrest, are a progenitor to AML. Enasidenib is a selective small-molecule inhibitor of IDH2; mutant IDH2 inhibition promotes leukemic cell differentiation. Many patients treated with enasidenib relapse due to selection for resistant clones and subsequent clonal evolution. Quek et al. recently used enasidenib as a model pathway to develop a method to reveal how cancerous subclones respond to therapeutic agents via clonal mapping throughout disease progression (Quek et al., 2018). Additionally, combined therapy using various IDH1 inhibitors has shown promise in resetting the epigenetic landscape to a baseline hematopoietic state (Shih et al., 2017).

However, there is evidence that epigenetic-targeted therapies can help to decrease epiallele diversity in acute myeloid leukemia. This is a likely reason for the efficacy of epigenetic treatments in AML, as overall AML epigenetic diversity decreases, so does overall AML fitness (Li et al., 2017). There exists difficulty in developing a method to target all possible subclones, so as to eliminate the possibility for resistance via clonal evolution (Roboz, 2014). This indicates that continued patient monitoring will likely be required, across both the genetic and epigenetic dimensions, to gauge the impact of therapy and optimize changes in patient treatments.

Future Directions

Acute myeloid leukemia is a complex disease characterized not only by significant genetic mutation but also by deleterious epigenetic differentiation, which has motivated the development of therapies that target epigenetic modifications in AML. However, as these epigenetic modifications are heritable and highly heterogeneous, acute myeloid leukemia undergoes significant clonal evolution, culminating in therapy resistance and eventual relapse. New animal models are being explored that can begin to tease out aspects of these diseases, such as in canine and primates (Pipes et al., 2013; Peng et al., 2015), and these can accelerate testing of the epigenetic hypotheses described here. Also, new sequencing technologies can improve both the accuracy and speed of characterizing AML, such as those that leverage hybrid assembly approaches (e.g., short read and long read) to study the complexity of the genome (Rosenfeld et al., 2016) and rapid, real-time sequencing methods that work in almost any environment (McIntyre et al., 2016; Castro-Wallace et al., 2017) can expand access to diagnostic approaches in NGS. Finally, efforts to explore and describe the heterogeneous epigenetic landscape of acute myeloid leukemia (e.g., via single-cell next generation sequencing) offer promising avenues for the optimization of AML therapies by taking the characteristic heterogeneity of acute myeloid leukemia into account, thereby improving outcomes for this malignancy.

Author Contributions

SG, CH, and MK equally authored the significant portions of the work, and SG and CH directed the development of the manuscript. AS, YJ, and EA contributed to individual sections of the work. EA provided feedback. CM contributed to various sections throughout the manuscript and provided guidance and feedback regarding the overall direction of the work.

Funding

The study received funding from the Irma T. Hirschl and Monique Weill-Caulier Charitable Trusts, the Bert L and N Kuggie Vallee Foundation, the WorldQuant Foundation, the Pershing Square Sohn Cancer Research Alliance, NASA (NNX14AH50G, NNX17AB26G), the National Institutes of Health (R25EB020393, R01NS076465, R01AI125416, R01ES021006, 1R21AI129851, 1R01MH117406), the Bill and Melinda Gates Foundation (OPP1151054), the Leukemia and Lymphoma Society (LLS) grants (LLS 923816, Mak, LLS-MCL-982, Chen-Kiang).

Conflict of Interest Statement

CM is a cofounder and a board member of Biotia and Onegevity Health, as well as an advisor or a compensated speaker of Abbvie, Acuamark Diagnostics, ArcBio, BioRad, DNA Genotek, Genialis, Genpro, Karius, Illumina, New England Biolabs, QIAGEN, Whole Biome, and Zymo Research.

The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Acknowledgments

We would like to thank the Epigenomics Core Facility at Weill Cornell Medicine as well as the Starr Cancer Consortium (I9-A9-071).

References

Afshinnekoo, A., and Mason, C. E. (2016). Epigenetic therapy in a new era of medicine: creating and integrating molecular profiles of patients. Ann. Transl. Med. 4:436. doi: 10.21037/atm.2016.11.19

PubMed Abstract | CrossRef Full Text | Google Scholar

Agirre, X., Meydan, C., Jiang, Y., Garate, L., Doane, A. S., Li, Z., et al. (2018). Long non-coding RNAs discriminate the stages and gene regulatory states of human humoral immune response. doi: 10.1038/s41467-019-08679-z

CrossRef Full Text | Google Scholar

Agrawal, S., Hofmann, W. -K., Tidow, N., Ehrich, M., Boom, D. v. d., Koschmieder, S., et al. (2007). The C/EBP tumor suppressor is silenced by hypermethylation in acute myeloid leukemia. Blood 109, 3895–3905. doi: 10.1182/blood-2006-08-040147

PubMed Abstract | CrossRef Full Text | Google Scholar

Allis, C. D., and Jenuwein, T. (2016). The molecular hallmarks of epigenetic control. Nat. Rev. Genet. 17, 487–500. doi: 10.1038/nrg.2016.59

PubMed Abstract | CrossRef Full Text | Google Scholar

Asou, N. (2003). The role of a Runt domain transcription factor AML1/RUNX1 in leukemogenesis and its clinical implications. Crit. Rev. Oncol. Hematol. 45, 129–150. doi: 10.1016/S1040-8428(02)00003-3

PubMed Abstract | CrossRef Full Text | Google Scholar

Bakshi, R., Hassan, M. Q., Pratap, J., Lian, J. B., Montecino, M. A., van Wijnen, A. J., et al. (2010). The human SWI/SNF complex associates with RUNX1 to control transcription of hematopoietic target genes. J. Cell. Physiol. 225, 569–576. doi: 10.1002/jcp.22240

PubMed Abstract | CrossRef Full Text | Google Scholar

Bender, C. M., Gonzalgo, M. L., Gonzales, F. A., Nguyen, C. T., Robertson, K. D., and Jones, P. A. (1999). Roles of cell division and gene transcription in the methylation of CpG islands. Mol. Cell. Biol. 19, 6690–6698. doi: 10.1128/MCB.19.10.6690

PubMed Abstract | CrossRef Full Text | Google Scholar

Berenguer-Daizé, C., Astorgues-Xerri, L., Odore, E., Cayol, M., Cvitkovic, E., Noel, K., et al. (2016). OTX015 (MK-8628), a novel BET inhibitor, displays in vitro and in vivo antitumor effects alone and in combination with conventional therapies in glioblastoma models. Int. J. Cancer 139, 2047–2055. doi: 10.1002/ijc.30256

PubMed Abstract | CrossRef Full Text | Google Scholar

Bernstein, B. E., Humphrey, E. L., Erlich, R. L., Schneider, R., Bouman, P., Liu, J. S., et al. (2002). Methylation of histone H3 Lys4 in coding regions of active genes. Proc. Natl. Acad. Sci. U.S.A. 99, 8695–8700. doi: 10.1073/pnas.082249499

CrossRef Full Text | Google Scholar

Borkin, D., He, S., Miao, H., Kempinska, K., Pollock, J., Chase, J., et al. (2015). Pharmacologic inhibition of the menin-MLL interaction blocks progression of MLL leukemia in vivo. Cancer Cell 27, 589–602. doi: 10.1016/j.ccell.2015.02.016

PubMed Abstract | CrossRef Full Text | Google Scholar

Bots, M., Verbrugge, I., Martin, B. P., Salmon, J. M., Ghisi, M., Baker, A., et al. (2014). Differentiation therapy for the treatment of t(8;21) acute myeloid leukemia using histone deacetylase inhibitors. Blood 123, 1341–1352. doi: 10.1182/blood-2013-03-488114

PubMed Abstract | CrossRef Full Text | Google Scholar

Brutovsky, B., and Horváth, D. (2012). Structure of intratumor heterogeneity: is cancer hedging its bets? arXiv:1307.0607

Google Scholar

Bullinger, L., Ehrich, M., Dohner, K., Schlenk, R. F., Dohner, H., Nelson, M. R., et al. (2009). Quantitative DNA methylation predicts survival in adult acute myeloid leukemia. Blood 115, 636–642. doi: 10.1182/blood-2009-03-211003

CrossRef Full Text | Google Scholar

Busque, L., Patel, J. P., Figueroa, M. E., Vasanthakumar, A., Provost, S., Hamilou, Z., et al. (2012). Recurrent somatic TET2 mutations in normal elderly individuals with clonal hematopoiesis. Nat. Genet. 44, 1179–1181. doi: 10.1038/ng.2413

PubMed Abstract | CrossRef Full Text | Google Scholar

Byvoet, P., Shepherd, G. R., Hardin, J. M., and Noland, B. J. (1972). The distribution and turnover of labelled methyl groups in histone fractions of cultured mammalian cells. Arch. Biochem. Biophys. 148, 558–567. doi: 10.1016/0003-9861(72)90174-9

PubMed Abstract | CrossRef Full Text | Google Scholar

The Cancer Genome Atlas Research Network. (2013). Genomic and epigenomic landscapes of adult de novo acute myeloid leukemia. N. Engl. J. Med. 368, 2059–2074. doi: 10.1056/NEJMoa1301689

PubMed Abstract | CrossRef Full Text | Google Scholar

Cairns, R. A., and Mak, T. W. (2013). Oncogenic isocitrate dehydrogenase mutations: mechanisms, models, and clinical opportunities. Cancer Discov. 3, 730–741. doi: 10.1158/2159-8290.CD-13-0083

PubMed Abstract | CrossRef Full Text | Google Scholar

Cao, R., Wang, L., Wang, H., Xia, L., Erdjument-Bromage, H., Tempst, P., et al. (2002). Role of histone H3 lysine 27 methylation in Polycomb-group silencing. Science 298, 1039–1043. doi: 10.1126/science.1076997

PubMed Abstract | CrossRef Full Text | Google Scholar

Caslini, C., Yang, Z., El-Osta, M., Milne, T. A., Slany, R. K., and Hess, J. L. (2007). Interaction of MLL amino terminal sequences with menin is required for transformation. Cancer Res. 67, 7275–7283. doi: 10.1158/0008-5472.CAN-06-2369

PubMed Abstract | CrossRef Full Text | Google Scholar

Castilla, L. H., Garrett, L., Adya, N., Orlic, D., Dutra, A., Anderson, S., et al. (1999). The fusion gene Cbfb-MYH11 blocks myeloid differentiation and predisposes mice to acute myelomonocytic leukaemia. Nat. Genet. 23, 144–146. doi: 10.1038/13776

PubMed Abstract | CrossRef Full Text | Google Scholar

Castilla, L. H., Perrat, P., Martinez, N. J., Landrette, S. F., Keys, R., Oikemus, S., et al. (2004). Identification of genes that synergize with Cbfb-MYH11 in the pathogenesis of acute myeloid leukemia. Proc. Natl. Acad. Sci. U.S.A. 101, 4924–4929. doi: 10.1073/pnas.0400930101

PubMed Abstract | CrossRef Full Text | Google Scholar

Castro-Wallace, S. L., Chiu, C. Y., John, K. K., Stahl, S. E., Rubins, K. H., McIntyre, A., et al. (2017). Nanopore DNA sequencing and genome assembly on the international space station. Sci. rep. 7:18022. doi: 10.1038/s41598-017-18364-0

CrossRef Full Text | Google Scholar

Cerulus, B., New, A. M., Pougach, K., and Verstrepen, K. J. (2016). Noise and epigenetic inheritance of single-cell division times influence population fitness. Curr. Biol. 26, 1138–1147. doi: 10.1016/j.cub.2016.03.010

PubMed Abstract | CrossRef Full Text | Google Scholar

Chabot, G. G., Bouchard, J., and Momparler, R. L. (1983). Kinetics of deamination of 5-aza-2′-deoxycytidine and cytosine arabinoside by human liver cytidine deaminase and its inhibition by 3-deazauridine, thymidine or uracil arabinoside. Biochem. Pharmacol. 32, 1327–1328. doi: 10.1016/0006-2952(83)90293-9

PubMed Abstract | CrossRef Full Text | Google Scholar

Chen, Y. -X., Yan, J., Keeshan, K., Tubbs, A. T., Wang, H., Silva, A., et al. (2006). The tumor suppressor menin regulates hematopoiesis and myeloid transformation by influencing Hox gene expression. Proc. Natl. Acad. Sci. U.S.A. 103, 1018–1023. doi: 10.1073/pnas.0510347103

PubMed Abstract | CrossRef Full Text | Google Scholar

Chhangawala, S., Rudy, G., Mason, C. E., and Rosenfeld, J. A. (2015). The impact of read length on quantification of differentially expressed genes and splice junction detection. Genome Biol. 16:131. doi: 10.1186/s13059-015-0697-y

CrossRef Full Text | Google Scholar

Chiron, D., Martin, P., Di Liberto, M., Huang, X., Ely, S., Lannutti, B., et al. (2013). Induction of prolonged early G1 arrest by CDK4/CDK6 inhibition reprogram lymphoma cells for durable PI3Kδ inhibition through PIK3IP1. Cell Cycle 12, 1892–1900. doi: 10.4161/cc.24928

CrossRef Full Text | Google Scholar

Chuang, J. C., Warner, S. L., Vollmer, D., Vankayalapati, H., Redkar, S., Bearss, D. J., et al. (2010). S110, a 5-aza-2′-deoxycytidine-containing dinucleotide, is an effective DNA methylation inhibitor in vivo and can reduce tumor growth. Mol. Cancer Ther. 9, 1443–1450. doi: 10.1158/1535-7163.MCT-09-1048

PubMed Abstract | CrossRef Full Text | Google Scholar

Cierpicki, T., and Grembecka, J. (2014). Challenges and opportunities in targeting the menin–MLL interaction. Future Med. Chem. 6, 447–462. doi: 10.4155/fmc.13.214

PubMed Abstract | CrossRef Full Text | Google Scholar

Coombs, C. C., Tavakkoli, M., and Tallman, M. S. (2015). Acute promyelocytic leukemia: where did we start, where are we now and the future. Blood Cancer J. 5, e304–e304. doi: 10.1038/bcj.2015.25

PubMed Abstract | CrossRef Full Text | Google Scholar

Coudé, M. -M., Braun, T., Berrou, J., Dupont, M., Bertrand, S., Masse, A., et al. (2015). BET inhibitor OTX015 targets BRD2 and BRD4 and decreases c-MYC in acute leukemia cells. Oncotarget 6, 17698–17712. doi: 10.18632/oncotarget.4131

PubMed Abstract | CrossRef Full Text | Google Scholar

Craddock, C. F., Houlton, A. E., Quek, L. S., Ferguson, P., Gbandi, E., Roberts, C., et al. (2017). Outcome of azacitidine therapy in acute myeloid leukemia is not improved by concurrent vorinostat therapy but is predicted by a diagnostic molecular signature. Clin. Cancer Res. 23, 6430–6440. doi: 10.1158/1078-0432.ccr-17-1423

PubMed Abstract | CrossRef Full Text | Google Scholar

Cullen, S. M., Mayle, A., Rossi, L., and Goodell, M. A. (2014). “Hematopoietic stem cell development” in Current topics in developmental biology. (Elsevier), 107, 39–75.

Google Scholar

Daigle, S., Olhava, E. J., Therkelsen, C. A., Majer, C. R., Sneeringer, C. J., Song, J., et al. (2011). Selective killing of mixed lineage leukemia cells by a potent small-molecule DOT1L inhibitor. Cancer Cell 20, 53–65. doi: 10.1016/j.ccr.2011.06.009

PubMed Abstract | CrossRef Full Text | Google Scholar

Dang, L., White, D. W., Gross, S., Bennett, B. D., Bittinger, M. A., Driggers, E. M., et al. (2009). Cancer-associated IDH1 mutations produce 2-hydroxyglutarate. Nature 462, 739–744. doi: 10.1038/nature08617

PubMed Abstract | CrossRef Full Text | Google Scholar

Daskalakis, M., Nguyen, T. T., Nguyen, C., Guldberg, P., Kohler, G., Wijermans, P., et al. (2002). Demethylation of a hypermethylated P15/INK4B gene in patients with myelodysplastic syndrome by 5-Aza-2′-deoxycytidine (decitabine) treatment. Blood 100, 2957–2964. doi: 10.1182/blood.V100.8.2957

PubMed Abstract | CrossRef Full Text | Google Scholar

De Braekeleer, E., Douet-Guilbert, N., and De Braekeleer, M. (2014). RARAfusion genes in acute promyelocytic leukemia: a review. Expert. Rev. Hematol. 7, 347–357. doi: 10.1586/17474086.2014.903794

PubMed Abstract | CrossRef Full Text | Google Scholar

De Kouchkovsky, I., and Abdul-Hay, M. (2016). Acute myeloid leukemia: a comprehensive review and 2016 update. Blood Cancer J. 6, e441–e441. doi: 10.1038/bcj.2016.50

PubMed Abstract | CrossRef Full Text | Google Scholar

Deshpande, A., Chen, L., Fazio, M., Sinha, A. U., Bernt, K. M., Banka, D., et al. (2013). Leukemic transformation by the MLL-AF6 fusion oncogene requires the H3K79 methyltransferase Dot1l. Blood 121, 2533–2541. doi: 10.1182/blood-2012-11-465120

PubMed Abstract | CrossRef Full Text | Google Scholar

de Thé, H., and Chen, Z. (2010). Acute promyelocytic leukaemia: novel insights into the mechanisms of cure. Nat. Rev. Cancer 10, 775–783. doi: 10.1038/nrc2943

CrossRef Full Text | Google Scholar

Delaunay, J. (2003). Prognosis of inv(16)/t(16;16) acute myeloid leukemia (AML): a survey of 110 cases from the French AML Intergroup. Blood 102, 462–469. doi: 10.1182/blood-2002-11-3527

PubMed Abstract | CrossRef Full Text | Google Scholar

Delmore, J. E., Issa, G. C., Lemieux, M. E., Rahl, P. B., Shi, J., Jacobs, H. M., et al. (2011). BET bromodomain inhibition as a therapeutic strategy to target c-Myc. Cell 146, 904–917. doi: 10.1016/j.cell.2011.08.017

PubMed Abstract | CrossRef Full Text | Google Scholar

Dhillon, S. (2018). Ivosidenib: first global approval. Drugs 78, 1509–1516. doi: 10.1007/s40265-018-0978-3

PubMed Abstract | CrossRef Full Text | Google Scholar

Di Croce, L. (2002). Methyltransferase recruitment and DNA hypermethylation of target promoters by an oncogenic transcription factor. Science 295, 1079–1082. doi: 10.1126/science.1065173

CrossRef Full Text | Google Scholar

Dillon, S., Zhang, X., Trievel, R. C., and Cheng, X. (2005). The SETdomain protein superfamily: protein lysine methyltransferases. Genome Biol. 6:227. doi: 10.1186/gb-2005-6-8-227

PubMed Abstract | CrossRef Full Text | Google Scholar

DiNardo, C. D., Stein, E. M., de Botton, S., Roboz, G. J., Altman, J. K., Mims, A. S., et al. (2018). Durable remissions with ivosidenib in IDH1-mutated relapsed or refractory AML. N. Engl. J. Med. 378, 2386–2398. doi: 10.1056/NEJMoa1716984

PubMed Abstract | CrossRef Full Text | Google Scholar

DiNardo, C. D., Stein, A. S., Fathi, A. T., Montesinos, P., Odenike, O., Kantarjian, H. M., et al. (2017). Mutant isocitrate dehydrogenase (mIDH) inhibitors, enasidenib or ivosidenib, in combination with azacitidine (AZA): preliminary results of a phase 1b/2 study in patients with newly diagnosed acute myeloid leukemia (AML). Blood 130(Suppl 1), 639. doi: 10.1200/JCO.2018.36.15_suppl.7042

CrossRef Full Text | Google Scholar

Ding, L., Ley, T. J., Larson, D. E., Miller, C. A., Koboldt, D. C., Welch, J. S., et al. (2012). Clonal evolution in relapsed acute myeloid leukemia revealed by whole genome sequencing. Nature 481, 506–510. doi: 10.1038/nature10738

PubMed Abstract | CrossRef Full Text | Google Scholar

Dombret, H., Preudhomme, C., Berthon, C., Raffoux, E., Thomas, X., Vey, N., et al. (2014). A phase 1 study of the BET-bromodomain inhibitor OTX015 in patients with advanced acute leukemia. Blood 124, 117. http://www.bloodjournal.org/content/124/21/117

Google Scholar

D’Oto, A., Tian, Q., Davidoff, A. M., and Yang, J. (2016). Histone demethylases and their roles in cancer epigenetics. J. Med. Oncol. Ther. 1, 34–40.

PubMed Abstract | Google Scholar

Eghtedar, A., Borthakur, G., Ravandi, F., Jabbour, E., Cortes, J., Pierce, S., et al. (2012). Characteristics of translocation (16;16)(p13;q22) acute myeloid leukemia. Am. J. Hematol. 87, 317–318. doi: 10.1002/ajh.22258

PubMed Abstract | CrossRef Full Text | Google Scholar

Feng, Q., Wang, H., Ng, H. H., Erdjument-Bromage, H., Tempst, P., Struhl, K., et al. (2002). Methylation of H3-lysine 79 is mediated by a new family of HMTases without a SET domain. Curr. Biol. 12, 1052–1058. doi: 10.1016/S0960-9822(02)00901-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Figueroa, M. E., Lugthart, S., Li, Y., Erpelinck-Verschueren, C., Deng, X., Christos, P. J., et al. (2010). DNA methylation signatures identify biologically distinct subtypes in acute myeloid leukemia. Cancer Cell 17, 13–27. doi: 10.1016/j.ccr.2009.11.020

PubMed Abstract | CrossRef Full Text | Google Scholar

Fisher, R., Pusztai, L., and Swanton, C. (2013). Cancer heterogeneity: implications for targeted therapeutics. Br. J. Cancer 108, 479–485. doi: 10.1038/bjc.2012.581

PubMed Abstract | CrossRef Full Text | Google Scholar

Fu, S., Consoli, U., Hanania, E. G., Zu, Z., Claxton, D. F., Andreeff, M., et al. (1995). PML/RARalpha, a Fusion Protein in Acute Promyelocytic Leukemia, Prevents Growth Factor Withdrawal-Induced Apoptosis in Tf-1 Cells. Clin. Cancer Res. 1, 583–590.

PubMed Abstract | Google Scholar

Gallagher, R. E., Li, Y. P., Rao, S., Paietta, E., Andersen, J., Etkind, P., et al. (1995). Characterization of Acute Promyelocytic Leukemia Cases with Pml-Rar Alpha Break/Fusion Sites in Pml Exon 6: Identification of a Subgroup with Decreased in Vitro Responsiveness to All-Trans Retinoic Acid. Blood 86, 1540–1547.

PubMed Abstract | Google Scholar

Garcia-Manero, G., Yang, H., Bueso-Ramos, C., Ferrajoli, A., Cortes, J., Wierda, W. G., et al. (2008). Phase 1 study of the histone deacetylase inhibitor vorinostat (suberoylanilide hydroxamic acid [SAHA]) in patients with advanced leukemias and myelodysplastic syndromes. Blood 111, 1060–1066. doi: 10.1182/blood-2007-06-098061

PubMed Abstract | CrossRef Full Text | Google Scholar

Garrett-Bakelman, F. E., Li, S., Chung, S. S., Hricik, T., Franck, R., Patel, J., et al. (2015). Divergent dynamics of epigenetic and genetic heterogeneity in relapsed acute myeloid leukemia. Blood 126, 306. http://www.bloodjournal.org/content/126/23/306?sso-checked=true

Google Scholar

Gaudet, F., Hodgson, J. G., Eden, A., Jackson-Grusby, L., Dausman, J., Gray, J. W., et al. (2003). Induction of tumors in mice by genomic hypomethylation. Science 300, 489–492. doi: 10.1126/science.1083558

PubMed Abstract | CrossRef Full Text | Google Scholar

Gebhard, C., Glatz, D., Schwarzfischer, L., Wimmer, J., Stasik, S., Nuetzel, M., et al. (2018). Profiling of aberrant DNA methylation in acute myeloid leukemia reveals subclasses of CG-rich regions with epigenetic or genetic association. Leukemia 33, 26–36. doi: 10.1038/s41375-018-0165-2

CrossRef Full Text | Google Scholar

Glaser, K. B., Staver, M. J., Waring, J. F., Stender, J., Ulrich, R. G., and Davidsen, S. K. (2003). Gene expression profiling of multiple histone deacetylase (HDAC) inhibitors: defining a common gene set produced by HDAC inhibition in T24 and MDA carcinoma cell lines. Mol. Cancer Ther. 2, 151–163. http://mct.aacrjournals.org/content/2/2/151.long

PubMed Abstract | Google Scholar

Greaves, M., and Maley, C. C. (2012). Clonal evolution in cancer. Nature 481, 306–313. doi: 10.1038/nature10762

PubMed Abstract | CrossRef Full Text | Google Scholar

Grembecka, J., He, S., Shi, A., Purohit, T., Muntean, A. G., Sorenson, R. J., et al. (2012). Menin-MLL inhibitors reverse oncogenic activity of MLL fusion proteins in leukemia. Nat. Chem. Biol. 8, 277–284. doi: 10.1038/nchembio.773

PubMed Abstract | CrossRef Full Text | Google Scholar

Grignani, F., Ferrucci, P. F., Testa, U., Talamo, G., Fagioli, M., Alcalay, M., et al. (1993). The acute promyelocytic leukemia-specific PML-RARα fusion protein inhibits differentiation and promotes survival of myeloid precursor cells. Cell 74, 423–431. doi: 10.1016/0092-8674(93)80044-F

PubMed Abstract | CrossRef Full Text | Google Scholar

Grossmann, V., Bacher, U., Kohlmann, A., Butschalowski, K., Roller, A., Jeromin, S., et al. (2012). Expression of CEBPA is reduced in RUNX1-mutated acute myeloid leukemia. Blood Cancer J. 2:e86. doi: 10.1038/bcj.2012.34

PubMed Abstract | CrossRef Full Text | Google Scholar

Guillamot, M., Cimmino, L., and Aifantis, I. (2016). The impact of DNA methylation in hematopoietic malignancies. Trends Cancer. 2, 70–83. doi: 10.1016/j.trecan.2015.12.006

PubMed Abstract | CrossRef Full Text | Google Scholar

Haberland, M., Montgomery, R. L., and Olson, E. N. (2009). The many roles of histone deacetylases in development and physiology: implications for disease and therapy. Nat. Rev. Genet. 10, 32–42. doi: 10.1038/nrg2485

PubMed Abstract | CrossRef Full Text | Google Scholar

Hackl, H., Astanina, K., and Wieser, R. (2017). Molecular and genetic alterations associated with therapy resistance and relapse of acute myeloid leukemia. J. Hematol. Oncol. 10:51. doi: 10.1186/s13045-017-0416-0

CrossRef Full Text | Google Scholar

Harris, W. J., Huang, X., Lynch, J. T., Spencer, G. J., Hitchin, J. R., Li, Y., et al. (2012). The histone demethylase KDM1A sustains the oncogenic potential of MLL-AF9 leukemia stem cells. Cancer Cell 21, 473–487. doi: 10.1016/j.ccr.2012.03.014

CrossRef Full Text | Google Scholar

Hassan, C., Afshinnekoo, E., Li, S., Wu, S., and Mason, C. E. (2017). Genetic and epigenetic heterogeneity and the impact on cancer relapse. Exp. Hematol. 54, 26–30. doi: 10.1016/j.exphem.2017.07.002

PubMed Abstract | CrossRef Full Text | Google Scholar

Hasserjian, R. P. (2013). Acute myeloid leukemia: advances in diagnosis and classification. Int. J. Lab. Hematol. 35, 358–366. doi: 10.1111/ijlh.12081

PubMed Abstract | CrossRef Full Text | Google Scholar

Hayami, S., Kelly, J. D., Cho, H. -S., Yoshimatsu, M., Unoki, M., Tsunoda, T., et al. (2010). Overexpression of LSD1 contributes to human carcinogenesis through chromatin regulation in various cancers. Int. J. Cancer 128, 574–586. doi: 10.1002/ijc.25349

CrossRef Full Text | Google Scholar

He, S., Senter, T. J., Pollock, J., Han, C., Upadhyay, S. K., Purohit, T., et al. (2014). High-affinity small-molecule inhibitors of the menin-mixed lineage leukemia (MLL) interaction closely mimic a natural protein–protein interaction. J. Med. Chem. 57, 1543–1556. doi: 10.1021/jm401868d

PubMed Abstract | CrossRef Full Text | Google Scholar

Hohmann, A. F., Martin, L. J., Minder, J. L., Roe, J. -S., Shi, J., Steurer, S., et al. (2016). Sensitivity and engineered resistance of myeloid leukemia cells to BRD9 inhibition. Nat. Chem. Biol. 12, 672–679. doi: 10.1038/nchembio.2115

PubMed Abstract | CrossRef Full Text | Google Scholar

Hou, H. -A., Lin, C. -C., Chou, W. -C., Liu, C. -Y., Chen, C. -Y., Tang, J. -L., et al. (2013). Integration of cytogenetic and molecular alterations in risk stratification of 318 patients with de novo non-M3 acute myeloid leukemia. Leukemia 28, 50–58. doi: 10.1038/leu.2013.236

CrossRef Full Text | Google Scholar

Howard, G., Eiges, R., Gaudet, F., Jaenisch, R., and Eden, A. (2008). Activation and transposition of endogenous retroviral elements in hypomethylation induced tumors in mice. Oncogene 27, 404–408. doi: 10.1038/sj.onc.1210631

PubMed Abstract | CrossRef Full Text | Google Scholar

Hu, D., and Shilatifard, A. (2016). Epigenetics of hematopoiesis and hematological malignancies. Genes Dev. 30, 2021–2041. doi: 10.1101/gad.284109.116

PubMed Abstract | CrossRef Full Text | Google Scholar

Hu, W., Dooley, J., Chung, S. S., Chandramohan, D., Cimmino, L., Mukherjee, S., et al. (2015). MiR-29a maintains mouse hematopoietic stem cell self-renewal by regulating Dnmt3a. Blood 125, 2206–2216. doi: 10.1182/blood-2014-06-585273

PubMed Abstract | CrossRef Full Text | Google Scholar

Huang, Y., Anderle, P., Bussey, K. J., Barbacioru, C., Shankavaram, U., Dai, Z., et al. (2004). Membrane transporters and channels: role of the transportome in cancer chemosensitivity and chemoresistance. Cancer Res. 64, 4294–4301. doi: 10.1158/0008-5472.CAN-03-3884

PubMed Abstract | CrossRef Full Text | Google Scholar

Huang, Y., Dong, G., Li, H., Liu, N., Zhang, W., and Sheng, C. (2018). Discovery of janus kinase 2 (JAK2) and histone deacetylase (HDAC) dual inhibitors as a novel strategy for the combinational treatment of leukemia and invasive fungal infections. J. Med. Chem. 61, 6056–6074. doi: 10.1021/acs.jmedchem.8b00393

PubMed Abstract | CrossRef Full Text | Google Scholar

Hubeek, I., Stam, R. W., Peters, G. J., Broekhuizen, R., Meijerink, J. P., van Wering, E. R., et al. (2005). The human equilibrative nucleoside transporter 1 mediates in vitro cytarabine sensitivity in childhood acute myeloid leukaemia. Br. J. Cancer 93, 1388–1394. doi: 10.1038/sj.bjc.6602881

PubMed Abstract | CrossRef Full Text | Google Scholar

Iland, H. J., and Seymour, J. F. (2013). Role of arsenic trioxide in acute promyelocytic leukemia. Curr. Treat. Options in Oncol. 14, 170–184. doi: 10.1007/s11864-012-0223-3

PubMed Abstract | CrossRef Full Text | Google Scholar

Ishikawa, Y., Nakayama, K., Morimoto, M., Mizutani, A., Nakayama, A., Toyoshima, K., et al. (2017). Synergistic anti-AML effects of the LSD1 inhibitor T-3775440 and the NEDD8-activating enzyme inhibitor pevonedistat via transdifferentiation and DNA rereplication. Oncogene 6:e377. doi: 10.1038/oncsis.2017.76

PubMed Abstract | CrossRef Full Text | Google Scholar

Jones, P. A., and Taylor, S. M. (1980). Cellular differentiation, cytidine analogs and DNA methylation. Cell 20, 85–93. doi: 10.1016/0092-8674(80)90237-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Joseph, J. W., Jensen, M. V., Ilkayeva, O., Palmieri, F., Alárcon, C., Rhodes, C. J., et al. (2006). The mitochondrial citrate/isocitrate carrier plays a regulatory role in glucose-stimulated insulin secretion. J. Biol. Chem. 281, 35624–35632. doi: 10.1074/jbc.M602606200

PubMed Abstract | CrossRef Full Text | Google Scholar

Kaushik, S., Liu, F., Veazey, K. J., Gao, G., Das, P., Neves, L. F., et al. (2017). Genetic deletion or small-molecule inhibition of the arginine methyltransferase PRMT5 exhibit anti-tumoral activity in mouse models of MLL-rearranged AML. Leukemia 32, 499–509. doi: 10.1038/leu.2017.206

CrossRef Full Text | Google Scholar

Kleer, C. G., Cao, Q., Varambally, S., Shen, R., Ota, I., Tomlins, S. A., et al. (2003). EZH2 is a marker of aggressive breast cancer and promotes neoplastic transformation of breast epithelial cells. Proc. Natl. Acad. Sci. U.S.A. 100, 11606–11611. doi: 10.1073/pnas.1933744100

CrossRef Full Text | Google Scholar

Koh, H. -J., Lee, S. -M., Son, B. -G., Lee, S. -H., Ryoo, Z. Y., Chang, K. -T., et al. (2004). Cytosolic NADP+-dependent isocitrate dehydrogenase plays a key role in lipid metabolism. J. Biol. Chem. 279, 39968–39974. doi: 10.1074/jbc.M402260200

PubMed Abstract | CrossRef Full Text | Google Scholar

Kramer, A., and Challen, G. A. (2017). The epigenetic basis of hematopoietic stem cell aging. Semin. Hematol. 54, 19–24. doi: 10.1053/j.seminhematol.2016.10.006

PubMed Abstract | CrossRef Full Text | Google Scholar

Krauth, M. -T., Eder, C., Alpermann, T., Bacher, U., Nadarajah, N., Kern, W., et al. (2014). High number of additional genetic lesions in acute myeloid leukemia with t(8;21)/RUNX1-RUNX1T1: frequency and impact on clinical outcome. Leukemia 28, 1449–1458. doi: 10.1038/leu.2014.4

PubMed Abstract | CrossRef Full Text | Google Scholar

Lafage-Pochitaloff, M., Alcalay, M., Brunel, V., Longo, L., Sainty, D., Simonetti, J., et al. (1995). Acute promyelocytic leukemia cases with nonreciprocal PML/RARa or RARa/PML fusion genes. Blood 85, 1169–1174.

PubMed Abstract | Google Scholar

Larmonie, N. S. D., Arentsen-Peters, T. C. J. M., Obulkasim, A., Valerio, D., Sonneveld, E., Danen-van Oorschot, A. A., et al. (2017). MN1 overexpression is driven by loss of DNMT3B methylation activity in inv(16) pediatric AML. Oncogene 37, 107–115. doi: 10.1038/onc.2017.293

CrossRef Full Text | Google Scholar

Larson, R. A., Williams, S. F., Le beau, M. M., Bitter, M. A., Vardiman, J. W., and Rowley, J. D. (1986). Acute myelomonocytic leukemia with abnormal eosinophils and inv(16) or t(16;16) has a favorable prognosis. Blood 68, 1242–1249.

PubMed Abstract | Google Scholar

Lavau, C., and Dejean, A. (1994). The t(15;17) translocation in acute promyelocytic leukemia. Leukemia 8, 1615–1621.

PubMed Abstract | Google Scholar

Lazarus, H. M., and Litzow, M. R. (2012). AML cytogenetics: the complex just got simpler. Blood 120, 2357–2358. doi: 10.1182/blood-2012-08-448555

PubMed Abstract | CrossRef Full Text | Google Scholar

Li, L., Paz, A. C., Wilky, B. A., Johnson, B., Galoian, K., Rosenberg, A., et al. (2015). Treatment with a small molecule mutant IDH1 inhibitor suppresses tumorigenic activity and decreases production of the oncometabolite 2-hydroxyglutarate in human chondrosarcoma cells. PLoS One 10:e0133813. doi: 10.1371/journal.pone.0139377

PubMed Abstract | CrossRef Full Text | Google Scholar

Li, S., Chen, X., Wang, J., Meydan, C., Glass, J. L., Shih, A. H., et al. (2017). Genetically-defined AML subgroups represent diverse epigenetic heterogeneity landscape. Blood 130(Suppl. 1), 243. http://www.bloodjournal.org/content/130/Suppl_1/243

Google Scholar

Li, S., Garrett-Bakelman, F. E., Chung, S. S., Sanders, M. A., Hricik, T., Rapaport, F., et al. (2016a). Distinct evolution and dynamics of epigenetic and genetic heterogeneity in acute myeloid leukemia. Nat. Med. 22, 792–799. doi: 10.1038/nm.4125

CrossRef Full Text | Google Scholar

Li, S., Mason, C., and Melnick, A. (2016b). Genetic and epigenetic heterogeneity in acute myeloid leukemia. Curr. Opin. Genet. Dev. 36, 100–106. doi: 10.1016/j.gde.2016.03.011

CrossRef Full Text | Google Scholar

Li, Y., Zhang, M., Sheng, M., Zhang, P., Chen, Z., Xing, W., et al. (2018). Therapeutic potential of GSK-J4, a histone demethylase KDM6B/JMJD3 inhibitor, for acute myeloid leukemia. J. Cancer Res. Clin. Oncol. 144, 1065–1077. doi: 10.1007/s00432-018-2631-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Li, Y., and Woster, P. M. (2015). Discovery of a new class of histone deacetylase inhibitors with a novel zinc binding group. Med. Chem. Commun. 6, 613–618. doi: 10.1039/c4md00401a

PubMed Abstract | CrossRef Full Text | Google Scholar

Link, K. A., Lin, S., Shrestha, M., Bowman, M., Wunderlich, M., Bloomfield, C. D., et al. (2016). Supraphysiologic levels of the AML1-ETO isoform AE9a are essential for transformation. Proc. Natl. Acad. Sci. U.S.A. 113, 9075–9080. doi: 10.1073/pnas.1524225113

PubMed Abstract | CrossRef Full Text | Google Scholar

Liu, S., Klisovic, R. B., Vukosavljevic, T., Yu, J., Paschka, P., Huynh, L., et al. (2007). Targeting AML1/ETO-histone deacetylase repressor complex: a novel mechanism for valproic acid-mediated gene expression and cellular differentiation in AML1/ETO-positive acute myeloid leukemia cells. J. Pharmacol. Exp. Ther. 321, 953–960. doi: 10.1124/jpet.106.118406

PubMed Abstract | CrossRef Full Text | Google Scholar

Loke, J., Assi, S. A., Imperato, M. R., Ptasinska, A., Cauchy, P., Grabovska, Y., et al. (2017). RUNX1-ETO and RUNX1-EVI1 differentially reprogram the chromatin landscape in t(8;21) and t(3;21) AML. Cell Rep. 19, 1654–1668. doi: 10.1016/j.celrep.2017.05.005

PubMed Abstract | CrossRef Full Text | Google Scholar

Losman, J. -A., Looper, R. E., Koivunen, P., Lee, S., Schneider, R. K., McMahon, C., et al. (2013). (R)-2-hydroxyglutarate is sufficient to promote leukemogenesis and its effects are reversible. Science 339, 1621–1625. doi: 10.1126/science.1231677

PubMed Abstract | CrossRef Full Text | Google Scholar

Luo, H. (2005). c-Myc rapidly induces acute myeloid leukemia in mice without evidence of lymphoma-associated antiapoptotic mutations. Blood 106, 2452–2461. doi: 10.1182/blood-2005-02-0734

PubMed Abstract | CrossRef Full Text | Google Scholar

Marlton, P., Claxton, D. F., Liu, P., Estey, E. H., Beran, M., LeBeau, M., et al. (1995). Molecular characterization of 16p deletions associated with inversion 16 defines the critical fusion for leukemogenesis. Blood 85, 772–779.

PubMed Abstract | Google Scholar

Martin, L. J., Koegl, M., Bader, G., Cockcroft, X. -L., Fedorov, O., Fiegen, D., et al. (2016). Structure-based design of an in vivo active selective BRD9 inhibitor. J. Med. Chem. 59, 4462–4475. doi: 10.1021/acs.jmedchem.5b01865

PubMed Abstract | CrossRef Full Text | Google Scholar

McIntyre, A. B. R., Rizzardi, L., Yu, A. M., Alexander, N., Rosen, G. L., Botkin, D. J., et al. (2016). Nanopore sequencing in microgravity. npj Microgravity 2:16035.

Google Scholar

McLean, C., Karemaker, I., and Van, L. (2014). The emerging roles of dot1l in leukemia and normal development. Leukemia 28, 2131–2138. doi: 10.1038/leu.2014.169

PubMed Abstract | CrossRef Full Text | Google Scholar

Melnick, A., and Licht, J. D. (1999). Deconstructing a disease: RARalpha, its fusion partners, and their roles in the pathogenesis of acute promyelocytic leukemia. Blood 93, 3167–3215.

PubMed Abstract | Google Scholar

Miller, D. M., Thomas, S. D., Islam, A., Muench, D., and Sedoris, K. (2012). c-Myc and cancer metabolism. Clin. Cancer Res. 18, 5546–5553. doi: 10.1158/1078-0432.CCR-12-0977

PubMed Abstract | CrossRef Full Text | Google Scholar

Mims, A. S., Mishra, A., Orwick, S., Blachly, J., Klisovic, R. S., Garzon, R., et al. (2018). A novel regimen for relapsed/refractory adult acute myeloid leukemia using a KMT2A partial tandem duplication targeted therapy: results of phase 1 study NCI 8485. Haematologica 103, 982–987. doi: 10.3324/haematol.2017.186890

PubMed Abstract | CrossRef Full Text | Google Scholar

Momparler, R. L., and Derse, D. (1979). Kinetics of phosphorylation of 5-aza-2′-deoxyycytidine by deoxycytidine kinase. Biochem. Pharmacol. 28, 1443–1444. doi: 10.1016/0006-2952(79)90454-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Mund, C., Brueckner, B., and Lyko, F. (2006). Reactivation of epigenetically silenced genes by DNA methyltransferase inhibitors: basic concepts and clinical applications. Epigenetics 1, 7–13. doi: 10.4161/epi.1.1.2375

PubMed Abstract | CrossRef Full Text | Google Scholar

Murray, K. (1964). The occurrence of ε-N-methyl lysine in histones. Biochemistry 3, 10–15.

Google Scholar

Nassereddine, S., Lap, C., Haroun, F., and Tabbara, I. (2017). The role of mutant idh1 and idh2 inhibitors in the treatment of acute myeloid leukemia. Ann. Hematol. 96, 1983–1991. doi: 10.1007/s00277-017-3161-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Neff, T., and Armstrong, S. A. (2013). Recent progress toward epigenetic therapies: the example of mixed lineage leukemia. Blood 121, 4847–4853. doi: 10.1182/blood-2013-02-474833

PubMed Abstract | CrossRef Full Text | Google Scholar

Nishii, K., Usui, E., Katayama, N., Lorenzo, F., Nakase, K., Kobayashi, T., et al. (2003). Characteristics of t(8;21) acute myeloid leukemia (AML) with additional chromosomal abnormality: concomitant trisomy 4 may constitute a distinctive subtype of t(8;21) AML. Leukemia 17, 731–737. doi: 10.1038/sj.leu.2402871

PubMed Abstract | CrossRef Full Text | Google Scholar

Novoa, E. M., Mason, C. E., and Mattick, J. S. (2017). Charting the unknown human epitranscriptome. Nat. Rev. Mol. Cell Biol. 18, 339–340. doi: 10.1038/nrm.2017.49

PubMed Abstract | CrossRef Full Text | Google Scholar

Nowell, P. C. (1976). The clonal evolution of tumor cell populations. Science 194, 23–28. doi: 10.1126/science.959840

PubMed Abstract | CrossRef Full Text | Google Scholar

Okada, Y., Feng, Q., Lin, Y., Jiang, Q., Li, Y., Coffield, V. M., et al. (2005). hDOT1L Links Histone Methylation to Leukemogenesis. Cell 121, 167–178. doi: 10.1016/j.cell.2005.02.020

PubMed Abstract | CrossRef Full Text | Google Scholar

Okuda, T., Nishimura, M., Nakao, M., and Fujitaa, Y. (2001). Int. J. Hematol. 74, 252. doi: 10.1007/BF02982057

CrossRef Full Text | Google Scholar

Pabst, T., Mueller, B. U., Harakawa, N., Schoch, C., Haferlach, T., Behre, G., et al. (2001). AML1–ETO downregulates the granulocytic differentiation factor C/EBPα in t(8;21) myeloid leukemia. Nat. Med. 7, 444–451. doi: 10.1038/86515

CrossRef Full Text | Google Scholar

Pabst, T., and Mueller, B. U. (2009). Complexity of CEBPA dysregulation in human acute myeloid leukemia. Clin. Cancer Res. 15, 5303–5307. doi: 10.1158/1078-0432.CCR-08-2941

PubMed Abstract | CrossRef Full Text | Google Scholar

Paguirigan, A. L., Smith, J., Meshinchi, S., Carroll, M., Maley, C., and Radich, J. P. (2015). Single-cell genotyping demonstrates complex clonal diversity in acute myeloid leukemia. Sci. Transl. Med. 7:281re2. doi: 10.1126/scitranslmed.aaa0763

PubMed Abstract | CrossRef Full Text | Google Scholar

Pérez-Salvia, M., and Esteller, M. (2017). Bromodomain inhibitors and cancer therapy: from structures to applications. Epigenetics 12, 323–339. doi: 10.1080/15592294.2016.1265710

PubMed Abstract | CrossRef Full Text | Google Scholar

Pipes, L., Li, S., Bozinoski, M., Palermo, R., Peng, X., Blood, P., et al. (2013). The nonhuman primate reference transcriptome resource (NHPRTR) for comparative functional genomics. Nucleic Acids Res. 41, D906–D914. doi: 10.1093/nar/gks1268

PubMed Abstract | CrossRef Full Text | Google Scholar

Peng, X., Thierry-Mieg, J., Thierry-Mieg, D., Nishida, A., Pipes, L., Bozinoski, M., et al. (2015). Tissue-specific transcriptome sequencing analysis expands the non-human primate reference transcriptome resource (NHPRTR). Nucleic Acids Res. 43, D737–D742. doi: 10.1093/nar/gku1110

PubMed Abstract | CrossRef Full Text | Google Scholar

Plass, C., Oakes, C., Blum, W., and Marcucci, G. (2008). Epigenetics in acute myeloid leukemia. Semin. Oncol. 35, 378–387. doi: 10.1053/j.seminoncol.2008.04.008

PubMed Abstract | CrossRef Full Text | Google Scholar

Prange, K. H. M., Mandoli, A., Kuznetsova, T., Wang, S. -Y., Sotoca, A. M., Marneth, A. E., et al. (2017). MLL-AF9 and MLL-AF4 oncofusion proteins bind a distinct enhancer repertoire and target the RUNX1 program in 11q23 acute myeloid leukemia. Oncogene 36, 3346–3356. doi: 10.1038/onc.2016.488

PubMed Abstract | CrossRef Full Text | Google Scholar

Qi, J., Singh, S., Hua, W. -K., Cai, Q., Chao, S. -W., Li, L., et al. (2015). HDAC8 inhibition specifically targets Inv(16) acute myeloid leukemic stem cells by restoring p53 acetylation. Cell Stem Cell 17, 597–610. doi: 10.1016/j.stem.2015.08.004

PubMed Abstract | CrossRef Full Text | Google Scholar

Quek, L., David, M., Kennedy, A., Metzner, M., Amatangelo, M., Shih, A., et al. (2018). Clonal heterogeneity of acute myeloid leukemia treated with the idh2 inhibitor enasidenib. Nat. Med. 24, 1167–1177. doi: 10.1038/s41591-018-0115-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Richter, L., Wang, Y., and Hyde, R. K. (2016). Targeting binding partners of the CBFβ-SMMHC fusion protein for the treatment of inversion 16 acute myeloid leukemia. Oncotarget 7, 66255–66266. doi: 10.18632/oncotarget.11357

PubMed Abstract | CrossRef Full Text | Google Scholar

Roboz, G. (2014). Epigenetic targeting and personalized approaches for AML. Hematology Am. Soc. Hematol. Educ. Program 2014, 44–51. doi: 10.1182/asheducation-2014.1.44

PubMed Abstract | CrossRef Full Text | Google Scholar

Rosenfeld, J., Reeves, D., Brugler, M. R., Narechania, A., Simon, S., Kolokotronis, S., et al. (2016). Genome assembly, annotation, and urban phylogenomics of the bedbug (Cimex lectularius). Nat. Commun. 7, 10164. doi: 10.1038/ncomms10164

CrossRef Full Text | Google Scholar

Saletore, Y., Meyer, K., Korlach, J., Vilfan, I., Jaffrey, S., and Mason, C. E. (2012). The birth of the epitranscriptome: deciphering the function of RNA modifications. Genome Biol. 13, 175. doi: 10.1186/gb-2012-13-10-175

PubMed Abstract | CrossRef Full Text | Google Scholar

Salomoni, P., and Pandolfi, P. P. (2002). The role of PML in tumor suppression. Cell 108, 165–170. doi: 10.1016/S0092-8674(02)00626-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Samlowski, W. E., Leachman, S. A., Wade, M., Cassidy, P., Porter-Gill, P., Busby, L., et al. (2005). Evaluation of a 7-day continuous intravenous infusion of decitabine: inhibition of promoter-specific and global genomic DNA methylation. J. Clin. Oncol. 23, 3897–3905. doi: 10.1200/JCO.2005.06.118

PubMed Abstract | CrossRef Full Text | Google Scholar

Santi, D. V., Norment, A., and Garrett, C. E. (1984). Covalent bond formation between a DNA-cytosine methyltransferase and DNA containing 5-azacytosine. Proc. Natl. Acad. Sci. U.S.A. 81, 6993–6997.

Google Scholar

Saultz, J. N., and Garzon, R. (2016). Acute myeloid leukemia: a concise review. J. Clin. Med. 5:33. doi: 10.3390/jcm5030033

PubMed Abstract | CrossRef Full Text | Google Scholar

Sawada, K., Yang, Z., Horton, J. R., Collins, R. E., Zhang, X., and Cheng, X. (2004). Structure of the conserved core of the yeast Dot1p, a nucleosomal histone H3 lysine 79 methyltransferase. J. Biol. Chem. 279, 43296–43306. doi: 10.1074/jbc.M405902200

PubMed Abstract | CrossRef Full Text | Google Scholar

Schenk, T., Chen, W. C., Göllner, S., Howell, L., Jin, L., Hebestreit, K., et al. (2012). Inhibition of the LSD1 (KDM1A) demethylase reactivates the all-trans-retinoic acid differentiation pathway in acute myeloid leukemia. Nat. Med. 18, 605–611. doi: 10.1038/nm.2661

PubMed Abstract | CrossRef Full Text | Google Scholar

Schoofs, T., Rohde, C., Hebestreit, K., Klein, H. -U., Gollner, S., Schulze, I., et al. (2012). DNA methylation changes are a late event in acute promyelocytic leukemia and coincide with loss of transcription factor binding. Blood 121, 178–187. doi: 10.1182/blood-2012-08-448860

CrossRef Full Text | Google Scholar

Schulte, J. H., Lim, S., Schramm, A., Friedrichs, N., Koster, J., Versteeg, R., et al. (2009). Lysine-specific demethylase 1 is strongly expressed in poorly differentiated neuroblastoma: implications for therapy. Cancer Res. 69, 2065–2071. doi: 10.1158/0008-5472.CAN-08-1735

PubMed Abstract | CrossRef Full Text | Google Scholar

Schulze, J., Jackson, J., Nakanishi, S., Gardner, J. M., Hentrich, T., Haug, J., et al. (2009). Linking cell cycle to histone modifications: SBF and H2B monoubiquitination machinery and cell-cycle regulation of H3K79 dimethylation. Mol. Cell 35, 626–641. doi: 10.1016/j.molcel.2009.07.017

PubMed Abstract | CrossRef Full Text | Google Scholar

Shi, J., Whyte, W. A., Zepeda-Mendoza, C. J., Milazzo, J. P., Shen, C., Roe, J. -S., et al. (2013). Role of SWI/SNF in acute leukemia maintenance and enhancer-mediated Myc regulation. Genes Dev. 27, 2648–2662. doi: 10.1101/gad.232710.113

PubMed Abstract | CrossRef Full Text | Google Scholar

Shi, Y., Lan, F., Matson, C., Mulligan, P., Whetstine, J. R., Cole, P. A., et al. (2004). Histone demethylation mediated by the nuclear amine oxidase homolog LSD1. Cell 119, 941–953. doi: 10.1016/j.cell.2004.12.012

CrossRef Full Text | Google Scholar

Shia, W. -J., Okumura, A. J., Yan, M., Sarkeshik, A., Lo, M. -C., Matsuura, S., et al. (2012). PRMT1 interacts with AML1-ETO to promote its transcriptional activation and progenitor cell proliferative potential. Blood 119, 4953–4962. doi: 10.1182/blood-2011-04-347476

PubMed Abstract | CrossRef Full Text | Google Scholar

Shih, A. H., Jiang, Y., Meydan, C., Shank, K., Pandey, S., Barreyro, L., et al. (2015). Mutational cooperativity linked to combinatorial epigenetic gain of function in acute myeloid leukemia. Cancer Cell 27, 502–515. doi: 10.1016/j.ccell.2015.03.009

PubMed Abstract | CrossRef Full Text | Google Scholar

Shih, A. H., Meydan, C., Shank, K., Garrett-Bakelman, F. E., Ward, P. S., Intlekofer, A. M., et al. (2017). Combination targeted therapy to disrupt aberrant oncogenic signaling and reverse epigenetic dysfunction in IDH2- and TET2-mutant acute myeloid leukemia. Cancer Discov. 7, 494–505. doi: 10.1158/2159-8290.CD-16-1049

PubMed Abstract | CrossRef Full Text | Google Scholar

Singh, A. A., Petraglia, F., Nebbioso, A., Yi, G., Conte, M., Valente, S., et al. (2018). Multi-omics profiling reveals a distinctive epigenome signature for high-risk acute promyelocytic leukemia. Oncotarget 9, 25647–25660. doi: 10.18632/oncotarget.25429

PubMed Abstract | CrossRef Full Text | Google Scholar

Snowden, A., Gregory, P. D., Case, C. C., and Pabo, C. O. (2002). Gene-specific targeting of H3K9 methylation is sufficient for initiating repression in vivo. Curr. Biol. 12, 2159–2166. doi: 10.1016/S0960-9822(02)01391-X

PubMed Abstract | CrossRef Full Text | Google Scholar

Stegmann, A. P., Honders, M. W., Willemze, R., and Landegent, J. E. (1995). De novo induced mutations in the deoxycytidine kinase (dck) gene in rat leukemic clonal cell lines confer resistance to cytarabine (AraC) and 5-aza-2′-deoxycytidine (DAC). Leukemia 9, 1032–1038.

PubMed Abstract | Google Scholar

Steensma, D. P., Bejar, R., Jaiswal, S., Lindsley, R. C., Sekeres, M. A., Hasserjian, R. P., et al. (2015). Clonal hematopoiesis of indeterminate potential and its distinction from myelodysplastic syndromes. Blood 126, 9–16. doi: 10.1182/blood-2015-03-631747

PubMed Abstract | CrossRef Full Text | Google Scholar

Stein, E. M., Garcia-Manero, G., Rizzieri, D. A., Tibes, R., Berdeja, J. G., Savona, M. R., et al. (2018). The DOT1L inhibitor pinometostat reduces H3K79 methylation and has modest clinical activity in adult acute leukemia. Blood 131, 2661–2669. doi: 10.1182/blood-2017-12-818948

PubMed Abstract | CrossRef Full Text | Google Scholar

Stresemann, C., and Lyko, F. (2008). Modes of action of the DNA methyltransferase inhibitors azacytidine and decitabine. Int. J. Cancer 123, 8–13. doi: 10.1002/ijc.23607

PubMed Abstract | CrossRef Full Text | Google Scholar

Subramanyam, D., Belair, C. D., Barry-Holson, K. Q., Lin, H., Kogan, S. C., Passegué, E., et al. (2010). PML-RAR and Dnmt3a1 cooperate in vivo to promote acute promyelocytic leukemia. Cancer Res. 70, 8792–8801. doi: 10.1158/0008-5472.can-08-4481

PubMed Abstract | CrossRef Full Text | Google Scholar

Toyota, M., Ahuja, N., Ohe-Toyota, M., Herman, J. G., Baylin, S. B., and Issa, J. P. (1999). CpG island methylator phenotype in colorectal cancer. Proc. Natl. Acad. Sci. U.S.A. 96, 8681–8686.

Google Scholar

Turcan, S., Rohle, D., Goenka, A., Walsh, L. A., Fang, F., Yilmaz, E., et al. (2012). IDH1 mutation is sufficient to establish the glioma hypermethylator phenotype. Nature 483, 479–483. doi: 10.1038/nature10866

PubMed Abstract | CrossRef Full Text | Google Scholar

Turner, N. C., and Reis-Filho, J. S. (2012). Genetic heterogeneity and cancer drug resistance. Lancet Oncol. 13, e178–e185. doi: 10.1016/S1470-2045(11)70335-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Vu, L. P., Perna, F., Wang, L., Voza, F., Figueroa, M. E., Tempst, P., et al. (2013). PRMT4 blocks myeloid differentiation by assembling a methyl-RUNX1-dependent repressor complex. Cell Rep. 5, 1625–1638. doi: 10.1016/j.celrep.2013.11.025

PubMed Abstract | CrossRef Full Text | Google Scholar

Vu, L., Pickering, B. F., Cheng, Y., Zaccara, S., Nguyen, D., Minuesa, G., et al. (2017). The N6-methyladenosine (m6A)-forming enzyme METTL3 controls myeloid differentiation of normal and leukemia cells. Nat. Med. 23, 1369–1376. doi: 10.1038/nm.4416

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, L., Gural, A., Sun, X. -J., Zhao, X., Perna, F., Huang, G., et al. (2011). The leukemogenicity of AML1-ETO is dependent on site-specific lysine acetylation. Science 333, 765–769. doi: 10.1126/science.1201662

PubMed Abstract | CrossRef Full Text | Google Scholar

Welch, J. S., Ley, T. J., Link, D. C., Miller, C. A., Larson, D. E., Koboldt, D. C., et al. (2012). The origin and evolution of mutations in acute myeloid leukemia. Cell 150, 264–278. doi: 10.1016/j.cell.2012.06.023

PubMed Abstract | CrossRef Full Text | Google Scholar

Wichmann, C., Quagliano-Lo Coco, I., Yildiz, Ö., Chen-Wichmann, L., Weber, H., Syzonenko, T., et al. (2015). Activating c-KIT mutations confer oncogenic cooperativity and rescue RUNX1/ETO-induced DNA damage and apoptosis in human primary CD34+ hematopoietic progenitors. Leukemia 29, 279–289. doi: 10.1038/leu.2014.179

PubMed Abstract | CrossRef Full Text | Google Scholar

Wigle, T., Knutson, S. K., Jin, L., Kuntz, K. W., Pollock, R. M., Richon, V. M., et al. (2011). The Y641C mutation of EZH2 alters substrate specificity for histone H3 lysine 27 methylation states. FEBS Lett. 585, 3011–3014. doi: 10.1016/j.febslet.2011.08.018

PubMed Abstract | CrossRef Full Text | Google Scholar

Xu, W. S., Parmigiani, R. B., and Marks, P. A. (2007). Histone deacetylase inhibitors: molecular mechanisms of action. Oncogene 26, 5541–5552. doi: 10.1038/sj.onc.1210620

PubMed Abstract | CrossRef Full Text | Google Scholar

Xu, K., Wu, Z. J., Groner, A. C., He, H. H., Cai, C., Lis, R. T., et al. (2012). EZH2 oncogenic activity in castration-resistant prostate cancer cells is Polycomb-independent. Science 338, 1465–1469. doi: 10.1126/science.1227604

PubMed Abstract | CrossRef Full Text | Google Scholar

Xu, W., Yang, H., Liu, Y., Yang, Y., Wang, P., Kim, S. -H., et al. (2011). Oncometabolite 2-hydroxyglutarate is a competitive inhibitor of α-ketoglutarate-dependent dioxygenases. Cancer Cell 19, 17–30. doi: 10.1016/j.ccr.2010.12.014

PubMed Abstract | CrossRef Full Text | Google Scholar

Yan, M., Burel, S. A., Peterson, L. F., Kanbe, E., Iwasaki, H., Boyapati, A., et al. (2004). Deletion of an AML1-ETO C-terminal NcoR/SMRT-interacting region strongly induces leukemia development. Proc. Natl. Acad. Sci. U.S.A. 101, 17186–17191. doi: 10.1073/pnas.0406702101

PubMed Abstract | CrossRef Full Text | Google Scholar

Yang, Z., Yik, J. H. N., Chen, R., He, N., Jang, M. K., Ozato, K., et al. (2005). Recruitment of P-TEFb for stimulation of transcriptional elongation by the bromodomain protein Brd4. Mol. Cell 19, 535–545. doi: 10.1016/j.molcel.2005.06.029

PubMed Abstract | CrossRef Full Text | Google Scholar

Yokoyama, A., Somervaille, T. C. P., Smith, K. S., Rozenblatt-Rosen, O., Meyerson, M., and Cleary, M. L. (2005). The menin tumor suppressor protein is an essential oncogenic cofactor for MLL-associated leukemogenesis. Cell 123, 207–218. doi: 10.1016/j.cell.2005.09.025

PubMed Abstract | CrossRef Full Text | Google Scholar

Yoo, C. B., Jeong, S., Egger, G., Liang, G., Phiasivongsa, P., Tang, C., et al. (2007). Delivery of 5-aza-2′-deoxycytidine to cells using oligodeoxynucleotides. Cancer Res. 67, 6400–6408. doi: 10.1158/0008-5472.CAN-07-0251

PubMed Abstract | CrossRef Full Text | Google Scholar

Yu, W., Chory, E. J., Wernimont, A. K., Tempel, W., Scopton, A., Federation, A., et al. (2012). Catalytic site remodelling of the DOT1L methyltransferase by selective inhibitors. Nat. Commun. 3, 1288. doi: 10.1038/ncomms2304

CrossRef Full Text | Google Scholar

Zampini, M., Tregnago, C., Bisio, V., Simula, L., Borella, G., Manara, E., et al. (2018). Epigenetic heterogeneity affects the risk of relapse in children with t(8;21)RUNX1-RUNX1T1-rearranged AML. Leukemia 32, 1124–1134. doi: 10.1038/s41375-017-0003-y

PubMed Abstract | CrossRef Full Text | Google Scholar

Zeisig, B. B., Milne, T., Garcia-Cuellar, M. -P., Schreiner, S., Martin, M. -E., Fuchs, U., et al. (2003). Hoxa9 and Meis1 are key targets for MLL-ENL-mediated cellular immortalization. Mol. Cell. Biol. 24, 617–628. doi: 10.1128/mcb.24.2.617-628.2004

CrossRef Full Text | Google Scholar

Zhang, H., Chen, Z., He, J., Miao, Y., Cen, J., and Qiu, Q. (2007). The coexpression of AML1-ETO9a isoform and AML1-ETO in acute myeloid leukemia M2 subtype and its clinical significance. Blood 110, 2390.

Google Scholar

Zhang, K., Siino, J. S., Jones, P. R., Yau, P. M., and Bradbury, E. M. (2004). A mass spectrometric "Western blot" to evaluate the correlations between histone methylation and histone acetylation. Proteomics 4, 3765–3775. doi: 10.1002/pmic.200400819

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhou, L., Fu, L., Lv, N., Liu, J., Li, Y., Chen, X., et al. (2018). Methylation-associated silencing of BASP1 contributes to leukemogenesis in t(8;21) acute myeloid leukemia. Exp. Mol. Med. 50. doi: 10.1038/s12276-018-0067-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Zuber, J., Shi, J., Wang, E., Rappaport, A. R., Herrmann, H., Sison, E. A., et al. (2011). RNAi screen identifies Brd4 as a therapeutic target in acute myeloid leukaemia. Nature 478, 524–528. doi: 10.1038/nature10334

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: acute myeloid leukemia, AML, epigenetics, heterogeneity, epialleles

Citation: Goldman SL, Hassan C, Khunte M, Soldatenko A, Jong Y, Afshinnekoo E and Mason CE (2019) Epigenetic Modifications in Acute Myeloid Leukemia: Prognosis, Treatment, and Heterogeneity. Front. Genet. 10:133. doi: 10.3389/fgene.2019.00133

Received: 04 October 2018; Accepted: 08 February 2019;
Published: 01 March 2019.

Edited by:

Asma Amleh, American University in Cairo, Egypt

Reviewed by:

Aniruddha Deshpande, Sanford Burnham Prebys Medical Discovery Institute (SBP), United States
Gokul Gopinath, Texas A&M University Baylor College of Dentistry, United States

Copyright © 2019 Goldman, Hassan, Khunte, Soldatenko, Jong, Afshinnekoo and Mason. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Christopher E. Mason, chm2042@med.cornell.edu

These authors have contributed equally to this work

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.