Next Article in Journal
Sperm Energy Restriction and Recovery (SER) Alters Epigenetic Marks during the First Cell Cycle of Development in Mice
Next Article in Special Issue
First Report of FARSA in the Regulation of Cell Cycle and Survival in Mantle Cell Lymphoma Cells via PI3K-AKT and FOXO1-RAG1 Axes
Previous Article in Journal
Effect of Quercetin on mitoBKCa Channel and Mitochondrial Function in Human Bronchial Epithelial Cells Exposed to Particulate Matter
Previous Article in Special Issue
PARP Inhibitor Inhibits the Vasculogenic Mimicry through a NF-κB-PTX3 Axis Signaling in Breast Cancer Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Targeting DNA Methylation in Leukemia, Myelodysplastic Syndrome, and Lymphoma: A Potential Diagnostic, Prognostic, and Therapeutic Tool

Department of Genetics, Cancer Research Institute, Biomedical Research Center of Slovak Academy of Sciences, 845 05 Bratislava, Slovakia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(1), 633; https://doi.org/10.3390/ijms24010633
Submission received: 5 December 2022 / Revised: 19 December 2022 / Accepted: 20 December 2022 / Published: 30 December 2022

Abstract

:
DNA methylation represents a crucial mechanism of epigenetic regulation in hematologic malignancies. The methylation process is controlled by specific DNA methyl transferases and other regulators, which are often affected by genetic alterations. Global hypomethylation and hypermethylation of tumor suppressor genes are associated with hematologic cancer development and progression. Several epi-drugs have been successfully implicated in the treatment of hematologic malignancies, including the hypomethylating agents (HMAs) decitabine and azacytidine. However, combinations with other treatment modalities and the discovery of new molecules are still the subject of research to increase sensitivity to anti-cancer therapies and improve patient outcomes. In this review, we summarized the main functions of DNA methylation regulators and genetic events leading to changes in methylation landscapes. We provide current knowledge about target genes with aberrant methylation levels in leukemias, myelodysplastic syndromes, and malignant lymphomas. Moreover, we provide an overview of the clinical trials, focused mainly on the combined therapy of HMAs with other treatments and its impact on adverse events, treatment efficacy, and survival rates among hematologic cancer patients. In the era of precision medicine, a transition from genes to their regulation opens up the possibility of an epigenetic-based approach as a diagnostic, prognostic, and therapeutic tool.

1. Introduction

Hematologic malignancies form a heterogeneous group of acute and chronic diseases, clonally expanding into the blood, bone marrow, and lymph nodes. Due to their highly aggressive manner, blood cancers are characterized by rapid progress, affecting both pediatric and adult patients. Improved treatment response and outcomes have been documented in pediatric patients, who achieve remission in the overwhelming majority. Genetic and epigenetic changes lead to the clonal proliferation of stem and progenitor cells. Alterations in downstream signaling pathways contribute to the disruption of the self-renewal ability of hematopoietic cells and their differentiation into other lineages.
Although epigenetic mechanisms have been largely evaluated in embryonic development, differentiation, and organogenesis [1,2], mounting evidence supports their critical role in cancer development and treatment. In a very recent concept, aberrant epigenetic programming is one of the hallmarks of malignant progression [3]. The process of DNA methylation is the most studied epigenetic mechanism in both normal and cancer cells, and its correct regulation is crucial for the transcription of different regulating genes, maintaining genome integrity, and proper immune responses [4]. Genetic changes in DNA methylation enzymes and regulators together with abnormal DNA methylation are associated with cancer-promoting changes [5]. In addition to gene silencing by promoter hypermethylation at CpG islands, DNA genome hypomethylation is a typical sign of human cancers [4].
DNA methylation plays a critical role in hematopoiesis and hematopoietic stem cell differentiation and proliferation [6]. Epigenetic modulation by DNA methylation is involved in several stages, including blood cell lineaging and the formation of the final cell types [7]. According to the findings, aberrant DNA methylation patterns during hematopoiesis linked to the dysfunction of DNA methylation-related enzymes often lead to blood cancer development [8]. With heterogeneity and different clinical severity of individual subtypes, it is more than inevitable that investigations into new diagnostic and therapeutic possibilities are performed.
In this review, we provide current data focusing on the emerging role of DNA methylation-related changes in hematologic malignancies, including leukemias, myelodysplastic syndromes, and lymphomas. The studies concerning the aberrant expression of DNA methylation enzymes and target genes are analyzed. Moreover, we will discuss the therapeutic implication of DNA methylation status and provide an overview of the clinical trials, evaluating the clinical significance of DNA methylation mechanisms in hematological malignancies.

2. Enzymes and Regulators of DNA Methylation in Hematologic Malignancies

The process of DNA methylation is catalyzed by enzymes called DNA methyltransferases (DNMTs) that transfer a methyl group from S-adenosine methionine (SAM) to the fifth carbon of a cytosine. Five DNMT genes are known in the human genome, including DNMT1, DNMT2, DNMT3A, DNMT3B, and DNMT3L, but only de novo DNMT3A, DNMT3B, and DNMT1 have a catalytic domain with methylation activity. There was much evidence proving that the regulation of DNA methylation by DNMTs is critical for normal hematopoiesis [9]. Changes in the methylation level can be the result of mutations causing gain or loss of function of the various enzymes involved in the methylation, including DNMTs, and other methylation modifiers such as ten-eleven translocation methylcytosine dioxygenases (TETs), isocitrate dehydrogenase 1/2; (IDH1/2), or others [10].

2.1. DNA Methyltransferases

During normal hematopoiesis, DNMT1 has an essential role in hematopoietic stem cell (HSC) self-renewal and multilineage hematopoietic differentiation, because the lack of DNMT1 in HSCs leads to a reduction in myeloid lineages compared to lymphoid lineages [11]. Together with other DNMTs, DNMT1 was substantially overexpressed in leukemia cells specifically according to leukemia type; likewise, acute myeloid leukemia (AML) cells with methylated p15(INAK4B) expressed higher levels of DNMT1 and DNMT3B [12]. Overexpression of DNMT1 was observed in other hematologic malignancies, including T-cell acute lymphoblastic leukemia (ALL), T-cell lymphomas, and diffuse large B-cell lymphoma (DLBCL) [13,14,15]. Nevertheless, in humans, only a few mutations at very low frequencies in the DNMT1 gene have been detected in AML cells [10].
Since DNMT1 represents one of the key methylation regulators, it could serve as a potential therapeutic target for lymphoid and myeloid malignancies. In chronic myeloid leukemia (CML), aberrant DNA hypermethylation contributes to low tyrosine kinase inhibitor (TKI) sensitivity and the persistence of leukemic stem cells (LSCs). Targeting DNMT1-mediated aberrant DNA hypermethylation using OR21 (orally available single-compound prodrug of decitabine) showed anti-tumor activity and impaired LSCs, suggesting it is a potential therapeutic target for CML [16]. DNMT3a and DNMT3b represent de novo DNMTs, which functionally cooperate in hematopoiesis and enable HSC differentiation. Although low-frequency mutations in the DNMT3B gene are found in human hematological malignancies, the role of DNMT3B was well documented in mouse hematopoiesis.
Lopusna et al. observed that 43% of Dnmt3b+/− mice developed T-cell lymphomas, chronic lymphocytic leukemia (CLL), and myeloproliferative disorders. Dnmt3b+/CI (CI, catalytically inactive) and Dnmt3bCI/CI mice were affected by B-cell rather than T-cell malignancies. These observations could suggest that Dnmt3b is critical to prevent B-cell transformation in vivo [17]. Moreover, the overexpression of DNMT3B and its isoforms has been associated with poor prognosis in older AML patients [18]. However, contradictory studies were published regarding DNMT3b expression in hematological malignancies [19]. As documented, DNMT3a is essential for HSC differentiation. Challen et al. reported a decline in the differentiation capacity of DNMT3a-null HSCs, resulting in the accumulation of undifferentiated HSCs in the bone marrow [20].
In hematopoietic disorders and malignancies, a high prevalence of DNMT3A mutations has been reported, mainly in patients with AML [21,22]. Somatic mutations decrease DNMT3A activity, induce CpG hypomethylation, and may promote the transformation of hematopoietic to cancer cells [23,24]. The most frequent DNMT3A hot-spot mutation R882H was observed in up to 60% of AML [21,25], 18% of ALL [26], and 10% of myelodysplastic syndrome (MDS) patients [27]. This alteration is associated with an adverse prognosis since it reduces DNA binding at the homodimeric interface. Spencer et al. observed a focal methylation loss in non-leukemic hematopoietic cells with a DNMT3A mutation R882H, suggesting that this mechanism could precede AML. As noted, CpG island hypermethylation in patients with wild-type DNMT3A did not correlate with gene silencing. According to the authors, DNMT3A-mediated hypermethylation might be a consequence of AML progression [28]. Importantly, DNMT3A mutations in AML patients are associated with shorter overall survival (OS) and predict an improved outcome with high-dose induction chemotherapy, representing a promising prognostic and predictive tool [29,30].

2.2. Modifiers and Regulators of DNA Methylation

TET enzymes represent DNA demethylation regulators, which are involved in the removal of methyl marks from DNA and play a role in hematopoietic differentiation. TET1 and TET3 genetic changes are very rare in leukemia. However, TET2 is one of the most mutated genes in AML, chronic myelomonocytic leukemia (CMML), and MDS [8]. In a cohort of 27 DLBCL patients, only 1 patient (3.7%) with a frameshift mutation in TET2 was detected, suggesting a low mutation rate in DLBCL patients [31]. Patients with low TET2 expression had markedly shorter OS in both non-M3 (hypergranular AML) and cytogenetically normal AML (CN-AML), with a significant correlation between TET2 expression and OS in CN-AML patients [32]. Cimmino et al. documented a tumor-suppressor role of TET1 and found alternative mechanisms of TET1 inactivation in human B-cell non-Hodgkin lymphomas (NHL) [33]. In contrast, high TET1 expression was independently associated with poor OS and complete remission (CR) in CN-AML patients [34].
IDH1/2 are enzymes that convert isocitrate to α-ketoglutarate, which could serve as a substrate for other enzymes, e.g., TET2. In a study on AML patient cohort, high-throughput resequencing revealed genetic alterations in IDH1, IDH2, and TET2 with frequencies of 6.2%, 8.6%, and 7.3%, respectively [35]. As noted, all 45 differentially methylated regions (DMRs) were hypermethylated in AML patients harboring IDH1/2 mutations. AML cells with mutations in TET2 or IDH1/2 displayed hypermethylation, suggesting its important function in hematopoietic cells among AML patients [35]. In contrast to AML, the incidence and prognostic value of IDH1/2 mutations in T-cell acute lymphoblastic leukemia (T-ALL) are poorly reported. Simonin et al. provided the first analysis of the oncogenetic landscape of IDH1/2 mutations in 1085 T-ALL patients. The results showed that almost 4% of IDH1/2 mutations correlated with poor prognosis [36].
Additionally, other regulators of DNA hypermethylation, including CCAAT/enhancer binding protein α (CEBPA) and Wilms tumor protein 1 (WT1), were reported to increase methylation in AML patients. CEBPA physically interacts with DNMT3A, and the N321D mutation in CEBPA decreases the DNA-binding ability of DNMT3. Furthermore, CEBPA mutations are responsible for promoter hypermethylation of various genes, including Polycomb Repression Complex (PCR2) target genes active in histone H3 methylation [37].

3. DNA Methylation of Target Genes in Leukemias, Myelodysplastic Syndromes, and Lymphomas

Genetic changes in regulators of the DNA methylation process can cause alterations in genome-methylation levels in the form of a loss of global methylation. On the other hand, almost intact global methylation was found in several hematologic malignancies. Genome-wide methylation studies in AML, CML, CLL, ALL, and lymphomas described different regions with aberrantly methylated genes, which could help to elucidate the malignant transformation process and find potential targets for diagnosis, prognosis, or therapy of hematological cancers [38,39,40,41,42,43].

3.1. Aberrant Methylation in MDS and AML

MDS and AML belong to the group of hematologic malignancies characterized by clonal hematopoiesis. Both diagnoses share similar clinical and pathologic features, but they differ in the percentage of blasts in peripheral blood and bone marrow. DNA methylation of different target genes has been described in patients with MDS and acute/chronic leukemia (Figure 1).
The HOX gene family represents the most studied genes, which are regulated by DNA methylation in AML. In an extensive study, Gao et al. identified 29 genes whose expression correlated with differently methylated CpG sites. Within the HOX family, high methylation was found in HOXA7, HOXA9, HOXA10, and HOXB3 genes [38]. Similar results were found in mesenchymal stromal cells from bone marrow in MDS and AML patients, where preferentially aberrant methylated genes were HOXA1, HOXA4, HOXA5, HOXA9, HOXA10, HOXA11, HOXB5, HOXC4, and HOXC6 [44].
Promoters of many tumor suppressor genes (TSGs) are most commonly inhibited by methylation. Cyclin-dependent kinase inhibitors p15 and p16, important in the regulation of proliferation, are frequently methylated in AML patients. Their higher methylation was associated with lower OS, recurrence-free survival (RFS), and frequency of CR [45,46,47]. In AML and MDS patients, the relationship between gene methylation status and shorter OS was reported for ERα, OLIG2, ITGBL1, SCIN, DLX5, MSH2, RAD50, and SOCS1 genes [48,49,50,51,52,53].
The DOK protein family, characterized as phosphotyrosine adapters with several functions in cell biology, is commonly expressed in myeloid cells. In a study comprising AML patients, He et al. observed that hypermethylation-mediated decreased expression of DOK1 and DOK2 genes was associated with lower OS [54]. Similarly, the higher methylation status of another member DOK6 gene was found in AML patients, but with the opposite effect on OS [55]. Longer OS was associated with higher promoter methylation in other genes, including C1R and DNA repair genes MLH1 and RAD51 [52,56].
Sestakova et al. performed an extensive validation study for 27 genes from 14 studies to verify the predictive role of aberrant DNA methylation in AML. The results showed that hypermethylation of CEBPA, PBX3, LZTS2, and NRGA1 serves as a predictor for longer survival [57]. In addition, higher methylation of GPX3 and DLX4 correlated with a favorable treatment impact, which is in contrast to previously reported studies documenting their correlation with lower OS [58,59].
In some cases, mutations in important transcription factors can lead to alterations in epigenetic modifications. Genetic changes in Runt-related transcription factor 1 (RUNX1), a key player in hematopoiesis, led to aberrant methylation in target genes. The study on the RUNX1-mutated AML cohort identified 51 differential methylated genes. As shown, the changes in expression profiles were found in ten of them. Hypermethylation of CD96, LTK, and MS4A3 correlated with poor prognosis, which could be related to their effects on cell cycle regulation and differentiation [60]. Higher methylation of SOX7, TCF21, CTNNA1, and CAMK4, as well as hypomethylation of DDX43, LZTS2, or NRGA1, was observed in MDS/AML patients. Due to the lack of prognostic potential in numerous studies, complex evaluation requires further investigation [61,62,63,64,65,66].

3.2. Aberrant Methylation in CML

CML is characterized by the presence of translocation-creating oncoprotein BCR-ABL1, and the development of this disease consists of three phases, namely, the chronic phase (CP), accelerated phase (AP), and blast crisis phase (BP) [67]. The high content of aberrantly methylated CpG compared to healthy donors was found through a methylome analysis of CML patients. Moreover, the number of abnormally methylated CpG increased from CP to the BP phase. Most of the CpG sites with increased methylation (88%) were located in GpG islands or in the very close region, which overlapped with 348 genes in the peripheral blood of BP patients [39].
Jelinek et al. observed that among 10 selected genes analyzed by pyrosequencing, ABL1, CDH3, and NPM2 presented the highest methylation in all phases of CML. Furthermore, increased methylation of CDKN2B (p15), OSCP1, PGRA, PGRB, and TFAP2E genes was described during CML progression [68]. Another extensive study defined 33 highly methylation-affected regions with hypermethylation of ABL1, WT1, ZNF577, and hypomethylation of G6B and TRIM15 [69]. DDX43 represents a frequently overexpressed gene in hematological malignancies. Epigenetic regulation of DDX43 by promoter methylation and a negative correlation between hypomethylation and higher expression of DDX43 was observed in CML patients. Moreover, the frequency of DDX43 hypomethylation increased in CP, AP, and BP by 23.4%, 25.0%, and 33.3%, respectively [70].
Promoter hypermethylation is mostly associated with TSGs involved in crucial cell functions, including the regulation of differentiation, proliferation, apoptosis, cell cycle, and growth. Similar to AML, HOXA4 and HOXA5 were hypermethylated in CML patients, and the presence of promoter methylation correlated with resistance to imatinib, belonging to a group of TKIs. Patients with HOXA4 and HOXA5 methylation levels higher than 63% showed 3.78- and 3.95-times-higher risk for imatinib resistance, respectively [71]. Imatinib resistance was also recorded in the case of higher methylation of OSCP1 and NPM2 genes [68]. Aberrant methylation of TSGs, including PLCD1, DLX4, DDIT3, PU.1, DAPK1, BIM, and GPX3, could represent potential prognostic or therapeutic targets in CML [72,73,74,75,76,77,78]. Accordingly, the downregulation of SHP-1 and relevant protein expression was associated with the presence of promoter methylation in advanced CML patients. SHP-1, a protein tyrosine phosphatase, is expressed mainly in HSC and plays a critical role in the regulation of JAK/STAT and MYC, AKT, and MAPK pathways. Alterations in SHP-1 methylation status could lead to the deregulation of included pathways and blastic transformation in CML patients [79].

3.3. Aberrant Methylation in ALL

ALL represents hematological cancer of immature T or B cells. T-ALL and B-ALL represent approximately 15–20% and 85% of all cases, respectively [80]. A genome-wide methylation analysis, based on nine genes with the identified CpG methylator phenotype, was capable of predicting a poor outcome subgroup of adult T-ALL. Patients with low methylation levels reported shorter OS and a higher risk of death in univariate and multivariate analyses. According to the results, the lowest methylation levels in patients were significantly associated with gender, younger age, and a higher count of white blood cells [81]. Similar to AML, CDKN2B (p15) presented decreased expression in most T-ALL cases, originating from deletion and promoter hypermethylation. CDKN2B hypermethylation frequently occurred together with mutations in DNMT3A and NRAS genes. In addition to the association with an older age of onset, the results showed a relatively early presence of T-cell precursors of ALL, causing the quick arrest of T-cell differentiation [82]. DLX3, belonging to the DLX gene family with a wide range of functions during hematopoiesis, could be active in the resistance to apoptosis. In pediatric B-cell ALL, Campo Dell Orto et al. observed aberrant methylation of DLX3 with reduced gene expression in patients with MLL-AF4 fusion, while no methylation was found in the subgroup with the TEL-AML1 fusion protein. The results suggested a potential role of DLX3 methylation in B-cell acute leukemias [83]. Some epigenetic promoter alterations can be lineage-specific. Higher methylation and methylation-mediated downregulation of RUNDC3B expression are typical for lymphoid but not myeloid malignancies [84]. RUNDC3B participates in the MAPK cascade in the role of Rap2-MAPK signaling mediator. Silencing by promoter methylation could disrupt the MAPK signaling pathway and promote leukemogenesis of lymphoid cells [84].
Several specific methylation profiles were found to be diagnostic, prognostic, or therapeutic markers for ALL. Chatterton et al. monitored the methylation of FOXE3 and TLX3 genes, showing their ability to discriminate between cancerous and healthy bone marrow samples with high specificity and sensitivity, which indicates their potential as diagnostic markers. Furthermore, TLX3 methylation correlated with minimal residual disease (MRD) in pediatric ALL patients [85]. In B-ALL and T-ALL-, RASSF6 and RASSF10 genes were frequently methylated and associated with the MRD in peripheral blood samples of adult ALL patients. In addition, the hypermethylation of RASSF6 is significantly associated with shorter OS in precursor B-ALL patients [86,87]. In a series of further studies, Roman-Gomez et al. described the role of promoter hypermethylation in a prognostic manner while observing the association of higher methylation of p21, WNT5A, Dkk-3, and NES1 genes with shorter disease-free survival (DFS) and OS [88,89,90,91]. Higher methylation of the PCDH17 gene was frequently observed in both B-ALL and T-ALL with a relationship with lower OS and increased risk for relapse and death [92,93].
Importantly, the specific gene methylation profiles can represent a predictive therapeutic marker because aberrant methylation can be responsible for chemotherapeutic resistance. TWIST2 hypermethylation and its inactivation were observed in more than 50% of ALL patients and 91% of samples from relapsed patients. In vitro experiments showed that the re-expression of TWIST2 increased apoptosis and sensitivity to chemotherapeutics [94]. A comparison of chemo-resistant and sensitive B-ALL pediatric patients detected higher levels of methylation in ADAMTSL5 (93% vs. 38%) and CDH11 (79% vs. 40%) in chemo-resistant vs. chemo-sensitive patients, respectively [95]. On the other hand, hypomethylation of the ASNS gene in T-ALL childhood patients was associated with poor outcomes and resistance to asparaginase, which is a high-dose drug involved in T-ALL therapy [96].

3.4. Aberrant Methylation in CLL

CLL belongs to the most common leukemias in the adult population, characterized by the clonal expansion of malignant B cells [97]. Similar to some other hematological malignancies, global hypomethylation is a characteristic sign of CLL. However, aberrantly methylated regions were characterized for this diagnosis in previous years [98]. In a genome-wide methylation study, Pei et al. identified approximately 1764 known genes with different methylation at 5’ regulatory regions. Among them, the results showed the presence of aberrant methylation in all four HOX gene clusters [40]. Previously, several studies found higher promoter methylation of genes, including TWIST2, DAPK1, SLIT2, or ZAP70 [99,100,101,102]. Increased expression of ZAP70 predicts poor outcomes for CLL patients, and CpG sites important for the regulation of transcription were identified in the 5’ regulatory region. According to these findings, a decreased methylation level in this specific CpG dinucleotide is a predictive biomarker for poor prognosis [102]. The WNT signaling pathway is generally involved in carcinogenesis and leukemogenesis. The constantly activated WNT pathway associated with the detection of hypermethylation of the seven WNT antagonist genes WIF1, DKK3, APC, SFRP1, SFRP2, SFRP4, and SFRP5 was observed in the peripheral blood of CLL patients. However, no association between methylation and clinical parameters was confirmed [103,104].
The hormone peptide Endothelin-1 (ET-1) plays a role in various cell functions, including proliferation. Microenvironment stimuli activated downstream receptors, leading to increased ET-1 expression in CLL. The unmethylated ET-1 gene was observed in healthy donors, while CLL patient samples exhibited 32% unmethylated and 68% methylated profiles. As shown, high methylation of the first ET-1 intron decreased its expression, suggesting the importance of epigenetic regulation [105]. The same authors detected that low methylation levels of neoangiogenic factor ANGPT2 correlated with increased expression and are associated with shorter OS and poor prognosis in CLL patients [106]. The prognostic potential of aberrant methylation in CLL was also observed in other genes, including PAX9, DUSP22, RPRM, SASH1, and CRY1 [107,108,109]. In the low-risk CLL subgroup, the hypermethylated CRY1 promoter inactivated its expression and was associated with better outcomes [109]. For CLL diagnosis, the most differentially methylated gene SHANK1 positively correlated with the absolute lymphocyte count. Increased SHANK1 methylation was found in samples during the pre-CLL diagnosis period, suggesting that epigenetic modification of SHANK1 occurred early in CLL carcinogenesis [110]. In comparison with SHANK1, the NFATC1 gene belongs to the most hypomethylated genes identified in CLL. A decreased methylation level is strongly associated with NFATC1 upregulation, resulting in the deregulation of target gene expression. Inactivation of NFAT regulator calcineurin by ibrutinib increased apoptosis in leukemic cells. Thus, NFACT1 might represent a potential therapeutic target in CLL diagnosis [111].
Two subgroups of CLL patients are classified according to the presence of somatic mutations in immunoglobulin (Ig) genes. The Ig heavy chain variable region (IGHV) with high mutational prevalence (IGHV-M) correlated with a more favorable prognosis compared to IGHV unmutated (IGHV-UM) CLL patients [112]. In the IGHV-UM subgroup, VHL and ABI3, acting as TSGs, were preferentially methylated and correlated with decreasing expression [113]. In contrast, the expression of WNT5A distinguished patients with worse outcomes in the IGHV-M subgroup. Reduced WNT5A expression through hypermethylation preferentially in three CpG dinucleotides within the regulatory region correlated with good prognoses [114].

3.5. Aberrant Methylation in Malignant Lymphomas

Malignant lymphomas (MLs) represent a heterogeneous group of hematologic malignancies in primary or secondary lymphatic organs arising from various types of B and T lymphocytes or NK cells. Generally, MLs cover classical Hodgkin lymphomas (HL) and diverse groups of NHL [115]. Pathogenesis of HL is presumably associated with family anamnesis or infection with Epstein–Barr virus (EBV). According to the previous findings in HL, DNA methylation led to the silencing of RASSF1A, p16INK4a, p18INK4c, p15INK4b, SYK, BOB.1/OBF.1, and CD79B [116,117,118]. Recently, differences in DNA methylation signatures were detected in a study with monozygotic triplets with HL. Two of the triplets with HL shared DNA methylation changes in naive B-cells and marginal zone-like B-cells compared to a healthy non-HL-triplet. Hypermethylation of one region within chromosome 18 in naive B-cells was found exclusively in HL triplets [119].
Bethge et al. identified 233 downregulated genes in a cohort of B-cell NHL patients. From the analyzed gene panel, DSP, FZD8, KCNH2, and PPP1R14A exhibited promoter methylation in 28%, 67%, 22%, and 78%, respectively. In addition, the highest methylation level after treatment with demethylating agents was detected in LRP12 and CDH1 genes, presenting 94% and 92%, respectively [120,121]. Another study evaluated eight genes associated with lymphoma pathogenesis and found decreased SIRT1 and increased KLF4, DAPK1, and SPG20 gene methylation levels. In vitro analysis revealed that DNMT1 did not affect hypermethylation maintenance of KLF4, DAPK1, and SPG20 genes [122,123].
Most DNA methylation studies have been performed in NHL, specifically in DLBCL (Figure 2). The results from the genome-wide methylation study reported approximately 200 differentially methylated genes in DLBCL patients with the aberrant methylation of p16/CDKN2A, p21/CDKN1A, and p27/CDKN1B [124]. However, only 37% of DLBCL patients had p16 methylation higher than 5%. According to the results, patients younger than 65 years manifested better progression-free survival (PFS) when the p16 methylation level reached more than 25% [125]. Aberrant methylation of another from CDK inhibitors, p57/KIP2, suggested that epigenetic modification of p57 could be established as a biomarker for MRD in DLBCL [126]. Shawky et al. analyzed the panel of 20 TSGs for promoter hypermethylation and correlation with clinical characteristics and patient outcomes in the DLBCL group. The methylation of several studied genes associated with survival and chemoresistance, specifically RUNX3, DAPK1, and MT16, represent prognostic factors for DFS. Moreover, hypermethylation of RUNX3 and CDH1 was shown to be an independent prognostic factor for OS [127].
Promoter methylation analysis of genes coding cadherins and protocadherins uncovered the association of CDH23 and PCDH10 hypermethylation and downregulated expression with worse outcomes in DLBCL patients. As noted, methylation of these genes could serve as a risk marker or a potential therapeutic target [128,129]. Several methylation studies investigated the DAPK1 gene showing higher methylation significantly associated with lower OS, disease-specific survival, and 5- year survival in the DLBCL patient cohort [130,131]. In addition, a prognostic and predictive potential of increased DAPK1 methylation in plasma samples was revealed in DLBCL. Patients with decreased methylation levels survived longer than patients with unchanged or regained DAPK1 methylation [132]. Specific promoter methylation with prognostic significance for DLBCL patients was discovered in several other genes, including SLIT2, DUSP4, and MGMT [133,134,135]. Importantly, Clozel et al. documented the link between aberrant DNA methylation and resistance to chemotherapeutics. Methylation analysis in chemoresistant DLBCL patients found nine hypermethylated genes. Among them, SMAD1 was a critical player. In a clinical trial on DLBCL patients, treatment with azacitidine followed by chemoimmunotherapy showed demethylation of SMAD1 and increased chemosensitivity [136].

4. Therapeutic Implication of DNA Methylation-Based Approach

Both global hypomethylation and specific promoter hypermethylation, primarily leading to the downregulation of TSGs, are important players in hematologic malignancies. Clinical investigations focused on drugs reversing altered epigenetic events, namely, epi-drugs. Two different types of DNMT inhibitors (DNMTi) represent analogs of nucleoside and non-nucleoside agents. Hypomethylating agents (HMAs) derived from cytidine, namely azacitidine and decitabine, were approved by the FDA for AML and MDS treatment [137]. Guadecitabine, a dinucleotide of decitabine and deoxyguanosine, belongs to the second generation of HMAs. Due to its better incorporation into DNA, guadecitabine exhibits a higher stabilization against cytidine deamination compared to azacitidine and decitabine.
In addition, the novel molecule GSK368503 was identified for the potential reversible and selective inactivation of DNMT1 in preclinical models. The GSK3685032 molecule showed an increased anti-proliferative effect and presented less cytotoxicity compared to decitabine (80% vs. 29% for the negative growth death index) in 51 cell lines derived from leukemia, lymphoma, and multiple myeloma. The application of GSK3685032 resulted in robust DNA hypomethylation and activation of transcription, leading to better survival in mouse models [138].
Mounting evidence documented that a combination of HMAs with chemotherapy, immunotherapy, or other treatment modalities can improve targeting, drug processing, and cytotoxicity (Figure 3).

4.1. HMAs Monotherapy

In the DACO-016 randomized phase III study for decitabine monotherapy, the median OS was 7.7 months in patients with AML treated with decitabine compared to the control arm with OS of 5.0 months (hazard ratio (HR), 0.85; 95% confidence interval (CI), 0.69–1.04; p = 0.1079). The survival rate was similar after an additional year, with a median OS in decitabine-intervened patients compared to controls at 7.7 and 5.0 months, respectively (HR, 0.82; 95% CI, 0.68–0.99; p = 0.0373). Furthermore, PFS and response rates were more favorable for decitabine than for other treatments [139]. Accordingly, in a phase III randomized study with azacitidine monotherapy, a longer OS was observed for azacitidine-treated patients compared to standard treatment (10.4 vs. 6.5 months) (HR 0.85; 95% CI, 0.69–1.03; p = 0.1009) [140].
In addition to standard subcutaneous and intravenous application of decitabine and azacitidine, oral azacitidine (CC-486) intervention was approved as maintenance therapy. In phase III of the QUAZAR AML-001 study, CC-486 administration to older AML patients previously treated with chemotherapy presented significantly longer median OS (24.7 vs. 14.8 months, p < 0.001) and relapse-free survival (10.2 vs. 4.8 months, p < 0.001) compared to the placebo group. In addition, adverse effects associated with CC-486 treatment rarely required discontinuation and were successfully managed with a change of dose or using a medication, suggesting a safety profile of CC-486 therapy for AML [141,142]. Due to rapid deamination, oral decitabine can be applicated only in combination with cedazuridine (cytidine deaminase inhibitor). In the phase I study comprising MDS and CMML patients, this combination showed a similar clinical response, safety profile, and demethylation to intravenous decitabine [143].
In phase II of the dose-expansion study, patients with AML received 60 or 90 mg/m2 of guadecitabine via subcutaneous administration in the five-day regimen or 60 mg/m2 in a ten-day regimen. As noted, higher doses did not show a benefit in terms of clinical outcomes for patients. Although complete response rates were higher in the ten-day regimen compared to the five-day regimen (18.9% vs. 8%; p = 0.15), the differences were not statistically significant. Adverse events during treatment occurred mainly during the ten-day regimen [144]. The response rates to guadecitabine in phase I and II studies were 17% and 23%, respectively [145]. In the guadecitabine study on MDS patients, the number of patients who achieved an overall response was similar between dose-differentiated groups (60 vs. 90 mg/m2). Moreover, 43% of relapsed or refractory MDS patients previously treated with HMAs achieved an overall response [146].
An overview of ongoing clinical trials focusing on the role of DNA methylation in leukemias and MDS, containing the study of decitabine/cedazuridine treatment with BMS-986253 in MDS patients, is shown in Table 1.

4.2. Combined HMAs Therapy with Other Treatments

Azacitidine pretreatment followed by high-dose chemotherapy with cytarabine and mitoxantrone showed an improved response rate in 61% of older AML patients with poor prognoses [147]. Furthermore, a combination of decitabine with vorinostat and chemotherapy led to a 54% overall response rate in pediatric patients with relapsed and refractory AML. The absence of MRD observed in 90% of patients showing 2-year OS for patients without or with adverse effects was 75.6% and 17.9% (p < 0.001), respectively [148]. Treatment with azacitidine and venetoclax (Bcl-2 inhibitor) represents a promising therapeutic approach since this combination showed significantly better survival and response rates compared to placebo or intensive chemotherapy-treated AML patients [149,150].
An active anti-cancer response by specific T-lymphocytes was observed in MDS patients who received an intervention with decitabine and the NY-ESO-1 vaccine [151]. However, Holmberg-Thydén et al. did not confirm an improved immune response when combining this vaccine and azacitidine application in high-risk MDS patients, and all patients progressed to AML [152]. Contradictory results were observed in clinical trials that studied a combination of HAMs with immune checkpoint inhibitors. While Saxena et al. and Daver et al. observed an anti-tumor immune response and better survival in AML [153,154], other studies did not find better clinical outcomes for MDS and AML patients compared to HMA treatment alone [155,156].
Nieto et al. studied a combination of azacitidine with high-dose chemotherapy (gemcitabine with busulfan and melphalan) and stem-cell transplantation in refractory/relapsed lymphomas, presenting that event-free survival and OS rates for HL and DLBCL patients were 76% vs. 95% and 65% vs. 77%, respectively [157]. Similar results were observed in a clinical trial with a combination of decitabine and a modified DHAP regimen (cisplatin, cytarabine, dexamethasone) in relapsed DLBCL patients, leading to improved responses and better prognosis [158]. In clinical trial phase II testing, CAR-T immunotherapy in relapsed DLBCL patients pretreated with decitabine, fludarabine, and cyclophosphamide showed a median follow-up of 10.9 months and overall response and CR at 90.9% and 63.6%, respectively. These findings suggested that the application of CAR-T therapy with decitabine pretreatment could be a potentially safe and efficient intervention for DLBCL patients [159].
Ongoing clinical trials involving HMAs in combined therapy aim to determine the optimal therapy dose and synergy effect in the treatment of lymphoma patients (Table 2).

5. Conclusions and Future Directions

Current clinical oncology and cancer research aims to find new biomarkers for precise diagnosis and new targets for more efficient anti-cancer treatment modalities. Due to the aggressive nature of hematologic malignancies, early and correct diagnosis is the key to improving prognosis, treatment results, and patient survival. The highly toxic treatment negatively affects the entire organism and develops severe side effects.
Enormous improvements in high-throughput sequencing methods and bioinformatic tools have helped to elucidate the impact of genetics and epigenetics on carcinogenesis. In addition to known genetic alterations, DNA methylation is one of the main epigenetic mechanisms that play a role in cancer initiation and progression. In the future, DNA methylation-based stratification of hematologic patients might lead to more personalized treatment with better outcomes. The reversibility of changes in DNA methylation landscapes enables broad clinical implications. However, adverse events associated with indiscriminate global hypomethylation with DNA methylation inhibitors are a matter of concern, and further investigations are highly warranted. According to the findings, DNA methylation processes in hematologic malignancies are usually associated with other mechanisms of epigenetic regulations, including histone modifications and miRNA regulation. Thus, evaluating the role of epigenetic modifications in a more complex matter would be beneficial for blood cancer patients.
Mounting evidence from clinical trials highlights the potential of therapeutic strategies focusing on reversing aberrant DNA methylation patterns in hematological malignancies. Importantly, a combination of epi-drugs with chemo- or immunotherapy remains an emerging research area. Recent findings suggest that the DNA methylation-based approach sensitizes other anti-cancer treatment modalities and may improve the clinical outcomes for hematologic cancer patients.

Author Contributions

Conceptualization, L.K. and S.C.; writing—original draft preparation L.K. and S.C. writing—review and editing, L.K., I.F., A.S. and S.C.; visualization, A.S., V.S., L.K. and S.C.; project administration, S.C.; funding acquisition, S.C. All authors have read and agreed to the published version of the manuscript.

Funding

The study was funded by the Scientific Grant Agency of the Ministry of Education, Science, Research, and Sport of the Slovak Republic and Slovak Academy of Sciences (VEGA) contract No. 2/0069/22 and 2/0036/19. The funding source had no influence on the writing of the manuscript.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We are very thankful to Michal Mego for critical reading and reviewing the Figures that were created with Biorender.com.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Goldberg, A.D.; Allis, C.D.; Bernstein, E. Epigenetics: A landscape takes shape. Cell 2007, 128, 635–638. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Zeng, Y.; Chen, T. DNA Methylation Reprogramming during Mammalian Development. Genes 2019, 10, 257. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Hanahan, D. Hallmarks of Cancer: New Dimensions. Cancer Discov. 2022, 12, 31–46. [Google Scholar] [CrossRef] [PubMed]
  4. Weisenberger, D.J.; Lakshminarasimhan, R.; Liang, G. The Role of DNA Methylation and DNA Methyltransferases in Cancer. Adv. Exp. Med. Biol. 2022, 1389, 317–348. [Google Scholar] [CrossRef] [PubMed]
  5. Dan, J.; Chen, T. Genetic Studies on Mammalian DNA Methyltransferases. Adv. Exp. Med. Biol. 2022, 1389, 111–136. [Google Scholar] [CrossRef] [PubMed]
  6. Cypris, O.; Frobel, J.; Rai, S.; Franzen, J.; Sontag, S.; Goetzke, R.; de Toledo, M.A.S.; Zenke, M.; Wagner, W. Tracking of epigenetic changes during hematopoietic differentiation of induced pluripotent stem cells. Clin. Epigenet. 2019, 11, 19. [Google Scholar] [CrossRef] [Green Version]
  7. Farlik, M.; Halbritter, F.; Muller, F.; Choudry, F.A.; Ebert, P.; Klughammer, J.; Farrow, S.; Santoro, A.; Ciaurro, V.; Mathur, A.; et al. DNA Methylation Dynamics of Human Hematopoietic Stem Cell Differentiation. Cell Stem Cell 2016, 19, 808–822. [Google Scholar] [CrossRef] [Green Version]
  8. Guillamot, M.; Cimmino, L.; Aifantis, I. The Impact of DNA Methylation in Hematopoietic Malignancies. Trends Cancer 2016, 2, 70–83. [Google Scholar] [CrossRef] [Green Version]
  9. Hoang, N.M.; Rui, L. DNA methyltransferases in hematological malignancies. J. Genet. Genom. 2020, 47, 361–372. [Google Scholar] [CrossRef]
  10. Blecua, P.; Martinez-Verbo, L.; Esteller, M. The DNA methylation landscape of hematological malignancies: An update. Mol. Oncol. 2020, 14, 1616–1639. [Google Scholar] [CrossRef]
  11. Trowbridge, J.J.; Snow, J.W.; Kim, J.; Orkin, S.H. DNA methyltransferase 1 is essential for and uniquely regulates hematopoietic stem and progenitor cells. Cell Stem Cell 2009, 5, 442–449. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Mizuno, S.; Chijiwa, T.; Okamura, T.; Akashi, K.; Fukumaki, Y.; Niho, Y.; Sasaki, H. Expression of DNA methyltransferases DNMT1, 3A, and 3B in normal hematopoiesis and in acute and chronic myelogenous leukemia. Blood 2001, 97, 1172–1179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Poole, C.J.; Zheng, W.; Lodh, A.; Yevtodiyenko, A.; Liefwalker, D.; Li, H.; Felsher, D.W.; van Riggelen, J. DNMT3B overexpression contributes to aberrant DNA methylation and MYC-driven tumor maintenance in T-ALL and Burkitt’s lymphoma. Oncotarget 2017, 8, 76898–76920. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Peters, S.L.; Hlady, R.A.; Opavska, J.; Klinkebiel, D.; Novakova, S.; Smith, L.M.; Lewis, R.E.; Karpf, A.R.; Simpson, M.A.; Wu, L.; et al. Essential role for Dnmt1 in the prevention and maintenance of MYC-induced T-cell lymphomas. Mol. Cell. Biol. 2013, 33, 4321–4333. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Amara, K.; Ziadi, S.; Hachana, M.; Soltani, N.; Korbi, S.; Trimeche, M. DNA methyltransferase DNMT3b protein overexpression as a prognostic factor in patients with diffuse large B-cell lymphomas. Cancer Sci. 2010, 101, 1722–1730. [Google Scholar] [CrossRef]
  16. Kamachi, K.; Ureshino, H.; Watanabe, T.; Yoshida, N.; Yamamoto, Y.; Kurahashi, Y.; Fukuda-Kurahashi, Y.; Hayashi, Y.; Hirai, H.; Yamashita, S.; et al. Targeting DNMT1 by demethylating agent OR-2100 increases tyrosine kinase inhibitors-sensitivity and depletes leukemic stem cells in chronic myeloid leukemia. Cancer Lett. 2022, 526, 273–283. [Google Scholar] [CrossRef]
  17. Lopusna, K.; Nowialis, P.; Opavska, J.; Abraham, A.; Riva, A.; Haney, S.L.; Opavsky, R. Decreases in different Dnmt3b activities drive distinct development of hematologic malignancies in mice. J. Biol. Chem. 2021, 296, 100285. [Google Scholar] [CrossRef]
  18. Niederwieser, C.; Kohlschmidt, J.; Volinia, S.; Whitman, S.P.; Metzeler, K.H.; Eisfeld, A.K.; Maharry, K.; Yan, P.; Frankhouser, D.; Becker, H.; et al. Prognostic and biologic significance of DNMT3B expression in older patients with cytogenetically normal primary acute myeloid leukemia. Leukemia 2015, 29, 567–575. [Google Scholar] [CrossRef]
  19. Zhang, H.; Ying, H.; Wang, X. Methyltransferase DNMT3B in leukemia. Leuk. Lymphoma 2020, 61, 263–273. [Google Scholar] [CrossRef]
  20. Challen, G.A.; Sun, D.; Jeong, M.; Luo, M.; Jelinek, J.; Berg, J.S.; Bock, C.; Vasanthakumar, A.; Gu, H.; Xi, Y.; et al. Dnmt3a is essential for hematopoietic stem cell differentiation. Nat. Genet. 2011, 44, 23–31. [Google Scholar] [CrossRef]
  21. Yan, X.J.; Xu, J.; Gu, Z.H.; Pan, C.M.; Lu, G.; Shen, Y.; Shi, J.Y.; Zhu, Y.M.; Tang, L.; Zhang, X.W.; et al. Exome sequencing identifies somatic mutations of DNA methyltransferase gene DNMT3A in acute monocytic leukemia. Nat. Genet. 2011, 43, 309–315. [Google Scholar] [CrossRef] [PubMed]
  22. Park, D.J.; Kwon, A.; Cho, B.S.; Kim, H.J.; Hwang, K.A.; Kim, M.; Kim, Y. Characteristics of DNMT3A mutations in acute myeloid leukemia. Blood Res. 2020, 55, 17–26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Zhang, Z.M.; Lu, R.; Wang, P.; Yu, Y.; Chen, D.; Gao, L.; Liu, S.; Ji, D.; Rothbart, S.B.; Wang, Y.; et al. Structural basis for DNMT3A-mediated de novo DNA methylation. Nature 2018, 554, 387–391. [Google Scholar] [CrossRef] [PubMed]
  24. Anteneh, H.; Fang, J.; Song, J. Structural basis for impairment of DNA methylation by the DNMT3A R882H mutation. Nat. Commun. 2020, 11, 2294. [Google Scholar] [CrossRef]
  25. Ley, T.J.; Ding, L.; Walter, M.J.; McLellan, M.D.; Lamprecht, T.; Larson, D.E.; Kandoth, C.; Payton, J.E.; Baty, J.; Welch, J.; et al. DNMT3A mutations in acute myeloid leukemia. N. Engl. J. Med. 2010, 363, 2424–2433. [Google Scholar] [CrossRef] [Green Version]
  26. Aref, S.; El Menshawy, N.; El-Ghonemy, M.S.; Zeid, T.A.; El-Baiomy, M.A. Clinicopathologic Effect of DNMT3A Mutation in Adult T-Cell Acute Lymphoblastic Leukemia. Clin. Lymphoma Myeloma Leuk. 2016, 16, 43–48. [Google Scholar] [CrossRef]
  27. Walter, M.J.; Ding, L.; Shen, D.; Shao, J.; Grillot, M.; McLellan, M.; Fulton, R.; Schmidt, H.; Kalicki-Veizer, J.; O’Laughlin, M.; et al. Recurrent DNMT3A mutations in patients with myelodysplastic syndromes. Leukemia 2011, 25, 1153–1158. [Google Scholar] [CrossRef] [Green Version]
  28. Spencer, D.H.; Russler-Germain, D.A.; Ketkar, S.; Helton, N.M.; Lamprecht, T.L.; Fulton, R.S.; Fronick, C.C.; O’Laughlin, M.; Heath, S.E.; Shinawi, M.; et al. CpG Island Hypermethylation Mediated by DNMT3A Is a Consequence of AML Progression. Cell 2017, 168, 801–816.e13. [Google Scholar] [CrossRef] [Green Version]
  29. Wang, R.Q.; Chen, C.J.; Jing, Y.; Qin, J.Y.; Li, Y.; Chen, G.F.; Zhou, W.; Li, Y.H.; Wang, J.; Li, D.W.; et al. Characteristics and prognostic significance of genetic mutations in acute myeloid leukemia based on a targeted next-generation sequencing technique. Cancer Med. 2020, 9, 8457–8467. [Google Scholar] [CrossRef]
  30. Patel, J.P.; Gonen, M.; Figueroa, M.E.; Fernandez, H.; Sun, Z.; Racevskis, J.; van Vlierberghe, P.; Dolgalev, I.; Thomas, S.; Aminova, O.; et al. Prognostic relevance of integrated genetic profiling in acute myeloid leukemia. N. Engl. J. Med. 2012, 366, 1079–1089. [Google Scholar] [CrossRef]
  31. Kubuki, Y.; Yamaji, T.; Hidaka, T.; Kameda, T.; Shide, K.; Sekine, M.; Kamiunten, A.; Akizuki, K.; Shimoda, H.; Tahira, Y.; et al. TET2 mutation in diffuse large B-cell lymphoma. J. Clin. Exp. Hematop. 2017, 56, 145–149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Zhang, T.J.; Zhou, J.D.; Yang, D.Q.; Wang, Y.X.; Wen, X.M.; Guo, H.; Yang, L.; Lian, X.Y.; Lin, J.; Qian, J. TET2 expression is a potential prognostic and predictive biomarker in cytogenetically normal acute myeloid leukemia. J. Cell. Physiol. 2018, 233, 5838–5846. [Google Scholar] [CrossRef] [PubMed]
  33. Cimmino, L.; Dawlaty, M.M.; Ndiaye-Lobry, D.; Yap, Y.S.; Bakogianni, S.; Yu, Y.; Bhattacharyya, S.; Shaknovich, R.; Geng, H.; Lobry, C.; et al. TET1 is a tumor suppressor of hematopoietic malignancy. Nat. Immunol. 2015, 16, 653–662. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Wang, J.; Li, F.; Ma, Z.; Yu, M.; Guo, Q.; Huang, J.; Yu, W.; Wang, Y.; Jin, J. High Expression of TET1 Predicts Poor Survival in Cytogenetically Normal Acute Myeloid Leukemia From Two Cohorts. eBioMedicine 2018, 28, 90–96. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Figueroa, M.E.; Abdel-Wahab, O.; Lu, C.; Ward, P.S.; Patel, J.; Shih, A.; Li, Y.; Bhagwat, N.; Vasanthakumar, A.; Fernandez, H.F.; et al. Leukemic IDH1 and IDH2 mutations result in a hypermethylation phenotype, disrupt TET2 function, and impair hematopoietic differentiation. Cancer Cell 2010, 18, 553–567. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Simonin, M.; Schmidt, A.; Bontoux, C.; Dourthe, M.E.; Lengline, E.; Andrieu, G.P.; Lhermitte, L.; Graux, C.; Grardel, N.; Cayuela, J.M.; et al. Oncogenetic landscape and clinical impact of IDH1 and IDH2 mutations in T-ALL. J. Hematol. Oncol. 2021, 14, 74. [Google Scholar] [CrossRef]
  37. Chen, X.; Zhou, W.; Song, R.H.; Liu, S.; Wang, S.; Chen, Y.; Gao, C.; He, C.; Xiao, J.; Zhang, L.; et al. Tumor suppressor CEBPA interacts with and inhibits DNMT3A activity. Sci. Adv. 2022, 8, eabl5220. [Google Scholar] [CrossRef]
  38. Gao, H.; He, X.; Li, Q.; Wang, Y.; Tian, Y.; Chen, X.; Wang, J.; Guo, Y.; Wang, W.; Li, X. Genome-wide DNA methylome analysis reveals methylation subtypes with different clinical outcomes for acute myeloid leukemia patients. Cancer Med. 2020, 9, 6296–6305. [Google Scholar] [CrossRef]
  39. Heller, G.; Topakian, T.; Altenberger, C.; Cerny-Reiterer, S.; Herndlhofer, S.; Ziegler, B.; Datlinger, P.; Byrgazov, K.; Bock, C.; Mannhalter, C.; et al. Next-generation sequencing identifies major DNA methylation changes during progression of Ph+ chronic myeloid leukemia. Leukemia 2016, 30, 1861–1868. [Google Scholar] [CrossRef] [Green Version]
  40. Pei, L.; Choi, J.H.; Liu, J.; Lee, E.J.; McCarthy, B.; Wilson, J.M.; Speir, E.; Awan, F.; Tae, H.; Arthur, G.; et al. Genome-wide DNA methylation analysis reveals novel epigenetic changes in chronic lymphocytic leukemia. Epigenetics 2012, 7, 567–578. [Google Scholar] [CrossRef]
  41. Hetzel, S.; Mattei, A.L.; Kretzmer, H.; Qu, C.; Chen, X.; Fan, Y.; Wu, G.; Roberts, K.G.; Luger, S.; Litzow, M.; et al. Acute lymphoblastic leukemia displays a distinct highly methylated genome. Nat. Cancer 2022, 3, 768–782. [Google Scholar] [CrossRef] [PubMed]
  42. Haider, Z.; Landfors, M.; Golovleva, I.; Erlanson, M.; Schmiegelow, K.; Flaegstad, T.; Kanerva, J.; Noren-Nystrom, U.; Hultdin, M.; Degerman, S. DNA methylation and copy number variation profiling of T-cell lymphoblastic leukemia and lymphoma. Blood Cancer J. 2020, 10, 45. [Google Scholar] [CrossRef]
  43. Liu, P.; Jiang, W.; Zhao, J.; Zhang, H. Integrated analysis of genome-wide gene expression and DNA methylation microarray of diffuse large B-cell lymphoma with TET mutations. Mol. Med. Rep. 2017, 16, 3777–3782. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Roux, B.; Picou, F.; Debeissat, C.; Koubi, M.; Gallay, N.; Hirsch, P.; Ravalet, N.; Bene, M.C.; Maigre, M.; Hunault, M.; et al. Aberrant DNA methylation impacts HOX genes expression in bone marrow mesenchymal stromal cells of myelodysplastic syndromes and de novo acute myeloid leukemia. Cancer Gene Ther. 2022, 29, 1263–1275. [Google Scholar] [CrossRef] [PubMed]
  45. Sampath, S.; Misra, P.; Yadav, S.K.; Sharma, S.; Somasundaram, V. A study on DNA methylation status in promoter region of p15 gene in patients of acute myeloid leukemia and myelodysplastic syndrome. Med. J. Armed Forces India 2021, 77, 337–342. [Google Scholar] [CrossRef] [PubMed]
  46. Fu, H.Y.; Wu, D.S.; Zhou, H.R.; Shen, J.Z. CpG island methylator phenotype and its relationship with prognosis in adult acute leukemia patients. Hematology 2014, 19, 329–337. [Google Scholar] [CrossRef]
  47. Kurtovic, N.K.; Krajnovic, M.; Bogdanovic, A.; Suvajdzic, N.; Jovanovic, J.; Dimitrijevic, B.; Colovic, M.; Krtolica, K. Concomitant aberrant methylation of p15 and MGMT genes in acute myeloid leukemia: Association with a particular immunophenotype of blast cells. Med. Oncol. 2012, 29, 3547–3556. [Google Scholar] [CrossRef]
  48. Hiller, J.K.; Schmoor, C.; Gaidzik, V.I.; Schmidt-Salzmann, C.; Yalcin, A.; Abdelkarim, M.; Blagitko-Dorfs, N.; Dohner, K.; Bullinger, L.; Duyster, J.; et al. Evaluating the impact of genetic and epigenetic aberrations on survival and response in acute myeloid leukemia patients receiving epigenetic therapy. Ann. Hematol. 2017, 96, 559–565. [Google Scholar] [CrossRef]
  49. Lian, X.Y.; Ma, J.C.; Zhou, J.D.; Zhang, T.J.; Wu, D.H.; Deng, Z.Q.; Zhang, Z.H.; Li, X.X.; He, P.F.; Yan, Y.; et al. Hypermethylation of ITGBL1 is associated with poor prognosis in acute myeloid leukemia. J. Cell. Physiol. 2019, 234, 9438–9446. [Google Scholar] [CrossRef]
  50. Zhang, Z.H.; Zhang, W.; Zhou, J.D.; Zhang, T.J.; Ma, J.C.; Xu, Z.J.; Lian, X.Y.; Wu, D.H.; Wen, X.M.; Deng, Z.Q.; et al. Decreased SCIN expression, associated with promoter methylation, is a valuable predictor for prognosis in acute myeloid leukemia. Mol. Carcinog. 2018, 57, 735–744. [Google Scholar] [CrossRef]
  51. Zhang, T.J.; Xu, Z.J.; Gu, Y.; Wen, X.M.; Ma, J.C.; Zhang, W.; Deng, Z.Q.; Leng, J.Y.; Qian, J.; Lin, J.; et al. Identification and validation of prognosis-related DLX5 methylation as an epigenetic driver in myeloid neoplasms. Clin. Transl. Med. 2020, 10, e29. [Google Scholar] [CrossRef] [PubMed]
  52. Park, S.; So, M.K.; Huh, J. Methylation of DNA Repair Genes as a Prognostic Biomarker in AML of a TCGA-LAML Cohort. Clin. Lab. 2022, 68, 7. [Google Scholar] [CrossRef] [PubMed]
  53. Chaubey, R.; Sazawal, S.; Mahapatra, M.; Chhikara, S.; Saxena, R. Prognostic relevance of aberrant SOCS-1 gene promoter methylation in myelodysplastic syndromes patients. Int. J. Lab. Hematol. 2015, 37, 265–271. [Google Scholar] [CrossRef] [PubMed]
  54. He, P.F.; Xu, Z.J.; Zhou, J.D.; Li, X.X.; Zhang, W.; Wu, D.H.; Zhang, Z.H.; Lian, X.Y.; Yao, X.Y.; Deng, Z.Q.; et al. Methylation-associated DOK1 and DOK2 down-regulation: Potential biomarkers for predicting adverse prognosis in acute myeloid leukemia. J. Cell. Physiol. 2018, 233, 6604–6614. [Google Scholar] [CrossRef] [PubMed]
  55. Sun, G.K.; Tang, L.J.; Zhou, J.D.; Xu, Z.J.; Yang, L.; Yuan, Q.; Ma, J.C.; Liu, X.H.; Lin, J.; Qian, J.; et al. DOK6 promoter methylation serves as a potential biomarker affecting prognosis in de novo acute myeloid leukemia. Cancer Med. 2019, 8, 6393–6402. [Google Scholar] [CrossRef] [Green Version]
  56. Bozic, T.; Lin, Q.; Frobel, J.; Wilop, S.; Hoffmann, M.; Muller-Tidow, C.; Brummendorf, T.H.; Jost, E.; Wagner, W. DNA-methylation in C1R is a prognostic biomarker for acute myeloid leukemia. Clin. Epigenet. 2015, 7, 116. [Google Scholar] [CrossRef] [Green Version]
  57. Sestakova, S.; Cerovska, E.; Salek, C.; Kundrat, D.; Jeziskova, I.; Folta, A.; Mayer, J.; Racil, Z.; Cetkovsky, P.; Remesova, H. A validation study of potential prognostic DNA methylation biomarkers in patients with acute myeloid leukemia using a custom DNA methylation sequencing panel. Clin. Epigenet. 2022, 14, 22. [Google Scholar] [CrossRef]
  58. Zhou, J.D.; Yao, D.M.; Zhang, Y.Y.; Ma, J.C.; Wen, X.M.; Yang, J.; Guo, H.; Chen, Q.; Lin, J.; Qian, J. GPX3 hypermethylation serves as an independent prognostic biomarker in non-M3 acute myeloid leukemia. Am. J. Cancer Res. 2015, 5, 1786–1794. [Google Scholar]
  59. Zhou, J.D.; Zhang, T.J.; Wang, Y.X.; Yang, D.Q.; Yang, L.; Ma, J.C.; Wen, X.M.; Yang, J.; Lin, J.; Qian, J. DLX4 hypermethylation is a prognostically adverse indicator in de novo acute myeloid leukemia. Tumour Biol. 2016, 37, 8951–8960. [Google Scholar] [CrossRef]
  60. Li, J.; Jin, W.; Tan, Y.; Wang, B.; Wang, X.; Zhao, M.; Wang, K. Distinct gene expression pattern of RUNX1 mutations coordinated by target repression and promoter hypermethylation in acute myeloid leukemia. Front. Med. 2022, 16, 627–636. [Google Scholar] [CrossRef]
  61. Man, C.H.; Fung, T.K.; Wan, H.; Cher, C.Y.; Fan, A.; Ng, N.; Ho, C.; Wan, T.S.; Tanaka, T.; So, C.W.; et al. Suppression of SOX7 by DNA methylation and its tumor suppressor function in acute myeloid leukemia. Blood 2015, 125, 3928–3936. [Google Scholar] [CrossRef] [PubMed]
  62. Gu, Y.; Zhou, J.D.; Xu, Z.J.; Zhang, T.J.; Wen, X.M.; Ma, J.C.; Ji, R.B.; Yuan, Q.; Zhang, W.; Chen, Q.; et al. Promoter methylation of the candidate tumor suppressor gene TCF21 in myelodysplastic syndrome and acute myeloid leukemia. Am. J. Transl. Res. 2019, 11, 3450–3460. [Google Scholar] [PubMed]
  63. Chen, X.X.; Lin, J.; Qian, J.; Qian, W.; Yang, J.; Ma, J.C.; Deng, Z.Q.; An, C.; Tang, C.Y.; Qian, Z.; et al. Methylation of CTNNA1 promoter: Frequent but not an adverse prognostic factor in acute myeloid leukemia. Leuk. Res. 2014, 38, 613–618. [Google Scholar] [CrossRef] [PubMed]
  64. Chen, W.W.; Liu, D.B.; Xiao, H.X.; Zhou, L.J.; Qu, J. Identification of differentially expressed genes induced by aberrant methylation in acute myeloid leukemia using integrated bioinformatics analyses. Oncol. Lett. 2022, 24, 383. [Google Scholar] [CrossRef]
  65. Lin, J.; Chen, Q.; Yang, J.; Qian, J.; Deng, Z.Q.; Qian, W.; Chen, X.X.; Ma, J.C.; Xiong, D.S.; Ma, Y.J.; et al. DDX43 promoter is frequently hypomethylated and may predict a favorable outcome in acute myeloid leukemia. Leuk. Res. 2014, 38, 601–607. [Google Scholar] [CrossRef]
  66. Qu, X.; Othus, M.; Davison, J.; Wu, Y.; Yan, L.; Meshinchi, S.; Ostronoff, F.; Estey, E.H.; Radich, J.P.; Erba, H.P.; et al. Prognostic methylation markers for overall survival in cytogenetically normal patients with acute myeloid leukemia treated on SWOG trials. Cancer 2017, 123, 2472–2481. [Google Scholar] [CrossRef] [Green Version]
  67. Minciacchi, V.R.; Kumar, R.; Krause, D.S. Chronic Myeloid Leukemia: A Model Disease of the Past, Present and Future. Cells 2021, 10, 117. [Google Scholar] [CrossRef]
  68. Jelinek, J.; Gharibyan, V.; Estecio, M.R.; Kondo, K.; He, R.; Chung, W.; Lu, Y.; Zhang, N.; Liang, S.; Kantarjian, H.M.; et al. Aberrant DNA methylation is associated with disease progression, resistance to imatinib and shortened survival in chronic myelogenous leukemia. PLoS ONE 2011, 6, e22110. [Google Scholar] [CrossRef]
  69. Maupetit-Mehouas, S.; Court, F.; Bourgne, C.; Guerci-Bresler, A.; Cony-Makhoul, P.; Johnson, H.; Etienne, G.; Rousselot, P.; Guyotat, D.; Janel, A.; et al. DNA methylation profiling reveals a pathological signature that contributes to transcriptional defects of CD34+ CD15 cells in early chronic-phase chronic myeloid leukemia. Mol. Oncol. 2018, 12, 814–829. [Google Scholar] [CrossRef] [Green Version]
  70. Chen, Q.; Lin, J.; Qian, J.; Deng, Z.Q.; Qian, W.; Yang, J.; Li, Y.; Chen, X.X.; Ma, Y.J.; Ma, J.C.; et al. The methylation status of the DDX43 promoter in Chinese patients with chronic myeloid leukemia. Genet. Test. Mol. Biomarkers 2013, 17, 508–511. [Google Scholar] [CrossRef]
  71. Elias, M.H.; Azlan, H.; Sulong, S.; Baba, A.A.; Ankathil, R. Aberrant DNA methylation at HOXA4 and HOXA5 genes are associated with resistance to imatinib mesylate among chronic myeloid leukemia patients. Cancer Rep. 2018, 1, e1111. [Google Scholar] [CrossRef] [PubMed]
  72. Song, J.J.; Liu, Q.; Li, Y.; Yang, Z.S.; Yang, L.; Xiang, T.X.; Ren, G.S.; Chen, J.B. Epigenetic inactivation of PLCD1 in chronic myeloid leukemia. Int. J. Mol. Med. 2012, 30, 179–184. [Google Scholar] [CrossRef] [PubMed]
  73. Zhou, J.D.; Wang, Y.X.; Zhang, T.J.; Yang, D.Q.; Yao, D.M.; Guo, H.; Yang, L.; Ma, J.C.; Wen, X.M.; Yang, J.; et al. Epigenetic inactivation of DLX4 is associated with disease progression in chronic myeloid leukemia. Biochem. Biophys. Res. Commun. 2015, 463, 1250–1256. [Google Scholar] [CrossRef]
  74. Wang, Y.L.; Qian, J.; Lin, J.; Yao, D.M.; Qian, Z.; Zhu, Z.H.; Li, J.Y. Methylation status of DDIT3 gene in chronic myeloid leukemia. J. Exp. Clin. Cancer Res. 2010, 29, 54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Yang, H.; Liang, H.; Yan, J.S.; Tao, R.; Hao, S.G.; Ma, L.Y. Down-regulation of hematopoiesis master regulator PU.1 via aberrant methylation in chronic myeloid leukemia. Int. J. Hematol. 2012, 96, 65–73. [Google Scholar] [CrossRef]
  76. Qian, J.; Wang, Y.L.; Lin, J.; Yao, D.M.; Xu, W.R.; Wu, C.Y. Aberrant methylation of the death-associated protein kinase 1 (DAPK1) CpG island in chronic myeloid leukemia. Eur. J. Haematol. 2009, 82, 119–123. [Google Scholar] [CrossRef]
  77. San Jose-Eneriz, E.; Agirre, X.; Jimenez-Velasco, A.; Cordeu, L.; Martin, V.; Arqueros, V.; Garate, L.; Fresquet, V.; Cervantes, F.; Martinez-Climent, J.A.; et al. Epigenetic down-regulation of BIM expression is associated with reduced optimal responses to imatinib treatment in chronic myeloid leukaemia. Eur. J. Cancer 2009, 45, 1877–1889. [Google Scholar] [CrossRef]
  78. Yao, D.M.; Zhou, J.D.; Zhang, Y.Y.; Yang, L.; Wen, X.M.; Yang, J.; Guo, H.; Chen, Q.; Lin, J.; Qian, J. GPX3 promoter is methylated in chronic myeloid leukemia. Int. J. Clin. Exp. Pathol. 2015, 8, 6450–6457. [Google Scholar]
  79. Li, Y.; Yang, L.; Pan, Y.; Yang, J.; Shang, Y.; Luo, J. Methylation and decreased expression of SHP-1 are related to disease progression in chronic myelogenous leukemia. Oncol. Rep. 2014, 31, 2438–2446. [Google Scholar] [CrossRef] [Green Version]
  80. Wenzinger, C.; Williams, E.; Gru, A.A. Updates in the Pathology of Precursor Lymphoid Neoplasms in the Revised Fourth Edition of the WHO Classification of Tumors of Hematopoietic and Lymphoid Tissues. Curr. Hematol. Malig. Rep. 2018, 13, 275–288. [Google Scholar] [CrossRef]
  81. Touzart, A.; Boissel, N.; Belhocine, M.; Smith, C.; Graux, C.; Latiri, M.; Lhermitte, L.; Mathieu, E.L.; Huguet, F.; Lamant, L.; et al. Low level CpG island promoter methylation predicts a poor outcome in adult T-cell acute lymphoblastic leukemia. Haematologica 2020, 105, 1575–1581. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Jang, W.; Park, J.; Kwon, A.; Choi, H.; Kim, J.; Lee, G.D.; Han, E.; Jekarl, D.W.; Chae, H.; Han, K.; et al. CDKN2B downregulation and other genetic characteristics in T-acute lymphoblastic leukemia. Exp. Mol. Med. 2019, 51, 1–15. [Google Scholar] [CrossRef] [PubMed]
  83. Dell’Orto, M.C.; Banelli, B.; Giarin, E.; Accordi, B.; Trentin, L.; Romani, M.; Te Kronnie, G.; Basso, G. Down-regulation of DLX3 expression in MLL-AF4 childhood lymphoblastic leukemias is mediated by promoter region hypermethylation. Oncol. Rep. 2007, 18, 417–423. [Google Scholar]
  84. Burmeister, D.W.; Smith, E.H.; Cristel, R.T.; McKay, S.D.; Shi, H.; Arthur, G.L.; Davis, J.W.; Taylor, K.H. The expression of RUNDC3B is associated with promoter methylation in lymphoid malignancies. Hematol. Oncol. 2017, 35, 25–33. [Google Scholar] [CrossRef]
  85. Chatterton, Z.; Burke, D.; Emslie, K.R.; Craig, J.M.; Ng, J.; Ashley, D.M.; Mechinaud, F.; Saffery, R.; Wong, N.C. Validation of DNA methylation biomarkers for diagnosis of acute lymphoblastic leukemia. Clin. Chem. 2014, 60, 995–1003. [Google Scholar] [CrossRef] [Green Version]
  86. Younesian, S.; Shahkarami, S.; Ghaffari, P.; Alizadeh, S.; Mehrasa, R.; Ghavamzadeh, A.; Ghaffari, S.H. DNA hypermethylation of tumor suppressor genes RASSF6 and RASSF10 as independent prognostic factors in adult acute lymphoblastic leukemia. Leuk. Res. 2017, 61, 33–38. [Google Scholar] [CrossRef]
  87. Younesian, S.; Shahkarami, S.; Ghaffari, P.; Alizadeh, S.; Mehrasa, R.; Ghaffari, S.H. Residual methylation of tumor suppressor gene promoters, RASSF6 and RASSF10, as novel biomarkers for minimal residual disease detection in adult acute lymphoblastic leukemia. Ann. Hematol. 2019, 98, 2719–2727. [Google Scholar] [CrossRef]
  88. Roman-Gomez, J.; Castillejo, J.A.; Jimenez, A.; Gonzalez, M.G.; Moreno, F.; Rodriguez, M.C.; Barrios, M.; Maldonado, J.; Torres, A. 5’ CpG island hypermethylation is associated with transcriptional silencing of the p21(CIP1/WAF1/SDI1) gene and confers poor prognosis in acute lymphoblastic leukemia. Blood 2002, 99, 2291–2296. [Google Scholar] [CrossRef]
  89. Roman-Gomez, J.; Jimenez-Velasco, A.; Cordeu, L.; Vilas-Zornoza, A.; San Jose-Eneriz, E.; Garate, L.; Castillejo, J.A.; Martin, V.; Prosper, F.; Heiniger, A.; et al. WNT5A, a putative tumour suppressor of lymphoid malignancies, is inactivated by aberrant methylation in acute lymphoblastic leukaemia. Eur. J. Cancer 2007, 43, 2736–2746. [Google Scholar] [CrossRef]
  90. Roman-Gomez, J.; Jimenez-Velasco, A.; Agirre, X.; Castillejo, J.A.; Navarro, G.; Barrios, M.; Andreu, E.J.; Prosper, F.; Heiniger, A.; Torres, A. Transcriptional silencing of the Dickkopfs-3 (Dkk-3) gene by CpG hypermethylation in acute lymphoblastic leukaemia. Br. J. Cancer 2004, 91, 707–713. [Google Scholar] [CrossRef]
  91. Roman-Gomez, J.; Jimenez-Velasco, A.; Agirre, X.; Castillejo, J.A.; Barrios, M.; Andreu, E.J.; Prosper, F.; Heiniger, A.; Torres, A. The normal epithelial cell-specific 1 (NES1) gene, a candidate tumor suppressor gene on chromosome 19q13.3-4, is downregulated by hypermethylation in acute lymphoblastic leukemia. Leukemia 2004, 18, 362–365. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Narayan, G.; Freddy, A.J.; Xie, D.; Liyanage, H.; Clark, L.; Kisselev, S.; Un Kang, J.; Nandula, S.V.; McGuinn, C.; Subramaniyam, S.; et al. Promoter methylation-mediated inactivation of PCDH10 in acute lymphoblastic leukemia contributes to chemotherapy resistance. Genes Chromosomes Cancer 2011, 50, 1043–1053. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Uyen, T.N.; Sakashita, K.; Al-Kzayer, L.F.; Nakazawa, Y.; Kurata, T.; Koike, K. Aberrant methylation of protocadherin 17 and its prognostic value in pediatric acute lymphoblastic leukemia. Pediatr. Blood Cancer 2017, 64, e26259. [Google Scholar] [CrossRef] [PubMed]
  94. Thathia, S.H.; Ferguson, S.; Gautrey, H.E.; van Otterdijk, S.D.; Hili, M.; Rand, V.; Moorman, A.V.; Meyer, S.; Brown, R.; Strathdee, G. Epigenetic inactivation of TWIST2 in acute lymphoblastic leukemia modulates proliferation, cell survival and chemosensitivity. Haematologica 2012, 97, 371–378. [Google Scholar] [CrossRef] [PubMed]
  95. Abdullah, M.; Choo, C.W.; Alias, H.; Rahman, E.J.A.; Ibrahim, H.M.; Jamal, R.; Hussin, N.H. ADAMTSL5 and CDH11: Putative epigenetic markers for therapeutic resistance in acute lymphoblastic leukemia. Hematology 2017, 22, 386–391. [Google Scholar] [CrossRef] [Green Version]
  96. Akahane, K.; Kimura, S.; Miyake, K.; Watanabe, A.; Kagami, K.; Yoshimura, K.; Shinohara, T.; Harama, D.; Kasai, S.; Goi, K.; et al. Association of allele-specific methylation of the ASNS gene with asparaginase sensitivity and prognosis in T-ALL. Blood Adv. 2022, 6, 212–224. [Google Scholar] [CrossRef]
  97. Hallek, M.; Shanafelt, T.D.; Eichhorst, B. Chronic lymphocytic leukaemia. Lancet 2018, 391, 1524–1537. [Google Scholar] [CrossRef]
  98. Cahill, N.; Rosenquist, R. Uncovering the DNA methylome in chronic lymphocytic leukemia. Epigenetics 2013, 8, 138–148. [Google Scholar] [CrossRef] [Green Version]
  99. Raval, A.; Lucas, D.M.; Matkovic, J.J.; Bennett, K.L.; Liyanarachchi, S.; Young, D.C.; Rassenti, L.; Kipps, T.J.; Grever, M.R.; Byrd, J.C.; et al. TWIST2 demonstrates differential methylation in immunoglobulin variable heavy chain mutated and unmutated chronic lymphocytic leukemia. J. Clin. Oncol. 2005, 23, 3877–3885. [Google Scholar] [CrossRef]
  100. Raval, A.; Tanner, S.M.; Byrd, J.C.; Angerman, E.B.; Perko, J.D.; Chen, S.S.; Hackanson, B.; Grever, M.R.; Lucas, D.M.; Matkovic, J.J.; et al. Downregulation of death-associated protein kinase 1 (DAPK1) in chronic lymphocytic leukemia. Cell 2007, 129, 879–890. [Google Scholar] [CrossRef] [Green Version]
  101. Dunwell, T.L.; Dickinson, R.E.; Stankovic, T.; Dallol, A.; Weston, V.; Austen, B.; Catchpoole, D.; Maher, E.R.; Latif, F. Frequent epigenetic inactivation of the SLIT2 gene in chronic and acute lymphocytic leukemia. Epigenetics 2009, 4, 265–269. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Claus, R.; Lucas, D.M.; Stilgenbauer, S.; Ruppert, A.S.; Yu, L.; Zucknick, M.; Mertens, D.; Buhler, A.; Oakes, C.C.; Larson, R.A.; et al. Quantitative DNA methylation analysis identifies a single CpG dinucleotide important for ZAP-70 expression and predictive of prognosis in chronic lymphocytic leukemia. J. Clin. Oncol. 2012, 30, 2483–2491. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Chim, C.S.; Pang, R.; Liang, R. Epigenetic dysregulation of the Wnt signalling pathway in chronic lymphocytic leukaemia. J. Clin. Pathol. 2008, 61, 1214–1219. [Google Scholar] [CrossRef] [PubMed]
  104. Bennett, L.B.; Taylor, K.H.; Arthur, G.L.; Rahmatpanah, F.B.; Hooshmand, S.I.; Caldwell, C.W. Epigenetic regulation of WNT signaling in chronic lymphocytic leukemia. Epigenomics 2010, 2, 53–70. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Martinelli, S.; Maffei, R.; Fiorcari, S.; Quadrelli, C.; Zucchini, P.; Benatti, S.; Potenza, L.; Luppi, M.; Marasca, R. The expression of endothelin-1 in chronic lymphocytic leukemia is controlled by epigenetic mechanisms and extracellular stimuli. Leuk. Res. 2017, 54, 17–24. [Google Scholar] [CrossRef] [PubMed]
  106. Martinelli, S.; Kanduri, M.; Maffei, R.; Fiorcari, S.; Bulgarelli, J.; Marasca, R.; Rosenquist, R. ANGPT2 promoter methylation is strongly associated with gene expression and prognosis in chronic lymphocytic leukemia. Epigenetics 2013, 8, 720–729. [Google Scholar] [CrossRef] [Green Version]
  107. Rani, L.; Mathur, N.; Gupta, R.; Gogia, A.; Kaur, G.; Dhanjal, J.K.; Sundar, D.; Kumar, L.; Sharma, A. Genome-wide DNA methylation profiling integrated with gene expression profiling identifies PAX9 as a novel prognostic marker in chronic lymphocytic leukemia. Clin. Epigenet. 2017, 9, 57. [Google Scholar] [CrossRef] [Green Version]
  108. Pan, H.; Renaud, L.; Chaligne, R.; Bloehdorn, J.; Tausch, E.; Mertens, D.; Fink, A.M.; Fischer, K.; Zhang, C.; Betel, D.; et al. Discovery of Candidate DNA Methylation Cancer Driver Genes. Cancer Discov. 2021, 11, 2266–2281. [Google Scholar] [CrossRef]
  109. Hanoun, M.; Eisele, L.; Suzuki, M.; Greally, J.M.; Huttmann, A.; Aydin, S.; Scholtysik, R.; Klein-Hitpass, L.; Duhrsen, U.; Durig, J. Epigenetic silencing of the circadian clock gene CRY1 is associated with an indolent clinical course in chronic lymphocytic leukemia. PLoS ONE 2012, 7, e34347. [Google Scholar] [CrossRef] [Green Version]
  110. Loi, E.; Moi, L.; Fadda, A.; Satta, G.; Zucca, M.; Sanna, S.; Amini Nia, S.; Cabras, G.; Padoan, M.; Magnani, C.; et al. Methylation alteration of SHANK1 as a predictive, diagnostic and prognostic biomarker for chronic lymphocytic leukemia. Oncotarget 2019, 10, 4987–5002. [Google Scholar] [CrossRef] [Green Version]
  111. Wolf, C.; Garding, A.; Filarsky, K.; Bahlo, J.; Robrecht, S.; Becker, N.; Zucknick, M.; Rouhi, A.; Weigel, A.; Claus, R.; et al. NFATC1 activation by DNA hypomethylation in chronic lymphocytic leukemia correlates with clinical staging and can be inhibited by ibrutinib. Int. J. Cancer 2018, 142, 322–333. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Fabbri, G.; Dalla-Favera, R. The molecular pathogenesis of chronic lymphocytic leukaemia. Nat. Rev. Cancer 2016, 16, 145–162. [Google Scholar] [CrossRef] [PubMed]
  113. Kanduri, M.; Cahill, N.; Goransson, H.; Enstrom, C.; Ryan, F.; Isaksson, A.; Rosenquist, R. Differential genome-wide array-based methylation profiles in prognostic subsets of chronic lymphocytic leukemia. Blood 2010, 115, 296–305. [Google Scholar] [CrossRef] [PubMed]
  114. Poppova, L.; Pavlova, S.; Gonzalez, B.; Kotaskova, J.; Plevova, K.; Dumbovic, G.; Janovska, P.; Bystry, V.; Panovska, A.; Bezdekova, L.; et al. Memory B-cell like chronic lymphocytic leukaemia is associated with specific methylation profile of WNT5A promoter and undetectable expression of WNT5A gene. Epigenetics 2022, 17, 1628–1635. [Google Scholar] [CrossRef]
  115. Rodriguez-Abreu, D.; Bordoni, A.; Zucca, E. Epidemiology of hematological malignancies. Ann. Oncol. 2007, 18 (Suppl. S1), i3–i8. [Google Scholar] [CrossRef]
  116. Murray, P.G.; Qiu, G.H.; Fu, L.; Waites, E.R.; Srivastava, G.; Heys, D.; Agathanggelou, A.; Latif, F.; Grundy, R.G.; Mann, J.R.; et al. Frequent epigenetic inactivation of the RASSF1A tumor suppressor gene in Hodgkin’s lymphoma. Oncogene 2004, 23, 1326–1331. [Google Scholar] [CrossRef] [Green Version]
  117. Garcia, M.J.; Martinez-Delgado, B.; Cebrian, A.; Martinez, A.; Benitez, J.; Rivas, C. Different incidence and pattern of p15INK4b and p16INK4a promoter region hypermethylation in Hodgkin’s and CD30-Positive non-Hodgkin’s lymphomas. Am. J. Pathol. 2002, 161, 1007–1013. [Google Scholar] [CrossRef]
  118. Ushmorov, A.; Leithauser, F.; Sakk, O.; Weinhausel, A.; Popov, S.W.; Moller, P.; Wirth, T. Epigenetic processes play a major role in B-cell-specific gene silencing in classical Hodgkin lymphoma. Blood 2006, 107, 2493–2500. [Google Scholar] [CrossRef] [Green Version]
  119. Xia, C.; Olsen, T.K.; Zirakzadeh, A.A.; Almamoun, R.; Sjoholm, L.K.; Dahlstrom, J.; Sjoberg, J.; Claesson, H.E.; Johnsen, J.I.; Winqvist, O.; et al. Hodgkin Lymphoma Monozygotic Triplets Reveal Divergences in DNA Methylation Signatures. Front. Oncol. 2020, 10, 598872. [Google Scholar] [CrossRef]
  120. Bethge, N.; Honne, H.; Hilden, V.; Troen, G.; Eknaes, M.; Liestol, K.; Holte, H.; Delabie, J.; Smeland, E.B.; Lind, G.E. Identification of highly methylated genes across various types of B-cell non-hodgkin lymphoma. PLoS ONE 2013, 8, e79602. [Google Scholar] [CrossRef] [Green Version]
  121. Bethge, N.; Honne, H.; Andresen, K.; Hilden, V.; Troen, G.; Liestol, K.; Holte, H.; Delabie, J.; Lind, G.E.; Smeland, E.B. A gene panel, including LRP12, is frequently hypermethylated in major types of B-cell lymphoma. PLoS ONE 2014, 9, e104249. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Frazzi, R.; Zanetti, E.; Pistoni, M.; Tamagnini, I.; Valli, R.; Braglia, L.; Merli, F. Methylation changes of SIRT1, KLF4, DAPK1 and SPG20 in B-lymphocytes derived from follicular and diffuse large B-cell lymphoma. Leuk. Res. 2017, 57, 89–96. [Google Scholar] [CrossRef] [PubMed]
  123. Frazzi, R.; Cusenza, V.Y.; Pistoni, M.; Canovi, L.; Cascione, L.; Bertoni, F.; Merli, F. KLF4, DAPK1 and SPG20 promoter methylation is not affected by DNMT1 silencing and hypomethylating drugs in lymphoma cells. Oncol. Rep. 2022, 47, 10. [Google Scholar] [CrossRef] [PubMed]
  124. Chambwe, N.; Kormaksson, M.; Geng, H.; De, S.; Michor, F.; Johnson, N.A.; Morin, R.D.; Scott, D.W.; Godley, L.A.; Gascoyne, R.D.; et al. Variability in DNA methylation defines novel epigenetic subgroups of DLBCL associated with different clinical outcomes. Blood 2014, 123, 1699–1708. [Google Scholar] [CrossRef] [Green Version]
  125. Zainuddin, N.; Kanduri, M.; Berglund, M.; Lindell, M.; Amini, R.M.; Roos, G.; Sundstrom, C.; Enblad, G.; Rosenquist, R. Quantitative evaluation of p16(INK4a) promoter methylation using pyrosequencing in de novo diffuse large B-cell lymphoma. Leuk. Res. 2011, 35, 438–443. [Google Scholar] [CrossRef]
  126. Hagiwara, K.; Li, Y.; Kinoshita, T.; Kunishma, S.; Ohashi, H.; Hotta, T.; Nagai, H. Aberrant DNA methylation of the p57KIP2 gene is a sensitive biomarker for detecting minimal residual disease in diffuse large B cell lymphoma. Leuk. Res. 2010, 34, 50–54. [Google Scholar] [CrossRef]
  127. Shawky, S.A.; El-Borai, M.H.; Khaled, H.M.; Guda, I.; Mohanad, M.; Abdellateif, M.S.; Zekri, A.N.; Bahanasy, A.A. The prognostic impact of hypermethylation for a panel of tumor suppressor genes and cell of origin subtype on diffuse large B-cell lymphoma. Mol. Biol. Rep. 2019, 46, 4063–4076. [Google Scholar] [CrossRef]
  128. Cao, B.; Guo, X.; Huang, L.; Wang, B.; Wang, W.; Han, D.; Zhang, W.; Zhong, K. Methylation silencing CDH23 is a poor prognostic marker in diffuse large B-cell lymphoma. Aging 2021, 13, 17768–17788. [Google Scholar] [CrossRef]
  129. Huang, W.; Xue, X.; Shan, L.; Qiu, T.; Guo, L.; Ying, J.; Lu, N. Clinical significance of PCDH10 promoter methylation in diffuse large B-cell lymphoma. BMC Cancer 2017, 17, 815. [Google Scholar] [CrossRef] [Green Version]
  130. Kristensen, L.S.; Asmar, F.; Dimopoulos, K.; Nygaard, M.K.; Aslan, D.; Hansen, J.W.; Ralfkiaer, E.; Gronbaek, K. Hypermethylation of DAPK1 is an independent prognostic factor predicting survival in diffuse large B-cell lymphoma. Oncotarget 2014, 5, 9798–9810. [Google Scholar] [CrossRef] [Green Version]
  131. Wang, H.; Zhou, L.Y.; Guan, Z.B.; Zeng, W.B.; Zhou, L.L.; Liu, Y.N.; Pan, X.Y. Prognostic significance of DAPK promoter methylation in lymphoma: A meta-analysis. PLoS ONE 2019, 14, e0210943. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Kristensen, L.S.; Hansen, J.W.; Kristensen, S.S.; Tholstrup, D.; Harslof, L.B.; Pedersen, O.B.; De Nully Brown, P.; Gronbaek, K. Aberrant methylation of cell-free circulating DNA in plasma predicts poor outcome in diffuse large B cell lymphoma. Clin. Epigenet. 2016, 8, 95. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Mohamed, G.; Talima, S.; Li, L.; Wei, W.; Rudzki, Z.; Allam, R.M.; Simmons, W.; Tao, Q.; Murray, P.G. Low Expression and Promoter Hypermethylation of the Tumour Suppressor SLIT2, are Associated with Adverse Patient Outcomes in Diffuse Large B Cell Lymphoma. Pathol. Oncol. Res. 2019, 25, 1223–1231. [Google Scholar] [CrossRef] [PubMed]
  134. Schmid, C.A.; Robinson, M.D.; Scheifinger, N.A.; Muller, S.; Cogliatti, S.; Tzankov, A.; Muller, A. DUSP4 deficiency caused by promoter hypermethylation drives JNK signaling and tumor cell survival in diffuse large B cell lymphoma. J. Exp. Med. 2015, 212, 775–792. [Google Scholar] [CrossRef] [PubMed]
  135. Lee, S.M.; Lee, E.J.; Ko, Y.H.; Lee, S.H.; Maeng, L.; Kim, K.M. Prognostic significance of O6-methylguanine DNA methyltransferase and p57 methylation in patients with diffuse large B-cell lymphomas. APMIS 2009, 117, 87–94. [Google Scholar] [CrossRef] [PubMed]
  136. Clozel, T.; Yang, S.; Elstrom, R.L.; Tam, W.; Martin, P.; Kormaksson, M.; Banerjee, S.; Vasanthakumar, A.; Culjkovic, B.; Scott, D.W.; et al. Mechanism-based epigenetic chemosensitization therapy of diffuse large B-cell lymphoma. Cancer Discov. 2013, 3, 1002–1019. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Xu, P.; Hu, G.; Luo, C.; Liang, Z. DNA methyltransferase inhibitors: An updated patent review (2012-2015). Expert Opin. Ther. Pat. 2016, 26, 1017–1030. [Google Scholar] [CrossRef]
  138. Pappalardi, M.B.; Keenan, K.; Cockerill, M.; Kellner, W.A.; Stowell, A.; Sherk, C.; Wong, K.; Pathuri, S.; Briand, J.; Steidel, M.; et al. Discovery of a first-in-class reversible DNMT1-selective inhibitor with improved tolerability and efficacy in acute myeloid leukemia. Nat. Cancer 2021, 2, 1002–1017. [Google Scholar] [CrossRef]
  139. Nieto, M.; Demolis, P.; Behanzin, E.; Moreau, A.; Hudson, I.; Flores, B.; Stemplewski, H.; Salmonson, T.; Gisselbrecht, C.; Bowen, D.; et al. The European Medicines Agency Review of Decitabine (Dacogen) for the Treatment of Adult Patients With Acute Myeloid Leukemia: Summary of the Scientific Assessment of the Committee for Medicinal Products for Human Use. Oncologist 2016, 21, 692–700. [Google Scholar] [CrossRef] [Green Version]
  140. Dombret, H.; Seymour, J.F.; Butrym, A.; Wierzbowska, A.; Selleslag, D.; Jang, J.H.; Kumar, R.; Cavenagh, J.; Schuh, A.C.; Candoni, A.; et al. International phase 3 study of azacitidine vs conventional care regimens in older patients with newly diagnosed AML with >30% blasts. Blood 2015, 126, 291–299. [Google Scholar] [CrossRef] [Green Version]
  141. Wei, A.H.; Dohner, H.; Pocock, C.; Montesinos, P.; Afanasyev, B.; Dombret, H.; Ravandi, F.; Sayar, H.; Jang, J.H.; Porkka, K.; et al. Oral Azacitidine Maintenance Therapy for Acute Myeloid Leukemia in First Remission. N. Engl. J. Med. 2020, 383, 2526–2537. [Google Scholar] [CrossRef] [PubMed]
  142. Ravandi, F.; Roboz, G.J.; Wei, A.H.; Dohner, H.; Pocock, C.; Selleslag, D.; Montesinos, P.; Sayar, H.; Musso, M.; Figuera-Alvarez, A.; et al. Management of adverse events in patients with acute myeloid leukemia in remission receiving oral azacitidine: Experience from the phase 3 randomized QUAZAR AML-001 trial. J. Hematol. Oncol. 2021, 14, 133. [Google Scholar] [CrossRef] [PubMed]
  143. Savona, M.R.; Odenike, O.; Amrein, P.C.; Steensma, D.P.; DeZern, A.E.; Michaelis, L.C.; Faderl, S.; Harb, W.; Kantarjian, H.; Lowder, J.; et al. An oral fixed-dose combination of decitabine and cedazuridine in myelodysplastic syndromes: A multicentre, open-label, dose-escalation, phase 1 study. Lancet Haematol. 2019, 6, e194–e203. [Google Scholar] [CrossRef] [PubMed]
  144. Roboz, G.J.; Kantarjian, H.M.; Yee, K.W.L.; Kropf, P.L.; O’Connell, C.L.; Griffiths, E.A.; Stock, W.; Daver, N.G.; Jabbour, E.; Ritchie, E.K.; et al. Dose, schedule, safety, and efficacy of guadecitabine in relapsed or refractory acute myeloid leukemia. Cancer 2018, 124, 325–334. [Google Scholar] [CrossRef]
  145. Chung, W.; Kelly, A.D.; Kropf, P.; Fung, H.; Jelinek, J.; Su, X.Y.; Roboz, G.J.; Kantarjian, H.M.; Azab, M.; Issa, J.J. Genomic and epigenomic predictors of response to guadecitabine in relapsed/refractory acute myelogenous leukemia. Clin. Epigenet. 2019, 11, 106. [Google Scholar] [CrossRef]
  146. Garcia-Manero, G.; Roboz, G.; Walsh, K.; Kantarjian, H.; Ritchie, E.; Kropf, P.; O’Connell, C.; Tibes, R.; Lunin, S.; Rosenblat, T.; et al. Guadecitabine (SGI-110) in patients with intermediate or high-risk myelodysplastic syndromes: Phase 2 results from a multicentre, open-label, randomised, phase 1/2 trial. Lancet Haematol. 2019, 6, e317–e327. [Google Scholar] [CrossRef]
  147. Cahill, K.E.; Karimi, Y.H.; Karrison, T.G.; Jain, N.; Green, M.; Weiner, H.; Fulton, N.; Kadri, S.; Godley, L.A.; Artz, A.S.; et al. A phase 1 study of azacitidine with high-dose cytarabine and mitoxantrone in high-risk acute myeloid leukemia. Blood Adv. 2020, 4, 599–606. [Google Scholar] [CrossRef]
  148. Pommert, L.; Schafer, E.S.; Malvar, J.; Gossai, N.; Florendo, E.; Pulakanti, K.; Heimbruch, K.; Stelloh, C.; Chi, Y.Y.; Sposto, R.; et al. Decitabine and vorinostat with FLAG chemotherapy in pediatric relapsed/refractory AML: Report from the therapeutic advances in childhood leukemia and lymphoma (TACL) consortium. Am. J. Hematol. 2022, 97, 613–622. [Google Scholar] [CrossRef]
  149. DiNardo, C.D.; Jonas, B.A.; Pullarkat, V.; Thirman, M.J.; Garcia, J.S.; Wei, A.H.; Konopleva, M.; Dohner, H.; Letai, A.; Fenaux, P.; et al. Azacitidine and Venetoclax in Previously Untreated Acute Myeloid Leukemia. N. Engl. J. Med. 2020, 383, 617–629. [Google Scholar] [CrossRef]
  150. Cherry, E.M.; Abbott, D.; Amaya, M.; McMahon, C.; Schwartz, M.; Rosser, J.; Sato, A.; Schowinsky, J.; Inguva, A.; Minhajuddin, M.; et al. Venetoclax and azacitidine compared with induction chemotherapy for newly diagnosed patients with acute myeloid leukemia. Blood Adv. 2021, 5, 5565–5573. [Google Scholar] [CrossRef]
  151. Griffiths, E.A.; Srivastava, P.; Matsuzaki, J.; Brumberger, Z.; Wang, E.S.; Kocent, J.; Miller, A.; Roloff, G.W.; Wong, H.Y.; Paluch, B.E.; et al. NY-ESO-1 Vaccination in Combination with Decitabine Induces Antigen-Specific T-lymphocyte Responses in Patients with Myelodysplastic Syndrome. Clin. Cancer Res. 2018, 24, 1019–1029. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Holmberg-Thyden, S.; Dufva, I.H.; Gang, A.O.; Breinholt, M.F.; Schejbel, L.; Andersen, M.K.; Kadivar, M.; Svane, I.M.; Gronbaek, K.; Hadrup, S.R.; et al. Epigenetic therapy in combination with a multi-epitope cancer vaccine targeting shared tumor antigens for high-risk myelodysplastic syndrome—A phase I clinical trial. Cancer Immunol. Immunother. 2022, 71, 433–444. [Google Scholar] [CrossRef] [PubMed]
  153. Saxena, K.; Herbrich, S.M.; Pemmaraju, N.; Kadia, T.M.; DiNardo, C.D.; Borthakur, G.; Pierce, S.A.; Jabbour, E.; Wang, S.A.; Bueso-Ramos, C.; et al. A phase 1b/2 study of azacitidine with PD-L1 antibody avelumab in relapsed/refractory acute myeloid leukemia. Cancer 2021, 127, 3761–3771. [Google Scholar] [CrossRef] [PubMed]
  154. Daver, N.; Garcia-Manero, G.; Basu, S.; Boddu, P.C.; Alfayez, M.; Cortes, J.E.; Konopleva, M.; Ravandi-Kashani, F.; Jabbour, E.; Kadia, T.; et al. Efficacy, Safety, and Biomarkers of Response to Azacitidine and Nivolumab in Relapsed/Refractory Acute Myeloid Leukemia: A Nonrandomized, Open-Label, Phase II Study. Cancer Discov. 2019, 9, 370–383. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Zeidan, A.M.; Boss, I.; Beach, C.L.; Copeland, W.B.; Thompson, E.; Fox, B.A.; Hasle, V.E.; Ogasawara, K.; Cavenagh, J.; Silverman, L.R.; et al. A randomized phase 2 trial of azacitidine with or without durvalumab as first-line therapy for higher-risk myelodysplastic syndromes. Blood Adv. 2022, 6, 2207–2218. [Google Scholar] [CrossRef]
  156. Zeidan, A.M.; Boss, I.; Beach, C.L.; Copeland, W.B.; Thompson, E.; Fox, B.A.; Hasle, V.E.; Hellmann, A.; Taussig, D.C.; Tormo, M.; et al. A randomized phase 2 trial of azacitidine with or without durvalumab as first-line therapy for older patients with AML. Blood Adv. 2022, 6, 2219–2229. [Google Scholar] [CrossRef]
  157. Nieto, Y.; Valdez, B.C.; Thall, P.F.; Jones, R.B.; Wei, W.; Myers, A.; Hosing, C.; Ahmed, S.; Popat, U.; Shpall, E.J.; et al. Double epigenetic modulation of high-dose chemotherapy with azacitidine and vorinostat for patients with refractory or poor-risk relapsed lymphoma. Cancer 2016, 122, 2680–2688. [Google Scholar] [CrossRef] [Green Version]
  158. Hu, J.; Wang, X.; Chen, F.; Ding, M.; Dong, M.; Yang, W.; Yin, M.; Wu, J.; Zhang, L.; Fu, X.; et al. Combination of Decitabine and a Modified Regimen of Cisplatin, Cytarabine and Dexamethasone: A Potential Salvage Regimen for Relapsed or Refractory Diffuse Large B-Cell Lymphoma After Second-Line Treatment Failure. Front. Oncol. 2021, 11, 687374. [Google Scholar] [CrossRef]
  159. Qu, C.; Zou, R.; Wang, P.; Zhu, Q.; Kang, L.; Ping, N.; Xia, F.; Liu, H.; Kong, D.; Yu, L.; et al. Decitabine-primed tandem CD19/CD22 CAR-T therapy in relapsed/refractory diffuse large B-cell lymphoma patients. Front. Immunol. 2022, 13, 969660. [Google Scholar] [CrossRef]
Figure 1. Methylation-affected genes associated with leukemias and myelodysplastic syndromes. DNA methylation plays a key role in the initiation and progression of hematological malignancies. DNMTs catalyze transferring of the methyl group (a red circle) to the 5-carbon position of cytosine within CpG dinucleotides in DNA sequence (each strand of double-stranded DNA is marked by different colors, yellow and blue), leading to the formation of 5-methyl-cytosine. Abnormal methylation patterns in bone marrow cells may predict responsiveness to the treatment. A panel of aberrantly methylated genes shown to have diagnostic and prognostic value. Targeting the hypermethylated promoters of tumor suppressor genes might represent a perspective trend for hypomethylating drug therapy alone or in combination. Abbreviation: DNMTs, DNA methyltransferases.
Figure 1. Methylation-affected genes associated with leukemias and myelodysplastic syndromes. DNA methylation plays a key role in the initiation and progression of hematological malignancies. DNMTs catalyze transferring of the methyl group (a red circle) to the 5-carbon position of cytosine within CpG dinucleotides in DNA sequence (each strand of double-stranded DNA is marked by different colors, yellow and blue), leading to the formation of 5-methyl-cytosine. Abnormal methylation patterns in bone marrow cells may predict responsiveness to the treatment. A panel of aberrantly methylated genes shown to have diagnostic and prognostic value. Targeting the hypermethylated promoters of tumor suppressor genes might represent a perspective trend for hypomethylating drug therapy alone or in combination. Abbreviation: DNMTs, DNA methyltransferases.
Ijms 24 00633 g001
Figure 2. Methylation-affected genes associated with lymphomas. Aberrant DNA methylation (a red circle) regulates gene expression in Hodgkin, Non-Hodgkin, and Diffuse large B-cell lymphomas. Abnormally methylated genes might be used as potential biomarkers for therapeutic decisions and as predictive markers for patient outcomes.
Figure 2. Methylation-affected genes associated with lymphomas. Aberrant DNA methylation (a red circle) regulates gene expression in Hodgkin, Non-Hodgkin, and Diffuse large B-cell lymphomas. Abnormally methylated genes might be used as potential biomarkers for therapeutic decisions and as predictive markers for patient outcomes.
Ijms 24 00633 g002
Figure 3. Clinical significance of combined treatment with hypomethylating agents and other therapies in hematologic cancer patients. DNA methyl transferases DNMT1, DNMT3A, and DNMT3B mediated DNA methylation resulting in aberrant gene silencing. In hematologic cancers, aberrant DNA methylation cause changes in the function of tumor suppressor genes, which play a role in cancer-associated pathways (a). Epi-drugs, including first-generation azacitidine, decitabine, and second-generation guadecitabine, are used as DNA methylation inhibitors in leukemias, myelodysplastic syndromes, and lymphomas. HMAs aim to reactivate silenced genes subsequently involved in apoptotic, repairing, and angiogenic pathways. Recently, a strategy of combined treatment with HMAs and dose-specific chemotherapy/immunotherapy/radiotherapy and other epi-drugs represents unique challenges in patient outcomes (b). Abbreviations: DNMT1, DNA methyl transferase 1; DNMT3A, DNA methyl transferase 3A; DNMT3B, DNA methyl transferase 3B; HMAs, hypomethylating agents; TSG, tumor suppressor gene.
Figure 3. Clinical significance of combined treatment with hypomethylating agents and other therapies in hematologic cancer patients. DNA methyl transferases DNMT1, DNMT3A, and DNMT3B mediated DNA methylation resulting in aberrant gene silencing. In hematologic cancers, aberrant DNA methylation cause changes in the function of tumor suppressor genes, which play a role in cancer-associated pathways (a). Epi-drugs, including first-generation azacitidine, decitabine, and second-generation guadecitabine, are used as DNA methylation inhibitors in leukemias, myelodysplastic syndromes, and lymphomas. HMAs aim to reactivate silenced genes subsequently involved in apoptotic, repairing, and angiogenic pathways. Recently, a strategy of combined treatment with HMAs and dose-specific chemotherapy/immunotherapy/radiotherapy and other epi-drugs represents unique challenges in patient outcomes (b). Abbreviations: DNMT1, DNA methyl transferase 1; DNMT3A, DNA methyl transferase 3A; DNMT3B, DNA methyl transferase 3B; HMAs, hypomethylating agents; TSG, tumor suppressor gene.
Ijms 24 00633 g003
Table 1. The effect of hypomethylating agents on outcomes of patients with leukemias and myelodysplastic syndromes. The table summarizes the list of active and recruiting clinical trials focused on adverse events, treatment efficacy, EFS, PFS, OS, and ORR (according to https://clinicaltrials.gov/ and https://beta.clinicaltrials.gov/, accessed on 1 December 2022).
Table 1. The effect of hypomethylating agents on outcomes of patients with leukemias and myelodysplastic syndromes. The table summarizes the list of active and recruiting clinical trials focused on adverse events, treatment efficacy, EFS, PFS, OS, and ORR (according to https://clinicaltrials.gov/ and https://beta.clinicaltrials.gov/, accessed on 1 December 2022).
StudyDesignDiseasePurposePatients InterventionStatus
NCT02935361An interventional open-label study
single group assignment
CMML
MDS
AML
To assess the potential risk of treatment-related adverse events and to evaluate ORR and OS33 adultsGuadecitabine and atezolizumab as a humanized monoclonal antibody every 28 days up to 2 years in the absence of disease toxicityActive, not recruiting
NCT04167917An interventional open-label study
single group assignment
AML
MDS
CML
To confirm the tolerability and safety of the oral HMAs and to evaluate the CBR, OS, PFS, and ORR after epi-drug therapy20 adultsDNMT1 inhibitor known as NTX-301 with previously shown anti-leukemic efficacyRecruiting
NCT05042531An interventional prospective randomized study
parallel assignment
AMLTo determine the incidence of adverse events developed during the first and second cycles of azacitidine treatment30 adultsLow-dose dasatinib with azacitidine as a maintenance therapy option in chemotherapy-treated patientsRecruiting
NCT02878785An interventional randomized open-label study
factorial assignment
AMLTo test the best dose and assess the treatment-related adverse events25 adultsA combination of decitabine (5 days every month) with talazoparib (daily oral administration)Active, not recruiting
NCT03164057An interventional randomized open-label study
parallel assignment
AML
MDS
To analyze OS, EFS, and non-hematologic toxic events in patients receiving DNMTi before the chemotherapy206 adults/
children
Intravenously administered azacitidine or decitabine before chemotherapy regimensActive, not recruiting
NCT04722952An interventional open-label study
single-group assignment
AMLTo define the efficacy of combined therapies in refractory/relapsed patients and evaluate CR, ORR, OS, EFS, and PFS30 adultsPD-1 monoclonal antibody visilizumab plus azacitidine, and a homoharringtonine regimenRecruiting
NCT01806116An interventional open-label study
single group assignment
MDS
AML
To investigate the impact on posttransplant outcomes and evaluate the incidence of GvHD 30 adults/
children
Intravenous administration of decitabine followed by allogeneic HSCTUnknown
NCT02684162An interventional open-label study
single-group assignment
AML
CML
MDS
To confirm CRR, RFS, safety, and toxicity of treatment with/without donor lymphocyte infusion 55 adultsGuadecitabine and donor lymphocyte infusion in case of the absence of disease progression or toxicityActive, not recruiting
NCT03080766An interventional open-label study
single-group assignment
AMLTo identify CR, DFS, and OS in patients with chromosomal abnormalities associated with bad prognosis20 adultsDecitabine treatment with the examination of bone marrow after each therapy course Afterward, patients will receive HSCT plus decitabine.Recruiting
NCT05426798An interventional open-label study
single-group assignment
AML MDSTo test the tolerability and anti-tumor efficacy of combined treatment in patients with recurrent/refractory disease73 adultsInjection of monoclonal antibody TQB2618 with azacytidine/decitabine Recruiting
NCT03298321An observational prospective studyAMLTo study the potential effect of epi-drug on arrhythmia and left atrium fibrosis14 adultsAzacitidine administrationUnknown
NCT03066648An interventional non-randomized study
parallel assignment
AMLTo characterize doses for monoclonal antibodies in combination with HMAs and to evaluate dose-limiting toxicities, ORR, PFS, and TTP243 adultsDecitabine in combination with MBG453 and/or PDR001 as monoclonal antibodies, or azacitidine together with MBG453Active, not recruiting
NCT03238248An interventional open-label study
single-group assignment
MDSTo reveal the efficacy of combined treatment in patients who failed primary therapy with DNMTi and to determine CR and PFS71 adultsAzacitidine infusion and antineoplastic pevonedistat as an inhibitor of the Nedd8 activating enzymeActive, not recruiting
NCT02131597An interventional open-label study
single-group assignment
MDSTo determine CRR, OS, EFS adverse events, and mortality rate in study participants 107 adultsAdministration of guadecitabine up to 24 courses in the absence of progression and toxicityActive, not recruiting
NCT05148234An interventional non-randomized open-label study
sequential assignment
MDSTo evaluate the efficacy, safety, and tolerability of treatment and determine ORR, CRR, and therapy-associated adverse events200 adultsBMS-986253 plus decitabine/cedazuridine Recruiting
NCT04187703An interventional open-label study
single group assignment
MDSTo analyze ORR, CRR, improvement, and specify treatment-associated adverse events 20 adultsLow doses of azacitidine and decitabineRecruiting
Abbreviations: AML, acute myeloid leukemia; CBR, clinical benefit rate; CML, chronic myelomonocytic leukemia; CMML, chronic myelomonocytic leukemia; CR, complete remission; CRR, complete response rate; DFS, disease-free survival; DNMT, DNA methyltransferase; DNMTi, DNA methyltransferase inhibitor; EFS, event-free survival; GvHD, graft-versus-host disease; HMAs, hypomethylating agents; HSCT, hematopoietic stem cell transplantation; MDS, myelodysplastic syndromes; OS, overall survival; ORR, objective response rate; PD-1, programmed cell death-1; PFS, progression-free survival, RFS, relapse-free survival; TTP, time of progression.
Table 2. The impact of combined therapy with hypomethylating agents and other treatments on the outcomes of lymphoma patients. The table summarizes the list of active and recruiting clinical trials focused on adverse events, treatment efficacy, and survival rates (according to https://clinicaltrials.gov/ and https://beta.clinicaltrials.gov/, accessed on 1 December 2022).
Table 2. The impact of combined therapy with hypomethylating agents and other treatments on the outcomes of lymphoma patients. The table summarizes the list of active and recruiting clinical trials focused on adverse events, treatment efficacy, and survival rates (according to https://clinicaltrials.gov/ and https://beta.clinicaltrials.gov/, accessed on 1 December 2022).
StudyStudy DesignDiseasePurposePatients InterventionStudy Status
NCT04510610An interventional open-label study
single-group assignment
HLTo define the response duration and PFS in patients with relapsed/refractory disease treated with camrelizumab plus decitabine100 adults/
children
Decitabine plus anti-PD-1 monoclonal antibody camrelizumab Recruiting
NCT04233294An interventional open-label study
single-group assignment
HLTo document the adverse effects of combined therapy and ORR in cancer patients100 adults/
children
HDAC inhibitor chidamide in combination with monoclonal antibody camrelizumab plus decitabineRecruiting
NCT03250962An interventional randomized
open-label study
parallel assignment
HLTo test the safety of decitabine with SHR-1210 therapy compared to SHR-1210 monotherapy and investigate treatment-related adverse events280 adults/
children
SHR-1210 as an anti-PD-1 monoclonal antibody with/without the decitabine in relapsed or refractory patients Recruiting
NCT04337606An interventional randomized open-label studyNHLTo assess the safety, treatment efficacy, and therapy-induced adverse events 100 adultsDecitabine plus HDAC inhibitor chidamide or monoclonal antibody camrelizumab.Recruiting
NCT04514081An interventional randomized open-label study
parallel assignment
HLTo compare ORR and adverse events between immunotherapy-resistant patients with chidamide+decitabine+camrelizumab treatment vs. decitabine+camrelizumab200 adults/
children
Decitabine with monoclonal antibody camrelizumab or as a combination with camrelizumab+HDAC inhibitor chidamide.Recruiting
NCT03494296An observational case-control prospective studyDLBCLTo determine the efficacy and safety of low-dose decitabine with a chemotherapy regiment in relapsed and refractory patients150 adultsA combination of chemotherapeutics cyclophosphamide+vindesine+bonisone with decitabineRecruiting
NCT03346642An interventional open-label study
single-group assignment
DLBCLTo document potential treatment-related adverse events, ORR, CR, PFS, and OS in cancer patients30 adultsCombined therapies with gemcitabine, vinorelbine, doxorubicin, and monoclonal antibody SHR-1210 with/without decitabineUnknown
NCT05320640An interventional open-label study
single-group assignment
NHL
solid tumors
To evaluate predictive biomarkers of response, adverse events, ORR, the safety, and efficacy of combined therapy in participants with relapsed/refractory NHL and advanced solid tumors100 adults/
children
Chemotherapy-free regimen with HDAC inhibitor chidamide, low-dose hypomethylating agent decitabine, and anti-PD-1/PD-L1/CTLA-4 antibodies Recruiting
NCT03445858An interventional open-label study
single-group assignment
Lymphoma
solid tumors
To detect feasibility, toxicities, tolerability, and anti-tumor effect of combined therapy 21 adults/
children
Pembrolizumab with decitabine and fixed-dose hypofractionated radiotherapyActive, not recruiting
NCT04553393An interventional randomized open-label study
parallel assignment
NHLTo assess the effect of combined therapies in relapsed/refractory participants and reveal adverse events, PFS, and OS after the intervention80 adults/ childrenA combination of decitabine-primed Tandem CAR19/20 engineered T cells with/without chidamide, decitabineRecruiting
Abbreviations: anti-PD-1, anti-programmed cell death-1; CR, complete response; CTLA-4, cytotoxic T-lymphocyte–associated antigen 4; DLBCL, Diffuse large B-cell lymphoma; HDAC, histone deacetylase; HL, Hodgkin lymphoma; NHL, non-Hodgkin lymphoma; OS, overall survival; PD-L1, programmed death-ligand 1; ORR, objective response rate; PFS, progression-free survival.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kalinkova, L.; Sevcikova, A.; Stevurkova, V.; Fridrichova, I.; Ciernikova, S. Targeting DNA Methylation in Leukemia, Myelodysplastic Syndrome, and Lymphoma: A Potential Diagnostic, Prognostic, and Therapeutic Tool. Int. J. Mol. Sci. 2023, 24, 633. https://doi.org/10.3390/ijms24010633

AMA Style

Kalinkova L, Sevcikova A, Stevurkova V, Fridrichova I, Ciernikova S. Targeting DNA Methylation in Leukemia, Myelodysplastic Syndrome, and Lymphoma: A Potential Diagnostic, Prognostic, and Therapeutic Tool. International Journal of Molecular Sciences. 2023; 24(1):633. https://doi.org/10.3390/ijms24010633

Chicago/Turabian Style

Kalinkova, Lenka, Aneta Sevcikova, Viola Stevurkova, Ivana Fridrichova, and Sona Ciernikova. 2023. "Targeting DNA Methylation in Leukemia, Myelodysplastic Syndrome, and Lymphoma: A Potential Diagnostic, Prognostic, and Therapeutic Tool" International Journal of Molecular Sciences 24, no. 1: 633. https://doi.org/10.3390/ijms24010633

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

Article Metrics

Back to TopTop