Abstract
Non-Hodgkin’s lymphomas (NHLs) comprise a diverse group of diseases, either of mature B-cell or of T-cell derivation, characterized by heterogeneous molecular features and clinical manifestations. While most of the patients are responsive to standard chemotherapy, immunotherapy, radiation and/or stem cell transplantation, relapsed and/or refractory cases still have a dismal outcome. Deep sequencing analysis have pointed out that epigenetic dysregulations, including mutations in epigenetic enzymes, such as chromatin modifiers and DNA methyltransferases (DNMTs), are prevalent in both B- cell and T-cell lymphomas. Accordingly, over the past decade, a large number of epigenetic-modifying agents have been developed and introduced into the clinical management of these entities, and a few specific inhibitors have already been approved for clinical use. Here we summarize the main epigenetic alterations described in B- and T-NHL, that further supported the clinical development of a selected set of epidrugs in determined diseases, including inhibitors of DNMTs, histone deacetylases (HDACs), and extra-terminal domain proteins (bromodomain and extra-terminal motif; BETs). Finally, we highlight the most promising future directions of research in this area, explaining how bioinformatics approaches can help to identify new epigenetic targets in B- and T-cell lymphoid neoplasms.
Keywords: BET inhibitors, bioinformatics, clinical testing, DNMT, drug combination, epigenetics, EZH2, HAT, HDAC, non-Hodgkin’s lymphoma
Introduction
Non-Hodgkin’s lymphoma (NHL) is a common type of hematological malignant tumor, composed of multiple subtypes that originate from B lymphocytes, T lymphocytes, and natural killer (NK) cells. These entities are sub-divided into distinct categories based on the differentiation stage of the malignant cells and on the presence of specific genetic alterations, being diffuse large B-cell lymphoma (DLBCL), follicular lymphoma (FL), T-cell lymphoma (TCL), and NK-T-cell lymphoma (NKTCL) the most common subtypes.
Characteristics of B-NHL
At the origin of roughly 90% of the cases of lymphoma, B-NHLs comprise a heterogeneous group of lymphoid neoplasms originated from either mature or immature B cells, the diagnosis of which is based on morphological, immunophenotypic, and genetic features, and most of them being dictated by the cell of origin 1 (see Figure 1). B-NHL subtypes range in their severity from well-controlled, indolent diseases, to extremely aggressive forms with urgent unmet medical needs that still require the development of novel therapeutic options.
Figure 1.
B- and T-cell lymphomagenesis.
(a) Emergence of the main subtypes of B-cell non-Hodgkin’s lymphoma (B-NHL). Naive B cells first participate in the formation of germinal centers (GCs) upon interacting with antigens. In the dark zone, centroblasts prolife and undergo somatic hypermutation (SMH), while in the light zone, centrocytes are sorted on B-cell receptor (BCR)-based affinity and undergo class switch recombination (CSR). GC cells are the normal counterparts of follicular lymphoma (FL), Burkitt’s lymphoma (BL), and diffuse large B-cell lymphoma (DLBCL) of GC subtype (GCB). DLBCL of the activated B-cell subtype (ABC) originates from post-GC cells, and multiple myeloma (MM) arises from differentiated plasma cells. Chronic lymphocytic leukemia (CLL) can originate from either naïve or differentiated memory B cells. Mantle cell (MCL) and marginal zone lymphoma (MZL) arise from B cells located in the mantle and marginal zone of lymphoid follicles, respectively. (b) Intrinsic or extrinsic factors may favor TCLs pathogenesis through immune evasion, alterations in T cell receptor (TCR) signaling pathways, activation of transcription factors and proto-oncogenes, that favor the emergence of different entities during T cell differentiation from the thymus to lymph nodes.
AITL, angioimmunoblastic T-cell Lymphoma; ALCL-ALK+, anaplastic large cell lymphoma, ALK-positive; EATL, enteropathy-associated T-cell lymphoma; ENKL, extranodal natural killer/T-cell lymphoma; FTCL, follicular T-cell lymphoma; HTCL, hepatosplenic γδ T-cell lymphoma; MEITL, monomorphic epitheliotropic intestinal T cell lymphoma; PTCL, peripheral T-cell lymphoma.
The most common B-NHL subtype is DLBCL, which accounts for 25–35% of all cases. This aggressive disease is characterized by a heterogeneous molecular pathogenesis that can lead to different clinical characteristics. 2 Despite high-throughput sequencing having recently identified up to seven genetic subtypes,3–5 gene expression profiling (GEP)-mediated identification of germinal center B-cell (GCB), activated B-cell (ABC), and unclassifiable subgroup 6 remains widely used in the clinical management of these patients. Patients with ABC-DLBCL or genetic alterations in MYC and BCL2 and/or BCL6, called double hit (DHLs) or triple hit lymphomas (THLs), have generally a poor survival prognosis.
The second most common B-NHL is FL, another germinal center (GC)-derived malignancy that accounts for 20% of all cases. FL is characterized by the t(14;18)(q32;q21) translocation that leads to the overexpression of the BCL2 anti-apoptotic gene. 7 FL represents a paradigm of dependence on the microenvironment. Seminal microarray studies in lymph node (LN) biopsies from the Leukemia and Lymphoma Profiling Project (LLMPP) series established for the first time that FL prognosis was not given by the tumor cell per se but by the composition of non-malignant cells. 8 Although the clinical course is mostly indolent, about 20% of patients may relapse or progress to a transformed (more aggressive) form of FL (t-FL).
Burkitt lymphoma (BL) is a third GC-derived lymphoma, which represents 1–5% of all NHL, and which is characterized by the deregulation of MYC due to translocations such as t(8;14)(q23;q32). Three subtypes have been described, namely endemic, sporadic, and immunodeficiency-associated form, which is mostly found in patients infected with the human immunodeficiency virus (HIV). 9
Originated from mature B-cells of the mantle zone of the LN, mantle cell lymphoma (MCL) accounts for 3–10% of B-NHL, being the t(11;14)(q13;q32) translocation and the expression of the cell cycle regulator cyclin D1 (CCND1), physiologically undetectable in normal B cells, the main characteristics of the disease. Supporting the notion that t(11;14) translocation is an epigenetic event, the Eμ and 3′ Cα cis IgH enhancer elements and the regions upstream of the CCND1 gene are hypomethylated on the translocated allele, and histones surrounding the translocation have shown hyperacetylation. 10 MCL is also characterized by a high genomic instability with a high number of secondary genetic alterations involving cell cycle regulation, DNA damage response, cell death, nuclear factor kappa B (NF-κB), or epigenetic modifiers, with a median number of six secondary events. ATM, CCND1, TP53, and RB1 are among the most recurrently mutated genes. 11 MCL has poor prognosis due to diagnosis often at a disseminated stage and an aggressive clinical evolution.
Physiopathology of T-cell lymphoma
Contrasting with B-cell lymphoma, T-cell lymphomas (TCLs) are characterized by numerous T-cell subsets, functional plasticity of which depends on the microenvironment. As a result, appearance of features that do not fit with the cell originating the tumor is common,12–14 thereby preventing tumor classification and a proper diagnosis of the disease and affecting the clinical management of the patients. Introduction of global GEP technologies are slowly improving our knowledge in the pathological origin of TCLs, but there is still room for improvement.
Mechanisms implicated in TCL development can be intrinsic and imply recurrent mutations leading to: (1) immune evasion, being the most clear example of the nucleophosmin-anaplastic lymphoma kinase (NPM-ALK) oncogene, expressed in ALK+ anaplastic large cell lymphoma (ALCL), which regulates the expression of the immunosuppressive protein programmed death-ligand 1 (PD-L1); 15 (2) alterations within T cell receptor (TCR) signaling pathway, such as mutations affecting the RhoA GTPase 16 or gain-of-function mutations in regulators of T-cell activation and function such as phospholipase C gamma 1 (PLCG1), phosphatidylinositol-3 kinase (PI3K) family, catenin beta-1 (CTNNB1), CD28, caspase recruitment domain family member 11 (CARD11), vav guanine nucleotide exchange factor 1 (VAV1), interferon regulatory factor 4 (IRF4), mainly observed in peripheral T-cell lymphomas (PTCLs), including adult T-cell leukemia/lymphoma (ATLL) and angioimmunoblastic T-cell lymphoma (AITL);14,17 (3) the implication of whole families of genes or signaling pathways, such as the PI3K-AKT-mTOR axis found to be constitutive activated during TCL pathogenesis,18,19 or the Janus kinase (JAK)-signal transducer and activator of transcription (STAT) pathway, in which mutations in kinases and transcription activators such as STAT3 or STAT5B, JAK1, JAK3, or JAK5, may not directly induce the lymphoma, but rather confer clinical aggressiveness due to increased cytokine-dependent signaling and consequent promotion of cell growth, immune response, tumor metabolism, and cell death blockade;17,18,20,21 (4) alteration in transcription factors such as activator protein-1 (AP-1), musculoaponeurotic fibrosarcoma (MAF) and basic leucine zipper ATF-like (BAF) transcription factor families, which are commonly involved in cutaneous T-cell lymphoma (CTCL) and CD30 + PTCL development, by activating the expression of the T-cell growth and survival genes, MYC and CD30, or driving the production of pro-tumoral cytokines;22,23 (5) PI3K-AKT-mTOR-mediated deregulation or glutamine, nucleotides, and lipid metabolism, and consequent enhancement of lymphoma cell growth;24,25 (6) altered mitochondrial functions, including a STAT3-mediated deregulation of mitochondrial respiration,26,27 and as recently described, a glyceraldehyde 3-phosphate dehydrogenase (GAPDH) overexpression-driven development of AITL-like disease; 28 and (7) epigenetic deregulation, leading to the silencing of tumor-suppressor genes or the overexpression of several proto-oncogenes (see Figure 1).
Beside these intrinsic mechanisms, a number of extrinsic factors also favor TCLs pathogenesis, such as pro-inflammatory cytokines and chemokines, which build a microenvironment that promote and sustain the survival of the tumor cells, or oncogenic Epstein–Barr virus (EBV)- and T-cell lymphotropic virus type 1 (HTLV-1)-derived viral factors that promote oncogenesis from indolent cutaneous lymphoproliferative diseases to extremely aggressive TCLs such as ATLL.29,30
Epigenetic deregulations in B- and T-cell lymphoma
A high level of alteration in chromatin state and a frequent deregulation of several epigenetic modulators, are common hallmarks that have been highlighted in most B-NHL subtypes in early 2000s.
Modification of DNA methylation patterns (either hypo- or hyper- DNA methylation status) have been associated for years to somatic mutations in epigenetic regulators of DNA methylation and to a generally repressed conformation of the chromatin in B-NHL.31,32 Loss-of-function (LOF) mutations (truncation or frameshift mutations) affecting the SET domain of the histone-lysine N-methyltransferase 2D (MLL2/KMT2D) gene is the most frequently epigenetic alteration in DLBCL and FL, occurring, respectively, in about 30% and 90% of the cases, respectively. Inactivation of KMT2D leads to the expansion of GC-derived B cells, impedes class switch recombination and cooperates with the deregulation of BCL2 to increase the incidence of tumors, in association with reduced H3K4 methylated levels.33–38
While LOF and/or deleterious mutations in the CREB binding protein (CREBBP) and E1A binding protein 300 (EP300) coding for two histone acetyltransferases (HATs) were detected in about 40% of DLBCL and FL cases,35,39 these two entities also share the presence of point mutations in the myocyte enhancer binding factor 2B (MEF2B) gene codifying for a HAT recruiter (13% of GCB-DLBCL cases and 15% of FL patients). 39 From one side, mutant CREBBP and EP300 proteins are deficient in acetylating BCL6 and p53 leading, respectively, to constitutive oncogenic properties and to decreased tumor suppressor activity, thus favoring an increased tolerance for DNA damage mediated by impaired apoptosis triggering and cell cycle arrest. 40 In parallel, MEF2B mutations decrease the transcriptional activity of this factor, thereby lowering B-cell lymphoma 6 (BCL6) expression levels and leading to decreased activity of the TGFB1 tumor suppressor gene, encoding for transforming growth factor β, together with an increased expression of MYC. The global biological impact of these modifications consists in a reduced inhibition of chemotaxis and in the promotion of B-cell lymphomagenesis. 41
In a lower extent, activating mutations of the enhancer of zeste homolog 2 (EZH2) histone methyltransferase (HMT) gene are specifically found in GC-derived NHL, such as GCB-DLBCL (22% of the patients) and FL (7% of the patients)42,43 (see Figure 2). Expression of the EZH2Y641F mutant was associated with the reprogramming of the immunological niche, a process by which the supporting activity of T cells toward GC-B cell development becomes ensured by follicular dendritic cells (FDCs). 44 Finally, despite genes coding for histone deacetylases (HDACs) remained unmutated in B-NHL patients, an overexpression of HDAC1, HDAC2, and HDAC6 with a consequent global decrease in the activity of the transcription machinery, has been reported in DLBCL cases. 45 In particular, HDAC6 exerts a crucial role in the response to proteotoxic stress, as it orchestrates the acetylation-dependent stability of heat shock protein of 90 kD (HSP0) and the recruitment of misfolded protein cargo to dynein motors for their transport to intracellular aggresomes. 46
Figure 2.
Alterations of chromatin states in B- and T-cell lymphoma.
(a) Loss-of-function mutations affecting epigenetic regulators such as the SET domain of the histone-lysine N-methyltransferase 2D (MLL2/KMT2D), CREB binding protein (CREBBP), and E1A binding protein 300 (EP300) genes, as well as activating mutations of the enhancer of zeste homolog 2 (EZH2) histone methyltransferase (HMT) gene and the overexpression of HDAC1/2/6 have been reported in B-cell lymphoma cases. (b) Genetic alterations in DNA methylation genes affect Tet methylcytosine dioxygenase 2 (TET2), DNA methyl transferase 3 A (DNMT3A) and isocitrate dehydrogenase 2 (IDH2), as well as in chromatin remodelers such as SETD2, SETD1B, INO80, ARID1B, ARID2, ARID5B, SMARCC1 and histone acetylation genes such as EP300/CREBBP are found in T-cell lymphoma.
TCLs present a similar mutation spectrum of epigenetic genes than B-cell lymphoma, which includes missense mutations, concomitantly to a general chemo-resistant phenotype.14,47 Genetic alterations in DNA methylation genes affect Tet methylcytosine dioxygenase 2 (TET2), DNA methyl transferase 3 A (DNMT3A) and isocitrate dehydrogenase 2 (IDH2). These alterations are mainly found in T follicular helper (Tfh) cell-derived PTCLs and constitute the hallmark of AITL, a disease where the three genes are affected simultaneously. 48 TET2 and DNMT3A deregulations are also found in CTCL. 47 Similar mutations of TET2 are found in the hematopoietic stem and progenitor cells (HSPCs) of some AITL patients. Notwithstanding, none of these mutations will lead to lymphoma appearance, but will rather favor a premalignant status characterized by a disrupted homeostasis within the hematopoietic stem/progenitor compartment, being the malignant transformation achieved upon the acquisition of further defects in key genes such as RHOA, PLCG1, or NOTCH2. 49
Modifications in chromatin remodelers such as SET domain containing 2 (SETD2), INO80, involved in DNA repair and cell cycle regulation, and AT-Rich Interaction Domain 1B (ARID1B), have been found in up to 62% of hepatosplenic γδ T-cell lymphoma (HTCL) patients. 50 SETD2, which acts as a tumor suppressor gene, was the most silenced gene in HTCL and is mutated in monomorphic epitheliotropic intestinal T-cell lymphoma (MEITL). Loss of SETD2 has been further associated with increased cell growth, altered expression of cell cycle genes, and compromised DNA damage response and repair. 51 ARID2, ARID1B, SETD1B mutations are also found in PTCL-NOS. 52 ARID1A, ARID5B, and SMARCC1 mutations are present in CTCL. 53 Finally, ARID1A mutations have been detected in NKTCL. 54 While ARID1A expression is sufficient to suppress cellular proliferation and tumor growth in mouse models of cancers, in-frame mutations of this gene lead to the nuclear retention of this tumor suppressor, due to its increased degradation and incapacity to stimulate the expression of the cell cycle inhibitor gene, CDKN1A. 55
Defects in histone methylation have been associated to mutations in KMT2C, KM2D, and KDM6A genes in PTCL-NOS. 52 NKTC lymphoma harbors mutations in KMT2D, EZH2, and also in ASXL transcriptional regulator 3 (ASXL3) genes, a phenomenon linked to altered histone deubiquitination. 54 Histone acetylation genes such as EP300 and CREBBP are mutated in PTCL-NOS 52 (see Figure 2). Deregulated expression and/or activity of these epigenetic writers, lead to genome-wide DNA and histone modification pattern alterations, modifying the chromatin structure, and thereby leading to cancer development. 56
Importantly, HDACs have become a promising therapeutic targets (see below) in the context of some TCLs such as CTCL, through their role in the silencing of different BCL-2 proapoptotic family members and in the regulation of autophagy. 57
Role of epigenetic drugs in B- and T-cell lymphoma
DNA methylation inhibitors
DNA methyltransferases (DNMTs) family includes three major members that have functional methylation activities, namely DNMT1, which mediates maintenance methylation during cell division, and DNMT3A and DNMT3B, that regulate de novo DNA methylation.58,59 Accumulating evidences have shown that DNMTs regulation of methylation have an important role in normal hematopoiesis.60,61 On the contrary, aberrant methylation is a key molecular event on the development of hematological malignancies. Indeed, hypermethylation of CDKN2A/p16 is associated with aggressive forms of B-NHL, and can drive progression of T-cell malignancies.62,63 Considering that an altered methylation pattern is frequently observed in most types of cancers, including B- and T-cell lymphoma, the development of epigenetic drugs to restore methylation levels seems to be an interesting strategy for the treatment of these diseases. 64
In the 1970s, the nucleoside analogs 5-azacytidine and 5-aza-20-deoxycytidine (or decitabine) appeared in the clinic as a chemotherapeutic option against acute leukemia. Mechanistically, these cytidine analogs are incorporated into DNA, promoting an irreversible blockage of DNMT1, which consequently leads to global demethylation.65–67 Although azacytidine and decitabine were considered as promising chemotherapeutic agents, the early clinical trials showed a disappointing response and a pronounced toxicity in myelodysplastic syndrome (MDS) and acute myeloid leukemia (AML) patients.68–70 Subsequently, in vitro data indicated that low doses of azacytidine can induce a decrease in DNA methylation level, supporting its use as an epigenetic drug. 71 Given the remarkable improvement of the overall survival (OS), and more generally the excellent clinical response of MDS patients to this agent, azacytidine became in 2004 the first drug approved by Food and Drug Administration (FDA) for this indication.72,73 In B-NHL patients, only two phase I studies using decitabine as a single agent have been completed so far.74,75 However, until now, the use of DNMT inhibitors (DNMTis) as monotherapy yielded disappointing results in lymphoid malignancies. 76 Currently, azacytidine and decitabine are being evaluated mostly in combination with other drugs in approximately 95 active clinical trials involving B- and T-cell lymphoma patients. In this review, we will present clinical data from completed or terminated clinical trials (see Table 1).
Table 1.
Clinical evaluation of epidrug and epidrug-based combination therapies in B- and T-cell lymphoma.
| Targets | Drug/Regimen | Trial ID | Phase/Status | No. of patients | Diseases | Response | Toxicity | Study |
|---|---|---|---|---|---|---|---|---|
| DNMT | Decitabine + Tetrahydrouridine | NCT02846935 | I Completed | 7 | R/R- HL, DLBCL, PTCL, AITL | OR 57% | Grade 3–4 AEs: TP 100% | Hill et al. 77 |
| DNMT | Azacitidine + Vorinostat +Gemcitabine + Busulfan + Melphalan | NCT01983969 | I/II Completed | 61 | DLBCL, HL, T-cell NHL, FL, MCL | EFS 75% | Not serious AEs: 100% | Clinical trial 78 |
| DNMT | Azacitidine + Lenalidomide | NCT01121757 | II Terminated | 9 | R/R – FL, MZL | CR 17% | Grade 3–4 AEs: Toxicities 100%, TP 33% | Clinical trial 79 |
| DNMT | Azacitidine + Vorinostat | NCT01120834 | I/II Completed | 18 | R/R DLBCL | ORR 6.7% | Grade 3–4 AEs: Toxicities 100%, Anemia 7%, TP 33% | Clinical trial 80 |
| DNMT | Azacitidine + RCHOP | NCT01004991 | I/II Completed | 12 | DLBCL | CR 91.7% | Grade 3–4 AEs: NP 100% | Clinical trial 81 |
| DNMT | Azacitidine + RCHOP | NCT02343536 | I Completed | 52 | DLBCL, FL, tFL | ORR 94.9%, CR 88.1% 1- and 2-year PSF 84.1% and 78.6% |
Grade 3–4 AEs: NP 62.7%, febrile NP 25.4% | Martin et al. 82 |
| DNMT | Azacitidine + Romidepsin | NCT01998035 | I/II Terminated | 23 | R/R–NHL, HL | ORR 61% CR 48% |
Grade 3–4 AEs: TP 48%, NP 40%, LP 32%, anemia 16% | Falchi et al. 83 O’Connor et al. 84 |
| DNMT | Azacitidine + Avelumab + Utomilumab + | NCT02951156 | III Terminated | 9 | R/R DLBCL | ORR 0 | No dose-limiting toxicities were reported | Hawkes et al. 85 |
| DNMT | Vincristine + Prednisone + Etoposide + Thioguanine, Cytarabine + Asparaginase + Methotrexate | NCT00002590 | II Completed | 86 | Children with lymphoma | 5-year OS 80% 5-year EFS 68% | Grade 4 AEs: NP 82%, TP 66%, anemia 38%, hepatotoxicity 4% |
Lowe et al. 86 |
| HDAC | Belinostat | NCT00274651 | II Terminated | 53 | R/R cutaneous and peripheral TCL | ORR 25% (PTCL) and 14% (CTCL) | Grade 3–4 AEs: 77% | Johnston et al. 87 |
| HDAC | Belinostat + CHOP | NCT01839097 | II Completed | 23 | PTCL first line | ORR 86% CR 71% |
No additional toxicity was observed with the combination SAEs 43% | Johnston et al. 88 |
| HDAC | Belinostat | NCT00865969 | II Completed | 129 | R/R PTL | ORR 25.8% | Grade 3–4 AEs 47.29% | Campbell and Thomas 89 |
| HDAC | Panobinostat + Bortezomib | NCT00901147 | II Completed | 25 | R/R–PTL lymphoma or NK/TCL | ORR 43% CR 22% |
Grade 3–4 AEs: TP 68%, NP 20%, diarrhea 20% | Tan et al. 90 |
| HDAC | Panobinostat | NCT01523834 | II Completed | 35 | DLBCL | ORR 17.1% CR 11.4% |
Grade 3–4 AEs: TP >5% |
Duvic et al. 91 |
| HDAC | Panobinostat | NCT00425555 | II Completed | 139 | Refractory CTL | PR 21.6% CR 0.7% |
Grade 3–4 AEs: 28.8 |
Duvic et al. 91 |
| HDAC | Panobinostat | NCT01034163 | III Completed | 41 | HL | Not Completed 100% | Grade 3–4 AEs: TP 27%, NP 27%, diarrhea 89% | Clinical trial 92 |
| HDAC | Panobinostat + Everolimus | NCT00967044 | I/II Completed | 31 | R/R lymphoma | ORR 10% (TCL, MCL, HL) ORR 43% (HL) CR 15% (HL) |
Grade 3–4 AEs: TP 64%, NP 47%, anemia 20% | Oki et al. 93 |
| HDAC | Valproic acid + RCHOP | NCT01622439 | I/II Completed | 495 | DLBCL | 1-year PFS 97.0%, 2-year PFS 84.7% 1-year OS 100% 2-year OS 96.8% |
No toxicity reported | Drott et al. 94 |
| HDAC | Abexinostat | NCT00724984 | I/II Completed | 87 | R/R lymphoma, NHL, and CLL | ORR 28% CR 5% |
Grade 3 AEs: TP 80%, NP 27%, anemia 12% |
Evens et al. 95 and Ribrag et al. 96 |
| HDAC | Mocetinostat | NCT00358982 | II Terminated | 51 | R/R HL | PR 30% CR 10% |
Grade 3–4 AEs: TP 21.5%, NP 13.7% |
Younes et al. 97 |
| HDAC | Fimepinostat + Rituximab, Venetoclax | NCT01742988 | I Completed | 44 | R/R DLBCL, MYC-altered DLBCL | DLBCL ORR 37% CR 11.2% PR 14.6 MYC-altered DLBCL ORR 64% CR 36% PR27% |
Grade 3–4 AEs: TP 20%, NP 7% |
Younes et al. 98 |
| HDAC | Vorinostat + Rituximab + Combination chemotherapy | NCT00601718 | I/II Completed | 29 | R/R lymphoma or untreated TNHL, MCL | ORR 70% CR 30% |
Grade 3 AEs: gastrointestinal toxicity 33% |
Budde et al. 99 |
| HDAC | Vorinostat + Niacinamide + Etoposide | NCT00691210 | I Completed | 40 | HL, NHL | ORR 24%, CR 10% PR 14% 7.5%. Twelve patients achieved SD (57%). |
Grade 3–4 AEs: 27.5% | Amengual et al. 100 |
| HDAC | Vorinostat + Chemotherapy + Rituximab | NCT01193842 | I/II Completed | 86 | HIV-related DLBCL or other aggressive BCL | CR 74% versus 68% for EPOCH versus EPOCH-vorinostat | Grade 4 AEs: NP (47% and 20%), LP (20% and 7%), and TP (29% and 2%) for vorinostat-EPOCH and EPOCH alone, respectively |
Ramos et al. 101 |
| HDAC | Vorinostat + Fludarabine phosphate + Cyclophosphamide + Rituximab | NCT00918723 | I/II Completed | 40 | CLL 32, SLL 8 | ORR 91% CR 74%, PR 11% |
Grade 4 AEs: hematologic 36%, electrolyte abnormalities 9%, gastrointestinal 8%, and cardiovascular 6% |
Shadman et al. 102 |
| HDAC | Vorinostat + Cladribine + Rituximab | NCT00764517 | I/II Completed | 57 | MCL, R/R-CLL, or BNHL | ORR 39% in R/R patients, ORR 97%, CR 80% in MCL |
Grade 2–3 AEs: NP 67%, TP 42%, anemia 14% | Spurgeon et al. 103 |
| HDAC | Vorinostat + Bortezomib | NCT00992446 | II Completed | 27 | NHL | ORR 74% CR 42% |
Grade 4 AEs: NP 30% |
Holmberg et al. 104 |
| HDAC | Vorinostat | NCT00253630 | II Completed | 35 | R/R-FL, MZL, MCL | FL-ORR 47%, CR 23.5%, PR 23.5%, MZL-ORR 22%, CR 11%, PR 11% | Acceptable safety profile | Kirschbaum et al. 105 |
| HDAC | Vorinostat | NCT00875056 | II Completed | 50 | R/R-FL, BNHL, MCL | FL-ORR 49%, CR 10%, PR 31%. | Grade 3–4 AEs: TP 93%, NP 68% |
Ogura et al. 106 |
| HDAC | Vorinostat + Rituximab + Cyclophosphamide + Etoposide + Prednisone | NCT00667615 | I/II Completed | 30 | R/R DLBCL | ORR 57%, CR 35%, PR 22% | Grade 3–4 AEs: LP 90%, leucopenia 66%, NP 52%, TP 41%, and anemia 31% |
Straus et al. 107 |
| HDAC | Vorinostat + Rituximab | NCT00720876 | II Completed | 28 | iNHL | ORR 41%, CR 27%, PFS 29.2 months | Grade 3–4 AEs: Fatigue 32%, LP 25% |
Chen et al. 108 |
| HDAC | Chidamide + CHOP | NCT02809573 | I/II Completed | 383 | PTCL | ORR 39% as chidamide monotherapy, ORR 51% combined therapy | Grade 3–4 AEs: TP 18.1%, NP 12.6%, anemia 7.1%, and fatigue 5.5% |
Shi et al. 109 |
| HDAC | Romidepsin | NCT00007345 | II Completed | 71 | TCL, PTCL | ORR 34%, CR 20% | Grade 3–4 AEs TP 47%, leucopenia 47%, granulocytopenia 45%, anemia 40%, LP 17% |
Piekarz et al. 110 |
| HDAC | Romidepsin + Alisertib | NCT01897012 | I Completed | 25 | R/R aggressive BCL and TCL | ORR 28% CR 12% |
Grade 3–4 AEs: TP 40%, NP 24%, anemia 28% |
Strati et al. 111 |
| HDAC | Romidepsin + 5-azacitidine | NCT01998035 | I/II Terminated | 25 | PTCL, FL | PTCL ORR 61%, CR 48%, FL ORR 80%, CR 67% | Grade 3–4 AEs: TP 48%, NP 40%, LP 32%, and anemia 16% |
Falchi et al. 83 |
| HDAC | Romidepsin | NCT00106431 | II Completed | 96 | R/R CTCL | ORR 34%, CR 8%, DOR of 15.0 months | AEs: 96.2% | Duvic et al. 112 |
| HDAC | Romidepsin | NCT01456039 | I/II Completed | 40 | R/R PTCL | ORR 43%, CR 25%, DFS 5.6 months | Grade 3–4 AEs: TP 98%, LP 88%, leukopenia 84%, NP 80%, and anemia 34% |
Maruyama et al. 113 |
| HDAC | Romidepsin | NCT00426764 | II Completed | 131 | Progressive or relapsed PTCL | ORR 25%, CR 15%, DOR 28 months | Grade 3–4 AEs: TP 40% |
Foss et al. 114 |
| HDAC | Romidepsin + Gemcitabine | NCT01822886 | II Completed | 20 | R/R PTCL | ORR 30%, CR 15%, 2-year OS 50%, PSF 11.2% | Grade 3 AEs: TP 60%, NP 50%, anemia 20% |
Pellegrini et al. 115 |
| HDAC | Entinostat | NCT00866333 | II Terminated | 59 | R/R HL | PR 12% | Grade 3 AEs: TP 63%, anemia 47%, NP 41%, leukopenia 10%, |
Tan et al. 116 |
| HDAC | Ricolinostat | NCT02091063 | I Completed | 21 | TCL, HL, PTLD, FL, DLBCL, CTCL | 56 days of disease stabilization for 50% of patients | Grade 3–4 AEs: anemia 9.5% and hypercalcemia 9.5% |
Amengual et al. 117 |
| EZH2 | Tazemetostat | NCT03009344 | I Completed | 10 | R/R NHL | ORR 57%, CR 14.3% | Grade 3–4 AEs: TP and dysgeusia 42.9% |
Munakata et al. 118 |
| EZH2 | Tazemetostat | NCT03456726 | II Completed | 165 | R/R BNHL EZH2mut | ORR 40% (DLBCL EZH2mut) 18% in patients DLBCL EZH2wt; ORR 63% ( FL EZH2mut) 28% (FL EZH2wt) | Grade 3–4 AEs: 18% | Izutsu et al. 119 |
| EZH2 | Tazemetostat + Atezolizumab | NCT02220842 | I Completed | 43 | R/R-FL and DLCBL | ORR 16%, CR 5%, PR 12% | Grade 3–4 AEs: anemia 12%, NP 9%, TP 9% |
Palomba et al. 120 |
| EZH2 | Tazemetostat | NCT03010982 | I Completed | 3 | BCL | Not reported | Grade 3–4 AEs: 66% |
Argon et al. 121 |
| EZH2 | Tazemetostat + Fluconazole + Omeprazole + Repaglinide | NCT03028103 | I Completed | 16 | BCL | ORR 15.4%, CR 7.7%, PR 7.7% | Grade 3–4 AEs: 25% tazemetostat-related | Clinical trial 122 |
| EZH2 | Tazemetostat | NCT01897571 | I/II Completed | 99 | DLBCL EZH2wt and mut | ORR 69% (EZH2mut) and 35% (EZH2wt), CR 13% (EZH2mut) and 4% (EZH2wt) | Grade 3–4 AEs: TP 3%, NP 3%, anemia 2% |
Morschhauser et al. 123 |
| EZH2 | GSK2816126 | NCT02082977 | I Terminated | 20 | R/R DBCL, tFL and NHL | Insufficient evidence of clinical activity | Grade 3–4 AEs: 32% | Yap et al. 124 |
| BET | Birabresib | NCT01713582 | I Completed | 17 | DLCBL | DOR 17% | Grade 3–4 AEs: TP 58% | Amorim et al. 125 |
| BET | RO6870810 + Venetoclax + Rituximab | NCT03255096 | I Completed | 39 | DLBCL | ORR 38.5%, CR 20.5, PR 17.9% | Grade 3–4 AEs: NP 28%, anemia and TP 23% each |
Dickinson et al. 126 |
| BET | INCB057643 | NCT02711137 | I/II Terminated | 16 | Lymphoma | PR 25%, CR 12.5% | Grade 3–4 AEs: TP 32% |
Falchook et al. 127 |
| BET | INCB054329 | NCT02431260 | I/II Terminated | 4 | Lymphoma | Not reported | Grade 3–4 AEs: TP 33% |
Falchook et al. 127 |
| BET | CPI-0610 | NCT01949883 | I Completed | 64 | Progressive lymphoma | ORR 7.8%, CR 3% | Grade 3–4 AEs: TP 45% |
Blum et al. 128 |
| BET | FT-1101 | NCT02543879 | I Completed | 10 | NHL | SD 33% | Grade 3–4 AEs: pleural effusion 20% |
Patel et al. 129 |
AEs, adverse effects; AITL, angioimmunoblastic T-cell lymphoma; BCL, B-cell lymphoma; BET, bromodomain and extra-terminal motif; CLL, chronic lymphocytic leukemia; CR, complete response; CTCL, cutaneous T-cell lymphoma; DLBCL, diffuse large B-cell lymphoma; DOR, duration of response; FL, follicular lymphoma; HDAC, histone deacetylase; MCL, mantle cell lymphoma; MM, multiple myeloma; MZL, marginal zone lymphoma; NA, not available; NHL, non-Hodgkin’s lymphoma; NP, neutropenia; ORR, overall response rate; OS, overall survival; PD, progressive disease; PFS, progression-free survival; PR, partial response; PTCL, peripheral T-cell lymphoma; R/R, relapsed and refractory; SD, stable disease; TP, thrombopenia.
Azacytidine was evaluated, in a phase II clinical study (NCT01998035), in combination with the HDAC inhibitor (HDACi) romidepsin. This combination was safe and demonstrated to be an effective regimen for treatment-naïve patients and R/R PTCL cases. The reported overall response (ORR) and complete response (CR) rates were 61% and 48%, respectively, therefore supporting the use of oral azacytidine and romidepsin in both first-line and R/R settings, including as a bridging therapy.83,84 The mechanism of action underlying this combinatorial activity might be related to the enhanced expression of several cancer-testis antigens involved in the promotion of tumor immunogenicity. 130 The combination of azacytidine with avelumab (anti-PD-L1 antibody) and utomilumab (4-1BB/CD137 agonist) was evaluated in R/R DLBCL patients. However, due to insufficient clinical activity, the study was prematurely discontinued. 85
High levels of cytidine deaminase (CDA) were reported to decrease the half-life of decitabine and azacytidine.131,132 Based on this observation, the effect of tetrahydrouridine (THU), a CDA inhibitor, was evaluated in relapsed B- and T-cell malignancies (NCT02846935, see Table 1). The overall results indicated a significant reduction in tumor burden, together with subjective improvements in symptoms in 57% of the patients; however, these effects were lost upon treatment interruptions consequent to neutropenia. 77 Decitabine was also evaluated in treatment-naïve DLBCL patients prior to receiving R-CHOP. The results from this clinical trial (NCT02951728) indicated that 86% of the patients responded to the treatment, with 74% CR and 11% PR. 133
Besides the above-described FDA-approved DNA methylation inhibitors, over the last decades several new DNMTis have been developed, showing promising results in pre-clinical models of hematological disorders. Among them it is worth to highlight the thioguanine [2-amino-1,7-dihydro-6H-purine-6-thione (6-tG)] that was FDA-approved for the treatment of AML patients. 134 Thioguanine is currently included in the NHL-BFM90 protocol. 135 The main findings reported by the large European network applying this protocol for the management of childhood and adolescent NHL, indicated that the 5-year event-free survival (EFS) was in the range 82–90% (NCT00275106 and NCT00004228, respectively, see Table 1), while 5-year OS was in the range 87–96%. 136
HDAC inhibitors
As commented above, HDACs play a crucial role in the control of cell proliferation and growth, and angiogenesis. Accordingly, their overregulation promotes the occurrence of cancer, making the use of HDACis an efficient anticancer approach as these agents can trigger apoptosis, cell cycle arrest or even regulate the immune system.57,137,138 Among the growing list of clinical trials exploring the possible therapeutic application of HDACi (see Table 1), the most successful ones are depicted below.
Vorinostat is a hydroxamic acid presenting inhibitory activity against classes I and II HDACs. 139 It has been used successfully in monotherapy in different subtypes of B-NHL, including FL and MCL,106,140 and has been approved in 2006 by FDA to treat R/R CTCL. 141 This agent has also been tested with success in combination with multiple drugs such as the anti-CD20 monoclonal antibody rituximab, cladribine, fludarabine phosphate, cyclophosphamide, niacinamide, prednisone, or etoposide, for the treatment of indolent or aggressive B-NHL, including relapsed cases, in R/R CLL cases and in previously untreated T-NHL.99,107,108,142–145
Mocetinostat, a selective HDAC class I and IV inhibitor developed by Methylgene has also been approved for R/R CTCL, 146 while its therapeutic effect in the context of R/R Hodgkin’s lymphoma (HL) managed to control the disease evolution in half of the treated patients. 97
Belinostat is a sulfonamide-hydroxamic acid acting as a pan-HDACi 147 developed by TopoTarget. While its use has been approved for the treatment of R/R PTCL, it also demonstrated remarkable activity when assessed in CTCL patients.87,89,148 Conversely, this agent exhibited poor activity as monotherapy in B-cell lymphoma patients. Interestingly, when combined to CHOP chemotherapeutic regimen as a first-line treatment for PTCL patient, it achieved a significant CR rate. 149
Chidamide is a selective HDAC class I inhibitor 150 currently approved by the China Food and Drug Administration (CFDA) as a treatment for R/R PTCL, alone or in combination with CHOP regimen. A new clinical trial involving refractory extranodal NKTCL patients, and the results of which will be released in 2023, has already shown promising results. 151 Its applicability to the management of AITL is also under evaluation. Several clinical trials aiming at evaluating chidamide as a therapeutic strategy for B-NHL treatment alone or in combination with R-GDP, DICE, R-CHOP have been registered and are in recruiting processes.
Ricolinostat, a selective HDAC6 inhibitor, 152 exerted a synergic antitumor effect in combination with the BTK inhibitor ibrutinib, the alkylating agent bendamustine, the ALK inhibitor crizotinib and the proteasome inhibitor carfilzomib, in preclinical models of DLBCL and MCL, but possible translation into human subjects is still under evaluation. 57 In a clinical study including patients with R/R B/T-lymphomas, this agent harbored lower toxicity compared with other pan-HDACis and achieved the stabilization of half of the patients, although in the absence of CR or partial response (PR). 145
Fimepinostat, which dually targets HDAC and PI3K, has been tested in combination with rituximab in relapsed DLBCL and MYC-altered DLBCL patients.98,153 Response to treatment was achieved most notably in MYC-altered DLBCL (64% versus 37% in non-MYC-altered).
Romidepsin is a depsipeptide from bacterial origin able to inhibit Class I HDACs. 154 While its use has been approved for R/R CTCL and PTCL,114,155–158 it also showed additive effect when combined with other agents employed for the treatment of PTCL, FL, or R/R T-cell lymphomas, such as the nucleoside analog gemcitabine, 5-azacitidine or the aurora kinase inhibitor alisertib.159–161
Panobinostat, a hidroxamic acid with therapeutic activity as pan-HDAC inhibitory, has been evaluated up to phase III clinical trials in R/R HL patients. Unfortunately, it failed to achieve the therapeutic effect expected from previous phases. 162 Similarly, phase II clinical trials involving R/R DLBCL patients demonstrated a modest activity of the drug. 163 In T-cell lymphoma and HL, some degree of clinical response was observed when associating this agent with the mTOR inhibitor everolimus. 93 Further drug assessment in HL, but not in CTCL patients, employing panobinostat in combination with ICE, also showed initial positive response.164,165 Panobinostat was also administered to R/R HL patients in combination with the immunomodulatory drug lenalidomide, showing good tolerances but low therapeutic activity. 166 Finally, this HDACi has been tested in combination with bortezomib for the treatment of R/R TCL, providing a good response rate, but also producing serious adverse effects (AEs) in almost half of the cases. 167
Entinostat is a HDAC I and III inhibitor, therapeutic effect of which has been tested in R/R HL. Different dosages were tested, achieving an ORR of 12%, and a mean tumor size reduction of 58%. 116
Abexinostat, a pan-HDAC inhibitor, has been tested in a subset of R/R NHL and CLL patients. A phase II trial demonstrated that it can be successfully considered for the treatment of FL, TCL, and DLBCL, 95 with ORRs in the 56–31% range (see Table 1).
Valproic acid, an organic weak acid previously used as anticonvulsant, also exhibits HDAC inhibition properties. Several clinical trials aimed at evaluating its effect in lymphoma patients have been carried out, showing that it significantly improved the efficacy of the R-CHOP chemo-immunotherapy as a first-line treatment for DLBCL, as shown by a reduced rate of relapsed disease and a 96.8% OS at 2 years. 94
HMT inhibitors
Currently, the development of HMT inhibitors is mainly focused on EZH2 targeting. This gene encodes an HMT that forms the catalytic subunit of the polycomb repressive complex 2, which mediates gene silencing via the addition of methyl groups to H3K27. EZH2 overexpression has been correlated with poor prognosis in several tumor types including hematological malignancies.168–171 Moreover, it has been shown that the gain-of-function mutation at Y641 residue of EZH2, leads to its enzymatic activity converting mono- or di-methylated H3K27 to the tri-methylated state, promoting the repression of cell differentiation and the downregulation of tumor suppressor genes. 172 Interestingly, these gain-of-function EZH2 mutations has been frequently associated with the presence of BCL2 rearrangement in both FL (28%) and GCB-DLBCL groups (33%), and it has been shown to be a crucial regulator of MYC-associated lymphomagenesis in B cells.173–175 Thus, EZH2 activity is currently a potential therapeutic strategy to treat B-NHL.
In this context, several selective EZH2 inhibitors have been developed in the last decade, for the treatment of hematological patients. Among these molecules, GSK126 demonstrated an improved inhibitory capacity toward mutant EZH2 in preclinical studies.176,177 However, its clinical evaluation in R/R patients, including cases with DLBCL, transformed tFL, MM, and other NHL, (NCT02082977) was stopped prematurely due to insufficient therapeutic activity. 124 Another small molecule called CPI-1205, which showed promising activity in xenograft mouse models and in human B-NHL cell lines, 178 is currently being evaluated in a phase I trial involving DLBCL patients (NCT02395601). At this time, no data have been released regarding the efficacy and the safety of this agent in these heavily pretreated patients. Antitumoral effects of EZH2 inhibitors have been also reported in TCL cell lines. 179 The use of EZH1/2 inhibitor, valemetostat, has been evaluated in clinical trial including B- and T-NHL subtypes (NCT02732275). The observed ORR was 53% in a phase I trial involving 15 patients, being the cases with TCL by far the best responders (80% ORR). 180
Tazemetostat, a potent and selective EZH2 inhibitor, first demonstrated encouraging preclinical activity, in both in vitro and in vivo models (including xenografts) of EZH2-mutant NHL. 181 An initial phase I study showed that tazemetostat, as single agent, was safe and exerted notable antitumor activity (OR = 38%) in patients with refractory B-NHL. 182 Based on pharmacodynamic and toxicity profiling, a phase II study including EZH2mut and EZH2wt R/R FL patients was carried out. Surprisingly, an ORR of 69% was observed among EZH2mut patients, and 35% in EZH2wt cohort. 123 Based on these data, FDA granted accelerated approval of this agent for R/R-FL patients with EZH2mut. Similarly, two independent cohorts of R/R DLBCL and FL patients corroborated that tazemetostat was effective in R/R EZH2mut FL patients.118,119 Tazemetostat was also evaluated in combination with the anti-PD-L1 antibody atezolizumab in R/R DLBCL patients (NCT02220842). However, this combination did not provide additional efficacy. 120 Thus, further clinical trials are needed to establish new drug combinations based on HMT inhibitors to improve the OS of the patients.
Although the clinical development of HMT inhibitors is centered on the targeting of EZH2 in B-NHL patients, other indirect strategies have shown encouraging results in preclinical models of the disease. Among these approaches, the carbocyclic adenosine analog 3-deazaneplanocin A (DZNep), is an inhibitor of S-adenosylhomocysteine hydrolase (AdoHcyase), that triggers intracellular accumulation of 5-adenosylhomocystein, followed by the blockade of several methyltransferases, including EZH2. Despite a strong apoptogenic effect in different preclinical cancer models, DZNep antitumor effect was not restricted to EZH2-mutated cases and may involve EZH2-independent targets, therefore, hinting as potential undesired effects in patients. 183 Blockade of the transfer of a methyl group from S-adenosyl methionine (SAM) to a lysine or arginine residue can also be achieved upon exposure to 2-chloro-2′-deoxyadenosine (2-CdA, cladribine), a drug that inactivates S-adenosyl-L-homocysteine hydrolase (SAH), thereby impairing the capacity of genomic DNA (gDNA) to accept methyl groups. Although cladribine was shown to efficiently reduce the level of methylated cytosines in myeloid cells, 184 its clinical activity as single agent in lymphoid malignancies, including CLL, MCL, hairy cell leukemia, and marginal zone lymphoma (MZL), was initially associated to its capacity, as a purine analog, to selectively target and suppress lymphocyte growth. 185 However, in MCL, induction therapy associating a purine analog to rituximab and cyclophosphamide demonstrated a lower response rate than standard R-CHOP regimen, 186 contrasting with the impressive 97% ORR observed in patients receiving a cladribine-containing immuno-epigenetic combo. 142 Therefore, it is highly plausible that, at least in MCL patients, the remarkable clinical activity of cladribine is derived from its epigenetic modulatory properties.
Bromodomain inhibitors
Bromodomain-containing family of proteins comprise 46 members, including nuclear proteins with HAT or HMT activity, chromatin remodelers, helicases, transcription co-activators, and mediators or scaffold proteins. Undoubtedly most of the scientific attention was focused on BRD2, BRD3, and BRD4. 187 Interestingly, these proteins have a bromodomain and extra-terminal motif (BET) domain, which is responsible for protein–protein interactions,187,188 that acts as DNA enhancers for oncogene regulation. 189 BET bromodomains are important regulators of several players with a central role in several cancers, including hematological malignancies, such as MYC, cell cycle regulators, and NF-κB-related genes. 190 Thus, in early 2000s, the development of BET inhibitors (BETi) rapidly gained a growing attention as a promising anticancer strategy.
The first BETi to enter into clinical development was birabresib. The data from the initial dose-escalation, open-label, phase I trial (NCT01713582) indicated that, when used as monotherapy, birabresib achieved a 47% CR rate in 17 R/R DLBCL patients, while durable OR was reported for only 18% of them. 125 A dose exploration study of this drug, involving DLBCL and AML patients (NCT02698189), was closed prematurely due to a lack of efficacy.
Others BETis, INCB054329 and INCB057643, were evaluated as monotherapy in a small cohort of patients with hematologic malignancies (NCT02431260 and NCT02711137, respectively). In both cases, the clinical trial was terminated due to a lack of responses and notable toxicity. 127
CPI-0610 is another BETi, which was evaluated in B-NHL patients, reaching a modest 7% OR rate (NCT01949883). Mechanistically, the analysis of BET target genes demonstrated a dose-dependent decrease (2–8 h post-dose) in interleukin-8 (IL8) and C-C motif chemokine receptor 1 (CCR1) transcript levels. 128
FT-1101, is a small molecule that presents a potent anti-proliferative activity in vitro, which was subsequently confirmed in B-NHL patients in a phase I clinical trial (NCT02543879). Although this agent showed an acceptable safety profile, a modest clinical activity was observed. 129
Given the strong synergy between BET and BCL-2 inhibitors in vitro, 191 a phase Ib trial evaluated the effects of RO6870810 combined to the BCL2 antagonist, venetoclax, with or without rituximab in R/R DLBCL patients. ORR was 38.5%, including 20.5% CR and 17.9% PR. A stable disease (SD) was observed in 15.4% of the patients, whereas 30.8% maintained a progressive disease (PD). Although the triple combination resulted in higher response rates when compared with single agents, authors concluded to a lack of synergistic effect of this combinatorial approach. 126
Future perspectives: paving the way for precision medicine with epidrugs in B- and T-cell lymphoma and implementing bioinformatics approaches to identify new epigenetic targets in hematological cancers
Personalized epigenome targeting in NHL
The identification of recurrent mutations in NHL provides a rationale to introduce epigenetic drugs for the design of selective and personalized therapies for these patients. Supported by promising results obtained in preclinical and clinical trials, the notion that epigenetic alterations can be used for lymphoma diagnosis and as valuable clinical biomarkers of response to therapy, represents a game changer for the clinical management of these patients.
Several preclinical studies have already shown that a genetic or epigenetic alteration could be specifically targeted by a particular drug. As described in the section ‘HMT inhibitors’, EZH2 inhibitors have demonstrated enhanced preclinical and clinical activity against EZH2-mutated B-NHL. Preclinical studies evaluating the safety and efficacy of a new class of LSD1 inhibitors showed that this class of agents had a very potent activity against proliferation of MLL-rearranged leukemia cells, with EC50 values in the nanomolar range. These cell-permeable molecules were found to significantly increase the levels of H3K4me2 in MLL-rearranged leukemic cells, to significantly reduce tumor burden, and to expand the life expectancy of leukemic mice, demonstrating the potential of this class of compounds to become clinically useful for MLL-rearranged leukemia. 192
In a different context, HDAC3 inhibition using the novel molecule BRD3308 led to the upregulation of the cell cycle inhibitor, p21, resulting in cell cycle blockade and in proliferation arrest exclusively in CBP/p300-mutated DLBCL cells, characterized by the presence of BCL6–HDAC3 complexes that repress the transcription of the p21-encoding gene, CDKN1A. 193
Despite these significant advances in the design of personalized treatments that could target a specific epigenetic alteration by means of one or several epidrugs, until recently efforts were still to be made to elaborate the best treatment approach for a determined patient with altered epigenome. The implementation of massive sequencing in early 2000s, followed by the standardization of big data processing, responded partly to these crucial needs.
In silico epigenetic profiling: basic concepts
Bioinformatics tools have enabled the interpretation of complex and dynamic phenomena, such as epigenetics, as well as the development of large-scale (big data), highly selective (e.g. single cell chip), and high-throughput modern molecular biology techniques. 194 These latter are aimed at facilitating the prediction of epigenetic events such as DNA methylation, chromatin conformation changes, histone modifications and non-coding RNA activity, and their relationship with other sources of biological information (e.g. phenotype, transcriptomic, and proteomic), with the final objective to improve the diagnosis and prognosis of different diseases, 195 and the design of effective and safe epidrugs, 196 and breeding strategies. 197 Epigenomic data, defined as the set of variations in certain genes, regions, or in the whole genome, 198 can be evaluated by methods that vary according to (1) the analytical platform used for data collection (e.g. array, sequencing, immunoprecipitation)199–201, (2) the stage of analysis (e.g. preprocessing of raw data, statistical comparison, variable selection, functional analysis, and drug-target structural analysis), 202 or (3) the objective of the study (e.g. multiomics, single cell, drug repurposing).203–205
Epigenetics-based drug therapy is primarily aimed at reversing aberrant gene expression that may be found in the development and progression of lymphoma and understanding the potential to affect multiple cellular processes. These bioinformatics targets can be found by directly affecting regulatory regions of genes, such as in the methylome or DNA methylation information sets, or by altering histone proteins and related structures. 206 The methylome can be profiled experimentally by bisulfite sequencing, while the main technique for identifying histone alteration profiles is immunoprecipitation of these proteins.
Computational validation of epigenetic alterations in B- and T-cell lymphoma
The first stages of bioinformatics analysis of epigenetic data are data acquisition, quality control, and preprocessing. At these levels, the characteristics of the bioinformatics methods depend on the particularities of the experimental techniques. Chromatin immunoprecipitation (ChIP), followed by DNA identification by microarray (ChIP-on-chip) or sequencing (ChIP-seq), and mass spectrometry (MS) are the most common experimental approaches to profile histone post-translational modifications. ChIP-on-chip data require the normalization of hybridized fragment intensities, followed by the selection of representative peaks, and the merging of overlapping overrepresented regions. 207 Many packages have been developed for this purpose such as Ringo, 208 Tilescope, 209 HMMTiling, 210 among others. ChIP-seq analyses require mapping short histone-associated DNA fragments to reference genomes. This step constitutes a challenge in both ChIP-on-chip and ChIP-seq. This so-called peak calling step needs more attention and quality control to avoid false overrepresentation and to map sequences that are actually in contact with our protein of interest (e.g. histones). Several open source (e.g. Bowtie, BLAT, or EpiChip)211–213 or proprietary (e.g. Broad Institute sequencing platform) 214 programs or packages have been developed for reference alignment. In addition, packages such as SAMtools can be used for removal of duplicate reads due to amplification artifacts, to finally perform peak calling using some of the available approaches that best fit the experimental design. 215 In this sense, Peakzilla uses predefined cutoff values, 216 while MACS uses Poisson model to analyze the distribution of reads, 217 and JAMM retrieves information from biological replicates to determine the amplitude of enriched sites. 218 There are several studies using the ChIP technique to confirm whether alterations in chromatin structure and expression of certain genes are restored after epidrug treatment.219–221
MS has been used to determine the altered structure of proteins of interest, although there are limitations in recognizing the large number of ions inherent to histones or other cellular proteins. Different algorithms have been developed depending on the integrity of the quantified protein (total, peptides or amino acids), THRASH being one of the most versioned although it still depends on the structural reference database. 222 These MS methods have recently been used to accurately target EZH2 and HDAC in lymphoma. 223
The methylome encompasses information on DNA methylation at the genomic level, a well-studied phenomenon for which there are analytical methods based on bisulfite modification followed by sequencing 200 or microarray 224 analysis, and enrichment methods such as methylated DNA immunoprecipitation (MeDIP) or methyl-binding domain proteins (MethylCap) followed by DNA sequencing.225,226 Bisulfite sequencing is based on the quantification of cytosines sequenced as thymines, after interaction with bisulfites that do not affect methylated bases. The bioinformatics algorithms used in the data generated by this technique have, as their main objective, the correct alignment to a reference genome and the calculation of the rate of cytosines and thymines that represent the methylation level of the genome studied. In addition to the previously mentioned methods for ChIP-seq, there are specific approaches that consider cytosine mismatches implemented, such as RRBSMAP 227 or Segemehl, 228 among others, that consider a three-base alignment (T, G, and A) such as MethylCoder or BRAT-nova.229,230 With the sequences already aligned, it is possible to quantify the frequency of cytosines and thymines in certain regions of the genomes using probabilistic algorithms such as Bis-SNP and MethylExtract.231,232
Different databases have compiled and organized the data of epigenetic modifications and the information related to each deposited project. These platforms provide information for planning future projects and facilitate meta-analyses that may lead to new conclusions not seen in individual studies. 233 Databases such as NCBI epigenomics and ENCODE collect data from several large-scale projects.234,235 MethDB and Methbank, among others, provide data exclusively on DNA methylation, while ChromatinDB and HHMD contain information on histone post-translational modifications for different species.236–239
The epigenetic alterations evaluated in previous steps can be used to define a differential profile that can be related to other types of omics data, especially gene expression data that are directly affected by such phenomena. As an example, the relationship between differential gene expression and binding analysis was used to point out the reversion of CREBBP mutations by HDAC3 selective inhibitors. 193 Programs such as diffReps 240 allow the detection of differential sites in ChIP-seq data where differences in the transcriptomic profile generated in those same regions are expected to be seen. Although it is often difficult to determine this association due to the variability of peak callers, some studies have shown its importance in, for example, defining genes and pathways associated with doxorubicin resistance. 241 Some studies have successfully used data on DNA sequence variations (e.g. in the differentiation of MCL subtypes), 242 protein or metabolite abundance, and clinical information. 243 This approach was also used to determine the association between metabolic and sensitivity to HDACis, 244 and to establish the relevance of phosphoproteomics data, gene expression and protein–protein interaction for the deciphering of a novel mechanism of regulation of KMT2D. 245 These types of associations could allow the reconstruction of multilevel regulatory networks. 246 The methods used to analyze this integration can vary greatly depending on the type of data used, between probabilistic models, 247 machine learning 248 (i.e. a model of protein-compound interaction with epigenetic targets using different machine learning methods), 249 coexpression networks 250 (i.e. differential coexpression and regulation networks to propose therapeutic factor for adult T-cell LL), 251 based on dimensionality reduction and clusterization, 252 integrated as networks based on quantitative relationships or described in previous literature, among others. Finally, we can use the structural information of the proteins responsible for modulating epigenetic marks in nucleosomes and their activity in the covalent bonds that generate these modifications. These data may help in the discovery of new epidrugs and epiprobes. 253 The main computer-aided drug design (CADD) methods in this area include druggability prediction, 254 virtual screening, 255 pharmacophore modeling, 256 and molecular dynamics simulations. 257
Conclusion and future perspectives
In the last decade, the advances in high-throughput sequencing technologies have provided multitude of information on the role of both genomic and epigenomic deregulations in malignant transformation and in the physiopathology of B and T lymphoid neoplasms. Besides helping the characterization of some lymphoma subtypes considered so far unclassifiable, the identification of recurrent mutations affecting epigenetic enzymes with crucial functions in the determination of B and T cell fate, fostered the development and evaluation of several epigenetic drugs, which rapidly showed high degree of efficacy, first in preclinical settings, and further in clinical trials essentially in relapsed NHL patients. Single agents and combination therapies involving HDAC inhibition or DNMT blockade have shown remarkable activity in specific subsets of B-cell and T-cell lymphoma, which remained unresponsive to or relapsed after chemo-immunotherapeutic regimens. However, beside general inhibitors of acetylation and demethylase, specific epigenetic drugs, including class-specific HDACi, EZH2 inhibitors or BET antagonists are showing significant clinical activity and have already been broadly implemented in the clinical management of determined subtypes of NHL patients. In combination therapy, some of these agents have demonstrated their ability to facilitate the response of B-NHL patients to anti-CD20 treatment. Among them, vorinostat was able to improve the outcomes of R/R FL, MCL, and MZL patients when associated to rituximab; 258 however, this effect seemed barely additive and the combined use of additional chemotherapeutic combo such as ICE or CHOP may be a prerequisite to achieve synergistic antitumor activity (NCT00601718, NCT00972478). Other HDACis such as panobinostat, fimepinostat, belinostat, or valproic acid either failed to exhibit such combinatorial activity with rituximab-based regimens or were associated with important AEs that limited their clinical development.94,153,259 Within the family of methylation modulators, cladribine also exhibits combinatorial activity with rituximab, as MCL patients receiving this combination presented a superior CR rate than the group of patients treated with cladribine alone (52% versus 42%, respectively). 260 In previously untreated patients, the observed responses were even higher when vorinostat was added to this doublet (97% ORR and 80% CR). 142 Importantly, these last studies showed a plateau in PFS and OS as well as a prolonged PFS, implying potential curative potential for cladribine–rituximab based regimens. In the case of EZH2 inhibitors, the first results of the phase Ib Epi-R-CHOP study combining tazemetostat to rituximab-CHOP suggested that this combo, although generally well tolerated, does not appear to substantially improve rituximab-CHOP toxicity and efficacy profiles. 261 Thus, in determined but limited conditions, epidrug therapy may significantly improve the outcome of NHL patients receiving anti-CD20-based regimens.
Despite these outstanding advances, further field of development include the identification of ad hoc biomarkers that may facilitate patient selection for their inclusion in clinical trials, and the design of rationally based combination regimens, may help reaching the final aim of establishing selective and personalized therapies for this subgroup of patients. One of the latest and most promising approach, which will help selecting the most relevant targeted therapies in the near future, is the multiomic characterization of each patient sample. However, the considerable volume of information provided by the latest technologies, including single-cell proteogenomics, requires a rigorous and reproducible analytical workflow that is nowadays within our reach, thanks to the recent improvements in bioinformatic data analysis. These approaches will allow us to be closer to the complete mapping of B- and T-cell lymphoma and, therefore, will empower the design of optimal therapeutic scheme, including new epigenetic drug combinations.
Acknowledgments
The authors thank Josep Carreras Leukaemia Research Institute and CERCA Programme/Generalitat de Catalunya for institutional support.
Footnotes
ORCID iDs: Salvador Sánchez Vinces
https://orcid.org/0000-0001-7041-6544
Gael Roué
https://orcid.org/0000-0003-0245-2257
Contributor Information
Marcelo Lima Ribeiro, Lymphoma Translational Group, Josep Carreras Leukaemia Research Institute, Badalona, Spain; Laboratory of Immunopharmacology and Molecular Biology, Sao Francisco University Medical School, Braganca Paulista, Brazil.
Salvador Sánchez Vinces, Laboratory of Immunopharmacology and Molecular Biology, Sao Francisco University Medical School, Braganca Paulista, Brazil.
Laura Mondragon, T Cell Lymphoma Group, Josep Carreras Leukaemia Research Institute, IJC. Ctra de Can Ruti, Camí de les Escoles s/n, 08916 Badalona, Barcelona, Spain.
Gael Roué, Lymphoma Translational Group, Josep Carreras Leukaemia Research Institute, IJC. Ctra de Can Ruti, Camí de les Escoles s/n, 08916 Badalona, Barcelona, Spain.
Declarations
Ethics approval and consent to participate: Not applicable.
Consent for publication: Not applicable.
Author contributions: Marcelo Lima Ribeiro: Investigation; Methodology; Resources; Validation; Writing – original draft; Writing – review & editing.
Salvador Sánchez Vinces: Methodology; Resources; Software; Writing – original draft; Writing – review & editing.
Laura Mondragon: Conceptualization; Investigation; Supervision; Writing – original draft; Writing – review & editing.
Gael Roué: Conceptualization; Funding acquisition; Investigation; Supervision; Writing – original draft; Writing – review & editing.
Funding: The authors disclosed receipt of the following financial support for the research, authorship, and/or publication of this article: Research at Roué lab was co-financed by the European Regional Development Fund (ERDF) through the Interreg V-A Spain-France-Andorra (POCTEFA) program (grant no. EFA360/19) and the Spanish Ministry of Economy and Competitiveness (grant no. PID2021-391230OB-C21).
The authors declared the following potential conflicts of interest with respect to the research, authorship, and/or publication of this article: G. Roué received research funding from TG Therapeutics, Inc. and Kancera AB, to support studies unrelated to this work. The remaining authors have no competing financial interests.
Availability of data and materials: Not applicable.
References
- 1.Seifert M, Scholtysik R, Küppers R. Origin and pathogenesis of B cell lymphomas. Methods Mol Biol 2013; 971: 1–25. [DOI] [PubMed] [Google Scholar]
- 2.Sehn LH, Salles G. Diffuse large B-cell lymphoma. N Engl J Med 2021; 384: 842. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Chapuy B, Stewart C, Dunford AJ, et al. Molecular subtypes of diffuse large B cell lymphoma are associated with distinct pathogenic mechanisms and outcomes. Nat Med 2018; 24: 679–690. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Schmitz R, Wright GW, Huang DW, et al. Genetics and pathogenesis of diffuse large B-cell lymphoma. N Engl J Med 2018; 378: 1396–1407. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Wright GW, Huang DW, Phelan JD, et al. A probabilistic classification tool for genetic subtypes of diffuse large B cell lymphoma with therapeutic implications. Cancer Cell 2020; 37: 551–568.e14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Alizadeh AA, Elsen MB, Davis RE, et al. Distinct types of diffuse large B-cell lymphoma identified by gene expression profiling. Nat 2000; 403: 503–511. [DOI] [PubMed] [Google Scholar]
- 7.Mozas P, Rivero A, López-Guillermo A. Past, present and future of prognostic scores in follicular lymphoma. Blood Rev 2021; 50: 100865. [DOI] [PubMed] [Google Scholar]
- 8.Kline J, Godfrey J, Ansell SM. The immune landscape and response to immune checkpoint blockade therapy in lymphoma. Blood 2020; 135: 523–533. [DOI] [PubMed] [Google Scholar]
- 9.Roschewski M, Staudt LM, Wilson WH. Burkitt’s lymphoma. N Engl J Med 2022; 387: 1111–1122. [DOI] [PubMed] [Google Scholar]
- 10.Hasanali Z, Sharma K, Epner E. Flipping the cyclin D1 switch in mantle cell lymphoma. Best Pract Res Clin Haematol 2012; 25: 143–152. [DOI] [PubMed] [Google Scholar]
- 11.Saleh K, Cheminant M, Chiron D, et al. Tumor microenvironment and immunotherapy-based approaches in mantle cell lymphoma. Cancers 2022; 14: 3229. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.De Leval L, Jaffe ES. Lymphoma classification. Cancer J 2020; 26: 176–185. [DOI] [PubMed] [Google Scholar]
- 13.Elenitoba-Johnson KSJ, Lim MS. New insights into lymphoma pathogenesis. Annu Rev Pathol 2018; 13: 193–217. [DOI] [PubMed] [Google Scholar]
- 14.Fiore D, Cappelli LV, Broccoli A, et al. Peripheral T cell lymphomas: from the bench to the clinic. Nat Rev Cancer 2020; 20: 323–342. [DOI] [PubMed] [Google Scholar]
- 15.Kasprzycka M, Marzec M, Liu X, et al. Nucleophosmin/anaplastic lymphoma kinase (NPM/ALK) oncoprotein induces the T regulatory cell phenotype by activating STAT3. Proc Natl Acad Sci USA 2006; 103: 9964–9969. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Cortes JR, Ambesi-Impiombato A, Couronne L, et al. RHOA G17V induces T follicular helper cell specification and promotes lymphomagenesis. Cancer Cell 2018; 33: 259–273.e7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Vallois D, Dobay MP, Morin RD, et al. Activating mutations in genes related to TCR signaling in angioimmunoblastic and other follicular helper T-cell-derived lymphomas. Blood 2016; 128: 1490–1502. [DOI] [PubMed] [Google Scholar]
- 18.Pechloff K, Holch J, Ferch U, et al. The fusion kinase ITK-SYK mimics a T cell receptor signal and drives oncogenesis in conditional mouse models of peripheral T cell lymphoma. J Exp Med 2010; 207: 1031–1044. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Heavican TB, Bouska A, Yu J, et al. Genetic drivers of oncogenic pathways in molecular subgroups of peripheral T-cell lymphoma. Blood 2019; 133: 1664–1676. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Witzig TE, Hu G, Offer SM, et al. Epigenetic mechanisms of protein tyrosine phosphatase 6 suppression in diffuse large B-cell lymphoma: implications for epigenetic therapy. Leukemia 2014; 28: 147–154. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Zhang Q, Wang HY, Wei F, et al. Cutaneous T cell lymphoma expresses immunosuppressive CD80 (B7-1) cell surface protein in a STAT5-dependent manner. J Immunol 2014; 192: 2913–2919. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Schleussner N, Merkel O, Costanza M, et al. The AP-1-BATF and -BATF3 module is essential for growth, survival and TH17/ILC3 skewing of anaplastic large cell lymphoma. Leukemia 2018; 32: 1994–2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Marzec M, Zhang Q, Goradia A, et al. Oncogenic kinase NPM/ALK induces through STAT3 expression of immunosuppressive protein CD274 (PD-L1, B7-H1). Proc Natl Acad Sci USA 2008; 105: 20852–20857. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.McDonnell SR, Hwang SR, Basrur V, et al. NPM-ALK signals through glycogen synthase kinase 3beta to promote oncogenesis. Oncogene 2012; 31: 3733–3740. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Lien EC, Lyssiotis CA, Cantley LC. Metabolic reprogramming by the PI3K-Akt-mTOR pathway in cancer. Recent Results Cancer Res 2016; 207: 39–72. [DOI] [PubMed] [Google Scholar]
- 26.McDonnell SR, Hwang SR, Rolland D, et al. Integrated phosphoproteomic and metabolomic profiling reveals NPM-ALK-mediated phosphorylation of PKM2 and metabolic reprogramming in anaplastic large cell lymphoma. Blood 2013; 122: 958–968. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Gough DJ, Corlett A, Schlessinger K, et al. Mitochondrial STAT3 supports Ras-dependent oncogenic transformation. Science 2009; 324: 1713–1716. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Mondragon L, Mhaidly R, De Donatis GM, et al. GAPDH overexpression in the T cell lineage promotes angioimmunoblastic T cell lymphoma through an NF-κB-dependent mechanism. Cancer Cell 2019; 36: 268–287.e10. [DOI] [PubMed] [Google Scholar]
- 29.Kim WY, Montes-Mojarro IA, Fend F, et al. Epstein-Barr virus-associated T and NK-cell lymphoproliferative diseases. Front Pediatr 2019; 7: 71. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Watanabe T. Adult T-cell leukemia: molecular basis for clonal expansion and transformation of HTLV-1-infected T cells. Blood 2017; 129: 1071–1081. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Hassler MR, Schiefer AI, Egger G. Combating the epigenome: epigenetic drugs against non-Hodgkin’s lymphoma. Epigenomics 2013; 5: 397–415. [DOI] [PubMed] [Google Scholar]
- 32.Blecua P, Martinez-Verbo L, Esteller M. The DNA methylation landscape of hematological malignancies: an update. Mol Oncol 2020; 14: 1616–1639. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Shilatifard A. Molecular implementation and physiological roles for histone H3 lysine 4 (H3K4) methylation. Curr Opin Cell Biol 2008; 20: 341–348. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Morin RD, Mendez-Lago M, Mungall AJ, et al. Frequent mutation of histone-modifying genes in non-Hodgkin lymphoma. Nature 2011; 476: 298–303. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Pasqualucci L, Trifonov V, Fabbri G, et al. Analysis of the coding genome of diffuse large B-cell lymphoma. Nat Genet 2011; 43: 830–837. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Lohr JG, Stojanov P, Lawrence MS, et al. Discovery and prioritization of somatic mutations in diffuse large B-cell lymphoma (DLBCL) by whole-exome sequencing. Proc Natl Acad Sci USA 2012; 109: 3879–3884. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Reddy A, Zhang J, Davis NS, et al. Genetic and functional drivers of diffuse large B cell lymphoma. Cell 2017; 171: 481–494.e15. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Ortega-Molina A, Boss IW, Canela A, et al. The histone lysine methyltransferase KMT2D sustains a gene expression program that represses B cell lymphoma development. Nat Med 2015; 21: 1199–1208. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Morin RD, Mendez-Lago M, Mungall AJ, et al. Frequent mutation of histone-modifying genes in non-Hodgkin lymphoma. Nature 2011; 476: 298–303. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Pasqualucci L, Dominguez-Sola D, Chiarenza A, et al. Inactivating mutations of acetyltransferase genes in B-cell lymphoma. Nature 2011; 471: 189. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Pon JR, Wong J, Saberi S, et al. MEF2B mutations in non-Hodgkin lymphoma dysregulate cell migration by decreasing MEF2B target gene activation. Nat Commun 2015; 6: 7953. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Morin RD, Johnson NA, Severson TM, et al. Somatic mutations altering EZH2 (Tyr641) in follicular and diffuse large B-cell lymphomas of germinal-center origin. Nat Genet 2010; 42: 181–185. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Serganova I, Chakraborty S, Yamshon S, et al. Epigenetic, metabolic, and immune crosstalk in germinal-center-derived B-cell lymphomas: unveiling new vulnerabilities for rational combination therapies. Front Cell Dev Biol 2022; 9: 3626. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Béguelin W, Teater M, Meydan C, et al. Mutant EZH2 induces a pre-malignant lymphoma Niche by reprogramming the immune response. Cancer Cell 2020; 37: 655–673.e11. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Marquard L, Poulsen CB, Gjerdrum LM, et al. Histone deacetylase 1, 2, 6 and acetylated histone H4 in B- and T-cell lymphomas. Histopathology 2009; 54: 688–698. [DOI] [PubMed] [Google Scholar]
- 46.Cosenza M, Pozzi S. The therapeutic strategy of HDAC6 inhibitors in lymphoproliferative disease. Int J Mol Sci 2018; 19: 2337. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Cortes JR, Palomero T. Biology and molecular pathogenesis of mature T-cell lymphomas. Cold Spring Harb Perspect Med 2021; 11: a035402. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Cortés JR, Palomero T. The curious origins of angioimmunoblastic T-cell lymphoma. Curr Opin Hematol 2016; 23: 434–443. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Quivoron C, Couronné L, Della Valle V, et al. TET2 inactivation results in pleiotropic hematopoietic abnormalities in mouse and is a recurrent event during human lymphomagenesis. Cancer Cell 2011; 20: 25–38. [DOI] [PubMed] [Google Scholar]
- 50.McKinney M, Moffitt AB, Gaulard P, et al. The genetic basis of hepatosplenic T-cell lymphoma. Cancer Discov 2017; 7: 369–379. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Mar BG, Chu SH, Kahn JD, et al. SETD2 alterations impair DNA damage recognition and lead to resistance to chemotherapy in leukemia. Blood 2017; 130: 2631–2641. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Palomero T, Couronné L, Khiabanian H, et al. Recurrent mutations in epigenetic regulators, RHOA and FYN kinase in peripheral T cell lymphomas. Nat Genet 2014; 46: 166–170. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Choi J, Goh G, Walradt T, et al. Genomic landscape of cutaneous T cell lymphoma. Nat Genet 2015; 47: 1011–1019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Choi S, Go JH, Kim EK, et al. Mutational analysis of extranodal NK/T-cell lymphoma using targeted sequencing with a comprehensive cancer panel. Genomics Inf 2016; 14: 78–84. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Guan B, Gao M, Wu CH, et al. Functional analysis of in-frame indel ARID1A mutations reveals new regulatory mechanisms of its tumor suppressor functions. Neoplasia 2012; 14: 986–993. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Yamagishi M. The role of epigenetics in T-cell lymphoma. Int J Hematol 2022; 116: 828. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Chen IC, Sethy B, Liou JP. Recent update of HDAC inhibitors in lymphoma. Front Cell Dev Biol 2020; 8: 576391. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Belinsky SA, Klinge DM, Stidley CA, et al. Inhibition of DNA methylation and histone deacetylation prevents murine lung cancer. Cancer Res 2003; 63: 7089–7093. [PubMed] [Google Scholar]
- 59.Hermann A, Goyal R, Jeltsch A. The Dnmt1 DNA-(cytosine-C5)-methyltransferase methylates DNA processively with high preference for hemimethylated target sites. J Biol Chem 2004; 279: 48350–48359. [DOI] [PubMed] [Google Scholar]
- 60.Shaknovich R, Cerchietti L, Tsikitas L, et al. DNA methyltransferase 1 and DNA methylation patterning contribute to germinal center B-cell differentiation. Blood 2011; 118: 3559–3569. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Lai AY, Mav D, Shah R, et al. DNA methylation profiling in human B cells reveals immune regulatory elements and epigenetic plasticity at Alu elements during B-cell activation. Genome Res 2013; 23: 2030–2041. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Nosaka K, Maeda M, Tamiya S, et al. Increasing methylation of the CDKN2A gene is associated with the progression of adult T-cell leukemia. Cancer Res 2000; 60: 1043–1048. [PubMed] [Google Scholar]
- 63.Martinez-Delgado B, Fernandez-Piqueras J, Garcia MJ, et al. Hypermethylation of a 5’ CpG island of p16 is a frequent event in non-Hodgkin’s lymphoma. Leukemia 1997; 11: 425–428. [DOI] [PubMed] [Google Scholar]
- 64.Azad N, Zahnow CA, Rudin CM, et al. The future of epigenetic therapy in solid tumours – lessons from the past. Nat Rev Clin Oncol 2013; 10: 256–266. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Ghoshal K, Datta J, Majumder S, et al. 5-Aza-deoxycytidine induces selective degradation of DNA methyltransferase 1 by a proteasomal pathway that requires the KEN box, bromo-adjacent homology domain, and nuclear localization signal. Mol Cell Biol 2005; 25: 4727–4741. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
- 66.Juttermann R, Li E, Jaenisch R. Toxicity of 5-aza-2’-deoxycytidine to mammalian cells is mediated primarily by covalent trapping of DNA methyltransferase rather than DNA demethylation. Proc Natl Acad Sci USA 1994; 91: 11797–11801. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Almstedt M, Blagitko-Dorfs N, Duque-Afonso J, et al. The DNA demethylating agent 5-aza-2’-deoxycytidine induces expression of NY-ESO-1 and other cancer/testis antigens in myeloid leukemia cells. Leuk Res 2010; 34: 899–905. [DOI] [PubMed] [Google Scholar]
- 68.Vogler WR, Miller DS, Keller JW. 5 Azacytidine (NSC102816): a new drug for the treatment of myeloblastic leukemia. Blood 1976; 48: 331–337. [PubMed] [Google Scholar]
- 69.Saiki JH, McCredie KB, Vietti TJ, et al. 5-azacytidine in acute leukemia. Cancer 1978; 42: 2111–2114. [DOI] [PubMed] [Google Scholar]
- 70.Levi JA, Wiernik PH. A comparative clinical trial of 5-azacytidine and guanazole in previously treated adults with acute nonlymphocytic leukemia. Cancer 1976; 38: 36–41. [DOI] [PubMed] [Google Scholar]
- 71.Jones PA, Taylor SM. Cellular differentiation, cytidine analogs and DNA methylation. Cell 1980; 20: 85–93. [DOI] [PubMed] [Google Scholar]
- 72.Silverman LR, Demakos EP, Peterson BL, et al. Randomized controlled trial of azacitidine in patients with the myelodysplastic syndrome: a study of the cancer and leukemia group B. J Clin Oncol 2002; 20: 2429–2440. [DOI] [PubMed] [Google Scholar]
- 73.Fenaux P, Mufti GJ, Hellstrom-Lindberg E, et al. Efficacy of azacitidine compared with that of conventional care regimens in the treatment of higher-risk myelodysplastic syndromes: a randomised, open-label, phase III study. Lancet Oncol 2009; 10: 223–232. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Stewart DJ, Issa JP, Kurzrock R, et al. Decitabine effect on tumor global DNA methylation and other parameters in a phase I trial in refractory solid tumors and lymphomas. Clin Cancer Res 2009; 15: 3881–3888. [DOI] [PubMed] [Google Scholar]
- 75.Blum KA, Liu Z, Lucas DM, et al. Phase I trial of low dose decitabine targeting DNA hypermethylation in patients with chronic lymphocytic leukaemia and non-Hodgkin lymphoma: dose-limiting myelosuppression without evidence of DNA hypomethylation. Br J Haematol 2010; 150: 189–195. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Kuendgen A, Müller-Thomas C, Lauseker M, et al. Efficacy of azacitidine is independent of molecular and clinical characteristics-an analysis of 128 patients with myelodysplastic syndromes or acute myeloid leukemia and a review of the literature. Oncotarget 2018; 9: 27882–27894. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Hill B, Jagadeesh D, Pohlman B, et al. A pilot clinical trial of oral tetrahydrouridine/decitabine for noncytotoxic epigenetic therapy of chemoresistant lymphoid malignancies. Semin Hematol 2021; 58: 35–44. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Aza-SAHA-GBM with AutoSCT for refractory lymphoma – study results. ClinicalTrials.gov, https://clinicaltrials.gov/ct2/show/results/NCT01983969 (accessed 16 January 2023).
- 79.Epigenetic modulation in relapsed/refractory follicular lymphoma and marginal zone lymphoma – study results. ClinicalTrials.gov, https://clinicaltrials.gov/ct2/show/results/NCT01121757 (accessed 16 January 2023).
- 80.Study of 5-azacitidine in combination with vorinostat in patients with relapsed or refractory diffuse large B cell lymphoma (DLBCL) – study results. ClinicalTrials.gov, https://clinicaltrials.gov/ct2/show/results/NCT01120834 (accessed 16 January 2023).
- 81.Phase I/II trial of R-CHOP + azacytidine in diffuse large B cell lymphoma – study results. ClinicalTrials.gov, https://clinicaltrials.gov/ct2/show/results/NCT01004991 (accessed 16 January 2023).
- 82.Martin P, Bartlett NL, Chavez JC, et al. Phase 1 study of oral azacitidine (CC-486) plus R-CHOP in previously untreated intermediate- to high-risk DLBCL. Blood 2022; 139: 1147–1159. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Falchi L, Ma H, Klein S, et al. Combined oral 5-azacytidine and romidepsin are highly effective in patients with PTCL: a multicenter phase 2 study. Blood 2021; 137: 2161–2170. [DOI] [PubMed] [Google Scholar]
- 84.O’Connor OA, Falchi L, Lue JK, et al. Oral 5-azacytidine and romidepsin exhibit marked activity in patients with PTCL: a multicenter phase 1 study. Blood 2019; 134: 1395–1405. [DOI] [PubMed] [Google Scholar]
- 85.Hawkes EA, Phillips T, Budde LE, et al. Avelumab in combination regimens for relapsed/refractory DLBCL: results from the phase Ib JAVELIN DLBCL study. Target Oncol 2021; 16: 761–771. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Lowe EJ, Sposto R, Perkins SL, et al. Intensive chemotherapy for systemic anaplastic large cell lymphoma in children and adolescents: final results of children’s cancer group study 5941. Pediatr Blood Cancer 2009; 52: 335–339. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Foss F, Advani R, Duvic M, et al. A phase II trial of belinostat (PXD101) in patients with relapsed or refractory peripheral or cutaneous T-cell lymphoma. Br J Haematol 2015; 168: 811–819. [DOI] [PubMed] [Google Scholar]
- 88.Johnston PB, Cashen AF, Nikolinakos PG, et al. Belinostat in combination with standard cyclophosphamide, doxorubicin, vincristine and prednisone as first-line treatment for patients with newly diagnosed peripheral T-cell lymphoma. Exp Hematol Oncol 2021; 10: 15. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Campbell P, Thomas CM. Belinostat for the treatment of relapsed or refractory peripheral T-cell lymphoma. J Oncol Pharm Pr 2017; 23: 143–147. [DOI] [PubMed] [Google Scholar]
- 90.Tan D, Phipps C, Hwang WYK, et al. Panobinostat in combination with bortezomib in patients with relapsed or refractory peripheral T-cell lymphoma: an open-label, multicentre phase 2 trial. Lancet Haematol 2015; 2: e326–e333. [DOI] [PubMed] [Google Scholar]
- 91.Duvic M, Dummer R, Becker JC, et al. Panobinostat activity in both bexarotene-exposed and -naïve patients with refractory cutaneous T-cell lymphoma: results of a phase II trial. Eur J Cancer 2013; 49: 386–394. [DOI] [PubMed] [Google Scholar]
- 92.A phase III randomized, double blind, placebo controlled multi-center study of panobinostat for maintenance of response in patients with Hodgkin’s Lymphoma (HL) – study results. ClinicalTrials.gov, https://clinicaltrials.gov/ct2/show/results/NCT01034163 (accessed 16 January 2023).
- 93.Oki Y, Buglio D, Fanale M, et al. Phase I study of panobinostat plus everolimus in patients with relapsed or refractory lymphoma. Clin Cancer Res 2013; 19: 6882–6890. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Drott K, Hagberg H, Papworth K, et al. Valproate in combination with rituximab and CHOP as first-line therapy in diffuse large B-cell lymphoma (VALFRID). Blood Adv 2018; 2: 1386–1392. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Evens AM, Balasubramanian S, Vose JM, et al. A phase I/II multicenter, open-label study of the oral histone deacetylase inhibitor abexinostat in relapsed/refractory lymphoma. Clin Cancer Res 2016; 22: 1059–1066. [DOI] [PubMed] [Google Scholar]
- 96.Ribrag V, Kim WS, Bouabdallah R, et al. Safety and efficacy of abexinostat, a pan-histone deacetylase inhibitor, in non-Hodgkin lymphoma and chronic lymphocytic leukemia: results of a phase II study. Haematologica 2017; 102: 903–909. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Younes A, Oki Y, Bociek RG, et al. Mocetinostat for relapsed classical Hodgkin’s lymphoma: an open-label, single-arm, phase 2 trial. Lancet Oncol 2011; 12: 1222–1228. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Younes A, Berdeja JG, Patel MR, et al. Safety, tolerability, and preliminary activity of CUDC-907, a first-in-class, oral, dual inhibitor of HDAC and PI3K, in patients with relapsed or refractory lymphoma or multiple myeloma: an open-label, dose-escalation, phase 1 trial. Lancet Oncol 2016; 17: 622–631. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Budde LE, Zhang MM, Shustov AR, et al. A phase I study of pulse high-dose vorinostat (V) plus rituximab (R), ifosphamide, carboplatin, and etoposide (ICE) in patients with relapsed lymphoma. Br J Haematol 2013; 161: 183–191. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Amengual JE, Clark-Garvey S, Kalac M, et al. Sirtuin and pan-class I/II deacetylase (DAC) inhibition is synergistic in preclinical models and clinical studies of lymphoma. Blood 2013; 122: 2104–2113. [DOI] [PubMed] [Google Scholar]
- 101.Ramos JC, Sparano JA, Chadburn A, et al. Impact of Myc in HIV-associated non-Hodgkin lymphomas treated with EPOCH and outcomes with vorinostat (AMC-075 trial). Blood 2020; 136: 1284–1297. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Shadman M, Gopal AK, Press O, et al. Fludarabine, cyclophosphamide, rituximab (FCR) and vorinostat followed by rituximab and vorinostat maintenance therapy in patients with previously untreated chronic lymphocytic leukemia (CLL) or small lymphocytic leukemia (SLL) – final results of a phase. Blood 2016; 128: 4396. [Google Scholar]
- 103.Spurgeon SE, Sharma K, Claxton DF, et al. Phase 1–2 study of vorinostat (SAHA), cladribine and rituximab (SCR) in relapsed B-cell non-Hodgkin lymphoma and previously untreated mantle cell lymphoma. Br J Haematol 2019; 186: 845–854. [DOI] [PubMed] [Google Scholar]
- 104.Holmberg LA, Green D, Libby E, et al. Bortezomib and vorinostat therapy as maintenance therapy after autologous transplant for multiple myeloma. Acta Haematol 2020; 143: 146–154. [DOI] [PubMed] [Google Scholar]
- 105.Kirschbaum M, Frankel P, Popplewell L, et al. Phase II study of vorinostat for treatment of relapsed or refractory indolent non-Hodgkin’s lymphoma and mantle cell lymphoma. J Clin Oncol 2011; 29: 1198–1203. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Ogura M, Ando K, Suzuki T, et al. A multicentre phase II study of vorinostat in patients with relapsed or refractory indolent B-cell non-Hodgkin lymphoma and mantle cell lymphoma. Br J Haematol 2014; 165: 768–776. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107.Straus DJ, Hamlin PA, Matasar MJ, et al. Phase I/II trial of vorinostat with rituximab, cyclophosphamide, etoposide and prednisone as palliative treatment for elderly patients with relapsed or refractory diffuse large B-cell lymphoma not eligible for autologous stem cell transplantation. Br J Haematol 2015; 168: 663–670. [DOI] [PubMed] [Google Scholar]
- 108.Chen R, Frankel P, Popplewell L, et al. A phase II study of vorinostat and rituximab for treatment of newly diagnosed and relapsed/refractory indolent non-Hodgkin lymphoma. Haematologica 2015; 100: 357–362. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Shi Y, Jia B, Xu W, et al. Chidamide in relapsed or refractory peripheral T cell lymphoma: a multicenter real-world study in China. J Hematol Oncol 2017; 10: 69. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Piekarz RL, Frye R, Prince HM, et al. Phase 2 trial of romidepsin in patients with peripheral T-cell lymphoma. Blood 2011; 117: 5827–5834. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Strati P, Nastoupil LJ, Davis RE, et al. A phase 1 trial of alisertib and romidepsin for relapsed/refractory aggressive B-cell and T-cell lymphomas. Haematologica 2020; 105: e26–e28. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Duvic M, Bates SE, Piekarz R, et al. Responses to romidepsin in patients with cutaneous T-cell lymphoma and prior treatment with systemic chemotherapy. Leuk Lymphoma 2018; 59: 880–887. [DOI] [PubMed] [Google Scholar]
- 113.Maruyama D, Tobinai K, Ogura M, et al. Romidepsin in Japanese patients with relapsed or refractory peripheral T-cell lymphoma: a phase I/II and pharmacokinetics study. Int J Hematol 2017; 106: 655–665. [DOI] [PubMed] [Google Scholar]
- 114.Foss F, Pro B, Miles Prince H, et al. Responses to romidepsin by line of therapy in patients with relapsed or refractory peripheral T-cell lymphoma. Cancer Med 2017; 6: 36–44. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Pellegrini C, Dodero A, Chiappella A, et al. A phase II study on the role of gemcitabine plus romidepsin (GEMRO regimen) in the treatment of relapsed/refractory peripheral T-cell lymphoma patients. J Hematol Oncol 2016; 9: 38. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116.Batlevi CL, Kasamon Y, Bociek RG, et al. ENGAGE-501: phase II study of entinostat (SNDX-275) in relapsed and refractory Hodgkin lymphoma. Haematologica 2016; 101: 968–975. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.Amengual JE, Lue JK, Ma H, et al. First-in-class selective HDAC6 inhibitor (ACY-1215) has a highly favorable safety profile in patients with relapsed and refractory lymphoma. Oncologist 2021; 26: 184-e366. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Munakata W, Shirasugi Y, Tobinai K, et al. Phase 1 study of tazemetostat in Japanese patients with relapsed or refractory B-cell lymphoma. Cancer Sci 2021; 112: 1123–1131. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.Izutsu K, Ando K, Nishikori M, et al. Phase II study of tazemetostat for relapsed or refractory B-cell non-Hodgkin lymphoma with EZH2 mutation in Japan. Cancer Sci 2021; 112: 3627–3635. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120.Palomba ML, Cartron G, Popplewell L, et al. Safety and clinical activity of atezolizumab in combination with tazemetostat in relapsed or refractory diffuse large B-cell lymphoma: primary analysis of a phase 1B study. Hematol Oncol 2019; 37: 517–519. [Google Scholar]
- 121.Argon EK, Nguyen F, Rajarethinam A, et al. Abstract CT170: phase 1 study to characterize the pharmacokinetics and safety of tazemetostat in b-cell lymphomas or advanced solid tumors. Cancer Res 2020; 80: CT170. [Google Scholar]
- 122.Open-label, multicenter, two-part, phase 1 study to characterize effects of a moderate CYP3A inhibitor on PK of tazemetostat, effects of tazemetostat on PK of CYP2C8 and CYP2C19 substrates, and effect of increased gastric pH on PK of tazemetostat in B-cell lymphoma or advanced solid tumor patients – study results. ClinicalTrials.gov, https://clinicaltrials.gov/ct2/show/results/NCT03028103 (accessed 16 January 2023).
- 123.Morschhauser F, Tilly H, Chaidos A, et al. Tazemetostat for patients with relapsed or refractory follicular lymphoma: an open-label, single-arm, multicentre, phase 2 trial. Lancet Oncol 2020; 21: 1433–1442. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Yap TA, Winter JN, Giulino-Roth L, et al. Phase I study of the novel enhancer of zeste homolog 2 (EZH2) inhibitor GSK2816126 in patients with advanced hematologic and solid tumors. Clin Cancer Res 2019; 25: 7331–7339. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Amorim S, Stathis A, Gleeson M, et al. Bromodomain inhibitor OTX015 in patients with lymphoma or multiple myeloma: a dose-escalation, open-label, pharmacokinetic, phase 1 study. Lancet Haematol 2016; 3: e196–e204. [DOI] [PubMed] [Google Scholar]
- 126.Dickinson M, Briones J, Herrera AF, et al. Phase 1b study of the BET protein inhibitor RO6870810 with venetoclax and rituximab in patients with diffuse large B-cell lymphoma. Blood Adv 2021; 5: 4762–4770. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127.Falchook G, Rosen S, LoRusso P, et al. Development of 2 bromodomain and extraterminal inhibitors with distinct pharmacokinetic and pharmacodynamic profiles for the treatment of advanced malignancies. Clin Cancer Res 2020; 26: 1247–1257. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128.Blum KA, Abramson J, Maris M, et al. 41OA phase I study of CPI-0610, a bromodomain and extra terminal protein (BET) inhibitor in patients with relapsed or refractory lymphoma. Ann Oncol 2018; 29: iii7–iii9. [Google Scholar]
- 129.Patel MR, Garcia-Manero G, Paquette R, et al. Phase 1 dose escalation and expansion study to determine safety, tolerability, pharmacokinetics, and pharmacodynamics of the BET inhibitor FT-1101 as a single agent in patients with relapsed or refractory hematologic malignancies. Blood 2019; 134: 3907. [Google Scholar]
- 130.Marchi E, Zullo KM, Amengual JE, et al. The combination of hypomethylating agents and histone deacetylase inhibitors produce marked synergy in preclinical models of T-cell lymphoma. Br J Haematol 2015; 171: 215–226. [DOI] [PubMed] [Google Scholar]
- 131.Ebrahem Q, Mahfouz R, Ng KP, et al. High cytidine deaminase expression in the liver provides sanctuary for cancer cells from decitabine treatment effects. Oncotarget 2012; 3: 1137–1145. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132.Liu Z, Liu S, Xie Z, et al. Characterization of in vitro and in vivo hypomethylating effects of decitabine in acute myeloid leukemia by a rapid, specific and sensitive LC-MS/MS method. Nucleic Acids Res 2007; 35: e31. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133.Zhang M, Xu P, Cheng S, et al. Decitabine plus R-chop in patients with newly diagnosed diffuse large B-cell lymphoma: interim results of a phase I/II study. Hematol Oncol 2019; S2: 380–382. [Google Scholar]
- 134.Munshi PN, Lubin M, Bertino JR. 6-thioguanine: a drug with unrealized potential for cancer therapy. Oncologist 2014; 19: 760–765. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 135.Reiter A, Schrappe M, Ludwig WD, et al. Intensive ALL-type therapy without local radiotherapy provides a 90% event-free survival for children with T-cell lymphoblastic lymphoma: a BFM group report. Blood 2000; 95: 416–421. [PubMed] [Google Scholar]
- 136.Landmann E, Burkhardt B, Zimmermann M, et al. Results and conclusions of the European intergroup EURO-LB02 trial in children and adolescents with lymphoblastic lymphoma. Haematologica 2017; 102: 2086–2096. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137.Zhang Q, Wang S, Chen J, et al. Histone deacetylases (HDACs) guided novel therapies for T-cell lymphomas. Int J Med Sci 2019; 16: 424–442. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 138.Wang L, Qin W, Huo YJ, et al. Advances in targeted therapy for malignant lymphoma. Signal Transduct Target Ther 2020; 5: 15. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 139.Marks PA, Breslow R. Dimethyl sulfoxide to vorinostat: development of this histone deacetylase inhibitor as an anticancer drug. Nat Biotechnol 2007; 25: 84–90. [DOI] [PubMed] [Google Scholar]
- 140.Kirschbaum M, Frankel P, Popplewell L, et al. Phase II study of vorinostat for treatment of relapsed or refractory indolent non-Hodgkin’s lymphoma and mantle cell lymphoma. J Clin Oncol 2011; 29: 1198–1203. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 141.Duvic M, Vu J. Vorinostat: a new oral histone deacetylase inhibitor approved for cutaneous T-cell lymphoma. Expert Opin Investig Drugs 2007; 16: 1111–1120. [DOI] [PubMed] [Google Scholar]
- 142.Spurgeon SE, Sharma K, Claxton DF, et al. Phase 1–2 study of vorinostat (SAHA), cladribine and rituximab (SCR) in relapsed B-cell non-Hodgkin lymphoma and previously untreated mantle cell lymphoma. Br J Haematol 2019; 186: 845–854. [DOI] [PubMed] [Google Scholar]
- 143.Shadman M, Gopal AK, Press O, et al. 642. CLL: therapy, excluding Transplantation: poster III. Fludarabine, cyclophosphamide, rituximab (FCR) and vorinostat followed by rituximab and vorinostat maintenance therapy in patients with previously untreated chronic lymphocytic leukemia (CLL) or SMA. Blood 2016; 128: 4396. [Google Scholar]
- 144.Ramos JC, Sparano JA, Chadburn A, et al. Impact of Myc in HIV-associated non-Hodgkin lymphomas treated with EPOCH and outcomes with vorinostat (AMC-075 trial). Blood 2020; 136: 1284–1297. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145.Amengual JE, Lue JK, Ma H, et al. First-in-class selective HDAC6 inhibitor (ACY-1215) has a highly favorable safety profile in patients with relapsed and refractory lymphoma. Oncologist 2021; 26: 184-e366. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 146.Boumber Y, Younes A, Garcia-Manero G. Mocetinostat (MGCD0103): a review of an isotype-specific histone deacetylase inhibitor. Exp Opin Investig Drugs 2011; 20: 823–829. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 147.Chowdhury S, Howell GM, Teggart CA, et al. Histone deacetylase inhibitor belinostat represses survivin expression through reactivation of transforming growth factor beta (TGFbeta) receptor II leading to cancer cell death. J Biol Chem 2011; 286: 30937–30948. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148.Poole RM. Belinostat: first global approval. Drugs 2014; 74: 1543–1554. [DOI] [PubMed] [Google Scholar]
- 149.Johnston PB, Cashen AF, Nikolinakos PG, et al. Belinostat in combination with standard cyclophosphamide, doxorubicin, vincristine and prednisone as first-line treatment for patients with newly diagnosed peripheral T-cell lymphoma. Exp Hematol Oncol 2021; 10: 15. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 150.Shi Y, Jia B, Xu W, et al. Chidamide in relapsed or refractory peripheral T cell lymphoma: a multicenter real-world study in China. J Hematol Oncol 2017; 10: 69. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 151.Yan G, Li P, Bing B, et al. 624. Hodgkin lymphoma and T/NK cell lymphoma-clinical studies: poster II. Chidamide, oral subtype-selective histone deacetylase inhibitor (HDACI) monotherapy was effective on the patients with relapsed or refractory extranodal natural killer (NK)/T-cell L. Blood 2017; 130: 2797. [Google Scholar]
- 152.Santo L, Hideshima T, Kung AL, et al. Preclinical activity, pharmacodynamic, and pharmacokinetic properties of a selective HDAC6 inhibitor, ACY-1215, in combination with bortezomib in multiple myeloma. Blood 2012; 119: 2579–2589. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 153.Oki Y, Kelly KR, Flinn I, et al. CUDC-907 in relapsed/refractory diffuse large B-cell lymphoma, including patients with MYC-alterations: results from an expanded phase I trial. Haematologica 2017; 102: 1923–1930. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 154.Ueda H, Nakajima H, Hori Y, et al. Action of FR901228, a novel antitumor bicyclic depsipeptide produced by Chromobacterium violaceum no. 968, on Ha-ras transformed NIH3T3 cells. Biosci Biotechnol Biochem 1994; 58: 1579–1583. [DOI] [PubMed] [Google Scholar]
- 155.Reddy SA. Romidepsin for the treatment of relapsed/refractory cutaneous T-cell lymphoma (mycosis fungoides/Sezary syndrome): use in a community setting. Crit Rev Oncol Hematol 2016; 106: 99–107. [DOI] [PubMed] [Google Scholar]
- 156.Maruyama D, Tobinai K, Ogura M, et al. Romidepsin in Japanese patients with relapsed or refractory peripheral T-cell lymphoma: a phase I/II and pharmacokinetics study. Int J Hematol 2017; 106: 655–665. [DOI] [PubMed] [Google Scholar]
- 157.Whittaker SJ, Demierre M-F, Kim EJ, et al. Final results from a multicenter, international, pivotal study of romidepsin in refractory cutaneous T-cell lymphoma. J Clin Oncol 2010; 28: 4485–4491. [DOI] [PubMed] [Google Scholar]
- 158.Piekarz RL, Frye R, Prince HM, et al. Phase 2 trial of romidepsin in patients with peripheral T-cell lymphoma. Blood 2011; 117: 5827–5834. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 159.Falchi L, Ma H, Klein S, et al. Combined oral 5-azacytidine and romidepsin are highly effective in patients with PTCL: a multicenter phase 2 study. Blood 2021; 137: 2161–2170. [DOI] [PubMed] [Google Scholar]
- 160.Pellegrini C, Dodero A, Chiappella A, et al. A phase II study on the role of gemcitabine plus romidepsin (GEMRO regimen) in the treatment of relapsed/refractory peripheral T-cell lymphoma patients. J Hematol Oncol 2016; 9: 38. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 161.Strati P, Nastoupil LJ, Davis RE, et al. A phase 1 trial of alisertib and romidepsin for relapsed/refractory aggressive B-cell and T-cell lymphomas. Haematologica 2020; 105: e26–e28. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 162.https://www.clinicaltrialsregister.eu/ctr-search/trial/2009-014846-26/results
- 163.Zaja F, Salvi F, Rossi M, et al. Single-agent panobinostat for relapsed/refractory diffuse large B-cell lymphoma: clinical outcome and correlation with genomic data. A phase 2 study of the Fondazione Italiana Linfomi. Leuk Lymphoma 2018; 59: 2904–2910. [DOI] [PubMed] [Google Scholar]
- 164.Panobinostat plus ifosfamide, carboplatin, and etoposide (ICE) compared with ICE for relapsed Hodgkin lymphoma – study results. ClinicalTrials.gov, https://clinicaltrials.gov/ct2/show/results/NCT01169636 (accessed 16 January 2023).
- 165.Duvic M, Dummer R, Becker JC, et al. Panobinostat activity in both bexarotene-exposed and -naive patients with refractory cutaneous T-cell lymphoma: results of a phase II trial. Eur J Cancer 2013; 49: 386–394. [DOI] [PubMed] [Google Scholar]
- 166.Maly JJ, Christian BA, Zhu X, et al. A phase I/II trial of panobinostat in combination with lenalidomide in patients with relapsed or refractory Hodgkin lymphoma. Clin Lymphoma Myeloma Leuk 2017; 17: 347. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 167.Tan D, Phipps C, Hwang WY, et al. Panobinostat in combination with bortezomib in patients with relapsed or refractory peripheral T-cell lymphoma: an open-label, multicentre phase 2 trial. Lancet Haematol 2015; 2: e326–e333. [DOI] [PubMed] [Google Scholar]
- 168.Varambally S, Dhanasekaran SM, Zhou M, et al. The polycomb group protein EZH2 is involved in progression of prostate cancer. Nature 419: 624–629. [DOI] [PubMed] [Google Scholar]
- 169.Kleer CG, Cao Q, Varambally S, et al. EZH2 is a marker of aggressive breast cancer and promotes neoplastic transformation of breast epithelial cells. Proc Natl Acad Sci USA 2003; 100: 11606–11611. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 170.Visser HPJ, Gunster MJ. The polycomb group protein EZH2 is upregulated in proliferating, cultured human mantle cell lymphoma. Br J Haematol 2001; 112: 950–958. [DOI] [PubMed] [Google Scholar]
- 171.Abd Al, Kader L, Oka T, Takata K, et al. In aggressive variants of non-Hodgkin lymphomas, Ezh2 is strongly expressed and polycomb repressive complex PRC1.4 dominates over PRC1.2. Virchows Arch 2013; 463: 697–711. [DOI] [PubMed] [Google Scholar]
- 172.Morin RD, Johnson NA, Severson TM, et al. Somatic mutations altering EZH2 (Tyr641) in follicular and diffuse large B-cell lymphomas of germinal-center origin. Nat Genet 2010; 42: 181–185. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 173.Zhang X, Zhao X, Fiskus W, et al. Coordinated silencing of MYC-mediated miR-29 by HDAC3 and EZH2 as a therapeutic target of histone modification in aggressive B-cell lymphomas. Cancer Cell 2012; 22: 506–523. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
- 174.Ryan RJH, Nitta M, Borger D, et al. EZH2 codon 641 mutations are common in BCL2-rearranged germinal center B cell lymphomas. PLoS ONE 2011; 6: 1–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 175.Zhao X, Lwin T, Zhang X, et al. Disruption of the MYC-miRNA-EZH2 loop to suppress aggressive B-cell lymphoma survival and clonogenicity. Leukemia 2013; 27: 2341–2350. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 176.Lue JK, Prabhu SA, Liu Y, et al. Precision targeting with EZH2 and HDAC inhibitors in epigenetically dysregulated lymphomas. Clin Cancer Res 2019; 25: 5271–5283. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 177.McCabe MT, Ott HM, Ganji G, et al. EZH2 inhibition as a therapeutic strategy for lymphoma with EZH2-activating mutations. Nature 2012; 492: 108–112. [DOI] [PubMed] [Google Scholar]
- 178.Vaswani RG, Gehling VS, Dakin LA, et al. Identification of (R)-N-((4-Methoxy-6-methyl-2-oxo-1,2-dihydropyridin-3-yl)methyl)-2-methyl-1-(1-(1-(2,2,2-trifluoroethyl)piperidin-4-yl)ethyl)-1H-indole-3-carboxamide (CPI-1205), a potent and selective inhibitor of histone methyltransferase EZH2, suitable for phase I clinical trials for B-cell lymphomas. J Med Chem 2016; 59: 9928–9941. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 179.Whittaker S. Global patterns of methylation in sézary syndrome provide insight into the role of epigenetics in cutaneous T-Cell lymphoma. J Invest Dermatol 2016; 136: 1753–1754. [DOI] [PubMed] [Google Scholar]
- 180.Maruyama D, Tobinai K, Makita S, et al. First-in-human study of the EZH1/2 dual inhibitor DS-3201b in patients with relapsed or refractory non-Hodgkin lymphomas – preliminary results. Blood 2017; 130: 4070–4070. [Google Scholar]
- 181.Knutson SK, Kawano S, Minoshima Y, et al. Selective inhibition of EZH2 by EPZ-6438 leads to potent antitumor activity in EZH2-mutant non-Hodgkin lymphoma. Mol Cancer Ther 2014; 13: 842–854. [DOI] [PubMed] [Google Scholar]
- 182.Italiano A, Soria J, Toulmonde M, et al. Tazemetostat, an EZH2 inhibitor, in relapsed or refractory B-cell non-Hodgkin lymphoma and advanced solid tumours: a first-in-human, open-label, phase 1 study. Lancet Oncol 2018; 19: 649–659. [DOI] [PubMed] [Google Scholar]
- 183.Akpa CA, Kleo K, Lenze D, et al. DZNep-mediated apoptosis in B-cell lymphoma is independent of the lymphoma type, EZH2 mutation status and MYC, BCL2 or BCL6 translocations. PLoS ONE 2019; 14: e0220681. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 184.Wyczechowska D, Fabianowska-Majewska K. The effects of cladribine and fludarabine on DNA methylation in K562 cells. Biochem Pharmacol 2003; 65: 219–225. [DOI] [PubMed] [Google Scholar]
- 185.Sigal DS, Miller HJ, Schram ED, et al. Beyond hairy cell: the activity of cladribine in other hematologic malignancies. Blood 2010; 116: 2884–2896. [DOI] [PubMed] [Google Scholar]
- 186.Kluin-Nelemans HC, Hoster E, Hermine O, et al. Treatment of older patients with mantle-cell lymphoma. N Engl J Med 2012; 367: 520–531. [DOI] [PubMed] [Google Scholar]
- 187.Padmanabhan B, Mathur S, Manjula R, et al. Bromodomain and extra-terminal (BET) family proteins: new therapeutic targets in major diseases. J Biosci 2016; 41: 295–311. [DOI] [PubMed] [Google Scholar]
- 188.Delmore JE, Issa GC, Lemieux ME, et al. BET bromodomain inhibition as a therapeutic strategy to target c-Myc. Cell 2011; 146: 904–917. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 189.Lovén J, Hoke HA, Lin CY, et al. Selective inhibition of tumor oncogenes by disruption of super-enhancers. Cell 2013; 153: 320–334. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 190.Yang Z, Yik JHN, Chen R, et al. Recruitment of P-TEFb for stimulation of transcriptional elongation by the bromodomain protein Brd4. Mol Cell 2005; 19: 535–545. [DOI] [PubMed] [Google Scholar]
- 191.Johnson-Farley N, Veliz J, Bhagavathi S, et al. ABT-199, a BH3 mimetic that specifically targets Bcl-2, enhances the antitumor activity of chemotherapy, bortezomib and JQ1 in ‘double hit’ lymphoma cells. Leuk Lymphoma 2015; 56: 2146–2152. [DOI] [PubMed] [Google Scholar]
- 192.Feng Z, Yao Y, Zhou C, et al. Pharmacological inhibition of LSD1 for the treatment of MLL-rearranged leukemia. J Hematol Oncol 2016; 9: 24. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 193.Mondello P, Tadros S, Teater M, et al. Selective inhibition of HDAC3 targets synthetic vulnerabilities and activates immune surveillance in lymphoma. Cancer Discov 2020; 10: 440–459. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 194.Yang HH, Lee MP. Application of bioinformatics in cancer epigenetics. Ann N Y Acad Sci 2004; 1020: 67–76. [DOI] [PubMed] [Google Scholar]
- 195.Li J, Ching T, Huang S, et al. Using epigenomics data to predict gene expression in lung cancer. BMC Bioinformatics 2015; 16(Suppl. 5): S10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 196.Mai A, Altucci L. Epi-drugs to fight cancer: from chemistry to cancer treatment, the road ahead. Int J Biochem Cell Biol 2009; 41: 199–213. [DOI] [PubMed] [Google Scholar]
- 197.Springer NM. Epigenetics and crop improvement. Trends Genet 2013; 29: 241–247. [DOI] [PubMed] [Google Scholar]
- 198.Mullard A. The roadmap epigenomics project opens new drug development avenues. Nat Rev Drug Discov 2015; 14: 223–225. [DOI] [PubMed] [Google Scholar]
- 199.Schumacher A, Kapranov P, Kaminsky Z, et al. Microarray-based DNA methylation profiling: technology and applications. Nucleic Acids Res 2006; 34: 528–542. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 200.Booth MJ, Branco MR, Ficz G, et al. Quantitative sequencing of 5-methylcytosine and 5-hydroxymethylcytosine at single-base resolution. Science 2012; 336: 934–937. [DOI] [PubMed] [Google Scholar]
- 201.Furey TS. ChIP–seq and beyond: new and improved methodologies to detect and characterize protein–DNA interactions. Nat Rev Genet 2012; 13: 840–852. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 202.Gautam B, Goswami K, Mishra NS, et al. The role of bioinformatics in epigenetics. In: Wadhwa G, Shanmughavel P, Singh AK, et al. (eds) Current trends in bioinformatics: an insight. Singapore: Springer Singapore, 2018, pp. 39–53. [Google Scholar]
- 203.Pettini F, Visibelli A, Cicaloni V, et al. Multi-omics model applied to cancer genetics. Int J Mol Sci 2021; 22: 5751. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 204.Meir Z, Mukamel Z, Chomsky E, et al. Single-cell analysis of clonal maintenance of transcriptional and epigenetic states in cancer cells. Nat Genet 2020; 52: 709–718. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 205.Montalvo-Casimiro M, González-Barrios R, Meraz-Rodriguez MA, et al. Epidrug repurposing: discovering new faces of old acquaintances in cancer therapy. Front Oncol 2020; 10: 605386. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 206.Liu J, Li JN, Wu H, et al. The status and prospects of epigenetics in the treatment of lymphoma. Front Oncol 2022; 12: 874645–874623. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 207.Bock C, Lengauer T. Computational epigenetics. Bioinformatics 2008; 24: 1–10. [DOI] [PubMed] [Google Scholar]
- 208.Toedling J, Sklyar O, Huber W. Ringo – an R/Bioconductor package for analyzing ChIP-chip readouts. BMC Bioinform 2007; 8: 221. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 209.Zhang ZD, Rozowsky J, Lam HY, et al. Tilescope: online analysis pipeline for high-density tiling microarray data. Genome Biol 2007; 8: R81. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 210.Li W, Meyer CA, Liu XS. A hidden Markov model for analyzing ChIP-chip experiments on genome tiling arrays and its application to p53 binding sequences. Bioinformatics 2005; 21(Suppl. 1): i274–i282. [DOI] [PubMed] [Google Scholar]
- 211.Langmead B, Trapnell C, Pop M, et al. Ultrafast and memory-efficient alignment of short DNA sequences to the human genome. Genome Biol 2009; 10: R25. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 212.Kent WJ. BLAT – the BLAST-like alignment tool. Genome Res 2002; 12: 656–664. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 213.Hebenstreit D, Gu M, Haider S, et al. EpiChIP: gene-by-gene quantification of epigenetic modification levels. Nucleic Acids Res 2011; 39: e27–e27. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 214.Mikkelsen TS, Ku M, Jaffe DB, et al. Genome-wide maps of chromatin state in pluripotent and lineage-committed cells. Nature 2007; 448: 553–560. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 215.Li H, Handsaker B, Wysoker A, et al. The sequence alignment/map format and SAMtools. Bioinformatics 2009; 25: 2078–2079. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 216.Bardet AF, Steinmann J, Bafna S, et al. Identification of transcription factor binding sites from ChIP-seq data at high resolution. Bioinformatics 2013; 29: 2705–2713. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 217.Zhang Y, Liu T, Meyer CA, et al. Model-based analysis of ChIP-seq (MACS). Genome Biol 2008; 9: R137. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 218.Ibrahim MM, Lacadie SA, Ohler U. JAMM: a peak finder for joint analysis of NGS replicates. Bioinformatics 2015; 31: 48–55. [DOI] [PubMed] [Google Scholar]
- 219.Wang C, Zhang L, Ke L, et al. Primary effusion lymphoma enhancer connectome links super-enhancers to dependency factors. Nat Commun 2020; 11: 6318. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 220.Xue K, Wu J-C, Li X-Y, et al. Chidamide triggers BTG1-mediated autophagy and reverses the chemotherapy resistance in the relapsed/refractory B-cell lymphoma. Cell Death Dis 2021; 12: 900. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 221.Li K, Wang F, Yang Z, et al. TRIB3 promotes MYC-associated lymphoma development through suppression of UBE3B-mediated MYC degradation. Nat Commun 2020; 11: 6316. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 222.Horn DM, Zubarev RA, McLafferty FW. Automated reduction and interpretation of. J Am Soc Mass Spectrom 2000; 11: 320–332. [DOI] [PubMed] [Google Scholar]
- 223.Lue JK, Prabhu SA, Liu Y, et al. Precision targeting with EZH2 and HDAC inhibitors in epigenetically dysregulated lymphomas. Clin Cancer Res 2019; 25: 5271–5283. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 224.Bibikova M, Barnes B, Tsan C, et al. High density DNA methylation array with single CpG site resolution. Genomics 2011; 98: 288–295. [DOI] [PubMed] [Google Scholar]
- 225.Down TA, Rakyan VK, Turner DJ, et al. A Bayesian deconvolution strategy for immunoprecipitation-based DNA methylome analysis. Nat Biotechnol 2008; 26: 779–785. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 226.Aberg KA, Xie L, Chan RF, et al. Evaluation of methyl-binding domain based enrichment approaches revisited. PLoS ONE 2015; 10: e0132205. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 227.Xi Y, Bock C, Müller F, et al. RRBSMAP: a fast, accurate and user-friendly alignment tool for reduced representation bisulfite sequencing. Bioinformatics 2012; 28: 430–432. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 228.Otto C, Stadler PF, Hoffmann S. Fast and sensitive mapping of bisulfite-treated sequencing data. Bioinformatics 2012; 28: 1698–1704. [DOI] [PubMed] [Google Scholar]
- 229.Pedersen B, Hsieh T-F, Ibarra C, et al. MethylCoder: software pipeline for bisulfite-treated sequences. Bioinformatics 2011; 27: 2435–2436. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 230.Harris EY, Ounit R, Lonardi S. BRAT-nova: fast and accurate mapping of bisulfite-treated reads. Bioinformatics 2016; 32: 2696–2698. [DOI] [PubMed] [Google Scholar]
- 231.Liu Y, Siegmund KD, Laird PW, et al. Bis-SNP: combined DNA methylation and SNP calling for Bisulfite-seq data. Genome Biol 2012; 13: R61. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 232.Barturen G, Rueda A, Oliver JL, et al. MethylExtract: high-quality methylation maps and SNV calling from whole genome bisulfite sequencing data. F1000Res 2014; 2: 217. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 233.Küpers LK, Monnereau C, Sharp GC, et al. Meta-analysis of epigenome-wide association studies in neonates reveals widespread differential DNA methylation associated with birthweight. Nat Commun 2019; 10: 1893. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 234.Fingerman IM, Zhang X, Ratzat W, et al. NCBI epigenomics: what’s new for 2013. Nucleic Acids Res 2012; 41: D221–D225. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 235.Dunham I, Kundaje A, Aldred SF, et al. An integrated encyclopedia of DNA elements in the human genome. Nature 2012; 489: 57–74. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 236.Grunau C. MethDB – a public database for DNA methylation data. Nucleic Acids Res 2001; 29: 270–274. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 237.Li R, Liang F, Li M, et al. MethBank 3.0: a database of DNA methylomes across a variety of species. Nucleic Acids Res 2018; 46: D288–D295. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 238.O’Connor TR, Wyrick JJ. ChromatinDB: a database of genome-wide histone modification patterns for Saccharomyces cerevisiae. Bioinformatics 2007; 23: 1828–1830. [DOI] [PubMed] [Google Scholar]
- 239.Zhang Y, Lv J, Liu H, et al. HHMD: the human histone modification database. Nucleic Acids Res 2010; 38(Database issue): D149–D154. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 240.Shen L, Shao N-Y, Liu X, et al. diffReps: detecting differential chromatin modification sites from ChIP-seq data with biological replicates. PLoS ONE 2013; 8: e65598. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 241.Hsu C-H, Tomiyasu H, Liao C-H, et al. Genome-wide DNA methylation and RNA-seq analyses identify genes and pathways associated with doxorubicin resistance in a canine diffuse large B-cell lymphoma cell line. PLoS ONE 2021; 16: e0250013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 242.Nadeu F, Martin-Garcia D, Clot G, et al. Genomic and epigenomic insights into the origin, pathogenesis, and clinical behavior of mantle cell lymphoma subtypes. Blood 2020; 136: 1419–1432. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 243.Zaghlool SB, Kühnel B, Elhadad MA, et al. Epigenetics meets proteomics in an epigenome-wide association study with circulating blood plasma protein traits. Nat Commun 2020; 11: 15. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 244.Shen F, Boccuto L, Pauly R, et al. Genome-scale network model of metabolism and histone acetylation reveals metabolic dependencies of histone deacetylase inhibitors. Genome Biol 2019; 20: 49. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 245.Saffie R, Zhou N, Rolland D, et al. FBXW7 triggers degradation of KMT2D to favor growth of diffuse large B-cell lymphoma cells. Cancer Res 2020; 80: 2498–2511. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 246.Evans JM, Parker HG, Rutteman GR, et al. Multi-omics approach identifies germline regulatory variants associated with hematopoietic malignancies in retriever dog breeds. PLoS Genet 2021; 17: e1009543. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 247.Schadt EE, Friend SH, Shaywitz DA. A network view of disease and compound screening. Nat Rev Drug Discov 2009; 8: 286–295. [DOI] [PubMed] [Google Scholar]
- 248.Cai Z, Poulos RC, Liu J, et al. Machine learning for multi-omics data integration in cancer. iScience 2022; 25: 103798. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 249.Sánchez-Cruz N, Medina-Franco JL. Epigenetic target fishing with accurate machine learning models. J Med Chem 2021; 64: 8208–8220. [DOI] [PubMed] [Google Scholar]
- 250.Langfelder P, Horvath S. WGCNA: an R package for weighted correlation network analysis. BMC Bioinformatics 2008; 9: 559. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 251.Ghobadi MZ, Emamzadeh R, Mozhgani SH. Deciphering microRNA-mRNA regulatory network in adult T-cell leukemia/lymphoma; the battle between oncogenes and anti-oncogenes. PLoS ONE 2021; 16: e0247713. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 252.Singh A, Shannon CP, Gautier B, et al. DIABLO: an integrative approach for identifying key molecular drivers from multi-omics assays. Bioinformatics 2019; 35: 3055–3062. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 253.Lu W, Zhang R, Jiang H, et al. Computer-aided drug design in epigenetics. Front Chem 2018; 6: 57. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 254.Fauman EB, Rai BK, Huang ES. Structure-based druggability assessment – identifying suitable targets for small molecule therapeutics. Curr Opin Chem Biol 2011; 15: 463–468. [DOI] [PubMed] [Google Scholar]
- 255.Chen S, Li L, Chen Y, et al. Identification of novel disruptor of telomeric silencing 1-like (DOT1L) inhibitors through structure-based virtual screening and biological assays. J Chem Inf Model 2016; 56: 527–534. [DOI] [PubMed] [Google Scholar]
- 256.Güner O, Clement O, Kurogi Y. Pharmacophore modeling and three dimensional database searching for drug design using catalyst: recent advances. Curr Med Chem 2004; 11: 2991–3005. [DOI] [PubMed] [Google Scholar]
- 257.Caulfield T, Medina-Franco JL. Molecular dynamics simulations of human DNA methyltransferase 3B with selective inhibitor nanaomycin A. J Struct Biol 2011; 176: 185–191. [DOI] [PubMed] [Google Scholar]
- 258.Chen R, Frankel P, Popplewell L, et al. A phase II study of vorinostat and rituximab for treatment of newly diagnosed and relapsed/refractory indolent non-Hodgkin lymphoma. Haematologica 2015; 100: 357–362. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 259.Assouline SE, Nielsen TH, Yu S, et al. Phase 2 study of panobinostat with or without rituximab in relapsed diffuse large B-cell lymphoma. Blood 2016; 128: 185–194. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 260.Inwards DJ, Fishkin PAS, Hillman DW, et al. Long-term results of the treatment of patients with mantle cell lymphoma with cladribine (2-CDA) alone (95-80-53) or 2-CDA and rituximab (N0189) in the North Central Cancer Treatment Group. Cancer 2008; 113: 108–116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 261.Sarkozy C, Morschhauser F, Dubois S, et al. A LYSA phase Ib study of tazemetostat (EPZ-6438) plus R-CHOP in patients with newly diagnosed diffuse large B-cell lymphoma (DLBCL) with poor prognosis features. Clin Cancer Res 2020; 26: 3145–3153. [DOI] [PubMed] [Google Scholar]


