Next Article in Journal
Is Maintaining Thyroid-Stimulating Hormone Effective in Patients Undergoing Thyroid Lobectomy for Low-Risk Differentiated Thyroid Cancer? A Systematic Review and Meta-Analysis
Next Article in Special Issue
Expression of PD-1, PD-L1 and PD-L2 in Lymphomas in Patients with Pre-Existing Rheumatic Diseases—A Possible Association with High Rheumatoid Arthritis Disease Activity
Previous Article in Journal
A Bout of High-Intensity Interval Training (HIIT) in Children and Adolescents during Acute Cancer Treatment—A Pilot Feasibility Study
Previous Article in Special Issue
PI3Kδ Inhibitors as Immunomodulatory Agents for the Treatment of Lymphoma Patients
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Tumor Immune Microenvironment in Lymphoma: Focus on Epigenetics

by
Daniel J. García-Domínguez
1,2,*,†,
Lourdes Hontecillas-Prieto
1,2,*,†,
Natalia Palazón-Carrión
2,
Carlos Jiménez-Cortegana
1,2,3,
Víctor Sánchez-Margalet
1,‡ and
Luis de la Cruz-Merino
2,‡
1
Clinical Laboratory, Department of Medical Biochemistry and Molecular Biology, School of Medicine, University of Seville, Virgen Macarena University Hospital, 41009 Seville, Spain
2
Oncology Service, Virgen Macarena University Hospital, Department of Medicines, School of Medicine, University of Seville, 41009 Seville, Spain
3
Department of Radiation Oncology, Weill Cornell Medical College, New York, NY 10065, USA
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work as first authors.
These authors contributed equally to this work as senior authors.
Cancers 2022, 14(6), 1469; https://doi.org/10.3390/cancers14061469
Submission received: 31 January 2022 / Revised: 23 February 2022 / Accepted: 11 March 2022 / Published: 13 March 2022
(This article belongs to the Special Issue Tumor Immune Microenvironment in Lymphoproliferative Syndromes)

Abstract

:

Simple Summary

Lymphoma and other cancers have been studied by mainly focusing on malignant cells. However, research findings about the role of the tumor microenvironment in the progression and immunosuppression of cancer, particularly in lymphomas, have increased considerably in recent years, allowing for a better understanding of the disease. Consequently, epigenetic mechanisms that are implicated in the interplay between tumor cells and the different components around them, promoting the survival and progression of the tumor, have been described. This review tries to summarize the complex interplay between lymphoma tumor cells and immune cells as well as the epigenetic alterations that result from this cross-talk, aiming at contributing towards underlining the value of epigenetic modifications as new biomarkers and the use of epigenetic drugs as an interesting therapeutic option.

Abstract

Lymphoma is a neoplasm arising from B or T lymphocytes or natural killer cells characterized by clonal lymphoproliferation. This tumor comprises a diverse and heterogeneous group of malignancies with distinct clinical, histopathological, and molecular characteristics. Despite advances in lymphoma treatment, clinical outcomes of patients with relapsed or refractory disease remain poor. Thus, a deeper understanding of molecular pathogenesis and tumor progression of lymphoma is required. Epigenetic alterations contribute to cancer initiation, progression, and drug resistance. In fact, over the past decade, dysregulation of epigenetic mechanisms has been identified in lymphomas, and the knowledge of the epigenetic aberrations has led to the emergence of the promising epigenetic therapy field in lymphoma tumors. However, epigenetic aberrations in lymphoma not only have been found in tumor cells, but also in cells from the tumor microenvironment, such as immune cells. Whereas the epigenetic dysregulation in lymphoma cells is being intensively investigated, there are limited studies regarding the epigenetic mechanisms that affect the functions of immune cells from the tumor microenvironment in lymphoma. Therefore, this review tries to provide a general overview of epigenetic alterations that affect both lymphoma cells and infiltrating immune cells within the tumor, as well as the epigenetic cross-talk between them.

1. Introduction

Cancer is a heterogeneous group of diseases that develop from normal cells. The successive abnormalities and changes in these normal cells drive their progressive transformation into malignant cells [1].
Although transformational events were traditionally associated with genetic changes [1], we currently know that cancer is caused by both genetic and epigenetic alterations. Each change provides advantageous properties to cancer cells, such as uncontrolled cell proliferation, resistance to cell death, evasion of apoptosis, invasiveness, and metastatic potential [2,3]. Thus, the interest in the epigenetic field has increased over the last decade in a wide variety of tumors. Given the importance of epigenetics in affecting cell function by providing disruptive properties, a better understanding of how epigenetic modifications regulate both tumor cells and the tumor microenvironment (TME) is mandatory. Tumors are a group of malignant cells that constantly interact with the TME, which contains different cell types, such as immune cells [4]. Lymphoma cells have been described to interact with the microenvironment to promote immune evasion [5]. In fact, the reciprocal interaction between lymphoma cells and immune cells (such as T and B cells) plays a key role in cancer initiation and progression [6,7]. In addition, immune cell function in the TME is affected by epigenetic alterations, which contributes to a favorable environment for tumor growth. As a result, tumor cells control immune cell functions through complex signaling networks, in which epigenetics plays an essential role.
Summarizing, the epigenetic cross-talk between tumor cells and immune cells promotes tumor growth and immune escape of malignant cells [8,9], as well as a deficient response to therapy [10] both in solid and hematological cancers, including (but not limited to) lymphomas. Therefore, the underlying epigenetic mechanisms that govern both tumor and immune cells can be used as a strategy to disrupt this interaction and contribute to developing therapeutic strategies against cancer.

2. Lymphoma: An Overview

Lymphoma represents a diverse and heterogeneous group of tumors and is among the ten most prevalent types of cancers worldwide [11]. This neoplasm arises from a clonal proliferation of lymphocytes. Specifically, lymphoma arises from B cell, T cell, or natural killer (NK) cell subsets at different stages of maturation [12,13].
Due to the diversity of lymphoid tumors, the first international classification of lymphoma malignancies was published in 2001 by the World Health Organization (WHO) [14]. Then, the classification was updated in 2008 and 2017 [14] to include important advances that had an impact on the diagnostic approach as well as on new therapeutic strategies for lymphoid neoplasms. The latest revision recognized more than 80 lymphoma subtypes classified into three main categories: B cell, T cell and NK cell neoplasms, and Hodgkin lymphomas. However, lymphomas have traditionally been classified as: (1) Hodgkin lymphoma (HL), which is characterized by the presence of Reed–Sternberg cells and accounts for 10–15% of cases, and (2) Non-Hodgkin lymphoma (NHL), which includes B cell NHLs, T cell NHLs (T NHLs), and natural killer (NK)/-cell NHLs and accounts for 80–85% of cases.
Establishing the lymphoma diagnosis is essential to evaluate the clinical and histopathological features with molecular and immunologic studies [14], which, altogether, will provide better knowledge about both the severity and prognosis of the disease. Thus, for therapeutic purposes, the general practice is to treat lymphoma patients based on limited/indolent (low grade) or advanced/aggressive (high grade) disease [15]. Although indolent lymphomas have favorable prognoses, they are not usually curable at advanced stages. For this reason, relapsed and/or refractory patients have poor prognosis.
There are multiple approved chemotherapy schedules in lymphoproliferative syndromes. Their use will depend mainly on the type of lymphoma, stage, disease situation (previously untreated or relapsed/refractory lymphoma), and characteristics of the patient (medically fit/unfit, comorbidities, age).
For untreated diffuse large B cell lymphoma (DLBCL), the most used schedule includes systemic chemotherapy (doxorubicin, cyclophosphamide, vincristine, and prednisone) plus immunotherapy with the recombinant anti-CD20 antibody rituximab (R-CHOP-21) given every 21 days, with estimated rates of progression-free survival (PFS) and overall survival (OS) of 70% and 75% for the advanced stage and rates of PFS and OS of 90% and 95% for the limited stage [16,17]. Although other chemoimmunotherapy regimens and schedules (dose-dense R-CHOP-14 or ACVBP—doxorubicin, cyclophosphamide, vindesine, bleomycin, prednisone followed by consolidation with methotrexate, etoposide, ifosfamide, cytarabine) have been used to treat DLBCL, none has a more favorable balance of outcomes and toxicity than R-CHOP [18,19]. For patients with advanced stage double-triple hit DLBCL (lymphomas with rearrangements of MYC and BCL2 and/or BCL6), acceptable intensive chemotherapy regimens include: dose-adjusted etoposide, doxorubicin, vincristine, cyclophosphamide, and prednisone plus rituximab (da-EPOCH-R), rituximab plus hyperfractionated cyclophosphamide, vincristine, doxorubicin and dexamethasone alternating with methotrexate and cytarabine (R-HyperCVAD/MA), rituximab plus cyclophosphamide, vincristine, doxorubicin, and high-dose methotrexate alternating with ifosfamide, etoposide, and cytarabine (R-CODOX-M/IVAC) given the worse prognosis of these patients. When compared with R-CHOP, more intensive front-line therapy (da-EPOCH-R, R-HyperCVAD/MA, R-CODOX-M/IVAC) was associated with superior three-year PFS (88 vs. 56%) [20]. Salvage chemotherapy (R-GDP, rituximab, gemcitabine, dexamethasone; R-ICE, rituximab, ifosfamide, carboplatin; R-DHAP, rituximab, dexamethasone, high dose cytarabine, cisplatin; R-GEMOX, rituximab, gemcitabine, oxaliplatin) followed by autologous hematopoietic cell transplantation (HCT) has long been the standard care in medically-fit patients with relapsed/refractory DLBCL (R/R DLBCL). It is associated with long-term survival in approximately half of patients with first relapse of DLBCL, and achieves superior OS (53 vs. 32%), overall response rate (ORR) (84 vs. 44%), and PFS (46 vs. 12%) compared with salvage therapy alone [21]. For patients unsuitable for autologous HCT, salvage treatments with rituximab combined with chemotherapy have a median OS and PFS of 10 and 6 months, respectively [22]. Other regimens in R/R DLBCL include polatuzumab–rituximab–bendamustine (median OS and PFS of 12 and 9 months) [23], tafasitamab–lenalidomide (median OS and PFS of 33 and 11 months) [24], and CAR T-cell therapies (ORR of 50–85% and CR of 47%) [25,26].
A treatment option for follicular, indolent, or mantle cell lymphoma is bendamustine plus rituximab (BR). Trials suggest similar efficacy and fewer side effects when compared with R-CHOP (median PFS 69.5 vs. 31.2 months; no difference in OS). ABVD (doxorubicin, bleomycin, vinblastine, and dacarbazine), the standard initial chemotherapy for patients with Hodgkin lymphoma (HL), shows OS rates at 5 and 10 years of approximately 85% and 55%, respectively, and CR rate of 80% in advanced stage [27].
Bleomycin, etoposide, doxorubicin, cyclophosphamide, vincristine, procarbazine, and prednisone (BEACOPP) and variants of BEACOPP have shown advantages in FPS (85% with BEACOPP and 73% with ABVD, p = 0.004), but not OS (89% vs. 84%, respectively, p = 0.39), when compared with ABVD in several randomized trials in HL; additionally, severe adverse events occur more frequently in the BEACOPP than in the ABVD schedule [28,29]. Short-term outcomes with BV + AVD (brentuximab vedotin, an anti-CD30 antibody-drug conjugate, doxorubicin, vinblastine, and dacarbazine) are at least as favorable as ABVD, but cause less pulmonary toxicity, so this treatment is an acceptable alternative to ABVD for patients at higher risk for pulmonary toxicity and HL [30]. Patients with relapse/refractory HL (R/R HL) are generally treated with targeted chemotherapy BV plus bendamustine or intensive combination chemotherapy (ICE, DHAP, or gemcitabine-containing regimens) and, if a response is demonstrated, start of autologous HCT, if eligible, with results in PFS rates from 30% to 50% [31].
Chemotherapy has improved the outcomes of lymphoma patients over the immunophenotypic subgroups and is the standard treatment for those patients, together with other treatments such as radiation, stem cell transplantation, targeted therapies, or immunotherapies [32,33,34]. Immunotherapy has become a tool for lymphoma treatment as it restores the anti-tumor response of the patient’s immune system. It is described that Programmed Death 1 (PD-1) ligands, PD-L1 and PD-L2, are overexpressed by Hodgkin Reed–Sternberg (HRS) cells in classic HL (cHL), leading to evasion of immune surveillance. Therefore, anti-PD-1 antibodies such as nivolumab and pembrolizumab are used in lymphoma patients. Prospective phase II studies have reported high response rates with nivolumab and pembrolizumab in R/R cHL, but further study is required to establish the durability of treatment [35,36,37]. Nivolumab is approved for patients with cHL who relapsed or progressed after autologous HCT and post-transplantation brentuximab vedotin. Nivolumab showed 69% OR, including 16% CR and median PFS of 15 months [35]. Pembrolizumab is approved in the United States for patients whose disease is refractory or has relapsed after two or more lines of therapy and in the European Union for patients with progression after autologous HCT and brentuximab vedotin, or who are transplant-ineligible and have failed brentuximab vedotin. Pembrolizumab reported 72% OR, including 28% CR, and 14-month median PFS in heavily pretreated patients with r/r HL [36,37].
Although all these treatments have improved the outcomes of lymphoma patients, it is important to develop and identify novel strategies and therapies that could reduce drug resistance and increase the survival rate by minimizing adverse events. In fact, preclinical data and clinical trials are showing epigenetic agents as promising and potential therapeutic tools.

3. Tumor Microenvironment: Interaction of Immune and Lymphoma Tumor Cells

Malignant cells need a dynamic and mutual communication with the components around the tumor to evade immune responses and progress. These components are a heterogeneous mix of immune cells, stromal cells, blood and lymphatic vessels, secreted factors, and extracellular matrix [38,39] which shape the TME. Thus, this interaction plays a pivotal role in supporting the survival and growth of the tumor, the invasion and metastasis, as well as the drug resistance and immune evasion [40,41]. As a result, the composition of the TME has been described to vary according to different cancer types and can predict the patients’ outcome in many cancers, including lymphomas [42]. For example, the TME constitutes more than 80% of the tumor mass in HL and several T-cell lymphoma subtypes, whereas it constitutes approximately 50% in indolent B-cell lymphomas or it is lower in aggressive lymphomas, DLBCL [42].
Immune cells are able to infiltrate tumors by promoting both pro- and anti-tumorigenic functions (Table 1). This dual relationship between immune cells and tumor cells depends on the context. Therefore, macrophages, dendritic cells (DCs), eosinophils, or tumor-infiltrated lymphocytes (TILs; T cells, B cells, and NK cells) have been demonstrated to be relevant for tumor control [43], whereas myeloid-derived suppressor cells (MDSCs), mast cells (MCs), regulatory T cells (Tregs), and type 2-polarized tumor-associated macrophages (TAMs) are important in immunosuppressing action [44].
Given the role of the immune cells in the behavior of the tumor, a deeper knowledge of interactions between them would be relevant to better understand the pathogenesis and prognosis of lymphomas as well as new therapeutic targets. For this reason, the interactions between lymphoma cells and surrounding immune cells are now being studied in much more detail. This section describes the interactions between lymphoma cells and some immune cells of the TME.
Lymphoid cells: Lymphoid cells or lymphocytes are a type of white blood cell that develops from lymphoid progenitor cells that include T cells, B cells, and NK cells. These three types of lymphocytes are commonly found in the TME in lymphomas.
(a) T cells are essential for both immune surveillance and disease progression. Therefore, they are being extensively studied in all types of cancers, including lymphoma subtypes [45,46]. The T cell compartment includes a variety of cell subpopulations with different immune functions: CD4+ T-helper cells (Ths), CD4+/FOXP3+ Tregs, and CD8+ cytotoxic T cells (CTLs). Additionally, there are two types of Th cells, which are Th1 and Th2 cells. Th1 cells produce interleukin (IL)-2, gamma-interferon (IFN-γ) and tumor necrosis factor-beta (TNF-beta), activating CTLs, antigen-presenting cells (APCs), and NK cells [47]. In the meantime, Th2 cells express IL-4, IL-5, IL-6, and IL-10 and support tumor cell growth through CD40–CD40 ligand (CD40L) interactions in HL and Helicobacter-induced MALT lymphomagenesis [6,48]. Accordingly, lymphoma tumor cells can increase IL-10 secretion to produce CD40L by T cells [49]. In fact, IL-10RA or IL-10RB gene amplifications are described in DLBCL, promoting the tumor’s survival [50,51]. In follicular lymphoma (FL), Th cells are predominant and secrete IL-2, IL-4, IFN-γ, and CD40L [52].
It has been described that T cells surround but do not destroy tumor cells in vitro in HL, probably because T cells are less responsive and are attenuated in this disease [53]. In line with this notion, lymphoma tumor cells induce the expression of programmed cell PD-L1, and both CD4+ and CD8+ T-cells express PD-1, thus, the PD-1/PD-L1/2 interaction could suppress the activity of T cells and promote their exhaustion [54]. As a consequence, lymphoma tumor cells drive T cells to apoptosis or promote their differentiation towards Th2 via PD-L1 expression [55], which also enables an immunosuppressive environment in HL through the interaction with PD-1 from macrophages [56].
Tregs are abundant in B-cell NHL, such as FL or DLBCL, showing an immunosuppressive capacity [57,58]. However, the relevance of tumor infiltrating Tregs on disease outcome is unclear. In FL, FOXP3-positive Tregs have been considered as a negative prognostic factor [59], either alone or in cooperation with M2 macrophages from the TME. Nevertheless, FOXP3+ Tregs have also been associated with better outcomes in HL and other lymphoma subtypes [60], despite the high PD-1+ T cells [61] and CD68+ macrophages [62] that have been found. In DLBCL, high levels of Tim-3+Foxp3+Tregs within the TME have been considered as a prognostic factor for poor survival [58], whereas the high intratumor infiltration of CD25+FOXP3+ Tregs was correlated with a favorable prognosis in a meta-analysis [63].
Given the controversy in the role of different Treg subtypes in the prognosis and their capacity to express co-inhibitory and co-stimulatory markers, a recent study performed an in-depth characterization of Treg heterogeneity in NHL. Three different Treg subsets have been identified up to now, revealing the heterogeneity within the Treg compartment. The first one displayed a low expression of PD-1, OX40, and CTLA-4 that corresponded to peripheral blood Tregs. The second subset had high levels of PD-1, OX40, ICOS, TIGIT, CTLA-4 and CD28, CD69, and CD95/Fas (activation markers), and the third subset had a lack of FOXP3 and a high expression of LAG3, CTLA-4, IL-10, CD38, and KLRB1 (immunosuppression-associated genes) [57].
CD8+ cytotoxic T cells (CTLs) play an important role in infectious and malignant diseases. Thus, the anti-tumor effect of CTLs can be reversed depending on the type of lymphoma [64,65]. In this sense, the progression of Epstein–Barr virus (EBV)+ HL [66] or B-cell NHL [67] is suppressed by CTLs, whereas CTLs have also been associated with progression of CD8+ lymphomas, such as nodal cytotoxic T-cell lymphoma and cutaneous T-cell lymphoma (CTCL), towards their malignant transformation [68,69].
(b) B cells are the main humoral immune cells that play a role in antitumor immunity. In indolent lymphomas, non-malignant B cells are frequently encountered in the TME, but their role remains poorly understood [38,45]. Regarding their relationship with lymphoma tumor cells, normal and activated B cells from the germinal center express CD95 (also called FAS) to induce apoptosis. However, the loss of CD95 has been described in many lymphomas, including FL, MALT, and DLCBL [70,71,72]. Additionally, the expression of CD47 has been found in lower levels on normal B cells than in malignant B cells [73] in NHL, which could contribute to immune evasion. Finally, in mature B-cell lymphoma, B cells allow the progression of the disease inducing anergy in CD4+ T cells by CTLA-4 up-regulation [7].
(c) NK cells play a role in the response against infection and tumor cells, so the activation and detection of NK cells in the lymphoma tumor confer a favorable prognosis [74]. Due to the antitumor function of NK cells, lymphomas contribute to the quantitative and functional deficiency of NK cells, which prevents tumor elimination [75]. Regardless of the HL subtype, the concentration of NK cells could be five times lower in affected tissues in comparison with normal tissues [75]. Moreover, a lower number of tumor infiltrating NK cells in HL was associated with poor prognosis and advanced clinical stages [76]. Many factors have been described to contribute in this reduction, such as IL-2 inhibition or the interruption of NKp30 and NKG2D surface receptors [77]. Indeed, a weak HL-derived NK cells cytolysis against the L428 HL cell line was observed by Reiners et al., in contrast to healthy donors where NK cells were able to efficiently eliminate lymphoma tumor cells [77]. This functional deficiency may be related to the significant reduction in NKG2D in NK cells of HL patients (p = 0.0001 for healthy versus untreated patients with HL) [77].
Tumor-associated macrophages (TAMs): Macrophages act against infection and injury via phagocytosis [78]. Due to their relevant role in TME, these cells are called TAMs, which have two different phenotypes depending on their activation states during tumor progression: M1 macrophages, which have anti-tumor effects because of their cytotoxic actions; and M2 macrophages, which promote tumor activity and support tumor growth [79]. The polarization towards M1 or M2 is influenced by cytokines secreted from both Th1 and Th2 cells, respectively [80].
Specifically in lymphomas, malignant cells were described to be able to induce macrophages to their M2 phenotype in vitro [81], which were associated with poor outcomes [82,83]. Although macrophages have been traditionally characterized by the expression of CD68 (a pan-macrophage marker that recognizes epitopes in a wide variety of tissue macrophages), several studies have revealed that M2 macrophages also express CD163 [84]. Accordingly, CD163+ TAMs have been associated with poor prognosis in many types of cancer, including lymphomas [85,86]. In DLBCL, a high number of CD68+/CD163+ or CD163+ M2 at diagnosis was significantly correlated with unfavorable prognosis [87,88].
In addition to CD68 and CD163 markers, it has also been described that MYC controls the expression of M2 specific genes, and its depletion in MYC+ macrophages inhibited tumor growth in melanoma and fibrosarcoma mouse models [89]. In line with this, a recent study evaluated the relevance of MYC-negative and -positive macrophages in HL patients based on positive CD68 and CD163 markers. The authors subdivided the patients into low, intermediate, and high groups based on the number of macrophages over the total number of CD68+ and CD163+ macrophages, the number of MYC+ macrophages, and the number of MYC−macrophages. The results showed that the worst outcomes were associated with a high number of CD163+/MYC+ cells, whereas the best outcomes were correlated with intermediate levels of CD68+/MYC+ macrophages [80].
Myeloid derived suppressor cells (MDSCs): MDSCs, originated from immature myeloid cells, are a group of heterogeneous cells characterized by suppressing immune responses [90]. Currently, MDSCs are generally defined by CD45+ CD3− CD19− CD20− CD56− CD16− HLADR− CD33+ CD11b+ through flow cytometry [91]. Besides, there are two different subpopulations of MDSCs that can either express CD14 or CD15: monocytic MDSCs (M-MDSCs), which express CD14 and may differentiate into monocytes (such as macrophages and dendritic cells), and granulocytic MDSCs (G-MDSCs), which express CD15 and may differentiate into granulocytes (polymorphonuclear leukocytes) [90,91,92]. MDSCs mainly inhibit T cell immune responses by using different mechanisms, such as reactive oxygen species or nitrogen oxide production [93] and Tregs induction, which contributes to tumor dissemination [94]. Through the production of angiogenic factors and proteases, MDSCs also promote angiogenesis [95]. Additionally, MDSCs inhibit the maturation and functions of NK cells and induce the polarization of macrophages into their M2 phenotype by secreting IL-10 and transforming growth factor (TGF)-β [95,96].
Regarding the prognosis in lymphoma, the increased number of MDSCs has been correlated with a more aggressive disease in both HL and NHL [97,98]. Thus, MDSC expansion has been associated with tumor growth and the promotion of a immunosuppressed TME in HL [99]. Moreover, higher levels of MDSCs reduced the response rate in lymphoma patients [99]. Thus, MDSC count was increased compared with healthy controls (12.5% vs. 11%, respectively), but was higher in non-responders (19.7%) [99]. For this reason, several studies support the hypothesis that MDSCs may be promising prognostic biomarkers in lymphomas such as DLBCL [93], and, even more, the results also support their relevance as potential therapeutic targets to overcome immunosuppression [93].
Other studies have also evaluated the relevance of M-MDSCs and G-MDSCs in patients with lymphoma. In B-cell NHL and T-NHL patients, an accumulation of M-MDSCs was described in peripheral blood in comparison with healthy donors [100,101,102], which was correlated with poor survival and advanced lymphoma stage [100,102,103]. It was also noted that M-MDSCs returned to normal levels after clinical tumor remission and T cells were reestablished after removing M-MDSCs [101,102,104,105]. This result has also been described for G-MDSCs, wherein the median percentage of G-MDSCs was higher in patients with HL and B-NHL when compared to healthy donors (2.18 (0.02–70.92) vs. 0.42 (0.04–2.97), p < 0.0001). Considering HL, indolent, and aggressive B-cell NHL patients separately, the percentage of G-MDSCs was higher in the second group of patients (1.54 (0.28–26.34), (2.15 (0.02–20.08), and 2.96 (0.25–70.92), respectively, p < 0.0001) [98]. Moreover, high G-MDSC levels were found in the duodenum related to enteropathy-associated T cell lymphoma [106].

4. Epigenetic

Epigenetic is a term introduced by developmental biologist Conrad Hal Waddington in 1942 in order to explain the dynamic interactions between the developmental environment and the genome [107]. Nowadays, epigenetic is generally accepted and defined as “the study of changes in gene function that are mitotically and/or meiotically heritable and that do not entail a change in DNA sequence” [108]. Epigenetics promotes changes in the chromatin structure, which regulates gene expression. These changes, unlike genetic changes, are reversible and comprise DNA methylation, histone modifications, and non-coding RNA [109]. Although the epigenetic role was first described as an essential mechanism for normal cell function, later, it was found that epigenetic disruptions promoted malignant cellular transformation leading to diverse human diseases, including cancer [110,111,112]. Pediatric tumors, such as Ewing sarcoma, are an example of the relevance of epigenetics in cancer. Indeed, pediatric tumors have been described to have a low frequency of somatic mutations together with hematological tumors [113]. The low frequency of mutations in these tumors does not explain the differences between those patients with different outcomes. Therefore, epigenetics constitutes a promising research area in order to understand the role of the epigenetic alterations in cancer and look for new alternatives to conventional strategies [114,115].

4.1. Epigenetic Alterations in Lymphoma

Increasing advances have allowed researchers to understand the relevance of epigenetic modifications, both in normal hematopoiesis and lymphoid development and in lymphoma. In fact, it has been described that changes in DNA methylation, histone modifications, and non-coding RNA are implicated in the development of B cells and in the maintenance of T-cell identity [116,117,118]. Moreover, it has been described that epigenetic alterations could be important to uncover the response to therapy and the prognosis in lymphoma malignancies [119,120] (Figure 1 and Table 2).
The most common epigenetic alterations in lymphoma are:
DNA methylation: The process catalyzed by DNA methyltransferases (DNMTs), in which a methyl group is transferred to the C-5 position of the cytosine ring of DNA [121], is called DNA methylation. This epigenetic mechanism occurs within promoter regions where methylation regulates gene expression by transcription silencing [122]. The DNMT family is comprised of DNMT1, DNMT2, DNMT3A, DNMT3B, and DNMT3L members [123,124]; however, DNMT1, DNMT3A, and DNMT3B are the only DNMTs implicated in lymphopoiesis and B-cell activation at early stages [125,126].
DNMT1 plays a relevant role in many types of lymphomas. In this regard, DNMT1 has been described to be frequently expressed and to promote cell cycle and DNA replication in DLBCL cells [127]. Moreover, the complex form by AID–DNMT1 inhibits BCL6 expression leading to cell apoptosis and the inhibition of tumor growth in DLBCL cell xenograft mice [128]. In T-cell lymphoma, Peters et al. reported that DNMT1 loss impaired tumor cell proliferation, induced apoptosis, and suppressed normal hematopoiesis. Thus, the authors suggested that DNMT1 was implicated in the maintenance of the tumor phenotype in MYC-induced T-cell lymphomas and in the de novo methylation during tumorigenesis [129]. Downregulation of DNMT1 by shRNA inhibited proliferation, cell cycle and clonal formation, as well as induced apoptosis in OCI-Ly10 and Granta-159 lymphoma cell lines. This downregulation significantly upregulated the expression of tumor suppressor genes such as SOCS3, BCL2L10, p16, p14, and SHP-1 [130].
In Burkitt’s lymphoma (BL), DNMT1 is overexpressed [131]. In fact, the MYC oncogene promotes overexpression of DNMT1, which results in tumor maintenance and progression in BL [132].
DNMT 3B overexpression plays a role in the progression in lymphomas, being predominantly reported in BL [131]. DNMT 3B is approximately overexpressed in 86% of BL patients, contributing to DNA methylation with DNMT1 [131]. This overexpression is due to the direct binding of the oncoprotein MYC to the DNMT 3B promoter region [132]. In addition, in DLBCL, the overexpression of DNMT 3B has been reported and is associated with worse prognosis, treatment resistance, and progression of the disease [132,133].
Even though DNMT1 and DNMT 3B are the most frequently overexpressed and mutated DNMTs in lymphoma, DNMT 3A has been found to be mutated in nearly 11–20% of patients with T-cell lymphoma [134]. DNMT 3A mutations promote its complete loss of function because of the biallelic mutations observed, which could be important in the development of these neoplasms [135]. This loss of DNMT 3A function is mainly the result of the missense mutations, and also, to a lesser extent, nonsense or frameshift mutations [136,137]. DNMT 3A mutation is correlated with Ten-Eleven Translocation 2 (TET2) mutation in peripheral T-cell lymphoma (PTCL) [134,135]. These two mutations would confer selective advantages through the activation of Notch1 signaling pathway by hypomethylation [138,139], or either by RhoA and/or IDH2 mutations [140,141].
Histone modifications: Acetylation, methylation, phosphorylation, and ubiquitination are histone modifications. In lymphomas, acetylation and methylation of histones are the most frequent aberrations.
(a)
Histone acetylation: Chromosome structural modification and regulation of gene expression are regulated by the functional balance between histone acetyltransferases (HATs) and histone deacetylases (HDACs), which contribute to acetylation and deacetylation of histones. Thus, an acetyl group is transferred from acetyl-CoA to the lysine residues NH2 group in proteins by HATs and is removed by HDACs [142,143].
Aberrant expression of HDACs in lymphomas has been reported to contribute to clinical outcome. In PTCL and DLBCL, HDAC1, HDAC2, and HDAC6 were overexpressed when compared to normal lymphoid tissue. Moreover, HDAC6 overexpression was correlated with favorable outcome in DLBCL patients, whereas it had the opposite effect in PTCL [144]. In line with this, the co-overexpression of EZH2 and HDAC2 in PTCL patients has been associated with a poorer survival rate [145].
Additionally, oncoproteins and non-histone proteins are also able to be modified by HDACs and HATs. For example, HDAC3 regulates the signal transducer and activator of transcription (STAT) 3 in DLBCL [146]. HDAC1 and HDAC6 upregulation promotes an increase in IL-15 secretion, contributing to inflammation in CTCL [147]. HDAC9 overexpression modulates BCL6 activity and the function of the tumor suppressor p53 in B-cell lymphomas [148]. In addition, HDACs’ inhibition can act on chaperone protein HSP90 in B-cell lymphoma [149], on the PI3K/AKT/mTOR pathway in PTCL patients [150], or on BIM (tumor suppressor in lymphomagenesis) through SIN3a/HDAC1/2 co-repressor complex in anaplastic large-cell lymphoma [151].
(b)
Histone methylation: Histone methyltransferase enzymes are responsible for transferring methyl groups to lysine and arginine of histones. Mono-, di-, and trimethylation affects gene transcription and promotes lymphoma pathogenesis [152]. Several methyltransferases, such as EZH2, MML2, or SETD2, among others, are associated with lymphoma tumors. SETD2 (SET-domain-containing 2), a methyltransferase responsible for H3 lysine 36 trimethylation (H3K36me3), is mostly silenced in enteropathy-associated T-cell lymphoma (EATL) [153], as well as monomorphic epitheliotropic intestinal T-cell lymphoma (MEITL) [154]. MLL2 (also known as KMT2D) is mutated in angioimmunoblastic T-cell lymphoma (AITL), PTCL not otherwise specified (PTCL-NOS) [155], DLBCL, FL [156], and NK/T-cell lymphoma (NK/TCL) [157,158], where approximately 91% of mutations result in silencing its histone methylation function [159]. The loss of MLL2 functions alters several genes such as ARID1A, TRAF3, or signaling pathways such as JAK-STAT [160] or MAPK. In fact, the combination of chidamide and decitabine improves the KMT2D-PU.1 transcription factor interaction, which inactivates the MAPK pathway [161], constitutively activated in T-cell lymphoma [162]. Likewise, the loss of MLL2 functions influences the poor prognosis in lymphoma patients [157,158,163].
The enhancer of Zeste Homolog 2 (EZH2) catalyzes histone H3 lysine 27 tri-methylation (H3K27me3), which is related to downregulation or suppression of gene transcription [164,165]. EZH2 SET domain is responsible for the transcriptional inhibition mediated by EZH2. A tyrosine deletion/mutation (Y641) at the EZH2 SET domain in DLBCL and FL patients increases H3K27me3 levels repressing tumor suppressor genes and cell differentiation [166]. In different human lymphoma cell lines, the mutation A677G in EZH2 showed analogous effects [166,167]. The relevance of EZH2 is highly reported in several lymphomas. In aggressive B-cell lymphoma, EZH2 is overexpressed in most cases. Additionally, a positive correlation with phosphorylated ERK1/2 (p-ERK1/2) in DLBCL and MYC expression in BL and double-hit lymphoma was observed [168]. EZH2 overexpression was also described in most T-cell lymphomas, such as NK/TCL [169], T-lymphoblastic lymphoma (T-LBL) [170], or adult T-cell leukaemia/lymphoma (ATLL) [171]. This overexpression has been associated with a high proliferation and poor prognosis in PTCL patients [145] and in NKTCL [172].
Non-coding RNA: Non-coding RNA are transcripts that do not encode proteins and are classified based on their length and structure in microRNA (miRNA), long noncoding RNA (lncRNA), and circular RNA (circRNA) [173]. In the last years, many studies have reported the non-coding RNA expression profile in lymphomas, especially B-cell NHLs.
In DLBCL, the most common and frequently aggressive NHL, numerous miRNAs with biological functions, such as miRNA-155 or miRNA-21, have been described to be deregulated [174,175,176]. In BL, which presents a translocation of c-Myc and immunoglobulin genes, a deregulation of some miRNAs controlled by c-Myc, such as let-7 family or miR-34b, among others, has been reported [177,178].
Additionally, different studies have focused on searching for the differences among lymphoma subtypes or between subgroups within a lymphoma. A study showed a miRNA signature that distinguishes between two groups of DLBCL, which are germinal center B-cell-like DLBCL (GCB-DLBCL) and activated B-cell-like DLBCL (ABC-DLBCL). This miRNA signature included the miRNAs 155, 21, and 221, which contributed to discriminating between GCB- and non-GCB DLBCL [174]. The miRNA-155 differential expression in ABC-DLBCL was confirmed in other independent studies [174,175,179]. A specific miRNA signature that could differentiate between BL and DLBCL has also been found [180,181]. In FL, 17 miRNAs associated with differentiation, apoptosis, and proliferation were differentially expressed with respect to t(14;18) (q32;q21) positive and negative FL patients [182].
Regarding lncRNA, a comparison between control B cells and DLBCL samples has shown 2632 novel lncRNAs differentially expressed in this disease [183], and 17 lncRNAs were able to make a distinction between GCB- and ABC-DLBCL subgroups [184]. In FL, 189 deregulated lncRNAs have been observed [185]. Additionally, a recent study has described that the lncRNA NKILA is frequently hypermethylated in DLBCL and its silencing promoted an increase in proliferation and a decrease in cell death in vitro by repression of NF-kB signaling in NHL [186].
There are only a few studies on lymphomas that evaluated the relevance of circRNA expression. Dahl M et al. provided a circRNA expression map in many B-cell malignancies. Their results showed circRNA profiles that could distinguish different B-cell malignancies [187]. Moreover, Hu et al. identified many differentially expressed circRNAs comparing three pairs of DLBCL tissues and adjacent tissues [188]. circ-APC (hsa_circ_0127621) was downregulated in DLBCL tissues, cell lines, and plasma, and the in vitro and in vivo studies verified that the proliferation of DLBCL cells and tumor growth was impaired by circ-APC [188].

4.2. Epigenetic Regulation of Immune Cells in the TME

To date, most research studies have been based on the role of epigenetic alterations on cancer cells. However, given the relevance of the immune cells in cancer, an increasing number of studies are showing that immune cells in the TME could be regulated by epigenetic modifications. Consequently, the study of the epigenetic alterations that affects immune cell function in the TME has become a growing area of investigation in the last few years.
Epigenetic modifications influence the function of lymphoid cells. DNA methylation by DNMT1 and H3K27 trimethylation by EZH2 have been found to impair T-cell infiltration in the TME [189]. Moreover, the DNA methylation patterns were different in exhausted CD8+ T cells. Thus, the gene LAG3 was methylated in naïve cells and demethylated during the activation of naïve CD8+ T cells [190]. LSD1, a lysine-specific demethylase 1A, also seems to be involved in the infiltration of T cells. Accordingly, the increase in H3K4me2 by LSD1 inhibition allowed more CD8+ T-cells recruitment [191], as shown in breast cancer [192]. Likewise, a DNA hypomethylation pattern is necessary for Treg specific gene expression and its immunosuppressive activity [193]. Furthermore, the pro-inflammatory activity mediated by Tregs in the TME has been associated with EZH2 inhibition [194].
Regarding NK cells, epigenetic alterations have been described to affect their maturation, differentiation, and activation [195]. Specifically, the receptors on the NK cell surface with effector function seem to be regulated epigenetically, which would link the epigenetic with the tumor activity of NK cells.
Different studies have reported an epigenetic regulation by histone acetylation and methylation in activating receptors. Therefore, histone acetylation has been shown to control the NKp30 and NKp46 receptors [196], and the NKG2D gene was unmethylated in NK cells and associated with high levels of H3K9 acetylation [197]. Additionally, an increase in the NKG2D receptor has been associated with a decrease in H3K27me3 promoted by an inhibition of EZH2 [74,198]. Initially, a decrease in NK cell activating receptors would be associated with an exhaustion of these cells, which would promote impaired anti-tumor immune responses. However, Yin et al. showed that the expansion and high cytotoxicity of NK cells against malignant cells were associated with an increase in the NKG2D receptor [74].
Since TAMs have been associated with worse prognosis in lymphoma [82,83], a wide variety of studies have tried to elucidate the network that regulates TAM polarization. Epigenetic modifications have described both epigenetic markers that positively regulate M2 polarization and those with the opposite effect. For example, DNMT 3B plays a role in macrophage polarization because its downregulation increases the M2 macrophage expression markers, such as Arg1 [199]. Furthermore, changes in H3K4 and H3K27 methylation seem to control the expression of M2 macrophage markers [200]. Thus, H3K4 methyltransferase SET and SMYD3 (MYND Domain 3) and Jumonji domain containing protein D3 (JMJD3), a H3K27 demethylase, have a role in M2 polarization [200,201,202]. JMJD3 would promote di- and tri-demethylation of H3K27, promoting an activation of Arg1, among other M2 markers [200,202]. Additionally, some HDACs have been associated with the regulation of M2 phenotype, such as HDAC4 and SIRT2, which are implicated in the regulation of Arg1 expression [203,204]. Concerning epigenetic modifications that negatively regulate M2 polarization, HDAC3 and HDAC9 have been proposed. A knockdown of HDAC3 showed the role of this HDAC in the decrease in inflammation [205]. Furthermore, a depletion of HDAC9 also caused a downregulation of inflammatory genes and accumulation of H3K9 and total acetylated H3 at ABCA1 (ATP-binding cassette transporter), ABCG1, and PPAR-γ (peroxisome proliferator-activated receptor) promoters in macrophages [206].
Regarding MDSCs, some studies indicate that HDACs would regulate their immunosuppressive function. Chen et al. described that MDSCs with HDAC11KO were more suppressive within the TME in a murine lymphoma microenvironment. This immunosuppressive activity was associated with the upregulation of expression and enzymatic activity of ARG1 and Nos2 [207]. By contrast, the inhibition of class I histone deacetylases by the epigenetic drug entinostat showed an inhibition of the immunosuppressive function of M-MDSCs. This effect was promoted by the reduction in ARG1, iNOS, and COX-2 levels [208]. Additionally, HDAC11 could be important in the regulation of IL-10 levels in myeloid cells as well as in the expansion of MDSCs [209]. The expression of IL-10 was also reported to potentially be regulated by HDAC6 acting as a transcriptional factor [210]. However, there is no evidence in lymphoma disease. In addition to the role of HDACs, miRNAs may have a role in the regulation of MDSC functions. Thus, a significant increase in miR494 has been identified in MDSCs derived from six different tumor models, at least by using T lymphoma (EG7) and B lymphoma (A20) cell lines. This miRNA promoted the accumulation and activity of MDSCs by Akt pathway activation and targeting the phosphatase and tensin homolog (PTEN) [211].

4.3. Epigenetic Cross-Talk between Lymphoma Tumor Cells and TME

Throughout the review, the method by which epigenetics is dysregulated both in lymphoma tumor cells and in cells surrounding the tumor has been described. However, little is known about the epigenetic modifications that participate in the link between lymphoma tumor cells and TME.
The reciprocal interplay between tumor cells and the TME components triggers activation of signal cascades that result in specific alterations in gene expression patterns driven by epigenetic mechanisms, as shown in Figure 2. For example, mantle cells lymphoma (MCL) and T cells are able to interact between them due to the expression of CD40 and CD40 ligand inducing PD-L1 on MCL cells [212,213]. PD-L1 expression level is, in part, controlled by epigenetic mechanisms [214]. As a consequence, the increase in PD-L1 on MCL cells enhances PD-L1 and PD1 interactions, which results in T-cell exhaustion and, therefore, the immune escape of lymphoma cells [8,9]. Moreover, recent studies have reported the contribution of micro-RNAs to PD-L1 expression. In DLBCL cells, the miR-155 overexpression increased PD-L1, promoting the immune evasion of lymphoma cells [215] by impairing the cytotoxic CD8+ T cells’ function.
The downregulation of miR-548m was also observed when lymphoma cells and stromal cells interacted with each other. The modification in miR-548m expression caused an alteration of HDAC6 expression. Thereby, there was a balance between miR-548m and HDAC6 to induce lymphoma formation and drug resistance [10].
Furthermore, both HATs and HDACs have been involved in the link between lymphoma tumor cells and the TME. Thus, Irina V. Tiper and Tonya J. Webb hypothesized that malignant cells used epigenetic modifications to dysregulate antigen presentation and to affect the ability of NK cells to recognize and kill cells [216]. They reported that the inhibition of HDACs enhanced the antigen presentation and NK cell responses to MCL cells, and inhibited STAT3 and the inflammatory cytokine secretion by MCL cells. All of this was due to the epigenetic regulation of CD1D by HDAC2, which binds to the CD1D promoter [216].
Regarding HATs, a recent study has shown that CREBBP/EP300 mutations contributed to both tumor progression and an aberrant TME in DLBCL [217]. For this purpose, the authors performed a genomic analysis in a cohort of DLBCL patients, in which it was shown that CREBBP/EP300 mutations were associated with tumor progression. In vitro, CREBBP/EP300 mutations in two B-lymphoma cell lines inhibited H3K27 acetylation and triggered a signaling cascade that resulted in the activation of NOTCH pathway. Moreover, these two mutated B-lymphoma cell lines in mononuclear cells co-cultured with peripheral blood promoted the proliferation of lymphoma cells and the macrophage activation and polarization to their M2 phenotype [217]. Hence, this study demonstrated that epigenetic alterations were implicated in the regulation of both tumor progression and TME in DLBCL.
Despite the studies described above, it is mandatory to obtain a deeper knowledge of the epigenetic cross-talk between lymphoma cells and TME, because tumor progression and drug resistance capacity are guided by the dynamic and mutual communication between tumor cells and TME, in which epigenetics plays a relevant role. Thus, this complex interplay would modify specific epigenetic markers that would be different in the absence of this interaction. Accordingly, this could complicate the use of traditional and epigenetic drugs in the treatment of neoplasms [142].

4.4. Epigenetic Therapies: Clinical Trials in Lymphoma Disease

Given the importance of developing novel therapies that can reduce drug resistance and increase the survival rate of lymphoma patients, the epigenetic drugs represent a promising and potentially therapeutic tool.
The first DNMT inhibitors (DNMTi) approved by the Food and Drug Administration (FDA) were azacitidine and decitabine [218]. These epigenetic drugs were used against myelodysplastic syndrome (MDS) and acute myeloid leukemia (AML), with moderate efficacy and high toxicity [219,220]. Moreover, the FDA has approved several epigenetic drugs for the treatment of diverse malignancies including the HDAC inhibitors (HDACi) vorinostat, romidepsin, belinostat, and panobinostat, with vorinostat being the first HDACi approved by the FDA to treat CTCL [221].
Currently, all these epigenetic drugs are being evaluated alone or in combination in several clinical trials involving lymphoma patients. Here, we report the active and recruiting clinical trials in lymphomas which use the previously approved epigenetic drugs (Table 3).
The previous results obtained in clinical trials have shown the effects of these epigenetic drugs in lymphoma patients. Azacitidine has been combined with chemotherapy and epigenetic drugs, among others. In a phase 1/2 trial (NCT01004991), the combination of azacitidine with R-CHOP reported a complete remission (CR) rate of 91.7% in untreated DLBCL patients. Moreover, the combination with vorinostat showed an ORR of 6.7% in patients with relapsed or refractory DLBCL in a phase 1/2 trial (NCT01120834). In B-NHL patients, decitabine decreased tumor DNMT1 and increased tumor apoptosis (p < 0.05), mitoses (p = 0.02), and copper/platinum transporter CTR1 [222]. Moreover, Blum et al. reported that the dose-limiting myelosuppression and infectious complications prevented dose escalation of decitabine to levels associated with changes in global methylation or gene re-expression in chronic lymphocytic leukaemia and NHL [223].
Regarding HDACis, vorinostat in monotherapy in a phase 1 trial conducted in FL, MCL, DLBL, and CTCL patients (NCT00127140) showed different responses in lymphoma patients, varying from CRs to worse responses [224]. A phase 2 trial (NCT00097929) testing vorinostat presented a CR of 5.6% in relapsed DLBCL. These studies with vorinostat in monotherapy seem to have a limited antitumor activity. However, vorinostat has sustained antitumor activity in patients with relapsed or refractory FL. In this phase 2 study, the ORR was 49% and the median PFS was 20 months for the 39 patients with FL [225]. In a combinatorial treatment, vorinostat plus rituximab in NHLs (phase 2 trial NCT00720876) showed an ORR of 50% in FL, 50% in MZL, 33% in MCL, and no response in LPL [226]. In a phase 1/2 study (NCT00972478), vorinostat combined with R-CHOP presented a tendency to improve R-CHOP efficacy in DLBCL (ORR 81% and CR 52%) [227]. Romidepsin or FK228 reported an ORR of 38% and CR 18% and an ORR of 34% and CR of 5.6% in relapsed or refractory PTCL and CTCL, respectively (phase 2 trial NCT00007345) [228,229]. Another HDACi, belinostat or PXD101, has shown an ORR of 10.5% (CR 0%) in relapsed or refractory aggressive B-NHLs (phase 2 trial NCT00303953) [230], and in another phase 2 trial (NCT00274651) involving relapsed or refractory PTCL or CTCL patients, belinostat presented an ORR of 25% (CR 8.3%) in PTCL and an ORR of 14% (CR 10.3%) in CTCL [231]. Finally, the combination of panobinostat and everolimus in relapsed HL and N-HL patients showed an ORR of 43% and CR of 15% [232].
With the clinical trial results, it might be argued that DNMT and HDAC inhibitors used as monotherapies seem to have limited efficacy in lymphoma in early phase studies. However, clinical trials that used combinatorial regimens have demonstrated potent efficacies in many lymphoma settings with an acceptable level of safety. Therefore, further research is essential to better know and understand the epigenetic mechanisms involved in the progression and prognosis in lymphomas, with the aim of developing new therapeutic approaches that significantly improve the clinical prognosis of patients.

5. Conclusions and Future Perspectives

The role of epigenetic alterations in lymphoma and other tumors has been predominantly focused on tumor cells. However, compelling evidence also highlights the relevant role of the TME in supporting the proliferation, migration, and survival of lymphoma cells. For this reason, an increasing number of studies are describing the epigenetic modifications that appear in the TME cells and their role in creating an optimal immunosuppressive environment for tumor development. Nevertheless, a deeper understanding of the interrelationship between lymphoma tumor cells, TME, and the epigenetic mechanisms involved is essential.
The extreme complexity of the interplay between the great variety of cells and the epigenetic alterations occurring in the context of the disease makes the challenge even greater. Nevertheless, a better understanding of this cross-talk would lead to advances in this field and would contribute to increasing our knowledge of the lymphoma disease. Accordingly, the epigenetic alterations could be used as new biomarkers in lymphoma, which would lead to patient stratification based on the expression of the epigenetic marks. Additionally, the use of epigenetic drugs might be an interesting therapeutic option [143] in order to not only address the elimination of lymphoma tumor cells but also counteract the modulating effect of the heterogeneous components that shape the TME. Altogether, it would help to disrupt the dynamic and mutual communication between malignant and immune cells and, consequently, it would contribute to inhibiting tumor progression and chemoresistance or to modifying the immunosuppressive TME activity.

Author Contributions

D.J.G.-D. and L.H.-P. conceived and wrote the manuscript and contributed to the figures’ design. N.P.-C., C.J.-C., V.S.-M. and L.d.l.C.-M. wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Baylin, S.B.; Jones, P.A. A decade of exploring the cancer epigenome—Biological and translational implications. Nat. Rev. Cancer 2011, 11, 726–734. [Google Scholar] [CrossRef] [PubMed]
  3. Sandoval, J.; Esteller, M. Cancer epigenomics: Beyond genomics. Curr. Opin. Genet. Dev. 2012, 22, 50–55. [Google Scholar] [CrossRef] [PubMed]
  4. Boedtkjer, E.; Pedersen, S.F. The Acidic Tumor Microenvironment as a Driver of Cancer. Annu. Rev. Physiol. 2020, 82, 103–126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Scott, D.W.; Gascoyne, R.D. The tumour microenvironment in B cell lymphomas. Nat. Rev. Cancer 2014, 14, 517–534. [Google Scholar] [CrossRef] [PubMed]
  6. Craig, V.J.; Cogliatti, S.B.; Arnold, I.; Gerke, C.; Balandat, J.E.; Wundisch, T.; Muller, A. B-cell receptor signaling and CD40 ligand-independent T cell help cooperate in Helicobacter-induced MALT lymphomagenesis. Leukemia 2010, 24, 1186–1196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Dalai, S.K.; Mirshahidi, S.; Morrot, A.; Zavala, F.; Sadegh-Nasseri, S. Anergy in memory CD4+ T cells is induced by B cells. J. Immunol. 2008, 181, 3221–3231. [Google Scholar] [CrossRef] [Green Version]
  8. Brusa, D.; Serra, S.; Coscia, M.; Rossi, D.; D’Arena, G.; Laurenti, L.; Jaksic, O.; Fedele, G.; Inghirami, G.; Gaidano, G.; et al. The PD-1/PD-L1 axis contributes to T-cell dysfunction in chronic lymphocytic leukemia. Haematologica 2013, 98, 953–963. [Google Scholar] [CrossRef]
  9. Xu-Monette, Z.Y.; Zhou, J.; Young, K.H. PD-1 expression and clinical PD-1 blockade in B-cell lymphomas. Blood 2018, 131, 68–83. [Google Scholar] [CrossRef] [Green Version]
  10. Lwin, T.; Zhao, X.; Cheng, F.; Zhang, X.; Huang, A.; Shah, B.; Zhang, Y.; Moscinski, L.C.; Choi, Y.S.; Kozikowski, A.P.; et al. A microenvironment-mediated c-Myc/miR-548m/HDAC6 amplification loop in non-Hodgkin B cell lymphomas. J. Clin. Investig. 2013, 123, 4612–4626. [Google Scholar] [CrossRef] [Green Version]
  11. Siegel, R.L.; Miller, K.D.; Jemal, A. Cancer statistics, 2018. CA A Cancer J. Clin. 2018, 68, 7–30. [Google Scholar] [CrossRef] [PubMed]
  12. Matasar, M.J.; Zelenetz, A.D. Overview of lymphoma diagnosis and management. Radiol. Clin. N. Am. 2008, 46, 175–198. [Google Scholar] [CrossRef] [PubMed]
  13. Mugnaini, E.N.; Ghosh, N. Lymphoma. Prim. Care 2016, 43, 661–675. [Google Scholar] [CrossRef] [PubMed]
  14. de Leval, L.; Jaffe, E.S. Lymphoma Classification. Cancer J. 2020, 26, 176–185. [Google Scholar] [CrossRef] [PubMed]
  15. The Non-Hodgkin’s Lymphoma Pathologic Classification Project. National Cancer Institute sponsored study of classifications of non-Hodgkin’s lymphomas: Summary and description of a working formulation for clinical usage. Cancer 1982, 49, 2112–2135. [Google Scholar] [CrossRef]
  16. Pfreundschuh, M.; Kuhnt, E.; Trumper, L.; Osterborg, A.; Trneny, M.; Shepherd, L.; Gill, D.S.; Walewski, J.; Pettengell, R.; Jaeger, U.; et al. CHOP-like chemotherapy with or without rituximab in young patients with good-prognosis diffuse large-B-cell lymphoma: 6-year results of an open-label randomised study of the MabThera International Trial (MInT) Group. Lancet. Oncol. 2011, 12, 1013–1022. [Google Scholar] [CrossRef]
  17. Scott, D.W.; Mottok, A.; Ennishi, D.; Wright, G.W.; Farinha, P.; Ben-Neriah, S.; Kridel, R.; Barry, G.S.; Hother, C.; Abrisqueta, P.; et al. Prognostic Significance of Diffuse Large B-Cell Lymphoma Cell of Origin Determined by Digital Gene Expression in Formalin-Fixed Paraffin-Embedded Tissue Biopsies. J. Clin. Oncol. 2015, 33, 2848–2856. [Google Scholar] [CrossRef]
  18. Pfreundschuh, M.; Schubert, J.; Ziepert, M.; Schmits, R.; Mohren, M.; Lengfelder, E.; Reiser, M.; Nickenig, C.; Clemens, M.; Peter, N.; et al. Six versus eight cycles of bi-weekly CHOP-14 with or without rituximab in elderly patients with aggressive CD20+ B-cell lymphomas: A randomised controlled trial (RICOVER-60). Lancet. Oncol. 2008, 9, 105–116. [Google Scholar] [CrossRef]
  19. Recher, C.; Coiffier, B.; Haioun, C.; Molina, T.J.; Ferme, C.; Casasnovas, O.; Thieblemont, C.; Bosly, A.; Laurent, G.; Morschhauser, F.; et al. Intensified chemotherapy with ACVBP plus rituximab versus standard CHOP plus rituximab for the treatment of diffuse large B-cell lymphoma (LNH03-2B): An open-label randomised phase 3 trial. Lancet 2011, 378, 1858–1867. [Google Scholar] [CrossRef]
  20. Landsburg, D.J.; Falkiewicz, M.K.; Maly, J.; Blum, K.A.; Howlett, C.; Feldman, T.; Mato, A.R.; Hill, B.T.; Li, S.; Medeiros, L.J.; et al. Outcomes of Patients With Double-Hit Lymphoma Who Achieve First Complete Remission. J. Clin. Oncol. 2017, 35, 2260–2267. [Google Scholar] [CrossRef] [Green Version]
  21. Philip, T.; Guglielmi, C.; Hagenbeek, A.; Somers, R.; Van der Lelie, H.; Bron, D.; Sonneveld, P.; Gisselbrecht, C.; Cahn, J.Y.; Harousseau, J.L.; et al. Autologous bone marrow transplantation as compared with salvage chemotherapy in relapses of chemotherapy-sensitive non-Hodgkin’s lymphoma. N. Engl. J. Med. 1995, 333, 1540–1545. [Google Scholar] [CrossRef] [PubMed]
  22. Cazelles, C.; Belhadj, K.; Vellemans, H.; Camus, V.; Poullot, E.; Gaulard, P.; Veresezan, L.; Itti, E.; Becker, S.; Carvalho, M.; et al. Rituximab plus gemcitabine and oxaliplatin (R-GemOx) in refractory/relapsed diffuse large B-cell lymphoma: A real-life study in patients ineligible for autologous stem-cell transplantation. Leuk. Lymphoma 2021, 62, 2161–2168. [Google Scholar] [CrossRef] [PubMed]
  23. Sehn, L.H.; Herrera, A.F.; Flowers, C.R.; Kamdar, M.K.; McMillan, A.; Hertzberg, M.; Assouline, S.; Kim, T.M.; Kim, W.S.; Ozcan, M.; et al. Polatuzumab Vedotin in Relapsed or Refractory Diffuse Large B-Cell Lymphoma. J. Clin. Oncol. 2020, 38, 155–165. [Google Scholar] [CrossRef] [PubMed]
  24. Salles, G.; Duell, J.; Gonzalez Barca, E.; Tournilhac, O.; Jurczak, W.; Liberati, A.M.; Nagy, Z.; Obr, A.; Gaidano, G.; Andre, M.; et al. Tafasitamab plus lenalidomide in relapsed or refractory diffuse large B-cell lymphoma (L-MIND): A multicentre, prospective, single-arm, phase 2 study. Lancet. Oncol. 2020, 21, 978–988. [Google Scholar] [CrossRef]
  25. Neelapu, S.S.; Locke, F.L.; Bartlett, N.L.; Lekakis, L.J.; Miklos, D.B.; Jacobson, C.A.; Braunschweig, I.; Oluwole, O.O.; Siddiqi, T.; Lin, Y.; et al. Axicabtagene Ciloleucel CAR T-Cell Therapy in Refractory Large B-Cell Lymphoma. N. Engl. J. Med. 2017, 377, 2531–2544. [Google Scholar] [CrossRef] [PubMed]
  26. Schuster, S.J.; Bishop, M.R.; Tam, C.S.; Waller, E.K.; Borchmann, P.; McGuirk, J.P.; Jager, U.; Jaglowski, S.; Andreadis, C.; Westin, J.R.; et al. Tisagenlecleucel in Adult Relapsed or Refractory Diffuse Large B-Cell Lymphoma. N. Engl. J. Med. 2019, 380, 45–56. [Google Scholar] [CrossRef] [PubMed]
  27. Bonadonna, G.; Zucali, R.; Monfardini, S.; De Lena, M.; Uslenghi, C. Combination chemotherapy of Hodgkin’s disease with adriamycin, bleomycin, vinblastine, and imidazole carboxamide versus MOPP. Cancer 1975, 36, 252–259. [Google Scholar] [CrossRef]
  28. Merli, F.; Luminari, S.; Gobbi, P.G.; Cascavilla, N.; Mammi, C.; Ilariucci, F.; Stelitano, C.; Musso, M.; Baldini, L.; Galimberti, S.; et al. Long-Term Results of the HD2000 Trial Comparing ABVD Versus BEACOPP Versus COPP-EBV-CAD in Untreated Patients With Advanced Hodgkin Lymphoma: A Study by Fondazione Italiana Linfomi. J. Clin. Oncol. 2016, 34, 1175–1181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Viviani, S.; Zinzani, P.L.; Rambaldi, A.; Brusamolino, E.; Levis, A.; Bonfante, V.; Vitolo, U.; Pulsoni, A.; Liberati, A.M.; Specchia, G.; et al. ABVD versus BEACOPP for Hodgkin’s lymphoma when high-dose salvage is planned. N. Engl. J. Med. 2011, 365, 203–212. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Connors, J.M.; Jurczak, W.; Straus, D.J.; Ansell, S.M.; Kim, W.S.; Gallamini, A.; Younes, A.; Alekseev, S.; Illes, A.; Picardi, M.; et al. Brentuximab Vedotin with Chemotherapy for Stage III or IV Hodgkin’s Lymphoma. N. Engl. J. Med. 2018, 378, 331–344. [Google Scholar] [CrossRef] [PubMed]
  31. Moskowitz, C.H.; Nimer, S.D.; Zelenetz, A.D.; Trippett, T.; Hedrick, E.E.; Filippa, D.A.; Louie, D.; Gonzales, M.; Walits, J.; Coady-Lyons, N.; et al. A 2-step comprehensive high-dose chemoradiotherapy second-line program for relapsed and refractory Hodgkin disease: Analysis by intent to treat and development of a prognostic model. Blood 2001, 97, 616–623. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Ansell, S.M.; Lin, Y. Immunotherapy of lymphomas. J. Clin. Investig. 2020, 130, 1576–1585. [Google Scholar] [CrossRef] [PubMed]
  33. Nieto, Y.; Gruschkus, S.; Valdez, B.C.; Jones, R.B.; Anderlini, P.; Hosing, C.; Popat, U.; Qazilbash, M.; Kebriaei, P.; Alousi, A.; et al. Improved outcomes of high-risk relapsed Hodgkin lymphoma patients after high-dose chemotherapy: A 15-year analysis. Haematologica 2021. [Google Scholar] [CrossRef] [PubMed]
  34. Reddy, J.P.; Akhtari, M.; Smith, G.L.; Pinnix, C.; Osborne, E.M.; Gunther, J.R.; Allen, P.K.; Yehia, Z.A.; Fanale, M.; Rodriguez, M.A.; et al. Outcomes After Chemotherapy Followed by Radiation for Stage IIB Hodgkin Lymphoma With Bulky Disease. Clin. Lymphoma Myeloma Leuk. 2015, 15, 664–670.e2. [Google Scholar] [CrossRef] [PubMed]
  35. Armand, P.; Engert, A.; Younes, A.; Fanale, M.; Santoro, A.; Zinzani, P.L.; Timmerman, J.M.; Collins, G.P.; Ramchandren, R.; Cohen, J.B.; et al. Nivolumab for Relapsed/Refractory Classic Hodgkin Lymphoma After Failure of Autologous Hematopoietic Cell Transplantation: Extended Follow-Up of the Multicohort Single-Arm Phase II CheckMate 205 Trial. J. Clin. Oncol. 2018, 36, 1428–1439. [Google Scholar] [CrossRef] [PubMed]
  36. Chen, R.; Zinzani, P.L.; Fanale, M.A.; Armand, P.; Johnson, N.A.; Brice, P.; Radford, J.; Ribrag, V.; Molin, D.; Vassilakopoulos, T.P.; et al. Phase II Study of the Efficacy and Safety of Pembrolizumab for Relapsed/Refractory Classic Hodgkin Lymphoma. J. Clin. Oncol. 2017, 35, 2125–2132. [Google Scholar] [CrossRef] [PubMed]
  37. Chen, R.; Zinzani, P.L.; Lee, H.J.; Armand, P.; Johnson, N.A.; Brice, P.; Radford, J.; Ribrag, V.; Molin, D.; Vassilakopoulos, T.P.; et al. Pembrolizumab in relapsed or refractory Hodgkin lymphoma: 2-year follow-up of KEYNOTE-087. Blood 2019, 134, 1144–1153. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Calabretta, E.; d’Amore, F.; Carlo-Stella, C. Immune and Inflammatory Cells of the Tumor Microenvironment Represent Novel Therapeutic Targets in Classical Hodgkin Lymphoma. Int. J. Mol. Sci. 2019, 20, 5503. [Google Scholar] [CrossRef] [Green Version]
  39. Joshi, M.; Ansell, S.M. Activating the Antitumor Immune Response in Non-Hodgkin Lymphoma Using Immune Checkpoint Inhibitors. J. Immunol. Res. 2020, 2020, 8820377. [Google Scholar] [CrossRef]
  40. Autio, M.; Leivonen, S.K.; Bruck, O.; Mustjoki, S.; Jorgensen, J.M.; Karjalainen-Lindsberg, M.L.; Beiske, K.; Holte, H.; Pellinen, T.; Leppa, S. Immune cell constitution in the tumor microenvironment predicts the outcome in diffuse large B-cell lymphoma. Haematologica 2021, 106, 718–729. [Google Scholar] [CrossRef] [Green Version]
  41. Fridman, W.H.; Pages, F.; Sautes-Fridman, C.; Galon, J. The immune contexture in human tumours: Impact on clinical outcome. Nat. Rev. Cancer 2012, 12, 298–306. [Google Scholar] [CrossRef] [PubMed]
  42. Menter, T.; Tzankov, A. Lymphomas and Their Microenvironment: A Multifaceted Relationship. Pathobiology 2019, 86, 225–236. [Google Scholar] [CrossRef] [PubMed]
  43. Gatti-Mays, M.E.; Balko, J.M.; Gameiro, S.R.; Bear, H.D.; Prabhakaran, S.; Fukui, J.; Disis, M.L.; Nanda, R.; Gulley, J.L.; Kalinsky, K.; et al. If we build it they will come: Targeting the immune response to breast cancer. NPJ Breast Cancer 2019, 5, 37. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Lorenzo-Sanz, L.; Munoz, P. Tumor-Infiltrating Immunosuppressive Cells in Cancer-Cell Plasticity, Tumor Progression and Therapy Response. Cancer Microenviron. 2019, 12, 119–132. [Google Scholar] [CrossRef] [PubMed]
  45. Mulder, T.A.; Wahlin, B.E.; Osterborg, A.; Palma, M. Targeting the Immune Microenvironment in Lymphomas of B-Cell Origin: From Biology to Clinical Application. Cancers 2019, 11, 915. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Wein, F.; Kuppers, R. The role of T cells in the microenvironment of Hodgkin lymphoma. J. Leukoc. Biol. 2016, 99, 45–50. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Steidl, C.; Connors, J.M.; Gascoyne, R.D. Molecular pathogenesis of Hodgkin’s lymphoma: Increasing evidence of the importance of the microenvironment. J. Clin. Oncol. 2011, 29, 1812–1826. [Google Scholar] [CrossRef]
  48. Aldinucci, D.; Gloghini, A.; Pinto, A.; Colombatti, A.; Carbone, A. The role of CD40/CD40L and interferon regulatory factor 4 in Hodgkin lymp.phoma microenvironment. Leuk. Lymphoma 2012, 53, 195–201. [Google Scholar] [CrossRef]
  49. Ohshima, K.; Karube, K.; Hamasaki, M.; Suefuji, H.; Tutiya, T.; Yamaguchi, T.; Suzumiya, J.; Kikuchi, M. Imbalances of chemokines, chemokine receptors and cytokines in Hodgkin lymphoma: Classical Hodgkin lymphoma vs. Hodgkin-like ATLL. Int. J. Cancer 2003, 106, 706–712. [Google Scholar] [CrossRef]
  50. Beguelin, W.; Sawh, S.; Chambwe, N.; Chan, F.C.; Jiang, Y.; Choo, J.W.; Scott, D.W.; Chalmers, A.; Geng, H.; Tsikitas, L.; et al. IL10 receptor is a novel therapeutic target in DLBCLs. Leukemia 2015, 29, 1684–1694. [Google Scholar] [CrossRef]
  51. Cha, Z.; Qian, G.; Zang, Y.; Gu, H.; Huang, Y.; Zhu, L.; Li, J.; Liu, Y.; Tu, X.; Song, H.; et al. Circulating CXCR5+CD4+ T cells assist in the survival and growth of primary diffuse large B cell lymphoma cells through interleukin 10 pathway. Exp. Cell Res. 2017, 350, 154–160. [Google Scholar] [CrossRef] [PubMed]
  52. Ame-Thomas, P.; Le Priol, J.; Yssel, H.; Caron, G.; Pangault, C.; Jean, R.; Martin, N.; Marafioti, T.; Gaulard, P.; Lamy, T.; et al. Characterization of intratumoral follicular helper T cells in follicular lymphoma: Role in the survival of malignant B cells. Leukemia 2012, 26, 1053–1063. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Payne, S.V.; Jones, D.B.; Wright, D.H. Reed-Sternberg-cell/lymphocyte interaction. Lancet 1977, 2, 768–769. [Google Scholar] [CrossRef]
  54. Goodman, A.; Patel, S.P.; Kurzrock, R. PD-1-PD-L1 immune-checkpoint blockade in B-cell lymphomas. Nat. Rev. Clin. Oncol. 2017, 14, 203–220. [Google Scholar] [CrossRef] [PubMed]
  55. Menter, T.; Tzankov, A. Mechanisms of Immune Evasion and Immune Modulation by Lymphoma Cells. Front. Oncol. 2018, 8, 54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Carey, C.D.; Gusenleitner, D.; Lipschitz, M.; Roemer, M.G.M.; Stack, E.C.; Gjini, E.; Hu, X.; Redd, R.; Freeman, G.J.; Neuberg, D.; et al. Topological analysis reveals a PD-L1-associated microenvironmental niche for Reed-Sternberg cells in Hodgkin lymphoma. Blood 2017, 130, 2420–2430. [Google Scholar] [CrossRef]
  57. Spasevska, I.; Josefsson, S.; Blaker, Y.N.; Steen, C.B.; Sharma, A.; Kushekhar, K.; Meyer, S.; Kolstad, A.; Beiske, K.; Holte, H.; et al. Heterogeneity of Regulatory T Cells in B-Cell Non-Hodgkin Lymphoma. Blood 2020, 136, 27–28. [Google Scholar] [CrossRef]
  58. Zhong, W.; Liu, X.; Zhu, Z.; Li, Q.; Li, K. High levels of Tim-3(+)Foxp3(+)Treg cells in the tumor microenvironment is a prognostic indicator of poor survival of diffuse large B cell lymphoma patients. Int. Immunopharmacol. 2021, 96, 107662. [Google Scholar] [CrossRef]
  59. Farinha, P.; Al-Tourah, A.; Gill, K.; Klasa, R.; Connors, J.M.; Gascoyne, R.D. The architectural pattern of FOXP3-positive T cells in follicular lymphoma is an independent predictor of survival and histologic transformation. Blood 2010, 115, 289–295. [Google Scholar] [CrossRef] [Green Version]
  60. Tzankov, A.; Meier, C.; Hirschmann, P.; Went, P.; Pileri, S.A.; Dirnhofer, S. Correlation of high numbers of intratumoral FOXP3+ regulatory T cells with improved survival in germinal center-like diffuse large B-cell lymphoma, follicular lymphoma and classical Hodgkin’s lymphoma. Haematologica 2008, 93, 193–200. [Google Scholar] [CrossRef] [Green Version]
  61. Muenst, S.; Hoeller, S.; Dirnhofer, S.; Tzankov, A. Increased programmed death-1+ tumor-infiltrating lymphocytes in classical Hodgkin lymphoma substantiate reduced overall survival. Hum. Pathol. 2009, 40, 1715–1722. [Google Scholar] [CrossRef] [PubMed]
  62. Tzankov, A.; Matter, M.S.; Dirnhofer, S. Refined prognostic role of CD68-positive tumor macrophages in the context of the cellular micromilieu of classical Hodgkin lymphoma. Pathobiology 2010, 77, 301–308. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Chang, C.; Chen, Y.P.; Medeiros, L.J.; Chen, T.Y.; Chang, K.C. Higher infiltration of intratumoral CD25+ FOXP3+ lymphocytes correlates with a favorable prognosis in patients with diffuse large B-cell lymphoma. Leuk. Lymphoma 2021, 62, 76–85. [Google Scholar] [CrossRef] [PubMed]
  64. Durgeau, A.; Virk, Y.; Corgnac, S.; Mami-Chouaib, F. Recent Advances in Targeting CD8 T-Cell Immunity for More Effective Cancer Immunotherapy. Front. Immunol. 2018, 9, 14. [Google Scholar] [CrossRef] [PubMed]
  65. Farhood, B.; Najafi, M.; Mortezaee, K. CD8(+) cytotoxic T lymphocytes in cancer immunotherapy: A review. J. Cell. Physiol. 2019, 234, 8509–8521. [Google Scholar] [CrossRef] [PubMed]
  66. Duffield, A.S.; Ascierto, M.L.; Anders, R.A.; Taube, J.M.; Meeker, A.K.; Chen, S.; McMiller, T.L.; Phillips, N.A.; Xu, H.; Ogurtsova, A.; et al. Th17 immune microenvironment in Epstein-Barr virus-negative Hodgkin lymphoma: Implications for immunotherapy. Blood Adv. 2017, 1, 1324–1334. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Willenbrock, K.; Roers, A.; Blohbaum, B.; Rajewsky, K.; Hansmann, M.L. CD8(+) T cells in Hodgkin’s disease tumor tissue are a polyclonal population with limited clonal expansion but little evidence of selection by antigen. Am. J. Pathol. 2000, 157, 171–175. [Google Scholar] [CrossRef]
  68. Mukai, H.Y.; Hasegawa, Y.; Kojima, H.; Okoshi, Y.; Takei, N.; Yamashita, Y.; Nagasawa, T.; Mori, N. Nodal CD8 positive cytotoxic T-cell lymphoma: A distinct clinicopathological entity. Mod. Pathol. 2002, 15, 1131–1139. [Google Scholar] [CrossRef] [Green Version]
  69. Wobser, M.; Reinartz, T.; Roth, S.; Goebeler, M.; Rosenwald, A.; Geissinger, E. Cutaneous CD8+ Cytotoxic T-Cell Lymphoma Infiltrates: Clinicopathological Correlation and Outcome of 35 Cases. Oncol. Ther. 2016, 4, 199–210. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Chatzitolios, A.; Venizelos, I.; Tripsiannis.s, G.; Anastassopoulos, G.; Papadopoulos, N. Prognostic significance of CD95, P53, and BCL2 expression in extranodal non-Hodgkin’s lymphoma. Ann. Hematol. 2010, 89, 889–896. [Google Scholar] [CrossRef] [Green Version]
  71. Kondo, E.; Yoshino, T.; Yamadori, I.; Matsuo, Y.; Kawasaki, N.; Minowada, J.; Akagi, T. Expression of Bcl-2 protein and Fas antigen in non-Hodgkin’s lymphomas. Am. J. Pathol. 1994, 145, 330–337. [Google Scholar] [PubMed]
  72. Zoi-Toli, O.; Meijer, C.J.; Oudejans, J.J.; de Vries, E.; van Beek, P.; Willemze, R. Expression of Fas and Fas ligand in cutaneous B-cell lymphomas. J. Pathol. 1999, 189, 533–538. [Google Scholar] [CrossRef]
  73. Chao, M.P.; Alizadeh, A.A.; Tang, C.; Myklebust, J.H.; Varghese, B.; Gill, S.; Jan, M.; Cha, A.C.; Chan, C.K.; Tan, B.T.; et al. Anti-CD47 antibody synergizes with rituximab to promote phagocytosis and eradicate non-Hodgkin lymphoma. Cell 2010, 142, 699–713. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Yin, J.; Leavenworth, J.W.; Li, Y.; Luo, Q.; Xie, H.; Liu, X.; Huang, S.; Yan, H.; Fu, Z.; Zhang, L.Y.; et al. Ezh2 regulates differentiation and function of natural killer cells through histone methyltransferase activity. Proc. Natl. Acad. Sci. USA 2015, 112, 15988–15993. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Gattringer, G.; Greil, R.; Radaszkiewicz, T.; Huber, H. In situ quantification of T-cell subsets, NK-like cells and macrophages in Hodgkin’s disease: Quantity and quality of infiltration density depends on histopathological subtypes. Blut 1986, 53, 49–58. [Google Scholar] [CrossRef] [PubMed]
  76. Alvaro-Naranjo, T.; Lejeune, M.; Salvado-Usach, M.T.; Bosch-Princep, R.; Reverter-Branchat, G.; Jaen-Martinez, J.; Pons-Ferre, L.E. Tumor-infiltrating cells as a prognostic factor in Hodgkin’s lymphoma: A quantitative tissue microarray study in a large retrospective cohort of 267 patients. Leuk. Lymphoma 2005, 46, 1581–1591. [Google Scholar] [CrossRef] [PubMed]
  77. Reiners, K.S.; Kessler, J.; Sauer, M.; Rothe, A.; Hansen, H.P.; Reusch, U.; Hucke, C.; Kohl, U.; Durkop, H.; Engert, A.; et al. Rescue of impaired NK cell activity in hodgkin lymphoma with bispecific antibodies in vitro and in patients. Mol. Ther. 2013, 21, 895–903. [Google Scholar] [CrossRef] [Green Version]
  78. Hirayama, D.; Iida, T.; Nakase, H. The Phagocytic Function of Macrophage-Enforcing Innate Immunity and Tissue Homeostasis. Int. J. Mol. Sci. 2017, 19, 92. [Google Scholar] [CrossRef] [Green Version]
  79. Pan, Y.; Yu, Y.; Wang, X.; Zhang, T. Tumor-Associated Macrophages in Tumor Immunity. Front. Immunol. 2020, 11, 583084. [Google Scholar] [CrossRef]
  80. Werner, L.; Dreyer, J.H.; Hartmann, D.; Barros, M.H.M.; Buttner-Herold, M.; Grittner, U.; Niedobitek, G. Tumor-associated macrophages in classical Hodgkin lymphoma: Hormetic relationship to outcome. Sci. Rep. 2020, 10, 9410. [Google Scholar] [CrossRef]
  81. Tudor, C.S.; Bruns, H.; Daniel, C.; Distel, L.V.; Hartmann, A.; Gerbitz, A.; Buettner, M.J. Macrophages and dendritic cells as actors in the immune reaction of classical Hodgkin lymphoma. PLoS ONE 2014, 9, e114345. [Google Scholar] [CrossRef] [Green Version]
  82. Wang, J.; Gao, K.; Lei, W.; Dong, L.; Xuan, Q.; Feng, M.; Wang, J.; Ye, X.; Jin, T.; Zhang, Z.; et al. Lymphocyte-to-monocyte ratio is associated with prognosis of diffuse large B-cell lymphoma: Correlation with CD163 positive M2 type tumor-associated macrophages, not PD-1 positive tumor-infiltrating lymphocytes. Oncotarget 2017, 8, 5414–5425. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Li, Y.L.; Shi, Z.H.; Wang, X.; Gu, K.S.; Zhai, Z.M. Tumor-associated macrophages predict prognosis in diffuse large B-cell lymphoma and correlation with peripheral absolute monocyte count. BMC Cancer 2019, 19, 1049. [Google Scholar] [CrossRef] [PubMed]
  84. Minami, K.; Hiwatashi, K.; Ueno, S.; Sakoda, M.; Iino, S.; Okumura, H.; Hashiguchi, M.; Kawasaki, Y.; Kurahara, H.; Mataki, Y.; et al. Prognostic significance of CD68, CD163 and Folate receptor-beta positive macrophages in hepatocellular carcinoma. Exp. Med. 2018, 15, 4465–4476. [Google Scholar] [CrossRef] [Green Version]
  85. Guo, B.; Cen, H.; Tan, X.; Ke, Q. Meta-analysis of the prognostic and clinical value of tumor-associated macrophages in adult classical Hodgkin lymphoma. BMC Med. 2016, 14, 159. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Steidl, C.; Lee, T.; Shah, S.P.; Farinha, P.; Han, G.; Nayar, T.; Delaney, A.; Jones, S.J.; Iqbal, J.; Weisenburger, D.D.; et al. Tumor-associated macrophages and survival in classic Hodgkin’s lymphoma. N. Engl. J. Med. 2010, 362, 875–885. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Marchesi, F.; Cirillo, M.; Bianchi, A.; Gately, M.; Olimpieri, O.M.; Cerchiara, E.; Renzi, D.; Micera, A.; Balzamino, B.O.; Bonini, S.; et al. High density of CD68+/CD163+ tumour-associated macrophages (M2-TAM) at diagnosis is significantly correlated to unfavorable prognostic factors and to poor clinical outcomes in patients with diffuse large B-cell lymphoma. Hematol. Oncol. 2015, 33, 110–112. [Google Scholar] [CrossRef] [PubMed]
  88. Nam, S.J.; Go, H.; Paik, J.H.; Kim, T.M.; Heo, D.S.; Kim, C.W.; Jeon, Y.K. An increase of M2 macrophages predicts poor prognosis in patients with diffuse large B-cell lymphoma treated with rituximab, cyclophosphamide, doxorubicin, vincristine and prednisone. Leuk. Lymphoma 2014, 55, 2466–2476. [Google Scholar] [CrossRef] [PubMed]
  89. Pello, O.M.; Chevre, R.; Laoui, D.; De Juan, A.; Lolo, F.; Andres-Manzano, M.J.; Serrano, M.; Van Ginderachter, J.A.; Andres, V. In vivo inhibition of c-MYC in myeloid cells impairs tumor-associated macrophage maturation and pro-tumoral activities. PLoS ONE 2012, 7, e45399. [Google Scholar] [CrossRef] [PubMed]
  90. Palazon-Carrion, N.; Jimenez-Cortegana, C.; Sanchez-Leon, M.L.; Henao-Carrasco, F.; Nogales-Fernandez, E.; Chiesa, M.; Caballero, R.; Rojo, F.; Nieto-Garcia, M.A.; Sanchez-Margalet, V.; et al. Circulating immune biomarkers in peripheral blood correlate with clinical outcomes in advanced breast cancer. Sci. Rep. 2021, 11, 14426. [Google Scholar] [CrossRef]
  91. Ma, P.; Beatty, P.L.; McKolanis, J.; Brand, R.; Schoen, R.E.; Finn, O.J. Circulating Myeloid Derived Suppressor Cells (MDSC) That Accumulate in Premalignancy Share Phenotypic and Functional Characteristics With MDSC in Cancer. Front. Immunol. 2019, 10, 1401. [Google Scholar] [CrossRef] [PubMed]
  92. Jimenez-Cortegana, C.; Liro, J.; Palazon-Carrion, N.; Salamanca, E.; Sojo-Dorado, J.; de la Cruz-Merino, L.; Pascual, A.; Rodriguez-Bano, J.; Sanchez-Margalet, V. Increased Blood Monocytic Myeloid Derived Suppressor Cells but Low Regulatory T Lymphocytes in Patients with Mild COVID-19. Viral Immunol. 2021, 34, 639–645. [Google Scholar] [CrossRef] [PubMed]
  93. Jimenez-Cortegana, C.; Palazon-Carrion, N.; Martin Garcia-Sancho, A.; Nogales-Fernandez, E.; Carnicero-Gonzalez, F.; Rios-Herranz, E.; de la Cruz-Vicente, F.; Rodriguez-Garcia, G.; Fernandez-Alvarez, R.; Rueda Dominguez, A.; et al. Circulating myeloid-derived suppressor cells and regulatory T cells as immunological biomarkers in refractory/relapsed diffuse large B-cell lymphoma: Translational results from the R2-GDP-GOTEL trial. J. Immunother. Cancer 2021, 9, e002323. [Google Scholar] [CrossRef] [PubMed]
  94. de Haas, N.; de Koning, C.; Spilgies, L.; de Vries, I.J.; Hato, S.V. Improving cancer immunotherapy by targeting the STATe of MDSCs. Oncoimmunology 2016, 5, e1196312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Qu, P.; Wang, L.Z.; Lin, P.C. Expansion and functions of myeloid-derived suppressor cells in the tumor microenvironment. Cancer Lett. 2016, 380, 253–256. [Google Scholar] [CrossRef] [PubMed]
  96. Gabrilovich, D.I. Myeloid-Derived Suppressor Cells. Cancer Immunol. Res. 2017, 5, 3–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Chen, B.J.; Dashnamoorthy, R.; Galera, P.; Makarenko, V.; Chang, H.; Ghosh, S.; Evens, A.M. The immune checkpoint molecules PD-1, PD-L1, TIM-3 and LAG-3 in diffuse large B-cell lymphoma. Oncotarget 2019, 10, 2030–2040. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Marini, O.; Spina, C.; Mimiola, E.; Cassaro, A.; Malerba, G.; Todeschini, G.; Perbellini, O.; Scupoli, M.; Carli, G.; Facchinelli, D.; et al. Identification of granulocytic myeloid-derived suppressor cells (G-MDSCs) in the peripheral blood of Hodgkin and non-Hodgkin lymphoma patients. Oncotarget 2016, 7, 27676–27688. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Novosad, O.; Gorbach, O.; Skachkova, O.; Khranovska, N.; Kryachok, P.I. Role of Circulating Myeloid-Derived Suppressor Cell (MDSC) in Hodgkin Lymphoma (HL) Progression: Updated Prospective Study. Blood 2020, 136, 31. [Google Scholar] [CrossRef]
  100. Khalifa, K.A.; Badawy, H.M.; Radwan, W.M.; Shehata, M.A.; Bassuoni, M.A. CD14(+) HLA-DR low/(-) monocytes as indicator of disease aggressiveness in B-cell non-Hodgkin lymphoma. Int. J. Lab. Hematol. 2014, 36, 650–655. [Google Scholar] [CrossRef] [PubMed]
  101. Xiu, B.; Lin, Y.; Grote, D.M.; Ziesmer, S.C.; Gustafson, M.P.; Maas, M.L.; Zhang, Z.; Dietz, A.B.; Porrata, L.F.; Novak, A.J.; et al. IL-10 induces the development of immunosuppressive CD14(+)HLA-DR(low/-) monocytes in B-cell non-Hodgkin lymphoma. Blood Cancer J. 2015, 5, e328. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Lin, Y.; Gustafson, M.P.; Bulur, P.A.; Gastineau, D.A.; Witzig, T.E.; Dietz, A.B. Immunosuppressive CD14+HLA-DR(low)/- monocytes in B-cell non-Hodgkin lymphoma. Blood 2011, 117, 872–881. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Tadmor, T.; Fell, R.; Polliack, A.; Attias, D. Absolute monocytosis at diagnosis correlates with survival in diffuse large B-cell lymphoma-possible link with monocytic myeloid-derived suppressor cells. Hematol. Oncol. 2013, 31, 65–71. [Google Scholar] [CrossRef] [PubMed]
  104. Wilcox, R.A.; Feldman, A.L.; Wada, D.A.; Yang, Z.Z.; Comfere, N.I.; Dong, H.; Kwon, E.D.; Novak, A.J.; Markovic, S.N.; Pittelkow, M.R.; et al. B7-H1 (PD-L1, CD274) suppresses host immunity in T-cell lymphoproliferative disorders. Blood 2009, 114, 2149–2158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Azzaoui, I.; Uhel, F.; Rossille, D.; Pangault, C.; Dulong, J.; Le Priol, J.; Lamy, T.; Houot, R.; Le Gouill, S.; Cartron, G.; et al. T-cell defect in diffuse large B-cell lymphomas involves expansion of myeloid-derived suppressor cells. Blood 2016, 128, 1081–1092. [Google Scholar] [CrossRef] [PubMed]
  106. Bontkes, H.J.; Jordanova, E.S.; Nijeboer, P.; Neefjes-Borst, E.A.; Cillessen, S.; Hayat, A.; Mulder, C.J.; Bouma, G.; von Blomberg, B.M.; de Gruijl, T.D. High myeloid-derived suppressor cell frequencies in the duodenum are associated with enteropathy associated T-cell lymphoma and its precursor lesions. Br. J. Haematol. 2017, 178, 988–991. [Google Scholar] [CrossRef] [PubMed]
  107. Waddington, C.H. The epigenotype. 1942. Int. J. Epidemiol. 2012, 41, 10–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Wu, C.; Morris, J.R. Genes, genetics, and epigenetics: A correspondence. Science 2001, 293, 1103–1105. [Google Scholar] [CrossRef] [Green Version]
  109. Rodriguez-Paredes, M.; Esteller, M. Cancer epigenetics reaches mainstream oncology. Nat. Med. 2011, 17, 330–339. [Google Scholar] [CrossRef]
  110. Ashraf, N.; Zino, S.; Macintyre, A.; Kingsmore, D.; Payne, A.P.; George, W.D.; Shiels, P.G. Altered sirtuin expression is associated with node-positive breast cancer. Br. J. Cancer 2006, 95, 1056–1061. [Google Scholar] [CrossRef] [Green Version]
  111. Adams, H.; Fritzsche, F.R.; Dirnhofer, S.; Kristiansen, G.; Tzankov, A. Class I histone deacetylases 1, 2 and 3 are highly expressed in classical Hodgkin’s lymphoma. Expert Opin. Ther. Targets 2010, 14, 577–584. [Google Scholar] [CrossRef] [PubMed]
  112. Choi, J.H.; Kwon, H.J.; Yoon, B.I.; Kim, J.H.; Han, S.U.; Joo, H.J.; Kim, D.Y. Expression profile of histone deacetylase 1 in gastric cancer tissues. Jpn. J. Cancer Res. 2001, 92, 1300–1304. [Google Scholar] [CrossRef] [PubMed]
  113. Lawrence, M.S.; Stojanov, P.; Polak, P.; Kryukov, G.V.; Cibulskis, K.; Sivachenko, A.; Carter, S.L.; Stewart, C.; Mermel, C.H.; Roberts, S.A.; et al. Mutational heterogeneity in cancer and the search for new cancer-associated genes. Nature 2013, 499, 214–218. [Google Scholar] [CrossRef] [PubMed]
  114. Baxter, E.; Windloch, K.; Gannon, F.; Lee, J.S. Epigenetic regulation in cancer progression. Cell Biosci. 2014, 4, 45. [Google Scholar] [CrossRef] [Green Version]
  115. Kovar, H.; Amatruda, J.; Brunet, E.; Burdach, S.; Cidre-Aranaz, F.; de Alava, E.; Dirksen, U.; van der Ent, W.; Grohar, P.; Grunewald, T.G.; et al. The second European interdisciplinary Ewing sarcoma research summit--A joint effort to deconstructing the multiple layers of a complex disease. Oncotarget 2016, 7, 8613–8624. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Lee, P.P.; Fitzpatrick, D.R.; Beard, C.; Jessup, H.K.; Lehar, S.; Makar, K.W.; Perez-Melgosa, M.; Sweetser, M.T.; Schlissel, M.S.; Nguyen, S.; et al. A critical role for Dnmt1 and DNA methylation in T cell development, function, and survival. Immunity 2001, 15, 763–774. [Google Scholar] [CrossRef] [Green Version]
  117. Thomas, R.M.; Gamper, C.J.; Ladle, B.H.; Powell, J.D.; Wells, A.D. De novo DNA methylation is required to restrict T helper lineage plasticity. J. Biol. Chem. 2012, 287, 22900–22909. [Google Scholar] [CrossRef] [Green Version]
  118. Zhang, J.A.; Mortazavi, A.; Williams, B.A.; Wold, B.J.; Rothenberg, E.V. Dynamic transformations of genome-wide epigenetic marking and transcriptional control establish T cell identity. Cell 2012, 149, 467–482. [Google Scholar] [CrossRef] [Green Version]
  119. Queiros, A.C.; Beekman, R.; Vilarrasa-Blasi, R.; Duran-Ferrer, M.; Clot, G.; Merkel, A.; Raineri, E.; Russinol, N.; Castellano, G.; Bea, S.; et al. Decoding the DNA Methylome of Mantle Cell Lymphoma in the Light of the Entire B Cell Lineage. Cancer Cell 2016, 30, 806–821. [Google Scholar] [CrossRef] [Green Version]
  120. Chambwe, N.; Kormaksson, M.; Geng, H.; De, S.; Michor, F.; Johnson, N.A.; Morin, R.D.; Scott, D.W.; Godley, L.A.; Gascoyne, R.D.; et al. Variability in DNA methylation defines novel epigenetic subgroups of DLBCL associated with different clinical outcomes. Blood 2014, 123, 1699–1708. [Google Scholar] [CrossRef] [Green Version]
  121. Robertson, K.D. DNA methylation and human disease. Nat. Rev. Genet. 2005, 6, 597–610. [Google Scholar] [CrossRef] [PubMed]
  122. Kulis, M.; Esteller, M. DNA methylation and cancer. Adv. Genet. 2010, 70, 27–56. [Google Scholar] [CrossRef] [PubMed]
  123. Bestor, T.H. The DNA methyltransferases of mammals. Hum. Mol. Genet. 2000, 9, 2395–2402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Okano, M.; Xie, S.; Li, E. Cloning and characterization of a family of novel mammalian DNA (cytosine-5) methyltransferases. Nat. Genet. 1998, 19, 219–220. [Google Scholar] [CrossRef] [PubMed]
  125. Lai, A.Y.; Mav, D.; Shah, R.; Grimm, S.A.; Phadke, D.; Hatzi, K.; Melnick, A.; Geigerman, C.; Sobol, S.E.; Jaye, D.L.; 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. [Google Scholar] [CrossRef] [Green Version]
  126. Shaknovich, R.; Cerchietti, L.; Tsikitas, L.; Kormaksson, M.; De, S.; Figueroa, M.E.; Ballon, G.; Yang, S.N.; Weinhold, N.; Reimers, M.; et al. DNA methyltransferase 1 and DNA methylation patterning contribute to germinal center B-cell differentiation. Blood 2011, 118, 3559–3569. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Loo, S.K.; Ab Hamid, S.S.; Musa, M.; Wong, K.K. DNMT1 is associated with cell cycle and DNA replication gene sets in diffuse large B-cell lymphoma. Pathol. Res. Pr. 2018, 214, 134–143. [Google Scholar] [CrossRef] [PubMed]
  128. Jiao, J.; Lv, Z.; Zhang, P.; Wang, Y.; Yuan, M.; Yu, X.; Otieno Odhiambo, W.; Zheng, M.; Zhang, H.; Ma, Y.; et al. AID assists DNMT1 to attenuate BCL6 expression through DNA methylation in diffuse large B-cell lymphoma cell lines. Neoplasia 2020, 22, 142–153. [Google Scholar] [CrossRef] [PubMed]
  129. Peters, S.L.; Hlady, R.A.; Opavska, J.; Klinkebiel, D.; Novakova, S.; Smith, L.M.; Lewis, R.E.; Karpf, A.R.; Simpson, M.A.; Wu, L.; et al. Essential role for Dnmt1 in the prevention and maintenance of MYC-induced T-cell lymphomas. Mol. Cell Biol. 2013, 33, 4321–4333. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Wang, Q.M.; Lian, G.Y.; Song, Y.; Peng, Z.D.; Xu, S.H.; Gong, Y. Downregulation of miR-152 contributes to DNMT1-mediated silencing of SOCS3/SHP-1 in non-Hodgkin lymphoma. Cancer Gene 2019, 26, 195–207. [Google Scholar] [CrossRef]
  131. Robaina, M.C.; Mazzoccoli, L.; Arruda, V.O.; Reis, F.R.; Apa, A.G.; de Rezende, L.M.; Klumb, C.E. Deregulation of DNMT1, DNMT3B and miR-29s in Burkitt lymphoma suggests novel contribution for disease pathogenesis. Exp. Mol. Pathol. 2015, 98, 200–207. [Google Scholar] [CrossRef] [PubMed]
  132. Poole, C.J.; Zheng, W.; Lodh, A.; Yevtodiyenko, A.; Liefwalker, D.; Li, H.; Felsher, D.W.; van Riggelen, J. DNMT3B overexpression contributes to aberrant DNA methylation and MYC-driven tumor maintenance in T-ALL and Burkitt’s lymphoma. Oncotarget 2017, 8, 76898–76920. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Amara, K.; Ziadi, S.; Hachana, M.; Soltani, N.; Korbi, S.; Trimeche, M. DNA methyltransferase DNMT3b protein overexpression as a prognostic factor in patients with diffuse large B-cell lymphomas. Cancer Sci. 2010, 101, 1722–1730. [Google Scholar] [CrossRef] [PubMed]
  134. Couronne, L.; Bastard, C.; Bernard, O.A. TET2 and DNMT3A mutations in human T-cell lymphoma. N. Engl. J. Med. 2012, 366, 95–96. [Google Scholar] [CrossRef] [PubMed]
  135. Yang, L.; Rau, R.; Goodell, M.A. DNMT3A in haematological malignancies. Nat. Rev. Cancer 2015, 15, 152–165. [Google Scholar] [CrossRef] [PubMed]
  136. Hamidi, T.; Singh, A.K.; Chen, T. Genetic alterations of DNA methylation machinery in human diseases. Epigenomics 2015, 7, 247–265. [Google Scholar] [CrossRef] [PubMed]
  137. Hou, H.A.; Kuo, Y.Y.; Liu, C.Y.; Chou, W.C.; Lee, M.C.; Chen, C.Y.; Lin, L.I.; Tseng, M.H.; Huang, C.F.; Chiang, Y.C.; et al. DNMT3A mutations in acute myeloid leukemia: Stability during disease evolution and clinical implications. Blood 2012, 119, 559–568. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Fukumoto, K.; Nguyen, T.B.; Chiba, S.; Sakata-Yanagimoto, M. Review of the biologic and clinical significance of genetic mutations in angioimmunoblastic T-cell lymphoma. Cancer Sci. 2018, 109, 490–496. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Scourzic, L.; Couronne, L.; Pedersen, M.T.; Della Valle, V.; Diop, M.; Mylonas, E.; Calvo, J.; Mouly, E.; Lopez, C.K.; Martin, N.; et al. DNMT3A(R882H) mutant and Tet2 inactivation cooperate in the deregulation of DNA methylation control to induce lymphoid malignancies in mice. Leukemia 2016, 30, 1388–1398. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Fujisawa, M.; Sakata-Yanagimoto, M.; Nishizawa, S.; Komori, D.; Gershon, P.; Kiryu, M.; Tanzima, S.; Fukumoto, K.; Enami, T.; Muratani, M.; et al. Activation of RHOA-VAV1 signaling in angioimmunoblastic T-cell lymphoma. Leukemia 2018, 32, 694–702. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  141. Vallois, D.; Dobay, M.P.; Morin, R.D.; Lemonnier, F.; Missiaglia, E.; Juilland, M.; Iwaszkiewicz, J.; Fataccioli, V.; Bisig, B.; Roberti, A.; 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. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Hajji, N.; Garcia-Dominguez, D.J.; Hontecillas-Prieto, L.; O’Neill, K.; de Alava, E.; Syed, N. The bitter side of epigenetics: Variability and resistance to chemotherapy. Epigenomics 2018, 13, 397–403. [Google Scholar] [CrossRef] [PubMed]
  143. Hontecillas-Prieto, L.; Flores-Campos, R.; Silver, A.; de Alava, E.; Hajji, N.; Garcia-Dominguez, D.J. Synergistic Enhancement of Cancer Therapy Using HDAC Inhibitors: Opportunity for Clinical Trials. Front. Genet. 2020, 11, 578011. [Google Scholar] [CrossRef] [PubMed]
  144. Marquard, L.; Poulsen, C.B.; Gjerdrum, L.M.; de Nully Brown, P.; Christensen, I.J.; Jensen, P.B.; Sehested, M.; Johansen, P.; Ralfkiaer, E. Histone deacetylase 1, 2, 6 and acetylated histone H4 in B- and T-cell lymphomas. Histopathology 2009, 54, 688–698. [Google Scholar] [CrossRef] [PubMed]
  145. Zhang, H.; Lv, H.; Jia, X.; Hu, G.; Kong, L.; Zhang, T.; Li, L.; Pan, Y.; Zhai, Q.; Meng, B.; et al. Clinical significance of enhancer of zeste homolog 2 and histone deacetylases 1 and 2 expression in peripheral T-cell lymphoma. Oncol. Lett. 2019, 18, 1415–1423. [Google Scholar] [CrossRef] [PubMed]
  146. Gupta, M.; Han, J.J.; Stenson, M.; Wellik, L.; Witzig, T.E. Regulation of STAT3 by histone deacetylase-3 in diffuse large B-cell lymphoma: Implications for therapy. Leukemia 2012, 26, 1356–1364. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Mishra, A.; La Perle, K.; Kwiatkowski, S.; Sullivan, L.A.; Sams, G.H.; Johns, J.; Curphey, D.P.; Wen, J.; McConnell, K.; Qi, J.; et al. Mechanism, Consequences, and Therapeutic Targeting of Abnormal IL15 Signaling in Cutaneous T-cell Lymphoma. Cancer Discov. 2016, 6, 986–1005. [Google Scholar] [CrossRef] [Green Version]
  148. Gil, V.S.; Bhagat, G.; Howell, L.; Zhang, J.; Kim, C.H.; Stengel, S.; Vega, F.; Zelent, A.; Petrie, K. Deregulated expression of HDAC9 in B cells promotes development of lymphoproliferative disease and lymphoma in mice. Dis. Model. Mech. 2016, 9, 1483–1495. [Google Scholar] [CrossRef] [Green Version]
  149. Ding, H.; Peterson, K.L.; Correia, C.; Koh, B.; Schneider, P.A.; Nowakowski, G.S.; Kaufmann, S.H. Histone deacetylase inhibitors interrupt HSP90*RASGRP1 and HSP90*CRAF interactions to upregulate BIM and circumvent drug resistance in lymphoma cells. Leukemia 2017, 31, 1593–1602. [Google Scholar] [CrossRef] [Green Version]
  150. Valdez, B.C.; Brammer, J.E.; Li, Y.; Murray, D.; Liu, Y.; Hosing, C.; Nieto, Y.; Champlin, R.E.; Andersson, B.S. Romidepsin targets multiple survival signaling pathways in malignant T cells. Blood Cancer J. 2015, 5, e357. [Google Scholar] [CrossRef]
  151. Piazza, R.; Magistroni, V.; Mogavero, A.; Andreoni, F.; Ambrogio, C.; Chiarle, R.; Mologni, L.; Bachmann, P.S.; Lock, R.B.; Collini, P.; et al. Epigenetic silencing of the proapoptotic gene BIM in anaplastic large cell lymphoma through an MeCP2/SIN3a deacetylating complex. Neoplasia 2013, 15, 511–522. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Jiang, Y.; Melnick, A. The epigenetic basis of diffuse large B-cell lymphoma. Semin. Hematol. 2015, 52, 86–96. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Moffitt, A.B.; Ondrejka, S.L.; McKinney, M.; Rempel, R.E.; Goodlad, J.R.; Teh, C.H.; Leppa, S.; Mannisto, S.; Kovanen, P.E.; Tse, E.; et al. Enteropathy-associated T cell lymphoma subtypes are characterized by loss of function of SETD2. J. Exp. Med. 2017, 214, 1371.1–1386. [Google Scholar] [CrossRef] [PubMed]
  154. Tomita, S.; Kikuti, Y.Y.; Carreras, J.; Sakai, R.; Takata, K.; Yoshino, T.; Bea, S.; Campo, E.; Missiaglia, E.; Bouilly, J.; et al. Monomorphic Epitheliotropic Intestinal T-Cell Lymphoma in Asia Frequently Shows SETD2 Alterations. Cancers 2020, 12, 3539. [Google Scholar] [CrossRef] [PubMed]
  155. Fernandez-Pol, S.; Ma, L.; Joshi, R.P.; Arber, D.A. A Survey of Somatic Mutations in 41 Genes in a Cohort of T-Cell Lymphomas Identifies Frequent Mutations in Genes Involved in Epigenetic Modification. Appl. Immunohistochem. Mol. Morphol. 2019, 27, 416–422. [Google Scholar] [CrossRef] [PubMed]
  156. Rao, R.C.; Dou, Y. Hijacked in cancer: The KMT2 (MLL) family of methyltransferases. Nat. Rev. Cancer 2015, 15, 334–346. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Lee, S.; Park, H.Y.; Kang, S.Y.; Kim, S.J.; Hwang, J.; Lee, S.; Kwak, S.H.; Park, K.S.; Yoo, H.Y.; Kim, W.S.; et al. Genetic alterations of JAK/STAT cascade and histone modification in extranodal NK/T-cell lymphoma nasal type. Oncotarget 2015, 6, 17764–17776. [Google Scholar] [CrossRef] [Green Version]
  158. Li, Q.; Zhang, W.; Li, J.; Xiong, J.; Liu, J.; Chen, T.; Wen, Q.; Zeng, Y.; Gao, L.; Zhang, C.; et al. Plasma circulating tumor DNA assessment reveals KMT2D as a potential poor prognostic factor in extranodal NK/T-cell lymphoma. Biomark. Res. 2020, 8, 27. [Google Scholar] [CrossRef] [PubMed]
  159. Morin, R.D.; Mendez-Lago, M.; Mungall, A.J.; Goya, R.; Mungall, K.L.; Corbett, R.D.; Johnson, N.A.; Severson, T.M.; Chiu, R.; Field, M.; et al. Frequent mutation of histone-modifying genes in non-Hodgkin lymphoma. Nature 2011, 476, 298–303. [Google Scholar] [CrossRef]
  160. Ortega-Molina, A.; Boss, I.W.; Canela, A.; Pan, H.; Jiang, Y.; Zhao, C.; Jiang, M.; Hu, D.; Agirre, X.; Niesvizky, I.; et al. The histone lysine methyltransferase KMT2D sustains a gene expression program that represses B cell lymphoma development. Nat. Med. 2015, 21, 1199–1208. [Google Scholar] [CrossRef]
  161. Jiang, L.; Gu, Z.H.; Yan, Z.X.; Zhao, X.; Xie, Y.Y.; Zhang, Z.G.; Pan, C.M.; Hu, Y.; Cai, C.P.; Dong, Y.; et al. Exome sequencing identifies somatic mutations of DDX3X in natural killer/T-cell lymphoma. Nat. Genet. 2015, 47, 1061–1066. [Google Scholar] [CrossRef] [PubMed]
  162. Ji, M.M.; Huang, Y.H.; Huang, J.Y.; Wang, Z.F.; Fu, D.; Liu, H.; Liu, F.; Leboeuf, C.; Wang, L.; Ye, J.; et al. Histone modifier gene mutations in peripheral T-cell lymphoma not otherwise specified. Haematologica 2018, 103, 679–687. [Google Scholar] [CrossRef] [PubMed]
  163. Ferrero, S.; Rossi, D.; Rinaldi, A.; Bruscaggin, A.; Spina, V.; Eskelund, C.W.; Evangelista, A.; Moia, R.; Kwee, I.; Dahl, C.; et al. KMT2D mutations and TP53 disruptions are poor prognostic biomarkers in mantle cell lymphoma receiving high-dose therapy: A FIL study. Haematologica 2020, 105, 1604–1612. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Bracken, A.P.; Helin, K. Polycomb group proteins: Navigators of lineage pathways led astray in cancer. Nat. Rev. Cancer 2009, 9, 773–784. [Google Scholar] [CrossRef] [PubMed]
  165. Ferrari, K.J.; Scelfo, A.; Jammula, S.; Cuomo, A.; Barozzi, I.; Stutzer, A.; Fischle, W.; Bonaldi, T.; Pasini, D. Polycomb-dependent H3K27me1 and H3K27me2 regulate active transcription and enhancer fidelity. Mol. Cell 2014, 53, 49–62. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. McCabe, M.T.; Graves, A.P.; Ganji, G.; Diaz, E.; Halsey, W.S.; Jiang, Y.; Smitheman, K.N.; Ott, H.M.; Pappalardi, M.B.; Allen, K.E.; et al. Mutation of A677 in histone methyltransferase EZH2 in human B-cell lymphoma promotes hypertrimethylation of histone H3 on lysine 27 (H3K27). Proc. Natl. Acad. Sci. USA 2012, 109, 2989–2994. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Morin, R.D.; Johnson, N.A.; Severson, T.M.; Mungall, A.J.; An, J.; Goya, R.; Paul, J.E.; Boyle, M.; Woolcock, B.W.; Kuchenbauer, F.; 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. [Google Scholar] [CrossRef] [PubMed]
  168. Tian, X.; Pelton, A.; Shahsafaei, A.; Dorfman, D.M. Differential expression of enhancer of zeste homolog 2 (EZH2) protein in small cell and aggressive B-cell non-Hodgkin lymphomas and differential regulation of EZH2 expression by p-ERK1/2 and MYC in aggressive B-cell lymphomas. Mod. Pathol. 2016, 29, 1050–1057. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Yan, J.; Ng, S.B.; Tay, J.L.; Lin, B.; Koh, T.L.; Tan, J.; Selvarajan, V.; Liu, S.C.; Bi, C.; Wang, S.; et al. EZH2 overexpression in natural killer/T-cell lymphoma confers growth advantage independently of histone methyltransferase activity. Blood 2013, 121, 4512–4520. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Ntziachristos, P.; Tsirigos, A.; Van Vlierberghe, P.; Nedjic, J.; Trimarchi, T.; Flaherty, M.S.; Ferres-Marco, D.; da Ros, V.; Tang, Z.; Siegle, J.; et al. Genetic inactivation of the polycomb repressive complex 2 in T cell acute lymphoblastic leukemia. Nat. Med. 2012, 18, 298–301. [Google Scholar] [CrossRef]
  171. Simon, C.; Chagraoui, J.; Krosl, J.; Gendron, P.; Wilhelm, B.; Lemieux, S.; Boucher, G.; Chagnon, P.; Drouin, S.; Lambert, R.; et al. A key role for EZH2 and associated genes in mouse and human adult T-cell acute leukemia. Genes Dev. 2012, 26, 651–656. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Liu, J.; Liang, L.; Huang, S.; Nong, L.; Li, D.; Zhang, B.; Li, T. Aberrant differential expression of EZH2 and H3K27me3 in extranodal NK/T-cell lymphoma, nasal type, is associated with disease progression and prognosis. Hum. Pathol. 2019, 83, 166–176. [Google Scholar] [CrossRef] [PubMed]
  173. Hombach, S.; Kretz, M. Non-coding RNAs: Classification, Biology and Functioning. Adv. Exp. Med. Biol. 2016, 937, 3–17. [Google Scholar] [CrossRef] [PubMed]
  174. Lawrie, C.H.; Soneji, S.; Marafioti, T.; Cooper, C.D.; Palazzo, S.; Paterson, J.C.; Cattan, H.; Enver, T.; Mager, R.; Boultwood, J.; et al. MicroRNA expression distinguishes between germinal center B cell-like and activated B cell-like subtypes of diffuse large B cell lymphoma. Int. J. Cancer 2007, 121, 1156–1161. [Google Scholar] [CrossRef] [PubMed]
  175. Roehle, A.; Hoefig, K.P.; Repsilber, D.; Thorns, C.; Ziepert, M.; Wesche, K.O.; Thiere, M.; Loeffler, M.; Klapper, W.; Pfreundschuh, M.; et al. MicroRNA signatures characterize diffuse large B-cell lymphomas and follicular lymphomas. Br. J. Haematol. 2008, 142, 732–744. [Google Scholar] [CrossRef] [PubMed]
  176. Song, J.; Shao, Q.; Li, C.; Liu, H.; Li, J.; Wang, Y.; Song, W.; Li, L.; Wang, G.; Shao, Z.; et al. Effects of microRNA-21 on apoptosis by regulating the expression of PTEN in diffuse large B-cell lymphoma. Medicine 2017, 96, e7952. [Google Scholar] [CrossRef] [PubMed]
  177. Leucci, E.; Cocco, M.; Onnis, A.; De Falco, G.; van Cleef, P.; Bellan, C.; van Rijk, A.; Nyagol, J.; Byakika, B.; Lazzi, S.; et al. MYC translocation-negative classical Burkitt lymphoma cases: An alternative pathogenetic mechanism involving miRNA deregulation. J. Pathol. 2008, 216, 440–450. [Google Scholar] [CrossRef] [PubMed]
  178. Sampson, V.B.; Rong, N.H.; Han, J.; Yang, Q.; Aris, V.; Soteropoulos, P.; Petrelli, N.J.; Dunn, S.P.; Krueger, L.J. MicroRNA let-7a down-regulates MYC and reverts MYC-induced growth in Burkitt lymphoma cells. Cancer Res. 2007, 67, 9762–9770. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Lim, E.L.; Trinh, D.L.; Scott, D.W.; Chu, A.; Krzywinski, M.; Zhao, Y.; Robertson, A.G.; Mungall, A.J.; Schein, J.; Boyle, M.; et al. Comprehensive miRNA sequence analysis reveals survival differences in diffuse large B-cell lymphoma patients. Genome Biol. 2015, 16, 18. [Google Scholar] [CrossRef] [Green Version]
  180. Iqbal, J.; Shen, Y.; Huang, X.; Liu, Y.; Wake, L.; Liu, C.; Deffenbacher, K.; Lachel, C.M.; Wang, C.; Rohr, J.; et al. Global microRNA expression profiling uncovers molecular markers for classification and prognosis in aggressive B-cell lymphoma. Blood 2015, 125, 1137–1145. [Google Scholar] [CrossRef]
  181. Lenze, D.; Leoncini, L.; Hummel, M.; Volinia, S.; Liu, C.G.; Amato, T.; De Falco, G.; Githanga, J.; Horn, H.; Nyagol, J.; et al. The different epidemiologic subtypes of Burkitt lymphoma share a homogenous micro RNA profile distinct from diffuse large B-cell lymphoma. Leukemia 2011, 25, 1869–1876. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Leich, E.; Zamo, A.; Horn, H.; Haralambieva, E.; Puppe, B.; Gascoyne, R.D.; Chan, W.C.; Braziel, R.M.; Rimsza, L.M.; Weisenburger, D.D.; et al. MicroRNA profiles of t(14;18)-negative follicular lymphoma support a late germinal center B-cell phenotype. Blood 2011, 118, 5550–5558. [Google Scholar] [CrossRef] [PubMed]
  183. Verma, A.; Jiang, Y.; Du, W.; Fairchild, L.L.; Melnick, A.; Elemento, O. Transcriptome sequencing reveals thousands of novel long non-coding RNAs in B cell lymphoma. Genome Med. 2015, 7, 110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Zhou, M.; Zhao, H.; Xu, W.; Bao, S.; Cheng, L.; Sun, J. Discovery and validation of immune-associated long non-coding RNA biomarkers associated with clinically molecular subtype and prognosis in diffuse large B cell lymphoma. Mol. Cancer 2017, 16, 16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Pan, Y.; Li, H.; Guo, Y.; Luo, Y.; Li, H.; Xu, Y.; Deng, J.; Sun, B. A pilot study of long noncoding RNA expression profiling by microarray in follicular lymphoma. Gene 2016, 577, 132–139. [Google Scholar] [CrossRef] [PubMed]
  186. Zhang, M.Y.; Calin, G.; Deng, M.D.; Au-Yeung, R.K.H.; Wang, L.Q.; Chim, C.S. Epigenetic Silencing of Tumor Suppressor lncRNA NKILA: Implication on NF-kappaB Signaling in Non-Hodgkin’s Lymphoma. Genes 2022, 13, 128. [Google Scholar] [CrossRef] [PubMed]
  187. Dahl, M.; Daugaard, I.; Andersen, M.S.; Hansen, T.B.; Gronbaek, K.; Kjems, J.; Kristensen, L.S. Enzyme-free digital counting of endogenous circular RNA molecules in B-cell malignancies. Lab. Investig. 2018, 98, 1657–1669. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Hu, Y.; Zhao, Y.; Shi, C.; Ren, P.; Wei, B.; Guo, Y.; Ma, J. A circular RNA from APC inhibits the proliferation of diffuse large B-cell lymphoma by inactivating Wnt/beta-catenin signaling via interacting with TET1 and miR-888. Aging 2019, 11, 8068–8084. [Google Scholar] [CrossRef]
  189. Peng, D.; Kryczek, I.; Nagarsheth, N.; Zhao, L.; Wei, S.; Wang, W.; Sun, Y.; Zhao, E.; Vatan, L.; Szeliga, W.; et al. Epigenetic silencing of TH1-type chemokines shapes tumour immunity and immunotherapy. Nature 2015, 527, 249–253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Yang, R.; Cheng, S.; Luo, N.; Gao, R.; Yu, K.; Kang, B.; Wang, L.; Zhang, Q.; Fang, Q.; Zhang, L.; et al. Distinct epigenetic features of tumor-reactive CD8+ T cells in colorectal cancer patients revealed by genome-wide DNA methylation analysis. Genome Biol. 2019, 21, 2. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. Qin, Y.; Vasilatos, S.N.; Chen, L.; Wu, H.; Cao, Z.; Fu, Y.; Huang, M.; Vlad, A.M.; Lu, B.; Oesterreich, S.; et al. Inhibition of histone lysine-specific demethylase 1 elicits breast tumor immunity and enhances antitumor efficacy of immune checkpoint blockade. Oncogene 2019, 38, 390–405. [Google Scholar] [CrossRef] [PubMed]
  192. Tu, W.J.; McCuaig, R.D.; Tan, A.H.Y.; Hardy, K.; Seddiki, N.; Ali, S.; Dahlstrom, J.E.; Bean, E.G.; Dunn, J.; Forwood, J.; et al. Targeting Nuclear LSD1 to Reprogram Cancer Cells and Reinvigorate Exhausted T Cells via a Novel LSD1-EOMES Switch. Front. Immunol. 2020, 11, 1228. [Google Scholar] [CrossRef] [PubMed]
  193. Ohkura, N.; Hamaguchi, M.; Morikawa, H.; Sugimura, K.; Tanaka, A.; Ito, Y.; Osaki, M.; Tanaka, Y.; Yamashita, R.; Nakano, N.; et al. T cell receptor stimulation-induced epigenetic changes and Foxp3 expression are independent and complementary events required for Treg cell development. Immunity 2012, 37, 785–799. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Wang, D.; Quiros, J.; Mahuron, K.; Pai, C.C.; Ranzani, V.; Young, A.; Silveria, S.; Harwin, T.; Abnousian, A.; Pagani, M.; et al. Targeting EZH2 Reprograms Intratumoral Regulatory T Cells to Enhance Cancer Immunity. Cell Rep. 2018, 23, 3262–3274. [Google Scholar] [CrossRef] [PubMed]
  195. Schenk, A.; Bloch, W.; Zimmer, P. Natural Killer Cells—An Epigenetic Perspective of Development and Regulation. Int. J. Mol. Sci. 2016, 17, 326. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Ogbomo, H.; Michaelis, M.; Kreuter, J.; Doerr, H.W.; Cinatl, J., Jr. Histone deacetylase inhibitors suppress natural killer cell cytolytic activity. FEBS Lett. 2007, 581, 1317–1322. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Fernandez-Sanchez, A.; Baragano Raneros, A.; Carvajal Palao, R.; Sanz, A.B.; Ortiz, A.; Ortega, F.; Suarez-Alvarez, B.; Lopez-Larrea, C. DNA demethylation and histone H3K9 acetylation determine the active transcription of the NKG2D gene in human CD8+ T and NK cells. Epigenetics 2013, 8, 66–78. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Sauvageau, M.; Sauvageau, G. Polycomb group proteins: Multi-faceted regulators of somatic stem cells and cancer. Cell Stem Cell 2010, 7, 299–313. [Google Scholar] [CrossRef] [Green Version]
  199. Yang, X.; Wang, X.; Liu, D.; Yu, L.; Xue, B.; Shi, H. Epigenetic regulation of macrophage polarization by DNA methyltransferase 3b. Mol. Endocrinol. 2014, 28, 565–574. [Google Scholar] [CrossRef] [Green Version]
  200. Ishii, M.; Wen, H.; Corsa, C.A.; Liu, T.; Coelho, A.L.; Allen, R.M.; Carson, W.F.T.; Cavassani, K.A.; Li, X.; Lukacs, N.W.; et al. Epigenetic regulation of the alternatively activated macrophage phenotype. Blood 2009, 114, 3244–3254. [Google Scholar] [CrossRef]
  201. Kittan, N.A.; Allen, R.M.; Dhaliwal, A.; Cavassani, K.A.; Schaller, M.; Gallagher, K.A.; Carson, W.F.T.; Mukherjee, S.; Grembecka, J.; Cierpicki, T.; et al. Cytokine induced phenotypic and epigenetic signatures are key to establishing specific macrophage phenotypes. PLoS ONE 2013, 8, e78045. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  202. Satoh, T.; Takeuchi, O.; Vandenbon, A.; Yasuda, K.; Tanaka, Y.; Kumagai, Y.; Miyake, T.; Matsushita, K.; Okazaki, T.; Saitoh, T.; et al. The Jmjd3-Irf4 axis regulates M2 macrophage polarization and host responses against helminth infection. Nat. Immunol. 2010, 11, 936–944. [Google Scholar] [CrossRef] [PubMed]
  203. Lo Sasso, G.; Menzies, K.J.; Mottis, A.; Piersigilli, A.; Perino, A.; Yamamoto, H.; Schoonjans, K.; Auwerx, J. SIRT2 deficiency modulates macrophage polarization and susceptibility to experimental colitis. PLoS ONE 2014, 9, e103573. [Google Scholar] [CrossRef] [PubMed]
  204. Yang, Q.; Wei, J.; Zhong, L.; Shi, M.; Zhou, P.; Zuo, S.; Wu, K.; Zhu, M.; Huang, X.; Yu, Y.; et al. Cross talk between histone deacetylase 4 and STAT6 in the transcriptional regulation of arginase 1 during mouse dendritic cell differentiation. Mol. Cell Biol. 2015, 35, 63–75. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Chen, X.; Barozzi, I.; Termanini, A.; Prosperini, E.; Recchiuti, A.; Dalli, J.; Mietton, F.; Matteoli, G.; Hiebert, S.; Natoli, G. Requirement for the histone deacetylase Hdac3 for the inflammatory gene expression program in macrophages. Proc. Natl. Acad. Sci. USA 2012, 109, E2865–E2874. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Cao, Q.; Rong, S.; Repa, J.J.; St Clair, R.; Parks, J.S.; Mishra, N. Histone deacetylase 9 represses cholesterol efflux and alternatively activated macrophages in atherosclerosis development. Arter. Thromb. Vasc. Biol. 2014, 34, 1871–1879. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Chen, J.; Cheng, F.; Wang, M.; Sahakian, E.; Powers, J.J.; Ibarz, J.P.; Smith, M.R.; Sotomayor, E. Loss of HDAC11 Promotes Myeloid-Derived Suppressor Cells Inhibition of T Cell Function in a Murine Lymphoma Microenvironment. Blood 2018, 132, 1105. [Google Scholar] [CrossRef]
  208. Orillion, A.; Hashimoto, A.; Damayanti, N.; Shen, L.; Adelaiye-Ogala, R.; Arisa, S.; Chintala, S.; Ordentlich, P.; Kao, C.; Elzey, B.; et al. Entinostat Neutralizes Myeloid-Derived Suppressor Cells and Enhances the Antitumor Effect of PD-1 Inhibition in Murine Models of Lung and Renal Cell Carcinoma. Clin. Cancer Res. 2017, 23, 5187–5201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Sahakian, E.; Powers, J.J.; Chen, J.; Deng, S.L.; Cheng, F.; Distler, A.; Woods, D.M.; Rock-Klotz, J.; Sodre, A.L.; Youn, J.I.; et al. Histone deacetylase 11: A novel epigenetic regulator of myeloid derived suppressor cell expansion and function. Mol. Immunol. 2015, 63, 579–585. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Cheng, F.; Lienlaf, M.; Perez-Villarroel, P.; Wang, H.W.; Lee, C.; Woan, K.; Woods, D.; Knox, T.; Bergman, J.; Pinilla-Ibarz, J.; et al. Divergent roles of histone deacetylase 6 (HDAC6) and histone deacetylase 11 (HDAC11) on the transcriptional regulation of IL10 in antigen presenting cells. Mol. Immunol. 2014, 60, 44–53. [Google Scholar] [CrossRef] [Green Version]
  211. Liu, Y.; Lai, L.; Chen, Q.; Song, Y.; Xu, S.; Ma, F.; Wang, X.; Wang, J.; Yu, H.; Cao, X.; et al. MicroRNA-494 is required for the accumulation and functions of tumor-expanded myeloid-derived suppressor cells via targeting of PTEN. J. Immunol. 2012, 188, 5500–5510. [Google Scholar] [CrossRef] [PubMed]
  212. Castillo, R.; Mascarenhas, J.; Telford, W.; Chadburn, A.; Friedman, S.M.; Schattner, E.J. Proliferative response of mantle cell lymphoma cells stimulated by CD40 ligation and IL-4. Leukemia 2000, 14, 292–298. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Harrington, B.K.; Wheeler, E.; Hornbuckle, K.; Shana’ah, A.Y.; Youssef, Y.; Smith, L.; Hassan, Q., II; Klamer, B.; Zhang, X.; Long, M.; et al. Modulation of immune checkpoint molecule expression in mantle cell lymphoma. Leuk. Lymphoma 2019, 60, 2498–2507. [Google Scholar] [CrossRef] [PubMed]
  214. Kumar, S.; Sharawat, S.K. Epigenetic regulators of programmed death-ligand 1 expression in human cancers. Transl. Res. J. Lab. Clin. Med. 2018, 202, 129–145. [Google Scholar] [CrossRef] [PubMed]
  215. Zheng, Z.; Sun, R.; Zhao, H.J.; Fu, D.; Zhong, H.J.; Weng, X.Q.; Qu, B.; Zhao, Y.; Wang, L.; Zhao, W.L. MiR155 sensitized B-lymphoma cells to anti-PD-L1 antibody via PD-1/PD-L1-mediated lymphoma cell interaction with CD8+T cells. Mol. Cancer 2019, 18, 54. [Google Scholar] [CrossRef] [PubMed]
  216. Tiper, I.V.; Webb, T.J. Histone deacetylase inhibitors enhance CD1d-dependent NKT cell responses to lymphoma. Cancer Immunol. Immunother. CII 2016, 65, 1411–1421. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Huang, Y.H.; Cai, K.; Xu, P.P.; Wang, L.; Huang, C.X.; Fang, Y.; Cheng, S.; Sun, X.J.; Liu, F.; Huang, J.Y.; et al. CREBBP/EP300 mutations promoted tumor progression in diffuse large B-cell lymphoma through altering tumor-associated macrophage polarization via FBXW7-NOTCH-CCL2/CSF1 axis. Signal. Transduct. Target. 2021, 6, 10. [Google Scholar] [CrossRef] [PubMed]
  218. Jones, P.A.; Issa, J.P.; Baylin, S. Targeting the cancer epigenome for therapy. Nat. Rev. Genet. 2016, 17, 630–641. [Google Scholar] [CrossRef]
  219. Li, L.H.; Olin, E.J.; Buskirk, H.H.; Reineke, L.M. Cytotoxicity and mode of action of 5-azacytidine on L1210 leukemia. Cancer Res. 1970, 30, 2760–2769. [Google Scholar]
  220. Vogler, W.R.; Miller, D.S.; Keller, J.W. 5-Azacytidine (NSC 102816): A new drug for the treatment of myeloblastic leukemia. Blood 1976, 48, 331–337. [Google Scholar] [CrossRef] [Green Version]
  221. Mann, B.S.; Johnson, J.R.; Cohen, M.H.; Justice, R.; Pazdur, R. FDA approval summary: Vorinostat for treatment of advanced primary cutaneous T-cell lymphoma. Oncologist 2007, 12, 1247–1252. [Google Scholar] [CrossRef] [PubMed]
  222. Stewart, D.J.; Issa, J.P.; Kurzrock, R.; Nunez, M.I.; Jelinek, J.; Hong, D.; Oki, Y.; Guo, Z.; Gupta, S.; Wistuba, I.I. 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. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Blum, K.A.; Liu, Z.; Lucas, D.M.; Chen, P.; Xie, Z.; Baiocchi, R.; Benson, D.M.; Devine, S.M.; Jones, J.; Andritsos, L.; 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. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Watanabe, T.; Kato, H.; Kobayashi, Y.; Yamasaki, S.; Morita-Hoshi, Y.; Yokoyama, H.; Morishima, Y.; Ricker, J.L.; Otsuki, T.; Miyagi-Maesima, A.; et al. Potential efficacy of the oral histone deacetylase inhibitor vorinostat in a phase I trial in follicular and mantle cell lymphoma. Cancer Sci. 2010, 101, 196–200. [Google Scholar] [CrossRef] [PubMed]
  225. Ogura, M.; Ando, K.; Suzuki, T.; Ishizawa, K.; Oh, S.Y.; Itoh, K.; Yamamoto, K.; Au, W.Y.; Tien, H.F.; Matsuno, Y.; 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. [Google Scholar] [CrossRef] [PubMed]
  226. Chen, R.; Frankel, P.; Popplewell, L.; Siddiqi, T.; Ruel, N.; Rotter, A.; Thomas, S.H.; Mott, M.; Nathwani, N.; Htut, M.; 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. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Persky, D.O.; Li, H.; Rimsza, L.M.; Barr, P.M.; Popplewell, L.L.; Bane, C.L.; Von Gehr, A.; LeBlanc, M.; Fisher, R.I.; Smith, S.M.; et al. A phase I/II trial of vorinostat (SAHA) in combination with rituximab-CHOP in patients with newly diagnosed advanced stage diffuse large B-cell lymphoma (DLBCL): SWOG S0806. Am. J. Hematol. 2018, 93, 486–493. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Piekarz, R.L.; Frye, R.; Prince, H.M.; Kirsch.hbaum, M.H.; Zain, J.; Allen, S.L.; Jaffe, E.S.; Ling, A.; Turner, M.; Peer, C.J.; et al. Phase 2 trial of romidepsin in patients with peripheral T-cell lymphoma. Blood 2011, 117, 5827–5834. [Google Scholar] [CrossRef] [Green Version]
  229. Piekarz, R.L.; Frye, R.; Turner, M.; Wright, J.J.; Allen, S.L.; Kirschbaum, M.H.; Zain, J.; Prince, H.M.; Leonard, J.P.; Geskin, L.J.; et al. Phase II multi-institutional trial of the histone deacetylase inhibitor romidepsin as monotherapy for patients with cutaneous T-cell lymphoma. J. Clin. Oncol. 2009, 27, 5410–5417. [Google Scholar] [CrossRef] [Green Version]
  230. Puvvada, S.D.; Li, H.; Rimsza, L.M.; Bernstein, S.H.; Fisher, R.I.; LeBlanc, M.; Schmelz, M.; Glinsmann-Gibson, B.; Miller, T.P.; Maddox, A.M.; et al. A phase II study of belinostat (PXD101) in relapsed and refractory aggressive B-cell lymphomas: SWOG S0520. Leuk. Lymphoma 2016, 57, 2359–2369. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  231. Foss, F.; Advani, R.; Duvic, M.; Hymes, K.B.; Intragumtornchai, T.; Lekhakula, A.; Shpilberg, O.; Lerner, A.; Belt, R.J.; Jacobsen, E.D.; 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. [Google Scholar] [CrossRef] [PubMed]
  232. Oki, Y.; Buglio, D.; Fanale, M.; Fayad, L.; Copeland, A.; Romaguera, J.; Kwak, L.W.; Pro, B.; de Castro Faria, S.; Neelapu, S.; et al. Phase I study of panobinostat plus everolimus in patients with relapsed or refractory lymphoma. Clin. Cancer Res. 2013, 19, 6882–6890. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. DNA methylation, non-coding RNA, and histone modifications as regulatory epigenetic mechanisms in lymphoma tumor cells.
Figure 1. DNA methylation, non-coding RNA, and histone modifications as regulatory epigenetic mechanisms in lymphoma tumor cells.
Cancers 14 01469 g001
Figure 2. Schematic representation of the epigenetic cross-talk between lymphoma tumor cells (TC) and the tumor microenvironment (TME) in lymphoma diseases.
Figure 2. Schematic representation of the epigenetic cross-talk between lymphoma tumor cells (TC) and the tumor microenvironment (TME) in lymphoma diseases.
Cancers 14 01469 g002
Table 1. Immune cells in tumor microenvironment of lymphoproliferative diseases.
Table 1. Immune cells in tumor microenvironment of lymphoproliferative diseases.
Immune Cells in TMEFunctionMechanismReferences
Th1 cellsAnti-tumor cellsProduce IL-2, IFN-gamma, and TNF-beta. Activate CTLs, antigen-presenting cells (APCs), and NK cells[47]
Th2 cellsPro-tumor cellsExpress IL-4, IL-5, IL-6, and IL-10 and support tumor cell growth through CD40–CD40L interactions[6,48]
TregsPro- and anti-tumor cellsExpress co-inhibitory and co-stimulatory markers. Three different Treg subsets have been identified in NHL. High levels of PD-1, OX40, ICOS, TIGIT, CTLA-4 and CD28, CD69, and CD95/Fas (activation markers). The third subset of Tregs had a lack of FOXP3 and a high expression of LAG3 and CTLA-4, IL10, CD38, KLRB1 (immunosuppression-associated genes)[57]
CTL cellsPro- and anti-tumor cellsSuppression of the progression of B-cell NHL [42]. Induce the progression in nodal cytotoxic T-cell lymphoma and cutaneous T-cell lymphoma [43,44][67,68,69]
B cellsPro- and anti-tumorContribute to immune evasion [48], B cells allowing the progression of the disease inducing anergy in CD4+ T cells by CTLA-4 up-regulation [49][7,73]
NK cellsAnti-tumor cellsIL-2 and NKp30 and NKG2D surface receptors[77]
M2 macrophagesPro-tumor cellsHigh number of CD163+/MYC+ cells in HL is associated with worst outcome[80]
MDSCsPro-tumor cellsInhibit T cell immune responses by reactive oxygen species or nitrogen oxide production. Inhibit the maturation and functions of NK cells and to induce macrophages polarization into M2 phenotype by secreting IL-10 and transforming growth factor (TGF)-β secretion[93,95,96]
Table 2. Epigenetic modifications involved in lymphoma tumors.
Table 2. Epigenetic modifications involved in lymphoma tumors.
Epigenetic SubgroupEpigenetic Modifications/MechanismFunction
DNA methylationDNMT1 overexpression: hipermethylation of tumor supressors promoters/impaired T-cell infiltration in the TMEPro-tumor cells
DNMT 3A mutation: hypomethylation and Notch1 pathway activationAnti-tumor cells (mutation pro-tumor cells)
DNMT 3B overexpression: hipermethylation MYC promoter; downregulation: increases M2 macrophage expression markers (Arg1)Pro-tumor cells
Histone acetylationHDAC1 overexpression: inhibits the tumor supressor BIM and increase in IL-15 secretionPro-tumor cells
HDAC2 and EZH2 overexpression: inhibits the tumor supressor BIM/mpaired T-cell infiltration in the TMEPro-tumor cells
HDAC3 overexpression: downregulation: downregulation of inflammatory genesPro-tumor cells (downregulation anti-tumor cells)
HDAC6 overexpression: increase in IL-15 secretionPro-tumor cells (PTCL) and anti-tumor cells (DLBCL)
HDAC9 overexpression: modulates BCL6 activity and inhibits the tumor supressor p53; downregulation: downregulation of inflammatory genesPro-tumor cells (downregulation pro-tumor cells)
HDAC11: impaires immunosuppressive–MDSC capacityAnti-tumor cells
Histone methylationEZH2 (catalyzes H3K27me3) overexpression: regulates MYC and induces high proliferation of/increases in NKG2D receptorPro-tumor cells
JMJD3 (promotes di- and tri-demethylation of H3K27) induces activation of Arg1, among other M2 markersPro-tumor cells
LSD1 (mono- or di-demethylase for H3K4) regulates infiltration of CD8+ T-cells into tumorsAnti-tumor cells (downregulation pro-tumor cells)
MLL2 (catalyzes H3K4 monomethyltransferase) mutation: alters genes expression (ARID1A, TRAF3) or signaling pathways (JAK-STAT or MAPK)Anti-tumor cells (mutation pro-tumor cells)
SETD2 (catalyzes H3K36me3) mutation: enteropathy-associated T-cell lymphomaAnti-tumor cells (mutation pro-tumor cells)
SMYD3 (catalyzes H3K4 and H4K5 methyltransferase) induces activation of Arg1, among other M2 markersPro-tumor cells
Non-coding RNAcirc-APC regulates tumor cell proliferationAnti-tumor cells (downregulation pro-tumor cells)
let-7 family inhibits MYC expressionAnti-tumor cells
miR-34b inhibits MYC expressionAnti-tumor cells
miR494 promotes the accumulation and activity of MDSCs by Akt pathway activationPro-tumor cells
miRNA-21 downregulation: increases PTEN expression and induces apoptosisAnti-tumor cells
Table 3. Clinical trials of FDA-approved epigenetic drugs in lymphoproliferative diseases.
Table 3. Clinical trials of FDA-approved epigenetic drugs in lymphoproliferative diseases.
NCT NumberTitleConditionsPhase
SAHA/Vorinostat (HDACi)
NCT00972478Vorinostat, Rituximab, and Combination Chemotherapy in Treating Patients With Newly Diagnosed Stage II, Stage III, or Stage IV Diffuse Large B-Cell Lymphoma
  • Ann Arbor Stage II Non-Hodgkin Lymphoma
  • Ann Arbor Stage III Non-Hodgkin Lymphoma
  • Ann Arbor Stage IV Non-Hodgkin Lymphoma
  • Phase 1
  • Phase 2
NCT00336063Vorinostat and Azacitidine in Treating Patients with Locally Recurrent or Metastatic Nasopharyngeal Cancer or Nasal Natural Killer T-Cell Lymphoma
  • Adult Nasal Type Extranodal NK/T-Cell Lymphoma Recurrent Nasopharyngeal Keratinizing Squamous Cell Carcinoma
  • Recurrent Nasopharyngeal Undifferentiated Carcinoma
  • Stage IV Nasopharyngeal Keratinizing Squamous Cell Carcinoma AJCC v7
  • Stage IV Nasopharyngeal Undifferentiated Carcinoma AJCC v7
Phase 1
NCT03150329Pembrolizumab and Vorinostat in Treating Patients with Relapsed or Refractory Diffuse Large B-Cell Lymphoma, Follicular Lymphoma, or Hodgkin Lymphoma
  • Grade 3b Follicular Lymphoma
  • Recurrent B-Cell Lymphoma, Unclassifiable, with Features Intermediate Between Diffuse Large B- Cell Lymphoma and Classic Hodgkin Lymphoma
  • Recurrent Classic Hodgkin Lymphoma
  • Recurrent Diffuse Large B-Cell Lymphoma
  • Recurrent Follicular Lymphoma
  • Recurrent Grade 1 Follicular Lymphoma
  • Recurrent Grade 2 Follicular Lymphoma
  • Recurrent Grade 3a Follicular Lymphoma
  • Recurrent Primary Mediastinal (Thymic) Large B-Cell Cell Lymphoma
  • Recurrent Transformed Non-Hodgkin Lymphoma
Phase 1
NCT01193842Vorinostat and Combination Chemotherapy with Rituximab in Treating Patients With HIV-Related Diffuse Large B-Cell Non-Hodgkin Lymphoma or Other Aggressive B-Cell Lymphomas
  • AIDS-Related Plasmablastic Lymphoma
  • AIDS-Related Primary Effusion Lymphoma
  • Ann Arbor Stage I Diffuse Large B-Cell Lymphoma
  • Ann Arbor Stage I Grade 3 Follicular Lymphoma
  • Ann Arbor Stage II Diffuse Large B-Cell Lymphoma
  • Ann Arbor Stage II Grade 3 Contiguous Follicular Lymphoma
  • Ann Arbor Stage II Grade 3 Non- Contiguous Follicular Lymphoma
  • Ann Arbor Stage III Diffuse Large B-Cell Lymphoma
  • Ann Arbor Stage III Grade 3 Follicular Lymphoma
  • Ann Arbor Stage IV Diffuse Large B-Cell Lymphoma
  • Phase 1
  • Phase 2
Romidepsin (HDACi)
NCT03547700Study of Ixazomib and Romidepsin in Peripheral T-cellLymphoma (PTCL)
  • Lymphoma, T-Cell, Peripheral
  • Phase 1
  • Phase 2
NCT03141203Evaluation of the Combination of Romidepsin and Carfilzomib in Relapsed/Refractory Peripheral T Cell Lymphoma Patients
  • Peripheral T-Cell Lymphoma
  • Phase 1
  • Phase 2
NCT01796002Efficacy and Safety of Romidepsin CHOP vs. CHOP inPatients with Untreated Peripheral T-Cell Lymphoma
  • Peripheral T-Cell Lymphoma
Phase 3
NCT02616965A Study to Assess the Feasibility of Romidepsin Combined with Brentuximab Vedotin in Cutaneous T-cell Lymphoma
  • Cutaneous T-Cell Lymphoma (CTCL)
Phase 1
NCT01908777A Phase 2 Multicenter Study of High Dose Chemotherapy with Autologous Stem Cell Transplant Followed by Maintenance Therapy with Romidepsin for the Treatment of T-Cell Non-Hodgkin Lymphoma
  • T-Cell Non-Hodgkin Lymphoma
Phase 2
NCT01755975Romidepsin in Combination with Lenalidomide in Adultswith Relapsed or Refractory Lymphomas and Myeloma
  • Multiple Myeloma
  • Non-Hodgkin’s Lymphoma
  • Phase 1
  • Phase 2
NCT03534180Venetoclax and Romidepsin in Treating Patients withRecurrent or Refractory Mature T-Cell Lymphoma
  • Anaplastic Large Cell Lymphoma
  • Recurrent Mature T-Cell and NK-Cell Non-Hodgkin Lymphoma
  • Refractory Mature T-Cell and NK-Cell Non-Hodgkin Lymphoma
Phase 2
Belinostat (HDACi)
NCT02737046Belinostat Therapy With Zidovudine for Adult T-CellLeukemia-Lymphoma
  • Adult T-Cell Leukemia–Lymphoma
  • ATLL
Phase 2
Panobinostat (HDACi)
NCT01261247Panobinostat in Treating Patients with Relapsed orRefractory Non-Hodgkin Lymphoma
  • Adult Nasal Type Extranodal NK/T-Cell Lymphoma
  • Anaplastic Large Cell Lymphoma
  • Angioimmunoblastic T-Cell Lymphoma
  • Extranodal Marginal Zone B-Cell Lymphoma of Mucosa-associated Lymphoid Tissue
  • Hepatosplenic T-Cell Lymphoma
  • Nodal Marginal Zone B-Cell Lymphoma
  • Peripheral T-Cell Lymphoma
  • Post-transplant Lymphoproliferative Disorder
  • Recurrent Adult Burkitt Lymphoma
  • Recurrent Adult Diffuse Large Cell Lymphoma
Phase 2
5-azacytidine (DNMTi)
NCT03450343Oral Azacitidine Plus Salvage Chemotherapy inRelapsed/Refractory Diffuse Large B-Cell Lymphoma
  • Large B-Cell Diffuse Lymphoma
Phase 1
NCT03703375Efficacy and Safety of Oral Azacitidine (CC-486) Compared to Investigator’s Choice Therapy in Patients with Relapsed or Refractory Angioimmunoblastic T-Cell Lymphoma
  • Lymphoma, T-Cell
Phase 3
NCT04897477Azacytidine, Bendamustine, Piamprizumab inRefractory/Relapsed B-Cell Non-Hodgkin Lymphoma
  • Non-Hodgkin Lymphoma, B-Cell
  • Phase 1
  • Phase 2
NCT04480125Azacitidine Combined with Chidamide in the Treatment of Newly Diagnosed PTCL Unfit for Conventional Chemotherapy
  • Peripheral T-Cell Lymphoma
Phase 2
NCT03593018Efficacy and Safety of Oral Azacitidine Compared to Investigator’s Choice Therapy in Patients with Relapsed or Refractory AITL
  • Relapsed Angioimmunoblastic T-Cell
  • Lymphoma
  • Refractory Angioimmunoblastic T-Cell
  • Lymphoma
Phase 3
NCT04747236A Randomized, Phase IIB, Multicenter, Trial of Oral Azacytidine Plus Romidepsin Versus Investigator’s Choice in Patients with Relapse or Refractory Peripheral T-Cell Lymphoma (PTCL)
  • PTCL
Phase 2
NCT05162976CC-486 and Nivolumab for the Treatment of HodgkinLymphoma Refractory to PD-1 Therapy or Relapsed
  • Recurrent Classic Hodgkin Lymphoma
  • Refractory Classic Hodgkin Lymphoma
Phase 1
NCT03542266CC486-CHOP in Patients with Previously UntreatedPeripheral T-Cell Lymphoma
  • Previously Untreated Peripheral T-Cell Lymphoma
Phase 2
NCT04578600CC-486, Lenalidomide, and Obinutuzumab for the Treatment of Recurrent or Refractory CD20 Positive B- Cell Lymphoma
  • Indolent B-Cell Non-Hodgkin Lymphoma
  • Recurrent B-Cell Non-Hodgkin Lymphoma
  • Recurrent Chronic Lymphocytic Leukemia
  • Recurrent Mucosa-associated Lymphoid Tissue Lymphoma
  • Recurrent Follicular LymphomaRecurrent Hairy Cell Leukemia
  • Recurrent Lymphoplasmacytic Lymphoma
  • Recurrent Mantle Cell Lymphoma
  • Recurrent Marginal Zone Lymphoma
  • Refractory B-Cell Non-Hodgkin Lymphoma
Phase 1
NCT02828358Azacitidine and Combination Chemotherapy in Treating Infants With Acute Lymphoblastic Leukemia and KMT2A Gene Rearrangement
  • Acute Leukemia of Ambiguous Lineage
  • B Acute Lymphoblastic Leukemia
  • Mixed Phenotype Acute Leukemia
Phase 2
5-Aza-2-deoxycytidine (DNMTi)
NCT02951728Decitabine Plus R-CHOP in Diffuse Large B-CellLymphoma
  • Diffuse Large B-Cell Lymphoma
  • Phase 1
  • Phase 2
NCT04697940Decitabine-primed Tandem CD19/CD20 CAR T-Cells’Treatment in r/r B-NHL
  • Relapase and Refractory B-Cell Non- Hodgkin Lymphoma
  • Decitabine-primed Tandem CD19/CD20 CAR T Cells
  • Phase 1
  • Phase 2
NCT04850560Sequential Low-dose Decitabine with PD-1/CD28 CD19CAR-T in Relapsed or Refractory B-Cell Lymphoma
  • Objective Response Rate
Phase 1
NCT03494296A Prospective Study of Low-dose Decitabine Combined with COP Regimen in the Treatment of Relapsed and Refractory DLBCL
  • Lymphoma
NCT04446130Study of Decitabine Combined with HAAG Regimen in Newly Diagnosed ETP-ALL/LBL, T/M-MPAL, and ALL/ LBL with Myeloid or Stem Cell Markers Patients
  • Induction Chemotherapy
  • Acute T-Lymphocytic Leukemia
  • T-Cell Lymphoblastic Lymphoma Leukemia
  • T-Cell/Myeloid Mixed Phenotype Acute Leukemia
Phase 3
NCT04553393Decitabine-primed Tandem CD19/CD20 CAR T-Cells Plus Epigenetic Agents in Aggressive r/r B-NHL with Huge Tumor Burden
  • Refractory or Relapsed Aggressive r/r B- NHL With Huge Tumor Burden
  • Phase 1
  • Phase 2
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

García-Domínguez, D.J.; Hontecillas-Prieto, L.; Palazón-Carrión, N.; Jiménez-Cortegana, C.; Sánchez-Margalet, V.; de la Cruz-Merino, L. Tumor Immune Microenvironment in Lymphoma: Focus on Epigenetics. Cancers 2022, 14, 1469. https://doi.org/10.3390/cancers14061469

AMA Style

García-Domínguez DJ, Hontecillas-Prieto L, Palazón-Carrión N, Jiménez-Cortegana C, Sánchez-Margalet V, de la Cruz-Merino L. Tumor Immune Microenvironment in Lymphoma: Focus on Epigenetics. Cancers. 2022; 14(6):1469. https://doi.org/10.3390/cancers14061469

Chicago/Turabian Style

García-Domínguez, Daniel J., Lourdes Hontecillas-Prieto, Natalia Palazón-Carrión, Carlos Jiménez-Cortegana, Víctor Sánchez-Margalet, and Luis de la Cruz-Merino. 2022. "Tumor Immune Microenvironment in Lymphoma: Focus on Epigenetics" Cancers 14, no. 6: 1469. https://doi.org/10.3390/cancers14061469

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

Article Metrics

Back to TopTop