Skip to main content

MINI REVIEW article

Front. Immunol., 22 September 2023
Sec. Cancer Immunity and Immunotherapy
This article is part of the Research Topic Understanding Convergent Evasion Mechanisms in Cancer and Chronic Infection: Implications for Immunotherapy View all 14 articles

Epigenetic targeting to enhance acute myeloid leukemia-directed immunotherapy

  • 1Department of Hematology and Medical Oncology, University Medical Center, Johannes Gutenberg-University, Mainz, Germany
  • 2German Cancer Consortium (DKTK) Partner Site Frankfurt/Mainz and German Cancer Research Center (DKFZ), Heidelberg, Germany
  • 3Children’s Hospital, Experimental Immunology, Johann Wolfgang Goethe University, Frankfurt, Germany
  • 4Frankfurt Cancer Institute, Goethe University, Frankfurt, Germany
  • 5University Cancer Center (UCT), Frankfurt, Germany

AML is a malignant disease of hematopoietic progenitor cells with unsatisfactory treatment outcome, especially in patients that are ineligible for intensive chemotherapy. Immunotherapy, comprising checkpoint inhibition, T-cell engaging antibody constructs, and cellular therapies, has dramatically improved the outcome of patients with solid tumors and lymphatic neoplasms. In AML, these approaches have been far less successful. Discussed reasons are the relatively low mutational burden of AML blasts and the difficulty in defining AML-specific antigens not expressed on hematopoietic progenitor cells. On the other hand, epigenetic dysregulation is an essential driver of leukemogenesis, and non-selective hypomethylating agents (HMAs) are the current backbone of non-intensive treatment. The first clinical trials that evaluated whether HMAs may improve immune checkpoint inhibitors’ efficacy showed modest efficacy except for the anti-CD47 antibody that was substantially more efficient against AML when combined with azacitidine. Combining bispecific antibodies or cellular treatments with HMAs is subject to ongoing clinical investigation, and efficacy data are awaited shortly. More selective second-generation inhibitors targeting specific chromatin regulators have demonstrated promising preclinical activity against AML and are currently evaluated in clinical trials. These drugs that commonly cause leukemia cell differentiation potentially sensitize AML to immune-based treatments by co-regulating immune checkpoints, providing a pro-inflammatory environment, and inducing (neo)-antigen expression. Combining selective targeted epigenetic drugs with (cellular) immunotherapy is, therefore, a promising approach to avoid unintended effects and augment efficacy. Future studies will provide detailed information on how these compounds influence specific immune functions that may enable translation into clinical assessment.

Introduction

Acute myeloid leukemia (AML) is a malignant neoplasm of hematopoietic progenitor cells driven by acquired genetic aberrations that mediate uncontrolled proliferation and a block in differentiation (1, 2).

Novel mechanism-based drugs have improved treatment options in recent years (2), but intensive chemotherapy is still the backbone of curative treatment and induces complete remissions in up to 70% of patients (3). However, relapse is common, and overall survival is generally unsatisfactory and heterogeneous based on two significant factors: the genetic alterations of individual AML blasts and the patient’s age at diagnosis (2, 4). Despite intensive treatment, most elderly patients will ultimately succumb to their disease (25). Survival for patients unfit for intensive treatment is dismal, with a 5-year overall survival (OS) below 10% with current standard of care options (36) underpinning the need for more efficient and less toxic treatment options.

Epigenetic dysregulation has been recognized as an essential driver for leukemogenesis, thereby providing a therapeutic opportunity. Hypomethylating agents (HMA) are non-selective first-generation epigenetic drugs and are considered a mainstay in treating unfit and elderly patients (7). Several more selective compounds targeting specific epigenetic dependencies have been developed in recent years with promising responses in clinical trials (810). Immunotherapy has revolutionized the treatment of solid tumors and lymphatic neoplasms (1130), but has been far less successful against AML. Mechanisms behind the limited efficacy remain obscure but have been attributed to difficulties in finding a target exclusively expressed on AML blasts, their relatively low mutational burden, and low neo-antigen expression (3134).

Epigenetic manipulation has been reported to induce immune modulatory effects, including an increased expression of tumor-associated antigens (35, 36) that may sensitize AML blasts for immunotherapy. Here we review the concept of combined epigenetic targeting with immunotherapeutic approaches against AML.

Epigenetic treatment in AML

Epigenetic dysregulation has been implicated in the pathogenesis of most cancer types, including AML. Sequencing efforts to characterize the genomic landscape of various cancer types have revealed recurrent mutations in epigenetic regulators, affecting AML in more than 60% of cases (37, 38). Epigenetic regulators determine the chromatin state by controlling regulatory regions and gene expression via chemical modifications, including DNA methylation and histone protein acetylation, methylation, or phosphorylation as reviewed elsewhere (3941). Therefore, epigenetic regulators were recognized as therapeutic opportunities for many cancers, particularly AML.

First-generation HMAs such as azacitidine and decitabine are non-selective drugs that reduce promotor hypermethylation to restore the expression of tumor suppressor genes (42). These drugs have built the backbone for non-intensive AML treatment (7), and their combination with the BCL2 inhibitor venetoclax is the current standard of care for unfit AML patients resulting in a median overall survival (OS) of 14.7 months (6). Histone deacetylase (HDAC) inhibitors, another class of non-selective epigenetic drugs that initially showed promising activity in preclinical models (43), failed to induce sustainable remissions in clinical trials in monotherapy (4446). Reasons for the low efficacy in clinical studies are not fully elucidated, however missing predictive biomarkers, the heterogeneous activity of different HDAC inhibitors, and dose-limiting off-target effects of pan-HDAC inhibitors remain an unsolved problem, especially in combination with other anti-neoplastic agents (4749).

Second-generation epigenetic inhibitors were developed to target specific chromatin modifiers and epigenetic dependencies in various cancers with potentially less off-target toxicity. Research has particularly focused on the development and clinical assessment of drugs targeting the following chromatin modifiers:

Bromodomain-containing transcriptional activators (BRDs) are recruited to histone-acetylated transcription sites to accelerate gene expression. BRD4 is a Bromodomain and extra-terminal (BET) protein, and its function is best characterized in AML (50, 51). Inhibitors of BET proteins, particularly BRD4, have shown promising preclinical activity (52) but demonstrated only modest activity as a single agent against AML with an overall response rate (ORR) of only 6% in relapsed refractory (R/R) AML (53).

The histone methyltransferase Disruptor of Telomeric Silencing 1-like (DOT1L) is the only histone 3 lysine 79 methyltransferase known to date. It maintains leukemic transcription in leukemias with Mixed-Lineage Leukemia (MLL, also known as KMT2A)-rearrangement (MLL-r) or partial tandem duplication and NPM1 mutant (NPM1mut) leukemia (54, 55). Similar to BET inhibitors, the first clinical trials with DOT1L inhibitors demonstrated limited activity with only two complete remissions (CR) in 52 patients in a phase I trial (56) despite promising preclinical activity (54, 55).

Protein Arginine Methyltransferase 5 (PRMT5) regulates gene expression by dimethylation of histone and non-histone proteins (e.g.,RNA splicing factors) (57, 58). Inhibition of PRMT5 has demonstrated anti-leukemic activity and induction of differentiation in preclinical MLL-r and FLT3-ITD AML models (59, 60), and several inhibitors are currently evaluated in early clinical trials for solid tumors, lymphomas, and leukemias, which was reviewed elsewhere (61). In brief, phase I studies have reported limited efficacy, with common adverse effects in solid tumors and primary myelofibrosis (6264). One phase I study is currently recruiting AML patients (65).

Enhancer of Zeste Homolog 2 (EZH2) is a lysine methyltransferase and the catalytic subunit of Polycomb Repressive Complex 2 (PRC2) that silences its target genes via H3K27 trimethylation (66, 67). EZH2 mutations are found in solid tumors and usually as gain-of-function events in lymphomas (68, 69). The inhibitor tazometestat induced durable and complete responses in Phase I/II trials in sarcomas and lymphomas (7072). EZH2 has been reported to act context-dependently as a tumor suppressor or sometimes as an oncogene in myeloid malignancies (66, 73). Its loss has been associated with poor prognosis and chemotherapy resistance, and mutations are more common in relapsed AML patients (7476). EZH1/2 inhibition has demonstrated in vitro and in vivo anti-leukemic activity (77, 78). Clinical outcome data for EZH2 inhibition in AML do not exist, also because a phase I trial was terminated due to insufficient patient recruitment (NCT03110354).

Lysine-Specific Demethylase-1 (LSD1, also known as KDM1A) is a histone 3 demethylase and is believed to participate in the control of leukemic gene expression programs (79). LSD1 inhibition had promising activity in preclinical leukemia models, and preliminary efficacy against AML has been reported from an ongoing clinical phase I/II trial (80, 81). Additional studies are needed to define the clinical activity in specific AML subtypes in detail.

Dramatic clinical responses in AML were observed with specific inhibitors of mutant isocitrate-dehydrogenase (IDH) 1 and 2 enzymes and are also explained by epigenetic mechanisms: Mutations in IDH1 and IDH2 lead to a neo-enzyme activity of both enzymes, accumulating the ordinarily absent oncometabolite 2-hydroxyglutarate (2-HG) (82). 2-HG inhibits ten-eleven translocation (TET) family enzymes responsible for DNA methylation, ultimately resulting in aberrant expression of leukemic genes (83). IDH1/2 inhibition induces cell differentiation of IDH-mutated AML blasts (84). The first phase I trial assessed the IDH2 inhibitor enasidenib as a single agent with an ORR of 40.3% and a median OS rate of 9.3 months in R/R AML patients (85). The combination of the IDH1 inhibitor ivosidenib with azacitidine was recently approved for newly diagnosed IDH1 mutated AML in Europe and the U.S. The approval was based on a randomized, placebo-controlled phase III trial where the combination significantly increased CR rates (47% vs. 15%, p<0.001) and survival (recently updated median OS: 29.3 vs. 7.9 months; HR 0.42, p-value <0.0001) compared to azacitidine plus placebo (9, 86).

A novel epigenetic target and auspicious therapeutic opportunity against specific AML subtypes is the protein interaction of the histone methyltransferase KMT2A (also known as MLL1) with its oncogenic adaptor protein menin (encoded by the MEN1 gene). While it was reported that menin is required for chromatin binding and target gene activation of oncogenic MLL1-fusion proteins in MLL1-rearranged leukemias (87), our group reported that the direct interaction of wildtype MLL with menin is a dependency in the most prevalent NPM1mut AML subtype (55). Characteristic leukemic gene expression programs, including high-level expression of MEIS1, PBX3, and various HOX transcription factor genes, also depend on the protein interaction (55). Pharmacological inhibition of the menin-MLL interaction has demonstrated profound in vitro and in vivo anti-leukemic activity inducing uniform transcriptional repression of MEIS1, PBX3, FLT3, and BCL2, and leading to differentiation and apoptosis in MLL-r and NPM1mut leukemias (55, 8790). These preclinical data translated into an ongoing clinical assessment of five different menin inhibitors against AML (NCT04067336, NCT04065399, NCT05153330, NCT04811560, NCT04988555) with astonishing first efficacy data from two phase I trials: The oral menin inhibitor revumenib induced complete remissions (combined; CRc) in 38% of heavily pretreated R/R AML with NPM1mut or MLL-r as a single agent, with responding patients exhibiting sustainable responses of more than 9.1 months (8). Ziftomenib also had promising clinical activity in NPM1mut or MLL-r R/R AML, with 35% of patients achieving CR/CRh or CRp rate in a phase I/II study (91). The single-agent evaluation of both drugs is currently ongoing. Combinatorial clinical trial assessment with intensive chemotherapy and specific small molecule inhibitors is also underway, as both inhibitors have exhibited synergistic in vitro and in vivo efficacy with various targeted cancer drugs (9294).

Targeting the immune system in AML

Within the last decade, similarly great excitement has greeted cancer immunotherapy, revolutionizing the treatment of many cancer types (1130). Concepts to guide the immune system in recognizing and fighting cancer cells comprise antibody-directed targeting, blockage of immune checkpoints, and adoptive transfer of immune cells. These approaches have led to sustainable responses, prolonged survival, and even cure of previously untreatable malignancies, but single-agent efficacy against AML has been limited.

Immune checkpoint blockade (ICB) with anti-CTLA-4 and anti-PD-L1/PD-1 antibodies dramatically improved overall survival in patients with advanced solid tumors as well as Hodgkin’s lymphoma (1318) and is now considered the standard of care for the treatment of many other cancer entities.

AML cells also have higher surface expression of inhibitory immune checkpoints (such as PD-L1) compared to normal hematopoietic stem (HSCs) and progenitor cells (HSPCs) and higher expression of PD-1 is observed on T-cells of AML patients compared to healthy donors (95102). Still, clinical trials assessing therapeutic checkpoint blockade yielded generally discouraging results in myeloid neoplasms. Only 1 out of 9 patients with AML or myelodysplastic syndrome (MDS) responded to the anti-PD-1 antibody pidilizumab in a first phase I trial (103). Also, ORR in studies assessing the anti-PD-1 antibody pembrolizumab and anti-PD-L1 antibody atezolizumab in R/R MDS patients were only 4% and 0%, respectively (104, 105). Responses to the anti-CTLA-4 antibody ipilimumab in early clinical trials assessing selected AML patients that relapsed following allogenic stem cell transplantation (SCT) were more promising, with 23% of patients achieving a CR. However, treatment was commonly associated with severe graft versus host disease (12).

CD47 is a checkpoint of the innate immune system that mediates a “do not eat me “ signal to macrophages (106, 107). Magrolimab, a monoclonal anti-CD47 antibody, demonstrated limited efficacy as a single-agent in AML with no objective responses (stable disease: 73%) (108), but might be more efficacious if added to established combination regimens (discussed below).

Bispecific T-cell engager (BiTE) or dual-affinity retargeting antibodies (DART) are artificial antibody constructs that contain two antigen binding sites, one directed against immune effector cells (mostly CD3 for T-cells) and the other against a specific surface antigen on tumor cells. The convergence leads to T- or NK-cell activation and killing (31). BiTEs targeting CD3 and CD19, such as blinatumomab, are efficient against and approved for treating B-cell neoplasms (28). Defining a unique leukemic target on myeloid blasts has yet limited efforts to extend this concept for successful AML treatment (discussed below), and so far, efficacy has been unsatisfactory. In a phase I trial assessing the anti-CD33xCD3 directed bispecific antibody AMG330 against R/R AML, CR/CRi rates were 17% (109) and 3 and 5% in ongoing phase I studies testing the anti-CD33xCD3 BiTE molecules AMV564 and AMG673 (110, 111). Reported ORR from a phase I/II trial exploring flotetuzumab, an anti-CD123xCD3 DART construct, against R/R AML was 30%. However, treatment was associated with high rates of severe cytokine release syndrome (CRS) (81%, 8% ≥3) (112), which was also commonly observed with the bispecific anti-CD123 antibody XmAb14045 (113). Other CD123-targeting antibodies are under clinical investigation (NCT03647800, NCT02715011).

Several reports suggest that the myeloid antigens WT1, PRAME, and CLL-1 (CLEC12A) are expressed only at low levels on HSCs, which may be associated with less hematologic toxicity if targeted by immunotherapy (32, 114117). A lower CRS rate was reported from a phase I trial exploring the first CLL-1xCD3-directed bispecific antibody MCLA-117 in R/R AML but with only 15% of patients achieving a partial response (118).

Cellular immunotherapy describes the adoptive transfer of genetically engineered autologous chimeric-antigen receptor (CAR)-T or -Natural Killer (NK) cells. Astonishing successes were reported from treatment of B-cell neoplasms with various CAR-T cell products and have led to their approval in the Europe and the U.S. (1928). As with BITEs and DARTs, CAR construct development against AML faces similar challenges in defining unique immunotargets on AML blasts. Lineage-specific antigens such as CD33 and CD123 are commonly expressed on AML blasts and evaluated as potential targets. Their expression on hematologic stem cells (HSCs) bears the risk of post-treatment bone marrow failure (32, 119, 120). As CAR-T cells commonly have a “memory effect”, hematologic toxicity might be even more severe compared to BITEs and DARTs.

One strategy to avoid the off-tumor toxicity is the development of AND-gated and NOT-gated CAR-T cells that engage two antigens to increase selectivity (121, 122). Perriello et al. developed cytokine-induced killer (CIK) cells with two CARs directed against CD123 and CD33. In this case, simultaneous binding of both CARs is necessary for a cytotoxic T-cell activation, because the CD33 CAR delivers the essential co-stimulatory signal (122). The authors also demonstrate that reduced binding activity of a CAR may increase selectivity by restricting reactivity to cells with high antigen expression. NOT-gates CARs represent an different approach to avoid off-tumor toxicity: Richards et al. developed CD93-directed CAR T-cells that express a second inhibitory CAR (iCAR) directed against an antigen present on endothelial cells but absent on myeloid blasts. This iCAR contains endodomains from ITIM-containing proteins including PD-1, TIM-3 or TIGIT delivering an inhibitory signal that interferes with the CAR T-cell activation signal (121).

So far, CAR-T-cells targeting CD33, CD123, or two antigens at once (e.g., CD33 and CLL-1; CD13 and TIM-3) are currently evaluated in early clinical trials (NCT03971799, NCT03795779, NCT03631576, NCT03190278, NCT03114670, NCT02159495, NCT04272125, NCT03222674, NCT04010877, NCT04097301). Three studies reported activity against heavily pretreated patients (123125), but longer follow-up efficacy data needed to draw more definitive conclusions are pending. For CD70, another immune target expressed on AML blast and low expression on HSCs, promising activity has been reported in preclinical AML models. Clinical trial evaluation is expected shortly (126, 127).

CAR-engineered NK cells may have potential advantages over CAR-T cells and be a promising alternative for two reasons: a) their HLA-class I independent tumor cell recognition allows maintaining intrinsic anti-tumor activity in case of antigen loss (128), and b) the lack of clonal expansion protects recipients from persistent graft versus host disease (GvHD) or long-term hematologic toxicity, reviewed in (129). First clinical applications have demonstrated encouraging anti-leukemic activity and tolerability with cord-blood-derived CD19-CAR NK cells against chronic lymphatic leukemia (130). CAR-NK cell products are effective against preclinical AML models in vitro and in vivo but clinical activity remains to be demonstrated (131).

While the efficacy of these concepts still needs improvement, the strong graft versus leukemia effect that has been observed over decades following allogenic SCT indicates that AML may still be prone to immunotherapy (132134). As mentioned above, one potential reason might be the particularly low mutational burden found in AML blasts compared to other cancers, which has been associated with generally lower responses to immune-based treatments (33, 34). Defining an AML-specific immunotarget that is not expressed on HSC is also an ongoing challenge for the development of potent immune-based treatments (32).

Combination of epigenetic treatment with immunotherapy

Epigenetic mechanisms have been implicated in contributing to the poor responses of AML to immunotherapy. One example is the silencing of HLA class II molecules observed in AML patients that relapsed after allogenic SCT (135137). This has been attributed to the DNA-hypermethylation of respective promotor regions (96). Therapeutic manipulation with HMAs to reverse promotor-methylation has successfully been used at relapse to boost graft-versus leukemia effects of donor lymphocyte infusions. However, this concept is less efficient with high leukemia burden (138140). Additional immune modulatory effects of HMA are currently being discussed. These include enhanced expression of tumor-associated antigens such as MAGE-1 and NY-ESO-1 (35, 36). Also, HMA-treatment is associated with tumor re-expression of endogenous retroviruses (ERVs) that is believed to improve T- and NK-cell activation via enhanced IFN-γ expression (141144), enhances tumor lymphocyte infiltration (145), and impairs expansion of regulatory T-cells (146), (Figure 1). The limited activity of HMAs commonly observed in the clinical setting may partly be explained by the upregulation of the immune inhibitory checkpoints (147, 148).

FIGURE 1
www.frontiersin.org

Figure 1 Epigenetic targeting in AML. Epigenetic regulators control transcription via chemical chromatin modifications, including histone protein and DNA (de-)methylation or histone (de-)acetylation that determine chromatin state. As therapeutic opportunities against AML, chromatin modifiers can alter leukemogenic gene expression, causing cell differentiation and proliferation arrest of the malignant blasts. Additional pro-immunogenic effects have recently been discussed, including an increased neoantigen-, immune checkpoint-, NK2GDL- and calreticulin expression on leukemic blasts and an augmented immune checkpoint expression and IFN-Y response of immune cells. The figure was created with BioRender.com.

HMA treatment has also been investigated in combination with immune checkpoint blockade in clinical trials. Encouraging results demonstrated a first phase II trial assessing the combination of PD-1 antibody nivolumab and azacitidine in R/R AML resulting in an ORR of 58% in HMA-naive and 22% in HMA-pretreated patients, respectively (149). Newly diagnosed and R/R patients achieved a CRc in 47% and 14% in a phase II trial assessing the combination of the PD-1 antibody pembrolizumab with azacitidine (150). Azacitidine combined with the anti-TIM-3 monoclonal antibody sabatolimab led to an ORR of 57% and a CRc of 30% in newly diagnosed AML in a phase Ib trial (151). The only randomized data available come from a trial assessing the anti-PD-L1 antibody durvalumab, Here, no significant benefit for the combination of durvalumab and azacitidine was observed over azacitidine alone in MDS/AML patients (152). Consistent with the data above, the authors of a recent meta-analysis concluded that the activity of checkpoint inhibitors is generally low in the relapsed/refractory AML setting (153). Further studies are currently ongoing (Table 1).

TABLE 1
www.frontiersin.org

Table 1 Current clinical trials evaluating combinations of epigenetic targeting and immunotherapy in AML.

HMAs in combination with immune checkpoint inhibitors were also assessed in the post-transplant setting, with only a few responses reported and increased immune-related toxicity (12, 154). This was demonstrated by the combination of avelumab and azacitidine, resulting in CR rates of only 10.5% and an increased risk of severe graft versus host disease (155). Several clinical trials are ongoing and will allow more definitive conclusions concerning efficacy and safety.

HDAC inhibitors can also induce tumor-associated antigens, improve antigen presentation, influence T-cell trafficking and activity but also increase PD-1 expression (156159). Several trials reported responses to HDAC inhibitors in combination with checkpoint blockade in solid tumors (160). However, in R/R MDS/AML patients, no activity of this concept has been reported in a recent phase 1b study assessing pembrolizumab plus entinostat with no responses in any of the patients (161).

In contrast, encouraging activity of combining the anti-CD47 antibody magrolimab with azacitidine and the BCL2-inhibitor venetoclax was reported from a phase I/II trial in the adverse TP53 mutated AML subtype. CRc rates were 63%, with an average one-year overall survival of 53% (162). Two randomized phase III trials are currently ongoing (NCT05079230, NCT04778397, Table 1).

HMAs and HDAC inhibitors were also reported to increase the expression of AML-associated antigens such as CD33 (163) and may therefore be a suitable combination partner for BiTEs, DARTs, and CAR-T, and -NK-cell treatment. Experimental in vitro and in vivo studies indicated improved T-cell activity for combined HMA or HDAC inhibitors with CD33-, CD123-, and CD70-directed CAR-T cells or bispecific antibodies (126, 164166).

Multiple lines of evidence support the view that epigenetic silencing of NKG2D-ligands (NKG2DL) contributes to impaired NK-cell function, which was reversed with HMA treatment in studies on cultured NK cells (167169). In preclinical AML models, decitabine enhanced the activity of BI836858, an anti-CD33 antibody that also engages NK cells via CD16 (170). In contrast, combining the NK-cell engaging and CD123 targeting monoclonal talacotuzumab with decitabine could not improve responses over decitabine alone in a phase II/III trial (171). Based on these data, combinations of HMAs with bispecific antibodies or CAR-T/CAR-NK cell treatment may also constitute an attractive combination. A comprehensive assessment of the biological effects of HMAs on cellular treatments is required before these combination treatments can be introduced into clinical testing.

Combining the more selective second-generation targeted epigenetic drugs with cancer immunotherapy appears attractive as it may be associated with fewer unintended effects and more efficacy. However, it also requires detailed studies before those concepts enter clinical trials. In particular, more data are needed on how these individual compounds may modulate effector and regulatory immune cell function in the context of substance-specific effects in leukemia cells. Most selective epigenetic compounds, for example, IDH or menin inhibitors, alter specific gene expression and induce differentiation (54, 55, 84, 93, 94), (Figure 1). These effects may represent a synergistic opportunity for combinatorial approaches as they commonly lead to the induction of surface antigen expression that may be utilized for immunotherapy, as reported with other targeted agents (172). Several other compound-specific effects may confer synergy with immunotherapeutic approaches: BET inhibitors, for instance, have been reported to impair PD-1 expression and T-cell exhaustion in vitro (173). Accordingly, improved T-cell expansion and anti-tumor efficacy have been observed in an adoptive T-cell transfer model upon JQ1 treatment (174). In a landmark study, it was observed that LSD1 inhibition stimulated T-cell-mediated anti-tumor responses by inducing endogenous ERV expression in cancer cells that resulted in type 1 interferon activation (175). Confirmative studies are needed before these approaches can be translated into clinical applications.

Summary and outlook

As outlined above, immunotherapy has dramatically improved treatment outcomes in patients with many cancers while these approaches have been far less successful in AML.

While the detailed mechanisms behind the relative resistance against immunotherapy remain obscure, the low immunogenicity of myeloid blasts for immune checkpoint blockade (31, 33, 34) and the difficulties in defining AML-specific antigens not expressed on HSCs for immune-directed treatment (32, 119, 120) remains an unsolved challenge. Epigenetic manipulation was shown to improve the responses to immunotherapy by inducing neoantigens, increasing antigen presentation, and co-regulating immune checkpoints (35, 36, 96, 141144, 146148). Clinical trials evaluating the combination of non-selective epigenetic drugs (such as HMAs) with checkpoint inhibitors have mainly reported modest activity in the R/R AML setting (12, 153, 176), while approaches combining the anti-CD47 antibody magrolimab with azacitidine with or without venetoclax resulted in very promising response rates in clinical trials (162, 177). Clinical data for the combination of HMAs with cellular immunotherapy is pending, while CAR-NK cell concepts seem auspicious due to their only temporary toxicity for the normal hematopoiesis (129). Promising strategies include the introduction of (second-generation) targeted epigenetic drugs into immunotherapeutic treatment regimens. These drugs commonly have less adverse effects and their common ability to release the differentiation block in AML blasts accompanied by antigen-induction may enhance cellular immunotherapy. Studies that define specific effects of these drugs on various immune cells are underway to enable translation of these concepts into clinical investigation.

Author contributions

JR: Writing – original draft, Writing – review & editing. EU: Writing – review & editing. MK: Writing – review & editing, Writing – original draft.

Funding

This work was supported by grants from the Deutsche Forschungsgemeinschaft (DFG) to MWMK and EU (SFB1292/2/ TP12) and MWMK (KU-2688/2-2).

Conflict of interest

MWMK receives honoraria and is a consultant for Pfizer, Kura Oncology, Jazz Pharmaceuticals, Bristol-Myers Squibb/Celgene Abbvie, and Servier; is on the speakers bureau of Gilead and receives travel support from Daiichi Sankyo. JR has received travel support from Abbvie. EU has a sponsored research project with Gilead and BMS.

EU declared that they were an editorial board member of Frontiers, at the time of submission. This had no impact on the peer review process and the final decision.

Publisher’s note

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

References

1. Arber DA, Orazi A, Hasserjian RP, Borowitz MJ, Calvo KR, Kvasnicka HM, et al. International Consensus Classification of Myeloid Neoplasms and Acute Leukemias: integrating morphologic, clinical, and genomic data. Blood (2022) 140(11):1200–28. doi: 10.1182/blood.2022015850

PubMed Abstract | CrossRef Full Text | Google Scholar

2. Döhner H, Wei AH, Appelbaum FR, Craddock C, DiNardo CD, Dombret H, et al. Diagnosis and management of AML in adults: 2022 recommendations from an international expert panel on behalf of the ELN. Blood (2022) 140(12):1345–77. doi: 10.1182/blood.2022016867

PubMed Abstract | CrossRef Full Text | Google Scholar

3. Röllig C, Kramer M, Schliemann C, Mikesch JH, Steffen B, Krämer A, et al. Does time from diagnosis to treatment affect the prognosis of patients with newly diagnosed acute myeloid leukemia? Blood (2020) 136(7):823–30. doi: 10.1182/blood.2019004583

PubMed Abstract | CrossRef Full Text | Google Scholar

4. Sasaki K, Ravandi F, Kadia TM, DiNardo CD, Short NJ, Borthakur G, et al. De novo acute myeloid leukemia: A population-based study of outcome in the United States based on the Surveillance, Epidemiology, and End Results (SEER) database, 1980 to 2017. Cancer (2021) 127(12):2049–61. doi: 10.1002/cncr.33458

PubMed Abstract | CrossRef Full Text | Google Scholar

5. Rausch C, Rothenberg-Thurley M, Dufour A, Schneider S, Gittinger H, Sauerland C, et al. Validation and refinement of the 2022 European LeukemiaNet genetic risk stratification of acute myeloid leukemia. Leukemia (2023) 37(6):1234–44. doi: 10.1038/s41375-023-01884-2

PubMed Abstract | CrossRef Full Text | Google Scholar

6. DiNardo CD, Jonas BA, Pullarkat V, Thirman MJ, Garcia JS, Wei AH, et al. Azacitidine and venetoclax in previously untreated acute myeloid leukemia. New Engl J Med (2020) 383(7):617–29. doi: 10.1056/NEJMoa2012971

PubMed Abstract | CrossRef Full Text | Google Scholar

7. Dombret H, Seymour JF, Butrym A, Wierzbowska A, Selleslag D, Jang JH, et al. International phase 3 study of azacitidine vs conventional care regimens in older patients with newly diagnosed AML with >30% blasts. Blood (2015) 126(3):291–9. doi: 10.1182/blood-2015-01-621664

PubMed Abstract | CrossRef Full Text | Google Scholar

8. Issa GC, Aldoss I, DiPersio J, Cuglievan B, Stone R, Arellano M, et al. The menin inhibitor revumenib in KMT2A-rearranged or NPM1-mutant leukaemia. Nature (2023) 615(7954):920–4. doi: 10.1038/s41586-023-05812-3

PubMed Abstract | CrossRef Full Text | Google Scholar

9. Montesinos P, Recher C, Vives S, Zarzycka E, Wang J, Bertani G, et al. Ivosidenib and azacitidine in IDH1-mutated acute myeloid leukemia. N Engl J Med (2022) 386(16):1519–31. doi: 10.1056/NEJMoa2117344

PubMed Abstract | CrossRef Full Text | Google Scholar

10. Dinardo CD, Schuh AC, Stein EM, Montesinos P, Wei A, De Botton S, et al. Effect of enasidenib (ENA) plus azacitidine (AZA) on complete remission and overall response versus AZA monotherapy in mutant-IDH2 (mIDH2) newly diagnosed acute myeloid leukemia (ND-AML). JCO (2020) 38(15_suppl):7501–1. doi: 10.1200/JCO.2020.38.15_suppl.7501

CrossRef Full Text | Google Scholar

11. Thieblemont C, Phillips T, Ghesquieres H, Cheah CY, Clausen MR, Cunningham D, et al. Epcoritamab, a novel, subcutaneous CD3xCD20 bispecific T-cell–engaging antibody, in relapsed or refractory large B-cell lymphoma: dose expansion in a phase I/II trial. J Clin Oncol (2023) 41(12):2238–47. doi: 10.1200/JCO.22.01725

PubMed Abstract | CrossRef Full Text | Google Scholar

12. Davids MS, Kim HT, Bachireddy P, Costello C, Liguori R, Savell A, et al. Ipilimumab for patients with relapse after allogeneic transplantation. N Engl J Med (2016) 375(2):143–53. doi: 10.1056/NEJMoa1601202

PubMed Abstract | CrossRef Full Text | Google Scholar

13. Robert C, Long GV, Brady B, Dutriaux C, Maio M, Mortier L, et al. Nivolumab in previously untreated melanoma without BRAF mutation. N Engl J Med (2015) 372(4):320–30. doi: 10.1056/NEJMoa1412082

PubMed Abstract | CrossRef Full Text | Google Scholar

14. Motzer RJ, Escudier B, McDermott DF, George S, Hammers HJ, Srinivas S, et al. Nivolumab versus everolimus in advanced renal-cell carcinoma. N Engl J Med (2015) 373(19):1803–13. doi: 10.1056/NEJMoa1510665

PubMed Abstract | CrossRef Full Text | Google Scholar

15. Brahmer JR, Tykodi SS, Chow LQM, Hwu WJ, Topalian SL, Hwu P, et al. Safety and activity of anti–PD-L1 antibody in patients with advanced cancer. New Engl J Med (2012) 366(26):2455–65. doi: 10.1056/NEJMoa1200694

PubMed Abstract | CrossRef Full Text | Google Scholar

16. Topalian SL, Sznol M, McDermott DF, Kluger HM, Carvajal RD, Sharfman WH, et al. Survival, durable tumor remission, and long-term safety in patients with advanced melanoma receiving nivolumab. JCO (2014) 53:0105. doi: 10.1200/JCO.2013.53.0105

CrossRef Full Text | Google Scholar

17. Ansell SM, Lesokhin AM, Borrello I, Halwani A, Scott EC, Gutierrez M, et al. PD-1 blockade with nivolumab in relapsed or refractory Hodgkin’s lymphoma. N Engl J Med (2015) 372(4):311–9. doi: 10.1056/NEJMoa1411087

PubMed Abstract | CrossRef Full Text | Google Scholar

18. Reck M, Rodríguez-Abreu D, Robinson AG, Hui R, Csőszi T, Fülöp A, et al. Five-year outcomes with pembrolizumab versus chemotherapy for metastatic non–small-cell lung cancer with PD-L1 tumor proportion score ≥ 50%. JCO (2021) 39(21):2339–49. doi: 10.1200/JCO.21.00174

CrossRef Full Text | Google Scholar

19. Grupp SA, Kalos M, Barrett D, Aplenc R, Porter DL, Rheingold SR, et al. Chimeric antigen receptor-modified T cells for acute lymphoid leukemia. N Engl J Med (2013) 368(16):1509–18. doi: 10.1056/NEJMoa1215134

PubMed Abstract | CrossRef Full Text | Google Scholar

20. Maude SL, Laetsch TW, Buechner J, Rives S, Boyer M, Bittencourt H, et al. Tisagenlecleucel in children and young adults with B-cell lymphoblastic leukemia. N Engl J Med (2018) 378(5):439–48. doi: 10.1056/NEJMoa1709866

PubMed Abstract | CrossRef Full Text | Google Scholar

21. Schuster SJ, Bishop MR, Tam CS, Waller EK, Borchmann P, McGuirk JP, et al. Tisagenlecleucel in adult relapsed or refractory diffuse large B-cell lymphoma. New Engl J Med (2019) 380(1):45–56. doi: 10.1056/NEJMoa1804980

PubMed Abstract | CrossRef Full Text | Google Scholar

22. Neelapu SS, Locke FL, Bartlett NL, Lekakis LJ, Miklos DB, Jacobson CA, et al. Axicabtagene ciloleucel CAR T-cell therapy in refractory large B-cell lymphoma. N Engl J Med (2017) 377(26):2531–44. doi: 10.1056/NEJMoa1707447

PubMed Abstract | CrossRef Full Text | Google Scholar

23. Wang M, Munoz J, Goy A, Locke FL, Jacobson CA, Hill BT, et al. KTE-X19 CAR T-cell therapy in relapsed or refractory mantle-cell lymphoma. N Engl J Med (2020) 382(14):1331–42. doi: 10.1056/NEJMoa1914347

PubMed Abstract | CrossRef Full Text | Google Scholar

24. Shah BD, Ghobadi A, Oluwole OO, Logan AC, Boissel N, Cassaday RD, et al. KTE-X19 for relapsed or refractory adult B-cell acute lymphoblastic leukaemia: phase 2 results of the single-arm, open-label, multicentre ZUMA-3 study. Lancet (2021) 398(10299):491–502. doi: 10.1016/S0140-6736(21)01222-8

PubMed Abstract | CrossRef Full Text | Google Scholar

25. Kamdar M, Solomon SR, Arnason J, Johnston PB, Glass B, Bachanova V, et al. Lisocabtagene maraleucel versus standard of care with salvage chemotherapy followed by autologous stem cell transplantation as second-line treatment in patients with relapsed or refractory large B-cell lymphoma (TRANSFORM): results from an interim analysis of an open-label, randomised, phase 3 trial. Lancet (2022) 399(10343):2294–308. doi: 10.1016/S0140-6736(22)00662-6

PubMed Abstract | CrossRef Full Text | Google Scholar

26. Munshi NC, Anderson LD, Shah N, Madduri D, Berdeja J, Lonial S, et al. Idecabtagene vicleucel in relapsed and refractory multiple myeloma. New Engl J Med (2021) 384(8):705–16. doi: 10.1056/NEJMoa2024850

PubMed Abstract | CrossRef Full Text | Google Scholar

27. Martin T, Usmani SZ, Berdeja JG, Agha M, Cohen AD, Hari P, et al. Ciltacabtagene autoleucel, an anti-B-cell maturation antigen chimeric antigen receptor T-cell therapy, for relapsed/refractory multiple myeloma: CARTITUDE-1 2-year follow-up. J Clin Oncol (2023) 41(6):1265–74. doi: 10.1200/JCO.22.00842

PubMed Abstract | CrossRef Full Text | Google Scholar

28. Kantarjian H, Stein A, Gökbuget N, Fielding AK, Schuh AC, Ribera JM, et al. Blinatumomab versus chemotherapy for advanced acute lymphoblastic leukemia. N Engl J Med (2017) 376(9):836–47. doi: 10.1056/NEJMoa1609783

PubMed Abstract | CrossRef Full Text | Google Scholar

29. Budde LE, Sehn LH, Matasar M, Schuster SJ, Assouline S, Giri P, et al. Safety and efficacy of mosunetuzumab, a bispecific antibody, in patients with relapsed or refractory follicular lymphoma: a single-arm, multicentre, phase 2 study. Lancet Oncol (2022) 23(8):1055–65. doi: 10.1016/S1470-2045(22)00335-7

PubMed Abstract | CrossRef Full Text | Google Scholar

30. Dickinson MJ, Carlo-Stella C, Morschhauser F, Bachy E, Corradini P, Iacoboni G, et al. Glofitamab for relapsed or refractory diffuse large B-cell lymphoma. New Engl J Med (2022) 387(24):2220–31. doi: 10.1056/NEJMoa2206913

PubMed Abstract | CrossRef Full Text | Google Scholar

31. Daver N, Alotaibi AS, Bücklein V, Subklewe M. T-cell-based immunotherapy of acute myeloid leukemia: current concepts and future developments. Leukemia (2021) 35(7):1843–63. doi: 10.1038/s41375-021-01253-x

PubMed Abstract | CrossRef Full Text | Google Scholar

32. Haubner S, Perna F, Köhnke T, Schmidt C, Berman S, Augsberger C, et al. Coexpression profile of leukemic stem cell markers for combinatorial targeted therapy in AML. Leukemia (2019) 33(1):64–74. doi: 10.1038/s41375-018-0180-3

PubMed Abstract | CrossRef Full Text | Google Scholar

33. Alexandrov LB, Nik-Zainal S, Wedge DC, Aparicio SAJR, Behjati S, Biankin AV, et al. Signatures of mutational processes in human cancer. Nature (2013) 500(7463):415–21. doi: 10.1038/nature12477

PubMed Abstract | CrossRef Full Text | Google Scholar

34. Yarchoan M, Hopkins A, Jaffee EM. Tumor mutational burden and response rate to PD-1 inhibition. N Engl J Med (2017) 377(25):2500–1. doi: 10.1056/NEJMc1713444

PubMed Abstract | CrossRef Full Text | Google Scholar

35. Almstedt M, Blagitko-Dorfs N, Duque-Afonso J, Karbach J, Pfeifer D, Jäger E, et al. The DNA demethylating agent 5-aza-2′-deoxycytidine induces expression of NY-ESO-1 and other cancer/testis antigens in myeloid leukemia cells. Leukemia Res (2010) 34(7):899–905. doi: 10.1016/j.leukres.2010.02.004

CrossRef Full Text | Google Scholar

36. Weber J, Salgaller M, Samid D, Johnson B, Herlyn M, Lassam N, et al. Expression of the MAGE-1 tumor antigen is up-regulated by the demethylating agent 5-aza-2’-deoxycytidine. Cancer Res (1994) 54(7):1766–71.

PubMed Abstract | Google Scholar

37. Papaemmanuil E, Gerstung M, Bullinger L, Gaidzik VI, Paschka P, Roberts ND, et al. Genomic classification and prognosis in acute myeloid leukemia. N Engl J Med (2016) 374(23):2209–21. doi: 10.1056/NEJMoa1516192

PubMed Abstract | CrossRef Full Text | Google Scholar

38. Kandoth C, McLellan MD, Vandin F, Ye K, Niu B, Lu C, et al. Mutational landscape and significance across 12 major cancer types. Nature (2013) 502(7471):333–9. doi: 10.1038/nature12634

PubMed Abstract | CrossRef Full Text | Google Scholar

39. Dawson MA, Kouzarides T, Huntly BJP. Targeting epigenetic readers in cancer. N Engl J Med (2012) 367(7):647–57. doi: 10.1056/NEJMra1112635

PubMed Abstract | CrossRef Full Text | Google Scholar

40. Baylin SB, Jones PA. Epigenetic determinants of cancer. Cold Spring Harb Perspect Biol (2016) 8(9):a019505. doi: 10.1101/cshperspect.a019505

PubMed Abstract | CrossRef Full Text | Google Scholar

41. Sasca D, Guezguez B, Kühn MWM. Next generation epigenetic modulators to target myeloid neoplasms. Curr Opin Hematol (2021) 28(5):356–63. doi: 10.1097/MOH.0000000000000673

PubMed Abstract | CrossRef Full Text | Google Scholar

42. Stomper J, Rotondo JC, Greve G, Lübbert M. Hypomethylating agents (HMA) for the treatment of acute myeloid leukemia and myelodysplastic syndromes: mechanisms of resistance and novel HMA-based therapies. Leukemia (2021) 35(7):1873–89. doi: 10.1038/s41375-021-01218-0

PubMed Abstract | CrossRef Full Text | Google Scholar

43. San José-Enériz E, Gimenez-Camino N, Agirre X, Prosper F. HDAC inhibitors in acute myeloid leukemia. Cancers (Basel) (2019) 11(11):1794. doi: 10.3390/cancers11111794

PubMed Abstract | CrossRef Full Text | Google Scholar

44. Kirschbaum MH, Foon KA, Frankel P, Ruel C, Pulone B, Tuscano JM, et al. A phase 2 study of belinostat (PXD101) in patients with relapsed or refractory acute myeloid leukemia or patients over the age of 60 with newly diagnosed acute myeloid leukemia: a California Cancer Consortium Study. Leuk Lymph. (2014) 55(10):2301–4. doi: 10.3109/10428194.2013.877134

CrossRef Full Text | Google Scholar

45. Schlenk RF, Krauter J, Raffoux E, Kreuzer KA, Schaich M, Noens L, et al. Panobinostat monotherapy and combination therapy in patients with acute myeloid leukemia: results from two clinical trials. Haematologica (2018) 103(1):e25–8. doi: 10.3324/haematol.2017.172411

PubMed Abstract | CrossRef Full Text | Google Scholar

46. Abaza YM, Kadia TM, Jabbour EJ, Konopleva MY, Borthakur G, Ferrajoli A, et al. Phase 1 dose escalation multicenter trial of pracinostat alone and in combination with azacitidine in patients with advanced hematologic Malignancies. Cancer (2017) 123(24):4851–9. doi: 10.1002/cncr.30949

PubMed Abstract | CrossRef Full Text | Google Scholar

47. Stahl M, Zeidan AM. Hypomethylating agents in combination with histone deacetylase inhibitors in higher risk myelodysplastic syndromes: Is there a light at the end of the tunnel? Cancer (2017) 123(6):911–4. doi: 10.1002/cncr.30532

PubMed Abstract | CrossRef Full Text | Google Scholar

48. Issa JP, Garcia-Manero G, Huang X, Cortes J, Ravandi F, Jabbour E, et al. Results of phase 2 randomized study of low-dose decitabine with or without valproic acid in patients with myelodysplastic syndrome and acute myelogenous leukemia. Cancer (2015) 121(4):556–61. doi: 10.1002/cncr.29085

PubMed Abstract | CrossRef Full Text | Google Scholar

49. Sekeres MA, Othus M, List AF, Odenike O, Stone RM, Gore SD, et al. Randomized phase II study of azacitidine alone or in combination with lenalidomide or with vorinostat in higher-risk myelodysplastic syndromes and chronic myelomonocytic leukemia: North American intergroup study SWOG S1117. J Clin Oncol (2017) 35(24):2745–53. doi: 10.1200/JCO.2015.66.2510

PubMed Abstract | CrossRef Full Text | Google Scholar

50. Jang MK, Mochizuki K, Zhou M, Jeong HS, Brady JN, Ozato K. The bromodomain protein Brd4 is a positive regulatory component of P-TEFb and stimulates RNA polymerase II-dependent transcription. Mol Cell (2005) 19(4):523–34. doi: 10.1016/j.molcel.2005.06.027

PubMed Abstract | CrossRef Full Text | Google Scholar

51. Roe JS, Mercan F, Rivera K, Pappin DJ, Vakoc CR. BET bromodomain inhibition suppresses the function of hematopoietic transcription factors in acute myeloid leukemia. Mol Cell (2015) 58(6):1028–39. doi: 10.1016/j.molcel.2015.04.011

PubMed Abstract | CrossRef Full Text | Google Scholar

52. Fiskus W, Cai T, DiNardo CD, Kornblau SM, Borthakur G, Kadia TM, et al. Superior efficacy of cotreatment with BET protein inhibitor and BCL2 or MCL1 inhibitor against AML blast progenitor cells. Blood Cancer J (2019) 9(2):1–13. doi: 10.1038/s41408-018-0165-5

CrossRef Full Text | Google Scholar

53. Borthakur G, Odenike O, Aldoss I, Rizzieri DA, Prebet T, Chen C, et al. A phase 1 study of the pan-bromodomain and extraterminal inhibitor mivebresib (ABBV-075) alone or in combination with venetoclax in patients with relapsed/refractory acute myeloid leukemia. Cancer (2021) 127(16):2943–53. doi: 10.1002/cncr.33590

PubMed Abstract | CrossRef Full Text | Google Scholar

54. Bernt KM, Zhu N, Sinha AU, Vempati S, Faber J, Krivtsov AV, et al. MLL-rearranged leukemia is dependent on aberrant H3K79 methylation by DOT1L. Cancer Cell (2011) 20(1):66–78. doi: 10.1016/j.ccr.2011.06.010

PubMed Abstract | CrossRef Full Text | Google Scholar

55. Kühn MWM, Song E, Feng Z, Sinha A, Chen CW, Deshpande AJ, et al. Targeting chromatin regulators inhibits leukemogenic gene expression in NPM1 mutant leukemia. Cancer Discovery (2016) 6(10):1166–81. doi: 10.1158/2159-8290.CD-16-0237

PubMed Abstract | CrossRef Full Text | Google Scholar

56. Stein EM, Garcia-Manero G, Rizzieri DA, Tibes R, Berdeja JG, Savona MR, et al. The DOT1L inhibitor pinometostat reduces H3K79 methylation and has modest clinical activity in adult acute leukemia. Blood (2018) 131(24):2661–9. doi: 10.1182/blood-2017-12-818948

PubMed Abstract | CrossRef Full Text | Google Scholar

57. Radzisheuskaya A, Shliaha PV, Grinev V, Lorenzini E, Kovalchuk S, Shlyueva D, et al. PRMT5 methylome profiling uncovers a direct link to splicing regulation in acute myeloid leukemia. Nat Struct Mol Biol (2019) 26(11):999–1012. doi: 10.1038/s41594-019-0313-z

PubMed Abstract | CrossRef Full Text | Google Scholar

58. Kim H, Ronai ZA. PRMT5 function and targeting in cancer. Cell Stress (2020) 4(8):199–215. doi: 10.15698/cst2020.08.228

PubMed Abstract | CrossRef Full Text | Google Scholar

59. Kaushik S, Liu F, Veazey KJ, Gao G, Das P, Neves LF, et al. Genetic deletion or small-molecule inhibition of the arginine methyltransferase PRMT5 exhibit anti-tumoral activity in mouse models of MLL-rearranged AML. Leukemia (2018) 32(2):499–509. doi: 10.1038/leu.2017.206

PubMed Abstract | CrossRef Full Text | Google Scholar

60. Tarighat SS, Santhanam R, Frankhouser D, Radomska HS, Lai H, Anghelina M, et al. The dual epigenetic role of PRMT5 in acute myeloid leukemia: gene activation and repression via histone arginine methylation. Leukemia (2016) 30(4):789–99. doi: 10.1038/leu.2015.308

PubMed Abstract | CrossRef Full Text | Google Scholar

61. Feustel K, Falchook GS. Protein arginine methyltransferase 5 (PRMT5) inhibitors in oncology clinical trials: A review. J Immunother Precis Oncol (2022) 5(3):58–67. doi: 10.36401/JIPO-22-1

PubMed Abstract | CrossRef Full Text | Google Scholar

62. Patel MR, Monga V, Jauhari S, Stevens D, Masarova L, McKean M, et al. A phase 1 dose escalation study of protein arginine methyltransferase 5 (PRMT5) inhibitor PRT543 in patients with myeloid malignancies. Blood (2021) 138(Supplement 1):2609. doi: 10.1182/blood-2021-150938

CrossRef Full Text | Google Scholar

63. El-Khoueiry AB, Clarke J, Neff T, Crossman T, Ratia N, Rathi C, et al. Phase 1 study of GSK3368715, a type I PRMT inhibitor, in patients with advanced solid tumors. Br J Cancer. (2023) 129(2):309–17. doi: 10.1038/s41416-023-02276-0

PubMed Abstract | CrossRef Full Text | Google Scholar

64. Monga V, Johanns TM, Stupp R, Chandra S, Falchook GS, Giglio P, et al. A phase 1 study of the protein arginine methyltransferase 5 (PRMT5) brain-penetrant inhibitor PRT811 in patients (pts) with recurrent high-grade glioma or uveal melanoma (UM). JCO (2023) 41(16_suppl):3008–8. doi: 10.1200/JCO.2023.41.16_suppl.3008

CrossRef Full Text | Google Scholar

65. Watts JM, Bradley TJ, Thomassen A, Brunner AM, Minden MD, Papadantonakis N, et al. A phase I/II study to investigate the safety and clinical activity of the protein arginine methyltransferase 5 inhibitor GSK3326595 in subjects with myelodysplastic syndrome and acute myeloid leukemia. Blood (2019) 134(Supplement_1):2656. doi: 10.1182/blood-2019-130337

CrossRef Full Text | Google Scholar

66. Rinke J, Chase A, Cross NCP, Hochhaus A, Ernst T. EZH2 in myeloid Malignancies. Cells (2020) 9(7):1639. doi: 10.3390/cells9071639

PubMed Abstract | CrossRef Full Text | Google Scholar

67. Margueron R, Reinberg D. The Polycomb complex PRC2 and its mark in life. Nature (2011) 469(7330):343–9. doi: 10.1038/nature09784

PubMed Abstract | CrossRef Full Text | Google Scholar

68. Morin RD, Johnson NA, Severson TM, Mungall AJ, An J, Goya R, et al. Somatic mutations altering EZH2 (Tyr641) in follicular and diffuse large B-cell lymphomas of germinal-center origin. Nat Genet (2010) 42(2):181–5. doi: 10.1038/ng.518

PubMed Abstract | CrossRef Full Text | Google Scholar

69. Kim KH, Roberts CWM. Targeting EZH2 in cancer. Nat Med (2016) 22(2):128–34. doi: 10.1038/nm.4036

PubMed Abstract | CrossRef Full Text | Google Scholar

70. Gounder M, Schöffski P, Jones RL, Agulnik M, Cote GM, Villalobos VM, et al. Tazemetostat in advanced epithelioid sarcoma with loss of INI1/SMARCB1: an international, open-label, phase 2 basket study. Lancet Oncol (2020) 21(11):1423–32. doi: 10.1016/S1470-2045(20)30451-4

PubMed Abstract | CrossRef Full Text | Google Scholar

71. Italiano A, Soria JC, Toulmonde M, Michot JM, Lucchesi C, Varga A, et al. Tazemetostat, an EZH2 inhibitor, in relapsed or refractory B-cell non-Hodgkin lymphoma and advanced solid tumours: a first-in-human, open-label, phase 1 study. Lancet Oncol (2018) 19(5):649–59. doi: 10.1016/S1470-2045(18)30145-1

PubMed Abstract | CrossRef Full Text | Google Scholar

72. Morschhauser F, Tilly H, Chaidos A, McKay P, Phillips T, Assouline S, et al. Tazemetostat for patients with relapsed or refractory follicular lymphoma: an open-label, single-arm, multicentre, phase 2 trial. Lancet Oncol (2020) 21(11):1433–42. doi: 10.1016/S1470-2045(20)30441-1

PubMed Abstract | CrossRef Full Text | Google Scholar

73. Basheer F, Giotopoulos G, Meduri E, Yun H, Mazan M, Sasca D, et al. Contrasting requirements during disease evolution identify EZH2 as a therapeutic target in AML. J Exp Med (2019) 216(4):966–81. doi: 10.1084/jem.20181276

PubMed Abstract | CrossRef Full Text | Google Scholar

74. Göllner S, Oellerich T, Agrawal-Singh S, Schenk T, Klein HU, Rohde C, et al. Loss of the histone methyltransferase EZH2 induces resistance to multiple drugs in acute myeloid leukemia. Nat Med (2017) 23(1):69–78. doi: 10.1038/nm.4247

PubMed Abstract | CrossRef Full Text | Google Scholar

75. Greif PA, Hartmann L, Vosberg S, Stief SM, Mattes R, Hellmann I, et al. Evolution of cytogenetically normal acute myeloid leukemia during therapy and relapse: an exome sequencing study of 50 patients. Clin Cancer Res (2018) 24(7):1716–26. doi: 10.1158/1078-0432.CCR-17-2344

PubMed Abstract | CrossRef Full Text | Google Scholar

76. Kempf JM, Weser S, Bartoschek MD, Metzeler KH, Vick B, Herold T, et al. Loss-of-function mutations in the histone methyltransferase EZH2 promote chemotherapy resistance in AML. Sci Rep (2021) 11(1):5838. doi: 10.1038/s41598-021-84708-6

PubMed Abstract | CrossRef Full Text | Google Scholar

77. Porazzi P, Petruk S, Pagliaroli L, De Dominici M, Deming D, Puccetti MV, et al. Targeting chemotherapy to decondensed H3K27me3-marked chromatin of AML cells enhances leukemia suppression. Cancer Res (2022) 82(3):458–71. doi: 10.1158/0008-5472.CAN-21-1297

PubMed Abstract | CrossRef Full Text | Google Scholar

78. Xu B, On DM, Ma A, Parton T, Konze KD, Pattenden SG, et al. Selective inhibition of EZH2 and EZH1 enzymatic activity by a small molecule suppresses MLL-rearranged leukemia. Blood (2015) 125(2):346–57. doi: 10.1182/blood-2014-06-581082

PubMed Abstract | CrossRef Full Text | Google Scholar

79. Cusan M, Cai SF, Mohammad HP, Krivtsov A, Chramiec A, Loizou E, et al. LSD1 inhibition exerts its antileukemic effect by recommissioning PU.1- and C/EBPα-dependent enhancers in AML. Blood (2018) 131(15):1730–42. doi: 10.1182/blood-2017-09-807024

PubMed Abstract | CrossRef Full Text | Google Scholar

80. Salamero O, Somervaille TCP, Molero A, Acuña-Cruz E, Perez-Simon JA, Coll R, et al. Iadademstat combination with azacitidine is a safe and effective treatment in first line acute myeloid leukemia. Final results of the alice trial. Blood (2022) 140(Supplement 1):1711–3. doi: 10.1182/blood-2022-168945

CrossRef Full Text | Google Scholar

81. Maes T, Mascaró C, Tirapu I, Estiarte A, Ciceri F, Lunardi S, et al. ORY-1001, a potent and selective covalent KDM1A inhibitor, for the treatment of acute leukemia. Cancer Cell (2018) 33(3):495–511.e12. doi: 10.1016/j.ccell.2018.02.002

PubMed Abstract | CrossRef Full Text | Google Scholar

82. Dang L, White DW, Gross S, Bennett BD, Bittinger MA, Driggers EM, et al. Cancer-associated IDH1 mutations produce 2-hydroxyglutarate. Nature (2009) 462(7274):739–44. doi: 10.1038/nature08617

PubMed Abstract | CrossRef Full Text | Google Scholar

83. Figueroa ME, Abdel-Wahab O, Lu C, Ward PS, Patel J, Shih A, et al. Leukemic IDH1 and IDH2 mutations result in a hypermethylation phenotype, disrupt TET2 function, and impair hematopoietic differentiation. Cancer Cell (2010) 18(6):553–67. doi: 10.1016/j.ccr.2010.11.015

PubMed Abstract | CrossRef Full Text | Google Scholar

84. Wang F, Travins J, DeLaBarre B, Penard-Lacronique V, Schalm S, Hansen E, et al. Targeted inhibition of mutant IDH2 in leukemia cells induces cellular differentiation. Science (2013) 340(6132):622–6. doi: 10.1126/science.1234769

PubMed Abstract | CrossRef Full Text | Google Scholar

85. Stein EM, DiNardo CD, Pollyea DA, Fathi AT, Roboz GJ, Altman JK, et al. Enasidenib in mutant IDH2 relapsed or refractory acute myeloid leukemia. Blood (2017) 130(6):722–31. doi: 10.1182/blood-2017-04-779405

PubMed Abstract | CrossRef Full Text | Google Scholar

86. De Botton S, Montesinos P, Vives Polo S, Zarzycka E, Wang J, Riva M, et al. Updated efficacy and safety data from the AGILE study in patients with newly diagnosed acute myeloid leukemia treated with ivosidenib + azacitidine compared to placebo + azacitidine. JCO (2023) 41(16_suppl):7012–2. doi: 10.1200/JCO.2023.41.16_suppl.7012

CrossRef Full Text | Google Scholar

87. Borkin D, He S, Miao H, Kempinska K, Pollock J, Chase J, et al. Pharmacologic inhibition of the menin-MLL interaction blocks progression of MLL leukemia in vivo. Cancer Cell (2015) 27(4):589–602. doi: 10.1016/j.ccell.2015.02.016

PubMed Abstract | CrossRef Full Text | Google Scholar

88. Klossowski S, Miao H, Kempinska K, Wu T, Purohit T, Kim E, et al. Menin inhibitor MI-3454 induces remission in MLL1-rearranged and NPM1-mutated models of leukemia. J Clin Invest (2020) 130(2):981–97. doi: 10.1172/JCI129126

PubMed Abstract | CrossRef Full Text | Google Scholar

89. Krivtsov AV, Evans K, Gadrey JY, Eschle BK, Hatton C, Uckelmann HJ, et al. A menin-MLL inhibitor induces specific chromatin changes and eradicates disease in models of MLL-rearranged leukemia. Cancer Cell (2019) 36(6):660–673.e11. doi: 10.1016/j.ccell.2019.11.001

PubMed Abstract | CrossRef Full Text | Google Scholar

90. Uckelmann HJ, Kim SM, Wong EM, Hatton C, Giovinazzo H, Gadrey JY, et al. Therapeutic targeting of preleukemia cells in a mouse model of NPM1 mutant acute myeloid leukemia. Science (2020) 367(6477):586–90. doi: 10.1126/science.aax5863

PubMed Abstract | CrossRef Full Text | Google Scholar

91. Fathi A, Wang E, Issa G, Altman J, Montesinos P, Botton SD, et al. P504 updated data for ziftomenib in patients with npm1-mutated relapsed or refractory acute myeloid leukemia. Hemasphere (2023) 7(Suppl). doi: 10.1097/01.HS9.0000968924.19161.da

CrossRef Full Text | Google Scholar

92. Fiskus W, Boettcher S, Daver N, Mill CP, Sasaki K, Birdwell CE, et al. Effective Menin inhibitor-based combinations against AML with MLL rearrangement or NPM1 mutation (NPM1c). Blood Cancer J (2022) 12(1):1–11. doi: 10.1038/s41408-021-00603-3

PubMed Abstract | CrossRef Full Text | Google Scholar

93. Rausch J, Dzama MM, Dolgikh N, Stiller HL, Bohl SR, Lahrmann C, et al. Menin inhibitor ziftomenib (KO-539) synergizes with drugs targeting chromatin regulation or apoptosis and sensitizes acute myeloid leukemia with MLL rearrangement or NPM1 mutation to venetoclax. Haematologica (2023) 27. doi: 10.3324/haematol.2022.282160

CrossRef Full Text | Google Scholar

94. Dzama MM, Steiner M, Rausch J, Sasca D, Schönfeld J, Kunz K, et al. Synergistic targeting of FLT3 mutations in AML via combined menin-MLL and FLT3 inhibition. Blood (2020) 136(21):2442–56. doi: 10.1182/blood.2020005037

PubMed Abstract | CrossRef Full Text | Google Scholar

95. Daver N, Basu S, Garcia-Manero G, Cortes JE, Ravandi F, Ning J, et al. Defining the immune checkpoint landscape in patients (pts) with acute myeloid leukemia (AML). Blood (2016) 128(22):2900. doi: 10.1182/blood.V128.22.2900.2900

CrossRef Full Text | Google Scholar

96. Toffalori C, Zito L, Gambacorta V, Riba M, Oliveira G, Bucci G, et al. Immune signature drives leukemia escape and relapse after hematopoietic cell transplantation. Nat Med (2019) 25(4):603–11. doi: 10.1038/s41591-019-0400-z

PubMed Abstract | CrossRef Full Text | Google Scholar

97. Zhou Q, Munger ME, Veenstra RG, Weigel BJ, Hirashima M, Munn DH, et al. Coexpression of Tim-3 and PD-1 identifies a CD8+ T-cell exhaustion phenotype in mice with disseminated acute myelogenous leukemia. Blood (2011) 117(17):4501–10. doi: 10.1182/blood-2010-10-310425

PubMed Abstract | CrossRef Full Text | Google Scholar

98. Wu Z, Ou J, Liu N, Wang Z, Chen J, Cai Z, et al. Upregulation of Tim-3 is associated with poor prognosis in acute myeloid leukemia. Cancer Med (2023) 12(7):8956–69. doi: 10.1002/cam4.5549

PubMed Abstract | CrossRef Full Text | Google Scholar

99. Wang M, Bu J, Zhou M, Sido J, Lin Y, Liu G, et al. CD8+T cells expressing both PD-1 and TIGIT but not CD226 are dysfunctional in acute myeloid leukemia (AML) patients. Clin Immunol (2018) 190:64–73. doi: 10.1016/j.clim.2017.08.021

PubMed Abstract | CrossRef Full Text | Google Scholar

100. Radwan SM, Elleboudy NS, Nabih NA, Kamal AM. The immune checkpoints Cytotoxic T lymphocyte antigen-4 and Lymphocyte activation gene-3 expression is up-regulated in acute myeloid leukemia. HLA (2020) 96(1):3–12. doi: 10.1111/tan.13872

PubMed Abstract | CrossRef Full Text | Google Scholar

101. Chen C, Liang C, Wang S, Chio CL, Zhang Y, Zeng C, et al. Expression patterns of immune checkpoints in acute myeloid leukemia. J Hematol Oncol (2020) 13(1):28. doi: 10.1186/s13045-020-00853-x

PubMed Abstract | CrossRef Full Text | Google Scholar

102. Williams P, Basu S, Garcia-Manero G, Hourigan CS, Oetjen KA, Cortes JE, et al. The distribution of T-cell subsets and the expression of immune checkpoint receptors and ligands in patients with newly diagnosed and relapsed acute myeloid leukemia. Cancer (2019) 125(9):1470–81. doi: 10.1002/cncr.31896

PubMed Abstract | CrossRef Full Text | Google Scholar

103. Berger R, Rotem-Yehudar R, Slama G, Landes S, Kneller A, Leiba M, et al. Phase I safety and pharmacokinetic study of CT-011, a humanized antibody interacting with PD-1, in patients with advanced hematologic Malignancies. Clin Cancer Res (2008) 14(10):3044–51. doi: 10.1158/1078-0432.CCR-07-4079

PubMed Abstract | CrossRef Full Text | Google Scholar

104. Gerds AT, Scott BL, Greenberg PL, Khaled SK, Lin TL, Pollyea DA, et al. PD-L1 blockade with atezolizumab in higher-risk myelodysplastic syndrome: an initial safety and efficacy analysis. Blood (2018) 132(Supplement 1):466. doi: 10.1182/blood-2018-99-118577

CrossRef Full Text | Google Scholar

105. Garcia-Manero G, Tallman MS, Martinelli G, Ribrag V, Yang H, Balakumaran A, et al. Pembrolizumab, a PD-1 inhibitor, in patients with myelodysplastic syndrome (MDS) after failure of hypomethylating agent treatment. Blood (2016) 128(22):345. doi: 10.1182/blood.V128.22.345.345

CrossRef Full Text | Google Scholar

106. Jaiswal S, Jamieson CHM, Pang WW, Park CY, Chao MP, Majeti R, et al. CD47 is upregulated on circulating hematopoietic stem cells and leukemia cells to avoid phagocytosis. Cell (2009) 138(2):271–85. doi: 10.1016/j.cell.2009.05.046

PubMed Abstract | CrossRef Full Text | Google Scholar

107. Wang C, Sun C, Li M, Xia B, Wang Y, Zhang L, et al. Novel fully human anti-CD47 antibodies stimulate phagocytosis and promote elimination of AML cells. J Cell Physiol (2021) 236(6):4470–81. doi: 10.1002/jcp.30163

PubMed Abstract | CrossRef Full Text | Google Scholar

108. Vyas P, Knapper S, Kelly R, Salim R, Lubowiecki M, Royston D, et al. Initial Phase 1 results of the first-in-class anti-CD47 antibody HU5F9- G4 in relapsed/refractory acute myeloid leukemia patients. EHA Library. (2018) 214718:PF232.

Google Scholar

109. Ravandi F, Walter RB, Subklewe M, Buecklein V, Jongen-Lavrencic M, Paschka P, et al. Updated results from phase I dose-escalation study of AMG 330, a bispecific T-cell engager molecule, in patients with relapsed/refractory acute myeloid leukemia (R/R AML). JCO (2020) 38(15_suppl):7508–8. doi: 10.1200/JCO.2020.38.15_suppl.7508

CrossRef Full Text | Google Scholar

110. Subklewe M, Stein A, Walter RB, Bhatia R, Wei AH, Ritchie D, et al. Preliminary results from a phase 1 first-in-human study of AMG 673, a novel half-life extended (HLE) anti-CD33/CD3 biTE® (Bispecific T-cell engager) in patients with relapsed/refractory (R/R) acute myeloid leukemia (AML). Blood (2019) 134:833. doi: 10.1182/blood-2019-127977

CrossRef Full Text | Google Scholar

111. Westervelt P, Roboz GJ, Cortes JE, Kantarjian HM, Lee S, Rettig MP, et al. Phase 1 first-in-human trial of AMV564, a bivalent bispecific (2x2) CD33/CD3 T-cell engager, in patients with relapsed/refractory acute myeloid leukemia (AML). Blood (2018) 132(Supplement 1):1455. doi: 10.1182/blood-2018-99-111529

PubMed Abstract | CrossRef Full Text | Google Scholar

112. Uy GL, Aldoss I, Foster MC, Sayre PH, Wieduwilt MJ, Advani AS, et al. Flotetuzumab as salvage immunotherapy for refractory acute myeloid leukemia. Blood (2021) 137(6):751–62. doi: 10.1182/blood.2020007732

PubMed Abstract | CrossRef Full Text | Google Scholar

113. Ravandi F, Bashey A, Foran JM, Stock W, Mawad R, Blum W, et al. Complete responses in relapsed/refractory acute myeloid leukemia (AML) patients on a weekly dosing schedule of xmAb14045, a CD123 x CD3 T cell-engaging bispecific antibody: initial results of a phase 1 study. Blood (2018) 132(Supplement 1):763. doi: 10.1182/blood-2018-99-119786

CrossRef Full Text | Google Scholar

114. Laborda E, Mazagova M, Shao S, Wang X, Quirino H, Woods AK, et al. Development of A chimeric antigen receptor targeting C-type lectin-like molecule-1 for human acute myeloid leukemia. Int J Mol Sci (2017) 18(11):2259. doi: 10.3390/ijms18112259

PubMed Abstract | CrossRef Full Text | Google Scholar

115. Steger B, Floro L, Amberger DC, Kroell T, Tischer J, Kolb HJ, et al. WT1, PRAME, and PR3 mRNA expression in acute myeloid leukemia (AML). J Immunother. (2020) 43(6):204–15. doi: 10.1097/CJI.0000000000000322

PubMed Abstract | CrossRef Full Text | Google Scholar

116. Tawara I, Kageyama S, Miyahara Y, Fujiwara H, Nishida T, Akatsuka Y, et al. Safety and persistence of WT1-specific T-cell receptor gene–transduced lymphocytes in patients with AML and MDS. Blood (2017) 130(18):1985–94. doi: 10.1182/blood-2017-06-791202

PubMed Abstract | CrossRef Full Text | Google Scholar

117. Rosenfeld C, Cheever MA, Gaiger A. WT1 in acute leukemia, chronic myelogenous leukemia and myelodysplastic syndrome: therapeutic potential of WT1 targeted therapies. Leukemia (2003) 17(7):1301–12. doi: 10.1038/sj.leu.2402988

PubMed Abstract | CrossRef Full Text | Google Scholar

118. Mascarenhas J, Cortes J, Huls G, Venditti A, Breems D, Botton SD, et al. Update from the ongoing Phase I multinational study of MCLA-117, a bispecific CLEC12AxCD3 T-cell engager in patients with acute myelogenous leukemia. Vol. 294456. EHA Library Mascarenhas J (2020).

Google Scholar

119. Kenderian SS, Ruella M, Shestova O, Klichinsky M, Aikawa V, Morrissette JJD, et al. CD33-specific chimeric antigen receptor T cells exhibit potent preclinical activity against human acute myeloid leukemia. Leukemia (2015) 29(8):1637–47. doi: 10.1038/leu.2015.52

PubMed Abstract | CrossRef Full Text | Google Scholar

120. Gill S, Tasian SK, Ruella M, Shestova O, Li Y, Porter DL, et al. Preclinical targeting of human acute myeloid leukemia and myeloablation using chimeric antigen receptor–modified T cells. Blood (2014) 123(15):2343–54. doi: 10.1182/blood-2013-09-529537

PubMed Abstract | CrossRef Full Text | Google Scholar

121. Richards RM, Zhao F, Freitas KA, Parker KR, Xu P, Fan A, et al. NOT-gated CD93 CAR T cells effectively target AML with minimized endothelial cross-reactivity. Blood Cancer Discovery (2021) 2(6):648–65. doi: 10.1158/2643-3230.BCD-20-0208

PubMed Abstract | CrossRef Full Text | Google Scholar

122. Perriello VM, Rotiroti MC, Pisani I, Galimberti S, Alberti G, Pianigiani G, et al. IL-3-zetakine combined with a CD33 costimulatory receptor as a dual CAR approach for safer and selective targeting of AML. Blood Adv (2023) 7(12):2855–71. doi: 10.1182/bloodadvances.2022008762

PubMed Abstract | CrossRef Full Text | Google Scholar

123. Liu F, Cao Y, Pinz K, Ma Y, Wada M, Chen K, et al. First-in-human CLL1-CD33 compound CAR T cell therapy induces complete remission in patients with refractory acute myeloid leukemia: update on phase 1 clinical trial. Blood (2018) 132(Supplement 1):901. doi: 10.1182/blood-2018-99-110579

CrossRef Full Text | Google Scholar

124. Naik S, Madden RM, Lipsitt A, Lockey T, Bran J, Rubnitz JE, et al. Safety and anti-leukemic activity of CD123-CAR T cells in pediatric patients with AML: preliminary results from a phase 1 trial. Blood (2022) 140(Supplement 1):4584–5. doi: 10.1182/blood-2022-170201

CrossRef Full Text | Google Scholar

125. Tambaro FP, Singh H, Jones E, Rytting M, Mahadeo KM, Thompson P, et al. Autologous CD33-CAR-T cells for treatment of relapsed/refractory acute myelogenous leukemia. Leukemia (2021) 35(11):3282–6. doi: 10.1038/s41375-021-01232-2

PubMed Abstract | CrossRef Full Text | Google Scholar

126. Cheng J, Ge T, Zhu X, Wang J, Zeng Y, Mu W, et al. Preclinical development and evaluation of nanobody-based CD70-specific CAR T cells for the treatment of acute myeloid leukemia. Cancer Immunol Immunother. (2023) 72(7):2331–46. doi: 10.1007/s00262-023-03422-6

PubMed Abstract | CrossRef Full Text | Google Scholar

127. Sauer T, Parikh K, Sharma S, Omer B, Sedloev D, Chen Q, et al. CD70-specific CAR T cells have potent activity against acute myeloid leukemia without HSC toxicity. Blood (2021) 138(4):318–30. doi: 10.1182/blood.2020008221

PubMed Abstract | CrossRef Full Text | Google Scholar

128. Ljunggren HG, Kärre K. In search of the “missing self”: MHC molecules and NK cell recognition. Immunol Today (1990) 11(7):237–44. doi: 10.1016/0167-5699(90)90097-S

PubMed Abstract | CrossRef Full Text | Google Scholar

129. Laskowski TJ, Biederstädt A, Rezvani K. Natural killer cells in antitumour adoptive cell immunotherapy. Nat Rev Cancer. (2022) 22(10):557–75. doi: 10.1038/s41568-022-00491-0

PubMed Abstract | CrossRef Full Text | Google Scholar

130. Liu E, Marin D, Banerjee P, Macapinlac HA, Thompson P, Basar R, et al. Use of CAR-transduced natural killer cells in CD19-positive lymphoid tumors. N Engl J Med (2020) 382(6):545–53. doi: 10.1056/NEJMoa1910607

PubMed Abstract | CrossRef Full Text | Google Scholar

131. Albinger N, Pfeifer R, Nitsche M, Mertlitz S, Campe J, Stein K, et al. Primary CD33-targeting CAR-NK cells for the treatment of acute myeloid leukemia. Blood Cancer J (2022) 12(4):61. doi: 10.1055/s-0042-1748710

PubMed Abstract | CrossRef Full Text | Google Scholar

132. Schmid C, Labopin M, Schaap N, Veelken H, Schleuning M, Stadler M, et al. Prophylactic donor lymphocyte infusion after allogeneic stem cell transplantation in acute leukaemia - a matched pair analysis by the Acute Leukaemia Working Party of EBMT. Br J Haematol (2019) 184(5):782–7. doi: 10.1111/bjh.15691

PubMed Abstract | CrossRef Full Text | Google Scholar

133. Horowitz MM, Gale RP, Sondel PM, Goldman JM, Kersey J, Kolb HJ, et al. Graft-versus-leukemia reactions after bone marrow transplantation. Blood (1990) 75(3):555–62. doi: 10.1182/blood.V75.3.555.555

PubMed Abstract | CrossRef Full Text | Google Scholar

134. Mathé G, Amiel JL, Schwarzenberg L, Cattan A, Schneider M. Adoptive immunotherapy of acute leukemia: experimental and clinical results. Cancer Res (1965) 25(9):1525–31.

PubMed Abstract | Google Scholar

135. Vago L, Perna SK, Zanussi M, Mazzi B, Barlassina C, Stanghellini MTL, et al. Loss of mismatched HLA in leukemia after stem-cell transplantation. New Engl J Med (2009) 361(5):478–88. doi: 10.1056/NEJMoa0811036

PubMed Abstract | CrossRef Full Text | Google Scholar

136. Christopher MJ, Petti AA, Rettig MP, Miller CA, Chendamarai E, Duncavage EJ, et al. Immune escape of relapsed AML cells after allogeneic transplantation. New Engl J Med (2018) 379(24):2330–41. doi: 10.1056/NEJMoa1808777

PubMed Abstract | CrossRef Full Text | Google Scholar

137. Jan M, Leventhal MJ, Morgan EA, Wengrod JC, Nag A, Drinan SD, et al. Recurrent genetic HLA loss in AML relapsed after matched unrelated allogeneic hematopoietic cell transplantation. Blood Adv (2019) 3(14):2199–204. doi: 10.1182/bloodadvances.2019000445

PubMed Abstract | CrossRef Full Text | Google Scholar

138. Steinmann J, Bertz H, Wäsch R, Marks R, Zeiser R, Bogatyreva L, et al. 5-Azacytidine and DLI can induce long-term remissions in AML patients relapsed after allograft. Bone Marrow Transplant. (2015) 50(5):690–5. doi: 10.1038/bmt.2015.10

PubMed Abstract | CrossRef Full Text | Google Scholar

139. Schroeder C, Gani C, Lamprecht U, von Weyhern CH, Weinmann M, Bamberg M, et al. Pathological complete response and sphincter-sparing surgery after neoadjuvant radiochemotherapy with regional hyperthermia for locally advanced rectal cancer compared with radiochemotherapy alone. Int J Hyperthermia. (2012) 28(8):707–14. doi: 10.3109/02656736.2012.722263

PubMed Abstract | CrossRef Full Text | Google Scholar

140. Craddock C, Labopin M, Robin M, Finke J, Chevallier P, Yakoub-Agha I, et al. Clinical activity of azacitidine in patients who relapse after allogeneic stem cell transplantation for acute myeloid leukemia. Haematologica (2016) 101(7):879–83. doi: 10.3324/haematol.2015.140996

PubMed Abstract | CrossRef Full Text | Google Scholar

141. Kübler A, Woiterski J, Witte KE, Bühring HJ, Hartwig UF, Ebinger M, et al. Both mature KIR+ and immature KIR– NK cells control pediatric acute B-cell precursor leukemia in NOD.Cg-Prkdcscid IL2rgtmWjl/Sz mice. Blood (2014) 124(26):3914–23. doi: 10.1182/blood-2014-05-572743

PubMed Abstract | CrossRef Full Text | Google Scholar

142. Chiappinelli KB, Strissel PL, Desrichard A, Li H, Henke C, Akman B, et al. Inhibiting DNA Methylation Causes an Interferon Response in Cancer via dsRNA Including Endogenous Retroviruses. Cell (2015) 162(5):974–86. doi: 10.1016/j.cell.2015.07.011

PubMed Abstract | CrossRef Full Text | Google Scholar

143. Roulois D, Loo Yau H, Singhania R, Wang Y, Danesh A, Shen SY, et al. DNA-demethylating agents target colorectal cancer cells by inducing viral mimicry by endogenous transcripts. Cell (2015) 162(5):961–73. doi: 10.1016/j.cell.2015.07.056

PubMed Abstract | CrossRef Full Text | Google Scholar

144. Sohlberg E, Pfefferle A, Andersson S, Baumann BC, Hellström-Lindberg E, Malmberg KJ. Imprint of 5-azacytidine on the natural killer cell repertoire during systemic treatment for high-risk myelodysplastic syndrome. Oncotarget (2015) 6(33):34178–90. doi: 10.18632/oncotarget.6213

PubMed Abstract | CrossRef Full Text | Google Scholar

145. Peng D, Kryczek I, Nagarsheth N, Zhao L, Wei S, Wang W, et al. Epigenetic silencing of TH1-type chemokines shapes tumour immunity and immunotherapy. Nature (2015) 527(7577):249–53. doi: 10.1038/nature15520

PubMed Abstract | CrossRef Full Text | Google Scholar

146. Costantini B, Kordasti SY, Kulasekararaj AG, Jiang J, Seidl T, Abellan PP, et al. The effects of 5-azacytidine on the function and number of regulatory T cells and T-effectors in myelodysplastic syndrome. Haematologica (2013) 98(8):1196–205. doi: 10.3324/haematol.2012.074823

PubMed Abstract | CrossRef Full Text | Google Scholar

147. Ørskov AD, Treppendahl MB, Skovbo A, Holm MS, Friis LS, Hokland M, et al. Hypomethylation and up-regulation of PD-1 in T cells by azacytidine in MDS/AML patients: A rationale for combined targeting of PD-1 and DNA methylation. Oncotarget (2015) 6(11):9612–26. doi: 10.18632/oncotarget.3324

PubMed Abstract | CrossRef Full Text | Google Scholar

148. Yang H, Bueso-Ramos C, DiNardo C, Estecio MR, Davanlou M, Geng QR, et al. Expression of PD-L1, PD-L2, PD-1 and CTLA4 in myelodysplastic syndromes is enhanced by treatment with hypomethylating agents. Leukemia (2014) 28(6):1280–8. doi: 10.1038/leu.2013.355

PubMed Abstract | CrossRef Full Text | Google Scholar

149. Daver N, Garcia-Manero G, Basu S, Boddu PC, Alfayez M, Cortes JE, et al. Efficacy, safety, and biomarkers of response to azacitidine and nivolumab in relapsed/refractory acute myeloid leukemia: A nonrandomized, open-label, phase II study. Cancer Discovery. (2019) 9(3):370–83. doi: 10.1158/2159-8290.CD-18-0774

PubMed Abstract | CrossRef Full Text | Google Scholar

150. Gojo I, Stuart RK, Webster J, Blackford A, Varela JC, Morrow J, et al. Multi-center phase 2 study of pembroluzimab (Pembro) and azacitidine (AZA) in patients with relapsed/refractory acute myeloid leukemia (AML) and in newly diagnosed (≥65 years) AML patients. Blood (2019) 134(Supplement_1):832. doi: 10.1182/blood-2019-127345

CrossRef Full Text | Google Scholar

151. Brunner AM, Esteve J, Porkka K, Knapper S, Traer E, Scholl S, et al. Efficacy and safety of sabatolimab (MBG453) in combination with hypomethylating agents (HMAs) in patients (Pts) with very high/high-risk myelodysplastic syndrome (vHR/HR-MDS) and acute myeloid leukemia (AML): final analysis from a phase ib study. Blood (2021) 138(Supplement 1):244. doi: 10.1182/blood-2021-146039

CrossRef Full Text | Google Scholar

152. Zeidan AM, Cavenagh J, Voso MT, Taussig D, Tormo M, Boss I, et al. Efficacy and safety of azacitidine (AZA) in combination with the anti-PD-L1 durvalumab (durva) for the front-line treatment of older patients (pts) with acute myeloid leukemia (AML) who are unfit for intensive chemotherapy (IC) and pts with higher-risk myelodysplastic syndromes (HR-MDS): results from a large, international, randomized phase 2 study. Blood (2019) 134(Supplement_1):829. doi: 10.1182/blood-2019-122896

CrossRef Full Text | Google Scholar

153. Gómez-Llobell M, Raíndo AP, Medina JC, Centurión IG, Orgueira AM. Immune checkpoint inhibitors in acute myeloid leukemia: A meta-analysis. Front Oncol (2022) 12:882531. doi: 10.3389/fonc.2022.882531

PubMed Abstract | CrossRef Full Text | Google Scholar

154. Gros FX, Cazaubiel T, Forcade E, Lechevalier N, Leguay T, Servant V, et al. Severe acute GvHD following administration of ipilimumab for early relapse of AML after haploidentical stem cell transplantation. Bone Marrow Transplant. (2017) 52(7):1047–8. doi: 10.1038/bmt.2017.78

PubMed Abstract | CrossRef Full Text | Google Scholar

155. Saxena K, Herbrich SM, Pemmaraju N, Kadia TM, DiNardo CD, Borthakur G, et al. A phase 1b/2 study of azacitidine with PD-L1 antibody avelumab in relapsed/refractory acute myeloid leukemia. Cancer (2021) 127(20):3761–71. doi: 10.1002/cncr.33690

PubMed Abstract | CrossRef Full Text | Google Scholar

156. Woods DM, Sodré AL, Villagra A, Sarnaik A, Sotomayor EM, Weber J. HDAC inhibition upregulates PD-1 ligands in melanoma and augments immunotherapy with PD-1 blockade. Cancer Immunol Res (2015) 3(12):1375–85. doi: 10.1158/2326-6066.CIR-15-0077-T

PubMed Abstract | CrossRef Full Text | Google Scholar

157. Wischnewski F, Pantel K, Schwarzenbach H. Promoter demethylation and histone acetylation mediate gene expression of MAGE-A1, -A2, -A3 , and -A12 in human cancer cells. Mol Cancer Res (2006) 4(5):339–49. doi: 10.1158/1541-7786.MCR-05-0229

PubMed Abstract | CrossRef Full Text | Google Scholar

158. Magner WJ, Kazim AL, Stewart C, Romano MA, Catalano G, Grande C, et al. Activation of MHC class I, II, and CD40 gene expression by histone deacetylase inhibitors. J Immunol (2000) 165(12):7017–24. doi: 10.4049/jimmunol.165.12.7017

PubMed Abstract | CrossRef Full Text | Google Scholar

159. Zheng H, Zhao W, Yan C, Watson CC, Massengill M, Xie M, et al. HDAC inhibitors enhance T-cell chemokine expression and augment response to PD-1 immunotherapy in lung adenocarcinoma. Clin Cancer Res (2016) 22(16):4119–32. doi: 10.1158/1078-0432.CCR-15-2584

PubMed Abstract | CrossRef Full Text | Google Scholar

160. Borcoman E, Kamal M, Marret G, Dupain C, Castel-Ajgal Z, Le Tourneau C. HDAC inhibition to prime immune checkpoint inhibitors. Cancers (Basel). (2021) 14(1):66. doi: 10.3390/cancers14010066

PubMed Abstract | CrossRef Full Text | Google Scholar

161. Bewersdorf JP, Shallis RM, Sharon E, Caldwell A, Wei W, Yacoub A, et al. A multi-center phase ib trial of the histone deactylase inhibitor (HDACi) entinostat in combination with anti-PD1 antibody pembrolizumab in patients with refractory/relapsed myelodysplastic syndromes (RR-MDS) or oligoblastic acute myeloid leukemia (RR-AML) after hypomethylating agent (HMA) failure. Blood (2022) 140(Supplement 1):9084–6. doi: 10.1182/blood-2022-158626

CrossRef Full Text | Google Scholar

162. Daver N, Senapati J, Maiti A, Loghavi S, Kadia TM, DiNardo CD, et al. Phase I/II study of azacitidine (AZA) with venetoclax (VEN) and magrolimab (Magro) in patients (pts) with newly diagnosed (ND) older/unfit or high-risk acute myeloid leukemia (AML) and relapsed/refractory (R/R) AML. Blood (2022) 140(Supplement 1):141–4. doi: 10.1182/blood-2022-170188

CrossRef Full Text | Google Scholar

163. Laszlo GS, Gudgeon CJ, Harrington KH, Dell’Aringa J, Newhall KJ, Means GD, et al. Cellular determinants for preclinical activity of a novel CD33/CD3 bispecific T-cell engager (BiTE) antibody, AMG 330, against human AML. Blood (2014) 123(4):554–61. doi: 10.1182/blood-2013-09-527044

PubMed Abstract | CrossRef Full Text | Google Scholar

164. El Khawanky N, Hughes A, Yu W, Myburgh R, Matschulla T, Taromi S, et al. Demethylating therapy increases anti-CD123 CAR T cell cytotoxicity against acute myeloid leukemia. Nat Commun (2021) 12(1):6436. doi: 10.1038/s41467-021-26683-0

PubMed Abstract | CrossRef Full Text | Google Scholar

165. Riether C, Pabst T, Höpner S, Bacher U, Hinterbrandner M, Banz Y, et al. Targeting CD70 with cusatuzumab eliminates acute myeloid leukemia stem cells in patients treated with hypomethylating agents. Nat Med (2020) 26(9):1459–67. doi: 10.1038/s41591-020-0910-8

PubMed Abstract | CrossRef Full Text | Google Scholar

166. Tang L, Kong Y, Wang H, Zou P, Sun T, Liu Y, et al. Demethylating therapy increases cytotoxicity of CD44v6 CAR-T cells against acute myeloid leukemia. Front Immunol (2023) 14:1145441. doi: 10.3389/fimmu.2023.1145441

PubMed Abstract | CrossRef Full Text | Google Scholar

167. Baragaño Raneros A, Martín-Palanco V, Fernandez AF, Rodriguez RM, Fraga MF, Lopez-Larrea C, et al. Methylation of NKG2D ligands contributes to immune system evasion in acute myeloid leukemia. Genes Immun (2015) 16(1):71–82. doi: 10.1038/gene.2014.58

PubMed Abstract | CrossRef Full Text | Google Scholar

168. Paczulla AM, Rothfelder K, Raffel S, Konantz M, Steinbacher J, Wang H, et al. Absence of NKG2D ligands defines leukaemia stem cells and mediates their immune evasion. Nature (2019) 572(7768):254–9. doi: 10.1038/s41586-019-1410-1

PubMed Abstract | CrossRef Full Text | Google Scholar

169. Rohner A, Langenkamp U, Siegler U, Kalberer CP, Wodnar-Filipowicz A. Differentiation-promoting drugs up-regulate NKG2D ligand expression and enhance the susceptibility of acute myeloid leukemia cells to natural killer cell-mediated lysis. Leuk Res (2007) 31(10):1393–402. doi: 10.1016/j.leukres.2007.02.020

PubMed Abstract | CrossRef Full Text | Google Scholar

170. Vasu S, He S, Cheney C, Gopalakrishnan B, Mani R, Lozanski G, et al. Decitabine enhances anti-CD33 monoclonal antibody BI 836858–mediated natural killer ADCC against AML blasts. Blood (2016) 127(23):2879–89. doi: 10.1182/blood-2015-11-680546

PubMed Abstract | CrossRef Full Text | Google Scholar

171. Montesinos P, Roboz GJ, Bulabois CE, Subklewe M, Platzbecker U, Ofran Y, et al. Safety and efficacy of talacotuzumab plus decitabine or decitabine alone in patients with acute myeloid leukemia not eligible for chemotherapy: results from a multicenter, randomized, phase 2/3 study. Leukemia (2021) 35(1):62–74. doi: 10.1038/s41375-020-0773-5

PubMed Abstract | CrossRef Full Text | Google Scholar

172. Jetani H, Garcia-Cadenas I, Nerreter T, Thomas S, Rydzek J, Meijide JB, et al. CAR T-cells targeting FLT3 have potent activity against FLT3-ITD+ AML and act synergistically with the FLT3-inhibitor crenolanib. Leukemia (2018) 32(5):1168–79. doi: 10.1038/s41375-018-0009-0

PubMed Abstract | CrossRef Full Text | Google Scholar

173. Zhong M, Gao R, Zhao R, Huang Y, Chen C, Li K, et al. BET bromodomain inhibition rescues PD-1-mediated T-cell exhaustion in acute myeloid leukemia. Cell Death Dis (2022) 13(8):671. doi: 10.1038/s41419-022-05123-x

PubMed Abstract | CrossRef Full Text | Google Scholar

174. Kagoya Y, Nakatsugawa M, Yamashita Y, Ochi T, Guo T, Anczurowski M, et al. BET bromodomain inhibition enhances T cell persistence and function in adoptive immunotherapy models. J Clin Invest. (2016) 126(9):3479–94. doi: 10.1172/JCI86437

PubMed Abstract | CrossRef Full Text | Google Scholar

175. Sheng W, LaFleur MW, Nguyen TH, Chen S, Chakravarthy A, Conway JR, et al. LSD1 ablation stimulates anti-tumor immunity and enables checkpoint blockade. Cell (2018) 174(3):549–563.e19. doi: 10.1016/j.cell.2018.05.052

PubMed Abstract | CrossRef Full Text | Google Scholar

176. Daver N, Boddu P, Garcia-Manero G, Yadav SS, Sharma P, Allison J, et al. Hypomethylating agents in combination with immune checkpoint inhibitors in acute myeloid leukemia and myelodysplastic syndromes. Leukemia (2018) 32(5):1094–105. doi: 10.1038/s41375-018-0070-8

PubMed Abstract | CrossRef Full Text | Google Scholar

177. Daver N, Vyas P, Chao M, Xing G, Renard C, Ramsingh G, et al. A phase 3, randomized, open-label study evaluating the safety and efficacy of magrolimab in combination with azacitidine in previously untreated patients with TP53-mutant acute myeloid leukemia. Blood (2021) 138:3426. doi: 10.1182/blood-2021-145208

CrossRef Full Text | Google Scholar

Keywords: acute myeloid leukemia, hypomethylating agents, immunotherapy, epigenetics, checkpoint inhibition, cellular therapy, chromatin modifiers, combination therapy

Citation: Rausch J, Ullrich E and Kühn MWM (2023) Epigenetic targeting to enhance acute myeloid leukemia-directed immunotherapy. Front. Immunol. 14:1269012. doi: 10.3389/fimmu.2023.1269012

Received: 28 July 2023; Accepted: 05 September 2023;
Published: 22 September 2023.

Edited by:

Matthieu Perreau, Centre Hospitalier Universitaire Vaudois (CHUV), Switzerland

Reviewed by:

Vincenzo Maria Perriello, University of Perugia, Italy

Copyright © 2023 Rausch, Ullrich and Kühn. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Michael W.M. Kühn, mickuehn@uni-mainz.de

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