Next Article in Journal
Targeting STAT3 and NF-κB Signaling Pathways in Cancer Prevention and Treatment: The Role of Chalcones
Previous Article in Journal
Evaluating Outcome Prediction via Baseline, End-of-Treatment, and Delta Radiomics on PET-CT Images of Primary Mediastinal Large B-Cell Lymphoma
Previous Article in Special Issue
Mechanisms of Resistance to Small Molecules in Acute Myeloid Leukemia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Venetoclax Resistance in Acute Myeloid Leukemia

by
Sylvain Garciaz
*,
Marie-Anne Hospital
,
Yves Collette
and
Norbert Vey
Aix-Marseille University, Inserm, CNRS, Institut Paoli-Calmettes, CRCM, 13009 Marseille, France
*
Author to whom correspondence should be addressed.
Cancers 2024, 16(6), 1091; https://doi.org/10.3390/cancers16061091
Submission received: 28 January 2024 / Revised: 27 February 2024 / Accepted: 3 March 2024 / Published: 8 March 2024
(This article belongs to the Special Issue Resistance to Targeted Therapies in Hematological Malignancies)

Abstract

:

Simple Summary

Venetoclax–azacitidine is a new standard for elderly or unfit acute myeloid leukemia patients. Nevertheless, resistance remains a matter of concern. The main genetic alterations associated with venetoclax sensitivity are IDH mutations, whereas TP53, signaling mutations, and BAX mutations are associated with venetoclax resistance. Non-genetic resistance mechanisms have also been described, including changes in apoptotic proteins, differentiation states, metabolic status, and mitochondrial machinery. Venetoclax-based triplet therapies including IDH and FLT3 inhibitors or innovative therapies are under investigation to target resistances.

Abstract

Venetoclax is a BH3-mimetics agent interacting with the anti-apoptotic protein BCL2, facilitating cytochrome c release from mitochondria, subsequent caspases activation, and cell death. Venetoclax combined with azacitidine (VEN-AZA) has become a new standard treatment for AML patients unfit for intensive chemotherapy. In the phase III VIALE-A study, VEN-AZA showed a 65% overall response rate and 14.7 months overall survival in comparison with 22% and 8 months in the azacitidine monotherapy control arm. Despite these promising results, relapses and primary resistance to venetoclax are frequent and remain an unmet clinical need. Clinical and preclinical studies have been conducted to identify factors driving resistance. Among them, the most documented are molecular alterations including IDH, FLT3, TP53, and the newly described BAX mutations. Several non-genetic factors are also described such as metabolic plasticity, changes in anti-apoptotic protein expression, and dependencies, as well as monocytic differentiation status. Strategies to overcome venetoclax resistance are being developed in clinical trials, including triplet therapies with targeted agents targeting IDH, FLT3, as well as the recently developed menin inhibitors or immunotherapies such as antibody–drug conjugated or monoclonal antibodies. A better understanding of the molecular factors driving venetoclax resistance by single-cell analyses will help the discovery of new therapeutic strategies in the future.

1. Introduction

The combination of the BCL2 inhibitor venetoclax (VEN) and azacitidine (AZA) has changed the paradigm of treatment for elderly or unfit acute myeloid leukemia (AML) patients. Indeed, in this hard-to-treat population, the use of hypomethylating agents (HMAs) such as decitabine or AZA was associated with a 20–30% complete response rate and 10 months survival [1,2]. In comparison, response rates with VEN-AZA are approximately 65%, and overall survival rate was 14.7 months in the pivotal VIALE-A study [3]. The long-term analyses of patients treated in this study have been recently updated with a 2-year OS of 37.5%, representing a gain of 20.6% compared to AZA monotherapy [4]. Nonetheless, primary resistance to VEN-AZA represents approximately one-third of treated AML patients. Moreover, the median duration of remission for responding patients is approximately 18 months, and virtually all the patients will finally relapse, as shown in the absence of a “plateau” on the survival curves, indicating that the regimen is not a curative therapy.
VEN-AZA relapsed or refractory AML (R/R-AML) patients have very poor outcomes, with a response rate and OS of about 20% and 2.4 months, respectively [5,6]. Therefore, a better understanding of VEN resistance is crucial to developing new therapeutic strategies and improving the survival of VEN-AZA R/R AML [6]. Recent reviews have pinpointed the role of genetic and non-genetic factors in the apparition of resistance [7,8,9,10,11,12,13,14]. In this review, we will describe the main mechanisms of resistance and the therapeutic strategies to overcome them. Finally, we will describe the most promising triplet therapies using VEN-AZA as a backbone for the treatment of unfit/elderly AML patients.

2. Rationale of BCL2 Inhibition

2.1. Mechanism of Intrinsic Apoptosis and Developmlent of BH3 Mimetics

Intrinsic apoptosis relies on the balance between pro- and anti-apoptotic proteins inducing mitochondrial outer membrane permeabilization (MOMP), ultimately leading to caspase-dependent cell death. MOMP is regulated by the members of the BCL2 protein family containing a single BH3 domain named “BH3-only proteins”, which have a pro-apoptotic role. Conversely, MOMP is blocked by a series of proteins that have an anti-apoptotic action, including BCL2, BCL-XL, and MCL1. When the balance between pro- and anti-apoptotic signals turns toward cell death due to internal or external stressors, the BAX and BAK proteins are activated and form pores in the mitochondrial outer membrane, subsequently releasing cytochrome c and inducing caspase activation [12].
Efforts have been made to develop BH3-mimetic agents that induce intrinsic apoptosis by blocking anti-apoptotic proteins. The first synthesized BH3-mimetics, ABT-737 and ABT-263, bound to both BCL2 and BCL-XL. Although ABT-263 (navitoclax) showed promising results in phase I studies, the dose limiting toxicity of the treatment was thrombocytopenia, as BCL-XL is essential for mature platelet survival [15]. ABT-199/VEN is a potent, selective, and orally available inhibitor of BCL2 [16], and its mechanism of action is shown in Figure 1. Preclinical studies have shown that VEN induced rapid apoptosis in a variety of AML cell line models at nanomolar concentrations and inhibited in vivo leukemia progression [17].

2.2. Main Clinical Studies Using Venetoclax in AML

VEN efficacy was assessed in a phase II single-agent study in R/R AML patients, showing a 19% overall response rate (ORR) and a median duration of remission of 48 days. Phase I/II studies combining VEN plus HMA or low-dose ARAC also showed a high response close to 65% and a favorable tolerance profile [18,19,20,21,22]. These results led to the registration randomized phase III trial “VIALE-A”, in which 431 patients were enrolled. Among them, 286 were allocated to the VEN-AZA group and 145 to the AZA group. The incidence of CR was higher with VEN-AZA than with the control regimen (36.7% vs. 17.9%; p < 0.001), as was the ORR = 66.4% vs. 28.3%; p < 0.001). After a median follow-up of 20.5 months, the median OS was 14.7 months in the VEN-AZA group and 9.6 months in the control group. Long-term follow-up confirmed the better OS in the VEN-AZA group [3,4]. The VEN-AZA regimen was approved for unfit or older AML patients in France and in most European countries in 2020. Since then, real-life experiences have confirmed the high response rate of VEN-AZA treatment in the context of newly diagnosed AML as well as refractory AML [23,24,25,26]. The main clinical studies published on VEN-HMA are summarized in Table 1.

3. Molecular Factors Modulating Venetoclax Efficacy

Current biological factors that are commonly used for AML in younger patients in the context of intensive chemotherapy (ICT) are questionable in the context of VEN-AZA treatment, as seen for instance in European Leukemia Network (ELN) 2022 cytogenetics and molecular classification [27]. Therefore, new molecular classifications are needed. In this section, we will review studies that have helped better patients’ stratification in the context of VEN-based regimens. The influence of the main AML gene mutations on VEN resistance are summarized in Figure 2.

3.1. IDH Mutations

Isocitrate dehydrogenase 1 and 2 (IDH1/2) are enzymes that perform key roles in various cellular functions, including the regulation of carbohydrate metabolism, epigenetics, and cell differentiation. Wild-type IDH1 and IDH2 oxidize and decarboxylate isocitrate to α-ketoglutarate (α-KG) in the cytoplasm and mitochondria, respectively, and simultaneously reduce NADP+ to NADPH. IDH1/2 gene mutations (IDH1/2mut) are almost always heterozygous and occur in enzymatically active sites. Hotspot mutations in IDH1R132, IDH2R140, and IDH2R172, enable a neomorphic reaction that converts α kg into the oncometabolite D-2-hydroxyglutarate (D-2-HG), which competitively inhibits α-KG-dependent dioxygenases and induces an AML blast differentiation blockade [28].
The favorable impact of IDH1/2 mutations in VEN-AZA-treated AML was already noticed in the phase II efficacy trial. Further analyses pooling results from VIALE-A and phase II studies showed 79% CR/CRi rates in IDH1/2mut patients. Interestingly, the response rates in the IDH2mut subgroup were higher (89%) than in the IDH1mut group (66.7%). IDH1/2mut were also associated with a longer median OS (24 months for IDH1/2 patients vs. 14 months in general). Again, OS was better in the IDH2mut subgroup compared to the IDH1mut [29]. The median OS was improved in the group of patients with negative minimal residual disease (MRD) assessed by flow cytometry (42 months) [30]. Mechanistically, the effect of VEN on IDH1/2mut preclinical models has been related to the inhibition of mitochondrial cytochrome c oxidase by D-2-HG, causing a reduction in the mitochondrial threshold for apoptosis induction [31] as well as a decrease in mitochondrial respiration [11,32].

3.2. TP53 Mutation

In a recent actualization of the International Consensus Classification (ICC) 2022 AML and the World Health Organization (WHO) 2022 classification, TP53 mutant (TP53mut) AML and myelodysplastic syndromes (MDS) have been individualized based on their poor prognosis [33,34]. Mutations in the TP53 genes are presents in 10–15% of AML cases and are enriched in poor-risk AML because of frequent additional cytogenetics or molecular alterations [35]. Most TP53mut are missense mutations that involve the DNA-binding domain. Mutant premature termination codons and frameshift mutations as well as most single amino-acid substitutions or deletions result in a strong disruption of TP53 function.
TP53mut diseases are known to poorly respond to ICT, as reported in recent large cohort studies. The overall survival rates of 6–9 months are among the worst within AML subsets [36]. The results seen with VEN-AZA have also been deceiving. In a pooled analysis of phase II and III studies, the response rate was 70% in the absence of TP53mut vs. 41% in the presence of TP53mut, and median OS was 23.4 months in the absence of TP53mut vs. 5.2 months in the presence of TP53mut. Patients treated with VEN-AZA were not doing better than those treated with AZA monotherapy. These data were confirmed in a real-life study from the MD Anderson Cancer Center showing a 6–9-month OS [37,38] and also in a study with VEN and decitabine, in which patients with TP53mut had a 5.2-month OS [39]. Taken together, these data question the rationale of using VEN in this setting, in particular for fragile or elderly patients [40].
From a mechanistic perspective, loss of function mutations in the TP53 gene decreased the expressions of pro-apoptotic genes such as NOXA or PUMA, which are known to be a transcriptional target of TP53 [41], as well as induced a less intense functional activation of the pro-apoptotic BAX and BAK proteins [42]. Consistently, studies using Crispr-Cas9 screens identified BAK, BAX, PUMA, and NOXA, as well as TP53 target genes as crucial regulators of VEN activity in vitro [41,42,43,44,45,46]. A recent study showed that interactions between TP53 and increased MYC transcriptional program led to VEN resistance that could be targeted [47].

3.3. Signaling Mutations (Including FLT3 Mutations)

Fms-like tyrosine kinase 3 (FLT3) is mutated in approximately 30% of adult AML cases, including internal tandem duplication (FLT3-ITD) and tyrosine kinase domain (TKD) amino acid substitution (FLT3-TKD), leading to constitutive tyrosine kinase activity and constitutive activation of the downstream proliferative signaling cascade [48]. The negative impact of FLT3-ITD on AML patients’ survival when treated with ICT is well established, as are other signaling mutations such as RAS mutations [49].
The impact of signaling mutations is unclear in the context of VEN-AZA treatment [14,20,21,50,51]. The response rates and OS were similar in FLT3-ITD versus FLT3wt cohorts in a pooled analyses of 353 patients from the VIALE-A and phase II studies, including 42 with FLT3 mutations [50]. Nevertheless, refined prognostic risk signatures based on N/KRAS, FLT3-ITD, and TP53 mutations have shown that patients with signaling mutations have at least an intermediate risk [27,52,53]. Multiple small-molecule tyrosine kinase inhibitors that target FLT3 are in development in combination with VEN for the treatment of patients with elderly/unfit AML patients [48]. Some of these molecules represent interesting candidates to test in triplet therapies based on the VEN-AZA backbone.

3.4. Secondary Type Mutations

The WHO 2022 and ICC 2022 classifications include a new AML subtype classifying AML with the presence of any of eight gene mutations, including ASXL1, BCOR, EZH2, STAG2, SF3B1, SRSF2, ZRSR2, and U2AF1 (as well as RUNX1 in the ICC 2022 classification only), as AML with myelodysplasia-related gene mutations (MR genes) [33,34,54]. MR gene mutations are associated with a poor response to ICT. As a result, the ELN 2022 classifies these AML cases in the poor-risk group [55]. Nevertheless, the impact of these gene mutations are less clearly defined in the context of VEN-AZA treatment [53,56]. For example, ASXL1 mutations confer VEN sensitivity in preclinical models but not in clinical studies [57,58]. Mutations in splicing factors, including SRSF2, U2AF1, SF3B1, and ZRSR2, have been implicated in the pathogenesis of MDS and AML [59,60]. These mutations are encountered in approximately 50% of secondary AML cases and usually correlate with inferior outcomes to ICT [49,61,62]. In the context of VEN-treatment, the negative impact of these genes on outcomes is less clear, as some studies report the absence of a negative impact [63,64], while others report that SRSF2 mutation was associated with a poorer response in preclinical [65] and clinical studies [53,66].

3.5. Mutations in Apoptotic Genes

Mutations in the BCL2 gene, in particular the BCL-2-binding groove variants G101V or D103Y, were shown to confer resistance to chronic lymphocytic leukemia (CLL) patients treated with VEN [67]. Contrary to CLL patients, BCL-2 variants were not found in VEN R/R AML cohorts. Nonetheless, BAX deficiency was identified in in vitro studies based on genome-wide Crispr screens exploring VEN resistance [44]. BAX mutations were identified in normal hematopoietic stem and progenitor cells of CLL patients treated with VEN [68]. Consistently, mutations in the BAX gene were found in 7 of 41 patients (17%) at relapse after VEN-AZA treatment, with no cases identified among patients with primary refractory disease [69]. The BAX variants that emerged at relapse included frameshift, nonsense, or donor splice-site abnormalities upstream of the BAX COOH-terminal domain or just before the BAX COOH-terminal domain, disrupting the COOH-terminal α9 helix necessary for BAX translocation from the cytosol into the mitochondrial outer membrane. Further studies will help us to understand the differential mechanisms of resistance between CLL and AML patients treated with VEN.

4. Non-Genetic Factors Driving Venetoclax Efficacy

4.1. Regulators of Intrinsic Apoptosis

Intrinsic apoptosis induction relies on the balance between the anti-apoptotic proteins (BCL-2, BCL-XL, MCL-1, BCLW, and BFL-1) and the pro-apoptotic proteins, which are divided into two groups: the effectors (BAX, BAK, and BOK) and the BH3-only proteins (BIM, BAD, NOXA, PUMA, BID, BIK, HRK) [70]. Consequently, both the upregulation of anti-apoptotic and the downregulation of pro-apoptotic signals can modulate AML cells’ sensitivity to BH3-mimetics and induce VEN resistance.
(a)
Increase in anti-apoptotic proteins.
  • BCL2
Several studies have shown that BCL2 is often overexpressed in AML cells, and is associated with poor prognosis and resistance to chemotherapy [71,72,73]. This phenomenon has also been found with VEN in vitro. Mechanistically, overexpression of BCL2 is related to its phosphorylation both in myeloid and lymphoid malignancies [74,75,76]. As a matter of fact, BCL2 dependency seems to be more important than expression level. Such a dependency on anti-apoptotic proteins can be revealed by performing a BH3 profiling”, assessing cancer cell susceptibility to induce MOMP. In this assay, AML cells are treated in vitro with synthetic BH3 peptides that selectively interact with pro-survival BCL-2 family proteins. For example, dependencies on BCL2 or MCL1 are inferred if MOMP occurs in response to BAD or MS-1 peptides, respectively [77,78]. An elegant study from the Letai’s lab showed that BH3 mimetic resistance is characterized by decreased mitochondrial apoptotic priming, both in PDX models and human clinical samples, due to alterations in BCL2 family proteins that vary among cases independently of acquired mutations in leukemia genes [79].
  • MCL1
Upregulation of the other anti-apoptotic MCL1 have also been identified as major determinants of resistance to intensive chemotherapy [80] as well as VEN resistance in both lymphoid and myeloid leukemias [76,81]. A strong correlation between MCL1 expression/dependency and abnormalities in signalization pathways, such as FLT3 mutations, AKT, or RAS/MAP deregulation, has been observed [82,83,84,85,86]. Moreover, differentiation status influences both signalization and MCL1 dependency. For instance, differentiation state and BH3 family change in protein expression are highly dependent [87,88,89,90,91]. Several studies have shown that monocytic AML is more resistant to VEN, and treatment with VEN selected differentiated monocytic cells at relapse [88,90].
  • BCLXL and BCL2A1
A recent study spotlighted on the functional dependence of AML subtypes’ erythroid or megakaryocytic differentiation, including acute erythroid leukemia (AEL), acute megakaryoblastic leukemia (AMKL), and pro-survival B-cell lymphoma-extra-large (BCL-xL) dependencies. Interestingly, BCLXL expression in relation to erythroid differentiation is highly connected with TP53 status, which is frequently mutated in these diseases [92,93]. In other studies, transcriptomic analysis of AML patients’ samples showed differential expressions of BCL2A1 in VEN-resistant cells. Upregulation of BCL2A1 expression has been correlated with higher VEN resistance. Knockdown of BCL2A1 restored apoptosis and reduced cell growth in the resistant AML cells without any substantial effect on the CD34+ hematopoietic stem and progenitor cells [65,94].
(b)
Decrease in pro-apoptotic signals.
Crispr screens studies have identified NOXA and PUMA as critical regulators of VEN activity in vitro [42,44,45,46]. Consistently, resistant AML cells have often been associated with a decrease in the expression of pro-apoptotic BAX and NOXA [75,95]. Mechanistically, it is noticeable that TP53 transcriptionally regulates the expression of BAX and NOXA proteins [41,43,96,97,98], partially explaining the role of TP53 mutation in mediating VEN resistance. As a proof of concept, NOXA-deficient cells were shown to be resistant to the combination of VEN-AZA and PUMA in vitro [99], which epigenetically mediated deregulation, causing VEN resistance in lymphoid [100] and myeloid leukemias [101].

4.2. Differentiation Status

Classical models have shown normal and pathologic hematopoiesis as a hematopoietic differentiation cascade with distinct cell states between stem/progenitor and terminally differentiated cells. It is now admitted that leukemic differentiation is more complex, based on single-cell analyses showing clonal heterogeneity [10,102]. An analysis of the BEAT AML data set (NCT03013998) demonstrated increased VEN resistance among leukemic patients with M4 and M5 AML [103]. Single-cell multiomics gave a better understanding of how monocytic differentiated subclones impact resistance to BCL2 inhibition in AML [90]. Differentiation status is somehow related to various expressions of pro- and anti-apoptotic protein. A recent study found that AML leukemic stem cells (LSC), based on flow cytometry sorting, were eliminated by VEN-AZA and determined the therapy outcome. Authors have developed and validated a score linking the ratio of protein expression of anti-apoptotic proteins BCL2, BCLXL, and MCL1 in LSCs to VEN response [104]. Together with ex vivo sensitivity profiling testing [91,92,105], these new tools would help to better identify resistance at a single-cell level.

4.3. Metabolism and Beyond

Mitochondria are organelles closely linked to cell death and cellular respiration, as well as other metabolic pathways. Crispr screen studies have found many genes that related to mitochondrial metabolism or structure critically regulate drugs targeting mitochondria and VEN sensitivity in vitro [45,46,106]. A recent example is CLPB, a gene maintaining the mitochondrial cristae structure via its interaction with the cristae-shaping protein OPA1. The CLPB protein is upregulated in human AML and is further induced upon acquisition of VEN resistance. Consequently, its ablation sensitizes AML to VEN [45].
Work from our lab has found that targeting mitochondrial iron can induce a dramatic loss of mitochondrial respiration, inducing BAX/BAK activation through pathways independent of BH3 mimetics. This new cell death modality on the frontier between apoptosis and ferroptosis was highly synergistic with VEN in vitro and in vivo [106]. Other studies have underpinned the important role of mitochondrial chaperones such as MARCH5 E3 ubiquitin ligase [107,108] or DRP1 [109].
Finally, several studies have tried to better understand the role of OXPHOS [110], mitochondrial respiration [111], amino acids [112], and lipids such as fatty acids [113,114,115,116], nicotinamide [117], and ceramide [118,119], in VEN response. It is likely that a better understanding of the integrated stress response (ISR) mediated by the ATF4 transcription factor will help to fill the gap between cell death, metabolism, and VEN resistance [46,91,106,120].
Another layer of complexity involves the role of the microenvironment. It is noteworthy that VEN enhances T cell-mediated antileukemic activity by increasing ROS production [121] and that MCL1 targeting can modulate leukemia cell metabolism, cell adhesion proteins, and leukemia–stromal interactions [122].

5. Overcoming Resistance with New Therapies

5.1. Hypomethylating Agents

The rationale for the synergistic activity between BCL2 inhibitors and AZA has been explored in preclinical works. It has been shown that AZA does not affect the expression of anti-apoptotic proteins but induces the expression of the pro-apoptotic BH3 proteins NOXA and PUMA only, resulting in a priming of AML cells to intrinsic apoptosis. Interestingly, this phenomenon is not epigenetically regulated, as shown by the absence of methylation changes in DNA promoters and the rapid kinetics of apoptosis induction [99,123]. This synergy may also be related to the metabolic activity of VEN-AZA on the leukemia stem cell [7,112]. The optimization of current non-intensive regimens such as weekly decitabine [124] and the use of alternative oral hypomethylating drugs may possibly modulate the synergistic effects obtained with VEN [125,126,127].

5.2. Innovative BH3 Mimetics

  • MCL1
There is a strong rationale for targeting MCL1 anti-apoptotic protein in preclinical models based on dependency of AML cells for sustaining blast proliferation [42,75,79,128,129]. Nevertheless, clinical development was halted by cardiac toxicity (mainly troponin elevations). Current MCL1 trials assess the safety of these MCL1 inhibitors as single agents, with the provision of combination therapy with VEN (NCT03672695, NCT02675452, NCT03218683) [130]. Recent reviews have covered the field of MCL1 inhibitors currently in development [131,132,133].
  • Dual BCL2 and BCLXL inhibitors
Despite its toxicity to platelets, navitoclax has seen renewed interest in targeting resistant AML blasts. The phase II TRANSFORM II study (NCT04472598) is currently recruiting relapsed or refractory myelofibrosis patients. New dual BCL2/BCLXL inhibitors are in development, such as AZD4320 [134]. VHL-recruiting proteolysis-targeting chimeras (PROTACs) cause BC2 and BCLXL degradation via an interaction with the von Hippel–Lindau E3 ubiquitin ligase. As platelets lack VHL expression, the drug spares the on-target platelet toxicity caused by navitoclax [135].

5.3. Tyrosine Kinase Inhibitor (TKI)

Indirect MCL-1 inhibition can be achieved by targeting cyclin-dependent kinases, including 7 and 9. Some early trials show promising responses [136]. The CDK9 inhibitor AZD4573 was tested in R/R AML (NCT05140382) and showed a 15% response rate in this heavily pretreated AML patients population [137]. Other cyclins may be of interest, such as CDK7/12/13 [138]

5.4. FLT3 Inhibitors

There is extensive preclinical evidence in combining VEN and FLT3 inhibitors (FLT3inh), including gilteritinib, midostaurin, and quizartinib [139,140,141]. Gilteritinib is a type 1 FLT3 inhibitor which has activity against FLT3, ALK, and AXL [48]. Mechanistically, VEN with gilteritinib decreased phosphorylation of ERK and GSK3B via combined AXL and FLT3 inhibition, with subsequent suppression of the anti-apoptotic protein MCL1 [142]. The combination of VEN plus gilteritinib in FLT3mut patients was shown to induce a CR rate of 75% and median OS of 10 months in relapsed or refractory AML patients who had received prior FLT3 inhibitor therapy in 64% [143]. A study exploring giteritinib combined with VEN-AZA is also recruiting patients in a phase I/II trial [144].
Midostaurin acts both on FLT3-ITD and FLT3-TKD, and has a broad activity against tyrosine kinase receptors including PKC-α, FLT3, c-KIT, VEGFR, and PDGFR [48]. Midostaurin combined with VEN-AZA seemed to be feasible based on subset analyses from multiple investigator-initiated clinical studies, with a high response rate but significant hematological toxicities [145,146].
Quizartinib shows high selectivity to FLT3, KIT, colony-stimulating factor-1 receptor, platelet-derived growth factor receptor, and RET kinase [48]. In a phase I/II study presented at the ASH meeting in 2023, 50 patients were treated with a triplet combining VEN, decitabine, and quizartinib. This series included 40 R/R and 10 newly FLT3-ITD-mutated AML patients. Among the R/R patients, the composite CR rate (including CR and CR with incomplete hematological recovery) was 68%. The 60-day mortality rate was 0%. With a median follow-up of 33 months, the median OS was 7.1 months. Non-hematologic toxicities included pneumonia (73%), neutropenic fever (53%), sepsis (18%), bacteremia (15%), and other infections involving the skin (18%) and the gastrointestinal tract (13%) [147].

5.5. IDH Inhibitors

The efficacy of combining AZA with the IDH1 inhibitor ivosidenib was shown in the phase III AGILE study. This regimen is now approved for newly diagnosed IDH1mut AML patients [148]. Given the high response rate of IDHmut patients, the triplet regimen was further evaluated. A phase I/II trial showed a promising 90% composite complete response rate and 42 months median overall survival [149]. The safety and efficacy of combining the IDH2 inhibitor enasidenib with VEN-AZA was also evaluated in AML patients with IDH2mut. The response rates were also very high [150]. The benefit from adding IDH inhibitors to VEN-AZA must be confirmed in randomized trials. The main classes of drugs found to synergize with VEN in clinical studies are summarized in Figure 3.

5.6. Future Therapies

Targeting the interaction between tumor suppressor P53 and the E3 ligase MDM2 represents an attractive treatment approach for cancers with wild-type or functional TP53. Indeed, several small molecules have been developed and evaluated in various malignancies [151]. The MDM2 inhibitors developed in clinics include idasanutlin and milademetan (DS3032b). These drugs are currently in phase I trials in combination with VEN for TP53wt patients [152,153].
For TP53mut patients, few therapeutic advances have been made so far. Eprenetapopt (APR-246) is a first-in-class, small-molecule p53 reactivator. A phase I study combining the drug with VEN-AZA has associated it with an acceptable safety profile and encouraging activity. Indeed, 25 (64%) patients had an overall response and 15 (38%, 23–55) had a complete response [154].
Menin inhibitors constitute a novel class of agents targeting the underlying biology of nucleophosmin (NPM1) mutant and KMT2A rearranged (KMT2A acute leukemias [155,156]. Several menin inhibitors have been developed for AML, including revumenib, ziftomenib, and JNJ-6617. This last drug is currently being evaluated in monotherapy and in combination with VEN-AZA in two phase I clinical trials, 75276617ALE1001 (NCT04811560) and 1002, respectively [157].
Innovative immunotherapies are under evaluation, including the monoclonal antibody magrolimab against CD47. Development of the drug is currently on hold because of its lack of efficacy [158]. Nevertheless, the Enhance-3 study including newly diagnosed AML cases displayed promising results (80% overall response rate) in the frontline setting [159]. The antibody–drug conjugate targeting CD123 IMGN632, also known as pivekimab sunirine (or PVEK), is currently being tested in monotherapy or with VEN-AZA for patients with CD123-positive AML (NCT04086264). The first results were presented both in R/R and frontline cohorts and showed a tolerable safety profile and promising efficacy [160].
The main clinical studies of triplet therapies, including VEN-AZA plus other drugs, are summarized in Table 2.

6. Conclusions

Non-intensive approaches are rarely curative in AML. Historically, HMA monotherapy had been associated with a 20–30% response rate and a 10-month median OS. The development of VEN-HMA-based combination regimens has revolutionized the treatment of these aggressive diseases, as it has led to a doubling of response rates and a significant increase in survival. Despite these good results, relapses and primary refractory AML are still a matter of concern.
It is now clear that TP53 mutations are associated with the lack of efficacy of VEN-AZA regimens contrary to IDH1/2 mutations (in particular IDH2), which confers the high efficacy of VEN. For most patients who are not harboring these molecular alterations or for those who harbor multiple alterations, the efficacy of VEN-AZA regimen is hard to anticipate. The utilization of minimal residual disease (MRD) assessments by flow cytometry and molecular analyses, although well described in the context of ICT [161,162], have to be clearly defined to guide therapeutics and prevent relapse upon VEN-AZA use [163,164,165]. It is likely that single-cell DNA sequencing using, for instance, the Tapestri® plateform developed by mission bio [166] will help us to better understand the role of clonal evolutions in the context of resistance.
Triplet therapies based on the combination of VEN/HMA + targeted agents may represent good options for patients with targeted molecular alterations, including IDH, FLT3, and NPM1 or KMT2A alterations; however, to date, evidence is lacking for supporting the use of triplets instead of VEN-AZA bi-therapy. It is likely that a combination with IDH or menin inhibitors will be feasible in the context of elderly AML patients given the low apparently hematological toxicities that have been reported with these two classes. On the other hand, combinations with FLT3 inhibitors may be difficult given the high additional toxicity of these classes of drugs with VEN-AZA. Therefore, these types of regimens may be reserved for fit patients failing intensive chemotherapy.
Finally, in the context of elderly AML, quality of life preservation is a major goal to achieve. Some retrospective studies have shown the feasibility of stopping VEN-AZA in remission [167,168]. It is likely that in the future, VEN-AZA regimens will be a backbone for new triplet therapies and off-treatment periods based on precise molecular alteration assessments and single-cell MRD follow-up.

Author Contributions

Conceptualization, S.G.; writing—original draft preparation, S.G.; writing—review and editing, S.G., Y.C. and M.-A.H.; supervision, N.V. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

S.G. declares a consulting or advisory role with Abbvie, Astellas, BMS-Celgene, Jazz Pharmaceuticals as well as Servier and received travel grants from Gilead. Y.C., M.-A.H. and N.V. declare no conflicts of interest.

References

  1. Dombret, H.; Seymour, J.F.; Butrym, A.; Wierzbowska, A.; Selleslag, D.; Jang, J.H.; Kumar, R.; Cavenagh, J.; Schuh, A.C.; Candoni, A.; et al. International phase 3 study of azacitidine vs conventional care regimens in older patients with newly diagnosed AML with >30% blasts. Blood 2015, 126, 291–299. [Google Scholar] [CrossRef]
  2. Kantarjian, H.M.; Thomas, X.G.; Dmoszynska, A.; Wierzbowska, A.; Mazur, G.; Mayer, J.; Gau, J.-P.; Chou, W.-C.; Buckstein, R.; Cermak, J.; et al. Multicenter, Randomized, Open-Label, Phase III Trial of Decitabine Versus Patient Choice, With Physician Advice, of Either Supportive Care or Low-Dose Cytarabine for the Treatment of Older Patients With Newly Diagnosed Acute Myeloid Leukemia. J. Clin. Oncol. 2012, 30, 2670–2677. [Google Scholar] [CrossRef]
  3. Dinardo, C.D.; Jonas, B.A.; Pullarkat, V.; Thirman, M.J.; Garcia, J.S.; Wei, A.H.; Konopleva, M.; Döhner, H.; Letai, A.; Fenaux, P.; et al. Azacitidine and Venetoclax in Previously Untreated Acute Myeloid Leukemia. N. Engl. J. Med. 2020, 383, 617–629. [Google Scholar] [CrossRef]
  4. Pratz, K.W.; Jonas, B.A.; Pullarkat, V.A.; Thirman, M.J.; Garcia, J.S.; Fiedler, W.; Yamamoto, K.; Wang, J.; Yoon, S.-S.; Wolach, O.; et al. Long-Term Follow-up of the Phase 3 Viale-a Clinical Trial of Venetoclax Plus Azacitidine for Patients with Untreated Acute Myeloid Leukemia Ineligible for Intensive Chemotherapy. Blood 2022, 140, 529–531. [Google Scholar] [CrossRef]
  5. Maiti, A.; Rausch, C.R.; Cortes, J.E.; Pemmaraju, N.; Daver, N.G.; Ravandi, F.; Garcia-Manero, G.; Borthakur, G.; Naqvi, K.; Ohanian, M.; et al. Outcomes of relapsed or refractory acute myeloid leukemia after frontline hypomethylating agent and venetoclax regimens. Haematologica 2021, 106, 894–898. [Google Scholar] [CrossRef] [PubMed]
  6. Maiti, A.; Carter, B.Z.; Andreeff, M.; Konopleva, M.Y. SOHO State of the Art Updates and Next Questions|Beyond BCL-2 Inhibition in Acute Myeloid Leukemia: Other Approaches to Leverage the Apoptotic Pathway. Clin. Lymphoma Myeloma Leuk. 2022, 22, 652–658. [Google Scholar] [CrossRef] [PubMed]
  7. Garciaz, S.; Saillard, C.; Hicheri, Y.; Hospital, M.-A.; Vey, N. Venetoclax in Acute Myeloid Leukemia: Molecular Basis, Evidences for Preclinical and Clinical Efficacy and Strategies to Target Resistance. Cancers 2021, 13, 5608. [Google Scholar] [CrossRef] [PubMed]
  8. Sullivan, G.P.; Flanagan, L.; Rodrigues, D.A.; Chonghaile, T.N. The path to venetoclax resistance is paved with mutations, metabolism, and more. Sci. Transl. Med. 2022, 14, eabo6891. [Google Scholar] [CrossRef] [PubMed]
  9. Xu, Y.; Ye, H. Progress in understanding the mechanisms of resistance to BCL-2 inhibitors. Exp. Hematol. Oncol. 2022, 11, 31. [Google Scholar] [CrossRef] [PubMed]
  10. Stubbins, R.J.; Maksakova, I.A.; Sanford, D.S.; Rouhi, A.; Kuchenbauer, F. Mitochondrial metabolism: Powering new directions in acute myeloid leukemia. Leuk. Lymphoma 2021, 62, 2331–2341. [Google Scholar] [CrossRef] [PubMed]
  11. Ong, F.; Kim, K.; Konopleva, M.Y. Venetoclax resistance: Mechanistic insights and future strategies. Cancer Drug Resist. 2022, 5, 380–400. [Google Scholar] [CrossRef]
  12. Garciaz, S.; Miller, T.; Collette, Y.; Vey, N. Targeting regulated cell death pathways in acute myeloid leukemia. Cancer Drug Resist. 2023, 6, 151–168. [Google Scholar] [CrossRef]
  13. Griffioen, M.S.; de Leeuw, D.C.; Janssen, J.J.W.M.; Smit, L. Targeting Acute Myeloid Leukemia with Venetoclax; Biomarkers for Sensitivity and Rationale for Venetoclax-Based Combination Therapies. Cancers 2022, 14, 3456. [Google Scholar] [CrossRef]
  14. DiNardo, C.D.; Tiong, I.S.; Quaglieri, A.; MacRaild, S.; Loghavi, S.; Brown, F.C.; Thijssen, R.; Pomilio, G.; Ivey, A.; Salmon, J.M.; et al. Molecular patterns of response and treatment failure after frontline venetoclax combinations in older patients with AML. Blood 2020, 135, 791–803. [Google Scholar] [CrossRef] [PubMed]
  15. Wilson, W.H.; O’Connor, O.A.; Czuczman, M.S.; LaCasce, A.S.; Gerecitano, J.F.; Leonard, J.P.; Tulpule, A.; Dunleavy, K.; Xiong, H.; Chiu, Y.-L.; et al. Navitoclax, a targeted high-affinity inhibitor of BCL-2, in lymphoid malignancies: A phase 1 dose-escalation study of safety, pharmacokinetics, pharmacodynamics, and antitumour activity. Lancet Oncol. 2010, 11, 1149–1159. [Google Scholar] [CrossRef]
  16. Souers, A.J.; Leverson, J.D.; Boghaert, E.R.; Ackler, S.L.; Catron, N.D.; Chen, J.; Dayton, B.D.; Ding, H.; Enschede, S.H.; Fairbrother, W.J.; et al. ABT-199, a potent and selective BCL-2 inhibitor, achieves antitumor activity while sparing platelets. Nat. Med. 2013, 19, 202–208. [Google Scholar] [CrossRef] [PubMed]
  17. Pan, R.; Hogdal, L.J.; Benito, J.M.; Bucci, D.; Han, L.; Borthakur, G.; Cortes, J.; DeAngelo, D.J.; Debose, L.; Mu, H.; et al. Selective BCL-2 Inhibition by ABT-199 Causes On-Target Cell Death in Acute Myeloid Leukemia. Cancer Discov. 2014, 4, 362–375. [Google Scholar] [CrossRef] [PubMed]
  18. DiNardo, C.D.; Pratz, K.W.; Letai, A.; Jonas, B.A.; Wei, A.H.; Thirman, M.; Arellano, M.; Frattini, M.G.; Kantarjian, H.; Popovic, R.; et al. Safety and preliminary efficacy of venetoclax with decitabine or azacitidine in elderly patients with previously untreated acute myeloid leukaemia: A non-randomised, open-label, phase 1b study. Lancet Oncol. 2018, 19, 216–228. [Google Scholar] [CrossRef]
  19. Dinardo, C.D.; Pratz, K.; Pullarkat, V.; Jonas, B.A.; Arellano, M.; Becker, P.S.; Frankfurt, O.; Konopleva, M.; Wei, A.H.; Kantarjian, H.M.; et al. Venetoclax combined with decitabine or azacitidine in treatment-naive, elderly patients with acute myeloid leukemia. Blood 2019, 133, 7–17. [Google Scholar] [CrossRef]
  20. DiNardo, C.D.; Maiti, A.; Rausch, C.R.; Pemmaraju, N.; Naqvi, K.; Daver, N.G.; Kadia, T.M.; Borthakur, G.; Ohanian, M.; Alvarado, Y.; et al. 10-day decitabine with venetoclax for newly diagnosed intensive chemotherapy ineligible, and relapsed or refractory acute myeloid leukaemia: A single-centre, phase 2 trial. Lancet Haematol. 2020, 7, e724–e736. [Google Scholar] [CrossRef]
  21. Wei, A.H.; Montesinos, P.; Ivanov, V.; DiNardo, C.D.; Novak, J.; Laribi, K.; Kim, I.; Stevens, D.A.; Fiedler, W.; Pagoni, M.; et al. Venetoclax plus LDAC for newly diagnosed AML ineligible for intensive chemotherapy: A phase 3 randomized placebo-controlled trial. Blood 2020, 135, 2137–2145. [Google Scholar] [CrossRef] [PubMed]
  22. Konopleva, M.; Pollyea, D.A.; Potluri, J.; Chyla, B.; Hogdal, L.; Busman, T.; McKeegan, E.; Salem, A.H.; Zhu, M.; Ricker, J.L.; et al. Efficacy and Biological Correlates of Response in a Phase II Study of Venetoclax Monotherapy in Patients with Acute Myelogenous Leukemia. Cancer Discov. 2016, 6, 1106–1117. [Google Scholar] [CrossRef] [PubMed]
  23. Garciaz, S.; Hospital, M.-A.; Alary, A.-S.; Saillard, C.; Hicheri, Y.; Mohty, B.; Rey, J.; D’incan, E.; Charbonnier, A.; Villetard, F.; et al. Azacitidine Plus Venetoclax for the Treatment of Relapsed and Newly Diagnosed Acute Myeloid Leukemia Patients. Cancers 2022, 14, 2025. [Google Scholar] [CrossRef] [PubMed]
  24. Petit, C.; Saillard, C.; Mohty, B.; Hicheri, Y.; Villetard, F.; Maisano, V.; Charbonnier, A.; Rey, J.; D‘incan, E.; Rouzaud, C.; et al. Azacitidine–venetoclax versus azacitidine salvage treatment for primary induction failure or first relapsed acute myeloid leukaemia patients. Eur. J. Haematol. 2023. [Google Scholar] [CrossRef] [PubMed]
  25. Todisco, E.; Papayannidis, C.; Fracchiolla, N.; Petracci, E.; Zingaretti, C.; Vetro, C.; Martelli, M.P.; Zappasodi, P.; Di Renzo, N.; Gallo, S.; et al. AVALON: The Italian cohort study on real-life efficacy of hypomethylating agents plus venetoclax in newly diagnosed or relapsed/refractory patients with acute myeloid leukemia. Cancer 2023, 129, 992–1004. [Google Scholar] [CrossRef] [PubMed]
  26. Shimony, S.; Bewersdorf, J.P.; Shallis, R.M.; Liu, Y.; Schaefer, E.J.; Zeidan, A.M.; Goldberg, A.D.; Stein, E.M.; Marcucci, G.; Lindsley, R.C.; et al. Hypomethylating agents plus venetoclax compared with intensive induction chemotherapy regimens in molecularly defined secondary AML. Leukemia 2024. [Google Scholar] [CrossRef]
  27. Döhner, H.; Pratz, K.W.; DiNardo, C.D.; Jonas, B.A.; Pullarkat, V.A.; Thirman, M.J.; Recher, C.; Schuh, A.C.; Babu, S.; Dail, M.; et al. ELN Risk Stratification Is Not Predictive of Outcomes for Treatment-Naïve Patients with Acute Myeloid Leukemia Treated with Venetoclax and Azacitidine. Blood 2022, 140, 1441–1444. [Google Scholar] [CrossRef]
  28. Molenaar, R.J.; Wilmink, J.W. IDH1/2 Mutations in Cancer Stem Cells and Their Implications for Differentiation Therapy. J. Histochem. Cytochem. 2022, 70, 83–97. [Google Scholar] [CrossRef]
  29. Pollyea, D.A.; DiNardo, C.D.; Arellano, M.L.; Pigneux, A.; Fiedler, W.; Konopleva, M.; Rizzieri, D.A.; Smith, B.D.; Shinagawa, A.; Lemoli, R.M.; et al. Impact of Venetoclax and Azacitidine in Treatment-Naïve Patients with Acute Myeloid Leukemia and IDH1/2 Mutations. Clin. Cancer Res. 2022, 28, 2753–2761. [Google Scholar] [CrossRef]
  30. Hammond, D.; Loghavi, S.; Wang, S.A.; Konopleva, M.Y.; Kadia, T.M.; Daver, N.G.; Ohanian, M.; Issa, G.C.; Alvarado, Y.; Short, N.J.; et al. Response patterns and impact of MRD in patients with IDH1/2-mutated AML treated with venetoclax and hypomethylating agents. Blood Cancer J. 2023, 13, 148. [Google Scholar] [CrossRef]
  31. Chan, S.M.; Thomas, D.; Corces-Zimmerman, M.R.; Xavy, S.; Rastogi, S.; Hong, W.-J.; Zhao, F.; Medeiros, B.C.; Tyvoll, D.A.; Majeti, R. Isocitrate dehydrogenase 1 and 2 mutations induce BCL-2 dependence in acute myeloid leukemia. Nat. Med. 2015, 21, 178–184. [Google Scholar] [CrossRef]
  32. Stuani, L.; Sabatier, M.; Saland, E.; Cognet, G.; Poupin, N.; Bosc, C.; Castelli, F.A.; Gales, L.; Turtoi, E.; Montersino, C.; et al. Mitochondrial metabolism supports resistance to IDH mutant inhibitors in acute myeloid leukemia. J. Exp. Med. 2021, 218, e20200924. [Google Scholar] [CrossRef]
  33. Arber, D.A.; Orazi, A.; Hasserjian, R.P.; Borowitz, M.J.; Calvo, K.R.; Kvasnicka, H.-M.; Wang, S.A.; Bagg, A.; Barbui, T.; Branford, S.; et al. International Consensus Classification of Myeloid Neoplasms and Acute Leukemias: Integrating morphologic, clinical, and genomic data. Blood 2022, 140, 1200–1228. [Google Scholar] [CrossRef]
  34. Khoury, J.D.; Solary, E.; Abla, O.; Akkari, Y.; Alaggio, R.; Apperley, J.F.; Bejar, R.; Berti, E.; Busque, L.; Chan, J.K.C.; et al. The 5th edition of the World Health Organization Classification of Haematolymphoid Tumours: Myeloid and Histiocytic/Dendritic Neoplasms. Leukemia 2022, 36, 1703–1719. [Google Scholar] [CrossRef]
  35. Fleming, S.; Tsai, X.C.-H.; Morris, R.; Hou, H.-A.; Wei, A.H. TP53 status and impact on AML prognosis within the ELN 2022 risk classification. Blood 2023, 142, 2029–2033. [Google Scholar] [CrossRef] [PubMed]
  36. Grob, T.; Al Hinai, A.S.A.; Sanders, M.A.; Kavelaars, F.G.; Rijken, M.; Gradowska, P.L.; Biemond, B.J.; Breems, D.A.; Maertens, J.; Kooy, M.v.M.; et al. Molecular characterization of mutant TP53 acute myeloid leukemia and high-risk myelodysplastic syndrome. Blood 2022, 139, 2347–2354. [Google Scholar] [CrossRef] [PubMed]
  37. Daver, N.G.; Iqbal, S.; Huang, J.; Renard, C.; Lin, J.; Pan, Y.; Williamson, M.; Ramsingh, G. Clinical characteristics and overall survival among acute myeloid leukemia patients with TP53 gene mutation or chromosome 17p deletion. Am. J. Hematol. 2023, 98, 1176–1184. [Google Scholar] [CrossRef] [PubMed]
  38. Daver, N.G.; Iqbal, S.; Renard, C.; Chan, R.J.; Hasegawa, K.; Hu, H.; Tse, P.; Yan, J.; Zoratti, M.J.; Xie, F.; et al. Treatment outcomes for newly diagnosed, treatment-naïve TP53-mutated acute myeloid leukemia: A systematic review and meta-analysis. J. Hematol. Oncol. 2023, 16, 19. [Google Scholar] [CrossRef] [PubMed]
  39. Kim, K.; Maiti, A.; Loghavi, S.; Pourebrahim, R.; Kadia, T.M.; Rausch, C.R.; Furudate, K.; Daver, N.G.; Alvarado, Y.; Ohanian, M.; et al. Outcomes of TP53 -mutant acute myeloid leukemia with decitabine and venetoclax. Cancer 2021, 127, 3772–3781. [Google Scholar] [CrossRef] [PubMed]
  40. Pollyea, D.A.; Pratz, K.W.; Wei, A.H.; Pullarkat, V.; Jonas, B.A.; Recher, C.; Babu, S.; Schuh, A.C.; Dail, M.; Sun, Y.; et al. Outcomes in Patients with Poor-Risk Cytogenetics with or without TP53 Mutations Treated with Venetoclax and Azacitidine. Clin. Cancer Res. 2022, 28, 5272–5279. [Google Scholar] [CrossRef]
  41. Aubrey, B.J.; Kelly, G.L.; Janic, A.; Herold, M.J.; Strasser, A. How does p53 induce apoptosis and how does this relate to p53-mediated tumour suppression? Cell Death Differ. 2018, 25, 104–113. [Google Scholar] [CrossRef]
  42. Thijssen, R.; Diepstraten, S.T.; Moujalled, D.; Chew, E.; Flensburg, C.; Shi, M.X.; Dengler, M.A.; Litalien, V.; MacRaild, S.; Chen, M.; et al. Intact TP-53 function is essential for sustaining durable responses to BH3-mimetic drugs in leukemias. Blood 2021, 137, 2721–2735. [Google Scholar] [CrossRef]
  43. Fischer, M. Census and evaluation of p53 target genes. Oncogene 2017, 36, 3943–3956. [Google Scholar] [CrossRef] [PubMed]
  44. Nechiporuk, T.; Kurtz, S.E.; Nikolova, O.; Liu, T.; Jones, C.L.; D’Alessandro, A.; Culp-Hill, R.; D’Almeida, A.; Joshi, S.K.; Rosenberg, M.; et al. The TP53 Apoptotic Network Is a Primary Mediator of Resistance to BCL2 Inhibition in AML Cells. Cancer Discov. 2019, 9, 910–925. [Google Scholar] [CrossRef] [PubMed]
  45. Chen, X.; Glytsou, C.; Zhou, H.; Narang, S.; Reyna, D.E.; Lopez, A.; Sakellaropoulos, T.; Gong, Y.; Kloetgen, A.; Yap, Y.S.; et al. Targeting Mitochondrial Structure Sensitizes Acute Myeloid Leukemia to Venetoclax Treatment. Cancer Discov. 2019, 9, 890–909. [Google Scholar] [CrossRef] [PubMed]
  46. Sharon, D.; Cathelin, S.; Mirali, S.; Di Trani, J.M.; Yanofsky, D.J.; Keon, K.A.; Rubinstein, J.L.; Schimmer, A.D.; Ketela, T.; Chan, S.M. Inhibition of mitochondrial translation overcomes venetoclax resistance in AML through activation of the integrated stress response. Sci. Transl. Med. 2019, 11, eaax2863. [Google Scholar] [CrossRef] [PubMed]
  47. Nishida, Y.; Ishizawa, J.; Ayoub, E.; Montoya, R.H.; Ostermann, L.B.; Muftuoglu, M.; Ruvolo, V.R.; Patsilevas, T.; Scruggs, D.A.; Khazaei, S.; et al. Enhanced TP53 reactivation disrupts MYC transcriptional program and overcomes venetoclax resistance in acute myeloid leukemias. Sci. Adv. 2023, 9, eadh1436. [Google Scholar] [CrossRef]
  48. Garciaz, S.; Hospital, M.-A. FMS-Like Tyrosine Kinase 3 Inhibitors in the Treatment of Acute Myeloid Leukemia: An Update on the Emerging Evidence and Safety Profile. OncoTargets Ther. 2023, 16, 31–45. [Google Scholar] [CrossRef]
  49. Papaemmanuil, E.; Gerstung, M.; Bullinger, L.; Gaidzik, V.I.; Paschka, P.; Roberts, N.D.; Potter, N.E.; Heuser, M.; Thol, F.; Bolli, N.; et al. Genomic Classification and Prognosis in Acute Myeloid Leukemia. N. Engl. J. Med. 2016, 374, 2209–2221. [Google Scholar] [CrossRef]
  50. Konopleva, M.; Thirman, M.J.; Pratz, K.W.; Garcia, J.S.; Recher, C.; Pullarkat, V.; Kantarjian, H.M.; DiNardo, C.D.; Dail, M.; Duan, Y.; et al. Impact of F LT3 Mutation on Outcomes after Venetoclax and Azacitidine for Patients with Treatment-Naïve Acute Myeloid Leukemia. Clin. Cancer Res. 2022, 28, 2744–2752. [Google Scholar] [CrossRef]
  51. Stahl, M.; Menghrajani, K.; Derkach, A.; Chan, A.; Xiao, W.; Glass, J.; King, A.C.; Daniyan, A.F.; Famulare, C.; Cuello, B.M.; et al. Clinical and molecular predictors of response and survival following venetoclax therapy in relapsed/refractory AML. Blood Adv. 2021, 5, 1552–1564. [Google Scholar] [CrossRef]
  52. Bataller, A.; Bazinet, A.; DiNardo, C.D.; Maiti, A.; Borthakur, G.; Daver, N.G.; Short, N.J.; Jabbour, E.J.; Issa, G.C.; Pemmaraju, N.; et al. Prognostic risk signature in patients with acute myeloid leukemia treated with hypomethylating agents and venetoclax. Blood Adv. 2023, 8, 927–935. [Google Scholar] [CrossRef] [PubMed]
  53. Jahn, E.; Saadati, M.; Fenaux, P.; Gobbi, M.; Roboz, G.J.; Bullinger, L.; Lutsik, P.; Riedel, A.; Plass, C.; Jahn, N.; et al. Clinical impact of the genomic landscape and leukemogenic trajectories in non-intensively treated elderly acute myeloid leukemia patients. Leukemia 2023, 37, 2187–2196. [Google Scholar] [CrossRef] [PubMed]
  54. McCarter, J.G.W.; Nemirovsky, D.; Famulare, C.A.; Farnoud, N.; Mohanty, A.S.; Stone-Molloy, Z.S.; Chervin, J.; Ball, B.J.; Epstein-Peterson, Z.D.; Arcila, M.E.; et al. Interaction between myelodysplasia-related gene mutations and ontogeny in acute myeloid leukemia. Blood Adv. 2023, 7, 5000–5013. [Google Scholar] [CrossRef] [PubMed]
  55. Döhner, H.; Wei, A.H.; Appelbaum, F.R.; Craddock, C.; DiNardo, C.D.; Dombret, H.; Ebert, B.L.; Fenaux, P.; Godley, L.A.; Hasserjian, R.P.; et al. Diagnosis and management of AML in adults: 2022 recommendations from an international expert panel on behalf of the ELN. Blood 2022, 140, 1345–1377. [Google Scholar] [CrossRef]
  56. Doyel, M.; Bouligny, I.M.; Murray, G.; Patel, T.; Boron, J.; Tran, V.; Gor, J.; Hang, Y.; Ho, T.; Zacholski, K.; et al. The impact of secondary-type mutations in newly diagnosed acute myeloid leukemia treated with venetoclax and decitabine or azacitidine. J. Clin. Oncol. 2023, 41, e19048. [Google Scholar] [CrossRef]
  57. Gangat, N.; Johnson, I.; McCullough, K.; Farrukh, F.; Al-Kali, A.; Alkhateeb, H.; Begna, K.; Mangaonkar, A.; Litzow, M.; Hogan, W.; et al. Molecular predictors of response to venetoclax plus hypomethylating agent in treatment-naïve acute myeloid leukemia. Haematologica 2022, 107, 2501–2505. [Google Scholar] [CrossRef]
  58. Rahmani, N.E.; Ramachandra, N.; Sahu, S.; Gitego, N.; Lopez, A.; Pradhan, K.; Bhagat, T.D.; Gordon-Mitchell, S.; Pena, B.R.; Kazemi, M.; et al. ASXL1 mutations are associated with distinct epigenomic alterations that lead to sensitivity to venetoclax and azacytidine. Blood Cancer J. 2021, 11, 157. [Google Scholar] [CrossRef]
  59. Shiozawa, Y.; Malcovati, L.; Gallì, A.; Sato-Otsubo, A.; Kataoka, K.; Sato, Y.; Watatani, Y.; Suzuki, H.; Yoshizato, T.; Yoshida, K.; et al. Aberrant splicing and defective mRNA production induced by somatic spliceosome mutations in myelodysplasia. Nat. Commun. 2018, 9, 3649. [Google Scholar] [CrossRef]
  60. Seiler, M.; Yoshimi, A.; Darman, R.; Chan, B.; Keaney, G.; Thomas, M.; Agrawal, A.A.; Caleb, B.; Csibi, A.; Sean, E.; et al. H3B-8800, an orally available small-molecule splicing modulator, induces lethality in spliceosome-mutant cancers. Nat. Med. 2018, 24, 497–504. [Google Scholar] [CrossRef]
  61. Caprioli, C.; Lussana, F.; Salmoiraghi, S.; Cavagna, R.; Buklijas, K.; Elidi, L.; Zanghi’, P.; Michelato, A.; Delaini, F.; Oldani, E.; et al. Clinical significance of chromatin-spliceosome acute myeloid leukemia: A report from the Northern Italy Leukemia Group (NILG) randomized trial 02/06. Haematologica 2020, 106, 2578–2587. [Google Scholar] [CrossRef] [PubMed]
  62. van der Werf, I.; Wojtuszkiewicz, A.; Meggendorfer, M.; Hutter, S.; Baer, C.; Heymans, M.; Valk, P.J.M.; Kern, W.; Haferlach, C.; Janssen, J.J.W.M.; et al. Splicing factor gene mutations in acute myeloid leukemia offer additive value if incorporated in current risk classification. Blood Adv. 2021, 5, 3254–3265. [Google Scholar] [CrossRef] [PubMed]
  63. Senapati, J.; Urrutia, S.; Loghavi, S.; Short, N.J.; Issa, G.C.; Maiti, A.; Abbas, H.A.; Daver, N.G.; Pemmaraju, N.; Pierce, S.; et al. Venetoclax Abrogates the Prognostic Impact of Splicing Factor Gene Mutations in Newly Diagnosed Acute Myeloid Leukemia. Blood 2023, 142, 1647–1657. [Google Scholar] [CrossRef] [PubMed]
  64. Lachowiez, C.A.; Loghavi, S.; Furudate, K.; Montalban-Bravo, G.; Maiti, A.; Kadia, T.; Daver, N.; Borthakur, G.; Pemmaraju, N.; Sasaki, K.; et al. Impact of splicing mutations in acute myeloid leukemia treated with hypomethylating agents combined with venetoclax. Blood Adv. 2021, 5, 2173–2183. [Google Scholar] [CrossRef] [PubMed]
  65. Zhang, H.; Nakauchi, Y.; Köhnke, T.; Stafford, M.; Bottomly, D.; Thomas, R.; Wilmot, B.; McWeeney, S.K.; Majeti, R.; Tyner, J.W. Integrated analysis of patient samples identifies biomarkers for venetoclax efficacy and combination strategies in acute myeloid leukemia. Nat. Cancer 2020, 1, 826–839. [Google Scholar] [CrossRef] [PubMed]
  66. Berton, G. Poor Prognosis of SRSF2 Gene Mutations in Patients Treated with Venetoclax-Azacitidine (VEN-AZA) for Newly Diagnosed Acute Myeloid Leukemia. a Multicentric Real-Life Study of 117 Patients. In Proceedings of the 65th ASH Annual Meeting & Exposition, ASH, San Diego, CA, USA, 9–12 December 2023; Available online: https://ash.confex.com/ash/2023/webprogram/Paper180733.html (accessed on 24 December 2023).
  67. Blombery, P.; Anderson, M.A.; Gong, J.-N.; Thijssen, R.; Birkinshaw, R.W.; Thompson, E.R.; Teh, C.E.; Nguyen, T.; Xu, Z.; Flensburg, C.; et al. Acquisition of the Recurrent Gly101Val Mutation in BCL2 Confers Resistance to Venetoclax in Patients with Progressive Chronic Lymphocytic Leukemia. Cancer Discov. 2019, 9, 342–353. [Google Scholar] [CrossRef] [PubMed]
  68. Blombery, P.; Lew, T.E.; Dengler, M.A.; Thompson, E.R.; Lin, V.S.; Chen, X.; Nguyen, T.; Panigrahi, A.; Handunnetti, S.M.; Carney, D.A.; et al. Clonal hematopoiesis, myeloid disorders and BAX-mutated myelopoiesis in patients receiving venetoclax for CLL. Blood 2022, 139, 1198–1207. [Google Scholar] [CrossRef] [PubMed]
  69. Moujalled, D.M.; Brown, F.C.; Chua, C.C.; Dengler, M.A.; Pomilio, G.; Anstee, N.S.; Litalien, V.; Thompson, E.; Morley, T.; MacRaild, S.; et al. Acquired mutations in BAX confer resistance to BH3-mimetic therapy in acute myeloid leukemia. Blood 2023, 141, 634–644. [Google Scholar] [CrossRef]
  70. Hafezi, S.; Rahmani, M. Targeting BCL-2 in Cancer: Advances, Challenges, and Perspectives. Cancers 2021, 13, 1292. [Google Scholar] [CrossRef]
  71. Campos, L.; Rouault, J.P.; Sabido, O.; Oriol, P.; Roubi, N.; Vasselon, C.; Archimbaud, E.; Magaud, J.P.; Guyotat, D. High expression of bcl-2 protein in acute myeloid leukemia cells is associated with poor response to chemotherapy. Blood 1993, 81, 3091–3096. [Google Scholar] [CrossRef]
  72. Delia, D.; Aiello, A.; Soligo, D.; Fontanella, E.; Melani, C.; Pezzella, F.; Pierotti, M.A.; Della Porta, G. bcl-2 proto-oncogene expression in normal and neoplastic human myeloid cells. Blood 1992, 79, 1291–1298. [Google Scholar] [CrossRef]
  73. Karakas, T.; Maurer, U.; Weidmann, E.; Miething, C.C.; Hoelzer, D.; Bergmann, L. High expression of bcl-2 mRNA as a determinant of poor prognosis in acute myeloid leukemia. Ann. Oncol. 1998, 9, 159–165. [Google Scholar] [CrossRef] [PubMed]
  74. Chong, S.J.F.; Zhu, F.; Dashevsky, O.; Mizuno, R.; Lai, J.X.; Hackett, L.; Ryan, C.E.; Collins, M.C.; Iorgulescu, J.B.; Guièze, R.; et al. Hyperphosphorylation of BCL-2 family proteins underlies functional resistance to venetoclax in lymphoid malignancies. J. Clin. Investig. 2023, 133, e170169. [Google Scholar] [CrossRef] [PubMed]
  75. Carter, B.Z.; Mak, P.Y.; Tao, W.; Zhang, Q.; Ruvolo, V.; Kuruvilla, V.M.; Wang, X.; Mak, D.H.; Battula, V.L.; Konopleva, M.; et al. Maximal Activation of Apoptosis Signaling by Cotargeting Antiapoptotic Proteins in BH3 Mimetic–Resistant AML and AML Stem Cells. Mol. Cancer Ther. 2022, 21, 879–889. [Google Scholar] [CrossRef] [PubMed]
  76. Konopleva, M.; Contractor, R.; Tsao, T.; Samudio, I.; Ruvolo, P.P.; Kitada, S.; Deng, X.; Zhai, D.; Shi, Y.-X.; Sneed, T.; et al. Mechanisms of apoptosis sensitivity and resistance to the BH3 mimetic ABT-737 in acute myeloid leukemia. Cancer Cell 2006, 10, 375–388. [Google Scholar] [CrossRef] [PubMed]
  77. Brunelle, J.K.; Ryan, J.; Yecies, D.; Opferman, J.T.; Letai, A. MCL-1–dependent leukemia cells are more sensitive to chemotherapy than BCL-2–dependent counterparts. J. Cell Biol. 2009, 187, 429–442. [Google Scholar] [CrossRef] [PubMed]
  78. Certo, M.; Del Gaizo Moore, V.; Nishino, M.; Wei, G.; Korsmeyer, S.; Armstrong, S.A.; Letai, A. Mitochondria primed by death signals determine cellular addiction to antiapoptotic BCL-2 family members. Cancer Cell 2006, 9, 351–365. [Google Scholar] [CrossRef] [PubMed]
  79. Bhatt, S.; Pioso, M.S.; Olesinski, E.A.; Yilma, B.; Ryan, J.A.; Mashaka, T.; Leutz, B.; Adamia, S.; Zhu, H.; Kuang, Y.; et al. Reduced Mitochondrial Apoptotic Priming Drives Resistance to BH3 Mimetics in Acute Myeloid Leukemia. Cancer Cell 2020, 38, 872–890.e6. [Google Scholar] [CrossRef]
  80. Kaufmann, S.H.; Karp, J.E.; Svingen, P.A.; Krajewski, S.; Burke, P.J.; Gore, S.D.; Reed, J.C. Elevated expression of the apoptotic regulator Mcl-1 at the time of leukemic relapse. Blood 1998, 91, 991–1000. [Google Scholar] [CrossRef]
  81. Lin, K.H.; Winter, P.S.; Xie, A.; Roth, C.; Martz, C.A.; Stein, E.M.; Anderson, G.R.; Tingley, J.P.; Wood, K.C. Targeting MCL-1/BCL-XL Forestalls the Acquisition of Resistance to ABT-199 in Acute Myeloid Leukemia. Sci. Rep. 2016, 6, 27696. [Google Scholar] [CrossRef]
  82. Kasper, S.; Breitenbuecher, F.; Heidel, F.; Hoffarth, S.; Markova, B.; Schuler, M.; Fischer, T. Targeting MCL-1 sensitizes FLT3-ITD-positive leukemias to cytotoxic therapies. Blood Cancer J. 2012, 2, e60. [Google Scholar] [CrossRef]
  83. Yoshimoto, G.; Miyamoto, T.; Jabbarzadeh-Tabrizi, S.; Iino, T.; Rocnik, J.L.; Kikushige, Y.; Mori, Y.; Shima, T.; Iwasaki, H.; Takenaka, K.; et al. FLT3-ITD up-regulates MCL-1 to promote survival of stem cells in acute myeloid leukemia via FLT3-ITD–specific STAT5 activation. Blood 2009, 114, 5034–5043. [Google Scholar] [CrossRef] [PubMed]
  84. Satta, T.; Li, L.; Chalasani, S.L.; Hu, X.; Nkwocha, J.; Sharma, K.; Kmieciak, M.; Rahmani, M.; Zhou, L.; Grant, S. Dual mTORC1/2 Inhibition Synergistically Enhances AML Cell Death in Combination with the BCL2 Antagonist Venetoclax. Clin. Cancer Res. 2023, 29, 1332–1343. [Google Scholar] [CrossRef] [PubMed]
  85. Bouligny, I.M.; Maher, K.R.; Grant, S. Augmenting Venetoclax Activity Through Signal Transduction in AML. J. Cell. Signal. 2023, 4, 1–12. [Google Scholar] [CrossRef] [PubMed]
  86. Zhang, Q.; Riley-Gillis, B.; Han, L.; Jia, Y.; Lodi, A.; Zhang, H.; Ganesan, S.; Pan, R.; Konoplev, S.N.; Sweeney, S.R.; et al. Activation of RAS/MAPK pathway confers MCL-1 mediated acquired resistance to BCL-2 inhibitor venetoclax in acute myeloid leukemia. Signal Transduct. Target. Ther. 2022, 7, 51. [Google Scholar] [CrossRef] [PubMed]
  87. Cherry, E.M.; Abbott, D.; Amaya, M.; McMahon, C.; Schwartz, M.; Rosser, J.; Sato, A.; Schowinsky, J.T.; Inguva, A.; Minhajuddin, M.; et al. Venetoclax and azacitidine compared with induction chemotherapy for newly diagnosed patients with acute myeloid leukemia. Blood Adv. 2021, 5, 5565–5573. [Google Scholar] [CrossRef]
  88. Kuusanmäki, H.; Leppä, A.-M.; Pölönen, P.; Kontro, M.; Dufva, O.; Deb, D.; Yadav, B.; Brück, O.; Kumar, A.; Everaus, H.; et al. Phenotype-based drug screening reveals association between venetoclax response and differentiation stage in acute myeloid leukemia. Haematologica 2020, 105, 708–720. [Google Scholar] [CrossRef] [PubMed]
  89. White, B.S.; Khan, S.A.; Mason, M.J.; Ammad-Ud-Din, M.; Potdar, S.; Malani, D.; Kuusanmäki, H.; Druker, B.J.; Heckman, C.; Kallioniemi, O.; et al. Bayesian multi-source regression and monocyte-associated gene expression predict BCL-2 inhibitor resistance in acute myeloid leukemia. npj Precis. Oncol. 2021, 5, 71. [Google Scholar] [CrossRef] [PubMed]
  90. Pei, S.; Pollyea, D.A.; Gustafson, A.; Stevens, B.M.; Minhajuddin, M.; Fu, R.; Riemondy, K.A.; Gillen, A.E.; Sheridan, R.M.; Kim, J.; et al. Monocytic Subclones Confer Resistance to Venetoclax-Based Therapy in Patients with Acute Myeloid Leukemia. Cancer Discov. 2020, 10, 536–551. [Google Scholar] [CrossRef]
  91. Kurtz, S.E.; Eide, C.A.; Kaempf, A.; Long, N.; Bottomly, D.; Nikolova, O.; Druker, B.J.; McWeeney, S.K.; Chang, B.H.; Tyner, J.W.; et al. Associating drug sensitivity with differentiation status identifies effective combinations for acute myeloid leukemia. Blood Adv. 2022, 6, 3062–3067. [Google Scholar] [CrossRef]
  92. Kuusanmäki, H.; Dufva, O.; Vähä-Koskela, M.; Leppä, A.-M.; Huuhtanen, J.; Vänttinen, I.; Nygren, P.; Klievink, J.; Bouhlal, J.; Pölönen, P.; et al. Erythroid/megakaryocytic differentiation confers BCL-XL dependency and venetoclax resistance in acute myeloid leukemia. Blood 2023, 141, 1610–1625. [Google Scholar] [CrossRef]
  93. Brown, F.C.; Wei, A.H. Is BCL-xL the Achilles’ heel of AEL and AMKL? Blood 2023, 141, 1505–1506. [Google Scholar] [CrossRef]
  94. Bisaillon, R.; Moison, C.; Thiollier, C.; Krosl, J.; Bordeleau, M.-E.; Lehnertz, B.; Lavallée, V.-P.; MacRae, T.; Mayotte, N.; Labelle, C.; et al. Genetic characterization of ABT-199 sensitivity in human AML. Leukemia 2020, 34, 63–74. [Google Scholar] [CrossRef]
  95. Carter, J.L.; Su, Y.; Qiao, X.; Zhao, J.; Wang, G.; Howard, M.; Edwards, H.; Bao, X.; Li, J.; Hüttemann, M.; et al. Acquired resistance to venetoclax plus azacitidine in acute myeloid leukemia: In vitro models and mechanisms. Biochem. Pharmacol. 2023, 216, 115759. [Google Scholar] [CrossRef] [PubMed]
  96. Miyashita, T.; Krajewski, S.; Krajewska, M.; Wang, H.G.; Lin, H.K.; Liebermann, D.A.; Hoffman, B.; Reed, J.C. Tumor suppressor p53 is a regulator of bcl-2 and bax gene expression in vitro and in vivo. Oncogene 1994, 9, 1799–1805. [Google Scholar] [PubMed]
  97. Romanovsky, E.; Kluck, K.; Ourailidis, I.; Menzel, M.; Beck, S.; Ball, M.; Kazdal, D.; Christopoulos, P.; Schirmacher, P.; Stiewe, T.; et al. Homogenous TP53mut-associated tumor biology across mutation and cancer types revealed by transcriptome analysis. Cell Death Discov. 2023, 9, 126. [Google Scholar] [CrossRef] [PubMed]
  98. Shibue, T.; Takeda, K.; Oda, E.; Tanaka, H.; Murasawa, H.; Takaoka, A.; Morishita, Y.; Akira, S.; Taniguchi, T.; Tanaka, N. Integral role of Noxa in p53-mediated apoptotic response. Minerva Anestesiol. 2003, 17, 2233–2238. [Google Scholar] [CrossRef]
  99. Jin, S.; Cojocari, D.; Purkal, J.J.; Popovic, R.; Talaty, N.N.; Xiao, Y.; Solomon, L.R.; Boghaert, E.R.; Leverson, J.D.; Phillips, D.C. 5-Azacitidine Induces NOXA to Prime AML Cells for Venetoclax-Mediated Apoptosis. Clin. Cancer Res. 2020, 26, 3371–3383. [Google Scholar] [CrossRef]
  100. Thomalla, D.; Beckmann, L.; Grimm, C.; Oliverio, M.; Meder, L.; Herling, C.D.; Nieper, P.; Feldmann, T.; Merkel, O.; Lorsy, E.; et al. Deregulation and epigenetic modification of BCL2-family genes cause resistance to venetoclax in hematologic malignancies. Blood 2022, 140, 2113–2126. [Google Scholar] [CrossRef]
  101. Odinius, T.O.; Buschhorn, L.; Wagner, C.; Hauch, R.T.; Dill, V.; Dechant, M.; Buck, M.C.; Shoumariyeh, K.; Moog, P.; Schwaab, J.; et al. Comprehensive characterization of central BCL-2 family members in aberrant eosinophils and their impact on therapeutic strategies. J. Cancer Res. Clin. Oncol. 2022, 148, 331–340. [Google Scholar] [CrossRef]
  102. Stubbins, R.J.; Karsan, A. Differentiation therapy for myeloid malignancies: Beyond cytotoxicity. Blood Cancer J. 2021, 11, 193. [Google Scholar] [CrossRef]
  103. Tyner, J.W.; Tognon, C.E.; Bottomly, D.; Wilmot, B.; Kurtz, S.E.; Savage, S.L.; Long, N.; Schultz, A.R.; Traer, E.; Abel, M.; et al. Functional genomic landscape of acute myeloid leukaemia. Nat. Cell Biol. 2018, 562, 526–531. [Google Scholar] [CrossRef]
  104. Waclawiczek, A.; Leppä, A.-M.; Renders, S.; Stumpf, K.; Reyneri, C.; Betz, B.; Janssen, M.; Shahswar, R.; Donato, E.; Karpova, D.; et al. Combinatorial BCL2 Family Expression in Acute Myeloid Leukemia Stem Cells Predicts Clinical Response to Azacitidine/Venetoclax. Cancer Discov. 2023, 13, 1408–1427. [Google Scholar] [CrossRef]
  105. Collignon, A.; Hospital, M.A.; Montersino, C.; Courtier, F.; Charbonnier, A.; Saillard, C.; D’incan, E.; Mohty, B.; Guille, A.; Adelaïde, J.; et al. A chemogenomic approach to identify personalized therapy for patients with relapse or refractory acute myeloid leukemia: Results of a prospective feasibility study. Blood Cancer J. 2020, 10, 64. [Google Scholar] [CrossRef]
  106. Garciaz, S.; Guirguis, A.A.; Müller, S.; Brown, F.C.; Chan, Y.-C.; Motazedian, A.; Rowe, C.L.; Kuzich, J.A.; Chan, K.L.; Tran, K.; et al. Pharmacologic Reduction of Mitochondrial Iron Triggers a Noncanonical BAX/BAK-Dependent Cell Death. Cancer Discov. 2021, 12, 774–791. [Google Scholar] [CrossRef]
  107. Nakao, F.; Setoguchi, K.; Semba, Y.; Yamauchi, T.; Nogami, J.; Sasaki, K.; Imanaga, H.; Terasaki, T.; Miyazaki, M.; Hirabayashi, S.; et al. Targeting a mitochondrial E3 ubiquitin ligase complex to overcome AML cell-intrinsic Venetoclax resistance. Leukemia 2023, 37, 1028–1038. [Google Scholar] [CrossRef]
  108. AHuang, A.S.; Chin, H.S.; Reljic, B.; Djajawi, T.M.; Tan, I.K.L.; Gong, J.-N.; Stroud, D.A.; Huang, D.C.S.; van Delft, M.F.; Dewson, G. Mitochondrial E3 ubiquitin ligase MARCHF5 controls BAK apoptotic activity independently of BH3-only proteins. Cell Death Differ. 2022, 30, 632–646. [Google Scholar] [CrossRef]
  109. Jang, J.E.; Hwang, D.Y.; Eom, J.-I.; Cheong, J.-W.; Jeung, H.-K.; Cho, H.; Chung, H.; Kim, J.S.; Min, Y.H. DRP1 Inhibition Enhances Venetoclax-Induced Mitochondrial Apoptosis in TP53-Mutated Acute Myeloid Leukemia Cells through BAX/BAK Activation. Cancers 2023, 15, 745. [Google Scholar] [CrossRef] [PubMed]
  110. Hurrish, K.H.; Qiao, X.; Li, X.; Su, Y.; Carter, J.; Ma, J.; Kalpage, H.A.; Hüttemann, M.; Edwards, H.; Wang, G.; et al. Co-targeting of HDAC, PI3K, and Bcl-2 results in metabolic and transcriptional reprogramming and decreased mitochondrial function in acute myeloid leukemia. Biochem. Pharmacol. 2022, 205, 115283. [Google Scholar] [CrossRef]
  111. Roca-Portoles, A.; Rodriguez-Blanco, G.; Sumpton, D.; Cloix, C.; Mullin, M.; Mackay, G.M.; O’neill, K.; Lemgruber, L.; Luo, X.; Tait, S.W.G. Venetoclax causes metabolic reprogramming independent of BCL-2 inhibition. Cell Death Dis. 2020, 11, 616. [Google Scholar] [CrossRef] [PubMed]
  112. Jones, C.L.; Stevens, B.M.; D’Alessandro, A.; Reisz, J.A.; Culp-Hill, R.; Nemkov, T.; Pei, S.; Khan, N.; Adane, B.; Ye, H.; et al. Inhibition of Amino Acid Metabolism Selectively Targets Human Leukemia Stem Cells. Cancer Cell 2018, 34, 724–740.e4. [Google Scholar] [CrossRef]
  113. Culp-Hill, R.; Stevens, B.M.; Jones, C.L.; Pei, S.; Dzieciatkowska, M.; Minhajuddin, M.; Jordan, C.T.; D’alessandro, A. Therapy-Resistant Acute Myeloid Leukemia Stem Cells Are Resensitized to Venetoclax + Azacitidine by Targeting Fatty Acid Desaturases 1 and 2. Metabolites 2023, 13, 467. [Google Scholar] [CrossRef] [PubMed]
  114. Stevens, B.M.; Jones, C.L.; Pollyea, D.A.; Culp-Hill, R.; D’alessandro, A.; Winters, A.; Krug, A.; Abbott, D.; Goosman, M.; Pei, S.; et al. Fatty acid metabolism underlies venetoclax resistance in acute myeloid leukemia stem cells. Nat. Cancer 2020, 1, 1176–1187. [Google Scholar] [CrossRef] [PubMed]
  115. Hoang, D.H.; Morales, C.; Rodriguez, I.R.; Valerio, M.; Guo, J.; Chen, M.-H.; Wu, X.; Horne, D.; Gandhi, V.; Chen, L.S.; et al. Synergy of Venetoclax and 8-Chloro-Adenosine in AML: The Interplay of rRNA Inhibition and Fatty Acid Metabolism. Cancers 2022, 14, 1446. [Google Scholar] [CrossRef] [PubMed]
  116. Schroeder, B.; Steen, T.V.; Espinoza, I.; Venkatapoorna, C.M.K.; Hu, Z.; Silva, F.M.; Regan, K.; Cuyàs, E.; Meng, X.W.; Verdura, S.; et al. Fatty acid synthase (FASN) regulates the mitochondrial priming of cancer cells. Cell Death Dis. 2021, 12, 977. [Google Scholar] [CrossRef] [PubMed]
  117. Jones, C.L.; Stevens, B.M.; Pollyea, D.A.; Culp-Hill, R.; Reisz, J.A.; Nemkov, T.; Gehrke, S.; Gamboni, F.; Krug, A.; Winters, A.; et al. Nicotinamide Metabolism Mediates Resistance to Venetoclax in Relapsed Acute Myeloid Leukemia Stem Cells. Cell Stem Cell 2020, 27, 748–764.e4. [Google Scholar] [CrossRef] [PubMed]
  118. Khokhlatchev, A.V.; Sharma, A.; Deering, T.G.; Shaw, J.J.P.; Costa-Pinheiro, P.; Golla, U.; Annageldiyev, C.; Cabot, M.C.; Conaway, M.R.; Tan, S.; et al. Ceramide nanoliposomes augment the efficacy of venetoclax and cytarabine in models of acute myeloid leukemia. FASEB J. 2022, 36, e22514. [Google Scholar] [CrossRef] [PubMed]
  119. Lewis, A.C.; Pope, V.S.; Tea, M.N.; Li, M.; Nwosu, G.O.; Nguyen, T.M.; Wallington-Beddoe, C.T.; Moretti, P.A.B.; Anderson, D.; Creek, D.J.; et al. Ceramide-induced integrated stress response overcomes Bcl-2 inhibitor resistance in acute myeloid leukemia. Blood 2022, 139, 3737–3751. [Google Scholar] [CrossRef] [PubMed]
  120. Cojocari, D.; Smith, B.N.; Purkal, J.J.; Arrate, M.P.; Huska, J.D.; Xiao, Y.; Gorska, A.; Hogdal, L.J.; Ramsey, H.E.; Boghaert, E.R.; et al. Pevonedistat and azacitidine upregulate NOXA (PMAIP1) to increase sensitivity to venetoclax in preclinical models of acute myeloid leukemia. Haematologica 2022, 107, 825–835. [Google Scholar] [CrossRef]
  121. Lee, J.B.; Khan, D.H.; Hurren, R.; Xu, M.; Na, Y.; Kang, H.; Mirali, S.; Wang, X.; Gronda, M.V.; Jitkova, Y.; et al. Venetoclax enhances T cell-mediated anti-leukemic activity by increasing ROS production. Blood 2021, 138, 234–245. [Google Scholar] [CrossRef]
  122. Carter, B.Z.; Mak, P.Y.; Tao, W.; Warmoes, M.; Lorenzi, P.L.; Mak, D.; Ruvolo, V.; Tan, L.; Cidado, J.; Drew, L.; et al. Targeting MCL-1 dysregulates cell metabolism and leukemia-stroma interactions and re-sensitizes acute myeloid leukemia to BCL-2 inhibition. Haematologica 2022, 107, 58–76. [Google Scholar] [CrossRef] [PubMed]
  123. Pakos-Zebrucka, K.; Koryga, I.; Mnich, K.; Ljujic, M.; Samali, A.; Gorman, A.M. The integrated stress response. EMBO Rep. 2016, 17, 1374–1395. [Google Scholar] [CrossRef] [PubMed]
  124. Levitz, D.; Saunthararajah, Y.; Fedorov, K.; Shapiro, L.C.; Mantzaris, I.; Shastri, A.; Kornblum, N.; Sica, R.A.; Shah, N.; Konopleva, M.; et al. A Metabolically Optimized, Noncytotoxic Low-Dose Weekly Decitabine/Venetoclax in MDS and AML. Clin. Cancer Res. 2023, 29, 2774–2780. [Google Scholar] [CrossRef] [PubMed]
  125. Garcia-Manero, G.; Griffiths, E.A.; Steensma, D.P.; Roboz, G.J.; Wells, R.; McCloskey, J.; Odenike, O.; DeZern, A.E.; Yee, K.; Busque, L.; et al. Oral cedazuridine/decitabine for MDS and CMML: A phase 2 pharmacokinetic/pharmacodynamic randomized crossover study. Blood 2020, 136, 674–683. [Google Scholar] [CrossRef]
  126. Ravandi, F.; Abuasab, T.; Valero, Y.A.; Issa, G.C.; Islam, R.; Short, N.J.; Yilmaz, M.; Jain, N.; Masarova, L.; Kornblau, S.M.; et al. Phase 2 study of ASTX727 (cedazuridine/decitabine) plus venetoclax (ven) in patients with relapsed/refractory acute myeloid leukemia (AML) or previously untreated, elderly patients (pts) unfit for chemotherapy. J. Clin. Oncol. 2022, 40, 7037. [Google Scholar] [CrossRef]
  127. Wei, A.H.; Döhner, H.; Pocock, C.; Montesinos, P.; Afanasyev, B.; Dombret, H.; Ravandi, F.; Sayar, H.; Jang, J.-H.; Porkka, K.; et al. Oral Azacitidine Maintenance Therapy for Acute Myeloid Leukemia in First Remission. N. Engl. J. Med. 2020, 383, 2526–2537. [Google Scholar] [CrossRef] [PubMed]
  128. Carter, B.Z.; Mak, P.Y.; Tao, W.; Ayoub, E.; Ostermann, L.B.; Huang, X.; Loghavi, S.; Boettcher, S.; Nishida, Y.; Ruvolo, V.; et al. Combined inhibition of BCL-2 and MCL-1 overcomes BAX deficiency-mediated resistance of TP53-mutant acute myeloid leukemia to individual BH3 mimetics. Blood Cancer J. 2023, 13, 57. [Google Scholar] [CrossRef]
  129. Teh, T.-C.; Nguyen, N.-Y.; Moujalled, D.M.; Segal, D.; Pomilio, G.; Rijal, S.; Jabbour, A.; Cummins, K.; Lackovic, K.; Blombery, P.; et al. Enhancing venetoclax activity in acute myeloid leukemia by co-targeting MCL1. Leukemia 2018, 32, 303–312. [Google Scholar] [CrossRef]
  130. Saxena, K.; Carter, B.Z.; Konopleva, M. EXABS-147-AML How Do We Overcome Resistance to Venetoclax. Clin. Lymphoma Myeloma Leuk. 2022, 22, S55–S57. [Google Scholar] [CrossRef]
  131. Bolomsky, A.; Vogler, M.; Köse, M.C.; Heckman, C.A.; Ehx, G.; Ludwig, H.; Caers, J. MCL-1 inhibitors, fast-lane development of a new class of anti-cancer agents. J. Hematol. Oncol. 2020, 13, 173. [Google Scholar] [CrossRef]
  132. Forsberg, M.; Konopleva, M. SOHO State of the Art Updates and Next Questions: Understanding and Overcoming Venetoclax Resistance in Hematologic Malignancies. Clin. Lymphoma Myeloma Leuk. 2024, 24, 1–14. [Google Scholar] [CrossRef]
  133. Wei, A.H.; Roberts, A.W.; Spencer, A.; Rosenberg, A.S.; Siegel, D.; Walter, R.B.; Caenepeel, S.; Hughes, P.; McIver, Z.; Mezzi, K.; et al. Targeting MCL-1 in hematologic malignancies: Rationale and progress. Blood Rev. 2020, 44, 100672. [Google Scholar] [CrossRef] [PubMed]
  134. Balachander, S.B.; Criscione, S.W.; Byth, K.F.; Cidado, J.; Adam, A.; Lewis, P.; Macintyre, T.; Wen, S.; Lawson, D.; Burke, K.; et al. AZD4320, A Dual Inhibitor of Bcl-2 and Bcl-xL, Induces Tumor Regression in Hematologic Cancer Models without Dose-limiting Thrombocytopenia. Clin. Cancer Res. 2020, 26, 6535–6549. [Google Scholar] [CrossRef] [PubMed]
  135. Lv, D.; Pal, P.; Liu, X.; Jia, Y.; Thummuri, D.; Zhang, P.; Hu, W.; Pei, J.; Zhang, Q.; Zhou, S.; et al. Development of a BCL-xL and BCL-2 dual degrader with improved anti-leukemic activity. Nat. Commun. 2021, 12, 6896. [Google Scholar] [CrossRef]
  136. Patel, H.; Periyasamy, M.; Sava, G.P.; Bondke, A.; Slafer, B.W.; Kroll, S.H.B.; Barbazanges, M.; Starkey, R.; Ottaviani, S.; Harrod, A.; et al. ICEC0942, an Orally Bioavailable Selective Inhibitor of CDK7 for Cancer Treatment. Mol. Cancer Ther. 2018, 17, 1156–1166. [Google Scholar] [CrossRef]
  137. Cidado, J.; Boiko, S.; Proia, T.; Ferguson, D.; Criscione, S.W.; Martin, M.S.; Pop-Damkov, P.; Su, N.; Franklin, V.N.R.; Chilamakuri, C.S.R.; et al. AZD4573 Is a Highly Selective CDK9 Inhibitor That Suppresses MCL-1 and Induces Apoptosis in Hematologic Cancer Cells. Clin. Cancer Res. 2020, 26, 922–934. [Google Scholar] [CrossRef]
  138. He, L.; Arnold, C.; Thoma, J.; Rohde, C.; Kholmatov, M.; Garg, S.; Hsiao, C.; Viol, L.; Zhang, K.; Sun, R.; et al. CDK7/12/13 inhibition targets an oscillating leukemia stem cell network and synergizes with venetoclax in acute myeloid leukemia. EMBO Mol. Med. 2022, 14, e14990. [Google Scholar] [CrossRef] [PubMed]
  139. Seipel, K.; Graber, C.; Flückiger, L.; Bacher, U.; Pabst, T. Rationale for a Combination Therapy with the STAT5 Inhibitor AC-4-130 and the MCL1 Inhibitor S63845 in the Treatment of FLT3-Mutated or TET2-Mutated Acute Myeloid Leukemia. Int. J. Mol. Sci. 2021, 22, 8092. [Google Scholar] [CrossRef]
  140. Ma, J.; Zhao, S.; Qiao, X.; Knight, T.; Edwards, H.; Polin, L.; Kushner, J.; Dzinic, S.H.; White, K.; Wang, G.; et al. Inhibition of Bcl-2 Synergistically Enhances the Antileukemic Activity of Midostaurin and Gilteritinib in Preclinical Models of FLT3-Mutated Acute Myeloid Leukemia. Clin. Cancer Res. 2019, 25, 6815–6826. [Google Scholar] [CrossRef]
  141. Jalte, M.; Abbassi, M.; Belghiti, H.D.; Ahakoud, M.; Bekkari, H.; EL Mouhi, H. FLT3 Mutations in Acute Myeloid Leukemia: Unraveling the Molecular Mechanisms and Implications for Targeted Therapies. Cureus 2023, 15, e45765. [Google Scholar] [CrossRef]
  142. Janssen, M.; Schmidt, C.; Bruch, P.-M.; Blank, M.F.; Rohde, C.; Waclawiczek, A.; Heid, D.; Renders, S.; Göllner, S.; Vierbaum, L.; et al. Venetoclax synergizes with gilteritinib in FLT3 wild-type high-risk acute myeloid leukemia by suppressing MCL-1. Blood 2022, 140, 2594–2610. [Google Scholar] [CrossRef] [PubMed]
  143. Daver, N.; Perl, A.E.; Maly, J.; Levis, M.; Ritchie, E.; Litzow, M.; McCloskey, J.; Smith, C.C.; Schiller, G.; Bradley, T.; et al. Venetoclax Plus Gilteritinib for FLT3-Mutated Relapsed/Refractory Acute Myeloid Leukemia. J. Clin. Oncol. 2022, 40, 4048–4059. [Google Scholar] [CrossRef]
  144. Short, N.; Macaron, W.; Dinardo, C.; Daver, N.; Yilmaz, M.; Borthakur, G.; Montalban-Bravo, G.; Garcia-Manero, G.; Issa, G.; Sasaki, K.; et al. P485: Azacitidine, venetoclax and gilteritinib for patients with newly diagnosed flt3-mutated acute myeloid leukemia: A subgroup analysis from a phase II study. Hemasphere 2023, 7, e535260f. [Google Scholar] [CrossRef]
  145. Maiti, A.; DiNardo, C.D.; Daver, N.G.; Rausch, C.R.; Ravandi, F.; Kadia, T.M.; Pemmaraju, N.; Borthakur, G.; Bose, P.; Issa, G.C.; et al. Triplet therapy with venetoclax, FLT3 inhibitor and decitabine for FLT3-mutated acute myeloid leukemia. Blood Cancer J. 2021, 11, 25. [Google Scholar] [CrossRef] [PubMed]
  146. Yilmaz, M.; Kantarjian, H.; Short, N.J.; Reville, P.; Konopleva, M.; Kadia, T.; DiNardo, C.; Borthakur, G.; Pemmaraju, N.; Maiti, A.; et al. Hypomethylating agent and venetoclax with FLT3 inhibitor “triplet” therapy in older/unfit patients with FLT3 mutated AML. Blood Cancer J. 2022, 12, 77. [Google Scholar] [CrossRef]
  147. Yilmaz, M.; Muftuoglu, M.; Kantarjian, H.M.; Dinardo, C.D.; Kadia, T.M.; Konopleva, M.; Borthakur, G.; Pemmaraju, N.; Short, N.J.; Valero, Y.A.; et al. Quizartinib (QUIZ) with decitabine (DAC) and venetoclax (VEN) is active in patients (pts) with FLT3-ITD mutated acute myeloid leukemia (AML): A phase I/II clinical trial. J. Clin. Oncol. 2022, 40, 7036. [Google Scholar] [CrossRef]
  148. Montesinos, P.; Recher, C.; Vives, S.; Zarzycka, E.; Wang, J.; Bertani, G.; Heuser, M.; Calado, R.T.; Schuh, A.C.; Yeh, S.-P.; et al. Ivosidenib and Azacitidine inIDH1-Mutated Acute Myeloid Leukemia. N. Engl. J. Med. 2022, 386, 1519–1531. [Google Scholar] [CrossRef]
  149. Lachowiez, C.A.; Loghavi, S.; Zeng, Z.; Tanaka, T.; Kim, Y.J.; Uryu, H.; Turkalj, S.; Jakobsen, N.A.; Luskin, M.R.; Duose, D.Y.; et al. A Phase Ib/II Study of Ivosidenib with Venetoclax ± Azacitidine in IDH1-Mutated Myeloid Malignancies. Blood Cancer Discov. 2023, 4, 276–293. [Google Scholar] [CrossRef]
  150. Venugopal, S.; Takahashi, K.; Daver, N.; Maiti, A.; Borthakur, G.; Loghavi, S.; Short, N.J.; Ohanian, M.; Masarova, L.; Issa, G.; et al. Efficacy and safety of enasidenib and azacitidine combination in patients with IDH2 mutated acute myeloid leukemia and not eligible for intensive chemotherapy. Blood Cancer J. 2022, 12, 10. [Google Scholar] [CrossRef]
  151. Konopleva, M.; Martinelli, G.; Daver, N.; Papayannidis, C.; Wei, A.; Higgins, B.; Ott, M.; Mascarenhas, J.; Andreeff, M. MDM2 inhibition: An important step forward in cancer therapy. Leukemia 2020, 34, 2858–2874. [Google Scholar] [CrossRef]
  152. Daver, N.G.; Dail, M.; Garcia, J.S.; Jonas, B.A.; Yee, K.W.L.; Kelly, K.R.; Vey, N.; Assouline, S.; Roboz, G.J.; Paolini, S.; et al. Venetoclax and idasanutlin in relapsed/refractory AML: A nonrandomized, open-label phase 1b trial. Blood 2023, 141, 1265–1276. [Google Scholar] [CrossRef]
  153. Senapati, J.; Muftuoglu, M.; Ishizawa, J.; Abbas, H.A.; Loghavi, S.; Borthakur, G.; Yilmaz, M.; Issa, G.C.; Dara, S.I.; Basyal, M.; et al. A Phase I study of Milademetan (DS3032b) in combination with low dose cytarabine with or without venetoclax in acute myeloid leukemia: Clinical safety, efficacy, and correlative analysis. Blood Cancer J. 2023, 13, 101. [Google Scholar] [CrossRef] [PubMed]
  154. Garcia-Manero, G.; Goldberg, A.D.; Winer, E.S.; Altman, J.K.; Fathi, A.T.; Odenike, O.; Roboz, G.J.; Sweet, K.; Miller, C.; Wennborg, A.; et al. Eprenetapopt combined with venetoclax and azacitidine in TP53-mutated acute myeloid leukaemia: A phase 1, dose-finding and expansion study. Lancet Haematol. 2023, 10, e272–e283. [Google Scholar] [CrossRef] [PubMed]
  155. Issa, G.C.; Aldoss, I.; DiPersio, J.; Cuglievan, B.; Stone, R.; Arellano, M.; Thirman, M.J.; Patel, M.R.; Dickens, D.S.; Shenoy, S.; et al. The menin inhibitor revumenib in KMT2A-rearranged or NPM1-mutant leukaemia. Nature 2023, 615, 920–924. [Google Scholar] [CrossRef] [PubMed]
  156. Swaminathan, M.; Bourgeois, W.; Armstrong, S.A.; Wang, E.S. Menin Inhibitors in Acute Myeloid Leukemia—What Does the Future Hold? Cancer J. 2022, 28, 62–66. [Google Scholar] [CrossRef]
  157. Jabbour, E.; Searle, E.; Abdul-Hay, M.; Abedin, S.; Aldoss, I.; Piérola, A.A.; Alonso-Dominguez, J.M.; Chevallier, P.; Cost, C.; Daskalakis, N.; et al. A First-in-Human Phase 1 Study of the Menin-KMT2A (MLL1) Inhibitor JNJ-75276617 in Adult Patients with Relapsed/Refractory Acute Leukemia Harboring KMT2A or NPM1 Alterations. Blood 2023, 142, 57. [Google Scholar] [CrossRef]
  158. Gilead Announces Partial Clinical Hold for Magrolimab Studies in AML. Available online: https://www.gilead.com/news-and-press/press-room/press-releases/2023/8/gilead-announces-partial-clinical-hold-for-magrolimab-studies-in-aml (accessed on 28 January 2024).
  159. Daver, N.; Konopleva, M.; Maiti, A.; Kadia, T.M.; DiNardo, C.D.; Loghavi, S.; Pemmaraju, N.; Jabbour, E.J.; Montalban-Bravo, G.; Tang, G.; et al. Phase I/II Study of Azacitidine (AZA) with Venetoclax (VEN) and Magrolimab (Magro) in Patients (pts) with Newly Diagnosed Older/Unfit or High-Risk Acute Myeloid Leukemia (AML) and Relapsed/Refractory (R/R) AML. Blood 2021, 138, 371. [Google Scholar] [CrossRef]
  160. Daver, N. Pivekimab Sunirine (PVEK, IMGN632), a CD123-Targeting Antibody-Drug Conjugate, in Combination with Azacitidine and Venetoclax in Patients with Newly Diagnosed Acute Myeloid Leukemia. In Proceedings of the 65th ASH Annual Meeting & Exposition, ASH, San Diego, CA, USA, 9–12 December 2023; Available online: https://ash.confex.com/ash/2023/webprogram/Paper173413.html (accessed on 28 January 2024).
  161. Ivey, A.; Hills, R.K.; Simpson, M.A.; Jovanovic, J.V.; Gilkes, A.; Grech, A.; Patel, Y.; Bhudia, N.; Farah, H.; Mason, J.; et al. Assessment of Minimal Residual Disease in Standard-Risk AML. N. Engl. J. Med. 2016, 374, 422–433. [Google Scholar] [CrossRef]
  162. Heuser, M.; Freeman, S.D.; Ossenkoppele, G.J.; Buccisano, F.; Hourigan, C.S.; Ngai, L.L.; Tettero, J.M.; Bachas, C.; Baer, C.; Béné, M.-C.; et al. 2021 Update on MRD in acute myeloid leukemia: A consensus document from the European LeukemiaNet MRD Working Party. Blood 2021, 138, 2753–2767. [Google Scholar] [CrossRef]
  163. Winters, A.C.; Minhajuddin, M.; Stevens, B.M.; Major, A.; Bosma, G.; Abbott, D.; Miltgen, N.; Yuan, J.; Treece, A.L.; Siegele, B.J.; et al. Multi-gene measurable residual disease assessed by digital polymerase chain reaction has clinical and biological utility in acute myeloid leukemia patients receiving venetoclax/azacitidine. Haematologica 2023. [Google Scholar] [CrossRef]
  164. Maiti, A.; DiNardo, C.D.; Wang, S.A.; Jorgensen, J.; Kadia, T.M.; Daver, N.G.; Short, N.J.; Yilmaz, M.; Pemmaraju, N.; Borthakur, G.; et al. Prognostic value of measurable residual disease after venetoclax and decitabine in acute myeloid leukemia. Blood Adv. 2021, 5, 1876–1883. [Google Scholar] [CrossRef] [PubMed]
  165. Othman, J.; Tiong, I.S.; O’Nions, J.; Dennis, M.; Mokretar, K.; Ivey, A.; Austin, M.J.; Latif, A.-L.; Amer, M.; Chan, W.Y.; et al. Molecular MRD is strongly prognostic in patients with NPM1-mutated AML receiving venetoclax-based nonintensive therapy. Blood 2023, 143, 336–341. [Google Scholar] [CrossRef] [PubMed]
  166. Single-Cell Multi-Omics|Mission Bio. Available online: https://missionbio.com/capabilities/dna-protein/?gad_source=1&gclid=CjwKCAiArfauBhApEiwAeoB7qKxJJBI0pFtA47lwoW5BNyLcHWRO6uvfse-fZ_60Tdbf8DUtfQN5QhoCOXEQAvD_BwE (accessed on 27 February 2024).
  167. Garciaz, S.; Bertoli, S.; Sallman, D.A.; Decroocq, J.; Dumas, P.-Y.; Belhabri, A.; Orvain, C.; Requena, G.A.; Simand, C.; Laribi, K.; et al. Acute Myeloid Leukemia Patients Who Stopped Venetoclax or/and Azacytidine for Other Reasons Than Progression Have a Prolonged Treatment Free Remission and Overall Survival. a Filo Study. Blood 2023, 142, 161. [Google Scholar] [CrossRef]
  168. Chua, C.C.; Hammond, D.; Kent, A.; Tiong, I.S.; Konopleva, M.Y.; Pollyea, D.A.; DiNardo, C.D.; Wei, A.H. Treatment-free remission after ceasing venetoclax-based therapy in patients with acute myeloid leukemia. Blood Adv. 2022, 6, 3879–3883. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Mechanism of action of venetoclax. Venetoclax acts as a protein/protein interaction inhibitor that specifically binds the anti-apoptotic protein BCL2 and releases the pro-apoptotic proteins including BAX and BAK that induce mitochondrial outer membrane permeabilization (MOMP) and cell death by intrinsic apoptosis.
Figure 1. Mechanism of action of venetoclax. Venetoclax acts as a protein/protein interaction inhibitor that specifically binds the anti-apoptotic protein BCL2 and releases the pro-apoptotic proteins including BAX and BAK that induce mitochondrial outer membrane permeabilization (MOMP) and cell death by intrinsic apoptosis.
Cancers 16 01091 g001
Figure 2. Main genetic alterations influencing VEN resistance. Intrinsic apoptosis relies on the balance between pro- and anti-apoptotic proteins that will induce mitochondrial outer membrane permeabilization (MOMP). In response to apoptotic stimuli, BAX and BAK form pores in the mitochondrial membrane, inducing mitochondrial outer membrane permeabilization (MOMP). This phenomenon is increased by the members of the BCL2 protein family containing a single BH3 domain named “BH3-only proteins”, which have a pro-apoptotic role (BIM, NOXA, PUMA, BID). Conversely, MOMP is blocked by a series of proteins that have an anti-apoptotic role including BCL2, BCL-XL, and MCL1. IDH mutations are associated with a higher sensitivity to VEN mostly by their role in mitochondrial metabolism and the tricarboxylic acid cycle (TCA). TP53 mutations confer resistance to VEN by transcriptionally decreasing pro-apoptotic proteins such as NOXA or PUMA. FLT3 mutations are associated with VEN resistance because of increased signalization unbalancing BCL2 and MCL1 dependencies. BAX mutations disrupt the COOH-terminal α9 helix necessary to BAX translocation from the cytosol into the mitochondrial outer membrane, decreasing MOMP.
Figure 2. Main genetic alterations influencing VEN resistance. Intrinsic apoptosis relies on the balance between pro- and anti-apoptotic proteins that will induce mitochondrial outer membrane permeabilization (MOMP). In response to apoptotic stimuli, BAX and BAK form pores in the mitochondrial membrane, inducing mitochondrial outer membrane permeabilization (MOMP). This phenomenon is increased by the members of the BCL2 protein family containing a single BH3 domain named “BH3-only proteins”, which have a pro-apoptotic role (BIM, NOXA, PUMA, BID). Conversely, MOMP is blocked by a series of proteins that have an anti-apoptotic role including BCL2, BCL-XL, and MCL1. IDH mutations are associated with a higher sensitivity to VEN mostly by their role in mitochondrial metabolism and the tricarboxylic acid cycle (TCA). TP53 mutations confer resistance to VEN by transcriptionally decreasing pro-apoptotic proteins such as NOXA or PUMA. FLT3 mutations are associated with VEN resistance because of increased signalization unbalancing BCL2 and MCL1 dependencies. BAX mutations disrupt the COOH-terminal α9 helix necessary to BAX translocation from the cytosol into the mitochondrial outer membrane, decreasing MOMP.
Cancers 16 01091 g002
Figure 3. Main classes of drug synergizing with venetoclax in clinical studies. Hypomethylating agents such as azacitidine have been shown to induce an integrated stress response, upregulating the transcription of pro-apoptotic proteins NOXA and PUMA, which increases the efficacy of BCL2 inhibition. IDH inhibitors act by blocking the formation of the oncometabolite D-2-HG in IDH-mutated AML. IDH-mutated AML has been shown to be more susceptible to intrinsic apoptosis due to defects in the TCA cycle. FLT3 mutations induce MCL1 protein upregulation through the RAS/AKT signalization cascade. FLT3 inhibitors are deemed to cause an indirect MCL1 inhibition by degrading MCL1.
Figure 3. Main classes of drug synergizing with venetoclax in clinical studies. Hypomethylating agents such as azacitidine have been shown to induce an integrated stress response, upregulating the transcription of pro-apoptotic proteins NOXA and PUMA, which increases the efficacy of BCL2 inhibition. IDH inhibitors act by blocking the formation of the oncometabolite D-2-HG in IDH-mutated AML. IDH-mutated AML has been shown to be more susceptible to intrinsic apoptosis due to defects in the TCA cycle. FLT3 mutations induce MCL1 protein upregulation through the RAS/AKT signalization cascade. FLT3 inhibitors are deemed to cause an indirect MCL1 inhibition by degrading MCL1.
Cancers 16 01091 g003
Table 1. Main studies evaluating VEN-HMA in the context of newly diagnosed AML patients ineligible for intensive chemotherapy.
Table 1. Main studies evaluating VEN-HMA in the context of newly diagnosed AML patients ineligible for intensive chemotherapy.
Phase; Name; IDStudy PopulationVenetoclax DosesOther Agent Administration RegimensResponse RateSurvival Rate
I; (M14-358);
NCT02203773
ND AML
ineligible for chemotherapy
400–800–1200 mg
D1 to D28
AZA 75 mg/m2, D1 to D7
or DEC 20 mg/m2, D1 to D5
37% (CR),
68% (ORR), 83% (cCR)
17.5 months (median OS)
II;
NCT03404193
ND patients with AML > 60 years400 mg D1 to D28
(D1 to D21 if blasts < 5%)
DEC 20 mg/m2, D1 to D10 (induction)
DEC 20 mg/m2, D1 to D5 (consolidation)
63% (CR + CRi), 74% (ORR)18.1 months (median OS)
III; (VIALE-A);
NCT02993523
ND AML ineligible for chemotherapy400 mg D1 to D28 vs. placeboAZA 75 mg/m2, D1 to D736.7% vs. 17.9% (CR),
64.7% vs. 22.8% (CR + CRi)
14.7 months vs. 9.6 months
(median OS)
Abbreviations: AZA, azacitidine, cCR, composite complete response; CR, complete response; CRi, CR with incomplete hematological recovery, DEC, decitabine; MLFS, morphologic leukemia free state; ND, newly diagnosed; ORR, overall response rate; OS, overall survival.
Table 2. Main studies evaluating VEN-AZA-based triplet therapies.
Table 2. Main studies evaluating VEN-AZA-based triplet therapies.
TreatmentStudy IDPhase Number of PatientsResponse Treatment
Quizartinib
(FLT3 inhibitor)
NCT03661307I/II28 (23 R/R AML FLT3mut)78% achieved cCR (3 CR, 15 CRi) [147]
Ivosidenib
(IDH1 inhibitor)
NCT03471260Ib31 (all R/R, IDHmut)94% had a CR + CRh + CRi + PR + MLFS [149]
Magrolimab
(anti CD47)
NCT04435691I/II60 (41 ND, 29)ORR [CR, CRi, MLFS] Frontline = 80%[159]
Pivekimab sunirine (PVEK, anti CD123)NCT04086264I/II50 ND patients CR rate was 52% (26/50), and cCR rate was 66% (33/50)[160]
Abbreviations: cCR, composite complete response; CR, complete response; CRi, CR with incomplete hematological recovery, MLFS, morphologic leukemia free state; ND, newly diagnosed; ORR, overall response rate, PR, partial response.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Garciaz, S.; Hospital, M.-A.; Collette, Y.; Vey, N. Venetoclax Resistance in Acute Myeloid Leukemia. Cancers 2024, 16, 1091. https://doi.org/10.3390/cancers16061091

AMA Style

Garciaz S, Hospital M-A, Collette Y, Vey N. Venetoclax Resistance in Acute Myeloid Leukemia. Cancers. 2024; 16(6):1091. https://doi.org/10.3390/cancers16061091

Chicago/Turabian Style

Garciaz, Sylvain, Marie-Anne Hospital, Yves Collette, and Norbert Vey. 2024. "Venetoclax Resistance in Acute Myeloid Leukemia" Cancers 16, no. 6: 1091. https://doi.org/10.3390/cancers16061091

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