Skip to main content

p53 biology and reactivation for improved therapy in MDS and AML

Abstract

Myelodysplastic syndrome (MDS) and acute myeloid leukemia (AML) originate from preleukemic hematopoietic conditions, such as clonal hematopoiesis of indeterminate potential (CHIP) or clonal cytopenia of undetermined significance (CCUS) and have variable outcomes despite the successful implementation of targeted therapies. The prognosis differs depending on the molecular subgroup. In patients with TP53 mutations, the most inferior outcomes across independent studies were observed. Myeloid malignancies with TP53 mutations have complex cytogenetics and extensive structural variants. These factors contribute to worse responses to induction therapy, demethylating agents, or venetoclax-based treatments. Survival of patients with biallelic TP53 gene mutations is often less than one year but this depends on the type of treatment applied. It is still controversial whether the allelic state of mutant TP53 impacts the outcomes in patients with AML and high-risk MDS. Further studies are needed to justify estimating TP53 LOH status for better risk assessment. Yet, TP53-mutated MDS, MDS/AML and AML are now classified separately in the International Consensus Classification (ICC). In the clinical setting, the wild-type p53 protein is reactivated pharmacologically by targeting p53/MDM2/MDM4 interactions and mutant p53 reactivation is achieved by refolding the DNA binding domain to wild-type-like conformation or via targeted degradation of the mutated protein. This review discusses our current understanding of p53 biology in MDS and AML and the promises and failures of wild-type and mutant p53 reactivation in the clinical trial setting.

Introduction

Myelodysplastic syndromes (MDS) and acute myeloid leukemia (AML) are cognate, clonal hematological neoplasms and originate from pre-malignant, mutated hematopoietic stem cells (HSCs) that undergo clonal expansion after selection pressure, in a process called clonal hematopoiesis (CH) [1]. Extensive studies showed HSCs are constricted to the lineage − CD34 + CD38 − CD90 + CD45RA − compartment and bear driver mutations in CH [2]. CH results in an accumulation of large numbers of abnormal, immature myeloid cells in the bone marrow and peripheral blood called leukemic stem cells (Fig. 1a). Clonal hematopoiesis often occurs due to aging and is associated with a higher risk of hematological cancers. The rate of CH progression to hematologic neoplasm is 0.5%—1% per year [3, 4].

Fig. 1
figure 1

Origin of hematological neoplasms from clonal hematopoiesis (a) and TP53-mutated clonal hematopoiesis (b). With age, hematopoietic stem cells (HSCs) acquire somatic mutations and the mutated clones are expanded during HSCs renewal. Early genetic events lead to clonal hematopoiesis (CH) and the origin of clonal hematopoiesis of indeterminate potential (CHIP) or clonal cytopenias of undetermined significance (CCUS), which have different genetic backgrounds, differ in cytopenia status but possess same variant allele frequencies of VAF ≥ 2%. Among these, CCUS have a higher potential of transformation to myelodysplastic syndrome (MDS) with a range of incidence 18–95%. Fitness variants occurring later, confer a growth advantage in the presence of selective pressures such as inflammation, cytotoxic treatment, radiotherapy, bone marrow microenvironment or aging. The consequent selection of high-risk variants, accumulation of the co-occurring mutations, increase in the VAF > 10% and the co-existing cytopenia are predictive of hematopoietic malignancy development and thus, accelerate the progression to hematological neoplasm; myelodysplastic syndrome (MDS) and MDS/AML. TP53-mutated CH occurs in about 2–6% of cancer patients. TP53-mutated MDS and AML account for up to 13% of all de novo cases. In therapy-related myeloid neoplasm, TP53 gene mutations are present in 20–40% of cases [5]. The model shows the pathogenesis of MDS and AML in hereditary cancer syndromes; Li-Fraumeni syndrome (LFS) with congenital TP53 mutations and Schwachman syndrome (SDS) with congenital mutations in Schwachman–Bodian–Diamond syndrome (SBDS) gene and in therapy-related myeloid neoplasm (t-MN); diseases in which TP53 monoallelic mutations play a role in the progression from CH to myeloid malignancy. The incidence of malignant transformation is higher in the presence of selective pressure as delineated for SDS and t-MN. →  →  → tandem of arrows indicates a multi-step process. Modified from [1, 6, 7]. Created with BioRender.com

During their lifespan, hematopoietic stem cells (HSCs) experience functional decline due to accumulated mutations resulting from increased DNA damage or epigenetic reprogramming. They reside in the bone marrow as a genetically heterogeneous cell population [8]. Some HSCs that acquire somatic mutations in genetic modulators (DNMT3A, TET2, ASXL1) and in signaling molecules (JAK2V617F) gain a competitive fitness advantage in the presence of selective pressure and expand resulting in clonal hematopoiesis [9, 10].

MDS and AML originate during CH from clonal hematopoiesis of indeterminate potential (CHIP) or clonal cytopenia of undetermined significance (CCUS), which falls into the category of clonal cytopenia (Fig. 1a). CCUS can be distinguished from CHIP thanks to the advancements in NGS techniques. It represents a continuum with MDS to which it progresses faster than CHIP after acquiring additional mutations and dysplasia [6, 11]. In both, CHIP and CCUS the increased risk of progression to MDS and de novo AML occurs upon > 1 additional driver mutation, VAF > 10% and acquisition of additional mutations [12]. For patients with CCUS, the risk of progression to MDS/AML was reported to be 18% within 16 months and 95% in 10 years [3].

Many MDS and AML subtypes share common driver alterations which might occur at different frequencies but target the same pathways; for example DNA methylation (TET2, DNMT3A, and IDH1/IDH2), chromatin/histone modification (MLL2, EZH2 and ASXL1), RNA splicing (SF3B1, SRSF2, U2AF1, U2AF2, and SF3A1) or have mutations in TP53 gene, in RAS gene or in other signaling pathways [13, 14].

Somatic mutations in TP53 are considered early events in leukemogenesis and blood stem cells with TP53 mutations constitute a preleukemic niche. The pre-leukemic stem cells (pre-LSCs) with TP53 mutations retain the ability to differentiate which highlights the role of TP53 mutations in leukemogenesis [15] as described in more detail below.

Studies showed that in patients with therapy-related AML or secondary AML, the analysis of the material from antecedent hematological disorder displayed mutations in TP53 which manifests the role of low levels of the TP53 mutated clone in the disease origin, before the cytotoxic treatments for the primary malignancy [16]. The material from precedent disorder and/or CH or CCUS is not always available for MDS and AML patients, and early genetic alterations, like somatic TP53 mutations, cannot be tracked back, limiting the feasibility of applying mutant p53 as a prognostic biomarker in therapy-related AML or secondary AML.

MDS and AML are yet, heterogenous bone marrow disorders with common characteristics like expansion of clonal hematopoietic stem cells, cytopenia and marrow dysplasia.

Myelodysplastic syndrome (MDS)

MDS, especially high-grade MDS, has a high risk of transformation to a secondary AML [17]. MDS is the most common adult myeloid malignancy, has a blast count < 20%, and in about 30% transforms to AML, termed “secondary AML to MDS” or bone marrow failure. According to the International Consensus Classification (ICC), the novel intermediate state, MDS/AML, is characterized by the presence of ≥ 10% but < 20% of blasts which is applied to show a continuum of MDS to AML [9, 18]. The new, 2022, classification systems, ICC and WHO 2022, of MDS and AML, though accurate in genetic markers classification and in regression of blast threshold, is troublesome for healthcare providers due to existing discrepancies in subclassification and diagnostic criteria of acute myeloid leukemias which might affect the choice of standardized treatment or clinical trial eligibility [18]. Yet, it is outside the scope of this review to discuss the clinical weight and the shortcomings of the current classification system. Generally, MDS patients have a poor prognosis, with a median overall survival of only 5 years. MDS patients who progressed to AML, have higher-risk MDS, with myeloblast count ≥ 20% and acquired/expanded abnormalities in TP53, RUNX1, or RAS genes [19]. Phenotypically, high-grade MDS has lower cell death rates when compared to lower-risk patients [20], and typically has inferior rates of complete remission, relapse-free survival, and overall survival compared with patients with de novo AML [19].

Acute myeloid leukemia (AML)

Chromosomal abnormalities, copy number variations and translocations and inversions are common genetic events in both, MDS and AML. AML is on average diagnosed in older patients, 68 years old or older, and has poor outcomes with the five-year overall survival of less than 30% and up to 50% in younger patients [21]. Regardless of the age of diagnosis, patients who have not responded to induction therapy, have dismaying outcomes [6]. The criterion for AML diagnosis might differ depending on the driver mutation, yet in the majority, AML patients have ≥ 10% or 20%, if misdiagnosis with chronic myeloid leukemia (CML) might occur. Secondary AML accounts for up to 25% to 35% of total AML cases [19] with most (60–80%) arising from MDS. Due to clinical impact, in the new ICC system, a new subgroup was generated which constitutes a separate entity within the group of myeloid neoplasms with mutated TP53 which includes MDS, MDS/AML and AML with mutated TP53 [9, 22, 23]. TP53 mutations underly the aggressiveness of AML and, even though in MDS multihit TP53 mutations are required for diagnosis of MDS with mutated TP53, in AML and MDS/AML with mutated TP53, any pathogenic TP53 mutation VAF of ≥ 10% is sufficient for diagnosis [24]. TP53 mutations, both clonal and subclonal (VAF < 20%), have been firmly associated with an adverse outcome, as supported by significantly inferior complete remission, overall survival, and event-free survival rates [25].

Patients with AML having subclonal TP53 mutations are less likely to have complex karyotype or chromosomal losses. However, the responses to treatment were poor regardless of the mutation type. It was observed that the TP53-mutated cohort had a significantly lower complete remission rate (below 50%) compared to TP53 wild-type patients (above 80%). This rate was equally distributed among the VAF-based groups (TP53 VAF > 40%, 20%-40%, and < 20%). The 3-year overall survival rate had notable differences between TP53 wild-type and TP53-mutated patients (49.1% vs 8.3%) [25].

In 2017 and later, eleven, new drugs or combinations were approved for AML by the Food and Drug Administration [26]. Among the new approvals, five drugs target known AML vulnerabilities; FLT3 (midostaurin [27], gilteritinib [28]), IDH1 (ivosidenib) [29], IDH2 (enasidenib) [30] and BCL2 (venetoclax) [31]. Yet, the targeted treatments are not effective in high-risk TP53-mutated AML patients, who have dismaying outcomes as assessed for the frontline treated patient group [32].

p53 tumor suppressor protein

P53 is a tumor suppressor and is encoded by the TP53 gene located at 17p13.1, a site undergoing chromosomal aberrations resulting in cytogenetic deletion at 17p13.1; loss of heterozygosity (LOH) at the 17p TP53 locus or mutations largely of missense type [22]. TP53 gene (coding for p53 protein) is often mutated in human cancers, in the majority in the DNA binding domain. Cancers with a high incidence of TP53 mutations are high-grade serous ovarian cancer, lung, colon, brain or pancreatic cancer [33]. The TP53 gene mutations of missense, nonsense, frameshift or in/dels types, may result in loss-of-function, gain-of-function or in the dominant negative effect [34] propensities of the mutated protein (reviewed in [35]). Yet, our understanding of the biology behind the multiple pathogenic variants is limited.

In cancers with intact TP53 gene, the functional protein is inactivated by overexpressed mouse double minute 2 (MDM2) and/or MDM4 which bind to the N-terminal domain and inhibit the transcriptional function of p53, or promote p53 mono- and/or polyubiquitination and nuclear export and proteasomal degradation [33]. p53 is a transcription factor which directly or indirectly activates or represses a range of target genes involved in a multitude of cellular processes like; cell cycle regulation, DNA repair, senescence, pro- and anti-oxidant response, apoptosis, ferroptosis, pyroptosis, cuproptosis, autophagy, immune response, inflammation, metabolism or fertility and stem cells' renewal. The decision by which p53 drives the cell response to stress stimuli is complex and depends on the post-translational modifications and tissue/cell context (reviewed in [36]). The non-transcriptional functions of p53 are currently under debate and will not be discussed in this review.

In MDS and AML, p53 protein inactivation and TP53 gene mutations represent a clinically important predictive biomarker to DNA damaging chemotherapy [37] or venetoclax [38, 39] and are thus, underlying adverse prognoses which may dependent on complex karyotype.

In a clinical study called VIALE-A, TP53 mutant and wild-type AML patients who were new to treatment were given venetoclax and azacytidine (AZA) or AZA alone. Results showed improved remission rates but no significant difference in duration of response (DoR) or overall survival (OS) with the combination therapy in patients with poor-risk cytogenetics and mutant TP53. TP53 wild-type and complex karyotype patients showed comparable remission rates, DoR, and OS to the patients with intermediate-risk cytogenetics.

This review will highlight the contemporary status of the p53, p53-activating drugs and emerging new, investigational therapies targeting the p53 in MDS and AML.

TP53 gene alterations in MDS and AML

MDS and MDS/AML develop from mutated clones present in the hematopoietic compartment [40]. Yet, the presence of a TP53-mutated clone alone is not sufficient for the development of effective leukemogenesis (Fig. 1b).

Within European Leukemia Net (ELN) 2017, TP53 mutations are associated with unfavorable risk category and decreased overall survival (OS < 2 years). Patients with AML with TP53 mutations and complex karyotype (CK) have inferior OS of 161 days vs 374 days compared with wild-type TP53 [41]. Even though, patients with mutant TP53 AML after complete remission receive allogeneic hematopoietic stem cell transplants, yet are among the group with high relapse rates [42]. Patients with MDS with excess blasts-2 (MDS-EB-2) and TP53 mutations, share similar characteristics and clinical outcomes with mutant TP53, de novo AML patients. According to recent reports, both, TP53 mutated AML and MDS-EB-2, have practically undistinguishable biology; have blast count 15%-20% (20% cutoff is not considered specific any longer), in the majority (50%—70%), have no co-existing driver mutations or rareness of NPM1 or FLT3 alterations [24, 43], and possess high incidence of complex/monosomal karyotypes (80%–90%), which include abnormalities in chromosomes 5, 7, and 17 [24]. In therapy-related myeloid malignancies, TP53 mutations are not induced by the treatment itself but by existing progenitor clones with mutant TP53, that are resistant to DNA-damaging therapy, and expand in clonal hematopoiesis to give raise to TP53-mutated MDS/AML (Fig. 1b) [1].

Congenital cancer predisposition syndromes predisposing to myeloid neoplasm are a separate entity according to WHO [44]. Hereditary cancer predisposition syndrome, Li- Fraumeni (LFS), is an autosomal dominant condition connected with a high risk of a broad range of childhood- and adult-onset cancers. Patients with LFS develop multiple tumors during their lifespan; predominantly soft tissue sarcomas, osteosarcomas, pre-menopausal breast cancers, brain tumors, adrenocortical tumors and less frequently, pancreatic, ovarian or gastrointestinal cancers and other [45, 46]. The incidence of leukemias is < 5% [47, 48]. LFS is described by the heterozygous germline mutations in the TP53 gene [49]. Family history of inherited mutant (pathogenic variant) TP53 is a key criterion for the consideration of LFS yet, de novo mutations occur in 10%–20% of LFS cases [50]. Recent whole-genome sequence analysis combined with clock-like mutational signatures and MutationTimeR algorithm revealed that in LFS patients TP53 LOH occurs many years before tumor diagnosis, likely already in utero. It has been concluded that the copy number gains of mutant TP53 occur spontaneously in LFS patient cells and can readily outcompete diploid clones in a small number of generations [51]. In LFS patients all hematopoietic progenitor stem cells (HSPCs) carry TP53 mutations. Yet, the patients may mainly develop treatment-related myeloid neoplasm (t-MN) later in life (Fig. 1b) and prognosticate a poor prognosis with standard therapies and even allogeneic stem cell transplant [47, 52]. Thus, the risk of the development of t-MN in LFS patients should be taken into consideration when administering radiation treatment or myelosuppressive therapy.

Schwachman syndrome (SDS) with congenital mutations in Schwachman–Bodian–Diamond syndrome (SBDS) gene is a condition of high risk of developing myeloid neoplasms (MN) early in life [53]. In SDS, the SBDS protein which promotes the formation of the mature, translationally active 80S ribosome, is mutated, resulting in decreased ribosomal subunit joining and reduced translation efficiency and ribosomal stress (Fig. 1b) [54]. Survival is poor in SDS patients who develop MDS or AML originating from CH [55, 56]. It has been reported that the presence, number, persistence, and allele abundance of somatic TP53 mutations were not predictive of leukemia risk in SDS patients with CH, yet, the progression of TP53-mutated clones was found to be driven by the development of bi-allelic alterations of the TP53 locus via deletion, copy number (CN)-LOH, or point mutation (Fig. 1b) [56]. It is hypothesized that continued ribosome stress in SDS HSPCs carrying a heterozygous TP53 mutation selects for clones that inactivate the second TP53 allele and lead to the development of TP53-mutated CH, but also to its progression to myeloid malignancy (Fig. 1b) [1].

Because of the adverse outcomes for patients with TP53-altered AML/MDS as per norm, it should be strongly encouraged to enrol the patients into clinical trials. Such a strategy would let patients access promising treatments or new combinations with the potential to improve outcomes since the current gloom scenario shows dismal median survival only up to 10 months, irrespective of therapies used [57].

Targeting mutant p53 for improved therapy in MDS and AML

In TP53-mutated MDS, the “multihit” involvement with other genomic or chromosomal alterations is observed [58]. TP53 copy-number loss is prevalent in 70% of AML cases with a concomitant TP53 gene abnormality [32]. The recently investigated cohort of five hundred de novo and refractory AML patients shows that around 80% of patients harbor missense substitutions in the TP53 gene. Nonsense or in/del mutations are less common. In frontline patients, the predominant missense variations were R248, R273, R175 and Y220. CN loss with concomitant hot-spot TP53 variants is more deleterious in comparison with those with normal CN. In the cited work, the authors conducted an integrative multidimensional evaluation. The analysis was based on the interplay of somatic mutations, CN alterations, and protein expression patterns assessed using digital image-assisted immunohistochemistry. The study showed that TP53 mutation and CN status correlate with p53 protein expression patterns, such as p53high and p53truncated, which predict mutant TP53 [32]. Also, it was demonstrated that the contribution to MDS pathogenesis and AML transformation is more likely to involve TP53 hotspots that differ from those involved in de novo AML pathogenesis [32].

The most common hot-spot mutations in TP53 can be classified into two groups; structural mutants which have altered conformation of the DNA binding domain and DNA-contact mutants; which have alterations in amino acids responsible for direct interactions with DNA. Contact mutants like p53-R273H or p53-R248Q display aberrant interactions with DNA as reported by several labs [59, 60], yet, unlike conformational mutants (e.g. p53-R175H), have a structure similar to wild-type protein [61].

Wild-type p53 is involved in multiple processes enabling tumor suppression and efficiently drives cell death in a transcription-dependent manner under stress conditions (Fig. 2). Cancer cells typically contain abundant mutant p53 protein resulting from disruption of the p53-MDM2 negative feedback loop (Fig. 2). In addition, mutant p53 is stabilized in cells by interacting with heat-shock proteins (HSP) which halt the degradation by MDM2 and other E3 ubiquitin ligases [62]. Thus, mutant p53 seems a plausible target for the development of targeted therapeutics (reviewed in [63]). Mutant p53 has three oncogenic properties that contribute to tumorigenesis. Loss-of-function, caused by the inability to bind the wild-type p53 target genes essential for tumor suppression; dominant negative effect (DNE) by forming hetero-tetramers with wild-type p53 [34], p73 or p63 [64] and the gain-of-function (GOF) propensity [65,66,67,68] which heavily depends on the context and include, cooperation with other oncogenes like HIF1a to withstand the hostile hypoxic environment, promoting cytokine secretion, angiogenesis and persistent cell cycling [69, 70], escaping cell death, immune evasion and enabling DNA damage repair and fueling nutrients [71]. The mechanisms and the significance of the GOF of mutant p53 propensities in tumor progression and proliferation have been debated. For example, the expression of mutant trp53 proteins (including two hot-spot equivalent to human R248 and R273) in HSPCs derived from TRP53−/− mice did not accelerate tumorigenesis when compared to mice with reconstituted HSPCs from TRP53−/− mice [72]. It has been concluded that mutant trp53 proteins do not hasten lymphoma development in TRP53−/− and TRP53+/−. Next, a recent study showed that the genetic removal of twelve different p53 mutants previously reported to exhibit GOF activity did not impact proliferation, or response to chemotherapeutics, nor had the effect on growth or metastasis [73].

Fig. 2
figure 2

A simplified scheme of p53 regulation in cells upon stress and the drugs currently evaluated in clinical trials for p53 reactivation for improved therapy in MDS and AML. In normal cells, p53 protein is activated by numerous stress conditions like, oncogene activation, telomere shortening, replication stress, DNA damage, or elevated levels of reactive oxygen species (ROS). Wild-type p53 is released from MDM2 and MDM4 by phosphorylation by ataxia telangiectasia mutated (ATM) or ataxia telangiectasia and Rad3-related protein (ATR). (ATM-mediated phosphorylation of c-Abl mediates the release of p73 from the complex with MDM2, not shown). Consequent p53 acetylation and protein accumulation allow activation of p53 (and p73) transcriptional function. Activated ARF binds directly to MDM2 upon oncogenic stress and shifts the conformation so that p53 is released from the complex and becomes activated. ROS-activated c-Jun N-terminal kinase (JNK) was also shown to phosphorylate and activate p53 and p73. In normal cells, the negative feedback loop between p53-MDM2 is responsible for p53 protein turnover. In cancer cells amplified MDM2 prevents p53 accumulation and activation. Mutant p53 protein accumulates in large quantities in cancer cells and escapes the regulation by the p53-MDM2 feedback loop. Drugs targeting wild-type and mutant p53 appraised in clinical studies for MDS and AML are marked in bold. Modified from [63, 70, 74,75,76,77,78]. →  →  → tandem of arrows indicates a multi-step process. Created with BioRender.com

The documented prevalence of TP53 mutant variants in MDS and AML urges the development of a therapeutic approach that aims at the reactivation of mutant p53 to reinstate the wild-type p53 tumor suppression function in malignant cells. The evidence from pre-clinical studies supports the feasibility of mutant p53 reactivation in cancer cells for improved cancer therapy. The approaches which advanced to clinical trials include; refolding mutant p53 to wild-type-like conformation with small molecules, stabilizing the DNA core domain with Zn2+ chelators, mutant p53 degradation and gene therapies [79]. At the same time, reactivation of wild-type p53 through inhibiting the p53-MDM2 interactions is equally feasible, and several compounds are now tested in clinical studies [80, 81].

Mutant p53 conformation correctors

PRIMA-1MET/APR-246/Eprenetapopt

The first compound reported to act as a mutant p53 conformation corrector discovered in the protein-based screen, was CP-31398. It belongs to the Michael acceptors group of compounds [79, 82]. In 2002 Wiman and colleagues discovered a small molecule PRIMA-1 which killed tumor cells dependent on missense mutant p53 [83]. PRIMA-1 is a soft electrophile used as a scaffold to synthezise PRIMA-1MET (eprenetapopt, APR-246), a methylated analog [84]. Both PRIMA-1 and APR-246 are converted to methylene quinuclidinone (MQ) [85]. MQ binds to cysteines in the p53 core domain via Michael addition and corrects the conformation of mutant p53 to wild-type-like as detected using conformation-specific antibodies (Fig. 2) [86,87,88]. PRIMA-1/PRIMA-1MET reactivates wild-type p53 protein activity and induces cell death in multiple cancer cell lines with different contact and structural p53 mutants, including R110L, V157F, R175H, L194F, R213Q/Y234H, G245V, R248Q, R273C, R273H/P309S, R280K, and R282W and in in vivo tumor models [89,90,91]. In addition to mutant p53 per se, APR-246 also targets cancer cells’ redox balance [79, 91, 92]. Eprenetapopt was also recently shown to have mutant p53-independent functions and to kill tumor cells regardless of TP53 mutant status [93]. The outcomes observed in in vitro studies may depend on various factors, such as the number of cells used, the concentration of the drug applied, the time course of the proliferation/viability assay, and the type of tumor being studied. The authors of the study discuss these factors, which are also applicable to many drugs, even those that have already been approved.

The first-in-human, phase Ib clinical trial, in hematological and prostate cancers (NCT00900614), allowed to estimate the maximal tolerated dose and clinical response was observed in several patients and one patient with TP53-mutated AML showed a reduction of blast percentage from 46% to 26% in the bone marrow [94] (Table 1).

Table 1 p53 reactivating drugs in clinical studies in MDS and AML (for complete list of clinical trials refer to the text)

Two, phase Ib/II clinical trials with APR-246 have been concluded so far. Both studies were designed to recruit patients with TP53 gene mutations based on the primary mechanism of action of the drug. One study, for the safety and efficacy of APR-246 in combination with azacitidine and to assess complete remission (CR) of the patients with TP53-mutated myeloid neoplasm alone and in combination with azacitidine (AZA, vidaza) (NCT03588078) [95] (Table 1), and one to determine the safety and recommended dose of APR-246 in combination with azacitidine as well as to see if this combination of therapy improves overall survival (OS) (NCT03072043) [96]. In the NCT03588078 trial, fifty-two TP53-mutated patients (34 MDS, 18 AML) were recruited. 80% of the patients had complex karyotype and median baseline mutant TP53 VAF was 20%. In MDS patients an overall response rate (ORR) was 62%, including 47% CR, with a median duration of response at 10.4 months. In AML patients the ORR was 33% including 17% CR. Of the patients who responded, 73% achieved mutant TP53 VAF < 5% determined by negativity of next-generation sequencing (NGS). The median follow-up was 9.7 months, median OS was 12.1 months in MDS patients, and 13.9. The combination was well tolerated and showed potentially higher ORR and CR rates, and longer OS than reported with AZA alone [95]. In the NCT03072043 trial, fifty-five patients with at least one TP53 mutation were treated. 89% of patients had a complex karyotype and/or multihit, e.g. > 1 TP53 mutation or deletion 17p/-17. The mutant TP53 median VAF in peripheral blood was 21%. 96% of patients had at least one mutation in the DNA binding domain. Azacitidine and eprenetapopt resulted in a 71% ORR and 44% CRR in the intention-to-treat population (50% for patients with MDS) with a median OS of 10.8 months, comparing favorably with single-agent azacytidine [96].

In the follow-up, a phase II study was performed (NCT03931291) to investigate the efficacy and safety of APR-246 in combination with azacitidine for TP53-mutated MDS or AML patients as post-hematopoietic stem-cell transplantation (HSCT) maintenance therapy [97]. Patients were screened pre-HSCT and from fifty-five patients screened post-HCT, thirty-three were enrolled and were treated with eprenetapopt in combination with azacitidine. In total, thirty patients had mutant TP53 detectable in the pre-HSCT sample. Among ten patients who completed all 12 treatment cycles and did not relapse, pre-HSCT mutant TP53 was detected in four patients and VAFs remained low during the treatment. At a median follow-up of 14.5 months, the median relapse-free survival (RFS) was 12.5 months. With a median follow-up of 17.0 months, the median OS was 20.6 months. It has been concluded that post-HSCT maintenance with eprenetapopt plus azacitidine was well tolerated with acceptable safety and may improve outcomes in mutant TP53 MDS or AML [97].

Phase III clinical trial (NCT03745716) was conducted to compare the rate of CR and duration of CR, in patients with TP53-mutated MDS who will receive APR-246 and azacitidine or azacitidine alone. In total 154 patients were recruited. At a median follow-up of 12 months, Aprea Therapeutics, the study sponsor, reported that the CR rate was 37% higher in eprenetapopt with AZA arm compared to AZA alone but did not reach statistical significance and the study failed to meet the primary endpoint [98, 99].

Phase I trial (NCT04214860) was designed for dose-finding and cohort expansion study to determine the safety and preliminary efficacy of APR-246 in combination with venetoclax and azacitidine in patients with TP53-mutated myeloid malignancies. In total forty-nine patients were enrolled on the trial. Of the 49 patients who received study treatment, 20 (41%) had therapy-related acute myeloid leukaemia or had therapy-related secondary acute myeloid leukaemia, 24% of patients had more than one mutation of TP53 and 80% of patients had VAF > 50%. The overall response rate among patients receiving eprenetapopt and venetoclax with azacitidine was 64%, and the CRR was 38%. In NGS-tested patients the clearance of TP53 VAF < 5% was achieved in 26% of patients. The study showed that the combination of eprenetapopt and venetoclax with azacitidine had an acceptable safety profile in patients with previously untreated TP53-mutated acute myeloid leukemia [100].

Arsenic trioxide/ATO/Trisenox

Arsenic trioxide (ATO) is a standard of care in acute promyelocytic leukemia. ATO refolds p53 structural mutants to wild-type-like conformation and induces p53 transcription activity and cell death. The crystal structure of ATO-bound mutant p53 proteins showed, that alike APR-246, ATO binds covalently cysteine residues in the DNA binding domain, yet specifically targets residues in the allosteric cryptic site composed of three cysteines C124, C135 and C141 [88]. Phase I clinical study (NCT03855371), a combination of decitabine and ATO to treat AML/MDS expressing a classified type of mutant p53, evaluates the side effect and treatment potential of DAC + ATO in TP53 mutated high-risk MDS patients. According to the trial description, about two hundred AML/MDS patients will be recruited for TP53 sequencing. The mutant p53-positive AML/MDS patients will be treated with the combination.

Phase II study (NCT03381781) decitabine, cytarabine (Ara-C) and arsenic trioxide (ATO) in the treatment of acute myeloid leukemia with p53 mutations is designed to sequence one thousand five hundred MDS/AML patients and randomize around one hundred patients with TP53 mutations for the treatment. The outcomes of the studies remain to be reported.

Mutant p53 degraders

The most clinically investigated mutant p53 degraders are HSP90 inhibitor ganetespib and the FDA-approved histone deacetylase inhibitor, vorinostat. Yet, neither of the drugs is studied in patients based on stratification dependent on the presence of mutant TP53. Only one trial reported outcomes for vorinostat and decitabine in patients with acute myeloid leukaemia or myelodysplastic syndrome, but did not report the status of TP53 [101].

Atorvastatin, is a statin approved by the FDA to prevent cardiovascular disease in patients with abnormal levels of lipids. Statins block the key enzyme in the mevalonate pathway (sterol synthesis pathway) and lower cholesterol levels. Blocking the 3-hydroxy-3-methylglutaryl-CoA reductase (HMG-CR) prevents the synthesis of mevalonate, a cholesterol precursor, which when inhibited prevents protein prenylation, G proteins signaling and epithelial-to-mesenchymal transition (EMT) [102]. Atorvastatin, and other statins, were shown to promote degradation of misfolded mutant p53 proteins by releasing it from the complex with HSPs and the consequent degradation by E3 ligase CHIP (Fig. 2) [103]. Even though statins have demonstrated multiple mechanisms by which they might affect tumor cells’ viability, Phase I trial (NCT03560882), a pilot trial of atorvastatin in TP53-mutant and TP53 wild-type malignancies, will determine if atorvastatin will decrease the levels of conformational mutant p53 in solid tumors and relapsed AML. The trial is currently ongoing.

Other p53 structure correctors which reached the clinical trials’ testing include COTI-2 [87] or PC14586, a small molecule p53 reactivator that is selective for the p53 Y220C mutation [104], yet these drugs are not being evaluated in MDS or AML patients and therefore will not be discussed in detail in this review.

APR-246/eprenetapopt has been tested or is tested in thirteen clinical trials (clinicaltrials.gov, accessed 10.09.2023) in cancer and is, so far, the most clinically advanced and promising drug reactivating mutant p53 in myeloid malignancies to date. The clinical trials showed that eprenetapopt has a favorable safety profile. It exhibits antitumor activity even against tumors that carry biallelic TP53-mutant clones. The clearance of p53 protein by immunohistochemistry correlates with clone size, which is measured by its VAF. This correlation can be used as a surrogate marker to determine the clinical activity of the tested compound [98].

The phase III clinical study with frontline APR-246 in combination with azacitidine in TP53 mutant MDS showed good CR rates, yet the study did not meet the primary endpoint. The lack of difference between the arms could be attributed to the small effect size for APR-246 + AZA arm. This means that to see a statistically significant therapeutic effect, more patients may need to be treated. Additionally, patient selection based on factors beyond just the presence of a TP53 mutation may be necessary. For instance, the type of TP53 mutation could be a consideration, as discussed previously [98].

Targeting p53/MDM2/MDM4 axis with small molecules

In tumors which retain wild-type TP53, p53 protein is inactivated through two major routes: through binding of MDM2/MDM4 oncoproteins to the N-terminal domain and inhibition of p53 transcription function, and MDM2-, ubiquitin-mediated proteasomal degradation (Fig. 2) [105].

In hematological malignancies, amplification of MDM2 was reported for AML, CML ALL with no concomitant mutations in exons 4–10 of the TP53 gene [106] and overexpression of MDM2 is associated with poor prognosis in AML [107]. MDM4 negatively regulates p53 and it was assessed in AML and MDS only in one study. Immunohistochemistry showed overexpression of MDM4 in 78 AML cases (92%) and 12 MDS cases (52%) and might be a therapeutic target in AML and MDS [108]. Targeting the interactions between p53-MDMD2 has become a feasible strategy after the identification of the key p53 residues fitting into the MDM2 hydrophobic pocket in the crystal structure analysis which showed three sub-pockets within the MDM2 hydrophobic cleft that are occupied by the Leu26, Trp23, and Phe19 amino acid side chains of p53 [109]. The first drug ever developed to target the p53-MDM2 protein complex was nutlin, a cis-imidazoline [110].

RG7112 and Idasanutlin/RG7388/RO5503781

RG7112 (RO5045337), is a cis-imidazoline, a derivative of nutlin and the first MDM2 antagonist tested in clinical trials. It showed clinical activity in AML patients in the Phase I trial (NCT00623870) (Table 1). A total of 116 patients were enrolled and at least 16 patients with wild-type p53 AML were treated. TP53 mutational analysis showed TP53 mutations in 19 of 96 patients tested and most mutant TP53 patients failed to show evidence of response. Ten genes, all p53 targets, were induced in wild-type p53 patients after treatment [111]. Yet, gastrointestinal toxicity, myelosuppression, and related complications resulted in the discontinuation of RG7112 clinical trials (reviewed in [112]).

Idasanutlin (RG7388, a selective MDM2 inhibitor) is widely studied in clinical trials. Idasanutlin is a pyrrolidine with enhanced potency, selectivity, and bioavailability compared to RG7112. It has been tested in Phase I/Ib trial (NCT01773408), a study of RO5503781 as a single agent or in combination with cytarabine in participants with acute myelogenous leukemia, yet, the final outcomes of the study were not published apart from the abstract [113]. Marker analysis of the patients enrolled in the study, using flow cytometry data for sixty-three evaluable patients showed that MDM2 expression in leukemic blasts was significantly associated with patients exhibiting a composite complete remission, and CR with incomplete hematologic recovery vs. no response. MDM2 per cent cell positivity in CD45dim/CD34 + /CD117 + leukemic blasts also showed an association with clinical outcomes. Overall, the analysis supports improved MDM2 antagonist clinical outcomes in AML patients with higher levels of MDM2 protein expression and thus, MDM2 protein expression from blasts may serve as a stratification biomarker for AML patients likely to benefit from idasanutlin-based therapy [114].

Phase Ib/II study (NCT03850535), evaluating the safety and efficacy of idasanutlin in combination with cytarabine and daunorubicin in patients newly diagnosed with acute myeloid leukemia (AML) and the safety and efficacy of idasanutlin in the maintenance of first AML complete remission, is designed, to evaluate the safety, efficacy, and the sponsor decided not to continue the study based on the overall company strategy in AML and a too-small group of patients enrolled.

In Phase III study (NCT02545283), study of idasanutlin with cytarabine versus cytarabine plus placebo in participants with relapsed or refractory acute myeloid leukemia (AML) (MIRROS), evaluated efficacy and safety of the treatment. A total of 447 patients were enrolled, 81% of patients had wild-type TP53. At the median duration of follow-up 6.7 months in both arms (drugs vs placebo), no subgroup showed a different outcome for OS. The median duration of CR was 13.9 months in the group with idasanutlin and 29.4 months in the placebo group. The myelosuppressive effect of idasanutlin was observed, yet prolonged neutropenia affected the response rates. The study did not meet the primary endpoint [115].

Navtemadlin/AMG-232/KRT-232

AMG-232 is an improved derivative of piperidinone [116] and was evaluated in relapsed/refractory AML in a completed phase I study (NCT02016729) (Table 1) [117]. A study evaluated the safety and efficacy of AMG-232 alone and in combination with MEK inhibitor, trametinib. In the trial thirty-six patients with relapsed/refractory AML were enrolled, TP53 mutational status was known for 44% patients at enrollment. Expression of BAX, PUMA, P21, and MDM2 increased in the leukemic bone marrow and four patients achieved remission [117]. Two more studies are currently recruiting participants for the treatment of AML patients with navtemadlin; one phase Ib in combination with decitabine and venetoclax (NCT03041688) and one phase Ib testing the addition of an anti-cancer drug, navtemadlin, to the usual treatments (cytarabine and idarubicin) in patients with acute myeloid leukemia (NCT04190550).

Milademetan/DS-3032b

Milademetan (DS-3032) tosylate hydrate is a specific and orally active MDM2 inhibitor [118]. Phase I study (NCT02319369) (Table 1) evaluated the safety, tolerability and pharmacokinetics of milademetan alone and with 5-azacitidine in acute myelogenous leukemia (AML) or high-risk myelodysplastic syndrome (MDS). Among 38 patients, two with AML and one with myelodysplastic syndrome had complete remission [112, 119].

In the phase I study (NCT03552029), patients were enrolled to milademetan plus quizartinib combination study in FLT3-ITD mutant acute myeloid leukemia (AML). Ten patients were only recruited, and the study was terminated based on a business decision by the Sponsor.

Phase I/II study (NCT03634228) (Table 1), milademetan tosylate and low dose cytarabine with or without venetoclax in treating participants with recurrent or refractory acute myeloid leukemia enrolled a total of 21 patients. Combination of MDM2 inhibition with inhibition of BCL2 potentiated apoptotic response however, no meaningful clinical responses were reported. Noticeable gastrointestinal toxicities were reported, and the study was terminated early due to futility [120].

Siremadlin/ HDM201/CGM097

Siremadlin is a dihydroisoquinolinone derivative, the next-generation MDM2 inhibitor, evaluated in clinical studies [121]. Phase I (NCT02143635) (Table 1), a first-in-human dose-escalation study to determine and evaluate a safe and tolerated dose of HDM201 in patients with selected advanced tumors that are TP53wt, enrolled 115 patients with solid tumors and 93 patients with hematologic tumors (99% AML) [122]. A clear trend was observed for increases in serum GDF-15 (growth/differentiation factor-15), a biomarker for p53 transcription activity. Thirty-three per cent of evaluated patients had MDM2 amplification and fifty-three per cent of MDM2amp patients achieved either partial response or stable disease. The drug showed an acceptable safety profile [122].

Two trials with siremadlin are currently recruiting participants with AML; one, phase I/II (NCT05447663) to evaluate siremadlin alone and in combination with donor lymphocyte infusion in acute myeloid leukemia post-allogeneic stem cell transplant, second, phase I/II (NCT05155709) a study of siremadlin in combination with venetoclax plus azacitidine in adult participants with acute myeloid leukemia (AML) who are ineligible for chemotherapy. The results remain to be published.

APG-115

APG-115 belongs to spirooxindoles, a class of potent MDM2 inhibitors of Ki < 1 nM [123]. Phase Ib study (NCT04275518), of APG-115 single agent or in combination with azacitidine or cytarabine in patients with AML and MDS is recruiting one hundred two patients with relapsed/refractory AML and relapsed/progressed high/very high-risk MDS. Phase Ib/II study (NCT04358393) of APG-115 alone or in combination with azacitidine in patients with relapsed/refractory AML, CMML or MDS will enrol sixty-nine patients (Table 1). The outcomes of the studies remain to be published.

Sulanemadlin/ALRN-6924

Sulanemadlin (ALRN-6924), is the first cell-permeating, stabilized α-helical peptide which mimics the N-terminal domain of the p53 and binds with high affinity to both MDM2 and MDM4 to activate p53 signaling in cancer cells [124].

Phase I/Ib (NCT02909972) (Table 1) safety study of ALN-6924 in patients with acute myeloid leukemia or advanced myelodysplastic syndrome has recruited fifty-five patients and evaluates anti-tumor effects of ALRN-6924 alone or in combination with cytarabine. The outcome of the study remains to be published.

The phase 1b breast cancer chemoprotection trial with ALRN-6924 in patients with p53-mutated breast cancer (NCT05622058) was terminated due to severe grade 4 neutropenia and alopecia, failing to meet the trial's main endpoints. Aileron Therapeutics, the drug owner, is running several other trials with ALRN-6924 and just recently offered the strategy to strengthen the NCT05622058 trial. The new strategy foresees using a lower dose of ALRN-6924 for chemoprotection than originally planned.

Currently, other MDM2 inhibitors are under investigation in clinical trials, like BI-907828 (brigimadlin) [125] but are not evaluated in MDS or AML. Yet, the list of clinical trials discussed above may urge us to conclude that targeting MDM2 with high-affinity inhibitors has not so far delivered the expected clinical benefit in patients with AML and MDS. Likely, other strategies are needed to overcome the persevering problem of the insufficient response of patients due to persistent neutropenia and adverse events related to the gastrointestinal tract.

One way to improve the patients’ outcomes with high-affinity MDM2 inhibitors might be their conjugation with antibodies to improve drugs’ selectivity or a design of less toxic combination strategy based on repurposed drugs like metformin. This well-known antidiabetic drug is currently being tested in a pilot, MILI clinical study in nondiabetic Li-Fraumeni Syndrome (LFS) patients with germline TP53 variants (NCT01981525) [126]. The trial aims to assess the tolerability of daily oral metformin in LFS patients with the endpoint of reducing cancer incidence. The mechanism by which metformin might prevent or slow down cancer development in LFS patients is unclear. The mechanism may rely on inhibiting the pre-cancerous niche through a metabolic switch by activating AMPK, inhibiting mTOR, and improving insulin sensitivity in liver tissues of TP53 mutant carriers [127]. Results from murine models of LFS showed significantly prolonged median overall survival after metformin administration, providing grounds for clinical trial initiation [128].

Drug repurposing emerges as a promising therapeutic approach in oncology since the drugs of potential anti-cancer activities have already been approved by the FDA for another indication and the safety profiles are known [129]. Multiple clinical studies, like the biomarker prevention trial studying the combination of metformin with aspirin in stage I-III colorectal cancer patients (ASAMET, NCT03047837) are ongoing. The outcomes of these trials have not been reported yet.

Beyond MDM2/MDM4 inhibitors

Another promising strategy to target MDM2 is to promote its degradation using protein degrader, PROTAC. A recent pre-clinical study in breast cancer demonstrated the feasibility of reconstituting the p53 tumor suppressor pathway in the presence of mutant p53, through activation of p73 [130]. p73 belongs to the p53 protein family and together with p63 are ancestors of p53 in multicellular organisms. TP73 gene is rarely mutated in cancers. Due to high structure and function homology, p73 protein recognizes a plethora of p53 target genes involved in tumor suppression but has also p53-independent functions. Stress-induced p73 post-transcriptional regulation appears to be similar to that of p53 [74, 131] and both proteins undergo regulation by MDM2/MDM4 (Fig. 2). In the case of p73 protein, ITCH E3 ligase and MDM2 are responsible for protein turnover [132]. p73 is emerging as a promising therapeutic target for improved cancer therapy in cancers with TP53 gene mutations [129, 133]. Pharmacological reactivation of p73 in mouse models showed promising results with no apparent toxicity [130, 134]. We have shown that both, p53 and p73 proteins are reactivated by a repurposed drug, protoporphyrin IX, through targeting p53/MDMD2/MDM4 and p73/MDM2 interactions [135, 136]. While potently inducing cell death in tumor cells in pre-clinical studies, PpIX had only a mild effect on the proliferation of several normal cell lines/cells including human fibroblasts or peripheral blood mononuclear cells. Thus, PpIX or its analogs, might have better safety profiles in prospective clinical studies compared to MDM2 inhibitors. It remains to be investigated whether reactivating p73 for tumor suppression would yield comparable outcomes to wild-type p53 reactivation in terms of tumor suppression but with improved safety profile.

The complexity of p53 protein biology in AML

Tuval et al., recently reported that the pre-leukemic clones with DNMT3A mutations have a selective advantage and an intrinsic chemoresistance as they pre-dominantly express pseudo-mutant p53[137]. The pseudo-mutant p53 protein is a misfolded wild-type p53 protein and has a limited transcriptional activity [138]. The protein exists in the equilibrium state in pre-leukemic blasts and was predominantly found in DNMT3A— mutated (wild-type TP53) AML enabling the clones’ enhanced self-renewal.

Refolding of pseudo-mutant with a structure-correcting peptide, pCAP-250, resulted in conformation refolding and restoration of p53 transcription activity in vitro and in vivo [137]. This implies that some sub-group of AML patients harbouring pseudo-mutant might profit from the therapy with p53 structure correctors rather than from the treatment with MDM2 inhibitors. Yet, due to the limited application of conformation-specific antibodies at the diagnosis, the stratification strategy allowing to distinguish between p53 wild-type-like conformation and unfolded conformation is not applied in the clinical study design.

The emerging importance of pseudo-mutant p53 in CH, requires modifications to the current model of TP53-mutated CH (Fig. 1b). Further studies are needed to evaluate the biology of pseudo-mutant in the pre-leukemic niche and to comprehend the co-existing factors contributing to clone evolution and fitness advantage in CH.

In conclusion, the role of p53 alterations in clonal hematopoiesis is still not fully depicted. The most advanced clinical drug, targeting p53 in hematological malignancies, is the mutant p53 reactivating compound, eprenetapopt (APR-246). So far, variable outcomes have been reported for MDM2 inhibitors and rational combination strategies may be crucial to enhanced efficacy with these compounds.

Conclusions

There are other therapeutic strategies which are employed in the treatment of TP53 mutated myeloid neoplasm and show promising outcomes. A recent multicenter observational study conducted on a large group of patients with the mutant TP53 AML, has shown that the response to the standard induction therapy was only modest, and the overall survival rate was poor. The outcomes in this group of patients have not improved even with the application of novel therapies. However, the study revealed that those patients who responded well to induction therapy and proceeded to undergo allogeneic HCT had significantly better overall survival rates [139].

For eligible patients with mutant TP53 AML, alloHCT in the first remission is recommended until better therapies are developed for improving long-term outcomes. Taking the current outcomes of the clinical trials of the p53-targeting therapies, it can be concluded that the management of the TP53-mutated myelodysplastic syndrome (MDS) and acute myeloid leukemia (AML) remains a therapeutic challenge despite significant recent advancements. Also, clinical studies on wild-type and mutant p53 reactivation have yielded inconsistent results in both MDS and AML.

Thus, further research is necessary to explore the biology of p53 in the pre-leukemic population of hematopoietic stem cells (HSCs).

The remaining open questions in this field are:

  • Can the effectiveness of mutant p53 structure correctors in the TP53-mutated MDS and AML be determined through predictive biomarker stratification?

  • Is there a possibility of high-affinity MDM2 inhibitors being approved as standalone treatments for wild-type TP53 MDS and AML?

  • Could the utilization of p53 reactivating compounds in combination therapies to impede key drivers potentially result in enhanced outcomes in MDS and AML?

  • Is it possible to develop a biomarker discovery test that uses conformation-specific antibodies to stratify patients for mutant p53 reactivating drugs?

Availabiliy of data and materials

No datasets were generated or analysed during the current study.

References

  1. Warren JT, Link DC. Clonal hematopoiesis and risk for hematologic malignancy. Blood. 2020;136:1599–605.

    PubMed  PubMed Central  Google Scholar 

  2. Arends CM, Galan-Sousa J, Hoyer K, Chan W, Jäger M, Yoshida K, et al. Hematopoietic lineage distribution and evolutionary dynamics of clonal hematopoiesis. Leukemia. 2018;32:1908–19.

    Article  CAS  PubMed  Google Scholar 

  3. Vobugari N, Heuston C, Lai C. Clonal cytopenias of undetermined significance: potential predictor of myeloid malignancies? Clin Adv Hematol Oncol. 2022;20:375–83.

    PubMed  Google Scholar 

  4. Bullinger L, Döhner K, Döhner H. Genomics of acute myeloid leukemia diagnosis and pathways. J Clin Oncol. 2017;35:934–46.

    Article  CAS  PubMed  Google Scholar 

  5. Shih AH, Chung SS, Dolezal EK, Zhang S-J, Abdel-Wahab OI, Park CY, et al. Mutational analysis of therapy-related myelodysplastic syndromes and acute myelogenous leukemia. Haematologica. 2013;98:908–12.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Falini B, Martelli MP. Comparison of the International Consensus and 5th WHO edition classifications of adult myelodysplastic syndromes and acute myeloid leukemia. Am J Hematol. 2023;98:481–92.

    Article  PubMed  Google Scholar 

  7. Gaulin C, Kelemen K, Arana Yi C. Molecular pathways in clonal hematopoiesis: from the acquisition of somatic mutations to transformation into hematologic neoplasm. Life (Basel). 2022;12:1135.

    ADS  CAS  PubMed  Google Scholar 

  8. Welch JS, Ley TJ, Link DC, Miller CA, Larson DE, Koboldt DC, et al. The origin and evolution of mutations in acute myeloid leukemia. Cell. 2012;150:264–78.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Arber DA, Orazi A, Hasserjian RP, Borowitz MJ, Calvo KR, Kvasnicka H-M, et al. International Consensus Classification of Myeloid Neoplasms and Acute Leukemias: integrating morphologic, clinical, and genomic data. Blood. 2022;140:1200–28.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Bolton KL, Ptashkin RN, Gao T, Braunstein L, Devlin SM, Kelly D, et al. Cancer therapy shapes the fitness landscape of clonal hematopoiesis. Nat Genet. 2020;52:1219–26.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Jajosky AN, Sadri N, Meyerson HJ, Oduro KA, Kelkar A, Fitzgerald B, et al. Clonal cytopenia of undetermined significance (CCUS) with dysplasia is enriched for MDS-type molecular findings compared to CCUS without dysplasia. Eur J Haematol. 2021;106:500–7.

    Article  CAS  PubMed  Google Scholar 

  12. Steensma DP. Clinical consequences of clonal hematopoiesis of indeterminate potential. Hematology Am Soc Hematol Educ Program. 2018;2018:264–9.

    Article  PubMed  PubMed Central  Google Scholar 

  13. Haferlach T, Nagata Y, Grossmann V, Okuno Y, Bacher U, Nagae G, et al. Landscape of genetic lesions in 944 patients with myelodysplastic syndromes. Leukemia. 2014;28:241–7.

    Article  CAS  PubMed  Google Scholar 

  14. Ogawa S. Genetics of MDS. Blood. 2019;133:1049–59.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Lal R, Lind K, Heitzer E, Ulz P, Aubell K, Kashofer K, et al. Somatic TP53 mutations characterize preleukemic stem cells in acute myeloid leukemia. Blood. 2017;129:2587–91.

    Article  CAS  PubMed  Google Scholar 

  16. Wong TN, Ramsingh G, Young AL, Miller CA, Touma W, Welch JS, et al. Role of TP53 mutations in the origin and evolution of therapy-related acute myeloid leukaemia. Nature. 2015;518:552–5.

    Article  ADS  CAS  PubMed  Google Scholar 

  17. Woll PS, Yoshizato T, Hellström-Lindberg E, Fioretos T, Ebert BL, Jacobsen SEW. Targeting stem cells in myelodysplastic syndromes and acute myeloid leukemia. J Intern Med. 2022;292:262–77.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Shallis RM, Daver N, Altman JK, Komrokji RS, Pollyea DA, Badar T, et al. Standardising acute myeloid leukaemia classification systems: a perspective from a panel of international experts. Lancet Haematol. 2023;10:e767–76.

    Article  CAS  PubMed  Google Scholar 

  19. Menssen AJ, Walter MJ. Genetics of progression from MDS to secondary leukemia. Blood. 2020;136:50–60.

    Article  PubMed  PubMed Central  Google Scholar 

  20. Allampallam K, Shetty V, Mundle S, Dutt D, Kravitz H, Reddy PL, et al. Biological significance of proliferation, apoptosis, cytokines, and monocyte/macrophage cells in bone marrow biopsies of 145 patients with myelodysplastic syndrome. Int J Hematol. 2002;75:289–97.

    Article  CAS  PubMed  Google Scholar 

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

    Article  PubMed  Google Scholar 

  22. Weinberg OK, Siddon A, Madanat YF, Gagan J, Arber DA, Dal Cin P, et al. TP53 mutation defines a unique subgroup within complex karyotype de novo and therapy-related MDS/AML. Blood Adv. 2022;6:2847–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Khoury JD, Solary E, Abla O, Akkari Y, Alaggio R, Apperley JF, et al. The 5th edition of the World Health Organization Classification of haematolymphoid tumours myeloid and histiocytic/dendritic neoplasms. Leukemia. 2022;36:1703–19.

    Article  PubMed  PubMed Central  Google Scholar 

  24. Grob T, Al Hinai ASA, Sanders MA, Kavelaars FG, Rijken M, Gradowska PL, et al. Molecular characterization of mutant TP53 acute myeloid leukemia and high-risk myelodysplastic syndrome. Blood. 2022;139:2347–54.

    Article  CAS  PubMed  Google Scholar 

  25. Prochazka KT, Pregartner G, Rücker FG, Heitzer E, Pabst G, Wölfler A, et al. Clinical implications of subclonal TP53 mutations in acute myeloid leukemia. Haematologica. 2019;104:516–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Shimony S, Stahl M, Stone RM. Acute myeloid leukemia: 2023 update on diagnosis, risk-stratification, and management. Am J Hematol. 2023;98:502–26.

    Article  PubMed  Google Scholar 

  27. Stone RM, Mandrekar SJ, Sanford BL, Laumann K, Geyer S, Bloomfield CD, et al. Midostaurin plus chemotherapy for acute myeloid leukemia with a FLT3 Mutation. N Engl J Med. 2017;377:454–64.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Pulte ED, Norsworthy KJ, Wang Y, Xu Q, Qosa H, Gudi R, et al. FDA approval summary: gilteritinib for relapsed or refractory acute myeloid leukemia with a FLT3 mutation. Clin Cancer Res. 2021;27:3515–21.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Norsworthy KJ, Luo L, Hsu V, Gudi R, Dorff SE, Przepiorka D, et al. FDA approval summary: Ivosidenib for relapsed or refractory acute myeloid leukemia with an Isocitrate Dehydrogenase-1 Mutation. Clin Cancer Res. 2019;25:3205–9.

    Article  CAS  PubMed  Google Scholar 

  30. Kim ES. Enasidenib: first global approval. Drugs. 2017;77:1705–11.

    Article  CAS  PubMed  Google Scholar 

  31. Wei AH, Strickland SA, Hou J-Z, Fiedler W, Lin TL, Walter RB, et al. Venetoclax combined with low-dose cytarabine for previously untreated patients with acute myeloid leukemia: results from a phase Ib/II Study. J Clin Oncol. 2019;37:1277–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Tashakori M, Kadia T, Loghavi S, Daver N, Kanagal-Shamanna R, Pierce S, et al. TP53 copy number and protein expression inform mutation status across risk categories in acute myeloid leukemia. Blood. 2022;140:58–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. de Andrade KC, Lee EE, Tookmanian EM, Kesserwan CA, Manfredi JJ, Hatton JN, et al. The TP53 database: transition from the International Agency for Research on Cancer to the US National Cancer Institute. Cell Death Differ. 2022;29:1071–3.

    Article  PubMed  PubMed Central  Google Scholar 

  34. Willis A, Jung EJ, Wakefield T, Chen X. Mutant p53 exerts a dominant negative effect by preventing wild-type p53 from binding to the promoter of its target genes. Oncogene. 2004;23:2330–8.

    Article  CAS  PubMed  Google Scholar 

  35. Wang Z, Strasser A, Kelly GL. Should mutant TP53 be targeted for cancer therapy? Cell Death Differ. 2022;29:911–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Peuget S, Selivanova G. p53-Dependent Repression: DREAM or Reality? Cancers (Basel). 2021;13:4850.

    Article  CAS  PubMed  Google Scholar 

  37. Kadia TM, Jain P, Ravandi F, Garcia-Manero G, Andreef M, Takahashi K, et al. TP53 mutations in newly diagnosed acute myeloid leukemia: Clinicomolecular characteristics, response to therapy, and outcomes. Cancer. 2016;122:3484–91.

    Article  CAS  PubMed  Google Scholar 

  38. Nechiporuk T, Kurtz SE, Nikolova O, Liu T, Jones CL, D’Alessandro A, et al. The TP53 apoptotic network is a primary mediator of resistance to BCL2 inhibition in AML cells. Cancer Discov. 2019;9:910–25.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Pollyea DA, Pratz KW, Wei AH, Pullarkat V, Jonas BA, Recher C, et al. Outcomes in patients with poor-risk cytogenetics with or without TP53 mutations treated with Venetoclax and Azacitidine. Clin Cancer Res. 2022;28:5272–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Rose D, Haferlach T, Schnittger S, Perglerová K, Kern W, Haferlach C. Subtype-specific patterns of molecular mutations in acute myeloid leukemia. Leukemia. 2017;31:11–7.

    Article  CAS  PubMed  Google Scholar 

  41. Sallman DA, Komrokji R, Vaupel C, Cluzeau T, Geyer SM, McGraw KL, et al. Impact of TP53 mutation variant allele frequency on phenotype and outcomes in myelodysplastic syndromes. Leukemia. 2016;30:666–73.

    Article  CAS  PubMed  Google Scholar 

  42. Ciurea SO, Chilkulwar A, Saliba RM, Chen J, Rondon G, Patel KP, et al. Prognostic factors influencing survival after allogeneic transplantation for AML/MDS patients with TP53 mutations. Blood. 2018;131:2989–92.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Chen X, Othus M, Wood BL, Walter RB, Becker PS, Percival M-E, et al. Comparison of myeloid blast counts and variant allele frequencies of gene mutations in myelodysplastic syndrome with excess blasts and secondary acute myeloid leukemia. Leuk Lymphoma. 2021;62:1226–33.

    Article  CAS  PubMed  Google Scholar 

  44. Arber DA, Orazi A, Hasserjian R, Thiele J, Borowitz MJ, Le Beau MM, et al. The 2016 revision to the World Health Organization classification of myeloid neoplasms and acute leukemia. Blood. 2016;127:2391–405.

    Article  CAS  PubMed  Google Scholar 

  45. Gonzalez KD, Noltner KA, Buzin CH, Gu D, Wen-Fong CY, Nguyen VQ, et al. Beyond Li Fraumeni Syndrome: clinical characteristics of families with p53 germline mutations. J Clin Oncol. 2009;27:1250–6.

    Article  CAS  PubMed  Google Scholar 

  46. McBride KA, Ballinger ML, Killick E, Kirk J, Tattersall MHN, Eeles RA, et al. Li-Fraumeni syndrome: cancer risk assessment and clinical management. Nat Rev Clin Oncol. 2014;11:260–71.

    Article  CAS  PubMed  Google Scholar 

  47. Swaminathan M, Bannon SA, Routbort M, Naqvi K, Kadia TM, Takahashi K, et al. Hematologic malignancies and Li-Fraumeni syndrome. Cold Spring Harb Mol Case Stud. 2019;5:a003210.

    Article  PubMed  PubMed Central  Google Scholar 

  48. Guha T, Malkin D. Inherited TP53 Mutations and the Li-Fraumeni Syndrome. Cold Spring Harb Perspect Med. 2017;7:a026187.

    Article  PubMed  PubMed Central  Google Scholar 

  49. Nichols KE, Malkin D, Garber JE, Fraumeni JF, Li FP. Germ-line p53 mutations predispose to a wide spectrum of early-onset cancers. Cancer Epidemiol Biomarkers Prev. 2001;10:83–7.

    CAS  PubMed  Google Scholar 

  50. Correa H. Li-Fraumeni Syndrome. J Pediatr Genet. 2016;5:84–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Light N, Layeghifard M, Attery A, Subasri V, Zatzman M, Anderson ND, et al. Germline TP53 mutations undergo copy number gain years prior to tumor diagnosis. Nat Commun. 2023;14:77.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  52. Zebisch A, Lal R, Müller M, Lind K, Kashofer K, Girschikofsky M, et al. Acute myeloid leukemia with TP53 germline mutations. Blood. 2016;128:2270–2.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Nelson AS, Myers KC. Diagnosis, treatment, and molecular pathology of Shwachman-Diamond Syndrome. Hematol Oncol Clin North Am. 2018;32:687–700.

    Article  PubMed  Google Scholar 

  54. Warren AJ. Molecular basis of the human ribosomopathy Shwachman-Diamond syndrome. Adv Biol Regul. 2018;67:109–27.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Myers KC, Furutani E, Weller E, Siegele B, Galvin A, Arsenault V, et al. Clinical features and outcomes of patients with Shwachman-Diamond syndrome and myelodysplastic syndrome or acute myeloid leukaemia: a multicentre, retrospective, cohort study. Lancet Haematol. 2020;7:e238–46.

    Article  PubMed  Google Scholar 

  56. Kennedy AL, Myers KC, Bowman J, Gibson CJ, Camarda ND, Furutani E, et al. Distinct genetic pathways define pre-malignant versus compensatory clonal hematopoiesis in Shwachman-Diamond syndrome. Nat Commun. 2021;12:1334.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  57. Shallis RM, Daver NG, Altman JK, Hasserjian RP, Kantarjian HM, Platzbecker U, et al. TP53-altered acute myeloid leukemia and myelodysplastic syndrome with excess blasts should be approached as a single entity. Cancer. 2023;129:175–80.

    Article  CAS  PubMed  Google Scholar 

  58. Bernard E, Nannya Y, Hasserjian RP, Devlin SM, Tuechler H, Medina-Martinez JS, et al. Implications of TP53 allelic state for genome stability, clinical presentation and outcomes in myelodysplastic syndromes. Nat Med. 2020;26:1549–56.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Ang HC, Joerger AC, Mayer S, Fersht AR. Effects of common cancer mutations on stability and DNA binding of full-length p53 compared with isolated core domains. J Biol Chem. 2006;281:21934–41.

    Article  CAS  PubMed  Google Scholar 

  60. Wong KB, DeDecker BS, Freund SM, Proctor MR, Bycroft M, Fersht AR. Hot-spot mutants of p53 core domain evince characteristic local structural changes. Proc Natl Acad Sci USA. 1999;96:8438–42.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  61. Cho Y, Gorina S, Jeffrey PD, Pavletich NP. Crystal structure of a p53 tumor suppressor-DNA complex: understanding tumorigenic mutations. Science. 1994;265:346–55.

    Article  ADS  CAS  PubMed  Google Scholar 

  62. Alexandrova EM, Moll UM. Depleting stabilized GOF mutant p53 proteins by inhibiting molecular folding chaperones: a new promise in cancer therapy. Cell Death Differ. 2017;24:3–5.

    Article  CAS  PubMed  Google Scholar 

  63. Bykov VJN, Eriksson SE, Bianchi J, Wiman KG. Targeting mutant p53 for efficient cancer therapy. Nat Rev Cancer. 2018;18:89–102.

    Article  CAS  PubMed  Google Scholar 

  64. Billant O, Léon A, Le Guellec S, Friocourt G, Blondel M, Voisset C. The dominant-negative interplay between p53, p63 and p73: A family affair. Oncotarget. 2016;7:69549–64.

    Article  PubMed  PubMed Central  Google Scholar 

  65. Hanel W, Marchenko N, Xu S, Yu SX, Weng W, Moll U. Two hot spot mutant p53 mouse models display differential gain of function in tumorigenesis. Cell Death Differ. 2013;20:898–909.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Stein Y, Aloni-Grinstein R, Rotter V. Mutant p53 oncogenicity: dominant-negative or gain-of-function? Carcinogenesis. 2020;41:1635–47.

    Article  CAS  PubMed  Google Scholar 

  67. Lieschke E, Wang Z, Kelly GL, Strasser A. Discussion of some “knowns” and some “unknowns” about the tumour suppressor p53. J Mol Cell Biol. 2019;11:212–23.

    Article  CAS  PubMed  Google Scholar 

  68. Ferraiuolo M, Di Agostino S, Blandino G, Strano S. Oncogenic intra-p53 family member interactions in human cancers. Front Oncol. 2016;6:77.

    Article  PubMed  PubMed Central  Google Scholar 

  69. Fontemaggi G, Dell’Orso S, Trisciuoglio D, Shay T, Melucci E, Fazi F, et al. The execution of the transcriptional axis mutant p53, E2F1 and ID4 promotes tumor neo-angiogenesis. Nat Struct Mol Biol. 2009;16:1086–93.

    Article  CAS  PubMed  Google Scholar 

  70. Amelio I, Melino G. Context is everything: extrinsic signalling and gain-of-function p53 mutants. Cell Death Discov. 2020;6:16.

    Article  PubMed  PubMed Central  Google Scholar 

  71. Cooks T, Pateras IS, Jenkins LM, Patel KM, Robles AI, Morris J, et al. Mutant p53 cancers reprogram macrophages to tumor supporting macrophages via exosomal miR-1246. Nat Commun. 2018;9:771.

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  72. Aubrey BJ, Janic A, Chen Y, Chang C, Lieschke EC, Diepstraten ST, et al. Mutant TRP53 exerts a target gene-selective dominant-negative effect to drive tumor development. Genes Dev. 2018;32:1420–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Wang Z, Burigotto M, Ghetti S, Vaillant F, Tan T, Capaldo BD, et al. Loss-of-function but not gain-of-function properties of mutant TP53 are critical for the proliferation, survival and metastasis of a broad range of cancer cells. Cancer Discov. 2023;14:362–79.

    Article  PubMed Central  Google Scholar 

  74. Kostecka A, Sznarkowska A, Meller K, Acedo P, Shi Y, Mohammad Sakil HA, et al. JNK-NQO1 axis drives TAp73-mediated tumor suppression upon oxidative and proteasomal stress. Cell Death Dis. 2014;5: e1484.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Daver NG, Maiti A, Kadia TM, Vyas P, Majeti R, Wei AH, et al. TP53-mutated myelodysplastic syndrome and acute myeloid leukemia: biology, current therapy, and future directions. Cancer Discov. 2022;12:2516–29.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Hassin O, Oren M. Drugging p53 in cancer: one protein, many targets. Nat Rev Drug Discov. 2023;22:127–44.

    Article  CAS  PubMed  Google Scholar 

  77. Hernández Borrero LJ, El-Deiry WS. Tumor suppressor p53: biology, signaling pathways, and therapeutic targeting. Biochim Biophys Acta Rev Cancer. 2021;1876:188556.

    Article  PubMed  Google Scholar 

  78. Vaseva AV, Moll UM. The mitochondrial p53 pathway. Biochim Biophys Acta. 2009;1787:414–20.

    Article  CAS  PubMed  Google Scholar 

  79. Nishikawa S, Iwakuma T. Drugs targeting p53 mutations with FDA approval and in clinical trials. Cancers (Basel). 2023;15:429.

    Article  CAS  PubMed  Google Scholar 

  80. Kojima K, Konopleva M, Samudio IJ, Shikami M, Cabreira-Hansen M, McQueen T, et al. MDM2 antagonists induce p53-dependent apoptosis in AML: implications for leukemia therapy. Blood. 2005;106:3150–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Sekiguchi N, Kasahara S, Miyamoto T, Kiguchi T, Ohno H, Takagi T, et al. Phase I dose-escalation study of milademetan in patients with relapsed or refractory acute myeloid leukemia. Int J Hematol. 2023;117:68–77.

    Article  CAS  PubMed  Google Scholar 

  82. Foster BA, Coffey HA, Morin MJ, Rastinejad F. Pharmacological rescue of mutant p53 conformation and function. Science. 1999;286:2507–10.

    Article  CAS  PubMed  Google Scholar 

  83. Bykov VJN, Issaeva N, Shilov A, Hultcrantz M, Pugacheva E, Chumakov P, et al. Restoration of the tumor suppressor function to mutant p53 by a low-molecular-weight compound. Nat Med. 2002;8:282–8.

    Article  CAS  PubMed  Google Scholar 

  84. Bykov VJN, Zache N, Stridh H, Westman J, Bergman J, Selivanova G, et al. PRIMA-1(MET) synergizes with cisplatin to induce tumor cell apoptosis. Oncogene. 2005;24:3484–91.

    Article  CAS  PubMed  Google Scholar 

  85. Lambert JMR, Gorzov P, Veprintsev DB, Söderqvist M, Segerbäck D, Bergman J, et al. PRIMA-1 reactivates mutant p53 by covalent binding to the core domain. Cancer Cell. 2009;15:376–88.

    Article  CAS  PubMed  Google Scholar 

  86. Zhang Q, Bykov VJN, Wiman KG, Zawacka-Pankau J. APR-246 reactivates mutant p53 by targeting cysteines 124 and 277. Cell Death Dis. 2018;9:439.

    Article  PubMed  PubMed Central  Google Scholar 

  87. Synnott NC, O’Connell D, Crown J, Duffy MJ. COTI-2 reactivates mutant p53 and inhibits growth of triple-negative breast cancer cells. Breast Cancer Res Treat. 2020;179:47–56.

    Article  CAS  PubMed  Google Scholar 

  88. Chen S, Wu J-L, Liang Y, Tang Y-G, Song H-X, Wu L-L, et al. Arsenic trioxide rescues structural p53 mutations through a cryptic allosteric site. Cancer Cell. 2021;39:225-239.e8.

    Article  PubMed  Google Scholar 

  89. Liu DSH, Read M, Cullinane C, Azar WJ, Fennell CM, Montgomery KG, et al. APR-246 potently inhibits tumour growth and overcomes chemoresistance in preclinical models of oesophageal adenocarcinoma. Gut. 2015;64:1506–16.

    Article  CAS  PubMed  Google Scholar 

  90. Mohell N, Alfredsson J, Fransson Å, Uustalu M, Byström S, Gullbo J, et al. APR-246 overcomes resistance to cisplatin and doxorubicin in ovarian cancer cells. Cell Death Dis. 2015;6: e1794.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Ceder S, Eriksson SE, Cheteh EH, Dawar S, Corrales Benitez M, Bykov VJN, et al. A thiol-bound drug reservoir enhances APR-246-induced mutant p53 tumor cell death. EMBO Mol Med. 2021;13:e10852.

    Article  CAS  PubMed  Google Scholar 

  92. Peng X, Zhang MQZ, Conserva F, Hosny G, Selivanova G, Bykov VJN, et al. APR-246/PRIMA-1MET inhibits thioredoxin reductase 1 and converts the enzyme to a dedicated NADPH oxidase. Cell Death Dis. 2013;4:e881.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Wang Z, Hu H, Heitink L, Rogers K, You Y, Tan T, et al. The anti-cancer agent APR-246 can activate several programmed cell death processes to kill malignant cells. Cell Death Differ. 2023;30:1033–46.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Lehmann S, Bykov VJN, Ali D, Andrén O, Cherif H, Tidefelt U, et al. Targeting p53 in vivo: a first-in-human study with p53-targeting compound APR-246 in refractory hematologic malignancies and prostate cancer. J Clin Oncol. 2012;30:3633–9.

    Article  CAS  PubMed  Google Scholar 

  95. Cluzeau T, Sebert M, Rahmé R, Cuzzubbo S, Lehmann-Che J, Madelaine I, et al. Eprenetapopt plus azacitidine in TP53-mutated myelodysplastic syndromes and acute myeloid leukemia: A phase II study by the groupe Francophone des Myélodysplasies (GFM). J Clin Oncol. 2021;39:1575–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Sallman DA, DeZern AE, Garcia-Manero G, Steensma DP, Roboz GJ, Sekeres MA, et al. Eprenetapopt (APR-246) and azacitidine in TP53-mutant myelodysplastic syndromes. J Clin Oncol. 2021;39:1584–94.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Mishra A, Tamari R, DeZern AE, Byrne MT, Gooptu M, Chen Y-B, et al. Eprenetapopt plus azacitidine after allogeneic hematopoietic stem-cell transplantation for TP53-mutant acute myeloid leukemia and myelodysplastic syndromes. J Clin Oncol. 2022;40:3985–93.

    Article  CAS  PubMed  Google Scholar 

  98. Tuval A, Strandgren C, Heldin A, Palomar-Siles M, Wiman KG. Pharmacological reactivation of p53 in the era of precision anticancer medicine. Nat Rev Clin Oncol. 2024;21:106–20.

    Article  CAS  PubMed  Google Scholar 

  99. US National Library of Medicine. ClinicalTrials.gov. 2020;Available from: http://clinicaltrials.gov/study/NCT03745716?tab=results

  100. Garcia-Manero G, Goldberg AD, Winer ES, Altman JK, Fathi AT, Odenike O, 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:e272–83.

    Article  CAS  PubMed  Google Scholar 

  101. Kirschbaum M, Gojo I, Goldberg SL, Bredeson C, Kujawski LA, Yang A, et al. A phase 1 clinical trial of vorinostat in combination with decitabine in patients with acute myeloid leukaemia or myelodysplastic syndrome. Br J Haematol. 2014;167:185–93.

    Article  CAS  PubMed  Google Scholar 

  102. Juarez D, Fruman DA. Targeting the mevalonate pathway in cancer. Trends Cancer. 2021;7:525–40.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Parrales A, Ranjan A, Iyer SV, Padhye S, Weir SJ, Roy A, et al. DNAJA1 controls the fate of misfolded mutant p53 through the mevalonate pathway. Nat Cell Biol. 2016;18:1233–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Dumble M, Xu L, Dominique R, Liu B, Yang H, McBrayer M-K, et al. Abstract LB006: PC14586: The first orally bioavailable small molecule reactivator of Y220C mutant p53 in clinical development. In: Experimental and Molecular Therapeutics. Am Association Cancer Res; 2021;81:LB006–LB006.

  105. Oren M. Regulation of the p53 tumor suppressor protein. J Biol Chem. 1999;274:36031–4.

    Article  CAS  PubMed  Google Scholar 

  106. Quesnel B, Preudhomme C, Oscier D, Lepelley P, Collyn-d’Hooghe M, Facon T, et al. Over-expression of the MDM2 gene is found in some cases of haematological malignancies. Br J Haematol. 1994;88:415–8.

    Article  CAS  PubMed  Google Scholar 

  107. Quintás-Cardama A, Hu C, Qutub A, Qiu YH, Zhang X, Post SM, et al. p53 pathway dysfunction is highly prevalent in acute myeloid leukemia independent of TP53 mutational status. Leukemia. 2017;31:1296–305.

    Article  PubMed  Google Scholar 

  108. Han X, Medeiros LJ, Zhang YH, You MJ, Andreeff M, Konopleva M, et al. High expression of human homologue of Murine Double Minute 4 and the short splicing variant, HDM4-S, in bone marrow in patients with acute myeloid leukemia or myelodysplastic syndrome. Clin Lymphoma Myeloma Leuk. 2016;16(Suppl):S30–8.

    Article  PubMed  Google Scholar 

  109. Kussie PH, Gorina S, Marechal V, Elenbaas B, Moreau J, Levine AJ, et al. Structure of the MDM2 oncoprotein bound to the p53 tumor suppressor transactivation domain. Science. 1996;274:948–53.

    Article  ADS  CAS  PubMed  Google Scholar 

  110. Vassilev LT, Vu BT, Graves B, Carvajal D, Podlaski F, Filipovic Z, et al. In vivo activation of the p53 pathway by small-molecule antagonists of MDM2. Science. 2004;303:844–8.

    Article  ADS  CAS  PubMed  Google Scholar 

  111. Andreeff M, Kelly KR, Yee K, Assouline S, Strair R, Popplewell L, et al. Results of the phase I trial of RG7112, a small-molecule MDM2 antagonist in leukemia. Clin Cancer Res. 2016;22:868–76.

    Article  CAS  PubMed  Google Scholar 

  112. Konopleva M, Martinelli G, Daver N, Papayannidis C, Wei A, Higgins B, et al. MDM2 inhibition: an important step forward in cancer therapy. Leukemia. 2020;34:2858–74.

    Article  PubMed  Google Scholar 

  113. Yee K, Martinelli G, Vey N, Dickinson MJ, Seiter K, Assouline S, et al. Phase 1/1b study of RG7388, a potent MDM2 antagonist, in acute myelogenous leukemia (AML) patients (pts). Blood. 2014;124:116–116.

    Article  Google Scholar 

  114. Reis B, Jukofsky L, Chen G, Martinelli G, Zhong H, So WV, et al. Acute myeloid leukemia patients’ clinical response to idasanutlin (RG7388) is associated with pre-treatment MDM2 protein expression in leukemic blasts. Haematologica. 2016;101:e185–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Konopleva MY, Röllig C, Cavenagh J, Deeren D, Girshova L, Krauter J, et al. Idasanutlin plus cytarabine in relapsed or refractory acute myeloid leukemia: results of the MIRROS trial. Blood Adv. 2022;6:4147–56.

    CAS  PubMed  PubMed Central  Google Scholar 

  116. Sun D, Li Z, Rew Y, Gribble M, Bartberger MD, Beck HP, et al. Discovery of AMG 232, a potent, selective, and orally bioavailable MDM2-p53 inhibitor in clinical development. J Med Chem. 2014;57:1454–72.

    Article  CAS  PubMed  Google Scholar 

  117. Erba HP, Becker PS, Shami PJ, Grunwald MR, Flesher DL, Zhu M, et al. Phase 1b study of the MDM2 inhibitor AMG 232 with or without trametinib in relapsed/refractory acute myeloid leukemia. Blood Adv. 2019;3:1939–49.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  118. Arnhold V, Schmelz K, Proba J, Winkler A, Wünschel J, Toedling J, et al. Reactivating TP53 signaling by the novel MDM2 inhibitor DS-3032b as a therapeutic option for high-risk neuroblastoma. Oncotarget. 2018;9:2304–19.

    Article  PubMed  Google Scholar 

  119. DiNardo CD, Rosenthal J, Andreeff M, Zernovak O, Kumar P, Gajee R, et al. Phase 1 Dose Escalation Study of MDM2 Inhibitor DS-3032b in Patients with Hematological Malignancies - Preliminary Results. Blood. 2016;128:593–593.

    Article  Google Scholar 

  120. Senapati J, Muftuoglu M, Ishizawa J, Abbas HA, Loghavi S, Borthakur G, 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.

    Article  PubMed  PubMed Central  Google Scholar 

  121. Ferretti S, Rebmann R, Berger M, Santacroce F, Albrecht G, Pollehn K, et al. Abstract 1224 Insights into the mechanism of action of NVP-HDM201, a differentiated and versatile Next-Generation small-molecule inhibitor of Mdm2, under evaluation in phase I clinical trials In Experimental and Molecular Therapeutics. Am Assoc Cancer Res. 2016;76:1224–1224.

    Article  Google Scholar 

  122. Stein EM, DeAngelo DJ, Chromik J, Chatterjee M, Bauer S, Lin C-C, et al. Results from a First-in-Human Phase I Study of Siremadlin (HDM201) in Patients with Advanced Wild-Type TP53 Solid Tumors and Acute Leukemia. Clin Cancer Res. 2022;28:870–81.

    Article  CAS  PubMed  Google Scholar 

  123. Aguilar A, Lu J, Liu L, Du D, Bernard D, McEachern D, et al. Discovery of 4-((3’R,4’S,5’R)-6″-Chloro-4’-(3-chloro-2-fluorophenyl)-1’-ethyl-2″-oxodispiro[cyclohexane-1,2’-pyrrolidine-3’,3″-indoline]-5’-carboxamido)bicyclo[2.2.2]octane-1-carboxylic Acid (AA-115/APG-115): A potent and orally active murine Double Minute 2 (MDM2) inhibitor in clinical development. J Med Chem. 2017;60:2819–39.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Guerlavais V, Sawyer TK, Carvajal L, Chang YS, Graves B, Ren J-G, et al. Discovery of Sulanemadlin (ALRN-6924), the first cell-permeating, stabilized α-helical peptide in clinical development. J Med Chem. 2023;66:9401–17.

    Article  CAS  PubMed  Google Scholar 

  125. LoRusso P, Yamamoto N, Patel MR, Laurie SA, Bauer TM, Geng J, et al. The MDM2-p53 antagonist Brigimadlin (BI 907828) in patients with advanced or metastatic solid tumors: results of a phase Ia, first-in-human, dose-escalation study. Cancer Discov. 2023;13:1802–13.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Walcott FL, Wang P-Y, Bryla CM, Huffstutler RD, Singh N, Pollak MN, et al. Pilot study assessing tolerability and metabolic effects of metformin in patients with Li-Fraumeni Syndrome. JNCI Cancer Spectr. 2020;4:pkaa063.

    Article  PubMed  PubMed Central  Google Scholar 

  127. Pantziarka P, Blagden S. Inhibiting the priming for cancer in Li-Fraumeni Syndrome. Cancers (Basel). 2022;14:1621.

    Article  CAS  PubMed  Google Scholar 

  128. Buzzai M, Jones RG, Amaravadi RK, Lum JJ, DeBerardinis RJ, Zhao F, et al. Systemic treatment with the antidiabetic drug metformin selectively impairs p53-deficient tumor cell growth. Cancer Res. 2007;67:6745–52.

    Article  CAS  PubMed  Google Scholar 

  129. Zawacka-Pankau JE. The undervalued avenue to reinstate tumor suppressor functionality of the p53 protein family for improved cancer therapy-drug repurposing. Cancers (Basel). 2020;12:2717.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. Adams CM, Mitra R, Xiao Y, Michener P, Palazzo J, Chao A, et al. Targeted MDM2 degradation reveals a new vulnerability for p53-inactivated triple-negative breast cancer. Cancer Discov. 2023;13:1210–29.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Tiwary R, Yu W, Sanders BG, Kline K. α-TEA cooperates with chemotherapeutic agents to induce apoptosis of p53 mutant, triple-negative human breast cancer cells via activating p73. Breast Cancer Res. 2011;13:R1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. Wu H, Leng RP. MDM2 mediates p73 ubiquitination: a new molecular mechanism for suppression of p73 function. Oncotarget. 2015;6:21479–92.

    Article  PubMed  PubMed Central  Google Scholar 

  133. Venkatanarayan A, Raulji P, Norton W, Flores ER. Novel therapeutic interventions for p53-altered tumors through manipulation of its family members, p63 and p73. Cell Cycle. 2016;15:164–71.

    Article  CAS  PubMed  Google Scholar 

  134. Kobayashi T, Makino T, Yamashita K, Saito T, Tanaka K, Takahashi T, et al. APR-246 induces apoptosis and enhances chemo-sensitivity via activation of ROS and TAp73-Noxa signal in oesophageal squamous cell cancer with TP53 missense mutation. Br J Cancer. 2021;125:1523–32.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Jiang L, Malik N, Acedo P, Zawacka-Pankau J. Protoporphyrin IX is a dual inhibitor of p53/MDM2 and p53/MDM4 interactions and induces apoptosis in B-cell chronic lymphocytic leukemia cells. Cell Death Discov. 2019;5:77.

    Article  PubMed  PubMed Central  Google Scholar 

  136. Jiang L, Zawacka-Pankau J. The p53/MDM2/MDMX-targeted therapies-a clinical synopsis. Cell Death Dis. 2020;11:237.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Tuval A, Brilon Y, Azogy H, Moskovitz Y, Leshkowitz D, Salame TM, et al. Pseudo-mutant P53 is a unique phenotype of DNMT3A-mutated pre-leukemia. Haematologica. 2022;107:2548–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  138. Trinidad AG, Muller PAJ, Cuellar J, Klejnot M, Nobis M, Valpuesta JM, et al. Interaction of p53 with the CCT complex promotes protein folding and wild-type p53 activity. Mol Cell. 2013;50:805–17.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  139. Badar T, Atallah E, Shallis RM, Goldberg AD, Patel A, Abaza Y, et al. Outcomes of TP53-mutated AML with evolving frontline therapies: Impact of allogeneic stem cell transplantation on survival. Am J Hematol. 2022;97:E232–E235.

Download references

Acknowledgements

The author apologizes to other colleagues whose works were not cited due to space restrictions. The author would like to express special thank you note to Klas G. Wiman from Karolinska Institute for helpful discussions.

Funding

Open access funding provided by Karolinska Institute. The work was supported by Cathrine Everts forsknigsstiftelse (to JZ-P) and Polish National Science Center (Narodowe Centrum Nauki), grant OPUS 20 no. 2020/39/B/NZ7/00757 (to JZ-P).

Author information

Authors and Affiliations

Authors

Contributions

J.E.Z. conceptualisation, wrote a draft, drafted figures, collected data, led the project, prepared the final version, revised the manuscript.

Corresponding author

Correspondence to Joanna E. Zawacka.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zawacka, J.E. p53 biology and reactivation for improved therapy in MDS and AML. Biomark Res 12, 34 (2024). https://doi.org/10.1186/s40364-024-00579-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40364-024-00579-9

Keywords