Next Article in Journal
Fibrosis of Peritoneal Membrane, Molecular Indicators of Aging and Frailty Unveil Vulnerable Patients in Long-Term Peritoneal Dialysis
Next Article in Special Issue
Special Issue “Advances in Molecular Pathogenesis and Targeted Therapies for Myeloid Neoplasms”
Previous Article in Journal
Function and Evolution of the Loop Extrusion Machinery in Animals
Previous Article in Special Issue
Targeting Measurable Residual Disease (MRD) in Acute Myeloid Leukemia (AML): Moving beyond Prognostication
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Understanding the Continuum between High-Risk Myelodysplastic Syndrome and Acute Myeloid Leukemia

by
Phaedon D. Zavras
,
Ilias Sinanidis
,
Panagiotis Tsakiroglou
and
Theodoros Karantanos
*
Division of Hematologic Malignancies and Bone Marrow Transplantation, Sidney Kimmel Comprehensive Cancer Center, Johns Hopkins University, Baltimore, MD 21231, USA
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2023, 24(5), 5018; https://doi.org/10.3390/ijms24055018
Submission received: 31 January 2023 / Revised: 1 March 2023 / Accepted: 3 March 2023 / Published: 6 March 2023

Abstract

:
Myelodysplastic syndrome (MDS) is a clonal hematopoietic neoplasm characterized by bone marrow dysplasia, failure of hematopoiesis and variable risk of progression to acute myeloid leukemia (AML). Recent large-scale studies have demonstrated that distinct molecular abnormalities detected at earlier stages of MDS alter disease biology and predict progression to AML. Consistently, various studies analyzing these diseases at the single-cell level have identified specific patterns of progression strongly associated with genomic alterations. These pre-clinical results have solidified the conclusion that high-risk MDS and AML arising from MDS or AML with MDS-related changes (AML-MRC) represent a continuum of the same disease. AML-MRC is distinguished from de novo AML by the presence of certain chromosomal abnormalities, such as deletion of 5q, 7/7q, 20q and complex karyotype and somatic mutations, which are also present in MDS and carry crucial prognostic implications. Recent changes in the classification and prognostication of MDS and AML by the International Consensus Classification (ICC) and the World Health Organization (WHO) reflect these advances. Finally, a better understanding of the biology of high-risk MDS and the mechanisms of disease progression have led to the introduction of novel therapeutic approaches, such as the addition of venetoclax to hypomethylating agents and, more recently, triplet therapies and agents targeting specific mutations, including FLT3 and IDH1/2. In this review, we analyze the pre-clinical data supporting that high-risk MDS and AML-MRC share the same genetic abnormalities and represent a continuum, describe the recent changes in the classification of these neoplasms and summarize the advances in the management of patients with these neoplasms.

1. Introduction

Myelodysplastic syndrome (MDS) is a heterogeneous group of clonal hematopoietic stem cell disorders characterized by bone marrow dysplasia, hematopoiesis failure and high risk of progression to acute myeloid leukemia (AML) [1,2,3]. One out of three patients diagnosed with MDS will progress to AML, characterized by an increased percentage of blasts over 20% [4]. Based on the 2016 world classification of hematologic malignancies, AML arising from MDS was classified as a distinct clinicopathologic entity entitled “AML with MDS-related changes” (AML-MRC) [5]. AML-MRC is associated with an overall poor response to both induction chemotherapy and low-intensity therapy, a high incidence of relapse and worse overall survival compared to patients with de novo AML [6]. Due to the poor outcomes of individuals with this disease, it has been the focus of intense research at both the pre-clinical and clinical levels.
Recent large-scale studies have demonstrated that various cytogenetic abnormalities and gene mutations, such as deletion of 5q or 7q/7, mutations in spliceosome genes and genes encoding epigenetic modifiers, are common in MDS and AML-MRC but appear to be less frequent in de novo AML [7,8,9]. On the contrary, other genomic alterations, such as mutations in transcriptional factors or genes encoding proteins involved in signal transduction, are identified more frequently in AML-MRC compared to MDS [8,10,11], suggesting that these are probably later biological events during disease progression. Single-cell multi-omic studies have recently revealed specific patterns of clonal evolution driving the progression of MDS to AML-MRC [12,13], further highlighting that high-risk MDS and AML-MRC probably represent different stages of the same myeloid disease.
These research advances have significantly improved our understanding of the biology of high-risk MDS and its transformation to AML-MRC, leading to significant improvements in the classification of these presentations and a more comprehensive therapeutic approach incorporating novel agents and broadening opportunities for clinical trials. These changes are reflected in the recently updated classification of myeloid neoplasms by the International Consensus Classification (ICC) and the World Health Organization (WHO) [14,15]. Notably, the recognition of a continuum between high-risk MDS and AML-MRC has yielded the enrollment of high-risk MDS patients into AML clinical trials and the enrollment of AML-MRC patients into MDS clinical trials, which is expected to tremendously improve our understanding of this disease’s biology and the survival of these patients.
In this review, we summarize the existing pre-clinical studies that support the notion that high-risk MDS and AML-MRC represent a continuum and provide a novel understanding of the mechanisms of disease progression. Moreover, we describe the most recent changes in the classification of MDS and AML-MRC, as well as the implications of these advances in the management of patients with these myeloid neoplasms.

2. Genetic Alterations Underlie MDS Progression to AML-MRC

2.1. Chromosomal Abnormalities

Chromosomal abnormalities and copy number alterations are common in MDS and AML-MRC and have important prognostic implications. MDS and AML-MRC share various chromosomal abnormalities, such as deletion of 5q, 7/7q, 20q and complex karyotype, that typically lead to copy number alterations as opposed to de novo AML, which is frequently characterized by balanced re-arrangements, such as translocations 15;17 and 8;21 and inversion 16 [7]. Most of the chromosomal abnormalities identified in patients with MDS have critical prognostic implications and are associated with variable risk of disease progression.
Deletion of 5q is the most common chromosomal abnormality in MDS, particularly among patients with lower-risk disease, and has been associated with an overall good prognosis [16,17,18]. Hematopoietic cells with 5q deletion have defective ribosomal biogenesis, resulting in elevated levels of free ribosomal proteins in their cytoplasm, which bind and promote the degradation of MDM2, a key P53 regulator [19]. This leads to P53 activation, which promotes cell-cycle arrest and induces apoptosis of erythroid progenitors and ineffective erythropoiesis [20]. These molecular events render P53 activity critical for the prevention of this disease’s progression to AML. Indeed, analysis of 55 individuals with low or intermediate-1 MDS with 5q deletion showed that TP53 mutation is associated with an increased risk of AML transformation [21]. Interestingly, TP53 mutations were identified in the majority of patients at an early disease stage and were always detectable before AML transformation [21]. Thus, the presence of a TP53 mutation should be an alarm for a high risk of progression in MDS with 5q deletion, even if identified at an earlier disease stage. It should be also noted that MDS patients with a complex karyotype, including 5q deletion, have a poor response to chemotherapy or lenalidomide and a dismal prognosis, which are linked to genomic instability due to P53 dysfunction [22]. Thus, MDS with 5q deletion in the context of a complex karyotype should not be managed as a low-risk disease; it needs to be approached as a high-risk disease. Consistently, 5q deletion is associated with poor survival, high genomic complexity and biallelic TP53 variants among patients with high-risk MDS and AML-MRC [23].
Loss of chromosome 7 or 7q is the second most common chromosomal abnormality among MDS patients. It is associated with overall poor prognosis and a higher risk of progression to AML-MRC with worse outcomes for patients with loss of the entire chromosome 7 [24]. Similarly to 5q, the loss of 7q is associated with high genomic complexity and particularly short survival among patients with AML-MRC [23]. A number of genes with a critical role in the pathogenesis of myeloid malignancies, such as SAMD9, SAMD9L, EZH2, CUX1 and MLL3, are located on chromosome 7 [25]. Most of these genes encode proteins that act as tumor suppressors, and deletion or missense mutations provide survival and growth benefits to hematopoietic stem and progenitor cells [26,27]. Recent data also support the notion that loss of chromosome 7 or 7q is associated with significant alterations in the expression of interferon-gamma pathway genes, promoting an immunosuppressive microenvironment enriched in T regulatory immune cells, potentially conferring with the inferior outcomes of patients with myeloid neoplasms with these abnormalities [28]. Overall, these results support that loss of chromosome 7 or 7q contributes to the pathogenesis and, more importantly, to the progression of MDS to AML-MRC.
The presence of three or more chromosomal abnormalities is defined as a complex karyotype (CK) and is always considered a predictor of adverse outcomes, independent of actual specific abnormalities [29]. CK, particularly monosomal karyotype (MK), is strongly associated with one or more mutations in TP53 and frequently with loss of heterozygosity (LOH) of this gene [30,31,32]. A number of studies have highlighted that CK is associated with particularly poor survival and a high risk of transformation to AML among MDS patients, independently of blast percentage and other clinical features such as cytopenias and transfusion dependence [33,34,35]. These findings further support that the notion that the biology of MDS and its progression is defined by genetic alterations, probably at the earlier stages of this disease.

2.2. Somatic Mutations

Next-generation sequencing studies have demonstrated that MDS and AML-MRC share mutations in genes implicated in cellular processes, such as RNA splicing, epigenetic and transcriptional regulation, and DNA damage repair [8]. Based on an analysis of 299 AML patients, mutations in spliceosome genes such as SRSF2, SF3B1, ZRSF2 and U2AF1, EZH2, BCOR and STAG2 are >95% specific for AML-MRC and can distinguish AML-MRC from de novo AML, even in the absence of an antecedent MDS diagnosis [8]. The identification of mutations in genes implicated in different cellular processes highlights the progression of clonal myeloid neoplasms from earlier stages, such as clonal hematopoiesis of indeterminate potential (CHIP) to MDS and AML-MRC. Particularly, mutations in epigenetic modifiers, such as TET2 and DNMT3A and TP53, are common across a spectrum of clonal hematopoietic disorders, such as CHIP, clonal cytopenias, MDS and AML-MRC [36]. On the contrary, mutations in spliceosome genes are more common in MDS and AML-MRC [37,38], while mutations in transcriptional factors, such as RUNX1 and CEBPA, and activating signaling genes, such as NRAS and FLT3, are more common in AML-MRC [8,9]. These findings suggest that clonal evolution through the acquisition of additional somatic mutations probably drives the progression of these clonal myeloid neoplasms.
SF3B1 is commonly mutated in MDS [39], with the majority of mutations being missense substitutions affecting the spliceosome machinery, resulting in altered proteome [40]. SF3B1 mutations have been associated with MDS with ring sideroblasts and a relatively low risk of AML transformation [41]. However, specific substitutions, such as K666N hotspot mutation, result in distinct patterns of RNA splicing and increased risk of progression to AML-MRC [42]. SRSF2 mutations are found in 10–15% of patients with MDS and are associated with older age, higher levels of hemoglobin and a normal karyotype [43]. SRSF2 mutations frequently co-occur with RUNX1, IDH2 and ASXL1 mutations and are associated with a higher risk of transformation to AML-MRC [43,44]. The adverse biology of MDS with SRSF2 mutations could be associated with the altered splicing of genes such as EZH2 [45] and their implications in DNA damage accumulation through alterations of the P53 function [46]. U2AF1 mutations occur in about 5–10% of MDS patients and are associated with a high risk of disease progression to AML-MRC [47]. Aberrant splicing due to U2AF1 mutations results in suppression of ATG7 levels and reduced autophagy, causing increased oxidative stress and chromosomal instability [48]. A recent study also revealed that U2AF1 mutations cause an increase in the expression of the long isoform of interleukin-1 receptor-associated kinase 4 (IRAK4), which promotes the activation of nuclear factor kappa-light-chain-enhancer of B cells (NF-kB) that have been implicated in leukemic growth [49]. ZRSR2 is mutated in about 5% of patients with MDS, affect predominantly men and cause aberrant splicing by retaining U12-dependent introns [31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50]. While most studies indicate a negative impact of ZRSR2 mutations on the survival of MDS patients, the impact is not as strong compared to SRSF2 and U2AF1 mutations [51]. These results highlight that mutations in spliceosome genes can mediate leukemic growth and promote the progression of MDS to AML-MRC.
STAG2 encodes a subunit of the cohesin complex, which regulates the separation of sister chromatids during mitosis [52] and is mutated in close to 10% of patients with MDS [53] with male predominance [54]. STAG2 mutations have been associated with poor overall survival among MDS patients [53] but have not been directly linked to progression to AML-MRC. It has been hypothesized that the haploinsufficiency of genes encoding subunits of the cohesin complex affects the expression of genes that are essential for lineage priming and differentiation [55]. Thus, mutations in these genes likely cause abnormal hematopoietic stem cell maturation contributing to dysplasia, but cooperating oncogenic mutations are required for progression to AML-MRC [55]. Indeed, RUNX1, CEBPA, NPM1, and RAS mutations co-occur with STAG2 mutations in AML [9,56,57].
EZH2 is a core component of polycomb group complex 2 (PRC2), which is responsible for trimethylation of lysine 27 of histone 3 and plays a critical role in epigenetic gene silencing [58]. EZH2 is mutated in 4–5% of MDS patients [59] and is frequently co-mutated with ASXL1 and RUNX1 [60,61]. Concurrent TET2 and EZH2 deletion is sufficient to induce an MDS/myeloproliferative phenotype in mice [62], and in a RUNX1 mutant model, EZH2 loss promotes myelodysplasia development but inhibits transformation to AML [63]. In fact, EZH2 is upregulated in AML with CK [64] and EZH2 deletion suppresses the proliferation of MLL-AF9-transformed cells and delays AML transformation in transgenic mice [65]. Consistently, EZH2 mutations have been associated with poor overall survival in patients with chronic myeloid neoplasms [30,54,66] and poor response to hypomethylating agents [67] but not with a higher risk of AML-MRC transformation. Thus, similarly to STAG2 mutations, it can be hypothesized that EZH2 mutations are early events in MDS [68], and additional oncogenic mutations are required for transformation to AML-MRC.
The BCOR gene is located on the chromosome X and encodes transcription factor that is essential for normal embryonic development [69]. BCOR mutations are detected in less than 5% of patients with MDS and are typically frameshift insertions or deletions, or stopgain or non-sense mutations, causing a suppression of function of the BCOR protein in hematopoietic cells [70,71]. Recent studies have demonstrated that loss of BCOR function is associated with altered activity of polycomb group repressive complex 1 (PRC1), resulting in induced myeloid cells’ proliferation [71]. Most of the studies on large cohorts have demonstrated that BCOR mutations have a neutral impact on overall survival of MDS patients [72,73].
RUNX1 encodes the alpha subunit of the core binding transcription factor and is implicated in the differentiation of hematopoietic stem cells [74]. Germline point mutations of RUNX1 have been identified in an autosomal dominant platelet disorder with a high risk of transformation to AML [75]. RUNX1 mutations occur in 10–15% of MDS and have been associated with thrombocytopenia, poor survival with a high risk of progression to AML-MRC, co-occurrence of RAS mutations and loss of chromosome 7/7q [76,77,78]. Recent data have also associated RUNX1 mutations with poor response to hypomethylating agents [67,68,69,70,71,72,73,74,75,76,77,78,79]. Recently, it was reported that RUNX1 mutation is associated with rapid progression of low-risk MDS and that mutations in this gene in CD34+ cells from low-risk MDS patients result in dysregulated DNA damage repair and cellular senescence [78]. The authors also found that the transcriptional profiles of CD34+ cells from low-risk MDS patients with RUNX1 mutation resemble those of high-risk MDS at diagnosis [78]. These findings further suggest that specific genomic alterations identified even at the earlier stage of MDS can alter disease biology and induce disease rapid progression.
NRAS and KRAS mutations occur in about 4–5% of patients with MDS and are associated with higher white blood cell counts and a higher risk of transformation to AML-MRC [80,81,82,83]. Mutations in RAS genes result in RAS/Raf/MEK and PI3K/AKT signaling activation and are sufficient to induce a myeloproliferative phenotype or even AML in mice [84,85]. Similarly, FLT3-ITD mutations are relatively uncommon among MDS patients with a frequency of 0.6–6% [86] and have been associated with higher blasts percentage, increased risk of leukemic transformation and worse overall survival [87]. FLT3-ITD mutations in hematopoietic stem and progenitor cells cause constitutive autophosphorylation of the FLT3 receptor, leading to the activation of downstream signaling, such as STAT3/5, MAPK and PI3K/AKT [88]. Analysis of 278 MDS patients with low- or intermediate-1-risk MDS revealed that the detection of a RAS or FLT3-ITD mutation is associated with a particularly high risk of progression to AML-MRC with very poor response to treatment with hypomethylating agents [11]. Similarly, the detection of RAS or FLT3-ITD mutations at the time of transformation of MDS to AML-MRC is correlated with significantly worse survival [89]. Thus, independently of the disease stage at diagnosis, the acquisition of these signaling genes mutations significantly alters disease biology and is connected with poor outcomes.
TP53 is a tumor-suppressor gene located on the short arm of chromosome 17 and encodes transcription factor P53, which is critical for cell-cycle arrest, DNA repair and apoptosis in the setting of DNA damage [90]. As a result, cancers with TP53 mutation accumulate DNA damage and demonstrate a poor response to cytotoxic therapy [91]. TP53 alterations are discovered in about 5–10% of MDS patients [34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92], with the majority of cases having “multi-hit” involvement with more than one genomic and/or chromosome 17 abnormalities and fewer alterations in other genes [92]. A recent analysis of 3324 MDS patients revealed that high-risk presentation with CK, high incidence of transformation to AML-MRC and poor overall survival are associated with multi-hit TP53 and not monoallelic TP53 alterations [92]. It should be noted that these associations were independent of the Revised International Prognostic Scoring System (IPSS-R) [92], once more indicating that, independent of the disease stage at diagnosis, the molecular alterations that affect malignant cell biology define disease course and outcomes.
The incidence, biological mechanisms and clinical significance of genetic abnormalities in MDS and AML-MRC are summarized in Table 1.

2.3. Single-Cell Analysis Reveals Biologic Mechanisms Implicated in MDS Progression to AML-MRC

Single-cell DNA and RNA sequencing has significantly improved our understanding of heterogeneity and clonal evolution in myeloid neoplasms [94,95,96]. Chen et al. utilized longitudinal, paired samples from seven MDS patients who progressed to AML-MRC and performed ensemble deep sequencing and single-cell sequencing of sorted pre-malignant and malignant stem cells based on the expression of previously identified stem cell markers and blasts [95]. The authors demonstrated a significantly higher complexity of sub-clonal mutations in stem cells at the MDS stage compared to blast cells [95]. Moreover, the authors presented data that support a model of parallel clonal evolution at the stem cell level during MDS progression [95]. Particularly, all the patients studied demonstrated a highly diverse pool of pre-malignant MDS stem cells with most of the patients showing relatively early branching of these pre-malignant MDS stem cells towards progression to AML stem cells [95]. These findings suggest that the drivers of progression are probably present in MDS stem cells even at the earliest stages and dictate the natural history of this disease. Finally, the identification of actionable mutations in MDS stem cells could provide a particularly promising opportunity for the early prevention of disease progression to AML-MRC.
Guess et al. performed single-cell DNA sequencing of paired MDS and AML-MRC to better characterize clonal shifts upon MDS progression [12]. The authors found two different patterns of clonal evolution during MDS progression to AML-MRC [12]. The first pattern, called the static group, was characterized by relatively stable clonal architecture during progression [12]. In this group, patients had founder mutations in DNA methylation genes, such as DNMT3A, TET2 and IDH1/2, and potentially epigenetic alterations drive progression to AML-MRC through blast growth [12]. On the contrary, the second pattern, called the dynamic group, was characterized by significant clonal architectural changes defined by new chromosomal abnormalities and TP53 mutations or genomic alterations with more prominent mutations in signaling genes [12]. Single-cell transcriptomic analysis demonstrated that pathways such as cell-cycle regulation via E2F activation and inflammatory signaling mediated by cytokine receptors are amongst the top upregulated cellular pathways associated with transformation to AML-MRC [12]. These results are consistent with mechanistic studies highlighting that inflammatory signaling is linked with cell-cycle progression promoting the progression of MDS to AML-MRC and affecting the disease biology and sensitivity to current therapies [93,97,98,99,100,101].

3. Changes in the Definition and Classification of MDS and AML-MRC

The International Consensus Classification (ICC) and the World Health Organization (WHO) have recently updated the classification of myeloid neoplasms that come with a better understanding of the biology of the disease and therapeutic advancements. Compared to the WHO 2016 classification, the ICC introduced changes in the predefining blast cut-offs, and both committees incorporated genetic alterations of the diseases.

3.1. MDS and AML with Myelodysplasia-Related Gene Mutations as a Disease Continuum

The new guidelines recognize clonal hematopoiesis (CH) as the precursor of cytopenias, ranging from clonal cytopenia of undetermined significance (CCUS) to MDS. Cases with cytopenia but without dysplasia lack three cytogenetic abnormalities, namely del(5q), –7/del(7q) and complex karyotype, and are now classified under CCUS. The definitions of the pre-malignant clonal hematopoiesis disease spectrum are summarized in Table 2. The ICC and the WHO recognize MDS as myeloid neoplasms characterized by clonal hematopoiesis, dysplastic changes in the BM of PB and cytopenia in at least one cell line [15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102]. Clonality is established via next-generation sequencing (NGS) or karyotype. In cases where clonality cannot be proven, the diagnosis of MDS can still be made with qualifying dysplasia and cytopenia(s). On the other hand, irrespective of the presence of dysplasia, MDS-defining genetic abnormalities are sufficient for diagnosis, especially in the setting of persistent cytopenia.
Emphasis has been placed on the molecular categorization of MDS and AML in the new ICC guidelines because the genetic footprint seems to be a stronger predictor of prognosis than the degree of dysplasia [103].
The previously defined categories of MDS with excess blasts (EBs) have been refined. Recent data show that MDS with a higher blast percentage is more likely to transform to AML and carries a worse prognosis, similar to overt AML [104,105]. To more accurately reflect the continuum between MDS and AML, a new entity “MDS/AML” has been suggested to define MDS with blasts between 10 and 19%, lacking AML-defining genetics. Patients with MDS/AML should be eligible for either MDS or AML treatment modalities or clinical trials. MDS/AML is further subclassified into cases with mutated TP53, cases with myelodysplasia-related gene mutations, myelodysplasia-related cytogenetic abnormalities and MDS/AML-NOS. Blasts of greater than 20% of define AML in these cases. AML positive for the Philadelphia chromosome (Ph+) or BCR:ABL1/t(9;22)(q34.1;q11.2) is an exception and is still defined as blasts ≥20%. The MDS/AML category is not applicable in this case so as to distinguish it from the progression of Ph+ CML. For cases with 5–9% and 2–9% blasts in BM and PB, respectively, the term MDS with EB is used.
The previously defined category of AML-MRC has been eliminated by the ICC. Genetic characteristics, rather than clinical history (i.e., de novo AML, or AML arising from MDS, or therapy-related AML), are shown to be better primary determinants of disease classification [8]. Two new categories have been suggested instead, irrespective of any history of MDS, namely AML with mutated TP53 and AML with myelodysplasia-related gene mutations. The latter encompasses the prior entity of AML with mutated RUNX1, expanded to include the other eight genes shown to be related to AML arising from MDS or MDS/MPN as follows: ASXL1; BCOR; EZH2; SF3B1; SRSF2; STAG2; U2AF1; and ZRSR2. These mutations are strongly associated with a history of MDS or MDS/MPN and carry a poor prognosis [8].
In the absence of class-defining genetic abnormalities, a separate category of AML with myelodysplasia-related cytogenetic abnormalities incorporates cases with a complex karyotype (defined as ≥3 unrelated clonal chromosomal abnormalities) and/or other unbalanced chromosomal changes that previously fell into the prior AML-MRC category but lack a TP53 mutation or myelodysplasia-related gene mutations, i.e., del(5q)/t(5q)/add(5q), –7/del(7q), +8, del(12p)/t(12p)/add(12p), i(17q), –17/add(17p)/del(17p), del(20q), or idic(X)(q13) [106].
The WHO has retained the 20% blast cut-off between MDS and AML for disease without AML-defining mutations because altering the threshold was felt to be arbitrary and carries the risk of overtreatment. Hence, the family of MDS with increased blasts (IBs) includes disease with <20% blasts, and this is further subclassified into IB1 (5–9% BM and/or 2–4% PB blasts), IB2 (10–19% BM and/or 5–19% BM blasts), and MDS with fibrosis (in which case 5–19% BM and/or 2–19% PB blasts are allowed) [15]. The AML-MRC category is maintained in the WHO guidelines, and it encompasses de novo AML or AML transformed from MDS or MDS/MPN. A key change is that diagnosis can no longer be made solely based on morphology, and it requires the identification of genetic abnormalities and/or history of MDS or MDS/MPN. The group of single-gene somatic mutations again includes the same group of myelodysplasia-associated genes mentioned above, present in the vast majority of AML arising from MDS or MDS/MPN with the exception of RUNX1, claiming insufficient unifying characteristics [15].

3.2. TP53-Mutated Disease

As described above, MDS and AML with TP53 mutation are usually associated with a complex karyotype, carry a much worse prognosis and are recognized as separate disease entities by the ICC [9,102,108,109]. The TP53 mutation is an independent predicting factor of poor response to cytarabine or hypomethylating agent (HMA)/Venetoclax-based therapy [110]. MDS with mutated TP53 (i.e., blasts < 10%) requires the presence of a multi-hit TP53 mutation or TP53 mutations with variant allele frequency (VAF) > 10% and a complex karyotype, often with a loss of 17p. Cases with a complex karyotype but not a TP53 mutation do not qualify for this category as they have different, more favorable outcomes [111]. Multi-hit TP53 mutation is defined as two distinct TP53 mutations (each with VAF > 10%), or a single TP53 mutation with either 17p deletion (VAF > 50%) or loss of heterozygosity (LOH) in the 17p TP53 locus [92]. In the absence of LOH information, a single TP53 mutation in the context of a complex karyotype also qualifies for diagnosis. Monoallelic TP53-mutated MDS has different disease biology to multi-hit disease; thus, it is not included in this category and is instead categorized under MDS-NOS (if no EBs) or MDS with EB. On the contrary, monoallelic TP53-mutated MDS/AML or AML have a worse prognosis and are allowed in this category. TP53-mutated AML is not defined separately by the WHO [15].

3.3. Therapy-Related Disease

The therapy-related subgroup of MDS and AML is eliminated by the ICC, and therapy-related cases are now subclassified following primary diagnosis [102]. For example, MDS-NOS with single lineage dysplasia is therapy-related. This definition applies to therapy-related MDS, AML or MDS/AML, and AML progressed from MDS or MDS/MPN. Although it is recognized that secondary disease carries a worse prognosis, the priority is to first define the disease based on its morphologic and genetic features. The secondary myeloid neoplasm category has been preserved by the WHO and it includes therapy-related diseases (renamed to “myeloid neoplasm, post cytotoxic therapy or MN-pCT”), and diseases associated with germline predisposition (Down-syndrome-related AML falls into this category) [15]. Exposure to PARP inhibitors has been incorporated as a qualifying criterion for MN-pCT, whereas methotrexate exposure has been excluded, based on a lack of a significant association that has been reported [112,113]. Of note, AML transformed from another myeloid neoplasm is no longer classified as secondary by the WHO. AML arising from MDS or MDS/MPN is now classified under AML-MR (as per above) and AML arising from MPN is retained under the MPN category.

3.4. Other Changes Introduced by the ICC and the WHO

In the fourth edition of the WHO classification of myeloid neoplasms released in 2016, there was a significant overlap between the various AML subcategories [5]. A 20% blast cut-off was required for a diagnosis of AML, apart from cases of APL or AML with mutated CBF. Another major change introduced by both the ICC and the WHO groups is the expansion of AML-defining recurring genetic abnormalities to include t(9;11)(p21.3;q23.3)/MLLT3::KMT2A (or other KMT2A rearrangements), t(6;9)(p22.3;q34.1)/DEK::NUP214, inv(3)(q21.3q26.2)/MECOM(EVI1) (or other MECOM rearrangements) and NPM1, in-frame basic leucine zipper-region (bZIP) CEBPA mutations [114,115,116,117]. Biallelic CEBPA mutations are no longer required, with several studies showing favorable prognosis with monoallelic bZIP mutations [118,119]. In the case of NMP1-mutated AML, the WHO allows diagnosis irrespective of the blast percentage, highlighting the fact that NMP1-mutated neoplasms with blasts <20%, namely MDS and MDS/MPN cases, are associated with aggressive disease and rapid progression to overt AML within 12 months of diagnosis [120]. As newer, targeted therapies have emerged, early identification of the molecular subtype of the disease can lead to earlier initiation of treatment.
The classification algorithm of MDS and AML based on the new changes can be seen in Figure 1 and Figure 2.

4. Risk Stratification Changes in Myeloid Neoplasms

4.1. MDS

The IPSS-R has been used for years as the gold standard method for MDS risk stratification, as well as the numerous clinical implications this entails, such as clinical trial design and treatment modalities [121]. The IPSS-R is solely based on the hematologic and cytogenetic abnormalities of the disease and does not take into account individual mutations leading to clonality. As new knowledge around the significance of gene mutations in the prognostication of myeloid neoplasms has emerged, a new risk stratification scoring system has been recently suggested [30,31,73]. The IPSS-Molecular (IPSS-M) prognostic model was recently developed [122]. This new evidence is based on data from 2957 patients, from whom mutations of 152 different genes that were implicated in the development of myeloid neoplasms were assessed. The variables of the model were selected based on a stability selection algorithm that was applied to three different clinical outcomes as follows: overall survival (OS), leukemia-free survival (LFS) and AML transformation. As opposed to the IPSS-R, the absolute neutrophil count (ANC) is not considered part of the IPSS-M. At least one gene abnormality was found in 94% of the patients and the increasing number of mutations was negatively correlated with leukemia-free survival and thus disease severity [122].
This model uses hematologic (i.e., medullary blast percentage, hemoglobin, and platelet counts) and cytogenetic data (the five clusters that were included in the IPSS-R scoring system) and, in addition, includes molecular data, namely mutations identified in 16 “main effect” genes and 15 additional “residual” genes. Mutations in three of the main effect genes, TP53 multihit (but not mono-allelic TP53 mutations), MLLPTD (partial tandem duplication) and FLT3 (tyrosine kinase domain and internal tandem duplication) strongly correlated with adverse outcomes, highlighting the importance of screening for those genes at the time of diagnosis [122]. FLT3 and MLLPTD mutations exhibited the strongest correlation with AML transformation [122]. Other main effect genes included ASXL1, CBL, DNMT3A, ETV6, EZH2, IDH2, KRAS, NPM1, NRAS, RUNX1 and SF3B1 (computational pattern SF3B1a) or SF3B15q, SRSF2, and U2AF [122]. In the case of SF3B1 mutations, there are three different clusters with different prognostic outcomes based on the co-mutation pattern identified. SF3B15q (7% of SF3B1 mutant cases) refers to the concomitant presence of isolated 5q deletion; SF3B1β (15% of the cases) refers to co-mutation of SF3B1 along with any of the BCOR, BCORL1, NRAS, RUNX1, SRSF2, or STAG2 genes; and SF3B1α (78% of the cases) is defined as any other mutant SF3B1, in which case 37% of the patients were found to have simple co-mutation patterns, including one of the DNMT3A, TET2, and/or ASXL1 genes [122]. Favorable outcomes were observed for the SF3B1α group, whereas they were not for the other two groups [122]. Fifteen residual genes were also determined (BCOR, BCORL1, CEBPA, ETNK1, GATA2, GNB1, IDH1, NF1, PHF6, PPM1D, PRPF8, PTPN11, SETBP1, STAG2 and WT1) based on adverse effects identified by univariate analysis and greater than 1% recurrence among all patients with MDS [122].
Six disease risk strata were identified based on this new model (instead of five per IPSS-R), which correlated with leukemia-free survival, overall survival and transformation to AML, as follows: very low (14%); low (33%); moderate low (11%); moderate-high (11%); high (14%); and very high (17%). Almost half (46%) of the patients of the cohort were re-classified based on IPSS-M. Of these, 74% were upstaged and 26% were downstaged [122]. The IPSS-M was applied to a separate Japanese cohort of 718 patients, and its discriminative power in terms of LFS was found to be superior to the IPSS-R across all key endpoints. The IPSS-M is applicable to both de novo disease and treatment-related MDS. It is a dynamic metric, i.e., it can be applied at any point during treatment. Its use in predicting treatment outcomes is yet to be determined.

4.2. AML

Both risk stratification and treatment management of AML are based primarily on the age, the molecular and cytogenetic footprint, and early measurable residual disease (MRD) status [123]. MRD positivity following intensive or non-intensive induction therapy is a major, independent, post-diagnosis poor prognostic factor [124,125,126].
The European Leukemia Net (ELN) has introduced some changes in regard to the genetic risk stratification of AML [127].
  • FLT3-ITD-mutated AML irrespective of the allelic ratio (formerly FLT3-ITDhigh for biallelic mutation and FLT3-ITDlow for mono-allelic mutation) or the concurrent NMP1 mutation status is now considered intermediate risk disease. This is based on evidence of disease modifying impact of midostaurin-based therapy [128].
  • The previously recognized RUNX1 and ASXL1-mutated AML categories are expanded to include the whole AML category with gene-related mutations classified under high-risk disease. This new category encompasses the nine above-mentioned gene mutations (ASXL1, RUNX1, BCOR, EZH2, SF3B1, SRSF2, STAG2, U2AF1, ZRSR2) associated with AML arising from MDS.
  • NMP1-mutated AML with concurrent adverse-risk-associated cytogenetic abnormalities has been reclassified as a high-risk disease, based on new data that shows an association with worse outcomes [129].
  • AML with in-frame monoallelic mutations in the basic leucine zipper region (bZIP) of CEBPA is now also classified under favorable-risk disease, along with biallelic mutated disease [130].
  • The following cytogenetic abnormalities are added to the adverse risk group: t(3q26.2;v), involving the MECOM gene, and t(8;16)(p11.2;p13.3), associated with KAT6A::CREBBP gene fusion (39, 40). Last, hyperploidy with multiple trisomies or polysomies is no longer considered a complex karyotype and is therefore not classified as an adverse risk [14].

5. Current Management of Patients with High-Risk MDS and AML-MRC

Understanding the biology of MDS progression to AML-MRC is critical for the discovery of effective treatments and the development of a personalized therapeutic approach for patients with high-risk MDS and AML-MRC. Hypomethylating agents (HMAs) followed by allogeneic bone marrow transplantation (AlloBMT) remains the most frequently used approach [2]. Despite the requirement for further interpretation and exploitation of the new biological findings, significant progress has been made in the therapeutic approach of patients with MDS and AML-MRC. Venetoclax has now been approved in combination with hypomethylating agents by the United States Food and Drug Administration (FDA) [131], and targeted therapies, such as IDH1/2 inhibitors, have been effectively used in high-risk MDS and AML-MRC patients with IDH1/2 mutations [131]. Clinical trials testing novel agents, such as magrolimab and epretanoport, and anti-CD123 approaches in high-risk MDS and AML-MRC are currently ongoing [131]. Importantly, the new classifications of WHO and ICC have allowed clinicians to enroll high-risk MDS patients into AML clinical trials and AML-MRC patients into MDS clinical trials [15,132], which will undoubtedly expedite the results of the numerous ongoing clinical trials.

5.1. Current Standard Approaches

The current therapeutic approach for patients with MDS relies on (1) symptoms and clinical manifestation; (2) risk assessment using IPSS or Revised-IPSS or the most recently added IPSS-M calculator [122]; and (3) overall performance status and patient’s preference.

5.1.1. AlloBMT

AlloBMT constitutes the only potentially curative approach for patients in the clinical spectrum of high-risk MDS and AML-MRC. The decision of AlloBMT candidacy is based on factors such as age, physical and mental status, and comorbidities and should be accessed individually [133]. Specifically, age is not an absolute exclusive criterion but a relative one. Although younger patients (<65 years old) tend to have longer overall survival, there is no significant difference in survival in older patients (>65 years old) [134].
The timing of the AlloBMT remains a significant challenge in clinical practice. Cutler et al. demonstrated that the optimal point of alloBMT is the time of progression to intermediate risk as per the IPSS-R. Particularly, low- and very-low-risk patients should be monitored closely and they should undergo delayed transplantation, but intermediate to very-high-risk patients benefit from early transplantation [135].
It is not yet clear which regimen is more effective at inducing remission in patients as a bridge to alloBMT. HMAs or induction chemotherapy have been used for this purpose [136] with an ongoing clinical trial comparing these two approaches before AlloBMT in patients with MDS (NCT01812252). The RICMAC and MDS200 trial failed to show any superiority of intensive chemotherapy over low-dose chemotherapy in patients with high-risk MDS or in AML-MRC patients [137]. Interestingly, another important consideration for MDS and AML-MRC patients undergoing alloBMT is the conditioning regimen. Scott et al. showed that even though a MAC regimen leads to increased transplant-related mortality, 4-year overall and relapse-free survival were significantly better compared to RIC, with AML-MRC patients deriving increased benefits compared to MDS patients [138].
Despite initial enthusiasm, maintenance strategies following AlloBMT for patients with high-risk MDS and AML-MRC have not led to significant survival benefits [139,140,141], except for specific subsets, such as TP53-mutated diseases [142].

5.1.2. HMAs

The use of HMAs as first-line therapy is well established for patients with high-risk MDS and AML-MRC, and it is overall considered superior to intensive chemotherapy, partially due to a more tolerant and modest toxicity profile [143]. HMAs reduce the risk of death by 50% in high-risk MDS when used as first-line therapy in comparison with conventional care regimens, while most of the patients respond within 6 months of treatment [143]. Unfortunately, the HMA effect is transient, with responses usually maintained for less than 12 months, particularly in patients with adverse disease biology [144]. The prognosis of patients who progress to HMA is particularly dismal [144].
High-risk MDS and AML-MRC patients with poor response to HMAs showed a response rate of 41% and 32% when treated with chemotherapy and consequently underwent AlloBMT at a rate of 40% and 42%, respectively [145]. Despite the mechanisms of HMA resistance being the subject of intense pre-clinical research [101,146,147,148,149,150], novel combinational therapies based on these mechanisms have not yielded clinically significant results.
As opposed to post-AlloBMT maintenance, maintenance therapy with HMAs and particularly the oral formulation of azacitidine, CC-486 improve the outcomes of AML patients including a number of AML-MRC patients in first remission who did not receive AlloBMT [151].

5.1.3. Venetoclax

Venetoclax (VEN, ABT-199) is a BH3 mimetic that inhibits the anti-apoptotic effects of BCL-2 proteins, inducing cell death in AML blasts, and was approved by the FDA in 2018 for the treatment of older patients with AML who are not eligible for intensive chemotherapy [152]. Azacitidine/venetoclax combination outperformed azacitidine alone in both de novo AML and AML-MRC based on overall survival (14.7 months vs. 9.6 months) and complete-remission rates (66.4% vs. 28.3%) [153]. AML with IDH1/2 mutations demonstrated the highest sensitivity to the combination of azacitidine/venetoclax compared to azacitidine alone. Despite a more prolonged neutropenia and higher incidence of neutropenic fevers with combination therapy, the rate of discontinuation was similar between the two groups [153].
Recent evidence also supports that azacitidine/venetoclax is an effective combination in treatment-naïve high-risk MDS [154,155]. Liu et al. analyzed 1057 patients with high-risk MDS and AML-MRC treated with HMAs and venetoclax [156]. Fifty-six (56%) patients had complete remission and the overall response rate was 68% [156]. Subgroup analysis revealed that the complete-remission rate was 61% among MDS patients, 68% among patients with newly diagnosed AML-MRC and 39% for relapsed/refractory AML-MRC [156]. The combination of azacitidine/venetoclax is being evaluated in a phase 3 clinical trial for treatment-naïve high-risk MDS patients (NCT04401748).
The combination of azacitidine/venetoclax is also effective in relapsed/refractory MDS, as it outperformed azacitidine monotherapy and led to marrow response and complete-remission rates of 38.6% and 6.8%, respectively, within 1.5 months of treatment for approximately 9 months [157]. Of note, the response rate was stratified based on molecular profile rather than the conventional predicting tools (IPSS-R and blast percentage), with IDH2 and DNMT3A mutations showing the best response rate and TP53 mutations being associated with worse response rates.
Venetoclax-based regimens appear to be promising approaches as a bridge to alloBMT. It has been reported that high-risk MDS and AML-MRC patients responding to venetoclax-based therapy and then receiving AlloBMT have a 1-year overall survival rate of 75–79% with a relapse rate of 16.7% within the first 12–14 months [158,159].
It remains unclear if azacitidine/venetoclax is a better approach compared to induction chemotherapy for fit patients with high-risk MDS and AML-MRC. Ongoing clinical trials focus on comparing induction chemotherapy to the azacitidine/venetoclax combination, with emphasis on the outcomes of patients with AML-MRC (NCT04801797).

5.1.4. Chemotherapy

Low-dose cytarabine and induction chemotherapy have been evaluated for the management of patients with AML-MRC with responses of approximately 50–60% but a median survival of 6.5 months [160]. Prior treatment with HMAs or lenalidomide and longer time to transform to AML-MRC were associated with worse response [160]. A recent retrospective analysis showed high response rates of induction chemotherapy, followed by AlloBMT in MDS and MDS/MPN patients with NPM1 mutations [116], suggesting that molecular profile may be critical for the identification of MDS/AML-MRC patients that may benefit from induction therapy. Lancet et al. compared the activity of CPX-351, a dual-drug liposomal encapsulation of cytarabine and daunorubicin that delivers a synergistic 5:1 drug ratio into leukemia cells to a greater extent than normal bone marrow cells, to the activity of 7 + 3 in older individuals with AML-MRC [161]. CPX-351 demonstrated significantly improved remission rates and overall survival, despite a prolonged time to neutrophil and platelet engraftment [161]. This survival benefit of CPX-351 over 7 + 3 was maintained after 5 years of follow-up based on a subsequent study [162]. Overall, these results indicate that chemotherapy may be an option for fit individuals with high-risk MDS with elevated blasts or AML-MRC.

5.2. Specific Molecular Subtypes

As described above, the molecular profile of MDS and AML-MRC define the disease biology and affect its natural history independent of the disease stage at diagnosis. Consistently, treatment approaches for specific molecular subtypes are currently the focus of intense research.

5.2.1. TP53 Mutated High-Risk MDS and AML-MRC

Biallelic TP53 alteration is one of the most established adverse genomic subtypes of high-risk MDS and AML-MRC and, as described above is associated with particularly poor outcomes. HMAs and HMAs/venetoclax combinations show some efficacy in TP53-mutated high-risk MDS and AML-MCR [153,163], but the incidence of relapse and long-term overall survival remains disappointing [164]. Similarly, it remains unclear if chemotherapy provides any significant benefit for TP53-mutated high-risk MDS and AML-MRC patients [91].
AlloBMT is the only approach that can prolong the survival of TP53-mutated MDS and AML-MRC patients [110]. However, TP53-mutated patients, compared with TP53 wild-type patients, showed a higher relapse rate after alloBMT [34]. The achievement of a lower burden of TP53 mutation before alloBMT has been associated with improved survival [34].
These results clearly support that TP53-mutated MDS and AML-MRC patients represent an unmet need. Recent studies have led to the introduction of two novel agents with promising efficacy for these patients.
Eprenetapopt (ARP-246) is a small molecule that stabilizes when converted to methylene quinuclidinone, the misfolded P53 core domain, and restores the function of the mutated protein [165,166,167]. Pre-clinical studies supported that APR-246 has synergistic effects when combined with azacitidine, ara-c and daunorubicin in TP53-mutated AML cells [168,169]. Despite the encouraging findings of two phase I/II trials showing that the APR-246/azacitidine combination resulted in high response and complete-remission rates, as well as high rates of TP53-mutation clearance [170,171], the survival benefit of this combination over azacitidine alone for TP53-mutated high-risk MDS/AML-MRC has not been confirmed. However, recent data support that the APR-246/azacitidine combination has promising efficacy as a maintenance therapy following AlloBMT for TP53-mutated MDS and AML patients [142]. Finally, evaluation of the triplet combination azacitidine/venetoclax/APR-246 is ongoing in high-risk MDS and AML-MRC with adverse cytogenetic and molecular features [172].
Magrolimab is a monoclonal antibody that targets the CD47 surface marker of tumor cells, also known as the “don’t eat me” signal, promoting their phagocytosis by macrophages [173]. Among TP53-mutated MDS and AML-MRC patients, the combination of azacitidine/magrolimab showed a response rate of about 40%, which is similar to TP53-unmutated patients with anemia due to hemolysis being a well-described side effect [174]. Of note, the combination of magrolimab/azacitidine/venetoclax demonstrated complete-remission rates higher than 80% in newly diagnosed AML patients with ELN adverse risk, suggesting that this approach may become the new standard of care for high-risk MDS/AML-MRC patients who are ineligible for induction chemotherapy but can tolerate this triple therapy [175]. Magrolimab-based therapies are currently being compared to standard of care, such as azacitidine and azacitidine/venetoclax, in various ongoing phase three clinical trials (NCT05079230, NCT04313881, NCT04435691, NCT04778397, NCT04435691).

5.2.2. IDH1/2 Mutated High-Risk MDS and AML-MRC

IDH1/2 mutations are relatively uncommon in the chronic phase of MDS, but their incidence increases with the transformation to AML-MRC and are associated with overall poor outcomes [176,177]. The IDH1 inhibitor ivosidenib results in high complete-remission rates in patients with IDH1-mutated relapsed/refractory AML [178] and newly diagnosed AML [179]. Phase III clinical trials have shown that the addition of ivosidenib to azacitidine significantly increases remission rates and prolongs the survival of patients with IDH1-mutated AML, including AML-MRC [180]. This led to the FDA approval of this combination for patients with IDH1-mutated AML who are ineligible for induction chemotherapy. Enasidenib is an IDH2 inhibitor that leads to response rates of 30–50% in relapsed/refractory IDH2-mutated AML, including AML-MRC [181,182]. Enasidenib/azacitidine was evaluated for high-risk IDH2-mutated MDS, demonstrating a response rate of 74% and a complete remission of 26% in 27% of patients proceeding to alloBMT and survival after 2 years [182]. Finally, enasidenib has been evaluated as maintenance therapy for high-risk MDS and AML-MRC patients with IDH2 mutation with overall, progression-free and relapse rates being 74%, 69% and 16%, respectively, and relatively rare severe adverse effects [183]. Ongoing clinical trials evaluate the efficacy of IDH1/2 inhibitors in MDS and AML-MRC patients (NCT03503409, NCT04603001, NCT04493164, NCT03471260) and their combination with venetoclax, azacitidine and intensive chemotherapy (NCT03471260, NCT03839771).

6. Conclusions

High-risk MDS is a clonal myeloid neoplasm characterized by genetic abnormalities in stem and progenitor hematopoietic cells with a high incidence of disease progression to a disease stage with a high percentage of blasts defined as AML-MRC or AML with myelodysplasia-related gene mutations or MDS/AML. Recent pre-clinical data support that genetic alterations, such as chromosomal abnormalities and somatic mutations, are detectable in hematopoietic cells even at the earlier stages of the disease and predict disease progression defining its natural history and biology. These results suggest that MDS and AML-MRC represent a continuum of the same disease and suggest that early-risk stratification and intervention may be a particularly promising approach for patients with MDS and a high risk of progression. These advances are reflected in the new changes in the classification of MDS and AML and are utilized in the evaluation of novel therapies.

Author Contributions

Conceptualization, T.K.; literature Review P.D.Z., I.S., P.T. and T.K.; writing—original draft preparation, P.D.Z., I.S., P.T. and T.K.; writing—review and editing, supervision, T.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the ASH Research Training Award and the MacMillan Pathway to Independence Award.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Nimer, S.D. Myelodysplastic syndromes. Blood 2008, 111, 4841–4851. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Cazzola, M. Myelodysplastic Syndromes. N. Engl. J. Med. 2020, 383, 1358–1374. [Google Scholar] [CrossRef]
  3. Karantanos, T.; DeZern, A.E. Biology and clinical management of hypoplastic MDS: MDS as a bone marrow failure syndrome. Best Pract. Res. Clin. Haematol. 2021, 34, 101280. [Google Scholar] [CrossRef] [PubMed]
  4. Menssen, A.J.; Walter, M.J. Genetics of progression from MDS to secondary leukemia. Blood 2020, 136, 50–60. [Google Scholar] [CrossRef]
  5. Arber, D.A.; Orazi, A.; Hasserjian, R.; Thiele, J.; Borowitz, M.J.; Le Beau, M.M.; Bloomfield, C.D.; Cazzola, M.; Vardiman, J.W. The 2016 revision to the World Health Organization classification of myeloid neoplasms and acute leukemia. Blood 2016, 127, 2391–2405. [Google Scholar] [CrossRef] [PubMed]
  6. Granfeldt Østgård, L.S.; Medeiros, B.C.; Sengeløv, H.; Nørgaard, M.; Andersen, M.K.; Dufva, I.H.; Friis, L.S.; Kjeldsen, E.; Marcher, C.W.; Preiss, B.; et al. Epidemiology and Clinical Significance of Secondary and Therapy-Related Acute Myeloid Leukemia: A National Population-Based Cohort Study. J. Clin. Oncol. 2015, 33, 3641–3649. [Google Scholar] [CrossRef] [PubMed]
  7. Le Beau, M.M.; Espinosa, R.; Davis, E.M.; Eisenbart, J.D.; Larson, R.; Green, E.D. Cytogenetic and molecular delineation of a region of chromosome 7 commonly deleted in malignant myeloid diseases. Blood 1996, 88, 1930–1935. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Lindsley, R.C.; Mar, B.G.; Mazzola, E.; Grauman, P.V.; Shareef, S.; Allen, S.L.; Pigneux, A.; Wetzler, M.; Stuart, R.K.; Erba, H.P.; et al. Acute myeloid leukemia ontogeny is defined by distinct somatic mutations. Blood 2015, 125, 1367–1376. [Google Scholar] [CrossRef] [Green Version]
  9. 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]
  10. Shiozawa, Y.; Malcovati, L.; Galli’, A.; Pellagatti, A.; Karimi, M.; Sato-Otsubo, A.; Satoru, M.; Suzuki, H.; Yoshizato, T.; Yoshida, K.; et al. Gene expression and risk of leukemic transformation in myelodysplasia. Blood 2017, 130, 2642–2653. [Google Scholar] [CrossRef]
  11. Takahashi, K.; Jabbour, E.; Wang, X.; Luthra, R.; Buesoramos, C.; Patel, K.K.; Pierce, S.; Yang, H.; Wei, Y.; Daver, N.; et al. Dynamic acquisition of FLT3 or RAS alterations drive a subset of patients with lower risk MDS to secondary AML. Leukemia 2013, 27, 2081–2083. [Google Scholar] [CrossRef] [Green Version]
  12. Guess, T.; Potts, C.R.; Bhat, P.; Cartailler, J.A.; Brooks, A.; Holt, C.; Yenamandra, A.; Wheeler, F.C.; Savona, M.R.; Cartailler, J.-P.; et al. Distinct Patterns of Clonal Evolution Drive Myelodysplastic Syndrome Progression to Secondary Acute Myeloid Leukemia. Blood Cancer Discov. 2022, 3, 316–329. [Google Scholar] [CrossRef] [PubMed]
  13. Ganan-Gomez, I.; Yang, H.; Ma, F.; Montalban-Bravo, G.; Thongon, N.; Marchica, V.; Richard-Carpentier, G.; Chien, K.; Manyam, G.; Wang, F.; et al. Stem cell architecture drives myelodysplastic syndrome progression and predicts response to venetoclax-based therapy. Nat. Med. 2022, 28, 557–567. [Google Scholar] [CrossRef]
  14. Chilton, L.; Hills, R.; Harrison, C.; Burnett, A.K.; Grimwade, D.; Moorman, A.V. Hyperdiploidy with 49–65 chromosomes represents a heterogeneous cytogenetic subgroup of acute myeloid leukemia with differential outcome. Leukemia 2013, 28, 321–328. [Google Scholar] [CrossRef]
  15. 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]
  16. Giagounidis, A.A.; Germing, U.; Aul, C. Biological and prognostic significance of chromosome 5q deletions in myeloid malignancies. Clin. Cancer Res. 2006, 12, 5–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Solé, F.; Espinet, B.; Sanz, G.F.; Cervera, J.; Calasanz, M.J.; Luño, E.; Prieto, F.; Granada, I.; Hernández, J.M.; Cigudosa, J.C.; et al. Incidence, characterization and prognostic significance of chromosomal abnormalities in 640 patients with primary myelodysplastic syndromes. Br. J. Haematol. 2000, 108, 346–356. [Google Scholar] [CrossRef]
  18. Ebert, B.L. Genetic deletions in AML and MDS. Best Pract. Res. Clin. Haematol. 2010, 23, 457–461. [Google Scholar] [CrossRef] [Green Version]
  19. Barlow, J.L.; Drynan, L.F.; Hewett, D.R.; Holmes, L.; Lorenzo-Abalde, S.; Lane, A.L.; E Jolin, H.; Pannell, R.; Middleton, A.J.; Wong, S.H.; et al. A p53-dependent mechanism underlies macrocytic anemia in a mouse model of human 5q– syndrome. Nat. Med. 2009, 16, 59–66. [Google Scholar] [CrossRef] [Green Version]
  20. Boultwood, J. The role of haploinsufficiency of RPS14 and p53 activation in the molecular pathogenesis of the 5q- syndrome. Pediatr. Rep. 2011, 3, 25–27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Jädersten, M.; Saft, L.; Smith, A.; Kulasekararaj, A.; Pomplun, S.; Göhring, G.; Hedlund, A.; Hast, R.; Schlegelberger, B.; Porwit, A.; et al. TP53 mutations in low-risk myelodysplastic syndromes with del(5q) predict disease progression. J. Clin. Oncol. 2011, 29, 1971–1979. [Google Scholar] [CrossRef] [PubMed]
  22. Rasmussen, B.; Göhring, G.; Bernard, E.; Nilsson, L.; Tobiasson, M.; Jädersten, M.; Garelius, H.; Dybedal, I.; Grønbaek, K.; Ejerblad, E.; et al. “Randomized phase II study of azacitidine ± lenalidomide in higher-risk myelodysplastic syndromes and acute myeloid leukemia with a karyotype including Del(5q)”. Leukemia 2022, 36, 1436–1439. [Google Scholar] [CrossRef] [PubMed]
  23. Pitel, B.A.; Sharma, N.; Zepeda-Mendoza, C.; Smadbeck, J.B.; Pearce, K.E.; Cook, J.M.; Vasmatzis, G.; Sachs, Z.; Kanagal-Shamanna, R.; Viswanatha, D.; et al. Myeloid malignancies with 5q and 7q deletions are associated with extreme genomic complexity, biallelic TP53 variants, and very poor prognosis. Blood Cancer J. 2021, 11, 18. [Google Scholar] [CrossRef]
  24. Cordoba, I.; González-Porras, J.R.; Nomdedeu, B.; Luño, E.; de Paz, R.; Such, E.; Tormo, M.; Vallespi, T.; Collado, R.; Xicoy, B.; et al. Better prognosis for patients with del(7q) than for patients with monosomy 7 in myelodysplastic syndrome. Cancer 2011, 118, 127–133. [Google Scholar] [CrossRef]
  25. Inaba, T.; Honda, H.; Matsui, H. The enigma of monosomy 7. Blood 2018, 131, 2891–2898. [Google Scholar] [CrossRef]
  26. Nagamachi, A.; Matsui, H.; Asou, H.; Ozaki, Y.; Aki, D.; Kanai, A.; Takubo, K.; Suda, T.; Nakamura, T.; Wolff, L.; et al. Haploinsufficiency of SAMD9L, an endosome fusion facilitator, causes myeloid malignancies in mice mimicking human diseases with monosomy 7. Cancer Cell 2013, 24, 305–317. [Google Scholar] [CrossRef] [Green Version]
  27. Wong, C.C.; Chronic Myeloid Disorders Working Group of the International Cancer Genome Consortium; Martincorena, I.; Rust, A.; Rashid, M.; Alifrangis, C.; Alexandrov, L.B.; Tiffen, J.C.; Kober, C.; Green, A.; et al. Inactivating CUX1 mutations promote tumorigenesis. Nat. Genet. 2013, 46, 33–38. [Google Scholar] [CrossRef] [PubMed]
  28. Abbas, H.A.; Hao, D.; Tomczak, K.; Barrodia, P.; Im, J.S.; Reville, P.K.; Alaniz, Z.; Wang, W.; Wang, R.; Wang, F.; et al. Single cell T cell landscape and T cell receptor repertoire profiling of AML in context of PD-1 blockade therapy. Nat. Commun. 2021, 12, 6071. [Google Scholar] [CrossRef]
  29. Schanz, J.; Steidl, C.; Fonatsch, C.; Pfeilstöcker, M.; Nösslinger, T.; Tuechler, H.; Valent, P.; Hildebrandt, B.; Giagounidis, A.; Aul, C.; et al. Coalesced multicentric analysis of 2,351 patients with myelodysplastic syndromes indicates an underestimation of poor-risk cytogenetics of myelodysplastic syndromes in the international prognostic scoring system. J. Clin. Oncol. 2011, 29, 1963–1970. [Google Scholar] [CrossRef] [Green Version]
  30. Bejar, R.; Stevenson, K.; Abdel-Wahab, O.; Galili, N.; Nilsson, B.; Garcia-Manero, G.; Kantarjian, H.; Raza, A.; Levine, R.L.; Neuberg, D.; et al. Clinical effect of point mutations in myelodysplastic syndromes. N. Engl. J. Med. 2011, 364, 2496–2506. [Google Scholar] [CrossRef] [Green Version]
  31. Papaemmanuil, E.; Gerstung, M.; Malcovati, L.; Tauro, S.; Gundem, G.; Van Loo, P.; Yoon, C.J.; Ellis, P.; Wedge, D.C.; Pellagatti, A.; et al. Clinical and biological implications of driver mutations in myelodysplastic syndromes. Blood 2013, 122, 3616–3627. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Haferlach, C.; Dicker, F.; Herholz, H.; Schnittger, S.; Kern, W. Mutations of the TP53 gene in acute myeloid leukemia are strongly associated with a complex aberrant karyotype. Leukemia 2008, 22, 1539–1541. [Google Scholar] [CrossRef]
  33. Bejar, R.; Stevenson, K.E.; Caughey, B.; Lindsley, R.C.; Mar, B.G.; Stojanov, P.; Getz, G.; Steensma, D.P.; Ritz, J.; Soiffer, R.; et al. Somatic mutations predict poor outcome in patients with myelodysplastic syndrome after hematopoietic stem-cell transplantation. J. Clin. Oncol. 2014, 32, 2691–2698. [Google Scholar] [CrossRef] [Green Version]
  34. Lindsley, R.C.; Saber, W.; Mar, B.G.; Redd, R.; Wang, T.; Haagenson, M.D.; Grauman, P.V.; Hu, Z.-H.; Spellman, S.R.; Lee, S.J.; et al. Prognostic Mutations in Myelodysplastic Syndrome after Stem-Cell Transplantation. N. Engl. J. Med. 2017, 376, 536–547. [Google Scholar] [CrossRef] [Green Version]
  35. Haase, D.; Stevenson, K.E.; Neuberg, D.; Maciejewski, J.P.; Nazha, A.; Sekeres, M.A.; Ebert, B.L.; Garcia-Manero, G.; Haferlach, C.; Haferlach, T.; et al. TP53 mutation status divides myelodysplastic syndromes with complex karyotypes into distinct prognostic subgroups. Leukemia 2019, 33, 1747–1758. [Google Scholar] [CrossRef] [Green Version]
  36. Makishima, H.; Yoshizato, T.; Yoshida, K.; Sekeres, M.A.; Radivoyevitch, T.; Suzuki, H.; Przychodzen, B.; Nagata, Y.; Meggendorfer, M.; Sanada, M.; et al. Dynamics of clonal evolution in myelodysplastic syndromes. Nat. Genet. 2017, 49, 204–212. [Google Scholar] [CrossRef]
  37. Saez, B.; Walter, M.J.; Graubert, T.A. Splicing factor gene mutations in hematologic malignancies. Blood 2017, 129, 1260–1269. [Google Scholar] [CrossRef] [Green Version]
  38. Sperling, A.; Gibson, C.J.; Ebert, A.S.S.C.J.G.B.L. The genetics of myelodysplastic syndrome: From clonal haematopoiesis to secondary leukaemia. Nat. Rev. Cancer 2016, 17, 5–19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Haferlach, T.; Nagata, Y.; Grossmann, V.; Okuno, Y.; Bacher, U.; Nagae, G.; Schnittger, S.; Sanada, M.; Kon, A.; Alpermann, T.; et al. Landscape of genetic lesions in 944 patients with myelodysplastic syndromes. Leukemia 2014, 28, 241–247. [Google Scholar] [CrossRef] [Green Version]
  40. Dalton, W.B.; Helmenstine, E.; Walsh, N.; Gondek, L.P.; Kelkar, D.S.; Read, A.; Natrajan, R.; Christenson, E.S.; Roman, B.; Das, S.; et al. Hotspot SF3B1 mutations induce metabolic reprogramming and vulnerability to serine deprivation. J. Clin. Investig. 2019, 129, 4708–4723. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Papaemmanuil, E.; Cazzola, M.; Boultwood, J.; Malcovati, L.; Vyas, P.; Bowen, D.; Pellagatti, A.; Wainscoat, J.S.; Hellstrom-Lindberg, E.; Gambacorti-Passerini, C.; et al. SomaticSF3B1Mutation in myelodysplasia with ring sideroblasts. N. Engl. J. Med. 2011, 365, 1384–1395. [Google Scholar] [CrossRef] [Green Version]
  42. Dalton, W.B.; Helmenstine, E.; Pieterse, L.; Li, B.; Gocke, C.D.; Donaldson, J.; Xiao, Z.; Gondek, L.P.; Ghiaur, G.; Gojo, I.; et al. The K666N mutation in SF3B1 is associated with increased progression of MDS and distinct RNA splicing. Blood Adv. 2020, 4, 1192–1196. [Google Scholar] [CrossRef]
  43. Hou, H.-A.; Tsai, C.-H.; Lin, C.-C.; Chou, W.-C.; Kuo, Y.-Y.; Liu, C.-Y.; Tseng, M.-H.; Peng, Y.-L.; Liu, M.-C.; Liu, C.-W.; et al. Incorporation of mutations in five genes in the revised International Prognostic Scoring System can improve risk stratification in the patients with myelodysplastic syndrome. Blood Cancer J. 2018, 8, 39. [Google Scholar] [CrossRef] [Green Version]
  44. Zheng, X.; Zhan, Z.; Naren, D.; Li, J.; Yan, T.; Gong, Y. Prognostic value of SRSF2 mutations in patients with de novo myelodysplastic syndromes: A meta-analysis. PLoS ONE 2017, 12, e0185053. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Kim, E.; Ilagan, J.O.; Liang, Y.; Daubner, G.M.; Lee, S.C.W.; Ramakrishnan, A.; Li, Y.; Chung, Y.R.; Micol, J.-B.; Murphy, M.E.; et al. SRSF2 Mutations Contribute to Myelodysplasia by Mutant-Specific Effects on Exon Recognition. Cancer Cell 2015, 27, 617–630. [Google Scholar] [CrossRef] [Green Version]
  46. Kon, A.; Yamazaki, S.; Nannya, Y.; Kataoka, K.; Ota, Y.; Nakagawa, M.M.; Yoshida, K.; Shiozawa, Y.; Morita, M.; Yoshizato, T.; et al. Physiological Srsf2 P95H expression causes impaired hematopoietic stem cell functions and aberrant RNA splicing in mice. Blood 2018, 131, 621–635. [Google Scholar] [CrossRef] [PubMed]
  47. Wu, S.-J.; Tang, J.-L.; Lin, C.-T.; Kuo, Y.-Y.; Li, L.-Y.; Tseng, M.-H.; Huang, C.-F.; Lai, Y.-J.; Lee, F.-Y.; Liu, M.-C.; et al. Clinical implications of U2AF1 mutation in patients with myelodysplastic syndrome and its stability during disease progression. Am. J. Hematol. 2013, 88, E277–E282. [Google Scholar] [CrossRef] [PubMed]
  48. Park, S.M.; Ou, J.; Chamberlain, L.; Simone, T.M.; Yang, H.; Virbasius, C.-M.; Ali, A.M.; Zhu, L.J.; Mukherjee, S.; Raza, A.; et al. U2AF35(S34F) Promotes Transformation by Directing Aberrant ATG7 Pre-mRNA 3′ End Formation. Mol. Cell 2016, 62, 479–490. [Google Scholar] [CrossRef] [Green Version]
  49. Smith, M.A.; Choudhary, G.S.; Pellagatti, A.; Choi, K.; Bolanos, L.C.; Bhagat, T.D.; Gordon-Mitchell, S.; Von Ahrens, D.; Pradhan, K.; Steeples, V.; et al. U2AF1 mutations induce oncogenic IRAK4 isoforms and activate innate immune pathways in myeloid malignancies. Nature 2019, 21, 640–650. [Google Scholar] [CrossRef]
  50. Madan, V.; Kanojia, D.; Li, J.; Okamoto, R.; Sato-Otsubo, A.; Kohlmann, A.; Sanada, M.; Grossmann, V.; Sundaresan, J.; Shiraishi, Y.; et al. Aberrant splicing of U12-type introns is the hallmark of ZRSR2 mutant myelodysplastic syndrome. Nat. Commun. 2015, 6, 6042. [Google Scholar] [CrossRef] [Green Version]
  51. Wang, X.; Song, X.; Yan, X. Effect of RNA splicing machinery gene mutations on prognosis of patients with MDS: A meta-analysis. Medicine 2019, 98, e15743. [Google Scholar] [CrossRef]
  52. Haering, C.H.; Löwe, J.; Hochwagen, A.; Nasmyth, K. Molecular Architecture of SMC Proteins and the Yeast Cohesin Complex. Mol. Cell 2002, 9, 773–788. [Google Scholar] [CrossRef] [PubMed]
  53. Thota, S.; Viny, A.D.; Makishima, H.; Spitzer, B.; Radivoyevitch, T.; Przychodzen, B.; Sekeres, M.A.; Levine, R.L.; Maciejewski, J.P. Genetic alterations of the cohesin complex genes in myeloid malignancies. Blood 2014, 124, 1790–1798. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Karantanos, T.; Gondek, L.P.; Varadhan, R.; Moliterno, A.R.; DeZern, A.E.; Jones, R.J.; Jain, T. Gender-related differences in the outcomes and genomic landscape of patients with myelodysplastic syndrome/myeloproliferative neoplasm overlap syndromes. Br. J. Haematol. 2021, 193, 1142–1150. [Google Scholar] [CrossRef]
  55. Viny, A.D.; Levine, R.L. Cohesin mutations in myeloid malignancies made simple. Curr. Opin. Hematol. 2018, 25, 61–66. [Google Scholar] [CrossRef]
  56. Patel, J.L.; Schumacher, J.A.; Frizzell, K.; Sorrells, S.; Shen, W.; Clayton, A.; Jattani, R.; Kelley, T.W. Coexisting and cooperating mutations in NPM1-mutated acute myeloid leukemia. Leuk. Res. 2017, 56, 7–12. [Google Scholar] [CrossRef]
  57. Dolnik, A.; Engelmann, J.C.; Scharfenberger-Schmeer, M.; Mauch, J.; Kelkenberg-Schade, S.; Haldemann, B.; Fries, T.; Krönke, J.; Kühn, M.W.M.; Paschka, P.; et al. Commonly altered genomic regions in acute myeloid leukemia are enriched for somatic mutations involved in chromatin remodeling and splicing. Blood 2012, 120, e83–e92. [Google Scholar] [CrossRef] [Green Version]
  58. Christofides, A.; Karantanos, T.; Bardhan, K.; Boussiotis, V.A. Epigenetic regulation of cancer biology and anti-tumor immunity by EZH2. Oncotarget 2016, 7, 85624–85640. [Google Scholar] [CrossRef] [Green Version]
  59. Stasik, S.; Middeke, J.M.; Kramer, M.; Röllig, C.; Krämer, A.; Scholl, S.; Hochhaus, A.; Crysandt, M.; Brümmendorf, T.H.; Naumann, R.; et al. EZH2 mutations and impact on clinical outcome: An analysis in 1,604 patients with newly diagnosed acute myeloid leukemia. Haematologica 2019, 105, e228–e231. [Google Scholar] [CrossRef] [Green Version]
  60. Triviai, I.; Zeschke, S.; Rentel, J.; Spanakis, M.; Scherer, T.; Gabdoulline, R.; Panagiota, V.; Thol, F.; Heuser, M.; Stocking, C.; et al. ASXL1/EZH2 mutations promote clonal expansion of neoplastic HSC and impair erythropoiesis in PMF. Leukemia 2018, 33, 99–109. [Google Scholar] [CrossRef] [PubMed]
  61. Gaidzik, V.I.; Teleanu, V.; Papaemmanuil, E.; Weber, D.; Paschka, P.; Hahn, J.; Wallrabenstein, T.; Kolbinger, B.; Köhne, C.H.; Horst, H.A.; et al. RUNX1 mutations in acute myeloid leukemia are associated with distinct clinico-pathologic and genetic features. Leukemia 2016, 30, 2160–2168. [Google Scholar] [CrossRef]
  62. Muto, T.; Sashida, G.; Oshima, M.; Wendt, G.R.; Mochizuki-Kashio, M.; Nagata, Y.; Sanada, M.; Miyagi, S.; Saraya, A.; Kamio, A.; et al. Concurrent loss of Ezh2 and Tet2 cooperates in the pathogenesis of myelodysplastic disorders. J. Exp. Med. 2013, 210, 2627–2639. [Google Scholar] [CrossRef]
  63. Sashida, G.; Harada, H.; Matsui, H.; Oshima, M.; Yui, M.; Harada, Y.; Tanaka, S.; Mochizuki-Kashio, M.; Wang, C.; Saraya, A.; et al. Ezh2 loss promotes development of myelodysplastic syndrome but attenuates its predisposition to leukaemic transformation. Nat. Commun. 2014, 5, 4177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Grubach, L.; Juhl-Christensen, C.; Rethmeier, A.; Olesen, L.H.; Aggerholm, A.; Hokland, P.; Østergaard, M. Gene expression profiling of Polycomb, Hox and Meis genes in patients with acute myeloid leukaemia. Eur. J. Haematol. 2008, 81, 112–122. [Google Scholar] [CrossRef]
  65. Tanaka, S.; Miyagi, S.; Sashida, G.; Chiba, T.; Yuan, J.; Mochizuki-Kashio, M.; Suzuki, Y.; Sugano, S.; Nakaseko, C.; Yokote, K.; et al. Ezh2 augments leukemogenicity by reinforcing differentiation blockage in acute myeloid leukemia. Blood 2012, 120, 1107–1117. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Bejar, R.; Stevenson, K.E.; Caughey, B.A.; Abdel-Wahab, O.; Steensma, D.P.; Galili, N.; Raza, A.; Kantarjian, H.; Levine, R.L.; Neuberg, D.; et al. Validation of a prognostic model and the impact of mutations in patients with lower-risk myelodysplastic syndromes. J. Clin. Oncol. 2012, 30, 3376–3382. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Karantanos, T.; Tsai, H.-L.; Gondek, L.P.; DeZern, A.E.; Ghiaur, G.; Dalton, W.B.; Gojo, I.; Prince, G.T.; Webster, J.; Ambinder, A.; et al. Genomic landscape of myelodysplastic/myeloproliferative neoplasm can predict response to hypomethylating agent therapy. Leuk. Lymphoma 2022, 63, 1942–1948. [Google Scholar] [CrossRef] [PubMed]
  68. Rinke, J.; Müller, J.P.; Blaess, M.F.; A Chase, A.; Meggendorfer, M.; Schäfer, V.; Winkelmann, N.; Haferlach, C.; Cross, N.C.P.; A Hochhaus, A.; et al. Molecular characterization of EZH2 mutant patients with myelodysplastic/myeloproliferative neoplasms. Leukemia 2017, 31, 1936–1943. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Ng, D.; Thakker, N.; Corcoran, C.M.; Donnai, D.; Perveen, R.; Schneider, A.; Hadley, D.W.; Tifft, C.; Zhang, L.; Wilkie, A.O.M.; et al. Oculofaciocardiodental and Lenz microphthalmia syndromes result from distinct classes of mutations in BCOR. Nat. Genet. 2004, 36, 411–416. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Abuhadra, N.; Mukherjee, S.; Al-Issa, K.; Adema, V.; Hirsch, C.M.; Advani, A.; Przychodzen, B.; Makhoul, A.; Awada, H.; Maciejewski, J.P.; et al. BCOR and BCORL1 mutations in myelodysplastic syndromes (MDS): Clonal architecture and impact on outcomes. Leuk. Lymphoma 2019, 60, 1587–1590. [Google Scholar] [CrossRef] [PubMed]
  71. Cao, Q.; Gearhart, M.D.; Gery, S.; Shojaee, S.; Yang, H.; Sun, H.; Lin, D.-C.; Bai, J.-W.; Mead, M.; Zhao, Z.; et al. BCOR regulates myeloid cell proliferation and differentiation. Leukemia 2016, 30, 1155–1165. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Nazha, A.; Narkhede, M.; Radivoyevitch, T.; Seastone, D.J.; Patel, B.J.; Gerds, A.T.; Mukherjee, S.; Kalaycio, M.; Advani, A.; Przychodzen, B.; et al. Incorporation of molecular data into the Revised International Prognostic Scoring System in treated patients with myelodysplastic syndromes. Leukemia 2016, 30, 2214–2220. [Google Scholar] [CrossRef] [PubMed]
  73. Nazha, A.; Al-Issa, K.; Hamilton, B.K.; Radivoyevitch, T.; Gerds, A.; Mukherjee, S.; Adema, V.; Zarzour, A.; Abuhadra, N.; Patel, B.J.; et al. Adding molecular data to prognostic models can improve predictive power in treated patients with myelodysplastic syndromes. Leukemia 2017, 31, 2848–2850. [Google Scholar] [CrossRef]
  74. Ichikawa, M.; Yoshimi, A.; Nakagawa, M.; Nishimoto, N.; Watanabe-Okochi, N.; Kurokawa, M. A role for RUNX1 in hematopoiesis and myeloid leukemia. Int. J. Hematol. 2013, 97, 726–734. [Google Scholar] [CrossRef]
  75. Song, W.-J.; Sullivan, M.G.; Legare, R.D.; Hutchings, S.; Tan, X.; Kufrin, D.; Ratajczak, J.; Resende, I.C.; Haworth, C.; Hock, R.; et al. Haploinsufficiency of CBFA2 causes familial thrombocytopenia with propensity to develop acute myelogenous leukaemia. Nat. Genet. 1999, 23, 166–175. [Google Scholar] [CrossRef]
  76. Chen, C.-Y.; Lin, L.-I.; Tang, J.-L.; Ko, B.-S.; Tsay, W.; Chou, W.-C.; Yao, M.; Wu, S.-J.; Tseng, M.-H.; Tien, H.-F. RUNX1 gene mutation in primary myelodysplastic syndrome—The mutation can be detected early at diagnosis or acquired during disease progression and is associated with poor outcome. Br. J. Haematol. 2007, 139, 405–414. [Google Scholar] [CrossRef] [PubMed]
  77. Wang, K.; Zhou, F.; Cai, X.; Chao, H.; Zhang, R.; Chen, S. Mutational landscape of patients with acute myeloid leukemia or myelodysplastic syndromes in the context of RUNX1 mutation. Hematology 2020, 25, 211–218. [Google Scholar] [CrossRef]
  78. Kaisrlikova, M.; Vesela, J.; Kundrat, D.; Votavova, H.; Merkerova, M.D.; Krejcik, Z.; Divoky, V.; Jedlicka, M.; Fric, J.; Klema, J.; et al. RUNX1 mutations contribute to the progression of MDS due to disruption of antitumor cellular defense: A study on patients with lower-risk MDS. Leukemia 2022, 36, 1898–1906. [Google Scholar] [CrossRef]
  79. Wu, P.; Weng, J.; Li, M.; Lu, Z.; Deng, C.; Sun, Q.; Xu, R.; Geng, S.; Du, X. Co-occurrence of RUNX1 and ASXL1 mutations underlie poor response and outcome for MDS patients treated with HMAs. Am. J. Transl. Res. 2019, 11, 3651–3658. [Google Scholar]
  80. Nakagawa, T.; Saitoh, S.; Imoto, S.; Itoh, M.; Tsutsumi, M.; Hikiji, K.; Nakamura, H.; Matozaki, S.; Ogawa, R.; Nakao, Y.; et al. Multiple point mutation of N-ras and K-ras oncogenes in myelodysplastic syndrome and acute myelogenous leukemia. Oncology 1992, 49, 114–122. [Google Scholar] [CrossRef]
  81. Paquette, R.L.; Landaw, E.M.; Pierre, R.V.; Kahan, J.; Lübbert, M.; Lazcano, O.; Isaac, G.; McCormick, F.; Koeffler, H.P. N-ras mutations are associated with poor prognosis and increased risk of leukemia in myelodysplastic syndrome. Blood 1993, 82, 590–599. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Al-Kali, A.; Quintás-Cardama, A.; Luthra, R.; Bueso-Ramos, C.; Pierce, S.; Kadia, T.; Borthakur, G.; Estrov, Z.; Jabbour, E.; Faderl, S.; et al. Prognostic impact of RAS mutations in patients with myelodysplastic syndrome. Am. J. Hematol. 2013, 88, 365–369. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Padua, R.; Guinn, B.-A.; Al-Sabah, A.; Smith, M.; Taylor, C.; Pettersson, T.; Ridge, S.; Carter, G.; White, D.; Oscier, D.; et al. RAS, FMS and p53 mutations and poor clinical outcome in myelodysplasias: A 10-year follow-up. Leukemia 1998, 12, 887–892. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Zhang, J.; Wang, J.; Liu, Y.; Sidik, H.; Young, K.H.; Lodish, H.F.; Fleming, M.D. Oncogenic Kras-induced leukemogeneis: Hematopoietic stem cells as the initial target and lineage-specific progenitors as the potential targets for final leukemic transformation. Blood 2009, 113, 1304–1314. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Parikh, C.; Subrahmanyam, R.; Ren, R. Oncogenic NRAS rapidly and efficiently induces CMML- and AML-like diseases in mice. Blood 2006, 108, 2349–2357. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Bacher, U.; Haferlach, T.; Kern, W.; Haferlach, C.; Schnittger, S. A comparative study of molecular mutations in 381 patients with myelodysplastic syndrome and in 4130 patients with acute myeloid leukemia. Haematologica 2007, 92, 744–752. [Google Scholar] [CrossRef]
  87. Daver, N.; Strati, P.; Jabbour, E.; Kadia, T.; Luthra, R.; Wang, S.; Patel, K.; Ravandi, F.; Cortes, J.; Dong, X.Q.; et al. FLT3 mutations in myelodysplastic syndrome and chronic myelomonocytic leukemia. Am. J. Hematol. 2012, 88, 56–59. [Google Scholar] [CrossRef] [Green Version]
  88. Levis, M.; Small, D. FLT3: ITDoes matter in leukemia. Leukemia 2003, 17, 1738–1752. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Badar, T.; Patel, K.P.; Thompson, P.A.; DiNardo, C.; Takahashi, K.; Cabrero, M.; Borthakur, G.; Cortes, J.; Konopleva, M.; Kadia, T.; et al. Detectable FLT3-ITD or RAS mutation at the time of transformation from MDS to AML predicts for very poor outcomes. Leuk. Res. 2015, 39, 1367–1374. [Google Scholar] [CrossRef] [Green Version]
  90. Lakin, N.D.; Jackson, S.P. Regulation of p53 in response to DNA damage. Oncogene 1999, 18, 7644–7655. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Daver, N.G.; Maiti, A.; Kadia, T.M.; Vyas, P.; Majeti, R.; Wei, A.H.; Garcia-Manero, G.; Craddock, C.; Sallman, D.A.; Kantarjian, H.M. TP53-Mutated Myelodysplastic Syndrome and Acute Myeloid Leukemia: Biology, Current Therapy, and Future Directions. Cancer Discov. 2022, 12, 2516–2529. [Google Scholar] [CrossRef]
  92. Bernard, E.; Nannya, Y.; Hasserjian, R.P.; Devlin, S.M.; Tuechler, H.; Medina-Martinez, J.S.; Yoshizato, T.; Shiozawa, Y.; Saiki, R.; Malcovati, L.; et al. Implications of TP53 allelic state for genome stability, clinical presentation and outcomes in myelodysplastic syndromes. Nat. Med. 2020, 26, 1549–1556. [Google Scholar] [CrossRef]
  93. Muto, T.; Walker, C.S.; Choi, K.; Hueneman, K.; Smith, M.A.; Gul, Z.; Garcia-Manero, G.; Ma, A.; Zheng, Y.; Starczynowski, D.T. Adaptive response to inflammation contributes to sustained myelopoiesis and confers a competitive advantage in myelodysplastic syndrome HSCs. Nat. Immunol. 2020, 21, 535–545. [Google Scholar] [CrossRef]
  94. Kong, T.; Laranjeira, A.B.A.; Yang, K.; Fisher, D.A.C.; Yu, L.; De La Frégonnière, L.P.; Wang, A.Z.; Ruzinova, M.B.; Fowles, J.S.; Fulbright, M.C.; et al. DUSP6 mediates resistance to JAK2 inhibition and drives leukemic progression. Nat. Cancer 2022, 4, 108–127. [Google Scholar] [CrossRef]
  95. Chen, J.; Kao, Y.-R.; Sun, D.; Todorova, T.I.; Reynolds, D.; Narayanagari, S.-R.; Montagna, C.; Will, B.; Verma, A.; Steidl, U. Myelodysplastic syndrome progression to acute myeloid leukemia at the stem cell level. Nat. Med. 2018, 25, 103–110. [Google Scholar] [CrossRef] [PubMed]
  96. Dillon, L.W.; Ghannam, J.; Nosiri, C.; Gui, G.; Goswami, M.; Calvo, K.R.; Lindblad, K.E.; Oetjen, K.A.; Wilkerson, M.D.; Soltis, A.R.; et al. Personalized Single-Cell Proteogenomics to Distinguish Acute Myeloid Leukemia from Nonmalignant Clonal Hematopoiesis. Blood Cancer Discov. 2021, 2, 319–325. [Google Scholar] [CrossRef] [PubMed]
  97. Barreyro, L.; Sampson, A.M.; Ishikawa, C.; Hueneman, K.M.; Choi, K.; Pujato, M.A.; Chutipongtanate, S.; Wyder, M.; Haffey, W.D.; O’Brien, E.; et al. Blocking UBE2N abrogates oncogenic immune signaling in acute myeloid leukemia. Sci. Transl. Med. 2022, 14. [Google Scholar] [CrossRef]
  98. Agarwal, P.; Li, H.; Choi, K.; Hueneman, K.; He, J.; Welner, R.S.; Starczynowski, D.T.; Bhatia, R. TNF-α-induced alterations in stromal progenitors enhance leukemic stem cell growth via CXCR2 signaling. Cell Rep. 2021, 36, 109386. [Google Scholar] [CrossRef] [PubMed]
  99. Schinke, C.; Giricz, O.; Carolina, S.; Shastri, A.; Gordon, S.; Barreyro, L.; Bhagat, T.; Bhattacharyya, S.; Ramachandra, N.; Bartenstein, M.; et al. IL8-CXCR2 pathway inhibition as a therapeutic strategy against MDS and AML stem cells. Blood 2015, 125, 3144–3152. [Google Scholar] [CrossRef] [Green Version]
  100. Karantanos, T.; Teodorescu, P.; Perkins, B.; Christodoulou, I.; Esteb, C.; Varadhan, R.; Helmenstine, E.; Rajkhowa, T.; Paun, B.C.; Bonifant, C.; et al. The role of the atypical chemokine receptor CCRL2 in myelodysplastic syndrome and secondary acute myeloid leukemia. Sci. Adv. 2022, 8, eabl8952. [Google Scholar] [CrossRef]
  101. Karantanos, T.; Teodorescu, P.; Arvanitis, M.; Perkins, B.; Jain, T.; DeZern, A.E.; Dalton, W.B.; Christodoulou, I.; Paun, B.C.; Varadhan, R.; et al. CCRL2 affects the sensitivity of myelodysplastic syndrome and secondary acute myeloid leukemia cells to azacytidine. Haematologica 2022. [Google Scholar] [CrossRef]
  102. 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]
  103. Ogawa, S. Genetics of MDS. Blood 2019, 133, 1049–1059. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Chen, X.; Othus, M.; Wood, B.L.; Walter, R.B.; Becker, P.S.; Percival, M.-E.; Abkowitz, J.L.; Appelbaum, F.R.; Estey, E.H. 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 2020, 62, 1226–1233. [Google Scholar] [CrossRef] [PubMed]
  105. DiNardo, C.D.; Garcia-Manero, G.; Pierce, S.; Nazha, A.; Bueso-Ramos, C.; Jabbour, E.; Ravandi, F.; Cortes, J.; Kantarjian, H. Interactions and relevance of blast percentage and treatment strategy among younger and older patients with acute myeloid leukemia (AML) and myelodysplastic syndrome (MDS). Am. J. Hematol. 2016, 91, 227–232. [Google Scholar] [CrossRef] [Green Version]
  106. Fang, H.; He, R.; Chiu, A.; Viswanatha, D.S.; Ketterling, R.P.; Patnaik, M.S.; Reichard, K.K. Genetic Factors in Acute Myeloid Leukemia with Myelodysplasia-Related Changes. Am. J. Clin. Pathol. 2020, 153, 656–663. [Google Scholar] [CrossRef]
  107. Malcovati, L.; Galli’, A.; Travaglino, E.; Ambaglio, I.; Rizzo, E.; Molteni, E.; Elena, C.; Ferretti, V.V.; Catricalà, S.; Bono, E.; et al. Clinical significance of somatic mutation in unexplained blood cytopenia. Blood 2017, 129, 3371–3378. [Google Scholar] [CrossRef]
  108. Rücker, F.G.; Schlenk, R.F.; Bullinger, L.; Kayser, S.; Teleanu, V.; Kett, H.; Habdank, M.; Kugler, C.-M.; Holzmann, K.; Gaidzik, V.I.; et al. TP53 alterations in acute myeloid leukemia with complex karyotype correlate with specific copy number alterations, monosomal karyotype, and dismal outcome. Blood 2012, 119, 2114–2121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. 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]
  110. Short, N.J.; Montalban-Bravo, G.; Hwang, H.; Ning, J.; Franquiz, M.J.; Kanagal-Shamanna, R.; Patel, K.P.; DiNardo, C.D.; Ravandi, F.; Garcia-Manero, G.; et al. Prognostic and therapeutic impacts of mutant TP53 variant allelic frequency in newly diagnosed acute myeloid leukemia. Blood Adv. 2020, 4, 5681–5689. [Google Scholar] [CrossRef]
  111. Weinberg, O.K.; Siddon, A.J.; Madanat, Y.F.; Gagan, J.; Arber, D.A.; Cin, P.D.; Narayanan, D.; Ouseph, M.M.; Kurzer, J.H.; Hasserjian, R.P. TP53 mutation defines a unique subgroup within complex karyotype de novo and therapy-related MDS/AML. Blood Adv. 2022, 6, 2847–2853. [Google Scholar] [CrossRef] [PubMed]
  112. Csizmar, C.M.; Saliba, A.N.; Swisher, E.M.; Kaufmann, S.H. PARP Inhibitors and Myeloid Neoplasms: A Double-Edged Sword. Cancers 2021, 13, 6385. [Google Scholar] [CrossRef]
  113. Ertz-Archambault, N.; Kosiorek, H.; Taylor, G.E.; Kelemen, K.; Dueck, A.; Castro, J.; Marino, R.; Gauthier, S.; Finn, L.; Sproat, L.Z.; et al. Association of Therapy for Autoimmune Disease with Myelodysplastic Syndromes and Acute Myeloid Leukemia. JAMA Oncol. 2017, 3, 936–943. [Google Scholar] [CrossRef] [PubMed]
  114. Cui, W.; Sun, J.; Cotta, C.V.; Medeiros, L.J.; Lin, P.; Singh, S.; Sharma, S.; Jain, M.; Chauhan, R. Myelodysplastic syndrome with inv(3)(q21q26.2) or t(3;3)(q21;q26.2) has a high risk for progression to acute myeloid leukemia. Am. J. Clin. Pathol. 2011, 136, 282–288. [Google Scholar] [CrossRef] [Green Version]
  115. Fang, H.; Yabe, M.; Zhang, X.; Kim, Y.; Wu, X.; Wei, P.; Chi, S.; Zheng, L.; Garcia-Manero, G.; Shao, L.; et al. Myelodysplastic syndrome with t(6;9)(p22;q34.1)/DEK-NUP214 better classified as acute myeloid leukemia? A multicenter study of 107 cases. Mod. Pathol. 2021, 34, 1143–1152. [Google Scholar] [CrossRef]
  116. Montalban-Bravo, G.; Kanagal-Shamanna, R.; Sasaki, K.; Patel, K.; Gañán-Gómez, I.; Jabbour, E.; Kadia, T.; Ravandi, F.; Dinardo, C.; Borthakur, G.; et al. NPM1 mutations define a specific subgroup of MDS and MDS/MPN patients with favorable outcomes with intensive chemotherapy. Blood Adv. 2019, 3, 922–933. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Issa, G.C.; Zarka, J.; Sasaki, K.; Qiao, W.; Pak, D.; Ning, J.; Short, N.J.; Haddad, F.; Tang, Z.; Patel, K.P.; et al. Predictors of outcomes in adults with acute myeloid leukemia and KMT2A rearrangements. Blood Cancer J. 2021, 11, 162. [Google Scholar] [CrossRef]
  118. Taube, F.; Georgi, J.A.; Kramer, M.; Stasik, S.; Middeke, J.M.; Röllig, C.; Krug, U.; Krämer, A.; Scholl, S.; Hochhaus, A.; et al. CEBPA mutations in 4708 patients with acute myeloid leukemia: Differential impact of bZIP and TAD mutations on outcome. Blood 2022, 139, 87–103. [Google Scholar] [CrossRef]
  119. Wakita, S.; Sakaguchi, M.; Oh, I.; Kako, S.; Toya, T.; Najima, Y.; Doki, N.; Kanda, J.; Kuroda, J.; Mori, S.; et al. Prognostic impact of CEBPA bZIP domain mutation in acute myeloid leukemia. Blood Adv. 2022, 6, 238–247. [Google Scholar] [CrossRef]
  120. Forghieri, F.; Nasillo, V.; Paolini, A.; Bettelli, F.; Pioli, V.; Giusti, D.; Gilioli, A.; Colasante, C.; Acquaviva, G.; Riva, G.; et al. NPM1-Mutated Myeloid Neoplasms with <20% Blasts: A Really Distinct Clinico-Pathologic Entity? Int. J. Mol. Sci. 2020, 21, 8975. [Google Scholar] [CrossRef]
  121. Greenberg, P.L.; Tuechler, H.; Schanz, J.; Sanz, G.; Garcia-Manero, G.; Solé, F.; Bennett, J.M.; Bowen, D.; Fenaux, P.; Dreyfus, F.; et al. Revised International Prognostic Scoring System for Myelodysplastic Syndromes. Blood 2012, 120, 2454–2465. [Google Scholar] [CrossRef] [Green Version]
  122. Bernard, E.; Tuechler, H.; Greenberg, P.L.; Hasserjian, R.P.; Ossa, J.E.A.; Nannya, Y.; Devlin, S.M.; Creignou, M.; Pinel, P.; Monnier, L.; et al. Molecular International Prognostic Scoring System for Myelodysplastic Syndromes. NEJM Évid. 2022, 1, EVIDoa2200008. [Google Scholar] [CrossRef]
  123. Schuurhuis, G.J.; Heuser, M.; Freeman, S.; Béné, M.-C.; Buccisano, F.; Cloos, J.; Grimwade, D.; Haferlach, T.; Hills, R.K.; Hourigan, C.S.; et al. Minimal/measurable residual disease in AML: A consensus document from the European LeukemiaNet MRD Working Party. Blood 2018, 131, 1275–1291. [Google Scholar] [CrossRef] [Green Version]
  124. Short, N.J.; Zhou, S.; Fu, C.; Berry, D.A.; Walter, R.B.; Freeman, S.D.; Hourigan, C.S.; Huang, X.; Gonzalez, G.N.; Hwang, H.; et al. Association of Measurable Residual Disease with Survival Outcomes in Patients with Acute Myeloid Leukemia: A Systematic Review and Meta-analysis. JAMA Oncol. 2020, 6, 1890. [Google Scholar] [CrossRef]
  125. 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]
  126. Simoes, C.; Paiva, B.; Martínez-Cuadrón, D.; Bergua, J.-M.; Vives, S.; Algarra, L.; Tormo, M.; Martinez, P.; Serrano, J.; Herrera, P.; et al. Measurable residual disease in elderly acute myeloid leukemia: Results from the PETHEMA-FLUGAZA phase 3 clinical trial. Blood Adv. 2021, 5, 760–770. [Google Scholar] [CrossRef] [PubMed]
  127. 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]
  128. Döhner, K.; Thiede, C.; Jahn, N.; Panina, E.; Gambietz, A.; Larson, R.A.; Prior, T.W.; Marcucci, G.; Jones, D.; Krauter, J.; et al. Impact of NPM1/FLT3-ITD genotypes defined by the 2017 European LeukemiaNet in patients with acute myeloid leukemia. Blood 2020, 135, 371–380. [Google Scholar] [CrossRef] [PubMed]
  129. Angenendt, L.; Röllig, C.; Montesinos, P.; Martínez-Cuadrón, D.; Barragan, E.; García, R.; Botella, C.; Martínez, P.; Ravandi, F.; Kadia, T.; et al. Chromosomal Abnormalities and Prognosis in NPM1-Mutated Acute Myeloid Leukemia: A Pooled Analysis of Individual Patient Data from Nine International Cohorts. J. Clin. Oncol. 2019, 37, 2632–2642. [Google Scholar] [CrossRef] [PubMed]
  130. Sierra, J.; Nomdedeu, J.F. CEBPA bZip mutations: Just a single shot. Blood 2021, 138, 1091–1092. [Google Scholar] [CrossRef]
  131. Madanat, Y.F.; Xie, Z.; Zeidan, A.M. Advances in myelodysplastic syndromes: Promising novel agents and combination strategies. Expert Rev. Hematol. 2023, 16, 51–63. [Google Scholar] [CrossRef]
  132. Estey, E.H.; Hasserjian, R.P.; Döhner, H. Distinguishing AML from MDS: A fixed blast percentage may no longer be optimal. Blood 2022, 139, 323–332. [Google Scholar] [CrossRef] [PubMed]
  133. Heidenreich, S.; Ziagkos, D.; de Wreede, L.C.; van Biezen, A.; Finke, J.; Platzbecker, U.; Niederwieser, D.; Einsele, H.; Bethge, W.; Schleuning, M.; et al. Allogeneic Stem Cell Transplantation for Patients Age ≥ 70 Years with Myelodysplastic Syndrome: A Retrospective Study of the MDS Subcommittee of the Chronic Malignancies Working Party of the EBMT. Biol. Blood Marrow Transplant. 2016, 23, 44–52. [Google Scholar] [CrossRef] [Green Version]
  134. Atallah, E.; Logan, B.; Chen, M.; Cutler, C.; Deeg, J.; Jacoby, M.; Champlin, R.; Nishihori, T.; Confer, D.; Gajewski, J.; et al. Comparison of Patient Age Groups in Transplantation for Myelodysplastic Syndrome: The Medicare Coverage with Evidence Development Study. JAMA Oncol. 2020, 6, 486. [Google Scholar] [CrossRef]
  135. Cutler, C.S.; Lee, S.J.; Greenberg, P.; Deeg, H.J.; Pérez, W.S.; Anasetti, C.; Bolwell, B.J.; Cairo, M.S.; Gale, R.P.; Klein, J.P.; et al. A decision analysis of allogeneic bone marrow transplantation for the myelodysplastic syndromes: Delayed transplantation for low-risk myelodysplasia is associated with improved outcome. Blood 2004, 104, 579–585. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Franke, G.-N.; Lückemeier, P.; Platzbecker, U. Allogeneic Stem-Cell Transplantation in Patients with Myelodysplastic Syndromes and Prevention of Relapse. Clin. Lymphoma Myeloma Leuk. 2020, 21, 1–7. [Google Scholar] [CrossRef]
  137. Kröger, N.; Iacobelli, S.; Franke, G.-N.; Platzbecker, U.; Uddin, R.; Hübel, K.; Scheid, C.; Weber, T.; Robin, M.; Stelljes, M.; et al. Dose-Reduced Versus Standard Conditioning Followed by Allogeneic Stem-Cell Transplantation for Patients with Myelodysplastic Syndrome: A Prospective Randomized Phase III Study of the EBMT (RICMAC Trial). J. Clin. Oncol. 2017, 35, 2157–2164. [Google Scholar] [CrossRef]
  138. Scott, B.L.; Pasquini, M.C.; Fei, M.; Fraser, R.; Wu, J.; Devine, S.M.; Porter, D.L.; Maziarz, R.T.; Warlick, E.; Fernandez, H.F.; et al. Myeloablative versus Reduced-Intensity Conditioning for Hematopoietic Cell Transplantation in Acute Myelogenous Leukemia and Myelodysplastic Syndromes—Long-Term Follow-Up of the BMT CTN 0901 Clinical Trial. Transplant. Cell. Ther. 2021, 27, 483.e1–483.e6. [Google Scholar] [CrossRef]
  139. Webster, J.A.; Yogarajah, M.; Zahurak, M.; Symons, H.; Dezern, A.E.; Gojo, I.; Prince, G.T.; Morrow, J.; Jones, R.J.; Smith, B.D.; et al. A phase II study of azacitidine in combination with granulocyte-macrophage colony-stimulating factor as maintenance treatment, after allogeneic blood or marrow transplantation in patients with poor-risk acute myeloid leukemia (AML) or myelodysplastic syndrome (MDS). Leuk. Lymphoma 2021, 62, 3181–3191. [Google Scholar] [CrossRef] [PubMed]
  140. Bewersdorf, J.P.; Allen, C.; Mirza, A.-S.; Grimshaw, A.A.; Giri, S.; Podoltsev, N.A.; Gowda, L.; Cho, C.; Tallman, M.S.; Zeidan, A.M.; et al. Hypomethylating Agents and FLT3 Inhibitors as Maintenance Treatment for Acute Myeloid Leukemia and Myelodysplastic Syndrome After Allogeneic Hematopoietic Stem Cell Transplantation–A Systematic Review and Meta-Analysis. Transplant. Cell. Ther. 2021, 27, 997.e1–997.e11. [Google Scholar] [CrossRef]
  141. Oran, B.; de Lima, M.; Garcia-Manero, G.; Thall, P.F.; Lin, R.; Popat, U.; Alousi, A.M.; Hosing, C.; Giralt, S.; Rondon, G.; et al. A phase 3 randomized study of 5-azacitidine maintenance vs observation after transplant in high-risk AML and MDS patients. Blood Adv. 2020, 4, 5580–5588. [Google Scholar] [CrossRef]
  142. Mishra, A.; Tamari, R.; DeZern, A.E.; Byrne, M.T.; Gooptu, M.; Chen, Y.-B.; Deeg, H.J.; Sallman, D.; Gallacher, P.; Wennborg, A.; 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, 3985–3993. [Google Scholar] [CrossRef]
  143. Fenaux, P.; Mufti, G.J.; Hellstrom-Lindberg, E.; Santini, V.; Finelli, C.; Giagounidis, A.; Schoch, R.; Gattermann, N.; Sanz, G.; List, A.; et al. Efficacy of azacitidine compared with that of conventional care regimens in the treatment of higher-risk myelodysplastic syndromes: A randomised, open-label, phase III study. Lancet Oncol. 2009, 10, 223–232. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Prébet, T.; Gore, S.D.; Esterni, B.; Gardin, C.; Itzykson, R.; Thepot, S.; Dreyfus, F.; Rauzy, O.B.; Recher, C.; Adès, L.; et al. Outcome of High-Risk Myelodysplastic Syndrome after Azacitidine Treatment Failure. J. Clin. Oncol. 2011, 29, 3322–3327. [Google Scholar] [CrossRef]
  145. Ball, B.; Komrokji, R.S.; Adès, L.; Sekeres, M.A.; DeZern, A.E.; Pleyer, L.; Vey, N.; Almeida, A.; Germing, U.; Cluzeau, T.; et al. Evaluation of induction chemotherapies after hypomethylating agent failure in myelodysplastic syndromes and acute myeloid leukemia. Blood Adv. 2018, 2, 2063–2071. [Google Scholar] [CrossRef] [PubMed]
  146. Zhao, G.; Wang, Q.; Li, S.; Wang, X. Resistance to Hypomethylating Agents in Myelodysplastic Syndrome and Acute Myeloid Leukemia from Clinical Data and Molecular Mechanism. Front. Oncol. 2021, 11. [Google Scholar] [CrossRef] [PubMed]
  147. Leung, K.K.; Nguyen, A.; Shi, T.; Tang, L.; Ni, X.; Escoubet, L.; MacBeth, K.J.; DiMartino, J.; Wells, J.A. Multiomics of azacitidine-treated AML cells reveals variable and convergent targets that remodel the cell-surface proteome. Proc. Natl. Acad. Sci. USA 2019, 116, 695–700. [Google Scholar] [CrossRef] [Green Version]
  148. Bowler, E.H.; Bell, J.; Divecha, N.; Skipp, P.; Ewing, R.M. Proteomic Analysis of Azacitidine-Induced Degradation Profiles Identifies Multiple Chromatin and Epigenetic Regulators Including Uhrf1 and Dnmt1 as Sensitive to Azacitidine. J. Proteome Res. 2019, 18, 1032–1042. [Google Scholar] [CrossRef]
  149. Unnikrishnan, A.; Papaemmanuil, E.; Beck, D.; Deshpande, N.P.; Verma, A.; Kumari, A.; Woll, P.S.; Richards, L.A.; Knezevic, K.; Chandrakanthan, V.; et al. Integrative Genomics Identifies the Molecular Basis of Resistance to Azacitidine Therapy in Myelodysplastic Syndromes. Cell Rep. 2017, 20, 572–585. [Google Scholar] [CrossRef] [Green Version]
  150. Valencia, A.; Masala, E.; Rossi, A.; Martino, A.; Sanna, A.; Buchi, F.; Canzian, F.; Cilloni, D.; Gaidano, V.; Voso, M.T.; et al. Expression of nucleoside-metabolizing enzymes in myelodysplastic syndromes and modulation of response to azacitidine. Leukemia 2013, 28, 621–628. [Google Scholar] [CrossRef] [Green Version]
  151. 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]
  152. 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] [Green Version]
  153. 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]
  154. Bazinet, A.; Darbaniyan, F.; Jabbour, E.; Montalban-Bravo, G.; Ohanian, M.; Chien, K.; Kadia, T.; Takahashi, K.; Masarova, L.; Short, N.; et al. Azacitidine plus venetoclax in patients with high-risk myelodysplastic syndromes or chronic myelomonocytic leukaemia: Phase 1 results of a single-centre, dose-escalation, dose-expansion, phase 1–2 study. Lancet Haematol. 2022, 9, e756–e765. [Google Scholar] [CrossRef] [PubMed]
  155. Pullarkat, V.; Pratz, K.; Dohner, H.; Recher, C.; Thirman, M.J.; Dinardo, C.D.; Fenaux, P.; Schuh, A.C.; Wei, A.H.; Pigneux, A.; et al. Venetoclax and azacitidine combination in chemotherapy ineligible untreated patients with therapy-related myeloid neoplasms, antecedent myelodysplastic syndromes, or myelodysplastic/myeloproliferative neoplasms. In Proceedings of the ASCO Annual Meeting, Chicago, IL, USA, 4–8 June 2021. [Google Scholar]
  156. Guo, Y.; Liu, B.; Deng, L.; Qiao, Y.; Jian, J. The efficacy and adverse events of venetoclax in combination with hypomethylating agents treatment for patients with acute myeloid leukemia and myelodysplastic syndrome: A systematic review and meta-analysis. Hematology 2020, 25, 414–423. [Google Scholar] [CrossRef] [PubMed]
  157. Zeidan, A.M.; Borate, U.; Pollyea, D.A.; Brunner, A.M.; Roncolato, F.; Garcia, J.S.; Filshie, R.; Odenike, O.; Watson, A.M.; Krishnadasan, R.; et al. A phase 1b study of venetoclax and azacitidine combination in patients with relapsed or refractory myelodysplastic syndromes. Am. J. Hematol. 2022, 98, 272–281. [Google Scholar] [CrossRef] [PubMed]
  158. Bewersdorf, J.P.; Derkach, A.; Gowda, L.; Menghrajani, K.; DeWolf, S.; Ruiz, J.D.; Ponce, D.M.; Shaffer, B.C.; Tamari, R.; Young, J.W.; et al. Venetoclax-based combinations in AML and high-risk MDS prior to and following allogeneic hematopoietic cell transplant. Leuk. Lymphoma 2021, 62, 3394–3401. [Google Scholar] [CrossRef]
  159. Yang, T.-T.; Song, X.-L.; Zhao, Y.-M.; Ye, B.-D.; Luo, Y.; Xiao, H.-W.; Chen, Y.; Fu, H.-R.; Yu, J.; Liu, L.-Z.; et al. Outcome after allogeneic hematopoietic stem cell transplantation following Venetoclax-based therapy among AML and MDS patients. Ann. Hematol. 2022, 101, 2731–2741. [Google Scholar] [CrossRef]
  160. Bello, C.; Yu, D.; Komrokji, R.S.; Zhu, W.; Wetzstein, G.A.; List, A.F.; Lancet, J.E. Outcomes after induction chemotherapy in patients with acute myeloid leukemia arising from myelodysplastic syndrome. Cancer 2010, 117, 1463–1469. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  161. Lancet, J.E.; Uy, G.L.; Cortes, J.E.; Newell, L.F.; Lin, T.L.; Ritchie, E.K.; Stuart, R.K.; Strickland, S.A.; Hogge, D.; Solomon, S.R.; et al. CPX-351 (cytarabine and daunorubicin) Liposome for Injection Versus Conventional Cytarabine Plus Daunorubicin in Older Patients with Newly Diagnosed Secondary Acute Myeloid Leukemia. J. Clin. Oncol. 2018, 36, 2684–2692. [Google Scholar] [CrossRef]
  162. E Lancet, J.; Uy, G.L.; Newell, L.F.; Lin, T.L.; Ritchie, E.K.; Stuart, R.K.; A Strickland, S.; Hogge, D.; Solomon, S.R.; Bixby, D.L.; et al. CPX-351 versus 7+3 cytarabine and daunorubicin chemotherapy in older adults with newly diagnosed high-risk or secondary acute myeloid leukaemia: 5-year results of a randomised, open-label, multicentre, phase 3 trial. Lancet Haematol. 2021, 8, e481–e491. [Google Scholar] [CrossRef] [PubMed]
  163. Welch, J.S.; Petti, A.A.; Miller, C.A.; Fronick, C.C.; O’Laughlin, M.; Fulton, R.S.; Wilson, R.K.; Baty, J.D.; Duncavage, E.J.; Tandon, B.; et al. TP53 and Decitabine in Acute Myeloid Leukemia and Myelodysplastic Syndromes. N. Engl. J. Med. 2016, 375, 2023–2036. [Google Scholar] [CrossRef]
  164. Ball, S.; Loghavi, S.; Zeidan, A.M. TP53-altered higher-risk myelodysplastic syndromes/neoplasms and acute myeloid leukemia: A distinct genetic entity with unique unmet needs. Leuk. Lymphoma 2022, 1–11. [Google Scholar] [CrossRef] [PubMed]
  165. Lambert, J.M.; Gorzov, P.; Veprintsev, D.B.; Söderqvist, M.; Segerbäck, D.; Bergman, J.; Fersht, A.R.; Hainaut, P.; Wiman, K.G.; Bykov, V.J. PRIMA-1 reactivates mutant p53 by covalent binding to the core domain. Cancer Cell 2009, 15, 376–388. [Google Scholar] [CrossRef] [Green Version]
  166. Bykov, V.N.; Issaeva, N.; Shilov, A.; Hultcrantz, M.; Pugacheva, E.; Chumakov, P.; Bergman, J.; Wiman, K.; Selivanova, G. Restoration of the tumor suppressor function to mutant p53 by a low-molecular-weight compound. Nat. Med. 2002, 8, 282–288. [Google Scholar] [CrossRef] [PubMed]
  167. Zhang, Q.; Bykov, V.J.N.; Wiman, K.G.; Zawacka-Pankau, J. APR-246 reactivates mutant p53 by targeting cysteines 124 and 277. Cell Death Dis. 2018, 9, 439. [Google Scholar] [CrossRef] [Green Version]
  168. Ali, D.; Jönsson-Videsäter, K.; Deneberg, S.; Bengtzén, S.; Nahi, H.; Paul, C.; Lehmann, S. APR-246 exhibits anti-leukemic activity and synergism with conventional chemotherapeutic drugs in acute myeloid leukemia cells. Eur. J. Haematol. 2010, 86, 206–215. [Google Scholar] [CrossRef]
  169. Maslah, N.; Salomao, N.; Drevon, L.; Verger, E.; Partouche, N.; Ly, P.; Aubin, P.; Naoui, N.; Schlageter, M.-H.; Bally, C.; et al. Synergistic effects of PRIMA-1Met (APR-246) and 5-azacitidine in TP53-mutated myelodysplastic syndromes and acute myeloid leukemia. Haematologica 2019, 105, 1539–1551. [Google Scholar] [CrossRef] [Green Version]
  170. Cluzeau, T.; Sebert, M.; Rahmé, R.; Cuzzubbo, S.; Lehmann-Che, J.; Madelaine, I.; Peterlin, P.; Bève, B.; Attalah, H.; Chermat, F.; 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, 1575–1583. [Google Scholar] [CrossRef]
  171. Sallman, D.A.; DeZern, A.E.; Garcia-Manero, G.; Steensma, D.P.; Roboz, G.J.; Sekeres, M.A.; Cluzeau, T.; Sweet, K.L.; McLemore, A.; McGraw, K.L.; et al. Eprenetapopt (APR-246) and Azacitidine in TP53-Mutant Myelodysplastic Syndromes. J. Clin. Oncol. 2021, 39, 1584–1594. [Google Scholar] [CrossRef]
  172. Garcia-Manero, G.; Goldberg, A.D.; Winer, E.S.; Altman, J.K.; Fathi, A.T.; Odenike, O.; Roboz, G.J.; Gallacher, P.; Wennborg, A.; Hickman, D.K.; et al. Phase I and Expansion Study of Eprenetapopt (APR-246) in Combination with Venetoclax (VEN) and Azacitidine (AZA) in TP53-Mutant Acute Myeloid Leukemia (AML). Blood 2021, 138, 3409. [Google Scholar] [CrossRef]
  173. Chao, M.P.; Takimoto, C.H.; Feng, D.D.; McKenna, K.; Gip, P.; Liu, J.; Volkmer, J.-P.; Weissman, I.L.; Majeti, R. Therapeutic Targeting of the Macrophage Immune Checkpoint CD47 in Myeloid Malignancies. Front. Oncol. 2020, 9, 1380. [Google Scholar] [CrossRef]
  174. Daver, N.G.; Vyas, P.; Kambhampati, S.; Al Malki, M.M.; Larson, R.A.; Asch, A.S.; Mannis, G.N.; Chai-Ho, W.; Tanaka, T.N.; Bradley, T.J.; et al. Tolerability and efficacy of the first-in-class anti-CD47 antibody magrolimab combined with azacitidine in frontline TP53m AML patients: Phase 1b results. J. Clin. Oncol. 2022, 40, 7020. [Google Scholar] [CrossRef]
  175. Naval, G.; 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. In Proceedings of the ASH Annual Meeting, Atlanta, GA, USA, 11–14 December 2021. [Google Scholar]
  176. Jin, J.; Hu, C.; Yu, M.; Chen, F.; Ye, L.; Yin, X.; Zhuang, Z.; Tong, H. Prognostic value of isocitrate dehydrogenase mutations in myelodysplastic syndromes: A retrospective cohort study and meta-analysis. PLoS ONE 2014, 9, e100206. [Google Scholar] [CrossRef] [Green Version]
  177. Patnaik, M.M.; A Hanson, C.; Hodnefield, J.M.; Lasho, T.L.; Finke, C.M.; A Knudson, R.; Ketterling, R.; Pardanani, A.; Tefferi, A. Differential prognostic effect of IDH1 versus IDH2 mutations in myelodysplastic syndromes: A Mayo Clinic Study of 277 patients. Leukemia 2011, 26, 101–105. [Google Scholar] [CrossRef] [PubMed]
  178. DiNardo, C.D.; Stein, E.M.; de Botton, S.; Roboz, G.J.; Altman, J.K.; Mims, A.S.; Swords, R.; Collins, R.H.; Mannis, G.N.; Pollyea, D.A.; et al. Durable Remissions with Ivosidenib in IDH1-Mutated Relapsed or Refractory AML. N. Engl. J. Med. 2018, 378, 2386–2398. [Google Scholar] [CrossRef]
  179. Roboz, G.J.; DiNardo, C.D.; Stein, E.M.; de Botton, S.; Mims, A.S.; Prince, G.T.; Altman, J.K.; Arellano, M.L.; Donnellan, W.; Erba, H.P.; et al. Ivosidenib induces deep durable remissions in patients with newly diagnosed IDH1-mutant acute myeloid leukemia. Blood 2020, 135, 463–471. [Google Scholar] [CrossRef] [Green Version]
  180. Dinardo, C.D.; Stein, A.S.; Stein, E.M.; Fathi, A.T.; Frankfurt, O.; Schuh, A.C.; Döhner, H.; Martinelli, G.; Patel, P.A.; Raffoux, E.; et al. Mutant Isocitrate Dehydrogenase 1 Inhibitor Ivosidenib in Combination with Azacitidine for Newly Diagnosed Acute Myeloid Leukemia. J. Clin. Oncol. 2021, 39, 57–65. [Google Scholar] [CrossRef] [PubMed]
  181. Stein, E.M.; Fathi, A.T.; DiNardo, C.D.; A Pollyea, D.; Roboz, G.J.; Collins, R.; A Sekeres, M.; Stone, R.M.; Attar, E.C.; Frattini, M.G.; et al. Enasidenib in patients with mutant IDH2 myelodysplastic syndromes: A phase 1 subgroup analysis of the multicentre, AG221-C-001 trial. Lancet Haematol. 2020, 7, e309–e319. [Google Scholar] [CrossRef]
  182. DiNardo, C.D.; Venugopal, S.; Lachowiez, C.A.; Takahashi, K.; Loghavi, S.; Montalban-Bravo, G.; Wang, X.; E Carraway, H.; Sekeres, M.A.; Sukkur, A.; et al. Targeted therapy with the mutant IDH2 inhibitor enasidenib for high-risk IDH2-mutant myelodysplastic syndrome. Blood Adv. 2022. [Google Scholar] [CrossRef]
  183. Fathi, A.T.; Kim, H.T.; Soiffer, R.J.; Levis, M.J.; Li, S.; Kim, A.S.; Mims, A.S.; DeFilipp, Z.; El-Jawahri, A.; McAfee, S.L.; et al. Enasidenib as maintenance following allogeneic hematopoietic cell transplantation for IDH2-mutated myeloid malignancies. Blood Adv. 2022, 6, 5857–5865. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The classification algorithm of MDS based on the new changes. 1 Diagnosis of MDS/AML or AML with mutated TP53 can be made with any somatic TP53 mutation with VAF > 10%, whereas MDS with mutated TP53 requires the presence of a multi-hit TP53 mutation or a TP53 mutation (VAF > 10%) and a complex karyotype, often with loss of 17p if LOH information is not available. A multi-hit TP53 mutation is defined as either two distinct TP53 mutations, each with VAF > 10%, or a single TP53 mutation with either 17p deletion on cytogenetics, VAF ≥ 50%, or copy-neutral LOH at the 17p locus. Abbreviations: MDS, myelodysplastic syndromes; AML, acute myeloid leukemia; EBs, excess blasts; NOS, not otherwise specified; CCUS, clonal cytopenia of undetermined significance.
Figure 1. The classification algorithm of MDS based on the new changes. 1 Diagnosis of MDS/AML or AML with mutated TP53 can be made with any somatic TP53 mutation with VAF > 10%, whereas MDS with mutated TP53 requires the presence of a multi-hit TP53 mutation or a TP53 mutation (VAF > 10%) and a complex karyotype, often with loss of 17p if LOH information is not available. A multi-hit TP53 mutation is defined as either two distinct TP53 mutations, each with VAF > 10%, or a single TP53 mutation with either 17p deletion on cytogenetics, VAF ≥ 50%, or copy-neutral LOH at the 17p locus. Abbreviations: MDS, myelodysplastic syndromes; AML, acute myeloid leukemia; EBs, excess blasts; NOS, not otherwise specified; CCUS, clonal cytopenia of undetermined significance.
Ijms 24 05018 g001
Figure 2. The classification algorithm of AML based on the new changes. (A) AML with recurrent genetic abnormalities takes precedence in the diagnostic algorithm, and the blast cut-off for diagnosis is lowered to 10%, with the exception of BCR::ABL-mutated AML. Within this group, PML::RARA mutations supersede RUNX1::RUNX1T1 mutations, followed by CBFB:MYH11, then MLLT3:KMT2A, then DEK::NUP214, then GATA2;MECOM(EV11), followed by other rare recurring translocations as depicted below, then BCR.:ABL1, then NPM1, and last in-frame bZIP CEBPA mutations. Among the remaining categories, (B) TP53-mutated AML supersedes (C) AML with myelodysplasia-related gene mutations and the latter supersedes (D) AML with myelodysplasia-related cytogenetic abnormalities. If none of the aforementioned genetic abnormalities are present, then AML is defined as (E) not-otherwise specified. 1 Acute myeloid leukemia (AML) with other rare recurring translocations includes the following: t(1;3)(p36.3;q21.3)/PRDM16::RPN1, t(3;5)(q25.3;q35.1)/NPM1::MLF1, t(8;16)(p11.2;p13.3)/KAT6A::CREBBP, t(1;22)(p13.3;q13.1)/RBM15::MRTF1, t(5;11)(q35.2;p15.4/NUP98::NSD1, t(11;12)(p15.4;p13.3)/NUP98::KMD5A, NUP98 and other partners, t(7;12)(q36.3;p13.2)/ETV6::MNX1, t(10;11)(p12.3;q14.2)/PICALM::MLLT10, t(16;21)(p11.2;q22.2)/FUS::ERG, t(16;21)(q24.3;q22.1)/RUNX1::CBFA2T3, inv(16)(p13.3q24.3)/CBFA2T3::GLIS2. 2 The MDS/AML category is not applicable due to its overlap with progression of BCR::ABL1-positive CML. Abbreviations: MDS, myelodysplastic syndromes; AML, acute myeloid leukemia; NOS, not otherwise specified.
Figure 2. The classification algorithm of AML based on the new changes. (A) AML with recurrent genetic abnormalities takes precedence in the diagnostic algorithm, and the blast cut-off for diagnosis is lowered to 10%, with the exception of BCR::ABL-mutated AML. Within this group, PML::RARA mutations supersede RUNX1::RUNX1T1 mutations, followed by CBFB:MYH11, then MLLT3:KMT2A, then DEK::NUP214, then GATA2;MECOM(EV11), followed by other rare recurring translocations as depicted below, then BCR.:ABL1, then NPM1, and last in-frame bZIP CEBPA mutations. Among the remaining categories, (B) TP53-mutated AML supersedes (C) AML with myelodysplasia-related gene mutations and the latter supersedes (D) AML with myelodysplasia-related cytogenetic abnormalities. If none of the aforementioned genetic abnormalities are present, then AML is defined as (E) not-otherwise specified. 1 Acute myeloid leukemia (AML) with other rare recurring translocations includes the following: t(1;3)(p36.3;q21.3)/PRDM16::RPN1, t(3;5)(q25.3;q35.1)/NPM1::MLF1, t(8;16)(p11.2;p13.3)/KAT6A::CREBBP, t(1;22)(p13.3;q13.1)/RBM15::MRTF1, t(5;11)(q35.2;p15.4/NUP98::NSD1, t(11;12)(p15.4;p13.3)/NUP98::KMD5A, NUP98 and other partners, t(7;12)(q36.3;p13.2)/ETV6::MNX1, t(10;11)(p12.3;q14.2)/PICALM::MLLT10, t(16;21)(p11.2;q22.2)/FUS::ERG, t(16;21)(q24.3;q22.1)/RUNX1::CBFA2T3, inv(16)(p13.3q24.3)/CBFA2T3::GLIS2. 2 The MDS/AML category is not applicable due to its overlap with progression of BCR::ABL1-positive CML. Abbreviations: MDS, myelodysplastic syndromes; AML, acute myeloid leukemia; NOS, not otherwise specified.
Ijms 24 05018 g002
Table 1. The incidence, biologic mechanisms and clinical significance of genetic abnormalities in MDS and AML-MRC.
Table 1. The incidence, biologic mechanisms and clinical significance of genetic abnormalities in MDS and AML-MRC.
Genetic AlterationIncidence in MDS/AMLPossible Biologic MechanismClinical SignificanceReferences
Chromosomal abnormalities
Deletion of 5q10–15%Defective ribosomal biogenesis leads to degradation of MDM2Overall good prognosis[16,17,18,19,20,21]
Deletion of chromosome 7/7q~20%Loss of chromosome 7 is linked with compromised expression of interferon gamma pathway genes.Overall poor prognosis[23,24,25,26,27,28]
Complex karyotype (CK) 30%Mutations in TP53 and frequent LOH of this genePoor prognosis/survival[29,30,31,32,33,34,35]
Somatic mutations
SF3B110–15%Alterations in the spliceosome that frequently lead to non-sense-mediated decay (NMD)Overall good prognosis but specific variants may have different impact[39,40,41,42]
SRSF210–15%Alteration of splicing and DNA damage accumulation.Poor prognosis[43,44,45,46]
U2AF15–15%Suppression of ATG7 levels is associated with increased oxidative stress and chromosomal instabilityPoor prognosis/survival[47,48,49]
ZRSR25%Retention of U12-dependent introns causing aberrant splicingNegative impact on survival less strong compared to SRSF2 and U2AF1[31,50,51]
STAG2~10%Dysplasia caused by abnormal hematopoietic stem cell maturation along with additional oncogenic mutationsPoor prognosis[53,54,55,56,57]
EZH24–5%Gain of function mutations lead to improper control of cell proliferationPoor prognosis/survival[58,59,60,63,65,93]
BCOR<5%Loss of function mutations causing increased myeloid cell proliferationNeutral impact on survival[69,70,71,72]
RUNX110–15%Mutated gene transcripts block gene activation responsible proper hematopoiesisPoor prognosis/survival[67,74,76,77,78,79]
NRAS/KRAS4–5%RAS gene mutations activate RAS/Raf/MEK and PI3K/AKT pathwaysPoor prognosis/survival[80,81,82,83,84,85]
FLT3-ITD0.6–6%Mutations cause constitutive auto-phosphorylation of the FLT3 receptor leading to activation of downstream pathwaysPoor prognosis/survival[11,88,89]
TP535–10%Mutations lead to impaired DNA damage repairs, LOH is critical for the prognosisPoor prognosis/survival[90,91,92]
Table 2. Pre-malignant myeloid clonal conditions.
Table 2. Pre-malignant myeloid clonal conditions.
DescriptionClonalityCytopeniaDysplasiaIncreased BlastsGenetic Testing
CHLineage derived from a single cloneYesNoNoNoGeneral term to define somatic mutations or cytogenetic abnormalities or copy number mutations 1
CHIPLineage derived from a single cloneYesNoNoNoSomatic mutation in a myeloid neoplasm driver gene with VAF ≥ 2%, or
Non-MDS defining clonal cytogenetic abnormality in the absence of a myeloid neoplasm or cytopenia
CCUSLineage derived from a single clone + CytopeniaYesYesNoNoSomatic mutation in a myeloid neoplasm driver gene with VAF ≥ 2% 2
VEXAS syndromeCH + anemiaYesYesNoNoSomatic mutation in UBA1 gene
1 Can be seen post-treatment of myeloid neoplasms or even solid tumors, in which case the clinical implications might differ. 2 Higher VAF is associated with higher progression to MDS [107]. Abbreviations: CH, clonal hematopoiesis; CHIP, clonal hematopoiesis of indeterminate potential; CCUS, clonal cytopenia of undetermined significance; VEXAS syndrome (vacuoles, E1 enzyme, X-linked, autoinflammatory, somatic).
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

Zavras, P.D.; Sinanidis, I.; Tsakiroglou, P.; Karantanos, T. Understanding the Continuum between High-Risk Myelodysplastic Syndrome and Acute Myeloid Leukemia. Int. J. Mol. Sci. 2023, 24, 5018. https://doi.org/10.3390/ijms24055018

AMA Style

Zavras PD, Sinanidis I, Tsakiroglou P, Karantanos T. Understanding the Continuum between High-Risk Myelodysplastic Syndrome and Acute Myeloid Leukemia. International Journal of Molecular Sciences. 2023; 24(5):5018. https://doi.org/10.3390/ijms24055018

Chicago/Turabian Style

Zavras, Phaedon D., Ilias Sinanidis, Panagiotis Tsakiroglou, and Theodoros Karantanos. 2023. "Understanding the Continuum between High-Risk Myelodysplastic Syndrome and Acute Myeloid Leukemia" International Journal of Molecular Sciences 24, no. 5: 5018. https://doi.org/10.3390/ijms24055018

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