ﺑﺎﺯﮔﺸﺖ ﺑﻪ ﺻﻔﺤﻪ ﻗﺒﻠﯽ
خرید پکیج
تعداد آیتم قابل مشاهده باقیمانده : 3 مورد
نسخه الکترونیک
medimedia.ir

Acute myeloid leukemia: Molecular genetics

Acute myeloid leukemia: Molecular genetics
Literature review current through: Jan 2024.
This topic last updated: Apr 18, 2022.

INTRODUCTION — Acute myeloid leukemia (AML) develops as the consequence of a series of genetic changes in a hematopoietic precursor cell. These changes alter normal hematopoietic growth and differentiation, resulting in an accumulation of large numbers of abnormal, immature myeloid cells in the bone marrow and peripheral blood. These cells are capable of dividing and proliferating, but cannot differentiate into mature hematopoietic cells (ie, neutrophils).

Similar to other malignancies, the genetic alterations in AML include mutation of oncogenes as well as the loss of tumor suppressor genes. In contrast to most solid tumors, many hematologic malignancies are associated with a single characteristic cytogenetic abnormality (eg, the Philadelphia chromosome [t(9;22)] in chronic myeloid leukemia and t(15;17) in acute promyelocytic leukemia).

Advances in the identification of recurring chromosomal abnormalities and translocations have provided major insight into the pathobiology of AML. (See "Tools for genetics and genomics: Cytogenetics and molecular genetics".)

Specific cytogenetic abnormalities identified by karyotype analysis have considerable prognostic significance for patients with AML and affect treatment planning (table 1). Abnormalities in certain genes (eg, mutations in FLT3, nucleophosmin, KIT) as well as gene expression profiles confer prognostic significance in adult patients with AML. Even those patients without obvious abnormalities detected by karyotypic analysis or gene expression profiles have acquired copy number alterations that may help to identify genes important for the pathogenesis of AML [1,2]. (See "Acute myeloid leukemia: Risk factors and prognosis".)

The focus of this topic review will be on the various molecular genetic events involved in the pathogenesis of AML. More general discussions of genetic events in hematologic malignancies are presented separately. (See "Genetic abnormalities in hematologic and lymphoid malignancies" and "General aspects of cytogenetic analysis in hematologic malignancies".)

GENE MUTATIONS — Gene sequencing studies have shown that, on average, de novo AML cases contain more than 10 significant gene mutations, many of which can be broadly grouped into nine categories of genes thought to participate in leukemogenesis: DNA methylation, tumor suppressors, transcription factor fusions, nucleophosmin, signaling molecules, chromatin modification, myeloid transcription factors, the cohesin complex, and the spliceosome complex [3]. The most common genes mutated are: FLT3 (28 percent), NPM1 (27 percent), DNMT3A (26 percent), IDH1 or IDH2 (20 percent), NRAS or KRAS (12 percent), RUNX1 (10 percent), TET2 (8 percent), TP53 (8 percent), CEBPA (6 percent), and WT1 (6 percent).

The nature of the genetic lesions that are required to establish AML is not well defined, but the combination of mutations in any given AML is not random [4,5]. As examples, mutations in genes that encode signaling molecules (eg, FLT3, RAS) often accompany chromosomal rearrangements that target hematopoietic transcription factors (eg, PML-RARalpha, CBFbeta-MYH11) or key hematopoietic transcription factors (eg, CEBPA, RUNX1). Conversely, mutations of genes in a single physiologic pathway are often mutually exclusive; as an example, it is uncommon to find mutations of more than one gene that regulates DNA demethylation (eg, TET2, IDH) in a given AML sample.

The pathogenetic effects of some of these mutations will be discussed here. The prognostic value of gene mutations in patients with AML is discussed separately. (See "Acute myeloid leukemia: Risk factors and prognosis".)

Mutations affecting DNA methylation — Gene mutations that affect DNA methylation have been identified in patients with AML. These mutations may result in the expression of genes that are normally silenced or in the silencing of genes that are normally expressed. DNA methylation patterns may also have prognostic significance [6].

DNA methylation involves the addition of a methyl group to cytosine or adenine residues within DNA. Hypermethylated genes are generally not expressed (ie, silenced); as such, DNA methylation provides an epigenetic method for altering gene expression. This process is an essential component of normal development and cell differentiation, and alterations in DNA methylation may impair cell differentiation, as is seen in AML [7].

Mutations associated with AML are found in genes that encode enzymes involved in both DNA methylation (eg, DNMTs) and DNA demethylation (eg, IDH, WT1, TET2). Deregulation of the IDH/WT1/TET2 axis is seen in as many as one-third of patients with AML, but mutations in genes associated with DNA demethylation are usually mutually exclusive in a given AML sample. The critical downstream targets of these changes in DNA methylation have not yet been identified.

DNMT mutations — DNA methyltransferases (DNMTs) are enzymes responsible for initiating and/or maintaining DNA methylation in mammalian cells. Without DNMTs, cells would gradually lose methyl groups from their DNA, leading to the expression of genes that are usually silenced. In one study, DNMT3A mutations were found in 22 percent of 281 patients with de novo AML [8]. When compared with patients without such mutations, patients with DNMT3A mutations had a significantly shorter median survival (12 versus 41 months). Another DNA sequencing study suggested that the rate of DNMT3A mutations varies by AML subtype, with the highest rate (20 percent) seen among cases of acute monocytic leukemia (ie, FAB M5) [9].

TET mutations — The TET (ten-eleven translocation) genes encode proteins with enzymatic activity that initiate the multistep process of DNA demethylation [10]. Somatic mutations in TET2 occur in approximately 15 percent of myeloid cancers, including AML [11]. Mutations of TET2 result in increased methylation and silencing of genes that are normally expressed. In mice, ablation of TET2 alters the function of hematopoietic stem cells and induces myeloid leukemia [12]. WT1 (Wilms tumor 1) is an interaction partner of TET2, and regulates TET2 binding to DNA, thereby providing another mechanism that can alter DNA methylation [13,14]. (See 'Tumor suppressor genes' below.)

IDH mutations — IDH (isocitrate dehydrogenase) enzymes catalyze the conversion of isocitrate to alpha-ketoglutarate. Mutated IDH enzymes generate the "oncometabolite" 2-hydroxyglutarate, which inhibits TET2 function. Mutations of IDH1 and IDH2 are found in approximately 7 and 9 percent, respectively, of patients with AML, and are generally found in AML with a normal karyotype [15,16]. (See "Acute myeloid leukemia: Risk factors and prognosis", section on 'IDH1'.)

ASXL mutations — The function of ASXL (additional sex combs-like) proteins is poorly understood, but they may act as epigenetic regulators by serving as scaffolds for transcription activators, repressors, and related proteins [17]. (See "Acute myeloid leukemia: Risk factors and prognosis", section on 'ASXL1'.)

Tumor suppressor genes — Abnormalities of several tumor suppressor genes have been well characterized in myeloid leukemias. These are described in the following sections.

TP53 mutations — Progression to blast crisis in CML is commonly associated with acquisition of mutations in TP53 [18-20]. In AML, only approximately 7 percent of cases have known TP53 mutations at the time of diagnosis, although the incidence of TP53 abnormalities may be higher in patients with AML evolving from a prior myelodysplastic syndrome (MDS) [21], in AML associated with 11q23 translocations [22,23], and in AML with complex karyotype [24]. (See "Cellular and molecular biology of chronic myeloid leukemia", section on 'Progression to acute phase CML' and "Acute myeloid leukemia: Risk factors and prognosis", section on 'TP53'.)

WT1 mutations — The Wilms tumor suppressor gene (WT1) is overexpressed in the leukemic blasts of approximately 75 percent of cases of newly diagnosed AML [25,26]. The WT1 gene, located at chromosome 11p13, encodes a transcription factor expressed mainly in kidneys and gonads [27,28], but it is also expressed in normal CD34+/CD38– hematopoietic progenitor cells [29]. In AML, high WT1 levels at diagnosis have, in some studies, been shown to correlate with lower remission rates and overall survival [25,26].

NF1 mutations — Patients with neurofibromatosis and mutations of the NF1 tumor suppressor gene have a high rate of developing neoplasms, including progression of MDS into AML [30,31].

Other tumor suppressor mutations — Some tumor suppressors that are commonly associated with other malignancies (eg, lymphomas) are infrequently found in AML, whereas certain as-yet-unidentified regulators may be located at chromosomal regions that are frequently deleted in AML.

Examples include the following:

INK4 (ie, p16) is a cyclin dependent kinase inhibitor that is frequently deleted in acute lymphoblastic leukemia, but is only very rarely mutated in AML [32].

BCL-2 (B cell leukemia/lymphoma-2) is overexpressed in a wide variety of lymphoid malignancies as a result of t(14;18), which juxtaposes BCL-2 with the Ig heavy chain locus. BCL-2 overexpression in AML has been reported only rarely, but in a study of patients with newly diagnosed AML, a multivariate analysis showed that BCL-2 overexpression was associated with lower complete remission rates and a significantly shorter survival [33]. (See "Clinical manifestations, pathologic features, diagnosis, and prognosis of follicular lymphoma", section on 'Cytogenetics'.)

Certain chromosomal deletions are closely associated with AML (eg, loss of the long arms of chromosomes 5, 7, and 20). These abnormalities occur commonly in therapy-related AML and AML associated with prior MDS, and are associated with a poor prognosis (table 2) [34-36]. The involved genes have not yet been identified, although novel tumor suppressor genes or other regulatory molecules are likely to be encoded at these loci.

Activated signaling — Mutations impacting signal activation are seen in approximately 60 percent of AML cases and include point mutations in FLT3, KIT, other tyrosine kinases, serine-threonine kinases, KRAS/NRAS, or protein-tyrosine phosphatases [3].

FLT3 — FLT3 is a transmembrane tyrosine kinase receptor that stimulates cell proliferation upon activation. Mutations in the FMS-like tyrosine kinase 3 gene producing internal transmembrane duplications (FLT3/ITD) and constitutive activation of the FLT3 receptor tyrosine kinase are quite common in AML, particularly in patients with normal karyotypes.

There are two main types of FLT3 mutations. The most common are internal tandem duplications of different length that result in ligand-independent activation of the FLT3 receptor and a proliferative signal. Alternatively, point mutations in the activating loop of the kinase domain of FLT3 may result in tyrosine kinase activation of FLT3 in 5 to 10 percent of patients [37,38].

The prognostic impact of FLT3 mutations is discussed in more detail separately. (See "Acute myeloid leukemia: Risk factors and prognosis", section on 'FLT3'.)

RAS oncogene mutations — Mutations of three closely related RAS (HRAS, KRAS, NRAS) proto-oncogenes occur in human malignancies. Each of these genes encodes a structurally similar 21 kilodalton protein which localizes to the inner plasma membrane and plays an important role in signal transduction [39]. Specific point mutations occur characteristically in each of the three genes, although no consistent karyotypic abnormalities have been associated with RAS mutation and activation [40].

Mutations of RAS have been identified in approximately 25 percent of cases of AML [40,41], and are observed with higher frequency in MDS (35 percent) [42,43]. The majority of the mutations in AML and MDS are in the NRAS gene; KRAS and HRAS mutations occur less frequently. In both AML and MDS, RAS mutations have been reported more frequently in cases with a monocytic morphology (eg, FAB M4 and chronic myelomonocytic leukemia) [44]. An association between AML in patients with a variety of occupational exposures and RAS mutations has also been observed [45].

The significance of RAS mutations in AML remains unclear. RAS mutations have been reported at diagnosis, but not at relapse in some AML cases [46]. Conversely, RAS mutations have been acquired at the time of relapse or disease progression [44,47]. However, the presence of RAS mutations has been associated with improved survival [41], possibly related to a lower leukemic cell burden at presentation, or an increased sensitivity to chemotherapy with cytarabine [48].

NPM1 mutations — Nucleophosmin (NPM1) is a ubiquitously-expressed phosphoprotein involved in ribosomal protein assembly and transport and regulation the tumor suppressor ARF (cyclin-dependent kinase inhibitor 2A). NPM1 shuttles between the nucleus and cytoplasm, but mutations associated with AML (typically four base pair insertions that cause a frameshift) eliminate the C-terminus amino acids responsible for nucleolar localization; as a result, mutant NPM1 is retained in the cytoplasm [49,50]. HOXA and HOX B cluster genes are consistently overexpressed in NPM1-mutated AML.

NPM1 mutations are associated with improved outcomes in younger and older adults and in children with AML, although the mechanism for increased chemosensitivity is not known. The prognostic impact of NPM1 mutations in AML is discussed in more detail separately. (See "Acute myeloid leukemia: Risk factors and prognosis", section on 'NPM1'.)

Splicing factor mutations — Mutations involving the mRNA assembly system that cause aberrant mRNA splicing are highly specific to myeloid diseases that exhibit dysplastic features.

Splicing factor mutations are found in 85 percent of MDS with increased ring sideroblasts; 45 percent of MDS without increased ring sideroblasts; 55 percent of chronic myelomonocytic leukemia (CMML); and in 25 percent of secondary AML, therapy-related myeloid neoplasms (t-MN), or AML with myelodysplasia-related changes; these mutations are rare in de novo AML (5 to 10 percent) [51]. These mutations affect a number of genes that are components of the pre–messenger RNA splicing machinery including SF3B1, SRSF2, U2AF1, U2AF35, ZRSR2, U2AF65, SF1, SF3A1, and PRPF40B and have, in several studies, been shown to mirror the clinical course of patients with secondary AML or t-MN (eg, inferior response to standard induction chemotherapy).

Distinct molecular subsets with specific alternative spliced isoforms of genes that are involved in innate immune and NF-kB signaling; these findings may provide insight into AML pathogenesis and provide therapeutic targets. As an example, IRAK4-L was the dominant isoform in AML with U2AF1 mutations [52,53]. IRAK4-L interacts with MyD88 to facilitate assembly of the myddosome complex and maximally activate NF-kB and MAPK. IRAK4-L was shown to be essential for leukemic cell function, and inhibition of IRAK4-L suppressed leukemic growth, particularly in AML cells with higher expression of the IRAK4-L isoform. Early phase trials of IRAK-4 inhibitors in MDS and AML have been initiated.

CHROMOSOMAL TRANSLOCATIONS — Chromosomal translocations are common in AML. They can affect the cell by one of two mechanisms [54]:

Juxtaposition of an intact transcription unit from one chromosome to an enhancer element from a gene on another chromosome. As an example, in t(14;18), the BCL-2 gene translocates into the immunoglobulin heavy chain (IgH) locus, leading to the inappropriate expression of a normal BCL-2 gene product. (See "Clinical manifestations, pathologic features, diagnosis, and prognosis of follicular lymphoma", section on 'Cytogenetics'.)

Formation of chimeric fusion proteins. Chromosomal translocations can disrupt two different genes within their coding sequences, leading to the creation of a chimeric protein. For example, the t(9;22) in CML results in the formation of a chimeric BCR/ABL1 fusion protein. (See "Cellular and molecular biology of chronic myeloid leukemia", section on 'The BCR-ABL1 fusion protein'.)

In AML, the vast majority of chromosomal translocations result in the generation of chimeric fusion genes that are never expressed in wild type cells. Characterization of the genes involved in these translocations has led to the elucidation of many transcriptional regulatory pathways in hematopoiesis [55-58]. For example, the alpha and beta subunits of core-binding factor (CBF) have been found to be required for definitive hematopoiesis; both of these subunits are involved in chromosomal translocations in AML.

The involvement in chromosomal translocations of several other classes of transcription factors has been recognized, including those that contain zinc fingers, ring fingers, leucine zippers, basic helix-loop-helix motifs, and ETS domains [59]. Homologues of these transcription factors have been well characterized in animal models such as Drosophila, including the RUNX1 (AML1) homologue "runt," and the KMT2A (MLL) homologue "trithorax." Several other proteins with well-defined hematopoietic functions have been recognized at chromosomal translocation breakpoints, including retinoic acid receptors and Hox genes. However, structural proteins that do not appear to be specific to myeloid cells have also been identified, including members of the nuclear pore complex and nucleolar phosphoproteins. The functions of many other genes involved in chromosomal translocations remain to be elucidated.

Many genes are capable of participating in translocations that involve multiple alternative partner genes. As an example, KMT2A (MLL) forms chimeric fusion genes with >80 known partners. These various fusions provide an opportunity to contrast the contributions of each component of the fusion gene and thus dissect the motifs within these genes that are critical to tumorigenesis. The fusion of nucleophosmin (NPM) to ALK is a frequent recurring abnormality in anaplastic large cell lymphoma. NPM also fuses to RARalpha in acute promyelocytic leukemia and to MLF1 in other subtypes of AML. (See "Clinical manifestations, pathologic features, and diagnosis of systemic anaplastic large cell lymphoma (sALCL)", section on 'Pathogenesis'.)

In the following sections, we will discuss specific fusion genes in the context of families of fusions that share common features. The recurring involvement of the same genes in multiple different translocations has two important implications:

The genomic structure of these genes may contain "hot spots" for chromosomal breakage.

Disruption of the pathways regulated by these genes results in malignant transformation, rather than cell death.

Involvement of core binding factor — Chromosomal aberrations in AML commonly involve subunits of the core-binding factor (CBF) group of transcription factors [60,61]. Members of the CBF group function as heterodimeric complexes regulating diverse target genes involved in tissue differentiation [62]. CBF factors consist of an alpha subunit that binds DNA, and a beta subunit that does not bind DNA directly but enhances binding by the alpha subunit. Three alpha subunits have been identified: RUNX1 (AML1/CBFA2/PEBP2aB), RUNX2 (AML3/CBFA1/PEBP2aA), and RUNX3 (AML2/CBFA3/PEBP2aC). As will be described below, both the alpha and beta subunits of CBF are disrupted in common subtypes of AML and form chimeric fusion transcripts with genes not normally expressed in hematopoietic cells [63].

In hematopoietic cells, CBF binding sites have been observed in genes specific to both the lymphoid lineage (T-cell receptor enhancers, CD3epsilon, and LCK proximal promoter) and myeloid lineage (M-CSF receptor, IL-3, GM-CSF, myeloperoxidase, granzyme B, and neutrophil elastase) [55,64].

The RUNX1 (AML1) subunit has been studied extensively in hematopoiesis. Multiple splice variants have been identified; AML1B contains the runt domain and a transactivation domain, whereas AML1A lacks the transactivation domain [65]. Disruption of RUNX1 by gene targeting in embryonic stem cells results in embryonic lethality due to an absence of fetal liver hematopoiesis and to central nervous system hemorrhage [66]. The knockout of CBFbeta results in a similar phenotype, indicating that both the alpha and beta subunits of CBF are essential for definitive but not primitive hematopoiesis [60,67]. The role of CBF in later stages of hematopoietic differentiation is not yet clear.

Familial AML — Abnormalities in several genes are known to cause inherited forms of myelodysplastic syndrome (MDS) or acute leukemia that can present throughout the lifespan. These include abnormalities in RUNX1, ANKRD26, CEBPA, GATA2, TERC, TERT, SRP72, PAX5, SH2B3, and TP53. These are discussed in more detail separately. (See "Familial disorders of acute leukemia and myelodysplastic syndromes".)

AML and MDS and myeloproliferative neoplasms — Point mutations of RUNX1 (AML1) have been infrequently reported in de novo AML or MDS, at incidences varying from zero to 12 percent [68-70].

In one study of 132 patients with primary MDS, the presence of these mutations was associated with a significantly higher frequency of chromosome 7/7q deletion as well as shortened overall survival [70].

In a study from Japan, RUNX1 (AML1) mutations were found in 6 of 13 Atom-bomb survivors who developed MDS, and in 5 of 13 patients who developed AML/MDS following alkylating agent treatment with or without supplementary radiation [68].

In a study of 18 patients with myeloproliferative neoplasms (eg, polycythemia vera, essential thrombocythemia, primary myelofibrosis) who progressed from chronic phase (CP) to acute leukemia, point mutations of RUNX1 were detected in five at the time of leukemic transformation but not while they were still in CP [71]. Consistent with the leukemogenic potential of RUNX1 point mutations, when the AML1D171N mutation was transduced into CD34+ cells from patients in the CP of MPNs, this resulted in proliferation of immature myeloid cells, enhanced self-renewal capacity, and proliferation of primitive progenitors.

Abnormalities in RUNX1 are also seen in patients with familial platelet disorder with propensity to myeloid malignancies. (See "Familial disorders of acute leukemia and myelodysplastic syndromes", section on 'Familial platelet disorder with propensity to myeloid malignancies (FPD)'.)

Cooperating mutations in AML — The t(8;21)(q22;q22) is associated with the FAB-M2 subtype of AML (picture 1), and was the first recurring chromosomal translocation in AML to be identified [72]. The clinical and morphologic correlations of this subtype of AML include a high propensity to develop extramedullary collections of leukemic blasts (granulocytic sarcomas). Another hallmark of t(8;21) leukemia is the relatively high rate of responsiveness to chemotherapy, particularly to high dose Ara-C [73].

Leukemic myeloblasts in this subtype have a characteristic morphology, including prominent Auer rods, large cytoplasmic granules, and vacuoles. When the characteristic histologic appearance of these cells is identified in AML blasts that do not contain a t(8;21), the RUNX1/RUNX1T1 (ETO) fusion can be identified by polymerase chain reaction techniques, indicating that the formation of the molecular fusion is responsible for the morphologic and clinical features of this subtype of AML [74].

In t(8;21), RUNX1 fuses to the RUNX1T1 (ETO) gene on the der(8) chromosome to form a chimeric fusion protein [75]. The chimeric protein fuses the N-terminus of RUNX1 including the runt homology domain, but excluding the transactivation domain, to nearly all of RUNX1T1. The RUNX1/RUNX1T1 fusion gene has pleiotropic effects in reporter gene assays as it appears to repress several RUNX1 target genes [76,77], but to activate the M-CSF receptor promoter [78,79]. (See "Regulation of myelopoiesis", section on 'Transcription factors'.)

By engineering a "knock-in" of the RUNX1T1 cDNA into the murine RUNX1 locus, a mouse model of t(8;21) has been generated. These mice exhibit a mid-gestation embryonic lethality with an absence of fetal liver hematopoiesis and central nervous system hemorrhage that is similar to the RUNX1 and CBFbeta knockout mice [80]. This result suggests that the RUNX1/RUNX1T1 fusion gene can act as a dominant negative of normal RUNX1 function.

RUNX1/EAP, MDS1, EVI1 in t(3;21) — The t(3;21)(q26;q22) has been observed in the blast crisis of CML and therapy-related MDS and AML [81]. The RUNX1 (AML1) gene has been found to be rearranged in these cases, leading to the formation of alternative fusion transcripts involving the EAP, MDS1, and EVI1 genes [82,83]. These fusions contain the N-terminus of RUNX1, including its runt homology domain, fused to alternative C-terminal sequences derived from partner genes on 3q26. EVI1 is also involved in the inv(3)(q21q26) and t(3;3)(q21;q26) in AML [84]. However, the consequence of these chromosome aberrations is the inappropriate expression of the EVI1 transcription unit, whereas in the t(3;21), EVI1 participates in the formation of a fusion gene with RUNX1. Sequences from another transcription unit, MDS1, are also observed in t(3;21). Although mapping data indicate that the exons of MDS1 and EVI1 are separated by several hundred kilobases, they have been found to be spliced together not only in cases with the t(3;21), but also in normal tissues. The RUNX1-MDS1 fusion has been shown to transform the Rat1A fibroblast cell line [85].

TEL(ETV6) and RUNX1 in t(12;21) — RUNX1 (AML1) is involved in the t(12;21)(p13;q22), resulting in the generation of a chimeric protein with the TEL(ETV6) gene [86]. The t(12;21) fusion is observed commonly in pre-B-cell ALL. The TEL (ETV6) gene was originally cloned from a t(5;12)(q33;p13) in CMML where it fuses to the PDGFRbeta gene and it has also been found to fuse to the ABL1 gene in t(12;22) in AML [87,88].

CBFb/MYH11 in inv(16) and t(16;16) — The inv(16)(p13q22) and t(16;16)(p13;q22) are observed primarily in the FAB-M4Eo subtype of AML (picture 2). These patients, similar to those with the t(8;21), have a relatively good prognosis and a high likelihood to respond to high dose cytosine arabinoside-containing chemotherapy regimens [89]. The CBFbeta subunit located at 16q22 is disrupted as a result of either the chromosomal inversion or translocation, leading to the formation of a fusion transcript with the gene for smooth muscle myosin heavy chain, MYH11, located at 16p13 [90].

A mouse model for inv(16) leukemia has been generated through a "knock-in" approach [91]. These mice exhibit a mid-gestation embryonic lethality, an absence of fetal liver hematopoiesis, and central nervous system hemorrhage. The chimeric mice generated from embryonic stem cells containing the CBFb/MYH11 fusion did not develop leukemia. However, there was no contribution by the embryonic stem cells to hematopoiesis. Thus, a striking similarity has been observed for the knockouts of RUNX1 (AML1) and CBFbeta, as well as the RUNX1/RUNX1T1 and CBFb/MYH11 "knock-in" mice. This convergence in phenotypes from seemingly disparate gene targeting experiments indicates that RUNX1 and CBFbeta are both essential for definitive hematopoiesis, and that both are required for the function of the CBF heterodimer.

Involvement of the KMT2A (MLL) locus — The mixed lineage leukemia gene, KMT2A (also known as MLL), located at chromosome band 11q23, is frequently involved in translocations in both AML and ALL and is associated with a poor prognosis. Translocations involving 11q23 are the single most common cytogenetic abnormality in infants with acute leukemia, regardless of phenotype, and occur in approximately 70 to 80 percent of cases. 11q23 translocations are also observed frequently in therapy-related leukemias in patients who have previously been treated with drugs that inhibit topoisomerase II, especially the epipodophyllotoxins. This syndrome differs from the therapy-related myeloid leukemias that occur in patients exposed to alkylating agents, which are commonly associated with deletional events on chromosomes 5 and 7. (See "Acute myeloid leukemia: Cytogenetic abnormalities", section on 'Therapy-related myeloid neoplasms'.)

Partner chromosomes — There were at least 121 different partner loci involved in recurring reciprocal 11q23 translocations in a molecular analysis of 1590 KMT2A (MLL)-rearranged biopsy samples from adults and children with AML and ALL with KMT2A gene rearrangements [92-96]. This exceeds the number of known translocations affecting the immunoglobulin loci, suggesting that the 11q23 breakpoint region may contain genomically unstable sequences leading to recombination events.

The chromosomal partners in 11q23 translocations are usually lineage specific. In AML, t(9;11)(p22;q23), t(11;19)(q23;p13.1) and t(6;11)(q27;q23) are the most common, and in ALL, t(4;11)(q21;q23) and t(11;19)(q23;p13.3) occur predominantly. Translocations at 11q23 have been observed in several FAB subtypes, but occur most commonly in the FAB-M4 myelomonocytic and FAB-M5 monoblastic leukemias [93,97,98]. Myeloid leukemias with 11q23 translocations often coexpress lymphoid markers, whereas 11q23 lymphoid leukemias often express myeloid or monocytoid markers in addition to B-cell markers. These observations suggest that rearrangements of KMT2A may affect a pluripotent stem cell or, alternatively, that disruption of KMT2A may affect a common differentiation pathway shared by lymphoid and myeloid progenitor cells.

In a subset of patients with AML and either trisomy 11 or a normal karyotype, a unique pattern of rearrangement of the KMT2A gene has been observed. As in the translocations that affect KMT2A, a fusion occurs involving one of the exons in its breakpoint cluster region. However, rather than fusing to a partner gene, the fusion is to 5' sequences from KMT2A itself [99,100]. These partial duplications of KMT2A appear to occur primarily in older patients and are infrequent in childhood and therapy-related leukemias. The morphology of the leukemias also differs in that the partial duplication patients usually are classified as FAB M1 or M2, rather than the M4 or M5 typically observed in cases with KMT2A translocations.

Mechanism of action — The KMT2A (MLL) gene was isolated from the 11q23 breakpoint by several groups and has been referred to by other names including HRX, ALL-1, and Htrx [101-104]. KMT2A encodes a large protein (predicted molecular weight of 430 kD) that contains two regions of extensive homology to the Drosophila trithorax gene [105]. KMT2A is a histone methyltransferase that plays a critical role in normal embryonic development and hematopoiesis, mainly regulating transcription via epigenetic mechanisms [106]. Translocations create KMT2A fusion proteins that induce aberrant expression of downstream mediators, including HOXA proteins and myeloid ecotropic viral integration site 1 (MEIS1) [107-110]. KMT2A fusion proteins also appear to deregulate the RNA polymerase II elongation factor, thereby disrupting checkpoints for transcription elongation [111,112]. Both Menin and lens epithelium-derived growth factor (LEDGF) appear to be key elements in the interaction between the KMT2A histone methyltransferase and the nonmethylated CpG DNA to which it binds [113]. A selective inhibitor of the KMT2A-Menin interaction was shown to inhibit growth of cell lines harboring KMT2A fusions and prolong survival in mouse models of KMT2A leukemia [114].

Several animal and cell culture models of KMT2A leukemia have been developed. To create an KMT2A-AF9 "knock-in" mouse model of 11q23 leukemias, the AF9 cDNA was targeted into the murine KMT2A locus by homologous recombination in embryonic stem cells [115]. The chimeric mice generated from these embryonic stem cells all developed acute myeloid leukemia within 6 to 9 months. Retroviral bone marrow infection has been used to express another 11q23 fusion gene, KMT2A-ENL, in hematopoietic cells [116]. After transplantation of KMT2A-ENL transduced bone marrow, recipient mice developed acute myeloid leukemia. Retroviral transduction/transplantation models using several KMT2A partner genes have been established [117,118]. These models have established the proof of principle that the expression of KMT2A fusion genes contributes to leukemogenesis, and that the KMT2A partner genes are essential to this process [94,119].

Involvement of the retinoic acid receptor — The cytogenetic hallmark of acute promyelocytic leukemia (APL, FAB-M3), is a translocation involving the retinoic acid receptor-alpha (RARalpha) locus on chromosome 17 [120]. The vast majority of these cases contain a t(15;17)(q22;q11-12), although several variant translocations involving RARalpha have been identified. These include PLZF in t(11;17)(q23;q11-12) and NPM in t(5;17)(q35;q11-12). These translocations lead to the formation of chimeric fusion transcripts from each of the derivative chromosomes. The pathobiology of APL is discussed in detail separately. (See "Molecular biology of acute promyelocytic leukemia".)

Other fusion genes in AML

DEK/CAN in t(6;9) and SET-CAN — The t(6;9)(p23;q34) is observed usually in the M2 and M4 subtypes of AML but has also been identified in the M1 subtype and MDS. This translocation is associated with a younger age and a poor prognosis. The CAN gene was also found to be rearranged in a patient with acute undifferentiated leukemia and a normal karyotype. A chimeric transcript containing an alternative CAN partner gene, named SET, was isolated in this case [121]. The SET gene was also localized to 9q34, suggesting that this fusion results from an inversion of chromosome 9 or a cryptic translocation.

TLS/FUS/ERG and t(16;21) — The t(16;21)(p11;q22) has been observed in several FAB subtypes of AML, blast crisis of CML, and MDS [122,123]. In this translocation, the TLS/FUS gene at 16p11 fuses to the ERG gene at 21q22 to generate a chimeric protein. Both TLS/FUS and ERG are involved in chromosomal translocations with other partner genes in sarcomas [124-126]. The striking similarity of the TLS/FUS/ERG fusion gene in AML to the sarcoma associated fusion genes suggests that disruption of a common differentiation pathway may lead to transformation in multiple cell types.

NPM/MLF1 in t(3;5) — The t(3;5)(q25.1;q35) has been observed in several FAB subtypes of AML, especially in M6 and MDS [127]. The NPM gene at 5q35 was found to be rearranged in patients with this translocation, resulting in the formation of a fusion gene with the MLF1 (myelodysplasia-myeloid leukemia factor 1) gene at 3q25.1. Previously, the involvement of NPM had been identified in the t(2;5)(p23;q35) in anaplastic large cell lymphoma, where it fuses to the ALK gene [128]. NPM is also rearranged in the t(5;17)(q35;q21) in APL, where it fuses to RARalpha.

The MLF1 protein normally localizes to the cytoplasm, whereas in leukemia cells with t(3;5), the NPM-MLF1 fusion protein is observed primarily in nucleoli. The involvement of NPM in chromosome translocations involving three different partner genes, resulting in three discrete types of hematologic malignancies, suggests that NPM plays an important role in normal hematopoietic differentiation.

EVI1 in inv(3) and t(3;3) — EVI1 (Ecotropic Virus Integration 1) was initially cloned as the gene at a common site of retroviral integration in murine myeloid leukemia. The EVI1 gene is normally expressed only in kidney and ovary. In mice, the retroviral integration at the murine Evi1 locus leads to its inappropriate expression in hematopoietic cells. A similar phenomenon occurs in human AML with inv(3)(q21q26) and t(3;3)(q21;q26) [129]. As a result of these cytogenetic aberrations, the EVI1 gene is juxtaposed to enhancer elements of the Ribophorin gene leading to inappropriate activation of the EVI1 transcription unit, a component of the complex fusion transcripts that involve the RUNX1 gene in t(3;21).

MOZ/CBP in t(8;16) — The t(8;16)(p11;p13) has been recognized in the M4 and M5 subtypes of AML that exhibit a characteristic erythrophagocytosis [130]. This translocation has been observed in both de novo and therapy-related leukemias. Several variant translocations involving 8p11 have been recognized in monocytic leukemias with erythrophagocytosis, but the genes involved in these translocations have not yet been cloned. CBP is also involved with a fusion to KMT2A in t(11;16)(q23;p13) [131].

CALM/AF10 in t(10;11) — The t(10;11)(p13;q14) has been recognized in both ALL and AML. In AML, the cases are primarily of the M0 and M1 subtypes [132].

NUP98/HOXA9 in t(7;11) and NUP98/DDX10 in inv(11) — The t(7;11)(p15;p15) is observed in the M2 and M4 subtypes of AML and CML. The HOXA9 gene was identified at the 7p15 breakpoint and the NUP98 gene at the 11p15 break [133]. In BXH-2 mice, the Hoxa9 gene has been implicated in myeloid leukemia as a result of its activation via proviral insertion. NUP98 has been found to be involved in the inv(11)(p15q22) in AML. This cytogenetic aberration has been recognized in both de novo and therapy-related MDS and AML.

In addition, a t(10;11)(q23;p15) translocation involving the NUP98 gene at 11q15 and the hematopoietically expressed homeobox gene (HHEX) at 10q23 has been implicated in AML [134]. Other cases of AML have been identified that demonstrate a NUP98-KMT2A fusion protein with low levels of HOXA gene expression [95].

ANIMAL MODELS OF AML — There are several examples of naturally occurring animal models of AML that are caused by retroviruses via one of two mechanisms. The retrovirus can:

Encode for an oncogene that leads to transformation, or

It can inappropriately activate the expression of a gene adjacent to its integration site

The characterization of these processes has revealed a number of genes critical to normal hematopoiesis and myeloid leukemias. These animal models have also demonstrated the multistage nature of the evolution of leukemia.

Myb leukemias — The v-Myb gene is involved in two different avian retroviruses that induce leukemia in chickens. The v-Myb gene was originally identified as the transforming element in the avian myeloblastosis virus (AMV). In the E26 retrovirus, a fusion gene consisting of gag, v-Myb, and v-Ets has been identified [135]. The formation of a fusion gene containing sequences of two transcription factors is reminiscent of the fusions that result from chromosomal translocations in human acute leukemia. In chickens, AMV causes monoblastic leukemias and transforms only myelomonocytic cells in vitro, whereas E26 induces erythroleukemias, but is capable of transforming myeloid, erythroid, and megakaryocytic lineages in vitro [136].

C-Myb, the normal cellular counterpart of v-Myb, is essential for definitive hematopoiesis [137]. Promonocytic leukemias have been induced by priming mice with intraperitoneal injections of pristane, followed by infection with the Moloney murine leukemia virus [138]. The majority of the mice that develop leukemia have undergone insertional mutagenesis at the c-Myb locus [139]. The role of pristane is unknown, but it appears to promote leukemogenesis by mediating an inflammatory response. At this time, there is no evidence for the rearrangement of c-Myb in human leukemias.

Friend virus/SFFV erythroleukemia — The erythroleukemia induced by infection with the Friend viral complex is the culmination of a three step process that has become a paradigm for multistage neoplastic transformation [140]:

Polyclonal proliferation induced by a cytokine

Inactivation of a tumor suppressor gene

Oncogenic activation of a gene involved in normal hematopoiesis

Infection of susceptible mice with the Friend murine leukemia virus/spleen focus-forming virus (SFFV) complex induces a polyclonal proliferation of erythroid progenitors within 48 hours. This expansion of erythroid precursors is secondary to the SFFV envelope gene product, the gp55 glycoprotein, which has the capacity to bind and activate the erythropoietin receptor. These erythroid cells are not transformed and retain the ability to undergo terminal differentiation.

After approximately two weeks, a clonal population of transformed erythroblasts emerges which depends on two additional independent genetic events. One of these is the inactivation of the p53 tumor suppressor gene by deletion, mutation, or proviral insertion [141,142]. In addition, the activation of either the PU.1 gene, also referred to as Spi-1 (SFFV proviral integration 1), or the Fli-1 gene occur as a result of proviral integration adjacent to these loci [143,144]. PU.1 and Fli-1 are members of the ETS family of transcription factors; PU.1 is essential for both myeloid and lymphoid development [145].

BXH-2 leukemias — BXH-2 mice are derived from a cross of C57BL/6J and C3H/He mice. Although the parental strains have a low incidence of leukemia, more than 95 percent of BXH-2 mice develop myeloid leukemia by one year of age, caused by expression of a horizontally transmitted Ecotropic murine leukemia virus [146]. These retroviruses induce leukemia by insertional activation or mutation of proto-oncogenes adjacent to the viral integration site. One of the most frequent viral integration sites is Evi-2 (Ecotropic viral integration site 2), localized to a large intron within the Nf1 tumor suppressor gene [147]. The consequence of this integration event is the disruption of the normal expression of Nf1 in the affected mice.

In humans, the neurofibromatosis type I autosomal dominant disorder is caused by mutation of the NF1 gene [148]. In addition to neurofibromas, patients have an increased risk of developing several types of solid tumors and malignant myeloid disorders, especially juvenile chronic myeloid leukemia (JCML) and the monosomy 7 syndrome [149].

Mice heterozygous for the Nf1 allele develop myeloid leukemia with loss of the wild type allele, indicating that Nf1 acts as a tumor suppressor [150]. In JCML patients without neurofibromatosis, activating RAS mutations are frequently identified. The consequence of either the disruption of NF1 or the activation of RAS has been shown to be an inability to negatively regulate GM-CSF signaling in hematopoietic cells [151,152].

Because viral integration within the Nf1 gene has been observed in only 15 percent of BXH-2 mice, the identification of additional mutations within other genes associated with myeloid leukemias has been pursued. A proviral tagging approach has been used to identify additional viral integration sites that alter the expression of three genes, Hoxa7, Hoxa9, and Meis-1 [153,154]. The HOXA9 gene is involved in t(7;11)(p15;p15) in AML, resulting in the formation of a fusion gene with NUP98. HOX gene overexpression has also been demonstrated in mice with acquired nucleophosmin mutations that result in the cytoplasmic dislocation of nucleophosmin (ie, NPM1c) [155]. PBX1 is involved in t(1;19) in ALL, where it fuses to the E2A gene. This underscores a recurring theme of the involvement of similar families of genes in both human and spontaneous animal models of AML.

INFORMATION FOR PATIENTS — UpToDate offers two types of patient education materials, "The Basics" and "Beyond the Basics." The Basics patient education pieces are written in plain language, at the 5th to 6th grade reading level, and they answer the four or five key questions a patient might have about a given condition. These articles are best for patients who want a general overview and who prefer short, easy-to-read materials. Beyond the Basics patient education pieces are longer, more sophisticated, and more detailed. These articles are written at the 10th to 12th grade reading level and are best for patients who want in-depth information and are comfortable with some medical jargon.

Here are the patient education articles that are relevant to this topic. We encourage you to print or e-mail these topics to your patients. (You can also locate patient education articles on a variety of subjects by searching on "patient info" and the keyword(s) of interest.)

Beyond the Basics topics (see "Patient education: Acute myeloid leukemia (AML) treatment in adults (Beyond the Basics)")

SUMMARY

Acute myeloid leukemia (AML) – AML develops as the consequence of genetic changes in a hematopoietic precursor cell that alters normal growth and differentiation, resulting in an accumulation of large numbers of abnormal, immature myeloid cells in the bone marrow and blood. These aberrant cells can divide and proliferate, but they cannot differentiate normally into mature, fully functional hematopoietic cells.

Prognostic significance – Many of the chromosomal abnormalities and gene mutations are associated with prognosis in AML (table 1 and table 2).

Mutations of oncogenes and tumor suppressors – Cases of AML may demonstrate mutations in oncogenes (eg, NRAS, FLT3) and/or loss of tumor suppressor genes (eg, TP53, WT1 [Wilms tumor suppressor gene]). (See 'Gene mutations' above.)

Chromosomal translocations – Many hematologic malignancies are associated with specific chromosomal translocations that generate chimeric fusion genes that are never expressed in wild type cells. Chromosomal aberrations in AML commonly involve transcription factors that regulate hematopoiesis. (See 'Chromosomal translocations' above.)

Core binding factor abnormalities – The chromosomal aberrations t(8;21)(q22;q22) and the inv(16)(p13q22) in AML involve subunits of the core-binding factor (CBF) group of transcription factors. Members of the CBF group function as heterodimeric complexes regulating diverse target genes involved in tissue differentiation. (See 'Involvement of core binding factor' above.)

KMT2A (MLL) – The mixed lineage leukemia gene, KMT2A (MLL), located at chromosome band 11q23, is frequently involved in translocations in AML with different partner chromosomes. KMT2A is a histone methyltransferase that plays a critical role in normal embryonic development and hematopoiesis, mainly regulating transcription via epigenetic mechanisms. Translocations involving KMT2A result in KMT2A fusion proteins that induce the aberrant expression of downstream mediators. (See 'Involvement of the KMT2A (MLL) locus' above.)

Acute promyelocytic leukemia (APL) – The cytogenetic hallmark of APL is a translocation involving the retinoic acid receptor-alpha (RARA) locus on chromosome 17. These translocations lead to the formation of chimeric fusion transcripts from each of the derivative chromosomes. The pathobiology of acute promyelocytic leukemia is discussed in detail separately. (See "Molecular biology of acute promyelocytic leukemia".)

  1. Walter MJ, Payton JE, Ries RE, et al. Acquired copy number alterations in adult acute myeloid leukemia genomes. Proc Natl Acad Sci U S A 2009; 106:12950.
  2. Radtke I, Mullighan CG, Ishii M, et al. Genomic analysis reveals few genetic alterations in pediatric acute myeloid leukemia. Proc Natl Acad Sci U S A 2009; 106:12944.
  3. Cancer Genome Atlas Research Network, Ley TJ, Miller C, et al. Genomic and epigenomic landscapes of adult de novo acute myeloid leukemia. N Engl J Med 2013; 368:2059.
  4. Grimwade D, Ivey A, Huntly BJ. Molecular landscape of acute myeloid leukemia in younger adults and its clinical relevance. Blood 2016; 127:29.
  5. Coombs CC, Tallman MS, Levine RL. Molecular therapy for acute myeloid leukaemia. Nat Rev Clin Oncol 2016; 13:305.
  6. Marcucci G, Yan P, Maharry K, et al. Epigenetics meets genetics in acute myeloid leukemia: clinical impact of a novel seven-gene score. J Clin Oncol 2014; 32:548.
  7. Cheung N, Fung TK, Zeisig BB, et al. Targeting Aberrant Epigenetic Networks Mediated by PRMT1 and KDM4C in Acute Myeloid Leukemia. Cancer Cell 2016; 29:32.
  8. Ley TJ, Ding L, Walter MJ, et al. DNMT3A mutations in acute myeloid leukemia. N Engl J Med 2010; 363:2424.
  9. Yan XJ, Xu J, Gu ZH, et al. Exome sequencing identifies somatic mutations of DNA methyltransferase gene DNMT3A in acute monocytic leukemia. Nat Genet 2011; 43:309.
  10. Ito S, D'Alessio AC, Taranova OV, et al. Role of Tet proteins in 5mC to 5hmC conversion, ES-cell self-renewal and inner cell mass specification. Nature 2010; 466:1129.
  11. Delhommeau F, Dupont S, Della Valle V, et al. Mutation in TET2 in myeloid cancers. N Engl J Med 2009; 360:2289.
  12. Li Z, Cai X, Cai CL, et al. Deletion of Tet2 in mice leads to dysregulated hematopoietic stem cells and subsequent development of myeloid malignancies. Blood 2011; 118:4509.
  13. Wang Y, Xiao M, Chen X, et al. WT1 recruits TET2 to regulate its target gene expression and suppress leukemia cell proliferation. Mol Cell 2015; 57:662.
  14. Rampal R, Alkalin A, Madzo J, et al. DNA hydroxymethylation profiling reveals that WT1 mutations result in loss of TET2 function in acute myeloid leukemia. Cell Rep 2014; 9:1841.
  15. Green CL, Evans CM, Zhao L, et al. The prognostic significance of IDH2 mutations in AML depends on the location of the mutation. Blood 2011; 118:409.
  16. Boissel N, Nibourel O, Renneville A, et al. Differential prognosis impact of IDH2 mutations in cytogenetically normal acute myeloid leukemia. Blood 2011; 117:3696.
  17. Katoh M. Functional and cancer genomics of ASXL family members. Br J Cancer 2013; 109:299.
  18. Ahuja H, Bar-Eli M, Arlin Z, et al. The spectrum of molecular alterations in the evolution of chronic myelocytic leukemia. J Clin Invest 1991; 87:2042.
  19. Kelman Z, Prokocimer M, Peller S, et al. Rearrangements in the p53 gene in Philadelphia chromosome positive chronic myelogenous leukemia. Blood 1989; 74:2318.
  20. Mashal R, Shtalrid M, Talpaz M, et al. Rearrangement and expression of p53 in the chronic phase and blast crisis of chronic myelogenous leukemia. Blood 1990; 75:180.
  21. Mori N, Hidai H, Yokota J, et al. Mutations of the p53 gene in myelodysplastic syndrome and overt leukemia. Leuk Res 1995; 19:869.
  22. Lanza C, Gaidano G, Cimino G, et al. p53 gene inactivation in acute lymphoblastic leukemia of B cell lineage associates with chromosomal breakpoints at 11q23 and 8q24. Leukemia 1995; 9:955.
  23. Lanza C, Gaidano G, Cimino G, et al. Distribution of TP53 mutations among acute leukemias with MLL rearrangements. Genes Chromosomes Cancer 1996; 15:48.
  24. Rücker FG, Schlenk RF, Bullinger L, 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.
  25. Inoue K, Sugiyama H, Ogawa H, et al. WT1 as a new prognostic factor and a new marker for the detection of minimal residual disease in acute leukemia. Blood 1994; 84:3071.
  26. Bergmann L, Miething C, Maurer U, et al. High levels of Wilms' tumor gene (wt1) mRNA in acute myeloid leukemias are associated with a worse long-term outcome. Blood 1997; 90:1217.
  27. Call KM, Glaser T, Ito CY, et al. Isolation and characterization of a zinc finger polypeptide gene at the human chromosome 11 Wilms' tumor locus. Cell 1990; 60:509.
  28. Haber DA, Buckler AJ, Glaser T, et al. An internal deletion within an 11p13 zinc finger gene contributes to the development of Wilms' tumor. Cell 1990; 61:1257.
  29. Baird PN, Simmons PJ. Expression of the Wilms' tumor gene (WT1) in normal hemopoiesis. Exp Hematol 1997; 25:312.
  30. Stiller CA, Chessells JM, Fitchett M. Neurofibromatosis and childhood leukaemia/lymphoma: a population-based UKCCSG study. Br J Cancer 1994; 70:969.
  31. Shannon KM, O'Connell P, Martin GA, et al. Loss of the normal NF1 allele from the bone marrow of children with type 1 neurofibromatosis and malignant myeloid disorders. N Engl J Med 1994; 330:597.
  32. Hirama T, Koeffler HP. Role of the cyclin-dependent kinase inhibitors in the development of cancer. Blood 1995; 86:841.
  33. Campos L, Rouault JP, Sabido O, et al. High expression of bcl-2 protein in acute myeloid leukemia cells is associated with poor response to chemotherapy. Blood 1993; 81:3091.
  34. Le Beau MM, Albain KS, Larson RA, et al. Clinical and cytogenetic correlations in 63 patients with therapy-related myelodysplastic syndromes and acute nonlymphocytic leukemia: further evidence for characteristic abnormalities of chromosomes no. 5 and 7. J Clin Oncol 1986; 4:325.
  35. Pedersen-Bjergaard J, Philip P. Two different classes of therapy-related and de-novo acute myeloid leukemia? Cancer Genet Cytogenet 1991; 55:119.
  36. Thirman MJ, Larson RA. Therapy-related myeloid leukemia. Hematol Oncol Clin North Am 1996; 10:293.
  37. Yamamoto Y, Kiyoi H, Nakano Y, et al. Activating mutation of D835 within the activation loop of FLT3 in human hematologic malignancies. Blood 2001; 97:2434.
  38. Bacher U, Haferlach C, Kern W, et al. Prognostic relevance of FLT3-TKD mutations in AML: the combination matters--an analysis of 3082 patients. Blood 2008; 111:2527.
  39. Zeleznik-Le NJ, Nucifora G, Rowley JD. The molecular biology of myeloproliferative disorders as revealed by chromosomal abnormalities. Semin Hematol 1995; 32:201.
  40. Ahuja HG, Foti A, Bar-Eli M, Cline MJ. The pattern of mutational involvement of RAS genes in human hematologic malignancies determined by DNA amplification and direct sequencing. Blood 1990; 75:1684.
  41. Neubauer A, Dodge RK, George SL, et al. Prognostic importance of mutations in the ras proto-oncogenes in de novo acute myeloid leukemia. Blood 1994; 83:1603.
  42. Bartram CR. Molecular genetic aspects of myelodysplastic syndromes. Hematol Oncol Clin North Am 1992; 6:557.
  43. Nakagawa T, Saitoh S, Imoto S, et al. Multiple point mutation of N-ras and K-ras oncogenes in myelodysplastic syndrome and acute myelogenous leukemia. Oncology 1992; 49:114.
  44. Bos JL. ras oncogenes in human cancer: a review. Cancer Res 1989; 49:4682.
  45. Taylor JA, Sandler DP, Bloomfield CD, et al. ras oncogene activation and occupational exposures in acute myeloid leukemia. J Natl Cancer Inst 1992; 84:1626.
  46. Lübbert M, Mirro J Jr, Kitchingman G, et al. Prevalence of N-ras mutations in children with myelodysplastic syndromes and acute myeloid leukemia. Oncogene 1992; 7:263.
  47. Paquette RL, Landaw EM, Pierre RV, et al. N-ras mutations are associated with poor prognosis and increased risk of leukemia in myelodysplastic syndrome. Blood 1993; 82:590.
  48. Liu ET, Dodge R, Myer A, et al. De novo acute myeloid leukemias harboring ras mutations exhibit preferential sensitivity to dose intensive cytarabine as consolidation therapy. Blood 1995; 86:598a.
  49. Falini B, Mecucci C, Tiacci E, et al. Cytoplasmic nucleophosmin in acute myelogenous leukemia with a normal karyotype. N Engl J Med 2005; 352:254.
  50. Brunetti L, Gundry MC, Goodell MA. New insights into the biology of acute myeloid leukemia with mutated NPM1. Int J Hematol 2019; 110:150.
  51. Yoshida K, Sanada M, Shiraishi Y, et al. Frequent pathway mutations of splicing machinery in myelodysplasia. Nature 2011; 478:64.
  52. Pellagatti A, Armstrong RN, Steeples V, et al. Impact of spliceosome mutations on RNA splicing in myelodysplasia: dysregulated genes/pathways and clinical associations. Blood 2018; 132:1225.
  53. Smith MA, Choudhary GS, Pellagatti A, et al. U2AF1 mutations induce oncogenic IRAK4 isoforms and activate innate immune pathways in myeloid malignancies. Nat Cell Biol 2019; 21:640.
  54. Rabbitts TH. Chromosomal translocations in human cancer. Nature 1994; 372:143.
  55. Tenen DG, Hromas R, Licht JD, Zhang DE. Transcription factors, normal myeloid development, and leukemia. Blood 1997; 90:489.
  56. Sawyers CL. Molecular genetics of acute leukaemia. Lancet 1997; 349:196.
  57. Shivdasani RA, Orkin SH. The transcriptional control of hematopoiesis. Blood 1996; 87:4025.
  58. Alcalay M, Meani N, Gelmetti V, et al. Acute myeloid leukemia fusion proteins deregulate genes involved in stem cell maintenance and DNA repair. J Clin Invest 2003; 112:1751.
  59. Nichols J, Nimer SD. Transcription factors, translocations, and leukemia. Blood 1992; 80:2953.
  60. Wang Q, Stacy T, Miller JD, et al. The CBFbeta subunit is essential for CBFalpha2 (AML1) function in vivo. Cell 1996; 87:697.
  61. Marcucci G, Caligiuri MA, Bloomfield CD. Molecular and clinical advances in core binding factor primary acute myeloid leukemia: a paradigm for translational research in malignant hematology. Cancer Invest 2000; 18:768.
  62. Rodan GA, Harada S. The missing bone. Cell 1997; 89:677.
  63. Michaud J, Scott HS, Escher R. AML1 interconnected pathways of leukemogenesis. Cancer Invest 2003; 21:105.
  64. Wang S, Wang Q, Crute BE, et al. Cloning and characterization of subunits of the T-cell receptor and murine leukemia virus enhancer core-binding factor. Mol Cell Biol 1993; 13:3324.
  65. Tanaka T, Tanaka K, Ogawa S, et al. An acute myeloid leukemia gene, AML1, regulates hemopoietic myeloid cell differentiation and transcriptional activation antagonistically by two alternative spliced forms. EMBO J 1995; 14:341.
  66. Okuda T, van Deursen J, Hiebert SW, et al. AML1, the target of multiple chromosomal translocations in human leukemia, is essential for normal fetal liver hematopoiesis. Cell 1996; 84:321.
  67. Sasaki K, Yagi H, Bronson RT, et al. Absence of fetal liver hematopoiesis in mice deficient in transcriptional coactivator core binding factor beta. Proc Natl Acad Sci U S A 1996; 93:12359.
  68. Harada H, Harada Y, Tanaka H, et al. Implications of somatic mutations in the AML1 gene in radiation-associated and therapy-related myelodysplastic syndrome/acute myeloid leukemia. Blood 2003; 101:673.
  69. Nakao M, Horiike S, Fukushima-Nakase Y, et al. Novel loss-of-function mutations of the haematopoiesis-related transcription factor, acute myeloid leukaemia 1/runt-related transcription factor 1, detected in acute myeloblastic leukaemia and myelodysplastic syndrome. Br J Haematol 2004; 125:709.
  70. Chen CY, Lin LI, Tang JL, et al. 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.
  71. Ding Y, Harada Y, Imagawa J, et al. AML1/RUNX1 point mutation possibly promotes leukemic transformation in myeloproliferative neoplasms. Blood 2009; 114:5201.
  72. Nucifora G, Rowley JD. AML1 and the 8;21 and 3;21 translocations in acute and chronic myeloid leukemia. Blood 1995; 86:1.
  73. Bloomfield CD, Lawrence D, Arthur DC, et al. Curative impact of intensification with high-dose cytarabine (HiDAC) in acute myeloid leukemia (AML) varies by cytogenetic group. Blood 1994; 84:111a.
  74. Nucifora G, Dickstein JI, Torbenson V, et al. Correlation between cell morphology and expression of the AML1/ETO chimeric transcript in patients with acute myeloid leukemia without the t(8;21). Leukemia 1994; 8:1533.
  75. Erickson P, Gao J, Chang KS, et al. Identification of breakpoints in t(8;21) acute myelogenous leukemia and isolation of a fusion transcript, AML1/ETO, with similarity to Drosophila segmentation gene, runt. Blood 1992; 80:1825.
  76. Linggi B, Müller-Tidow C, van de Locht L, et al. The t(8;21) fusion protein, AML1 ETO, specifically represses the transcription of the p14(ARF) tumor suppressor in acute myeloid leukemia. Nat Med 2002; 8:743.
  77. Vangala RK, Heiss-Neumann MS, Rangatia JS, et al. The myeloid master regulator transcription factor PU.1 is inactivated by AML1-ETO in t(8;21) myeloid leukemia. Blood 2003; 101:270.
  78. Rhoades KL, Hetherington CJ, Rowley JD, et al. Synergistic up-regulation of the myeloid-specific promoter for the macrophage colony-stimulating factor receptor by AML1 and the t(8;21) fusion protein may contribute to leukemogenesis. Proc Natl Acad Sci U S A 1996; 93:11895.
  79. Pabst T, Mueller BU, Harakawa N, et al. AML1-ETO downregulates the granulocytic differentiation factor C/EBPalpha in t(8;21) myeloid leukemia. Nat Med 2001; 7:444.
  80. Yergeau DA, Hetherington CJ, Wang Q, et al. Embryonic lethality and impairment of haematopoiesis in mice heterozygous for an AML1-ETO fusion gene. Nat Genet 1997; 15:303.
  81. Yin CC, Cortes J, Barkoh B, et al. t(3;21)(q26;q22) in myeloid leukemia: an aggressive syndrome of blast transformation associated with hydroxyurea or antimetabolite therapy. Cancer 2006; 106:1730.
  82. Nucifora G, Birn DJ, Espinosa R 3rd, et al. Involvement of the AML1 gene in the t(3;21) in therapy-related leukemia and in chronic myeloid leukemia in blast crisis. Blood 1993; 81:2728.
  83. Mitani K, Ogawa S, Tanaka T, et al. Generation of the AML1-EVI-1 fusion gene in the t(3;21)(q26;q22) causes blastic crisis in chronic myelocytic leukemia. EMBO J 1994; 13:504.
  84. Suzukawa K, Parganas E, Gajjar A, et al. Identification of a breakpoint cluster region 3' of the ribophorin I gene at 3q21 associated with the transcriptional activation of the EVI1 gene in acute myelogenous leukemias with inv(3)(q21q26). Blood 1994; 84:2681.
  85. Zent CS, Mathieu C, Claxton DF, et al. The chimeric genes AML1/MDS1 and AML1/EAP inhibit AML1B activation at the CSF1R promoter, but only AML1/MDS1 has tumor-promoter properties. Proc Natl Acad Sci U S A 1996; 93:1044.
  86. Golub TR, Barker GF, Bohlander SK, et al. Fusion of the TEL gene on 12p13 to the AML1 gene on 21q22 in acute lymphoblastic leukemia. Proc Natl Acad Sci U S A 1995; 92:4917.
  87. Golub TR, Barker GF, Lovett M, Gilliland DG. Fusion of PDGF receptor beta to a novel ets-like gene, tel, in chronic myelomonocytic leukemia with t(5;12) chromosomal translocation. Cell 1994; 77:307.
  88. Golub TR, Goga A, Barker GF, et al. Oligomerization of the ABL tyrosine kinase by the Ets protein TEL in human leukemia. Mol Cell Biol 1996; 16:4107.
  89. Bloomfield CD, Lawrence D, Byrd JC, et al. Frequency of prolonged remission duration after high-dose cytarabine intensification in acute myeloid leukemia varies by cytogenetic subtype. Cancer Res 1998; 58:4173.
  90. Liu P, Tarlé SA, Hajra A, et al. Fusion between transcription factor CBF beta/PEBP2 beta and a myosin heavy chain in acute myeloid leukemia. Science 1993; 261:1041.
  91. Castilla LH, Wijmenga C, Wang Q, et al. Failure of embryonic hematopoiesis and lethal hemorrhages in mouse embryos heterozygous for a knocked-in leukemia gene CBFB-MYH11. Cell 1996; 87:687.
  92. Thirman MJ, Gill HJ, Burnett RC, et al. Rearrangement of the MLL gene in acute lymphoblastic and acute myeloid leukemias with 11q23 chromosomal translocations. N Engl J Med 1993; 329:909.
  93. Schoch C, Schnittger S, Klaus M, et al. AML with 11q23/MLL abnormalities as defined by the WHO classification: incidence, partner chromosomes, FAB subtype, age distribution, and prognostic impact in an unselected series of 1897 cytogenetically analyzed AML cases. Blood 2003; 102:2395.
  94. Basecke J, Whelan JT, Griesinger F, Bertrand FE. The MLL partial tandem duplication in acute myeloid leukaemia. Br J Haematol 2006; 135:438.
  95. Kaltenbach S, Soler G, Barin C, et al. NUP98-MLL fusion in human acute myeloblastic leukemia. Blood 2010; 116:2332.
  96. Meyer C, Hofmann J, Burmeister T, et al. The MLL recombinome of acute leukemias in 2013. Leukemia 2013; 27:2165.
  97. Blum W, Mrózek K, Ruppert AS, et al. Adult de novo acute myeloid leukemia with t(6;11)(q27;q23): results from Cancer and Leukemia Group B Study 8461 and review of the literature. Cancer 2004; 101:1420.
  98. Jourdan E, Dastugue N. Adult de novo acute myeloid leukemia with t(6;11)(q27;q23): results from Cancer and Leukemia Group B Study 8461 and review of the literature. Cancer 2005; 103:1316; author reply 1316.
  99. Schichman SA, Canaani E, Croce CM. Self-fusion of the ALL1 gene. A new genetic mechanism for acute leukemia. JAMA 1995; 273:571.
  100. Caligiuri MA, Strout MP, Schichman SA, et al. Partial tandem duplication of ALL1 as a recurrent molecular defect in acute myeloid leukemia with trisomy 11. Cancer Res 1996; 56:1418.
  101. Ziemin-van der Poel S, McCabe NR, Gill HJ, et al. Identification of a gene, MLL, that spans the breakpoint in 11q23 translocations associated with human leukemias. Proc Natl Acad Sci U S A 1991; 88:10735.
  102. Gu Y, Nakamura T, Alder H, et al. The t(4;11) chromosome translocation of human acute leukemias fuses the ALL-1 gene, related to Drosophila trithorax, to the AF-4 gene. Cell 1992; 71:701.
  103. Tkachuk DC, Kohler S, Cleary ML. Involvement of a homolog of Drosophila trithorax by 11q23 chromosomal translocations in acute leukemias. Cell 1992; 71:691.
  104. Djabali M, Selleri L, Parry P, et al. A trithorax-like gene is interrupted by chromosome 11q23 translocations in acute leukaemias. Nat Genet 1993; 4:431.
  105. Mazo AM, Huang DH, Mozer BA, Dawid IB. The trithorax gene, a trans-acting regulator of the bithorax complex in Drosophila, encodes a protein with zinc-binding domains. Proc Natl Acad Sci U S A 1990; 87:2112.
  106. Liedtke M, Cleary ML. Therapeutic targeting of MLL. Blood 2009; 113:6061.
  107. Milne TA, Kim J, Wang GG, et al. Multiple interactions recruit MLL1 and MLL1 fusion proteins to the HOXA9 locus in leukemogenesis. Mol Cell 2010; 38:853.
  108. Wang Z, Song J, Milne TA, et al. Pro isomerization in MLL1 PHD3-bromo cassette connects H3K4me readout to CyP33 and HDAC-mediated repression. Cell 2010; 141:1183.
  109. Muntean AG, Tan J, Sitwala K, et al. The PAF complex synergizes with MLL fusion proteins at HOX loci to promote leukemogenesis. Cancer Cell 2010; 17:609.
  110. Wang Z, Iwasaki M, Ficara F, et al. GSK-3 promotes conditional association of CREB and its coactivators with MEIS1 to facilitate HOX-mediated transcription and oncogenesis. Cancer Cell 2010; 17:597.
  111. Mohan M, Lin C, Guest E, Shilatifard A. Licensed to elongate: a molecular mechanism for MLL-based leukaemogenesis. Nat Rev Cancer 2010; 10:721.
  112. Lin C, Smith ER, Takahashi H, et al. AFF4, a component of the ELL/P-TEFb elongation complex and a shared subunit of MLL chimeras, can link transcription elongation to leukemia. Mol Cell 2010; 37:429.
  113. Jin S, Zhao H, Yi Y, et al. c-Myb binds MLL through menin in human leukemia cells and is an important driver of MLL-associated leukemogenesis. J Clin Invest 2010; 120:593.
  114. Krivtsov AV, Evans K, Gadrey JY, et al. A Menin-MLL Inhibitor Induces Specific Chromatin Changes and Eradicates Disease in Models of MLL-Rearranged Leukemia. Cancer Cell 2019; 36:660.
  115. Corral J, Lavenir I, Impey H, et al. An Mll-AF9 fusion gene made by homologous recombination causes acute leukemia in chimeric mice: a method to create fusion oncogenes. Cell 1996; 85:853.
  116. Lavau C, Szilvassy SJ, Slany R, Cleary ML. Immortalization and leukemic transformation of a myelomonocytic precursor by retrovirally transduced HRX-ENL. EMBO J 1997; 16:4226.
  117. Lavau C, Luo RT, Du C, Thirman MJ. Retrovirus-mediated gene transfer of MLL-ELL transforms primary myeloid progenitors and causes acute myeloid leukemias in mice. Proc Natl Acad Sci U S A 2000; 97:10984.
  118. Luo RT, Lavau C, Du C, et al. The elongation domain of ELL is dispensable but its ELL-associated factor 1 interaction domain is essential for MLL-ELL-induced leukemogenesis. Mol Cell Biol 2001; 21:5678.
  119. DiMartino JF, Miller T, Ayton PM, et al. A carboxy-terminal domain of ELL is required and sufficient for immortalization of myeloid progenitors by MLL-ELL. Blood 2000; 96:3887.
  120. de Thé H, Chomienne C, Lanotte M, et al. The t(15;17) translocation of acute promyelocytic leukaemia fuses the retinoic acid receptor alpha gene to a novel transcribed locus. Nature 1990; 347:558.
  121. von Lindern M, Breems D, van Baal S, et al. Characterization of the translocation breakpoint sequences of two DEK-CAN fusion genes present in t(6;9) acute myeloid leukemia and a SET-CAN fusion gene found in a case of acute undifferentiated leukemia. Genes Chromosomes Cancer 1992; 5:227.
  122. Ichikawa H, Shimizu K, Hayashi Y, Ohki M. An RNA-binding protein gene, TLS/FUS, is fused to ERG in human myeloid leukemia with t(16;21) chromosomal translocation. Cancer Res 1994; 54:2865.
  123. Panagopoulos I, Aman P, Fioretos T, et al. Fusion of the FUS gene with ERG in acute myeloid leukemia with t(16;21)(p11;q22). Genes Chromosomes Cancer 1994; 11:256.
  124. Rabbitts TH, Forster A, Larson R, Nathan P. Fusion of the dominant negative transcription regulator CHOP with a novel gene FUS by translocation t(12;16) in malignant liposarcoma. Nat Genet 1993; 4:175.
  125. Sorensen PH, Lessnick SL, Lopez-Terrada D, et al. A second Ewing's sarcoma translocation, t(21;22), fuses the EWS gene to another ETS-family transcription factor, ERG. Nat Genet 1994; 6:146.
  126. Panagopoulos I, Höglund M, Mertens F, et al. Fusion of the EWS and CHOP genes in myxoid liposarcoma. Oncogene 1996; 12:489.
  127. Yoneda-Kato N, Look AT, Kirstein MN, et al. The t(3;5)(q25.1;q34) of myelodysplastic syndrome and acute myeloid leukemia produces a novel fusion gene, NPM-MLF1. Oncogene 1996; 12:265.
  128. Morris SW, Kirstein MN, Valentine MB, et al. Fusion of a kinase gene, ALK, to a nucleolar protein gene, NPM, in non-Hodgkin's lymphoma. Science 1994; 263:1281.
  129. Lugthart S, Gröschel S, Beverloo HB, et al. Clinical, molecular, and prognostic significance of WHO type inv(3)(q21q26.2)/t(3;3)(q21;q26.2) and various other 3q abnormalities in acute myeloid leukemia. J Clin Oncol 2010; 28:3890.
  130. Hui CH, Suttle J. Erythrophagocytosis associated with AML-M5 and t(8;16). Br J Haematol 2001; 113:845.
  131. Sobulo OM, Borrow J, Tomek R, et al. MLL is fused to CBP, a histone acetyltransferase, in therapy-related acute myeloid leukemia with a t(11;16)(q23;p13.3). Proc Natl Acad Sci U S A 1997; 94:8732.
  132. Caudell D, Aplan PD. The role of CALM-AF10 gene fusion in acute leukemia. Leukemia 2008; 22:678.
  133. Borrow J, Shearman AM, Stanton VP Jr, et al. The t(7;11)(p15;p15) translocation in acute myeloid leukaemia fuses the genes for nucleoporin NUP98 and class I homeoprotein HOXA9. Nat Genet 1996; 12:159.
  134. Jankovic D, Gorello P, Liu T, et al. Leukemogenic mechanisms and targets of a NUP98/HHEX fusion in acute myeloid leukemia. Blood 2008; 111:5672.
  135. Metz T, Graf T. Fusion of the nuclear oncoproteins v-Myb and v-Ets is required for the leukemogenicity of E26 virus. Cell 1991; 66:95.
  136. Graf T, McNagny K, Brady G, Frampton J. Chicken "erythroid" cells transformed by the Gag-Myb-Ets-encoding E26 leukemia virus are multipotent. Cell 1992; 70:201.
  137. Mucenski ML, McLain K, Kier AB, et al. A functional c-myb gene is required for normal murine fetal hepatic hematopoiesis. Cell 1991; 65:677.
  138. Nazarov V, Hilbert D, Wolff L. Susceptibility and resistance to Moloney murine leukemia virus-induced promonocytic leukemia. Virology 1994; 205:479.
  139. Nazarov V, Wolff L. Novel integration sites at the distal 3' end of the c-myb locus in retrovirus-induced promonocytic leukemias. J Virol 1995; 69:3885.
  140. Ben-David Y, Bernstein A. Friend virus-induced erythroleukemia and the multistage nature of cancer. Cell 1991; 66:831.
  141. Munroe DG, Peacock JW, Benchimol S. Inactivation of the cellular p53 gene is a common feature of Friend virus-induced erythroleukemia: relationship of inactivation to dominant transforming alleles. Mol Cell Biol 1990; 10:3307.
  142. Lavigueur A, Bernstein A. p53 transgenic mice: accelerated erythroleukemia induction by Friend virus. Oncogene 1991; 6:2197.
  143. Moreau-Gachelin F, Ray D, Mattei MG, et al. The putative oncogene Spi-1: murine chromosomal localization and transcriptional activation in murine acute erythroleukemias. Oncogene 1989; 4:1449.
  144. Paul R, Schuetze S, Kozak SL, et al. The Sfpi-1 proviral integration site of Friend erythroleukemia encodes the ets-related transcription factor Pu.1. J Virol 1991; 65:464.
  145. Scott EW, Simon MC, Anastasi J, Singh H. Requirement of transcription factor PU.1 in the development of multiple hematopoietic lineages. Science 1994; 265:1573.
  146. Buchberg AM, Bedigian HG, Jenkins NA, Copeland NG. Evi-2, a common integration site involved in murine myeloid leukemogenesis. Mol Cell Biol 1990; 10:4658.
  147. Largaespada DA, Shaughnessy JD Jr, Jenkins NA, Copeland NG. Retroviral integration at the Evi-2 locus in BXH-2 myeloid leukemia cell lines disrupts Nf1 expression without changes in steady-state Ras-GTP levels. J Virol 1995; 69:5095.
  148. Cho BC, Shaughnessy JD Jr, Largaespada DA, et al. Frequent disruption of the Nf1 gene by a novel murine AIDS virus-related provirus in BXH-2 murine myeloid lymphomas. J Virol 1995; 69:7138.
  149. Shannon KM, Watterson J, Johnson P, et al. Monosomy 7 myeloproliferative disease in children with neurofibromatosis, type 1: epidemiology and molecular analysis. Blood 1992; 79:1311.
  150. Jacks T, Shih TS, Schmitt EM, et al. Tumour predisposition in mice heterozygous for a targeted mutation in Nf1. Nat Genet 1994; 7:353.
  151. Bollag G, Clapp DW, Shih S, et al. Loss of NF1 results in activation of the Ras signaling pathway and leads to aberrant growth in haematopoietic cells. Nat Genet 1996; 12:144.
  152. Largaespada DA, Brannan CI, Jenkins NA, Copeland NG. Nf1 deficiency causes Ras-mediated granulocyte/macrophage colony stimulating factor hypersensitivity and chronic myeloid leukaemia. Nat Genet 1996; 12:137.
  153. Nakamura T, Largaespada DA, Shaughnessy JD Jr, et al. Cooperative activation of Hoxa and Pbx1-related genes in murine myeloid leukaemias. Nat Genet 1996; 12:149.
  154. Nakamura T, Jenkins NA, Copeland NG. Identification of a new family of Pbx-related homeobox genes. Oncogene 1996; 13:2235.
  155. Vassiliou GS, Cooper JL, Rad R, et al. Mutant nucleophosmin and cooperating pathways drive leukemia initiation and progression in mice. Nat Genet 2011; 43:470.
Topic 4502 Version 22.0

References

آیا می خواهید مدیلیب را به صفحه اصلی خود اضافه کنید؟