Acute megakaryoblastic leukemia (AMKL) comprises between 4% and 15% of newly diagnosed pediatric acute myeloid leukemia patients. AMKL in children with Down syndrome (DS) is characterized by a founding GATA1 mutation that cooperates with trisomy 21, followed by the acquisition of additional somatic mutations. In contrast, non–DS-AMKL is characterized by chimeric oncogenes consisting of genes known to play a role in normal hematopoiesis. CBFA2T3-GLIS2 is the most frequent chimeric oncogene identified to date in this subset of patients and confers a poor prognosis.

Acute megakaryoblastic leukemia (AMKL) is a subtype of acute myeloid leukemia (AML) characterized by abnormal megakaryoblasts that express platelet-specific surface glycoprotein. Bone marrow biopsy frequently demonstrates extensive myelofibrosis, often making aspiration in these patients difficult. AMKL is extremely rare in adults, occurring in only 1% of AML patients.1  This is in contrast to children, where it comprises between 4% and 15% of AML patients.2,3  In pediatrics, the disease is divided into 2 major subgroups: AMKL in patients with Down syndrome (DS-AMKL) and AMKL in patients without DS (non–DS-AMKL). AMKL is the most frequent type of AML in children with DS, and the incidence in these patients is 500-fold higher than in the general population.4  In contrast to non–DS-AMKL, leukemic cells carry not only megakaryocytic cell-surface markers but also erythroid markers, resulting in the distinct World Health Organization classification “myeloid leukemia in Down syndrome”5 . Somatic mutations in GATA1 are found in almost all cases of DS-AMKL and precede the development of leukemia, as indicated by their presence in patients with transient myeloproliferative disease (TMD) in the neonatal period.6-11  DS-AMKL is both biologically and clinically distinct, with superior outcomes compared with non–DS-AMKL.12-15  Pediatric non–DS-AMKL is a heterogenous group of patients, a significant proportion of whom carry chimeric oncogenes including RBM15-MKL1, CBFA2T3-GLIS2, NUP98-KDM5A, and MLL gene rearrangements.16,17  Unfortunately, the outcome of non–DS-AMKL is generally poor, with lower event-free survival than DS-AMKL and pediatric AML, even in the face of intensified treatment.2,18 

TMD

DS-AMKL is associated with TMD, a hematologic disorder in infancy. In this disorder, a clonal population of megakaryoblasts accumulates in the peripheral blood. These blasts are phenotypically indistinguishable from AMKL leukemic blasts, and in the majority of cases, remission is spontaneous within 3 months in the absence of treatment. In ∼20% of TMD cases, patients will go on to develop myelodysplastic syndrome and/or AMKL.19  TMD is thought to originate in utero, as an identical mutation in GATA1, the genetic lesion associated with TMD, was found to be present at birth in twins with TMD.20  Further evidence came with the analysis of archived autopsy specimens from DS patients that identified GATA1 mutations in 2 fetal liver specimens.21  A subsequent study screening Guthrie cards from 585 DS infants identified GATA1 mutations in 3.8% of their cohort, confirming the presence of this lesion in a subset of patients at birth.22  The frequency of this lesion in newborn DS patients was significantly higher in a study that used next-generation sequencing, which has a greater sensitivity, to screen 200 neonates with DS.23  In this analysis, GATA1 mutations were detected in 29% of patients. The spontaneous resolution of TMD suggests that despite the presence of blasts in the peripheral blood that appear phenotypically indistinguishable from full-blown leukemia, they are in fact functionally different as they fail to persist. When TMD and AMKL blasts from patients with DS are injected into immunodeficient mice, this difference becomes apparent. Approximately 50% of DS-AMKL engraft into NOD/SCID mice, leading to widespread dissemination and the ability to propagate in secondary and tertiary recipients.24  In contrast, blasts from TMD patients very rarely engraft, fail to disseminate outside the bone marrow, and are unable to propagate disease in secondary and tertiary recipients.24  Exome sequencing of TMD has revealed that non–silent mutations in these blasts are primarily limited to the GATA1 gene.25  In contrast, AMKL blasts carry a higher burden of mutations, with additional lesions in epigenetic and kinase-signaling genes leading to progression of the disease. Collectively, these findings support a model whereby TMD blasts arise secondary to GATA1 mutations in the setting of trisomy 21, acquiring this so-called first hit, and persist in the bone marrow. Additional lesions can then occur providing the cooperating events that are necessary for full-blown leukemia to develop (Figure 1). Although sequencing studies have demonstrated the genetic lesions that are required for progression of TMD to AMKL, they do not provide any information on how to predict the 20% of patients that will go on to develop AMKL. An extensive analysis of germline DNA, including pathologic mutations in cancer-predisposition genes as well as genome-wide association studies to identify polymorphisms that may predispose an individual to developing AMKL, may provide clues. If predisposing factor(s) are identified, they have the potential to significantly impact clinical care, as the identification of those patients at high risk of developing AMKL would allow for early treatment of the premalignant cells with decreased intensity chemotherapy while maintaining the high cure rates.

Figure 1

DS-AMKL pathogenesis. In utero truncating mutations in GATA1 lead to a TMD in the neonatal period that resolves in the absence of treatment. Residual cells either undergo apoptosis or acquire additional cooperating mutations leading to overt AMKL with an average latency of 3 years. Recurrently targeted genes include but are not limited to cohesin complex components, CTCF, the PRC2 complex, and kinase-signaling genes. Of the 26 sequenced DS-AMKL cases that carry mutations in cohesin, 6 contained mutations in a PRC2 complex gene as well as a kinase as shown in this example.25  Cohesin mutation, ●; GATA1 mutation, ★; kinase mutation, ▲; PRC2 mutation, ▪; Trisomy 21, ×××.

Figure 1

DS-AMKL pathogenesis. In utero truncating mutations in GATA1 lead to a TMD in the neonatal period that resolves in the absence of treatment. Residual cells either undergo apoptosis or acquire additional cooperating mutations leading to overt AMKL with an average latency of 3 years. Recurrently targeted genes include but are not limited to cohesin complex components, CTCF, the PRC2 complex, and kinase-signaling genes. Of the 26 sequenced DS-AMKL cases that carry mutations in cohesin, 6 contained mutations in a PRC2 complex gene as well as a kinase as shown in this example.25  Cohesin mutation, ●; GATA1 mutation, ★; kinase mutation, ▲; PRC2 mutation, ▪; Trisomy 21, ×××.

Close modal

GATA1

The GATA family of proteins consists of transcription factors, 3 of which are expressed principally in hematopoietic cells (GATA1, GATA2, and GATA3). The GATA1 protein is typically present in cells of erythroid, megakaryocytic, mast, and eosinophilic lineages, whereas GATA2 is expressed in early hematopoietic progenitors.26  GATA1 is required for the development of erythrocytes, megakaryocytes, eosinophils, and mast cells. Mutations in GATA1 have been associated with thrombocytopenia, familial dyserythropoietic anemia, thalassemia, porphyria, Diamond-Blackfan anemia, TMD, and DS-AMKL.26-31  The mutations found in nonmalignant diseases either weaken or eliminate the interaction of GATA1 with its cofactor FOG1 or interfere with DNA binding.28-32  In contrast, the mutations detected in DS patients consist of short deletions, insertions, and point mutations within exon 2 that introduce a premature stop codon.7  This shorter mutant protein retains the ability to bind DNA and interact with its cofactor, but it lacks the transcriptional activation domain and hence has reduced transactivation potential.7  To model TMD, a knockin line of mice expressing a truncated form of GATA1 was generated and found to result in hyperproliferative megakaryocytic progenitors in the yolk sac and fetal liver that disappeared by the end of gestation.33  A separate group crossed mice transgenic for a truncated form of GATA1 to the GATA1 knockout strain.34  During the neonatal period, mice accumulate immature megakaryocytic progenitors in the liver that disappear during weaning of the pups. Regardless of the difference in timing, these models serve to validate that a truncated GATA1 protein is able to confer a proliferative advantage, generating a pool of precursors that have the potential to develop into a leukemic population. The mechanism whereby truncated GATA1 is able to induce a preleukemic state is not fully elucidated, although genome-wide chromatin immunoprecipitation sequencing of genes bound by GATA1 merged with expression profiling revealed a large number of activated and repressed genes, respectively, that were occupied by the GATA1 protein.35  Further studies have shown that GATA1 is able to activate lineage specific genes and repress progenitor maintenance genes depending on the cofactors present.36  It is therefore plausible that deregulation of these targets contributes to the differentiation arrest seen with the truncated GATA1 that is no longer able to transactivate transcription of lineage specific genes. A second mechanism proposed is the upregulation of genes by mutant GATA1 that promote self-renewal, as has been demonstrated for the microRNA miR-486-5p.37  Additionally, it is possible that the extra gene dosage of chromosome 21 contributes to this process; in fact, trisomy 21 has an impact on fetal hematopoiesis in and of itself.38-40  Fetal livers from DS patients have a two- to threefold increase in megakaryocyte erythroid progenitors, and trisomic stem cells exhibit alterations of hematopoiesis in vitro with an increase in multilineage colony-forming potential, an indicator of increased self-renewal.39-41  Supporting this cooperativity between GATA1 mutations and trisomy 21 is the specificity of GATA1 mutations: almost without exception, GATA1 mutations are not found outside the context of trisomy 21.26  Even in rare cases of non-DS-AMKL that carry GATA1 mutations, somatic copy number amplifications in the DS critical region of chromosome 21 are found to be present.16 

Patients with trisomy 21 have, in essence, an extra copy of many genes on chromosome 21 (chr21), and overexpression of one or more has been hypothesized to provide the cellular setting that is permissible for persistence and eventual transformation of GATA1 mutant cells. Candidate genes on chr21 that contribute to a preleukemic phenotype include but are not limited to ERG, RUNX1, DYRK1A, and MIR125B2.42-45 ERG is a member of the ETS transcription gene family. Increased expression of ERG is seen in some cases of AML and it is also a translocation partner in t(16;21) myeloid leukemia.46,47  ERG has been recently shown to play a role in hematopoietic stem cells as well as the development of the megakaryocytic lineage, and furthermore, transgenic expression of ERG and a mutant GATA1 protein in murine fetal liver cells results in a TMD like disease.48-50  Additionally, overexpression of ERG in hematopoietic progenitor cells by retroviral transduction and subsequent transplantation into mice results in megakaryoblastic leukemia.44  Another candidate is the RUNX1 gene, also found on chr21. Perhaps counterintuitively, RUNX1 expression was found to be lower in DS-AMKL cases in comparison with non–DS-AMKL in 2 separate cohorts despite the increase number of genomic copies.51,52  Although the mechanism of this downregulation is not clear, in core binding factor leukemias, a decrease in RUNX1 activity either by mutation or the transdominant effect of a translocation involving RUNX1 is associated with increased leukemic potential. Thus, a downregulation of RUNX1 in DS-AMKL would be consistent with previous data that a loss of RUNX1 wild-type function enhances self-renewal and blocks differentiation. In line with this hypothesis, RUNX1 upregulation was found to precede megakaryocyte differentiation in human hematopoietic cells and downregulation was seen when cells underwent erythroid differentiation, suggesting that it functions in megakaryocytic lineage commitment.45  A decrease in RUNX1 could therefore impair differentiation allowing persistence of GATA1 mutant cells in a more immature state.

Cooperating mutations

Given that only 20% of TMD progresses to leukemia, what then are the subsequent events or alterations that promote the preleukemic state to that of a fully transformed malignancy? Exome and targeted sequencing of 46 genes has provided insight to this question, identifying recurrently mutated genes in three major categories: cohesin, epigenetic regulators, and signaling molecules.25  Core cohesin complex components including STAG2, RAD21, SMC3, SMC1A, and the cohesin complex loading protein NIPBL were mutated in 53% of the 49 DS-AMKL cases and none of the 41 TMD cases interrogated. This is significantly higher than the reported frequency of 6% to 12% in AML, suggesting these mutations may play a specific role in promoting megakaryocytic disease.53-55  Additionally, 6 cases carried mutations in CTCF, a transcriptional repressor and insulator protein. Cohesin maintains sister chromatid cohesion, allowing for faithful chromosome segregation and DNA repair.56  In addition, the complex also functions in transcriptional regulation through DNA looping. CTCF and cohesin have been found to co-localize extensively throughout mammalian genomes.57  It has been suggested that together, they play a role in the establishment and maintenance of topological domains.58  Their disruption thus has the potential to significantly disrupt chromatin architecture and, in doing so, gene expression. Interestingly, GATA1 has been found to co-occupy genes with the RAD21 cohesin component as well as CTCF in adult proerythrocytes (796 and 656 target genes, respectively), providing direct evidence for cooperative effects between these genes.59 

EZH2, the catalytic subunit of the Polycomb repressive complex 2 (PRC2) was the most frequently targeted epigenetic regulator in DS-AMKL. Combined with SUZ12, PRC2 mutations were mutually exclusive and collectively occurred in 17 of 49 cases (35%), the majority of which also contained alterations in CTCF or cohesin. In erythroid cells, PRC2 is involved in epigenetic silencing of a subset of GATA1-repressed genes, some of which are associated with progenitor cells such as KIT and GATA2.60  Disruption of the repression may therefore enhance the self-renewal of cells, contributing to the differentiation block provided by the truncated GATA1 protein.

Close to 50% of DS-AMKL cases carry activating kinase mutations in JAK1, JAK2, JAK3, MPL, KRAS, or NRAS or loss-of-function mutations in SH2B3. These kinase genes fall broadly into 2 categories: JAK/signal transducer and activator of transcription (STAT) and RAS signaling, both of which play a role in megakaryopoiesis (Figure 2).61,62  Mutations between these 2 signaling cascades are, for the most part, mutually exclusive, although occasional cases carry a lesion in both. They result in constituitively activated signaling, leading to a gain of function as demonstrated by cytokine-independent growth in laboratory assays.63-65  Overexpression of one of the DS-AMKL–associated JAK3-activating mutations has been shown to result in a lethal megakaryocyte progenitor expansion in a subset of mice, further supporting this signaling pathway in AMKL.64 

Figure 2

JAK signaling in megakaryopoiesis. Cytokine binding to its cellular receptor leads to dimerization and phosphorylation that in turn binds and activates JAK, leading to downstream activation of RAS signaling and phosphorylation of STAT transcription factors. Receptors and kinases with activating mutations identified in AMKL include MPL, PDGFRB, JAK1, JAK2, JAK3, NRAS, and KRAS. Mutations in SH2B3 have been identified in DS-AMKL. SHC1, adapter molecule; SH2B3, inhibitor of JAK2.

Figure 2

JAK signaling in megakaryopoiesis. Cytokine binding to its cellular receptor leads to dimerization and phosphorylation that in turn binds and activates JAK, leading to downstream activation of RAS signaling and phosphorylation of STAT transcription factors. Receptors and kinases with activating mutations identified in AMKL include MPL, PDGFRB, JAK1, JAK2, JAK3, NRAS, and KRAS. Mutations in SH2B3 have been identified in DS-AMKL. SHC1, adapter molecule; SH2B3, inhibitor of JAK2.

Close modal

RBM15-MKL1

The t(1;22) translocation and its association with AMKL in infants was initially identified in a cohort of 252 children with AML accrued over a 24-month period.66  In this report, no cases of t(1;22) were identified in a concurrent pediatric ALL cohort of 2382 cases, and the translocation was exclusively found in patients with AMKL, all of whom were <1 year of age. This fusion was very specific for infant AMKL, as the 22 other infants with AML who lacked the translocation had a different phenotypic subtype. Further, the remaining 12 non–DS-AMKL cases carried no recurring chromosomal abnormalities and were all older. Others have since confirmed this association, but it was not until 10 years after the initial report that the genes involved in the translocation were characterized.67-70  Two groups simultaneously identified the genes on chromosomes 1 and 22 involved in the translocation: RBM15 (also known as OTT) and MKL1 (also known as MAL), respectively.67,70  Since their initial cloning, much has been learned about the function of the genes, and a role of the translocation in inducing leukemia has been demonstrated in a knockin mouse model.71 

MKL1 is a transcriptional coactivator for serum response factor (SRF), a transcription factor that regulates the expression of genes involved in cell growth, proliferation, and differentiation, as well as genes that control the actin cytoskeleton.72  In serum-starved cells, MKL1 associates with G actin monomers and is retained in the cytoplasm. Following serum stimulation and Rho-mediated actin polymerization, G actin pools are depleted and MKL1 translocates to the nucleus, associating with SRF to activate gene transcription.73,74  During murine megakaryocyte differentiation, Mkl1 is upregulated. Consistent with this, Mkl1-knockout mice have an increased percentage of megakaryocytic progenitors and a decrease in mature megakaryocytes as well as dysplastic megakaryocytes.75,76  RBM15 belongs to the Spen family of proteins and encodes a protein containing 3 amino-terminal RNA recognition motifs that bind to nucleic acids and a C-terminal SPOC domain that is thought to interact with the SMRT and NCoR corepressor complexes, as well as RBPJ, a transcription factor downstream of Notch signaling.77,78 Rbm15-knockout mice are embryonic lethal; thus, to evaluate the effect of this protein on hematopoiesis, conditional-knockout mice have been generated.79,80  These mice have a block in B lymphopoiesis and expansion of the myeloid, megakaryocytic, and progenitor compartments.75,79  The fusion of MKL1 to RBM15 deregulates the normal intracellular localization of MKL1 such that it is becomes constitutively localized to the nucleus, resulting in SRF activation even in the absence of stimuli.81  In addition to the SRF transcriptional program, the fusion also aberrantly activates RBPJ transcriptional targets. Although both transcription programs have been shown to be deregulated by the fusion gene, the degree to which they contribute to transformation is still unclear.

In studies done to address the role of the RBM15-MKL1 chimeric gene in AMKL, knockin mice were engineered to express the chimeric oncogene under control of the endogenous Rbm15 promoter.71  These mice display abnormal fetal and adult hematopoiesis, with a small fraction developing AMKL between 18 and 24 months of age.71  Using this mouse model, the authors present data to support RBM15-MKL1–activated RBPJ mediated transcriptional activity that leads to upregulation of the Notch pathway.71  Consistent with this, Rbm15 has been shown to modulate Notch-induced transcription in a cell-type–specific manner.82  Given that only a fraction of mice developed overt AMKL at a late age, the authors reasoned that cooperating oncogenic events were required to induce AMKL. The identification of such cooperating mutations has proved elusive due to a paucity of clinical samples with high tumor content for next-generation sequencing analysis. Nonetheless, careful analysis of one patient specimen along with a matched germline specimen revealed 12 high confidence mutations, one of which occurred in MMP8, a matrix metalloproteinase gene that is expressed in megakaryocyte-erythroid progenitors.83  Further studies are necessary to determine if this mutation is able to cooperate with the RBM15-MKL1 oncogene.

CBFA2T3-GLIS2

Until recently, with the exception of the RBM15-MKL1 fusion, the genetic etiology of non–DS-AMKL had remained elusive. A high-resolution study of DNA copy-number abnormalities and loss of heterozygosity on pediatric de novo AML samples demonstrated a very low burden of genomic alterations in all pediatric AML subtypes with the exception of AMKL.84  AMKL cases were characterized by complex chromosomal rearrangements and a high number of copy-number abnormalities. We predicted that these lesions would have functional consequences and therefore performed transcriptome and exome sequencing on diagnostic leukemia samples from 14 pediatric non–DS-AMKL cases as part of the St. Jude Children’s Research Hospital–Washington University Pediatric Cancer Genome Project.16  Indeed, we detected structural variations that resulted in the expression of chimeric transcripts in 12 of 14 samples. Remarkably, in 7 of 14 cases, a cryptic inversion on chromosome 16 [inv(16)(p13.3q24.3)] was detected that resulted in the joining of CBFA2T3, a member of the ETO family of nuclear corepressors, to GLIS2, a member of the GLI family of transcription factors.16  The gene expression profile of CBFA2T3-GLIS2 AMKL was distinct from that of AMKL cells lacking this chimeric transcript and from other genetic subtypes of pediatric AML.16  Furthermore, the CBFA2T3-GLIS2 fusion gene conferred a poor prognosis, a finding that has since been confirmed.16,17,85  This fusion was subsequently reported to also occur at a low frequency in pediatric cytogenetically normal AML.85  Expression of CBFA2T3-GLIS2 in Drosophila and murine hematopoietic cells induced bone morphogenic protein (BMP) signaling, a pathway not previously implicated in AML, and resulted in a marked increase in the self-renewal capacity of hematopoietic progenitors.16  The contribution of BMP signaling to self-renewal in CBFA2T3-GLIS2 modified murine hematopoietic cells has since been confirmed in colony-formation assays utilizing Bmp2 and Bmp4 conditional-knockout marrow (unpublished data).

CBFA2T3-GLIS2–expressing cells remained growth factor dependent in vitro, suggesting that cooperating mutations in growth factor signaling pathways are likely required for full leukemic transformation. Moreover, transplantation of CBFA2T3-GLIS2–transduced bone marrow cells into syngeneic recipients failed to induce overt leukemia, consistent with a requirement for cooperative mutations. Failure to induce leukemia in mice as a single lesion has been previously reported for other chimeric genes that confer the ability to serially replate in colony-forming assays, including AML1-ETO.86  Overall, the total burden of somatic mutations in our cohort was significantly lower in the CBFA2T3-GLIS2–expressing cases for which germline DNA was available than in non–DS-AMKL that lacked this fusion gene (7.2 ± 3.6 vs 16.6 ± 5.1, P = .009).16  Of the 15 CBFA2T3-GLIS2–positive cases analyzed to date, 5 carried lesions in either a Janus kinase (JAK) gene and/or a somatic amplification of the DS critical region on chromosome 21. However, the majority of cases do not contain an identifiable cooperating lesion (unpublished data).16  As these cases have been interrogated by single-nucleotide polymorphism arrays, exome, and/or transcriptome sequencing, a more thorough whole-genome approach may help to further delineate the additional events required by this fusion oncogene. Whole-genome sequencing would allow the identification of somatic mutations in noncoding intergenic regions that are oncogenic. Examples of these types of lesions include TERT promoter mutations and superenhancer formation upstream of the TAL1 oncogene, as identified in melanoma and T-cell acute lymphoblastic leukemia, respectively.87,88 

Lower-frequency fusion events

In addition to CBFA2T3-GLIS2, ∼8% of our pediatric cohort carried the previously described NUP98-KDM5A fusion gene (Figure 3).16  In parallel with our efforts, de Rooij and colleagues evaluated a separate non–DS-AMKL cohort for NUP98 fusion events by split-signal fluorescence in situ hybridization and found a similar frequency of 11%.17 NUP98, a nucleoporin family member with transactivation activity, fused to KDM5A, an H3K4me3-binding PHD finger, was initially described in adult AML.89,90  When introduced into murine bone marrow, this fusion oncogene induces a myeloid differentiation arrest and mice develop AML with an average latency of 69 days.91  Wang and colleagues demonstrated this fusion to be bound to H3K4me3 mononucleosomes, showing the PHD finger plays a role in targeting the fusion to the genome.91  Interestingly, microarray analysis identified several polycomb proteins carrying H3K4me3 marks to be transcriptionally upregulated in response to the fusion, whereas housekeeping genes with constitutive H3K4me3 marks remained unchanged. Affected polycomb targets confirmed by chromatin immunoprecipitation include genes upregulated in MLL rearranged leukemia such as HOXA5, HOXA7, HOXA9, HOXA10, MEIS1, and PBX1.91  Furthermore, the authors demonstrate a block in PRC2 binding, the complex that antagonizes polycomb proteins through transcriptional repression of target genes. Therefore, the NUP98-KDM5A fusion is able to prevent silencing of critical transcription factors that play a role in maintaining hematopoietic progenitor status, similar to MLL gene rearrangements. It is perhaps not surprising then that MLL-AF9 and MLL-AF10 fusion events have also been detected in non–DS-AMKL.17  As these lesions are also found in other subtypes of AML, there are likely additional factors contributing to the development of megakaryoblastic disease. Cooperating mutations, the target cell, and the microenvironment all have the potential to direct lineage during the process of transformation.

Figure 3

Key genomic events in non–DS-AMKL. A total of 142 pediatric non–DS-AMKL cases were analyzed for the presence of fusion gene events by transcriptome sequencing, reverse-transcription polymerase chain reaction (RT-PCR), or split-signal fluorescence in situ hybridization. A total of 96 samples were evaluated for the presence of the MLL-PTD by RT-PCR and 46 samples were evaluated for the presence of somatic GATA1 single-nucleotide variations and insertion/deletion by exome and/or Sanger sequencing. The proportion of MLL-PTD and GATA1 mutant cases was calculated based on the total number of samples evaluated for each of the lesions. Patients carrying GATA1 mutations did not have stigmata of DS or evidence of mutant reads in germline DNA, suggesting they are not mosaics. Cases that did not undergo transcriptome sequencing and were negative by RT-PCR for CBFA2T3-GLIS2, NUP98-KDM5A, RBM15-MKL1, and MLL rearrangements (MLLr) are designated as unknown. “Other fusion” includes single cases of each of the following: GATA2-HOXA9, NIPBL-HOXB9, MN1-FLI1, HLXB9-ETV6, FUS-ERG, and RUNX1-CBFA2T3. Data compiled from Gruber et al16  and de Rooij et al.17 

Figure 3

Key genomic events in non–DS-AMKL. A total of 142 pediatric non–DS-AMKL cases were analyzed for the presence of fusion gene events by transcriptome sequencing, reverse-transcription polymerase chain reaction (RT-PCR), or split-signal fluorescence in situ hybridization. A total of 96 samples were evaluated for the presence of the MLL-PTD by RT-PCR and 46 samples were evaluated for the presence of somatic GATA1 single-nucleotide variations and insertion/deletion by exome and/or Sanger sequencing. The proportion of MLL-PTD and GATA1 mutant cases was calculated based on the total number of samples evaluated for each of the lesions. Patients carrying GATA1 mutations did not have stigmata of DS or evidence of mutant reads in germline DNA, suggesting they are not mosaics. Cases that did not undergo transcriptome sequencing and were negative by RT-PCR for CBFA2T3-GLIS2, NUP98-KDM5A, RBM15-MKL1, and MLL rearrangements (MLLr) are designated as unknown. “Other fusion” includes single cases of each of the following: GATA2-HOXA9, NIPBL-HOXB9, MN1-FLI1, HLXB9-ETV6, FUS-ERG, and RUNX1-CBFA2T3. Data compiled from Gruber et al16  and de Rooij et al.17 

Close modal

In addition to the previously described NUP98-KDM5A fusion, we identified 3 novel fusion genes expressed in a single case each: GATA2-HOXA9, MN1-FLI1, and NIPBL-HOXB9 (Figure 3). Each of these chimeric transcripts are predicted to encode a fusion protein that would alter signaling pathways known to play a role in normal hematopoiesis, suggesting that these lesions are “driver” mutations that directly contribute to the development of leukemia. Several of the genes involved in these translocations play a direct role in normal megakaryocytic differentiation (GATA2 and FLI1), have been previously shown to be involved in leukemogenesis (HOXA9, MN1, and HOXB9), or are highly expressed in hematopoietic stem cells or myeloid/megakaryocytic progenitors.91-96  Genome-wide approaches in a larger AMKL cohort are necessary to determine if these fusion genes are recurrent. Current efforts in our laboratories include experiments to determine the ability of these fusion genes to enhance self-renewal, block differentiation, and induce leukemia in murine model(s) with a focus on the mechanism whereby these processes take place.

Pediatric AMKL is a heterogeneous disease comprising chimeric oncogenes or truncating GATA1 mutations that enhance self-renewal and block myeloid differentiation. Cooperating mutations that contribute to transformation include amplifications of chromosome 21 (either somatic or constitutional) as well as single-nucleotide variations and insertion/deletion in cohesin complex genes, CTCF, epigenetic regulators, and kinase genes. In ∼35% of pediatric non–DS-AMKL cases, the genetic alterations leading to the malignancy are unknown, warranting further comprehensive genomic studies (Figure 3). CBFA2T3-GLIS2 is the most frequent fusion event with a distinct biology in addition to a poor prognosis, occurring in 18% of patients. Development of targeted agents that inhibit the fusion directly, or critical self-renewal pathways upregulated as a result of the fusion, such as BMP, may provide therapeutic benefit. The diversity of CBFA2T3-GLIS2–negative nonDS-AMKL cases suggest that alternative less targeted approaches, such as the promotion of megakaryoblast differentiation, should be evaluated in an attempt to improve outcomes across patients with a wide spectrum of mutations.97,98  The presence of JAK/STAT- and RAS-pathwayactivating mutations provides a rationale for the use of kinase inhibitors, although their role as cooperating hits warrants caution, as these agents may be additive to existing treatment but not sufficient to eliminate disease on their own.

This work was supported by grants from the Eric Trump Foundation, Gabrielle Angel Foundation, and the American Lebanese Syrian Associated Charities of St. Jude Children’s Research Hospital.

Contribution: T.A.G. and J.R.D. wrote the manuscript.

Conflict-of-interest disclosure: The authors declare no competing financial interests.

Correspondence: Tanja A. Gruber, Department of Oncology, St. Jude Children’s Research Hospital, Memphis, TN 38105; e-mail: tanja.gruber@stjude.org.

1
Pagano
 
L
Pulsoni
 
A
Vignetti
 
M
, et al. 
Acute megakaryoblastic leukemia: experience of GIMEMA trials.
Leukemia
2002
, vol. 
16
 
9
(pg. 
1622
-
1626
)
2
Athale
 
UH
Razzouk
 
BI
Raimondi
 
SC
, et al. 
Biology and outcome of childhood acute megakaryoblastic leukemia: a single institution’s experience.
Blood
2001
, vol. 
97
 
12
(pg. 
3727
-
3732
)
3
Barnard
 
DR
Alonzo
 
TA
Gerbing
 
RB
Lange
 
B
Woods
 
WG
Children’s Oncology Group
Comparison of childhood myelodysplastic syndrome, AML FAB M6 or M7, CCG 2891: report from the Children’s Oncology Group.
Pediatr Blood Cancer
2007
, vol. 
49
 
1
(pg. 
17
-
22
)
4
Hitzler
 
JK
Zipursky
 
A
Origins of leukaemia in children with Down syndrome.
Nat Rev Cancer
2005
, vol. 
5
 
1
(pg. 
11
-
20
)
5
Arber
 
ABR
Orazi
 
A
Swerdlow
 
SHCE
Harris
 
NL
Acute myeloid leukaemia with myelodysplasia-related changes.
WHO Classification of Tumours of Haematopoietic and Lymphoid Tissues.
2008
Lyon, France
International Agency for Research on Cancer
(pg. 
124
-
126
)
6
Hitzler
 
JK
Cheung
 
J
Li
 
Y
Scherer
 
SW
Zipursky
 
A
GATA1 mutations in transient leukemia and acute megakaryoblastic leukemia of Down syndrome.
Blood
2003
, vol. 
101
 
11
(pg. 
4301
-
4304
)
7
Wechsler
 
J
Greene
 
M
McDevitt
 
MA
, et al. 
Acquired mutations in GATA1 in the megakaryoblastic leukemia of Down syndrome.
Nat Genet
2002
, vol. 
32
 
1
(pg. 
148
-
152
)
8
Hirose
 
Y
Kudo
 
K
Kiyoi
 
H
Hayashi
 
Y
Naoe
 
T
Kojima
 
S
Comprehensive analysis of gene alterations in acute megakaryoblastic leukemia of Down’s syndrome.
Leukemia
2003
, vol. 
17
 
11
(pg. 
2250
-
2252
)
9
Mundschau
 
G
Gurbuxani
 
S
Gamis
 
AS
Greene
 
ME
Arceci
 
RJ
Crispino
 
JD
Mutagenesis of GATA1 is an initiating event in Down syndrome leukemogenesis.
Blood
2003
, vol. 
101
 
11
(pg. 
4298
-
4300
)
10
Groet
 
J
McElwaine
 
S
Spinelli
 
M
, et al. 
Acquired mutations in GATA1 in neonates with Down’s syndrome with transient myeloid disorder.
Lancet
2003
, vol. 
361
 
9369
(pg. 
1617
-
1620
)
11
Rainis
 
L
Bercovich
 
D
Strehl
 
S
, et al. 
Mutations in exon 2 of GATA1 are early events in megakaryocytic malignancies associated with trisomy 21.
Blood
2003
, vol. 
102
 
3
(pg. 
981
-
986
)
12
Gamis
 
AS
Acute myeloid leukemia and Down syndrome evolution of modern therapy--state of the art review.
Pediatr Blood Cancer
2005
, vol. 
44
 
1
(pg. 
13
-
20
)
13
Gamis
 
AS
Woods
 
WG
Alonzo
 
TA
, et al. 
Children’s Cancer Group Study 2891
Increased age at diagnosis has a significantly negative effect on outcome in children with Down syndrome and acute myeloid leukemia: a report from the Children’s Cancer Group Study 2891.
J Clin Oncol
2003
, vol. 
21
 
18
(pg. 
3415
-
3422
)
14
Creutzig
 
U
Reinhardt
 
D
Diekamp
 
S
Dworzak
 
M
Stary
 
J
Zimmermann
 
M
AML patients with Down syndrome have a high cure rate with AML-BFM therapy with reduced dose intensity.
Leukemia
2005
, vol. 
19
 
8
(pg. 
1355
-
1360
)
15
Rao
 
A
Hills
 
RK
Stiller
 
C
, et al. 
Treatment for myeloid leukaemia of Down syndrome: population-based experience in the UK and results from the Medical Research Council AML 10 and AML 12 trials.
Br J Haematol
2006
, vol. 
132
 
5
(pg. 
576
-
583
)
16
Gruber
 
TA
Larson Gedman
 
A
Zhang
 
J
, et al. 
An Inv(16)(p13.3q24.3)-encoded CBFA2T3-GLIS2 fusion protein defines an aggressive subtype of pediatric acute megakaryoblastic leukemia.
Cancer Cell
2012
, vol. 
22
 
5
(pg. 
683
-
697
)
17
de Rooij
 
JD
Hollink
 
IH
Arentsen-Peters
 
ST
, et al. 
NUP98/JARID1A is a novel recurrent abnormality in pediatric acute megakaryoblastic leukemia with a distinct HOX gene expression pattern.
Leukemia
2013
, vol. 
27
 
12
(pg. 
2280
-
2288
)
18
Reinhardt
 
D
Diekamp
 
S
Langebrake
 
C
, et al. 
Acute megakaryoblastic leukemia in children and adolescents, excluding Down's syndrome: improved outcome with intensified induction treatment.
Leukemia
2005
, vol. 
19
 
8
(pg. 
1495
-
1496
)
19
Homans
 
AC
Verissimo
 
AM
Vlacha
 
V
Transient abnormal myelopoiesis of infancy associated with trisomy 21.
Am J Pediatr Hematol Oncol
1993
, vol. 
15
 
4
(pg. 
392
-
399
)
20
Shimada
 
A
Xu
 
G
Toki
 
T
Kimura
 
H
Hayashi
 
Y
Ito
 
E
Fetal origin of the GATA1 mutation in identical twins with transient myeloproliferative disorder and acute megakaryoblastic leukemia accompanying Down syndrome.
Blood
2004
, vol. 
103
 
1
pg. 
366
 
21
Taub
 
JW
Mundschau
 
G
Ge
 
Y
, et al. 
Prenatal origin of GATA1 mutations may be an initiating step in the development of megakaryocytic leukemia in Down syndrome.
Blood
2004
, vol. 
104
 
5
(pg. 
1588
-
1589
)
22
Pine
 
SR
Guo
 
Q
Yin
 
C
Jayabose
 
S
Druschel
 
CM
Sandoval
 
C
Incidence and clinical implications of GATA1 mutations in newborns with Down syndrome.
Blood
2007
, vol. 
110
 
6
(pg. 
2128
-
2131
)
23
Roberts
 
I
Alford
 
K
Hall
 
G
, et al. 
Oxford-Imperial Down Syndrome Cohort Study Group
GATA1-mutant clones are frequent and often unsuspected in babies with Down syndrome: identification of a population at risk of leukemia.
Blood
2013
, vol. 
122
 
24
(pg. 
3908
-
3917
)
24
Chen
 
J
Li
 
Y
Doedens
 
M
, et al. 
Functional differences between myeloid leukemia-initiating and transient leukemia cells in Down’s syndrome.
Leukemia
2010
, vol. 
24
 
5
(pg. 
1012
-
1017
)
25
Yoshida
 
K
Toki
 
T
Okuno
 
Y
, et al. 
The landscape of somatic mutations in Down syndrome-related myeloid disorders.
Nat Genet
2013
, vol. 
45
 
11
(pg. 
1293
-
1299
)
26
Shimizu
 
R
Engel
 
JD
Yamamoto
 
M
GATA1-related leukaemias.
Nat Rev Cancer
2008
, vol. 
8
 
4
(pg. 
279
-
287
)
27
Sankaran
 
VG
Ghazvinian
 
R
Do
 
R
, et al. 
Exome sequencing identifies GATA1 mutations resulting in Diamond-Blackfan anemia.
J Clin Invest
2012
, vol. 
122
 
7
(pg. 
2439
-
2443
)
28
Nichols
 
KE
Crispino
 
JD
Poncz
 
M
, et al. 
Familial dyserythropoietic anaemia and thrombocytopenia due to an inherited mutation in GATA1.
Nat Genet
2000
, vol. 
24
 
3
(pg. 
266
-
270
)
29
Freson
 
K
Devriendt
 
K
Matthijs
 
G
, et al. 
Platelet characteristics in patients with X-linked macrothrombocytopenia because of a novel GATA1 mutation.
Blood
2001
, vol. 
98
 
1
(pg. 
85
-
92
)
30
Mehaffey
 
MG
Newton
 
AL
Gandhi
 
MJ
Crossley
 
M
Drachman
 
JG
X-linked thrombocytopenia caused by a novel mutation of GATA-1.
Blood
2001
, vol. 
98
 
9
(pg. 
2681
-
2688
)
31
Yu
 
C
Niakan
 
KK
Matsushita
 
M
Stamatoyannopoulos
 
G
Orkin
 
SH
Raskind
 
WH
X-linked thrombocytopenia with thalassemia from a mutation in the amino finger of GATA-1 affecting DNA binding rather than FOG-1 interaction.
Blood
2002
, vol. 
100
 
6
(pg. 
2040
-
2045
)
32
Phillips
 
JD
Steensma
 
DP
Pulsipher
 
MA
Spangrude
 
GJ
Kushner
 
JP
Congenital erythropoietic porphyria due to a mutation in GATA1: the first trans-acting mutation causative for a human porphyria.
Blood
2007
, vol. 
109
 
6
(pg. 
2618
-
2621
)
33
Li
 
Z
Godinho
 
FJ
Klusmann
 
JH
Garriga-Canut
 
M
Yu
 
C
Orkin
 
SH
Developmental stage-selective effect of somatically mutated leukemogenic transcription factor GATA1.
Nat Genet
2005
, vol. 
37
 
6
(pg. 
613
-
619
)
34
Shimizu
 
R
Kobayashi
 
E
Engel
 
JD
Yamamoto
 
M
Induction of hyperproliferative fetal megakaryopoiesis by an N-terminally truncated GATA1 mutant.
Genes Cells
2009
, vol. 
14
 
9
(pg. 
1119
-
1131
)
35
Fujiwara
 
T
O’Geen
 
H
Keles
 
S
, et al. 
Discovering hematopoietic mechanisms through genome-wide analysis of GATA factor chromatin occupancy.
Mol Cell
2009
, vol. 
36
 
4
(pg. 
667
-
681
)
36
Chlon
 
TM
Crispino
 
JD
Combinatorial regulation of tissue specification by GATA and FOG factors.
Development
2012
, vol. 
139
 
21
(pg. 
3905
-
3916
)
37
Shaham
 
L
Vendramini
 
E
Ge
 
Y
, et al. 
MicroRNA-486-5p is an erythroid oncomiR of the myeloid leukemias of Down syndrome.
Blood
2015
, vol. 
125
 
8
(pg. 
1292
-
1301
)
38
Malinge
 
S
Izraeli
 
S
Crispino
 
JD
Insights into the manifestations, outcomes, and mechanisms of leukemogenesis in Down syndrome.
Blood
2009
, vol. 
113
 
12
(pg. 
2619
-
2628
)
39
Chou
 
ST
Opalinska
 
JB
Yao
 
Y
, et al. 
Trisomy 21 enhances human fetal erythro-megakaryocytic development.
Blood
2008
, vol. 
112
 
12
(pg. 
4503
-
4506
)
40
Tunstall-Pedoe
 
O
Roy
 
A
Karadimitris
 
A
, et al. 
Abnormalities in the myeloid progenitor compartment in Down syndrome fetal liver precede acquisition of GATA1 mutations.
Blood
2008
, vol. 
112
 
12
(pg. 
4507
-
4511
)
41
Maclean
 
GA
Menne
 
TF
Guo
 
G
, et al. 
Altered hematopoiesis in trisomy 21 as revealed through in vitro differentiation of isogenic human pluripotent cells.
Proc Natl Acad Sci USA
2012
, vol. 
109
 
43
(pg. 
17567
-
17572
)
42
Klusmann
 
JH
Li
 
Z
Böhmer
 
K
, et al. 
miR-125b-2 is a potential oncomiR on human chromosome 21 in megakaryoblastic leukemia.
Genes Dev
2010
, vol. 
24
 
5
(pg. 
478
-
490
)
43
Malinge
 
S
Bliss-Moreau
 
M
Kirsammer
 
G
, et al. 
Increased dosage of the chromosome 21 ortholog Dyrk1a promotes megakaryoblastic leukemia in a murine model of Down syndrome.
J Clin Invest
2012
, vol. 
122
 
3
(pg. 
948
-
962
)
44
Salek-Ardakani
 
S
Smooha
 
G
de Boer
 
J
, et al. 
ERG is a megakaryocytic oncogene.
Cancer Res
2009
, vol. 
69
 
11
(pg. 
4665
-
4673
)
45
Elagib
 
KE
Racke
 
FK
Mogass
 
M
Khetawat
 
R
Delehanty
 
LL
Goldfarb
 
AN
RUNX1 and GATA-1 coexpression and cooperation in megakaryocytic differentiation.
Blood
2003
, vol. 
101
 
11
(pg. 
4333
-
4341
)
46
Marcucci
 
G
Maharry
 
K
Whitman
 
SP
, et al. 
Cancer and Leukemia Group B Study
High expression levels of the ETS-related gene, ERG, predict adverse outcome and improve molecular risk-based classification of cytogenetically normal acute myeloid leukemia: a Cancer and Leukemia Group B Study.
J Clin Oncol
2007
, vol. 
25
 
22
(pg. 
3337
-
3343
)
47
Shimizu
 
K
Ichikawa
 
H
Tojo
 
A
, et al. 
An ets-related gene, ERG, is rearranged in human myeloid leukemia with t(16;21) chromosomal translocation.
Proc Natl Acad Sci USA
1993
, vol. 
90
 
21
(pg. 
10280
-
10284
)
48
Birger
 
Y
Goldberg
 
L
Chlon
 
TM
, et al. 
Perturbation of fetal hematopoiesis in a mouse model of Down syndrome’s transient myeloproliferative disorder.
Blood
2013
, vol. 
122
 
6
(pg. 
988
-
998
)
49
Stankiewicz
 
MJ
Crispino
 
JD
ETS2 and ERG promote megakaryopoiesis and synergize with alterations in GATA-1 to immortalize hematopoietic progenitor cells.
Blood
2009
, vol. 
113
 
14
(pg. 
3337
-
3347
)
50
Loughran
 
SJ
Kruse
 
EA
Hacking
 
DF
, et al. 
The transcription factor Erg is essential for definitive hematopoiesis and the function of adult hematopoietic stem cells.
Nat Immunol
2008
, vol. 
9
 
7
(pg. 
810
-
819
)
51
Bourquin
 
JP
Subramanian
 
A
Langebrake
 
C
, et al. 
Identification of distinct molecular phenotypes in acute megakaryoblastic leukemia by gene expression profiling.
Proc Natl Acad Sci USA
2006
, vol. 
103
 
9
(pg. 
3339
-
3344
)
52
Edwards
 
H
Xie
 
C
LaFiura
 
KM
, et al. 
RUNX1 regulates phosphoinositide 3-kinase/AKT pathway: role in chemotherapy sensitivity in acute megakaryocytic leukemia.
Blood
2009
, vol. 
114
 
13
(pg. 
2744
-
2752
)
53
Kon
 
A
Shih
 
LY
Minamino
 
M
, et al. 
Recurrent mutations in multiple components of the cohesin complex in myeloid neoplasms.
Nat Genet
2013
, vol. 
45
 
10
(pg. 
1232
-
1237
)
54
Thota
 
S
Viny
 
AD
Makishima
 
H
, et al. 
Genetic alterations of the cohesin complex genes in myeloid malignancies.
Blood
2014
, vol. 
124
 
11
(pg. 
1790
-
1798
)
55
Thol
 
F
Bollin
 
R
Gehlhaar
 
M
, et al. 
Mutations in the cohesin complex in acute myeloid leukemia: clinical and prognostic implications.
Blood
2014
, vol. 
123
 
6
(pg. 
914
-
920
)
56
Losada
 
A
Cohesin in cancer: chromosome segregation and beyond.
Nat Rev Cancer
2014
, vol. 
14
 
6
(pg. 
389
-
393
)
57
Wendt
 
KS
Yoshida
 
K
Itoh
 
T
, et al. 
Cohesin mediates transcriptional insulation by CCCTC-binding factor.
Nature
2008
, vol. 
451
 
7180
(pg. 
796
-
801
)
58
Dixon
 
JR
Selvaraj
 
S
Yue
 
F
, et al. 
Topological domains in mammalian genomes identified by analysis of chromatin interactions.
Nature
2012
, vol. 
485
 
7398
(pg. 
376
-
380
)
59
Xu
 
J
Shao
 
Z
Glass
 
K
, et al. 
Combinatorial assembly of developmental stage-specific enhancers controls gene expression programs during human erythropoiesis.
Dev Cell
2012
, vol. 
23
 
4
(pg. 
796
-
811
)
60
Yu
 
M
Riva
 
L
Xie
 
H
, et al. 
Insights into GATA-1-mediated gene activation versus repression via genome-wide chromatin occupancy analysis.
Mol Cell
2009
, vol. 
36
 
4
(pg. 
682
-
695
)
61
Oh
 
ST
Simonds
 
EF
Jones
 
C
, et al. 
Novel mutations in the inhibitory adaptor protein LNK drive JAK-STAT signaling in patients with myeloproliferative neoplasms.
Blood
2010
, vol. 
116
 
6
(pg. 
988
-
992
)
62
Geddis
 
AE
Megakaryopoiesis.
Semin Hematol
2010
, vol. 
47
 
3
(pg. 
212
-
219
)
63
Malinge
 
S
Ragu
 
C
Della-Valle
 
V
, et al. 
Activating mutations in human acute megakaryoblastic leukemia.
Blood
2008
, vol. 
112
 
10
(pg. 
4220
-
4226
)
64
Walters
 
DK
Mercher
 
T
Gu
 
TL
, et al. 
Activating alleles of JAK3 in acute megakaryoblastic leukemia.
Cancer Cell
2006
, vol. 
10
 
1
(pg. 
65
-
75
)
65
Sato
 
T
Toki
 
T
Kanezaki
 
R
, et al. 
Functional analysis of JAK3 mutations in transient myeloproliferative disorder and acute megakaryoblastic leukaemia accompanying Down syndrome.
Br J Haematol
2008
, vol. 
141
 
5
(pg. 
681
-
688
)
66
Carroll
 
A
Civin
 
C
Schneider
 
N
, et al. 
The t(1;22) (p13;q13) is nonrandom and restricted to infants with acute megakaryoblastic leukemia: a Pediatric Oncology Group Study.
Blood
1991
, vol. 
78
 
3
(pg. 
748
-
752
)
67
Ma
 
Z
Morris
 
SW
Valentine
 
V
, et al. 
Fusion of two novel genes, RBM15 and MKL1, in the t(1;22)(p13;q13) of acute megakaryoblastic leukemia.
Nat Genet
2001
, vol. 
28
 
3
(pg. 
220
-
221
)
68
Bernstein
 
J
Dastugue
 
N
Haas
 
OA
, et al. 
Nineteen cases of the t(1;22)(p13;q13) acute megakaryblastic leukaemia of infants/children and a review of 39 cases: report from a t(1;22) study group.
Leukemia
2000
, vol. 
14
 
1
(pg. 
216
-
218
)
69
Baruchel
 
A
Daniel
 
MT
Schaison
 
G
Berger
 
R
Nonrandom t(1;22)(p12-p13;q13) in acute megakaryocytic malignant proliferation.
Cancer Genet Cytogenet
1991
, vol. 
54
 
2
(pg. 
239
-
243
)
70
Mercher
 
T
Coniat
 
MB
Monni
 
R
, et al. 
Involvement of a human gene related to the Drosophila spen gene in the recurrent t(1;22) translocation of acute megakaryocytic leukemia.
Proc Natl Acad Sci USA
2001
, vol. 
98
 
10
(pg. 
5776
-
5779
)
71
Mercher
 
T
Raffel
 
GD
Moore
 
SA
, et al. 
The OTT-MAL fusion oncogene activates RBPJ-mediated transcription and induces acute megakaryoblastic leukemia in a knockin mouse model.
J Clin Invest
2009
, vol. 
119
 
4
(pg. 
852
-
864
)
72
Halene
 
S
Gao
 
Y
Hahn
 
K
, et al. 
Serum response factor is an essential transcription factor in megakaryocytic maturation.
Blood
2010
, vol. 
116
 
11
(pg. 
1942
-
1950
)
73
Miralles
 
F
Posern
 
G
Zaromytidou
 
AI
Treisman
 
R
Actin dynamics control SRF activity by regulation of its coactivator MAL.
Cell
2003
, vol. 
113
 
3
(pg. 
329
-
342
)
74
Smith
 
EC
Teixeira
 
AM
Chen
 
RC
, et al. 
Induction of megakaryocyte differentiation drives nuclear accumulation and transcriptional function of MKL1 via actin polymerization and RhoA activation.
Blood
2013
, vol. 
121
 
7
(pg. 
1094
-
1101
)
75
Cheng
 
EC
Luo
 
Q
Bruscia
 
EM
, et al. 
Role for MKL1 in megakaryocytic maturation.
Blood
2009
, vol. 
113
 
12
(pg. 
2826
-
2834
)
76
Gilles
 
L
Bluteau
 
D
Boukour
 
S
, et al. 
MAL/SRF complex is involved in platelet formation and megakaryocyte migration by regulating MYL9 (MLC2) and MMP9.
Blood
2009
, vol. 
114
 
19
(pg. 
4221
-
4232
)
77
Oswald
 
F
Kostezka
 
U
Astrahantseff
 
K
, et al. 
SHARP is a novel component of the Notch/RBP-Jkappa signalling pathway.
EMBO J
2002
, vol. 
21
 
20
(pg. 
5417
-
5426
)
78
Ariyoshi
 
M
Schwabe
 
JW
A conserved structural motif reveals the essential transcriptional repression function of Spen proteins and their role in developmental signaling.
Genes Dev
2003
, vol. 
17
 
15
(pg. 
1909
-
1920
)
79
Raffel
 
GD
Mercher
 
T
Shigematsu
 
H
, et al. 
Ott1(Rbm15) has pleiotropic roles in hematopoietic development.
Proc Natl Acad Sci USA
2007
, vol. 
104
 
14
(pg. 
6001
-
6006
)
80
Niu
 
C
Zhang
 
J
Breslin
 
P
Onciu
 
M
Ma
 
Z
Morris
 
SW
c-Myc is a target of RNA-binding motif protein 15 in the regulation of adult hematopoietic stem cell and megakaryocyte development.
Blood
2009
, vol. 
114
 
10
(pg. 
2087
-
2096
)
81
Descot
 
A
Rex-Haffner
 
M
Courtois
 
G
, et al. 
OTT-MAL is a deregulated activator of serum response factor-dependent gene expression.
Mol Cell Biol
2008
, vol. 
28
 
20
(pg. 
6171
-
6181
)
82
Ma
 
X
Renda
 
MJ
Wang
 
L
, et al. 
Rbm15 modulates Notch-induced transcriptional activation and affects myeloid differentiation.
Mol Cell Biol
2007
, vol. 
27
 
8
(pg. 
3056
-
3064
)
83
Kim
 
Y
Schulz
 
VP
Satake
 
N
, et al. 
Whole-exome sequencing identifies a novel somatic mutation in MMP8 associated with a t(1;22)-acute megakaryoblastic leukemia.
Leukemia
2014
, vol. 
28
 
4
(pg. 
945
-
948
)
84
Radtke
 
I
Mullighan
 
CG
Ishii
 
M
, et al. 
Genomic analysis reveals few genetic alterations in pediatric acute myeloid leukemia.
Proc Natl Acad Sci USA
2009
, vol. 
106
 
31
(pg. 
12944
-
12949
)
85
Masetti
 
R
Pigazzi
 
M
Togni
 
M
, et al. 
CBFA2T3-GLIS2 fusion transcript is a novel common feature in pediatric, cytogenetically normal AML, not restricted to FAB M7 subtype.
Blood
2013
, vol. 
121
 
17
(pg. 
3469
-
3472
)
86
Higuchi
 
M
O’Brien
 
D
Kumaravelu
 
P
Lenny
 
N
Yeoh
 
EJ
Downing
 
JR
Expression of a conditional AML1-ETO oncogene bypasses embryonic lethality and establishes a murine model of human t(8;21) acute myeloid leukemia.
Cancer Cell
2002
, vol. 
1
 
1
(pg. 
63
-
74
)
87
Huang
 
FW
Hodis
 
E
Xu
 
MJ
Kryukov
 
GV
Chin
 
L
Garraway
 
LA
Highly recurrent TERT promoter mutations in human melanoma.
Science
2013
, vol. 
339
 
6122
(pg. 
957
-
959
)
88
Mansour
 
MR
Abraham
 
BJ
Anders
 
L
, et al. 
Oncogene regulation. An oncogenic super-enhancer formed through somatic mutation of a noncoding intergenic element.
Science
2014
, vol. 
346
 
6215
(pg. 
1373
-
1377
)
89
van Zutven
 
LJ
Onen
 
E
Velthuizen
 
SC
, et al. 
Identification of NUP98 abnormalities in acute leukemia: JARID1A (12p13) as a new partner gene.
Genes Chromosomes Cancer
2006
, vol. 
45
 
5
(pg. 
437
-
446
)
90
Reader
 
JC
Meekins
 
JS
Gojo
 
I
Ning
 
Y
A novel NUP98-PHF23 fusion resulting from a cryptic translocation t(11;17)(p15;p13) in acute myeloid leukemia.
Leukemia
2007
, vol. 
21
 
4
(pg. 
842
-
844
)
91
Wang
 
GG
Song
 
J
Wang
 
Z
, et al. 
Haematopoietic malignancies caused by dysregulation of a chromatin-binding PHD finger.
Nature
2009
, vol. 
459
 
7248
(pg. 
847
-
851
)
92
Kawada
 
H
Ito
 
T
Pharr
 
PN
Spyropoulos
 
DD
Watson
 
DK
Ogawa
 
M
Defective megakaryopoiesis and abnormal erythroid development in Fli-1 gene-targeted mice.
Int J Hematol
2001
, vol. 
73
 
4
(pg. 
463
-
468
)
93
Visvader
 
JE
Crossley
 
M
Hill
 
J
Orkin
 
SH
Adams
 
JM
The C-terminal zinc finger of GATA-1 or GATA-2 is sufficient to induce megakaryocytic differentiation of an early myeloid cell line.
Mol Cell Biol
1995
, vol. 
15
 
2
(pg. 
634
-
641
)
94
Argiropoulos
 
B
Humphries
 
RK
Hox genes in hematopoiesis and leukemogenesis.
Oncogene
2007
, vol. 
26
 
47
(pg. 
6766
-
6776
)
95
Buijs
 
A
van Rompaey
 
L
Molijn
 
AC
, et al. 
The MN1-TEL fusion protein, encoded by the translocation (12;22)(p13;q11) in myeloid leukemia, is a transcription factor with transforming activity.
Mol Cell Biol
2000
, vol. 
20
 
24
(pg. 
9281
-
9293
)
96
Heuser
 
M
Yun
 
H
Berg
 
T
, et al. 
Cell of origin in AML: susceptibility to MN1-induced transformation is regulated by the MEIS1/AbdB-like HOX protein complex.
Cancer Cell
2011
, vol. 
20
 
1
(pg. 
39
-
52
)
97
Wen
 
Q
Goldenson
 
B
Silver
 
SJ
, et al. 
Identification of regulators of polyploidization presents therapeutic targets for treatment of AMKL.
Cell
2012
, vol. 
150
 
3
(pg. 
575
-
589
)
98
Krause
 
DS
Crispino
 
JD
Molecular pathways: induction of polyploidy as a novel differentiation therapy for leukemia.
Clin Cancer Res
2013
, vol. 
19
 
22
(pg. 
6084
-
6088
)
Sign in via your Institution