martes, 8 de octubre de 2019

Childhood Acute Myeloid Leukemia/Other Myeloid Malignancies Treatment (PDQ®)–Health Professional Version - National Cancer Institute 2/8

Childhood Acute Myeloid Leukemia/Other Myeloid Malignancies Treatment (PDQ®)–Health Professional Version - National Cancer Institute

National Cancer Institute



Childhood Acute Myeloid Leukemia/Other Myeloid Malignancies Treatment (PDQ®)–Health Professional Version



Histochemical, Immunophenotypic, and Molecular Evaluation for Childhood AML

Histochemical Evaluation

The treatment for children with acute myeloid leukemia (AML) differs significantly from that for acute lymphoblastic leukemia (ALL). As a consequence, it is critical to distinguish AML from ALL. Special histochemical stains performed on bone marrow specimens of children with acute leukemia can be helpful to confirm their diagnosis. The stains most commonly used include myeloperoxidase, periodic acid-Schiff, Sudan Black B, and esterase. In most cases, the staining pattern with these histochemical stains will distinguish AML from acute myelomonocytic leukemia (AMML) and ALL (refer to Table 5). Histochemical stains have been mostly replaced by flow cytometric immunophenotyping.
Table 5. Histochemical Staining Patternsa
ENLARGE
 M0AML, APL (M1-M3)AMML (M4)AMoL (M5)AEL (M6)AMKL (M7)ALL
AEL = acute erythroid leukemia; ALL = acute lymphoblastic leukemia; AML = acute myeloid leukemia; AMKL = acute megakaryocytic leukemia; AMML = acute myelomonocytic leukemia; AMoL = acute monocytic leukemia; APL = acute promyelocytic leukemia; PAS = periodic acid-Schiff.
aRefer to the French-American-British (FAB) Classification for Childhood Acute Myeloid Leukemia section of this summary for more information about the morphologic-histochemical classification system for AML.
bThese reactions are inhibited by fluoride.
Myeloperoxidase-++----
Nonspecific esterases       
 Chloracetate-++±---
 Alpha-naphthol acetate--bb-± b-
Sudan Black B-++----
PAS--±±+-+

Immunophenotypic Evaluation

The use of monoclonal antibodies to determine cell-surface antigens of AML cells is helpful to reinforce the histologic diagnosis. Various lineage-specific monoclonal antibodies that detect antigens on AML cells should be used at the time of initial diagnostic workup, along with a battery of lineage-specific T-lymphocyte and B-lymphocyte markers to help distinguish AML from ALL and acute leukemias of ambiguous lineage. The expression of various cluster determinant (CD) proteins that are relatively lineage-specific for AML include CD33, CD13, CD14, CDw41 (or platelet antiglycoprotein IIb/IIIa), CD15, CD11B, CD36, and antiglycophorin A. Lineage-associated B-lymphocytic antigens CD10, CD19, CD20, CD22, and CD24 may be present in 10% to 20% of AML cases, but monoclonal surface immunoglobulin and cytoplasmic immunoglobulin heavy chains are usually absent; similarly, CD2, CD3, CD5, and CD7 lineage-associated T-lymphocytic antigens are present in 20% to 40% of AML cases.[1-3] The aberrant expression of lymphoid-associated antigens by AML cells is relatively common but generally has no prognostic significance.[1,2]
Immunophenotyping can also be helpful in distinguishing the following French-American-British (FAB) classification subtypes of AML:
  • Testing for the presence of HLA-antigen D related (HLA-DR) can be helpful in identifying acute promyelocytic leukemia (APL). Overall, HLA-DR is expressed on 75% to 80% of AML cells but rarely expressed on APL cells.[4,5] In addition, APL is characterized by bright CD33 expression and by CD117 (c-KIT) expression in most cases, heterogeneous expression of CD13 with CD34, CD11a, and CD18 often negative or low.[4,5] The APL microgranular variant M3v more commonly expresses CD34 along with CD2.[4,6]
  • Testing for the presence of glycoprotein Ib, glycoprotein IIb/IIIa, or Factor VIII antigen expression is helpful in making the diagnosis of M7 (megakaryocytic leukemia).
  • Glycophorin expression is helpful in making the diagnosis of M6 (erythroid leukemia).[7]
Less than 5% of cases of acute leukemia in children are of ambiguous lineage, expressing features of both myeloid and lymphoid lineage.[8-10] These cases are distinct from ALL with myeloid coexpression in that the predominant lineage cannot be determined by immunophenotypic and histochemical studies. The definition of leukemia of ambiguous lineage varies among studies, although most investigators now use criteria established by the European Group for the Immunological Characterization of Leukemias (EGIL) or the more stringent World Health Organization (WHO) criteria.[11-13] In the WHO classification, the presence of myeloperoxidase (MPO) is required to establish myeloid lineage. This is not the case for the EGIL classification. The 2016 revision to the WHO classification also denotes that in some cases, leukemia with otherwise classic B-cell ALL immunophenotype may also express low-intensity MPO without other myeloid features, and the clinical significance of that finding is unclear such that one should be cautious before designating these cases as mixed phenotype acute leukemia (MPAL).[14]

Molecular Evaluation

Comprehensive molecular profiling of pediatric and adult AML has shown that AML is a disease demonstrating both commonalities and differences across the age spectrum.[15,16]
  • Pediatric AML, in contrast to AML in adults, is typically a disease of recurring chromosomal alterations (refer to Table 6 for a list of common gene fusions).[15,17] Within the pediatric age range, certain gene fusions occur primarily in children younger than 5 years (e.g., NUP98 gene fusions, KMT2A gene fusions, and CBFA2T3-GLIS2), while others occur primarily in children aged 5 years and older (e.g., RUNX1-RUNX1T1CBFB-MYH11, and NPM1-RARA).
  • Pediatric patients with AML have low rates of mutations, with most cases showing less than one somatic change in protein-coding regions per megabase.[16] This mutation rate is somewhat lower than that observed in adult AML and is much lower than the mutation rate for cancers that respond to checkpoint inhibitors (e.g., melanoma).[16]
  • The pattern of gene mutations differs between pediatric and adult AML cases. For example, IDH1/IDH2TP53RUNX1, and DNMT3A mutations are more common in adult AML than in pediatric AML, while NRAS and WT1 mutations are significantly more common in pediatric AML.[15,16]
Genetic analysis of leukemia blast cells (using both conventional cytogenetic methods and molecular methods) is performed on children with AML because both chromosomal and molecular abnormalities are important diagnostic and prognostic markers.[17-23] Clonal chromosomal abnormalities are identified in the blasts of about 75% of children with AML and are useful in defining subtypes with both prognostic and therapeutic significance.
Detection of molecular abnormalities can also aid in risk stratification and treatment allocation. For example, mutations of NPM and CEBPA are associated with favorable outcomes while certain mutations of FLT3 portend a high risk of relapse, and identifying the latter mutations may allow for targeted therapy.[24-27]
The 2016 revision to the World Health Organization (WHO) classification of myeloid neoplasms and acute leukemia emphasizes that recurrent chromosomal translocations in pediatric AML may be unique or have a different prevalence than in adult AML.[14] The pediatric AML chromosomal translocations that are found by conventional chromosome analysis and those that are cryptic (identified only with fluorescence in situ hybridization or molecular techniques) occur at higher rates than in adults. These recurrent translocations are summarized in Table 6.[14] Table 6 also shows, in the bottom three rows, additional relatively common recurrent translocations observed in children with AML.[21,22,28]
Table 6. Common Pediatric Acute Myeloid Leukemia (AML) Chromosomal Translocations
Gene Fusion ProductChromosomal TranslocationPrevalence in Pediatric AML (%)
aCryptic chromosomal translocation.
KMT2A (MLL) translocated11q23.325.0
NUP98-NSD1at(5;11)(q35.3;p15.5)7.0
CBFA2T3-GLIS2ainv(16)(p13.3;q24.3)3.0
NUP98-KDM5A4at(11;12)(p15.5;p13.5)3.0
DEK-NUP214t(6;9)(p23;q34.1)1.7
RBM15(OTT)-MKL1(MAL)t(1;22)(p13.3;q13.1)0.8
MNX1-ETV6t(7;12)(q36.3;p13.2)0.8
KAT6A-CREBBPt(8;16)(p11.2;p13.3)0.5
RUNX1-RUNX1T1t(8;21)(q22;q22)13–14
CBFB-MYH11inv(16)(p13.1;q22) or t(16;16)(p13.1;q22)4–9
PML-RARAt(15;17)(q24;q21)6–11
The genomic landscape of pediatric AML cases can change from diagnosis to relapse, with mutations detectable at diagnosis dropping out at relapse and, conversely, with new mutations appearing at relapse. In a study of 20 cases for which sequencing data were available at diagnosis and relapse, a key finding was that the variant allele frequency at diagnosis strongly correlated with persistence of mutations at relapse.[29] Approximately 90% of the diagnostic variants with variant allele frequency greater than 0.4 persisted to relapse, compared with only 28% with variant allele frequency less than 0.2 (P < .001). This observation is consistent with previous results showing that presence of the FLT3-ITD mutation predicted for poor prognosis only when there was a high FLT3-ITD allelic ratio.
Specific recurring cytogenetic and molecular abnormalities are briefly described below. The abnormalities are listed by those in clinical use that identify patients with favorable or unfavorable prognosis, followed by other abnormalities. The nomenclature of the 2016 revision to the WHO classification of myeloid neoplasms and acute leukemia is incorporated for disease entities where relevant.

Molecular abnormalities associated with a favorable prognosis

Molecular abnormalities associated with a favorable prognosis include the following:
  • Core-binding factor (CBF) AML includes cases with RUNX1-RUNX1T1 and CBFB-MYH11 fusion genes that disrupt the activity of core-binding factor, which contains RUNX1 and CBFB. These are specific entities in the 2016 revision to the WHO classification of myeloid neoplasms and acute leukemia.
    • AML with t(8;21)(q22;q22.1); RUNX1-RUNX1T1: In leukemias with t(8;21), the RUNX1 (AML1) gene on chromosome 21 is fused with the RUNX1T1 (ETO) gene on chromosome 8. The t(8;21) translocation is associated with the FAB M2 subtype and with granulocytic sarcomas.[30,31] Adults with t(8;21) have a more favorable prognosis than do adults with other types of AML.[18,32] Children with t(8;21) have a more favorable outcome than do children with AML characterized by normal or complex karyotypes,[18,33-35] with 5-year overall survival (OS) of 74% to 90%.[21,22,36] The t(8;21) translocation occurs in approximately 12% of children with AML.[21,22,36]
    • AML with inv(16)(p13.1;q22) or t(16;16)(p13.1;q22); CBFB-MYH11: In leukemias with inv(16), the CBFB gene at chromosome band 16q22 is fused with the MYH11 gene at chromosome band 16p13. The inv(16) translocation is associated with the FAB M4Eo subtype.[37] Inv(16) confers a favorable prognosis for both adults and children with AML,[18,33-35] with a 5-year OS of about 85%.[21,22] Inv(16) occurs in 7% to 9% of children with AML.[21,22,36] As noted above, cases with CBFB-MYH11 and cases with RUNX1-RUNX1T1 have distinctive secondary mutations; CBFB-MYH11 secondary mutations are primarily restricted to genes that activate receptor tyrosine kinase signaling (NRASFLT3, and KIT).[38,39]
    • AML with t(16;21)(q24;q22); RUNX1-CBFA2T3: In leukemias with t(16;21)(q24;q22), the RUNX1 gene is fused with the CBFA2T3 gene, and the gene expression profile is closely related to that of AML cases with t(8;21) and RUNX1-RUNX1T1.[40] These patients present at a median age of 7 years and are rare, representing approximately 0.1% to 0.3% of pediatric AML cases. Among 23 patients with RUNX1-CBFA2T3, five presented with secondary AML, including two patients who had a primary diagnosis of Ewing sarcoma. Outcome for the cohort of 23 patients was favorable, with a 4-year EFS of 77% and a cumulative incidence of relapse of 0%.[40]
    Both RUNX1-RUNX1T1 and CBFB-MYH11 subtypes commonly show mutations in genes that activate receptor tyrosine kinase signaling (e.g., NRASFLT3, and KIT); NRAS and KIT are the most commonly mutated genes for both subtypes. KIT mutations may indicate increased risk of treatment failure for patients with core-binding factor AML, although the prognostic significance of KIT mutations may be dependent on the mutant-allele ratio (high ratio unfavorable) and/or the specific type of mutation (exon 17 mutations unfavorable).[38,39] A study of children with RUNX1-RUNX1T1 AML observed KIT mutations in 24% of cases (79% being exon 17 mutations) and RAS mutations in 15%, but neither were significantly associated with outcome.[36]
    Although both RUNX1-RUNX1T1 and CBFB-MYH11 fusion genes disrupt the activity of core-binding factor, cases with these genomic alterations have distinctive secondary mutations.[38,39]
    • RUNX1-RUNX1T1 cases also have frequent mutations in genes regulating chromatin conformation (e.g., ASXL1 and ASXL2) (40% of cases) and genes encoding members of the cohesin complex (20% of cases). Mutations in ASXL1 and ASXL2 and mutations in members of the cohesin complex are rare in CBFB-MYH11 leukemias.[38,39]
    • A study of 204 adults with RUNX1-RUNX1T1 AML found that ASXL2 mutations (present in 17% of cases) and ASXL1 or ASXL2 mutations (present in 25% of cases) lacked prognostic significance.[41] Similar results, albeit with smaller numbers, were reported for children with RUNX1-RUNX1T1 AML and ASXL1 and ASXL2 mutations.[42]
  • Acute promyelocytic leukemia (APL) with PML-RARA: APL represents about 7% of children with AML.[22,43] AML with t(15;17) is invariably associated with APL, a distinct subtype of AML that is treated differently than other types of AML because of its marked sensitivity to arsenic trioxide and the differentiating effects of all-trans retinoic acid. The t(15;17) translocation or other more complex chromosomal rearrangements may lead to the production of a fusion protein involving the retinoid acid receptor alpha and PML.[44] The WHO 2016 revision does not include the t(15;17) cytogenetic designation to stress the significance of the PML-RARA fusion, which may be cryptic or result from complex karyotypic changes.[14]
    Utilization of quantitative reverse transcriptase–polymerase chain reaction (RT-PCR) for PML-RARA transcripts has become standard practice.[45] Quantitative RT-PCR allows identification of the three common transcript variants and is used for monitoring response on treatment and early detection of molecular relapse.[46] Other much less common translocations involving the retinoic acid receptor alpha can also result in APL (e.g., t(11;17)(q23;q21) involving the PLZF gene).[47-49] Identification of cases with the t(11;17)(q23;q21) is important because of their decreased sensitivity to all-trans retinoic acid.[44,47]
  • AML with mutated NPM1NPM1 is a protein that has been linked to ribosomal protein assembly and transport as well as being a molecular chaperone involved in preventing protein aggregation in the nucleolus. Immunohistochemical methods can be used to accurately identify patients with NPM1 mutations by the demonstration of cytoplasmic localization of NPM.[50] Mutations in the NPM1 protein that diminish its nuclear localization are primarily associated with a subset of AML with a normal karyotype, absence of CD34 expression,[51] and an improved prognosis in the absence of FLT3–internal tandem duplication (ITD) mutations in adults and younger adults.[51-56]
    Studies of children with AML suggest a lower rate of occurrence of NPM1 mutations in children compared with adults with normal cytogenetics. NPM1 mutations occur in approximately 8% of pediatric patients with AML and are uncommon in children younger than 2 years.[24,25,57,58NPM1 mutations are associated with a favorable prognosis in patients with AML characterized by a normal karyotype.[24,25,58] For the pediatric population, conflicting reports have been published regarding the prognostic significance of an NPM1 mutation when a FLT3-ITD mutation is also present. One study reported that an NPM1 mutation did not completely abrogate the poor prognosis associated with having a FLT3-ITD mutation,[24,59] but other studies showed no impact of a FLT3-ITD mutation on the favorable prognosis associated with an NPM1 mutation.[16,25,58]
  • AML with biallelic mutations of CEBPA: Mutations in the CEBPA gene occur in a subset of children and adults with cytogenetically normal AML.[60] In adults younger than 60 years, approximately 15% of cytogenetically normal AML cases have mutations in CEBPA.[55] Outcomes for adults with AML with CEBPA mutations appear to be relatively favorable and similar to that of patients with core-binding factor leukemias.[55,61] Studies in adults with AML have demonstrated that CEBPA double-mutant, but not single-mutant, AML is independently associated with a favorable prognosis,[62-65] leading to the WHO 2016 revision that requires biallelic mutations for the disease definition.[14]
    CEBPA mutations occur in 5% to 8% of children with AML and have been preferentially found in the cytogenetically normal subtype of AML with FAB M1 or M2; 70% to 80% of pediatric patients have double-mutant alleles, which is predictive of a significantly improved survival, similar to the effect observed in adult studies.[26,66] Although both double-mutant and single-mutant alleles of CEBPA were associated with a favorable prognosis in children with AML in one large study,[26] a second study observed inferior outcome for patients with single CEBPA mutations.[66] However, very low numbers of children with single-allele mutants were included in these two studies (only 13 total patients), which makes a conclusion regarding the prognostic significance of single-allele CEBPA mutations in children premature.[26] In newly diagnosed patients with double-mutant CEBPA AML, germline screening should be considered in addition to usual family history queries, because 5% to 10% of these patients are reported to have a germline CEBPA mutation.[60]
  • Myeloid leukemia associated with Down syndrome (GATA1 mutations): GATA1 mutations are present in most, if not all, Down syndrome children with either transient abnormal myelopoiesis (TAM) or acute megakaryoblastic leukemia (AMKL).[67-70GATA1 mutations were also observed in 9% of non–Down syndrome children and 4% of adults with AMKL (with coexistence of amplification of the Down syndrome Critical Region on chromosome 21 in 9 of 10 cases).[71GATA1 is a transcription factor that is required for normal development of erythroid cells, megakaryocytes, eosinophils, and mast cells.[72]
    GATA1 mutations confer increased sensitivity to cytarabine by down-regulating cytidine deaminase expression, possibly providing an explanation for the superior outcome of children with Down syndrome and M7 AML when treated with cytarabine-containing regimens.[73]

Molecular abnormalities associated with an unfavorable prognosis

Molecular abnormalities associated with an unfavorable prognosis include the following:
  • Chromosomes 5 and 7: Chromosomal abnormalities associated with poor prognosis in adults with AML include those involving chromosome 5 (monosomy 5 and del(5q)) and chromosome 7 (monosomy 7).[18,32,74] These cytogenetic subgroups represent approximately 2% and 4% of pediatric AML cases, respectively, and are also associated with poor prognosis in children.[21,32,74-78]
    In the past, patients with del(7q) were also considered to be at high risk of treatment failure, and data from adults with AML support a poor prognosis for both del(7q) and monosomy 7.[23] However, outcome for children with del(7q), but not monosomy 7, appears comparable to that of other children with AML.[22,77] The presence of del(7q) does not abrogate the prognostic significance of favorable cytogenetic characteristics (e.g., inv(16) and t(8;21)).[18,77,79]
    Chromosome 5 and 7 abnormalities appear to lack prognostic significance in AML patients with Down syndrome who are aged 4 years and younger.[80]
  • AML with inv(3)(q21.3;q26.2) or t(3;3)(q21.3;q26.2); GATA2MECOMMECOM at chromosome 3q26 codes for two proteins, EVI1 and MDS1-EVI1, both of which are transcription regulators. The inv(3) and t(3;3) abnormalities lead to overexpression of EVI1 and to reduced expression of GATA2.[81,82] These abnormalities are associated with poor prognosis in adults with AML,[18,32,83] but are very uncommon in children (<1% of pediatric AML cases).[21,34,84]
    Abnormalities involving MECOM can be detected in some AML cases with other 3q abnormalities and are also associated with poor prognosis.
  • FLT3 mutations: Presence of a FLT3-ITD mutation appears to be associated with poor prognosis in adults with AML,[85] particularly when both alleles are mutated or there is a high ratio of the mutant allele to the normal allele.[86,87FLT3-ITD mutations also convey a poor prognosis in children with AML.[27,59,88-91] The frequency of FLT3-ITD mutations in children is lower than that observed in adults, especially for children younger than 10 years, for whom 5% to 10% of cases have the mutation (compared with approximately 30% in adults).[90-92]
    The prognostic significance of FLT3-ITD is modified by the presence of other recurring genomic alterations. The prevalence of FLT3-ITD is increased in certain genomic subtypes of pediatric AML, including those with the NUP98-NSD1 fusion gene, of which 80% to 90% have FLT3-ITD.[93,94] Approximately 15% of patients with FLT3-ITD have NUP98-NSD1, and patients with both FLT3-ITD and NUP98-NSD1 have a poorer prognosis than do patients who have FLT3-ITD without NUP98-NSD1.[94] For patients who have FLT3-ITD, the presence of either WT1 mutations or NUP98-NSD1 fusions is associated with poorer outcome (EFS rates below 25%) than for patients who have FLT3-ITD without these alterations.[16] Conversely, when FLT3-ITD is accompanied by NPM1 mutations, the outcome is relatively favorable and is similar to that of pediatric AML cases without FLT3-ITD.[16]
    For APL, FLT3-ITD and point mutations occur in 30% to 40% of children and adults.[86,89,90,95-99] Presence of the FLT3-ITD mutation is strongly associated with the microgranular variant (M3v) of APL and with hyperleukocytosis.[89,97,100,101] It remains unclear whether FLT3 mutations are associated with poorer prognosis in patients with APL who are treated with modern therapy that includes all-trans retinoic acid and arsenic trioxide.[95,96,99,100,102-105]
    Activating point mutations of FLT3 have also been identified in both adults and children with AML, although the clinical significance of these mutations is not clearly defined. Some of these point mutations appear to be specific to pediatric patients.[16]
  • AML with t(16;21)(p11;q22); FUS-ERGIn leukemias with t(16;21)(p11;q22), the FUS gene is joined with the ERG gene, producing a distinctive AML subtype with a gene expression profile that clusters separately from other cytogenetic subgroups.[40] These patients present at a median age of 8 to 9 years and are rare, representing approximately 0.3% to 0.5% of pediatric AML cases. For a cohort of 31 patients with FUS-ERG AML, outcome was poor, with a 4-year EFS of 7% and a cumulative incidence of relapse of 74%.[40]

Other molecular abnormalities observed in pediatric AML

Other molecular abnormalities observed in pediatric AML include the following:
  • KMT2A (MLL) gene rearrangements: KMT2A gene rearrangement occurs in approximately 20% of children with AML.[21,22] These cases, including most AMLs secondary to epipodophyllotoxin,[106] are generally associated with monocytic differentiation (FAB M4 and M5). KMT2A rearrangements are also reported in approximately 10% of FAB M7 (AMKL) patients (see below).[71,107]
    The most common translocation, representing approximately 50% of KMT2A-rearranged cases in the pediatric AML population, is t(9;11)(p22;q23), in which the KMT2A gene is fused with MLLT3(AF9) gene.[108] The WHO 2016 revision defined AML with t(9;11)(p21.3;q23.3); MLLT3-KMT2A as a distinctive disease entity. However, more than 50 different fusion partners have been identified for the KMT2A gene in patients with AML.
    The median age for 11q23/KMT2A-rearranged cases in children is approximately 2 years, and most translocation subgroups have a median age at presentation of younger than 5 years.[108] However, significantly older median ages are seen at presentation of pediatric cases with t(6;11)(q27;q23) (12 years) and t(11;17)(q23;q21) (9 years).[108]
    Outcome for patients with de novo AML and KMT2A gene rearrangement is generally reported as being similar to that for other patients with AML.[18,21,108,109] However, as the KMT2A gene can participate in translocations with many different fusion partners, the specific fusion partner appears to influence prognosis, as demonstrated by a large international retrospective study evaluating outcome for 756 children with 11q23- or KMT2A-rearranged AML.[108] For example, cases with t(1;11)(q21;q23), representing 3% of all 11q23/KMT2A-rearranged AML, showed a highly favorable outcome, with a 5-year event-free survival (EFS) of 92%.
    While reports from single clinical trial groups have variably described more favorable prognosis for patients with AML who have t(9;11)(p21.3;q23.3)/MLLT3-KMT2A, the international retrospective study did not confirm the favorable prognosis for this subgroup.[18,21,108,110-112] An international collaboration evaluating pediatric AMKL patients observed that the presence of t(9;11), which was seen in approximately 5% of AMKL cases, was associated with an inferior outcome compared with other AMKL cases.[107]
    KMT2A-rearranged AML subgroups that appear to be associated with poor outcome include the following:
    • Cases with the t(10;11) translocation are a group at high risk of relapse in bone marrow and the CNS.[18,22,113] Some cases with the t(10;11) translocation have fusion of the KMT2A gene with the AF10-MLLT10 at 10p12, while others have fusion of KMT2A with ABI1 at 10p11.2.[114,115] An international retrospective study found that these cases, which present at a median age of approximately 1 year, have a 5-year EFS of 20% to 30%.[108]
    • Patients with t(6;11)(q27;q23) have a poor outcome, with a 5-year EFS of 11%.
    • Patients with t(4;11)(q21;q23) also have a poor outcome, with a 5-year EFS of 29%.[108]
    • A follow-up study by the international collaborative group demonstrated that additional cytogenetic abnormalities further influenced outcome of children with KMT2A translocations, with complex karyotypes and trisomy 19 predicting poor outcome and trisomy 8 predicting a more favorable outcome.[116]
  • AML with t(6;9)(p23;q34.1); DEK-NUP214: t(6;9) leads to the formation of a leukemia-associated fusion protein DEK-NUP214.[117,118] This subgroup of AML has been associated with a poor prognosis in adults with AML,[117,119,120] and occurs infrequently in children (less than 1% of AML cases). The median age of children with DEK-NUP214 AML is 10 to 11 years, and approximately 40% of pediatric patients have FLT3-ITD.[121,122]
    t(6;9) AML appears to be associated with a high risk of treatment failure in children, particularly for those not proceeding to allogeneic stem cell transplantation.[21,118,121,122]
  • Molecular subgroups of non–Down syndrome acute megakaryoblastic leukemia (AMKL): AMKL accounts for approximately 10% of pediatric AML and includes substantial heterogeneity at the molecular level. Molecular subtypes of AMKL are listed below.
    • CBFA2T3-GLIS2: CBFA2T3-GLIS2 is a fusion resulting from a cryptic chromosome 16 inversion (inv(16)(p13.3q24.3)).[123-127] It occurs almost exclusively in non–Down syndrome AMKL, representing 16% to 27% of pediatric AMKL and presenting with a median age of 1 year.[71,125,128,129] It appears to be associated with unfavorable outcome,[71,123,127-129] with EFS at 2 years less than 20% in two reports that included 28 patients.[71,127,129]
    • KMT2A-rearranged: Cases with KMT2A translocations represent 10% to 17% of pediatric AMKL, with MLLT3 (AF9) being the most common KMT2A fusion partner.[71,107,128KMT2A-rearranged cases appear to be associated with inferior outcome among children with AMKL, with OS rates at 4 to 5 years of approximately 30%.[71,107,128] An international collaboration evaluating pediatric AMKL observed that the presence of t(9;11)/MLLT3-KMT2A, which was seen in approximately 5% of AMKL cases (n = 21), was associated with an inferior outcome (5-year OS, approximately 20%) compared with other AMKL cases and other KMT2A-rearrangements (n = 17), each with a 5-year OS of 50% to 55%.[107] Inferior outcome was not observed for patients (n = 17) with other KMT2A-rearrangements.
    • NUP98-KDM5A4: NUP98-KDM5A4 is observed in approximately 10% of pediatric AMKL cases [71,128] and is observed at much lower rates in non-AMKL cases.[129NUP98-KDM5A4 cases showed a trend towards inferior prognosis, although the small number of cases studied limits confidence in this assessment.[71,128]
    • RBM15-MKL1: The t(1;22)(p13;q13) translocation that produces RBM15-MKL1 is uncommon (<1% of pediatric AML) and is restricted to acute megakaryocytic leukemia (AMKL).[21,129-134] Studies have found that t(1;22)(p13;q13) is observed in 10% to 18% of children with AMKL who have evaluable cytogenetics or molecular genetics.[71,107,128] Most AMKL cases with t(1;22) occur in infants, with the median age at presentation (4–7 months) being younger than that for other children with AMKL.[107,125,135] Cases with detectable RBM15-MKL1 fusion transcripts in the absence of t(1;22) have also been reported because these young patients usually have hypoplastic bone marrow.[132]
      An international collaborative retrospective study of 51 t(1;22) cases reported that patients with this abnormality had a 5-year EFS of 54.5% and an OS of 58.2%, similar to the rates for other children with AMKL.[107] In another international retrospective analysis of 153 cases with non–Down syndrome AMKL who had samples available for molecular analysis, the 4-year EFS for patients with t(1;22) was 59% and OS was 70%, significantly better than AMKL patients with other specific genetic abnormalities (CBFA2T3/GUS2NUP98/KDM5A4KMT2A rearrangements, monosomy 7).[128]
    • HOX-rearranged: Cases with a gene fusion involving a HOX cluster gene represented 15% of pediatric AMKL in one report.[71] This report observed that these patients appear to have a relatively favorable prognosis, although the small number of cases studied limits confidence in this assessment.
    • GATA1 mutated: GATA1-truncating mutations in non–Down syndrome AMKL arise in young children (median age, 1–2 years) and are associated with amplification of the Down syndrome critical region on chromosome 21.[71] These patients represented approximately 10% of non–Down syndrome AMKL and appeared to have a favorable outcome if there were no prognostically unfavorable fusion genes also present, although the number of patients studied was small (n = 8).[71]
  • t(8;16) (MYST3-CREBBP): The t(8;16) translocation fuses the MYST3 gene on chromosome 8p11 to CREBBP on chromosome 16p13. t(8;16) AML rarely occurs in children. In an international Berlin-Frankfurt-Münster (BFM) AML study of 62 children, presence of this translocation was associated with younger age at diagnosis (median, 1.2 years), FAB M4/M5 phenotype, erythrophagocytosis, leukemia cutis, and disseminated intravascular coagulation.[136] Outcome for children with t(8;16) AML appears similar to other types of AML.
    A substantial proportion of infants diagnosed with t(8;16) AML in the first month of life show spontaneous remission, although AML recurrence may occur months to years later.[136-142] These observations suggest that a watch and wait policy could be considered in cases of t(8;16) AML diagnosed in the neonatal period if close long-term monitoring can be ensured.[136]
  • t(7;12)(q36;p13): The t(7;12)(q36;p13) translocation involves ETV6 on chromosome 12p13 and variable breakpoints on chromosome 7q36 in the region of MNX1 (HLXB9).[143] The translocation may be cryptic by conventional karyotyping and in some cases may be confirmed only by FISH.[144-146] This alteration occurs virtually exclusively in children younger than 2 years, is mutually exclusive with the KMT2A (MLL) rearrangement, and is associated with a high risk of treatment failure.[21,22,58,144,145,147]
  • NUP98 gene fusions: NUP98 has been reported to form leukemogenic gene fusions with more than 20 different partners.[148] In the pediatric AML setting, the two most common fusion genes are NUP98-NSD1 and NUP98-KDM5A4 (JARID1A), with the former observed in one report in approximately 15% of cytogenetically normal pediatric AML and the latter observed in approximately 10% of pediatric AMKL (see above).[71,93,125] AML cases with either NUP98 fusion gene show high expression of HOXA and HOXB genes, indicative of a stem cell phenotype.[118,125]
    The NUP98-NSD1 fusion gene, which is often cytogenetically cryptic, results from the fusion of NUP98 (chromosome 11p15) with NSD1 (chromosome 5q35).[93,94,118,149-152] This alteration occurs in approximately 4% to 7% of pediatric AML cases.[14,28,93,118,151] The highest frequency in the pediatric population is in the 5- to 9-year age group (approximately 8%), with lower frequency in younger children (approximately 2% in children younger than 2 years). NUP98-NSD1 cases present with high WBC count (median, 147 × 109/L in one study).[93,94] Most NUP98-NSD1 AML cases do not show cytogenetic aberrations.[93,118,149] A high percentage of NUP98-NSD1 cases (74% to 90%) have FLT3-ITD.[28,93,94]
    A study that included 12 children with NUP98-NSD1 AML reported that although all patients achieved CR, presence of NUP98-NSD1 independently predicted poor prognosis, and children with NUP98-NSD1 AML had a high risk of relapse, with a resulting 4-year EFS of approximately 10%.[93] In another study that included children (n = 38) and adults (n = 7) with NUP98-NSD1 AML, presence of both NUP98-NSD1 and FLT3-ITD independently predicted poor prognosis; patients with both lesions had a low CR rate (approximately 30%) and a low 3-year EFS rate (approximately 15%).[94]
  • RUNX1 mutations: AML with mutated RUNX1, which is a provisional entity in the 2016 WHO classification of AML and related neoplasms, is more common in adults than in children. In adults, the RUNX1 mutation is associated with a high risk of treatment failure. In a study of children with AML, RUNX1 mutations were observed in 11 of 503 patients (approximately 2%). Six of 11 patients with RUNX1-mutated AML failed to achieve remission and their 5-year EFS was 9%, suggesting that the RUNX1 mutation confers a poor prognosis in both children and adults.[153]
  • RAS mutations: Although mutations in RAS have been identified in 20% to 25% of patients with AML, the prognostic significance of these mutations has not been clearly shown.[58,154-156] Mutations in NRAS are observed more commonly than mutations in KRAS in pediatric AML cases.[58,157RAS mutations occur with similar frequency for all Type II alteration subtypes, with the exception of APL, for which RAS mutations are seldom observed.[58]
  • KIT mutations: Mutations in KIT occur in approximately 5% of AML, but in 10% to 40% of AML with core-binding factor abnormalities.[58,157-159]
    The presence of activating KIT mutations in adults with this AML subtype appears to be associated with a poorer prognosis compared with core-binding factor AML without KIT mutations.[158,160,161] The prognostic significance of KIT mutations occurring in pediatric core-binding factor AML remains unclear,[162-165] although the largest pediatric study reported to date observed no prognostic significance for KIT mutations.[166]
  • WT1 mutations: WT1, a zinc-finger protein regulating gene transcription, is mutated in approximately 10% of cytogenetically normal cases of AML in adults.[167-170] The WT1 mutation has been shown in some,[167,168,170] but not all studies [169] to be an independent predictor of worse disease-free survival, EFS, and OS of adults.
    In children with AML, WT1 mutations are observed in approximately 10% of cases.[171,172] Cases with WT1 mutations are enriched among children with normal cytogenetics and FLT3-ITD, but are less common among children younger than 3 years.[171,172] AML cases with NUP98-NSD1 are enriched for both FLT3-ITD and WT1 mutations.[93] In univariate analyses, WT1 mutations are predictive of poorer outcome in pediatric patients, but the independent prognostic significance of WT1 mutation status is unclear because of its strong association with FLT3-ITD and its association with NUP98-NSD1.[93,171,172] The largest study of WT1 mutations in children with AML observed that children with WT1 mutations in the absence of FLT3-ITD had outcomes similar to that of children without WT1 mutations, while children with both WT1 mutation and FLT3-ITD had survival rates less than 20%.[171]
  • DNMT3A mutations: Mutations of the DNMT3A gene have been identified in approximately 20% of adult AML patients and are uncommon in patients with favorable cytogenetics but occur in one-third of adult patients with intermediate-risk cytogenetics.[173] Mutations in this gene are independently associated with poor outcome.[173-175DNMT3A mutations are virtually absent in children.[176]
  • IDH1 and IDH2 mutations: Mutations in IDH1 and IDH2, which code for isocitrate dehydrogenase, occur in approximately 20% of adults with AML,[177-181] and they are enriched in patients with NPM1 mutations.[178,179,182] The specific mutations that occur in IDH1 and IDH2 create a novel enzymatic activity that promotes conversion of alpha-ketoglutarate to 2-hydroxyglutarate.[183,184] This novel activity appears to induce a DNA hypermethylation phenotype similar to that observed in AML cases with loss of function mutations in TET2.[182]
    Mutations in IDH1 and IDH2 are rare in pediatric AML, occurring in 0% to 4% of cases.[176,185-189] There is no indication of a negative prognostic effect for IDH1 and IDH2 mutations in children with AML.[185]
  • CSF3R mutations: CSF3R is the gene encoding the granulocyte colony-stimulating factor (G-CSF) receptor, and activating mutations in CSF3R are observed in 2% to 3% of pediatric AML cases.[190] These mutations lead to enhanced signaling through the G-CSF receptor, and they are primarily observed in AML with either CEBPA mutations or with core-binding factor abnormalities (RUNX1-RUNX1T1 and CBFB-MYH11).[190] The clinical characteristics of and prognosis for patients with CSF3R mutations do not seem to be significantly different from those of patients without CSF3R mutations.
    Activating mutations in CSF3R are also observed in patients with severe congenital neutropenia. These mutations are not the cause of severe congenital neutropenia, but rather arise as somatic mutations and can represent an early step in the pathway to AML.[191] In one study of patients with severe congenital neutropenia, 34% of patients who had not developed a myeloid malignancy had CSF3R mutations detectable in peripheral blood neutrophils and mononuclear cells, while 78% of patients who had developed a myeloid malignancy showed CSF3R mutations.[191] A study of 31 patients with severe congenital neutropenia who developed AML or MDS observed CSF3R mutations in approximately 80%, and also observed a high frequency of RUNX1 mutations (approximately 60%), suggesting cooperation between CSF3R and RUNX1 mutations for leukemia development within the context of severe congenital neutropenia.[192]
References
  1. Kuerbitz SJ, Civin CI, Krischer JP, et al.: Expression of myeloid-associated and lymphoid-associated cell-surface antigens in acute myeloid leukemia of childhood: a Pediatric Oncology Group study. J Clin Oncol 10 (9): 1419-29, 1992. [PUBMED Abstract]
  2. Smith FO, Lampkin BC, Versteeg C, et al.: Expression of lymphoid-associated cell surface antigens by childhood acute myeloid leukemia cells lacks prognostic significance. Blood 79 (9): 2415-22, 1992. [PUBMED Abstract]
  3. Dinndorf PA, Andrews RG, Benjamin D, et al.: Expression of normal myeloid-associated antigens by acute leukemia cells. Blood 67 (4): 1048-53, 1986. [PUBMED Abstract]
  4. Zhou Y, Jorgensen JL, Wang SA, et al.: Usefulness of CD11a and CD18 in flow cytometric immunophenotypic analysis for diagnosis of acute promyelocytic leukemia. Am J Clin Pathol 138 (5): 744-50, 2012. [PUBMED Abstract]
  5. Paietta E, Goloubeva O, Neuberg D, et al.: A surrogate marker profile for PML/RAR alpha expressing acute promyelocytic leukemia and the association of immunophenotypic markers with morphologic and molecular subtypes. Cytometry B Clin Cytom 59B (1): 1-9, 2004. [PUBMED Abstract]
  6. Lin P, Hao S, Medeiros LJ, et al.: Expression of CD2 in acute promyelocytic leukemia correlates with short form of PML-RARalpha transcripts and poorer prognosis. Am J Clin Pathol 121 (3): 402-7, 2004. [PUBMED Abstract]
  7. Creutzig U, Ritter J, Schellong G: Identification of two risk groups in childhood acute myelogenous leukemia after therapy intensification in study AML-BFM-83 as compared with study AML-BFM-78. AML-BFM Study Group. Blood 75 (10): 1932-40, 1990. [PUBMED Abstract]
  8. Gerr H, Zimmermann M, Schrappe M, et al.: Acute leukaemias of ambiguous lineage in children: characterization, prognosis and therapy recommendations. Br J Haematol 149 (1): 84-92, 2010. [PUBMED Abstract]
  9. Rubnitz JE, Onciu M, Pounds S, et al.: Acute mixed lineage leukemia in children: the experience of St Jude Children's Research Hospital. Blood 113 (21): 5083-9, 2009. [PUBMED Abstract]
  10. Al-Seraihy AS, Owaidah TM, Ayas M, et al.: Clinical characteristics and outcome of children with biphenotypic acute leukemia. Haematologica 94 (12): 1682-90, 2009. [PUBMED Abstract]
  11. Bene MC, Castoldi G, Knapp W, et al.: Proposals for the immunological classification of acute leukemias. European Group for the Immunological Characterization of Leukemias (EGIL). Leukemia 9 (10): 1783-6, 1995. [PUBMED Abstract]
  12. Vardiman JW, Thiele J, Arber DA, et al.: The 2008 revision of the World Health Organization (WHO) classification of myeloid neoplasms and acute leukemia: rationale and important changes. Blood 114 (5): 937-51, 2009. [PUBMED Abstract]
  13. Borowitz MJ, Béné MC, Harris NL: Acute leukaemias of ambiguous lineage. In: Swerdlow SH, Campo E, Harris NL, et al., eds.: WHO Classification of Tumours of Haematopoietic and Lymphoid Tissues. 4th ed. Lyon, France: International Agency for Research on Cancer, 2008, pp 150-5.
  14. Arber DA, Orazi A, Hasserjian R, et al.: The 2016 revision to the World Health Organization classification of myeloid neoplasms and acute leukemia. Blood 127 (20): 2391-405, 2016. [PUBMED Abstract]
  15. Tarlock K, Meshinchi S: Pediatric acute myeloid leukemia: biology and therapeutic implications of genomic variants. Pediatr Clin North Am 62 (1): 75-93, 2015. [PUBMED Abstract]
  16. Bolouri H, Farrar JE, Triche T, et al.: The molecular landscape of pediatric acute myeloid leukemia reveals recurrent structural alterations and age-specific mutational interactions. Nat Med 24 (1): 103-112, 2018. [PUBMED Abstract]
  17. Creutzig U, van den Heuvel-Eibrink MM, Gibson B, et al.: Diagnosis and management of acute myeloid leukemia in children and adolescents: recommendations from an international expert panel. Blood 120 (16): 3187-205, 2012. [PUBMED Abstract]
  18. Grimwade D, Walker H, Oliver F, et al.: The importance of diagnostic cytogenetics on outcome in AML: analysis of 1,612 patients entered into the MRC AML 10 trial. The Medical Research Council Adult and Children's Leukaemia Working Parties. Blood 92 (7): 2322-33, 1998. [PUBMED Abstract]
  19. Gilliland DG: Targeted therapies in myeloid leukemias. Ann Hematol 83 (Suppl 1): S75-6, 2004. [PUBMED Abstract]
  20. Avivi I, Rowe JM: Prognostic factors in acute myeloid leukemia. Curr Opin Hematol 12 (1): 62-7, 2005. [PUBMED Abstract]
  21. Harrison CJ, Hills RK, Moorman AV, et al.: Cytogenetics of childhood acute myeloid leukemia: United Kingdom Medical Research Council Treatment trials AML 10 and 12. J Clin Oncol 28 (16): 2674-81, 2010. [PUBMED Abstract]
  22. von Neuhoff C, Reinhardt D, Sander A, et al.: Prognostic impact of specific chromosomal aberrations in a large group of pediatric patients with acute myeloid leukemia treated uniformly according to trial AML-BFM 98. J Clin Oncol 28 (16): 2682-9, 2010. [PUBMED Abstract]
  23. Grimwade D, Hills RK, Moorman AV, et al.: Refinement of cytogenetic classification in acute myeloid leukemia: determination of prognostic significance of rare recurring chromosomal abnormalities among 5876 younger adult patients treated in the United Kingdom Medical Research Council trials. Blood 116 (3): 354-65, 2010. [PUBMED Abstract]
  24. Brown P, McIntyre E, Rau R, et al.: The incidence and clinical significance of nucleophosmin mutations in childhood AML. Blood 110 (3): 979-85, 2007. [PUBMED Abstract]
  25. Hollink IH, Zwaan CM, Zimmermann M, et al.: Favorable prognostic impact of NPM1 gene mutations in childhood acute myeloid leukemia, with emphasis on cytogenetically normal AML. Leukemia 23 (2): 262-70, 2009. [PUBMED Abstract]
  26. Ho PA, Alonzo TA, Gerbing RB, et al.: Prevalence and prognostic implications of CEBPA mutations in pediatric acute myeloid leukemia (AML): a report from the Children's Oncology Group. Blood 113 (26): 6558-66, 2009. [PUBMED Abstract]
  27. Meshinchi S, Alonzo TA, Stirewalt DL, et al.: Clinical implications of FLT3 mutations in pediatric AML. Blood 108 (12): 3654-61, 2006. [PUBMED Abstract]
  28. Struski S, Lagarde S, Bories P, et al.: NUP98 is rearranged in 3.8% of pediatric AML forming a clinical and molecular homogenous group with a poor prognosis. Leukemia 31 (3): 565-572, 2017. [PUBMED Abstract]
  29. Farrar JE, Schuback HL, Ries RE, et al.: Genomic Profiling of Pediatric Acute Myeloid Leukemia Reveals a Changing Mutational Landscape from Disease Diagnosis to Relapse. Cancer Res 76 (8): 2197-205, 2016. [PUBMED Abstract]
  30. Rubnitz JE, Raimondi SC, Halbert AR, et al.: Characteristics and outcome of t(8;21)-positive childhood acute myeloid leukemia: a single institution's experience. Leukemia 16 (10): 2072-7, 2002. [PUBMED Abstract]
  31. Tallman MS, Hakimian D, Shaw JM, et al.: Granulocytic sarcoma is associated with the 8;21 translocation in acute myeloid leukemia. J Clin Oncol 11 (4): 690-7, 1993. [PUBMED Abstract]
  32. Mrózek K, Heerema NA, Bloomfield CD: Cytogenetics in acute leukemia. Blood Rev 18 (2): 115-36, 2004. [PUBMED Abstract]
  33. Creutzig U, Zimmermann M, Ritter J, et al.: Definition of a standard-risk group in children with AML. Br J Haematol 104 (3): 630-9, 1999. [PUBMED Abstract]
  34. Raimondi SC, Chang MN, Ravindranath Y, et al.: Chromosomal abnormalities in 478 children with acute myeloid leukemia: clinical characteristics and treatment outcome in a cooperative pediatric oncology group study-POG 8821. Blood 94 (11): 3707-16, 1999. [PUBMED Abstract]
  35. Lie SO, Abrahamsson J, Clausen N, et al.: Treatment stratification based on initial in vivo response in acute myeloid leukaemia in children without Down's syndrome: results of NOPHO-AML trials. Br J Haematol 122 (2): 217-25, 2003. [PUBMED Abstract]
  36. Klein K, Kaspers G, Harrison CJ, et al.: Clinical Impact of Additional Cytogenetic Aberrations, cKIT and RAS Mutations, and Treatment Elements in Pediatric t(8;21)-AML: Results From an International Retrospective Study by the International Berlin-Frankfurt-Münster Study Group. J Clin Oncol 33 (36): 4247-58, 2015. [PUBMED Abstract]
  37. Larson RA, Williams SF, Le Beau MM, et al.: Acute myelomonocytic leukemia with abnormal eosinophils and inv(16) or t(16;16) has a favorable prognosis. Blood 68 (6): 1242-9, 1986. [PUBMED Abstract]
  38. Duployez N, Marceau-Renaut A, Boissel N, et al.: Comprehensive mutational profiling of core binding factor acute myeloid leukemia. Blood 127 (20): 2451-9, 2016. [PUBMED Abstract]
  39. Faber ZJ, Chen X, Gedman AL, et al.: The genomic landscape of core-binding factor acute myeloid leukemias. Nat Genet 48 (12): 1551-1556, 2016. [PUBMED Abstract]
  40. Noort S, Zimmermann M, Reinhardt D, et al.: Prognostic impact of t(16;21)(p11;q22) and t(16;21)(q24;q22) in pediatric AML: a retrospective study by the I-BFM Study Group. Blood 132 (15): 1584-1592, 2018. [PUBMED Abstract]
  41. Jahn N, Agrawal M, Bullinger L, et al.: Incidence and prognostic impact of ASXL2 mutations in adult acute myeloid leukemia patients with t(8;21)(q22;q22): a study of the German-Austrian AML Study Group. Leukemia 31 (4): 1012-1015, 2017. [PUBMED Abstract]
  42. Yamato G, Shiba N, Yoshida K, et al.: ASXL2 mutations are frequently found in pediatric AML patients with t(8;21)/ RUNX1-RUNX1T1 and associated with a better prognosis. Genes Chromosomes Cancer 56 (5): 382-393, 2017. [PUBMED Abstract]
  43. Smith MA, Ries LA, Gurney JG, et al.: Leukemia. In: Ries LA, Smith MA, Gurney JG, et al., eds.: Cancer incidence and survival among children and adolescents: United States SEER Program 1975-1995. Bethesda, Md: National Cancer Institute, SEER Program, 1999. NIH Pub.No. 99-4649, pp 17-34. Also available online. Last accessed October 04, 2019.
  44. Mistry AR, Pedersen EW, Solomon E, et al.: The molecular pathogenesis of acute promyelocytic leukaemia: implications for the clinical management of the disease. Blood Rev 17 (2): 71-97, 2003. [PUBMED Abstract]
  45. Sanz MA, Grimwade D, Tallman MS, et al.: Management of acute promyelocytic leukemia: recommendations from an expert panel on behalf of the European LeukemiaNet. Blood 113 (9): 1875-91, 2009. [PUBMED Abstract]
  46. Grimwade D, Lo Coco F: Acute promyelocytic leukemia: a model for the role of molecular diagnosis and residual disease monitoring in directing treatment approach in acute myeloid leukemia. Leukemia 16 (10): 1959-73, 2002. [PUBMED Abstract]
  47. Licht JD, Chomienne C, Goy A, et al.: Clinical and molecular characterization of a rare syndrome of acute promyelocytic leukemia associated with translocation (11;17). Blood 85 (4): 1083-94, 1995. [PUBMED Abstract]
  48. Yan W, Zhang G: Molecular Characteristics and Clinical Significance of 12 Fusion Genes in Acute Promyelocytic Leukemia: A Systematic Review. Acta Haematol 136 (1): 1-15, 2016. [PUBMED Abstract]
  49. Grimwade D, Biondi A, Mozziconacci MJ, et al.: Characterization of acute promyelocytic leukemia cases lacking the classic t(15;17): results of the European Working Party. Groupe Français de Cytogénétique Hématologique, Groupe de Français d'Hematologie Cellulaire, UK Cancer Cytogenetics Group and BIOMED 1 European Community-Concerted Action "Molecular Cytogenetic Diagnosis in Haematological Malignancies". Blood 96 (4): 1297-308, 2000. [PUBMED Abstract]
  50. Falini B, Martelli MP, Bolli N, et al.: Immunohistochemistry predicts nucleophosmin (NPM) mutations in acute myeloid leukemia. Blood 108 (6): 1999-2005, 2006. [PUBMED Abstract]
  51. Falini B, Mecucci C, Tiacci E, et al.: Cytoplasmic nucleophosmin in acute myelogenous leukemia with a normal karyotype. N Engl J Med 352 (3): 254-66, 2005. [PUBMED Abstract]
  52. Döhner K, Schlenk RF, Habdank M, et al.: Mutant nucleophosmin (NPM1) predicts favorable prognosis in younger adults with acute myeloid leukemia and normal cytogenetics: interaction with other gene mutations. Blood 106 (12): 3740-6, 2005. [PUBMED Abstract]
  53. Verhaak RG, Goudswaard CS, van Putten W, et al.: Mutations in nucleophosmin (NPM1) in acute myeloid leukemia (AML): association with other gene abnormalities and previously established gene expression signatures and their favorable prognostic significance. Blood 106 (12): 3747-54, 2005. [PUBMED Abstract]
  54. Schnittger S, Schoch C, Kern W, et al.: Nucleophosmin gene mutations are predictors of favorable prognosis in acute myelogenous leukemia with a normal karyotype. Blood 106 (12): 3733-9, 2005. [PUBMED Abstract]
  55. Schlenk RF, Döhner K, Krauter J, et al.: Mutations and treatment outcome in cytogenetically normal acute myeloid leukemia. N Engl J Med 358 (18): 1909-18, 2008. [PUBMED Abstract]
  56. Gale RE, Green C, Allen C, et al.: The impact of FLT3 internal tandem duplication mutant level, number, size, and interaction with NPM1 mutations in a large cohort of young adult patients with acute myeloid leukemia. Blood 111 (5): 2776-84, 2008. [PUBMED Abstract]
  57. Cazzaniga G, Dell'Oro MG, Mecucci C, et al.: Nucleophosmin mutations in childhood acute myelogenous leukemia with normal karyotype. Blood 106 (4): 1419-22, 2005. [PUBMED Abstract]
  58. Balgobind BV, Hollink IH, Arentsen-Peters ST, et al.: Integrative analysis of type-I and type-II aberrations underscores the genetic heterogeneity of pediatric acute myeloid leukemia. Haematologica 96 (10): 1478-87, 2011. [PUBMED Abstract]
  59. Staffas A, Kanduri M, Hovland R, et al.: Presence of FLT3-ITD and high BAALC expression are independent prognostic markers in childhood acute myeloid leukemia. Blood 118 (22): 5905-13, 2011. [PUBMED Abstract]
  60. Tawana K, Wang J, Renneville A, et al.: Disease evolution and outcomes in familial AML with germline CEBPA mutations. Blood 126 (10): 1214-23, 2015. [PUBMED Abstract]
  61. Marcucci G, Maharry K, Radmacher MD, et al.: Prognostic significance of, and gene and microRNA expression signatures associated with, CEBPA mutations in cytogenetically normal acute myeloid leukemia with high-risk molecular features: a Cancer and Leukemia Group B Study. J Clin Oncol 26 (31): 5078-87, 2008. [PUBMED Abstract]
  62. Wouters BJ, Löwenberg B, Erpelinck-Verschueren CA, et al.: Double CEBPA mutations, but not single CEBPA mutations, define a subgroup of acute myeloid leukemia with a distinctive gene expression profile that is uniquely associated with a favorable outcome. Blood 113 (13): 3088-91, 2009. [PUBMED Abstract]
  63. Dufour A, Schneider F, Metzeler KH, et al.: Acute myeloid leukemia with biallelic CEBPA gene mutations and normal karyotype represents a distinct genetic entity associated with a favorable clinical outcome. J Clin Oncol 28 (4): 570-7, 2010. [PUBMED Abstract]
  64. Taskesen E, Bullinger L, Corbacioglu A, et al.: Prognostic impact, concurrent genetic mutations, and gene expression features of AML with CEBPA mutations in a cohort of 1182 cytogenetically normal AML patients: further evidence for CEBPA double mutant AML as a distinctive disease entity. Blood 117 (8): 2469-75, 2011. [PUBMED Abstract]
  65. Fasan A, Haferlach C, Alpermann T, et al.: The role of different genetic subtypes of CEBPA mutated AML. Leukemia 28 (4): 794-803, 2014. [PUBMED Abstract]
  66. Hollink IH, van den Heuvel-Eibrink MM, Arentsen-Peters ST, et al.: Characterization of CEBPA mutations and promoter hypermethylation in pediatric acute myeloid leukemia. Haematologica 96 (3): 384-92, 2011. [PUBMED Abstract]
  67. Groet J, McElwaine S, Spinelli M, et al.: Acquired mutations in GATA1 in neonates with Down's syndrome with transient myeloid disorder. Lancet 361 (9369): 1617-20, 2003. [PUBMED Abstract]
  68. Hitzler JK, Cheung J, Li Y, et al.: GATA1 mutations in transient leukemia and acute megakaryoblastic leukemia of Down syndrome. Blood 101 (11): 4301-4, 2003. [PUBMED Abstract]
  69. 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 102 (3): 981-6, 2003. [PUBMED Abstract]
  70. Wechsler J, Greene M, McDevitt MA, et al.: Acquired mutations in GATA1 in the megakaryoblastic leukemia of Down syndrome. Nat Genet 32 (1): 148-52, 2002. [PUBMED Abstract]
  71. de Rooij JD, Branstetter C, Ma J, et al.: Pediatric non-Down syndrome acute megakaryoblastic leukemia is characterized by distinct genomic subsets with varying outcomes. Nat Genet 49 (3): 451-456, 2017. [PUBMED Abstract]
  72. Gurbuxani S, Vyas P, Crispino JD: Recent insights into the mechanisms of myeloid leukemogenesis in Down syndrome. Blood 103 (2): 399-406, 2004. [PUBMED Abstract]
  73. Ge Y, Stout ML, Tatman DA, et al.: GATA1, cytidine deaminase, and the high cure rate of Down syndrome children with acute megakaryocytic leukemia. J Natl Cancer Inst 97 (3): 226-31, 2005. [PUBMED Abstract]
  74. Johnston DL, Alonzo TA, Gerbing RB, et al.: Outcome of pediatric patients with acute myeloid leukemia (AML) and -5/5q- abnormalities from five pediatric AML treatment protocols: a report from the Children's Oncology Group. Pediatr Blood Cancer 60 (12): 2073-8, 2013. [PUBMED Abstract]
  75. Stevens RF, Hann IM, Wheatley K, et al.: Marked improvements in outcome with chemotherapy alone in paediatric acute myeloid leukemia: results of the United Kingdom Medical Research Council's 10th AML trial. MRC Childhood Leukaemia Working Party. Br J Haematol 101 (1): 130-40, 1998. [PUBMED Abstract]
  76. Wells RJ, Arthur DC, Srivastava A, et al.: Prognostic variables in newly diagnosed children and adolescents with acute myeloid leukemia: Children's Cancer Group Study 213. Leukemia 16 (4): 601-7, 2002. [PUBMED Abstract]
  77. Hasle H, Alonzo TA, Auvrignon A, et al.: Monosomy 7 and deletion 7q in children and adolescents with acute myeloid leukemia: an international retrospective study. Blood 109 (11): 4641-7, 2007. [PUBMED Abstract]
  78. Rasche M, von Neuhoff C, Dworzak M, et al.: Genotype-outcome correlations in pediatric AML: the impact of a monosomal karyotype in trial AML-BFM 2004. Leukemia 31 (12): 2807-2814, 2017. [PUBMED Abstract]
  79. Swansbury GJ, Lawler SD, Alimena G, et al.: Long-term survival in acute myelogenous leukemia: a second follow-up of the Fourth International Workshop on Chromosomes in Leukemia. Cancer Genet Cytogenet 73 (1): 1-7, 1994. [PUBMED Abstract]
  80. Blink M, Zimmermann M, von Neuhoff C, et al.: Normal karyotype is a poor prognostic factor in myeloid leukemia of Down syndrome: a retrospective, international study. Haematologica 99 (2): 299-307, 2014. [PUBMED Abstract]
  81. Gröschel S, Sanders MA, Hoogenboezem R, et al.: A single oncogenic enhancer rearrangement causes concomitant EVI1 and GATA2 deregulation in leukemia. Cell 157 (2): 369-81, 2014. [PUBMED Abstract]
  82. Yamazaki H, Suzuki M, Otsuki A, et al.: A remote GATA2 hematopoietic enhancer drives leukemogenesis in inv(3)(q21;q26) by activating EVI1 expression. Cancer Cell 25 (4): 415-27, 2014. [PUBMED Abstract]
  83. 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 28 (24): 3890-8, 2010. [PUBMED Abstract]
  84. Balgobind BV, Lugthart S, Hollink IH, et al.: EVI1 overexpression in distinct subtypes of pediatric acute myeloid leukemia. Leukemia 24 (5): 942-9, 2010. [PUBMED Abstract]
  85. Schnittger S, Schoch C, Dugas M, et al.: Analysis of FLT3 length mutations in 1003 patients with acute myeloid leukemia: correlation to cytogenetics, FAB subtype, and prognosis in the AMLCG study and usefulness as a marker for the detection of minimal residual disease. Blood 100 (1): 59-66, 2002. [PUBMED Abstract]
  86. Thiede C, Steudel C, Mohr B, et al.: Analysis of FLT3-activating mutations in 979 patients with acute myelogenous leukemia: association with FAB subtypes and identification of subgroups with poor prognosis. Blood 99 (12): 4326-35, 2002. [PUBMED Abstract]
  87. Whitman SP, Archer KJ, Feng L, et al.: Absence of the wild-type allele predicts poor prognosis in adult de novo acute myeloid leukemia with normal cytogenetics and the internal tandem duplication of FLT3: a cancer and leukemia group B study. Cancer Res 61 (19): 7233-9, 2001. [PUBMED Abstract]
  88. Iwai T, Yokota S, Nakao M, et al.: Internal tandem duplication of the FLT3 gene and clinical evaluation in childhood acute myeloid leukemia. The Children's Cancer and Leukemia Study Group, Japan. Leukemia 13 (1): 38-43, 1999. [PUBMED Abstract]
  89. Arrigoni P, Beretta C, Silvestri D, et al.: FLT3 internal tandem duplication in childhood acute myeloid leukaemia: association with hyperleucocytosis in acute promyelocytic leukaemia. Br J Haematol 120 (1): 89-92, 2003. [PUBMED Abstract]
  90. Meshinchi S, Stirewalt DL, Alonzo TA, et al.: Activating mutations of RTK/ras signal transduction pathway in pediatric acute myeloid leukemia. Blood 102 (4): 1474-9, 2003. [PUBMED Abstract]
  91. Zwaan CM, Meshinchi S, Radich JP, et al.: FLT3 internal tandem duplication in 234 children with acute myeloid leukemia: prognostic significance and relation to cellular drug resistance. Blood 102 (7): 2387-94, 2003. [PUBMED Abstract]
  92. Chang P, Kang M, Xiao A, et al.: FLT3 mutation incidence and timing of origin in a population case series of pediatric leukemia. BMC Cancer 10: 513, 2010. [PUBMED Abstract]
  93. Hollink IH, van den Heuvel-Eibrink MM, Arentsen-Peters ST, et al.: NUP98/NSD1 characterizes a novel poor prognostic group in acute myeloid leukemia with a distinct HOX gene expression pattern. Blood 118 (13): 3645-56, 2011. [PUBMED Abstract]
  94. Ostronoff F, Othus M, Gerbing RB, et al.: NUP98/NSD1 and FLT3/ITD coexpression is more prevalent in younger AML patients and leads to induction failure: a COG and SWOG report. Blood 124 (15): 2400-7, 2014. [PUBMED Abstract]
  95. Shih LY, Kuo MC, Liang DC, et al.: Internal tandem duplication and Asp835 mutations of the FMS-like tyrosine kinase 3 (FLT3) gene in acute promyelocytic leukemia. Cancer 98 (6): 1206-16, 2003. [PUBMED Abstract]
  96. Noguera NI, Breccia M, Divona M, et al.: Alterations of the FLT3 gene in acute promyelocytic leukemia: association with diagnostic characteristics and analysis of clinical outcome in patients treated with the Italian AIDA protocol. Leukemia 16 (11): 2185-9, 2002. [PUBMED Abstract]
  97. Gale RE, Hills R, Pizzey AR, et al.: Relationship between FLT3 mutation status, biologic characteristics, and response to targeted therapy in acute promyelocytic leukemia. Blood 106 (12): 3768-76, 2005. [PUBMED Abstract]
  98. Abu-Duhier FM, Goodeve AC, Wilson GA, et al.: Identification of novel FLT-3 Asp835 mutations in adult acute myeloid leukaemia. Br J Haematol 113 (4): 983-8, 2001. [PUBMED Abstract]
  99. Kutny MA, Moser BK, Laumann K, et al.: FLT3 mutation status is a predictor of early death in pediatric acute promyelocytic leukemia: a report from the Children's Oncology Group. Pediatr Blood Cancer 59 (4): 662-7, 2012. [PUBMED Abstract]
  100. Tallman MS, Kim HT, Montesinos P, et al.: Does microgranular variant morphology of acute promyelocytic leukemia independently predict a less favorable outcome compared with classical M3 APL? A joint study of the North American Intergroup and the PETHEMA Group. Blood 116 (25): 5650-9, 2010. [PUBMED Abstract]
  101. Sung L, Aplenc R, Alonzo TA, et al.: Predictors and short-term outcomes of hyperleukocytosis in children with acute myeloid leukemia: a report from the Children's Oncology Group. Haematologica 97 (11): 1770-3, 2012. [PUBMED Abstract]
  102. Callens C, Chevret S, Cayuela JM, et al.: Prognostic implication of FLT3 and Ras gene mutations in patients with acute promyelocytic leukemia (APL): a retrospective study from the European APL Group. Leukemia 19 (7): 1153-60, 2005. [PUBMED Abstract]
  103. Schnittger S, Bacher U, Haferlach C, et al.: Clinical impact of FLT3 mutation load in acute promyelocytic leukemia with t(15;17)/PML-RARA. Haematologica 96 (12): 1799-807, 2011. [PUBMED Abstract]
  104. Breccia M, Loglisci G, Loglisci MG, et al.: FLT3-ITD confers poor prognosis in patients with acute promyelocytic leukemia treated with AIDA protocols: long-term follow-up analysis. Haematologica 98 (12): e161-3, 2013. [PUBMED Abstract]
  105. Poiré X, Moser BK, Gallagher RE, et al.: Arsenic trioxide in front-line therapy of acute promyelocytic leukemia (C9710): prognostic significance of FLT3 mutations and complex karyotype. Leuk Lymphoma 55 (7): 1523-32, 2014. [PUBMED Abstract]
  106. Pui CH, Relling MV, Rivera GK, et al.: Epipodophyllotoxin-related acute myeloid leukemia: a study of 35 cases. Leukemia 9 (12): 1990-6, 1995. [PUBMED Abstract]
  107. Inaba H, Zhou Y, Abla O, et al.: Heterogeneous cytogenetic subgroups and outcomes in childhood acute megakaryoblastic leukemia: a retrospective international study. Blood 126 (13): 1575-84, 2015. [PUBMED Abstract]
  108. Balgobind BV, Raimondi SC, Harbott J, et al.: Novel prognostic subgroups in childhood 11q23/MLL-rearranged acute myeloid leukemia: results of an international retrospective study. Blood 114 (12): 2489-96, 2009. [PUBMED Abstract]
  109. Swansbury GJ, Slater R, Bain BJ, et al.: Hematological malignancies with t(9;11)(p21-22;q23)--a laboratory and clinical study of 125 cases. European 11q23 Workshop participants. Leukemia 12 (5): 792-800, 1998. [PUBMED Abstract]
  110. Rubnitz JE, Raimondi SC, Tong X, et al.: Favorable impact of the t(9;11) in childhood acute myeloid leukemia. J Clin Oncol 20 (9): 2302-9, 2002. [PUBMED Abstract]
  111. Mrózek K, Heinonen K, Lawrence D, et al.: Adult patients with de novo acute myeloid leukemia and t(9; 11)(p22; q23) have a superior outcome to patients with other translocations involving band 11q23: a Cancer and Leukemia Group B study. Blood 90 (11): 4532-8, 1997. [PUBMED Abstract]
  112. Martinez-Climent JA, Espinosa R, Thirman MJ, et al.: Abnormalities of chromosome band 11q23 and the MLL gene in pediatric myelomonocytic and monoblastic leukemias. Identification of the t(9;11) as an indicator of long survival. J Pediatr Hematol Oncol 17 (4): 277-83, 1995. [PUBMED Abstract]
  113. Casillas JN, Woods WG, Hunger SP, et al.: Prognostic implications of t(10;11) translocations in childhood acute myelogenous leukemia: a report from the Children's Cancer Group. J Pediatr Hematol Oncol 25 (8): 594-600, 2003. [PUBMED Abstract]
  114. Morerio C, Rosanda C, Rapella A, et al.: Is t(10;11)(p11.2;q23) involving MLL and ABI-1 genes associated with congenital acute monocytic leukemia? Cancer Genet Cytogenet 139 (1): 57-9, 2002. [PUBMED Abstract]
  115. Taki T, Shibuya N, Taniwaki M, et al.: ABI-1, a human homolog to mouse Abl-interactor 1, fuses the MLL gene in acute myeloid leukemia with t(10;11)(p11.2;q23). Blood 92 (4): 1125-30, 1998. [PUBMED Abstract]
  116. Coenen EA, Raimondi SC, Harbott J, et al.: Prognostic significance of additional cytogenetic aberrations in 733 de novo pediatric 11q23/MLL-rearranged AML patients: results of an international study. Blood 117 (26): 7102-11, 2011. [PUBMED Abstract]
  117. Ageberg M, Drott K, Olofsson T, et al.: Identification of a novel and myeloid specific role of the leukemia-associated fusion protein DEK-NUP214 leading to increased protein synthesis. Genes Chromosomes Cancer 47 (4): 276-87, 2008. [PUBMED Abstract]
  118. Shiba N, Ichikawa H, Taki T, et al.: NUP98-NSD1 gene fusion and its related gene expression signature are strongly associated with a poor prognosis in pediatric acute myeloid leukemia. Genes Chromosomes Cancer 52 (7): 683-93, 2013. [PUBMED Abstract]
  119. Slovak ML, Gundacker H, Bloomfield CD, et al.: A retrospective study of 69 patients with t(6;9)(p23;q34) AML emphasizes the need for a prospective, multicenter initiative for rare 'poor prognosis' myeloid malignancies. Leukemia 20 (7): 1295-7, 2006. [PUBMED Abstract]
  120. Alsabeh R, Brynes RK, Slovak ML, et al.: Acute myeloid leukemia with t(6;9) (p23;q34): association with myelodysplasia, basophilia, and initial CD34 negative immunophenotype. Am J Clin Pathol 107 (4): 430-7, 1997. [PUBMED Abstract]
  121. Sandahl JD, Coenen EA, Forestier E, et al.: t(6;9)(p22;q34)/DEK-NUP214-rearranged pediatric myeloid leukemia: an international study of 62 patients. Haematologica 99 (5): 865-72, 2014. [PUBMED Abstract]
  122. Tarlock K, Alonzo TA, Moraleda PP, et al.: Acute myeloid leukaemia (AML) with t(6;9)(p23;q34) is associated with poor outcome in childhood AML regardless of FLT3-ITD status: a report from the Children's Oncology Group. Br J Haematol 166 (2): 254-9, 2014. [PUBMED Abstract]
  123. 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 22 (5): 683-97, 2012. [PUBMED Abstract]
  124. Thiollier C, Lopez CK, Gerby B, et al.: Characterization of novel genomic alterations and therapeutic approaches using acute megakaryoblastic leukemia xenograft models. J Exp Med 209 (11): 2017-31, 2012. [PUBMED Abstract]
  125. 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 27 (12): 2280-8, 2013. [PUBMED Abstract]
  126. 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 121 (17): 3469-72, 2013. [PUBMED Abstract]
  127. Masetti R, Rondelli R, Fagioli F, et al.: Infants with acute myeloid leukemia treated according to the Associazione Italiana di Ematologia e Oncologia Pediatrica 2002/01 protocol have an outcome comparable to that of older children. Haematologica 99 (8): e127-9, 2014. [PUBMED Abstract]
  128. de Rooij JD, Masetti R, van den Heuvel-Eibrink MM, et al.: Recurrent abnormalities can be used for risk group stratification in pediatric AMKL: a retrospective intergroup study. Blood 127 (26): 3424-30, 2016. [PUBMED Abstract]
  129. Hara Y, Shiba N, Ohki K, et al.: Prognostic impact of specific molecular profiles in pediatric acute megakaryoblastic leukemia in non-Down syndrome. Genes Chromosomes Cancer 56 (5): 394-404, 2017. [PUBMED Abstract]
  130. 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 78 (3): 748-52, 1991. [PUBMED Abstract]
  131. Lion T, Haas OA: Acute megakaryocytic leukemia with the t(1;22)(p13;q13). Leuk Lymphoma 11 (1-2): 15-20, 1993. [PUBMED Abstract]
  132. Duchayne E, Fenneteau O, Pages MP, et al.: Acute megakaryoblastic leukaemia: a national clinical and biological study of 53 adult and childhood cases by the Groupe Français d'Hématologie Cellulaire (GFHC). Leuk Lymphoma 44 (1): 49-58, 2003. [PUBMED Abstract]
  133. 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 28 (3): 220-1, 2001. [PUBMED Abstract]
  134. 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 U S A 98 (10): 5776-9, 2001. [PUBMED Abstract]
  135. 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 14 (1): 216-8, 2000. [PUBMED Abstract]
  136. Coenen EA, Zwaan CM, Reinhardt D, et al.: Pediatric acute myeloid leukemia with t(8;16)(p11;p13), a distinct clinical and biological entity: a collaborative study by the International-Berlin-Frankfurt-Munster AML-study group. Blood 122 (15): 2704-13, 2013. [PUBMED Abstract]
  137. Wong KF, Yuen HL, Siu LL, et al.: t(8;16)(p11;p13) predisposes to a transient but potentially recurring neonatal leukemia. Hum Pathol 39 (11): 1702-7, 2008. [PUBMED Abstract]
  138. Wu X, Sulavik D, Roulston D, et al.: Spontaneous remission of congenital acute myeloid leukemia with t(8;16)(p11;13). Pediatr Blood Cancer 56 (2): 331-2, 2011. [PUBMED Abstract]
  139. Terui K, Sato T, Sasaki S, et al.: Two novel variants of MOZ-CBP fusion transcripts in spontaneously remitted infant leukemia with t(1;16;8)(p13;p13;p11), a new variant of t(8;16)(p11;p13). Haematologica 93 (10): 1591-3, 2008. [PUBMED Abstract]
  140. Sainati L, Bolcato S, Cocito MG, et al.: Transient acute monoblastic leukemia with reciprocal (8;16)(p11;p13) translocation. Pediatr Hematol Oncol 13 (2): 151-7, 1996 Mar-Apr. [PUBMED Abstract]
  141. Weintraub M, Kaplinsky C, Amariglio N, et al.: Spontaneous regression of congenital leukaemia with an 8;16 translocation. Br J Haematol 111 (2): 641-3, 2000. [PUBMED Abstract]
  142. Classen CF, Behnisch W, Reinhardt D, et al.: Spontaneous complete and sustained remission of a rearrangement CBP (16p13)-positive disseminated congenital myelosarcoma. Ann Hematol 84 (4): 274-5, 2005. [PUBMED Abstract]
  143. Beverloo HB, Panagopoulos I, Isaksson M, et al.: Fusion of the homeobox gene HLXB9 and the ETV6 gene in infant acute myeloid leukemias with the t(7;12)(q36;p13). Cancer Res 61 (14): 5374-7, 2001. [PUBMED Abstract]
  144. Slater RM, von Drunen E, Kroes WG, et al.: t(7;12)(q36;p13) and t(7;12)(q32;p13)--translocations involving ETV6 in children 18 months of age or younger with myeloid disorders. Leukemia 15 (6): 915-20, 2001. [PUBMED Abstract]
  145. von Bergh AR, van Drunen E, van Wering ER, et al.: High incidence of t(7;12)(q36;p13) in infant AML but not in infant ALL, with a dismal outcome and ectopic expression of HLXB9. Genes Chromosomes Cancer 45 (8): 731-9, 2006. [PUBMED Abstract]
  146. Tosi S, Harbott J, Teigler-Schlegel A, et al.: t(7;12)(q36;p13), a new recurrent translocation involving ETV6 in infant leukemia. Genes Chromosomes Cancer 29 (4): 325-32, 2000. [PUBMED Abstract]
  147. Park J, Kim M, Lim J, et al.: Three-way complex translocations in infant acute myeloid leukemia with t(7;12)(q36;p13): the incidence and correlation of a HLXB9 overexpression. Cancer Genet Cytogenet 191 (2): 102-5, 2009. [PUBMED Abstract]
  148. Takeda A, Yaseen NR: Nucleoporins and nucleocytoplasmic transport in hematologic malignancies. Semin Cancer Biol 27: 3-10, 2014. [PUBMED Abstract]
  149. Brown J, Jawad M, Twigg SR, et al.: A cryptic t(5;11)(q35;p15.5) in 2 children with acute myeloid leukemia with apparently normal karyotypes, identified by a multiplex fluorescence in situ hybridization telomere assay. Blood 99 (7): 2526-31, 2002. [PUBMED Abstract]
  150. Panarello C, Rosanda C, Morerio C: Cryptic translocation t(5;11)(q35;p15.5) with involvement of the NSD1 and NUP98 genes without 5q deletion in childhood acute myeloid leukemia. Genes Chromosomes Cancer 35 (3): 277-81, 2002. [PUBMED Abstract]
  151. Cerveira N, Correia C, Dória S, et al.: Frequency of NUP98-NSD1 fusion transcript in childhood acute myeloid leukaemia. Leukemia 17 (11): 2244-7, 2003. [PUBMED Abstract]
  152. Jaju RJ, Fidler C, Haas OA, et al.: A novel gene, NSD1, is fused to NUP98 in the t(5;11)(q35;p15.5) in de novo childhood acute myeloid leukemia. Blood 98 (4): 1264-7, 2001. [PUBMED Abstract]
  153. Yamato G, Shiba N, Yoshida K, et al.: RUNX1 mutations in pediatric acute myeloid leukemia are associated with distinct genetic features and an inferior prognosis. Blood 131 (20): 2266-2270, 2018. [PUBMED Abstract]
  154. Radich JP, Kopecky KJ, Willman CL, et al.: N-ras mutations in adult de novo acute myelogenous leukemia: prevalence and clinical significance. Blood 76 (4): 801-7, 1990. [PUBMED Abstract]
  155. Farr C, Gill R, Katz F, et al.: Analysis of ras gene mutations in childhood myeloid leukaemia. Br J Haematol 77 (3): 323-7, 1991. [PUBMED Abstract]
  156. Berman JN, Gerbing RB, Alonzo TA, et al.: Prevalence and clinical implications of NRAS mutations in childhood AML: a report from the Children's Oncology Group. Leukemia 25 (6): 1039-42, 2011. [PUBMED Abstract]
  157. Kühn MW, Radtke I, Bullinger L, et al.: High-resolution genomic profiling of adult and pediatric core-binding factor acute myeloid leukemia reveals new recurrent genomic alterations. Blood 119 (10): e67-75, 2012. [PUBMED Abstract]
  158. Schnittger S, Kohl TM, Haferlach T, et al.: KIT-D816 mutations in AML1-ETO-positive AML are associated with impaired event-free and overall survival. Blood 107 (5): 1791-9, 2006. [PUBMED Abstract]
  159. Tokumasu M, Murata C, Shimada A, et al.: Adverse prognostic impact of KIT mutations in childhood CBF-AML: the results of the Japanese Pediatric Leukemia/Lymphoma Study Group AML-05 trial. Leukemia 29 (12): 2438-41, 2015. [PUBMED Abstract]
  160. Cairoli R, Beghini A, Grillo G, et al.: Prognostic impact of c-KIT mutations in core binding factor leukemias: an Italian retrospective study. Blood 107 (9): 3463-8, 2006. [PUBMED Abstract]
  161. Paschka P, Marcucci G, Ruppert AS, et al.: Adverse prognostic significance of KIT mutations in adult acute myeloid leukemia with inv(16) and t(8;21): a Cancer and Leukemia Group B Study. J Clin Oncol 24 (24): 3904-11, 2006. [PUBMED Abstract]
  162. Shimada A, Taki T, Tabuchi K, et al.: KIT mutations, and not FLT3 internal tandem duplication, are strongly associated with a poor prognosis in pediatric acute myeloid leukemia with t(8;21): a study of the Japanese Childhood AML Cooperative Study Group. Blood 107 (5): 1806-9, 2006. [PUBMED Abstract]
  163. Shih LY, Liang DC, Huang CF, et al.: Cooperating mutations of receptor tyrosine kinases and Ras genes in childhood core-binding factor acute myeloid leukemia and a comparative analysis on paired diagnosis and relapse samples. Leukemia 22 (2): 303-7, 2008. [PUBMED Abstract]
  164. Goemans BF, Zwaan CM, Miller M, et al.: Mutations in KIT and RAS are frequent events in pediatric core-binding factor acute myeloid leukemia. Leukemia 19 (9): 1536-42, 2005. [PUBMED Abstract]
  165. Boissel N, Leroy H, Brethon B, et al.: Incidence and prognostic impact of c-Kit, FLT3, and Ras gene mutations in core binding factor acute myeloid leukemia (CBF-AML). Leukemia 20 (6): 965-70, 2006. [PUBMED Abstract]
  166. Pollard JA, Alonzo TA, Gerbing RB, et al.: Prevalence and prognostic significance of KIT mutations in pediatric patients with core binding factor AML enrolled on serial pediatric cooperative trials for de novo AML. Blood 115 (12): 2372-9, 2010. [PUBMED Abstract]
  167. Paschka P, Marcucci G, Ruppert AS, et al.: Wilms' tumor 1 gene mutations independently predict poor outcome in adults with cytogenetically normal acute myeloid leukemia: a cancer and leukemia group B study. J Clin Oncol 26 (28): 4595-602, 2008. [PUBMED Abstract]
  168. Virappane P, Gale R, Hills R, et al.: Mutation of the Wilms' tumor 1 gene is a poor prognostic factor associated with chemotherapy resistance in normal karyotype acute myeloid leukemia: the United Kingdom Medical Research Council Adult Leukaemia Working Party. J Clin Oncol 26 (33): 5429-35, 2008. [PUBMED Abstract]
  169. Gaidzik VI, Schlenk RF, Moschny S, et al.: Prognostic impact of WT1 mutations in cytogenetically normal acute myeloid leukemia: a study of the German-Austrian AML Study Group. Blood 113 (19): 4505-11, 2009. [PUBMED Abstract]
  170. Renneville A, Boissel N, Zurawski V, et al.: Wilms tumor 1 gene mutations are associated with a higher risk of recurrence in young adults with acute myeloid leukemia: a study from the Acute Leukemia French Association. Cancer 115 (16): 3719-27, 2009. [PUBMED Abstract]
  171. Ho PA, Zeng R, Alonzo TA, et al.: Prevalence and prognostic implications of WT1 mutations in pediatric acute myeloid leukemia (AML): a report from the Children's Oncology Group. Blood 116 (5): 702-10, 2010. [PUBMED Abstract]
  172. Hollink IH, van den Heuvel-Eibrink MM, Zimmermann M, et al.: Clinical relevance of Wilms tumor 1 gene mutations in childhood acute myeloid leukemia. Blood 113 (23): 5951-60, 2009. [PUBMED Abstract]
  173. Ley TJ, Ding L, Walter MJ, et al.: DNMT3A mutations in acute myeloid leukemia. N Engl J Med 363 (25): 2424-33, 2010. [PUBMED Abstract]
  174. Yan XJ, Xu J, Gu ZH, et al.: Exome sequencing identifies somatic mutations of DNA methyltransferase gene DNMT3A in acute monocytic leukemia. Nat Genet 43 (4): 309-15, 2011. [PUBMED Abstract]
  175. Thol F, Damm F, Lüdeking A, et al.: Incidence and prognostic influence of DNMT3A mutations in acute myeloid leukemia. J Clin Oncol 29 (21): 2889-96, 2011. [PUBMED Abstract]
  176. Ho PA, Kutny MA, Alonzo TA, et al.: Leukemic mutations in the methylation-associated genes DNMT3A and IDH2 are rare events in pediatric AML: a report from the Children's Oncology Group. Pediatr Blood Cancer 57 (2): 204-9, 2011. [PUBMED Abstract]
  177. Green CL, Evans CM, Hills RK, et al.: The prognostic significance of IDH1 mutations in younger adult patients with acute myeloid leukemia is dependent on FLT3/ITD status. Blood 116 (15): 2779-82, 2010. [PUBMED Abstract]
  178. Paschka P, Schlenk RF, Gaidzik VI, et al.: IDH1 and IDH2 mutations are frequent genetic alterations in acute myeloid leukemia and confer adverse prognosis in cytogenetically normal acute myeloid leukemia with NPM1 mutation without FLT3 internal tandem duplication. J Clin Oncol 28 (22): 3636-43, 2010. [PUBMED Abstract]
  179. Abbas S, Lugthart S, Kavelaars FG, et al.: Acquired mutations in the genes encoding IDH1 and IDH2 both are recurrent aberrations in acute myeloid leukemia: prevalence and prognostic value. Blood 116 (12): 2122-6, 2010. [PUBMED Abstract]
  180. Marcucci G, Maharry K, Wu YZ, et al.: IDH1 and IDH2 gene mutations identify novel molecular subsets within de novo cytogenetically normal acute myeloid leukemia: a Cancer and Leukemia Group B study. J Clin Oncol 28 (14): 2348-55, 2010. [PUBMED Abstract]
  181. Wagner K, Damm F, Göhring G, et al.: Impact of IDH1 R132 mutations and an IDH1 single nucleotide polymorphism in cytogenetically normal acute myeloid leukemia: SNP rs11554137 is an adverse prognostic factor. J Clin Oncol 28 (14): 2356-64, 2010. [PUBMED Abstract]
  182. Figueroa ME, Abdel-Wahab O, Lu C, et al.: Leukemic IDH1 and IDH2 mutations result in a hypermethylation phenotype, disrupt TET2 function, and impair hematopoietic differentiation. Cancer Cell 18 (6): 553-67, 2010. [PUBMED Abstract]
  183. Ward PS, Patel J, Wise DR, et al.: The common feature of leukemia-associated IDH1 and IDH2 mutations is a neomorphic enzyme activity converting alpha-ketoglutarate to 2-hydroxyglutarate. Cancer Cell 17 (3): 225-34, 2010. [PUBMED Abstract]
  184. Dang L, White DW, Gross S, et al.: Cancer-associated IDH1 mutations produce 2-hydroxyglutarate. Nature 462 (7274): 739-44, 2009. [PUBMED Abstract]
  185. Damm F, Thol F, Hollink I, et al.: Prevalence and prognostic value of IDH1 and IDH2 mutations in childhood AML: a study of the AML-BFM and DCOG study groups. Leukemia 25 (11): 1704-10, 2011. [PUBMED Abstract]
  186. Oki K, Takita J, Hiwatari M, et al.: IDH1 and IDH2 mutations are rare in pediatric myeloid malignancies. Leukemia 25 (2): 382-4, 2011. [PUBMED Abstract]
  187. Pigazzi M, Ferrari G, Masetti R, et al.: Low prevalence of IDH1 gene mutation in childhood AML in Italy. Leukemia 25 (1): 173-4, 2011. [PUBMED Abstract]
  188. Ho PA, Alonzo TA, Kopecky KJ, et al.: Molecular alterations of the IDH1 gene in AML: a Children's Oncology Group and Southwest Oncology Group study. Leukemia 24 (5): 909-13, 2010. [PUBMED Abstract]
  189. Andersson AK, Miller DW, Lynch JA, et al.: IDH1 and IDH2 mutations in pediatric acute leukemia. Leukemia 25 (10): 1570-7, 2011. [PUBMED Abstract]
  190. Maxson JE, Ries RE, Wang YC, et al.: CSF3R mutations have a high degree of overlap with CEBPA mutations in pediatric AML. Blood 127 (24): 3094-8, 2016. [PUBMED Abstract]
  191. Germeshausen M, Kratz CP, Ballmaier M, et al.: RAS and CSF3R mutations in severe congenital neutropenia. Blood 114 (16): 3504-5, 2009. [PUBMED Abstract]
  192. Skokowa J, Steinemann D, Katsman-Kuipers JE, et al.: Cooperativity of RUNX1 and CSF3R mutations in severe congenital neutropenia: a unique pathway in myeloid leukemogenesis. Blood 123 (14): 2229-37, 2014. [PUBMED Abstract]

No hay comentarios:

Publicar un comentario