lunes, 23 de septiembre de 2019

Childhood Cancer Genomics (PDQ®) 2/9 –Health Professional Version - National Cancer Institute

Childhood Cancer Genomics (PDQ®)–Health Professional Version - National Cancer Institute

National Cancer Institute

Childhood Cancer Genomics (PDQ®)–Health Professional Version



Gene polymorphisms in drug metabolic pathways

A number of polymorphisms of genes involved in the metabolism of chemotherapeutic agents have been reported to have prognostic significance in childhood ALL.[144-146] For example, patients with mutant phenotypes of thiopurine methyltransferase (TPMT, a gene involved in the metabolism of thiopurines, such as mercaptopurine [6-MP]), appear to have more favorable outcomes,[147] although such patients may also be at higher risk of developing significant treatment-related toxicities, including myelosuppression and infection.[148,149] Patients with homozygosity for TPMT variants associated with low enzymatic activity tolerate only very low doses of mercaptopurine (approximately 10% of the standard dose) and are treated with reduced doses of mercaptopurine to avoid excessive toxicity. Patients who are heterozygous for this mutant enzyme gene generally tolerate mercaptopurine without serious toxicity, but they do require more frequent dose reductions for hematologic toxicity than do patients who are homozygous for the normal allele.[150,151]
Germline variants in nucleoside diphosphate–linked moiety X-type motif 15 (NUDT15) that reduce or abolish activity of this enzyme also lead to diminished tolerance to thiopurines.[150,152] The variants are most common in East Asians and Hispanics, and they are rare in Europeans and Africans. Patients homozygous for the risk variants tolerate only very low doses of mercaptopurine, while patients heterozygous for the risk alleles tolerate lower doses than do patients homozygous for the wild-type allele (approximately 25% dose reduction on average), but there is broad overlap in tolerated doses between the two groups.[150,153]
Gene polymorphisms may also affect the expression of proteins that play central roles in the cellular effects of anticancer drugs. As an example, patients who are homozygous for a polymorphism in the promoter region of CEP72 (a centrosomal protein involved in microtubule formation) are at increased risk of vincristine neurotoxicity.[154]
Genome-wide polymorphism analysis has identified specific single nucleotide polymorphisms associated with high end-induction MRD and risk of relapse. Polymorphisms of IL-15, as well as genes associated with the metabolism of etoposide and methotrexate, were significantly associated with treatment response in two large cohorts of ALL patients treated on SJCRH and COG protocols.[155] Polymorphic variants involving the reduced folate carrier and methotrexate metabolism have been linked to toxicity and outcome.[156,157] While these associations suggest that individual variations in drug metabolism can affect outcome, few studies have attempted to adjust for these variations; it is unknown whether individualized dose modification on the basis of these findings will improve outcome.
(Refer to the PDQ summary on Childhood Acute Lymphoblastic Leukemia Treatment for information about the treatment of childhood ALL.)

Acute Myeloid Leukemia (AML)

Comprehensive molecular profiling of pediatric and adult AML has shown that AML is a disease demonstrating both commonalities and differences across the age spectrum.[158,159]
  • Pediatric AML, in contrast to AML in adults, is typically a disease of recurring chromosomal alterations (refer to Table 1 for a list of common gene fusions).[158,160] 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.[159] 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).[159]
  • 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.[158,159]
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.[160-166] 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.[167-170]
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.[20] 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 1.[20] Table 1 also shows, in the bottom three rows, additional relatively common recurrent translocations observed in children with AML.[164,165,171]
Table 1. 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.[172] 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.[173,174] Adults with t(8;21) have a more favorable prognosis than do adults with other types of AML.[161,175] Children with t(8;21) have a more favorable outcome than do children with AML characterized by normal or complex karyotypes,[161,176-178] with 5-year overall survival (OS) of 74% to 90%.[164,165,179] The t(8;21) translocation occurs in approximately 12% of children with AML.[164,165,179]
    • 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.[180] Inv(16) confers a favorable prognosis for both adults and children with AML,[161,176-178] with a 5-year OS of about 85%.[164,165] Inv(16) occurs in 7% to 9% of children with AML.[164,165,179] 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).[181,182]
    • 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.[183] 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%.[183]
    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).[181,182] 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.[179]
    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.[181,182]
    • 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.[181,182]
    • 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.[184] Similar results, albeit with smaller numbers, were reported for children with RUNX1-RUNX1T1 AML and ASXL1 and ASXL2 mutations.[185]
  • Acute promyelocytic leukemia (APL) with PML-RARA: APL represents about 7% of children with AML.[165,186] 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.[187] 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.[20]
    Utilization of quantitative reverse transcriptase–polymerase chain reaction (RT-PCR) for PML-RARA transcripts has become standard practice.[188] 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.[189] 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).[190-192] Identification of cases with the t(11;17)(q23;q21) is important because of their decreased sensitivity to all-trans retinoic acid.[187,190]
  • 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.[193] 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,[194] and an improved prognosis in the absence of FLT3–internal tandem duplication (ITD) mutations in adults and younger adults.[194-199]
    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.[167,168,200,201NPM1 mutations are associated with a favorable prognosis in patients with AML characterized by a normal karyotype.[167,168,201] 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,[167,202] but other studies showed no impact of a FLT3-ITD mutation on the favorable prognosis associated with an NPM1 mutation.[159,168,201]
  • AML with biallelic mutations of CEBPA: Mutations in the CEBPA gene occur in a subset of children and adults with cytogenetically normal AML.[203] In adults younger than 60 years, approximately 15% of cytogenetically normal AML cases have mutations in CEBPA.[198] Outcomes for adults with AML with CEBPA mutations appear to be relatively favorable and similar to that of patients with core-binding factor leukemias.[198,204] Studies in adults with AML have demonstrated that CEBPA double-mutant, but not single-mutant, AML is independently associated with a favorable prognosis,[205-208] leading to the WHO 2016 revision that requires biallelic mutations for the disease definition.[20]
    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.[169,209] Although both double-mutant and single-mutant alleles of CEBPA were associated with a favorable prognosis in children with AML in one large study,[169] a second study observed inferior outcome for patients with single CEBPA mutations.[209] 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.[169] 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.[203]
  • 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).[210-213GATA1 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).[214GATA1 is a transcription factor that is required for normal development of erythroid cells, megakaryocytes, eosinophils, and mast cells.[215]
    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.[216]

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).[161,175,217] These cytogenetic subgroups represent approximately 2% and 4% of pediatric AML cases, respectively, and are also associated with poor prognosis in children.[164,175,217-221]
    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.[166] However, outcome for children with del(7q), but not monosomy 7, appears comparable to that of other children with AML.[165,220] The presence of del(7q) does not abrogate the prognostic significance of favorable cytogenetic characteristics (e.g., inv(16) and t(8;21)).[161,220,222]
    Chromosome 5 and 7 abnormalities appear to lack prognostic significance in AML patients with Down syndrome who are aged 4 years and younger.[223]
  • 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.[224,225] These abnormalities are associated with poor prognosis in adults with AML,[161,175,226] but are very uncommon in children (<1% of pediatric AML cases).[164,177,227]
    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,[228] particularly when both alleles are mutated or there is a high ratio of the mutant allele to the normal allele.[229,230FLT3-ITD mutations also convey a poor prognosis in children with AML.[170,202,231-234] 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).[233-235]
    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.[236,237] 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.[237] 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.[159] 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.[159]
    For APL, FLT3-ITD and point mutations occur in 30% to 40% of children and adults.[229,232,233,238-242] Presence of the FLT3-ITD mutation is strongly associated with the microgranular variant (M3v) of APL and with hyperleukocytosis.[232,240,243,244] 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.[238,239,242,243,245-248]
    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.[159]
  • 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.[183] 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%.[183]

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.[164,165] These cases, including most AMLs secondary to epipodophyllotoxin,[249] 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).[214,250]
    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.[251] 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.[251] 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).[251]
    Outcome for patients with de novo AML and KMT2A gene rearrangement is generally reported as being similar to that for other patients with AML.[161,164,251,252] 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.[251] 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.[161,164,251,253-255] 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.[250]
    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.[161,165,256] 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.[257,258] 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%.[251]
    • 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%.[251]
    • 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.[259]
  • 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.[260,261] This subgroup of AML has been associated with a poor prognosis in adults with AML,[260,262,263] 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.[264,265]
    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.[164,261,264,265]
  • 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)).[266-270] 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.[214,268,271,272] It appears to be associated with unfavorable outcome,[214,266,270-272] with EFS at 2 years less than 20% in two reports that included 28 patients.[214,270,272]
    • KMT2A-rearranged: Cases with KMT2A translocations represent 10% to 17% of pediatric AMKL, with MLLT3 (AF9) being the most common KMT2A fusion partner.[214,250,271KMT2A-rearranged cases appear to be associated with inferior outcome among children with AMKL, with OS rates at 4 to 5 years of approximately 30%.[214,250,271] 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%.[250] 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 [214,271] and is observed at much lower rates in non-AMKL cases.[272NUP98-KDM5A4 cases showed a trend towards inferior prognosis, although the small number of cases studied limits confidence in this assessment.[214,271]
    • 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).[164,272-277] 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.[214,250,271] 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.[250,268,278] 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.[275]
      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.[250] 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).[271]
    • HOX-rearranged: Cases with a gene fusion involving a HOX cluster gene represented 15% of pediatric AMKL in one report.[214] 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.[214] 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).[214]
  • 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.[279] 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.[279-285] 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.[279]
  • 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).[286] The translocation may be cryptic by conventional karyotyping and in some cases may be confirmed only by FISH.[287-289] 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.[164,165,201,287,288,290]
  • NUP98 gene fusions: NUP98 has been reported to form leukemogenic gene fusions with more than 20 different partners.[291] 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).[214,236,268] AML cases with either NUP98 fusion gene show high expression of HOXA and HOXB genes, indicative of a stem cell phenotype.[261,268]
    The NUP98-NSD1 fusion gene, which is often cytogenetically cryptic, results from the fusion of NUP98 (chromosome 11p15) with NSD1 (chromosome 5q35).[236,237,261,292-295] This alteration occurs in approximately 4% to 7% of pediatric AML cases.[20,171,236,261,294] 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).[236,237] Most NUP98-NSD1 AML cases do not show cytogenetic aberrations.[236,261,292] A high percentage of NUP98-NSD1 cases (74% to 90%) have FLT3-ITD.[171,236,237]
    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%.[236] 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%).[237]
  • 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.[296]
  • 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.[201,297-299] Mutations in NRAS are observed more commonly than mutations in KRAS in pediatric AML cases.[201,300RAS mutations occur with similar frequency for all Type II alteration subtypes, with the exception of APL, for which RAS mutations are seldom observed.[201]
  • KIT mutations: Mutations in KIT occur in approximately 5% of AML, but in 10% to 40% of AML with core-binding factor abnormalities.[201,300-302]
    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.[301,303,304] The prognostic significance of KIT mutations occurring in pediatric core-binding factor AML remains unclear,[305-308] although the largest pediatric study reported to date observed no prognostic significance for KIT mutations.[309]
  • WT1 mutations: WT1, a zinc-finger protein regulating gene transcription, is mutated in approximately 10% of cytogenetically normal cases of AML in adults.[310-313] The WT1 mutation has been shown in some,[310,311,313] but not all studies [312] 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.[314,315] Cases with WT1 mutations are enriched among children with normal cytogenetics and FLT3-ITD, but are less common among children younger than 3 years.[314,315] AML cases with NUP98-NSD1 are enriched for both FLT3-ITD and WT1 mutations.[236] 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.[236,314,315] 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%.[314]
  • 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.[316] Mutations in this gene are independently associated with poor outcome.[316-318DNMT3A mutations are virtually absent in children.[319]
  • IDH1 and IDH2 mutations: Mutations in IDH1 and IDH2, which code for isocitrate dehydrogenase, occur in approximately 20% of adults with AML,[320-324] and they are enriched in patients with NPM1 mutations.[321,322,325] The specific mutations that occur in IDH1 and IDH2 create a novel enzymatic activity that promotes conversion of alpha-ketoglutarate to 2-hydroxyglutarate.[326,327] This novel activity appears to induce a DNA hypermethylation phenotype similar to that observed in AML cases with loss of function mutations in TET2.[325]
    Mutations in IDH1 and IDH2 are rare in pediatric AML, occurring in 0% to 4% of cases.[319,328-332] There is no indication of a negative prognostic effect for IDH1 and IDH2 mutations in children with AML.[328]
  • 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.[333] 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).[333] 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.[334] 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.[334] 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.[335]
(Refer to the PDQ summary on Childhood Acute Myeloid Leukemia/Other Myeloid Malignancies Treatment for information about the treatment of childhood AML.)

Juvenile Myelomonocytic Leukemia (JMML)

The genomic landscape of JMML is characterized by mutations in one of five genes of the Ras pathway: NF1NRASKRASPTPN11, and CBL.[336-338] In a series of 118 consecutively diagnosed JMML cases with Ras pathway–activating mutations, PTPN11 was the most commonly mutated gene, accounting for 51% of cases (19% germline and 32% somatic) (refer to Figure 3).[336] Patients with mutated NRAS accounted for 19% of cases, and patients with mutated KRAS accounted for 15% of cases. NF1 mutations accounted for 8% of cases and CBL mutations accounted for 11% of cases. Although mutations among these five genes are generally mutually exclusive, 4% to 17% of cases have mutations in two of these Ras pathway genes,[336-338] a finding that is associated with poorer prognosis.[336,338]
The mutation rate in JMML leukemia cells is very low, but additional mutations beyond those of the five Ras pathway genes described above are observed.[336-338] Secondary genomic alterations are observed for genes of the transcriptional repressor complex PRC2 (e.g., ASXL1 was mutated in 7%–8% of cases). Some genes associated with myeloproliferative neoplasms in adults are also mutated at low rates in JMML (e.g., SETBP1 was mutated in 6%–9% of cases).[336-339JAK3 mutations are also observed in a small percentage (4%–12%) of JMML cases.[336-339] Cases with germline PTPN11 and germline CBL mutations showed low rates of additional mutations (refer to Figure 3).[336] The presence of mutations beyond disease-defining Ras pathway mutations is associated with an inferior prognosis.[336,337]
A report describing the genomic landscape of JMML found that 16 of 150 patients (11%) lacked canonical Ras pathway mutations. Among these 16 patients, 3 were observed to have in-frame fusions involving receptor tyrosine kinases (DCTN1-ALKRANBP2-ALK, and TBL1XR1-ROS1). These patients all had monosomy 7 and were aged 56 months or older. One patient with an ALK fusion was treated with crizotinib plus conventional chemotherapy and achieved a complete molecular remission and proceeded to allogeneic bone marrow transplantation.[338]
ENLARGEChart showing alteration profiles in individual JMML cases.
Figure 3. Alteration profiles in individual JMML cases. Germline and somatically acquired alterations with recurring hits in the RAS pathway and PRC2 network are shown for 118 patients with JMML who underwent detailed genetic analysis. Blast excess was defined as a blast count ≥10% but <20% of nucleated cells in the bone marrow at diagnosis. Blast crisis was defined as a blast count ≥20% of nucleated cells in the bone marrow. NS, Noonan syndrome. Reprinted by permission from Macmillan Publishers Ltd: Nature GeneticsExit Disclaimer (Caye A, Strullu M, Guidez F, et al.: Juvenile myelomonocytic leukemia displays mutations in components of the RAS pathway and the PRC2 network. Nat Genet 47 [11]: 1334-40, 2015), copyright (2015).

Prognosis (genomic and molecular factors)

Several genomic factors affect the prognosis of patients with JMML, including the following:
  1. Number of non–Ras pathway mutations. A predictor of prognosis for children with JMML is the number of mutations beyond the disease-defining Ras pathway mutations.[336,337]
    • One study observed that zero or one somatic alteration (pathogenic mutation or monosomy 7) was identified in 64 patients (65.3%) at diagnosis, whereas two or more alterations were identified in 34 patients (34.7%).[337] In multivariate analysis, mutation number (2 or more vs. 0 or 1) maintained significance as a predictor of inferior event-free survival (EFS) and overall survival (OS). A higher proportion of patients diagnosed with two or more alterations were older and male, and these patients also demonstrated a higher rate of monosomy 7 or somatic NF1 mutations.[337]
    • Another study observed that approximately 60% of patients had one or more additional mutations beyond their disease-defining Ras pathway mutation. These patients had an inferior OS compared with patients who had no additional mutations (3-year OS, 61% vs. 85%, respectively).[336]
    • A third study observed a trend for an inferior OS for patients with two or more mutations compared with patients with zero or one mutation.[338]
  2. Ras pathway double mutations. Although mutations in the five canonical Ras pathway genes associated with JMML (NF1NRASKRASPTPN11, and CBL) are generally mutually exclusive, 4% to 17% of cases have mutations in two of these Ras pathway genes,[336,337] a finding that has been associated with a poorer prognosis.[336,337]
    • Two Ras pathway mutations were identified in 11% of JMML patients in one report, and these patients had significantly inferior EFS (14%) compared with patients who had a single Ras pathway mutation (62%). Patients with Noonan syndrome were excluded from the analyses.[337]
    • Similar findings for Ras pathway mutations were reported in a second study that observed that patients with Ras pathway double mutations (15 of 96 patients) had lower survival rates than did patients with either no additional mutations or with additional mutations beyond the Ras pathway mutation.[336]
  3. DNA methylation profile.
    • One study applied DNA methylation profiling to a discovery cohort of 39 patients with JMML and to a validation cohort of 40 patients. Distinctive subsets of JMML with either high, intermediate, or low methylation levels were observed in both cohorts. Patients with the lowest methylation levels had the highest survival rates, and all but 1 of 15 patients experienced spontaneous resolution in the low methylation cohort. High methylation status was associated with lower EFS rates.[340]
    • Another study applied DNA methylation profiling to a cohort of 106 patients with JMML and observed one subgroup of patients with a hypermethylation profile and one subgroup of patients with a hypomethylation profile. Patients in the hypermethylation group had a significantly lower OS rate than did patients in the hypomethylation group (5-year OS, 46% vs. 73%, respectively). Patients in the hypermethylation group also had a significantly poorer 5-year transplant-free survival rate than did patients in the hypomethylation group (2.2%; 95% CI, 0.2%–10.1% vs. 41.2%; 95% CI, 27.1%–54.8%). Hypermethylation status was associated with two or more mutations, higher fetal hemoglobin levels, older age, and lower platelet count at diagnosis. All patients with Noonan syndrome were in the hypomethylation group.[338]
  4. LIN28B overexpression. LIN28B overexpression is present in approximately one-half of children with JMML and identifies a biologically distinctive subset of JMML. LIN28B is an RNA-binding protein that regulates stem cell renewal.[341]
    • LIN28B overexpression was positively correlated with high blood fetal hemoglobin level and age (both of which are associated with poor prognosis), and it was negatively correlated with presence of monosomy 7 (also associated with inferior prognosis). Although LIN28B overexpression identifies a subset of patients with increased risk of treatment failure, it was not found to be an independent prognostic factor when other factors such as age and monosomy 7 status are considered.[341]
    • Another study also observed a subset of JMML patients with elevated LIN28B expression and identified LIN28B as the gene for which expression was most strongly associated with hypermethylation status.[338]

Myelodysplastic Syndromes (MDS)

Pediatric myelodysplastic syndromes (MDS) are associated with a distinctive constellation of genetic alterations compared with MDS arising in adults. In adults, MDS often evolves from clonal hematopoiesis and is characterized by mutations in TET2DNMT3A, and TP53. In contrast, mutations in these genes are rare in pediatric MDS, while mutations in GATA2SAMD9/SAMD9LSETBP1ASXL1, and Ras/MAPK pathway genes are observed in subsets of pediatric MDS cases.[342,343]
A report of the genomic landscape of pediatric MDS described the results of whole-exome sequencing for 32 pediatric primary MDS patients and targeted sequencing for another 14 cases.[342] These 46 cases were equally divided between refractory cytopenia of childhood and MDS with excess blasts (MDS-EB). The results from the report include the following:
  • Mutations in Ras/MAPK pathway genes were observed in 43% of primary MDS cases, with mutations most commonly involving PTPN11 and NRAS but with mutations also observed in other pathway members (e.g., BRAF [non–BRAF V600E], CBL, and KRAS). Ras/MAPK mutations were more common in patients with MDS-EB (65%) than in patients with refractory cytopenia of childhood (17%).
  • Germline variants in SAMD9 (n = 4) or SAMD9L (n = 4) were observed in 17% of patients with primary MDS, with seven of eight mutations occurring in patients with refractory cytopenia of childhood. These cases all showed loss of material on chromosome 7. Approximately 40% of patients with deletions of part or all of chromosome 7 had germline SAMD9 or SAMD9L variants.
  • GATA2 mutations were observed in three cases (7%), and all cases were confirmed or presumed to be germline.
  • Deletions involving chromosome 7 were the most common copy number alteration and were observed in 41% of cases. Loss of part or all of chromosome 7 was most commonly observed in SAMD9/SAMD9L cases (100%) and in MDS-EB patients with a Ras/MAPK mutation (71%).
  • Other genes that were mutated in more than 1 of the 46 cases studied included SETBP1ETV6, and TP53.
A second report described the application of a targeted sequencing panel of 105 genes to 50 pediatric patients with MDS (refractory cytopenia of childhood = 31 and MDS-EB = 19) and was enriched for cases with monosomy 7 (48%).[342,343SAMD9 and SAMD9L were not included in the gene panel. The second report described the following results:
  • Germline GATA2 mutations were observed in 30% of patients, and RUNX1 mutations were observed in 6% of patients.
  • Somatic mutations were observed in 34% of patients and were more common in patients with MDS-EB than in patients with refractory cytopenia of childhood (68% vs. 13%).
  • The most commonly mutated gene was SETBP1 (18%); less commonly mutated genes included ASXL1RUNX1, and Ras/MAPK pathway genes (PTPN11NRASKRASNF1). Twelve percent of cases showed mutations in Ras/MAPK pathway genes.
Patients with germline GATA2 mutations, in addition to MDS, show a wide range of hematopoietic and immune defects as well as nonhematopoietic manifestations.[344] The former defects include monocytopenia with susceptibility to atypical mycobacterial infection and DCML deficiency (loss of dendritic cells, monocytes, and B and natural killer lymphoid cells). The resulting immunodeficiency leads to increased susceptibility to warts, severe viral infections, mycobacterial infections, fungal infections, and human papillomavirus–related cancers. The nonhematopoietic manifestations include deafness and lymphedema. Germline GATA2 mutations were studied in 426 pediatric patients with primary MDS and 82 cases with secondary MDS who were enrolled in consecutive studies of the European Working Group of MDS in Childhood (EWOG-MDS).[345] The study had the following results:
  • Germline GATA2 mutations were identified in 7% of pediatric patients with primary MDS. While the median age of patients presenting with GATA2 mutations was 12.3 years in the EWOG-MDS pediatric population, most cases of germline GATA2-related myeloid neoplasms occur during adulthood.[346]
  • GATA2 mutations were more common in patients with MDS-EB (15%) than in patients with refractory cytopenia of childhood (4%).
  • Among patients with GATA2 mutations, 46% presented with MDS-EB and 70% showed monosomy 7.
  • Familial MDS/AML was identified in 12 of 53 GATA2-mutated patients for whom detailed family histories were available.
  • Nonhematologic phenotypes of GATA2 deficiency were present in 51% of GATA2-mutated patients with MDS and included deafness (9%), lymphedema/hydrocele (23%), and immunodeficiency (39%).
SAMD9 and SAMD9L germline mutations are both associated with pediatric MDS cases in which there is an additional loss of all or part of chromosome 7.[347] In 2016, SAMD9 was identified as the cause of the MIRAGE syndrome (myelodysplasia, infection, restriction of growth, adrenal hypoplasia, genital phenotypes, and enteropathy), which is associated with early-onset MDS with monosomy 7.[348] Subsequently, mutations in SAMD9L were identified in patients with ataxia pancytopenia syndrome (ATXPC; OMIM 159550Exit Disclaimer). SAMD9 and SAMD9L mutations were also identified as the cause of myelodysplasia and leukemia syndrome with monosomy 7 (MLSM7; OMIM 252270Exit Disclaimer),[349] a syndrome first identified in phenotypically normal siblings who developed MDS or AML associated with monosomy 7 during childhood.[350]
  • Causative mutations in both SAMD9 and SAMD9L are gain-of-function mutations and enhance the growth-suppressing activity of SAMD9/SAMD9L.[348,350]
  • Both SAMD9 and SAMD9L are located at chromosome 7q21.2. Cases of MDS in patients with SAMD9 or SAMD9L mutations often show monosomy 7, with the remaining chromosome 7 having wild-type SAMD9/SAMD9L. This results in the loss of the enhanced growth-suppressing activity of the mutated gene.
  • Phenotypically normal patients with SAMD9/SAMD9L mutations and monosomy 7 may progress to MDS or AML or, alternatively, may show loss of their monosomy 7 with a return of normal hematopoiesis.[350] The former outcome is associated with the acquisition of mutations in genes associated with MDS/AML (e.g., ETV6 or SETBP1), while the latter is associated with genetic alterations (e.g., revertant mutations or copy-neutral loss of heterozygosity with retention of the wild-type allele) that result in normalization of SAMD9/SAMD9L activity. These observations suggest that monitoring of patients with SAMD9/SAMD9L-related monosomy 7 using clinical sequencing for acquired mutations in genes associated with progression to AML may identify patients at high risk of leukemic transformation who may benefit most from hematopoietic stem cell transplantation.[350]


References
  1. Mullighan CG: Genomic characterization of childhood acute lymphoblastic leukemia. Semin Hematol 50 (4): 314-24, 2013. [PUBMED Abstract]
  2. Mullighan CG, Goorha S, Radtke I, et al.: Genome-wide analysis of genetic alterations in acute lymphoblastic leukaemia. Nature 446 (7137): 758-64, 2007. [PUBMED Abstract]
  3. Mullighan CG, Miller CB, Radtke I, et al.: BCR-ABL1 lymphoblastic leukaemia is characterized by the deletion of Ikaros. Nature 453 (7191): 110-4, 2008. [PUBMED Abstract]
  4. Roberts KG, Li Y, Payne-Turner D, et al.: Targetable kinase-activating lesions in Ph-like acute lymphoblastic leukemia. N Engl J Med 371 (11): 1005-15, 2014. [PUBMED Abstract]
  5. Harvey RC, Mullighan CG, Wang X, et al.: Identification of novel cluster groups in pediatric high-risk B-precursor acute lymphoblastic leukemia with gene expression profiling: correlation with genome-wide DNA copy number alterations, clinical characteristics, and outcome. Blood 116 (23): 4874-84, 2010. [PUBMED Abstract]
  6. Clappier E, Auclerc MF, Rapion J, et al.: An intragenic ERG deletion is a marker of an oncogenic subtype of B-cell precursor acute lymphoblastic leukemia with a favorable outcome despite frequent IKZF1 deletions. Leukemia 28 (1): 70-7, 2014. [PUBMED Abstract]
  7. Zaliova M, Zimmermannova O, Dörge P, et al.: ERG deletion is associated with CD2 and attenuates the negative impact of IKZF1 deletion in childhood acute lymphoblastic leukemia. Leukemia 28 (1): 182-5, 2014. [PUBMED Abstract]
  8. Holmfeldt L, Wei L, Diaz-Flores E, et al.: The genomic landscape of hypodiploid acute lymphoblastic leukemia. Nat Genet 45 (3): 242-52, 2013. [PUBMED Abstract]
  9. Loh ML, Zhang J, Harvey RC, et al.: Tyrosine kinome sequencing of pediatric acute lymphoblastic leukemia: a report from the Children's Oncology Group TARGET Project. Blood 121 (3): 485-8, 2013. [PUBMED Abstract]
  10. Bercovich D, Ganmore I, Scott LM, et al.: Mutations of JAK2 in acute lymphoblastic leukaemias associated with Down's syndrome. Lancet 372 (9648): 1484-92, 2008. [PUBMED Abstract]
  11. Andersson AK, Ma J, Wang J, et al.: The landscape of somatic mutations in infant MLL-rearranged acute lymphoblastic leukemias. Nat Genet 47 (4): 330-7, 2015. [PUBMED Abstract]
  12. Ma X, Edmonson M, Yergeau D, et al.: Rise and fall of subclones from diagnosis to relapse in pediatric B-acute lymphoblastic leukaemia. Nat Commun 6: 6604, 2015. [PUBMED Abstract]
  13. Meyer JA, Wang J, Hogan LE, et al.: Relapse-specific mutations in NT5C2 in childhood acute lymphoblastic leukemia. Nat Genet 45 (3): 290-4, 2013. [PUBMED Abstract]
  14. Li B, Li H, Bai Y, et al.: Negative feedback-defective PRPS1 mutants drive thiopurine resistance in relapsed childhood ALL. Nat Med 21 (6): 563-71, 2015. [PUBMED Abstract]
  15. Mullighan CG, Zhang J, Kasper LH, et al.: CREBBP mutations in relapsed acute lymphoblastic leukaemia. Nature 471 (7337): 235-9, 2011. [PUBMED Abstract]
  16. Liu Y, Easton J, Shao Y, et al.: The genomic landscape of pediatric and young adult T-lineage acute lymphoblastic leukemia. Nat Genet 49 (8): 1211-1218, 2017. [PUBMED Abstract]
  17. Armstrong SA, Look AT: Molecular genetics of acute lymphoblastic leukemia. J Clin Oncol 23 (26): 6306-15, 2005. [PUBMED Abstract]
  18. Karrman K, Forestier E, Heyman M, et al.: Clinical and cytogenetic features of a population-based consecutive series of 285 pediatric T-cell acute lymphoblastic leukemias: rare T-cell receptor gene rearrangements are associated with poor outcome. Genes Chromosomes Cancer 48 (9): 795-805, 2009. [PUBMED Abstract]
  19. Moorman AV, Ensor HM, Richards SM, et al.: Prognostic effect of chromosomal abnormalities in childhood B-cell precursor acute lymphoblastic leukaemia: results from the UK Medical Research Council ALL97/99 randomised trial. Lancet Oncol 11 (5): 429-38, 2010. [PUBMED Abstract]
  20. 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]
  21. Paulsson K, Johansson B: High hyperdiploid childhood acute lymphoblastic leukemia. Genes Chromosomes Cancer 48 (8): 637-60, 2009. [PUBMED Abstract]
  22. Aricò M, Valsecchi MG, Rizzari C, et al.: Long-term results of the AIEOP-ALL-95 Trial for Childhood Acute Lymphoblastic Leukemia: insight on the prognostic value of DNA index in the framework of Berlin-Frankfurt-Muenster based chemotherapy. J Clin Oncol 26 (2): 283-9, 2008. [PUBMED Abstract]
  23. Dastugue N, Suciu S, Plat G, et al.: Hyperdiploidy with 58-66 chromosomes in childhood B-acute lymphoblastic leukemia is highly curable: 58951 CLG-EORTC results. Blood 121 (13): 2415-23, 2013. [PUBMED Abstract]
  24. Synold TW, Relling MV, Boyett JM, et al.: Blast cell methotrexate-polyglutamate accumulation in vivo differs by lineage, ploidy, and methotrexate dose in acute lymphoblastic leukemia. J Clin Invest 94 (5): 1996-2001, 1994. [PUBMED Abstract]
  25. Moorman AV, Richards SM, Martineau M, et al.: Outcome heterogeneity in childhood high-hyperdiploid acute lymphoblastic leukemia. Blood 102 (8): 2756-62, 2003. [PUBMED Abstract]
  26. Chilton L, Buck G, Harrison CJ, et al.: High hyperdiploidy among adolescents and adults with acute lymphoblastic leukaemia (ALL): cytogenetic features, clinical characteristics and outcome. Leukemia 28 (7): 1511-8, 2014. [PUBMED Abstract]
  27. Sutcliffe MJ, Shuster JJ, Sather HN, et al.: High concordance from independent studies by the Children's Cancer Group (CCG) and Pediatric Oncology Group (POG) associating favorable prognosis with combined trisomies 4, 10, and 17 in children with NCI Standard-Risk B-precursor Acute Lymphoblastic Leukemia: a Children's Oncology Group (COG) initiative. Leukemia 19 (5): 734-40, 2005. [PUBMED Abstract]
  28. Harris MB, Shuster JJ, Carroll A, et al.: Trisomy of leukemic cell chromosomes 4 and 10 identifies children with B-progenitor cell acute lymphoblastic leukemia with a very low risk of treatment failure: a Pediatric Oncology Group study. Blood 79 (12): 3316-24, 1992. [PUBMED Abstract]
  29. Heerema NA, Harbott J, Galimberti S, et al.: Secondary cytogenetic aberrations in childhood Philadelphia chromosome positive acute lymphoblastic leukemia are nonrandom and may be associated with outcome. Leukemia 18 (4): 693-702, 2004. [PUBMED Abstract]
  30. Nachman JB, Heerema NA, Sather H, et al.: Outcome of treatment in children with hypodiploid acute lymphoblastic leukemia. Blood 110 (4): 1112-5, 2007. [PUBMED Abstract]
  31. Raimondi SC, Zhou Y, Shurtleff SA, et al.: Near-triploidy and near-tetraploidy in childhood acute lymphoblastic leukemia: association with B-lineage blast cells carrying the ETV6-RUNX1 fusion, T-lineage immunophenotype, and favorable outcome. Cancer Genet Cytogenet 169 (1): 50-7, 2006. [PUBMED Abstract]
  32. Attarbaschi A, Mann G, König M, et al.: Incidence and relevance of secondary chromosome abnormalities in childhood TEL/AML1+ acute lymphoblastic leukemia: an interphase FISH analysis. Leukemia 18 (10): 1611-6, 2004. [PUBMED Abstract]
  33. Lemez P, Attarbaschi A, Béné MC, et al.: Childhood near-tetraploid acute lymphoblastic leukemia: an EGIL study on 36 cases. Eur J Haematol 85 (4): 300-8, 2010. [PUBMED Abstract]
  34. Paulsson K, Lilljebjörn H, Biloglav A, et al.: The genomic landscape of high hyperdiploid childhood acute lymphoblastic leukemia. Nat Genet 47 (6): 672-6, 2015. [PUBMED Abstract]
  35. Harrison CJ, Moorman AV, Broadfield ZJ, et al.: Three distinct subgroups of hypodiploidy in acute lymphoblastic leukaemia. Br J Haematol 125 (5): 552-9, 2004. [PUBMED Abstract]
  36. Mullighan CG, Jeha S, Pei D, et al.: Outcome of children with hypodiploid ALL treated with risk-directed therapy based on MRD levels. Blood 126 (26): 2896-9, 2015. [PUBMED Abstract]
  37. Pui CH, Rebora P, Schrappe M, et al.: Outcome of Children With Hypodiploid Acute Lymphoblastic Leukemia: A Retrospective Multinational Study. J Clin Oncol 37 (10): 770-779, 2019. [PUBMED Abstract]
  38. McNeer JL, Devidas M, Dai Y, et al.: Hematopoietic Stem-Cell Transplantation Does Not Improve the Poor Outcome of Children With Hypodiploid Acute Lymphoblastic Leukemia: A Report From Children's Oncology Group. J Clin Oncol 37 (10): 780-789, 2019. [PUBMED Abstract]
  39. Irving J, Matheson E, Minto L, et al.: Ras pathway mutations are prevalent in relapsed childhood acute lymphoblastic leukemia and confer sensitivity to MEK inhibition. Blood 124 (23): 3420-30, 2014. [PUBMED Abstract]
  40. Qian M, Cao X, Devidas M, et al.: TP53 Germline Variations Influence the Predisposition and Prognosis of B-Cell Acute Lymphoblastic Leukemia in Children. J Clin Oncol 36 (6): 591-599, 2018. [PUBMED Abstract]
  41. Rubnitz JE, Wichlan D, Devidas M, et al.: Prospective analysis of TEL gene rearrangements in childhood acute lymphoblastic leukemia: a Children's Oncology Group study. J Clin Oncol 26 (13): 2186-91, 2008. [PUBMED Abstract]
  42. Kanerva J, Saarinen-Pihkala UM, Niini T, et al.: Favorable outcome in 20-year follow-up of children with very-low-risk ALL and minimal standard therapy, with special reference to TEL-AML1 fusion. Pediatr Blood Cancer 42 (1): 30-5, 2004. [PUBMED Abstract]
  43. Aldrich MC, Zhang L, Wiemels JL, et al.: Cytogenetics of Hispanic and White children with acute lymphoblastic leukemia in California. Cancer Epidemiol Biomarkers Prev 15 (3): 578-81, 2006. [PUBMED Abstract]
  44. Loh ML, Goldwasser MA, Silverman LB, et al.: Prospective analysis of TEL/AML1-positive patients treated on Dana-Farber Cancer Institute Consortium Protocol 95-01. Blood 107 (11): 4508-13, 2006. [PUBMED Abstract]
  45. Borowitz MJ, Devidas M, Hunger SP, et al.: Clinical significance of minimal residual disease in childhood acute lymphoblastic leukemia and its relationship to other prognostic factors: a Children's Oncology Group study. Blood 111 (12): 5477-85, 2008. [PUBMED Abstract]
  46. Madzo J, Zuna J, Muzíková K, et al.: Slower molecular response to treatment predicts poor outcome in patients with TEL/AML1 positive acute lymphoblastic leukemia: prospective real-time quantitative reverse transcriptase-polymerase chain reaction study. Cancer 97 (1): 105-13, 2003. [PUBMED Abstract]
  47. Bhojwani D, Pei D, Sandlund JT, et al.: ETV6-RUNX1-positive childhood acute lymphoblastic leukemia: improved outcome with contemporary therapy. Leukemia 26 (2): 265-70, 2012. [PUBMED Abstract]
  48. Enshaei A, Schwab CJ, Konn ZJ, et al.: Long-term follow-up of ETV6-RUNX1 ALL reveals that NCI risk, rather than secondary genetic abnormalities, is the key risk factor. Leukemia 27 (11): 2256-9, 2013. [PUBMED Abstract]
  49. Barbany G, Andersen MK, Autio K, et al.: Additional aberrations of the ETV6 and RUNX1 genes have no prognostic impact in 229 t(12;21)(p13;q22)-positive B-cell precursor acute lymphoblastic leukaemias treated according to the NOPHO-ALL-2000 protocol. Leuk Res 36 (7): 936-8, 2012. [PUBMED Abstract]
  50. Forestier E, Heyman M, Andersen MK, et al.: Outcome of ETV6/RUNX1-positive childhood acute lymphoblastic leukaemia in the NOPHO-ALL-1992 protocol: frequent late relapses but good overall survival. Br J Haematol 140 (6): 665-72, 2008. [PUBMED Abstract]
  51. Seeger K, Stackelberg AV, Taube T, et al.: Relapse of TEL-AML1--positive acute lymphoblastic leukemia in childhood: a matched-pair analysis. J Clin Oncol 19 (13): 3188-93, 2001. [PUBMED Abstract]
  52. Gandemer V, Chevret S, Petit A, et al.: Excellent prognosis of late relapses of ETV6/RUNX1-positive childhood acute lymphoblastic leukemia: lessons from the FRALLE 93 protocol. Haematologica 97 (11): 1743-50, 2012. [PUBMED Abstract]
  53. Zuna J, Ford AM, Peham M, et al.: TEL deletion analysis supports a novel view of relapse in childhood acute lymphoblastic leukemia. Clin Cancer Res 10 (16): 5355-60, 2004. [PUBMED Abstract]
  54. van Delft FW, Horsley S, Colman S, et al.: Clonal origins of relapse in ETV6-RUNX1 acute lymphoblastic leukemia. Blood 117 (23): 6247-54, 2011. [PUBMED Abstract]
  55. Aricò M, Schrappe M, Hunger SP, et al.: Clinical outcome of children with newly diagnosed Philadelphia chromosome-positive acute lymphoblastic leukemia treated between 1995 and 2005. J Clin Oncol 28 (31): 4755-61, 2010. [PUBMED Abstract]
  56. Schrappe M, Aricò M, Harbott J, et al.: Philadelphia chromosome-positive (Ph+) childhood acute lymphoblastic leukemia: good initial steroid response allows early prediction of a favorable treatment outcome. Blood 92 (8): 2730-41, 1998. [PUBMED Abstract]
  57. Ribeiro RC, Broniscer A, Rivera GK, et al.: Philadelphia chromosome-positive acute lymphoblastic leukemia in children: durable responses to chemotherapy associated with low initial white blood cell counts. Leukemia 11 (9): 1493-6, 1997. [PUBMED Abstract]
  58. Biondi A, Schrappe M, De Lorenzo P, et al.: Imatinib after induction for treatment of children and adolescents with Philadelphia-chromosome-positive acute lymphoblastic leukaemia (EsPhALL): a randomised, open-label, intergroup study. Lancet Oncol 13 (9): 936-45, 2012. [PUBMED Abstract]
  59. Schultz KR, Bowman WP, Aledo A, et al.: Improved early event-free survival with imatinib in Philadelphia chromosome-positive acute lymphoblastic leukemia: a children's oncology group study. J Clin Oncol 27 (31): 5175-81, 2009. [PUBMED Abstract]
  60. Schultz KR, Carroll A, Heerema NA, et al.: Long-term follow-up of imatinib in pediatric Philadelphia chromosome-positive acute lymphoblastic leukemia: Children's Oncology Group study AALL0031. Leukemia 28 (7): 1467-71, 2014. [PUBMED Abstract]
  61. Pui CH, Chessells JM, Camitta B, et al.: Clinical heterogeneity in childhood acute lymphoblastic leukemia with 11q23 rearrangements. Leukemia 17 (4): 700-6, 2003. [PUBMED Abstract]
  62. Johansson B, Moorman AV, Haas OA, et al.: Hematologic malignancies with t(4;11)(q21;q23)--a cytogenetic, morphologic, immunophenotypic and clinical study of 183 cases. European 11q23 Workshop participants. Leukemia 12 (5): 779-87, 1998. [PUBMED Abstract]
  63. Raimondi SC, Peiper SC, Kitchingman GR, et al.: Childhood acute lymphoblastic leukemia with chromosomal breakpoints at 11q23. Blood 73 (6): 1627-34, 1989. [PUBMED Abstract]
  64. Harrison CJ, Moorman AV, Barber KE, et al.: Interphase molecular cytogenetic screening for chromosomal abnormalities of prognostic significance in childhood acute lymphoblastic leukaemia: a UK Cancer Cytogenetics Group Study. Br J Haematol 129 (4): 520-30, 2005. [PUBMED Abstract]
  65. Pui CH, Pei D, Campana D, et al.: A revised definition for cure of childhood acute lymphoblastic leukemia. Leukemia 28 (12): 2336-43, 2014. [PUBMED Abstract]
  66. Pieters R, Schrappe M, De Lorenzo P, et al.: A treatment protocol for infants younger than 1 year with acute lymphoblastic leukaemia (Interfant-99): an observational study and a multicentre randomised trial. Lancet 370 (9583): 240-50, 2007. [PUBMED Abstract]
  67. Pui CH, Gaynon PS, Boyett JM, et al.: Outcome of treatment in childhood acute lymphoblastic leukaemia with rearrangements of the 11q23 chromosomal region. Lancet 359 (9321): 1909-15, 2002. [PUBMED Abstract]
  68. Rubnitz JE, Camitta BM, Mahmoud H, et al.: Childhood acute lymphoblastic leukemia with the MLL-ENL fusion and t(11;19)(q23;p13.3) translocation. J Clin Oncol 17 (1): 191-6, 1999. [PUBMED Abstract]
  69. Hunger SP: Chromosomal translocations involving the E2A gene in acute lymphoblastic leukemia: clinical features and molecular pathogenesis. Blood 87 (4): 1211-24, 1996. [PUBMED Abstract]
  70. Uckun FM, Sensel MG, Sather HN, et al.: Clinical significance of translocation t(1;19) in childhood acute lymphoblastic leukemia in the context of contemporary therapies: a report from the Children's Cancer Group. J Clin Oncol 16 (2): 527-35, 1998. [PUBMED Abstract]
  71. Fischer U, Forster M, Rinaldi A, et al.: Genomics and drug profiling of fatal TCF3-HLF-positive acute lymphoblastic leukemia identifies recurrent mutation patterns and therapeutic options. Nat Genet 47 (9): 1020-9, 2015. [PUBMED Abstract]
  72. Pui CH, Sandlund JT, Pei D, et al.: Results of therapy for acute lymphoblastic leukemia in black and white children. JAMA 290 (15): 2001-7, 2003. [PUBMED Abstract]
  73. Crist WM, Carroll AJ, Shuster JJ, et al.: Poor prognosis of children with pre-B acute lymphoblastic leukemia is associated with the t(1;19)(q23;p13): a Pediatric Oncology Group study. Blood 76 (1): 117-22, 1990. [PUBMED Abstract]
  74. Andersen MK, Autio K, Barbany G, et al.: Paediatric B-cell precursor acute lymphoblastic leukaemia with t(1;19)(q23;p13): clinical and cytogenetic characteristics of 47 cases from the Nordic countries treated according to NOPHO protocols. Br J Haematol 155 (2): 235-43, 2011. [PUBMED Abstract]
  75. Pui CH, Campana D, Pei D, et al.: Treating childhood acute lymphoblastic leukemia without cranial irradiation. N Engl J Med 360 (26): 2730-41, 2009. [PUBMED Abstract]
  76. Jeha S, Pei D, Raimondi SC, et al.: Increased risk for CNS relapse in pre-B cell leukemia with the t(1;19)/TCF3-PBX1. Leukemia 23 (8): 1406-9, 2009. [PUBMED Abstract]
  77. Minson KA, Prasad P, Vear S, et al.: t(17;19) in Children with Acute Lymphocytic Leukemia: A Report of 3 Cases and a Review of the Literature. Case Rep Hematol 2013: 563291, 2013. [PUBMED Abstract]
  78. Lilljebjörn H, Henningsson R, Hyrenius-Wittsten A, et al.: Identification of ETV6-RUNX1-like and DUX4-rearranged subtypes in paediatric B-cell precursor acute lymphoblastic leukaemia. Nat Commun 7: 11790, 2016. [PUBMED Abstract]
  79. Zhang J, McCastlain K, Yoshihara H, et al.: Deregulation of DUX4 and ERG in acute lymphoblastic leukemia. Nat Genet 48 (12): 1481-1489, 2016. [PUBMED Abstract]
  80. Gu Z, Churchman M, Roberts K, et al.: Genomic analyses identify recurrent MEF2D fusions in acute lymphoblastic leukaemia. Nat Commun 7: 13331, 2016. [PUBMED Abstract]
  81. Liu YF, Wang BY, Zhang WN, et al.: Genomic Profiling of Adult and Pediatric B-cell Acute Lymphoblastic Leukemia. EBioMedicine 8: 173-83, 2016. [PUBMED Abstract]
  82. Suzuki K, Okuno Y, Kawashima N, et al.: MEF2D-BCL9 Fusion Gene Is Associated With High-Risk Acute B-Cell Precursor Lymphoblastic Leukemia in Adolescents. J Clin Oncol 34 (28): 3451-9, 2016. [PUBMED Abstract]
  83. Lilljebjörn H, Ågerstam H, Orsmark-Pietras C, et al.: RNA-seq identifies clinically relevant fusion genes in leukemia including a novel MEF2D/CSF1R fusion responsive to imatinib. Leukemia 28 (4): 977-9, 2014. [PUBMED Abstract]
  84. Hirabayashi S, Ohki K, Nakabayashi K, et al.: ZNF384-related fusion genes define a subgroup of childhood B-cell precursor acute lymphoblastic leukemia with a characteristic immunotype. Haematologica 102 (1): 118-129, 2017. [PUBMED Abstract]
  85. Qian M, Zhang H, Kham SK, et al.: Whole-transcriptome sequencing identifies a distinct subtype of acute lymphoblastic leukemia with predominant genomic abnormalities of EP300 and CREBBP. Genome Res 27 (2): 185-195, 2017. [PUBMED Abstract]
  86. Shago M, Abla O, Hitzler J, et al.: Frequency and outcome of pediatric acute lymphoblastic leukemia with ZNF384 gene rearrangements including a novel translocation resulting in an ARID1B/ZNF384 gene fusion. Pediatr Blood Cancer 63 (11): 1915-21, 2016. [PUBMED Abstract]
  87. Yao L, Cen J, Pan J, et al.: TAF15-ZNF384 fusion gene in childhood mixed phenotype acute leukemia. Cancer Genet 211: 1-4, 2017. [PUBMED Abstract]
  88. Hogan TF, Koss W, Murgo AJ, et al.: Acute lymphoblastic leukemia with chromosomal 5;14 translocation and hypereosinophilia: case report and literature review. J Clin Oncol 5 (3): 382-90, 1987. [PUBMED Abstract]
  89. Grimaldi JC, Meeker TC: The t(5;14) chromosomal translocation in a case of acute lymphocytic leukemia joins the interleukin-3 gene to the immunoglobulin heavy chain gene. Blood 73 (8): 2081-5, 1989. [PUBMED Abstract]
  90. Meeker TC, Hardy D, Willman C, et al.: Activation of the interleukin-3 gene by chromosome translocation in acute lymphocytic leukemia with eosinophilia. Blood 76 (2): 285-9, 1990. [PUBMED Abstract]
  91. Sutton R, Lonergan M, Tapp H, et al.: Two cases of hypereosinophilia and high-risk acute lymphoblastic leukemia. Leukemia 22 (7): 1463-5, 2008. [PUBMED Abstract]
  92. Heerema NA, Carroll AJ, Devidas M, et al.: Intrachromosomal amplification of chromosome 21 is associated with inferior outcomes in children with acute lymphoblastic leukemia treated in contemporary standard-risk children's oncology group studies: a report from the children's oncology group. J Clin Oncol 31 (27): 3397-402, 2013. [PUBMED Abstract]
  93. Moorman AV, Robinson H, Schwab C, et al.: Risk-directed treatment intensification significantly reduces the risk of relapse among children and adolescents with acute lymphoblastic leukemia and intrachromosomal amplification of chromosome 21: a comparison of the MRC ALL97/99 and UKALL2003 trials. J Clin Oncol 31 (27): 3389-96, 2013. [PUBMED Abstract]
  94. Harrison CJ, Moorman AV, Schwab C, et al.: An international study of intrachromosomal amplification of chromosome 21 (iAMP21): cytogenetic characterization and outcome. Leukemia 28 (5): 1015-21, 2014. [PUBMED Abstract]
  95. Schwab C, Nebral K, Chilton L, et al.: Intragenic amplification of PAX5: a novel subgroup in B-cell precursor acute lymphoblastic leukemia? Blood Adv 1 (19): 1473-7, 2017.
  96. Den Boer ML, van Slegtenhorst M, De Menezes RX, et al.: A subtype of childhood acute lymphoblastic leukaemia with poor treatment outcome: a genome-wide classification study. Lancet Oncol 10 (2): 125-34, 2009. [PUBMED Abstract]
  97. Mullighan CG, Su X, Zhang J, et al.: Deletion of IKZF1 and prognosis in acute lymphoblastic leukemia. N Engl J Med 360 (5): 470-80, 2009. [PUBMED Abstract]
  98. Reshmi SC, Harvey RC, Roberts KG, et al.: Targetable kinase gene fusions in high-risk B-ALL: a study from the Children's Oncology Group. Blood 129 (25): 3352-3361, 2017. [PUBMED Abstract]
  99. Roberts KG, Morin RD, Zhang J, et al.: Genetic alterations activating kinase and cytokine receptor signaling in high-risk acute lymphoblastic leukemia. Cancer Cell 22 (2): 153-66, 2012. [PUBMED Abstract]
  100. van der Veer A, Waanders E, Pieters R, et al.: Independent prognostic value of BCR-ABL1-like signature and IKZF1 deletion, but not high CRLF2 expression, in children with B-cell precursor ALL. Blood 122 (15): 2622-9, 2013. [PUBMED Abstract]
  101. Roberts KG, Reshmi SC, Harvey RC, et al.: Genomic and outcome analyses of Ph-like ALL in NCI standard-risk patients: a report from the Children's Oncology Group. Blood 132 (8): 815-824, 2018. [PUBMED Abstract]
  102. Roberts KG, Pei D, Campana D, et al.: Outcomes of children with BCR-ABL1–like acute lymphoblastic leukemia treated with risk-directed therapy based on the levels of minimal residual disease. J Clin Oncol 32 (27): 3012-20, 2014. [PUBMED Abstract]
  103. Harvey RC, Mullighan CG, Chen IM, et al.: Rearrangement of CRLF2 is associated with mutation of JAK kinases, alteration of IKZF1, Hispanic/Latino ethnicity, and a poor outcome in pediatric B-progenitor acute lymphoblastic leukemia. Blood 115 (26): 5312-21, 2010. [PUBMED Abstract]
  104. Mullighan CG, Collins-Underwood JR, Phillips LA, et al.: Rearrangement of CRLF2 in B-progenitor- and Down syndrome-associated acute lymphoblastic leukemia. Nat Genet 41 (11): 1243-6, 2009. [PUBMED Abstract]
  105. Cario G, Zimmermann M, Romey R, et al.: Presence of the P2RY8-CRLF2 rearrangement is associated with a poor prognosis in non-high-risk precursor B-cell acute lymphoblastic leukemia in children treated according to the ALL-BFM 2000 protocol. Blood 115 (26): 5393-7, 2010. [PUBMED Abstract]
  106. Ensor HM, Schwab C, Russell LJ, et al.: Demographic, clinical, and outcome features of children with acute lymphoblastic leukemia and CRLF2 deregulation: results from the MRC ALL97 clinical trial. Blood 117 (7): 2129-36, 2011. [PUBMED Abstract]
  107. Schmäh J, Fedders B, Panzer-Grümayer R, et al.: Molecular characterization of acute lymphoblastic leukemia with high CRLF2 gene expression in childhood. Pediatr Blood Cancer 64 (10): , 2017. [PUBMED Abstract]
  108. Schwab CJ, Chilton L, Morrison H, et al.: Genes commonly deleted in childhood B-cell precursor acute lymphoblastic leukemia: association with cytogenetics and clinical features. Haematologica 98 (7): 1081-8, 2013. [PUBMED Abstract]
  109. Chen IM, Harvey RC, Mullighan CG, et al.: Outcome modeling with CRLF2, IKZF1, JAK, and minimal residual disease in pediatric acute lymphoblastic leukemia: a Children's Oncology Group study. Blood 119 (15): 3512-22, 2012. [PUBMED Abstract]
  110. Palmi C, Vendramini E, Silvestri D, et al.: Poor prognosis for P2RY8-CRLF2 fusion but not for CRLF2 over-expression in children with intermediate risk B-cell precursor acute lymphoblastic leukemia. Leukemia 26 (10): 2245-53, 2012. [PUBMED Abstract]
  111. Iacobucci I, Li Y, Roberts KG, et al.: Truncating Erythropoietin Receptor Rearrangements in Acute Lymphoblastic Leukemia. Cancer Cell 29 (2): 186-200, 2016. [PUBMED Abstract]
  112. Clappier E, Grardel N, Bakkus M, et al.: IKZF1 deletion is an independent prognostic marker in childhood B-cell precursor acute lymphoblastic leukemia, and distinguishes patients benefiting from pulses during maintenance therapy: results of the EORTC Children's Leukemia Group study 58951. Leukemia 29 (11): 2154-61, 2015. [PUBMED Abstract]
  113. Buitenkamp TD, Pieters R, Gallimore NE, et al.: Outcome in children with Down's syndrome and acute lymphoblastic leukemia: role of IKZF1 deletions and CRLF2 aberrations. Leukemia 26 (10): 2204-11, 2012. [PUBMED Abstract]
  114. Krentz S, Hof J, Mendioroz A, et al.: Prognostic value of genetic alterations in children with first bone marrow relapse of childhood B-cell precursor acute lymphoblastic leukemia. Leukemia 27 (2): 295-304, 2013. [PUBMED Abstract]
  115. Feng J, Tang Y: Prognostic significance of IKZF1 alteration status in pediatric B-lineage acute lymphoblastic leukemia: a meta-analysis. Leuk Lymphoma 54 (4): 889-91, 2013. [PUBMED Abstract]
  116. Dörge P, Meissner B, Zimmermann M, et al.: IKZF1 deletion is an independent predictor of outcome in pediatric acute lymphoblastic leukemia treated according to the ALL-BFM 2000 protocol. Haematologica 98 (3): 428-32, 2013. [PUBMED Abstract]
  117. Olsson L, Castor A, Behrendtz M, et al.: Deletions of IKZF1 and SPRED1 are associated with poor prognosis in a population-based series of pediatric B-cell precursor acute lymphoblastic leukemia diagnosed between 1992 and 2011. Leukemia 28 (2): 302-10, 2014. [PUBMED Abstract]
  118. Boer JM, van der Veer A, Rizopoulos D, et al.: Prognostic value of rare IKZF1 deletion in childhood B-cell precursor acute lymphoblastic leukemia: an international collaborative study. Leukemia 30 (1): 32-8, 2016. [PUBMED Abstract]
  119. Tran TH, Harris MH, Nguyen JV, et al.: Prognostic impact of kinase-activating fusions and IKZF1 deletions in pediatric high-risk B-lineage acute lymphoblastic leukemia. Blood Adv 2 (5): 529-533, 2018. [PUBMED Abstract]
  120. van der Veer A, Zaliova M, Mottadelli F, et al.: IKZF1 status as a prognostic feature in BCR-ABL1-positive childhood ALL. Blood 123 (11): 1691-8, 2014. [PUBMED Abstract]
  121. Stanulla M, Dagdan E, Zaliova M, et al.: IKZF1plus Defines a New Minimal Residual Disease-Dependent Very-Poor Prognostic Profile in Pediatric B-Cell Precursor Acute Lymphoblastic Leukemia. J Clin Oncol 36 (12): 1240-1249, 2018. [PUBMED Abstract]
  122. Yeoh AEJ, Lu Y, Chin WHN, et al.: Intensifying Treatment of Childhood B-Lymphoblastic Leukemia With IKZF1 Deletion Reduces Relapse and Improves Overall Survival: Results of Malaysia-Singapore ALL 2010 Study. J Clin Oncol 36 (26): 2726-2735, 2018. [PUBMED Abstract]
  123. Bergeron J, Clappier E, Radford I, et al.: Prognostic and oncogenic relevance of TLX1/HOX11 expression level in T-ALLs. Blood 110 (7): 2324-30, 2007. [PUBMED Abstract]
  124. van Grotel M, Meijerink JP, Beverloo HB, et al.: The outcome of molecular-cytogenetic subgroups in pediatric T-cell acute lymphoblastic leukemia: a retrospective study of patients treated according to DCOG or COALL protocols. Haematologica 91 (9): 1212-21, 2006. [PUBMED Abstract]
  125. Cavé H, Suciu S, Preudhomme C, et al.: Clinical significance of HOX11L2 expression linked to t(5;14)(q35;q32), of HOX11 expression, and of SIL-TAL fusion in childhood T-cell malignancies: results of EORTC studies 58881 and 58951. Blood 103 (2): 442-50, 2004. [PUBMED Abstract]
  126. Baak U, Gökbuget N, Orawa H, et al.: Thymic adult T-cell acute lymphoblastic leukemia stratified in standard- and high-risk group by aberrant HOX11L2 expression: experience of the German multicenter ALL study group. Leukemia 22 (6): 1154-60, 2008. [PUBMED Abstract]
  127. Ferrando AA, Neuberg DS, Dodge RK, et al.: Prognostic importance of TLX1 (HOX11) oncogene expression in adults with T-cell acute lymphoblastic leukaemia. Lancet 363 (9408): 535-6, 2004. [PUBMED Abstract]
  128. Mansour MR, Abraham BJ, Anders L, et al.: Oncogene regulation. An oncogenic super-enhancer formed through somatic mutation of a noncoding intergenic element. Science 346 (6215): 1373-7, 2014. [PUBMED Abstract]
  129. Petit A, Trinquand A, Chevret S, et al.: Oncogenetic mutations combined with MRD improve outcome prediction in pediatric T-cell acute lymphoblastic leukemia. Blood 131 (3): 289-300, 2018. [PUBMED Abstract]
  130. Burmeister T, Gökbuget N, Reinhardt R, et al.: NUP214-ABL1 in adult T-ALL: the GMALL study group experience. Blood 108 (10): 3556-9, 2006. [PUBMED Abstract]
  131. Graux C, Stevens-Kroef M, Lafage M, et al.: Heterogeneous patterns of amplification of the NUP214-ABL1 fusion gene in T-cell acute lymphoblastic leukemia. Leukemia 23 (1): 125-33, 2009. [PUBMED Abstract]
  132. Hagemeijer A, Graux C: ABL1 rearrangements in T-cell acute lymphoblastic leukemia. Genes Chromosomes Cancer 49 (4): 299-308, 2010. [PUBMED Abstract]
  133. Quintás-Cardama A, Tong W, Manshouri T, et al.: Activity of tyrosine kinase inhibitors against human NUP214-ABL1-positive T cell malignancies. Leukemia 22 (6): 1117-24, 2008. [PUBMED Abstract]
  134. Clarke S, O'Reilly J, Romeo G, et al.: NUP214-ABL1 positive T-cell acute lymphoblastic leukemia patient shows an initial favorable response to imatinib therapy post relapse. Leuk Res 35 (7): e131-3, 2011. [PUBMED Abstract]
  135. Deenik W, Beverloo HB, van der Poel-van de Luytgaarde SC, et al.: Rapid complete cytogenetic remission after upfront dasatinib monotherapy in a patient with a NUP214-ABL1-positive T-cell acute lymphoblastic leukemia. Leukemia 23 (3): 627-9, 2009. [PUBMED Abstract]
  136. Crombet O, Lastrapes K, Zieske A, et al.: Complete morphologic and molecular remission after introduction of dasatinib in the treatment of a pediatric patient with t-cell acute lymphoblastic leukemia and ABL1 amplification. Pediatr Blood Cancer 59 (2): 333-4, 2012. [PUBMED Abstract]
  137. Seki M, Kimura S, Isobe T, et al.: Recurrent SPI1 (PU.1) fusions in high-risk pediatric T cell acute lymphoblastic leukemia. Nat Genet 49 (8): 1274-1281, 2017. [PUBMED Abstract]
  138. Weng AP, Ferrando AA, Lee W, et al.: Activating mutations of NOTCH1 in human T cell acute lymphoblastic leukemia. Science 306 (5694): 269-71, 2004. [PUBMED Abstract]
  139. Gallo Llorente L, Luther H, Schneppenheim R, et al.: Identification of novel NOTCH1 mutations: increasing our knowledge of the NOTCH signaling pathway. Pediatr Blood Cancer 61 (5): 788-96, 2014. [PUBMED Abstract]
  140. Trinquand A, Tanguy-Schmidt A, Ben Abdelali R, et al.: Toward a NOTCH1/FBXW7/RAS/PTEN-based oncogenetic risk classification of adult T-cell acute lymphoblastic leukemia: a Group for Research in Adult Acute Lymphoblastic Leukemia study. J Clin Oncol 31 (34): 4333-42, 2013. [PUBMED Abstract]
  141. Zhang J, Ding L, Holmfeldt L, et al.: The genetic basis of early T-cell precursor acute lymphoblastic leukaemia. Nature 481 (7380): 157-63, 2012. [PUBMED Abstract]
  142. Gutierrez A, Dahlberg SE, Neuberg DS, et al.: Absence of biallelic TCRgamma deletion predicts early treatment failure in pediatric T-cell acute lymphoblastic leukemia. J Clin Oncol 28 (24): 3816-23, 2010. [PUBMED Abstract]
  143. Yang YL, Hsiao CC, Chen HY, et al.: Absence of biallelic TCRγ deletion predicts induction failure and poorer outcomes in childhood T-cell acute lymphoblastic leukemia. Pediatr Blood Cancer 58 (6): 846-51, 2012. [PUBMED Abstract]
  144. Davies SM, Bhatia S, Ross JA, et al.: Glutathione S-transferase genotypes, genetic susceptibility, and outcome of therapy in childhood acute lymphoblastic leukemia. Blood 100 (1): 67-71, 2002. [PUBMED Abstract]
  145. Krajinovic M, Costea I, Chiasson S: Polymorphism of the thymidylate synthase gene and outcome of acute lymphoblastic leukaemia. Lancet 359 (9311): 1033-4, 2002. [PUBMED Abstract]
  146. Krajinovic M, Lemieux-Blanchard E, Chiasson S, et al.: Role of polymorphisms in MTHFR and MTHFD1 genes in the outcome of childhood acute lymphoblastic leukemia. Pharmacogenomics J 4 (1): 66-72, 2004. [PUBMED Abstract]
  147. Schmiegelow K, Forestier E, Kristinsson J, et al.: Thiopurine methyltransferase activity is related to the risk of relapse of childhood acute lymphoblastic leukemia: results from the NOPHO ALL-92 study. Leukemia 23 (3): 557-64, 2009. [PUBMED Abstract]
  148. Relling MV, Hancock ML, Boyett JM, et al.: Prognostic importance of 6-mercaptopurine dose intensity in acute lymphoblastic leukemia. Blood 93 (9): 2817-23, 1999. [PUBMED Abstract]
  149. Stanulla M, Schaeffeler E, Flohr T, et al.: Thiopurine methyltransferase (TPMT) genotype and early treatment response to mercaptopurine in childhood acute lymphoblastic leukemia. JAMA 293 (12): 1485-9, 2005. [PUBMED Abstract]
  150. Yang JJ, Landier W, Yang W, et al.: Inherited NUDT15 variant is a genetic determinant of mercaptopurine intolerance in children with acute lymphoblastic leukemia. J Clin Oncol 33 (11): 1235-42, 2015. [PUBMED Abstract]
  151. Relling MV, Hancock ML, Rivera GK, et al.: Mercaptopurine therapy intolerance and heterozygosity at the thiopurine S-methyltransferase gene locus. J Natl Cancer Inst 91 (23): 2001-8, 1999. [PUBMED Abstract]
  152. Moriyama T, Nishii R, Perez-Andreu V, et al.: NUDT15 polymorphisms alter thiopurine metabolism and hematopoietic toxicity. Nat Genet 48 (4): 367-73, 2016. [PUBMED Abstract]
  153. Tanaka Y, Kato M, Hasegawa D, et al.: Susceptibility to 6-MP toxicity conferred by a NUDT15 variant in Japanese children with acute lymphoblastic leukaemia. Br J Haematol 171 (1): 109-15, 2015. [PUBMED Abstract]
  154. Diouf B, Crews KR, Lew G, et al.: Association of an inherited genetic variant with vincristine-related peripheral neuropathy in children with acute lymphoblastic leukemia. JAMA 313 (8): 815-23, 2015. [PUBMED Abstract]
  155. Yang JJ, Cheng C, Yang W, et al.: Genome-wide interrogation of germline genetic variation associated with treatment response in childhood acute lymphoblastic leukemia. JAMA 301 (4): 393-403, 2009. [PUBMED Abstract]
  156. Gregers J, Christensen IJ, Dalhoff K, et al.: The association of reduced folate carrier 80G>A polymorphism to outcome in childhood acute lymphoblastic leukemia interacts with chromosome 21 copy number. Blood 115 (23): 4671-7, 2010. [PUBMED Abstract]
  157. Radtke S, Zolk O, Renner B, et al.: Germline genetic variations in methotrexate candidate genes are associated with pharmacokinetics, toxicity, and outcome in childhood acute lymphoblastic leukemia. Blood 121 (26): 5145-53, 2013. [PUBMED Abstract]
  158. 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]
  159. 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]
  160. 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]
  161. 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]
  162. Gilliland DG: Targeted therapies in myeloid leukemias. Ann Hematol 83 (Suppl 1): S75-6, 2004. [PUBMED Abstract]
  163. Avivi I, Rowe JM: Prognostic factors in acute myeloid leukemia. Curr Opin Hematol 12 (1): 62-7, 2005. [PUBMED Abstract]
  164. 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]
  165. 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]
  166. 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]
  167. 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]
  168. 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]
  169. 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]
  170. Meshinchi S, Alonzo TA, Stirewalt DL, et al.: Clinical implications of FLT3 mutations in pediatric AML. Blood 108 (12): 3654-61, 2006. [PUBMED Abstract]
  171. 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]
  172. 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]
  173. 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]
  174. 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]
  175. Mrózek K, Heerema NA, Bloomfield CD: Cytogenetics in acute leukemia. Blood Rev 18 (2): 115-36, 2004. [PUBMED Abstract]
  176. 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]
  177. 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]
  178. 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]
  179. 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]
  180. 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]
  181. 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]
  182. 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]
  183. 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]
  184. 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]
  185. 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]
  186. 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 January 31, 2019.
  187. 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]
  188. 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]
  189. 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]
  190. 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]
  191. 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]
  192. 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]
  193. 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]
  194. 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]
  195. 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]
  196. 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]
  197. 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]
  198. 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]
  199. 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]
  200. 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]
  201. 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]
  202. 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]
  203. 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]
  204. 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]
  205. 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]
  206. 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]
  207. 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]
  208. 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]
  209. 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]
  210. 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]
  211. 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]
  212. 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]
  213. 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]
  214. 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]
  215. 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]
  216. 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]
  217. 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]
  218. 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]
  219. 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]
  220. 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]
  221. 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]
  222. 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]
  223. 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]
  224. 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]
  225. 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]
  226. 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]
  227. 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]
  228. 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]
  229. 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]
  230. 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]
  231. 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]
  232. 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]
  233. 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]
  234. 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]
  235. 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]
  236. 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]
  237. 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]
  238. 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]
  239. 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]
  240. 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]
  241. 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]
  242. 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]
  243. 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]
  244. 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]
  245. 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]
  246. 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]
  247. 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]
  248. 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]
  249. 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]
  250. 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]
  251. 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]
  252. 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]
  253. 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]
  254. 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]
  255. 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]
  256. 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]
  257. 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]
  258. 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]
  259. 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]
  260. 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]
  261. 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]
  262. 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]
  263. 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]
  264. 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]
  265. 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]
  266. 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]
  267. 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]
  268. 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]
  269. 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]
  270. 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]
  271. 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]
  272. 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]
  273. 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]
  274. Lion T, Haas OA: Acute megakaryocytic leukemia with the t(1;22)(p13;q13). Leuk Lymphoma 11 (1-2): 15-20, 1993. [PUBMED Abstract]
  275. 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]
  276. 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]
  277. 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]
  278. 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]
  279. 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]
  280. 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]
  281. 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]
  282. 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]
  283. 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]
  284. 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]
  285. 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]
  286. 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]
  287. 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]
  288. 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]
  289. 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]
  290. 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]
  291. Takeda A, Yaseen NR: Nucleoporins and nucleocytoplasmic transport in hematologic malignancies. Semin Cancer Biol 27: 3-10, 2014. [PUBMED Abstract]
  292. 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]
  293. 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]
  294. 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]
  295. 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]
  296. 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]
  297. 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]
  298. 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]
  299. 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]
  300. 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]
  301. 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]
  302. 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]
  303. 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]
  304. 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]
  305. 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]
  306. 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]
  307. 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]
  308. 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]
  309. 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]
  310. 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]
  311. 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]
  312. 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]
  313. 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]
  314. 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]
  315. 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]
  316. 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]
  317. 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]
  318. 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]
  319. 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]
  320. 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]
  321. 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]
  322. 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]
  323. 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]
  324. 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]
  325. 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]
  326. 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]
  327. Dang L, White DW, Gross S, et al.: Cancer-associated IDH1 mutations produce 2-hydroxyglutarate. Nature 462 (7274): 739-44, 2009. [PUBMED Abstract]
  328. 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]
  329. 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]
  330. 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]
  331. 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]
  332. Andersson AK, Miller DW, Lynch JA, et al.: IDH1 and IDH2 mutations in pediatric acute leukemia. Leukemia 25 (10): 1570-7, 2011. [PUBMED Abstract]
  333. 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]
  334. Germeshausen M, Kratz CP, Ballmaier M, et al.: RAS and CSF3R mutations in severe congenital neutropenia. Blood 114 (16): 3504-5, 2009. [PUBMED Abstract]
  335. 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]
  336. Caye A, Strullu M, Guidez F, et al.: Juvenile myelomonocytic leukemia displays mutations in components of the RAS pathway and the PRC2 network. Nat Genet 47 (11): 1334-40, 2015. [PUBMED Abstract]
  337. Stieglitz E, Taylor-Weiner AN, Chang TY, et al.: The genomic landscape of juvenile myelomonocytic leukemia. Nat Genet 47 (11): 1326-33, 2015. [PUBMED Abstract]
  338. Murakami N, Okuno Y, Yoshida K, et al.: Integrated molecular profiling of juvenile myelomonocytic leukemia. Blood 131 (14): 1576-1586, 2018. [PUBMED Abstract]
  339. Sakaguchi H, Okuno Y, Muramatsu H, et al.: Exome sequencing identifies secondary mutations of SETBP1 and JAK3 in juvenile myelomonocytic leukemia. Nat Genet 45 (8): 937-41, 2013. [PUBMED Abstract]
  340. Stieglitz E, Mazor T, Olshen AB, et al.: Genome-wide DNA methylation is predictive of outcome in juvenile myelomonocytic leukemia. Nat Commun 8 (1): 2127, 2017. [PUBMED Abstract]
  341. Helsmoortel HH, Bresolin S, Lammens T, et al.: LIN28B overexpression defines a novel fetal-like subgroup of juvenile myelomonocytic leukemia. Blood 127 (9): 1163-72, 2016. [PUBMED Abstract]
  342. Schwartz JR, Ma J, Lamprecht T, et al.: The genomic landscape of pediatric myelodysplastic syndromes. Nat Commun 8 (1): 1557, 2017. [PUBMED Abstract]
  343. Pastor V, Hirabayashi S, Karow A, et al.: Mutational landscape in children with myelodysplastic syndromes is distinct from adults: specific somatic drivers and novel germline variants. Leukemia 31 (3): 759-762, 2017. [PUBMED Abstract]
  344. Collin M, Dickinson R, Bigley V: Haematopoietic and immune defects associated with GATA2 mutation. Br J Haematol 169 (2): 173-87, 2015. [PUBMED Abstract]
  345. Wlodarski MW, Hirabayashi S, Pastor V, et al.: Prevalence, clinical characteristics, and prognosis of GATA2-related myelodysplastic syndromes in children and adolescents. Blood 127 (11): 1387-97; quiz 1518, 2016. [PUBMED Abstract]
  346. Wlodarski MW, Collin M, Horwitz MS: GATA2 deficiency and related myeloid neoplasms. Semin Hematol 54 (2): 81-86, 2017. [PUBMED Abstract]
  347. Davidsson J, Puschmann A, Tedgård U, et al.: SAMD9 and SAMD9L in inherited predisposition to ataxia, pancytopenia, and myeloid malignancies. Leukemia 32 (5): 1106-1115, 2018. [PUBMED Abstract]
  348. Narumi S, Amano N, Ishii T, et al.: SAMD9 mutations cause a novel multisystem disorder, MIRAGE syndrome, and are associated with loss of chromosome 7. Nat Genet 48 (7): 792-7, 2016. [PUBMED Abstract]
  349. Chen DH, Below JE, Shimamura A, et al.: Ataxia-Pancytopenia Syndrome Is Caused by Missense Mutations in SAMD9L. Am J Hum Genet 98 (6): 1146-1158, 2016. [PUBMED Abstract]
  350. Wong JC, Bryant V, Lamprecht T, et al.: Germline SAMD9 and SAMD9L mutations are associated with extensive genetic evolution and diverse hematologic outcomes. JCI Insight 3 (14): , 2018. [PUBMED Abstract]

No hay comentarios:

Publicar un comentario