lunes, 26 de agosto de 2019

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

Childhood Acute Lymphoblastic Leukemia Treatment (PDQ®)–Health Professional Version - National Cancer Institute

National Cancer Institute



Childhood Acute Lymphoblastic Leukemia Treatment (PDQ®)–Health Professional Version



World Health Organization (WHO) Classification System for Childhood ALL

The 2016 revision to the WHO classification of tumors of the hematopoietic and lymphoid tissues lists the following entities for acute lymphoid leukemias:[1]

2016 WHO Classification of B-Lymphoblastic Leukemia/Lymphoma

  • B-lymphoblastic leukemia/lymphoma, not otherwise specified (NOS).
  • B-lymphoblastic leukemia/lymphoma with recurrent genetic abnormalities.
  • B-lymphoblastic leukemia/lymphoma with t(9;22)(q34.1;q11.2); BCR-ABL1.
  • B-lymphoblastic leukemia/lymphoma with t(v;11q23.3); KMT2A rearranged.
  • B-lymphoblastic leukemia/lymphoma with t(12;21)(p13.2;q22.1); ETV6-RUNX1.
  • B-lymphoblastic leukemia/lymphoma with hyperdiploidy.
  • B-lymphoblastic leukemia/lymphoma with hypodiploidy.
  • B-lymphoblastic leukemia/lymphoma with t(5;14)(q31.1;q32.3); IL3-IGH.
  • B-lymphoblastic leukemia/lymphoma with t(1;19)(q23;p13.3); TCF3-PBX1.
  • Provisional entity: B-lymphoblastic leukemia/lymphoma, BCR-ABL1–like.
  • Provisional entity: B-lymphoblastic leukemia/lymphoma with iAMP21.

2016 WHO Classification of T-Lymphoblastic Leukemia/Lymphoma

  • Provisional entity: Early T-cell precursor lymphoblastic leukemia.

2016 WHO Classification of Acute Leukemias of Ambiguous Lineage

For acute leukemias of ambiguous lineage, the group of acute leukemias that have characteristics of both acute myeloid leukemia (AML) and acute lymphoblastic leukemia (ALL), the WHO classification system is summarized in Table 1.[2,3] The criteria for lineage assignment for a diagnosis of mixed phenotype acute leukemia (MPAL) are provided in Table 2.[1]
Table 1. Acute Leukemias of Ambiguous Lineage According to the World Health Organization Classification of Tumors of Hematopoietic and Lymphoid Tissuesa
ConditionDefinition
NOS = not otherwise specified.
aBéné MC: Biphenotypic, bilineal, ambiguous or mixed lineage: strange leukemias! Haematologica 94 (7): 891-3, 2009.[2] Obtained from Haematologica/the Hematology Journal website http://www.haematologica.orgExit Disclaimer.
Acute undifferentiated leukemiaAcute leukemia that does not express any marker considered specific for either lymphoid or myeloid lineage
Mixed phenotype acute leukemia with t(9;22)(q34;q11.2); BCR-ABL1Acute leukemia meeting the diagnostic criteria for mixed phenotype acute leukemia in which the blasts also have the (9;22) translocation or the BCR-ABL1 rearrangement
Mixed phenotype acute leukemia with t(v;11q23); KMT2A (MLL) rearrangedAcute leukemia meeting the diagnostic criteria for mixed phenotype acute leukemia in which the blasts also have a translocation involving the KMT2A gene
Mixed phenotype acute leukemia, B/myeloid, NOSAcute leukemia meeting the diagnostic criteria for assignment to both B and myeloid lineage, in which the blasts lack genetic abnormalities involving BCR-ABL1 or KMT2A
Mixed phenotype acute leukemia, T/myeloid, NOSAcute leukemia meeting the diagnostic criteria for assignment to both T and myeloid lineage, in which the blasts lack genetic abnormalities involving BCR-ABL1 or KMT2A
Mixed phenotype acute leukemia, B/myeloid, NOS—rare typesAcute leukemia meeting the diagnostic criteria for assignment to both B and T lineage
Other ambiguous lineage leukemiasNatural killer–cell lymphoblastic leukemia/lymphoma
Table 2. Lineage Assignment Criteria for Mixed Phenotype Acute Leukemia According to the 2016 Revision to the World Health Organization Classification of Myeloid Neoplasms and Acute Leukemiaa
LineageCriteria
aAdapted from Arber et al.[1]
bStrong defined as equal to or brighter than the normal B or T cells in the sample.
Myeloid lineageMyeloperoxidase (flow cytometry, immunohistochemistry, or cytochemistry); ormonocytic differentiation (at least two of the following: nonspecific esterase cytochemistry, CD11c, CD14, CD64, lysozyme)
T lineageStrongb cytoplasmic CD3 (with antibodies to CD3 epsilon chain); or surface CD3
B lineageStrongb CD19 with at least one of the following strongly expressed: CD79a, cytoplasmic CD22, or CD10; or weak CD19 with at least two of the following strongly expressed: CD79a, cytoplasmic CD22, or CD10
Leukemias of mixed phenotype may be seen in various presentations, including the following:
  1. Bilineal leukemias in which there are two distinct populations of cells, usually one lymphoid and one myeloid.
  2. Biphenotypic leukemias in which individual blast cells display features of both lymphoid and myeloid lineage.
Biphenotypic cases represent the majority of mixed phenotype leukemias.[4] Patients with B-myeloid biphenotypic leukemias lacking the TEL-AML1 fusion have lower rates of complete remission (CR) and significantly worse event-free survival (EFS) rates compared with patients with precursor B-cell ALL.[4] Some studies suggest that patients with biphenotypic leukemia may fare better with a lymphoid, as opposed to a myeloid, treatment regimen.[5-8] A large retrospective study from the international Berlin-Frankfurt-Münster (BFM) group demonstrated that initial therapy with an ALL-type regimen was associated with a superior outcome compared with AML-type or combined ALL/AML regimens, particularly in cases with CD19 positivity or other lymphoid antigen expression. In this study, hematopoietic stem cell transplantation (HSCT) in first CR was not beneficial, with the possible exception of cases with morphologic evidence of persistent marrow disease (≥5% blasts) after the first month of treatment.[8]
Key clinical and biological characteristics, as well as the prognostic significance for these entities, are discussed in the Cytogenetics/Genomics of Childhood ALL section of this summary.
References
  1. 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]
  2. Béné MC: Biphenotypic, bilineal, ambiguous or mixed lineage: strange leukemias! Haematologica 94 (7): 891-3, 2009. [PUBMED Abstract]
  3. Borowitz MJ, Béné MC, Harris NL: Acute leukaemias of ambiguous lineage. In: Swerdlow SH, Campo E, Harris NL, et al., eds.: WHO Classification of Tumours of Haematopoietic and Lymphoid Tissues. 4th ed. Lyon, France: International Agency for Research on Cancer, 2008, pp 150-5.
  4. Gerr H, Zimmermann M, Schrappe M, et al.: Acute leukaemias of ambiguous lineage in children: characterization, prognosis and therapy recommendations. Br J Haematol 149 (1): 84-92, 2010. [PUBMED Abstract]
  5. Rubnitz JE, Onciu M, Pounds S, et al.: Acute mixed lineage leukemia in children: the experience of St Jude Children's Research Hospital. Blood 113 (21): 5083-9, 2009. [PUBMED Abstract]
  6. Al-Seraihy AS, Owaidah TM, Ayas M, et al.: Clinical characteristics and outcome of children with biphenotypic acute leukemia. Haematologica 94 (12): 1682-90, 2009. [PUBMED Abstract]
  7. Matutes E, Pickl WF, Van't Veer M, et al.: Mixed-phenotype acute leukemia: clinical and laboratory features and outcome in 100 patients defined according to the WHO 2008 classification. Blood 117 (11): 3163-71, 2011. [PUBMED Abstract]
  8. Hrusak O, de Haas V, Stancikova J, et al.: International cooperative study identifies treatment strategy in childhood ambiguous lineage leukemia. Blood 132 (3): 264-276, 2018. [PUBMED Abstract]

Cytogenetics/Genomics of Childhood ALL

The genomics of childhood ALL has been extensively investigated and multiple distinctive subtypes based on cytogenetic and molecular characterizations have been defined, each with its own pattern of clinical and prognostic characteristics.[1] Figure 2 illustrates the distribution of ALL cases by cytogenetic/molecular subtype.[1]
ENLARGEPie chart showing subclassification of childhood ALL.
Figure 2. Subclassification of childhood ALL. Blue wedges refer to B-progenitor ALL, yellow to recently identified subtypes of B-ALL, and red wedges to T-lineage ALL. Reprinted from Seminars in HematologyExit Disclaimer, Volume 50, Charles G. Mullighan, Genomic Characterization of Childhood Acute Lymphoblastic Leukemia, Pages 314–324, Copyright (2013), with permission from Elsevier.
The genomic landscape of precursor B-cell ALL is typified by a range of genomic alterations that disrupt normal B-cell development and in some cases by mutations in genes that provide a proliferation signal (e.g., activating mutations in RAS family genes or mutations/translocations leading to kinase pathway signaling). Genomic alterations leading to blockage of B-cell development include translocations (e.g., TCF3-PBX1 and ETV6-RUNX1), point mutations (e.g., IKZF1 and PAX5), and intragenic/intergenic deletions (e.g., IKZF1PAX5EBF, and ERG).[2]
The genomic alterations in precursor B-cell ALL tend not to occur at random, but rather to cluster within subtypes that can be delineated by biological characteristics such as their gene expression profiles. Cases with recurring chromosomal translocations (e.g., TCF3-PBX1 and ETV6-RUNX1, and MLL (KMT2A)-rearranged ALL) have distinctive biological features and illustrate this point, as do the examples below of specific genomic alterations within distinctive biological subtypes:
  • IKZF1 deletions and mutations are most commonly observed within cases of Philadelphia chromosome–positive (Ph+) ALL and Ph-like ALL.[3,4]
  • Intragenic ERG deletions occur within a distinctive subtype characterized by this alteration and lacking other recurring cytogenetic alterations associated with pediatric precursor B-cell ALL.[5-7]
  • TP53 mutations occur at high frequency in patients with low hypodiploid ALL with 32 to 39 chromosomes, and the TP53 mutations in these patients are often germline.[8TP53mutations are uncommon in other patients with precursor B-cell ALL.
Activating point mutations in kinase genes are uncommon in high-risk precursor B-cell ALL, and JAK genes are the primary kinases that are found to be mutated. These mutations are generally observed in patients with Ph-like ALL that have CRLF2 abnormalities, although JAK2 mutations are also observed in approximately 15% of children with Down syndrome ALL.[4,9,10] Several kinase genes and cytokine receptor genes are activated by translocation as described below in the discussion of Ph-positive ALL and Ph-like ALL. FLT3mutations occur in a minority of cases (approximately 10%) of hyperdiploid ALL and MLL(KMT2A)-rearranged ALL, and are rare in other subtypes.[11]
Understanding of the genomics of precursor B-cell ALL at relapse is less advanced than understanding of ALL genomics at diagnosis. Childhood ALL is often polyclonal at diagnosis and under the selective influence of therapy, some clones may be extinguished and new clones with distinctive genomic profiles may arise.[12] Of particular importance are new mutations that arise at relapse that may be selected by specific components of therapy. As an example, mutations in NT5C2 are not found at diagnosis whereas specific mutations in NT5C2 were observed in 7 of 44 (16%) and 9 of 20 (45%) cases of precursor B-cell ALL with early relapse that were evaluated for this mutation.[12,13NT5C2 mutations are uncommon in patients with late relapse, and they appear to induce resistance to mercaptopurine and thioguanine.[13] Another gene that is found mutated only at relapse is PRSP1, a gene involved purine biosynthesis.[14] Mutations were observed in 13.0% of a Chinese cohort and 2.7% of a German cohort, and were observed in patients with on-treatment relapses. The PRSP1 mutations observed in relapsed cases induce resistance to thiopurines in leukemia cell lines. CREBBP mutations are also enriched at relapse and appear to be associated with increased resistance to glucocorticoids.[12,15] With increased understanding of the genomics of relapse, it may be possible to tailor upfront therapy to avoid relapse or detect resistance-inducing mutations early and intervene before a frank relapse.
Specific genomic and chromosomal alterations are described below, with a focus on their prognostic significance.
T-cell ALL is characterized by genomic alterations leading to activation of transcriptional programs related to T-cell development and by a high frequency of cases (approximately 60%) with mutations in NOTCH1 and/or FBXW7 that result in activation of the NOTCH1 pathway.[16] In contrast to B-cell ALL, the prognostic significance of T-cell ALL genomic alterations is less well-defined. Cytogenetic abnormalities common in B-lineage ALL (e.g., hyperdiploidy, 51–65 chromosomes) are rare in T-cell ALL.[17,18]

B-cell ALL cytogenetics/genomics

A number of recurrent chromosomal abnormalities have been shown to have prognostic significance, especially in precursor B-cell ALL. Some chromosomal alterations are associated with more favorable outcomes, such as high hyperdiploidy (51–65 chromosomes) and the ETV6-RUNX1 fusion. Others historically have been associated with a poorer prognosis, including the Philadelphia chromosome (t(9;22)(q34;q11.2)), rearrangements of the MLL (KMT2A) gene, hypodiploidy, and intrachromosomal amplification of the AML1 gene (iAMP21).[19]
In recognition of the clinical significance of many of these genomic alterations, the 2016 revision of the World Health Organization classification of tumors of the hematopoietic and lymphoid tissues lists the following entities for precursor B-cell ALL:[20]
  • B-lymphoblastic leukemia/lymphoma, not otherwise specified (NOS).
  • B-lymphoblastic leukemia/lymphoma with recurrent genetic abnormalities.
  • B-lymphoblastic leukemia/lymphoma with t(9;22)(q34.1;q11.2); BCR-ABL1.
  • B-lymphoblastic leukemia/lymphoma with t(v;11q23.3); KMT2A rearranged.
  • B-lymphoblastic leukemia/lymphoma with t(12;21)(p13.2;q22.1); ETV6-RUNX1.
  • B-lymphoblastic leukemia/lymphoma with hyperdiploidy.
  • B-lymphoblastic leukemia/lymphoma with hypodiploidy.
  • B-lymphoblastic leukemia/lymphoma with t(5;14)(q31.1;q32.3); IL3-IGH.
  • B-lymphoblastic leukemia/lymphoma with t(1;19)(q23;p13.3); TCF3-PBX1.
  • Provisional entity: B-lymphoblastic leukemia/lymphoma, BCR-ABL1–like.
  • Provisional entity: B-lymphoblastic leukemia/lymphoma with iAMP21.
These and other chromosomal and genomic abnormalities for childhood ALL are described below.
  1. Chromosome number
    • High hyperdiploidy (51–65 chromosomes)
      High hyperdiploidy, defined as 51 to 65 chromosomes per cell or a DNA index greater than 1.16, occurs in 20% to 25% of cases of precursor B-cell ALL, but very rarely in cases of T-cell ALL.[21] Hyperdiploidy can be evaluated by measuring the DNA content of cells (DNA index) or by karyotyping. In cases with a normal karyotype or in which standard cytogenetic analysis was unsuccessful, interphase fluorescence in situ hybridization (FISH) may detect hidden hyperdiploidy. High hyperdiploidy generally occurs in cases with clinically favorable prognostic factors (patients aged 1 to <10 years with a low white blood cell [WBC] count) and is itself an independent favorable prognostic factor.[21-23] Within the hyperdiploid range of 51 to 65 chromosomes, patients with higher modal numbers (58–66) appeared to have a better prognosis in one study.[23] Hyperdiploid leukemia cells are particularly susceptible to undergoing apoptosis and accumulate higher levels of methotrexate and its active polyglutamate metabolites,[24] which may explain the favorable outcome commonly observed in these cases.
      While the overall outcome of patients with high hyperdiploidy is considered to be favorable, factors such as age, WBC count, specific trisomies, and early response to treatment have been shown to modify its prognostic significance.[25,26]
      Patients with trisomies of chromosomes 4, 10, and 17 (triple trisomies) have been shown to have a particularly favorable outcome as demonstrated by both Pediatric Oncology Group (POG) and Children's Cancer Group analyses of National Cancer Institute (NCI) standard-risk ALL.[27] POG data suggest that NCI standard-risk patients with trisomies of 4 and 10, without regard to chromosome 17 status, have an excellent prognosis.[28]
      Chromosomal translocations may be seen with high hyperdiploidy, and in those cases, patients are more appropriately risk-classified based on the prognostic significance of the translocation. For instance, in one study, 8% of patients with the Philadelphia chromosome (t(9;22)(q34;q11.2)) also had high hyperdiploidy,[29] and the outcome of these patients (treated without tyrosine kinase inhibitors) was inferior to that observed in non-Philadelphia chromosome–positive (Ph+) high hyperdiploid patients.
      Certain patients with hyperdiploid ALL may have a hypodiploid clone that has doubled (masked hypodiploidy).[30] These cases may be interpretable based on the pattern of gains and losses of specific chromosomes. These patients have an unfavorable outcome, similar to those with hypodiploidy.[30]
      Near triploidy (68–80 chromosomes) and near tetraploidy (>80 chromosomes) are much less common and appear to be biologically distinct from high hyperdiploidy.[31] Unlike high hyperdiploidy, a high proportion of near tetraploid cases harbor a cryptic ETV6-RUNX1 fusion.[31-33] Near triploidy and tetraploidy were previously thought to be associated with an unfavorable prognosis, but later studies suggest that this may not be the case.[31,33]
      The genomic landscape of hyperdiploid ALL is represented by mutations in genes of the receptor tyrosine kinase (RTK)/RAS pathway in approximately one-half of cases. Genes encoding histone modifiers are also present in a recurring manner in a minority of cases. Analysis of mutation profiles demonstrates that chromosomal gains are early events in the pathogenesis of hyperdiploid ALL.[34]
    • Hypodiploidy (<44 chromosomes)
      Precursor B-cell ALL cases with fewer than the normal number of chromosomes have been subdivided in various ways, with one report stratifying based on modal chromosome number into the following four groups:[30]
      • Near-haploid: 24 to 29 chromosomes (n = 46).
      • Low-hypodiploid: 33 to 39 chromosomes (n = 26).
      • High-hypodiploid: 40 to 43 chromosomes (n = 13).
      • Near-diploid: 44 chromosomes (n = 54).
      Most patients with hypodiploidy are in the near-haploid and low-hypodiploid groups, and both of these groups have an elevated risk of treatment failure compared with nonhypodiploid cases.[30,35] Patients with fewer than 44 chromosomes have a worse outcome than do patients with 44 or 45 chromosomes in their leukemic cells.[30] A number of studies have shown that patients with high minimal residual disease (MRD) (≥0.01%) after induction do very poorly, with 5-year event-free survival (EFS) rates ranging from 25% to 47%. Although hypodiploid patients with low MRD after induction fare better (5-year EFS, 64%–75%), their outcomes are still inferior to most children with other types of ALL.[36-38]
      The recurring genomic alterations of near-haploid and low-hypodiploid ALL appear to be distinctive from each other and from other types of ALL.[8] In near-haploid ALL, alterations targeting RTK signaling, RAS signaling, and IKZF3are common.[39] In low-hypodiploid ALL, genetic alterations involving TP53RB1, and IKZF2 are common. Importantly, the TP53 alterations observed in low-hypodiploid ALL are also present in nontumor cells in approximately 40% of cases, suggesting that these mutations are germline and that low-hypodiploid ALL represents, in some cases, a manifestation of Li-Fraumeni syndrome.[8]
      Approximately two-thirds of patients with ALL and germline pathogenic TP53variants have hypodiploid ALL.[40]
  2. Chromosomal translocations and gains/deletions of chromosomal segments
    • t(12;21)(p13.2;q22.1); ETV6-RUNX1 (formerly known as TEL-AML1)
      Fusion of the ETV6 gene on chromosome 12 to the RUNX1 gene on chromosome 21 is present in 20% to 25% of cases of precursor B-cell ALL but is rarely observed in T-cell ALL.[32] The t(12;21)(p12;q22) produces a cryptic translocation that is detected by methods such as FISH, rather than conventional cytogenetics, and it occurs most commonly in children aged 2 to 9 years.[41,42] Hispanic children with ALL have a lower incidence of t(12;21)(p13;q22) than do white children.[43]
      Reports generally indicate favorable EFS and overall survival (OS) in children with the ETV6-RUNX1 fusion; however, the prognostic impact of this genetic feature is modified by the following factors:[44-48]
      • Early response to treatment.
      • NCI risk category (age and WBC count at diagnosis).
      • Treatment regimen.
      In one study of the treatment of newly diagnosed children with ALL, multivariate analysis of prognostic factors found age and leukocyte count, but not ETV6-RUNX1, to be independent prognostic factors.[44] It does not appear that the presence of secondary cytogenetic abnormalities, such as deletion of ETV6 (12p) or CDKN2A/B (9p), impacts the outcome of patients with the ETV6-RUNX1 fusion.[48,49] There is a higher frequency of late relapses in patients with ETV6-RUNX1 fusion compared with other precursor B-cell ALL.[44,50] Patients with the ETV6-RUNX1 fusion who relapse seem to have a better outcome than other relapse patients,[51] with an especially favorable prognosis for patients who relapse more than 36 months from diagnosis.[52] Some relapses in patients with t(12;21)(p13;q22) may represent a new independent second hit in a persistent preleukemic clone (with the first hit being the ETV6-RUNX1 translocation).[53,54]
    • t(9;22)(q34.1;q11.2); BCR-ABL1 (Ph+)
      The Philadelphia chromosome t(9;22)(q34;q11.2) is present in approximately 3% of children with ALL and leads to production of a BCR-ABL1 fusion protein with tyrosine kinase activity (refer to Figure 3).
      ENLARGEPhiladelphia chromosome; three-panel drawing shows a piece of chromosome 9 and a piece of chromosome 22 breaking off and trading places, creating a changed chromosome 22 called the Philadelphia chromosome. In the left panel, the drawing shows a normal chromosome 9 with the ABL gene and a normal chromosome 22 with the BCR gene. In the center panel, the drawing shows chromosome 9 breaking apart in the ABL gene and chromosome 22 breaking apart below the BCR gene. In the right panel, the drawing shows chromosome 9 with the piece from chromosome 22 attached and chromosome 22 with the piece from chromosome 9 containing part of the ABL gene attached. The changed chromosome 22 with the BCR-ABLgene is called the Philadelphia chromosome.
      Figure 3. The Philadelphia chromosome is a translocation between the ABL-1 oncogene (on the long arm of chromosome 9) and the breakpoint cluster region (BCR) (on the long arm of chromosome 22), resulting in the fusion gene BCR-ABL1BCR-ABL1 encodes an oncogenic protein with tyrosine kinase activity.
      This subtype of ALL is more common in older children with precursor B-cell ALL and high WBC count, with the incidence of the t(9;22)(q34;q11.2) increasing to about 25% in young adults with ALL.
      Historically, the Philadelphia chromosome t(9;22)(q34;q11.2) was associated with an extremely poor prognosis (especially in those who presented with a high WBC count or had a slow early response to initial therapy), and its presence had been considered an indication for allogeneic hematopoietic stem cell transplantation (HSCT) in patients in first remission.[29,55-57] Inhibitors of the BCR-ABL1 tyrosine kinase, such as imatinib mesylate, are effective in patients with Ph+ ALL.[58] A study by the Children's Oncology Group (COG), which used intensive chemotherapy and concurrent imatinib mesylate given daily, demonstrated a 5-year EFS rate of 70% ± 12%, which was superior to the EFS rate of historical controls in the pre-tyrosine kinase inhibitor (imatinib mesylate) era.[59,60]
    • t(v;11q23.3); MLL (KMT2A)-rearranged
      Rearrangements involving the MLL (KMT2A) gene occur in approximately 5% of childhood ALL cases overall, but in up to 80% of infants with ALL. These rearrangements are generally associated with an increased risk of treatment failure.[61-64] The t(4;11)(q21;q23) is the most common rearrangement involving the MLL gene in children with ALL and occurs in approximately 1% to 2% of childhood ALL.[62,65]
      Patients with the t(4;11)(q21;q23) are usually infants with high WBC counts; they are more likely than other children with ALL to have central nervous system (CNS) disease and to have a poor response to initial therapy.[66] While both infants and adults with the t(4;11)(q21;q23) are at high risk of treatment failure, children with the t(4;11)(q21;q23) appear to have a better outcome than either infants or adults.[61,62] Irrespective of the type of MLL (KMT2A) gene rearrangement, infants with leukemia cells that have MLL gene rearrangements have a worse treatment outcome than older patients whose leukemia cells have an MLL gene rearrangement.[61,62] Whole-genome sequencing has determined that cases of infant ALL with MLL gene rearrangements have few additional genomic alterations, none of which have clear clinical significance.[11] Deletion of the MLL gene has not been associated with an adverse prognosis.[67]
      Of interest, the t(11;19)(q23;p13.3) involving MLL (KMT2A) and MLLT1/ENLoccurs in approximately 1% of ALL cases and occurs in both early B-lineage and T-cell ALL.[68] Outcome for infants with the t(11;19) is poor, but outcome appears relatively favorable in older children with T-cell ALL and the t(11;19).[68]
    • t(1;19)(q23;p13.3); TCF3-PBX1 and t(17;19)(q22;p13); TCF3-HLF
      The t(1;19) occurs in approximately 5% of childhood ALL cases and involves fusion of the TCF3 gene on chromosome 19 to the PBX1 gene on chromosome 1.[69,70] The t(1;19) may occur as either a balanced translocation or as an unbalanced translocation and is the primary recurring genomic alteration of the pre-B ALL immunophenotype (cytoplasmic Ig positive).[71] Black children are relatively more likely than white children to have pre-B ALL with the t(1;19).[72]
      The t(1;19) had been associated with inferior outcome in the context of antimetabolite-based therapy,[73] but the adverse prognostic significance was largely negated by more aggressive multiagent therapies.[70,74] However, in a trial conducted by St. Jude Children's Research Hospital (SJCRH) on which all patients were treated without cranial radiation, patients with the t(1;19) had an overall outcome comparable to children lacking this translocation, with a higher risk of CNS relapse and a lower rate of bone marrow relapse, suggesting that more intensive CNS therapy may be needed for these patients.[75,76]
      The t(17;19) resulting in the TCF3-HLF fusion occurs in less than 1% of pediatric ALL cases. ALL with the TCF3-HLF fusion is associated at diagnosis with disseminated intravascular coagulation and with hypercalcemia. Outcome is very poor for children with the t(17;19), with a literature review noting mortality for 20 of 21 cases reported.[77] In addition to the TCF3-HLF fusion, the genomic landscape of this ALL subtype was characterized by deletions in genes involved in B-cell development (PAX5BTG1, and VPREB1) and by mutations in RAS pathway genes (NRASKRAS, and PTPN11).[71]
    • DUX4-rearranged ALL with frequent ERG deletions
      Approximately 5% of standard-risk and 10% of high-risk pediatric precursor B-cell ALL patients have a rearrangement involving DUX4 that leads to its overexpression.[78,79] The frequency in older adolescents (aged >15 years) is approximately 10%. The most common rearrangement produces IGH-DUX4fusions, with ERG-DUX4 fusions also observed. Approximately 50% of DUX4-rearranged cases have focal intragenic deletions involving ERG that are not observed in other ALL subtypes,[78,79] and DUX4-rearranged cases show a distinctive gene expression pattern that was initially identified as being associated with these focal deletions in ERG.[5-7IKZF1 alterations are observed in 35% to 40% of DUX4-rearranged ALL.[78,79ERG deletion connotes an excellent prognosis, with OS exceeding 90%; even when the IZKF1 deletion is present, prognosis remains highly favorable.[5-7] Patients with DUX4rearrangements who lack ERG deletion also appear to have favorable prognosis.[79]
    • MEF2D-rearranged ALL
      Gene fusions involving MEF2D, a transcription factor that is expressed during B-cell development, are observed in approximately 4% of childhood ALL cases.[80,81] Although multiple fusion partners may occur, most cases involve BCL9, which is located on chromosome 1q21, as is MEF2D.[80,82] The interstitial deletion producing the MEF2D-BCL9 fusion is too small to be detected by conventional cytogenetic methods. Cases with MEF2D gene fusions show a distinctive gene expression profile, except for rare cases with MEF2D-CSFR1 that have a Philadelphia chromosome (Ph)–like gene expression profile.[80,83] The median age at diagnosis for cases of MEF2D-rearranged ALL in studies that included both adult and pediatric patients was 12 to 14 years.[80,81] For 22 children with MEF2D-rearranged ALL enrolled in a high-risk ALL clinical trial, the 5-year EFS was 72% (standard error, ±10%), which was inferior to that for other patients.[80]
    • ZNF384-rearranged ALL
      ZNF384 is a transcription factor that is rearranged in approximately 4% to 5% of pediatric B-cell ALL cases.[80,84,85] Multiple fusion partners for ZNF384 have been reported, including ARID1BCREBBPEP300SMARCA2TAF15, and TCF3. Regardless of the fusion partner, ZNF384-rearranged ALL cases show a distinctive gene expression profile.[80,84,85ZNF384 rearrangement does not appear to confer independent prognostic significance.[80,84,85] The immunophenotype of B-cell ALL with ZNF384 rearrangement is characterized by weak or negative CD10 expression, with expression of CD13 and/or CD33 commonly observed.[84,85] Cases of mixed phenotype B/myeloid acute leukemia that have ZNF384 gene fusions have been reported, but it is unclear whether the clinical behavior of these cases is the same as that of ZNF384-rearranged B-cell ALL.[86,87]
    • t(5;14)(q31.1;q32.3); IL3-IGH
      This entity is included in the 2016 revision of the WHO classification of tumors of the hematopoietic and lymphoid tissues.[20] The finding of t(5;14)(q31.1;q32.3) in patients with ALL and hypereosinophilia in the 1980s was followed by the identification of the IL3-IGH fusion as the underlying genetic basis for the condition.[88,89] The joining of the IGH locus to the promoter region of the interleukin-3 gene (IL3) leads to dysregulation of IL3 expression.[90] Cytogenetic abnormalities in children with ALL and eosinophilia are variable, with only a subset resulting from the IL3-IGH fusion.[91]
      The number of cases of IL3-IGH ALL described in the published literature is too small to assess the prognostic significance of the IL3-IGH fusion.
    • Intrachromosomal amplification of chromosome 21 (iAMP21)
      iAMP21 with multiple extra copies of the RUNX1 (AML1) gene at 21q22 occurs in approximately 2% of precursor B-cell ALL cases and is associated with older age (median, approximately 10 years), presenting WBC of less than 50 × 109/L, a slight female preponderance, and high end-induction MRD.[92-94]
      The United Kingdom (UK)–ALL clinical trials group initially reported that the presence of iAMP21 conferred a poor prognosis in patients treated in the MRC ALL 97/99 trial (5-year EFS, 29%).[19] In their subsequent trial (UKALL2003 [NCT00222612]), patients with iAMP21 were assigned to a more intensive chemotherapy regimen and had a markedly better outcome (5-year EFS, 78%).[93] Similarly, the COG has reported that iAMP21 was associated with a significantly inferior outcome in NCI standard-risk patients (4-year EFS, 73% for iAMP21 vs. 92% in others), but not in NCI high-risk patients (4-year EFS, 73% vs. 80%).[92] On multivariate analysis, iAMP21 was an independent predictor of inferior outcome only in NCI standard-risk patients.[92] The results of the UKALL2003 and COG studies suggest that treatment of iAMP21 patients with high-risk chemotherapy regimens abrogates its adverse prognostic significance and obviates the need for SCT in first remission.[94]
    • Amplification of PAX5
      PAX5 amplification was identified in approximately 1% of B-cell ALL cases, and it was usually detected in cases lacking known leukemia-driver genomic alterations.[95] Cases with PAX5 amplification show male predominance (66%), with most (55%) having NCI high-risk status. For a cohort of patients with PAX5amplification diagnosed between 1993 and 2015, the 5-year EFS rate was 49% (95% confidence interval [CI], 36%–61%), and the OS rate was 67% (95% CI, 54%–77%), suggesting a relatively poor prognosis for this B-cell ALL subtype.
    • BCR-ABL1–like (Ph-like)
      BCR-ABL1–negative patients with a gene expression profile similar to BCR-ABL1–positive patients have been referred to as BCR-ABL1–like.[96-98] This occurs in 10% to 20% of pediatric ALL patients, increasing in frequency with age, and has been associated with IKZF1 deletion or mutation.[9,96,97,99,100]
      Retrospective analyses have indicated that patients with BCR-ABL1–like ALL have a poor prognosis.[4,96] In one series, the 5-year EFS for NCI high-risk children and adolescents with BCR-ABL1–like ALL was 58% and 41%, respectively.[4] While it is more frequent in older and higher-risk patients, the BCR-ABL1–like subtype has also been identified in NCI standard-risk patients. In a COG study, 13.6% of 1,023 NCI standard-risk B-cell ALL patients were found to have BCR-ABL1–like ALL; these patients had an inferior EFS compared with non-BCR-ABL1–like standard-risk patients (82% vs. 91%), although no difference in overall survival (93% vs. 96%) was noted.[101] In one study of 40 BCR-ABL1–like patients, the adverse prognostic significance of this subtype appeared to be abrogated when patients were treated with risk-directed therapy on the basis of MRD levels.[102]
      The hallmark of BCR-ABL1–like ALL is activated kinase signaling, with 50% containing CRLF2 genomic alterations [98,103] and half of those cases containing concomitant JAK mutations.[104] Additional information about BCR-ABL1–like ALL cases with CRLF2 genomic alterations is provided below.
      Many of the remaining cases of BCR-ABL1–like ALL have been noted to have a series of translocations with a common theme of involvement of kinases, including ABL1ABL2CSF1RJAK2, and PDGFRB.[4,99] Fusion proteins from these gene combinations have been noted in some cases to be transformative and have responded to tyrosine kinase inhibitors both in vitro and in vivo,[99] suggesting potential therapeutic strategies for these patients. The prevalence of targetable kinase fusions in BCR-ABL1–like ALL is lower in NCI standard-risk patients (3.5%) than in NCI high-risk patients (approximately 30%).[101] Point mutations in kinase genes, aside from those in JAK1 and JAK2, are uncommon in Ph-like ALL cases.[9]
      Genomic alterations in CRLF2, a cytokine receptor gene located on the pseudoautosomal regions of the sex chromosomes, have been identified in 5% to 10% of cases of precursor B-cell ALL; they represent approximately 50% of cases of BCR-ABL1–like ALL.[105-107] The chromosomal abnormalities that commonly lead to CRLF2 overexpression include translocations of the IgH locus (chromosome 14) to CRLF2 and interstitial deletions in pseudoautosomal regions of the sex chromosomes, resulting in a P2RY8-CRLF2 fusion.[9,103,105,106CRLF2 abnormalities are strongly associated with the presence of IKZF1 deletions and JAK mutations;[9,103,104,106,108] they are also more common in children with Down syndrome.[106] Point mutations in tyrosine kinase genes other than JAK1 and JAK2 are uncommon in CRLF2-overexpressing cases.[9]
      Although the results of several retrospective studies suggest that CRLF2abnormalities may have adverse prognostic significance on univariate analyses, most do not find this abnormality to be an independent predictor of outcome.[103,105,106,109,110] For example, in a large European study, increased expression of CRLF2 was not associated with unfavorable outcome in multivariate analysis, while IKZF1 deletion and BCR-ABL1–like expression signatures were associated with unfavorable outcome.[100] Controversy exists about whether the prognostic significance of CRLF2 abnormalities should be analyzed based on CRLF2 overexpression or on the presence of CRLF2 genomic alterations.[109,110]
      Approximately 9% of BCR-ABL1–like ALL cases result from rearrangements that lead to overexpression of a truncated erythropoietin receptor (EPOR).[111] The C-terminal region of the receptor that is lost is the region that is mutated in primary familial congenital polycythemia and that controls stability of the EPOR. The portion of the EPOR remaining is sufficient for JAK-STAT activation and for driving leukemia development.
    • IKZF1 deletions
      IKZF1 deletions, including deletions of the entire gene and deletions of specific exons, are present in approximately 15% of precursor B-cell ALL cases. Less commonly, IKZF1 can be inactivated by deleterious point mutations.[97] Cases with IKZF1 deletions tend to occur in older children, have a higher WBC count at diagnosis, and are therefore, more common in NCI high-risk patients than in NCI standard-risk patients.[2,97,108,112] A high proportion of BCR-ABL1 cases have a deletion of IKZF1,[3,108] and ALL arising in children with Down syndrome appears to have elevated rates of IKZF1 deletions.[113IKZF1deletions are also common in cases with CRLF2 genomic alterations and in Ph-like (BCR-ABL1–like) ALL (see above).[5,96,108]
      Multiple reports have documented the adverse prognostic significance of an IKZF1 deletion, and most studies have reported that this deletion is an independent predictor of poor outcome on multivariate analyses.[5,96,97,100,108,114-119]; [120][Level of evidence: 2Di] However, the prognostic significance of IKZF1 may not apply equally across ALL biological subtypes, as illustrated by the apparent lack of prognostic significance in patients with ERG deletion.[7] The Associazione Italiana di Ematologia e Oncologia Pediatrica (AIEOP)–Berlin-Frankfurt-Münster (BFM) group reported that IKZF1 deletions were significant adverse prognostic factors only in B-cell ALL patients with high end-induction MRD and in whom co-occurrence of deletions of CDKN2ACDKN2BPAX5, or PAR1 (in the absence of ERG deletion) were identified.[121]
      There are few published results of changing therapy on the basis of IKZF1 gene status. The Malaysia-Singapore group published results of two consecutive trials. In the first trial (MS2003), IKZF1 status was not considered in risk stratification, while in the subsequent trial (MS2010), IKZF1-deleted patients were excluded from the standard-risk group. Thus, more IKZF1-deleted patients in the MS2010 trial received intensified therapy. Patients with IKZF1-deleted ALL had improved outcomes in MS2010 compared with patients in MS2003, but interpretation of this observation is limited by other changes in risk stratification and therapeutic differences between the two trials.[122][Level of evidence: 2A]

T-cell ALL cytogenetics/genomics

Multiple chromosomal translocations have been identified in T-cell ALL that lead to deregulated expression of the target genes. These chromosome rearrangements fuse genes encoding transcription factors (e.g., TAL1/TAL2LMO1 and LMO2LYL1TLX1TLX3NKX2-IHOXA, and MYB) to one of the T-cell receptor loci (or to other genes) and result in deregulated expression of these transcription factors in leukemia cells.[16,17,123-127] These translocations are often not apparent by examining a standard karyotype, but can be identified using more sensitive screening techniques, including fluorescence in situhybridization (FISH) or polymerase chain reaction (PCR).[17] Mutations in a noncoding region near the TAL1 gene that produce a super-enhancer upstream of TAL1 represent nontranslocation genomic alterations that can also activate TAL1 transcription to induce T-cell ALL.[128]
Translocations resulting in chimeric fusion proteins are also observed in T-cell ALL.[129]
  • NUP214-ABL1 fusion has been noted in 4% to 6% of T-cell ALL cases and is observed in both adults and children, with a male predominance.[130-132] The fusion is cytogenetically cryptic and is seen in FISH on amplified episomes or, more rarely, as a small homogeneous staining region.[132] T-cell ALL may also uncommonly show ABL1 fusion proteins with other gene partners (e.g., ETV6BCR, and EML1).[132ABL tyrosine kinase inhibitors, such as imatinib or dasatinib, may demonstrate therapeutic benefits in this T-cell ALL subtype,[130,131,133] although clinical experience with this strategy is very limited.[134-136]
  • Gene fusions involving SPI1 (encoding the transcription factor PU.1) were reported in 4% of Japanese children with T-cell ALL.[137] Fusion partners included STMN1 and TCF7. T-cell ALL cases with SPI1 fusions had a particularly poor prognosis; six of seven affected individuals died within 3 years of diagnosis of early relapse.
  • Other recurring gene fusions in T-cell ALL patients include those involving MLLT10KMT2A, and NUP98.[16]
Notch pathway signaling is commonly activated by NOTCH1 and FBXW7 gene mutations in T-cell ALL, and these are the most commonly mutated genes in pediatric T-cell ALL.[16,138NOTCH1-activating gene mutations occur in approximately 50% to 60% of T-cell ALL cases, and FBXW7-inactivating gene mutations occur in approximately 15% of cases, with the result that approximately 60% of cases have Notch pathway activation by mutations in at least one of these genes.[139]
The prognostic significance of NOTCH1/FBXW7 mutations may be modulated by genomic alterations in RAS and PTEN. The French Acute Lymphoblastic Leukaemia Study Group (FRALLE) and the Group for Research on Adult Acute Lymphoblastic Leukemia groups reported that patients having mutated NOTCH1/FBXW7 and wild-type PTEN/RAS constituted a favorable-risk group while patients with PTEN or RAS mutations, regardless of NOTCH1/FBXW7 status, have a significantly higher risk of treatment failure.[129,140] In the FRALLE study, 5-year cumulative incidence of relapse and disease-free survival (DFS) were 50% and 46% for patients with mutated NOTCH1/FBXW7 and mutated PTEN/RAS versus 13% and 87% for patients with mutated NOTCH1/FBXW7 and wild-type PTEN/RAS.[129] The overall 5-year DFS in the FRALLE study was 73%, and additional research is needed to determine whether the same prognostic significance for NOTCH1/FBXW7 and PTEN/RASmutations will apply to current treatment regimens, which produce overall 5-year DFS rates that approach 90%.

Early T-cell precursor ALL

Detailed molecular characterization of early T-cell precursor ALL showed this entity to be highly heterogeneous at the molecular level, with no single gene affected by mutation or copy number alteration in more than one-third of cases.[141] Compared with other T-cell ALL cases, the early T-cell precursor group had a lower rate of NOTCH1 mutations and significantly higher frequencies of alterations in genes regulating cytokine receptors and RAS signaling, hematopoietic development, and histone modification. The transcriptional profile of early T-cell precursor ALL shows similarities to that of normal hematopoietic stem cells and myeloid leukemia stem cells.[141]
Studies have found that the absence of biallelic deletion of the TCRgamma locus (ABGD), as detected by comparative genomic hybridization and/or quantitative DNA-PCR, was associated with early treatment failure in patients with T-cell ALL.[142,143] ABGD is characteristic of early thymic precursor cells, and many of the T-cell ALL patients with ABGD have an immunophenotype consistent with the diagnosis of early T-cell precursor phenotype.

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.
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]

No hay comentarios:

Publicar un comentario