Skip to main content
  • AACR Publications
    • Blood Cancer Discovery
    • Cancer Discovery
    • Cancer Epidemiology, Biomarkers & Prevention
    • Cancer Immunology Research
    • Cancer Prevention Research
    • Cancer Research
    • Clinical Cancer Research
    • Molecular Cancer Research
    • Molecular Cancer Therapeutics

AACR logo

  • Register
  • Log in
  • My Cart
Advertisement

Main menu

  • Home
  • About
    • The Journal
    • AACR Journals
    • Subscriptions
    • Permissions and Reprints
    • Reviewing
    • CME
  • Articles
    • OnlineFirst
    • Current Issue
    • Past Issues
    • CCR Focus Archive
    • Meeting Abstracts
    • Collections
      • COVID-19 & Cancer Resource Center
      • Breast Cancer
      • Clinical Trials
      • Immunotherapy: Facts and Hopes
      • Editors' Picks
      • "Best of" Collection
  • For Authors
    • Information for Authors
    • Author Services
    • Best of: Author Profiles
    • Submit
  • Alerts
    • Table of Contents
    • Editors' Picks
    • OnlineFirst
    • Citation
    • Author/Keyword
    • RSS Feeds
    • My Alert Summary & Preferences
  • News
    • Cancer Discovery News
  • COVID-19
  • Webinars
  • Search More

    Advanced Search

  • AACR Publications
    • Blood Cancer Discovery
    • Cancer Discovery
    • Cancer Epidemiology, Biomarkers & Prevention
    • Cancer Immunology Research
    • Cancer Prevention Research
    • Cancer Research
    • Clinical Cancer Research
    • Molecular Cancer Research
    • Molecular Cancer Therapeutics

User menu

  • Register
  • Log in
  • My Cart

Search

  • Advanced search
Clinical Cancer Research
Clinical Cancer Research
  • Home
  • About
    • The Journal
    • AACR Journals
    • Subscriptions
    • Permissions and Reprints
    • Reviewing
    • CME
  • Articles
    • OnlineFirst
    • Current Issue
    • Past Issues
    • CCR Focus Archive
    • Meeting Abstracts
    • Collections
      • COVID-19 & Cancer Resource Center
      • Breast Cancer
      • Clinical Trials
      • Immunotherapy: Facts and Hopes
      • Editors' Picks
      • "Best of" Collection
  • For Authors
    • Information for Authors
    • Author Services
    • Best of: Author Profiles
    • Submit
  • Alerts
    • Table of Contents
    • Editors' Picks
    • OnlineFirst
    • Citation
    • Author/Keyword
    • RSS Feeds
    • My Alert Summary & Preferences
  • News
    • Cancer Discovery News
  • COVID-19
  • Webinars
  • Search More

    Advanced Search

CCR Focus

Advances in the Genetics of High-Risk Childhood B-Progenitor Acute Lymphoblastic Leukemia and Juvenile Myelomonocytic Leukemia: Implications for Therapy

Mignon L. Loh and Charles G. Mullighan
Mignon L. Loh
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Charles G. Mullighan
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
DOI: 10.1158/1078-0432.CCR-11-1936 Published May 2012
  • Article
  • Figures & Data
  • Info & Metrics
  • PDF
Loading

Abstract

Hematologic malignancies of childhood comprise the most common childhood cancers. These neoplasms derive from the pathologic clonal expansion of an abnormal cancer-initiating cell and span a diverse spectrum of phenotypes, including acute lymphoblastic leukemia (ALL), acute myeloid leukemia (AML), myeloproliferative neoplasms (MPN), and myelodysplastic syndromes (MDS). Expansion of immature lymphoid or myeloid blasts with suppression of normal hematopoiesis is the hallmark of ALL and AML, whereas MPN is associated with proliferation of 1 or more lineages that retain the ability to differentiate, and MDS is characterized by abnormal hematopoiesis and cytopenias. The outcomes for children with the most common childhood cancer, B-progenitor ALL (B-ALL), in general, is quite favorable, in contrast to children affected by myeloid malignancies. The advent of highly sensitive genomic technologies reveals the remarkable genetic complexity of multiple subsets of high-risk B-progenitor ALL, in contrast to a somewhat simpler model of myeloid neoplasms, although a number of recently discovered alterations displayed by both types of malignancies may lead to common therapeutic approaches. This review outlines recent advances in our understanding of the genetic underpinnings of high-risk B-ALL and juvenile myelomonocytic leukemia, an overlap MPN/MDS found exclusively in children, and we also discuss novel therapeutic approaches that are currently being tested in clinical trials. Recent insights into the clonal heterogeneity of leukemic samples and the implications for diagnostic and therapeutic approaches are also discussed. Clin Cancer Res; 18(10); 2754–67. ©2012 AACR.

Acute Lymphoblastic Leukemia

Acute lymphoblastic leukemia (ALL) is the commonest childhood malignancy and, despite progressive improvements in the outcome of therapy, remains the leading cause of cancer-related death in children and young adults (1). ALL is characterized by accumulation of immature lymphoid progenitors in the bone marrow, peripheral blood, and central nervous system and results in death from the consequences of bone marrow failure: anemia, neutropenia and infection, and thrombocytopenia and bleeding. Although these features are common to all patients, ALL represents a remarkably diverse range of subtypes characterized by distinct constellations of gross and submicroscopic genetic alterations, including chromosomal translocations, structural variants, and DNA sequence mutations. Many of these lesions are central events in leukemogenesis, and several are also important determinants of the risk of treatment failure and relapse. General overviews of the genetic basis of ALL may be found in several recent excellent reviews (2–6). This review is part of a series of articles in this issue reviewing new genetic insights and therapeutic opportunities in childhood malignancies (7–11). Here, we briefly review key chromosomal rearrangements in ALL and then focus on several high-risk subtypes, which have hitherto been poorly understood and for which recent detailed genomic-profiling approaches are providing important insights into the genetic basis of disease, with potentially important implications for therapy.

Approximately three quarters of childhood ALL cases exhibit aneuploidy (hyperdiploidy or, less commonly, low hypodiploidy) or chromosomal rearrangements, such as t(12;21) ETV6-RUNX1, t(1;19) TCF3-PBX1, t(9;22) BCR-ABL1, rearrangement of MLL to a diverse range of fusion partners, and rearrangement of the enhancer elements of the immunoglobulin heavy chain locus (IGH@) to a range of partner genes, including the cytokine receptor gene CRLF2 (Fig. 1; refs. 2, 12). Several of these alterations are associated with favorable (e.g., hyperdiploidy with greater than 50 chromosomes, ETV6-RUNX1) or poor outcome (BCR-ABL1, MLL rearrangement, hypodiploidy), yet many of these alterations are insufficient to induce leukemia in experimental models, suggesting the presence of cooperating genetic lesions. Moreover, a substantial number of childhood ALL samples, including many from those with high-risk features (such as older age, male, and high-presentation blood leukocyte count) and those who experience relapse, lack one of these known chromosomal alterations. Consequently, the past 5 years have witnessed great activity in the use of genomic approaches to identify genetic alterations not evident on cytogenetic analysis (cryptic or submicroscopic lesions; ref. 13). These approaches include microarray-based profiling of structural genetic alterations, such as array-based comparative genomic hybridization and single-nucleotide polymorphism (SNP) microarrays, candidate gene sequencing, and more recently, next-generation sequencing of all coding exons (exome sequencing), the expressed genome (transcriptome sequencing or RNA-seq), and whole-genome sequencing. Importantly, these studies have often incorporated integration with gene expression profiling data derived from microarrays and clinical data.

Figure 1.
  • Download figure
  • Open in new tab
  • Download powerpoint
Figure 1.

Spectrum of recurring chromosomal rearrangements in childhood ALL. Common recurring numerical and structural genetic alterations in childhood B-progenitor and T-lineage ALL. Alterations specific to T-lineage ALL are shown at the bottom of the pie chart in magenta. Reprinted with permission of Elsevier from Mullighan (13), with original data from Pui et al. (3, 107).

ALL genomes typically acquire fewer large structural genetic alterations than many solid tumors, but they harbor more than 50 recurring regions of genetic alteration, most commonly focal deletions that involve few or a limited number of genes (14–16). Many of the genes targeted encode proteins with roles in key cellular pathways, including lymphoid development and differentiation (e.g., PAX5, IKZF1, EBF1, LEF1, and VPREB), cell-cycle regulation and tumor suppression [CDKN2A, CDKN2B (INK4/ARF), TP53, PTEN, RB1], lymphoid signaling (BTLA, CD200, TOX), transcription factors and transcriptional coregulators (ERG, TBL1XR1, CREBBP), regulation of apoptosis (BTG1), and drug receptor genes (NR3C1). Importantly, several of these genes are targeted by multiple mechanisms of mutation, most notably the lymphoid transcription factor PAX5, which is involved by loss-of-function and dominant negative deletions, sequence mutations, and translocations. Furthermore, many lesions are significantly associated with the presence of chromosomal rearrangements and risk of treatment failure in ALL (Table 1), including several subtypes of ALL. These studies have also identified genetic lesions that define novel subtypes of ALL with distinct gene expression profiles, such as subtypes characterized by rearrangement of CRLF2 (17, 18), intrachromosomal amplification of chromosome 21 (19), and focal deletions of ERG (20).

View this table:
  • View inline
  • View popup
Table 1.

Recently identified genetic alterations in B-progenitor acute lymphoblastic leukemia

BCR-ABL1 leukemia

Expression of the constitutively active tyrosine kinase BCR-ABL1 is a hallmark of chronic myeloid leukemia (CML), an expansion of relatively mature granulocytes, and a subset of ALL, which is typically an aggressive form of leukemia that responds poorly to therapy. Genomic profiling has shown that BCR-ABL1–positive lymphoid leukemia, either de novo ALL or CML at the progression to lymphoid blast crisis, is commonly accompanied by deletion or, less frequently, sequence mutation of the IKZF1 gene (21, 22), which encodes a lymphoid transcription factor required for normal lymphoid development. These alterations are loss of function or dominant negative, usually mono-allelic, and accelerate the onset of leukemia in murine models (23). In contrast, IKZF1 alterations are rare in chronic phase CML and at myeloid blast crisis, indicating that these alterations are important determinants of lymphoid disease lineage and progression. IKZF1 alterations are also associated with poor outcome in BCR-ABL1–negative lymphoid leukemia (24, 25), suggesting an important role in treatment failure, and provide a potential explanation for the inferior outcome of BCR-ABL1 ALL compared with CML. The mechanistic basis for IKZF1 mutation–induced treatment failure is poorly understood, but it may relate to loss of IKZF1 activity, resulting in acquisition of a primitive, stem cell–like phenotype that is more resistant to therapy (24). These findings suggest that modulation of IKZF1 activity, or the downstream pathways regulated by this gene, might further improve the outcome of this subtype of ALL.

“BCR-ABL1–like” acute lymphoblastic leukemia

Two years ago, 2 groups, the Children's Oncology Group (COG)/National Cancer Institute Therapeutically Applicable Research to Generate Effective Treatments (TARGET) consortium (24) and a Dutch group led by Monique Den Boer and Rob Pieters (26), identified a subgroup of BCR-ABL1–negative childhood ALL cases, characterized by a gene expression profile similar to that of BCR-ABL1 ALL, which has poor outcome. These cases usually have alterations of IKZF1, but until recently, the genetic alterations substituting for BCR-ABL1 were unknown. These “BCR-ABL–like” or “Ph-like” ALL cases have at least 2 broad groups of genetic changes, resulting in constitutive activation of cytokine receptor and/or tyrosine kinase signaling.

Up to 50% of Ph-like cases harbor rearrangements of CRLF2 [encoding cytokine receptor-like factor 2 or the thymic stromal lymphopoietin receptor (TSLPR); Fig. 2; ref. 27]. CRLF2 forms a heterodimer with interleukin receptor 7α (IL-7R) for the cytokine TSLP. CRLF2 is located at the pseudoautosomal region of Xp/Yp, and the rearrangements are either translocation to IGH@ at 14q33 (18) or a focal deletion upstream of CRLF2 that results in a novel fusion, P2RY8-CRLF2, in which the first noncoding exon of P2RY8 is fused to the entire open reading frame of CRLF2 (17). Both result in aberrant expression of CRLF2 by leukemic cells that may be detected by routine flow cytometric immunophenotyping at the time of diagnosis (17, 18). Less common is a point mutation, CRLF2 p.Phe232Cys, which also results in receptor overexpression (28, 29). Half of CRLF2-rearranged cases have activating mutations in Janus kinases (17, 18, 29), most commonly at or near R683 in the pseudokinase domain of JAK2, but also in the kinase domain of JAK2, and in both the kinase and pseudokinase domains of JAK1 (30). Coexpression of CRLF2 and JAK mutant alleles transforms model cell lines (17, 29), and human CRLF2-rearranged leukemic cells exhibit JAK-STAT activation (18). This JAK-STAT activation, in either experimental models or primary human leukemic cells, is attenuated by the use of commercially available selective JAK inhibitors (17). Several studies of high-risk ALL cohorts have shown that CRLF2/JAK alterations are associated with IKZF1 deletion and/or mutation and that these cases have poor outcome (27, 31). Thus, the addition of JAK inhibitor therapy to conventional multiagent chemotherapy is attractive and is being explored in a dose-finding phase I trial of the JAK inhibitor ruxolitinib in relapsed and refractory childhood tumors (the COG ADVL1011 trial). These important findings represent one of the first novel targeted therapies to be implemented in ALL since the dramatically effective incorporation of imatinib to the treatment of BCR-ABL1–positive ALL (32).

Figure 2.
  • Download figure
  • Open in new tab
  • Download powerpoint
Figure 2.

CRLF2 alterations in B-ALL. A, deletions in the pseudoautosomal 1 region of Xp/Yp. SNP microarray copy number data for 6 ALL cases with matched tumor (T) and normal (N) data are shown. Dark blue shows a focal PAR1 deletion is present, the extent of which is identical in all cases. B, mapping of the PAR1 deletion (probe-level data are shown as vertical red lines) to between intron 1 of P2RY8 and upstream of CRLF2. C, the PAR1 deletion results in a P2RY8-CRLF2 fusion containing the entire CRLF2 open reading frame. D, the P2RY8-CRLF2 fusion results in CRLF2 overexpression that may be detected by immunophenotyping of leukemic cells. Data from Mullighan et al. (17).

Several questions about the role of CRLF2 in leukemogenesis remain unanswered. Although CRLF2 rearrangement is present in only 5% to 7% of general ALL cohorts, it is exceptionally common in B-ALL arising in children with Down syndrome (DS-ALL; refs. 17, 18). The basis for this enrichment of CRLF2 alterations in this unique setting is unknown. Moreover, although data are limited, CRLF2 alterations are not significantly associated with poor outcome in DS-ALL (17), despite a similar frequency of associated JAK2 mutations to that observed in non-DS-ALL (17, 18), perhaps, in part, because of the lack of IKZF1 alterations in this subgroup (17). Moreover, in both DS- and non-DS ALL, half of CRLF2-rearranged cases lack activating JAK mutations and, indeed, seem not to have any additional sequence variants or structural DNA variations activating tyrosine kinases (M.L. Loh and C.G. Mullighan; unpublished observations), despite exhibiting similar activities of downstream signaling pathways to CRLF2-rearranged and/or JAK-mutated cases. The alterations substituting for JAK2 mutations or the reasons why JAK mutations are dispensable for transformation in these cases are unknown, but they should be answered by ongoing genome sequencing efforts and the development of experimental models that carefully model the effect and interaction of each class of lesion (CRLF2, JAK, and IKZF1) in transformation.

A second subgroup of Ph-like cases harbors a remarkably diverse range of cytokine receptor and kinase signaling alterations. RNA-seq and whole-genome sequencing of 12 Ph-like ALL cases, done by the COG/TARGET initiative and St. Jude Children's Research Hospital (Memphis, TN), has shown that cryptic rearrangements, sequence variations, and submicroscopic structural variants are a hallmark of CRLF2 wild-type Ph-like ALL (33, 34). These include chimeric fusions encoding constitutively active tyrosine kinases (EBF1-PDGFRB, NUP214-ABL1, STRN3-JAK2, and BCR-JAK2); dysregulating cytokine receptors (IGH@-EPOR); activating complex sequence mutations in the transmembrane domain of IL7R, encoding the IL-7R chain (also recently reported by other groups in B- and T-lineage ALL; refs. 34–36); and deletions of SH2B3 (LNK), which encodes a negative regulator of JAK2 (37, 38). Recurrence screening of large cohorts of high-risk B-ALL has shown that several of these alterations (NUP214-ABL1, EBF1-PDGFRB, IL7R mutations, and SH2B3 alterations) are recurrent in Ph-like ALL (34). Expression of these alterations in model cell lines and primary hematopoietic progenitors has shown that EBF1-PDGFRB and IL-7R mutant alleles are transforming and result in Jak-Stat activation that is attenuated by Abl1 or Jak inhibitors. Similarly, primary leukemic cells harboring these alterations exhibit activation of downstream signaling pathways that are sensitive to tyrosine kinase inhibitors (34). Thus, although these studies have identified a diverse range of genetic alterations, these converge on common signaling pathways that are potentially amenable to kinase inhibitor therapy.

These findings are important, as Ph-like ALL is common (15%, or at least 3 times as common as BCR-ABL1–positive ALL) and is also present at a similar or higher frequency in adolescents and young adults (M.L. Loh and C.G. Mullighan; unpublished observations), in whom the outcome of ALL therapy is inferior to that of younger children. Moreover, although the genetic landscape of Ph-like is complex, the Ph-like phenotype is readily detectable by expression profiling of a limited set of genes or by flow cytometric analysis of pathway activation (e.g., CRKL phosphorylation for ABL1-rearranged cases and total tyrosine phosphorylation for JAK2-rearranged cases). Implementation of these approaches and the development of trials to treat Ph-like ALL with tyrosine kinase inhibitor therapy are being vigorously pursued.

Hypodiploid acute lymphoblastic leukemia

Hypodiploid ALL is an uncommon subtype of ALL that is associated with a very high risk of treatment failure (39–43). These cases have multiple whole chromosomal losses and have been subclassified according to the degree of aneuploidy: near-haploid cases with 24 to 31 chromosomes and low-hypodiploid cases with 32 to 44 chromosomes. Apart from the documentation of aneuploidy, the genetic basis of this subtype of ALL has been very poorly understood. To investigate this, in conjunction with the COG, we performed detailed genomic profiling of a large cohort of hypodiploid ALL cases, incorporating microarray analysis of structural genetic variation and gene expression, extensive candidate gene sequencing, and next-generation sequencing (44).

These studies have shown that near-haploid ALL cases have a very high frequency of deletions and sequence variations, known or predicted to activate Ras signaling, and that near-haploid and low-hypodiploid cases have distinct and mutually exclusive lesions inactivating lymphoid transcription factors. Several of the lesions, such as alterations of the IKAROS gene family, have only been rarely observed in other leukemias. Conversely, several of the genes involved in regulation of Ras signaling, including NF1, KRAS, NRAS, and PTPN11, are mutated in other disorders, such as juvenile myelomonocytic leukemia (JMML). Moreover, careful analysis of the frequency and nature of these lesions and analysis of leukemic transcriptome have shown that near haploid and low hypodiploid are distinct diseases at the genetic level. As hypodiploidy confers a high risk of treatment failure and satisfactory therapies are lacking, the high frequency of Ras pathway activation suggests that targeted therapies designed to disrupt downstream Ras signaling should be pursued in this disease.

Clonal Heterogeneity in Leukemia

Most genomic-profiling studies in leukemia have examined unfractionated leukemic cell samples, most commonly obtained at the time of diagnosis. These studies have clearly identified multiple novel lesions that are present in the majority of leukemic cells that are important determinants of leukemogenesis and the outcome of treatment. However, as these profiling techniques become increasingly sophisticated, it has become evident that many patients harbor multiple leukemic clones that have distinct constellations of genetic alterations that influence responsiveness to therapy. Many of these insights have been obtained from sequential genomic profiling of leukemic cells obtained at diagnosis and subsequent disease progression.

Cytogenetic analyses performed many years ago showed that many leukemic samples acquire additional chromosomal alterations at the time of relapse, suggesting that the leukemic clone is not static but evolves over time, possibly in response to the selective pressure of treatment (45). Support for this notion has been obtained from multiple studies that did SNP array analysis of DNA copy number alterations of matched samples obtained at diagnosis and relapse (25, 46–49). Genetically identical diagnosis and relapse clones are uncommon (Fig. 3). Similarly, the emergence of a completely distinct leukemia at relapse is infrequent, although it may be more common in very late relapse and in T-lineage ALL (50). In contrast, most samples retain at least some of the genetic lesions from diagnosis to relapse but also exhibit differences, either as the acquisition of new alterations, suggestive of simple clonal evolution, or more commonly, a more complex picture in which some lesions present at diagnosis are lost, and others are gained. This latter scenario suggests that a “prediagnosis” or “ancestral” clone that acquires some, but not all, lesions required for leukemogenesis undergoes divergent evolution, in which one clone acquires additional lesions and expands to form the predominant clone present at diagnosis (Fig. 3). Other clone(s) acquire different lesions, comprise a minority of cells at diagnosis, but may expand during therapy and form the predominant clone at relapse. The genes targeted by novel lesions at relapse include genes previously implicated in poor outcome and leukemogenesis, including CDKN2A/B, IKZF1, and ETV6. Notably, some copy number alterations, such as those involving PAX5 and EBF1, are lost from diagnosis to relapse, whereas others, such as IKZF1, are consistently retained, suggesting that they are acquired early during leukemogenesis and that they are important determinants of resistance to therapy.

Figure 3.
  • Download figure
  • Open in new tab
  • Download powerpoint
Figure 3.

Clonal relationship of diagnosis and relapse samples in ALL. Most relapse samples have a clear relationship to the presenting diagnosis leukemic clone, either arising through the acquisition of additional genetic lesions or, more commonly, arising from an ancestral (prediagnosis) clone. In the latter scenario, the relapse clone retains some, but not all, of the lesions found in the diagnostic sample, while acquiring new lesions. Lesion-specific backtracking studies have shown that, in most cases, the relapse clone exists as a minor subclone within the diagnostic sample prior to the initiation of therapy. In only a minority of ALL cases does the relapse clone represent the emergence of a genetically distinct and thus unrelated second leukemia. Figure originally published in Mullighan et al. (46).

These genetic data are supported by recent studies that have modeled clonal heterogeneity by detailed FISH of leukemic cells, which have shown that individual leukemic cells have distinct genetic alterations and that the patterns of lesions between cells can be used to construct a branching pattern of clonal evolution in individual ALL cases (51). Transplantation of diagnostic BCR-ABL1 ALL samples into immunocompromised mice, followed by comparative genetic profiling of the engrafted tumors with the primary diagnostic samples, has also shown that distinct subclones engraft different mice, and as for the FISH analyses, the patterns of genetic lesions between individual engrafted samples can be used to infer patterns of clonal evolution (52).

These findings have important implications for diagnostic approaches in ALL, as they suggest that a subset of patients have genetic alterations present in minor subclones at diagnosis that will not be detected by profiling of the unfractionated tumor sample by conventional approaches. Strategies to detect lesions present at very low levels will be required, such as quantitative PCR approaches for individual lesions or next-generation sequencing approaches capable of detecting lesions at very low levels.

Sequence analysis of relapsed acute lymphoblastic leukemia

Genome-wide studies of sequence variation have lagged behind those of DNA copy number variation, but they are now feasible with advances in the yield and cost-effectiveness of next-generation sequencing. Although whole-genome sequencing studies of ALL have not yet been reported, candidate gene sequencing efforts using Sanger sequencing have identified a number of novel targets of sequence mutation in high-risk B-ALL (Fig. 4; ref. 53). To extend our understanding of the genetic basis of treatment failure in ALL, we sequenced 300 genes in 23 matched diagnosis-relapse ALL samples. A key finding of this study was the presence of deleterious mutations in CREBBP, encoding CREB-binding protein (CREBBP), in almost 20% of relapsed ALL cases. CREBBP is a multifunctional protein with roles in histone and nonhistone acetylation and transcriptional regulation, by acting as a scaffold of a diverse range of transcription factors. Many of the mutations were located in the histone acetyltransferase domain and impaired the normal acetylation activity of CREBBP. We also showed that the CREBBP mutations impaired the normal transcriptional response to glucocorticoids, which are widely used in ALL therapy, suggesting that CREBBP mutations may have an important role in determining resistance to these agents. Moreover, like IKZF1 alterations, CREBBP mutations were preserved from diagnosis to relapse, or they were present in subclones at diagnosis and emerged in the predominant clone at relapse, supporting the notion that these mutations provide a selective advantage during treatment. Histone deacetylase inhibitors are being explored as potential therapeutic agents in a range of tumors, and preliminary studies using leukemic cell lines showed that many, including those with CREBBP mutations, are exquisitely sensitive to these agents. These findings suggest that histone deacetylase inhibitors may represent a novel therapeutic approach in high-risk leukemia.

Figure 4.
  • Download figure
  • Open in new tab
  • Download powerpoint
Figure 4.

Recurring sequence mutations in high-risk B-ALL. Sequencing of 125 genes in 187 patients with high-risk B-ALL identified multiple novel targets of mutation. This research was originally published by Zhang et al. (53).

Juvenile Myelomonocytic Leukemia

JMML is a rare, but highly aggressive, myeloid malignancy of childhood (54–56). JMML can be a difficult diagnosis to make, as many patients present with nonspecific clinical features that mimic infections or other conditions. However, it is now known that up to 85% of patients harbor a molecular alteration in 1 of 5 genes, making it easier for clinicians to accurately identify patients with JMML. The World Health Organization (WHO) diagnostic criteria still comprise a range of clinical and laboratory findings, but the addition of genetic mutations creates an easier diagnostic algorithm for clinicians to follow (Table 2). These 5 genes (NF1, NRAS, KRAS, PTPN11, and CBL) encode proteins that when mutated are predicted to activate the Ras/mitogen—activated protein kinase (MAPK) pathway (Fig. 5) and, thus, provide additional avenues for therapeutic approaches, although to date, only hematopoietic stem cell transplant (HSCT) is known to be curative.

Figure 5.
  • Download figure
  • Open in new tab
  • Download powerpoint
Figure 5.

Ligand-stimulated Ras activation, the Ras/MAPK pathway, and the gene mutations reported to date, contributing to the neuro-cardio-facio-cutaneous congenital disorders, a newly defined Cbl germline syndrome, and JMML. CFC, cardio-facio-cutaneousa; MGCL, multiple giant cell lesion. Reprinted with permission from Loh (56).

View this table:
  • View inline
  • View popup
Table 2.

Current diagnostic criteria for juvenile myelomonocytic leukemia

One of the hallmark laboratory features of JMML, hypersensitivity of myeloid progenitor cells to granulocyte macrophage colony-stimulating factor (GM-CSF) in colony forming assays (57), can also be present in certain viral infections, thus providing a sensitive, but nonspecific and laborious, adjunctive diagnostic tool (58). However, the presence of the colony assay to assess GM-CSF hypersensitivity has proven invaluable as a laboratory tool to validate newly discovered JMML mutations in either mouse or human cells and can also provide a platform upon which to screen new compounds.

Patients with genetic syndromes provided important clues to unraveling the molecular pathogenesis of juvenile myelomonocytic leukemia

Children affected by specific congenital syndromes, including neurofibromatosis type I (NF1), Noonan syndrome, or Cbl syndrome are at a higher risk of developing JMML or a related transient myeloproliferative disorder in infancy, and the identification of genes associated with these syndromes has provided unique insights into the pathogenesis of this myeloid malignancy.

In 1978, patients with NF1 were noted to be particularly prone to developing a CML or acute myelomonocytic leukemia in the first decade of life (59). Subsequently, NF1 was cloned and found to encode neurofibromin, which hydrolyzes the active, GTP-bound conformation of Ras to the inactive, GDP-bound conformation and, thus, functions as a GTPase-activating protein (60). In 1994, Shannon and colleagues (61) showed loss of the wild-type allele (later shown to identify acquired isodisomy of the mutant allele) in the diseased bone marrows of children with JMML affected by NF1, thus establishing NF1 as a tumor suppressor gene (61, 62). In subsequent mouse models engineered to conditionally delete the Nf1 allele, elevated levels of Ras-GTP occur, and a myeloproliferative neoplasm (MPN) develops (63, 64).

Somatic, oncogenic mutations in NRAS and KRAS2 (RAS) were subsequently identified that year in independent cohorts of patients with JMML, thus providing additional genetic evidence to support that hyperactivation of the Ras/MAPK pathway was essential for the pathogenesis of this disease (Fig. 5; refs. 65, 66). Additional animal models have subsequently revealed that Kras mutant mice also develop fatal myeloid disorders that resemble JMML in vivo and in vitro (67, 68).

Noonan syndrome is another congenital disorder that is inherited in an autosomal dominant fashion 50% of the time (reviewed in ref. 69). Interestingly, some infants with Noonan syndrome also display a hematologic phenotype, including a self-resolving myeloproliferative disorder that resembles JMML (70, 71). In 2001, Tartaglia and colleagues identified germline mutations in PTPN11, a gene encoding the nonreceptor tyrosine phosphatase protein SHP-2, as causative in approximately 50% of children with Noonan syndrome (72). The observations of a JMML-like MPN in patients with Noonan syndrome (primarily those arising as spontaneous mono-allelic germline events) led to the identification of mono-allelic, activating somatic mutations in PTPN11, occurring in as many as 35% of nonsyndromic JMML (73, 74). A comparison between the PTPN11 mutations in de novo and syndromic JMML (Noonan syndrome) reveals that many of the same amino acids in exons 3, 4, and 13 are affected (75), but with different codon substitutions, leading investigators to hypothesize that the transforming ability of mutations in Noonan syndrome are “weaker” and, thus, may be tolerated as germline events. For example, the most common PTPN11 mutation in de novo JMML is the c. 226G > A, resulting in p. Glu76Lys, an alteration that has never been documented as a germline lesion in Noonan syndrome.

To identify additional novel mutations in patients with JMML, we did Affymetrix SNP 6.0 array analysis on samples from JMML patients with and without known Ras pathway abnormalities (76). We identified a region of 11q that acquired uniparental isodisomy in 5 of 27 cases. Importantly, recent reports had uncovered similar lesions in adults with MPN and myelodysplastic syndrome, and subsequent molecular analysis identified homozygous c-CBL (CBL) mutations (77–79). All 5 of our patients also displayed homozygous CBL lesions. The most common CBL mutation in JMML is the c.1111T > C transition, resulting in the substitution of a histidine for a tyrosine residue at codon 371 in the alpha linker region of the protein (76, 80, 81).

A number of children with JMML harboring these homozygous CBL mutations shared phenotypic features, suggesting that these lesions might first occur as germline events (82–84). Additional studies indicate that these mutations are autosomally inherited in a dominant fashion approximately 50% of the time (82). Importantly, genotyping of myeloid colonies grown from 1 patient's cord blood revealed the presence of rare homozygous clones at birth, supporting our previous finding that oncogenic JMML mutations can arise in utero (75, 76).

One striking phenomenon about patients with JMML and homozygous CBL mutations is their high rate of spontaneous resolution (82–84). However, of 5 patients reported with spontaneously resolving JMML and the germline Cbl syndrome, 4 developed signs consistent with serious vasculopathies by the end of their second decade of life, as evidenced by optic atrophy, hypertension, cardiomyopathy, or arteritis (82).

The relationship of Cbl to Ras/MAPK signaling is poorly understood, and furthermore, the function of mutant Cbl proteins in malignancy is likely complex. There are 3 members of the Cbl family of E3 ubiquitin ligases (CBL, CBLB, and CBLC), all of which are known to mark activated protein tyrosine kinases and other associated proteins for degradation by ubiquitination, but Cbl proteins also retain many important adaptor functions (reviewed in ref. 85). Most of the mutant Cbl proteins reported to occur thus far in myeloid malignancies result from missense mutations or splice site variants in CBL, in which the majority have been shown to result in an encoded protein that lacks E3 ligase activity only in the absence of the wild-type protein (79, 82). Two recently generated mouse models indicate possible mechanisms by which mutant Cbl proteins exert their phenotypic effects (86, 87). The first is a Cbl, Cblb double-deficient mouse that develops a rapidly fatal MPN in comparison with a Cbl null mouse, which does not develop disease, suggesting that mutant Cbl proteins may confer a dominant negative effect on Cbl b (86). The second is a heterozygous Cbl RING finger mutant knock-in mouse that develops a more latent, but equally fatal, MPN only on a Cbl-null background (87). This suggests that mutant Cbl proteins may act as gain-of-function oncogenes that continue association with certain proteins but that cannot function to degrade protein tyrosine kinases because of the loss of the negative regulatory E3 ligase function, thus constitutively activating growth pathways. Further work to fully interrogate the consequences of mutant Cbl proteins in myeloid malignancies is under way.

Other mutations have been recently reported to occur rarely in JMML, such as ASXL1 or FLT3 (88, 89). In contrast to other MPNs of adulthood, neither JAK2 nor TET2 mutations occur in JMML (M.L. Loh; unpublished data; ref. 81). Taken together, these data would suggest that the current spectrum of commonly occurring mutations in NF1, RAS, PTPN11, and CBL converge on the Ras/MAPK pathway in a highly specific way in JMML, which remains to be fully elucidated.

Approaches to therapy

Although the current standard of care for JMML is allogeneic HSCT, approximately 50% of children will still suffer from relapse. No standard chemotherapy regimen used prior to HSCT has been shown to affect the incidence of relapse after HSCT (90). Based on the spectrum of mutations that were first characterized by loss of NF1 and oncogenic lesions in RAS, it would be reasonable to hypothesize that suppression of Ras/MAPK signaling might have therapeutic efficacy in JMML. However, 3 issues should be mentioned. First, direct inhibition of hyperactive Ras has been problematic for several reasons. For example, Ras is activated when localized to the cellular membrane, which requires covalent binding of a farnesyl group to a conserved CAAX motif by farnesyl transferases (FTase). As prenylation of Ras is catalyzed by FTase, FTase inhibitors (FTI) were developed to prevent Ras localization to the plasma membrane and inhibit Ras signaling (91). However, alternative geranylation can substitute for prenylation, and thus the clinical promise of FTI therapy as Ras inhibitors has been disappointing. The COG sponsored a phase II trial, AAML0122, incorporating a window phase of an FTI followed by chemotherapy only and HSCT (92). The event-free survival on AAML0122 was 40%, with an overall survival of 55%, with a median follow-up of 4 years (T. Cooper; personal communication). Second, the relationships of SHP2 and CBL to Ras signaling are less well understood, and it may be that the biochemical consequences of these lesions will render them more sensitive to alternative therapeutic approaches. Finally, the importance of Ras signaling in diverse physiologic processes in normal cells raises the concern of nonspecific systemic toxicities with any therapies in which normal Ras proteins or their effectors are inhibited. Ideally, one hopes to capitalize on the existence of “oncogene addiction” and identify inhibitors that preferentially treat mutant rather than normal cells. However, amplification of signals upstream or parallel to Ras may occur, raising the possibility that combination therapy to overcome resistance will be necessary to achieve therapeutic effects (93).

One example of a Ras effector pathway involved in JMML is the Raf/MAP–extracellular signal—regulated kinase (ERK) kinase (MEK)/ERK pathway. Ras-GTP induces a cascade of activation from the Raf kinases, to MAPK kinase (MEK1 and MEK2), to ERK, which stimulates diverse nuclear and cytosolic effects. Raf mutations are found in melanoma and renal cell carcinoma, and it is reasonable to hypothesize that inhibiting the Raf/MEK/ERK cascade has therapeutic potential in JMML. To date, the largest clinical experience with Raf inhibitors is with sorafenib (Bay 43-9006), which is U.S. Food and Drug Administration (FDA) approved for treatment of advanced renal cell carcinoma and hepatocellular carcinoma (94, 95). Sorafenib is a novel bi-aryl, multikinase inhibitor with activity against VEGF receptor-2 and receptor-3, platelet—derived growth factor receptor b, fibroblast growth factor receptor 1, FLT-3, and c-KIT (96), as well as Raf-1 and B-raf (97), and it is currently being used in conjunction with chemotherapy for acute myeloid leukemia (AML) in the COG. However, the recent data to support the emergence of skin tumors in patients treated with Raf inhibitors highlight the complexity of signaling networks in cancer cells (93).

MEK is also an attractive target, based on the frequency of Ras and Raf mutations in human cancer. Selective inhibition of MEK, resulting in inhibition of its only known substrate, ERK1/2, can be achieved with non-ATP competitive inhibitors. In recent reports, the use of a MEK inhibitor, PD0325901 (PD901; Pfizer), in a model of JMML and chronic myelomonocytic leukemia, characterized by the conditional expression of the KrasG12D mutant allele, had encouraging results. PD901 treatment resulted in a sustained reduction of peripheral white blood count, improved anemia, and enhanced survival in the treated mutants compared with the wild-type controls (98). Importantly, PD901 abrogated the classic GM-CSF hypersensitivity exhibited by myeloid progenitors in in vitro assays. MEK inhibition may also have a role for patients with refractory JMML who transform to AML. This is suggested by the inhibition of Nf1 mutant acute myeloid leukemic colony growth at low doses of CI-1040 and prolonged survival of mice transplanted with Nf1 mutant AML when treated with CI-1040 (99). Given the recent description of Ras pathway mutations in hypodiploid ALL, it would be reasonable to assess the clinical utility of MEK inhibition for both of these diseases.

Finally, we and others described the phosphorylation of STAT5 to low-dose GM-CSF in subsets of JMML cells (100, 101). The discovery of activating JAK2 mutations in other MPNs, including the majority of adults with polycythemia vera, essential thrombocythemia, and primary myelofibrosis (102–104), led to the development of a number of specific and nonspecific agents that inhibit wild-type and mutant JAK2. Ruxolitinib (formally INCB018424; Incyte/Novartis) has recently been FDA approved for primary myelofibrosis patients, showing excellent tolerability and clinical responses (105, 106). In collaboration with Incyte/Novartis, the COG is determining the maximum tolerated dose of ruxolitinib in children (ADVL1011) on the basis of recent data showing the presence of JAK lesions in approximately 10% of children with high-risk ALL (30), the known JAK lesions in polycythemia vera and essential thrombocythemia, and the documented signaling abnormalities in JMML (100). Correlating the effect on colony formation, cell growth, and signaling networks in the presence of the inhibitor will help evaluate JAK-STAT signaling in disease pathogenesis and may reveal previously unappreciated relationships between the JAK/STAT5 and Ras/MAPK pathways.

Summary and directions for future work

Advances in genomic technologies have identified a number of lesions that have transformed our understanding of the molecular pathogenesis of leukemia, revealing a number of targetable lesions in both lymphoid and myeloid diseases, particularly in subsets of patients for whom cure is currently elusive or for whom cure is associated with substantial morbidity.

Results from next-generation sequencing efforts have indicated that identification of all genetic alterations contributing to leukemogenesis and treatment failure will require comprehensive analysis of sequence and structural variations with these technologies and deep sequencing of samples at multiple disease time points to identify critical lesions in tumor subclones that confer resistance to therapy.

Ongoing investigations of targeted therapies will require continued study of the biochemical and functional consequences of the genetic lesions occurring in these high-risk diseases and will also require rigorous clinical trials to assess both target inhibition and outcome. However, recent efforts will undoubtedly continue to provide fruitful and productive collaborations among basic, translational, and clinical scientists.

Disclosure of Potential Conflicts of Interest

M.L. Loh is a consultant for Bristol-Myers Squibb. No potential conflicts of interest were disclosed by the other author.

Authors' Contributions

Conception and design: C.G. Mullighan, M.L. Loh

Development of methodology: C.G. Mullighan, M.L. Loh

Acquisition of data (provided animals, acquired and managed patients, provided facilities, etc.): C.G. Mullighan, M.L. Loh

Analysis and interpretation of data (e.g., statistical analysis, biostatistics, computational analysis): C.G. Mullighan, M.L. Loh

Writing, review, and/or revision of the manuscript: C.G. Mullighan, M.L. Loh

Administrative, technical, or material support (i.e., reporting or organizing data, constructing databases): M.L. Loh

Study supervision: M.L. Loh

Grant Support

C.G. Mullighan is supported by the American Lebanese Syrian Associated Charities of St. Jude Children's Research Hospital, is a Pew Scholar in the Biomedical Sciences, and is a St. Baldrick's Scholar. M.L. Loh is a Clinical Scholar of the Leukemia Lymphoma Society.

Acknowledgments

We thank our colleagues at St. Jude Children's Research Hospital, the Children's Oncology Group, the National Cancer Institute, the University of New Mexico, and the British Columbia Genome Sciences Center who have contributed to the studies described in this review.

  • Received December 13, 2011.
  • Revision received February 26, 2102.
  • Accepted March 5, 2102.
  • ©2012 American Association for Cancer Research.

References

  1. 1.↵
    1. Pui CH,
    2. Robison LL,
    3. Look AT
    . Acute lymphoblastic leukaemia. Lancet 2008;371:1030–43.
    OpenUrlCrossRefPubMed
  2. 2.↵
    1. Harrison CJ
    . Cytogenetics of paediatric and adolescent acute lymphoblastic leukaemia. Br J Haematol 2009;144:147–56.
    OpenUrlCrossRefPubMed
  3. 3.↵
    1. Pui CH,
    2. Carroll WL,
    3. Meshinchi S,
    4. Arceci RJ
    . Biology, risk stratification, and therapy of pediatric acute leukemias: an update. J Clin Oncol 2011;29:551–65.
    OpenUrlAbstract/FREE Full Text
  4. 4.↵
    1. Mullighan CG,
    2. Downing JR
    . Genome-wide profiling of genetic alterations in acute lymphoblastic leukemia: recent insights and future directions. Leukemia 2009;23:1209–18.
    OpenUrlCrossRefPubMed
  5. 5.↵
    1. Collins-Underwood JR,
    2. Mullighan CG
    . Genomic profiling of high-risk acute lymphoblastic leukemia. Leukemia 2010;24:1676–85. doi:10.1038/leu.2010.177.
    OpenUrlCrossRefPubMed
  6. 6.↵
    1. Aifantis I,
    2. Raetz E,
    3. Buonamici S
    . Molecular pathogenesis of T-cell leukaemia and lymphoma. Nat Rev Immunol 2008;8:380–90.
    OpenUrlCrossRefPubMed
  7. 7.↵
    1. Lawlor ER,
    2. Thiele CJ
    . Epigenetic changes in pediatric solid tumors: promising new targets. Clin Cancer Res 2012;18:2768–79.
    OpenUrlAbstract/FREE Full Text
  8. 8.↵
    1. Lee DW,
    2. Barrett DM,
    3. Mackall C,
    4. Orentas R,
    5. Grupp SA
    . The future is now: chimeric antigen receptors as new targeted therapies for childhood cancer. Clin Cancer Res 2012;18:2780–90.
    OpenUrlAbstract/FREE Full Text
  9. 9.↵
    1. Matthay KK,
    2. George RE,
    3. Yu AL
    . Promising therapeutic targets in neuroblastoma. Clin Cancer Res 2012;18:2740–53.
    OpenUrlAbstract/FREE Full Text
  10. 10.↵
    1. Pinto N,
    2. Cohn SL,
    3. Dolan ME
    . Using germline genomics to individualize pediatric cancer treatments. Clin Cancer Res 2012;18:2791–800.
    OpenUrlAbstract/FREE Full Text
  11. 11.↵
    1. Thiele CJ,
    2. Cohn SL
    . Genetically InFormed Therapies—a “GIFT” for children with cancer. Clin Cancer Res 2012;18:2735–9.
    OpenUrlAbstract/FREE Full Text
  12. 12.↵
    1. Dyer MJ,
    2. Akasaka T,
    3. Capasso M,
    4. Dusanjh P,
    5. Lee YF,
    6. Karran EL,
    7. et al.
    Immunoglobulin heavy chain (IGH) locus chromosomal translocations in B-cell precursor acute lymphoblastic leukemia (BCP-ALL): rare clinical curios or potent genetic drivers? Blood 2009;115:1490–9.
    OpenUrlPubMed
  13. 13.↵
    1. Mullighan CG
    . Genomic profiling of B-progenitor acute lymphoblastic leukemia. Best Pract Res Clin Haematol 2011;24:489–503.
    OpenUrlCrossRefPubMed
  14. 14.↵
    1. Mullighan CG,
    2. Goorha S,
    3. Radtke I,
    4. Miller CB,
    5. Coustan-Smith E,
    6. Dalton JD,
    7. et al.
    Genome-wide analysis of genetic alterations in acute lymphoblastic leukaemia. Nature 2007;446:758–64.
    OpenUrlCrossRefPubMed
  15. 15.↵
    1. Kuiper RP,
    2. Schoenmakers EF,
    3. van Reijmersdal SV,
    4. Hehir-Kwa JY,
    5. van Kessel AG,
    6. van Leeuwen FN,
    7. et al.
    High-resolution genomic profiling of childhood ALL reveals novel recurrent genetic lesions affecting pathways involved in lymphocyte differentiation and cell cycle progression. Leukemia 2007;21:1258–66.
    OpenUrlCrossRefPubMed
  16. 16.↵
    1. Kawamata N,
    2. Ogawa S,
    3. Zimmermann M,
    4. Kato M,
    5. Sanada M,
    6. Hemminki K,
    7. et al.
    Molecular allelokaryotyping of pediatric acute lymphoblastic leukemias by high-resolution single nucleotide polymorphism oligonucleotide genomic microarray. Blood 2008;111:776–84.
    OpenUrlAbstract/FREE Full Text
  17. 17.↵
    1. Mullighan CG,
    2. Collins-Underwood JR,
    3. Phillips LA,
    4. Loudin MG,
    5. Liu W,
    6. Zhang J,
    7. et al.
    Rearrangement of CRLF2 in B-progenitor- and Down syndrome-associated acute lymphoblastic leukemia. Nat Genet 2009;41:1243–6.
    OpenUrlCrossRefPubMed
  18. 18.↵
    1. Russell LJ,
    2. Capasso M,
    3. Vater I,
    4. Akasaka T,
    5. Bernard OA,
    6. Calasanz MJ,
    7. et al.
    Deregulated expression of cytokine receptor gene, CRLF2, is involved in lymphoid transformation in B-cell precursor acute lymphoblastic leukemia. Blood 2009;114:2688–98.
    OpenUrlAbstract/FREE Full Text
  19. 19.↵
    1. Strefford JC,
    2. van Delft FW,
    3. Robinson HM,
    4. Worley H,
    5. Yiannikouris O,
    6. Selzer R,
    7. et al.
    Complex genomic alterations and gene expression in acute lymphoblastic leukemia with intrachromosomal amplification of chromosome 21. Proc Natl Acad Sci U S A 2006;103:8167–72.
    OpenUrlAbstract/FREE Full Text
  20. 20.↵
    1. Mullighan CG,
    2. Miller CB,
    3. Su X,
    4. Radtke I,
    5. Dalton J,
    6. Song G,
    7. et al.
    ERG deletions define a novel subtype of B-progenitor acute lymphoblastic leukemia. Blood [ASH Annual Meeting Abstracts] 2007;110:691.
    OpenUrl
  21. 21.↵
    1. Mullighan CG,
    2. Miller CB,
    3. Radtke I,
    4. Phillips LA,
    5. Dalton J,
    6. Ma J,
    7. et al.
    BCR-ABL1 lymphoblastic leukaemia is characterized by the deletion of Ikaros. Nature 2008;453:110–4.
    OpenUrlCrossRefPubMed
  22. 22.↵
    1. Iacobucci I,
    2. Storlazzi CT,
    3. Cilloni D,
    4. Lonetti A,
    5. Ottaviani E,
    6. Soverini S,
    7. et al.
    Identification and molecular characterization of recurrent genomic deletions on 7p12 in the IKZF1 gene in a large cohort of BCR-ABL1-positive acute lymphoblastic leukemia patients: on behalf of Gruppo Italiano Malattie Ematologiche dell'Adulto Acute Leukemia Working Party (GIMEMA AL WP). Blood 2009;114:2159–67.
    OpenUrlAbstract/FREE Full Text
  23. 23.↵
    1. Virely C,
    2. Moulin S,
    3. Cobaleda C,
    4. Lasgi C,
    5. Alberdi A,
    6. Soulier J,
    7. et al.
    Haploinsufficiency of the IKZF1 (IKAROS) tumor suppressor gene cooperates with BCR-ABL in a transgenic model of acute lymphoblastic leukemia. Leukemia 2010;24:1200–4.
    OpenUrlCrossRefPubMed
  24. 24.↵
    1. Mullighan CG,
    2. Su X,
    3. Zhang J,
    4. Radtke I,
    5. Phillips LA,
    6. Miller CB,
    7. et al.,
    8. Children's Oncology Group
    . Deletion of IKZF1 and prognosis in acute lymphoblastic leukemia. N Engl J Med 2009;360:470–80.
    OpenUrlCrossRefPubMed
  25. 25.↵
    1. Kuiper RP,
    2. Waanders E,
    3. van der Velden VH,
    4. van Reijmersdal SV,
    5. Venkatachalam R,
    6. Scheijen B,
    7. et al.
    IKZF1 deletions predict relapse in uniformly treated pediatric precursor B-ALL. Leukemia 2010;24:1258–64.
    OpenUrlCrossRefPubMed
  26. 26.↵
    1. Den Boer ML,
    2. van Slegtenhorst M,
    3. De Menezes RX,
    4. Cheok MH,
    5. Buijs-Gladdines JG,
    6. Peters ST,
    7. et al.
    A subtype of childhood acute lymphoblastic leukaemia with poor treatment outcome: a genome-wide classification study. Lancet Oncol 2009;10:125–34.
    OpenUrlCrossRefPubMed
  27. 27.↵
    1. Harvey RC,
    2. Mullighan CG,
    3. Chen IM,
    4. Wharton W,
    5. Mikhail FM,
    6. Carroll AJ,
    7. 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 2010;115:5312–21.
    OpenUrlAbstract/FREE Full Text
  28. 28.↵
    1. Yoda A,
    2. Yoda Y,
    3. Chiaretti S,
    4. Bar-Natan M,
    5. Mani K,
    6. Rodig SJ,
    7. et al.
    Functional screening identifies CRLF2 in precursor B-cell acute lymphoblastic leukemia. Proc Natl Acad Sci U S A 2010;107:252–7.
    OpenUrlAbstract/FREE Full Text
  29. 29.↵
    1. Hertzberg L,
    2. Vendramini E,
    3. Ganmore I,
    4. Cazzaniga G,
    5. Schmitz M,
    6. Chalker J,
    7. et al.
    Down syndrome acute lymphoblastic leukemia, a highly heterogeneous disease in which aberrant expression of CRLF2 is associated with mutated JAK2: a report from the International BFM Study Group. Blood 2010;115:1006–17.
    OpenUrlAbstract/FREE Full Text
  30. 30.↵
    1. Mullighan CG,
    2. Zhang J,
    3. Harvey RC,
    4. Collins-Underwood JR,
    5. Schulman BA,
    6. Phillips LA,
    7. et al.
    JAK mutations in high-risk childhood acute lymphoblastic leukemia. Proc Natl Acad Sci U S A 2009;106:9414–8.
    OpenUrlAbstract/FREE Full Text
  31. 31.↵
    1. Cario G,
    2. Zimmermann M,
    3. Romey R,
    4. Gesk S,
    5. Vater I,
    6. Harbott J,
    7. 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 2010;115:5393–7.
    OpenUrlAbstract/FREE Full Text
  32. 32.↵
    1. Schultz KR,
    2. Bowman WP,
    3. Aledo A,
    4. Slayton WB,
    5. Sather H,
    6. Devidas M,
    7. 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 2009;27:5175–81.
    OpenUrlAbstract/FREE Full Text
  33. 33.↵
    1. Mullighan CG,
    2. Morin RD,
    3. Zhang J,
    4. Hirst M,
    5. Zhao Y,
    6. Yan C,
    7. et al.
    Next generation transcriptomic resequencing identifies novel genetic alterations in high-risk (HR) childhood acute lymphoblastic leukemia (ALL): a report from the Children's Oncology Group (COG) HR ALL TARGET Project. Blood [ASH Annual Meeting Abstracts] 2009;114:704.
    OpenUrl
  34. 34.↵
    1. Roberts KG,
    2. Morin RD,
    3. Zhang J,
    4. Hirst M,
    5. Harvey RC,
    6. Kasap C,
    7. et al.
    Novel chromosomal rearrangements and sequence mutations in high-risk Ph-like acute lymphoblastic leukemia. Blood [ASH Annual Meeting Abstracts] 2011;118:67.
    OpenUrl
  35. 35.↵
    1. Shochat C,
    2. Tal N,
    3. Bandapalli OR,
    4. Palmi C,
    5. Ganmore I,
    6. te Kronnie G,
    7. et al.
    Gain-of-function mutations in interleukin-7 receptor-a (IL7R) in childhood acute lymphoblastic leukemias. J Exp Med 2011;208:901–8.
    OpenUrlAbstract/FREE Full Text
  36. 36.↵
    1. Zenatti PP,
    2. Ribeiro D,
    3. Li W,
    4. Zuurbier L,
    5. Silva MC,
    6. Paganin M,
    7. et al.
    Oncogenic IL7R gain-of-function mutations in childhood T-cell acute lymphoblastic leukemia. Nat Genet 2011;43:932–9.
    OpenUrlCrossRefPubMed
  37. 37.↵
    1. Lasho TL,
    2. Pardanani A,
    3. Tefferi A
    . LNK mutations in JAK2 mutation-negative erythrocytosis. N Engl J Med 2010;363:1189–90.
    OpenUrlCrossRefPubMed
  38. 38.↵
    1. Baran-Marszak F,
    2. Magdoud H,
    3. Desterke C,
    4. Alvarado A,
    5. Roger C,
    6. Harel S,
    7. et al.
    Expression level and differential JAK2-V617F-binding of the adaptor protein Lnk regulates JAK2-mediated signals in myeloproliferative neoplasms. Blood 2010;116:5961–71.
    OpenUrlAbstract/FREE Full Text
  39. 39.↵
    1. Harrison CJ,
    2. Moorman AV,
    3. Broadfield ZJ,
    4. Cheung KL,
    5. Harris RL,
    6. Reza Jalali G,
    7. et al.,
    8. Childhood and Adult Leukaemia Working Parties
    . Three distinct subgroups of hypodiploidy in acute lymphoblastic leukaemia. Br J Haematol 2004;125:552–9.
    OpenUrlCrossRefPubMed
  40. 40.↵
    1. Heerema NA,
    2. Nachman JB,
    3. Sather HN,
    4. Sensel MG,
    5. Lee MK,
    6. Hutchinson R,
    7. et al.
    Hypodiploidy with less than 45 chromosomes confers adverse risk in childhood acute lymphoblastic leukemia: a report from the children's cancer group. Blood 1999;94:4036–45.
    OpenUrlAbstract/FREE Full Text
  41. 41.↵
    1. Nachman JB,
    2. Heerema NA,
    3. Sather H,
    4. Camitta B,
    5. Forestier E,
    6. Harrison CJ,
    7. et al.
    Outcome of treatment in children with hypodiploid acute lymphoblastic leukemia. Blood 2007;110:1112–5.
    OpenUrlAbstract/FREE Full Text
  42. 42.↵
    1. Pui CH,
    2. Williams DL,
    3. Raimondi SC,
    4. Rivera GK,
    5. Look AT,
    6. Dodge RK,
    7. et al.
    Hypodiploidy is associated with a poor prognosis in childhood acute lymphoblastic leukemia. Blood 1987;70:247–53.
    OpenUrlAbstract/FREE Full Text
  43. 43.↵
    1. Raimondi SC,
    2. Zhou Y,
    3. Mathew S,
    4. Shurtleff SA,
    5. Sandlund JT,
    6. Rivera GK,
    7. et al.
    Reassessment of the prognostic significance of hypodiploidy in pediatric patients with acute lymphoblastic leukemia. Cancer 2003;98:2715–22.
    OpenUrlCrossRefPubMed
  44. 44.↵
    1. Holmfeldt L,
    2. Zhang J,
    3. Ma J,
    4. Devidas M,
    5. Carroll AJ,
    6. Heerema NA,
    7. et al.
    Genome-wide analysis of genetic alterations in hypodiploid acute lymphoblastic leukemia identifies a high frequency of mutations targeting the IKAROS gene family and Ras signaling. Blood [ASH Annual Meeting Abstracts] 2010;116:411.
    OpenUrl
  45. 45.↵
    1. Raimondi SC,
    2. Pui CH,
    3. Head DR,
    4. Rivera GK,
    5. Behm FG
    . Cytogenetically different leukemic clones at relapse of childhood acute lymphoblastic leukemia. Blood 1993;82:576–80.
    OpenUrlAbstract/FREE Full Text
  46. 46.↵
    1. Mullighan CG,
    2. Phillips LA,
    3. Su X,
    4. Ma J,
    5. Miller CB,
    6. Shurtleff SA,
    7. et al.
    Genomic analysis of the clonal origins of relapsed acute lymphoblastic leukemia. Science 2008;322:1377–80.
    OpenUrlAbstract/FREE Full Text
  47. 47.↵
    1. Yang JJ,
    2. Bhojwani D,
    3. Yang W,
    4. Cai X,
    5. Stocco G,
    6. Crews K,
    7. et al.
    Genome-wide copy number profiling reveals molecular evolution from diagnosis to relapse in childhood acute lymphoblastic leukemia. Blood 2008;112:4178–83.
    OpenUrlAbstract/FREE Full Text
  48. 48.↵
    1. Kawamata N,
    2. Ogawa S,
    3. Seeger K,
    4. Kirschner-Schwabe R,
    5. Huynh T,
    6. Chen J,
    7. et al.
    Molecular allelokaryotyping of relapsed pediatric acute lymphoblastic leukemia. Int J Oncol 2009;34:1603–12.
    OpenUrlPubMed
  49. 49.↵
    1. van Delft FW,
    2. Horsley S,
    3. Colman S,
    4. Anderson K,
    5. Bateman C,
    6. Kempski H,
    7. et al.
    Clonal origins of relapse in ETV6-RUNX1 acute lymphoblastic leukemia. Blood 2011;117:6247–54.
    OpenUrlAbstract/FREE Full Text
  50. 50.↵
    1. Szczepanski T,
    2. van der Velden VH,
    3. Waanders E,
    4. Kuiper RP,
    5. Van Vlierberghe P,
    6. Gruhn B,
    7. et al.
    Late recurrence of childhood T-cell acute lymphoblastic leukemia frequently represents a second leukemia rather than a relapse: first evidence for genetic predisposition. J Clin Oncol 2011;29:1643–9.
    OpenUrlAbstract/FREE Full Text
  51. 51.↵
    1. Anderson K,
    2. Lutz C,
    3. van Delft FW,
    4. Bateman CM,
    5. Guo Y,
    6. Colman SM,
    7. et al.
    Genetic variegation of clonal architecture and propagating cells in leukaemia. Nature 2011;469:356–61.
    OpenUrlCrossRefPubMed
  52. 52.↵
    1. Notta F,
    2. Mullighan CG,
    3. Wang JC,
    4. Poeppl A,
    5. Doulatov S,
    6. Phillips LA,
    7. et al.
    Evolution of human BCR-ABL1 lymphoblastic leukaemia-initiating cells. Nature 2011;469:362–7.
    OpenUrlCrossRefPubMed
  53. 53.↵
    1. Zhang J,
    2. Mullighan CG,
    3. Harvey RC,
    4. Wu G,
    5. Chen X,
    6. Edmonson M,
    7. et al.
    Key pathways are frequently mutated in high-risk childhood acute lymphoblastic leukemia: a report from the Children's Oncology Group. Blood 2011;118:3080–7.
    OpenUrlAbstract/FREE Full Text
  54. 54.↵
    1. Hasle H,
    2. Wadsworth LD,
    3. Massing BG,
    4. McBride M,
    5. Schultz KR
    . A population-based study of childhood myelodysplastic syndrome in British Columbia, Canada. Br J Haematol 1999;106:1027–32.
    OpenUrlCrossRefPubMed
  55. 55.↵
    1. Passmore SJ,
    2. Chessells JM,
    3. Kempski H,
    4. Hann IM,
    5. Brownbill PA,
    6. Stiller CA
    . Paediatric myelodysplastic syndromes and juvenile myelomonocytic leukaemia in the UK: a population-based study of incidence and survival. Br J Haematol 2003;121:758–67.
    OpenUrlCrossRefPubMed
  56. 56.↵
    1. Loh ML
    . Childhood myelodysplastic syndrome: focus on the approach to diagnosis and treatment of juvenile myelomonocytic leukemia. Hematology Am Soc Hematol Educ Program 2010;2010:357–62.
    OpenUrlAbstract/FREE Full Text
  57. 57.↵
    1. Emanuel PD,
    2. Bates LJ,
    3. Castleberry RP,
    4. Gualtieri RJ,
    5. Zuckerman KS
    . Selective hypersensitivity to granulocyte-macrophage colony-stimulating factor by juvenile chronic myeloid leukemia hematopoietic progenitors. Blood 1991;77:925–9.
    OpenUrlAbstract/FREE Full Text
  58. 58.↵
    1. Pinkel D
    . Differentiating juvenile myelomonocytic leukemia from infectious disease. Blood 1998;91:365–7.
    OpenUrlFREE Full Text
  59. 59.↵
    1. Bader JL,
    2. Miller RW
    . Neurofibromatosis and childhood leukemia. J Pediatr 1978;92:925–9.
    OpenUrlCrossRefPubMed
  60. 60.↵
    1. Xu GF,
    2. Lin B,
    3. Tanaka K,
    4. Dunn D,
    5. Wood D,
    6. Gesteland R,
    7. et al.
    The catalytic domain of the neurofibromatosis type 1 gene product stimulates ras GTPase and complements ira mutants of S. cerevisiae. Cell 1990;63:835–41.
    OpenUrlCrossRefPubMed
  61. 61.↵
    1. Shannon KM,
    2. O'Connell P,
    3. Martin GA,
    4. Paderanga D,
    5. Olson K,
    6. Dinndorf P,
    7. et al.
    Loss of the normal NF1 allele from the bone marrow of children with type 1 neurofibromatosis and malignant myeloid disorders. N Engl J Med 1994;330:597–601.
    OpenUrlCrossRefPubMed
  62. 62.↵
    1. Stephens K,
    2. Weaver M,
    3. Leppig KA,
    4. Maruyama K,
    5. Emanuel PD,
    6. Le Beau MM,
    7. et al.
    Interstitial uniparental isodisomy at clustered breakpoint intervals is a frequent mechanism of NF1 inactivation in myeloid malignancies. Blood 2006;108:1684–9.
    OpenUrlAbstract/FREE Full Text
  63. 63.↵
    1. Largaespada DA,
    2. Brannan CI,
    3. Jenkins NA,
    4. Copeland NG
    . Nf1 deficiency causes Ras-mediated granulocyte/macrophage colony stimulating factor hypersensitivity and chronic myeloid leukaemia. Nat Genet 1996;12:137–43.
    OpenUrlCrossRefPubMed
  64. 64.↵
    1. Bollag G,
    2. Clapp DW,
    3. Shih S,
    4. Adler F,
    5. Zhang YY,
    6. Thompson P,
    7. et al.
    Loss of NF1 results in activation of the Ras signaling pathway and leads to aberrant growth in haematopoietic cells. Nat Genet 1996;12:144–8.
    OpenUrlCrossRefPubMed
  65. 65.↵
    1. Kalra R,
    2. Paderanga DC,
    3. Olson K,
    4. Shannon KM
    . Genetic analysis is consistent with the hypothesis that NF1 limits myeloid cell growth through p21ras . Blood 1994;84:3435–9.
    OpenUrlAbstract/FREE Full Text
  66. 66.↵
    1. Miyauchi J,
    2. Asada M,
    3. Sasaki M,
    4. Tsunematsu Y,
    5. Kojima S,
    6. Mizutani S
    . Mutations of the N-ras gene in juvenile chronic myelogenous leukemia. Blood 1994;83:2248–54.
    OpenUrlAbstract/FREE Full Text
  67. 67.↵
    1. Braun BS,
    2. Tuveson DA,
    3. Kong N,
    4. Le DT,
    5. Kogan SC,
    6. Rozmus J,
    7. et al.
    Somatic activation of oncogenic Kras in hematopoietic cells initiates a rapidly fatal myeloproliferative disorder. Proc Natl Acad Sci U S A 2004;101:597–602.
    OpenUrlAbstract/FREE Full Text
  68. 68.↵
    1. Chan IT,
    2. Kutok JL,
    3. Williams IR,
    4. Cohen S,
    5. Kelly L,
    6. Shigematsu H,
    7. et al.
    Conditional expression of oncogenic K-ras from its endogenous promoter induces a myeloproliferative disease. J Clin Invest 2004;113:528–38.
    OpenUrlCrossRefPubMed
  69. 69.↵
    1. Rauen KA,
    2. Schoyer L,
    3. McCormick F,
    4. Lin AE,
    5. Allanson JE,
    6. Stevenson DA,
    7. et al
    . Proceedings from the 2009 genetic syndromes of the Ras/MAPK pathway: from bedside to bench and back. Am J Med Genet Part A 2010;152A:4–24.
    OpenUrl
  70. 70.↵
    1. Bader-Meunier B,
    2. Tchernia G,
    3. Miélot F,
    4. Fontaine JL,
    5. Thomas C,
    6. Lyonnet S,
    7. et al.
    Occurrence of myeloproliferative disorder in patients with Noonan syndrome. J Pediatr 1997;130:885–9.
    OpenUrlCrossRefPubMed
  71. 71.↵
    1. Side LE,
    2. Shannon KM
    . Myeloid disorders in infants with Noonan syndrome and a resident's “rule” recalled. J Pediatr 1997;130:857–9.
    OpenUrlPubMed
  72. 72.↵
    1. Tartaglia M,
    2. Mehler EL,
    3. Goldberg R,
    4. Zampino G,
    5. Brunner HG,
    6. Kremer H,
    7. et al.
    Mutations in PTPN11, encoding the protein tyrosine phosphatase SHP-2, cause Noonan syndrome. Nat Genet 2001;29:465–8.
    OpenUrlCrossRefPubMed
  73. 73.↵
    1. Loh ML,
    2. Vattikuti S,
    3. Schubbert S,
    4. Reynolds MG,
    5. Carlson E,
    6. Lieuw KH,
    7. et al.
    Somatic mutations in PTPN11 implicate the protein tyrosine phosphatase SHP-2 in leukemogenesis. Blood 2004;103:2325–31.
    OpenUrlAbstract/FREE Full Text
  74. 74.↵
    1. Tartaglia M,
    2. Niemeyer CM,
    3. Fragale A,
    4. Song X,
    5. Buechner J,
    6. Jung A,
    7. et al.
    Somatic mutations in PTPN11 in juvenile myelomonocytic leukemia, myelodysplastic syndromes and acute myeloid leukemia. Nat Genet 2003;34:148–50.
    OpenUrlCrossRefPubMed
  75. 75.↵
    1. Kratz CP,
    2. Niemeyer CM,
    3. Castleberry RP,
    4. Cetin M,
    5. Bergsträsser E,
    6. Emanuel PD,
    7. et al.
    The mutational spectrum of PTPN11 in juvenile myelomonocytic leukemia and Noonan syndrome/myeloproliferative disease. Blood 2005;106:2183–5.
    OpenUrlAbstract/FREE Full Text
  76. 76.↵
    1. Loh ML,
    2. Sakai DS,
    3. Flotho C,
    4. Kang M,
    5. Fliegauf M,
    6. Archambeault S,
    7. et al.
    Mutations in CBL occur frequently in juvenile myelomonocytic leukemia. Blood 2009;114:1859–63.
    OpenUrlAbstract/FREE Full Text
  77. 77.↵
    1. Dunbar AJ,
    2. Gondek LP,
    3. O'Keefe CL,
    4. Makishima H,
    5. Rataul MS,
    6. Szpurka H,
    7. et al.
    250K single nucleotide polymorphism array karyotyping identifies acquired uniparental disomy and homozygous mutations, including novel missense substitutions of c-Cbl, in myeloid malignancies. Cancer Res 2008;68:10349–57.
    OpenUrlAbstract/FREE Full Text
  78. 78.↵
    1. Grand FH,
    2. Hidalgo-Curtis CE,
    3. Ernst T,
    4. Zoi K,
    5. Zoi C,
    6. McGuire C,
    7. et al.
    Frequent CBL mutations associated with 11q acquired uniparental disomy in myeloproliferative neoplasms. Blood 2009;113:6182–92.
    OpenUrlAbstract/FREE Full Text
  79. 79.↵
    1. Sanada M,
    2. Suzuki T,
    3. Shih LY,
    4. Otsu M,
    5. Kato M,
    6. Yamazaki S,
    7. et al.
    Gain-of-function of mutated C-CBL tumour suppressor in myeloid neoplasms. Nature 2009;460:904–8.
    OpenUrlCrossRefPubMed
  80. 80.↵
    1. Muramatsu H,
    2. Makishima H,
    3. Jankowska AM,
    4. Cazzolli H,
    5. O'Keefe C,
    6. Yoshida N,
    7. et al.
    Mutations of an E3 ubiquitin ligase c-Cbl but not TET2 mutations are pathogenic in juvenile myelomonocytic leukemia. Blood 2010;115:1969–75.
    OpenUrlAbstract/FREE Full Text
  81. 81.↵
    1. Makishima H,
    2. Cazzolli H,
    3. Szpurka H,
    4. Dunbar A,
    5. Tiu R,
    6. Huh J,
    7. et al.
    Mutations of e3 ubiquitin ligase cbl family members constitute a novel common pathogenic lesion in myeloid malignancies. J Clin Oncol 2009;27:6109–16.
    OpenUrlAbstract/FREE Full Text
  82. 82.↵
    1. Niemeyer CM,
    2. Kang MW,
    3. Shin DH,
    4. Furlan I,
    5. Erlacher M,
    6. Bunin NJ,
    7. et al.
    Germline CBL mutations cause developmental abnormalities and predispose to juvenile myelomonocytic leukemia. Nat Genet 2010;42:794–800.
    OpenUrlCrossRefPubMed
  83. 83.↵
    1. Matsuda K,
    2. Shimada A,
    3. Yoshida N,
    4. Ogawa A,
    5. Watanabe A,
    6. Yajima S,
    7. et al.
    Spontaneous improvement of hematologic abnormalities in patients having juvenile myelomonocytic leukemia with specific RAS mutations. Blood 2007;109:5477–80.
    OpenUrlAbstract/FREE Full Text
  84. 84.↵
    1. Pérez B,
    2. Mechinaud F,
    3. Galambrun C,
    4. Ben Romdhane N,
    5. Isidor B,
    6. Philip N,
    7. et al.
    Germline mutations of the CBL gene define a new genetic syndrome with predisposition to juvenile myelomonocytic leukaemia. J Med Genet 2010;47:686–91.
    OpenUrlAbstract/FREE Full Text
  85. 85.↵
    1. Schmidt MH,
    2. Dikic I
    . The Cbl interactome and its functions. Nat Rev Mol Cell Biol 2005;6:907–18.
    OpenUrlCrossRefPubMed
  86. 86.↵
    1. Naramura M,
    2. Nandwani N,
    3. Gu H,
    4. Band V,
    5. Band H
    . Rapidly fatal myeloproliferative disorders in mice with deletion of Casitas B-cell lymphoma (Cbl) and Cbl-b in hematopoietic stem cells. Proc Natl Acad Sci U S A 2010;107:16274–9.
    OpenUrlAbstract/FREE Full Text
  87. 87.↵
    1. Rathinam C,
    2. Thien CB,
    3. Flavell RA,
    4. Langdon WY
    . Myeloid leukemia development in c-Cbl RING finger mutant mice is dependent on FLT3 signaling. Cancer Cell 2010;18:341–52.
    OpenUrlCrossRefPubMed
  88. 88.↵
    1. Sugimoto Y,
    2. Muramatsu H,
    3. Makishima H,
    4. Prince C,
    5. Jankowska AM,
    6. Yoshida N,
    7. et al.
    Spectrum of molecular defects in juvenile myelomonocytic leukaemia includes ASXL1 mutations. Br J Haematol 2010;150:83–7.
    OpenUrlPubMed
  89. 89.↵
    1. Gratias EJ,
    2. Liu YL,
    3. Meleth S,
    4. Castleberry RP,
    5. Emanuel PD
    . Activating FLT3 mutations are rare in children with juvenile myelomonocytic leukemia. Pediatr Blood Cancer 2005;44:142–6.
    OpenUrlCrossRefPubMed
  90. 90.↵
    1. Bergstraesser E,
    2. Hasle H,
    3. Rogge T,
    4. Fischer A,
    5. Zimmermann M,
    6. Noellke P,
    7. et al.
    Non-hematopoietic stem cell transplantation treatment of juvenile myelomonocytic leukemia: a retrospective analysis and definition of response criteria. Pediatr Blood Cancer 2007;49:629–33.
    OpenUrlCrossRefPubMed
  91. 91.↵
    1. Boguski MS,
    2. McCormick F
    . Proteins regulating Ras and its relatives. Nature 1993;366:643–54.
    OpenUrlCrossRefPubMed
  92. 92.↵
    1. Castleberry RP,
    2. Loh ML,
    3. Jayaprakash N,
    4. Peterson A,
    5. Casey V,
    6. Chang M,
    7. et al.
    Phase II window study of the farnesyltransferase inhibitor R115777 (Zarnestra(R)) in untreated juvenile myelomonocytic leukemia (JMML): A Children's Oncology Group Study. Blood [ASH Annual Meeting Abstracts] 2005;106:2587.
    OpenUrl
  93. 93.↵
    1. Nazarian R,
    2. Shi H,
    3. Wang Q,
    4. Kong X,
    5. Koya RC,
    6. Lee H,
    7. et al.
    Melanomas acquire resistance to B-RAF(V600E) inhibition by RTK or N-RAS upregulation. Nature 2010;468:973–7.
    OpenUrlCrossRefPubMed
  94. 94.↵
    1. Escudier B,
    2. Eisen T,
    3. Stadler WM,
    4. Szczylik C,
    5. Oudard S,
    6. Siebels M,
    7. et al.,
    8. TARGET Study Group
    . Sorafenib in advanced clear-cell renal-cell carcinoma. N Engl J Med 2007;356:125–34.
    OpenUrlCrossRefPubMed
  95. 95.↵
    1. Liu L,
    2. Cao Y,
    3. Chen C,
    4. Zhang X,
    5. McNabola A,
    6. Wilkie D,
    7. et al.
    Sorafenib blocks the RAF/MEK/ERK pathway, inhibits tumor angiogenesis, and induces tumor cell apoptosis in hepatocellular carcinoma model PLC/PRF/5. Cancer Res 2006;66:11851–8.
    OpenUrlAbstract/FREE Full Text
  96. 96.↵
    1. Wilhelm S,
    2. Carter C,
    3. Lynch M,
    4. Lowinger T,
    5. Dumas J,
    6. Smith RA,
    7. et al.
    Discovery and development of sorafenib: a multikinase inhibitor for treating cancer. Nat Rev Drug Discov 2006;5:835–44.
    OpenUrlCrossRefPubMed
  97. 97.↵
    1. Adnane L,
    2. Trail PA,
    3. Taylor I,
    4. Wilhelm SM
    . Sorafenib (BAY 43-9006, Nexavar), a dual-action inhibitor that targets RAF/MEK/ERK pathway in tumor cells and tyrosine kinases VEGFR/PDGFR in tumor vasculature. Methods Enzymol 2006;407:597–612.
    OpenUrlCrossRefPubMed
  98. 98.↵
    1. Lyubynska N,
    2. Gorman MF,
    3. Lauchle JO,
    4. Hong WX,
    5. Akutagawa JK,
    6. Shannon K,
    7. et al.
    A MEK inhibitor abrogates myeloproliferative disease in Kras mutant mice. Sci Transl Med 2011;3:76ra27.
    OpenUrlAbstract/FREE Full Text
  99. 99.↵
    1. Lauchle JO,
    2. Kim D,
    3. Le DT,
    4. Akagi K,
    5. Crone M,
    6. Krisman K,
    7. et al.
    Response and resistance to MEK inhibition in leukaemias initiated by hyperactive Ras. Nature 2009;461:411–4.
    OpenUrlCrossRefPubMed
  100. 100.↵
    1. Kotecha N,
    2. Flores NJ,
    3. Irish JM,
    4. Simonds EF,
    5. Sakai DS,
    6. Archambeault S,
    7. et al.
    Single-cell profiling identifies aberrant STAT5 activation in myeloid malignancies with specific clinical and biologic correlates. Cancer Cell 2008;14:335–43.
    OpenUrlCrossRefPubMed
  101. 101.↵
    1. Gaipa G,
    2. Bugarin C,
    3. Longoni D,
    4. Cesana S,
    5. Molteni C,
    6. Faini A,
    7. et al.
    Aberrant GM-CSF signal transduction pathway in juvenile myelomonocytic leukemia assayed by flow cytometric intracellular STAT5 phosphorylation measurement. Leukemia 2009;23:791–3.
    OpenUrlCrossRefPubMed
  102. 102.↵
    1. Baxter EJ,
    2. Scott LM,
    3. Campbell PJ,
    4. East C,
    5. Fourouclas N,
    6. Swanton S,
    7. et al.,
    8. Cancer Genome Project
    . Acquired mutation of the tyrosine kinase JAK2 in human myeloproliferative disorders. Lancet 2005;365:1054–61.
    OpenUrlCrossRefPubMed
  103. 103.↵
    1. James C,
    2. Ugo V,
    3. Le Couédic JP,
    4. Staerk J,
    5. Delhommeau F,
    6. Lacout C,
    7. et al.
    A unique clonal JAK2 mutation leading to constitutive signalling causes polycythaemia vera. Nature 2005;434:1144–8.
    OpenUrlCrossRefPubMed
  104. 104.↵
    1. Levine RL,
    2. Wadleigh M,
    3. Cools J,
    4. Ebert BL,
    5. Wernig G,
    6. Huntly BJ,
    7. et al.
    Activating mutation in the tyrosine kinase JAK2 in polycythemia vera, essential thrombocythemia, and myeloid metaplasia with myelofibrosis. Cancer Cell 2005;7:387–97.
    OpenUrlCrossRefPubMed
  105. 105.↵
    1. Quintás-Cardama A,
    2. Vaddi K,
    3. Liu P,
    4. Manshouri T,
    5. Li J,
    6. Scherle PA,
    7. et al.
    Preclinical characterization of the selective JAK1/2 inhibitor INCB018424: therapeutic implications for the treatment of myeloproliferative neoplasms. Blood 2010;115:3109–17.
    OpenUrlAbstract/FREE Full Text
  106. 106.↵
    1. Verstovsek S,
    2. Kantarjian H,
    3. Mesa RA,
    4. Pardanani AD,
    5. Cortes-Franco J,
    6. Thomas DA,
    7. et al.
    Safety and efficacy of INCB018424, a JAK1 and JAK2 inhibitor, in myelofibrosis. N Engl J Med 2010;363:1117–27.
    OpenUrlCrossRefPubMed
  107. 107.↵
    1. Pui CH,
    2. Relling MV,
    3. Downing JR
    . Acute lymphoblastic leukemia. N Engl J Med 2004;350:1535–48.
    OpenUrlCrossRefPubMed
  108. 108.
    1. Martinelli G,
    2. Iacobucci I,
    3. Storlazzi CT,
    4. Vignetti M,
    5. Paoloni F,
    6. Cilloni D,
    7. et al.
    IKZF1 (Ikaros) deletions in BCR-ABL1-positive acute lymphoblastic leukemia are associated with short disease-free survival and high rate of cumulative incidence of relapse: a GIMEMA AL WP report. J Clin Oncol 2009;27:5202–7.
    OpenUrlAbstract/FREE Full Text
  109. 109.
    1. Treviño LR,
    2. Yang W,
    3. French D,
    4. Hunger SP,
    5. Carroll WL,
    6. Devidas M,
    7. et al.
    Germline genomic variants associated with childhood acute lymphoblastic leukemia. Nat Genet 2009;41:1001–5.
    OpenUrlCrossRefPubMed
  110. 110.
    1. Papaemmanuil E,
    2. Hosking FJ,
    3. Vijayakrishnan J,
    4. Price A,
    5. Olver B,
    6. Sheridan E,
    7. et al.
    Loci on 7p12.2, 10q21.2 and 14q11.2 are associated with risk of childhood acute lymphoblastic leukemia.Nat Genet 2009;41:1006–10.
    OpenUrlCrossRefPubMed
  111. 111.
    1. Bercovich D,
    2. Ganmore I,
    3. Scott LM,
    4. Wainreb G,
    5. Birger Y,
    6. Elimelech A,
    7. et al.
    Mutations of JAK2 in acute lymphoblastic leukaemias associated with Down's syndrome. Lancet 2008;372:1484–92.
    OpenUrlCrossRefPubMed
  112. 112.
    1. Gaikwad A,
    2. Rye CL,
    3. Devidas M,
    4. Heerema NA,
    5. Carroll AJ,
    6. Izraeli S,
    7. et al.
    Prevalence and clinical correlates of JAK2 mutations in Down syndrome acute lymphoblastic leukaemia. Br J Haematol 2009;144:930–2.
    OpenUrlCrossRefPubMed
  113. 113.
    1. Kearney L,
    2. Gonzalez De Castro D,
    3. Yeung J,
    4. Procter J,
    5. Horsley SW,
    6. Eguchi-Ishimae M,
    7. et al.
    Specific JAK2 mutation (JAK2R683) and multiple gene deletions in Down syndrome acute lymphoblastic leukemia. Blood 2009;113:646–8.
    OpenUrlAbstract/FREE Full Text
  114. 114.
    1. Mullighan CG,
    2. Zhang J,
    3. Kasper LH,
    4. Lerach S,
    5. Payne-Turner D,
    6. Phillips LA,
    7. et al.
    CREBBP mutations in relapsed acute lymphoblastic leukaemia. Nature 2011;471:235–9.
    OpenUrlCrossRefPubMed
  115. 115.
    1. Pasqualucci L,
    2. Dominguez-Sola D,
    3. Chiarenza A,
    4. Fabbri G,
    5. Grunn A,
    6. Trifonov V,
    7. et al.
    Inactivating mutations of acetyltransferase genes in B-cell lymphoma. Nature 2011;471:189–95.
    OpenUrlCrossRefPubMed
  116. 116.↵
    1. Chan RJ,
    2. Cooper T,
    3. Kratz CP,
    4. Weiss B,
    5. Loh ML
    . Juvenile myelomonocytic leukemia: a report from the 2nd International JMML Symposium. Leuk Res 2009;33:355–62.
    OpenUrlCrossRefPubMed
View Abstract
PreviousNext
Back to top
Clinical Cancer Research: 18 (10)
May 2012
Volume 18, Issue 10
  • Table of Contents
  • Table of Contents (PDF)
  • About the Cover

Sign up for alerts

View this article with LENS

Open full page PDF
Article Alerts
Sign In to Email Alerts with your Email Address
Email Article

Thank you for sharing this Clinical Cancer Research article.

NOTE: We request your email address only to inform the recipient that it was you who recommended this article, and that it is not junk mail. We do not retain these email addresses.

Enter multiple addresses on separate lines or separate them with commas.
Advances in the Genetics of High-Risk Childhood B-Progenitor Acute Lymphoblastic Leukemia and Juvenile Myelomonocytic Leukemia: Implications for Therapy
(Your Name) has forwarded a page to you from Clinical Cancer Research
(Your Name) thought you would be interested in this article in Clinical Cancer Research.
CAPTCHA
This question is for testing whether or not you are a human visitor and to prevent automated spam submissions.
Citation Tools
Advances in the Genetics of High-Risk Childhood B-Progenitor Acute Lymphoblastic Leukemia and Juvenile Myelomonocytic Leukemia: Implications for Therapy
Mignon L. Loh and Charles G. Mullighan
Clin Cancer Res May 15 2012 (18) (10) 2754-2767; DOI: 10.1158/1078-0432.CCR-11-1936

Citation Manager Formats

  • BibTeX
  • Bookends
  • EasyBib
  • EndNote (tagged)
  • EndNote 8 (xml)
  • Medlars
  • Mendeley
  • Papers
  • RefWorks Tagged
  • Ref Manager
  • RIS
  • Zotero
Share
Advances in the Genetics of High-Risk Childhood B-Progenitor Acute Lymphoblastic Leukemia and Juvenile Myelomonocytic Leukemia: Implications for Therapy
Mignon L. Loh and Charles G. Mullighan
Clin Cancer Res May 15 2012 (18) (10) 2754-2767; DOI: 10.1158/1078-0432.CCR-11-1936
del.icio.us logo Digg logo Reddit logo Twitter logo CiteULike logo Facebook logo Google logo Mendeley logo
  • Tweet Widget
  • Facebook Like
  • Google Plus One

Jump to section

  • Article
    • Abstract
    • Acute Lymphoblastic Leukemia
    • Clonal Heterogeneity in Leukemia
    • Juvenile Myelomonocytic Leukemia
    • Disclosure of Potential Conflicts of Interest
    • Authors' Contributions
    • Grant Support
    • Acknowledgments
    • References
  • Figures & Data
  • Info & Metrics
  • PDF
Advertisement

Related Articles

Cited By...

More in this TOC Section

  • Endpoints for Immuno-oncology Trials
  • Limitations and Challenges in Immuno-oncology Trials
  • Developing Early-Phase Combination Immunotherapy Trials
Show more CCR Focus
  • Home
  • Alerts
  • Feedback
  • Privacy Policy
Facebook  Twitter  LinkedIn  YouTube  RSS

Articles

  • Online First
  • Current Issue
  • Past Issues
  • CCR Focus Archive
  • Meeting Abstracts

Info for

  • Authors
  • Subscribers
  • Advertisers
  • Librarians

About Clinical Cancer Research

  • About the Journal
  • Editorial Board
  • Permissions
  • Submit a Manuscript
AACR logo

Copyright © 2021 by the American Association for Cancer Research.

Clinical Cancer Research
eISSN: 1557-3265
ISSN: 1078-0432

Advertisement