Skip to main content Accessibility help
×
Hostname: page-component-77c89778f8-7drxs Total loading time: 0 Render date: 2024-07-20T12:01:37.411Z Has data issue: false hasContentIssue false

7 - Molecular genetics of acute lymphoblastic leukemia

from Section 2 - Cell biology and pathobiology

Published online by Cambridge University Press:  05 April 2013

Ching-Hon Pui
Affiliation:
St Jude's Children's Research Hospital
Get access

Summary

Introduction

The majority of cases of acute lymphoblastic leukemia (ALL) are known to have a genetic basis. Most leukemic cells harbor gross chromosomal alterations that have key roles in initiating leukemogenesis. These alterations include gains and losses of whole chromosomes (hyper- and hypodiploidy), and structural alterations including translocations, inversions, deletions, and amplifications. Identification, cloning, and functional analysis of genes disrupted or rearranged by these alterations has led to characterization of numerous oncogenes and tumor suppressors with key roles in leukemogenesis. Detection of these alterations by cytogenetic techniques (karyotyping or fluorescence in situ hybridization [FISH]) and molecular analyses is extensively used in diagnosis and risk stratification in order to assign appropriate treatment. While the importance of these gross chromosomal alterations in leukemogenesis has been known for many years, it has also been long recognized that many of the chromosomal alterations are insufficient to cause leukemia in experimental models, or may be detected many years prior to the clinical onset of leukemia. Moreover, a substantial minority of children (and a higher proportion of adults) lack a gross chromosomal alteration on cytogenetic analysis. Together, these findings indicate that additional submicroscopic genetic alterations also contribute to leukemogenesis. With the advent of technology to interrogate genome-wide changes in DNA copy number (i.e., amplifications and deletions) at very high resolution, and to identify sequence mutations in a high-throughput manner, it has become clear that ALL genomes commonly harbor both gross and submicroscopic genetic alterations. These alterations commonly involve genes that regulate lymphoid development, cell cycle, tumor suppression, and a variety of other key cellular pathways, and emerging data indicate that many of these alterations influence leukemogenesis and treatment responsiveness.

Type
Chapter
Information
Childhood Leukemias , pp. 168 - 203
Publisher: Cambridge University Press
Print publication year: 2012

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Pui, CH, Robison, LL, Look, AT. Acute lymphoblastic leukaemia. Lancet 2008;371:1030–1043.CrossRefGoogle ScholarPubMed
Bishop, JM. The molecular genetics of cancer. Science 1987;235:305–311.CrossRefGoogle ScholarPubMed
Solomon, E, Borrow, J, Goddard, AD. Chromosome aberrations and cancer. Science 1991;254:1153–1160.CrossRefGoogle ScholarPubMed
Rowley, JD. Molecular cytogenetics: Rosetta stone for understanding cancer – Twenty-Ninth G. H. A. Clowes Memorial Award Lecture. Cancer Res 1990;50:3816–3825.Google Scholar
Rabbitts, TH. Chromosomal translocations in human cancer. Nature 1994;372:143–149.CrossRefGoogle ScholarPubMed
Ferrando, AA, Look, AT. Clinical implications of recurring chromosomal and associated molecular abnormalities in acute lymphoblastic leukemia. Semin Hematol 2000;37:381–395.CrossRefGoogle ScholarPubMed
Look, AT. Oncogenic transcription factors in the human acute leukemias. Science 1997;278:1059–1064.CrossRefGoogle ScholarPubMed
O'Neil, J, Look, AT. Mechanisms of transcription factor deregulation in lymphoid cell transformation. Oncogene 2007;26:6838–6849.CrossRefGoogle ScholarPubMed
Harrison, CJ. Cytogenetics of paediatric and adolescent acute lymphoblastic leukaemia. Br J Haematol 2009;144:147–156.CrossRefGoogle ScholarPubMed
Cleary, ML. Oncogenic conversion of transcription factors by chromosomal translocations. Cell 1991;66:619–622.CrossRefGoogle ScholarPubMed
Sawyers, CL. Molecular genetics of acute leukaemia. Lancet 1997;349:196–200.CrossRefGoogle ScholarPubMed
O'Connor, NT, Wainscoat, JS, Weatherall, DJ, et al. Rearrangement of the T-cell-receptor beta-chain gene in the diagnosis of lymphoproliferative disorders. Lancet 1985;i:1295–1297.CrossRefGoogle Scholar
Wiemels, J, Kang, M, Greaves, M. Backtracking of leukemic clones to birth. Methods Mol Biol 2009;538:7–27.CrossRefGoogle Scholar
Mullighan, CG. Single nucleotide polymorphism microarray analysis of genetic alterations in cancer. Methods Mol Biol 2011;730:235–258.CrossRefGoogle Scholar
Heinrichs, S, Li, C, Look, AT. SNP array analysis in hematologic malignancies: avoiding false discoveries. Blood 2010;115:4157–4161.CrossRefGoogle ScholarPubMed
Meyerson, M, Gabriel, S, Getz, G. Advances in understanding cancer genomes through second-generation sequencing. Nat Rev Genet 2010;11:685–696.CrossRefGoogle ScholarPubMed
Mullighan, CG, Goorha, S, Radtke, I, et al. Genome-wide analysis of genetic alterations in acute lymphoblastic leukaemia. Nature 2007;446:758–764.CrossRefGoogle ScholarPubMed
Treviño, LR, Yang, W, French, D, et al. Germline genomic variants associated with childhood acute lymphoblastic leukemia. Nat Genet 2009;41:1001–1005.CrossRefGoogle ScholarPubMed
Papaemmanuil, E, Hosking, FJ, Vijayakrishnan, J, et al. Loci on 7p12.2, 10q21.2 and 14q11.2 are associated with risk of childhood acute lymphoblastic leukemia. Nat Genet 2009;41:1006–1010.CrossRefGoogle ScholarPubMed
Prasad, RB, Hosking, FJ, Vijayakrishnan, J, et al. Verification of the susceptibility loci on 7p12.2, 10q21.2, and 14q11.2 in precursor B-cell acute lymphoblastic leukemia of childhood. Blood 2010;115:1765–1767.CrossRefGoogle ScholarPubMed
Sherborne, AL, Hosking, FJ, Prasad, RB, et al. Variation in CDKN2A at 9p21.3 influences childhood acute lymphoblastic leukemia risk. Nat Genet 2010;42:492–494.CrossRefGoogle ScholarPubMed
Yang, JJ, Cheng, C, Devidas, M, et al. Ancestry and pharmacogenomics of relapse in acute lymphoblastic leukemia. Nat Genet 2011;43:237–241.CrossRefGoogle ScholarPubMed
Borkhardt, A, Cazzaniga, G, Viehmann, S, et al. Incidence and clinical relevance of TEL/AML1 fusion genes in children with acute lymphoblastic leukemia enrolled in the German and Italian multicenter therapy trials. Associazione Italiana Ematologia Oncologia Pediatrica and the Berlin–Frankfurt–Münster Study Group. Blood 1997;90:571–577.Google ScholarPubMed
McLean, TW, Ringold, S, Neuberg, D, et al. TEL/AML-1 dimerizes and is associated with a favorable outcome in childhood acute lymphoblastic leukemia. Blood 1996;88:4252–4258.Google ScholarPubMed
Rubnitz, JE, Downing, JR, Pui, CH, et al. TEL gene rearrangement in acute lymphoblastic leukemia: a new genetic marker with prognostic significance. J Clin Oncol 1997;15:1150–1157.CrossRefGoogle ScholarPubMed
Rubnitz, JE, Shuster, JJ, Land, VJ, et al. Case–control study suggests a favorable impact of TEL rearrangement in patients with B-lineage acute lymphoblastic leukemia treated with antimetabolite-based therapy: a Pediatric Oncology Group study. Blood 1997;89:1143–1146.Google ScholarPubMed
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 2010;11:429–438.CrossRefGoogle ScholarPubMed
Romana, SP, Mauchauffe, M, Le Coniat, M, et al. The t(12;21) of acute lymphoblastic leukemia results in a TEL-AML1 gene fusion. Blood 1995;85:3662–3670.Google Scholar
Raimondi, SC, Behm, FG, Roberson, PK, et al. Cytogenetics of pre-B-cell acute lymphoblastic leukemia with emphasis on prognostic implications of the t(1;19). J Clin Oncol 1990;8:1380–1388.CrossRefGoogle Scholar
Raimondi, SC. Current status of cytogenetic research in childhood acute lymphoblastic leukemia. Blood 1993;81:2237–2251.Google ScholarPubMed
Raimondi, SC, Frestedt, JL, Pui, CH, et al. Acute lymphoblastic leukemias with deletion of 11q23 or a novel inversion (11)(p13q23) lack MLL gene rearrangements and have favorable clinical features. Blood 1995;86:1881–1886.Google ScholarPubMed
Liang, DC, Chou, TB, Chen, JS, et al. High incidence of TEL/AML1 fusion resulting from a cryptic t(12;21) in childhood B-lineage acute lymphoblastic leukemia in Taiwan. Leukemia 1996;10:991–993.Google Scholar
Nakao, M, Yokota, S, Horiike, S, et al. Detection and quantification of TEL/AML1 fusion transcripts by polymerase chain reaction in childhood acute lymphoblastic leukemia. Leukemia 1996;10:1463–1470.Google ScholarPubMed
Amor, DJ, Algar, EM, Slater, HR, et al. High frequency of t(12;21) in childhood acute lymphoblastic leukemia detected by RT-PCR. Pathology 1998;30:381–385.CrossRefGoogle Scholar
Rubnitz, JE, Behm, FG, Pui, CH, et al. Genetic studies of childhood acute lymphoblastic leukemia with emphasis on p16, MLL, and ETV6 gene abnormalities: results of St. Jude Total Therapy Study XII. Leukemia 1997;11:1201–1206.CrossRefGoogle ScholarPubMed
Raimondi, SC, Privitera, E, Williams, DL, et al. New recurring chromosomal translocations in childhood acute lymphoblastic leukemia. Blood 1991;77:2016–2022.Google ScholarPubMed
Bohlander, SK. ETV6: a versatile player in leukemogenesis. Semin Cancer Biol 2005;15:162–174.CrossRefGoogle ScholarPubMed
Golub, TR, Barker, GF, Lovett, M, et al. Fusion of PDGF receptor beta to a novel ets-like gene, tel, in chronic myelomonocytic leukemia with t(5;12) chromosomal translocation. Cell 1994;77:307–316.CrossRefGoogle Scholar
Golub, TR, Barker, GF, Stegmaier, K, et al. The TEL gene contributes to the pathogenesis of myeloid and lymphoid leukemias by diverse molecular genetic mechanisms. Curr Top Microbiol Immunol 1997;220:67–79.Google ScholarPubMed
Okuda, T, van Deursen, J, Hiebert, SW, et al. AML1, the target of multiple chromosomal translocations in human leukemia, is essential for normal fetal liver hematopoiesis. Cell 1996;84:321–330.CrossRefGoogle ScholarPubMed
Song, WJ, Sullivan, MG, Legare, RD, et al. Haploinsufficiency of CBFA2 causes familial thrombocytopenia with propensity to develop acute myelogenous leukaemia. Nat Genet 1999;23:166–175.CrossRefGoogle ScholarPubMed
Roumier, C, Fenaux, P, Lafage, M, et al. New mechanisms of AML1 gene alteration in hematological malignancies. Leukemia 2003;17:9–16.CrossRefGoogle ScholarPubMed
Hiebert, SW, Sun, W, Davis, JN, et al. The t(12;21) translocation converts AML-1B from an activator to a repressor of transcription. Mol Cell Biol 1996;16:1349–1355.CrossRefGoogle Scholar
Guidez, F, Petrie, K, Ford, AM, et al. Recruitment of the nuclear receptor corepressor N-CoR by the TEL moiety of the childhood leukemia-associated TEL-AML1 oncoprotein. Blood 2000;96:2557–2561.Google ScholarPubMed
Wang, LC, Kuo, F, Fujiwara, Y, et al. Yolk sac angiogenic defect and intra-embryonic apoptosis in mice lacking the Ets-related factor TEL. EMBO J 1997;16:4374–4383.CrossRefGoogle ScholarPubMed
Wang, Q, Stacy, T, Binder, M, et al. Disruption of the Cbfa2 gene causes necrosis and hemorrhaging in the central nervous system and blocks definitive hematopoiesis. Proc Natl Acad Sci USA 1996;93:3444–3449.CrossRefGoogle ScholarPubMed
Sasaki, K, Yagi, H, Bronson, RT, et al. Absence of fetal liver hematopoiesis in mice deficient in transcriptional coactivator core binding factor beta. Proc Natl Acad Sci USA 1996;93:12359–12363.CrossRefGoogle ScholarPubMed
Wang, Q, Stacy, T, Miller, JD, et al. The CBFbeta subunit is essential for CBFalpha2 (AML1) function in vivo. Cell 1996;87:697–708.CrossRefGoogle ScholarPubMed
Niki, M, Okada, H, Takano, H, et al. Hematopoiesis in the fetal liver is impaired by targeted mutagenesis of a gene encoding a non-DNA binding subunit of the transcription factor, polyomavirus enhancer binding protein 2/core binding factor. Proc Natl Acad Sci USA 1997;94:5697–5702.CrossRefGoogle ScholarPubMed
Andreasson, P, Schwaller, J, Anastasiadou, E, et al. The expression of ETV6/CBFA2 (TEL/AML1) is not sufficient for the transformation of hematopoietic cell lines in vitro or the induction of hematologic disease in vivo. Cancer Genet Cytogenet 2001;130:93–104.CrossRefGoogle ScholarPubMed
Morrow, M, Horton, S, Kioussis, D, et al. TEL-AML1 promotes development of specific hematopoietic lineages consistent with preleukemic activity. Blood 2004;103:3890–3896.CrossRefGoogle ScholarPubMed
Lausten-Thomsen, U, Madsen, HO, Vestergaard, TR, et al. Prevalence of t(12;21)[ETV6-RUNX1]-positive cells in healthy neonates. Blood 2011;117:186–189.CrossRefGoogle Scholar
Jousset, C, Carron, C, Boureux, A, et al. A domain of TEL conserved in a subset of ETS proteins defines a specific oligomerization interface essential to the mitogenic properties of the TEL-PDGFR beta oncoprotein. EMBO J 1997;16:69–82.CrossRefGoogle Scholar
Cave, H, Cacheux, V, Raynaud, S, et al. ETV6 is the target of chromosome 12p deletions in t(12;21) childhood acute lymphocytic leukemia. Leukemia 1997;11:1459–1464.CrossRefGoogle Scholar
Takeuchi, S, Seriu, T, Bartram, CR, et al. TEL is one of the targets for deletion on 12p in many cases of childhood B-lineage acute lymphoblastic leukemia. Leukemia 1997;11:1220–1223.CrossRefGoogle ScholarPubMed
Parker, H, An, Q, Barber, K, et al. The complex genomic profile of ETV6-RUNX1 positive acute lymphoblastic leukemia highlights a recurrent deletion of TBL1XR1. Genes Chromosomes Cancer 2008;47:1118–1125.CrossRefGoogle ScholarPubMed
Nowell, PC, Hungerford, DA. Chromosome studies on normal and leukemic human leukocytes. J Natl Cancer Inst 1960;25:85–109.Google ScholarPubMed
Rowley, JD. A new consistent chromosomal abnormality in chronic myelogenous leukaemia identified by quinacrine fluorescence and Giemsa staining. Nature 1973;243:290–293.CrossRefGoogle ScholarPubMed
Pui, CH. Childhood leukemias. N Engl J Med 1995;332:1618–1630.CrossRefGoogle ScholarPubMed
Ribeiro, RC, Abromowitch, M, Raimondi, SC, et al. Clinical and biologic hallmarks of the Philadelphia chromosome in childhood acute lymphoblastic leukemia. Blood 1987;70:948–953.Google ScholarPubMed
Bartram, CR, de Klein, A, Hagemeijer, A, et al. Translocation of c-ab1 oncogene correlates with the presence of a Philadelphia chromosome in chronic myelocytic leukaemia. Nature 1983;306:277–280.CrossRefGoogle ScholarPubMed
Gale, RP, Canaani, E. An 8-kilobase abl RNA transcript in chronic myelogenous leukemia. Proc Natl Acad Sci USA 1984;81:5648–5652.CrossRefGoogle ScholarPubMed
Collins, SJ, Kubonishi, I, Miyoshi, I, et al. Altered transcription of the c-abl oncogene in K-562 and other chronic myelogenous leukemia cells. Science 1984;225:72–74.CrossRefGoogle ScholarPubMed
Stam, K, Heisterkamp, N, Grosveld, G, et al. Evidence of a new chimeric bcr/c-abl mRNA in patients with chronic myelocytic leukemia and the Philadelphia chromosome. N Engl J Med 1985;313:1429–1433.CrossRefGoogle ScholarPubMed
Canaani, E, Gale, RP, Steiner-Saltz, D, et al. Altered transcription of an oncogene in chronic myeloid leukaemia. Lancet 1984;i:593–595.CrossRefGoogle Scholar
Shtivelman, E, Lifshitz, B, Gale, RP, et al. Fused transcript of abl and bcr genes in chronic myelogenous leukaemia. Nature 1985;315:550–554.CrossRefGoogle ScholarPubMed
Heisterkamp, N, Stephenson, JR, Groffen, J, et al. Localization of the c-ab1 oncogene adjacent to a translocation break point in chronic myelocytic leukaemia. Nature 1983;306:239–242.CrossRefGoogle ScholarPubMed
Leibowitz, D, Schaefer-Rego, K, Popenoe, DW, et al. Variable breakpoints on the Philadelphia chromosome in chronic myelogenous leukemia. Blood 1985;66:243–245.Google ScholarPubMed
Grosveld, G, Verwoerd, T, van Agthoven, T, et al. The chronic myelocytic cell line K562 contains a breakpoint in bcr and produces a chimeric bcr/c-abl transcript. Mol Cell Biol 1986;6:607–616.CrossRefGoogle ScholarPubMed
Groffen, J, Stephenson, JR, Heisterkamp, N, et al. Philadelphia chromosomal breakpoints are clustered within a limited region, bcr, on chromosome 22. Cell 1984;36:93–99.CrossRefGoogle ScholarPubMed
Heisterkamp, N, Stam, K, Groffen, J, et al. Structural organization of the bcr gene and its role in the Ph′ translocation. Nature 1985;315:758–761.CrossRefGoogle ScholarPubMed
Chan, LC, Karhi, KK, Rayter, SI, et al. A novel abl protein expressed in Philadelphia chromosome positive acute lymphoblastic leukaemia. Nature 1987;325:635–637.CrossRefGoogle ScholarPubMed
Clark, SS, McLaughlin, J, Crist, WM, et al. Unique forms of the abl tyrosine kinase distinguish Ph1-positive CML from Ph1-positive ALL. Science 1987;235:85–88.CrossRefGoogle ScholarPubMed
Kurzrock, R, Shtalrid, M, Romero, P, et al. A novel c-abl protein product in Philadelphia-positive acute lymphoblastic leukaemia. Nature 1987;325:631–635.CrossRefGoogle ScholarPubMed
van Etten, RA, Jackson, P, Baltimore, D. The mouse type IV c-abl gene product is a nuclear protein, and activation of transforming ability is associated with cytoplasmic localization. Cell 1989;58:669–678.CrossRefGoogle ScholarPubMed
Kharbanda, S, Ren, R, Pandey, P, et al. Activation of the c-Abl tyrosine kinase in the stress response to DNA-damaging agents. Nature 1995;376:785–788.CrossRefGoogle ScholarPubMed
Sawyers, CL, McLaughlin, J, Goga, A, et al. The nuclear tyrosine kinase c-Abl negatively regulates cell growth. Cell 1994;77:121–131.CrossRefGoogle ScholarPubMed
Mattioni, T, Jackson, PK, Bchini-Hooft van Huijsduijnen, O, et al. Cell cycle arrest by tyrosine kinase Abl involves altered early mitogenic response. Oncogene 1995;10:1325–1333.Google ScholarPubMed
Goga, A, Liu, X, Hambuch, TM, et al. p53 dependent growth suppression by the c-Abl nuclear tyrosine kinase. Oncogene 1995;11:791–799.Google ScholarPubMed
Wang, JY. Regulation of cell death by the Abl tyrosine kinase. Oncogene 2000;19:5643–5650.CrossRefGoogle ScholarPubMed
Tybulewicz, VL, Crawford, CE, Jackson, PK, et al. Neonatal lethality and lymphopenia in mice with a homozygous disruption of the c-abl proto-oncogene. Cell 1991;65:1153–1163.CrossRefGoogle ScholarPubMed
Schwartzberg, PL, Stall, AM, Hardin, JD, et al. Mice homozygous for the ablm1 mutation show poor viability and depletion of selected B and T cell populations. Cell 1991;65:1165–1175.CrossRefGoogle ScholarPubMed
Lugo, TG, Pendergast, AM, Muller, AJ, et al. Tyrosine kinase activity and transformation potency of bcr-abl oncogene products. Science 1990;247:1079–1082.CrossRefGoogle ScholarPubMed
Daley, GQ, McLaughlin, J, Witte, ON, et al. The CML-specific P210 bcr/abl protein, unlike v-abl, does not transform NIH/3T3 fibroblasts. Science 1987;237:532–535.CrossRefGoogle Scholar
Daley, GQ, Baltimore, D. Transformation of an interleukin 3-dependent hematopoietic cell line by the chronic myelogenous leukemia-specific P210bcr/abl protein. Proc Natl Acad Sci USA 1988;85:9312–9316.CrossRefGoogle ScholarPubMed
Elefanty, AG, Hariharan, IK, Cory, S. bcr-abl, the hallmark of chronic myeloid leukaemia in man, induces multiple haemopoietic neoplasms in mice. EMBO J 1990;9:1069–1078.Google ScholarPubMed
Gishizky, ML, Johnson-White, J, Witte, ON. Efficient transplantation of BCR-ABL-induced chronic myelogenous leukemia-like syndrome in mice. Proc Natl Acad Sci USA 1993;90:3755–3759.CrossRefGoogle ScholarPubMed
Kelliher, M, Knott, A, McLaughlin, J, et al. Differences in oncogenic potency but not target cell specificity distinguish the two forms of the BCR/ABL oncogene. Mol Cell Biol 1991;11:4710–4716.CrossRefGoogle Scholar
van Etten, RA. Studying the pathogenesis of BCR-ABL+ leukemia in mice. Oncogene 2002;21:8643–8651.CrossRefGoogle ScholarPubMed
Cortez, D, Reuther, G, Pendergast, AM. The Bcr–Abl tyrosine kinase activates mitogenic signaling pathways and stimulates G1-to-S phase transition in hematopoietic cells. Oncogene 1997;15:2333–2342.CrossRefGoogle ScholarPubMed
Varticovski, L, Daley, GQ, Jackson, P, et al. Activation of phosphatidylinositol 3-kinase in cells expressing abl oncogene variants. Mol Cell Biol 1991;11:1107–1113.CrossRefGoogle ScholarPubMed
Reuther, JY, Reuther, GW, Cortez, D, et al. A requirement for NF-kappaB activation in Bcr-Abl-mediated transformation. Genes Dev 1998;12:968–981.CrossRefGoogle ScholarPubMed
Carlesso, N, Frank, DA, Griffin, JD. Tyrosyl phosphorylation and DNA binding activity of signal transducers and activators of transcription (STAT) proteins in hematopoietic cell lines transformed by Bcr/Abl. J Exp Med 1996;183:811–820.CrossRefGoogle ScholarPubMed
Raitano, AB, Halpern, JR, Hambuch, TM, et al. The Bcr–Abl leukemia oncogene activates Jun kinase and requires Jun for transformation. Proc Natl Acad Sci USA 1995;92:11746–11750.CrossRefGoogle ScholarPubMed
Sawyers, CL, McLaughlin, J, Witte, ON. Genetic requirement for Ras in the transformation of fibroblasts and hematopoietic cells by the Bcr–Abl oncogene. J Exp Med 1995;181:307–313.CrossRefGoogle ScholarPubMed
Skorski, T, Kanakaraj, P, Nieborowska-Skorska, M, et al. Phosphatidylinositol-3 kinase activity is regulated by BCR/ABL and is required for the growth of Philadelphia chromosome-positive cells. Blood 1995;86:726–736.Google ScholarPubMed
Skorski, T, Bellacosa, A, Nieborowska-Skorska, M, et al. Transformation of hematopoietic cells by BCR/ABL requires activation of a PI-3k/Akt-dependent pathway. EMBO J 1997;16:6151–6161.CrossRefGoogle ScholarPubMed
Skorski, T. BCR/ABL regulates response to DNA damage: the role in resistance to genotoxic treatment and in genomic instability. Oncogene 2002;21:8591–8604.CrossRefGoogle ScholarPubMed
Duy, C, Hurtz, C, Shojaee, S, et al. BCL6 enables Ph+ acute lymphoblastic leukaemia cells to survive BCR-ABL1 kinase inhibition. Nature 2011;473:384–388.CrossRefGoogle ScholarPubMed
Mullighan, CG, Miller, CB, Radtke, I, et al. BCR-ABL1 lymphoblastic leukaemia is characterized by the deletion of Ikaros. Nature 2008;453:110–114.CrossRefGoogle ScholarPubMed
Iacobucci, I, Storlazzi, CT, Cilloni, D, 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;114:2159–2167.CrossRefGoogle Scholar
Georgopoulos, K, Bigby, M, Wang, JH, et al. The Ikaros gene is required for the development of all lymphoid lineages. Cell 1994;79:143–156.CrossRefGoogle ScholarPubMed
Molnar, A, Georgopoulos, K. The Ikaros gene encodes a family of functionally diverse zinc finger DNA-binding proteins. Mol Cell Biol 1994;14:8292–8303.CrossRefGoogle ScholarPubMed
Virely, C, Moulin, S, Cobaleda, C, et al. Haploinsufficiency of the IKZF1 (IKAROS) tumor suppressor gene cooperates with BCR–ABL in a transgenic model of acute lymphoblastic leukemia. Leukemia 2010;24:1200–1204.CrossRefGoogle Scholar
Collins-Underwood, JR, Boulos, N, Payne-Turner, D, et al. The role of dominant-negative IKAROS mutations in the pathogenesis and treatment responsiveness of BCR-ABL1 positive acute lymphoblastic leukemia. [ASH Annual Meeting Abstracts.]Blood 2010;116:540.Google Scholar
Collins-Underwood, R, Miller, CB, Downing, JR, et al. Ikzf1 haploinsufficiency contributes to the pathogenesis of BCR-ABL1 positive acute lymphoblastic leukemia. [ASH Annual Meeting Abstracts.]Blood 2009;114:678.Google Scholar
Martinelli, G, Iacobucci, I, Storlazzi, CT, 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;27:5202–5207.CrossRefGoogle ScholarPubMed
Crist, W, Carroll, A, Shuster, J, et al. Philadelphia chromosome positive childhood acute lymphoblastic leukemia: clinical and cytogenetic characteristics and treatment outcome. A Pediatric Oncology Group study. Blood 1990;76:489–494.Google ScholarPubMed
Fletcher, JA, Lynch, EA, Kimball, VM, et al. Translocation (9;22) is associated with extremely poor prognosis in intensively treated children with acute lymphoblastic leukemia. Blood 1991;77:435–439.Google ScholarPubMed
Arico, M, Schrappe, M, Hunger, S, et al. Clinical outcome of 640 children with newly diagnosed Philadelphia chromosome-positive acute lymphoblastic leukemia treated between 1995 and 2005. [ASH Annual Meeting Abstracts.]Blood, 2008;112:568.Google Scholar
Arico, 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 2010;28:4755–4761.CrossRefGoogle ScholarPubMed
Arico, M, Valsecchi, MG, Camitta, B, et al. Outcome of treatment in children with Philadelphia chromosome-positive acute lymphoblastic leukemia. N Engl J Med 2000;342:998–1006.CrossRefGoogle ScholarPubMed
Roberts, WM, Rivera, GK, Raimondi, SC, et al. Intensive chemotherapy for Philadelphia-chromosome-positive acute lymphoblastic leukaemia. Lancet 1994;343:331–332.CrossRefGoogle ScholarPubMed
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 1997;11:1493–1496.CrossRefGoogle ScholarPubMed
Buchdunger, E, Zimmermann, J, Mett, H, et al. Inhibition of the Abl protein-tyrosine kinase in vitro and in vivo by a 2-phenylaminopyrimidine derivative. Cancer Res 1996;56:100–104.Google ScholarPubMed
Druker, BJ, Lydon, NB. Lessons learned from the development of an abl tyrosine kinase inhibitor for chronic myelogenous leukemia. J Clin Invest 2000;105:3–7.CrossRefGoogle ScholarPubMed
Druker, BJ, Sawyers, CL, Kantarjian, H, et al. Activity of a specific inhibitor of the BCR–ABL tyrosine kinase in the blast crisis of chronic myeloid leukemia and acute lymphoblastic leukemia with the Philadelphia chromosome. N Engl J Med 2001;344:1038–1042.CrossRefGoogle ScholarPubMed
Druker, BJ, Talpaz, M, Resta, DJ, et al. Efficacy and safety of a specific inhibitor of the BCR–ABL tyrosine kinase in chronic myeloid leukemia. N Engl J Med 2001;344:1031–1037.CrossRefGoogle ScholarPubMed
Kantarjian, H, Sawyers, C, Hochhaus, A, et al. Hematologic and cytogenetic responses to imatinib mesylate in chronic myelogenous leukemia. N Engl J Med 2002;346:645–652.CrossRefGoogle ScholarPubMed
Talpaz, M, Silver, RT, Druker, BJ, et al. Imatinib induces durable hematologic and cytogenetic responses in patients with accelerated phase chronic myeloid leukemia: results of a phase 2 study. Blood 2002;99:1928–1937.CrossRefGoogle ScholarPubMed
Kantarjian, H, Shah, NP, Hochhaus, A, et al. Dasatinib versus imatinib in newly diagnosed chronic-phase chronic myeloid leukemia. N Engl J Med 2010;362:2260–2270.CrossRefGoogle ScholarPubMed
von Bubnoff, N, Peschel, C, Duyster, J. Resistance of Philadelphia-chromosome positive leukemia towards the kinase inhibitor imatinib (STI571, Glivec): a targeted oncoprotein strikes back. Leukemia 2003;17:829–838.CrossRefGoogle ScholarPubMed
Gorre, ME, Sawyers, CL. Molecular mechanisms of resistance to STI571 in chronic myeloid leukemia. Curr Opin Hematol 2002;9:303–307.CrossRefGoogle ScholarPubMed
Hochhaus, A, La Rosee, P, Muller, MC, et al. Impact of BCR-ABL mutations on patients with chronic myeloid leukemia. Cell Cycle 2011;10:250–260.CrossRefGoogle ScholarPubMed
Pui, CH, Relling, MV, Downing, JR. Acute lymphoblastic leukemia. N Engl J Med 2004;350:1535–1548.CrossRefGoogle ScholarPubMed
Pui, CH, Carroll, WL, Meshinchi, S, et al. Biology, risk stratification, and therapy of pediatric acute leukemias: an update. J Clin Oncol 2011;29:551–565.CrossRefGoogle ScholarPubMed
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 2009;27:5175–5181.CrossRefGoogle ScholarPubMed
Schultz, KR, Bowman, WP, Aledo, A, et al. Continuous dosing imatinib with intensive chemotherapy gives equivalent outcomes to allogeneic BMT for Philadelphia chromosome-positive (Ph+) acute lymphoblastic leukemia (ALL) with longer term follow up: updated results of Children's Oncology Group (COG) AALL0031. Pediatr Blood Cancer 2010;54:788.Google Scholar
Privitera, E, Kamps, MP, Hayashi, Y, et al. Different molecular consequences of the 1;19 chromosomal translocation in childhood B-cell precursor acute lymphoblastic leukemia. Blood 1992;79:1781–1788.Google ScholarPubMed
Hunger, SP. Chromosomal translocations involving the E2A gene in acute lymphoblastic leukemia: clinical features and molecular pathogenesis. Blood 1996;87:1211–1224.Google ScholarPubMed
Mellentin, JD, Murre, C, Donlon, TA, et al. The gene for enhancer binding proteins E12/E47 lies at the t(1;19) breakpoint in acute leukemias. Science 1989;246:379–382.CrossRefGoogle Scholar
Kamps, MP, Murre, C, Sun, XH, et al. A new homeobox gene contributes the DNA binding domain of the t(1;19) translocation protein in pre-B ALL. Cell 1990;60:547–555.CrossRefGoogle Scholar
Nourse, J, Mellentin, JD, Galili, N, et al. Chromosomal translocation t(1;19) results in synthesis of a homeobox fusion mRNA that codes for a potential chimeric transcription factor. Cell 1990;60:535–545.CrossRefGoogle Scholar
Sigvardsson, M, O'Riordan, M, Grosschedl, R. EBF and E47 collaborate to induce expression of the endogenous immunoglobulin surrogate light chain genes. Immunity 1997;7:25–36.CrossRefGoogle ScholarPubMed
Bain, G, Romanow, WJ, Albers, K, et al. Positive and negative regulation of V(D)J recombination by the E2A proteins. J Exp Med 1999;189:289–300.CrossRefGoogle Scholar
Zhuang, Y, Soriano, P, Weintraub, H. The helix-loop-helix gene E2A is required for B cell formation. Cell 1994;79:875–884.CrossRefGoogle ScholarPubMed
Bain, G, Engel, I, Robanus Maandag, EC, et al. E2A deficiency leads to abnormalities in alphabeta T-cell development and to rapid development of T-cell lymphomas. Mol Cell Biol 1997;17:4782–4791.CrossRefGoogle ScholarPubMed
Bain, G, Maandag, EC, Izon, DJ, et al. E2A proteins are required for proper B cell development and initiation of immunoglobulin gene rearrangements. Cell 1994;79:885–892.CrossRefGoogle Scholar
Rauskolb, C, Wieschaus, E. Coordinate regulation of downstream genes by extradenticle and the homeotic selector proteins. EMBO J 1994;13:3561–3569.Google ScholarPubMed
Flegel, WA, Singson, AW, Margolis, JS, et al. Dpbx, a new homeobox gene closely related to the human proto-oncogene pbx1 molecular structure and developmental expression. Mech Dev 1993;41:155–161.CrossRefGoogle ScholarPubMed
Rauskolb, C, Peifer, M, Wieschaus, E. extradenticle, a regulator of homeotic gene activity, is a homolog of the homeobox-containing human proto-oncogene pbx1. Cell 1993;74:1101–1112.CrossRefGoogle ScholarPubMed
Peifer, M, Wieschaus, E. Mutations in the Drosophila gene extradenticle affect the way specific homeo domain proteins regulate segmental identity. Genes Dev 1990;4:1209–1223.CrossRefGoogle ScholarPubMed
Kamps, MP. E2A–Pbx1 induces growth, blocks differentiation, and interacts with other homeodomain proteins regulating normal differentiation. Curr Top Microbiol Immunol 1997;220:25–43.Google ScholarPubMed
Sanyal, M, Tung, JW, Karsunky, H, et al. B-cell development fails in the absence of the Pbx1 proto-oncogene. Blood 2007;109:4191–4199.CrossRefGoogle ScholarPubMed
Chang, CP, Jacobs, Y, Nakamura, T, et al. Meis proteins are major in vivo DNA binding partners for wild-type but not chimeric Pbx proteins. Mol Cell Biol 1997;17:5679–5687.CrossRefGoogle Scholar
Lu, Q, Kamps, MP. Heterodimerization of Hox proteins with Pbx1 and oncoprotein E2a-Pbx1 generates unique DNA-binding specifities at nucleotides predicted to contact the N-terminal arm of the Hox homeodomain: demonstration of Hox-dependent targeting of E2a–Pbx1 in vivo. Oncogene 1997;14:75–83.CrossRefGoogle ScholarPubMed
Fu, X, Kamps, MP. E2a–Pbx1 induces aberrant expression of tissue-specific and developmentally regulated genes when expressed in NIH 3T3 fibroblasts. Mol Cell Biol 1997;17:1503–1512.CrossRefGoogle ScholarPubMed
Dedera, DA, Waller, EK, LeBrun, DP, et al. Chimeric homeobox gene E2A-PBX1 induces proliferation, apoptosis, and malignant lymphomas in transgenic mice. Cell 1993;74:833–843.CrossRefGoogle ScholarPubMed
LeBrun, DP, Cleary, ML. Fusion with E2A alters the transcriptional properties of the homeodomain protein PBX1 in t(1;19) leukemias. Oncogene 1994;9:1641–1647.Google Scholar
Kamps, MP, Baltimore, D. E2A–Pbx1, the t(1;19) translocation protein of human pre-B-cell acute lymphocytic leukemia, causes acute myeloid leukemia in mice. Mol Cell Biol 1993;13:351–357.CrossRefGoogle Scholar
Kamps, MP, Wright, DD. Oncoprotein E2A–Pbx1 immortalizes a myeloid progenitor in primary marrow cultures without abrogating its factor-dependence. Oncogene 1994;9:3159–3166.Google ScholarPubMed
Monica, K, LeBrun, DP, Dedera, DA, et al. Transformation properties of the E2a–Pbx1 chimeric oncoprotein: fusion with E2a is essential, but the Pbx1 homeodomain is dispensable. Mol Cell Biol 1994;14:8304–8314.CrossRefGoogle ScholarPubMed
LeBrun, DP, Matthews, BP, Feldman, BJ, et al. The chimeric oncoproteins E2A–PBX1 and E2A–HLF are concentrated within spherical nuclear domains. Oncogene 1997;15:2059–2067.CrossRefGoogle ScholarPubMed
Smith, KS, Jacobs, Y, Chang, CP, et al. Chimeric oncoprotein E2a–Pbx1 induces apoptosis of hematopoietic cells by a p53-independent mechanism that is suppressed by Bcl-2. Oncogene 1997;14:2917–2926.CrossRefGoogle ScholarPubMed
Bijl, J, Sauvageau, M, Thompson, A, et al. High incidence of proviral integrations in the Hoxa locus in a new model of E2a–PBX1-induced B-cell leukemia. Genes Dev 2005;19:224–233.CrossRefGoogle Scholar
Bijl, J, Krosl, J, Lebert-Ghali, CE, et al. Evidence for Hox and E2A–PBX1 collaboration in mouse T-cell leukemia. Oncogene 2008;27:6356–6364.CrossRefGoogle ScholarPubMed
Hunger, SP, Ohyashiki, K, Toyama, K, et al. Hlf, a novel hepatic bZIP protein, shows altered DNA-binding properties following fusion to E2A in t(17;19) acute lymphoblastic leukemia. Genes Dev 1992;6:1608–1620.CrossRefGoogle Scholar
Inaba, T, Roberts, WM, Shapiro, LH, et al. Fusion of the leucine zipper gene HLF to the E2A gene in human acute B-lineage leukemia. Science 1992;257:531–534.CrossRefGoogle ScholarPubMed
Landschulz, WH, Johnson, PF, McKnight, SL. The leucine zipper: a hypothetical structure common to a new class of DNA binding proteins. Science 1988;240:1759–1764.CrossRefGoogle ScholarPubMed
O'Shea, EK, Klemm, JD, Kim, PS, et al. X-ray structure of the GCN4 leucine zipper, a two-stranded, parallel coiled coil. Science 1991;254:539–544.CrossRefGoogle ScholarPubMed
Drolet, DW, Scully, KM, Simmons, DM, et al. TEF, a transcription factor expressed specifically in the anterior pituitary during embryogenesis, defines a new class of leucine zipper proteins. Genes Dev 1991;5:1739–1753.CrossRefGoogle ScholarPubMed
Wasylyk, B, Hahn, SL, Giovane, A. The Ets family of transcription factors. Eur J Biochem 1993;211:7–18.CrossRefGoogle ScholarPubMed
Falvey, E, Fleury-Olela, F, Schibler, U. The rat hepatic leukemia factor (HLF) gene encodes two transcriptional activators with distinct circadian rhythms, tissue distributions and target preferences. EMBO J 1995;14: 4307–4317.Google ScholarPubMed
Hunger, SP, Brown, R, Cleary, ML. DNA-binding and transcriptional regulatory properties of hepatic leukemia factor (HLF) and the t(17;19) acute lymphoblastic leukemia chimera E2A–HLF. Mol Cell Biol 1994;14:5986–5996.CrossRefGoogle Scholar
Inaba, T, Shapiro, LH, Funabiki, T, et al. DNA-binding specificity and trans-activating potential of the leukemia-associated E2A-hepatic leukemia factor fusion protein. Mol Cell Biol 1994;14:3403–3413.CrossRefGoogle ScholarPubMed
Yoshihara, T, Inaba, T, Shapiro, LH, et al. E2A-HLF-mediated cell transformation requires both the trans-activation domains of E2A and the leucine zipper dimerization domain of HLF. Mol Cell Biol 1995;15:3247–3255.CrossRefGoogle ScholarPubMed
Smith, KS, Rhee, JW, Naumovski, L, et al. Disrupted differentiation and oncogenic transformation of lymphoid progenitors in E2A-HLF transgenic mice. Mol Cell Biol 1999;19:4443–4451.CrossRefGoogle ScholarPubMed
Inukai, T, Inaba, T, Yoshihara, T, et al. Cell transformation mediated by homodimeric E2A–HLF transcription factors. Mol Cell Biol 1997;17: 1417–1424.CrossRefGoogle ScholarPubMed
Seidel, MG, Look, AT. E2A–HLF usurps control of evolutionarily conserved survival pathways. Oncogene 2001;20:5718–5725.CrossRefGoogle ScholarPubMed
Smith, KS, Rhee, JW, Cleary, ML. Transformation of bone marrow B-cell progenitors by E2a–Hlf requires coexpression of Bcl-2. Mol Cell Biol 2002;22:7678–7687.CrossRefGoogle ScholarPubMed
de Boer, J, Yeung, J, Ellu, J, et al. The E2A–HLF oncogenic fusion protein acts through Lmo2 and Bcl-2 to immortalize hematopoietic progenitors. Leukemia 2011;25:321–330.CrossRefGoogle ScholarPubMed
Hirose, K, Inukai, T, Kikuchi, J, et al. Aberrant induction of LMO2 by the E2A–HLF chimeric transcription factor and its implication in leukemogenesis of B-precursor ALL with t(17;19). Blood 2010;116:962–970.CrossRefGoogle Scholar
Hunger, SP, Devaraj, PE, Foroni, L, et al. Two types of genomic rearrangements create alternative E2A–HLF fusion proteins in t(17;19)-ALL. Blood 1994;83:2970–2977.Google Scholar
Ohyashiki, K, Fujieda, H, Miyauchi, J, et al. Establishment of a novel heterotransplantable acute lymphoblastic leukemia cell line with a t(17;19) chromosomal translocation the growth of which is inhibited by interleukin-3. Leukemia 1991;5:322–331.Google Scholar
Devaraj, PE, Foroni, L, Sekhar, M, et al. E2A/HLF fusion cDNAs and the use of RT-PCR for the detection of minimal residual disease in t(17;19)(q22;p13) acute lymphoblastic leukemia. Leukemia 1994;8:1131–1138.Google Scholar
Kaneko, Y, Maseki, N, Takasaki, N, et al. Clinical and hematologic characteristics in acute leukemia with 11q23 translocations. Blood 1986;67: 484–491.Google ScholarPubMed
Raimondi, SC, Kalwinsky, DK, Hayashi, Y, et al. Cytogenetics of childhood acute nonlymphocytic leukemia. Cancer Genet Cytogenet 1989;40:13–27.CrossRefGoogle ScholarPubMed
Pui, CH, Frankel, LS, Carroll, AJ, et al. Clinical characteristics and treatment outcome of childhood acute lymphoblastic leukemia with the t(4;11)(q21;q23): a collaborative study of 40 cases. Blood 1991;77:440–447.Google Scholar
Heerema, NA, Arthur, DC, Sather, H, et al. Cytogenetic features of infants less than 12 months of age at diagnosis of acute lymphoblastic leukemia: impact of the 11q23 breakpoint on outcome: a report of the Childrens Cancer Group. Blood 1994;83:2274–2284.Google ScholarPubMed
Pui, CH, Kane, JR, Crist, WM. Biology and treatment of infant leukemias. Leukemia 1995;9:762–769.Google ScholarPubMed
Chen, CS, Sorensen, PH, Domer, PH, et al. Molecular rearrangements on chromosome 11q23 predominate in infant acute lymphoblastic leukemia and are associated with specific biologic variables and poor outcome. Blood 1993;81:2386–2393.Google ScholarPubMed
Pui, CH, Behm, FG, Raimondi, SC, et al. Secondary acute myeloid leukemia in children treated for acute lymphoid leukemia. N Engl J Med 1989;321:136–142.CrossRefGoogle ScholarPubMed
Rubnitz, JE, Behm, FG, Downing, JR. 11q23 rearrangements in acute leukemia. Leukemia 1996;10:74–82.Google ScholarPubMed
Chessells, JM, Harrison, CJ, Kempski, H, et al. Clinical features, cytogenetics and outcome in acute lymphoblastic and myeloid leukaemia of infancy: report from the MRC Childhood Leukaemia working party. Leukemia 2002;16:776–784.CrossRefGoogle ScholarPubMed
Group Francais de Cytogenetique Hematologique. Cytogenetic abnormalities in adult acute lymphoblastic leukemia: correlations with hematologic findings outcome. A Collaborative Study of the Group Francais de Cytogenetique Hematologique. Blood 1996;87:3135–3142.Google Scholar
Chessels, JM, Swansbury, GJ, Reeves, B, et al. Cytogenetics and prognosis in childhood lymphoblastic leukaemia: results of MRC UKALL X. Medical Research Council Working Party in Childhood Leukaemia. Br J Haematol 1997;99:93–100.CrossRefGoogle ScholarPubMed
Secker-Walker, LM, Prentice, HG, Durrant, J, et al. Cytogenetics adds independent prognostic information in adults with acute lymphoblastic leukaemia on MRC trial UKALL XA. MRC Adult Leukaemia Working Party. Br J Haematol 1997;96:601–610.CrossRefGoogle ScholarPubMed
Wetzler, M, Dodge, RK, Mrozek, K, et al. Prospective karyotype analysis in adult acute lymphoblastic leukemia: the Cancer and Leukemia Group B experience. Blood 1999;93:3983–3993.Google ScholarPubMed
Mancini, M, Scappaticci, D, Cimino, G, et al. A comprehensive genetic classification of adult acute lymphoblastic leukemia (ALL): analysis of the GIMEMA 0496 protocol. Blood 2005;105:3434–3441.CrossRefGoogle ScholarPubMed
Moorman, AV, Harrison, CJ, Buck, GA, et al. Karyotype is an independent prognostic factor in adult acute lymphoblastic leukemia (ALL): analysis of cytogenetic data from patients treated on the Medical Research Council (MRC) UKALLXII/Eastern Cooperative Oncology Group (ECOG) 2993 trial. Blood 2007;109:3189–3197.CrossRefGoogle ScholarPubMed
Forestier, E, Johansson, B, Gustafsson, G, et al. Prognostic impact of karyotypic findings in childhood acute lymphoblastic leukaemia: a Nordic series comparing two treatment periods. The Nordic Society of Paediatric Haematology and Oncology (NOPHO) Leukaemia Cytogenetic Study Group. Br J Haematol 2000;110:147–153.CrossRefGoogle Scholar
Rowley, JD. The critical role of chromosome translocations in human leukemias. Annu Rev Genet 1998;32:495–519.CrossRefGoogle ScholarPubMed
Meyer, C, Schneider, B, Jakob, S, et al. The MLL recombinome of acute leukemias. Leukemia; 2006;20: 777–784.CrossRefGoogle ScholarPubMed
Behm, FG, Raimondi, SC, Frestedt, JL, et al. Rearrangement of the MLL gene confers a poor prognosis in childhood acute lymphoblastic leukemia, regardless of presenting age. Blood 1996;87:2870–2877.Google ScholarPubMed
Rubnitz, JE, Link, MP, Shuster, JJ, et al. Frequency and prognostic significance of HRX rearrangements in infant acute lymphoblastic leukemia: a Pediatric Oncology Group study. Blood 1994;84:570–573.Google ScholarPubMed
Thirman, MJ, Gill, HJ, Burnett, RC, et al. Rearrangement of the MLL gene in acute lymphoblastic and acute myeloid leukemias with 11q23 chromosomal translocations. N Engl J Med 1993;329:909–914.CrossRefGoogle ScholarPubMed
Downing, JR, Head, DR, Raimondi, SC, et al. The der(11)-encoded MLL/AF-4 fusion transcript is consistently detected in t(4;11)(q21;q23)-containing acute lymphoblastic leukemia. Blood 1994;83:330–335.Google Scholar
Yamamoto, K, Seto, M, Iida, S, et al. A reverse transcriptase-polymerase chain reaction detects heterogeneous chimeric mRNAs in leukemias with 11q23 abnormalities. Blood 1994;83:2912–2921.Google ScholarPubMed
Repp, R, Borkhardt, A, Haupt, E, et al. Detection of four different 11q23 chromosomal abnormalities by multiplex-PCR and fluorescence-based automatic DNA-fragment analysis. Leukemia 1995;9:210–215.Google ScholarPubMed
Biondi, A, Rambaldi, A, Rossi, V, et al. Detection of ALL-1/AF4 fusion transcript by reverse transcription-polymerase chain reaction for diagnosis and monitoring of acute leukemias with the t(4;11) translocation. Blood 1993;82:2943–2947.Google Scholar
Schichman, SA, Caligiuri, MA, Gu, Y, et al. ALL-1 partial duplication in acute leukemia. Proc Natl Acad Sci USA 1994;91:6236–6239.CrossRefGoogle ScholarPubMed
Schnittger, S, Kinkelin, U, Schoch, C, et al. Screening for MLL tandem duplication in 387 unselected patients with AML identify a prognostically unfavorable subset of AML. Leukemia 2000;14:796–804.CrossRefGoogle ScholarPubMed
Bernard, OA, Romana, SP, Schichman, SA, et al. Partial duplication of HRX in acute leukemia with trisomy 11. Leukemia 1995;9:1487–1490.Google ScholarPubMed
Andersen, MK, Christiansen, DH, Kirchhoff, M, et al. Duplication or amplification of chromosome band 11q23, including the unrearranged MLL gene, is a recurrent abnormality in therapy-related MDS and AML, and is closely related to mutation of the TP53 gene and to previous therapy with alkylating agents. Genes Chromosomes Cancer 2001;31:33–41.CrossRefGoogle ScholarPubMed
Poppe, B, Vandesompele, J, Schoch, C, et al. Expression analyses identify MLL as a prominent target of 11q23 amplification and support an etiologic role for MLL gain of function in myeloid malignancies. Blood 2004;103:229–235.CrossRefGoogle ScholarPubMed
Zatkova, A, Ullmann, R, Rouillard, JM, et al. Distinct sequences on 11q13.5 and 11q23-24 are frequently coamplified with MLL in complexly organized 11q amplicons in AML/MDS patients. Genes Chromosomes Cancer 2004;39:263–276.CrossRefGoogle ScholarPubMed
Whitman, SP, Strout, MP, Marcucci, G, et al. The partial nontandem duplication of the MLL (ALL1) gene is a novel rearrangement that generates three distinct fusion transcripts in B-cell acute lymphoblastic leukemia. Cancer Res 2001;61:59–63.Google ScholarPubMed
Dohner, K, Tobis, K, Ulrich, R, et al. Prognostic significance of partial tandem duplications of the MLL gene in adult patients 16 to 60 years old with acute myeloid leukemia and normal cytogenetics: a study of the Acute Myeloid Leukemia Study Group Ulm. J Clin Oncol 2002;20:3254–3261.CrossRefGoogle ScholarPubMed
Ziemin-van der Poel, S, McCabe, NR, Gill, HJ, et al. Identification of a gene, MLL, that spans the breakpoint in 11q23 translocations associated with human leukemias. Proc Natl Acad Sci USA 1991;88:10735–10739.CrossRefGoogle ScholarPubMed
Tkachuk, DC, Kohler, S, Cleary, ML. Involvement of a homolog of Drosophila trithorax by 11q23 chromosomal translocations in acute leukemias. Cell 1992;71:691–700.CrossRefGoogle ScholarPubMed
Gu, Y, Nakamura, T, Alder, H, et al. The t(4;11) chromosome translocation of human acute leukemias fuses the ALL-1 gene, related to Drosophila trithorax, to the AF-4 gene. Cell 1992;71:701–708.CrossRefGoogle Scholar
Ayton, PM, Cleary, ML. Molecular mechanisms of leukemogenesis mediated by MLL fusion proteins. Oncogene 2001;20:5695–5707.CrossRefGoogle ScholarPubMed
Mazo, AM, Huang, DH, Mozer, BA, et al. The trithorax gene, a trans-acting regulator of the bithorax complex in Drosophila, encodes a protein with zinc-binding domains. Proc Natl Acad Sci USA 1990;87:2112–2116.CrossRefGoogle ScholarPubMed
Reeves, R, Nissen, MS. The AT-DNA-binding domain of mammalian high mobility group I chromosomal proteins. A novel peptide motif for recognizing DNA structure. J Biol Chem 1990;265:8573–8582.Google Scholar
Ma, Q, Alder, H, Nelson, KK, et al. Analysis of the murine All-1 gene reveals conserved domains with human ALL-1 and identifies a motif shared with DNA methyltransferases. Proc Natl Acad Sci USA 1993;90:6350–6354.CrossRefGoogle ScholarPubMed
Milne, TA, Briggs, SD, Brock, HW, et al. MLL targets SET domain methyltransferase activity to Hox gene promoters. Mol Cell 2002;10: 1107–1117.CrossRefGoogle ScholarPubMed
Nakamura, T, Mori, T, Tada, S, et al. ALL-1 is a histone methyltransferase that assembles a supercomplex of proteins involved in transcriptional regulation. Mol Cell 2002;10:1119–1128.CrossRefGoogle ScholarPubMed
Yokoyama, A, Wang, Z, Wysocka, J, et al. Leukemia proto-oncoprotein MLL forms a SET1-like histone methyltransferase complex with menin to regulate Hox gene expression. Mol Cell Biol 2004;24:5639–5649.CrossRefGoogle ScholarPubMed
Yokoyama, A, Kitabayashi, I, Ayton, PM, et al. Leukemia proto-oncoprotein MLL is proteolytically processed into 2 fragments with opposite transcriptional properties. Blood 2002;100:3710–3718.CrossRefGoogle ScholarPubMed
Hsieh, JJ, Ernst, P, Erdjument-Bromage, H, et al. Proteolytic cleavage of MLL generates a complex of N- and C-terminal fragments that confers protein stability and subnuclear localization. Mol Cell Biol 2003;23:186–194.CrossRefGoogle ScholarPubMed
Moorman, AV, Raimondi, SC, Pui, CH, et al. No prognostic effect of additional chromosomal abnormalities in children with acute lymphoblastic leukemia and 11q23 abnormalities. Leukemia 2005;19:557–563.CrossRefGoogle ScholarPubMed
Pui, CH, Behm, FG, Downing, JR, et al. 11q23/MLL rearrangement confers a poor prognosis in infants with acute lymphoblastic leukemia. J Clin Oncol 1994;12:909–915.CrossRefGoogle ScholarPubMed
Cimino, G, Rapanotti, MC, Rivolta, A, et al. Prognostic relevance of ALL-1 gene rearrangement in infant acute leukemias. Leukemia 1995;9: 391–395.Google ScholarPubMed
Cimino, G, Lo-Coco, F, Biondi, A, et al. ALL-1 gene at chromosome 11q23 is consistently altered in acute leukemia of early infancy. Blood 1993;82:544–546.Google ScholarPubMed
Lavau, C, Szilvassy, SJ, Slany, R, et al. Immortalization and leukemic transformation of a myelomonocytic precursor by retrovirally transduced HRX-ENL. EMBO J 1997;16: 4226–4237.CrossRefGoogle ScholarPubMed
DiMartino, JF, Miller, T, Ayton, PM, et al. A carboxy-terminal domain of ELL is required and sufficient for immortalization of myeloid progenitors by MLL-ELL. Blood 2000;96: 3887–3893.Google ScholarPubMed
Lavau, C, Du, C, Thirman, M, et al. Chromatin-related properties of CBP fused to MLL generate a myelodysplastic-like syndrome that evolves into myeloid leukemia. EMBO J 2000;19:4655–4664.CrossRefGoogle ScholarPubMed
DiMartino, JF, Ayton, PM, Chen, EH, et al. The AF10 leucine zipper is required for leukemic transformation of myeloid progenitors by MLL-AF10. Blood 2002;99:3780–3785.CrossRefGoogle ScholarPubMed
So, CW, Lin, M, Ayton, PM, et al. Dimerization contributes to oncogenic activation of MLL chimeras in acute leukemias. Cancer Cell 2003;4:99–110.CrossRefGoogle ScholarPubMed
Dobson, CL, Warren, AJ, Pannell, R, et al. Tumorigenesis in mice with a fusion of the leukaemia oncogene MLL and the bacterial lacZ gene. EMBO J 2000;19:843–851.CrossRefGoogle ScholarPubMed
Martin, ME, Milne, TA, Bloyer, S, et al. Dimerization of MLL fusion proteins immortalizes hematopoietic cells. Cancer Cell 2003;4:197–207.CrossRefGoogle ScholarPubMed
Hsu, K, Look, AT. Turning on a dimer: new insights into MLL chimeras. Cancer Cell 2003;4:81–83.CrossRefGoogle ScholarPubMed
Yu, BD, Hess, JL, Horning, SE, et al. Altered Hox expression and segmental identity in MLL-mutant mice. Nature 1995;378:505–508.CrossRefGoogle ScholarPubMed
Yu, BD, Hanson, RD, Hess, JL, et al. MLL, a mammalian trithorax-group gene, functions as a transcriptional maintenance factor in morphogenesis. Proc Natl Acad Sci USA 1998;95:10632–10636.CrossRefGoogle ScholarPubMed
Hess, JL, Yu, BD, Li, B, et al. Defects in yolk sac hematopoiesis in MLL-null embryos. Blood 1997;90:1799–1806.Google ScholarPubMed
Corral, J, Lavenir, I, Impey, H, et al. An MLL-AF9 fusion gene made by homologous recombination causes acute leukemia in chimeric mice: a method to create fusion oncogenes. Cell 1996;85:853–861.CrossRefGoogle ScholarPubMed
Zeisig, BB, Garcia-Cuellar, MP, Winkler, TH, et al. The oncoprotein MLL–ENL disturbs hematopoietic lineage determination and transforms a biphenotypic lymphoid/myeloid cell. Oncogene 2003;22:1629–1637.CrossRefGoogle ScholarPubMed
Krivtsov, AV, Twomey, D, Feng, Z, et al. Transformation from committed progenitor to leukaemia stem cell initiated by MLL–AF9. Nature 2006;442:818–822.CrossRefGoogle ScholarPubMed
Wei, J, Wunderlich, M, Fox, C, et al. Microenvironment determines lineage fate in a human model of MLL-AF9 leukemia. Cancer Cell 2008;13:483–495.CrossRefGoogle Scholar
Forster, A, Pannell, R, Drynan, LF, et al. Engineering de novo reciprocal chromosomal translocations associated with Mll to replicate primary events of human cancer. Cancer Cell 2003;3:449–458.CrossRefGoogle ScholarPubMed
Drynan, LF, Pannell, R, Forster, A, et al. Mll fusions generated by Cre-loxP-mediated de novo translocations can induce lineage reassignment in tumorigenesis. EMBO J 2005;24:3136–3146.CrossRefGoogle ScholarPubMed
Metzler, M, Forster, A, Pannell, R, et al. A conditional model of MLL-AF4 B-cell tumourigenesis using invertor technology. Oncogene 2006;25:3093–3103.CrossRefGoogle ScholarPubMed
Coenen, EA, Raimondi, SC, Harbott, J, et al. Prognostic significance of additional cytogenetic aberrations in 733 de novo pediatric 11q23/MLL-rearranged AML patients: results of an international study. Blood 2011;117:7102–7111.CrossRefGoogle ScholarPubMed
Bardini, M, Spinelli, R, Bungaro, S, et al. DNA copy-number abnormalities do not occur in infant ALL with t(4;11)/MLL-AF4. Leukemia 2010;24:169–176.CrossRefGoogle Scholar
Rozovskaia, T, Feinstein, E, Mor, O, et al. Upregulation of Meis1 and HoxA9 in acute lymphocytic leukemias with the t(4;11) abnormality. Oncogene 2001;20:874–878.CrossRefGoogle Scholar
Yeoh, EJ, Ross, ME, Shurtleff, SA, et al. Classification, subtype discovery, and prediction of outcome in pediatric acute lymphoblastic leukemia by gene expression profiling. Cancer Cell 2002;1:133–143.CrossRefGoogle ScholarPubMed
Armstrong, SA, Staunton, JE, Silverman, LB, et al. MLL translocations specify a distinct gene expression profile that distinguishes a unique leukemia. Nat Genet 2002;30:41–47.CrossRefGoogle ScholarPubMed
Ross, ME, Zhou, X, Song, G, et al. Classification of pediatric acute lymphoblastic leukemia by gene expression profiling. Blood 2003;102:2951–2959.CrossRefGoogle ScholarPubMed
Ayton, PM, Cleary, ML. Transformation of myeloid progenitors by MLL oncoproteins is dependent on Hoxa7 and Hoxa9. Genes Dev 2003;17:2298–2307.CrossRefGoogle ScholarPubMed
Faber, J, Krivtsov, AV, Stubbs, MC, et al. HOXA9 is required for survival in human MLL-rearranged acute leukemias. Blood 2009;113:2375–2385.CrossRefGoogle ScholarPubMed
Orlovsky, K, Kalinkovich, A, Rozovskaia, T, et al. Down-regulation of homeobox genes MEIS1 and HOXA in MLL-rearranged acute leukemia impairs engraftment and reduces proliferation. Proc Natl Acad Sci USA, 2011;108:7956–7961.CrossRefGoogle ScholarPubMed
Armstrong, SA, Kung, AL, Mabon, ME, et al. Inhibition of FLT3 in MLL. Validation of a therapeutic target identified by gene expression based classification. Cancer Cell 2003;3:173–183.CrossRefGoogle ScholarPubMed
Stubbs, MC, Kim, YM, Krivtsov, AV, et al. MLL-AF9 and FLT3 cooperation in acute myelogenous leukemia: development of a model for rapid therapeutic assessment. Leukemia 2008;22:66–77.CrossRefGoogle ScholarPubMed
Schotte, D, Chau, JC, Sylvester, G, et al. Identification of new microRNA genes and aberrant microRNA profiles in childhood acute lymphoblastic leukemia. Leukemia 2009;23:313–322.CrossRefGoogle ScholarPubMed
Mi, S, Li, Z, Chen, P, et al. Aberrant overexpression and function of the miR-17-92 cluster in MLL-rearranged acute leukemia. Proc Natl Acad Sci USA 2010;107:3710–3715.CrossRefGoogle ScholarPubMed
Wong, P, Iwasaki, M, Somervaille, TC, et al. The miR-17-92 microRNA polycistron regulates MLL leukemia stem cell potential by modulating p21 expression. Cancer Res 2010;70:3833–3842.CrossRefGoogle ScholarPubMed
Schafer, E, Irizarry, R, Negi, S, et al. Promoter hypermethylation in MLL-r infant acute lymphoblastic leukemia: biology and therapeutic targeting. Blood 2010;115:4798–4809.CrossRefGoogle ScholarPubMed
Stumpel, DJ, Schneider, P, van Roon, EH, et al. Specific promoter methylation identifies different subgroups of MLL-rearranged infant acute lymphoblastic leukemia, influences clinical outcome, and provides therapeutic options. Blood 2009;114:5490–5498.CrossRefGoogle ScholarPubMed
Stumpel, DJ, Schotte, D, Lange-Turenhout, EA, et al. Hypermethylation of specific microRNA genes in MLL-rearranged infant acute lymphoblastic leukemia: major matters at a micro scale. Leukemia 2011;25:429–439.CrossRefGoogle Scholar
Krivtsov, AV, Feng, Z, Lemieux, ME, et al. H3K79 methylation profiles define murine and human MLL-AF4 leukemias. Cancer Cell 2008;14:355–368.CrossRefGoogle ScholarPubMed
Chang, MJ, Wu, H, Achille, NJ, et al. Histone H3 lysine 79 methyltransferase Dot1 is required for immortalization by MLL oncogenes. Cancer Res 2010;70:10234–10242.CrossRefGoogle ScholarPubMed
Nguyen, AT, Taranova, O, He, J, et al. DOT1L, the H3K79 methyltransferase, is required for MLL-AF9-mediated leukemogenesis. Blood 2011;117:6912–6922.CrossRef
Jo, SY, Granowicz, EM, Maillard, I, et al. Requirement for Dot1l in murine postnatal hematopoiesis and leukemogenesis by MLL translocation. Blood 2011;117:4759–4768.CrossRefGoogle ScholarPubMed
Klapproth, K, Wirth, T. Advances in the understanding of MYC-induced lymphomagenesis. Br J Haematol 2010;149:484–497.CrossRefGoogle ScholarPubMed
Dyer, MJ, Akasaka, T, Capasso, M, 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 2010;115:1490–1499.CrossRefGoogle ScholarPubMed
Dalla-Favera, R, Bregni, M, Erikson, J, et al. Human c-myc oncogene is located on the region of chromosome 8 that is translocated in Burkitt lymphoma cells. Proc Natl Acad Sci USA 1982;79:7824–7827.CrossRefGoogle ScholarPubMed
Taub, R, Kirsch, I, Morton, C, et al. Translocation of the c-myc gene into the immunoglobulin heavy chain locus in human Burkitt lymphoma and murine plasmacytoma cells. Proc Natl Acad Sci USA 1982;79:7837–7841.CrossRefGoogle ScholarPubMed
Adams, JM, Gerondakis, S, Webb, E, et al. Cellular myc oncogene is altered by chromosome translocation to an immunoglobulin locus in murine plasmacytomas and is rearranged similarly in human Burkitt lymphomas. Proc Natl Acad Sci USA 1983;80:1982–1986.CrossRefGoogle Scholar
Erikson, J, Nishikura, K, ar-Rushdi, A, et al. Translocation of an immunoglobulin kappa locus to a region 3′ of an unrearranged c-myc oncogene enhances c-myc transcription. Proc Natl Acad Sci USA 1983;80:7581–7585.CrossRefGoogle ScholarPubMed
Croce, CM, Thierfelder, W, Erikson, J, et al. Transcriptional activation of an unrearranged and untranslocated c-myc oncogene by translocation of a C lambda locus in Burkitt. Proc Natl Acad Sci USA 1983;80:6922–6926.CrossRefGoogle Scholar
Emanuel, BS, Selden, JR, Chaganti, RS, et al. The 2p breakpoint of a 2;8 translocation in Burkitt lymphoma interrupts the V kappa locus. Proc Natl Acad Sci USA 1984;81: 2444–2446.CrossRefGoogle ScholarPubMed
Hollis, GF, Mitchell, KF, Battey, J, et al. A variant translocation places the lambda immunoglobulin genes 3′ to the c-myc oncogene in Burkitt's lymphoma. Nature 1984;307:752–755.CrossRefGoogle ScholarPubMed
Rappold, GA, Hameister, H, Cremer, T, et al. c-myc and immunoglobulin kappa light chain constant genes are on the 8q+ chromosome of three Burkitt lymphoma lines with t(2;8) translocations. EMBO J 1984;3:2951–2955.Google Scholar
Taub, R, Kelly, K, Battey, J, et al. A novel alteration in the structure of an activated c-myc gene in a variant t(2;8) Burkitt lymphoma. Cell 1984;37:511–520.CrossRefGoogle Scholar
Amati, B, Dalton, S, Brooks, MW, et al. Transcriptional activation by the human c-Myc oncoprotein in yeast requires interaction with Max. Nature 1992;359:423–426.CrossRefGoogle ScholarPubMed
Kato, GJ, Lee, WM, Chen, LL, et al. Max: functional domains and interaction with c-Myc. Genes Dev 1992;6:81–92.CrossRefGoogle ScholarPubMed
Amati, B, Brooks, MW, Levy, N, et al. Oncogenic activity of the c-Myc protein requires dimerization with Max. Cell 1993;72:233–245.CrossRefGoogle ScholarPubMed
Blackwell, TK, Huang, J, Ma, A, et al. Binding of myc proteins to canonical and noncanonical DNA sequences. Mol Cell Biol 1993;13:5216–5224.CrossRefGoogle ScholarPubMed
Grandori, C, Mac, J, Siebelt, F, et al. Myc–Max heterodimers activate a DEAD box gene and interact with multiple E box-related sites in vivo. EMBO J 1996;15:4344–4357.Google Scholar
Ayer, DE, Kretzner, L, Eisenman, RN. Mad: a heterodimeric partner for Max that antagonizes Myc transcriptional activity. Cell 1993;72:211–222.CrossRefGoogle ScholarPubMed
Foley, KP, Eisenman, RN. Two MAD tails: what the recent knockouts of Mad1 and Mxi1 tell us about the MYC/MAX/MAD network. Biochim Biophys Acta 1999;1423:M37–M47.Google ScholarPubMed
Hurlin, PJ, Queva, C, Eisenman, RN. Mnt: a novel Max-interacting protein and Myc antagonist. Curr Top Microbiol Immunol 1997;224:115–121.Google ScholarPubMed
Coller, HA, Grandori, C, Tamayo, P, et al. Expression analysis with oligonucleotide microarrays reveals that MYC regulates genes involved in growth, cell cycle, signaling, and adhesion. Proc Natl Acad Sci USA 2000;97:3260–3265.CrossRefGoogle ScholarPubMed
Soucek, L, Evan, GI. The ups and downs of Myc biology. Curr Opin Genet Dev 2010;20:91–95.CrossRefGoogle ScholarPubMed
Chen, X, Xu, H, Yuan, P, et al. Integration of external signaling pathways with the core transcriptional network in embryonic stem cells. Cell 2008;133:1106–1117.CrossRefGoogle ScholarPubMed
Brenner, C, Deplus, R, Didelot, C, et al. Myc represses transcription through recruitment of DNA methyltransferase corepressor. EMBO J 2005;24:336–346.CrossRefGoogle ScholarPubMed
Mateyak, MK, Obaya, AJ, Sedivy, JM. c-Myc regulates cyclin D-Cdk4 and -Cdk6 activity but affects cell cycle progression at multiple independent points. Mol Cell Biol 1999;19:4672–4683.CrossRefGoogle ScholarPubMed
Muller, D, Bouchard, C, Rudolph, B, et al. Cdk2-dependent phosphorylation of p27 facilitates its Myc-induced release from cyclin E/cdk2 complexes. Oncogene 1997;15:2561–2576.CrossRefGoogle ScholarPubMed
Galaktionov, K, Chen, X, Beach, D. Cdc25 cell-cycle phosphatase as a target of c-myc. Nature 1996;382:511–517.CrossRefGoogle ScholarPubMed
Zindy, F, Eischen, CM, Randle, DH, et al. Myc signaling via the ARF tumor suppressor regulates p53-dependent apoptosis and immortalization. Genes Dev 1998;12:2424–2433.CrossRefGoogle ScholarPubMed
Wang, J, Xie, LY, Allan, S, et al. Myc activates telomerase. Genes Dev 1998;12:1769–1774.CrossRefGoogle ScholarPubMed
Wu, KJ, Grandori, C, Amacker, M, et al. Direct activation of TERT transcription by c-MYC. Nat Genet 1999;21:220–224.CrossRefGoogle ScholarPubMed
Bello-Fernandez, C, Packham, G, Cleveland, JL. The ornithine decarboxylase gene is a transcriptional target of c-Myc. Proc Natl Acad Sci USA 1993;90:7804–7808.CrossRefGoogle ScholarPubMed
Mai, S, Jalava, A. c-Myc binds to 5′ flanking sequence motifs of the dihydrofolate reductase gene in cellular extracts: role in proliferation. Nucleic Acids Res 1994;22:2264–2273.CrossRefGoogle ScholarPubMed
Miltenberger, RJ, Sukow, KA, Farnham, PJ. An E-box-mediated increase in cad transcription at the G1/S-phase boundary is suppressed by inhibitory c-Myc mutants. Mol Cell Biol 1995;15:2527–2535.CrossRefGoogle ScholarPubMed
Boyd, KE, Farnham, PJ. Myc versus USF: discrimination at the cad gene is determined by core promoter elements. Mol Cell Biol 1997;17:2529–2537.CrossRefGoogle ScholarPubMed
Pusch, O, Soucek, T, Hengstschlager-Ottnad, E, et al. Cellular targets for activation by c-Myc include the DNA metabolism enzyme thymidine kinase. DNA Cell Biol 1997;16:737–747.Google ScholarPubMed
Boyd, KE, Wells, J, Gutman, J, et al. c-Myc target gene specificity is determined by a post-DNA binding mechanism. Proc Natl Acad Sci USA 1998;95:13887–13892.CrossRefGoogle Scholar
Bush, A, Mateyak, M, Dugan, K, et al. c-myc null cells misregulate cad and gadd45 but not other proposed c-Myc targets. Genes Dev 1998;12:3797–3802.CrossRefGoogle Scholar
Rosenwald, IB, Rhoads, DB, Callanan, LD, et al. Increased expression of eukaryotic translation initiation factors eIF-4E and eIF-2 alpha in response to growth induction by c-myc. Proc Natl Acad Sci USA 1993;90:6175–6178.CrossRefGoogle ScholarPubMed
Schuldiner, O, Eden, A, Ben-Yosef, T, et al. ECA39, a conserved gene regulated by c-Myc in mice, is involved in G1/S cell cycle regulation in yeast. Proc Natl Acad Sci USA 1996;93:7143–7148.CrossRefGoogle Scholar
Jones, RM, Branda, J, Johnston, KA, et al. An essential E box in the promoter of the gene encoding the mRNA cap-binding protein (eukaryotic initiation factor 4E) is a target for activation by c-myc. Mol Cell Biol 1996;16:4754–4764.CrossRefGoogle Scholar
He, L, Thomson, JM, Hemann, MT, et al. A microRNA polycistron as a potential human oncogene. Nature 2005;435:828–833.CrossRefGoogle ScholarPubMed
O'Donnell, KA, Wentzel, EA, Zeller, KI, et al. c-Myc-regulated microRNAs modulate E2F1 expression. Nature 2005;435:839–843.CrossRefGoogle ScholarPubMed
Mu, P, Han, YC, Betel, D, et al. Genetic dissection of the miR-17~92 cluster of microRNAs in Myc-induced B-cell lymphomas. Genes Dev 2009;23:2806–2811.CrossRefGoogle ScholarPubMed
Chang, TC, Zeitels, LR, Hwang, HW, et al. Lin-28B transactivation is necessary for Myc-mediated let-7 repression and proliferation. Proc Natl Acad Sci USA, 2009;106:3384–3389.CrossRefGoogle ScholarPubMed
Russell, LJ, Capasso, M, Vater, I, et al. Deregulated expression of cytokine receptor gene, CRLF2, is involved in lymphoid transformation in B-cell precursor acute lymphoblastic leukemia. Blood 2009;114:2688–2698.CrossRefGoogle ScholarPubMed
Mullighan, CG, Phillips, LA, Su, X, et al. Genomic analysis of the clonal origins of relapsed acute lymphoblastic leukemia. Science 2008;322:1377–1380.CrossRefGoogle ScholarPubMed
Mullighan, CG, Collins-Underwood, JR, Phillips, LA, et al. Rearrangement of CRLF2 in B-progenitor- and Down syndrome-associated acute lymphoblastic leukemia. Nat Genet 2009;41:1243–1246.CrossRefGoogle ScholarPubMed
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 2010;115:5312–5321.CrossRefGoogle Scholar
Hertzberg, L, Vendramini, E, Ganmore, I, 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 iBFM Study Group. Blood 2010;115:1006–1017.CrossRefGoogle Scholar
Yoda, A, Yoda, Y, Chiaretti, S, et al. Functional screening identifies CRLF2 in precursor B-cell acute lymphoblastic leukemia. Proc Natl Acad Sci USA 2010;107:252–257.CrossRefGoogle ScholarPubMed
Flex, E, Petrangeli, V, Stella, L, et al. Somatically acquired JAK1 mutations in adult acute lymphoblastic leukemia. J Exp Med 2008;205:751–758.CrossRefGoogle ScholarPubMed
Levine, RL, Gilliland, DG. Myeloproliferative disorders. Blood 2008;112:2190–2198.CrossRefGoogle ScholarPubMed
Bercovich, D, Ganmore, I, Scott, LM, et al. Mutations of JAK2 in acute lymphoblastic leukaemias associated with Down's syndrome. Lancet 2008;372:1484–1492.CrossRefGoogle ScholarPubMed
Mullighan, CG, Zhang, J, Harvey, RC, et al. JAK mutations in high-risk childhood acute lymphoblastic leukemia. Proc Natl Acad Sci USA 2009;106:9414–9418.CrossRefGoogle ScholarPubMed
Kearney, L, Gonzalez De Castro, D, Yeung, J, et al. A specific JAK2 mutation (JAK2R683) and multiple gene deletions in Down syndrome acute lymphoblastic leukaemia. Blood 2008;113:646–648.CrossRefGoogle Scholar
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 2010;115:5393–5397.CrossRefGoogle ScholarPubMed
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 2011;117:2129–2136.CrossRefGoogle ScholarPubMed
Russell, LJ, De Castro, DG, Griffiths, M, et al. A novel translocation, t(14;19)(q32;p13), involving IGH@ and the cytokine receptor for erythropoietin. Leukemia 2009;23:614–617.CrossRefGoogle Scholar
Akasaka, T, Balasas, T, Russell, LJ, et al. Five members of the CEBP transcription factor family are targeted by recurrent IGH translocations in B-cell precursor acute lymphoblastic leukemia (BCP-ALL). Blood 2007;109:3451–3461.CrossRefGoogle Scholar
Kawamata, N, Sakajiri, S, Sugimoto, KJ, et al. A novel chromosomal translocation t(1;14)(q25;q32) in pre-B acute lymphoblastic leukemia involves the LIM homeodomain protein gene, Lhx4. Oncogene 2002;21:4983–4991.CrossRefGoogle Scholar
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 1989;73:2081–2085.Google Scholar
Meeker, TC, Hardy, D, Willman, C, et al. Activation of the interleukin-3 gene by chromosome translocation in acute lymphocytic leukemia with eosinophilia. Blood 1990;76:285–289.Google ScholarPubMed
Harewood, L, Robinson, H, Harris, R, et al. Amplification of AML1 on a duplicated chromosome 21 in acute lymphoblastic leukemia: a study of 20 cases. Leukemia 2003;17:547–553.CrossRefGoogle ScholarPubMed
Soulier, J, Trakhtenbrot, L, Najfeld, V, et al. Amplification of band q22 of chromosome 21, including AML1, in older children with acute lymphoblastic leukemia: an emerging molecular cytogenetic subgroup. Leukemia 2003;17:1679–1682.CrossRefGoogle ScholarPubMed
Robinson, HM, Harrison, CJ, Moorman, AV, et al. Intrachromosomal amplification of chromosome 21 (iAMP21) may arise from a breakage-fusion-bridge cycle. Genes Chromosomes Cancer 2007;46:318–326.CrossRefGoogle ScholarPubMed
Sinclair, PB, Parker, H, An, Q, et al. Analysis of a breakpoint cluster reveals insight into the mechanism of intrachromosomal amplification in a lymphoid malignancy. Hum Mol Genet 2011;20:2591–2602.CrossRef
Strefford, JC, van Delft, FW, Robinson, HM, et al. Complex genomic alterations and gene expression in acute lymphoblastic leukemia with intrachromosomal amplification of chromosome 21. Proc Natl Acad Sci USA, 2006;103:8167–8172.CrossRefGoogle ScholarPubMed
Moorman, AV, Richards, SM, Robinson, HM, et al. Prognosis of children with acute lymphoblastic leukemia (ALL) and intrachromosomal amplification of chromosome 21 (iAMP21). Blood 2007;109:2327–2330.CrossRefGoogle Scholar
Kuppers, R, Dalla-Favera, R. Mechanisms of chromosomal translocations in B cell lymphomas. Oncogene 2001;20:5580–5594.CrossRefGoogle ScholarPubMed
Nickoloff, JA, De Haro, LP, Wray, J, et al. Mechanisms of leukemia translocations. Curr Opin Hematol 2008;15:338–345.CrossRefGoogle ScholarPubMed
Cauwelier, B, Dastugue, N, Cools, J, et al. Molecular cytogenetic study of 126 unselected T-ALL cases reveals high incidence of TCRbeta locus rearrangements and putative new T-cell oncogenes. Leukemia 2006;20:1238–1244.CrossRefGoogle ScholarPubMed
McKeithan, TW, Shima, EA, Le Beau, MM, et al. Molecular cloning of the breakpoint junction of a human chromosomal 8;14 translocation involving the T-cell receptor alpha-chain gene and sequences on the 3′ side of MYC. Proc Natl Acad Sci USA 1986;83:6636–6640.CrossRefGoogle ScholarPubMed
Finger, LR, Harvey, RC, Moore, RC, et al. A common mechanism of chromosomal translocation in T- and B-cell neoplasia. Science 1986;234:982–985.CrossRefGoogle Scholar
Shima, EA, Le Beau, MM, McKeithan, TW, et al. Gene encoding the alpha chain of the T-cell receptor is moved immediately downstream of c-myc in a chromosomal 8;14 translocation in a cell line from a human T-cell leukemia. Proc Natl Acad Sci USA 1986;83:3439–3443.CrossRefGoogle Scholar
Chen, Q, Cheng, JT, Tasi, LH, et al. The tal gene undergoes chromosome translocation in T cell leukemia and potentially encodes a helix-loop-helix protein. EMBO J 1990;9:415–424.Google Scholar
Xia, Y, Brown, L, Yang, CY, et al. TAL2, a helix-loop-helix gene activated by the (7;9)(q34;q32) translocation in human T-cell leukemia. Proc Natl Acad Sci USA 1991;88:11416–11420.CrossRefGoogle ScholarPubMed
Mellentin, JD, Smith, SD, Cleary, ML. lyl-1, a novel gene altered by chromosomal translocation in T cell leukemia, codes for a protein with a helix-loop-helix DNA binding motif. Cell 1989;58:77–83.CrossRefGoogle Scholar
Wang, J, Jani-Sait, SN, Escalon, EA, et al. The t(14;21)(q11.2;q22) chromosomal translocation associated with T-cell acute lymphoblastic leukemia activates the BHLHB1 gene. Proc Natl Acad Sci USA 2000;97:3497–3502.CrossRefGoogle Scholar
McGuire, EA, Hockett, RD, Pollock, KM, et al. The t(11;14)(p15;q11) in a T-cell acute lymphoblastic leukemia cell line activates multiple transcripts, including Ttg-1, a gene encoding a potential zinc finger protein. Mol Cell Biol 1989;9:2124–2132.CrossRefGoogle Scholar
Boehm, T, Foroni, L, Kaneko, Y, et al. The rhombotin family of cysteine-rich LIM-domain oncogenes: distinct members are involved in T-cell translocations to human chromosomes 11p15 and 11p13. Proc Natl Acad Sci USA 1991;88:4367–4371.CrossRefGoogle ScholarPubMed
Royer-Pokora, B, Loos, U, Ludwig, WD. TTG-2, a new gene encoding a cysteine-rich protein with the LIM motif, is overexpressed in acute T-cell leukaemia with the t(11;14)(p13;q11). Oncogene 1991;6:1887–1893.Google Scholar
Hatano, M, Roberts, CW, Minden, M, et al. Deregulation of a homeobox gene, HOX11, by the t(10;14) in T cell leukemia. Science 1991;253:79–82.CrossRefGoogle Scholar
Kennedy, MA, Gonzalez-Sarmiento, R, Kees, UR, et al. HOX11, a homeobox-containing T-cell oncogene on human chromosome 10q24. Proc Natl Acad Sci USA, 1991;88:8900–8904.CrossRefGoogle ScholarPubMed
Bernard, OA, Busson-LeConiat, M, Ballerini, P, et al. A new recurrent and specific cryptic translocation, t(5;14)(q35;q32), is associated with expression of the Hox11L2 gene in T acute lymphoblastic leukemia. Leukemia 2001;15:1495–1504.CrossRefGoogle Scholar
Begley, CG, Aplan, PD, Davey, MP, et al. Chromosomal translocation in a human leukemic stem-cell line disrupts the T-cell antigen receptor delta-chain diversity region and results in a previously unreported fusion transcript. Proc Natl Acad Sci USA 1989;86:2031–2035.CrossRefGoogle Scholar
Brown, L, Cheng, JT, Chen, Q, et al. Site-specific recombination of the tal-1 gene is a common occurrence in human T cell leukemia. EMBO J 1990;9:3343–3351.Google ScholarPubMed
Aplan, PD, Lombardi, DP, Kirsch, IR. Structural characterization of SIL, a gene frequently disrupted in T-cell acute lymphoblastic leukemia. Mol Cell Biol 1991;11:5462–5469.CrossRefGoogle ScholarPubMed
Aplan, PD, Lombardi, DP, Reaman, GH, et al. Involvement of the putative hematopoietic transcription factor SCL in T-cell acute lymphoblastic leukemia. Blood 1992;79:1327–1333.Google ScholarPubMed
Bernard, O, Lecointe, N, Jonveaux, P, et al. Two site-specific deletions and t(1;14) translocation restricted to human T-cell acute leukemias disrupt the 5′ part of the tal-1 gene. Oncogene 1991;6:1477–1488.Google Scholar
Breit, TM, Mol, EJ, Wolvers-Tettero, IL, et al. Site-specific deletions involving the tal-1 and sil genes are restricted to cells of the T cell receptor alpha/beta lineage: T cell receptor delta gene deletion mechanism affects multiple genes. J Exp Med 1993;177:965–977.CrossRefGoogle Scholar
Baer, R. TAL1, TAL2 and LYL1: a family of basic helix-loop-helix proteins implicated in T cell acute leukaemia. Semin Cancer Biol 1993;4:341–347.Google ScholarPubMed
Bash, RO, Hall, S, Timmons, CF, et al. Does activation of the TAL1 gene occur in a majority of patients with T-cell acute lymphoblastic leukemia? A pediatric oncology group study. Blood 1995;86:666–676.Google Scholar
Aplan, PD, Lombardi, DP, Ginsberg, AM, et al. Disruption of the human SCL locus by “illegitimate” V-(D)-J recombinase activity. Science 1990;250:1426–1429.CrossRefGoogle ScholarPubMed
Ferrando, AA, Neuberg, DS, Staunton, J, et al. Gene expression signatures define novel oncogenic pathways in T cell acute lymphoblastic leukemia. Cancer Cell 2002;1:75–87.CrossRefGoogle ScholarPubMed
Hsu, HL, Cheng, JT, Chen, Q, et al. Enhancer-binding activity of the tal-1 oncoprotein in association with the E47/E12 helix-loop-helix proteins. Mol Cell Biol 1991;11:3037–3042.CrossRefGoogle ScholarPubMed
Miyamoto, A, Cui, X, Naumovski, L, et al. Helix-loop-helix proteins LYL1 and E2a form heterodimeric complexes with distinctive DNA-binding properties in hematolymphoid cells. Mol Cell Biol 1996;16: 2394–2401.CrossRefGoogle ScholarPubMed
Shivdasani, RA, Mayer, EL, Orkin, SH. Absence of blood formation in mice lacking the T-cell leukaemia oncoprotein tal-1/SCL. Nature 1995;373:432–434.CrossRefGoogle ScholarPubMed
Robb, L, Lyons, I, Li, R, et al. Absence of yolk sac hematopoiesis from mice with a targeted disruption of the scl gene. Proc Natl Acad Sci USA 1995;92:7075–7079.CrossRefGoogle ScholarPubMed
Begley, CG, Aplan, PD, Denning, SM, et al. The gene SCL is expressed during early hematopoiesis and encodes a differentiation-related DNA-binding motif. Proc Natl Acad Sci USA 1989;86:10128–10132.CrossRefGoogle ScholarPubMed
Porcher, C, Swat, W, Rockwell, K, et al. The T cell leukemia oncoprotein SCL/tal-1 is essential for development of all hematopoietic lineages. Cell 1996;86:47–57.CrossRefGoogle Scholar
Hsu, HL, Wadman, I, Tsan, JT, et al. Positive and negative transcriptional control by the TAL1 helix-loop-helix protein. Proc Natl Acad Sci USA 1994;91:5947–5951.CrossRefGoogle ScholarPubMed
Park, ST, Sun, XH. The Tal1 oncoprotein inhibits E47-mediated transcription. Mechanism of inhibition. J Biol Chem 1998;273:7030–7037.CrossRefGoogle ScholarPubMed
Yan, W, Young, AZ, Soares, VC, et al. High incidence of T-cell tumors in E2A-null mice and E2A/Id1 double-knockout mice. Mol Cell Biol 1997;17:7317–7327.CrossRefGoogle ScholarPubMed
Begley, CG, Green, AR. The SCL gene: from case report to critical hematopoietic regulator. Blood 1999;93:2760–2770.Google ScholarPubMed
Van Vlierberghe, P, van Grotel, M, Beverloo, HB, et al. The cryptic chromosomal deletion del(11)(p12p13) as a new activation mechanism of LMO2 in pediatric T-cell acute lymphoblastic leukemia. Blood 2006;108:3520–3529.CrossRefGoogle ScholarPubMed
Greenberg, JM, Boehm, T, Sofroniew, MV, et al. Segmental and developmental regulation of a presumptive T-cell oncogene in the central nervous system. Nature 1990;344:158–160.CrossRefGoogle ScholarPubMed
Valge-Archer, VE, Osada, H, Warren, AJ, et al. The LIM protein RBTN2 and the basic helix-loop-helix protein TAL1 are present in a complex in erythroid cells. Proc Natl Acad Sci USA 1994;91:8617–8621.CrossRefGoogle Scholar
Wadman, I, Li, J, Bash, RO, et al. Specific in vivo association between the bHLH and LIM proteins implicated in human T cell leukemia. EMBO J 1994;13:4831–4839.Google ScholarPubMed
Larson, RC, Lavenir, I, Larson, TA, et al. Protein dimerization between Lmo2 (Rbtn2) and Tal1 alters thymocyte development and potentiates T cell tumorigenesis in transgenic mice. EMBO J 1996;15:1021–1027.Google ScholarPubMed
McGuire, EA, Rintoul, CE, Sclar, GM, et al. Thymic overexpression of Ttg-1 in transgenic mice results in T-cell acute lymphoblastic leukemia/lymphoma. Mol Cell Biol 1992;12:4186–4196.CrossRefGoogle ScholarPubMed
Larson, RC, Fisch, P, Larson, TA, et al. T cell tumours of disparate phenotype in mice transgenic for Rbtn-2. Oncogene 1994;9:3675–3681.Google ScholarPubMed
Larson, RC, Osada, H, Larson, TA, et al. The oncogenic LIM protein Rbtn2 causes thymic developmental aberrations that precede malignancy in transgenic mice. Oncogene 1995;11:853–862.Google ScholarPubMed
Neale, GA, Rehg, JE, Goorha, RM. Ectopic expression of rhombotin-2 causes selective expansion of CD4−CD8− lymphocytes in the thymus and T-cell tumors in transgenic mice. Blood 1995;86:3060–3071.Google ScholarPubMed
Aplan, PD, Jones, CA, Chervinsky, DS, et al. An scl gene product lacking the transactivation domain induces bony abnormalities and cooperates with LMO1 to generate T-cell malignancies in transgenic mice. EMBO J 1997;16:2408–2419.CrossRefGoogle ScholarPubMed
Chervinsky, DS, Zhao, XF, Lam, DH, et al. Disordered T-cell development and T-cell malignancies in SCL LMO1 double-transgenic mice: parallels with E2A-deficient mice. Mol Cell Biol 1999;19:5025–5035.CrossRefGoogle ScholarPubMed
McCormack, MP, Young, LF, Vasudevan, S, et al. The Lmo2 oncogene initiates leukemia in mice by inducing thymocyte self-renewal. Science 2010;327:879–883.CrossRefGoogle ScholarPubMed
Argiropoulos, B, Humphries, RK. Hox genes in hematopoiesis and leukemogenesis. Oncogene 2007;26:6766–6776.CrossRefGoogle ScholarPubMed
Dear, TN, Sanchez-Garcia, I, Rabbitts, TH. The HOX11 gene encodes a DNA-binding nuclear transcription factor belonging to a distinct family of homeobox genes. Proc Natl Acad Sci USA 1993;90:4431–4435.CrossRefGoogle ScholarPubMed
Allen, JD, Lints, T, Jenkins, NA, et al. Novel murine homeo box gene on chromosome 1 expressed in specific hematopoietic lineages and during embryogenesis. Genes Dev 1991;5:509–520.CrossRefGoogle ScholarPubMed
McGinnis, W, Krumlauf, R. Homeobox genes and axial patterning. Cell 1992;68:283–302.CrossRefGoogle ScholarPubMed
Lu, M, Gong, ZY, Shen, WF, et al. The tcl-3 proto-oncogene altered by chromosomal translocation in T-cell leukemia codes for a homeobox protein. EMBO J 1991;10:2905–2910.Google ScholarPubMed
Dube, ID, Kamel-Reid, S, Yuan, CC, et al. A novel human homeobox gene lies at the chromosome 10 breakpoint in lymphoid neoplasias with chromosomal translocation t(10;14). Blood 1991;78:2996–3003.Google Scholar
Salvati, PD, Ranford, PR, Ford, J, et al. HOX11 expression in pediatric acute lymphoblastic leukemia is associated with T-cell phenotype. Oncogene 1995;11:1333–1338.Google ScholarPubMed
Ferrando, AA, Herblot, S, Palomero, T, et al. Biallelic transcriptional activation of oncogenic transcription factors in T-cell acute lymphoblastic leukemia. Blood 2004;103:1909–1911.CrossRefGoogle ScholarPubMed
Watt, PM, Kumar, R, Kees, UR. Promoter demethylation accompanies reactivation of the HOX11 proto-oncogene in leukemia. Genes Chromosomes Cancer 2000;29:371–377.3.0.CO;2-Y>CrossRefGoogle ScholarPubMed
Roberts, CW, Shutter, JR, Korsmeyer, SJ. Hox11 controls the genesis of the spleen. Nature 1994;368:747–749.CrossRefGoogle ScholarPubMed
Dear, TN, Colledge, WH, Carlton, MB, et al. The Hox11 gene is essential for cell survival during spleen development. Development 1995;121:2909–2915.Google ScholarPubMed
Hawley, RG, Hawley, TS, Cantor, AB. TLX1 (HOX11) immortalization of embryonic stem cell-derived and primary murine hematopoietic progenitors. Current Protocols in Stem Cell Biology, Ch. 1, Unit 1F 7. New York: Wiley, 2008.Google Scholar
Keller, G, Wall, C, Fong, AZ, et al. Overexpression of HOX11 leads to the immortalization of embryonic precursors with both primitive and definitive hematopoietic potential. Blood 1998;92:877–887.Google ScholarPubMed
Hawley, RG, Fong, AZ, Reis, MD, et al. Transforming function of the HOX11/TCL3 homeobox gene. Cancer Res 1997;57:337–345.Google ScholarPubMed
Cave, 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 2004;103:442–450.CrossRefGoogle Scholar
Ballerini, P, Blaise, A, Busson-Le Coniat, M, et al. HOX11L2 expression defines a clinical subtype of pediatric T-ALL associated with poor prognosis. Blood 2002;100:991–997.CrossRefGoogle ScholarPubMed
Berger, R, Dastugue, N, Busson, M, et al. t(5;14)/HOX11L2-positive T-cell acute lymphoblastic leukemia. A collaborative study of the Groupe Francais de Cytogenetique Hematologique (GFCH). Leukemia 2003;17:1851–1857.CrossRefGoogle Scholar
de Keersmaecker, K, Real, PJ, Gatta, GD, et al. The TLX1 oncogene drives aneuploidy in T cell transformation. Nat Med 2010;16:1321–1327.CrossRefGoogle ScholarPubMed
Nagel, S, Scherr, M, Kel, A, et al. Activation of TLX3 and NKX2–5 in t(5;14)(q35;q32) T-cell acute lymphoblastic leukemia by remote 3′-BCL11B enhancers and coregulation by PU.1 and HMGA1. Cancer Res 2007;67:1461–1471.CrossRefGoogle Scholar
Nagel, S, Kaufmann, M, Drexler, HG, et al. The cardiac homeobox gene NKX2-5 is deregulated by juxtaposition with BCL11B in pediatric T-ALL cell lines via a novel t(5;14)(q35.1;q32.2). Cancer Res 2003;63:5329–5334.Google Scholar
Su, XY, Busson, M, Della Valle, V, et al. Various types of rearrangements target TLX3 locus in T-cell acute lymphoblastic leukemia. Genes Chromosomes Cancer 2004;41:243–249.CrossRefGoogle ScholarPubMed
Van Vlierberghe, P, Homminga, I, Zuurbier, L, et al. Cooperative genetic defects in TLX3 rearranged pediatric T-ALL. Leukemia 2008;22:762–770.CrossRefGoogle ScholarPubMed
Hayette, S, Tigaud, I, Maguer-Satta, V, et al. Recurrent involvement of the MLL gene in adult T-lineage acute lymphoblastic leukemia. Blood 2002;99:4647–4649.CrossRefGoogle ScholarPubMed
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 1999;17:191–196.CrossRefGoogle Scholar
Ferrando, AA, Armstrong, SA, Neuberg, DS, et al. Gene expression signatures in MLL-rearranged T-lineage and B-precursor acute leukemias: dominance of HOX dysregulation. Blood 2003;102:262–268.CrossRefGoogle ScholarPubMed
Dreyling, MH, Martinez-Climent, JA, Zheng, M, et al. The t(10;11)(p13;q14) in the U937 cell line results in the fusion of the AF10 gene and CALM, encoding a new member of the AP-3 clathrin assembly protein family. Proc Natl Acad Sci USA 1996;93:4804–4809.CrossRefGoogle Scholar
Asnafi, V, Radford-Weiss, I, Dastugue, N, et al. CALM-AF10 is a common fusion transcript in T-ALL and is specific to the TCRgammadelta lineage. Blood 2003;102:1000–1006.CrossRefGoogle ScholarPubMed
Bohlander, SK, Muschinsky, V, Schrader, K, et al. Molecular analysis of the CALM/AF10 fusion: identical rearrangements in acute myeloid leukemia, acute lymphoblastic leukemia and malignant lymphoma patients. Leukemia 2000;14:93–99.CrossRefGoogle ScholarPubMed
Dreyling, MH, Schrader, K, Fonatsch, C, et al. MLL and CALM are fused to AF10 in morphologically distinct subsets of acute leukemia with translocation t(10;11): both rearrangements are associated with a poor prognosis. Blood 1998;91:4662–4667.Google Scholar
Deshpande, AJ, Cusan, M, Rawat, VP, et al. Acute myeloid leukemia is propagated by a leukemic stem cell with lymphoid characteristics in a mouse model of CALM/AF10-positive leukemia. Cancer Cell 2006;10:363–374.CrossRefGoogle Scholar
Caudell, D, Zhang, Z, Chung, YJ, et al. Expression of a CALM-AF10 fusion gene leads to Hoxa cluster overexpression and acute leukemia in transgenic mice. Cancer Res 2007;67:8022–8031.CrossRefGoogle ScholarPubMed
Coustan-Smith, E, Mullighan, CG, Onciu, M, et al. Early T-cell precursor leukaemia: a subtype of very high-risk acute lymphoblastic leukaemia. Lancet Oncol 2009;10:147–156.CrossRefGoogle ScholarPubMed
Homminga, I, Pieters, R, Langerak, AW, et al. Integrated transcript and genome analyses reveal NKX2-1 and MEF2C as potential oncogenes in T cell acute lymphoblastic leukemia. Cancer Cell 2011;19:484–497.CrossRefGoogle ScholarPubMed
Wiemels, JL, Cazzaniga, G, Daniotti, M, et al. Prenatal origin of acute lymphoblastic leukaemia in children. Lancet 1999;354:1499–1503.CrossRefGoogle ScholarPubMed
Wiemels, JL, Ford, AM, van Wering, ER, et al. Protracted and variable latency of acute lymphoblastic leukemia after TEL-AML1 gene fusion in utero. Blood 1999;94:1057–1062.Google ScholarPubMed
Mullighan, CG, Downing, JR. Genome-wide profiling of genetic alterations in acute lymphoblastic leukemia: recent insights and future directions. Leukemia 2009;23:1209–1218.CrossRefGoogle ScholarPubMed
Collins-Underwood, JR, Mullighan, CG. Genomic profiling of high-risk acute lymphoblastic leukemia. Leukemia 2010;24:1676–1685.CrossRefGoogle ScholarPubMed
Beroukhim, R, Mermel, CH, Porter, D, et al. The landscape of somatic copy-number alteration across human cancers. Nature 2010;463: 899–905.CrossRefGoogle ScholarPubMed
Kuiper, RP, Schoenmakers, EF, van Reijmersdal, SV, 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;21:1258–1266.CrossRefGoogle ScholarPubMed
Kawamata, N, Ogawa, S, Zimmermann, M, et al. Molecular allelokaryotyping of pediatric acute lymphoblastic leukemias by high-resolution single nucleotide polymorphism oligonucleotide genomic microarray. Blood 2008;111:776–784.CrossRefGoogle ScholarPubMed
Mullighan, CG, Downing, JR. Global genomic characterization of acute lymphoblastic leukemia. Semin Hematol 2009;46:3–15.CrossRefGoogle ScholarPubMed
Mullighan, CG, Su, X, Zhang, J, et al. Deletion of IKZF1 and prognosis in acute lymphoblastic leukemia. N Engl J Med 2009;360:470–480.CrossRefGoogle ScholarPubMed
Kuiper, RP, Waanders, E, van der Velden, VH, et al. IKZF1 deletions predict relapse in uniformly treated pediatric precursor B-ALL. Leukemia 2010;24:1258–1264.CrossRefGoogle ScholarPubMed
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 2009;10:125–134.CrossRefGoogle ScholarPubMed
Gaikwad, A, Rye, CL, Devidas, M, et al. Prevalence and clinical correlates of JAK2 mutations in Down syndrome acute lymphoblastic leukaemia. Br J Haematol 2009;144:930–932.CrossRefGoogle ScholarPubMed
Mullighan, CG, Zhang, J, Kasper, LH, et al. CREBBP mutations in relapsed acute lymphoblastic leukaemia. Nature 2011;471:235–239.CrossRefGoogle ScholarPubMed
Pasqualucci, L, Dominguez-Sola, D, Chiarenza, A, et al. Inactivating mutations of acetyltransferase genes in B-cell lymphoma. Nature 2011;471:189–195.CrossRef
Roberts, KG, Mullighan, CG. How new advances in genetic analysis are influencing the understanding and treatment of childhood acute leukemia. Curr Opin Pediatr 2011;23:34–40.CrossRefGoogle ScholarPubMed
Collins-Underwood, JR, Mullighan, CG. Genetic alterations targeting lymphoid development in acute lymphoblastic leukemia. Curr Top Dev Biol 2011;94:171–196.CrossRefGoogle ScholarPubMed
Nebral, K, Konig, M, Harder, L, et al. Identification of PML as novel PAX5 fusion partner in childhood acute lymphoblastic leukaemia. Br J Haematol 2007;139:269–274.CrossRefGoogle ScholarPubMed
Bousquet, M, Broccardo, C, Quelen, C, et al. A novel PAX5–ELN fusion protein identified in B-cell acute lymphoblastic leukemia acts as a dominant negative on wild-type PAX5. Blood 2007;109:3417–3423.CrossRefGoogle ScholarPubMed
An, Q, Wright, SL, Konn, ZJ, et al. Variable breakpoints target PAX5 in patients with dicentric chromosomes: a model for the basis of unbalanced translocations in cancer. Proc Natl Acad Sci USA 2008;105:17050–17054.CrossRefGoogle Scholar
Nebral, K, Denk, D, Attarbaschi, A, et al. Incidence and diversity of PAX5 fusion genes in childhood acute lymphoblastic leukemia. Leukemia 2009;23:134–143.CrossRefGoogle ScholarPubMed
Urbanek, P, Wang, ZQ, Fetka, I, et al. Complete block of early B cell differentiation and altered patterning of the posterior midbrain in mice lacking Pax5/BSAP. Cell 1994;79:901–912.CrossRefGoogle Scholar
Dang, J, Mullighan, CG, Phillips, LA, et al. Retroviral and chemical mutagenesis identifies Pax5 as a tumor suppressor in B-progenitor acute lymphoblastic leukemia. [ASH Annual Meeting Abstracts.]Blood 2008;112:1789.Google Scholar
Miller, CB, Mullighan, CG, Su, X, et al. Pax5 haploinsufficiency cooperates with BCR-ABL1 to induce acute lymphoblastic leukemia. [ASH Annual Meeting Abstracts.]Blood, 2008;112:293.Google Scholar
Waanders, E, van der Velden, VH, van der Schoot, CE, et al. Integrated use of minimal residual disease classification and IKZF1 alteration status accurately predicts 79% of relapses in pediatric acute lymphoblastic leukemia. Leukemia 2011;25:254–258.CrossRef
Mullighan, CG, Morin, RD, Zhang, J, 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. [ASH Annual Meeting Abstracts.]Blood 2009;114:704.Google Scholar
Graux, C, Cools, J, Melotte, C, et al. Fusion of NUP214 to ABL1 on amplified episomes in T-cell acute lymphoblastic leukemia. Nat Genet 2004;36:1084–1089.CrossRefGoogle ScholarPubMed
Mullighan, CG, Miller, CB, Su, X, et al. ERG deletions define a novel subtype of B-progenitor acute lymphoblastic leukemia. [ASH Annual Meeting Abstracts.]Blood 2007;110:691.Google Scholar
Yamamoto, T, Isomura, M, Xu, Y, et al. PTPN11, RAS and FLT3 mutations in childhood acute lymphoblastic leukemia. Leuk Res 2006;30:1085–1089.CrossRefGoogle ScholarPubMed
Liang, DC, Shih, LY, Fu, JF, et al. K-Ras mutations and N-Ras mutations in childhood acute leukemias with or without mixed-lineage leukemia gene rearrangements. Cancer 2006;106:950–956.CrossRefGoogle ScholarPubMed
Wiemels, JL, Zhang, Y, Chang, J, et al. RAS mutation is associated with hyperdiploidy and parental characteristics in pediatric acute lymphoblastic leukemia. Leukemia 2005;19:415–419.CrossRefGoogle ScholarPubMed
Perentesis, JP, Bhatia, S, Boyle, E, et al. RAS oncogene mutations and outcome of therapy for childhood acute lymphoblastic leukemia. Leukemia 2004;18:685–692.CrossRefGoogle ScholarPubMed
Paulsson, K, Horvat, A, Strombeck, B, et al. Mutations of FLT3, NRAS, KRAS, and PTPN11 are frequent and possibly mutually exclusive in high hyperdiploid childhood acute lymphoblastic leukemia. Genes Chromosomes Cancer 2008;47:26–33.CrossRefGoogle ScholarPubMed
Zhang, J, Mullighan, CG, Harvey, RC, et al. Key pathways are frequently mutated in high-risk childhood acute lymphoblastic leukemia: a report from the Children's Oncology Group. Blood 2011;118:3080–3087.CrossRef
Kamijo, T, Zindy, F, Roussel, MF, et al. Tumor suppression at the mouse INK4a locus mediated by the alternative reading frame product p19ARF. Cell 1997;91:649–659.CrossRefGoogle ScholarPubMed
Quelle, DE, Zindy, F, Ashmun, RA, et al. Alternative reading frames of the INK4a tumor suppressor gene encode two unrelated proteins capable of inducing cell cycle arrest. Cell 1995;83:993–1000.Google ScholarPubMed
Sherr, CJ, Roberts, JM. Inhibitors of mammalian G1 cyclin-dependent kinases. Genes Dev 1995;9:1149–1163.CrossRefGoogle ScholarPubMed
Sherr, CJ. The INK4a/ARF network in tumour suppression. Nat Rev Mol Cell Biol 2001;2:731–737.CrossRefGoogle ScholarPubMed
Hebert, J, Cayuela, JM, Berkeley, J, et al. Candidate tumor-suppressor genes MTS1 (p16INK4A) and MTS2 (p15INK4B) display frequent homozygous deletions in primary cells from T- but not from B-cell lineage acute lymphoblastic leukemias. Blood 1994;84:4038–4044.Google Scholar
Quesnel, B, Preudhomme, C, Fenaux, P. p16ink4a gene and hematological malignancies. Leuk Lymphoma, 1996;22:11–24.CrossRefGoogle ScholarPubMed
Haidar, MA, Cao, XB, Manshouri, T, et al. p16INK4A and p15INK4B gene deletions in primary leukemias. Blood 1995;86:311–315.Google ScholarPubMed
Fizzotti, M, Cimino, G, Pisegna, S, et al. Detection of homozygous deletions of the cyclin-dependent kinase 4 inhibitor (p16) gene in acute lymphoblastic leukemia and association with adverse prognostic features. Blood 1995;85:2685–2690.Google ScholarPubMed
Rasool, O, Heyman, M, Brandter, LB, et al. p15ink4B and p16ink4 gene inactivation in acute lymphocytic leukemia. Blood 1995;85: 3431–3436.Google ScholarPubMed
Okuda, T, Shurtleff, SA, Valentine, MB, et al. Frequent deletion of p16INK4a/MTS1 and p15INK4b/MTS2 in pediatric acute lymphoblastic leukemia. Blood 1995;85:2321–2330.Google ScholarPubMed
Hirama, T, Koeffler, HP. Role of the cyclin-dependent kinase inhibitors in the development of cancer. Blood 1995;86:841–854.Google Scholar
Iolascon, A, Faienza, MF, Coppola, B, et al. Homozygous deletions of cyclin-dependent kinase inhibitor genes, p16(INK4A) and p18, in childhood T cell lineage acute lymphoblastic leukemias. Leukemia 1996;10:255–260.Google ScholarPubMed
Nakao, M, Yokota, S, Kaneko, H, et al. Alterations of CDKN2 gene structure in childhood acute lymphoblastic leukemia: mutations of CDKN2 are observed preferentially in T lineage. Leukemia 1996;10:249–254.Google ScholarPubMed
Cayuela, JM, Madani, A, Sanhes, L, et al. Multiple tumor-suppressor gene 1 inactivation is the most frequent genetic alteration in T-cell acute lymphoblastic leukemia. Blood 1996;87:2180–2186.Google ScholarPubMed
Takeuchi, S, Bartram, CR, Seriu, T, et al. Analysis of a family of cyclin-dependent kinase inhibitors: p15/MTS2/INK4B, p16/MTS1/INK4A, and p18 genes in acute lymphoblastic leukemia of childhood. Blood 1995;86:755–760.Google ScholarPubMed
Takeuchi, S, Koike, M, Seriu, T, et al. Homozygous deletions at 9p21 in childhood acute lymphoblastic leukemia detected by microsatellite analysis. Leukemia 1997;11:1636–1640.CrossRefGoogle ScholarPubMed
Heyman, M, Rasool, O, Borgonovo Brandter, L, et al. Prognostic importance of p15INK4B and p16INK4 gene inactivation in childhood acute lymphocytic leukemia. J Clin Oncol 1996;14:1512–1520.CrossRefGoogle ScholarPubMed
Drexler, HG. Review of alterations of the cyclin-dependent kinase inhibitor INK4 family genes p15, p16, p18 and p19 in human leukemia-lymphoma cells. Leukemia 1998;12:845–859.CrossRefGoogle ScholarPubMed
Nakamura, M, Sugita, K, Inukai, T, et al. p16/MTS1/INK4A gene is frequently inactivated by hypermethylation in childhood acute lymphoblastic leukemia with 11q23 translocation. Leukemia 1999;13:884–890.CrossRefGoogle ScholarPubMed
Mirebeau, D, Acquaviva, C, Suciu, S, et al. The prognostic significance of CDKN2A, CDKN2B and MTAP inactivation in B-lineage acute lymphoblastic leukemia of childhood. Results of the EORTC studies 58881 and 58951. Haematologica 2006;91:881–885.Google ScholarPubMed
Sidransky, D. Two tracks but one race? Cancer genetics. Curr Biol 1996;6:523–525.CrossRefGoogle ScholarPubMed
Kamb, A, Gruis, NA, Weaver-Feldhaus, J, et al. A cell cycle regulator potentially involved in genesis of many tumor types. Science 1994;264:436–440.CrossRefGoogle ScholarPubMed
Nobori, T, Miura, K, Wu, DJ, et al. Deletions of the cyclin-dependent kinase-4 inhibitor gene in multiple human cancers. Nature 1994;368:753–756.CrossRefGoogle ScholarPubMed
Zhang, Y, Xiong, Y, Yarbrough, WG. ARF promotes MDM2 degradation and stabilizes p53: ARF-INK4a locus deletion impairs both the Rb and p53 tumor suppression pathways. Cell 1998;92:725–734.CrossRefGoogle ScholarPubMed
Pomerantz, J, Schreiber-Agus, N, Liegeois, NJ, et al. The Ink4a tumor suppressor gene product, p19Arf, interacts with MDM2 and neutralizes MDM2's inhibition of p53. Cell 1998;92:713–723.CrossRefGoogle ScholarPubMed
Ashcroft, M, Vousden, KH. Regulation of p53 stability. Oncogene 1999;18:7637–7643.CrossRefGoogle ScholarPubMed
Serrano, M, Lee, H, Chin, L, et al. Role of the INK4a locus in tumor suppression and cell mortality. Cell 1996;85:27–37.CrossRefGoogle ScholarPubMed
Krimpenfort, P, Quon, KC, Mooi, WJ, et al. Loss of p16Ink4a confers susceptibility to metastatic melanoma in mice. Nature 2001;413:83–86.CrossRefGoogle ScholarPubMed
Sharpless, NE, Bardeesy, N, Lee, KH, et al. Loss of p16Ink4a with retention of p19Arf predisposes mice to tumorigenesis. Nature 2001;413:86–91.CrossRefGoogle ScholarPubMed
Williams, RT, den Besten, W, Sherr, CJ. Cytokine-dependent imatinib resistance in mouse BCR-ABL+, Arf-null lymphoblastic leukemia. Genes Dev 2007;21:2283–2287.CrossRefGoogle ScholarPubMed
Williams, RT, Roussel, MF, Sherr, CJ. Arf gene loss enhances oncogenicity and limits imatinib response in mouse models of Bcr-Abl-induced acute lymphoblastic leukemia. Proc Natl Acad Sci USA 2006;103:6688–6693.CrossRefGoogle ScholarPubMed
Treanor, LM, Volanakis, EJ, Zhou, S, et al. Functional interactions between Lmo2, the Arf tumor suppressor, and Notch1 in murine T-cell malignancies. Blood 2011;117:5453–5462.CrossRefGoogle ScholarPubMed
Volanakis, EJ, Williams, RT, Sherr, CJ. Stage-specific Arf tumor suppression in Notch1-induced T-cell acute lymphoblastic leukemia. Blood 2009;114:4451–4459.CrossRefGoogle ScholarPubMed
Tycko, B, Smith, SD, Sklar, J. Chromosomal translocations joining LCK and TCRB loci in human T cell leukemia. J Exp Med 1991;174:867–873.CrossRefGoogle ScholarPubMed
Lacronique, V, Boureux, A, Valle, VD, et al. A TEL-JAK2 fusion protein with constitutive kinase activity in human leukemia. Science 1997;278:1309–1312.CrossRefGoogle ScholarPubMed
Hussey, DJ, Nicola, M, Moore, S, et al. The (4;11)(q21;p15) translocation fuses the NUP98 and RAP1GDS1 genes and is recurrent in T-cell acute lymphocytic leukemia. Blood 1999;94:2072–2079.Google ScholarPubMed
van Limbergen, H, Beverloo, HB, van Drunen, E, et al. Molecular cytogenetic and clinical findings in ETV6/ABL1-positive leukemia. Genes Chromosomes Cancer 2001;30:274–282.3.0.CO;2-1>CrossRefGoogle ScholarPubMed
de Keersmaecker, K, Graux, C, Odero, MD, et al. Fusion of EML1 to ABL1 in T-cell acute lymphoblastic leukemia with cryptic t(9;14)(q34;q32). Blood 2005;105:4849–4852.CrossRefGoogle Scholar
Van Vlierberghe, P, van Grotel, M, Tchinda, J, et al. The recurrent SET-NUP214 fusion as a new HOXA activation mechanism in pediatric T-cell acute lymphoblastic leukemia. Blood 2008;111:4668–4680.CrossRefGoogle ScholarPubMed
Maser, RS, Choudhury, B, Campbell, PJ, et al. Chromosomally unstable mouse tumours have genomic alterations similar to diverse human cancers. Nature 2007;447:966–971.CrossRefGoogle ScholarPubMed
Palomero, T, Sulis, ML, Cortina, M, et al. Mutational loss of PTEN induces resistance to NOTCH1 inhibition in T-cell leukemia. Nat Med 2007;13:1203–1210.CrossRefGoogle ScholarPubMed
O'Neil, J, Grim, J, Strack, P, et al. FBW7 mutations in leukemic cells mediate NOTCH pathway activation and resistance to gamma-secretase inhibitors. J Exp Med 2007;204:1813–1824.CrossRefGoogle ScholarPubMed
Tosello, V, Mansour, MR, Barnes, K, et al. WT1 mutations in T-ALL. Blood 2009;114:1038–1045.CrossRefGoogle ScholarPubMed
Kleppe, M, Lahortiga, I, El Chaar, T, et al. Deletion of the protein tyrosine phosphatase gene PTPN2 in T-cell acute lymphoblastic leukemia. Nat Genet 2010;42:530–535.CrossRefGoogle ScholarPubMed
Clappier, E, Cuccuini, W, Kalota, A, et al. The C-MYB locus is involved in chromosomal translocation and genomic duplications in human T-cell acute leukemia (T-ALL), the translocation defining a new T-ALL subtype in very young children. Blood 2007;110:1251–1261.CrossRefGoogle Scholar
Lahortiga, I, de Keersmaecker, K, Van Vlierberghe, P, et al. Duplication of the MYB oncogene in T cell acute lymphoblastic leukemia. Nat Genet 2007;39:593–595.CrossRefGoogle ScholarPubMed
Weng, AP, Ferrando, AA, Lee, W, et al. Activating mutations of NOTCH1 in human T cell acute lymphoblastic leukemia. Science 2004;306:269–271.CrossRefGoogle ScholarPubMed
Ellisen, LW, Bird, J, West, DC, et al. TAN-1, the human homolog of the Drosophila notch gene, is broken by chromosomal translocations in T lymphoblastic neoplasms. Cell 1991;66:649–661.CrossRefGoogle ScholarPubMed
Artavanis-Tsakonas, S, Rand, MD, Lake, RJ. Notch signaling: cell fate control and signal integration in development. Science 1999;284:770–776.CrossRefGoogle ScholarPubMed
Aster, JC, Blacklow, SC, Pear, WS. Notch signalling in T-cell lymphoblastic leukaemia/lymphoma and other haematological malignancies. J Pathol 2011;223:262–273.CrossRefGoogle ScholarPubMed
Rebay, I, Fleming, RJ, Fehon, RG, et al. Specific EGF repeats of Notch mediate interactions with Delta and Serrate: implications for Notch as a multifunctional receptor. Cell 1991;67:687–699.CrossRefGoogle ScholarPubMed
Sanchez-Irizarry, C, Carpenter, AC, Weng, AP, et al. Notch subunit heterodimerization and prevention of ligand-independent proteolytic activation depend, respectively, on a novel domain and the LNR repeats. Mol Cell Biol 2004;24:9265–9273.CrossRefGoogle ScholarPubMed
Nam, Y, Weng, AP, Aster, JC, et al. Structural requirements for assembly of the CSL.intracellular Notch1.Mastermind-like 1 transcriptional activation complex. J Biol Chem 2003;278:21232–21239.CrossRefGoogle ScholarPubMed
Maillard, I, Weng, AP, Carpenter, AC, et al. Mastermind critically regulates Notch-mediated lymphoid cell fate decisions. Blood 2004;104:1696–1702.CrossRefGoogle ScholarPubMed
Fryer, CJ, Lamar, E, Turbachova, I, et al. Mastermind mediates chromatin-specific transcription and turnover of the Notch enhancer complex. Genes Dev 2002;16:1397–1411.CrossRefGoogle ScholarPubMed
Hansson, ML, Popko-Scibor, AE, Saint, JustRibeiro, M, et al. The transcriptional coactivator MAML1 regulates p300 autoacetylation and HAT activity. Nucleic Acids Res 2009;37:2996–3006.CrossRefGoogle ScholarPubMed
Wallberg, AE, Pedersen, K, Lendahl, U, et al. p300 and PCAF act cooperatively to mediate transcriptional activation from chromatin templates by notch intracellular domains in vitro. Mol Cell Biol 2002;22:7812–7819.CrossRefGoogle ScholarPubMed
Palomero, T, Lim, WK, Odom, DT, et al. NOTCH1 directly regulates c-MYC and activates a feed-forward-loop transcriptional network promoting leukemic cell growth. Proc Natl Acad Sci USA 2006;103:18261–18266.CrossRefGoogle ScholarPubMed
Fryer, CJ, White, JB, Jones, KA. Mastermind recruits CycC:CDK8 to phosphorylate the Notch ICD and coordinate activation with turnover. Mol Cell 2004;16:509–520.CrossRefGoogle ScholarPubMed
Malyukova, A, Dohda, T, von der Lehr, N, et al. The tumor suppressor gene hCDC4 is frequently mutated in human T-cell acute lymphoblastic leukemia with functional consequences for Notch signaling. Cancer Res 2007;67:5611–5616.CrossRefGoogle ScholarPubMed
Asnafi, V, Buzyn, A, Le Noir, S, et al. NOTCH1/FBXW7 mutation identifies a large subgroup with favorable outcome in adult T-cell acute lymphoblastic leukemia (T-ALL): a Group for Research on Adult Acute Lymphoblastic Leukemia (GRAALL) study. Blood 2009;113:3918–3924.CrossRefGoogle Scholar
Kox, C, Zimmermann, M, Stanulla, M, et al. The favorable effect of activating NOTCH1 receptor mutations on long-term outcome in T-ALL patients treated on the ALL-BFM 2000 protocol can be separated from FBXW7 loss of function. Leukemia; 24:2005–2013.CrossRefGoogle Scholar
Breit, S, Stanulla, M, Flohr, T, et al. Activating NOTCH1 mutations predict favorable early treatment response and long-term outcome in childhood precursor T-cell lymphoblastic leukemia. Blood 2006;108:1151–1157.CrossRefGoogle ScholarPubMed
Mansour, MR, Sulis, ML, Duke, V, et al. Prognostic implications of NOTCH1 and FBXW7 mutations in adults with T-cell acute lymphoblastic leukemia treated on the MRC UKALLXII/ECOG E2993 protocol. J Clin Oncol 2009;27:4352–4356.CrossRefGoogle ScholarPubMed
Park, MJ, Taki, T, Oda, M, et al. FBXW7 and NOTCH1 mutations in childhood T cell acute lymphoblastic leukaemia and T cell non-Hodgkin lymphoma. Br J Haematol 2009;145:198–206.CrossRefGoogle ScholarPubMed
Grabher, C, von Boehmer, H, Look, AT. Notch 1 activation in the molecular pathogenesis of T-cell acute lymphoblastic leukaemia. Nat Rev Cancer 2006;6:347–359.CrossRefGoogle ScholarPubMed
Aster, J, Pear, W, Hasserjian, R, et al. Functional analysis of the TAN-1 gene, a human homolog of Drosophila notch. Cold Spring Harbor Symp Quant Biol 1994;59:125–136.CrossRefGoogle ScholarPubMed
Pear, WS, Aster, JC, Scott, ML, et al. Exclusive development of T cell neoplasms in mice transplanted with bone marrow expressing activated Notch alleles. J Exp Med 1996;183:2283–2291.CrossRefGoogle ScholarPubMed
Chen, J, Jette, C, Kanki, JP, et al. NOTCH1-induced T-cell leukemia in transgenic zebrafish. Leukemia 2007;21:462–471.CrossRefGoogle ScholarPubMed
Tsuji, H, Ishii-Ohba, H, Ukai, H, et al. Radiation-induced deletions in the 5′ end region of Notch1 lead to the formation of truncated proteins and are involved in the development of mouse thymic lymphomas. Carcinogenesis 2003;24:1257–1268.CrossRefGoogle ScholarPubMed
Dumortier, A, Jeannet, R, Kirstetter, P, et al. Notch activation is an early and critical event during T-cell leukemogenesis in Ikaros-deficient mice. Mol Cell Biol 2006;26: 209–220.CrossRefGoogle ScholarPubMed
O'Neil, J, Calvo, J, McKenna, K, et al. Activating Notch1 mutations in mouse models of T-ALL. Blood 2006;107:781–785.CrossRefGoogle ScholarPubMed
Lin, YW, Nichols, RA, Letterio, JJ, et al. Notch1 mutations are important for leukemic transformation in murine models of precursor-T leukemia/lymphoma. Blood 2006;107:2540–2543.CrossRefGoogle ScholarPubMed
Reschly, EJ, Spaulding, C, Vilimas, T, et al. Notch1 promotes survival of E2A-deficient T cell lymphomas through pre-T cell receptor-dependent and -independent mechanisms. Blood 2006;107:4115–4121.CrossRefGoogle ScholarPubMed
Mantha, S, Ward, M, McCafferty, J, et al. Activating Notch1 mutations are an early event in T-cell malignancy of Ikaros point mutant Plastic/+ mice. Leuk Res 2007;31:321–327.CrossRefGoogle ScholarPubMed
Real, PJ, Ferrando, AA. NOTCH inhibition and glucocorticoid therapy in T-cell acute lymphoblastic leukemia. Leukemia 2009;23: 1374–1377.CrossRefGoogle ScholarPubMed
Moellering, RE, Cornejo, M, Davis, TN, et al. Direct inhibition of the NOTCH transcription factor complex. Nature 2009;462:182–188.CrossRefGoogle ScholarPubMed
Wu, Y, Cain-Hom, C, Choy, L, et al. Therapeutic antibody targeting of individual Notch receptors. Nature 2010;464:1052–1057.CrossRefGoogle ScholarPubMed
Gutierrez, A, Sanda, T, Grebliunaite, R, et al. High frequency of PTEN, PI3K, and AKT abnormalities in T-cell acute lymphoblastic leukemia. Blood 2009;114:647–650.CrossRefGoogle ScholarPubMed
Kawamura, M, Ohnishi, H, Guo, SX, et al. Alterations of the p53, p21, p16, p15 and RAS genes in childhood T-cell acute lymphoblastic leukemia. Leuk Res 1999;23:115–126.CrossRefGoogle ScholarPubMed
Van Vlierberghe, P, Meijerink, JP, Stam, RW, et al. Activating FLT3 mutations in CD4+/CD8− pediatric T-cell acute lymphoblastic leukemias. Blood 2005;106:4414–4415.CrossRefGoogle ScholarPubMed
Paietta, E, Ferrando, AA, Neuberg, D, et al. Activating FLT3 mutations in CD117/KIT(+) T-cell acute lymphoblastic leukemias. Blood 2004;104:558–560.CrossRefGoogle ScholarPubMed
Balgobind, BV, Van Vlierberghe, P, van den Ouweland, AM, et al. Leukemia-associated NF1 inactivation in patients with pediatric T-ALL and AML lacking evidence for neurofibromatosis. Blood 2008;111:4322–4328.CrossRefGoogle ScholarPubMed
Van Vlierberghe, P, Palomero, T, Khiabanian, H, et al. PHF6 mutations in T-cell acute lymphoblastic leukemia. Nat Genet 2010;42:338–342.CrossRefGoogle ScholarPubMed
Meijerink, JP. Genetic rearrangements in relation to immunophenotype and outcome in T-cell acute lymphoblastic leukaemia. Best Pract Res Clin Haematol 2010;23:307–318.CrossRefGoogle ScholarPubMed
Raimondi, SC, Pui, CH, Head, DR, et al. Cytogenetically different leukemic clones at relapse of childhood acute lymphoblastic leukemia. Blood 1993;82:576–580.Google ScholarPubMed
Maloney, KW, McGavran, L, Odom, LF, et al. Acquisition of p16(INK4A) and p15(INK4B) gene abnormalities between initial diagnosis and relapse in children with acute lymphoblastic leukemia. Blood 1999;93: 2380–2385.Google ScholarPubMed
Irving, JA, Minto, L, Bailey, S, et al. Loss of heterozygosity and somatic mutations of the glucocorticoid receptor gene are rarely found at relapse in pediatric acute lymphoblastic leukemia but may occur in a subpopulation early in the disease course. Cancer Res 2005;65:9712–9718.CrossRefGoogle Scholar
Yang, JJ, Bhojwani, D, Yang, W, et al. Genome-wide copy number profiling reveals molecular evolution from diagnosis to relapse in childhood acute lymphoblastic leukemia. Blood 2008;112:4178–4183.CrossRefGoogle ScholarPubMed
van Delft, FW, Horsley, S, Colman, S, et al. Clonal origins of relapse in ETV6-RUNX1 acute lymphoblastic leukemia. Blood 2011;117:6247–6254.CrossRefGoogle ScholarPubMed
Kawamata, N, Ogawa, S, Seeger, K, et al. Molecular allelokaryotyping of relapsed pediatric acute lymphoblastic leukemia. Int J Oncol 2009;34:1603–1612.CrossRefGoogle ScholarPubMed
Kuster, L, Grausenburger, R, Fuka, G, et al. ETV6/RUNX1-positive relapses evolve from an ancestral clone and frequently acquire deletions of genes implicated in glucocorticoid signaling. Blood 2011;117:2658–2667.CrossRefGoogle ScholarPubMed
Szczepański, T, van der Velden, VH, Waanders, E, 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;29:1643–1649.CrossRefGoogle ScholarPubMed
Bateman, CM, Colman, SM, Chaplin, T, et al. Acquisition of genome-wide copy number alterations in monozygotic twins with acute lymphoblastic leukemia. Blood 2010;115:3553–3558.CrossRefGoogle ScholarPubMed
Notta, F, Mullighan, CG, Wang, JC, et al. Evolution of human BCR-ABL1 lymphoblastic leukaemia-initiating cells. Nature 2011;469:362–367.CrossRefGoogle ScholarPubMed
Anderson, K, Lutz, C, van Delft, FW, et al. Genetic variegation of clonal architecture and propagating cells in leukaemia. Nature 2011;469:356–361.CrossRefGoogle ScholarPubMed
Clappier, E, Gerby, B, Sigaux, F, et al. Clonal selection in xenografted human T cell acute lymphoblastic leukemia recapitulates gain of malignancy at relapse. J Exp Med 2011;208:653–661.CrossRefGoogle Scholar
Goodman, RH, Smolik, S. CBP/p300 in cell growth, transformation, and development. Genes Dev 2000;14:1553–1577.Google ScholarPubMed
Lahoud, MH, Ristevski, S, Venter, DJ, et al. Gene targeting of Desrt, a novel ARID class DNA-binding protein, causes growth retardation and abnormal development of reproductive organs. Genome Res 2001;11:1327–1334.CrossRefGoogle ScholarPubMed
Harrison, CJ, Haas, O, Harbott, J, et al. Detection of prognostically relevant genetic abnormalities in childhood B-cell precursor acute lymphoblastic leukaemia: recommendations from the Biology and Diagnosis Committee of the International Berlin–Frankfurt–Münster Study Group. Br J Haematol 2010;151:132–142.CrossRefGoogle ScholarPubMed
Balgobind, BV, Raimondi, SC, Harbott, J, et al. Novel prognostic subgroups in childhood 11q23/MLL-rearranged acute myeloid leukemia: results of an international retrospective study. Blood 2009;114:2489–2496.CrossRefGoogle ScholarPubMed
Pui, CH, Raimondi, SC, Hancock, ML, et al. Immunologic, cytogenetic, and clinical characterization of childhood acute lymphoblastic leukemia with the t(1;19) (q23;p13) or its derivative. J Clin Oncol 1994;12:2601–2606.CrossRefGoogle ScholarPubMed
Kager, L, Lion, T, Attarbaschi, A, et al. Incidence and outcome of TCF3-PBX1-positive acute lymphoblastic leukemia in Austrian children. Haematologica 2007;92:1561–1564.CrossRefGoogle ScholarPubMed
Hunger, SP, Li, S, Fall, MZ, et al. The proto-oncogene HLF and the related basic leucine zipper protein TEF display highly similar DNA-binding and transcriptional regulatory properties. Blood 1996;87:4607–4617.Google ScholarPubMed

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×