Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-p2v8j Total loading time: 0 Render date: 2024-06-08T08:07:58.959Z Has data issue: false hasContentIssue false

10 - Molecular genetics of acute lymphoblastic leukemia

from Part II - Cell biology and pathobiology

Published online by Cambridge University Press:  01 July 2010

Adolfo A. Ferrando
Affiliation:
Assistant Professor of Pathology and Pediatrics Institute for Cancer Genetics, Columbia University, Irving Cancer Research Center, New York, NY, USA
Jeffrey E. Rubnitz
Affiliation:
Associate Member, Department of Hematology/Oncology, Director of Fellowship Program, St. Jude Children's Research Hospital, Memphis, TN, USA
A. Thomas Look
Affiliation:
Professor of Pediatrics, Harvard Medical School, Vice Chair for Research, Department of Pediatric Oncology, Dana-Farber Cancer Institute, Boston, MA, USA
Ching-Hon Pui
Affiliation:
St. Jude Children's Research Hospital, Memphis
Get access

Summary

Introduction

Although most cases of acute lymphoblastic leukemia (ALL) do not arise from a recognized predisposing genetic condition, this disease is still considered to have a genetic basis. That is, somatically acquired genetic changes contribute in pivotal ways to the genesis of ALL and have important implications for its diagnosis and treatment. Such acquired lesions, restricted to the leukemic clone, include alterations in both the number (ploidy) and structure of the blast cell chromosomes, the latter consisting of reciprocal translocations, inversions, deletions, gene amplifications, and point mutations. Although many of these abnormalities can be detected by routine cytogenetic analysis, others require molecular assays. Molecular identification of the genetic targets of these alterations has led to the isolation and characterization of numerous oncogenes and tumor suppressors, providing valuable clues to the mechanisms of leukemogenesis. In this chapter, we review the molecular basis of childhood ALL, emphasizing the oncogenes, oncoproteins, and tumor suppressors that appear to be critical components of a limited number of transforming pathways.

Proto-oncogene activation

Standard karyotyping can identify chromosomal translocations, including both recurrent and random rearrangements, in about 50% of cases of childhood ALL. When molecular assays are used with cytogenetic analysis, evidence of a translocation is seen in about 70% of cases (Fig. 10.1). Translocations in ALL most often disrupt genes that encode transcription factors (Tables 10.1 and 10.2), defined as nuclear proteins that bind to DNA and regulate the expression of other cellular proteins.

Type
Chapter
Information
Childhood Leukemias , pp. 272 - 297
Publisher: Cambridge University Press
Print publication year: 2006

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

Bishop, J. M.The molecular genetics of cancer. Science, 1987; 235: 305–11.CrossRefGoogle ScholarPubMed
Solomon, E., Borrow, J., & Goddard, A. D.Chromosome aberrations and cancer. Science, 1991; 254: 1153–60.CrossRefGoogle ScholarPubMed
Rowley, J. D.Molecular cytogenetics: Rosetta stone for understanding cancer – twenty-ninth G. H. A. Clowes memorial award lecture. Cancer Res, 1990; 50: 3816–25.Google Scholar
Rabbitts, T. H.Chromosomal translocations in human cancer. Nature, 1994; 372: 143–9.CrossRefGoogle ScholarPubMed
Ferrando, A. A. & Look, A. T.Clinical implications of recurring chromosomal and associated molecular abnormalities in acute lymphoblastic leukemia. Semin Hematol, 2000; 37: 381–95.CrossRefGoogle ScholarPubMed
Look, A. T.Oncogenic transcription factors in the human acute leukemias. Science, 1997; 278: 1059–64.CrossRefGoogle ScholarPubMed
Pui, C.-H.Childhood leukemias. N Engl J Med, 1995; 332: 1618–30.CrossRefGoogle ScholarPubMed
Okuda, T., Fisher, R., & Downing, J. R.Molecular diagnostics in pediatric acute lymphoblastic leukemia. Mol Diagn, 1996; 1: 139–51.CrossRefGoogle ScholarPubMed
Cleary, M. L.Oncogenic conversion of transcription factors by chromosomal translocations. Cell, 1991; 66: 619–22.CrossRefGoogle ScholarPubMed
Sawyers, C. L.Molecular genetics of acute leukaemia. Lancet, 1997; 349: 196–200.CrossRefGoogle ScholarPubMed
O'Connor, N. T. J., Weatherall, D. J., Feller, A. C., et al.Re arrangement of the T cell-receptor beta-chain gene in the diagnosis of lymphoproliferative disorders. Lancet, 1985; 1: 1295–7.CrossRefGoogle Scholar
Ribeiro, R. C., Abromowitch, M., Raimondi, S. C., et al.Clinical and biologic hallmarks of the Philadelphia chromosome in childhood acute lymphoblastic leukemia. Blood, 1987; 70: 948–53.Google ScholarPubMed
Bartram, C. R., 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: 277–80.CrossRefGoogle ScholarPubMed
Gale, R. P. & Canaani, E.An 8-kilobase abl RNA transcript in chronic myelogenous leukemia. Proc Natl Acad Sci U S A, 1984; 81: 5648–52.CrossRefGoogle ScholarPubMed
Collins, S. J., Kubonishi, I., Miyoshi, I., & Groudine, M. T.Altered transcription of the c-abl oncogene in K562 and other chronic myelogenous leukemia cells. Science, 1984; 225: 72–4.CrossRefGoogle Scholar
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: 1429–33.CrossRefGoogle ScholarPubMed
Canaani, E., Gale, R. P., Steiner-Saltz, D., et al.Altered transcription of an oncogene in chronic myeloid leukemia. Lancet, 1984; 1(8377): 593–5.CrossRefGoogle Scholar
Shtivelman, E., Lifshitz, B., Gale, R. P., & Canaani, E.Fused transcript of abl and bcr genes in chronic myelogenous leukemia. Nature, 1985; 315: 550–4.CrossRefGoogle Scholar
Heisterkamp, N., Stephenson, J. R., Groffen, J., et al.Localization of the c-abl oncogene adjacent to a translocation breakpoint in chronic myelocytic leukaemia. Nature, 1983; 306: 239–42.CrossRefGoogle Scholar
Leibowitz, D., Schaefer-Rego, K., Popenoe, D. W., et al.Variable breakpoints on the Philadelphia chromosome in chronic myelogenous leukemia. Blood, 1985; 66: 243–5.Google ScholarPubMed
Grosveld, G., Verwoerd, T., 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: 607–16.CrossRefGoogle ScholarPubMed
Groffen, J., Stephenson, J. R., Heisterkamp, N., et al.Philadelphia chromosomal breakpoints are clustered within a limited region, bcr, on chromosome 22. Cell, 1984; 36: 93–9.CrossRefGoogle ScholarPubMed
Heisterkamp, N., Stam, K., Groffen, J., et al.Structural organization of the bcr gene and its role in Ph1 translocation. Nature, 1985; 315: 758.CrossRefGoogle Scholar
Chan, L. C., Karhi, K. K., Rayter, S. I., et al.A novel abl protein expressed in Philadelphia chromosome-positive acute lymphoblastic leukaemia. Nature, 1987; 325: 635–7.CrossRefGoogle ScholarPubMed
Clark, S. S., McLaughlin, J., Crist, W. M., et al.Unique forms of the abl tyrosine kinase distinguish Ph1-positive CML from Ph1-positive ALL. Science, 1987; 235: 85.CrossRefGoogle ScholarPubMed
Kurzrock, R., Shtalrid, M., Romero, P., et al.A novel c-abl protein product in Philadelphia-positive acute lymphoblastic leukemia. Nature, 1987; 325: 631–5.CrossRefGoogle Scholar
Etten, R. A., 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: 669–78.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: 785–8.CrossRefGoogle ScholarPubMed
Sawyers, C. L., McLaughlin, J., Goga, A., Havlik, M., & Witte, O.The nuclear tyrosine kinase c-Abl negatively regulates cell growth. Cell, 1994; 77: 121–31.CrossRefGoogle ScholarPubMed
Mattioni, T., Jackson, P. K., Bchini-Hooft van Huijsduijnen, O., & Picard, D.Cell cycle arrest by tyrosine kinase Abl involves altered early mitogenic response. Oncogene, 1995; 10: 1325–33.Google ScholarPubMed
Goga, A., Liu, X., Hambuch, T. M., et al.p53 dependent growth suppression by the c-Abl nuclear tyrosine kinase. Oncogene, 1995; 11: 791–9.Google ScholarPubMed
Wang, J. Y.Regulation of cell death by Abl tyrosine kinase. Oncogene, 2003; 19: 5643–50.CrossRefGoogle Scholar
Tybulewicz, V. L., Crawford, C. E., Jackson, P. K., Bronson, R. T., & Mulligan, R. C.Neonatal lethality and lymphopenia in mice with a homozygous disruption of the c-abl proto-oncogene. Cell, 1991; 65: 1153–63.CrossRefGoogle ScholarPubMed
Schwartzberg, P. L., Stall, A. M., Hardin, J. D., et al.Mice homozygous for the ablm1 mutation show poor viability and depletion of selected B and T cell populations. Cell, 1991; 65: 1165–75.CrossRefGoogle ScholarPubMed
Lugo, T. G., Pendergast, A. M., Muller, A. J., & Witte, O. N.Tyrosine kinase activity and transformation potency of bcr-abl oncogene products. Science, 1990; 247: 1079–82.CrossRefGoogle ScholarPubMed
Daley, G. Q., McLaughlin, J., Witte, O. N., & Baltimore, D.The CML-specific P210 bcr/abl protein, unlike v-abl, does not transform NIH/3T3 fibroblasts. Science, 1987; 237: 532–5.CrossRefGoogle Scholar
Daley, G. Q. & Baltimore, D.Transformation of an interleukin 3-dependent hematopoietic cell line by the chronic myelogenous leukemia-specific P210bcr/abl protein. Proc Natl Acad Sci U S A, 1988; 85: 9312–16.CrossRefGoogle ScholarPubMed
Elefanty, A. G., Hariharan, I. K., & Cory, S.bcr-abl, the hallmark of chronic myeloid leukaemia in man, induces multiple haemopoietic neoplasms in mice. EMBO J, 1990; 9: 1069–78.Google ScholarPubMed
Gishizky, M. L., Johnson-White, J., & Witte, O. N.Efficient transplantation of BCR-ABL-induced chronic myelogenous leukemia-like syndrome in mice. Proc Natl Acad Sci U S A, 1993; 90: 3755–9.CrossRefGoogle ScholarPubMed
Kelliher, M., Knott, A., McLaughlin, J., Witte, O. N., & Rosenberg, N.Differences in oncogenic potency but not target cell specificity distinguish the two forms of the BCR/ABL oncogene. Mol Cell Biol, 1991; 11: 4710–16.CrossRefGoogle Scholar
Etten, R. A.. Studying the pathogenesis of BCR-ABL+ leukemia in mice. Oncogene, 2002; 21: 8643–51.CrossRefGoogle ScholarPubMed
Cortez, D., Reuther, G., & Pendergast, A. M.The Bcr-Abl tyrosine activates mitogenic signaling pathways and stimulates G1-to-S phase transition in hematopoietic cells. Oncogene, 1997; 15: 2333–42.CrossRefGoogle ScholarPubMed
Varticovski, L., Daley, G. Q., Jackson, P., Baltimore, D., & Cantley, L. C.Activation of phosphatidylinositol 3-Kinase in cells expressing abl oncogene variants. Mol Cell Biol, 1991; 11: 11017–113.CrossRefGoogle ScholarPubMed
Reuther, J. Y., Reuther, G. W., Cortez, D., Pendergast, A. M., & Baldwin, A. S. Jr.A requirement for NF-κB activation in Bcr-Abl-mediated transformation. Genes Dev, 1998; 12: 968–81.CrossRefGoogle ScholarPubMed
Carlesso, N., Frank, D. A., & Griffin, J. D.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: 811–20.CrossRefGoogle ScholarPubMed
Raitano, A. B., Halpern, J. R., Hambuch, T. M., & Sawyers, C. L.The Bcr-Abl leukemia oncogene activates Jun kinase and requires Jun for transformation. Proc Natl Acad Sci U S A, 1995; 92: 11746–50.CrossRefGoogle ScholarPubMed
Sawyers, C. L., McLaughlin, J., & Witte, O. N.Genetic requirement for Ras in the transformation of fibroblasts and hematopoietic cells by the Bcr-Abl oncogene. J Exp Med, 1995; 181: 307–13.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: 726–36.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: 6151–61.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: 8591–604.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: 489–94.Google ScholarPubMed
Fletcher, J. A., Lynch, E. A., Kimball, V. M., et al.Translocation (9;22) is associated with extremely poor prognosis in intensively treated children with acute lymphoblastic leukemia. Blood, 1991; 77: 435–9.Google ScholarPubMed
Roberts, W. M., Rivera, G. K., Raimondi, S. C., et al.Intensive chemotherapy for Philadelphia-chromosome-positive acute lymphoblastic leukaemia. Lancet, 1994; 343: 331–2.CrossRefGoogle ScholarPubMed
Ribeiro, R. C., Broniscer, A., Rivera, G. K., 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: 1493–6.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: 100–4.Google ScholarPubMed
Druker, B. J. & Lydon, N. B.Lessons learned from the development of an abl tyrosine kinase inhibitor for chronic myelogen ous leukemia. J Clin Invest, 2000; 105: 3–7.CrossRefGoogle Scholar
Druker, B. J., Talpaz, M., Resta, D. J., 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: 1031–7.CrossRefGoogle ScholarPubMed
Druker, B. J., Sawyers, C. L., 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: 1038–42.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: 645–52.CrossRefGoogle ScholarPubMed
Sawyers, C. L., Hochhaus, A., Feldman, E., et al.Imatinib induces hematologic and cytogenetic responses in patients with chronic myelogenous leukemia in myeloid blast crisis: results of a phase II study. Blood, 2002; 99: 3530–9.CrossRefGoogle ScholarPubMed
Talpaz, M., Silver, R. T., Druker, B. J., 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: 1928–37.CrossRefGoogle ScholarPubMed
Bubnoff, N. von, 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: 829–38.CrossRefGoogle Scholar
Gorre, M. E. & Sawyers, C. L.Molecular mechanisms of resistance to STI571 in chronic myeloid leukemia. Curr Opin Hematol, 2002; 9: 303–7.CrossRefGoogle ScholarPubMed
Guilhot, F., Gardembas, M., Rousselot, P., et al.Imatinib in combination with cytarabine for the treatment of Philadelphia-positive chronic myelogenous leukemia chronic-phase patients: rationale and design of phase I/II trials. Semin Hematol, 2003; 40: 92–7.CrossRefGoogle ScholarPubMed
Thiesing, J. T., Ohno-Jones, S., Kolibaba, K. S., & Druker, B. J.Efficacy of STI571, an abl tyrosine kinase inhibitor, in conjunction with other antileukemic agents against bcr-abl-positive cells. Blood, 2000; 96: 3195–9.Google ScholarPubMed
Huron, D. R., Gorre, M. E., Kraker, A. J., et al.A novel pyridopyrimidine inhibitor of Abl kinase is a picomolar inhibitor of Bcr-abl-driven K562 cells and is effective against STI571-resistant Bcr-abl mutants. Clin Cancer Res, 2003; 9: 1267–73.Google ScholarPubMed
Wisniewski, D., Lambek, C. L., Liu, C., et al.Characterization of potent inhibitors of the Bcr-Abl and the c-kit receptor tyrosine kinases. Cancer Res, 2002; 62: 4244–55.Google ScholarPubMed
Privitera, E., Kamps, M. P., Hayashi, Y., et al.Different molecular consequences of the 1;19 chromosomal translocation in childhood B-cell precursor acute lymphoblastic leukemia. Blood, 1992; 79: 1781–8.Google ScholarPubMed
Crist, W. M., Carroll, A. J., Shuster, J. J., et al.Poor prognosis of children with pre-B acute lymphoblastic leukemia is associated with the t(1;19)(q23;p13): a Pediatric Oncology Group study. Blood, 1990; 76: 117–22.Google Scholar
Raimondi, S. C., Behm, F. G., Roberson, P. K., 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: 1380–8.CrossRefGoogle Scholar
Hunger, S. P.Chromosomal translocations involving the E2A gene in acute lymphoblastic leukemia: clinical features and molecular pathogenesis. Blood, 1996; 87: 1211–24.Google ScholarPubMed
Rauskolb, C. & Wieschaus, E.Coordinate regulation of downstream genes by extradenticle and the homeotic selector proteins. EMBO J, 1994; 13: 3561–9.Google ScholarPubMed
Flegel, W. A., Singson, A. W., Margolis, J. S., et al.Dpbx, a new homeobox gene closely related to the human proto-oncogene pbx1 molecular structure and developmental expression. Mech Dev, 1993; 41: 155–61.CrossRefGoogle ScholarPubMed
Rauskolb, C., Peifer, M., & Weischaus, E.Extradenticle, a regulator of homeotic gene activity, is a homolog of the homeobox-containing human proto-oncogene pbx1. Cell, 1993; 74: 1101–12.CrossRefGoogle ScholarPubMed
Dijk, M. A. & Murre, C.Extradenticle raises the DNA binding specificity of homeotic selector gene products. Cell, 1994; 78: 617–24.CrossRefGoogle ScholarPubMed
Mellentin, J. D., Murre, C., Donlon, T. A., et al.The gene for enhancer binding proteins E12/E47 lies at the t(1;19) breakpoint in acute leukemias. Science, 1989; 246: 379–82.CrossRefGoogle Scholar
Kamps, M. P., Murre, C., Sun, X. H., & Baltimore, D.A new homeobox gene contributes the DNA binding domain of the t(1;19) translocation protein in pre-B ALL. Cell, 1990; 60: 547–55.CrossRefGoogle Scholar
Nourse, J., Mellentin, J. D., 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: 535–45.CrossRefGoogle Scholar
Peifer, M. & Wieschaus, E.Mutations in the Drosophila gene extradenticle affect the way specific homeo domain proteins regulate segmental identity. Genes Dev, 1990; 4: 1209–23.CrossRefGoogle ScholarPubMed
Kamps, M. P.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
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
Murre, C., McCaw, P. S., & Baltimore, D.A new DNA binding and dimerization motif in immunoglobulin enhancer binding, daughterless, MyoD, and myc proteins. Cell, 1989; 56: 777–83.CrossRefGoogle ScholarPubMed
Murre, C., McCaw, P. S., Vaessin, H., et al.Interactions between heterologous helix-loop-helix proteins generate complexes that bind specifically to a common DNA sequence. Cell, 1989; 58: 537–44.CrossRefGoogle ScholarPubMed
Henthorn, P., Kiledjian, M., & Kadesch, T.Two distinct transcription factors that bind the immunoglobulin enhancer E5/E2 motif. Science, 1990; 247: 467–70.CrossRefGoogle Scholar
Sun, X. H. & Baltimore, D.An inhibitory domain of E12 transcription factor prevents DNA binding in E12 homodimers but not in E12 heterodimers. Cell, 1991; 64: 459–70.CrossRefGoogle ScholarPubMed
Aronheim, A., Shiran, R., Rosen, A., & Walker, M. D.The E2A gene product contains two separable and functionally distinct transcription activation domains. Proc Natl Acad Sci U S A, 1993; 90: 8063–7.CrossRefGoogle ScholarPubMed
Dedera, D. A., Waller, E. K., LeBrun, D. P., et al.Chimeric homeobox gene E2A-PBX1 induces proliferation, apoptosis, and malignant lymphomas in transgenic mice. Cell, 1993; 74: 833–43.CrossRefGoogle ScholarPubMed
LeBrun, D. L. & Cleary, M. L.Fusion with E2A alters the transcriptional properties of the homeodomain protein PBX1 in t(1;19) leukemias. Oncogene, 1994; 9: 1641–7.Google Scholar
Bain, G., Robanus-Maandag, E. C., Izon, D. J., et al.E2A proteins are required for proper B-cell development and initiation of immunoglobulin gene rearrangements. Cell, 1994; 79: 885–92.CrossRefGoogle ScholarPubMed
Bain, G., Romanow, W. J., Albers, K., Havran, W. L., & Murre, C.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: 875–84.CrossRefGoogle ScholarPubMed
Aspland, S. E., Bendall, H. H., & Murre, C.The role of E2A-PBX1 in leukemogenesis. Oncogene, 2001; 20: 5708–17.CrossRefGoogle ScholarPubMed
Kamps, M. P. & 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: 351–7.CrossRefGoogle Scholar
Kamps, M. P. & Wright, D. D.Oncoprotein E2A-Pbx1 immortalizes a myeloid progenitor in primary marrow cultures without abrogating its factor-dependence. Oncogene, 1994; 9: 3159–66.Google ScholarPubMed
Monica, K., LeBrun, D. P., Dedera, D. A., Brown, R., & Cleary, M. L.Transformation properties of the E2A-PBX1 chimeric oncoprotein: fusion with E2A is essential, but the PBX1 homeo domain is dispensable. Mol Cell Biol, 1994; 14: 8304–14.CrossRefGoogle Scholar
LeBrun, D. P., Matthews, B. P., Feldman, B. J., & Cleary, M. L.The chimeric oncoproteins E2A-PBX1 and E2A-HLF are concentrated within spherical nuclear domains. Oncogene, 1997; 15: 2059–67.CrossRefGoogle ScholarPubMed
Chang, C. P., Shen, W. F., Rozenfeld, S., et al.Pbx proteins display hexapeptide-dependent cooperative DNA binding with a subset of Hox proteins. Genes Dev, 1995; 9: 663–74.CrossRefGoogle ScholarPubMed
Kamps, M. P., Wright, D. D., & Lu, Q.DNA-binding by oncoprotein E2a-Pbx1 is important for blocking differentiation but dispensable for fibroblast transformation. Oncogene, 1996; 12: 19–30.Google ScholarPubMed
Chang, C. P., DeVivo, I., & Cleary, M. L.The Hox cooperativity motif of the chimeric oncoprotein E2a-Pbx1 is necessary and sufficient for oncogenesis. Mol Cell Biol, 1997; 17: 81–8.CrossRefGoogle ScholarPubMed
Lu, Q. & Kamps, M. P.Heterodimerization of Hox proteins with Pbx1 and oncoprotein E2a-Pbx1 generates unique DNA-binding specificities 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 Scholar
Chang, C. P., 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: 5679–87.CrossRefGoogle Scholar
Fu, X. & Kamps, M. P.E2a-Pbx1 induces aberrant expression of tissue-specific and developmentally regulated genes when expressed in NIH 3T3 fibroblasts. Mol Cell Biol, 1997; 17: 1503–12.CrossRefGoogle ScholarPubMed
Smith, K. S., Jacobs, Y., Chang, C. P., & Cleary, M. L.Chimeric oncoprotein E2a-Pbx1 induces apoptosis of hematopoietic cells by a p53-independent mechanism that is suppressed by Bcl-2. Oncogene, 1997; 14: 2917–26.CrossRefGoogle ScholarPubMed
McWhirter, J. R., Neuteboom, S. T. C., Wancewicz, E. V., et al.Oncogenic homeodomain transcription factor E2A-Pbx1 activates a novel WNT gene in pre-B acute lymphoblastoid leukemia. Proc Natl Acad Sci U S A, 1999; 96: 11464–9.CrossRefGoogle ScholarPubMed
Borowitz, M. J., Hunger, S. P., Carroll, A. J., et al.Predictability of the t(1;19)(q23;p13) from surface antigen phenotype. Implications for screening cases of childhood ALL for molecular analysis. A Pediatric Oncology Group study. Blood, 1993; 82: 1086–91.Google Scholar
Pui, C. H. & Crist, W. M.Cytogenetic abnormalities in childhood acute lymphoblastic leukemia correlates with clinical features and treatment outcome. Leuk Lymphoma, 1992; 7: 259–74.CrossRefGoogle ScholarPubMed
Rivera, G. K., Raimondi, S. C., Hancock, M. L., et al.Improved outcome in childhood acute lymphoblastic leukaemia with reinforced early treatment and rotational combination chemotherapy. Lancet, 1991; 337: 61–6.CrossRefGoogle ScholarPubMed
Hunger, S. P., Ohyashiki, K., Toyama, K., & Cleary, M. L.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: 1608–20.CrossRefGoogle Scholar
Inaba, T., Roberts, W. M., Shapiro, L. H., et al.Fusion of the leucine zipper gene HLF to the E2A gene in human acute B-lineage leukemia. Science, 1992; 257: 531–4.CrossRefGoogle ScholarPubMed
Landschulz, W. H., Johnson, P. F., & McKnight, S. L.The leucine zipper: a hypothetical structure common to a new class of DNA binding proteins. Science, 1988; 240: 1759–64.CrossRefGoogle ScholarPubMed
O'Shea, E. K., Klemm, J. D., Kim, P. S., & Alber, T.X-ray structure of the GCN4 leucine zipper, a two-stranded, parallel coiled coil. Science, 1991; 254: 539–44.CrossRefGoogle ScholarPubMed
Drolet, D. W., Scully, K. M., Simmons, D. M., 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: 1739–53.CrossRefGoogle ScholarPubMed
Wasylyk, B., Hahn, S. L., & Giovane, A.The Ets family of transcription factors [published erratum appears in Eur J Biochem, 1993; 215: 907]. Eur J Biochem, 1993; 211: 7–18.CrossRefGoogle Scholar
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: 4307–17.Google ScholarPubMed
Hunger, S. P., Brown, R., & Cleary, M. L.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: 5986–96.CrossRefGoogle Scholar
Inaba, T., Shapiro, L. H., 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: 3403–13.CrossRefGoogle ScholarPubMed
Yoshihara, T., Inaba, T., Shapiro, L. H., Kato, J., & Look, A. T.E2A-HLF-mediated cell transformation requires both the trans-activation domain of E2A and the leucine zipper dimerization domain of HLF. Mol Cell Biol, 1995; 15: 3247–55.CrossRefGoogle ScholarPubMed
Inukai, T., Inaba, T., Yoshihara, T., & Look, A. T.Cell transformation mediated by homodimeric E2A-HLF transcription factors. Mol Cell Biol, 1997; 17: 1417–24.CrossRefGoogle ScholarPubMed
Inaba, T., Inukai, T., Yoshihara, T., et al.Reversal of apoptosis by the leukaemia-associated E2A-HLF chimaeric transcription factor. Nature, 1996; 382: 541–4.CrossRefGoogle ScholarPubMed
Metzstein, M. M., Hengartner, M. O., Tsung, N., Ellis, R. E., & Horvitz, H. R.Transcriptional regulator of programmed cell death encoded by Caenorhabditis elegans gene ces-2. Nature, 1996; 382: 545–7.CrossRefGoogle ScholarPubMed
Metzstein, M. & Horvitz, H. R.The C. elegans cell-death specification gene ces-1 encodes a Snail-family zinc-finger protein. Mol Cell, 1999; 4: 309–19.CrossRefGoogle Scholar
Inukai, T., Inoue, A., Kurosawa, H., et al.SLUG, a ces-1-related zinc-finger transcription factor gene with antiapoptotic activity, is a downstream target of the E2A-HLF oncoprotein. Mol Cell, 1999; 4: 343–52.CrossRefGoogle ScholarPubMed
Inoue, A., Seidel, M. G., Wu, W., et al.Slug, a highly conserved zinc finger transcriptional repressor, protects hematopoietic progenitor cells from radiation-induced apoptosis in vivo. Cancer Cell, 2002; 2: 279–88.CrossRefGoogle ScholarPubMed
Hunger, S. P., Devaraj, P. E., Foroni, L., Secker-Walker, L. M., & Cleary, M. L.Two types of genomic rearrangements create alternative E2A-HLF fusion proteins in t(17;19)-ALL. Blood, 1994; 83: 2261–7.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: 322–31.Google Scholar
Devaraj, P. E., 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: 1131–8.Google Scholar
Kaneko, Y., Maseki, N., Takasaki, M., et al.Clinical and hematologic characteristics in acute leukemia with 11q23 translocations. Blood, 1986; 67: 484–8.Google ScholarPubMed
Raimondi, S. C., Kalwinsky, D. K., Hayashi, Y., et al.Cytogen etics of childhood acute nonlymphocytic leukemia. Cancer Genet Cytogenet, 1989; 40: 13–27.CrossRefGoogle Scholar
Pui, C.-H., Frankel, L. S., Carroll, A. J., 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: 440–7.Google Scholar
Raimondi, S. C.Current status of cytogenetic research in childhood acute lymphoblastic leukemia. Blood, 1993; 81: 2237–51.Google ScholarPubMed
Heerema, N. A., Arthur, D. C., 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 Children's Cancer Group. Blood, 1994; 83: 2274–84.Google Scholar
Pui, C. H., Kane, J. R., & Crist, W. M.Biology and treatment of infant leukemias. Leukemia, 1995; 9: 762–9.Google ScholarPubMed
Chen, C.-S., Sorensen, P. H. B., Domer, P. H., 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: 2386–93.Google ScholarPubMed
Pui, C.-H., Behm, F. G., Raimondi, S. C., et al.Secondary acute myeloid leukemia in children treated for acute lymphoid leukemia. N Engl J Med, 1989; 321: 136–42.CrossRefGoogle ScholarPubMed
Rubnitz, J. E., Behm, F. G., & Downing, J. R.11q23 rearrangements in acute leukemia. Leukemia, 1996; 10: 74–82.Google ScholarPubMed
Ziemin-van der Poel, S., McCabe, N. R., Gill, H. J., et al.Identification of a gene, MLL, that spans the breakpoint in 11q23 translocations associated with human leukemias. Proc Natl Acad Sci U S A, 1991; 88: 10735–9.CrossRefGoogle ScholarPubMed
Tkachuk, D. C., Kohler, S., & Cleary, M. L.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: 701–8.CrossRefGoogle Scholar
Ayton, P. M. & Cleary, M. L.Molecular mechanisms of leukemogenesis mediated by MLL fusion proteins. Oncogene, 2001; 20: 5695–707.CrossRefGoogle ScholarPubMed
Behm, F. G., Raimondi, S. C., Frestedt, J. L., et al.Rearrangement of the MLL gene confers a poor prognosis in childhood acute lymphoblastic leukemia, regardless of presenting age. Blood, 1996; 87: 2870–7.Google ScholarPubMed
Rubnitz, J. E., Link, M. P., Shuster, J. J., et al.Frequency and prognostic significance of HRX rearrangements in infant acute lymphoblastic leukemia: a Pediatric Oncology Group Study. Blood, 1994; 84: 570–3.Google ScholarPubMed
Pui, C.-H., Behm, F. G., Downing, J. R., et al.11q23/MLL re arrangment confers a poor prognosis in infants with acute lymphoblastic leukemia. J Clin Oncol, 1994; 12: 909–15.CrossRefGoogle Scholar
Cimino, G., Rapanotti, M. C., Rivolta, A., et al.Prognostic relevance of ALL-1 gene rearrangement in infant acute leukemias. Leukemia, 1995; 9: 391–5.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: 544–6.Google ScholarPubMed
Raimondi, S. C., Pui, C.-H., Downing, J. R., Head, D. R., & Behm, F. G.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: 1881–6.Google ScholarPubMed
Mazo, A. M., Huang, D. H., Mozer, B. A., & Dawid, I. B.The trithorax gene, a trans-acting regulator of the bithorax-complex in Drosophila, encodes a protein with zinc-binding domains. Proc Natl Acad Sci U S A, 1990; 87: 2112–16.CrossRefGoogle ScholarPubMed
Reeves, R. & Nissen, M. S.The A-T-DNA-binding domain of mammalian high mobility group I chromosomal proteins. J Biol Chem, 1990; 265: 8573–82.Google Scholar
Ma, Q., Alder, H., Nelson, K. K., et al.Analysis of the murine ALL-1 gene reveals conserved domains with human ALL-1 and identified a motif shared with DNA methyltransferases. Proc Natl Acad Sci U S A, 1993; 90: 6350–4.CrossRefGoogle ScholarPubMed
Yokoyama, A., Kitabayashi, I., Ayton, P. M., Cleary, M. L., & Ohki, M.Leukemia proto-oncoprotein MLL is proteolytically processed into 2 fragments with opposite transcriptional properties. Blood, 2002; 100: 3710–18.CrossRefGoogle ScholarPubMed
Hsieh, J. J., Ernst, P., Erdjument-Bromage, H., Tempst, P., & Korsmeyer, S. J.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: 186–94.CrossRefGoogle ScholarPubMed
Jones, R. S. & Gelbart, W. M.The Drosophila polycomb-group gene enhancer of zeste contains a region with sequence similarity to trithorax. Mol Cell Biol, 1993; 13: 6357–66.CrossRefGoogle ScholarPubMed
Cui, X., de Vivo, I., Slany, R., et al.Association of SET domain and myotubularin-related proteins modulates growth control. Nat Genet, 1998; 18: 331–7.CrossRefGoogle ScholarPubMed
Milne, T. A., Briggs, S. D., Brock, H. W., et al.MLL targets SET domain methyltransferase activity to Hox gene promoters. Mol Cell, 2002; 10: 1107–17.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: 1119–28.CrossRefGoogle ScholarPubMed
Yu, B. D., Hess, J. L., Horning, S. E., Brown, G. A., & Korsmeyer, S. J.Altered Hox expression and segmental identity in Mll-mutant mice. Nature, 1995; 378: 505–8.CrossRefGoogle ScholarPubMed
Hess, J. L., Yu, B. D., Li, B., Hanson, R., & Korsmeyer, S. J.Defects in yolk sac hematopoiesis in Mll-null embryos. Blood, 1997; 90: 1799–1806.Google ScholarPubMed
Rowley, J. D.The critical role of chromosome translocations in human leukemias. Annu Rev Genet, 1998; 32: 495–519.CrossRefGoogle ScholarPubMed
Morrissey, J., Tkachuk, D. C., Milatovich, A., et al.A serine/ proline-rich protein is fused to HRX in t(4;11) acute leukemias. Blood, 1993; 81: 1124–31.Google Scholar
Rubnitz, J. E., Morrissey, J., Savage, P. A., & Cleary, M. L.ENL, the gene fused with HRX in t(11;19) leukemias, encodes a nuclear protein with transcriptional activation potential in lymphoid and myeloid cells. Blood, 1994; 84: 1747–52.Google Scholar
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: 853–61.CrossRefGoogle ScholarPubMed
Lavau, C., Szilvassy, S. J., Slany, R., & Cleary, M. L.Immortalization and leukemic transformation of a myelomonocytic precursor by retrovirally transduced HRX-ENL. EMBO J, 1997; 16: 4426–37.CrossRefGoogle ScholarPubMed
Forster, A., Pannell, R., Drynan, L. F., et al.Engineering de novo reciprocal chromosomal translocations associated with Mll to replicate primary events of human cancer. Cancer Cell, 2003; 3: 449–58.CrossRefGoogle ScholarPubMed
DiMartino, J. F., Miller, T., Ayton, P. M., et al.A carboxy-terminal domain of ELL is required and sufficient for immortalization of myeloid progenitors by MLL-ELL. Blood, 2000; 96: 3887–93.Google ScholarPubMed
Lavau, C., Du, C., Thirman, M., & Zeleznik-Le, N.Chromatin-related properties of CBP fused to MLL generate a myelodysplastic-like syndrome that evolves into myeloid leukemia. EMBO J, 2000; 19: 4655–64.CrossRefGoogle ScholarPubMed
DiMartino, J. F., Ayton, P. M., Chen, E. H., et al.The AF10 leucine zipper is required for leukemic transformation of myeloid progenitors by MLL-AF10. Blood, 2002; 99: 3780–5.CrossRefGoogle ScholarPubMed
So, C. W., Lin, M., Ayton, P. M., Chen, E. H., & Cleary, M. L.Dimerization contributes to oncogenic activation of MLL chimeras in acute leukemias. Cancer Cell, 2003; 4: 99–110.CrossRefGoogle ScholarPubMed
Dobson, C. L., Warren, A. J., Pannell, R., Forster, A., & Rabbitts, T. H.Tumorigenesis in mice with a fusion of the leukaemia oncogene mll and the bacterial lacZ gene. EMBO J, 2000; 19: 843–51.CrossRefGoogle ScholarPubMed
Martin, M. E., Milne, T. A., Bloyer, S., et al.Dimerization of MLL fusion proteins immortalizes hematopoietic cells. Cancer Cell, 2003; 4: 197–207.CrossRefGoogle ScholarPubMed
Hsu, K. & Look, A. T.Turning on a dimer: new insights into MLL chimeras. Cancer Cell, 2003; 4: 81–3.CrossRefGoogle ScholarPubMed
Thirman, M. J., Gill, H. J., Burnett, R. C., et al.Rearrangement of the MLL gene in acute lymphoblastic and acute myeloid leukemias with 11q23 chromosomal translocations. N Engl J Med, 1993; 329: 909–14.CrossRefGoogle ScholarPubMed
Downing, J. R., Head, D. R., Raimondi, S. C., 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: 330–5.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: 2912–21.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: 210–15.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: 2943–7.Google Scholar
Rappaport, H. Tumors of the hematopoietic system. In Atlas of Tumor Pathology, Section III, Fascicle 8 (Washington, DC: Armed Forces Institute of Pathology, 1966), pp. 97–161.Google Scholar
Armstrong, S. A., Staunton, J. E., Silverman, L. B., et al.MLL translocations specify a distinct gene expression profile that distinguishes a unique leukemia. Nat Genet, 2002; 30: 41–7.CrossRefGoogle ScholarPubMed
Yeoh, E. J., Ross, M. E., Shurtleff, S. A., et al.Classification, subtype discovery, and prediction of outcome in pediatric acute lymphoblastic leukemia by gene expression profiling. Cancer Cell, 2002; 1: 133–43.CrossRefGoogle ScholarPubMed
Ferrando, A. A., Herblot, S., Palomero, T., et al.Biallelic transcription activation of oncogenic transcription factors in T-cell acute lymphoblastic leukemia. Blood, 2004; 103: 1909–11.CrossRefGoogle ScholarPubMed
Ayton, P. M. & Cleary, M. L.Transformation of myeloid progenitors by MLL oncoproteins is dependent on Hoxa7 and Hoxa9. Genes Dev, 2003; 17: 2298–307.CrossRefGoogle ScholarPubMed
Armstrong, S. A., Kung, A. L., Mabon, M. E., et al.Inhibition of FLT3 in MLL. Validation of a therapeutic target identified by gene expression based classification. Cancer Cell, 2003; 3: 173–83.CrossRefGoogle ScholarPubMed
Romana, S. P., 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: 3662–70.Google Scholar
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. Blood, 1997; 90: 571–7.Google ScholarPubMed
Liang, D.-C., Chou, T.-B., Chen, J.-S., 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: 991–3.Google Scholar
McLean, T. W., 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: 4252–8.Google ScholarPubMed
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: 1463–70.Google ScholarPubMed
Romana, S. P., Poirel, H., Leconiat, M., et al.High frequency of t(12;21) in childhood B-lineage acute lymphoblastic leukemia. Blood, 1995; 86: 4263–9.Google Scholar
Rubnitz, J. E., Behm, F. G., Pui, C. H., 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: 1201–6.CrossRefGoogle ScholarPubMed
Rubnitz, J. E., Downing, J. R., Pui, C.-H., et al.TEL gene rearrangement in acute lymphoblastic leukemia: a new genetic marker with prognostic significance. J Clin Oncol, 1997; 15: 1150–7.CrossRefGoogle ScholarPubMed
Rubnitz, J. E., Shuster, J. J., Land, V. J., 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: 1143–6.Google ScholarPubMed
Raimondi, S. C., Privitera, E., Williams, D. L., et al.New recurring chromosomal translocations in childhood acute lymphoblastic leukemia. Blood, 1991; 77: 2016–22.Google ScholarPubMed
Golub, T. R., Barker, G. F., Lovett, M., & Gilliland, D. G.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: 307–16.CrossRefGoogle Scholar
Golub, T. R., Barker, G. F., Stegmaier, K., & Gilliland, D. G.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., Deursen, J., Hiebert, S. W., Grosveld, G., & Downing, J. R.AML1, the target of multiple chromosomal translocations in human leukemia, is essential for normal fetal liver hematopoiesis. Cell, 1996; 84: 321–30.CrossRefGoogle ScholarPubMed
Song, W. J., Sullivan, M. G., Legare, R. D., et al.Haploinsufficiency of CBFA2 causes familial thrombocytopenia with propensity to develop acute myelogenous leukaemia. Nat Genet, 1999; 23: 166–75.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, S. W., Sun, W., Davis, J. N., et al.The t(12;21) translocation converts AML-1B from an activator to a repressor of transcription. Mol Cell Biol, 1996; 16: 1349–55.CrossRefGoogle Scholar
Jousset, C., Carron, C., Boureux, A., et al.A domain of TEL conserved in a subset of ETS proteins defines a specific oligomerizaton interface essential to the mitogenic properties of the TEL-PDGFRβ oncoprotein. EMBO J, 1997; 16: 69–82.CrossRefGoogle Scholar
Guidez, F., Petrie, K., Ford, A. M., 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: 2557–61.Google ScholarPubMed
Wang, L. C., 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: 4374–83.CrossRefGoogle ScholarPubMed
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: 1459–64.CrossRefGoogle Scholar
Takeuchi, S., Seriu, T., Bartram, C. R., 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: 1220–3.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 U S A, 1996; 93: 3444–9.CrossRefGoogle ScholarPubMed
Sasaki, K., Yagi, H., Bronson, R. T., et al.Absence of fetal liver hematopoiesis in mice deficient in transcriptional coactivator core binding factor beta. Proc Natl Acad Sci U S A, 1996; 93: 12359–63.CrossRefGoogle ScholarPubMed
Wang, Q., Stacy, T., Miller, J. D., et al.The CBFβ subunit is essential for CBFα2 (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 U S A, 1997; 94: 5697–702.CrossRefGoogle ScholarPubMed
Dalla-Favera, R., Bregni, M., Erikson, J., et al.Human c-myc onc gene is located on the region of chromosome 8 that is translocated in Burkitt lymphoma cells. Proc Natl Acad Sci U S A, 1982; 79: 7824–7.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 U S A, 1982; 79: 7837–41.CrossRefGoogle ScholarPubMed
Adams, J. M., 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 Burkitt lymphomas. Proc Natl Acad Sci U S A, 1983; 80: 1982–6.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 U S A, 1983; 80: 7581–5.CrossRefGoogle ScholarPubMed
Croce, C. M., 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 U S A, 1983; 80: 6922–6.CrossRefGoogle Scholar
Emanuel, B. S., Selden, J. R., Chaganti, R. S. K., et al.The 2p breakpoint of a 2;8 translocation in Burkitt lymphoma interrups the V kappa locus. Proc Natl Acad Sci U S A, 1984; 81: 2444.CrossRefGoogle ScholarPubMed
Hollis, G. F., Mitchell, K. F., 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: 752–5.CrossRefGoogle ScholarPubMed
Rappold, G. A., 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: 2951–5.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: 511–20.CrossRefGoogle Scholar
Amati, B., Dalton, S., Brooks, M. W., et al.Transcriptional activation by the human c-Myc oncoprotein in yeast requires interaction with Max. Nature, 1992; 359: 423–6.CrossRefGoogle ScholarPubMed
Kato, G. J., Lee, W. M., Chen, L. L., & Dang, C. V.Max: functional domains and interaction with c-Myc. Genes Dev, 1992; 6: 81–92.CrossRefGoogle ScholarPubMed
Amati, B., Brooks, M. W., Levy, N., et al.Oncogenic activity of the c-Myc protein requires dimerization with Max. Cell, 1993; 72: 233–45.CrossRefGoogle ScholarPubMed
Blackwell, T. K., Huang, J., Ma, A., et al.Binding of myc proteins to canonical and noncanonical DNA sequences. Mol Cell Biol, 1993; 13: 5216–24.CrossRefGoogle ScholarPubMed
Grandori, C., Mac, J., Siebelt, F., Ayer, D. E., & Eisenman, R. N.Myc-Max heterodimers activate a DEAD box gene and interact with multiple E box-related sites in vivo. EMBO J, 1996; 15: 4344–57.Google Scholar
Ayer, D. E., Kretzner, L., & Eisenman, R. N.Mad: a heterodimeric partner for Max that antagonizes Myc transcriptional activity. Cell, 1993; 72: 211–22.CrossRefGoogle ScholarPubMed
Foley, K. P. & Eisenman, R. N.Two MAD tails: what the recent knockouts of Mad1 and Mxi1 tell us about the MYC/MAX/MAD network. Biochim Biophys Acta, 1999; 1423: M37–47.Google ScholarPubMed
Hurlin, P. J., Queva, C., & Eisenman, R. N.Mnt: a novel Max-interacting protein and Myc antagonist. Curr Top Microbiol Immunol, 1997; 224: 115–21.Google ScholarPubMed
Schreiber-Agus, N., Chin, L., Chen, K., et al.An amino-terminal domain of Mxi1 mediates anti-Myc oncogenic activity and interacts with a homolog of the yeast transcriptional repressor SIN3. Cell, 1995; 80: 777–86.CrossRefGoogle ScholarPubMed
Ayer, D. E., Lawrence, Q. A., & Eisenman, R. N.Mad-Max transcriptional repression is mediated by ternary complex formation with mammalian homologs of yeast repressor Sin3. Cell, 1995; 80: 767–76.CrossRefGoogle ScholarPubMed
Alland, L., Muhle, R., Hou, H. J., et al.Role for N-CoR and histone deacetylase in Sin3-mediated transcriptional repression. Nature, 1997; 387: 49–55.CrossRefGoogle ScholarPubMed
Wolffe, A. P.Transcriptional control. Sinful repression. Nature, 1997; 387: 16–17.CrossRefGoogle ScholarPubMed
Laherty, C. D., Yang, W. M., Sun, J. M., et al.Histone deacetylases associated with the mSin3 corepressor mediate mad transcriptional repression. Cell, 1997; 89: 349–56.CrossRefGoogle ScholarPubMed
Hassig, C. A., Fleischer, T. C., Billin, A. N., Schreiber, S. L., & Ayer, D. E.Histone deacetylase activity is required for full transcriptional repression by mSin3A. Cell, 1997; 89: 341–7.CrossRefGoogle ScholarPubMed
Grignani, F., De Matteis, S., Nervi, C., et al.Fusion proteins of the retinoic acid receptor-alpha recruit histone deacetylase in promyelocytic leukaemia. Nature, 1998; 391: 815–18.CrossRefGoogle ScholarPubMed
Lin, R. J., Nagy, L., Inoue, S., et al.Role of the histone deacetylase complex in acute promyelocytic leukaemia. Nature, 1998; 391: 811–14.CrossRefGoogle ScholarPubMed
Guidez, F., Ivins, S., Zhu, J., et al.Reduced retinoic acid-sensitivities of nuclear receptor corepressor binding to PML- and PLZF-RARα underlie molecular pathogenesis and treatment of acute promyelocytic leukemia. Blood, 1998; 91: 2634–42.Google ScholarPubMed
Collins, S. J.Acute promyelocytic leukemia: relieving repression induces remission. Blood, 1998; 91: 2631–3.Google ScholarPubMed
Coller, H. A., 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 U S A, 2000; 97: 3260–5.CrossRefGoogle ScholarPubMed
Mateyak, M. K., Obaya, A. J., & Sedivy, J. M.c-Myc regulates cyclin D-Cdk4 and -Cdk6 activity but affects cell cycle progression at multiple independent points. Mol Cell Biol, 1999; 19: 4672–83.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: 2561–76.CrossRefGoogle ScholarPubMed
Galaktionov, K., Chen, X., & Beach, D.Cdc25 cell-cycle phosphatase as a target of c-myc. Nature, 1996; 382: 511–17.CrossRefGoogle ScholarPubMed
Zindy, F., Eischen, C. M., Randle, D. H., et al.MYC signaling via the ARF tumor suppressor regulates p53-dependent apoptosis and immortalization. Genes Dev, 1998; 12: 2424–33.CrossRefGoogle ScholarPubMed
Wang, J., Xie, L. Y., Allan, S., Beach, D., & Hannon, G. J.Myc activates telomerase. Genes Dev, 1998; 12: 1769–74.CrossRefGoogle ScholarPubMed
Wu, K. J., Grandori, C., Amacker, M., et al.Direct activation of TERT transcription by c-MYC. Nat Genet, 1999; 21: 220–4.CrossRefGoogle ScholarPubMed
Bello-Fernandez, C., Packham, G., & Cleveland, J. L.The ornithine decarboxylase gene is a transcriptional target of c-Myc. Proc Natl Acad Sci U S A, 1993; 90: 7804–8.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: 2264–73.CrossRefGoogle ScholarPubMed
Miltenberger, R. J., Sukow, K. A., & Farnham, P. J.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: 2527–35.CrossRefGoogle ScholarPubMed
Boyd, K. E. & Farnham, P. J.Myc versus USF: discrimination at the cad gene is determined by core promoter elements. Mol Cell Biol, 1997; 17: 2529–37.CrossRefGoogle ScholarPubMed
Pusch, O., Soucek, T., Hengstschlager-Ottnad, E., Bernaschek, G., & Hengstschlager, M.Cellular targets for activation by c-Myc include the DNA metabolism enzyme thymidine kinase. DNA Cell Biol, 1997; 16: 737–47.Google ScholarPubMed
Boyd, K. E., Wells, J., Gutman, J., Bartley, S. M., & Farnham, P. J.c-Myc target gene specificity is determined by a post-DNA binding mechanism. Proc Natl Acad Sci U S A, 1998; 95: 13887–92.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: 3797–802.CrossRefGoogle Scholar
Rosenwald, I. B., Rhoads, D. B., Callanan, L. D., Isselbacher, K. J., & Schmidt, E. V.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 U S A, 1993; 90: 6175–8.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 U S A, 1996; 93: 7143–8.CrossRefGoogle Scholar
Jones, R. M., Branda, J., Johnston, K. A., 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: 4754–64.CrossRefGoogle Scholar
McKeithan, T. W., Shima, E. A., Le Beau, M. M., 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 U S A, 1986; 83: 6636–40.CrossRefGoogle Scholar
Finger, L. R., Harvey, R. C., Moore, R. C., Showe, L. C., & Croce, C. M.A common mechanism of chromosomal translocation in T- and B-cell neoplasia. Science, 1986; 234: 982–5.CrossRefGoogle Scholar
Shima, E. A., Le Beau, M. M., Mckeithan, T. W., 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 U S A, 1986; 83: 3439–43.CrossRefGoogle Scholar
Chen, Q., Cheng, J. T., Tasi, L. H., et al.The tal gene undergoes chromosome translocation in T cell leukemia and potentially encodes a helix-loop-helix protein. EMBO J, 1990; 9: 415–24.Google Scholar
Xia, Y., Brown, L., Yang, C. Y., 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 U S A, 1991; 88: 11416–20.CrossRefGoogle ScholarPubMed
Mellentin, J. D., Smith, S. D., & Cleary, M. L.Lyl-l, 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 ScholarPubMed
Wang, J., Jani-Sait, S. N., Escalon, E. A., 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 U S A, 2000; 97: 3497–502.CrossRefGoogle Scholar
McGuire, E. A., Hockett, R. D., Pollock, K. M., et al.The t(11;14)(p15;q11) in a T-cell acute lymphoblastic leukemia cell line activates multiple transcripts, including tg-1, a gene encoding a potential zinc finger protein. Mol Cell Biol, 1989; 9: 2124–32.CrossRefGoogle Scholar
Boehm, T., Foroni, L., Kaneko, Y., Perutz, M. F., & Rabbitts, T. H.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 U S A, 1991; 88: 4367–71.CrossRefGoogle ScholarPubMed
Royer-Pokora, B., Loos, U., & Ludwig, W. D.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: 1887–93.Google Scholar
Hatano, M., Roberts, C. W., Minden, M., Crist, W. M., & Korsmeyer, S. J.Deregulation of a homeobox gene, HOX11, by the t(10;14) in T cell leukemia. Science, 1991; 253: 79–82.CrossRefGoogle Scholar
Kennedy, M. A., Gonzalez-Sarmiento, R., Kees, U. R., et al.HOX11, a homeobox-containing T-cell oncogene on human chromosome 10q24. Proc Natl Acad Sci U S A, 1991; 88: 8900–4.CrossRefGoogle ScholarPubMed
Bernard, O. A., 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: 1495–504.CrossRefGoogle Scholar
Brown, L., Cheng, J. T., 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: 3343–51.Google ScholarPubMed
Aplan, P. D., Lombardi, D. P., & Kirsch, I. R.Structural characterization of SIL, a gene frequently disrupted in T-cell acute lymphoblastic leukemia. Mol Cell Biol, 1991; 11: 5462–9.CrossRefGoogle ScholarPubMed
Aplan, P. D., Lombardi, D. P., Reaman, G. H., et al.Involvement of the putative hematopoietic transcription factor SCL in T-cell acute lymphoblastic leukemia. Blood, 1992; 79: 1327–33.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: 1477–88.Google Scholar
Breit, T. M., Mol, E. J., Wolvers-Tettero, I. L., 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: 965–77.CrossRefGoogle Scholar
Baer, R.TAL1, TAL2, and LYL1: a family of basic helix-loop-helix proteins implicated in T cell acute leukaemia. Sem Cancer Biol, 1993; 4: 341–7.Google ScholarPubMed
Bash, R. O., Hall, S., Timmons, C. F., 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: 666–76.Google Scholar
Hsu, H.-L., Cheng, J.-T., Chen, Q., & Baer, R.Enhancer-binding activity of the tal-1 oncoprotein in association with the E47/E12 helix-loop-helix proteins. Mol Cell Biol, 1991; 11: 3037–42.CrossRefGoogle ScholarPubMed
Miyamoto, A., Cui, X., Naumovski, L., & Cleary, M. L.Helix-loop-helix proteins LYL1 and E2a form heterodimeric complexes with distinctive DNA-binding properties in hematolymphoid cells. Mol Cell Biol, 1996; 16: 2394–401.CrossRefGoogle ScholarPubMed
Shivdasani, R. A., Mayer, E. L., & Orkin, S. H.Absence of blood formation in mice lacking the T-cell leukaemia oncoprotein tal-1/SCL. Nature, 1995; 373: 432–4.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 U S A, 1995; 92: 7075–9.CrossRefGoogle ScholarPubMed
Begley, C. G., Aplan, P. D., Denning, S. M., et al.The gene SCL is expressed during early hematopoiesis and encodes a differentiation-related DNA-binding motif. Proc Natl Acad Sci U S A, 1989; 86: 10128–32.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, H. L., Wadman, I., Tsan, J. T., & Baer, R.Positive and negative transcriptional control by the TAL1 helix-loop-helix protein. Proc Natl Acad Sci U S A, 1994; 91: 5947–51.CrossRefGoogle ScholarPubMed
Park, S. T. & Sun, X. H.The Tal1 oncoprotein inhibits E47-mediated transcription. Mechanism of inhibition. J Biol Chem, 1998; 273: 7030–7.CrossRefGoogle ScholarPubMed
Begley, C. G. & Green, A. R.The SCL gene: from case report to critical hematopoietic regulator. Blood, 1999; 93: 2760–70.Google ScholarPubMed
Bain, G., Engel, I., Maandag, E. C., et al.E2A deficiency leads to abnormalities in αβ T-cell development and to rapid development of T-cell lymphomas. Mol Cell Biol, 1997; 17: 4782–891.CrossRefGoogle ScholarPubMed
Yan, W., Young, A. Z., Soares, V. C., et al.High incidence of T-cell tumors in E2A-null mice and E2A/Id1 double-knockout mice. Mol Cell Biol, 1997; 17: 7317–27.CrossRefGoogle ScholarPubMed
Aplan, P. D., Jones, C. A., Chervinsky, D. S., 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: 2406–19.CrossRefGoogle ScholarPubMed
Greenberg, J. M., Boehm, T., Sofroniew, M. V., et al.Segmental and developmental regulation of a presumptive T-cell oncogene in the central nervous system. Nature, 1990; 344: 158–60.CrossRefGoogle ScholarPubMed
Valge-Archer, V. E., Osada, H., Warren, A. J., 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 U S A, 1994; 91: 8617–21.CrossRefGoogle Scholar
Wadman, I., Li, J., Bash, R. O., et al.Specific in vivo association between the bHLH and LIM proteins implicated in human T cell leukemia. EMBO J, 1994; 13: 4831–9.Google ScholarPubMed
Larson, R. C., Lavenir, I., Larson, T. A., et al.Protein dimerization between Lmo2 (Rbtn2) and Tal1 alters thymocyte development and potentiates T cell tumorigenesis in transgenic mice. EMBO J, 1996; 15: 1021–7.Google ScholarPubMed
McGuire, E. A., Rintoul, C. E., Sclar, G. M., & Korsmeyer, S. J.Thymic overexpression of Ttg-1 in transgenic mice results in T-cell acute lymphoblastic leukemia/lymphoma. Mol Cell Biol, 1992; 12: 4186–96.CrossRefGoogle ScholarPubMed
Larson, R. C., Fisch, P., Larson, T. A., et al.T cell tumours of disparate phenotype in mice transgenic for Rbtn-2. Oncogene, 1994; 9: 3675–81.Google ScholarPubMed
Larson, R. C., Osada, H., Larson, T. A., Lavenir, I., & Rabbitts, T. H.The oncogenic LIM protein Rbtn2 causes thymic developmental aberrations that precede malignancy in transgenic mice. Oncogene, 1995; 11: 853–62.Google ScholarPubMed
Neale, G. A., Rehg, J. E., & Goorha, R. M.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: 3060–71.Google ScholarPubMed
Chervinsky, D. S., Zhao, X.-F., Lam, D. U., 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: 5025–35.CrossRefGoogle ScholarPubMed
Dear, T. N., Sanchez-Garcia, I., & Rabbitts, T. H.The HOX11 gene encodes a DNA-binding nuclear transcription factor belonging to a distinct family of homeobox genes. Proc Natl Acad Sci, 1993; 90: 4431–5.CrossRefGoogle ScholarPubMed
Allen, J. D., Lints, T., Jenkins, N. A., et al.Novel murine homeo box gene on chromosome 1 expressed in specific hematopoietic lineages and during embryogenesis. Genes Dev, 1991; 5: 509–20.CrossRefGoogle ScholarPubMed
McGinnis, W. & Krumlauf, R.Homeobox genes and axial patterning. Cell, 1992; 68: 283–302.CrossRefGoogle ScholarPubMed
Roberts, C. W. M., Shutter, J. R., & Korsmeyer, S. J.Hox11 controls the genesis of the spleen. Nature, 1994; 368: 747–9.CrossRefGoogle ScholarPubMed
Dear, T. N., Colledge, W. H., Carlton, M. B., et al.The Hox11 gene is essential for cell survival during spleen development. Development, 1995; 121: 2909–15.Google ScholarPubMed
Shirasawa, S., Arata, A., Onimaru, H., et al.Rnx deficiency results in congenital central hypoventilation. Nat Genet, 2000; 24: 287–90.CrossRefGoogle ScholarPubMed
Ferrando, A. A., Neuberg, D. S., , 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
Grimaldi, J. C. & Meeker, T. C.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: 2081–5.Google Scholar
Meeker, T. C., Hardy, D., Willman, C., Hogan, T., & Abrams, J.Activation of the interleukin-3 gene by chromosome translocation in acute lymphocytic leukemia with eosinophilia. Blood, 1990; 76: 285–9.Google ScholarPubMed
Ellisen, L. W., Bird, J., West, D. C., et al.TAN-1, the human homolog of the Drosophila notch gene, is broken by chromosomal translocations in T lymphoblastic neoplasms. Cell, 1991; 66: 649–61.CrossRefGoogle ScholarPubMed
Weng, A. P., Ferrando, A. A., Lee, W., et al.Activating mutations of NOTCH1 in human T cell acute lymphoblastic leukemia. Science, 2004; 306: 269–71.CrossRefGoogle ScholarPubMed
Tycko, B., Smith, S. D., & Sklar, J.Chromosomal translocations joining LCK and TCRB loci in human T cell leukemia. J Exp Med, 1991; 174: 867–73.CrossRefGoogle ScholarPubMed
Knudson, A. G. Jr.Mutation and cancer: statistical study of retinoblastoma. Proc Natl Acad Sci U SA, 1971; 68: 820–3.CrossRefGoogle ScholarPubMed
Weinberg, R. A.The retinoblastoma protein and cell cycle control. Cell, 1995; 81: 323–30.CrossRefGoogle ScholarPubMed
Matlashewski, G., Lamb, P., Pim, D., et al.Isolation and characterization of a human p53 cDNA clone: expression of the human p53 gene. EMBO J, 1984; 3: 3257–62.Google ScholarPubMed
Zakut-Houri, R., Bienz-Tadmor, B., Givol, D., & Oren, M.Human p53 cellular tumor antigen: cDNA sequence and expression in COS cells. EMBO J, 1985; 4: 1251–5.Google ScholarPubMed
Baker, S. J., Fearon, E. R., Nigro, J. M., et al.Chromosome 17 deletions and p53 gene mutations in colorectal carcinomas. Science, 1989; 244: 217–21.CrossRefGoogle ScholarPubMed
Baker, S. J., Markowitz, S., Fearon, E. R., Willson, J. K. V., & Vogelstein, B.Suppression of human colorectal carcinoma cell growth by wild-type p53. Science, 1990; 249: 912–15.CrossRefGoogle ScholarPubMed
Levine, A. J.p53, the cellular gatekeeper for growth and division. Cell, 1997; 88: 323–31.CrossRefGoogle ScholarPubMed
King-Underwood, L., Renshaw, J., & Pritchard-Jones, K.Mutations in the Wilms' tumor gene WT1 in leukemias. Blood, 1996; 87: 2171–9.Google ScholarPubMed
Haber, D. A. & Buckler, A. J.WT1: a novel tumor suppressor gene inactivated in Wilms' tumor. New Biol, 1992; 4: 97–106.Google ScholarPubMed
Kamb, A., Gruis, N. A., Weaver-Feldhaus, J., et al.A cell cycle regulator potentially involved in genesis of many tumor types. Science, 1994; 264: 436–40.CrossRefGoogle ScholarPubMed
Nobori, T., Miura, K., Wu, D. J., et al.Deletions of the cyclin-dependent kinase-4 inhibitor gene in multiple human cancers. Nature, 1994; 368: 753–6.CrossRefGoogle ScholarPubMed
Funk, W. D., Pak, D. T., Karas, R. H., Wright, W. E., & Shay, J. W.A transcriptionally active DNA-binding site for human p53 protein complexes. Mol Cell Biol, 1992; 12: 2866–71.CrossRefGoogle ScholarPubMed
El-Deiry, W. S., Kern, S. E., Pietenpol, J. A., Kinzler, K. W., & Vogelstein, B.Definition of a consensus binding site for p53. Nat Genet, 1992; 1: 45–9.CrossRefGoogle ScholarPubMed
Farmer, G. E., Bargonetti, J., Zhu, H., et al.Wild-type p53 activates transcription in vitro. Nature, 1992; 358: 83–6.CrossRefGoogle ScholarPubMed
Fields, S. & Jang, S. K.Presence of a potent transcription activating sequence in the p53 protein. Science, 1990; 249: 1046–9.CrossRefGoogle ScholarPubMed
Kern, S. E., Kinzler, K. W., Bruskin, A., et al.Identification of p53 as a sequence-specific DNA-binding protein. Science, 1991; 252: 1708–11.CrossRefGoogle ScholarPubMed
Kastan, M. B., Onyekwere, O., Sidransky, D., Vogelstein, B., & Craig, R. W.Participation of p53 protein in the cellular response to DNA damage. Cancer Res, 1991; 51: 6304–11.Google ScholarPubMed
Oren, M.Decision making by p53: life, death and cancer. Cell Death Differ, 2003; 10: 431–42.CrossRefGoogle ScholarPubMed
Donehower, L. A., Harvey, M., Slagle, B. L., et al.Mice deficient in p53 are developmentally normal but susceptible to spontaneous tumours. Nature, 1992; 356: 215–21.CrossRefGoogle ScholarPubMed
Li, F. P., Fraumeni, J. F. Jr., Mulvihill, J. J., et al.A cancer family syndrome in twenty-four kindreds. Cancer Res, 1988; 48: 5358–62.Google ScholarPubMed
Malkin, D., Li, F. P., Strong, L. C., et al.Germ line p53 mutations in a familial syndrome of breast cancer, sarcomas, and other neoplasms. Science, 1990; 250: 1233–8.CrossRefGoogle Scholar
Srivastava, S., Zou, Z., Pirollo, K., Blattner, W., & Chang, E. H.Germ-line transmission of a mutated p53 gene in a cancer-prone family with Li-Fraumeni syndrome. Nature, 1990; 348: 747–9.CrossRefGoogle Scholar
Frebourg, T. & Friend, S. H.Cancer risks from germline p53 mutations. J Clin Invest, 1992; 90: 1637–41.CrossRefGoogle ScholarPubMed
Varley, J. M., Evans, D. G., & Birch, J. M.Li-Fraumeni syndrome – a molecular and clinical review. Br J Cancer, 1997; 76: 1–14.CrossRefGoogle ScholarPubMed
Krug, U., Ganser, A., & Koeffler, H. P.Tumor suppressor genes in normal and malignant hematopoiesis. Oncogene, 2002; 21: 3475–95.CrossRefGoogle ScholarPubMed
Hsiao, M. H., Yu, A. L., Yeargin, J., Ku, D., & Haas, M.Nonhereditary p53 mutations in T-cell acute lymphoblastic leukemia are associated with the relapse phase. Blood, 1994; 83: 2922–30.Google ScholarPubMed
Diccianni, M. B., Yu, J., Hsiao, M., Mukherjee, S., Shao, L. E., & Yu, A. L.Clinical significance of p53 mutations in relapsed T-cell acute lymphoblastic leukemia. Blood, 1994; 84: 3105–12.Google ScholarPubMed
Wada, M., Bartram, C. R., Nakamura, H., et al.Analysis of p53 mutations in a large series of lymphoid hematologic malignancies of childhood. Blood, 1993; 82: 3163–9.Google Scholar
Kawamura, M., Kikuchi, A., Kobayashi, S., et al.Mutations of the p53 and ras genes in childhood t(1;19)-acute lymphoblastic leukemia. Blood, 1995; 85: 2546–52.Google Scholar
Hangaishi, A., Ogawa, S., Imamura, N., et al.Inactivation of multiple tumor-suppressor genes involved in negative regulation of the cell cycle, MTS1/p16INK4A/CDKN2, MTS2/p15INK4B, p53, and Rb genes in primary lymphoid malignancies. Blood, 1996; 87: 4949–58.Google ScholarPubMed
Kawamura, M., Ohnishi, H., Guo, S. X., et al.Alterations of the p53, p21, p16, p15 and RAS genes in childhood T-cell acute lymphoblastic leukemia. Leuk Res, 1999; 23: 115–26.CrossRefGoogle ScholarPubMed
Marks, D. I., Kurz, B. W., Link, M. P., et al.High incidence of potential p53 inactivation in poor outcome childhood acute lymphoblastic leukemia at diagnosis. Blood, 1996; 87: 1155–61.Google ScholarPubMed
Marks, D. I., Kurz, B. W., Link, M. P., et al.Altered expression of p53 and mdm-2 proteins at diagnosis is associated with early treatment failure in childhood acute lymphoblastic leukemia. J Clin Oncol, 1997; 15: 1158–62.CrossRefGoogle ScholarPubMed
Sherr, C. J. & Roberts, J. M.Inhibitors of mammalian G1 cyclin-dependent kinases. Genes Dev, 1995; 9: 1149–63.CrossRefGoogle ScholarPubMed
Kamijo, T., Zindy, F., Roussel, M. F., et al.Tumor suppression at the mouse INK4a locus mediated by the alternative reading frame product p19ARF. Cell, 1997; 91: 649–59.CrossRefGoogle ScholarPubMed
Quelle, D. E., Zindy, F., Ashmun, R. A., & Sherr, C. J.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
Sidransky, D.Two tracks but one race ? Cancer genetics. Curr Biol, 1996; 6: 523–5.CrossRefGoogle ScholarPubMed
Kamb, A., Gruis, N. A., Weaver-Feldhaus, J., et al.A cell cycle regulator potentially involved in genesis of many tumor types. Science, 1994; 264: 436–40.CrossRefGoogle ScholarPubMed
Zhang, Y., Xiong, Y., & Yarbrough, W. G.ARF promotes MDM2 degradation and stabilizes p53: ARF-INK4a locus deletion impairs both the Rb and p53 tumor suppression pathways. Cell, 1998; 92: 725–34.CrossRefGoogle ScholarPubMed
Pomerantz, J., Schreiber-Agus, N., Liegeois, N. J., et al.The Ink4a tumor suppressor gene product, p19Arf, interacts with MDM2 and neutralizes MDM2's inhibition of p53. Cell, 1998; 92: 713–23.CrossRefGoogle ScholarPubMed
Ashcroft, M. & Vousden, K. H.Regulation of p53 stability. Oncogene, 1999; 18: 7637–43.CrossRefGoogle ScholarPubMed
Serrano, M., Lee, H. W., 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, K. C., Mooi, W. J., Loonstra, A., & Berns, A.Loss of p16Ink4a confers susceptibility to metastatic melanoma in mice. Nature, 2001; 413: 83–6.CrossRefGoogle ScholarPubMed
Sharpless, N. E., Bardeesy, N., Lee, K. H., et al.Loss of p16Ink4a with retention of the p19Arf predisposes mice to tumorigenesis. Nature, 2001; 413: 86–91.CrossRefGoogle ScholarPubMed
Hebert, J., Cayuela, J. M., Berkeley, J., & Sigaux, F.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: 4038–44.Google Scholar
Quesnel, B., Preudhomme, C., Philippe, N., et al.p16 gene homozygous deletions in acute lymphoblastic leukemia. Blood, 1995; 85: 657–63.Google ScholarPubMed
Haidar, M. A., Cao, X. B., Manshouri, T., et al.p16INK4A and p15INK4B gene deletions in primary leukemias. Blood, 1995; 86: 311–15.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: 2685–90.Google ScholarPubMed
Rasool, O., Heyman, M., Brandter, L. B., et al.p15ink4B and p16ink4 gene inactivation in acute lymphocytic leukemia. Blood, 1995; 85: 3431–6.Google ScholarPubMed
Okuda, T., Shurtleff, S. A., Valentine, M. B., et al.Frequent deletion of p16INK4a/MTS1 and p15INK4b/MTS2 in pediatric acute lymphoblastic leukemia. Blood, 1995; 85: 2321–30.Google ScholarPubMed
Hirama, T. & Koeffler, H. P.Role of the cyclin-dependent kinase inhibitors in the development of cancer. Blood, 1995; 86: 841–54.Google Scholar
Iolascon, A., Faienza, M. F., 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: 255–60.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: 249–54.Google ScholarPubMed
Cayuela, J. M., Madani, A., Sanhes, L., Stern, M. H., & Sigaux, F.Multiple tumor-suppressor gene 1 inactivation is the most frequent genetic alteration in T-cell acute lymphoblastic leukemia. Blood, 1996; 87: 2180–6.Google ScholarPubMed
Takeuchi, S., Bartram, C. R., 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: 755–60.Google ScholarPubMed
Heyman, M., Rasool, O., Borgonovo, B. L., et al.Prognostic importance of p15INK4B and p16INK4 gene inactivation in childhood acute lymphocytic leukemia. J Clin Oncol, 1996; 14: 1512–20.CrossRefGoogle ScholarPubMed
Drexler, H. G.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: 845–59.CrossRefGoogle ScholarPubMed
Tsujimoto, Y., Finger, L. R., Yunis, J., Nowell, P. C., & Croce, C. M.Cloning of the chromosome breakpoint of neoplastic B cells with the t(14;18) chromosome translocation. Science, 1984; 226: 1097–9.CrossRefGoogle Scholar
Tsujimoto, Y., Gorham, J., Cossman, J., Jaffe, E., & Croce, C. M.The t(14;18) chromosome translocations involved in B-cell neoplasms result from mistakes in VDJ joining. Science, 1985; 229: 1390–3.CrossRefGoogle Scholar
Bakhshi, A., Jensen, J. P., Goldman, P., et al.Cloning the chromosomal breakpoint of t(14;18) human lymphomas: clustering around JH on chromosome 14 and near a transcriptional unit on 18. Cell, 1985; 41: 899–906.CrossRefGoogle Scholar
Cleary, M. L. & Sklar, J.Nucleotide sequence of a t(14;18) chromosomal breakpoint in follicular lymphoma and demonstration of a breakpoint-cluster region near a transcriptionally active locus on chromosome 18. Proc Natl Acad Sci U S A, 1985; 82: 7439–43.CrossRefGoogle Scholar
Cleary, M. L., Smith, S. D., & Sklar, J.Cloning and structural analysis of cDNAs for bcl-2 and a hybrid bcl- 2/immunoglobulin transcript resulting from the t(14;18) translocation. Cell, 1986; 47: 19–28.CrossRefGoogle Scholar
Miyashita, T. & Reed, J. C.Bcl-2 oncoprotein blocks chemotherapy-induced apoptosis in a human leukemia cell line. Blood, 1993; 81: 151–7.Google Scholar
Naumovski, L. & Cleary, M. L.Bcl2 inhibits apoptosis associated with terminal differentiation of HL-60 myeloid leukemia cells. Blood, 1994; 83: 2261–7.Google ScholarPubMed
Yang, E. & Korsmeyer, S. J.Molecular thanatopsis: a discourse on the BCL2 family and cell death. Blood, 1996; 88: 386–401.Google ScholarPubMed
Vaux, D. L., Cory, S., & Adams, J. M.Bcl-2 gene promotes haemopoietic cell survival and cooperates with c-myc to immortalize pre-B cells. Nature, 1988; 335: 440–2.CrossRefGoogle ScholarPubMed
Oltvai, Z. N., Milliman, C. L., & Korsmeyer, S. J.Bcl-2 heterodimerizes in vivo with a conserved homolog, Bax, that accelerates programmed cell death. Cell, 1993; 74: 609–19.CrossRefGoogle ScholarPubMed
Kroemer, G.The proto-oncogene Bcl-2 and its role in regulating apoptosis. Nat Med, 1997; 3: 614–20.CrossRefGoogle ScholarPubMed
Coustan-Smith, E., Kitanaka, A., Pui, C. H., et al.Clinical relevance of BCL-2 overexpression in childhood acute lymphoblastic leukemia. Blood, 1996; 87: 1140–6.Google ScholarPubMed
Findley, H. W., Gu, L., Yeager, A. M., & Zhou, M.Expression and regulation of Bcl-2, Bcl-xl, and Bax correlate with p53 status and sensitivity to apoptosis in childhood acute lymphoblastic leukemia. Blood, 1997; 89: 2986–93.Google ScholarPubMed
Meijerink, J. P. P., Mensink, E. J., Wang, K., et al.Hematopoietic malignancies demonstrate loss-of-function mutations of BAX. Blood, 1998; 91: 2991–7.Google ScholarPubMed
Williams, D. L., Harber, J., Murphy, S. B., et al.Chromosomal translocation plays a unique role in influencing prognosis in childhood acute lymphoblastic leukemia. Blood, 1986; 68: 205–16.Google Scholar
Rubin, C. M., Le Beau, M. M., Mick, R., et al.Impact of chromosomal translocations on prognosis in childhood acute lymphoblastic leukemia. J Clin Oncol, 1991; 9: 2183–92.CrossRefGoogle ScholarPubMed
Fletcher, J. A., Kimball, V. M., Lynch, E., et al.Prognostic implications of cytogenetic studies in an intensively treated group of children with acute lymphoblastic leukemia. Blood, 1989; 74: 2130–5.Google Scholar
Trueworthy, R., Shuster, J., Look, T., et al.Ploidy of lymphoblasts is the strongest predictor of treatment outcome in B-progenitor cell acute lymphoblastic leukemia of childhood: a Pediatric Oncology Group study. J Clin Oncol, 1992; 10: 606–13.CrossRefGoogle ScholarPubMed
Look, A. T., Roberson, P. K., Williams, D. L., et al.Prognostic importance of blast cell DNA content in childhood acute lymphoblastic leukemia. Blood, 1979; 65: 1079–86.Google Scholar
Williams, D. L., Tsiatis, A., Brodeur, G. M., et al.Prognostic importance of chromosome number in 136 untreated children with acute lymphoblastic leukemia. Blood, 1982; 60: 864–71.Google ScholarPubMed
Harris, M. B., Shuster, J. J., Carroll, A., et al.Trisomy of leukemic cell chromosomes 4 and 10 identifies children with B-progenitor cell acute lymphoblastic leukemia with a very low risk of treatment failure: a Pediatric Oncology Group study. Blood, 1992; 79: 3316–24.Google ScholarPubMed
Pui, C. H. & Crist, W. M.Biology and treatment of acute lymphoblastic leukemia. J Pediatr, 1994; 124: 491–503.CrossRefGoogle ScholarPubMed
Smith, M., Arthur, D., Camitta, B., et al.Uniform approach to risk classification and treatment assignment for children with acute lymphoblastic leukemia. J Clin Oncol, 1996; 14: 18–24.CrossRefGoogle ScholarPubMed
Campana, D. & Pui, C. H.Detection of minimal residual disease in acute leukemia: methodologic advances and clinical significance. Blood, 1995; 85: 1416–34.Google ScholarPubMed
Moppett, J., Burke, G. A., Steward, C. G., Oakhill, A., & Goulden, N. J.The clinical relevance of detection of minimal residual disease in childhood acute lymphoblastic leukaemia. J Clin Pathol, 2003; 56: 249–53.CrossRefGoogle 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.

  • Molecular genetics of acute lymphoblastic leukemia
    • By Adolfo A. Ferrando, Assistant Professor of Pathology and Pediatrics Institute for Cancer Genetics, Columbia University, Irving Cancer Research Center, New York, NY, USA, Jeffrey E. Rubnitz, Associate Member, Department of Hematology/Oncology, Director of Fellowship Program, St. Jude Children's Research Hospital, Memphis, TN, USA, A. Thomas Look, Professor of Pediatrics, Harvard Medical School, Vice Chair for Research, Department of Pediatric Oncology, Dana-Farber Cancer Institute, Boston, MA, USA
  • Edited by Ching-Hon Pui
  • Book: Childhood Leukemias
  • Online publication: 01 July 2010
  • Chapter DOI: https://doi.org/10.1017/CBO9780511471001.011
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.

  • Molecular genetics of acute lymphoblastic leukemia
    • By Adolfo A. Ferrando, Assistant Professor of Pathology and Pediatrics Institute for Cancer Genetics, Columbia University, Irving Cancer Research Center, New York, NY, USA, Jeffrey E. Rubnitz, Associate Member, Department of Hematology/Oncology, Director of Fellowship Program, St. Jude Children's Research Hospital, Memphis, TN, USA, A. Thomas Look, Professor of Pediatrics, Harvard Medical School, Vice Chair for Research, Department of Pediatric Oncology, Dana-Farber Cancer Institute, Boston, MA, USA
  • Edited by Ching-Hon Pui
  • Book: Childhood Leukemias
  • Online publication: 01 July 2010
  • Chapter DOI: https://doi.org/10.1017/CBO9780511471001.011
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.

  • Molecular genetics of acute lymphoblastic leukemia
    • By Adolfo A. Ferrando, Assistant Professor of Pathology and Pediatrics Institute for Cancer Genetics, Columbia University, Irving Cancer Research Center, New York, NY, USA, Jeffrey E. Rubnitz, Associate Member, Department of Hematology/Oncology, Director of Fellowship Program, St. Jude Children's Research Hospital, Memphis, TN, USA, A. Thomas Look, Professor of Pediatrics, Harvard Medical School, Vice Chair for Research, Department of Pediatric Oncology, Dana-Farber Cancer Institute, Boston, MA, USA
  • Edited by Ching-Hon Pui
  • Book: Childhood Leukemias
  • Online publication: 01 July 2010
  • Chapter DOI: https://doi.org/10.1017/CBO9780511471001.011
Available formats
×