Skip to main content Accessibility help
×
Hostname: page-component-7c8c6479df-995ml Total loading time: 0 Render date: 2024-03-29T08:02:51.312Z Has data issue: false hasContentIssue false

4 - Nuclear Factors That Regulate Erythropoiesis

from SECTION ONE - THE MOLECULAR, CELLULAR, AND GENETIC BASIS OF HEMOGLOBIN DISORDERS

Published online by Cambridge University Press:  03 May 2010

Martin H. Steinberg
Affiliation:
Boston University
Bernard G. Forget
Affiliation:
Yale University, Connecticut
Douglas R. Higgs
Affiliation:
MRC Institute of Molecular Medicine, University of Oxford
David J. Weatherall
Affiliation:
Albert Einstein College of Medicine, New York
Get access

Summary

INTRODUCTION

Studies of erythroid transcription factors originate from efforts to identify and characterize the numerous tissue-specific and ubiquitous proteins that bind cis-regulatory motifs within the globin gene loci (Chapters 3 and 5). In addition to elucidating mechanisms of globin gene regulation and erythroid development, this approach has led to the discovery of nuclear proteins that function in a wide range of developmental processes. Experimental approaches and insights gained through studies of the globin loci have broad implications for understanding how transcription factors regulate the expression of individual genes and work together to coordinate cellular differentiation.

Erythrocyte formation in the vertebrate embryo occurs in several distinct waves (see also Chapter 1). The first erythrocytes, termed primitive (EryP), arise in the extraembryonic yolk sac at mouse embryonic day 7.5 (E7.5) and weeks 3–4 in the human embryo. Later, erythropoiesis shifts to the fetal liver where adult-type (EryD, definitive) erythrocytes are produced. Finally, at birth, blood formation shifts to the bone marrow, and also the spleen in mice. EryPs and EryDs are distinguished by their unique cellular morphology, cytokine responsiveness, transcription factor requirements, and patterns of gene expression. Most notably, the expression of individual globin genes is developmentally regulated (Chapter 3). Understanding how transcription factors regulate the temporal control of β-like globin genes during mammalian development is of general interest to the study of gene regulation in higher eukaryotes and could eventually lead to new approaches to reactivate the human fetal γ-globin genes in patients with β chain hemoglobinopathies, such as sickle cell anemia and β thalassemia.

Type
Chapter
Information
Disorders of Hemoglobin
Genetics, Pathophysiology, and Clinical Management
, pp. 62 - 85
Publisher: Cambridge University Press
Print publication year: 2009

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

Brotherton, TW, Chui, DHK, Gauldie, J, Patterson, M. Hemoglobin ontogeny during normal mouse fetal development. Proc Natl Acad Sci USA. 1979;76:2853–2857.CrossRefGoogle ScholarPubMed
Wood, WG. Erythropoiesis and haemoglobin production during development. In: Jones, CT, ed. Biochemical Development of the Fetus and Neonate. New York: Elsevier Biomedical Press; 1982:127–162.Google Scholar
Mucenski, ML, McLain, K, Kier, AB, et al. A functional c-myb gene is required for normal fetal hematopoiesis. Cell. 1991;65:677–689.CrossRefGoogle Scholar
Ogawa, M, Nishikawa, S, Yoshinaga, K, et al. Expression and function of c-Kit in fetal hemopoietic progenitor cells: transition from the early c-Kit-independent to the late c-Kit-dependent wave of hemopoiesis in the murine embryo. Development. 1993;117:1089–1098.Google ScholarPubMed
Wu, H, Liu, X, Jaenisch, R, Lodish, HF. Generation of committed erythroid BFU-E and CFU-E progenitors does not require erythropoietin or the erythropoietin receptor. Cell. 1995;83:59–67.CrossRefGoogle ScholarPubMed
Chyuan-Sheng, L, Lim, S-K, Agati, V, Costantini, F. Differential effects of an erythropoietin receptor gene disruption on primitive and definitive erythropoiesis. Genes Dev. 1996;10:154–164.Google Scholar
Okuda, T, Deursen, J, Hiebert, SW, Grosveld, G, Downing, JR. AML 1, the target of multiple chromosomal translocations in human leukemia, is essential for normal fetal liver hematopoiesis. Cell. 1996;84:321–330.CrossRefGoogle ScholarPubMed
Wang, Q, Stacy, T, Binder, M, Marin-Padilla, M, Sharpe, AH, Speck, N. 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
Kingsley, PD, Malik, J, Fantauzzo, KA, Palis, J. Yolk sac-derived primitive erythroblasts enucleate during mammalian embryogenesis. Blood. 2004;104(1):19–25.CrossRefGoogle ScholarPubMed
Kingsley, PD, Malik, J, Emerson, RL. “Maturational” globin switching in primary primitive erythroid cells. Blood. 2006;107(4):1665–1672.CrossRefGoogle ScholarPubMed
Tuan, D, Solomon, W, Li, Q, London, IM. The “beta-like-globin” gene domain in human erythroid cells. Proc Natl Acad Sci USA. 1985;82(19):6384–6388.CrossRefGoogle ScholarPubMed
Forrester, WC, Takegawa, S, Papayannopoulou, T, Stamatoyannopoulos, G, Groudine, M. Evidence for a locus activation region: the formation of developmentally stable hypersensitive sites in globin-expressing hybrids. Nucl Acids Res. 1987;15(24):10159–10177.CrossRefGoogle ScholarPubMed
Grosveld, F, Assendelft, GB, Greaves, DR, Kollias, G. Position-independent, high-level expression of the human beta-globin gene in transgenic mice. Cell. 1987;51(6):975–985.CrossRefGoogle ScholarPubMed
Tuan, DY, Solomon, WB, London, IM, Lee, DP. An erythroid-specific, developmental-stage-independent enhancer far upstream of the human “beta-like globin” genes. Proc Natl Acad Sci USA. 1989;86(8):2554–2558.CrossRefGoogle ScholarPubMed
Mitchell, PJ, Tjian, R. Transcriptional regulation in mammalian cells by sequence-specific DNA binding proteins. Science. 1989;245(4916):371–378.CrossRefGoogle ScholarPubMed
Ptashne, M, Gann, A. Transcriptional activation by recruitment. Nature. 1997;386(6625):569–577.CrossRefGoogle ScholarPubMed
Kadonaga, JT. Eukaryotic transcription: an interlaced network of transcription factors and chromatin-modifying machines. Cell. 1998;92(3):307–313.CrossRefGoogle ScholarPubMed
Wang, X, Crispino, JD, Letting, DL, Nakazawa, M, Poncz, M, Blobel, GA. Control of megakaryocyte-specific gene expression by GATA-1 and FOG-1: role of Ets transcription factors. EMBO J. 2002;21(19):5225–5234.CrossRefGoogle ScholarPubMed
Pang, L, Xue, HH, Szalai, G, et al. Maturation stage-specific regulation of megakaryopoiesis by pointed-domain Ets proteins. Blood. 2006;108(7):2198–2206.CrossRefGoogle ScholarPubMed
Drissen, R, Palstra, RJ, Gillemans, N, et al. The active spatial organization of the beta-globin locus requires the transcription factor EKLF. Genes Dev. 2004;18(20):2485–2490.CrossRefGoogle ScholarPubMed
Vakoc, CR, Letting, DL, Gheldof, N, et al. Proximity among distant regulatory elements at the beta-globin locus requires GATA-1 and FOG-1. Mol Cell. 2005;17(3):453–462.CrossRefGoogle ScholarPubMed
Friend, C, Scher, W, Holland, JG, Sato, T. Hemoglobin synthesis in murine virus-induced leukemic cells in vitro: stimulation of erythroid differentiation by dimethyl sulfoxide. Proc Natl Acad Sci USA. 1971;68(2):378–382.CrossRefGoogle ScholarPubMed
Friend, C, Scher, W, Holland, JG, Sato, T. Hemoglobin synthesis in murine virus-induced leukemic cells in vitro: stimulation of erythroid differentiation by dimethyl sulfoxide. Proc Natl Acad Sci USA. 1971;68:378–382.CrossRefGoogle ScholarPubMed
Rutherford, TR, Clegg, JB, Weatherall, DJ. K562 human leukaemic cells synthesise embryonic haemoglobin in response to haemin. Nature. 1979;280(5718):164–165.CrossRefGoogle ScholarPubMed
Weiss, MJ, Yu, C, Orkin, SH. Erythroid-cell-specific properties of transcription factor GATA-1 revealed by phenotypic rescue of a gene targeted cell line. Mol Cell Biol. 1997;17:1642–1651.CrossRefGoogle ScholarPubMed
Beug, H, Doederlein, G, Freudenstein, C, Graf, T. Erythroblast cell lines transformed by a temperature-sensitive mutant of avian erythroblastosis virus: a model system to study erythroid differentiation in vitro. J Cell Physiol. 1982;Suppl 1:195–207.CrossRefGoogle ScholarPubMed
Metz, T, Graf, T. v-myb and v-ets transform chicken erythroid cells and cooperate both in trans and in cis to induce distinct differentiation phenotypes. Genes Dev. 1991;5(3):369–380.CrossRefGoogle ScholarPubMed
Fibach, E, Manor, D, Oppenheim, A, Rachmilewitz, EA. Proliferation and maturation of human erythroid progenitors in liquid culture. Blood. 1989;73(1):100–103.Google ScholarPubMed
Lindern, M, Zauner, W, Mellitzer, G, et al. The glucocorticoid receptor cooperates with the erythropoietin receptor and c-Kit to enhance and sustain proliferation of erythroid progenitors in vitro. Blood. 1999;94(2):550–559.Google Scholar
Pope, SH, Fibach, E, Sun, J, Chin, K, Rodgers, GP. Two-phase liquid culture system models normal human adult erythropoiesis at the molecular level. Eur J Haematol. 2000;64(5):292–303.CrossRefGoogle ScholarPubMed
Lindern, M, Deiner, EM, Dolznig, H, et al. Leukemic transformation of normal murine erythroid progenitors: v- and c-ErbB act through signaling pathways activated by the EpoR and c-Kit in stress erythropoiesis. Oncogene. 2001;20(28):3651–3664.CrossRefGoogle Scholar
Wojda, U, Leigh, KR, Njoroge, JM, et al. Fetal hemoglobin modulation during human erythropoiesis: stem cell factor has “late” effects related to the expression pattern of CD117. Blood. 2003;101(2):492–497.CrossRefGoogle ScholarPubMed
Keller, G, Kennedy, M, Papayannopoulou, T, Wiles, MV. Hematopoietic differentiation during embryonic stem cell differentiation in culture. Mol Cell Biol. 1993;13(1):472–486.CrossRefGoogle ScholarPubMed
Carotta, S, Pilat, S, Mairhofer, A, et al. Directed differentiation and mass cultivation of pure erythroid progenitors from mouse embryonic stem cells. Blood. 2004;104(6):1873–1880.CrossRefGoogle ScholarPubMed
Orlando, V, Paro, R. Mapping Polycomb-repressed domains in the bithorax complex using in vivo formaldehyde cross-linked chromatin. Cell. 1993;75(6):1187–1198.CrossRefGoogle ScholarPubMed
Boyd, KE, Wells, J, Gutman, J, Bartley, SM, Farnham, PJ. c-Myc target gene specificity is determined by a post-DNA binding mechanism. Proc Natl Acad Sci USA. 1998;95(23):3887–13892.CrossRefGoogle Scholar
Metzker, ML. Emerging technologies in DNA sequencing. Genome Res. 2005;15(12):1767–1776.CrossRefGoogle ScholarPubMed
Bentley, DR. Whole-genome re-sequencing. Curr Opin Genet Dev. 2006;16(6):545–552.CrossRefGoogle ScholarPubMed
Welch, JJ, Watts, JA, Vakoc, CR, et al. Global regulation of erythroid gene expression by transcription factor GATA-1. Blood. 2004;104(10):3136–3147.CrossRefGoogle ScholarPubMed
Drissen, R, Lindern, M, Kolbus, A, et al. The erythroid phenotype of EKLF-null mice: defects in hemoglobin metabolism and membrane stability. Mol Cell Biol. 2005;25(12):5205–5214.CrossRefGoogle ScholarPubMed
Hodge, D, Coghill, E, Keys, J, et al. A global role for EKLF in definitive and primitive erythropoiesis. Blood. 2005.Google ScholarPubMed
Pilon, AM, Nilson, DG, Zhou, D, et al. Alterations in expression and chromatin configuration of the alpha hemoglobin-stabilizing protein gene in erythroid Kruppel-like factor-deficient mice. Mol Cell Biol. 2006;26(11):4368–4377.CrossRefGoogle ScholarPubMed
Gobbi, M, Anguita, E, Hughes, J, et al. Tissue-specific histone modification and transcription factor binding in α globin gene expression. Blood. 2007;110:4503–4510.CrossRefGoogle ScholarPubMed
Carter, D, Chakalova, L, Osborne, CS, Dai, YF, Fraser, P. Long–range chromatin regulatory interactions in vivo. Nat Genet. 2002;32(4):623–626.CrossRefGoogle ScholarPubMed
Tolhuis, B, Palstra, RJ, Splinter, E, Grosveld, F, Laat, W. Looping and interaction between hypersensitive sites in the active beta-globin locus. Mol Cell. 2002;10(6):1453–1465.CrossRefGoogle ScholarPubMed
Vernimmen, D, Gobbi, M, Sloane-Stanley, JA, Wood, WG, Higgs, DR. Long-range chromosomal interactions regulate the timing of the transition between poised and active gene expression. EMBO J. 2007;26(8):2041–2051.CrossRefGoogle ScholarPubMed
Kim, A, Dean, A. Developmental stage differences in chromatin subdomains of the beta-globin locus. Proc Natl Acad Sci USA. 2004;101(18):7028–7033.CrossRefGoogle ScholarPubMed
Cullen, KE, Kladde, MP, Seyfred, MA. Interaction between transcription regulatory regions of prolactin chromatin. Science. 1993;261(5118):203–206.CrossRefGoogle ScholarPubMed
Dekker, J, Rippe, K, Dekker, M, Kleckner, N. Capturing chromosome conformation. Science. 2002;295(5558):1306–1311.CrossRefGoogle ScholarPubMed
Kooren, J, Palstra, RJ, Klous, P, et al. Beta-globin active chromatin Hub formation in differentiating erythroid cells and in p45 NF-E2 knock-out mice. J Biol Chem. 2007;282(22):16544–16552.CrossRefGoogle ScholarPubMed
Patrinos, GP, Krom, M, Boer, E, et al. Multiple interactions between regulatory regions are required to stabilize an active chromatin hub. Genes Dev. 2004;18(12):1495–1509.CrossRefGoogle ScholarPubMed
Evans, MJ, Kaufman, MH. Establishment in culture of pluripotent cells from mouse embryos. Nature. 1981;292:154–156.CrossRefGoogle Scholar
Martin, GR. Isolation of a pluripotent cell line from early mouse embryos cultured in medium conditioned by teratocarcinoma cells. Proc Natl Acad Sci USA. 1981;78:7634–7638.CrossRefGoogle Scholar
Robertson, E. Pluripotential stem cell lines as a route into the mouse germ line. Trends Genet. 1986;2:9–13.CrossRefGoogle Scholar
Smithies, O, Gregg, RG, Boggs, SS, Kordewski, MA, Kucherlapati, RS. Insertion of DNA sequences into the human chromosomal beta-globin locus by homologous recombination. Nature. 1985;317:230–234.CrossRefGoogle ScholarPubMed
Capecchi, MR. Altering the genome by homologous recombination. Science. 1989;244:1288–1292.CrossRefGoogle ScholarPubMed
Gu, H, Marth, JD, Orban, PC, Mossman, H, Rajewsky, K. Deletion of a DNA polymerase ß gene segment in T cells using cell type-specific gene targeting. Science. 1994;265:103–106.CrossRefGoogle Scholar
Rossant, J, McMahon, A. “Cre”-ating mouse mutants – a meeting review on conditional mouse genetics. Genes Dev. 1999;13(2):142–145.CrossRefGoogle ScholarPubMed
Glaser, S, Anastassiadis, K, Stewart, AF. Current issues in mouse genome engineering. Nat Genet. 2005;37(11):1187–1193.CrossRefGoogle ScholarPubMed
Garcia-Otin, AL, Guillou, F. Mammalian genome targeting using site-specific recombinases. Front Biosci. 2006;11:1108–1136.CrossRefGoogle ScholarPubMed
te Riele, H, Maandag, ER, Clarke, A, Hooper, M, Berns, A. Consecutive inactivation of both alleles of the pim-1 proto-oncogene by homologous recombination in embryonic stem cells. Nature. 1990;348:649–651.CrossRefGoogle ScholarPubMed
Mortensen, RM, Conner, DA, Chao, S, Geisterfer-Lowrance, AAT, Seidman, JG. Production of homozygous mutant ES cells with a single targeting construct. Mol Cell Biol. 1992;12:2391–2395.CrossRefGoogle ScholarPubMed
Donahue, SL, Lin, Q, Cao, S, Ruley, HE. Carcinogens induce genome-wide loss of heterozygosity in normal stem cells without persistent chromosomal instability. Proc Natl Acad Sci USA. 2006;103(31):11642–11646.CrossRefGoogle ScholarPubMed
Doetschman, TC, Eistetter, H, Katz, M, Schmidt, W, Kemler, R. The in vitro development of blastocyst-derived embryonic stem cell lines: formation of visceral yolk sac, blood islands, and myocardium. J Embryol Exp Morphol. 1985;87:27–45.Google ScholarPubMed
Nakano, ,T, Kodama, H, Honjo, T. Generation of lymphohematopoietic cells from embryonic stem cells in culture. Science. 1994;265:1098–1101.CrossRefGoogle ScholarPubMed
Suwabe, N, Takahashi, S, Nakano, T, Yamamoto, M. GATA-1 regulates growth and differentiation of definitive erythroid lineage cells during in vitro ES cell differentiation. Blood. 1998;92(11):4108–4118.Google ScholarPubMed
Kaufman, DS, Hanson, ET, Lewis, RL, Auerbach, R, Thomson, JA. Hematopoietic colony-forming cells derived from human embryonic stem cells. Proc Natl Acad Sci USA. 2001;98(19):10716–10721.CrossRefGoogle ScholarPubMed
Qiu, C, Hanson, E, Olivier, E, et al. Differentiation of human embryonic stem cells into hematopoietic cells by coculture with human fetal liver cells recapitulates the globin switch that occurs early in development. Exp Hematol. 2005;33(12):1450–1458.CrossRefGoogle ScholarPubMed
Vodyanik, MA, Bork, JA, Thomson, JA, Slukvin, II. Human embryonic stem cell-derived CD34 +cells: efficient production in the coculture with OP9 stromal cells and analysis of lymphohematopoietic potential. Blood. 2005;105(2):617–626.CrossRefGoogle ScholarPubMed
Zambidis, ET, Peault, B, Park, TS, Bunz, F, Civin, CI. Hematopoietic differentiation of human embryonic stem cells progresses through sequential hematoendothelial, primitive, and definitive stages resembling human yolk sac development. Blood. 2005;106(3):860–870.CrossRefGoogle ScholarPubMed
Olivier, EN, Qiu, C, Velho, M, Hirsch, RE, Bouhassira, EE. Large-scale production of embryonic red blood cells from human embryonic stem cells. Exp Hematol. 2006;34(12):1635–1642.CrossRefGoogle ScholarPubMed
Kennedy, M, D'Souza, SL, Lynch-Kattman, M, Schwantz, S, Keller, G. Development of the hemangioblast defines the onset of hematopoiesis in human ES cell differentiation cultures. Blood. 2007;109(7):2679–2687.Google ScholarPubMed
Evans, T, Felsenfeld, G. The erythroid-specific transcription factor Eryf1: a new finger protein. Cell. 1989;58:877–885.CrossRefGoogle ScholarPubMed
Tsai, SF, Martin, DIK, Zon, LI, D'Andrea, AD, Wong, GG, Orkin, SH. Cloning of cDNA for the major DNA-binding protein of the erythroid lineage through expression in mammalian cells. Nature. 1989;339:446–451.CrossRefGoogle ScholarPubMed
Martin, DIK, Orkin, SH. Transcriptional activation and DNA binding by the erythroid factor GF-1/NF-E1/Eryf 1. Genes Dev. 1990;4:1886–1898.CrossRefGoogle ScholarPubMed
Yang, H-Y, Evans, T. Distinct roles for the two cGATA-1 finger domains. Mol Cell Biol. 1992;12:4562–4570.CrossRefGoogle ScholarPubMed
Whyatt, DJ, deBoer, E, Grosveld, F. The two zinc finger-like domains of GATA-1 have different DNA binding specifties. EMBO J. 1993;12:4993–5005.Google Scholar
Trainor, CD, Omichinski, JG, Vandergon, TL, Gronenborn, AM, Clore, GM, Felsenfeld, G. A Palindromic regulatory site within vertebrate GATA-1 promoters requires both zinc fingers of the GATA-1 DNA-binding domain for high-affinity interaction. Mol Cell Biol. 1996;16:2238–2247.CrossRefGoogle ScholarPubMed
Cantor, AB, Orkin, SH. Transcriptional regulation of erythropoiesis: an affair involving multiple partners. Oncogene. 2002;21(21):3368–3376.CrossRefGoogle ScholarPubMed
Ferreira, R, Ohneda, K, Yamamoto, M, Philipsen, S. GATA1 function, a paradigm for transcription factors in hematopoiesis. Mol Cell Biol. 2005;25(4):1215–1227.CrossRefGoogle ScholarPubMed
Pevny, L, Simon, MC, Robertson, E, et al. Erythroid differentiation in chimaeric mice blocked by a targeted mutation in the gene for transcription factor GATA-1. Nature. 1991;349:257–260.CrossRefGoogle ScholarPubMed
Simon, MC, Pevny, L, Wiles, M, Keller, G, Costantini, F, Orkin, SH. Rescue of erythroid development in gene targeted GATA-1-mouse embryonic stem cells. Nat Genet. 1992;1:92–98.CrossRefGoogle ScholarPubMed
Fujiwara, Y, Browne, CP, Cunniff, K, Goff, SC, Orkin, SH. Arrested development of embryonic red cell precursors in mouse embryos lacking transcription factor GATA-1. Proc Natl Acad Sci USA. 1996;93(22):12355–12358.CrossRefGoogle ScholarPubMed
Weiss, MJ, Keller, G, Orkin, SH. Novel insights into erythroid development revealed through in vitro differentiation of GATA-1 embryonic stem cells. Genes Dev. 1994;8:1184–1197.CrossRefGoogle ScholarPubMed
Weiss, MJ, Orkin, SH. Transcription factor GATA-1 permits survival and maturation of erythroid precursors by preventing apoptosis. Proc Natl Acad Sci USA. 1995;92:9623–9627.CrossRefGoogle ScholarPubMed
Pevny, L, Chyuan-Sheng, L, D'Agati, V, Simon, MC, Orkin, SH, Costantini, F. Development of hematopoietic cells lacking transcription factor GATA-1. Development. 1994;121:163–172.Google Scholar
Shivdasani, RA, Fujiwara, Y, McDevitt, MA, Orkin, SH. A lineage-selective knockout establishes the critical role of transcription factor GATA-1 in megakaryocyte growth and platelet development. EMBO J. 1997;16(13):3965–3973.CrossRefGoogle ScholarPubMed
Vyas, P, Ault, K, Jackson, CW, Orkin, S.H, Shivdasani, RA. Consequences of GATA-1 deficiency in megakaryocytes and platelets. Blood. 1999;93(9):2867–2875.Google ScholarPubMed
Yu, C, Cantor, AB, Yang, H, et al. Targeted deletion of a high-affinity GATA-binding site in the GATA-1 promoter leads to selective loss of the eosinophil lineage in vivo. J Exp Med. 2002;195(11):1387–1395.CrossRefGoogle ScholarPubMed
Migliaccio, AR, Rana, RA, Sanchez, M, et al. GATA-1 as a regulator of mast cell differentiation revealed by the phenotype of the GATA-1 low mouse mutant. J Exp Med. 2003;197(3):281–296.CrossRefGoogle ScholarPubMed
Gutierrez, L, Nikolic, T, Dijk, TB, et al. Gata1 regulates dendritic-cell development and survival. Blood. 2007;110(6):1933–1941.CrossRefGoogle Scholar
Yamamoto, M, Ko, LJ, Leonard, MW, Beug, H, Orkin, S, Engel, JD. Activity and tissue-specific expression of the transcription factor NF-E1 mutligene family. Genes Dev. 1990;4:1650–1662.CrossRefGoogle ScholarPubMed
Zon, LI, Mather, C, Burgess, S, Bolce, ME, Harland, RM, Orkin, SH. Expression of GATA-binding proteins during embryonic development in Xenopus laevis. Proc Natl Acad Sci USA. 1991;88:10642–10646.CrossRefGoogle ScholarPubMed
Arceci, RJ, King, AAJ, Simon, MC, Orkin, SH, Wilson, DB. Mouse GATA-4: a retinoic acid-inducible GATA-binding transcription factor expressed in endodermally derived tissues and heart. Mol Cell Biol. 1993;13(4):2235–2246.CrossRefGoogle ScholarPubMed
Kelley, C, Blumberg, H, Zon, LI, Evans, T. GATA-4 is a novel transcription factor expressed in endocardium of the developing heart. Development. 1993;118(3):817–827.Google ScholarPubMed
Laverriere, AC, MacNeill, C, Mueller, C, Poelman, RE, Burch, JBE, Evans, T. GATA-4/5/6, a subfamily of three transcription factors transcribed in developing heart and gut. J Biol Chem. 1994;269:23177–23184.Google ScholarPubMed
Detrich, HW, Kieran, MW, Chan, FY, et al. Intraembryonic hematopoietic cell migration during vertebrate development. Proc Natl Acad Sci USA. 1995;92(23):10713–10717.CrossRefGoogle ScholarPubMed
Jiang, Y, Evans, T. The Xenopus GATA-4/5/6 genes are associated with cardiac specification and can regulate cardiac-specific transcription during embryogenesis. Dev Biol. 1996;174(2):258–270.CrossRefGoogle ScholarPubMed
Orkin, SH. GATA-binding transcription factors in hematopoietic cells. Blood. 1992;80(3):575–581.Google ScholarPubMed
Weiss, MJ, Orkin, SH. GATA transcription factors: Key regulators of hematopoiesis. Exp Hematol. 1995;23:99–107.Google ScholarPubMed
Molkentin, JD. The zinc finger-containing transcription factors GATA-4, -5, and -6. Ubiquitously expressed regulators of tissue-specific gene expression. J Biol Chem. 2000;275(50):38949–38952.CrossRefGoogle ScholarPubMed
Patient, RK, McGhee, JD. The GATA family (vertebrates and invertebrates). Curr Opin Genet Dev. 2002;12(4):416–422.CrossRefGoogle Scholar
Sposi, NM, Zon, LI, Care, A, et al. Cycle-dependent initiation and lineage-dependent abrogation of GATA-1 expression in pure differentiating hematopoietic progenitors. Proc Natl Acad Sci USA. 1992;89:6353–6357.CrossRefGoogle ScholarPubMed
Leonard, M, Brice, M, Engel, JD, Papayannopoulou, T. Dynamics of GATA transcription factor expression during erythroid differentiation. Blood. 1993;82(4):1071–1079.Google ScholarPubMed
Tsai, F-Y, Keller, G, Kuo, FC, et al. An early hematopoietic defect in mice lacking the transcription factor GATA-2. Nature. 1994;371:221–226.CrossRefGoogle ScholarPubMed
Anguita, E, Hughes, J, Heyworth, C, Blobel, GA, Wood, WG, Higgs, DR. Globin gene activation during haemopoiesis is driven by protein complexes nucleated by GATA-1 and GATA-2. EMBO J. 2004;23(14):2841–2852.CrossRefGoogle ScholarPubMed
Grass, JA, Boyer, ME, Pal, S, Wu, J, Weiss, MJ, Bresnick, EH. GATA-1-dependent transcriptional repression of GATA-2 via disruption of positive autoregulation and domain-wide chromatin remodeling. Proc Natl Acad Sci USA. 2003;100(15):8811–8816.CrossRefGoogle ScholarPubMed
Raich, N, Clegg, CH, Grofti, J, Romeo, PH, Stamatoyannopoulos, G. GATA1 and YY1 are developmental repressors of the human epsilon-globin gene. EMBO J. 1995;14(4):801–809.Google ScholarPubMed
Li, J, Noguchi, CT, Miller, W, Hardison, R, Schechter, AN. Multiple regulatory elements in the 5′-flanking sequence of the human epsilon-globin gene. J Biol Chem. 1998;273(17):10202–10209.CrossRefGoogle ScholarPubMed
Berry, M, Grosveld, F, Dillon, N. A single point mutation is the cause of the Greek form of hereditary persistence of fetal haemoglobin. Nature. 1992;358(6386):499–502.CrossRefGoogle ScholarPubMed
Bartunek, P, Kralova, J, Blendinger, G, Dvorak, M, Zenke, M. GATA-1 and c-myb crosstalk during red blood cell differentiation through GATA-1 binding sites in the c-myb promoter. Oncogene. 2003;22(13):1927–1935.CrossRefGoogle ScholarPubMed
Rylski, M, Welch, JJ, Chen, YY, et al. GATA-1-mediated proliferation arrest during erythroid maturation. Mol Cell Biol. 2003;23(14):5031–5042.CrossRefGoogle ScholarPubMed
Munugalavadla, V, Dore, LC, Tan, BL, et al. Repression of c-kit and its downstream substrates by GATA-1 inhibits cell proliferation during erythroid maturation. Mol Cell Biol. 2005;25(15):6747–6759.CrossRefGoogle ScholarPubMed
Evans, T, Felsenfeld, G. trans-Activation of a globin promoter in nonerythroid cells. Mol Cell Biol. 1991;11:843–853.CrossRefGoogle ScholarPubMed
Visvader, JE, Crossley, M, Hill, J, Orkin, SH, Adams, JM. The C-terminal zinc finger of GATA-1 or GATA-2 is sufficient to induce megakaryocytic differentiation of an early myeloid cell line. Mol Cell Biol. 1995;15:634–641.CrossRefGoogle ScholarPubMed
Blobel, GA, Simon, MC, Orkin, SH. Rescue of GATA-1-deficient embryonic stem cells by heterologous GATA-binding proteins. Mol Cell Biol. 1995;15:626–633.CrossRefGoogle ScholarPubMed
Shimizu, R, Takahashi, S, Ohneda, K, Engel, JD, Yamamoto, M. In vivo requirements for GATA-1 functional domains during primitive and definitive erythropoiesis. EMBO J. 2001;20(18):5250–5260.CrossRefGoogle ScholarPubMed
Tsang, AP, Visvader, JE, Turner, CA, Fujiwara, Y, et al. FOG, a multitype zinc finger protein, acts as a cofactor for transcription factor GATA-1 in erythroid and megakaryocytic differentiation. Cell. 1997;90:109–119.CrossRefGoogle ScholarPubMed
Yu, C, Niakan, KK, Matsushita, M, Stamatoyannopoulos, G, Orkin, SH, Raskind, WH. X-linked thrombocytopenia with thalassemia from a mutation in the amino finger of GATA-1 affecting DNA binding rather than FOG-1 interaction. Blood. 2002;100(6):2040–2045.CrossRefGoogle ScholarPubMed
Crossley, M, Orkin, SH. Phosphorylation of the erythroid transcription factor GATA-1. J Biol Chem. 1994;269(24):16589–16596.Google ScholarPubMed
Partington, GA, Patient, RK. Phosphorylation of GATA-1 increases its DNAbinding affinity and is correlated with induction of human K562 erythroleukaemia cells. Nucl Acids Res. 1999;27(4):1168–1175.CrossRefGoogle Scholar
Ghaffari, S, Kitidis, C, Zhao, W, et al. AKT induces erythroid cell maturation of JAK2-deficient fetal liver progenitor cells and is required for epo regulation of erythroid cell differentiation. Blood. 2006;107:1888–1891.CrossRefGoogle ScholarPubMed
Kadri, Z, Maouche-Chretien, L, Rooke, HM, et al. Phosphatidylinositol 3–kinase/Akt induced by erythropoietin renders the erythroid differentiation factor GATA-1 competent for TIMP-1 gene transactivation. Mol Cell Biol. 2005;25(17):7412–7422.CrossRefGoogle ScholarPubMed
Zhao, W, Kitidis, C, Fleming, MD, Lodish, HF, Ghaffari, S. Erythropoietin stimulates phosphorylation and activation of GATA-1 via the PI3-kinase-AKT signaling pathway. Blood. 2005;107:907–915.CrossRefGoogle ScholarPubMed
Rooke, HM, Orkin, SH. Phosphorylation of Gata1 at serine residues 72, 142, and 310 is not essential for hematopoiesis in vivo. Blood. 2006;107(9):3527–3530.CrossRefGoogle Scholar
Boyes, J, Byfield, P, Nakatani, Y, Ogryzko, V. Regulation of activity of the transcription factor GATA-1 by acetylation. Nature. 1998;396(6711):594–598.CrossRefGoogle ScholarPubMed
Hung, HL, Lau, J, Kim, AY, Weiss, MJ, Blobel, GA. CREB-Binding protein acetylates hematopoietic transcription factor GATA-1 at functionally important sites. Mol Cell Biol. 1999;19(5):3496–3505.CrossRefGoogle ScholarPubMed
Lamonica, JM, Vakoc, CR, Blobel, GA. Acetylation of GATA-1 is required for chromatin occupancy. Blood. 2006;108(12):3736–3738.CrossRefGoogle ScholarPubMed
Collavin, L, Gostissa, M, Avolio, F, et al. Modification of the erythroid transcription factor GATA-1 by SUMO-1. Proc Natl Acad Sci USA. 2004;101(24):8870–8875.CrossRefGoogle ScholarPubMed
Hernandez-Hernandez, A, Ray, P, Litos, G, et al. Acetylation and MAPK phosphorylation cooperate to regulate the degradation of active GATA-1. EMBO J. 2006;25(14):3264–3274.CrossRefGoogle ScholarPubMed
Fischer, K-D, Haese, A, Nowock, J. Cooperation of GATA-1 and Sp1 can result in synergistic transcriptional activation or interference. J Biol Chem. 1993;268(32):23915–23923.Google ScholarPubMed
Crossley, M, Merika, M, Orkin, SH. Self association of the erythroid transcription factor GATA-1 mediated by its zinc finger domains. Mol Cell Biol. 1995;15:2448–2456.CrossRefGoogle ScholarPubMed
Merika, M, Orkin, SH. Functional synergy and physical interactions of the erythroid transcription factor GATA-1 with the Krüppel family proteins Sp1 and EKLF. Mol Cell Biol. 1995;15:2437–2447.CrossRefGoogle ScholarPubMed
Gregory, RC, Taxman, DJ, Seshasayee, D, Kensinger, MH, Bieker, JJ, Wojchowski, DM. Functional interaction of GATA1 with erythroid Kruppel-like factor and Sp1 at defined erythroid promoters. Blood. 1996;87(5):1793–1801.Google ScholarPubMed
Tsang, AP, Fujiwara, Y, Hom, DB, Orkin, SH. Failure of megakaryopoiesis and arrested erythropoiesis in mice lacking the GATA-1 transcriptional cofactor FOG. Genes Dev. 1998;12(8):1176–1188.CrossRefGoogle ScholarPubMed
Fox, AH, Liew, C, Holmes, M, Kowalski, K, Mackay, J, Crossley, M. Transcriptional cofactors of the FOG family interact with GATA proteins by means of multiple zinc fingers. EMBO J. 1999;18(10):2812–2822.CrossRefGoogle ScholarPubMed
Turner, J, Crossley, M. Cloning and characterization of mCtBP2, a corepressor that associates with basic Kruppel-like factor and other mammalian transcriptional regulators. EMBO J. 1998;17:5129–5140.CrossRefGoogle Scholar
Hong, W, Nakazawa, M, Chen, YY, et al. FOG-1 recruits the NuRD repressor complex to mediate transcriptional repression by GATA-1. EMBO J. 2005;24(13):2367–2378.CrossRefGoogle ScholarPubMed
Rodriguez, P, Bonte, E, Krijgsveld, J, et al. GATA-1 forms distinct activating and repressive complexes in erythroid cells. EMBO J. 2005;24(13):2354–2366.CrossRefGoogle ScholarPubMed
Katz, SG, Cantor, AB, Orkin, SH. Interaction between FOG-1 and the corepressor C-terminal binding protein is dispensable for normal erythropoiesis in vivo. Mol Cell Biol. 2002;22(9):3121–3128.CrossRefGoogle ScholarPubMed
Stumpf, M, Waskow, C, Krotschel, M, et al. The mediator complex functions as a coactivator for GATA-1 in erythropoiesis via subunit Med1/TRAP220. Proc Natl Acad Sci USA. 2006;103(49):18504–18509.CrossRefGoogle ScholarPubMed
Blobel, GA, Nakajima, T, Eckner, R, Montminy, M, Orkin, SH. CREB binding protein (CBP) cooperates with transcription factor GATA-1 and is required for erythroid differentiation. Proc Natl Acad Sci USA. 1998;95:2061–2066.CrossRefGoogle ScholarPubMed
Blobel, GA. CBP/p300: molecular integrators of hematopoietic transcription. Blood. 2000;95:745–755.Google Scholar
Letting, DL, Rakowski, C, Weiss, MJ, Blobel, GA. Formation of a tissue-specific histone acetylation pattern by the hematopoietic transcription factor GATA-1. Mol Cell Biol. 2003;23(4):1334–1340.CrossRefGoogle ScholarPubMed
Kiekhaefer, CM, Grass, JA, Johnson, KD, Boyer, ME, Bresnick, EH. Hematopoietic-specific activators establish an overlapping pattern of histone acetylation and methylation within a mammalian chromatin domain. Proc Natl Acad Sci USA. 2002;99(22):14309–14314.CrossRefGoogle ScholarPubMed
Koschmieder, S, Rosenbauer, F, Steidl, U, Owens, BM, Tenen, DG. Role of transcription factors C/EBPalpha and PU.1 in normal hematopoiesis and leukemia. Int J Hematol. 2005;81(5):368–377.CrossRefGoogle ScholarPubMed
Moreau-Gachelin, F, Ray, D, Mattei, MG, Tambourin, P, Tavitian, A. The putative oncogene Spi-1: murine chromosomal localization and transcriptional activation in murine acute erythroleukemias [published erratum appears in Oncogene. 1990;5(6):941]. Oncogene. 1989;4(12):1449–1456.Google ScholarPubMed
Moreau-Gachelin, F, Tavitian, A, Tambourin, P. Spi-1 is a putative oncogene in virally induced murine erythroleukaemias. Nature. 1988;331(6153):277–280.CrossRefGoogle ScholarPubMed
Moreau-Gachelin, F, Wendling, F, Molina, T, et al. Spi-1/PU.1 transgenic mice develop multistep erythroleukemias. Mol Cell Biol. 1996;16(5):2453–2463.CrossRefGoogle ScholarPubMed
Quang, CT, Pironin, M, Lindern, M, Beug, H, Ghysdael, J. Spi-1 and mutant p53 regulate different aspects of the proliferation and differentiation control of primary erythroid progenitors. Oncogene. 1995;11(7):1229–1239.Google ScholarPubMed
Rao, G, Rekhtman, N, Cheng, G, Krasikov, T, Skoultchi, AI. Deregulated expression of the PU.1 transcription factor blocks murine erythroleukemia cell terminal differentiation. Oncogene. 1997;14(1):123–131.CrossRefGoogle ScholarPubMed
Yamada, T, Kondoh, N, Matsumoto, M, Yoshida, M, Maekawa, A, Oikawa, T. Overexpression of PU.1 induces growth and differentiation inhibition and apoptotic cell death in murine erythroleukemia cells. Blood. 1997;89(4):1383–1393.Google ScholarPubMed
Delgado, MD, Gutierrez, P, Richard, C, Cuadrado, MA, Moreau-Gachelin, F, Leon, J. Spi-1/PU.1 proto-oncogene induces opposite effects on monocytic and erythroid differentiation of K562 cells. Biochem Biophys Res Commun. 1998;252(2):383–391.CrossRefGoogle ScholarPubMed
Yamada, T, Kihara-Negishi, F, Yamamoto, H, Yamamoto, M, Hashimoto, Y, Oikawa, T. Reduction of DNA binding activity of the GATA-1 transcription factor in the apoptotic process induced by overexpression of PU.1 in murine erythroleukemia cells. Exp Cell Res. 1998;245(1):186–194.CrossRefGoogle ScholarPubMed
Rekhtman, N, Radparvar, F, Evans, T, Skoultchi, AI. Direct interaction of hematopoietic transcription factors PU.1 and GATA- 1: functional antagonism in erythroid cells. Genes Dev. 1999;13(11):1398–1411.CrossRefGoogle ScholarPubMed
Rekhtman, N, Choe, KS, Matushansky, I, Murray, S, Stopka, T, Skoultchi, AI. PU.1 and pRB interact and cooperate to repress GATA-1 and block erythroid differentiation. Mol Cell Biol. 2003;23(21):7460–7474.CrossRefGoogle ScholarPubMed
Stopka, T, Amanatullah, DF, Papetti, M, Skoultchi, AI. PU.1 inhibits the erythroid program by binding to GATA-1on DNA and creating a repressive chromatin structure. EMBO J. 2005;24(21):3712–3723.CrossRefGoogle ScholarPubMed
Hong, W, Kim, AY, Ky, S, et al. Inhibition of CBP-mediated protein acetylation by the Ets family oncoprotein PU.1. Mol Cell Biol. 2002;22(11):3729–3743.CrossRefGoogle ScholarPubMed
Zhang, P, Behre, G, Pan, J, et al. Negative cross-talk between hematopoietic regulators:GATA proteins repress PU.1. Proc Natl Acad Sci USA. 1999;96(15):8705–8710.CrossRefGoogle ScholarPubMed
Wadman, IA, Osada, H, Grutz, GG, et al. The LIM-only protein Lmo2 is a bridging molecule assembling an erythroid, DNA-binding complex which includes the TAL1, E47, GATA-1 and Ldb1/NLI proteins. EMBO J. 1997;16(11):3145–3157.CrossRefGoogle ScholarPubMed
Lecuyer, E, Herblot, S, Saint-Denis, M, et al. The SCL complex regulates c-kit expression in hematopoietic cells through functional interaction with Sp1. Blood. 2002;100(7):2430–2440.CrossRefGoogle ScholarPubMed
Nichols, K, Crispino, JD, Poncz, M, et al. Familial dyserythropoietic anemia and thrombocytopenia due to an inherited mutation in GATA1. Nat Genet. 2000;24:266–270.CrossRefGoogle Scholar
Freson, K, Devriendt, K, Matthijs, G, et al.Platelet characteristics in patients with X-linked macrothrombocytopenia because of a novel GATA1 mutation. Blood. 2001;98(1):85–92.CrossRefGoogle ScholarPubMed
Mehaffey, MG, Newton, AL, Gandhi, MJ, Crossley, M, Drachman, JG. X-linked thrombocytopenia caused by a novel mutation of GATA-1. Blood. 2001;98(9):2681–2688.CrossRefGoogle ScholarPubMed
Freson, K, Matthijs, G, Thys, C, et al. Different substitutions at residue D218 of the X-linked transcription factor GATA1 lead to altered clinical severity of macrothrombocytopenia and anemia and are associated with variable skewed X inactivation. Hum Mol Genet. 2002;11(2):147–152.CrossRefGoogle ScholarPubMed
Balduini, CL, Pecci, A, Loffredo, G, et al. Effects of the R216Q mutation of GATA-1 on erythropoiesis and megakaryocytopoiesis. Thromb Haemost. 2004;91(1):129–140.Google ScholarPubMed
Del Vecchio, GC, Giordani, L, Santis, A, Mattia, D. Dyserythropoietic anemia and thrombocytopenia due to a novel mutation in GATA-1. Acta Haematol. 2005;114(2):113–116.CrossRefGoogle ScholarPubMed
Phillips, JD, Steensma, DP, Pulsipher, MA, Spangrude, GJ, Kushner, JP. Congenital erythropoietic porphyria due to a mutation in GATA1: the first transacting mutation causative for a human porphyria. Blood. 2007;109(6):2618–2621.CrossRefGoogle Scholar
Wechsler, J, Greene, M, McDevitt, MA, et al. Acquired mutations in GATA1 in the megakaryoblastic leukemia of Down syndrome. Nat Genet. 2002;32(1):148–152.CrossRefGoogle ScholarPubMed
Greene, ME, Mundschau, G, Wechsler, J, et al. Mutations in GATA1 in both transient myeloproliferative disorder and acute megakaryoblastic leukemia of Down syndrome. Blood Cells Mol Dis. 2003;31(3):351–356.CrossRefGoogle ScholarPubMed
Hitzler, JK, Cheung, J, Li, Y, Scherer, SW, Zipursky, A. GATA1 mutations in transient leukemia and acute megakaryoblastic leukemia of Down syndrome. Blood. 2003;101:4301–4304.CrossRefGoogle ScholarPubMed
Mundschau, G, Gurbuxani, S, Gamis, AS, Greene, ME, Arceci, RJ, Crispino, JD. Mutagenesis of GATA1 is an initiating event in Down syndrome leukemogenesis. Blood. 2003;101(11):4298–4300.CrossRefGoogle ScholarPubMed
Rainis, L, Bercovich, D, Strehl, S, et al. Mutations in exon 2 of GATA1 are early events in megakaryocytic malignancies associated with trisomy 21. Blood. 2003;102(3):981–986.CrossRefGoogle ScholarPubMed
Xu, G, Nagano, M, Kanezaki, R, et al. Frequent mutations in the GATA-1 gene in the transient myeloproliferative disorder of Down syndrome. Blood. 2003;102(8):2960–2968.CrossRefGoogle ScholarPubMed
Taub, JW, Mundschau, G, Ge, Y, et al. Prenatal origin of GATA1 mutations may be an initiating step in the development of megakaryocytic leukemia in Down syndrome. Blood. 2004;104(5):1588–1589.CrossRefGoogle ScholarPubMed
Hollanda, LM, Lima, CS, Cunha, AF, et al.An inherited mutation leading to production of only the short isoform of GATA-1 is associated with impaired erythropoiesis. Nat Genet. 2006;38(7):807–812.CrossRefGoogle ScholarPubMed
Begley, CG, Green, AR. The SCL gene: from case report to critical hematopoietic regulator. Blood. 1999;93(9):2760–2770.Google ScholarPubMed
Lecuyer, E, Hoang, T. SCL: from the origin of hematopoiesis to stem cells and leukemia. Exp Hematol. 2004;32(1):11–24.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
Shivdasani, RA, Mayer, EL, Orkin, SH. Absence of blood formation in mice lacking the T-cell leukemia oncoprotein tal-1/SCL. Nature. 1995;373:432–434.CrossRefGoogle ScholarPubMed
Porcher, C, Swat, W, Rockwell, K, Fujiwara, Y, Alt, F, Orkin, SH. The T cell leukemia oncoprotein SCL/tal-1 is essential for development of all hematopoietic lineages. Cell. 1996;86:47–57.CrossRefGoogle Scholar
Robb, L, Elwood, NJ, Elefanty, AG, et al. The scl gene is required for the generation of all hematopoietic lineages in the adult mouse. EMBO J. 1996;15:4123–4129.Google ScholarPubMed
Hall, MA, Curtis, DJ, Metcalf, D, et al. The critical regulator of embryonic hematopoiesis, SCL, is vital in the adult for megakaryopoiesis, erythropoiesis, and lineage choice in CFU-S12. Proc Natl Acad Sci USA. 2003;100(3):992–997.CrossRefGoogle ScholarPubMed
D'Souza, SL, Elefanty, AG, Keller, G. SCL/Tal-1 is essential for hematopoietic commitment of the hemangioblast but not for its development. Blood. 2005;105(10):3862–3870.CrossRefGoogle Scholar
Dooley, KA, Davidson, AJ, Zon, LI. Zebrafish scl functions independently in hematopoietic and endothelial development. Dev Biol. 2005;277(2):522–536.CrossRefGoogle ScholarPubMed
Patterson, LJ, Gering, M, Patient, R. Scl is required for dorsal aorta as well as blood formation in zebrafish embryos. Blood. 2005;105(9):3502–3511.CrossRefGoogle ScholarPubMed
Visvader, JE, Crossley, M, Hill, J, Orkin, SH, Adams, JM. The C-terminal zinc finger of GATA-1 or GATA-2 is sufficient to induce megakaryocytic differentiation of an early myeloid cell line. Mol Cell Biol. 1995;15:634–641.CrossRefGoogle ScholarPubMed
Mikkola, HK, Klintman, J, Yang, H, et al. Haematopoietic stem cells retain long-term repopulating activity and multipotency in the absence of stem-cell leukaemia SCL/tal-1 gene. Nature. 2003;421(6922):547–551.CrossRefGoogle ScholarPubMed
Curtis, DJ, Hall, MA, Stekelenburg, LJ, Robb, L, Jane, SM, Begley, CG. SCL is required for normal function of short-term repopulating hematopoietic stem cells. Blood. 2004;103(9):3342–3348.CrossRefGoogle ScholarPubMed
Green, AR, Salvaris, E, Begley, CG. Erythroid expression of the helix-loophelix gene, SCL. Oncogene. 1991;6:475–479.Google Scholar
Visvader, J, Begley, CG, Adams, JM. Differential expression of the Lyl, SCL, E2a helix-loop-helix genes within the hemopoietic system. Oncogene. 1991;6:187–194.Google ScholarPubMed
Mouthon, M-A, Bernard, O, Mitjavila, M-T, Romeo, PH, Vainchenker, W, Mathieu-Mahul, D. Expression of tal-1 and GATA-binding proteins during human hematopoiesis. Blood. 1993;81:647–655.Google ScholarPubMed
Aplan, PD, Nakahara, K, Orkin, SHO, Kirsch, IR. The SCL gene product: a positive regulator of erythroid differentiation. EMBO J. 1992;11(11):4073–4081.Google ScholarPubMed
Hoang, T, Paradis, E, Brady, G, et al. Opposing effects of the basic helix-loop-helix transcription factor SCL on erythroid and monocytic differentiation. Blood. 1996;87(1):102–111.Google ScholarPubMed
Elwood, NJ, Zogos, H, Pereira, DS, Dick, JE, Begley, CG. Enhanced megakaryocyte and erythroid development from normal human CD34(+) cells: consequence of enforced expression of SCL. Blood. 1998;91(10):3756–3765.Google ScholarPubMed
Valtieri, M, Tocci, A, Gabbianelli, M, et al. Enforced TAL-1 expression stimulates primitive, erythroid and megakaryocytic progenitors but blocks the granulopoietic differentiation program. Cancer Res. 1998;58(3):562–569.Google ScholarPubMed
Elnitski, L, Miller, W, Hardison, R. Conserved E boxes function as part of the enhancer in hypersensitive site 2 of the beta-globin locus control region. Role of basic helix- loop-helix proteins. J Biol Chem. 1997;272(1):369–378.CrossRefGoogle Scholar
Anderson, KP, Crable, SC, Lingrel, JB. Multiple proteins binding to a GATA-E box-GATA motif regulate the erythroid Kruppel-like factor (EKLF) gene. J Biol Chem. 1998;273(23):14347–14354.CrossRefGoogle ScholarPubMed
Vyas, P, McDevitt, MA, Cantor, AB, Katz, SG, Fujiwara, Y, Orkin, SH. Different sequence requirements for expression in erythroid and megakaryocytic cells within a regulatory element upstream of the GATA-1 gene. Development. 1999;126(12):2799–2811.Google ScholarPubMed
Anderson, KP, Crable, SC, Lingrel, JB. The GATA-E box-GATA motif in the EKLF promoter is required for in vivo expression. Blood. 2000;95(5):1652–1655.Google ScholarPubMed
Xu, Z, Huang, S, Chang, LS, Agulnick, AD, Brandt, SJ. Identification of a TAL1 target gene reveals a positive role for the LIM domain-binding protein Ldb1 in erythroid gene expression and differentiation. Mol Cell Biol. 2003;23(21):7585–7599.CrossRefGoogle Scholar
Lahlil, R, Lecuyer, E, Herblot, S, Hoang, T. SCL assembles a multifactorial complex that determines glycophorin A expression. Mol Cell Biol. 2004;24(4):1439–1452.CrossRefGoogle ScholarPubMed
Cohen-Kaminsky, S, Maouche-Chretien, L, Vitelli, L, et al. Chromatin immunoselection defines a TAL-1 target gene. EMBO J. 1998;17(17):5151–5160.CrossRefGoogle ScholarPubMed
Huang, S, Qiu, Y, Stein, RW, Brandt, SJ. p300 functions as a transcriptional coactivator for the TAL1/SCL oncoprotein. Oncogene. 1999;18(35):4958–4967.CrossRefGoogle ScholarPubMed
Huang, S, Brandt, SJ. mSin3A regulates murine erythroleukemia cell differentiation through association with the TAL1 (or SCL) transcription factor. Mol Cell Biol. 2000;20(6):2248–2259.CrossRefGoogle ScholarPubMed
Huang, S, Qiu, Y, Shi, Y, Xu, Z, Brandt, SJ. P/CAF-mediated acetylation regulates the function of the basic helix- loop-helix transcription factor TAL1/SCL. EMBO J. 2000;19(24):6792–6803.CrossRefGoogle ScholarPubMed
Schuh, AH, Tipping, AJ, Clark, AJ, et al. ETO-2 associates with SCL in erythroid cells and megakaryocytes and provides repressor functions in erythropoiesis. Mol Cell Biol. 2005;25(23):10235–10250.CrossRefGoogle ScholarPubMed
Goardon, N, Lambert, JA, Rodriguez, P, et al. ETO2 coordinates cellular proliferation and differentiation during erythropoiesis. EMBO J. 2006;25(2):357–366.CrossRefGoogle ScholarPubMed
Meier, N, Krpic, S, Rodriguez, P, et al. Novel binding partners of Ldb1 are required for haematopoietic development. Development. 2006;133(24):4913–4923.CrossRefGoogle ScholarPubMed
Orkin, SH, Kazazian, HHJ, Antonarakis, SE, et al. Linkage of beta-thalassaemia mutations and beta-globin gene polymorphisms with DNA polymorphisms in human beta-globin gene cluster. Nature. 1982;296:627–631.CrossRefGoogle ScholarPubMed
Orkin, SH, Antonarakis, SE, Kazazian, HHJ. Base substitution at position -88 in a beta-thalassemic globin gene. Further evidence for the role of distal promoter element ACACCC. J Biol Chem. 1984;259:8679–8681.Google Scholar
Kulozik, AE, Bellan-Koch, A, Bail, S, Kohne, E, Kleihauer, E. Thalassemia intermedia: moderate reduction of beta globin gene transcriptional activity by a novel mutation of the proximal CACCC promoter element. Blood. 1991;77:2054–2058.Google ScholarPubMed
Cook, T, Gebelein, B, Urrutia, R. Sp1 and its likes: biochemical and functional predictions for a growing family of zinc finger transcription factors. Ann NY Acad Sci. 1999;880:94–102.CrossRefGoogle ScholarPubMed
Philipsen, S, Suske, G. Survey and summary. A tale of three fingers: the family of mammalian Sp/XKLF transcription factors. Nucl Acids Res. 1999;27:2991–3000.CrossRefGoogle Scholar
Turner, J, Crossley, M. Mammalian Kruppel-like transcription factors: more than just a pretty finger. Trends Biochem Sci. 1999;24:236–240.CrossRefGoogle ScholarPubMed
Miller, IJ, Bieker, JJ. A novel, erythroid cell-specific murine transcription factor that binds to the CACCC element and is related to the Kruppel family of nuclear proteins. Mol Cell Biol. 1993;13(5):2776–2786.CrossRefGoogle ScholarPubMed
Southwood, CM, Downs, KM, Bieker, JJ. Erythroid Kruppel-like factor exhibits an early and sequentially localized pattern of expression during mammalian erythroid ontogeny. Dev Dyn. 1996;206(3):248–259.3.0.CO;2-I>CrossRefGoogle ScholarPubMed
Feng, WC, Southwood, CM, Bieker, JJ. Analyses of beta-thalassemia mutant DNA interactions with erythroid Kruppel-like factor (EKLF), an erythroid cell-specific transcription factor. J Biol Chem. 1994;269(2):1493–1500.Google Scholar
Nuez, B, Michalovich, D, Bygrave, A, Ploemacher, R, Grosveld, F. Defective haematopoiesis in fetal liver resulting from inactivation of the EKLF gene. Nature. 1995;375:316–318.CrossRefGoogle ScholarPubMed
Perkins, AC, Sharpe, AH, Orkin, SH. Lethal b-thalassaemia in mice lacking the erythroid CACCC-transcription factor EKLF. Nature. 1995;375:318–322.CrossRefGoogle Scholar
Zhou, D, Pawlik, KM, Ren, J, Sun, CW, Townes, TM. Differential binding of erythroid Krupple-like factor to embryonic/fetal globin gene promoters during development. J Biol Chem. 2006;281(23):16052–16057.CrossRefGoogle ScholarPubMed
Shyu, YC, Wen, SC, Lee, TL, et al. Chromatin-binding in vivo of the erythroid kruppellike factor, EKLF, in the murine globin loci. Cell Res. 2006;16(4):347–355.CrossRefGoogle Scholar
Perkins, AC, Gaensler, KM, Orkin, SH. Silencing of human fetal globin expression is impaired in the absence of the adult beta-globin gene activator protein EKLF. Proc Natl Acad Sci USA. 1996;93(22):12267–12271.CrossRefGoogle ScholarPubMed
Wijgerde, M, Gribnau, J, Trimborn, T, et al. The role of EKLF in human b–globin gene competition. Genes Dev. 1996;10:2894–2902.CrossRefGoogle Scholar
Tewari, R, Gillemans, N, Wijgerde, M, et al. Erythroid Kruppel-like factor (EKLF) is active in primitive and definitive erythroid cells and is required for the function of 5′HS3 of the betaglobin locus control region. EMBO J. 1998;8:2334–2341.CrossRefGoogle Scholar
Dillon, N, Grosveld, F. Human gamma-globin genes silenced independently of other genes in the beta-globin locus. Nature. 1991;350(6315):252–254.CrossRefGoogle ScholarPubMed
Lim, SK, Bieker, JJ, Lin, CS, Costantini, F. A shortened life span of EKLF−/− adult erythrocytes, due to a deficiency of beta-globin chains, is ameliorated by human gamma-globin chains. Blood. 1997;90(3):1291–1299.Google ScholarPubMed
Gallagher, PG, Pilon, AM, Arcasoy, MO, Bodine, DM. Multiple defects in erythroid gene expression in erythroid Kruppel-like factor (EKLF) target genes in EKLF-deficient mice. Blood. 2004;104(11):446a.Google Scholar
Nilson, DG, Sabatino, , Bodine, DM, Gallagher, PG. Major erythrocyte membrane protein genes in EKLF-deficient mice. Exp Hematol. 2006;34(6):705–712.CrossRefGoogle ScholarPubMed
Keys, JR, Tallack, M.R, Hodge, DJ, Cridland, SO, David, R, Perkins, AC. Genomic organisation and regulation of murine alpha haemoglobin stabilising protein by erythroid Kruppel-like factor. Br J Haematol. 2007;136(1):150–157.CrossRefGoogle ScholarPubMed
Frontelo, P, Manwani, D, Galdass, M, et al. 2007. Novel role for EKLF in megakaryocyte lineage commitment. Blood. 2007;110:3871–3880.CrossRefGoogle ScholarPubMed
Donze, D, Townes, TM, Bieker, JJ. Role of erythroid Kruppel-like factor in human gamma- to beta-globin gene switching. J Biol Chem. 1995;270(4):1955–1959.CrossRefGoogle ScholarPubMed
Bieker, JJ, Southwood, CM. The erythroid Kruppel-like factor transactivation domain is a critical component for cell-specific inducibility of a beta-globin promoter. Mol Cell Biol. 1995;15(2):852–860.CrossRefGoogle ScholarPubMed
Ouyang, L, Chen, X, Bieker, JJ. Regulation of erythroid Kruppel-like factor (EKLF) transcriptional activity by phosphorylation of a protein kinase casein kinase II site within its interaction domain. J Biol Chem. 1998;273(36):23019–23025.CrossRefGoogle ScholarPubMed
Zhang, W, Bieker, JJ. Acetylation and modulation of erythroid Kruppel-like factor (EKLF) activity by interaction with histone acetyltransferases. Proc Natl Acad Sci USA. 1998;95(17):9855–9860.CrossRefGoogle ScholarPubMed
Zhang, W, Kadam, S, Emerson, BM, Bieker, JJ. Site-specific acetylation by p300 or CREB binding protein regulates erythroid Kruppel-like factor transcriptional activity via its interaction with the SWI-SNF complex. Mol Cell Biol. 2001;21(7):2413–2422.CrossRefGoogle ScholarPubMed
Siatecka, M, Xue, L, Bieker, JJ. Sumoylation of EKLF promotes transcriptional repression and is involved in inhibition of megakaryopoiesis. Mol Cell Biol. 2007;27(24):8547–8560.CrossRefGoogle ScholarPubMed
Armstrong, J.A, Bieker, J.J, Emerson, BM. A SWI/SNF-related chromatin remodeling complex, E-RC1, is required for tissue-specific transcriptional regulation by EKLF in vitro. Cell. 1998;95(1):93–104.CrossRefGoogle ScholarPubMed
Asano, H, Li, XS, Stamatoyannopoulos, G. FKLF, a novel Kruppel-like factor that activates human embryonic and fetal beta-like globin genes. Mol Cell Biol. 1999;19(5):3571–3579.CrossRefGoogle ScholarPubMed
Asano, H, Li, XS, Stamatoyannopoulos, G. FKLF-2: a novel Kruppel-like transcriptional factor that activates globin and other erythroid lineage genes. Blood. 2000;95(11):3578–3584.Google ScholarPubMed
Funnell, AP, Maloney, CA, Thompson, LJ, et al. Erythroid Kruppel-like factor directly activates the basic Kruppel-like factor gene in erythroid cells. Mol Cell Biol. 2007;27(7):2777–2790.CrossRefGoogle ScholarPubMed
Perkins, AC, Yang, H, Crossley, PM, Fujiwara, Y, Orkin, SH. Deficiency of the CACC-element binding protein BKLF leads to a progressive myeloproliferative disease and impaired expression of SHP-1. Blood. 1997;90(Suppl 1):575a.Google Scholar
Marin, M, Karis, A, Visser, P, Grosveld, F, Philipsen, S. Transcription factor Sp1 is essential for early embryonic development but dispensable for cell growth and differentiation. Cell. 1997;89:619–628.CrossRefGoogle ScholarPubMed
Kruger, I, Vollmer, M, Simmons, D, Elsasser, HP, Philipsen, S, Suske, G. Sp1/Sp3 compound heterozygous mice are not viable: impaired erythropoiesis and severe placental defects. Dev Dyn. 2007;236(8):2235–2244.CrossRefGoogle Scholar
Talbot, D, Grosveld, F. The 5′HS2 of the globin locus control region enhances transcription through the interaction of a multimeric complex binding at two functionally distinct NF-E2 binding sites. EMBO J. 1991;10(6):1391–1398.Google ScholarPubMed
Stamatoyannopoulos, JA, Goodwin, A, Joyce, T, Lowrey, CH. NF-E2 and GATA binding motifs are required for the formation of DNase I hypersensitive site 4 of the human b-globin locus control region. EMBO J. 1995;14:106–116.Google Scholar
Boyes, J, Felsenfeld, G. Tissue-specific factors additively increase the probability of the all-or-none formation of a hypersensitive site. EMBO J. 1996;15: 2496–2507.Google ScholarPubMed
Gong, QH, McDowell, JC, Dean, A. Essential role of NF-E2 in remodeling of chromatin structure and transcriptional activation of the epsilon-globin gene in vivo by 5′ hypersensitive site 2 of the beta-globin locus control region. Mol Cell Biol. 1996;16(11):6055–6064.CrossRefGoogle ScholarPubMed
Pomerantz, O, Goodwin, AJ, Joyce, T, Lowrey, CH. Conserved elements containing NF-E2 and tandem GATA binding sites are required for erythroid-specific chromatin structure reorganization within the human b-globin locus control region. Nucl Acid Res. 1998;26:5684–5691.CrossRefGoogle Scholar
Mignotte, V, Eleouet, JF, Raich, N, Romeo, P-H. Cis- and trans-acting elements involved in the regulation of the erythroid promoter of the human porphobilinogen deaminase gene. Proc Natl Acad Sci USA. 1989;86:6548–6552.CrossRefGoogle ScholarPubMed
Taketani, S, Inazawa, J, Nakahashi, Y, Abe, T, Tokunaga, R. Structure of the human ferrochelatase gene. Exon/intron gene organization and location of the gene to chromosome 18. Eur J Biochem. 1992;205:217–222.CrossRefGoogle ScholarPubMed
Mignotte, V, Wall, L, Boer, E, Grosveld, F, Romeo, P-H. Two tissuespecific factors bind the erythroid promoter of the human porphobilinogen deaminase gene. Nucl Acids Res. 1989;17(1):37–54.CrossRefGoogle Scholar
Andrews, NC, Erdjument-Bromage, H, Davidson, M, Tempst, P, Orkin, SH. Erythroid transcription factor NF-E2 is a haematopoietic-specific basic leucine zipper protein. Nature. 1993;362:722–728.CrossRefGoogle ScholarPubMed
Andrews, NC, Kotkow, KJ, Ney, PA, Erdjument-Bromage, H, Tempst, P, Orkin, SH. The ubiquitous subunit of erythroid transcription factor NF-E2 is a small basic-leucine zipper protein related to the v-maf oncogene. Proc Natl Acad Sci USA. 1993;90:11488–11492.CrossRefGoogle ScholarPubMed
Ney, PA, Andrews, NC, Jane, SM, et al. Purification of the human NF-E2 complex: cDNA cloning of the hematopoietic cell-specific subunit and evidence for an associated partner. Mol Cell Biol. 1993;13:5604–5612.CrossRefGoogle ScholarPubMed
Blank, V, Andrews, NC. The Maf transcription factors: regulators of differentiation. Trends Biochem Sci. 1997;22(11):437–441.CrossRefGoogle Scholar
Motohashi, H, Shavit, JA, Igarashi, K, Yamamoto, M, Engel, JD. The world according to Maf. Nucl Acids Res. 1997;25(15):2953–2959.CrossRefGoogle ScholarPubMed
Lu, SJ, Rowan, S, Bani, MR, Ben-David, Y. Retroviral integration within the Fli-2 locus results in inactivation of the erythroid transcription factor NF-E2 in Friend erythroleukemias: evidence that NF-E2 is essential for globin gene expression. Proc Natl Acad Sci USA. 1994;91:8398–8402.CrossRefGoogle Scholar
Kotkow, K, Orkin, SH. Dependence of globin gene expression in mouse erythroleukemia cells on the NF-E2 heterodimer. Mol Cell Biol. 1995;15:4640–4647.CrossRefGoogle ScholarPubMed
Shivdasani, RA, Orkin, SH. Erythropoiesis and globin gene expression in mice lacking the transcription factor NF-E2. Proc Natl Acad Sci USA. 1995;92:8690–8694.CrossRefGoogle ScholarPubMed
Shivdasani, RA, Rosenblatt, MF, Zucker-Franklin, DC, et al. Transcription factor NF-E2 is required for platelet formation independent of the actions of thrombopoieitin/MGDF in megakaryocyte development. Cell. 1995;81:695–701.CrossRefGoogle ScholarPubMed
Chan, K, Lu, R, Chang, JC, Kan, YW. NRF2, a member of the NFE2 family of transcription factors, is not essential for murine erythropoiesis, growth, and development. Proc Natl Acad Sci USA. 1996;93:13943–13948.CrossRefGoogle Scholar
Kuroha, T, Takahashi, S, Komeno, T, Itoh, K, Nagasawa, T, Yamamoto, M. Ablation of Nrf2 function does not increase the erythroid or megakaryocytic cell lineage dysfunction caused by p45 NF-E2 gene disruption. J Biochem. 1998;123:376–379.CrossRefGoogle ScholarPubMed
Martin, F, Deursen, JM, Shivdasani, RA, Jackson, CW, Troutman, AG, Ney, PA. Erythroid maturation and globin gene expression in mice with combined deficiency of NF-E2 and nrf-2. Blood. 1998;91:3459–3466.Google ScholarPubMed
Farmer, SC, Sun, CW, Winnier, GE, Hogan, BL, Townes, TM. The bZIP transcription factor LCR-F1 is essential for mesoderm formation in mouse development. Genes Dev. 1997;11:786–798.CrossRefGoogle ScholarPubMed
Chan, JY, Kwong, M, Lu, R, et al. Targeted disruption of the ubiquitous CNC-bZIP transcription factor, Nrf-1, results in anemia and embryonic lethality in mice. EMBO J. 1998;17:1779–1787.CrossRefGoogle ScholarPubMed
Oyake, T, Itoh, K, Motohashi, H, et al. Bach proteins belong to a novel family of BTB-basic leucine zipper transcription factors that interact with MafK and regulate transcription through the NF-E2 site. Mol Cell Biol. 1996;16:6083–6095.CrossRefGoogle ScholarPubMed
Igarashi, K, Hoshino, H, Muto, A, et al. Multivalent DNA binding complex generated by small Maf and Bach1 as a possible biochemical basis for beta-globin locus control region complex. J Biol Chem. 1998;273:11783–11790.CrossRefGoogle ScholarPubMed
Ogawa, K, Sun, J, Taketani, S, Nakajima, O, et al. Heme mediates derepression of Maf recognition element through direct binding to transcription repressor Bach1. EMBO J. 2001;20(11):2835–2843.CrossRefGoogle ScholarPubMed
Suzuki, H, Tashiro, S, Hira, S, et al. Heme regulates gene expression by triggering Crm1-dependent nuclear export of Bach1. EMBO J. 2004;23(13):2544–2553.CrossRefGoogle ScholarPubMed
Zenke-Kawasaki, Y, Dohi, Y, Katoh, Y, et al. Heme induces ubiquitination and degradation of the transcription factor Bach1. Mol Cell Biol. 2007;27(19):6962–6971.CrossRefGoogle ScholarPubMed
Tahara, T, Sun, J, Igarashi, K, Taketani, S. Heme-dependent up-regulation of the alpha-globin gene expression by transcriptional repressor Bach1 in erythroid cells. Biochem Biophys Res Commun. 2004;324(1):77–85.CrossRefGoogle ScholarPubMed
Tahara, T, Sun, J, Nakanishi, K, et al. Heme positively regulates the expression of b-globin at the locus control region via the transcriptional factor Bach1 in erythroid cells. J Biol Chem. 2004;279:5480–5487.CrossRefGoogle Scholar
Igarashi, K, Kataoka, K, Itoh, K, Hayashi, N, Nishizawa, M, Yamamoto, M. Regulation of transcription by dimerization of erythroid factor NF-E2 p45 with small Maf proteins. Nature. 1994;367:568–572.CrossRefGoogle ScholarPubMed
Kotkow, KJ, Orkin, SH. Complexity of the erythroid transcription factor NFE2 as revealed by gene targeting of the mouse p18 NF-E2 locus. Proc Natl Acad Sci USA. 1996;93:3514–3518.CrossRefGoogle ScholarPubMed
Shavit, JA, Motohashi, H, Onodera, K, Akasaka, J-E, Yamamoto, M, Engel, JD. Impaired megakaryopoiesis and behavioral defects in mafG-null mutant mice. Genes Dev. 1998;12:2164–2174.CrossRefGoogle ScholarPubMed
Motohashi, H, Katsuoka, F, Miyoshi, C, et al. MafG sumoylation is required for active transcriptional repression. Mol Cell Biol. 2006;26(12):4652–4663.CrossRefGoogle ScholarPubMed
Bean, TL, Ney, PA. Multiple regions of p45 NF-E2 are required for b-globin gene expression in erythroid cells. Nucl Acids Res. 1997;25:2509–2515.CrossRefGoogle Scholar
Cheng, X, Reginato, MJ, Andrews, NC, Lazar, MA. The Transcriptional Integrator CREB- binding protein mediates positive cross talk between nuclear hormone receptors and the hematopoietic bZip protein p45/NF-E2. Mol Cell Biol. 1997;1:1407–1416.CrossRefGoogle Scholar
Gavva, NR, Gavva, R, Ermekova, K, Sudol, M, Shen, CJ. Interaction of WW domains with hematopoietic transcription factor p45/NF-E2 and RNA polymerase II. J Biol Chem. 1997;272:24105–24108.CrossRefGoogle ScholarPubMed
Mosser, EA, Kasanov, JD, Forsberg, EC, Kay, BK, Ney, PA, Bresnick, EH. Physical and functional interactions between the transactivation domain of the hematopoietic transcription factor NF-E2 and WW domains. Biochemistry. 1998;37:13686–13695.CrossRefGoogle ScholarPubMed
Amrolia, PJ, Ramamurthy, L, Saluja, D, Tanese, N, Jane, SM, Cunningham, JM. The activation domain of the enhancer binding protein p45NF-E2 interacts with TAFII130 and mediates long-range activation of the alpha- and beta-globin gene loci in an erythroid cell line. Proc Natl Acad Sci USA. 1997;94:10051–10056.CrossRefGoogle Scholar
Shikama, N, Lyon, J, LaThangue, NB. The p300/CBP family: integrating signals with transcription factors and chromatin. Trends Cell Biol. 1997;7:230–236.CrossRefGoogle Scholar
Armstrong, JA, Emerson, BM. NF-E2 disrupts chromatin structure at human beta-globin locus control region hypersensitive site 2 in vitro. Mol Cell Biol. 1996;16(10):5634–5644.CrossRefGoogle ScholarPubMed
Demers, C, Chaturvedi, CP, Ranish, JA, et al. Activator-mediated recruitment of the MLL2 methyltransferase complex to the beta-globin locus. Mol Cell. 2007;27(4):573–584.CrossRefGoogle ScholarPubMed
Brand, M, Ranish, JA, Kummer, NT, et al. Dynamic changes in transcription factor complexes during erythroid differentiation revealed by quantitative proteomics. Nat Struct Mol Biol. 2004;11(1):73–80.CrossRefGoogle ScholarPubMed
Nagai, T, Igarashi, K, Akasaka, J, et al. Regulation of NF-E2 activity in erythroleukemia cell differentiation. J Biol Chem. 1998;273:5358–5365.CrossRefGoogle ScholarPubMed
Versaw, WK, Blank, V, Andrews, NM, Bresnick, EH. Mitogen-activated protein kinases enhance long-range activation by the beta-globin locus control region. Proc Natl Acad Sci USA. 1998;95:8756–8760.CrossRefGoogle ScholarPubMed
Casteel, D, Suhasini, M, Gudi, T, Naima, R, Pilz, RB. Regulation of the erythroid transcription factor NF-E2 by cyclic adenosine monophosphate dependent protein kinase. Blood. 1998;91(9):3193–3201.Google ScholarPubMed
Shyu, YC, Lee, TL, Ting, CY, et al. Sumoylation of p45/NF-E2: nuclear positioning and transcriptional activation of the mammalian beta-like globin gene locus. Mol Cell Biol. 2005;25(23):10365–10378.CrossRefGoogle ScholarPubMed
Choi, O-RB, Engel, JD. Developmental regulation of b-globin gene switching. Cell. 1988;56:17–26.CrossRefGoogle Scholar
Wijgerde, M, Grosveld, F, Fraser, P. Transcription complex stability and chromatin dynamics in vivo. Nature. 1995;377(6546):209–213.CrossRefGoogle ScholarPubMed
Trimborn, T, Gribnau, J, Grosveld, F, Fraser, P. Mechanisms of developmental control of transcription in the murine alpha- and beta-globin loci. Genes Dev. 1999;13(1):112–124.CrossRefGoogle ScholarPubMed
Minie, ME, Kimura, T, Felsenfeld, G. The developmental switch in embryonic rho-globin expression is correlated with erythroid lineage–specific differences in transcription factor levels. Development. 1992;115(4):1149–1164.Google ScholarPubMed
Knezetic, JA, Felsenfeld, G. Mechanism of developmental regulation of alpha pi, the chicken embryonic alpha-globin gene. Mol Cell Biol. 1993;13(8):4632–4639.CrossRefGoogle ScholarPubMed
Jane, SM, Ney, PA, Vanin, EF, Gumucio, DL, Nienhuis, AW. Identification of a stage selector element in the human gamma-globin gene promoter that fosters preferential interaction with the 5′ HS2 enhancer when in competition with the beta-promoter. EMBO J. 1992;11(8):2961–2969.Google ScholarPubMed
Zhou, W, Zhao, Q, Sutton, R, et al. The role of p22 NF-E4 in human globin gene switching. J Biol Chem. 2004;279(25):26227–26232.CrossRefGoogle ScholarPubMed
Sargent, TG, Buller, AM, Teachey, DT, McCanna, KS, Lloyd, JA. The gamma-globin promoter has a major role in competitive inhibition of beta-globin gene expression in early erythroid development. DNA Cell Biol. 1999;18(4):293–303.CrossRefGoogle Scholar
Sargent, TG, DuBois, CC, Buller, AM, Lloyd, JA. The roles of 5′-HS2, 5′-HS3, and the gamma-globin TATA, CACCC, and stage selector elements in suppression of beta-globin expression in early development. J Biol Chem. 1999;274(16):11229–11236.CrossRefGoogle ScholarPubMed
Tanabe, O, Katsuoka, F, Campbell, AD, et al. An embryonic/fetal beta-type globin gene repressor contains a nuclear receptor TR2/TR4 heterodimer. EMBO J. 2002;21(13):3434–3442.CrossRefGoogle ScholarPubMed
Filipe, A, Li, Q, Deveaux, S, Godin, I, et al. Regulation of embryonic/fetal globin genes by nuclear hormone receptors: a novel perspective on hemoglobin switching. EMBO J. 1999;18(3):687–697.CrossRefGoogle ScholarPubMed
Huisman, THJ, Carver, MFH, Baysal, E. A Syllabus of Thalassemia Mutations. Augusta, GA: The Sickle Cell Anemia Foundation; 1997.Google Scholar
Crossley, M, Tsang, AP, Bieker, JJ, Orkin, SH. Regulation of the erythroid Kruppel-like factor (EKLF) gene promoter by the erythroid transcription factor GATA-1. J Biol Chem. 1994;269(22):15440–15444.Google ScholarPubMed
Candido, EP, Reeves, R, Davie, JR. Sodium butyrate inhibits histone deacetylation in cultured cells. Cell. 1978;14(1):105–113.CrossRefGoogle ScholarPubMed
McCaffrey, PG, Newsome, DA, Fibach, E, Yoshida, M, Su, MS. Induction of gamma-globin by histone deacetylase inhibitors. Blood. 1997;90(5):2075–2083.Google ScholarPubMed
Cao, H, Stamatoyannopoulos, G, Jung, M. Induction of human gamma globin gene expression by histone deacetylase inhibitors. Blood. 2004;103(2):701–709.CrossRefGoogle ScholarPubMed
Chen, WY, Bailey, EC, McCune, SL, Dong, JY, Townes, TM. Reactivation of silenced, virally transduced genes by inhibitors of histone deacetylase. Proc Natl Acad Sci USA. 1997;94(11):5798–5803.CrossRefGoogle ScholarPubMed
Sankaran, VG, Menns, TF, Xe, J. et al. Human fetal hemoglobin expression is regulated by the developmental stage-specific repressorr BCL11A. Science. 2008;322:1839–1842.CrossRefGoogle ScholarPubMed
Uda, M, Galanello, R, Sanna, S, et al. Genome-wide association study shows BCL11A associated with persistent fetal hemoglobin and amelioration of the phenotype of beta-thalassemia. Proc Nati Acad Sci USA. 2008;105:1620–1625.CrossRefGoogle ScholarPubMed
Lettre, G, Sankaran, VG, Bezerra, MA, et al. DNA polymorphisms at the BCL11A, HBSIL-MYB, and beta-globin loci associate with fetal hemoglobin levels and pain crises in sickle cell disease. Proc Nati Acad Sci USA. 2008;105:11869–11874.CrossRefGoogle Scholar
Sedgewick, AE, Timofeev, N, Sebastiani, P, et al. BCL11A is a major HbF quantitative trait locus in three different populations with beta-hemoglobinopathies. Blood Cells Mol Dis. 2008;41:255–258.CrossRefGoogle ScholarPubMed
Manzel, S, Garner, C, Gut, I, et al. A QTL, influencing F cell production maps to a gene recoding a zinc-finger protein on chromosome 2p15. Nat Genet. 2007;39:1197–1199.CrossRefGoogle Scholar

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
×