The discovery of the GATA binding protein (GATA factor) transcription factor family revolutionized hematology. Studies of GATA proteins have yielded vital contributions to our understanding of how hematopoietic stem and progenitor cells develop from precursors, how progenitors generate red blood cells, how hemoglobin synthesis is regulated, and the molecular underpinnings of nonmalignant and malignant hematologic disorders. This thrilling journey began with mechanistic studies on a β-globin enhancer- and promoter-binding factor, GATA-1, the founding member of the GATA family. This work ushered in the cloning of related proteins, GATA-2-6, with distinct and/or overlapping expression patterns. Herein, we discuss how the hematopoietic GATA factors (GATA-1-3) function via a battery of mechanistic permutations, which can be GATA factor subtype, cell type, and locus specific. Understanding this intriguing protein family requires consideration of how the mechanistic permutations are amalgamated into circuits to orchestrate processes of interest to the hematologist and more broadly.

Analyzing gene regulation in the current era almost invariably involves genome-wide studies. However, much of what we know about hematopoietic transcriptional mechanisms emerged from highly focused mechanistic studies at a single locus, β-globin. This work led to the cloning of a β-globin enhancer- and promoter-binding factor, GATA binding protein 1 (GATA-1),1,2  the founding member of the GATA transcription factor family and ushered in the cloning of the related proteins GATA-2-6.3-12  GATA-1, 2, and 3 are expressed in specific hematopoietic cell types, as well as in select nonhematopoietic cells, and regulate the development and function of diverse blood cell lineages.13,14  GATA-4, 5, and, 6 control the development and function of select nonhematopoietic cell types and organogenesis, including cardiovascular development.15,16 

The discovery of GATA transcription factors occurred at a time when the first mammalian sequence-specific DNA binding proteins were being purified and cloned.17  Whereas intensive efforts using a wide swath of molecular biological approaches inferred important functions for these factors, it was difficult to ascertain whether they had essential functions. Furthermore, even though millions of copies of the GATA factor-recognized DNA motif (WGATAR) reside in a genome,14  it was unclear whether GATA proteins control a restricted gene cohort or large genetic networks. Rigorous vetting of these foundational questions has yielded lucid answers. Whereas individual GATA factors are often essential for critical biological processes, redundancies can exist. GATA factors establish and maintain complex cell type–specific genetic networks involving dozens to hundreds of activated or repressed target genes,18-20  but these genes represent a very small fraction of those containing GATA motifs.14  The regulation of at least certain genes within the networks requires a specific GATA factor, rather than simultaneous actions of multiple GATA factors.

GATA-1 is expressed in erythroid, megakaryocytic, mast, eosinophil, basophil, and dendritic cells.13,21-23  GATA-1 and GATA-2 expression partially overlap (eg, in yolk sac–derived primitive erythroblasts, megakaryocytes, and eosinophils).24-26  In other contexts, their expression is mutually exclusive or anticorrelative. GATA-2 is uniquely expressed in hematopoietic stem and progenitor cells (HSPCs) and erythroid precursors prior to GATA-1 expression.21,27,28  The distinct patterns infer that these GATA factors control different biological processes. Extremely instructive gene targeting studies strongly supported this concept.

Targeted deletion of Gata1 in mice yields defective erythroid cell development and embryonic day 10.5 (E10.5) to E11.5 lethality.25,29,30  Subsequent studies established the importance of GATA-1 for megakaryocyte, mast cell, eosinophil, and basophil differentiation.31-33  Targeted deletion of Gata2 is lethal at ∼E10 because of a broad collapse of hematopoiesis,24,34  reflecting a GATA-2 requirement for HSPC genesis and function.35-40  Unlike GATA-1 and GATA-2, GATA-3 is expressed in T lymphocytes and controls T-cell lymphopoiesis.41-44  Nonhematopoietic functions include regulating prostate (GATA-2)45,46  and breast (GATA-3)47,48  development and tumorigenesis.

Herein, we discuss how the hematopoietic GATA factors function via a battery of mechanistic permutations to yield complex circuits, with some understood with molecular sophistication and others with only a cursory sketch of the mechanistic fabric. As GATA factor mechanisms can be GATA factor subtype, cell type, and/or locus specific, it is challenging to distill this diversity into a single mechanism that fits into a stringent mold. Understanding this intriguing protein family requires consideration of how the mechanistic permutations are amalgamated into circuits that drive processes of interest to the hematologist and more broadly.

GATA factors bind GATA sequence motifs via a highly conserved dual zinc finger domain (Figure 1A-B).14  This engenders GATA-1 and GATA-2 with similar, if not identical, sequence preferences. The genomic abundance of GATA motifs makes target gene predictions impossible based on sequence alone. Chromatin immunoprecipitation coupled with massively parallel sequencing revealed GATA-1 and GATA-2 occupancy genome-wide, redefined the designation of WGATAR as the GATA motif, and revealed additional motifs enriched at GATA-occupied sites.18-20,49-53  These studies indicated GATA-1 and GATA-2 occupy <1% of GATA motifs, illustrating an exquisite specificity of GATA motif selection.14,54,55 

Figure 1.

GATA-1 and GATA-2 protein attributes. (A) GATA-1 and GATA-2 protein attributes. N- and C-zinc fingers and posttranslational modification sites are indicated.84,87,88,90,117,122,124,125,170-172  (B) Amino acid sequence alignment of human GATA-1 and GATA-2. Protein domains and posttranslational modification sites are highlighted.

Figure 1.

GATA-1 and GATA-2 protein attributes. (A) GATA-1 and GATA-2 protein attributes. N- and C-zinc fingers and posttranslational modification sites are indicated.84,87,88,90,117,122,124,125,170-172  (B) Amino acid sequence alignment of human GATA-1 and GATA-2. Protein domains and posttranslational modification sites are highlighted.

Close modal

Combining chromatin occupancy with transcriptomic data can enable identification of direct GATA-1 and/or GATA-2 targets. Gene coregulation by different GATA factors can occur through binding identical or distinct GATA motifs within a locus. Multiple GATA factors in the same cell can compete for binding site occupancy, and replacement of one GATA factor by another is termed a GATA switch. GATA switches were initially described for GATA-1 and GATA-2 in erythroid differentiation models28,56-59  (Figure 2A). A genome-wide analysis of enhancer usage during HSPC to erythroid differentiation revealed switches at ∼30% of GATA-bound enhancers,60  and ∼30% of GATA-2-occupied sites are GATA-1-occupied upon megakaryopoiesis.51,61  GATA switching is facilitated by the shorter half-life of GATA-2 vs GATA-1.62,63  As GATA-1 and GATA-2 can exert opposing activities through the same chromatin site, GATA switches can change gene expression and may drive erythroid maturation.14,64  In certain contexts, GATA-2 might prime a chromatin site in preparation for subsequent GATA-1 function.14,64 

Figure 2.

GATA factor mechanistic principles. (A) GATA switch model. GATA switches involve replacement of one GATA factor by another at a chromatin target site. GATA switches can be associated with an altered transcriptional output. The GATA switch is illustrated at the Gata2 locus. In erythroblasts, friend of GATA-1 (FOG-1) promotes GATA-1-mediated replacement of chromatin-bound GATA-2, instigating repression.64,74  (B) Coregulator dependency matrix model. “Sensitive” and “insensitive” denote whether reductions in the endogenous coregulators impact expression of the GATA-1 target genes. Distinct coregulator ensembles mediate GATA-1-dependent transcription in a locus-specific and context-dependent manner.91,98,99,173 

Figure 2.

GATA factor mechanistic principles. (A) GATA switch model. GATA switches involve replacement of one GATA factor by another at a chromatin target site. GATA switches can be associated with an altered transcriptional output. The GATA switch is illustrated at the Gata2 locus. In erythroblasts, friend of GATA-1 (FOG-1) promotes GATA-1-mediated replacement of chromatin-bound GATA-2, instigating repression.64,74  (B) Coregulator dependency matrix model. “Sensitive” and “insensitive” denote whether reductions in the endogenous coregulators impact expression of the GATA-1 target genes. Distinct coregulator ensembles mediate GATA-1-dependent transcription in a locus-specific and context-dependent manner.91,98,99,173 

Close modal

GATA-1 and GATA-2 genomic occupancy sites and target genes can be cell type specific and/or overlapping.18,49,51  The milieu of GATA factor–associated coregulators almost certainly contributes to the context dependence65  (Figure 2B). The coregulator ensemble mediating GATA factor activity serves diverse functions, including catalysis of protein modifications (eg, acetylation, sumoylation, and phosphorylation) that directly or indirectly (via chromatin alterations) modulate GATA factor activity (Figure 1A-B). Perhaps the most important GATA-1 coregulator, FOG-1, mediates activation and repression of most GATA-1 target genes.66-68  Resembling Gata1 mice, targeted disruption of Zfpm1, encoding FOG-1, yields failure of primitive and definitive erythropoiesis and embryonic lethality.66,67  A small cohort of GATA-1 targets are insensitive to reduced FOG-1 levels and to GATA-1 mutations that inhibit FOG-1 binding.68-71  Despite its 9 zinc fingers, FOG-1 has not been demonstrated to bind DNA. Four FOG-1 zinc fingers can mediate binding to the GATA-1 N-terminal zinc finger (N-finger).72,73  The GATA-1/FOG-1 interaction promotes GATA-1 occupancy at numerous genomic sites,74,75  and N-finger mutations that inhibit FOG-1 binding can redistribute GATA-1 to ectopic sites.76  FOG-1 recruits the nucleosome remodeling and deacetylase complex77,78  and the repressor C-terminal binding protein (CtBP)73  to GATA-1-bound sites and promotes chromatin looping.79  Although FOG-1 sumoylation enhances binding to CtBP,80  this interaction may be dispensable for erythropoiesis based on studies with mice expressing a FOG-1 mutant lacking a CtBP binding sequence.81  Mice with knock-in mutations in both GATA-1 and GATA-2 that inhibit FOG-1 binding have severely defective megakaryopoiesis,82  resembling FOG-1−/− mice.67  These results led to the proposal that megakarypoiesis requires GATA-1 or GATA-2 to engage FOG-1. We are unaware of evidence indicating that FOG-1 mediates other GATA-2 activities (eg, in HSPCs or GATA-2-expressing erythroid precursors).

As regulation of GATA factor activity is multilayered, a major challenge is to identify and organize these layers into models that can be definitively tested. An intriguing piece of this puzzle involves the selective influence of GATA-1 posttranslational modifications on specific GATA-1 activities, rather than globally at all target loci. This is exemplified by sumoylation of GATA-1 K137 by the PIASy E3 ligase,83  which preferentially impacts FOG-1-sensitive vs FOG-1-insensitive targets.84  Tethering small ubiquitin-like modifier (SUMO) to GATA-1 bypasses the FOG-1 requirement to activate FOG-1-sensitive genes, although SUMO is not required for FOG-1 binding to GATA-1.84  Analysis of subnuclear locus positioning by 3-dimensional immunofluorescence in situ hybridization revealed differential behavior of FOG-1-sensitive and FOG-1-insensitive loci. Transcriptional activation resulted in expulsion of FOG-1/SUMO-sensitive genes from the nuclear periphery, whereas FOG-1-insensitive genes persisted at the periphery.84,85  As locus-specific mechanisms are not commonly elucidated in transcriptional analyses, these studies are likely to yield general principles.

Histone acetyltransferases acetylate histones and nonhistone proteins. The first coregulator interaction described for GATA-1 was with the broadly expressed histone acetyltransferase CREB binding protein and its paralog p300.86  CREB binding protein interacts with numerous transcription factors, binds GATA-1 zinc fingers, and acetylates lysines near the fingers.87,88  This was reported to enhance GATA-1 DNA binding89 ; however, another analysis revealed no impact on DNA binding, but increased chromatin occupancy.87  GATA-1 acetylation promotes Brd3 bromodomain protein binding recruitment to chromatin.90 

Another example of context-dependent regulation that impacts a subset of the GATA-1 target gene ensemble involves GATA-1 interactions with histone methyltransferases.65  SetD8, the only known histone H4 K20 monomethyltransferase (H4K20me1), mediates GATA-1-dependent repression of select target genes.91  SetD8 confers erythroblast survival,92  which may or may not reflect its GATA-1 corepressor function. The histone methyltransferases EZH1 and EZH2 catalyze histone H3 K27 trimethylation as part of the polycomb repressive complex (PRC2).93  During erythroid maturation, a switch from EZH2 to EZH1 in PRC2 redistributes the complex from active genes to bivalent or repressed genes.94  GATA-1 interacts with EZH2 and another component of PRC2, Suz12.19,94 

Chromatin remodeling impacts local and broad chromatin transitions that control gene activity. GATA-1 recruits BRG1, the catalytic subunit of the SWItch Sucrose Non-Fermentable (SWI/SNF) chromatin remodeling complex, to chromatin.71,95,96  BRG1 can mediate GATA-1-dependent chromatin looping and transcriptional activation of α- and β-globin loci.71,97  GATA-1 also recruits the multisubunit Mediator complex,98-100  which interacts with numerous transcription factors and basal transcriptional machinery and promotes RNA polymerase II transcription. Mediator-cohesin interactions promote chromatin looping.101  Targeted deletion of Med1 deregulates GATA-1 target genes and impairs erythroid maturation.102  Given the ubiquitous Mediator functions, it is likely integrated into diverse GATA-1 regulatory layers.

Distinct DNA binding protein combinations occupying chromatin can recruit unique coregulator assemblages. Pairing a GATA motif with an E-box (CANNTG) yields a composite element bound by GATA-1 or GATA-2 and the basic helix-loop-helix transcription factor Scl/TAL1.103  GATA-1–Scl/TAL1 recruits non-DNA binding proteins including LMO2 and LDB1, and additional factors are implicated.104-106  A detailed analysis of composite element mechanisms was described recently.107  Certain GATA factor target loci, Gata1, Gata2, Klf1, and Epb4.2, contain functional composite elements, and BRG1 can be recruited to composite elements.107  Although composite elements display a fourfold higher frequency of GATA occupancy vs classical GATA motifs (0.61% vs 0.14%),18,108  many questions remain regarding mechanistic similarities and differences. GATA-1–Scl/TAL1 complex occupancy does not require Scl/TAL1 DNA binding; GATA-1 can recruit Scl/TAL1 to sites lacking an E-box.109  In mouse HPC-7 and human CD34+ cells, GATA-2, Scl/TAL1, Lyl1, LMO2, Runx1, ERG, and Fli-1 co-occupy chromatin, including sites at genes important for HSPCs.53,110 

Multiple mechanisms governing GATA factor function are exemplified by Gata2 transcriptional regulation. The Gata2 locus represents the first, and most extensively studied, example of GATA switching. GATA-1-mediated replacement of GATA-2 represses Gata2 transcription during erythroid maturation via FOG-1- and SetD8-dependent mechanisms.65  GATA switching occurs at 5 highly conserved Gata2 sites, −77, −3.9, −2.8, −1.8, and +9.5 kb relative to the 1S promoter.28,56,57,64  As GATA-2 and GATA-1 occupy these sites at the active and repressed locus, respectively, predictions could not be made regarding whether sites contribute to Gata2 activation or repression or if they are dispensable. Knockout mouse strains lacking the individual sites revealed variable and unpredictable consequences vis-à-vis Gata2 expression and hematopoiesis. The −3.9, −2.8, and −1.8 homozygous mutant strains are viable, although Gata2 expression is modestly reduced in −2.8 mutant HSPCs.111-113  The −1.8 site maintains GATA-2 repression late in erythropoiesis.111  Deletion of the −3.9 site yields no alterations in Gata2 expression or hematopoiesis.113  The +9.5 and −77 sites are indispensable regulators of Gata2 transcription with distinct activities.35,40  Comparable results were obtained by targeted deletion of the +9.5 site in mice35  vs deleting Gata2 using Cre recombinase driven by a chromosomal segment containing the +9.5 site.114  The +9.5 site homozygous deletion inactivates the hematopoietic stem cell (HSC) generator in the aorta-gonad-mesonephros (AGM) region during embryogenesis.36  Of the Gata2 GATA switch sites, only the +9.5 site possesses an E-box/GATA composite element,115,116  and interrogation of the genome for “+9.5-like” composite elements revealed additional GATA-2-regulated enhancers active in HSPCs.107,108  The −77 homozygous deletion in mice decreases Gata2 expression in myeloid progenitors, strongly reducing granulocyte macrophage progenitor (GMP) and megakaryocyte erythrocyte progenitor (MEP) populations.40  The distinct +9.5 and −77 site activities provide an opportunity to compare how GATA factors engage regulatory factors/signals through these sites to confer unique transcriptional responses, which will yield principles governing GATA factor mechanisms.

Signal-dependent mechanisms

Although GATA factors are subjected to multiple posttranslational modifications, in most cases, it is unclear whether signaling mechanisms transduce vital regulation through these modifications. The entry point to this problem was underwhelming, as simultaneous mutation of multiple GATA-1 phosphorylation sites yielded little to no functional consequences.117,118  Akt stimulates GATA-1 serine 310 phosphorylation,119  and this is implicated in activation of a single GATA-1 target gene, Timp1.120  However, without IGF-1 signaling, Gata1S310A knock-in mice exhibit hemolytic anemia.121  Erythropoietin receptor (EpoR)-mediated serine 310 phosphorylation enhances GATA-1 binding to FOG-1 and disrupts a GATA-1/E2F-2/Rb complex that requires a GATA-1 LxCxE motif.121 

Multisite GATA-2 phosphorylation was discovered from analysis of the activity of a defective GATA-2 mutant (T354M) implicated in the primary immunodeficiency monocytopenia and mycobacterial infection syndrome (MonoMAC), which progresses to myelodysplastic syndrome (MDS) and acute myeloid leukemia (AML).122  The T354M mutant migrates with a slower mobility than GATA-2 on sodium dodecyl sulfate–polyacrylamide gel electrophoresis, and mass spectrometry revealed the slow mobility reflects hyperphosphorylation that requires S192 phosphorylation. Oncogenic Ras and p38 and/or extracellular signal-regulated kinase signaling stimulates GATA-2 hyperphosphorylation, thereby enhancing GATA-2 chromatin occupancy and activity at a target gene cohort. Unlike wild-type GATA-2, hyperphosphorylation cannot stimulate T354M activity, as this DNA binding domain mutation impairs chromatin occupancy. Thus, T354M hyperphosphorylation appears to be inconsequential and need not be reversed. A GATA-2 MAPK docking module (DEF motif) is essential for oncogenic Ras-induced GATA-2 hyperphosphorylation in AML cells.123  This mechanism increases interleukin-1β and CXCL2 expression, which can activate p38/extracellular signal-regulated kinase and further enhance GATA-2 activity (Figure 3A). As CXCL2 can stimulate AML cell proliferation, the MAPK-GATA-2-interleukin-1β/CXCL2 axis constitutes a positive-feedback circuit implicated in leukemic cell proliferation. GATA-2 is also regulated via insulin-dependent Akt activation, which phosphorylates GATA-2 S401 and suppresses GATA-2 function.124  CDK1 phosphorylates GATA-2 T176 and promotes Fbw7-mediated GATA-2 degradation.125 

Figure 3.

Emerging GATA factor-dependent mechanistic circuits. (A) Ras-MAPK signaling controls GATA-2 activity.123  (B) GATA-2-GPR65 circuit negatively regulates hematopoiesis.127  (C) GATA-1-heme circuit regulates erythroid differentiation.134  (D) GATA-1-FoxO3-exosome circuit controls erythroid maturation.146,148  FFL, feed forward loop.

Figure 3.

Emerging GATA factor-dependent mechanistic circuits. (A) Ras-MAPK signaling controls GATA-2 activity.123  (B) GATA-2-GPR65 circuit negatively regulates hematopoiesis.127  (C) GATA-1-heme circuit regulates erythroid differentiation.134  (D) GATA-1-FoxO3-exosome circuit controls erythroid maturation.146,148  FFL, feed forward loop.

Close modal

Can the signal-dependent GATA-2 mechanism be extrapolated to other GATA factors? P38 phosphorylates and regulates GATA-3, and this has been suggested to control nuclear translocation and Th2 cytokine gene expression.126  Essentially nothing is known about GATA factor subcellular transitions. Although the phosphorylated GATA-3 amino acid residues were not described, GATA-3 contains a threonine analogous to GATA-2 S192, and we predict GATA-3 also contains a DEF motif.

GATA-2-G-protein-coupled receptor circuit

Ensuring GATA-2 expression within a physiological window.

Gata2 +9.5−/− mouse embryos harbor very few HSPCs,35  as the HSC generator in their AGM is defective.36  This system permitted genomic analyses to establish how the +9.5 enhancer triggers the HSC generator. Although it was hypothesized that the enhancer ensures GATA-2 expression and establishes a genetic network consisting of positive mediators of GATA-2 function, the analysis yielded the discovery of a GATA-2-activated gene, Gpr65, that opposes GATA-2 function.127 

GPR65 downregulation in the AGM increases Gata2 expression and AGM-derived CD31+c-Kit+Sca1+ cells, which include HSCs. Mechanistically, GPR65 represses Gata2 transcription and involves SetD8-dependent establishment of repressive chromatin at the +9.5 enhancer. Whereas SetD8 can function as a GATA-1 corepressor, how H4K20me1 alters chromatin function is not understood. Furthermore, how GPR65 interfaces with SetD8 and Gata2 is unknown. Nevertheless, this GATA-2-GPR65 mechanism constitutes a negative-feedback circuit that contributes to maintenance of Gata2 expression within a physiological window (Figure 3B). Either too little or too much GATA-2 is linked to myeloid leukemogenesis,14  and therefore it is exceptionally important to ensure fidelity of the multilayered mechanism dictating normal GATA-2 levels.

GATA-1-heme circuit

Linking heme and globin chain synthesis to control hemoglobin production and erythrocyte development.

Heme biosynthesis in mitochondria128-130  requires GATA-1 target genes encoding heme biosynthesis enzymes. The first and rate-limiting step of heme biosynthesis is catalyzed by 5-aminolevulinic acid synthase (ALAS), which utilizes glycine and succinyl-CoA to form ALA.131  Mammals have 2 ALAS isoforms, ALAS-1 and ALAS-2. Whereas ALAS-1 mediates housekeeping functions in diverse cell types, ALAS-2 is expressed specifically in erythroid cells to generate heme for the developing erythroblast.132,133  GATA-1 strongly activates ALAS2 transcription.18,134  Posttranscriptionally, when iron is low, iron-responsive proteins (IRPs) bind the iron-responsive element (IRE) in the ALAS2 messenger RNA 5′ untranslated region to prevent its translation.135  Excessive heme also induces ALAS-2 protein ubiquitination and degradation,136,137  which establishes a negative feedback circuit to avoid heme toxicity. Furthermore, heme deficiency activates heme-regulated eIF2α kinase, which phosphorylates the eIF2α translational factor to decrease globin translation.138 

Considering that GATA-1 activates β- and α-globin gene transcription, as well as ALAS2 and other heme biosynthetic genes,18,139  and colocalizes with a key repressor of fetal hemoglobin expression BCL11A,140  GATA-1 is perhaps the most important factor mediating hemoglobin biosynthesis. Heme binds the transcriptional repressor Bach1, promoting its degradation via the ubiquitin-proteasome system.141  Bach1 only accumulates in a low-heme environment when GATA-1 activates Bach1 transcription.134  GATA-1-mediated Bach1 transcriptional activation is insufficient to elevate Bach1, as the posttranslational destruction system dominates when heme levels are normal. Bach1 represses β-globin transcription,134,142  which helps restore balance between globin chains and heme and prevents excessive globin chains from eliciting toxicity.

Dissecting mechanisms underlying GATA-1-mediated Alas2 transcription revealed an intriguing interconnectivity between heme and GATA-1 that constitutes a new GATA factor mechanism: heme amplifies GATA-1 activity at a cohort of target genes.134  Using CRISPR/Cas9 to delete 2 intronic GATA-1 binding regions at Alas2, GATA-1-null G1E proerythroblasts were engineered to express >10-fold lower heme. Although these mutant cells are viable and proliferate normally, a conditionally active GATA-1 allele (ER-GATA-1) is less effective in activating select GATA-1 target genes. The cell permeable metabolite 5-ALA bypasses this ALAS-2 requirement, rescuing heme levels and increasing GATA-1 target gene expression. This rescue involves Bach1-sensitive and Bach1-insensitive modes, as Bach1 downregulation increases expression of certain, but not all, GATA-1 target genes in the low-heme environment. Thus, GATA-1-mediated induction of Alas2 and Bach1 transcription, and heme downregulation of Bach1, conforms to a Type I incoherent feedforward loop143  (Figure 3C). This network motif is embedded in a complex circuit involving other constituents of the target gene ensemble. As heme functions extend beyond hemoglobin synthesis, this circuit may impact diverse physiological and pathological states. These studies illustrate how changes in a single cofactor (heme) can drastically impact GATA factor function.

GATA-1-exosome complex circuit

Balancing erythroblast proliferation and differentiation through amalgamated feed-forward loops.

Although many studies have described mechanisms promoting erythroid cell development, mechanisms counteracting prodifferentiation factors/signals are less well understood. The studies described previously and others144,145  demonstrate that low heme is not conducive to erythroid maturation. Are there physiological mechanisms that establish maturation roadblocks that are deconstructed by GATA-1 actions? An answer emerged from analyses of how the transcription factor FoxO3 amplifies GATA-1 activity to activate genes encoding autophagy components.50 

FoxO3 controls a subset of GATA-1-regulated genes, providing another example of context-dependent GATA-1 factor mechanisms.146  These GATA-1/FoxO3-coregulated genes include those encoding subunits of the exosome complex, an 11 subunit complex mediating the selective degradation of noncoding and coding RNAs.147  The GATA-1/FoxO3-mediated reduction of exosome complex components suggested the complex might create an impediment to erythroid maturation. Beyond controlling RNA degradation and processing, the complex is implicated in transcriptional regulation.147 

Loss-of-function studies provided evidence for an exosome complex-mediated erythroid maturation barricade.146  Downregulating exosome complex subunits disrupted the complex, which stimulated primary erythroid precursors to differentiate into more mature erythroblasts and even enucleated erythrocytes. Mechanistic studies revealed a circuit in which the exosome complex confers the barricade via at least 2 mechanisms.148  First, it sustains expression of the receptor tyrosine kinase c-Kit. Stem cell factor–mediated c-Kit activation can maintain erythroid precursors in a proliferating, undifferentiated state, and GATA-1-mediated repression of exosome complex subunits decreases c-Kit expression and attenuates c-Kit signaling.148  Moreover, GATA-1 directly represses Kit transcription.149  Second, GATA-1 is implicated in activating expression of the gene encoding EpoR, which generates prodifferentiation signals.150  In addition to sustaining c-Kit signaling, the exosome complex suppresses EpoR signaling.148  GATA-1-mediated transcriptional repression of exosome complex subunit genes and Kit, and exosome complex-mediated c-Kit expression conforms to a type II coherent feed-forward loop143  (Figure 3D). GATA-1-mediated repression of exosome complex subunits and exosome complex-mediated suppression of EpoR signaling conforms to a type IV coherent feed-forward loop143  (Figure 3D). The GATA-1-exosome complex circuit balances erythroid precursor proliferation vs differentiation. It is instructive to consider the interconnectivity of pro- and antidifferentiation circuits in various biological systems, and further probing into this problem will almost certainly yield new physiological and pathological insights.

Mechanisms governing GATA factor function have been informed by disease-causing mutations in humans with hematologic disorders. As GATA-1 dysregulation has been reviewed,151,152  our focus will be on mechanistic insights emerging from human GATA-2 mutations. Although the preponderance of known GATA2 mutations occur within the coding region,153-156  a minority reside in noncoding sequences.35,157  A GATA2-AML link emerged when GATA2 heterozygous mutations were detected in the clinical syndromes MonoMAC and DCML (dendritic cell, monocyte, B and NK lymphoid deficiency) primary immunodeficiency, familial MDS/AML and Emberger syndrome.153-156  Most missense mutations reside in or near the DNA binding GATA-2 C-finger, although function-disrupting frameshift and nonsense mutations also exist158  (Figure 4A). The GATA-2 T354M mutant detected in MonoMAC and familial MDS/AML, exhibited reduced chromatin occupancy and failed to activate the GATA-2 target gene Hdc.122  Reporter and DNA binding assays identified loss-of-function GATA2 mutations in MonoMAC.153,156  In addition to AML, GATA2 L359V mutation was detected in chronic myeloid leukemia patients in blast crisis.159  In AML patients with biallelic CEBPA mutations, GATA2 mutations are present with frequencies of ∼30% of patients, with most mutations being N-finger missense.160-162  N-finger mutations were described in 22% of patients with acute erythroid leukemia.163  Although N-finger mutations are known to disrupt GATA-1 function in humans164  and experimental models,165,166  how N-fingers mutations impact GATA-2 function is unknown (Figure 4A).

Figure 4.

GATA-2 mutations in human hematologic disorders inform GATA factor mechanisms. (A) Left, GATA-2 N-finger mutations in human AML patients with biallelic CEBPA mutations.160-163,174,175  V296 corresponds to GATA-1 V205, which enhances GATA-1 and FOG-1 binding. Right, C-finger mutations identified in AML-associated diseases.153-156,158,176-178  T354M is a loss-of-function mutation that inhibits chromatin occupancy and target gene activation.122,123  L359V was identified in chronic myeloid leukemia.159  (B) Mutations at and near the +9.5 GATA switch site enhancer in pediatric MDS169  and MonoMAC syndrome.35,157,158  del, deletion; ins, insertion.

Figure 4.

GATA-2 mutations in human hematologic disorders inform GATA factor mechanisms. (A) Left, GATA-2 N-finger mutations in human AML patients with biallelic CEBPA mutations.160-163,174,175  V296 corresponds to GATA-1 V205, which enhances GATA-1 and FOG-1 binding. Right, C-finger mutations identified in AML-associated diseases.153-156,158,176-178  T354M is a loss-of-function mutation that inhibits chromatin occupancy and target gene activation.122,123  L359V was identified in chronic myeloid leukemia.159  (B) Mutations at and near the +9.5 GATA switch site enhancer in pediatric MDS169  and MonoMAC syndrome.35,157,158  del, deletion; ins, insertion.

Close modal

A subset of acute AML is characterized by chromosomal rearrangements involving 3q21 and 3q26, which result in overexpression of MECOM encoding the proto-oncogene EVI1.167  Chromosomal rearrangements translocate the GATA2 −77 kb GATA switch site enhancer ∼4 megabases away into proximity of MECOM.167,168  As the −77 confers Gata2 expression to confer differentiation potential in mouse myeloid progenitors,40  its removal from human GATA2 likely decreases GATA-2 levels, while elevating EVI1 levels. Thus, co-opting a hematopoietic GATA switch site enhancer deregulates a proto-oncogene and underlies 3q21:q26 AML. These studies provided strong evidence that integrity of the −77 enhancer is crucial to suppress the development of a human hematologic malignancy.

Disruption of the GATA2 +9.5 GATA switch site enhancer can also cause MonoMAC.35  Typically, MonoMAC patients have mutations within the DNA binding region of one GATA2 allele, leading to haploinsufficiency. However, a MonoMAC patient with both GATA2 genes intact harbored a heterozygous deletion within the +9.5 site, which eliminated the E-box and several base pairs of upstream sequence, while preserving the GATA motif.35  Additional MonoMAC patients have +9.5 kb mutations in a neighboring E-twenty six (ETS) motif157  (Figure 4B). The +9.5 kb site mutations have also been detected in juvenile MDS patients.169  These studies indicate that multiple modules within the +9.5 composite element, which normally triggers the mouse HSC generator, control enhancer activity and hematopoiesis in humans. We expect additional GATA2 coding and noncoding mutations will provide vital mechanistic insights, and these insights may transform our understanding of how GATA factors are regulated, how they function physiologically, and how mechanistic corruption instigates pathologies including malignancy and anemia.

E.H.B. was supported by the National Institute of Diabetes and Digestive and Kidney Diseases (grants DK50107 and DK68634), the National Heart, Lung, and Blood Institute (grant HL116365), and the National Cancer Institute (Cancer Center Support Grant P30 CA014520), National Institutes of Health. Skye C. McIver and Kyle J. Hewitt were supported by American Heart Association postdoctoral fellowships. Alexandra A. Soukup was supported by the National Heart, Lung, and Blood Institute, National Institutes of Health (T32 HL07899 Training Grant in Hematology).

Contribution: This review was generated by a multidisciplinary team under the leadership of E.H.B.; and all authors contributed actively and significantly to the conceptualization and writing of the review.

Conflict-of-interest disclosure: The authors declare no competing financial interests.

A list of additional members of the GATA Factor Mechanisms Group appears in “Appendix.”

Correspondence: Emery H. Bresnick, Department of Cell and Regenerative Biology, University of Wisconsin School of Medicine and Public Health, 4009 WIMR, 1111 Highland Ave, Madison, WI 53705; e-mail: ehbresni@wisc.edu.

Additional members of the GATA Factor Mechanisms Group are: Xin Gao (University of Wisconsin School of Medicine and Public Health), Kyle J. Hewitt (University of Wisconsin School of Medicine and Public Health), Daniel R. Matson (University of Wisconsin School of Medicine and Public Health), Skye C. McIver (University of Wisconsin School of Medicine and Public Health), Charu Mehta (University of Wisconsin School of Medicine and Public Health), Alexandra A. Soukup (University of Wisconsin School of Medicine and Public Health), Nobuyuki Tanimura (University of Wisconsin School of Medicine and Public Health), Lihong Shi (State Key Laboratory of Experimental Hematology, Institute of Hematology and Blood Diseases Hospital, Chinese Academy of Medical Sciences & Peking Union Medical College), and Kirby D. Johnson (University of Wisconsin School of Medicine and Public Health).

1.
Evans
T
,
Felsenfeld
G
.
The erythroid-specific transcription factor Eryf1: a new finger protein
.
Cell
.
1989
;
58
(
5
):
877
-
885
.
2.
Tsai
SF
,
Martin
DI
,
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
(
6224
):
446
-
451
.
3.
Yamamoto
M
,
Ko
LJ
,
Leonard
MW
,
Beug
H
,
Orkin
SH
,
Engel
JD
.
Activity and tissue-specific expression of the transcription factor NF-E1 multigene family
.
Genes Dev
.
1990
;
4
(
10
):
1650
-
1662
.
4.
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
(
23
):
10642
-
10646
.
5.
Dorfman
DM
,
Wilson
DB
,
Bruns
GA
,
Orkin
SH
.
Human transcription factor GATA-2. Evidence for regulation of preproendothelin-1 gene expression in endothelial cells
.
J Biol Chem
.
1992
;
267
(
2
):
1279
-
1285
.
6.
Lee
ME
,
Temizer
DH
,
Clifford
JA
,
Quertermous
T
.
Cloning of the GATA-binding protein that regulates endothelin-1 gene expression in endothelial cells
.
J Biol Chem
.
1991
;
266
(
24
):
16188
-
16192
.
7.
Ko
LJ
,
Yamamoto
M
,
Leonard
MW
,
George
KM
,
Ting
P
,
Engel
JD
.
Murine and human T-lymphocyte GATA-3 factors mediate transcription through a cis-regulatory element within the human T-cell receptor delta gene enhancer
.
Mol Cell Biol
.
1991
;
11
(
5
):
2778
-
2784
.
8.
Joulin
V
,
Bories
D
,
Eléouet
JF
, et al
.
A T-cell specific TCR delta DNA binding protein is a member of the human GATA family
.
EMBO J
.
1991
;
10
(
7
):
1809
-
1816
.
9.
Ho
IC
,
Vorhees
P
,
Marin
N
, et al
.
Human GATA-3: a lineage-restricted transcription factor that regulates the expression of the T cell receptor alpha gene
.
EMBO J
.
1991
;
10
(
5
):
1187
-
1192
.
10.
Arceci
RJ
,
King
AA
,
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
.
11.
Morrisey
EE
,
Ip
HS
,
Lu
MM
,
Parmacek
MS
.
GATA-6: a zinc finger transcription factor that is expressed in multiple cell lineages derived from lateral mesoderm
.
Dev Biol
.
1996
;
177
(
1
):
309
-
322
.
12.
Morrisey
EE
,
Ip
HS
,
Tang
Z
,
Lu
MM
,
Parmacek
MS
.
GATA-5: a transcriptional activator expressed in a novel temporally and spatially-restricted pattern during embryonic development
.
Dev Biol
.
1997
;
183
(
1
):
21
-
36
.
13.
Weiss
MJ
,
Orkin
SH
.
GATA transcription factors: key regulators of hematopoiesis
.
Exp Hematol
.
1995
;
23
(
2
):
99
-
107
.
14.
Bresnick
EH
,
Katsumura
KR
,
Lee
HY
,
Johnson
KD
,
Perkins
AS
.
Master regulatory GATA transcription factors: mechanistic principles and emerging links to hematologic malignancies
.
Nucleic Acids Res
.
2012
;
40
(
13
):
5819
-
5831
.
15.
Charron
F
,
Nemer
M
.
GATA transcription factors and cardiac development
.
Semin Cell Dev Biol
.
1999
;
10
(
1
):
85
-
91
.
16.
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
.
17.
Kadonaga
JT
,
Carner
KR
,
Masiarz
FR
,
Tjian
R
.
Isolation of cDNA encoding transcription factor Sp1 and functional analysis of the DNA binding domain
.
Cell
.
1987
;
51
(
6
):
1079
-
1090
.
18.
Fujiwara
T
,
O’Geen
H
,
Keles
S
, et al
.
Discovering hematopoietic mechanisms through genome-wide analysis of GATA factor chromatin occupancy
.
Mol Cell
.
2009
;
36
(
4
):
667
-
681
.
19.
Yu
M
,
Riva
L
,
Xie
H
, et al
.
Insights into GATA-1-mediated gene activation versus repression via genome-wide chromatin occupancy analysis
.
Mol Cell
.
2009
;
36
(
4
):
682
-
695
.
20.
Cheng
Y
,
Wu
W
,
Kumar
SA
, et al
.
Erythroid GATA1 function revealed by genome-wide analysis of transcription factor occupancy, histone modifications, and mRNA expression
.
Genome Res
.
2009
;
19
(
12
):
2172
-
2184
.
21.
Leonard
M
,
Brice
M
,
Engel
JD
,
Papayannopoulou
T
.
Dynamics of GATA transcription factor expression during erythroid differentiation
.
Blood
.
1993
;
82
(
4
):
1071
-
1079
.
22.
Martin
DI
,
Zon
LI
,
Mutter
G
,
Orkin
SH
.
Expression of an erythroid transcription factor in megakaryocytic and mast cell lineages
.
Nature
.
1990
;
344
(
6265
):
444
-
447
.
23.
Gutiérrez
L
,
Nikolic
T
,
van Dijk
TB
, et al
.
Gata1 regulates dendritic-cell development and survival
.
Blood
.
2007
;
110
(
6
):
1933
-
1941
.
24.
Tsai
F-Y
,
Orkin
SH
.
Transcription factor GATA-2 is required for proliferation/survival of early hematopoietic cells and mast cell formation, but not for erythroid and myeloid terminal differentiation
.
Blood
.
1997
;
89
(
10
):
3636
-
3643
.
25.
Fujiwara
Y
,
Chang
AN
,
Williams
AM
,
Orkin
SH
.
Functional overlap of GATA-1 and GATA-2 in primitive hematopoietic development
.
Blood
.
2004
;
103
(
2
):
583
-
585
.
26.
Zon
LI
,
Yamaguchi
Y
,
Yee
K
, et al
.
Expression of mRNA for the GATA-binding proteins in human eosinophils and basophils: potential role in gene transcription
.
Blood
.
1993
;
81
(
12
):
3234
-
3241
.
27.
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
(
10
):
1184
-
1197
.
28.
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
.
29.
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
(
6306
):
257
-
260
.
30.
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
.
31.
Vyas
P
,
Ault
K
,
Jackson
CW
,
Orkin
SH
,
Shivdasani
RA
.
Consequences of GATA-1 deficiency in megakaryocytes and platelets
.
Blood
.
1999
;
93
(
9
):
2867
-
2875
.
32.
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
.
33.
Nei
Y
,
Obata-Ninomiya
K
,
Tsutsui
H
, et al
.
GATA-1 regulates the generation and function of basophils
.
Proc Natl Acad Sci USA
.
2013
;
110
(
46
):
18620
-
18625
.
34.
Tsai
FY
,
Keller
G
,
Kuo
FC
, et al
.
An early haematopoietic defect in mice lacking the transcription factor GATA-2
.
Nature
.
1994
;
371
(
6494
):
221
-
226
.
35.
Johnson
KD
,
Hsu
AP
,
Ryu
MJ
, et al
.
Cis-element mutated in GATA2-dependent immunodeficiency governs hematopoiesis and vascular integrity
.
J Clin Invest
.
2012
;
122
(
10
):
3692
-
3704
.
36.
Gao
X
,
Johnson
KD
,
Chang
YI
, et al
.
Gata2 cis-element is required for hematopoietic stem cell generation in the mammalian embryo
.
J Exp Med
.
2013
;
210
(
13
):
2833
-
2842
.
37.
Ling
KW
,
Ottersbach
K
,
van Hamburg
JP
, et al
.
GATA-2 plays two functionally distinct roles during the ontogeny of hematopoietic stem cells
.
J Exp Med
.
2004
;
200
(
7
):
871
-
882
.
38.
Rodrigues
NP
,
Janzen
V
,
Forkert
R
, et al
.
Haploinsufficiency of GATA-2 perturbs adult hematopoietic stem-cell homeostasis
.
Blood
.
2005
;
106
(
2
):
477
-
484
.
39.
Rodrigues
NP
,
Boyd
AS
,
Fugazza
C
, et al
.
GATA-2 regulates granulocyte-macrophage progenitor cell function
.
Blood
.
2008
;
112
(
13
):
4862
-
4873
.
40.
Johnson
KD
,
Kong
G
,
Gao
X
, et al
.
Cis-regulatory mechanisms governing stem and progenitor cell transitions
.
Sci Adv
.
2015
;
1
(
8
):
e1500503
.
41.
Hosoya
T
,
Kuroha
T
,
Moriguchi
T
, et al
.
GATA-3 is required for early T lineage progenitor development
.
J Exp Med
.
2009
;
206
(
13
):
2987
-
3000
.
42.
Pai
SY
,
Truitt
ML
,
Ho
IC
.
GATA-3 deficiency abrogates the development and maintenance of T helper type 2 cells
.
Proc Natl Acad Sci USA
.
2004
;
101
(
7
):
1993
-
1998
.
43.
Ting
CN
,
Olson
MC
,
Barton
KP
,
Leiden
JM
.
Transcription factor GATA-3 is required for development of the T-cell lineage
.
Nature
.
1996
;
384
(
6608
):
474
-
478
.
44.
Pai
SY
,
Truitt
ML
,
Ting
CN
,
Leiden
JM
,
Glimcher
LH
,
Ho
IC
.
Critical roles for transcription factor GATA-3 in thymocyte development
.
Immunity
.
2003
;
19
(
6
):
863
-
875
.
45.
Vidal
SJ
,
Rodriguez-Bravo
V
,
Quinn
SA
, et al
.
A targetable GATA2-IGF2 axis confers aggressiveness in lethal prostate cancer
.
Cancer Cell
.
2015
;
27
(
2
):
223
-
239
.
46.
Wu
D
,
Sunkel
B
,
Chen
Z
, et al
.
Three-tiered role of the pioneer factor GATA2 in promoting androgen-dependent gene expression in prostate cancer
.
Nucleic Acids Res
.
2014
;
42
(
6
):
3607
-
3622
.
47.
Kouros-Mehr
H
,
Slorach
EM
,
Sternlicht
MD
,
Werb
Z
.
GATA-3 maintains the differentiation of the luminal cell fate in the mammary gland
.
Cell
.
2006
;
127
(
5
):
1041
-
1055
.
48.
Kouros-Mehr
H
,
Bechis
SK
,
Slorach
EM
, et al
.
GATA-3 links tumor differentiation and dissemination in a luminal breast cancer model
.
Cancer Cell
.
2008
;
13
(
2
):
141
-
152
.
49.
Linnemann
AK
,
O’Geen
H
,
Keles
S
,
Farnham
PJ
,
Bresnick
EH
.
Genetic framework for GATA factor function in vascular biology
.
Proc Natl Acad Sci USA
.
2011
;
108
(
33
):
13641
-
13646
.
50.
Kang
YA
,
Sanalkumar
R
,
O’Geen
H
, et al
.
Autophagy driven by a master regulator of hematopoiesis
.
Mol Cell Biol
.
2012
;
32
(
1
):
226
-
239
.
51.
Doré
LC
,
Crispino
JD
.
Transcription factor networks in erythroid cell and megakaryocyte development
.
Blood
.
2011
;
118
(
2
):
231
-
239
.
52.
Doré
LC
,
Chlon
TM
,
Brown
CD
,
White
KP
,
Crispino
JD
.
Chromatin occupancy analysis reveals genome-wide GATA factor switching during hematopoiesis
.
Blood
.
2012
;
119
(
16
):
3724
-
3733
.
53.
Wilson
NK
,
Foster
SD
,
Wang
X
, et al
.
Combinatorial transcriptional control in blood stem/progenitor cells: genome-wide analysis of ten major transcriptional regulators
.
Cell Stem Cell
.
2010
;
7
(
4
):
532
-
544
.
54.
DeVilbiss
AW
,
Sanalkumar
R
,
Johnson
KD
,
Keles
S
,
Bresnick
EH
.
Hematopoietic transcriptional mechanisms: from locus-specific to genome-wide vantage points
.
Exp Hematol
.
2014
;
42
(
8
):
618
-
629
.
55.
Bresnick
EH
,
Johnson
KD
,
Kim
S-I
,
Im
H
.
Establishment and regulation of chromatin domains: mechanistic insights from studies of hemoglobin synthesis
.
Prog Nucleic Acid Res Mol Biol
.
2006
;
81
:
435
-
471
.
56.
Martowicz
ML
,
Grass
JA
,
Boyer
ME
,
Guend
H
,
Bresnick
EH
.
Dynamic GATA factor interplay at a multicomponent regulatory region of the GATA-2 locus
.
J Biol Chem
.
2005
;
280
(
3
):
1724
-
1732
.
57.
Grass
JA
,
Jing
H
,
Kim
S-I
, et al
.
Distinct functions of dispersed GATA factor complexes at an endogenous gene locus
.
Mol Cell Biol
.
2006
;
26
(
19
):
7056
-
7067
.
58.
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
.
59.
Lugus
JJ
,
Chung
YS
,
Mills
JC
, et al
.
GATA2 functions at multiple steps in hemangioblast development and differentiation
.
Development
.
2007
;
134
(
2
):
393
-
405
.
60.
Huang
J
,
Liu
X
,
Li
D
, et al
.
Dynamic control of enhancer repertoires drives lineage and stage-specific transcription during hematopoiesis
.
Dev Cell
.
2016
;
36
(
1
):
9
-
23
.
61.
Tijssen
MR
,
Cvejic
A
,
Joshi
A
, et al
.
Genome-wide analysis of simultaneous GATA1/2, RUNX1, FLI1, and SCL binding in megakaryocytes identifies hematopoietic regulators
.
Dev Cell
.
2011
;
20
(
5
):
597
-
609
.
62.
Lurie
LJ
,
Boyer
ME
,
Grass
JA
,
Bresnick
EH
.
Differential GATA factor stabilities: implications for chromatin occupancy by structurally similar transcription factors
.
Biochemistry
.
2008
;
47
(
3
):
859
-
869
.
63.
Minegishi
N
,
Suzuki
N
,
Kawatani
Y
,
Shimizu
R
,
Yamamoto
M
.
Rapid turnover of GATA-2 via ubiquitin-proteasome protein degradation pathway
.
Genes Cells
.
2005
;
10
(
7
):
693
-
704
.
64.
Bresnick
EH
,
Lee
HY
,
Fujiwara
T
,
Johnson
KD
,
Keles
S
.
GATA switches as developmental drivers
.
J Biol Chem
.
2010
;
285
(
41
):
31087
-
31093
.
65.
DeVilbiss
AW
,
Tanimura
N
,
McIver
SC
,
Katsumura
KR
,
Johnson
KD
,
Bresnick
EH
.
Navigating transcriptional coregulator ensembles to establish genetic networks: a GATA factor perspective
.
Curr Top Dev Biol
.
2016
;
118
:
205
-
244
.
66.
Tsang
AP
,
Visvader
JE
,
Turner
CA
, 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
(
1
):
109
-
119
.
67.
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
.
68.
Crispino
JD
,
Lodish
MB
,
MacKay
JP
,
Orkin
SH
.
Use of altered specificity mutants to probe a specific protein-protein interaction in differentiation: the GATA-1:FOG complex
.
Mol Cell
.
1999
;
3
(
2
):
219
-
228
.
69.
Johnson
KD
,
Boyer
ME
,
Kang
JA
,
Wickrema
A
,
Cantor
AB
,
Bresnick
EH
.
Friend of GATA-1-independent transcriptional repression: a novel mode of GATA-1 function
.
Blood
.
2007
;
109
(
12
):
5230
-
5233
.
70.
Pal
S
,
Nemeth
MJ
,
Bodine
D
, et al
.
Neurokinin-B transcription in erythroid cells: direct activation by the hematopoietic transcription factor GATA-1
.
J Biol Chem
.
2004
;
279
(
30
):
31348
-
31356
.
71.
Kim
S-I
,
Bultman
SJ
,
Jing
H
,
Blobel
GA
,
Bresnick
EH
.
Dissecting molecular steps in chromatin domain activation during hematopoietic differentiation
.
Mol Cell Biol
.
2007
;
27
(
12
):
4551
-
4565
.
72.
Cantor
AB
,
Katz
SG
,
Orkin
SH
.
Distinct domains of the GATA-1 cofactor FOG-1 differentially influence erythroid versus megakaryocytic maturation
.
Mol Cell Biol
.
2002
;
22
(
12
):
4268
-
4279
.
73.
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
.
74.
Pal
S
,
Cantor
AB
,
Johnson
KD
, et al
.
Coregulator-dependent facilitation of chromatin occupancy by GATA-1
.
Proc Natl Acad Sci USA
.
2004
;
101
(
4
):
980
-
985
.
75.
Letting
DL
,
Chen
YY
,
Rakowski
C
,
Reedy
S
,
Blobel
GA
.
Context-dependent regulation of GATA-1 by friend of GATA-1
.
Proc Natl Acad Sci USA
.
2004
;
101
(
2
):
476
-
481
.
76.
Chlon
TM
,
Doré
LC
,
Crispino
JD
.
Cofactor-mediated restriction of GATA-1 chromatin occupancy coordinates lineage-specific gene expression
.
Mol Cell
.
2012
;
47
(
4
):
608
-
621
.
77.
Miccio
A
,
Wang
Y
,
Hong
W
, et al
.
NuRD mediates activating and repressive functions of GATA-1 and FOG-1 during blood development
.
EMBO J
.
2010
;
29
(
2
):
442
-
456
.
78.
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
.
79.
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
.
80.
Snow
JW
,
Kim
J
,
Currie
CR
,
Xu
J
,
Orkin
SH
.
Sumoylation regulates interaction of FOG1 with C-terminal-binding protein (CTBP)
.
J Biol Chem
.
2010
;
285
(
36
):
28064
-
28075
.
81.
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
.
82.
Chang
AN
,
Cantor
AB
,
Fujiwara
Y
, et al
.
GATA-factor dependence of the multitype zinc-finger protein FOG-1 for its essential role in megakaryopoiesis
.
Proc Natl Acad Sci USA
.
2002
;
99
(
14
):
9237
-
9242
.
83.
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
.
84.
Lee
H-Y
,
Johnson
KD
,
Fujiwara
T
,
Boyer
ME
,
Kim
S-I
,
Bresnick
EH
.
Controlling hematopoiesis through sumoylation-dependent regulation of a GATA factor
.
Mol Cell
.
2009
;
36
(
6
):
984
-
995
.
85.
Lee
HY
,
Johnson
KD
,
Boyer
ME
,
Bresnick
EH
.
Relocalizing genetic loci into specific subnuclear neighborhoods
.
J Biol Chem
.
2011
;
286
(
21
):
18834
-
18844
.
86.
Blobel
GA
,
Nakajima
T
,
Eckner
R
,
Montminy
M
,
Orkin
SH
.
CREB-binding protein cooperates with transcription factor GATA-1 and is required for erythroid differentiation
.
Proc Natl Acad Sci USA
.
1998
;
95
(
5
):
2061
-
2066
.
87.
Lamonica
JM
,
Vakoc
CR
,
Blobel
GA
.
Acetylation of GATA-1 is required for chromatin occupancy
.
Blood
.
2006
;
108
(
12
):
3736
-
3738
.
88.
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
.
89.
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
.
90.
Lamonica
JM
,
Deng
W
,
Kadauke
S
, et al
.
Bromodomain protein Brd3 associates with acetylated GATA1 to promote its chromatin occupancy at erythroid target genes
.
Proc Natl Acad Sci USA
.
2011
;
108
(
22
):
E159
-
E168
.
91.
DeVilbiss
AW
,
Boyer
ME
,
Bresnick
EH
.
Establishing a hematopoietic genetic network through locus-specific integration of chromatin regulators
.
Proc Natl Acad Sci USA
.
2013
;
110
(
36
):
E3398
-
E3407
.
92.
DeVilbiss
AW
,
Sanalkumar
R
,
Hall
BD
,
Katsumura
KR
,
de Andrade
IF
,
Bresnick
EH
.
Epigenetic determinants of erythropoiesis: role of the histone methyltransferase SetD8 in promoting erythroid cell maturation and survival
.
Mol Cell Biol
.
2015
;
35
(
12
):
2073
-
2087
.
93.
Kuzmichev
A
,
Nishioka
K
,
Erdjument-Bromage
H
,
Tempst
P
,
Reinberg
D
.
Histone methyltransferase activity associated with a human multiprotein complex containing the Enhancer of Zeste protein
.
Genes Dev
.
2002
;
16
(
22
):
2893
-
2905
.
94.
Xu
J
,
Shao
Z
,
Li
D
, et al
.
Developmental control of polycomb subunit composition by GATA factors mediates a switch to non-canonical functions
.
Mol Cell
.
2015
;
57
(
2
):
304
-
316
.
95.
Kim
S-I
,
Bultman
SJ
,
Kiefer
CM
,
Dean
A
,
Bresnick
EH
.
BRG1 requirement for long-range interaction of a locus control region with a downstream promoter
.
Proc Natl Acad Sci USA
.
2009
;
106
(
7
):
2259
-
2264
.
96.
Kim
SI
,
Bresnick
EH
,
Bultman
SJ
.
BRG1 directly regulates nucleosome structure and chromatin looping of the alpha globin locus to activate transcription
.
Nucleic Acids Res
.
2009
;
37
(
18
):
6019
-
6027
.
97.
Im
H
,
Grass
JA
,
Johnson
KD
, et al
.
Chromatin domain activation via GATA-1 utilization of a small subset of dispersed GATA motifs within a broad chromosomal region
.
Proc Natl Acad Sci USA
.
2005
;
102
(
47
):
17065
-
17070
.
98.
Pope
NJ
,
Bresnick
EH
.
Differential coregulator requirements for function of the hematopoietic transcription factor GATA-1 at endogenous loci
.
Nucleic Acids Res
.
2010
;
38
(
7
):
2190
-
2200
.
99.
Pope
NJ
,
Bresnick
EH
.
Establishment of a cell-type-specific genetic network by the mediator complex component Med1
.
Mol Cell Biol
.
2013
;
33
(
10
):
1938
-
1955
.
100.
Stumpf
M
,
Waskow
C
,
Krötschel
M
, et al
.
The mediator complex functions as a coactivator for GATA-1 in erythropoiesis via subunit Med1/TRAP220 [published correction appears in Proc Natl Acad Sci USA. 2007;104(4):1442]
.
Proc Natl Acad Sci USA
.
2006
;
103
(
49
):
18504
-
18509
.
101.
Kagey
MH
,
Newman
JJ
,
Bilodeau
S
, et al
.
Mediator and cohesin connect gene expression and chromatin architecture
.
Nature
.
2010
;
467
(
7314
):
430
-
435
.
102.
Stumpf
M
,
Yue
X
,
Schmitz
S
,
Luche
H
,
Reddy
JK
,
Borggrefe
T
.
Specific erythroid-lineage defect in mice conditionally deficient for Mediator subunit Med1
.
Proc Natl Acad Sci USA
.
2010
;
107
(
50
):
21541
-
21546
.
103.
Wadman
IA
,
Osada
H
,
Grütz
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
.
104.
Xu
Z
,
Meng
X
,
Cai
Y
,
Liang
H
,
Nagarajan
L
,
Brandt
SJ
.
Single-stranded DNA-binding proteins regulate the abundance of LIM domain and LIM domain-binding proteins
.
Genes Dev
.
2007
;
21
(
8
):
942
-
955
.
105.
Hoang
T
,
Lambert
JA
,
Martin
R
.
SCL/TAL1 in Hematopoiesis and Cellular Reprogramming
.
Curr Top Dev Biol
.
2016
;
118
:
163
-
204
.
106.
Love
PE
,
Warzecha
C
,
Li
L
.
Ldb1 complexes: the new master regulators of erythroid gene transcription
.
Trends Genet
.
2014
;
30
(
1
):
1
-
9
.
107.
Hewitt
KJ
,
Johnson
KD
,
Gao
X
,
Keles
S
,
Bresnick
EH
.
The hematopoietic stem and progenitor cell cistrome: GATA factor-dependent cis-regulatory mechanisms
.
Curr Top Dev Biol
.
2016
;
118
:
45
-
76
.
108.
Hewitt
KJ
,
Kim
DH
,
Devadas
P
, et al
.
Hematopoietic signaling mechanism revealed from a stem/progenitor cell cistrome
.
Mol Cell
.
2015
;
59
(
1
):
62
-
74
.
109.
Tripic
T
,
Deng
W
,
Cheng
Y
, et al
.
SCL and associated proteins distinguish active from repressive GATA transcription factor complexes
.
Blood
.
2009
;
113
(
10
):
2191
-
2201
.
110.
Beck
D
,
Thoms
JA
,
Perera
D
, et al
.
Genome-wide analysis of transcriptional regulators in human HSPCs reveals a densely interconnected network of coding and noncoding genes
.
Blood
.
2013
;
122
(
14
):
e12
-
e22
.
111.
Snow
JW
,
Trowbridge
JJ
,
Fujiwara
T
, et al
.
A single cis element maintains repression of the key developmental regulator Gata2
.
PLoS Genet
.
2010
;
6
(
9
):
e1001103
.
112.
Snow
JW
,
Trowbridge
JJ
,
Johnson
KD
, et al
.
Context-dependent function of “GATA switch” sites in vivo
.
Blood
.
2011
;
117
(
18
):
4769
-
4772
.
113.
Sanalkumar
R
,
Johnson
KD
,
Gao
X
, et al
.
Mechanism governing a stem cell-generating cis-regulatory element
.
Proc Natl Acad Sci USA
.
2014
;
111
(
12
):
E1091
-
E1100
.
114.
Lim
KC
,
Hosoya
T
,
Brandt
W
, et al
.
Conditional Gata2 inactivation results in HSC loss and lymphatic mispatterning
.
J Clin Invest
.
2012
;
122
(
10
):
3705
-
3717
.
115.
Wozniak
RJ
,
Boyer
ME
,
Grass
JA
,
Lee
Y
,
Bresnick
EH
.
Context-dependent GATA factor function: combinatorial requirements for transcriptional control in hematopoietic and endothelial cells
.
J Biol Chem
.
2007
;
282
(
19
):
14665
-
14674
.
116.
Khandekar
M
,
Brandt
W
,
Zhou
Y
, et al
.
A Gata2 intronic enhancer confers its pan-endothelia-specific regulation
.
Development
.
2007
;
134
(
9
):
1703
-
1712
.
117.
Crossley
M
,
Orkin
SH
.
Phosphorylation of the erythroid transcription factor GATA-1
.
J Biol Chem
.
1994
;
269
(
24
):
16589
-
16596
.
118.
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
.
119.
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
.
2006
;
107
(
3
):
907
-
915
.
120.
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
.
121.
Kadri
Z
,
Lefevre
C
,
Goupille
O
, et al
.
Erythropoietin and IGF-1 signaling synchronize cell proliferation and maturation during erythropoiesis
.
Genes Dev
.
2015
;
29
(
24
):
2603
-
2616
.
122.
Katsumura
KR
,
Yang
C
,
Boyer
ME
,
Li
L
,
Bresnick
EH
.
Molecular basis of crosstalk between oncogenic Ras and the master regulator of hematopoiesis GATA-2
.
EMBO Rep
.
2014
;
15
(
9
):
938
-
947
.
123.
Katsumura
KR
,
Ong
IM
,
DeVilbiss
AW
,
Sanalkumar
R
,
Bresnick
EH
.
GATA factor-dependent positive-feedback circuit in acute myeloid leukemia cells
.
Cell Reports
.
2016
;
16
(
9
):
2428
-
2441
.
124.
Menghini
R
,
Marchetti
V
,
Cardellini
M
, et al
.
Phosphorylation of GATA2 by Akt increases adipose tissue differentiation and reduces adipose tissue-related inflammation: a novel pathway linking obesity to atherosclerosis
.
Circulation
.
2005
;
111
(
15
):
1946
-
1953
.
125.
Nakajima
T
,
Kitagawa
K
,
Ohhata
T
, et al
.
Regulation of GATA-binding protein 2 levels via ubiquitin-dependent degradation by Fbw7: involvement of cyclin B-cyclin-dependent kinase 1-mediated phosphorylation of THR176 in GATA-binding protein 2
.
J Biol Chem
.
2015
;
290
(
16
):
10368
-
10381
.
126.
Maneechotesuwan
K
,
Xin
Y
,
Ito
K
, et al
.
Regulation of Th2 cytokine genes by p38 MAPK-mediated phosphorylation of GATA-3
.
J Immunol
.
2007
;
178
(
4
):
2491
-
2498
.
127.
Gao
X
,
Wu
T
,
Johnson
KD
, et al
.
GATA factor-G-protein-coupled receptor circuit suppresses hematopoiesis
.
Stem Cell Rep
.
2016
;
6
(
3
):
368
-
382
.
128.
Dailey
HA
,
Meissner
PN
.
Erythroid heme biosynthesis and its disorders
.
Cold Spring Harb Perspect Med
.
2013
;
3
(
4
):
a011676
.
129.
Ajioka
RS
,
Phillips
JD
,
Kushner
JP
.
Biosynthesis of heme in mammals
.
Biochim Biophys Acta
.
2006
;
1763
(
7
):
723
-
736
.
130.
Barupala
DP
,
Dzul
SP
,
Riggs-Gelasco
PJ
,
Stemmler
TL
.
Synthesis, delivery and regulation of eukaryotic heme and Fe-S cluster cofactors
.
Arch Biochem Biophys
.
2016
;
592
:
60
-
75
.
131.
Hunter
GA
,
Ferreira
GC
.
Molecular enzymology of 5-aminolevulinate synthase, the gatekeeper of heme biosynthesis
.
Biochim Biophys Acta
.
2011
;
1814
(
11
):
1467
-
1473
.
132.
Bishop
DF
,
Henderson
AS
,
Astrin
KH
.
Human delta-aminolevulinate synthase: assignment of the housekeeping gene to 3p21 and the erythroid-specific gene to the X chromosome
.
Genomics
.
1990
;
7
(
2
):
207
-
214
.
133.
Medlock
AE
,
Dailey
HA
.
Regulation of mammalian heme biosynthesis
. In:
Medlock
AE
,
Dailey
HA
, eds.
Tetrapyrroles: Birth, Life, and Death
.
New York
:
Landes Bioscience and Springer Science Business Media
;
2009
:
116
-
127
.
134.
Tanimura
N
,
Miller
E
,
Igarashi
K
, et al
.
Mechanism governing heme synthesis reveals a GATA factor/heme circuit that controls differentiation
.
EMBO Rep
.
2016
;
17
(
2
):
249
-
265
.
135.
Rouault
TA
,
Tong
WH
.
Iron-sulphur cluster biogenesis and mitochondrial iron homeostasis
.
Nat Rev Mol Cell Biol
.
2005
;
6
(
4
):
345
-
351
.
136.
Iwai
K
,
Klausner
RD
,
Rouault
TA
.
Requirements for iron-regulated degradation of the RNA binding protein, iron regulatory protein 2
.
EMBO J
.
1995
;
14
(
21
):
5350
-
5357
.
137.
Iwai
K
,
Drake
SK
,
Wehr
NB
, et al
.
Iron-dependent oxidation, ubiquitination, and degradation of iron regulatory protein 2: implications for degradation of oxidized proteins
.
Proc Natl Acad Sci USA
.
1998
;
95
(
9
):
4924
-
4928
.
138.
Chen
JJ
.
Translational control by heme-regulated eIF2α kinase during erythropoiesis
.
Curr Opin Hematol
.
2014
;
21
(
3
):
172
-
178
.
139.
Katsumura
KR
,
DeVilbiss
AW
,
Pope
NJ
,
Johnson
KD
,
Bresnick
EH
.
Transcriptional mechanisms underlying hemoglobin synthesis
.
Cold Spring Harb Perspect Med
.
2013
;
3
(
9
):
a015412
.
140.
Xu
J
,
Sankaran
VG
,
Ni
M
, et al
.
Transcriptional silencing of gamma-globin by BCL11A involves long-range interactions and cooperation with SOX6
.
Genes Dev
.
2010
;
24
(
8
):
783
-
798
.
141.
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
.
142.
Sun
J
,
Brand
M
,
Zenke
Y
,
Tashiro
S
,
Groudine
M
,
Igarashi
K
.
Heme regulates the dynamic exchange of Bach1 and NF-E2-related factors in the Maf transcription factor network
.
Proc Natl Acad Sci USA
.
2004
;
101
(
6
):
1461
-
1466
.
143.
Shoval
O
,
Alon
U
.
SnapShot: network motifs
.
Cell
.
2010
;
143
(
2
):
326
-
326.e1
.
144.
Keel
SB
,
Doty
RT
,
Yang
Z
, et al
.
A heme export protein is required for red blood cell differentiation and iron homeostasis
.
Science
.
2008
;
319
(
5864
):
825
-
828
.
145.
Yang
Z
,
Keel
SB
,
Shimamura
A
, et al
.
Delayed globin synthesis leads to excess heme and the macrocytic anemia of Diamond Blackfan anemia and del(5q) myelodysplastic syndrome
.
Sci Transl Med
.
2016
;
8
(
338
):
338ra67
.
146.
McIver
SC
,
Kang
YA
,
DeVilbiss
AW
, et al
.
The exosome complex establishes a barricade to erythroid maturation
.
Blood
.
2014
;
124
(
14
):
2285
-
2297
.
147.
Kilchert
C
,
Wittmann
S
,
Vasiljeva
L
.
The regulation and functions of the nuclear RNA exosome complex
.
Nat Rev Mol Cell Biol
.
2016
;
17
(
4
):
227
-
239
.
148.
McIver
SC
,
Katsumura
KR
,
Davids
E
, et al
.
Exosome complex orchestrates developmental signaling to balance proliferation and differentiation during erythropoiesis
.
eLife
.
2016
;
5
:
e17877
.
149.
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
.
150.
Hattangadi
SM
,
Wong
P
,
Zhang
L
,
Flygare
J
,
Lodish
HF
.
From stem cell to red cell: regulation of erythropoiesis at multiple levels by multiple proteins, RNAs, and chromatin modifications
.
Blood
.
2011
;
118
(
24
):
6258
-
6268
.
151.
Shimizu
R
,
Yamamoto
M
.
GATA-related hematologic disorders
.
Exp Hematol
.
2016
;
44
(
8
):
696
-
705
.
152.
Malinge
S
,
Izraeli
S
,
Crispino
JD
.
Insights into the manifestations, outcomes, and mechanisms of leukemogenesis in Down syndrome
.
Blood
.
2009
;
113
(
12
):
2619
-
2628
.
153.
Hahn
CN
,
Chong
CE
,
Carmichael
CL
, et al
.
Heritable GATA2 mutations associated with familial myelodysplastic syndrome and acute myeloid leukemia
.
Nat Genet
.
2011
;
43
(
10
):
1012
-
1017
.
154.
Hsu
AP
,
Sampaio
EP
,
Khan
J
, et al
.
Mutations in GATA2 are associated with the autosomal dominant and sporadic monocytopenia and mycobacterial infection (MonoMAC) syndrome
.
Blood
.
2011
;
118
(
10
):
2653
-
2655
.
155.
Dickinson
RE
,
Griffin
H
,
Bigley
V
, et al
.
Exome sequencing identifies GATA-2 mutation as the cause of dendritic cell, monocyte, B and NK lymphoid deficiency
.
Blood
.
2011
;
118
(
10
):
2656
-
2658
.
156.
Ostergaard
P
,
Simpson
MA
,
Connell
FC
, et al
.
Mutations in GATA2 cause primary lymphedema associated with a predisposition to acute myeloid leukemia (Emberger syndrome)
.
Nat Genet
.
2011
;
43
(
10
):
929
-
931
.
157.
Hsu
AP
,
Johnson
KD
,
Falcone
EL
, et al
.
GATA2 haploinsufficiency caused by mutations in a conserved intronic element leads to MonoMAC syndrome
.
Blood
.
2013
;
121
(
19
):
3830
-
3837
.
158.
Spinner
MA
,
Sanchez
LA
,
Hsu
AP
, et al
.
GATA2 deficiency: a protean disorder of hematopoiesis, lymphatics and immunity
.
Blood
.
2014
;
123
(
6
):
809
-
821
.
159.
Zhang
SJ
,
Ma
LY
,
Huang
QH
, et al
.
Gain-of-function mutation of GATA-2 in acute myeloid transformation of chronic myeloid leukemia
.
Proc Natl Acad Sci USA
.
2008
;
105
(
6
):
2076
-
2081
.
160.
Greif
PA
,
Dufour
A
,
Konstandin
NP
, et al
.
GATA2 zinc finger 1 mutations associated with biallelic CEBPA mutations define a unique genetic entity of acute myeloid leukemia
.
Blood
.
2012
;
120
(
2
):
395
-
403
.
161.
Green
CL
,
Tawana
K
,
Hills
RK
, et al
.
GATA2 mutations in sporadic and familial acute myeloid leukaemia patients with CEBPA mutations
.
Br J Haematol
.
2013
;
161
(
5
):
701
-
705
.
162.
Theis
F
,
Corbacioglu
A
,
Gaidzik
VI
, et al
.
Clinical impact of GATA2 mutations in acute myeloid leukemia patients harboring CEBPA mutations: a study of the AML study group
.
Leukemia
.
2016
;
30
(
11
):
2248
-
2250
.
163.
Ping
N
,
Sun
A
,
Song
Y
, et al
.
Exome sequencing identifies highly recurrent somatic GATA2 and CEBPA mutations in acute erythroid leukemia
.
Leukemia
.
2017
;
31
(
1
):
195
-
202
.
164.
Nichols
KE
,
Crispino
JD
,
Poncz
M
, et al
.
Familial dyserythropoietic anaemia and thrombocytopenia due to an inherited mutation in GATA1
.
Nat Genet
.
2000
;
24
(
3
):
266
-
270
.
165.
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
(
5
):
2238
-
2247
.
166.
Hasegawa
A
,
Kaneko
H
,
Ishihara
D
, et al
.
GATA1 binding kinetics on conformation-specific binding sites elicit differential transcriptional regulation
.
Mol Cell Biol
.
2016
;
36
(
16
):
2151
-
2167
.
167.
Gröschel
S
,
Sanders
MA
,
Hoogenboezem
R
, et al
.
A single oncogenic enhancer rearrangement causes concomitant EVI1 and GATA2 deregulation in leukemia
.
Cell
.
2014
;
157
(
2
):
369
-
381
.
168.
Yamazaki
H
,
Suzuki
M
,
Otsuki
A
, et al
.
A remote GATA2 hematopoietic enhancer drives leukemogenesis in inv(3)(q21;q26) by activating EVI1 expression
.
Cancer Cell
.
2014
;
25
(
4
):
415
-
427
.
169.
Wlodarski
MW
,
Hirabayashi
S
,
Pastor
V
, et al
;
EWOG-MDS
.
Prevalence, clinical characteristics, and prognosis of GATA2-related myelodysplastic syndromes in children and adolescents
.
Blood
.
2016
;
127
(
11
):
1387
-
1397, quiz 1518
.
170.
Towatari
M
,
Ciro
M
,
Ottolenghi
S
,
Tsuzuki
S
,
Enver
T
.
Involvement of mitogen-activated protein kinase in the cytokine-regulated phosphorylation of transcription factor GATA-1
.
Hematol J
.
2004
;
5
(
3
):
262
-
272
.
171.
Towatari
M
,
May
GE
,
Marais
R
, et al
.
Regulation of GATA-2 phosphorylation by mitogen-activated protein kinase and interleukin-3
.
J Biol Chem
.
1995
;
270
(
8
):
4101
-
4107
.
172.
Hayakawa
F
,
Towatari
M
,
Ozawa
Y
,
Tomita
A
,
Privalsky
ML
,
Saito
H
.
Functional regulation of GATA-2 by acetylation
.
J Leukoc Biol
.
2004
;
75
(
3
):
529
-
540
.
173.
Fujiwara
T
,
Lee
HY
,
Sanalkumar
R
,
Bresnick
EH
.
Building multifunctionality into a complex containing master regulators of hematopoiesis
.
Proc Natl Acad Sci USA
.
2010
;
107
(
47
):
20429
-
20434
.
174.
Fasan
A
,
Eder
C
,
Haferlach
C
, et al
.
GATA2 mutations are frequent in intermediate-risk karyotype AML with biallelic CEBPA mutations and are associated with favorable prognosis
.
Leukemia
.
2013
;
27
(
2
):
482
-
485
.
175.
Grossmann
V
,
Haferlach
C
,
Nadarajah
N
, et al
.
CEBPA double-mutated acute myeloid leukaemia harbours concomitant molecular mutations in 76·8% of cases with TET2 and GATA2 alterations impacting prognosis
.
Br J Haematol
.
2013
;
161
(
5
):
649
-
658
.
176.
Pasquet
M
,
Bellanné-Chantelot
C
,
Tavitian
S
, et al
.
High frequency of GATA2 mutations in patients with mild chronic neutropenia evolving to MonoMac syndrome, myelodysplasia, and acute myeloid leukemia
.
Blood
.
2013
;
121
(
5
):
822
-
829
.
177.
Churpek
JE
,
Pyrtel
K
,
Kanchi
KL
, et al
.
Genomic analysis of germ line and somatic variants in familial myelodysplasia/acute myeloid leukemia
.
Blood
.
2015
;
126
(
22
):
2484
-
2490
.
178.
Dickinson
RE
,
Milne
P
,
Jardine
L
, et al
.
The evolution of cellular deficiency in GATA2 mutation
.
Blood
.
2014
;
123
(
6
):
863
-
874
.
Sign in via your Institution