The surge of human genetic information, enabled by increasingly facile and economically feasible genomic technologies, has accelerated discoveries on the relationship of germline genetic variation to hematologic diseases. For example, germline variation in GATA2, encoding a vital transcriptional regulator of multilineage hematopoiesis, creates a predisposition to bone marrow failure and acute myeloid leukemia termed GATA2 deficiency syndrome. More than 300 GATA2 variants representing missense, truncating, and noncoding enhancer mutations have been documented. Although these variants can diminish GATA2 expression and/or function, the functional ramifications of many variants are unknown. Studies using genetic rescue and knockin mouse systems have established that GATA2 mutations differentially affect molecular processes in distinct target genes and within a single target cell. Considering that target genes for a transcription factor can differ in sensitivity to altered levels of the factor, and transcriptional mechanisms are often cell type specific, the context-dependent consequences of GATA2 mutations in experimental systems portend the complex phenotypes and interindividual variation of GATA2 deficiency syndrome. This review documents GATA2 human genetics and the state of efforts to traverse from physiological insights to pathogenic mechanisms.

The discoveries of master regulatory transcription factors that control hematopoiesis have been rapidly followed, or even preceded, by evidence for causative roles in human blood diseases. Transcriptional circuitry governing hematopoietic system development and function generates almost infinite opportunities for hematopoiesis to go awry. This is exemplified by GATA1,1-4  Runx1,5,6  Etv6,7,8  and Scl/TAL1,9,10  as their dysregulation causes human diseases, including anemia, thrombocytopenia, bone marrow failure, and leukemia. After more than a decade of intensive study, strong evidence has emerged from human genetic analyses that GATA2 alterations also cause hematologic disease.

After discovery of the founding member of the GATA transcription factor family GATA1,1,2  GATA2 was shown by Orkin and colleagues in 1994 to be a vital determinant of multilineage hematopoiesis.11  Although high GATA2 expression correlates with disease severity in pediatric and adult acute myeloid leukemia (AML),12,13  analyses of GATA2 germline mutations definitively established its role in causing a complex disorder termed GATA2 deficiency syndrome.14-17  This syndrome involves immunodeficiency with monocytopenia; B cell, natural killer cell, and dendritic cell deficiencies; and common Mycobacterium, fungal, and viral infections. Patients with GATA2 deficiency syndrome may also exhibit lymphedema or monosomy 7.17-20 GATA2 mutations create an myelodysplastic syndrome/AML predisposition, and physiological GATA2 levels suppress bone marrow failure and leukemogenesis.13,21  The only potentially curative therapy is bone marrow transplantation.22,23  Chronic myelogenous leukemia and AML patients with somatic GATA2 mutations reportedly share select phenotypes with germline GATA2 deficiency syndrome,24,25 , albeit the differences between GATA2 deficiency syndrome and these leukemias are considerable.

GATA2 deficiency syndrome has a highly variable penetrance, with unpredictable presentations in children and adults.18,19,26,27  The enigmatic penetrance has profound therapeutic implications. GATA2 mutations in children and adults can be asymptomatic, despite having family members with GATA2 deficiency syndrome. Somatic mutations in a host of genes (RUNX1, ETV6, CEBPA, ASXL1, SETBP1, and STAG2) occur commonly with germline GATA2 mutations and may constitute pathogenic triggers.21,28-37  However, whether different constellations of somatic mutations, in concert with the GATA2 germline mutation, qualitatively or quantitatively influence pathogenesis is unknown. Infection or environmental stresses may establish or influence epigenetic mechanisms to trigger pathogenesis and dictate penetrance.37-39  Epigenetic plasticity is a double-edged sword, as it may promote disease progression40  or mitigate disease phenotypes.41 

Understanding GATA2 deficiency syndrome requires knowledge of GATA2-dependent mechanisms that sustain hematopoiesis and permit the hematopoietic system to adapt to diverse stresses. GATA2 promotes hematopoietic stem and progenitor cell (HSPC) generation and function11,42-48  and regulates erythroid and megakaryocytic precursors.13,49,50  In human embryonic stem cell systems, GATA2 overexpression promotes the conversion of hemogenic endothelium to hematopoietic precursors51  and maintains multipotent precursors.52  In addition to HSPCs, GATA2 is expressed in select differentiated cell types, including macrophages, endothelial, neuronal, and endocrine cells.49,51,53-57  Importantly, GATA2 regulation can vary greatly in different biological systems. For example, targeted ablation of the murine Gata2 +9.5 intronic enhancer strongly reduces Gata2 expression in HSPCs without affecting expression in brain.58  Targeted ablation of the far upstream Gata2 −77 enhancer abrogates the multilineage differentiation potential of fetal liver progenitor cells without affecting hematopoietic stem cell emergence,44  which is abrogated by the +9.5 enhancer deletion.43,45  Targeted ablation of the Gata2 −1.8 “GATA switch site” disrupts the mechanism that maintains Gata2 repression during erythroid differentiation, without affecting the initiation of Gata2 repression and Gata2 expression before repression.59  GATA2-instigated genetic networks vary in different cellular contexts and in the steady state vs stress.39,44-46,60  In progenitor cells, GATA2 activates expression of a cadre of mechanistically linked and diverse genes, including those encoding c-Kit receptor tyrosine kinase, c-Kit signaling facilitator Samd14, erythroid and megakaryocytic differentiation inducer GATA1, and histidine decarboxylase mediating histamine biosynthesis.13,61-64  During mouse embryogenesis, GATA2 suppresses expression of innate immune genes, endowing progenitors with multilineage differentiation potential.46  During erythrocyte development, a subset of GATA2-activated genes, including Gata2 itself, are repressed as GATA1 replaces GATA2 on chromatin, a process termed a GATA switch.49,53,65,66  Thus, in certain contexts, GATA1 and GATA2 function distinctly.

The foundation described here raises the question of whether GATA2 germline mutations uniformly disrupt GATA2 expression and function in all cellular contexts, or if the mutations compromise processes in select cellular contexts and/or a subset of GATA2-regulated processes within a single cell. In addition, does the phenotypic complexity of GATA2 deficiency syndrome reflect an amalgamation of hematopoietic cell–intrinsic and non–cell-autonomous defects arising solely from insufficient GATA2, or do certain GATA2 mutants exert ectopic activities? GATA2 germline and acquired mutations include coding and noncoding sequence deletions, frameshifts, and missense mutations. Of 343 GATA2 coding mutations designated “pathogenic,” “likely pathogenic,” and “uncertain significance” by ClinGen/ClinVar (Figure 1), 148 are deemed pathogenic, with 61% (90) altering sequences within the zinc finger domain. Although GATA2 N-finger function is unresolved, the C-finger mediates GATA motif binding.49  In GATA1, the N-finger mediates binding to the coregulator Friend of GATA167,68  and modulates DNA binding at complex sites containing composite GATA motifs.69  Pathogenic mutations external to the GATA2 zinc fingers (41) are predominantly frameshift or nonsense (28% of pathogenic mutations; 41 of 148). GATA2 zinc-finger domain mutations and external frameshift and nonsense mutations amount to 88% of the pathogenic mutations (131 of 148). Apart from nonsynonymous coding mutations, synonymous GATA2 coding mutations have also been described in a small number of patients. Despite unaltered protein composition, nucleotide changes in coding regions induced selective loss of messenger RNA expression and/or splicing errors.70  Phenotypic consequences resembled that of nonsynonymous mutations, and a single variant of synonymous germline mutation generated intrafamilial phenotypic variability.71 

Figure 1.

Human GATA2 coding mutations. GATA2 mutations that are designated “pathogenic,” “likely pathogenic,” and “uncertain significance” and their corresponding protein sequences. Top, GATA2 protein sequence depicting mutations external to the zinc finger domain. Bottom, enlargement of GATA2 zinc finger domain illustrating amino acids affected by mutations. Reported mutations are indicated in pink. N-ZF, N-finger; INT-ZF, inter-zinc finger spacer; C-ZF, C-finger. N- and C-fingers are also referred to as Zf1 and Zf2, respectively.

Figure 1.

Human GATA2 coding mutations. GATA2 mutations that are designated “pathogenic,” “likely pathogenic,” and “uncertain significance” and their corresponding protein sequences. Top, GATA2 protein sequence depicting mutations external to the zinc finger domain. Bottom, enlargement of GATA2 zinc finger domain illustrating amino acids affected by mutations. Reported mutations are indicated in pink. N-ZF, N-finger; INT-ZF, inter-zinc finger spacer; C-ZF, C-finger. N- and C-fingers are also referred to as Zf1 and Zf2, respectively.

Close modal

To develop principles underlying GATA2-dependent pathogenesis, it is instructive to consider the relationship between human mutations and posttranslational modifications that regulate GATA2 function. The MAPKs p38α and extracellular signal–regulated kinase can phosphorylate murine GATA2 at multiple sites (S73, S119, S192, S290, and S340), which are conserved between mouse and man (Figure 2A).62  The Ras-MAPK pathway mediates GATA2 multisite phosphorylation in a mechanism requiring S192 and promoted by amino acids 171-174,72  which conform to an extracellular signal–regulated kinase docking sequence termed a “DEF motif” (FXFP) (Figure 2).73,74  GATA2 multisite phosphorylation amplifies its activity to regulate select target genes. In Kasumi-1 AML cells, these genes include IL1β and CXCL2, which generate positive-feedback regulation of the RAS/MAPK-GATA2-IL1β/CXCL2 axis.72  CXCL2 promotes Kasumi-1 cell proliferation, and high CXCL2 expression correlates with poor prognosis of patients with AML. The DEF motif is also implicated in GATA2-mediated megakaryocyte generation from murine hemogenic endothelial cells ex vivo.75  In HEK293 and 3T3-F442A preadipocyte cells, insulin-dependent activation of the Akt kinase induces phosphorylation of GATA2 S401,76  and GRB10, a receptor tyrosine kinase–binding adapter protein, interferes with Akt-mediated GATA2 phosphorylation.77  Whether this signal-dependent mechanism operates in hematopoietic cells is unclear. CDK1 phosphorylates GATA2 at T176 and promotes GATA2 degradation by Fbw7.78  In vitro studies revealed that p300 and GCN5 acetylate GATA2 at K102, K281, K285, K334, K336, K389, K390, K399, K403, K405, K406, K408, and K409 residues (Figure 2B),79  and GATA1 was known to be acetylated.80  Functional analyses revealed that these residues enhance DNA binding in vitro and transactivation activity, and studies in Xenopus eggs implicated GATA2 acetylation in regulating primitive erythropoiesis.81  Two missense mutations reported in ClinGen/ClinVar occur at phosphorylation sites (T176P and S192F), one missense mutation is immediately at the 3′-side of the DEF motif (P175S), and 2 deletion mutations and 2 missense mutations occur at acetylation sites (K389_K390del, K390del, K390E, and K406M).

Figure 2.

Relationship of human clinical genetic mutations and GATA2 posttranslational modification sites. (A) Multisite GATA2 phosphorylation model. Extracellular signal–regulated kinase (ERK)/p38α docks on the GATA2 DEF motif, S192 is phosphorylated, and other serine residues (S73, S119, S290, and S340) are phosphorylated subsequently. (B) Top, GATA2 protein structure. N- and C-zinc fingers, DEF motif, phosphorylated residues, and acetylated residues are indicated. Bottom, amino acid sequence from 165 aa to 200 aa of GATA2 protein. Phosphorylated residues, DEF motif, and mutations reported by ClinVar/ClinGen are indicated.

Figure 2.

Relationship of human clinical genetic mutations and GATA2 posttranslational modification sites. (A) Multisite GATA2 phosphorylation model. Extracellular signal–regulated kinase (ERK)/p38α docks on the GATA2 DEF motif, S192 is phosphorylated, and other serine residues (S73, S119, S290, and S340) are phosphorylated subsequently. (B) Top, GATA2 protein structure. N- and C-zinc fingers, DEF motif, phosphorylated residues, and acetylated residues are indicated. Bottom, amino acid sequence from 165 aa to 200 aa of GATA2 protein. Phosphorylated residues, DEF motif, and mutations reported by ClinVar/ClinGen are indicated.

Close modal

Among the essential GATA2 noncoding sequences, GATA2 intron 5 (+9.5) enhancer can be mutated in GATA2 deficiency syndrome. These mutations disrupt +9.5 enhancer function, thus reducing GATA2 expression.13,38,39,58,82  The +9.5 enhancer,83-86  which is highly homologous to the human enhancer, is essential for hematopoiesis, HSPC generation/function, and hematopoietic regeneration upon myeloablative stress.39,43,45,58  The most commonly reported germline aberration in the +9.5 enhancer is a c.1017+572C>T transition mutation38,82  that destroys a motif binding members of the ETS transcription factor family.87,88  Although mice harboring this mutation develop into adults with largely normal steady-state hematopoiesis, the mutants are hypersensitive to 5-fluorouracil–mediated myeloablation.39  This phenotype reflects a hematopoietic cell–intrinsic defect in HSPC regeneration.38 

In aggregate, the frameshift, nonsense, and +9.5 enhancer mutations suggest a haploinsufficiency mechanism of pathogenesis. Certain missense mutations can generate proteins defective in naked DNA binding, transfection-based transactivation, and/or chromatin occupancy, also consistent with haploinsufficiency. However, transcription factor mutations can generate ectopic activities,89  which may not be revealed without comprehensive multi-omic analyses. Taken together with the granulopoiesis-inducing activity of GATA2 R307W90,91  and the many unanswered questions regarding GATA2 function at the genomic level, it is attractive to consider a model in which mutations inhibit certain, but not all, GATA2 activities, and mutation-induced ectopic activities further corrupt GATA2 mechanisms and cellular physiology. However, much more research is required to elucidate the molecular and cellular consequences of GATA2 disease mutations and, importantly, to determine if firm genotype–phenotype correlations can be established.

What are the most insightful assays to inform how human GATA2 mutations compromise GATA2 mechanisms to yield pathogenesis? Naked DNA binding has limitations, as affinities may or may not extrapolate to chromatin binding, based on influences from protein–protein interactions and protein posttranslational modifications. GATA factors assemble and integrate into multiprotein complexes on chromatin that may enable tethering of a DNA binding–defective mutant into the complex. Chromatin immunoprecipitation assays reveal whether a mutant loses capacity to occupy all or a subset of sites or acquires nonphysiological sites. Human GATA1 mutations can impair friend of GATA1 (FOG1) binding,4  and FOG1 facilitates chromatin occupancy in a locus-specific manner.49,92  Genome-wide analysis indicates that GATA1 mutations can impair occupancy at certain sites, concomitant with ectopic site acquisition.93  Some GATA2 disease mutants retain chromatin occupancy at certain sites.90  No evidence exists to suggest that transfection-based transactivation assays predict GATA factor–mediated activation or repression of target genes.

To surmount these complexities and limitations, a genetic rescue assay was developed in Gata2 −77 enhancer-mutant murine progenitor cells, in which GATA2 levels are ∼80% lower than wild-type progenitor cells.90,91  The −77 enhancer confers GATA2 expression in progenitors,44,46  and a 3q21q26 inversion in AML allows the −77 to induce MECOM1 as a leukemogenic mechanism.13,94,95  Using a retroviral-based expression strategy, GATA2 or a GATA2 disease mutant can be replaced at a near-physiological level, and the impact on target genes, differentiation, or any process of interest can be quantified. It is essential to compare wild-type and mutant proteins at normal expression levels, as transcription factor target genes can differ in sensitivity to reduced levels of the factor.96  In addition, in humans, there are multiple examples of disease phenotypes resulting from transcription factor haploinsufficiency.97  The C-finger germline T354M mutant can be stably expressed in cells, retains measurable naked DNA binding capacity15,98  and activity to regulate select target genes and to promote granulocytic differentiation in the rescue assay, albeit at reduced levels.90,91  An R307W N-finger mutation from patients with AML also can be stably expressed in cells and retains activity to regulate certain, but not all, GATA2 target genes. Although this mutant does not support erythroid progenitor function in the rescue assay, it retains the capacity to promote granulocytic differentiation at least as effectively as wild-type GATA2. A familial nine amino acid insertion between the zinc fingers abrogates target gene regulation and differentiation.91  The differential impact of GATA2 mutations on target gene expression is consistent with the existing paradigm that GATA factors function through multiple regulatory modes.49,99  Based on this mechanistic diversity, in which a cohort of GATA factor target genes may be regulated differently from another target gene cohort, a molecular aberration would not be expected to disrupt all mechanisms equivalently. This epitomizes the problem: how can a single assay inform pathogenic mechanisms that involve multiple genes controlled by distinct mechanisms? The differential functional consequences of GATA2 mutations illustrate the need to rigorously analyze relationships between GATA2 function in vivo with ex vivo and in vitro assay readouts; the most instructive assay readouts are unknown, and it is certainly possible that multiple assays will be required.

Although in vivo approaches are intrinsically low-throughput, they have considerable potential for modeling human GATA2-linked disease phenotypes, unveiling disease mechanisms, and potentially providing screening systems to achieve clinical/translational outcomes. The first Gata2 knockin mouse disease model was reported recently. This strain harbors a G320D mutant allele and biallelic Cebpa mutation, which recapitulates that detected in patients with acute erythroid leukemia, albeit the human mutations were somatic and not germline.100  Approximately 40% of the mice developed acute erythroid leukemia, illustrating the power of combining mutations to model disease phenotypes.

The aggregate data raise the question as to whether GATA2 disease mutations cause pathogenesis by attenuation of all GATA2-dependent processes, a subset of processes, acquisition of ectopic GATA2 activities, or a combination of these mechanisms. A mutation-sensitive process may be inextricably linked to a process not affected directly by the mutation, thus generating composite defects that can be difficult to deconvolute experimentally and therefore to fully understand. It will be crucial to systematically analyze a cohort of disease mutants with different assays and systems, including rescue assays ex vivo46,90,91  and knockin mice,101  to unveil GATA2 deficiency syndrome mechanisms. Furthermore, as mutations analogous to GATA2 disease mutations have been reported in genes encoding other GATA factors, and these mutations can be associated with human disease, insights from GATA2 mutations will almost certainly inform biology and pathology linked to other GATA factors.49  For example, the GATA1 R216Q mutation, detected in a patient with congenital erythropoietic porphyria, is analogous to the GATA2 R307W mutation.102 

Considering the thousands of GATA2 chromatin occupancy sites, cell type–specific GATA2 target genes, and context-dependent transcriptional regulatory modes, abrogating function at certain, but not all, target genes destroys stoichiometric relationships within the delicately balanced, GATA2-regulated genetic and protein networks. The aberrant networks may generate ectopic mechanisms that are deleterious to cellular physiology and not characteristic of normal hematopoiesis. Networks can be corrupted via multiple means, and the quantitative and qualitative impact of distinct lesions in network components may differ, creating phenotypic diversity. Alternatively, any aberration that creates a network imbalance might yield dominant phenotypes that are largely invariant across distinct molecular lesions. The impact of this problem extends beyond mechanism, as this may guide therapies.

From a therapeutic perspective, the objective to restore inadequate levels of GATA2 may be exceedingly challenging, without exacerbating pathological phenotypes. As noted earlier, high GATA2 correlates with poor-prognosis leukemia,12,103  and elevating GATA2 in murine bone marrow suppressed hematopoiesis in a transplant model.104  As an alternative or complement to bone marrow transplantation, it may be necessary to destroy the GATA2 mutant, repair the GATA2 mutation, supplant ectopically elevated mechanisms, or reinstate attenuated mechanisms. These promising approaches need to be intensively evaluated, yet each presents unique challenges that may or may not be surmountable.

While the depth of knowledge of GATA2 mechanisms and biological functions continues to increase, the phenotypic complexity and highly variable penetrance of GATA2 deficiency syndrome remains enigmatic. A simple explanation for this complexity is different GATA2 disease-linked mutations may create diverse molecular and cellular aberrations through quantitative differences in the severity of aberrations and/or qualitative differences. For example, certain aberrations may generate ectopic activity, whereas others diminish physiological activity, and there may be cases in which a given aberration elicits both consequences. Thus, multiple molecular paths to pathogenesis may exist. A common thread of distinct aberrations is their contribution to deconstruction of a finely balanced network, which reinforces the vital need to elucidate how such networks are established and maintained. Ascertaining whether GATA2 deficiency syndrome involves a predominant path or multiple paths to pathogenesis is an exceptionally high priority, as this mechanistic framework will guide the development of the most appropriate therapeutic options, and the outcomes will guide patient-specific precision medicine strategies.

This work was supported by the National Institutes of Health, National Institute of Diabetes and Digestive and Kidney Diseases (DK68634 and DK50107) (E.H.B.), the Carbone Cancer Center (P30CA014520), and the Edward P. Evans Foundation.

Contribution: All authors wrote the manuscript.

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

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

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.
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
.
4.
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
.
5.
Luo
X
,
Feurstein
S
,
Mohan
S
, et al
.
ClinGen Myeloid Malignancy Variant Curation Expert Panel recommendations for germline RUNX1 variants
.
Blood Adv
.
2019
;
3
(
20
):
2962
-
2979
.
6.
Song
WJ
,
Sullivan
MG
,
Legare
RD
, et al
.
Haploinsufficiency of CBFA2 causes familial thrombocytopenia with propensity to develop acute myelogenous leukaemia
.
Nat Genet
.
1999
;
23
(
2
):
166
-
175
.
7.
Zhang
MY
,
Churpek
JE
,
Keel
SB
, et al
.
Germline ETV6 mutations in familial thrombocytopenia and hematologic malignancy
.
Nat Genet
.
2015
;
47
(
2
):
180
-
185
.
8.
Noetzli
L
,
Lo
RW
,
Lee-Sherick
AB
, et al
.
Germline mutations in ETV6 are associated with thrombocytopenia, red cell macrocytosis and predisposition to lymphoblastic leukemia
.
Nat Genet
.
2015
;
47
(
5
):
535
-
538
.
9.
Hoang
T
,
Lambert
JA
,
Martin
R
.
SCL/TAL1 in hematopoiesis and cellular reprogramming
.
Curr Top Dev Biol
.
2016
;
118
:
163
-
204
.
10.
Begley
CG
,
Aplan
PD
,
Davey
MP
, et al
.
Chromosomal translocation in a human leukemic stem-cell line disrupts the T-cell antigen receptor delta-chain diversity region and results in a previously unreported fusion transcript
.
Proc Natl Acad Sci U S A
.
1989
;
86
(
6
):
2031
-
2035
.
11.
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
.
12.
Luesink
M
,
Hollink
IH
,
van der Velden
VH
, et al
.
High GATA2 expression is a poor prognostic marker in pediatric acute myeloid leukemia
.
Blood
.
2012
;
120
(
10
):
2064
-
2075
.
13.
Bresnick
EH
,
Johnson
KD
.
Blood disease-causing and -suppressing transcriptional enhancers: general principles and GATA2 mechanisms
.
Blood Adv
.
2019
;
3
(
13
):
2045
-
2056
.
14.
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
.
15.
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
.
16.
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
.
17.
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
.
18.
Spinner
MA
,
Sanchez
LA
,
Hsu
AP
, et al
.
GATA2 deficiency: a protean disorder of hematopoiesis, lymphatics, and immunity
.
Blood
.
2014
;
123
(
6
):
809
-
821
.
19.
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
.
20.
Wlodarski
MW
,
Sahoo
SS
,
Niemeyer
CM
.
Monosomy 7 in pediatric myelodysplastic syndromes
.
Hematol Oncol Clin North Am
.
2018
;
32
(
4
):
729
-
743
.
21.
Churpek
JE
,
Bresnick
EH
.
Transcription factor mutations as a cause of familial myeloid neoplasms
.
J Clin Invest
.
2019
;
129
(
2
):
476
-
488
.
22.
Parta
M
,
Shah
NN
,
Baird
K
, et al
.
Allogeneic hematopoietic stem cell transplantation for GATA2 deficiency using a busulfan-based regimen
.
Biol Blood Marrow Transplant
.
2018
;
24
(
6
):
1250
-
1259
.
23.
Hofmann
I
,
Avagyan
S
,
Stetson
A
, et al
.
Comparison of outcomes of myeloablative allogeneic stem cell transplantation for pediatric patients with bone marrow failure, myelodysplastic syndrome and acute myeloid leukemia with and without germline GATA2 mutations
.
Biol Blood Marrow Transplant
.
2020
;
26
(
6
):
1124
-
1130
.
24.
Alfayez
M
,
Wang
SA
,
Bannon
SA
, et al
.
Myeloid malignancies with somatic GATA2 mutations can be associated with an immunodeficiency phenotype
.
Leuk Lymphoma
.
2019
;
60
(
8
):
2025
-
2033
.
25.
Sekhar
M
,
Pocock
R
,
Lowe
D
, et al
.
Can somatic GATA2 mutation mimic germ line GATA2 mutation?
Blood Adv
.
2018
;
2
(
8
):
904
-
908
.
26.
McReynolds
LJ
,
Calvo
KR
,
Holland
SM
.
Germline GATA2 mutation and bone marrow failure
.
Hematol Oncol Clin North Am
.
2018
;
32
(
4
):
713
-
728
.
27.
Collin
M
,
Dickinson
R
,
Bigley
V
.
Haematopoietic and immune defects associated with GATA2 mutation
.
Br J Haematol
.
2015
;
169
(
2
):
173
-
187
.
28.
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
.
29.
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
.
30.
Drazer
MW
,
Kadri
S
,
Sukhanova
M
, et al
.
Prognostic tumor sequencing panels frequently identify germ line variants associated with hereditary hematopoietic malignancies
.
Blood Adv
.
2018
;
2
(
2
):
146
-
150
.
31.
Bödör
C
,
Renneville
A
,
Smith
M
, et al
.
Germ-line GATA2 p.THR354MET mutation in familial myelodysplastic syndrome with acquired monosomy 7 and ASXL1 mutation demonstrating rapid onset and poor survival
.
Haematologica
.
2012
;
97
(
6
):
890
-
894
.
32.
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
.
33.
Zhang
MY
,
Keel
SB
,
Walsh
T
, et al
.
Genomic analysis of bone marrow failure and myelodysplastic syndromes reveals phenotypic and diagnostic complexity
.
Haematologica
.
2015
;
100
(
1
):
42
-
48
.
34.
Keel
SB
,
Scott
A
,
Sanchez-Bonilla
M
, et al
.
Genetic features of myelodysplastic syndrome and aplastic anemia in pediatric and young adult patients
.
Haematologica
.
2016
;
101
(
11
):
1343
-
1350
.
35.
Fisher
KE
,
Hsu
AP
,
Williams
CL
, et al
.
Somatic mutations in children with GATA2-associated myelodysplastic syndrome who lack other features of GATA2 deficiency
.
Blood Adv
.
2017
;
1
(
7
):
443
-
448
.
36.
Bluteau
O
,
Sebert
M
,
Leblanc
T
, et al
.
A landscape of germ line mutations in a cohort of inherited bone marrow failure patients
.
Blood
.
2018
;
131
(
7
):
717
-
732
.
37.
Yamamoto
H
,
Hattori
H
,
Takagi
E
, et al
.
MonoMAC syndrome patient developing myelodysplastic syndrome following persistent EBV infection [in Japanese]
.
Rinsho Ketsueki
.
2018
;
59
(
3
):
315
-
322
.
38.
Soukup
AA
,
Bresnick
EH
.
GATA2 +9.5 enhancer: from principles of hematopoiesis to genetic diagnosis in precision medicine
.
Curr Opin Hematol
.
2020
;
27
(
3
):
163
-
171
.
39.
Soukup
AA
,
Zheng
Y
,
Mehta
C
, et al
.
Single-nucleotide human disease mutation inactivates a blood-regenerative GATA2 enhancer
.
J Clin Invest
.
2019
;
129
(
3
):
1180
-
1192
.
40.
Celton
M
,
Forest
A
,
Gosse
G
, et al
.
Epigenetic regulation of GATA2 and its impact on normal karyotype acute myeloid leukemia
.
Leukemia
.
2014
;
28
(
8
):
1617
-
1626
.
41.
Al Seraihi
AF
,
Rio-Machin
A
,
Tawana
K
, et al
.
GATA2 monoallelic expression underlies reduced penetrance in inherited GATA2-mutated MDS/AML
.
Leukemia
.
2018
;
32
(
11
):
2502
-
2507
.
42.
Tsai
FY
,
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
.
43.
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
.
44.
Johnson
KD
,
Kong
G
,
Gao
X
, et al
.
Cis-regulatory mechanisms governing stem and progenitor cell transitions
.
Sci Adv
.
2015
;
1
(
8
):
e1500503
.
45.
Mehta
C
,
Johnson
KD
,
Gao
X
, et al
.
Integrating enhancer mechanisms to establish a hierarchical blood development program
.
Cell Rep
.
2017
;
20
(
12
):
2966
-
2979
.
46.
Johnson
KD
,
Conn
DJ
,
Shishkova
E
, et al
.
Constructing and deconstructing GATA2-regulated cell fate programs to establish developmental trajectories
.
J Exp Med
.
2020
;
217
(
11
):
e20191526
.
47.
Rodrigues
NP
,
Janzen
V
,
Forkert
R
, et al
.
Haploinsufficiency of GATA-2 perturbs adult hematopoietic stem-cell homeostasis
.
Blood
.
2005
;
106
(
2
):
477
-
484
.
48.
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
.
49.
Katsumura
KR
,
Bresnick
EH
,
Group
GFM
;
GATA Factor Mechanisms Group
.
The GATA factor revolution in hematology
.
Blood
.
2017
;
129
(
15
):
2092
-
2102
.
50.
Huang
Z
,
Dore
LC
,
Li
Z
, et al
.
GATA-2 reinforces megakaryocyte development in the absence of GATA-1
.
Mol Cell Biol
.
2009
;
29
(
18
):
5168
-
5180
.
51.
Kang
H
,
Mesquitta
WT
,
Jung
HS
,
Moskvin
OV
,
Thomson
JA
,
Slukvin
II
.
GATA2 is dispensable for specification of hemogenic endothelium but promotes endothelial-to-hematopoietic transition
.
Stem Cell Reports
.
2018
;
11
(
1
):
197
-
211
.
52.
Zhou
Y
,
Zhang
Y
,
Chen
B
, et al
.
Overexpression of GATA2 enhances development and maintenance of human embryonic stem cell-derived hematopoietic stem cell-like progenitors
.
Stem Cell Reports
.
2019
;
13
(
1
):
31
-
47
.
53.
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
.
54.
Zhou
Y
,
Yamamoto
M
,
Engel
JD
.
GATA2 is required for the generation of V2 interneurons
.
Development
.
2000
;
127
(
17
):
3829
-
3838
.
55.
He
B
,
Lanz
RB
,
Fiskus
W
, et al
.
GATA2 facilitates steroid receptor coactivator recruitment to the androgen receptor complex
.
Proc Natl Acad Sci U S A
.
2014
;
111
(
51
):
18261
-
18266
.
56.
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
.
57.
Rubel
CA
,
Wu
SP
,
Lin
L
, et al
.
A Gata2-dependent transcription network regulates uterine progesterone responsiveness and endometrial function
.
Cell Rep
.
2016
;
17
(
5
):
1414
-
1425
.
58.
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
.
59.
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
.
60.
Linnemann
AK
,
O’Geen
H
,
Keles
S
,
Farnham
PJ
,
Bresnick
EH
.
Genetic framework for GATA factor function in vascular biology
.
Proc Natl Acad Sci U S A
.
2011
;
108
(
33
):
13641
-
13646
.
61.
Hewitt
KJ
,
Katsumura
KR
,
Matson
DR
, et al
.
GATA factor-regulated Samd14 enhancer confers red blood cell regeneration and survival in severe anemia
.
Dev Cell
.
2017
;
42
(
3
):
213
-
225.e214
.
62.
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
.
63.
Jing
H
,
Vakoc
CR
,
Ying
L
, et al
.
Exchange of GATA factors mediates transitions in looped chromatin organization at a developmentally regulated gene locus
.
Mol Cell
.
2008
;
29
(
2
):
232
-
242
.
64.
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
.
65.
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 U S A
.
2003
;
100
(
15
):
8811
-
8816
.
66.
Bresnick
EH
,
Lee
HY
,
Fujiwara
T
,
Johnson
KD
,
Keles
S
.
GATA switches as developmental drivers
.
J Biol Chem
.
2010
;
285
(
41
):
31087
-
31093
.
67.
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
.
68.
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
.
69.
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
.
70.
Kozyra
EJ
,
Pastor
VB
,
Lefkopoulos
S
, et al
.
Synonymous GATA2 mutations result in selective loss of mutated RNA and are common in patients with GATA2 deficiency [published online ahead of print 18 June 2020]
.
Leukemia
.
doi:10.1038/s41375-020-0899-5
.
71.
Fox
LC
,
Tan
M
,
Brown
AL
, et al
.
A synonymous GATA2 variant underlying familial myeloid malignancy with striking intrafamilial phenotypic variability [published online ahead of print 3 June 2020]
.
Br J Haematol
.
doi:10.1111/bjh.16819
.
72.
Katsumura
KR
,
Ong
IM
,
DeVilbiss
AW
,
Sanalkumar
R
,
Bresnick
EH
.
GATA factor-dependent positive-feedback circuit in acute myeloid leukemia cells
.
Cell Rep
.
2016
;
16
(
9
):
2428
-
2441
.
73.
Lee
T
,
Hoofnagle
AN
,
Kabuyama
Y
, et al
.
Docking motif interactions in MAP kinases revealed by hydrogen exchange mass spectrometry
.
Mol Cell
.
2004
;
14
(
1
):
43
-
55
.
74.
Murphy
LO
,
Smith
S
,
Chen
RH
,
Fingar
DC
,
Blenis
J
.
Molecular interpretation of ERK signal duration by immediate early gene products
.
Nat Cell Biol
.
2002
;
4
(
8
):
556
-
564
.
75.
Kitajima
K
,
Kanokoda
M
,
Nakajima
M
,
Hara
T
.
Domain-specific biological functions of the transcription factor Gata2 on hematopoietic differentiation of mouse embryonic stem cells
.
Genes Cells
.
2018
;
23
(
9
):
753
-
766
.
76.
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
.
77.
Bhattacharyya
S
,
Feferman
L
,
Tobacman
JK
.
Carrageenan inhibits insulin signaling through GRB10-mediated decrease in Tyr(P)-IRS1 and through inflammation-induced increase in Ser(P)307-IRS1
.
J Biol Chem
.
2015
;
290
(
17
):
10764
-
10774
.
78.
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
.
79.
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
.
80.
Lamonica
JM
,
Vakoc
CR
,
Blobel
GA
.
Acetylation of GATA-1 is required for chromatin occupancy
.
Blood
.
2006
;
108
(
12
):
3736
-
3738
.
81.
Dalgin
G
,
Goldman
DC
,
Donley
N
,
Ahmed
R
,
Eide
CA
,
Christian
JL
.
GATA-2 functions downstream of BMPs and CaM KIV in ectodermal cells during primitive hematopoiesis
.
Dev Biol
.
2007
;
310
(
2
):
454
-
469
.
82.
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
-
S7
.
83.
Grass
JA
,
Jing
H
,
Kim
SI
, et al
.
Distinct functions of dispersed GATA factor complexes at an endogenous gene locus
.
Mol Cell Biol
.
2006
;
26
(
19
):
7056
-
7067
.
84.
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
.
85.
Wozniak
RJ
,
Keles
S
,
Lugus
JJ
, et al
.
Molecular hallmarks of endogenous chromatin complexes containing master regulators of hematopoiesis
.
Mol Cell Biol
.
2008
;
28
(
21
):
6681
-
6694
.
86.
Khandekar
M
,
Brandt
W
,
Zhou
Y
, et al
.
A Gata2 intronic enhancer confers its pan-endothelia-specific regulation
.
Development
.
2007
;
134
(
9
):
1703
-
1712
.
87.
Hollenhorst
PC
,
McIntosh
LP
,
Graves
BJ
.
Genomic and biochemical insights into the specificity of ETS transcription factors
.
Annu Rev Biochem
.
2011
;
80
(
1
):
437
-
471
.
88.
Sumanas
S
,
Choi
K
.
ETS transcription factor ETV2/ER71/Etsrp in hematopoietic and vascular development
.
Curr Top Dev Biol
.
2016
;
118
:
77
-
111
.
89.
Brosh
R
,
Rotter
V
.
When mutants gain new powers: news from the mutant p53 field
.
Nat Rev Cancer
.
2009
;
9
(
10
):
701
-
713
.
90.
Katsumura
KR
,
Mehta
C
,
Hewitt
KJ
, et al
.
Human leukemia mutations corrupt but do not abrogate GATA-2 function
.
Proc Natl Acad Sci U S A
.
2018
;
115
(
43
):
E10109
-
E10118
.
91.
Cavalcante de Andrade Silva
M
,
Katsumura
KR
,
Mehta
C
,
Velloso
EDRP
,
Bresnick
EH
,
Godley
LA
.
Breaking the spatial constraint between neighboring zinc fingers: a new germline mutation in GATA2 deficiency syndrome. [published online ahead of print 14 April 2020]
.
Leukemia
.
doi:10.1038/s41375-020-0820-2
.
92.
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 U S A
.
2004
;
101
(
2
):
476
-
481
.
93.
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
.
94.
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
.
95.
Yamaoka
A
,
Suzuki
M
,
Katayama
S
,
Orihara
D
,
Engel
JD
,
Yamamoto
M
.
EVI1 and GATA2 misexpression induced by inv(3)(q21q26) contribute to megakaryocyte-lineage skewing and leukemogenesis
.
Blood Adv
.
2020
;
4
(
8
):
1722
-
1736
.
96.
Johnson
KD
,
Kim
SI
,
Bresnick
EH
.
Differential sensitivities of transcription factor target genes underlie cell type-specific gene expression profiles
.
Proc Natl Acad Sci USA
.
2006
;
103
(
43
):
15939
-
15944
.
97.
Seidman
JG
,
Seidman
C
.
Transcription factor haploinsufficiency: when half a loaf is not enough
.
J Clin Invest
.
2002
;
109
(
4
):
451
-
455
.
98.
Chong
CE
,
Venugopal
P
,
Stokes
PH
, et al
.
Differential effects on gene transcription and hematopoietic differentiation correlate with GATA2 mutant disease phenotypes
.
Leukemia
.
2018
;
32
(
1
):
194
-
202
.
99.
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
.
100.
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
.
101.
Di Genua
C
,
Valletta
S
,
Buono
M
, et al
.
C/EBPα and GATA-2 mutations induce bilineage acute erythroid leukemia through transformation of a neomorphic neutrophil-erythroid progenitor
.
Cancer Cell
.
2020
;
37
(
5
):
690
-
704.e8
.
102.
Tubman
VN
,
Levine
JE
,
Campagna
DR
, et al
.
X-linked gray platelet syndrome due to a GATA1 Arg216Gln mutation
.
Blood
.
2007
;
109
(
8
):
3297
-
3299
.
103.
Vicente
C
,
Vazquez
I
,
Conchillo
A
, et al
.
Overexpression of GATA2 predicts an adverse prognosis for patients with acute myeloid leukemia and it is associated with distinct molecular abnormalities
.
Leukemia
.
2012
;
26
(
3
):
550
-
554
.
104.
Persons
DA
,
Allay
JA
,
Allay
ER
, et al
.
Enforced expression of the GATA-2 transcription factor blocks normal hematopoiesis
.
Blood
.
1999
;
93
(
2
):
488
-
499
.