Next Article in Journal
Overcoming Sperm Cell Survival Challenges Cryopreserved in Nanoliter Volumes
Previous Article in Journal
Serum Cereblon (CRBN) Levels Predict Long Term Post- Lenalidomide-Dexamethasone Survival in Multiple Myeloma (MM) Patients and Correlate with Disease Characteristics
Previous Article in Special Issue
Resolution of RHCE Haplotype Ambiguities in Transfusion Settings
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Epigenetic Regulation of Erythropoiesis: From Developmental Programs to Therapeutic Targets

by
Ninos Ioannis Vasiloudis
1,2,
Kiriaki Paschoudi
1,2,
Christina Beta
1,2,
Grigorios Georgolopoulos
1 and
Nikoletta Psatha
1,*
1
Department of Genetics, Development and Molecular Biology, School of Biology, Aristotle University of Thessaloniki, 54124 Thessaloniki, Greece
2
Gene and Cell Therapy Center, Hematology Clinic-Bone Marrow Transplantation Unit, “George Papanikolaou” Hospital, 57010 Thessaloniki, Greece
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2025, 26(13), 6342; https://doi.org/10.3390/ijms26136342
Submission received: 2 May 2025 / Revised: 27 June 2025 / Accepted: 27 June 2025 / Published: 30 June 2025
(This article belongs to the Special Issue New Advances in Erythrocyte Biology and Functions)

Abstract

Erythropoiesis, the process driving the differentiation of hematopoietic stem and progenitor cells to mature erythrocytes, unfolds through tightly orchestrated developmental stages, each defined by profound epigenetic remodeling. From the initial commitment of hematopoietic progenitors to the terminal enucleation of erythrocytes, dynamic changes in chromatin accessibility, transcription factor occupancy, and three-dimensional genome architecture govern lineage specification and stage-specific gene expression. Advances in our understanding of the regulatory genome have uncovered how non-coding elements, including enhancers, silencers, and insulators, shape the transcriptional landscape of erythroid cells. These elements work in concert with lineage-determining transcription factors to establish and maintain erythroid identity. Disruption of these epigenetic programs—whether by inherited mutations, somatic alterations, or environmental stress—can lead to a wide range of hematologic disorders. Importantly, this growing knowledge base has opened new therapeutic avenues, enabling the development of precision tools that target regulatory circuits to correct gene expression. These include epigenetic drugs, enhancer-targeted genome editing, and lineage-restricted gene therapies that leverage endogenous regulatory logic. As our understanding of erythroid epigenomics deepens, so too does our ability to design rational, cell-type-specific interventions for red blood cell disorders.

1. Introduction

Erythropoiesis is the vital process responsible for the production of red blood cells (RBCs), ensuring the continuous supply of cells capable of transporting oxygen from the lungs to tissues throughout the body. To meet the changing oxygen demands and maintain adequate red cell numbers, erythropoiesis undergoes dynamic regulation during development, progressing through distinct embryonic, fetal, and adult stages. During the embryonic stage, primitive erythropoiesis originates in the yolk sac, producing large, nucleated RBCs that circulate briefly within the early embryo. These primitive cells mature within the vasculature and ultimately undergo enucleation to support embryonic growth. Conversely, definitive erythropoiesis, located in the fetal liver and later in the bone marrow, generates smaller, enucleated RBCs that maintain oxygen delivery throughout fetal development and after birth [1,2,3,4,5].
Both primitive and definitive erythroid lineages arise from committed progenitor cells that progress through distinct precursor stages, culminating in enucleated mature RBCs [6]. Alongside the developmental transitions in erythropoiesis, profound changes occur at each stage of erythroid differentiation, characterized by dramatic cellular and molecular shifts. These transitions include global alterations in chromatin structure, dynamic remodeling of the transcriptional landscape, and the progressive activation and silencing of specific gene networks, all of which are orchestrated by a tightly controlled and stage-specific epigenetic regulatory network [6,7,8]. Owing to the extensive and highly coordinated changes in gene expression and chromatin dynamics, erythropoiesis provides an exceptional model system for studying the epigenome and the gene regulatory networks (GRNs) that govern lineage commitment and differentiation. Insights gained from these studies have not only deepened our understanding of normal erythroid biology but have also revealed critical mechanisms underlying various erythroid disorders such as β-thalassemia, sickle cell disease, Diamond–Blackfan anemia, congenital dyserythropoietic anemias, and myelodysplastic syndromes with erythroid predominance. These findings offer novel diagnostic markers and therapeutic targets to combat these diseases. In this review, we sought to highlight key aspects of the epigenetic regulation of erythropoiesis, discuss how these mechanisms contribute to erythroid health and disease, and explore their potential therapeutic implications.

2. Dynamic Chromatin Remodeling During Development and Lineage Commitment

During development and lineage commitment, cells must undergo profound changes in gene expression to acquire and maintain their specific identities. A central mechanism driving these transitions is extensive chromatin remodeling [9] and dynamic changes in chromatin interactions [10], which regulate the accessibility of genomic regions to transcriptional machinery and orchestrate precise gene activation and repression programs. In the context of erythropoiesis, chromatin remodeling is essential for guiding multipotent hematopoietic stem and progenitor cells (HSPCs) towards a committed erythroid fate. We and others have previously shown that this process involves coordinated changes in chromatin structure coupled with transcription factor binding that progressively restricts cellular potential and reinforces erythroid-specific gene expression programs [8,11].

2.1. The β-Globin Locus: A Model of Developmental Gene Regulation

Among the many genomic regions undergoing developmental dynamic epigenetic remodeling in erythropoiesis, the β-globin locus remains the most thoroughly characterized, offering a paradigm for understanding how chromatin architecture and enhancer/promoter communication regulate gene expression throughout development. During mammalian development, shifting oxygen demands trigger two hemoglobin switches, governed in part by the human β-globin locus through the stage-specific activation of ε, Gγ, Aγ, δ, and/or β genes located on chromosome 11 [12,13]. The expression of globin genes is tightly regulated during the developmental stages through the locus control region (LCR). Beta-globin LCR is located 6 to 20 kb upstream of the beta-globin genes and consists of five DNase I hypersensitive sites (Figure 1).
Each of these regions has a core sequence of 250 nucleotides filled with transcription factor motifs [14]. It is known that the HS1-HS4 have cell-type-specific enhancer activity that increases the expression of beta-like genes, whereas the HS5 acts as an insulator and has a CTCF (CCCTC-binding factor) binding site [13,15]. In the embryonic stage, the ε-globin gene (HBE1) is predominantly expressed, facilitated by an open chromatin structure at the LCR, characterized both by DNase I hypersensitivity [15,16] and active histone modifications such as H3K9 acetylation and H3K4 methylation [17]. At this stage, the LCR engages in chromatin looping interactions with the ε-globin promoter, a mechanism mediated by transcription factors including GATA1 and NF-E2 [18,19,20,21,22,23,24,25,26,27] (Figure 1).
As development proceeds to the fetal stage, transcriptional activity shifts to the γ-globin genes (HBG1 and HBG2), while the LCR maintains an accessible configuration but reorients its enhancer/promoter interactions toward the γ-globin promoters [20]. This transition is marked by enriched H3K4me3 and histone acetylation at γ-globin regulatory regions, although repressive factors such as BCL11A begin to emerge later in fetal erythropoiesis, preparing for the eventual silencing of fetal globin expression [28,29,30].
In the adult stage, expression is predominantly restricted to the β-globin gene (HBB) (Figure 1), accompanied by repressive epigenetic modifications—such as H3K9me3 and DNA methylation—at embryonic and fetal gene promoters, and persistent activation marks at the β-globin promoter [28,30]. Notably, whether DNA methylation at embryonic and fetal globin promoters is a primary driver of gene silencing or a secondary consequence of transcriptional repression remains under active investigation; evidence supports both models, suggesting that promoter methylation may both reinforce and stabilize the hemoglobin switch once initiated by transcriptional regulators [31,32,33].

2.2. Chromatin Architecture and the Role of CTCF

In addition to transcriptional and epigenetic repression, higher-order chromatin organization, shaped by architectural proteins like CTCF, is crucial for the developmental regulation of the β-globin locus. CTCF is a key regulator of transcription, chromatin architecture, and genome organization. Notably, CTCF binding can be modulated by DNA methylation, linking changes in chromatin architecture to epigenetic state. Immortalized cells show widespread loss of CTCF binding linked to increased methylation despite upregulated CTCF expression, suggesting a compensatory mechanism that preserves global occupancy levels [34]. This methylation-dependent regulation of CTCF is particularly relevant to developmental gene control, such as in the hemoglobin switch. In this context, promoter methylation at embryonic and fetal globin genes may both initiate and reinforce transcriptional silencing. Furthermore, CTCF and cohesin complexes mediate higher-order chromatin architecture, promoting the formation of insulated domains that facilitate exclusive LCR-β-globin interactions [35,36,37].

2.3. Regulatory Mechanisms at the α-Globin Locus

Similar to the β-globins, the chromatin dynamics of the α-globin locus have been extensively studied. The alpha-globin locus is located at chromosome 16 in a G-C, Alu-dense, gene-dense segment of DNA. The alpha-globin locus encodes the alpha- and ζ-globin genes. The alpha-globin promoters have a NF-Y binding site (CAAT box) and binding sites for the transcription factors GATA1 and KLF1 [38,39]. Moreover, the promoters of all genes in alpha-globin locus contain GC-islands. The alpha-like genes are distal regulated by four cis-regulatory enhancer elements located 10 to 48kb upstream of the genes, known as alpha-globin LCR. There are four distinct DNaseI hypersensitive sites in the a-globin LCR, named MCS-R4 (multispecies conserved sequence R4), MCS-R3, MCS-R2, and MCS-R1 [40] (Figure 2). Among them the MCS-R2, known also as HS-40, is the most critical for alpha-globin expression.
During the erythroid differentiation, the alpha-globin expression is tightly regulated through strong interactions between the alpha-globin LCR and the promoter of the corresponding gene through looping formation [41]. The transcription factors GATA1, nuclear factor-erythroid 2 (NF-E2), stem cell leukemia (SCL) pentameric complex, and KLF1 play crucial roles in alpha-globin expression [39]. Notably, the alpha-globin expression is precisely regulated by specific epigenetics mechanisms. In non-erythroid cells, the polycomb repressive complex 2 (PRC2) binds to the alpha-globin promoter and induced H3K27me3 marks, promoting transcriptional silencing [42]. Conversely, during erythroid differentiation, the PRC2 is detached and the methylation marks are removed by demethylation enzymes, such as Jumonji domain-containing 3 and histone lysine demethylase (JMJD3) enzyme [43]. Capture-C experiments have shown that the α-globin locus also undergoes dramatic spatial re-arrangement during mouse erythroid differentiation [44]. Further studies have provided structural models of the different states of the region. The inactive α-globin locus in embryonic stem cells (ESCs) is characterized by a compact globular structure dominated by non-specific DNA-DNA contacts. In contrast, in the erythroid cells the region is re-arranged in a highly structured folder hairpin conformation which facilitates the contact between the α-globin genes and their enhancers in cis (i.e., on the same chromosome) [45], where increased frequency of enhancer/promoter interactions correlates with upregulation of α-globin expression [46].

2.4. Sequence Orientation and Cohesin Flow in Gene Regulation

Enhancers activate transcription when brought into spatial proximity with their target promoters. Unlike individual enhancers, which typically function in an orientation-independent manner, super-enhancers (SEs) and LCRs can exhibit orientation-dependent activity. In the case of the α-globin locus, active enhancers coincide with peaks of the cohesin complex and its loader Nipbl, indicating these regions may serve as cohesin entry sites in erythroid cells [47,48,49]. Cohesin translocation along chromatin can be impeded by elements like replication machinery or CTCF-bound insulators. The orientation of CTCF binding is critical; when the N-terminus of CTCF faces the enhancer, cohesin movement, and consequently enhancer function are preferentially blocked, leading to reduced gene expression [50]. Experimental inversion of CTCF sites and SEs within the α-globin cluster has demonstrated that the relative orientation of both elements influences transcriptional output [51]. Similar observations have been made at the β-globin locus, where inversion of the β-globin locus control region (LCR) significantly reduces β-globin expression [52]. Furthermore, the β-globin LCR, when inserted in different orientations among housekeeping genes, activates different sets of genes depending on its direction despite exerting bidirectional effects [53]. Recent findings demonstrated that forced chromatin looping of the β-globin locus control region (LCR) to the HBG promoters has shown to reactivate their expression in adult erythroid cells, demonstrating that physical juxtaposition of the LCR is sufficient to bypass promoter silencing. Deletions or inversions of non-insulating genomic regions can reduce linear distance between enhancers and promoters, effectively hijacking enhancer activity and enabling inappropriate gene activation [54]. Together, these findings highlight that while single enhancers function independently of orientation, SEs and LCRs may encode directional information that modulates their regulatory potential. Thus, the developmentally regulated chromatin dynamics at globin loci illustrate a sophisticated interplay between enhancer accessibility, chromatin looping, and stage-specific transcription factor recruitment.

2.5. Global DNA Methylation Dynamics in Erythropoiesis

Beyond chromatin accessibility and quaternary DNA structure dynamics, epigenetic regulation is facilitated through enzymatic DNA methylation, where changes in methylation patterns can activate or repress transcription. Genome-wide DNA methylation studies in both human and mouse erythropoiesis uncover global dynamic methylation patterns that govern erythroid development and differentiation [7,55,56]. Specifically, transition from HSPCs to mature erythroid precursors is characterized by gradual yet continuous global DNA demethylation involving dynamic interplay between DNA methyltrasferase DNMT1 and the de novo methylases DNMT3A and DNMT3B [55,56]. Interestingly, dramatic changes in methylation profiles occurred in promoters of genes that were not expressed throughout the erythropoiesis or newly silent genes which were mostly associated with myeloid and lymphoid fates, suggesting that methylation programs establish erythroid fate through lineage restriction [56]. In parallel, developmental comparisons of human erythroblasts derived from fetal liver versus adult bone marrow revealed over 5900 differentially methylated CpGs, many of which cluster within erythroid enhancers and transcription factor binding motifs [57]. Surprisingly, the majority of differentially methylated regions were hypermethylated in fetal erythroblasts, counter to the expectation of a globally more open fetal epigenome. These CpGs were strongly enriched in proximity to erythroid-specific enhancers and near genes with known or newly suggested roles in red blood cell differentiation and proliferation. Notably, many of these regions were associated with transcription factor binding motifs relevant to erythropoiesis, underscoring a developmentally staged rewiring of the epigenome that likely fine-tunes enhancer activity and gene expression during erythroid maturation.

3. Role of Erythroid Transcription Factors in Shaping Accessible Chromatin Landscapes

The temporal interplay between chromatin and transcription factors (TFs) plays a central role in shaping transcriptional programs during development and differentiation. In adult human erythropoiesis, this dynamic relationship has been comprehensively characterized using multiomic approaches, which profile chromatin accessibility and gene expression across erythroid lineage progression [8,11]. These studies revealed that erythroid differentiation is orchestrated by a series of discrete, temporally resolved regulatory modules composed of transcription factors and their target chromatin elements. These modules exhibit stage-specific activity and reflect the transition from hematopoietic stem and progenitor cells (HSPCs) through lineage restriction to terminal erythroid maturation. By integrating dense temporal chromatin accessibility and gene expression profiling, we recently identified a limited set of approximately 50 transcription factors that sequentially engage the regulatory landscape, underscoring a surprisingly compact regulatory architecture [8]. Single-cell proteomics along the ex vivo differentiation of red blood cells from HSPCs highlights gradients in protein abundance of co-expressed key transcription factors as the underlying mechanism of erythroid lineage specification. Specifically, gradual quantitative changes in the relative abundance of KLF1 and FLI1 can guide hematopoietic progenitors towards the erythroid or the megakaryocytic lineage, respectively [58].
The effects of regulatory variants on erythropoiesis ultimately converge on the gene regulatory networks controlled by erythroid transcription factors. These master regulators, such as GATA1, TAL1, and KLF1, not only activate lineage-specific genes but also orchestrate chromatin accessibility and three-dimensional genome organization. Understanding their roles is essential to decipher how the erythroid epigenome is established, maintained, and perturbed in health and disease.

3.1. The GATA2-to-GATA1 Switch: Coordinating Erythroid Lineage Progression

GATA1 and GATA2 transcription factors play key roles in gene regulation during erythropoiesis (Figure 3).
They are both zinc-finger DNA-binding proteins and upon their expression GATA1 and GATA2 bind to (A/T)GATA(A/G) motifs. GATA1 and GATA2 are strictly regulated in a cell-specific manner [59]. GATA2 supports HSCs proliferation and prevents premature differentiation. During the embryonic development GATA2 regulates embryonic blood formation in the yolk sac and liver [60,61]. GATA1 is more active at the later stages of differentiation. GATA1 regulates globin genes expression and transactivates the erythropoietin receptor gene [62,63], enzymes for heme synthesis, and proteins that make up the erythroid cell membrane. Also, is necessary for megakaryocyte maturation and platelet formation [64,65], as loss of GATA1 disrupts the proper maturation of erythroid and megakaryocyte cells, highlighting its essential role in ensuring correct blood cell lineage development [66,67,68]. Moreover, GATA1 interacts with several cofactors, including FOG1, PU.1, and the transcriptional coactivators p300/CBP, which enhance its regulatory function [67,69,70]. FOG1 (Friend of GATA1) is an essential cofactor that binds directly to GATA1, enabling it to activate or repress target genes required for erythroid and megakaryocyte (platelet precursor) development. The GATA1/FOG1 interaction is critical for normal blood cell maturation; disruption of this interaction impairs erythroid and megakaryocyte development and is linked to blood disorders [71,72,73,74,75]. GATA1 promotes erythroid differentiation, while PU.1 drives myeloid lineage development. These factors can physically interact and inhibit each other’s function, creating a genetic switch that determines cell fate [71,76,77]. FOG1 also represses PU.1 expression by recruiting GATA1 to the PU.1 promoter, thereby promoting erythroid over myeloid differentiation [78]. GATA1 undergoes acetylation by p300/CBP on several clusters of lysine residues adjacent to the two DNA-binding zinc fingers [79,80]. Mutating these sites dramatically impairs GATA1 function, suggesting that acetylation of GATA1 plays a role in its transcriptional activity. Notably, GATA1 plays crucial role to GATA2 downregulation. Genome-wide binding analysis in mouse fetal liver cells revealed that GATA1 strongly binds the GATA2 locus in both early and late erythroid stages and that this binding is associated with increased repressive histone marks, including H3K27me3, specifically in downregulated target gene clusters [81]. GATA2 promotes its own expression by binding to the upstream WGATAR motif. During erythroid differentiation, the increased production of GATA1 recruits the FOG1 and the NuRD complex, leading to the formation of the GATA1-FOG1-NuRD complex. This complex effectively occupies on WGATAR motif, resulting in GATA2 downregulation [82,83,84,85].

3.2. TAL1: Early Hematopoiesis and β-Globin Regulation

TAL1 (also known as SCL) is a basic helix–loop–helix transcription factor that plays a vital role in the early stages of blood cell formation (hematopoiesis). In mice lacking TAL1, embryonic development is halted due to a complete failure of hematopoiesis, and embryonic stem cells without TAL1 are unable to contribute to any blood cell lineages in adult chimeric models [86,87]. At the molecular level, TAL1 exerts its function by forming dimers with E proteins such as E12 and E47, enabling it to bind to E-box motifs in the regulatory regions (promoters and enhancers) of target genes, thereby influencing their transcription [88,89]. This mechanism is critical for regulating genes that control blood cell proliferation, survival, and differentiation. However, abnormal TAL1 expression, often due to chromosomal translocations or other genetic alterations, has been linked to blood cancers, particularly T-cell acute lymphoblastic leukemia (T-ALL) [90,91]. Understanding both TAL1’s role in normal blood development and its contribution to leukemia highlights its significance as a potential target for therapeutic intervention in hematologic disorders. TAL1 also collaborates with several cofactors—such as LMO2, GATA1, LDB1, and RUNX1—to form transcriptional complexes that enhance its regulatory precision and biological effectiveness in hematopoiesis [87,92,93]. In the context of hemoglobin regulation, TAL1 and GATA1, LMO2 complex forces long-range interaction between β-globin LCR and active globin genes. TAL1 knockdown in K562 cells decreased the binding of LDB1 and LMO2 in the LCR locus and resulted in a consequent gamma-globin depression. In contrast, overexpression of TAL1 increased the interaction between β-globin LCR and the Gγ-globin gene and increased gamma-globin expression [94].

3.3. KLF1: Terminal Erythroid Differentiation and Hemoglobin Switching

KLF1, also known as erythroid Krüppel-like factor (EKLF), is a transcription factor that binds specifically to the DNA sequence CCM-CRC-CCN via its C2H2-type zinc finger domains [95]. KLF1 is involved in several steps of hematopoiesis, including cell-cycle regulation, the central macrophages of erythroblastic island function, and the erythroid commitment [96,97,98,99,100]. It is found to be most essential for the final stages of erythroid cell development by regulating essential erythroid genes involved in erythroid cell proliferation, heme biosynthesis, and membrane integrity [100,101,102,103,104,105]. KLF1 has also been identified as a key player in globin regulation through simultaneously binding on the β-globin LCR and on the CACCC box at the β-globin promoter [106,107,108,109]. Moreover, KLF1 interacts with CBP/p300 and BRG1 components of SWI/SNF complex [110]. The loop formation between the β-globin LCR and the β-globin promoter induces β-globin production. Disruption of the KLF1 gene in mice impairs hemoglobin synthesis and causes a severe form of β-thalassemia, highlighting its critical function in erythropoiesis [111]. Studies using transcriptomic profiling of KLF1-deficient mice have revealed major alterations in gene networks responsible for controlling the cell cycle [100]. Specifically, the absence of KLF1 leads to defective progression into the S phase during red blood cell formation, reinforcing its essential role in coordinating proper cell cycle dynamics during erythropoiesis [112].

4. Non-Coding Variation and Erythroid Phenotypes

The dynamic chromatin remodeling that underlies erythropoiesis highlights the critical role of regulatory DNA elements in controlling gene expression. Genetic variation in these regulatory regions, particularly single-nucleotide polymorphisms (SNPs), can subtly alter transcription factor binding, chromatin accessibility or enhancer/promoter communication, thereby modulating erythroid gene expression programs [113].

4.1. Functional Consequences of Distal Regulatory Variants

Variants in distal regulatory elements (REs) have been less explored than those in promoters or 3′ UTRs, largely because their genomic locations are less predictable. The functional impact of mutations in the aforementioned regions depends on specific genomic contexts, including binding site alterations and enhancer sensitivity to sequence variation [114,115]. This indirect and cumulative influence makes it especially difficult to pinpoint which genes are affected, how regulation is altered, and what the biological consequences are, particularly when compared to coding variants [116]. CRISPR/Cas9-mediated loss-of-function studies offer a direct way to test the functional impact of non-coding mutations and regulatory SNPs associated with monogenic erythroid disorders. Combined with experimental and bioinformatic analyses, this approach has been used to generate mutational maps that predict the effects of variants across CREs. It has also helped reveal the functional hierarchy of constituent enhancers and provide critical insights into their complexity [117,118].

4.2. Therapeutic Relevance of Regulatory SNPs

Deciphering the functional consequences of regulatory SNPs is critical for advancing therapies that target gene regulatory networks. When regulatory SNPs disrupt gene expression or pathway activity, they can reveal potential therapeutic targets, enabling the development of treatments that can restore normal protein levels and ameliorate disease. A clear example of this is seen in the treatment of sickle cell disease and β-thalassemia, where genome-editing therapeutic strategies have emerged based on our understanding of fetal hemoglobin (HbF) regulation. Hereditary persistence of fetal hemoglobin (HPFH), a term introduced in 1958, refers to naturally occurring deletions or point mutations in the β-globin gene cluster that allow continued expression of HbF into adulthood [119]. The resulting distribution of HbF across red blood cells varies. When HbF is present only in a subset of cells, the condition is termed heterocellular HPFH; in contrast, pancellular HPFH refers to a more uniform expression across most or all red blood cells [120].
This difference in distribution likely reflects the strength of the γ-globin gene expression driven by a specific mutation [120]. A well-characterized non-coding SNP (ncSNP) example is the C > T polymorphism at rs7482144 in the HBG2 promoter, which is associated with elevated HbF levels through disruption of transcriptional repression. This variant interferes with the binding of BCL11A and ZBTB7A, which typically occupy motifs located approximately 115 bp and 200 bp upstream of the transcription start site, respectively [121,122,123,124]. Disruption of the HBG2 promoter at rs7482144 (158 bp upstream) increases HbF, though to a lesser extent than mutations directly affecting the ZBTB7A or BCL11A binding motifs, likely because it only impacts HBG2 rather than both γ-globin genes [125,126]. The diversity in HbF levels observed across different point mutations may depend on how they alter transcription factor binding or protein–complex interactions. Notably, when rs7482144 co-occurs in cis (i.e., on the same locus and allele) with other mutations (e.g., 175 T > C or 202 C > G), HbF levels can rise dramatically, up to 41% in some HbS heterozygotes, suggesting synergistic effects from the combined disruption of repressor binding and potential recruitment of activators such as GATA1 [120,127,128]. The HBG promoter is not the only region where mutations have an impact in gamma globin expression. Various quantitative trait loci (QTL) have been associated with persistence of elevated HbF, including BCL11A, HBS1L-MYB, and the HBB gene cluster in general [121]. Among these, ncSNPs implicated in elevated HbF levels include rs4671393 in the BCL11A locus, as well as rs9399137 and rs9494142 in the HBS1L-MYB region [121]. In addition to HbF levels, certain ncSNPs can also affect other hematological parameters, including RBC count and size, platelet count, and hemoglobin levels [129]. For instance, rs9399137—a well-known ncSNP in the HBS1L-MYB—is associated with fewer RBCs and monocytes but larger, hemoglobin-rich RBCs and higher platelet counts [130]. Overall, SNPs are crucial genetic markers for managing β-thalassemia and SCD. Primarily, they help predict clinical severity by their association with HbF levels, which directly informs patient management, such as the timing of blood transfusions or decisions about hematopoietic stem cell transplantation [131,132]. A key example of this is seen in the β-globin gene cluster haplotypes in SCD. The Arab-Indian and Senegal haplotypes, for instance, are associated with elevated HbF, leading to milder symptoms, whereas the Benin and Bantu haplotypes are linked to lower HbF and more severe disease [133].
In addition to predicting overall disease severity, certain SNPs also serve as markers for specific complications. For instance, variants in the TLR1/TANK and MALT1 genes have been associated with an increased risk of alloimmunization in some populations [134]. Others, such as rs7319269, are linked to a higher risk of acute chest syndrome, although the underlying mechanisms are still being investigated [135,136].
Finally, mapping these genetic variations enhances diagnostic accuracy and supports informed reproductive choices through early and noninvasive prenatal screening [137,138].

4.3. Non-Coding Variants in Transcription Factor Binding Sites

Non-coding SNPs can also give rise to both inherited and acquired erythroid disorders through mutations in genes encoding transcription factors and epigenetic regulators [138]. Mutations in GATA1, a master regulator of erythropoiesis and megakaryopoiesis, illustrate how disruptions in transcription factors or their DNA binding sites can impair blood cell development. For instance, missense mutations in GATA1 cause anemia and thrombocytopenia, highlighting its essential role in hematopoiesis. The SNP rs311103, located in the XG gene region (Xg blood group), disrupts a GATA1-binding motif upstream of the gene. This disruption abolishes the expression of the Xga blood group antigen on erythrocytes, resulting in the Xg(a−) phenotype [139]. Two SNPs within the GATA2, rs2335052 and rs78245253, are associated with an increased risk of acute myeloid leukemia (AML) in Chinese populations [140]. Pathogenic variants in GATA1-binding motifs have also been implicated in several inherited erythroid disorders, including congenital dyserythropoietic anemia type II, X-linked sideroblastic anemia, and pyruvate kinase deficiency [118,141,142]. In vitro, CRISPR-mediated disruption of non-coding variants associated with X-linked sideroblastic anemia and pyruvate kinase deficiency revealed that even small (2–4 nt) changes in GATA1 binding sites reduce GATA1 binding and impair recruitment of cofactors like TAL1 [118]. In contrast, mutations in TAL1 binding sites have milder effects on GATA1 binding but still alter gene expression, highlighting distinct roles for different transcription factors. In vitro, CRISPR-mediated disruption of non-coding variants associated with X-linked sideroblastic anemia and pyruvate kinase deficiency revealed that even small (2–4 nt) changes in GATA1 binding sites reduce GATA1 binding and impair recruitment of cofactors like TAL1 [118]. In contrast, mutations in TAL1 binding sites have milder effects on GATA1 binding but still alter gene expression, highlighting distinct roles for different transcription factors [93]. Further bioinformatic analysis of multiple regulatory elements near genes in RBC structure, metabolism, and function supports a model where independent elements coordinate gene expression, mitigating the impact of single-site mutations [118].
Non-coding mutations have also been implicated in diagnostically challenging cases of Diamond–Blackfan anemia syndrome (DBAS), a disorder defined by pure red cell aplasia and early-onset macrocytic or normocytic anemia, typically caused by coding mutations in ribosomal protein genes. In cases where conventional testing, including targeted panels and exome sequencing, failed to identify a causative variant, non-coding regions have revealed pathogenic variants, particularly through in-depth reanalysis of whole-genome sequencing data [122]. Additionally, germline N-terminal truncating mutations in GATA1, producing the short isoform GATA1s, result in an inherited erythroid failure that phenotypically resembles DBAS [143,144].
Somatic mutations in the GATA1 gene can lead to the production of a truncated protein called GATA1s that lacks the N-terminal domain. In children with Down syndrome (DS), the combination of trisomy 21 and GATA1s significantly increases the risk of developing transient abnormal myelopoiesis (TAM), a pre-leukemic disorder characterized by abnormal proliferation of immature megakaryoblasts due to blocked erythroid differentiation. While most TAM cases resolve spontaneously, about 10–30% progress to myeloid leukemia associated with Down syndrome (ML-DS) [145,146,147,148,149]. In the TAM cases that progress to ML-DS, additional somatic mutations occur in genes such as those encoding the cohesin complex (for example, STAG2), epigenetic regulators, or signaling molecules. These mutations enhance self-renewal and further block differentiation, ultimately cooperating with GATA1s to drive overt leukemia [148].
The miR-144/451 cluster, activated by the transcription factor GATA1, is essential for erythroid maturation and differentiation. miR-451 promotes red blood cell development by downregulating GATA2, a stem-cell-associated transcription factor, particularly in zebrafish [150,151]. Additionally, the cluster suppresses Myc to ensure proper erythroid differentiation [152]. Together, the GATA1–miR-144/451–GATA2/Myc regulatory axis plays a crucial role in erythropoiesis, and disruption of this pathway can result in impaired red blood cell formation and anemia.

4.4. Somatic Non-Coding Mutations in Hematologic Malignancies

Besides hereditary conditions, acquired mutations in erythroid REs can also contribute to hematologic malignancies [113]. Emerging evidence points to a critical role of non-coding regulatory variants as cancer drivers [153,154]. For instance, in the context of blood malignancies, somatic mutations can introduce de novo binding motifs for the MYB transcription factor at specific non-coding sites upstream of the TAL1 proto-oncogene. MYB binds to these newly created sites and recruits its H3K27 acetyltransferase coactivator CBP, along with core components of a major leukemogenic transcriptional complex, including RUNX1, GATA3, and TAL1 itself. This assembly drives the formation of an oncogenic super-enhancer that aberrantly activates TAL1 expression, promoting leukemogenesis [124]. One of the most typical myeloproliferative neoplasms (MPNs) is Polycythemia Vera (PV). PV is characterized by excessive proliferation across all three hematopoietic lineages as well as extramedullary hematopoiesis, leading to elevated red blood cells and hematocrit [155]. The onset of the PV is primary driven by somatic mutations in JAK2 coding regions [156]. Additionally, mutations in non-coding regulatory regions have been associated with the development and/or the severity of PV. Namely, mutations in the enhancer’s region near to the JAK2 locus may influence chromatin loop formation, altering JAK2 expression [157]. Common non-coding germline SNPs in the TERT locus have been associated with MPN predisposition. Among all, the rs2736100_C is the most implicated with PV and is associated with longer telomeres and high blood numbers [158]. Moreover, variants in non-coding regions of SH2B3, GATA2 and MECOM may alter hematopoietic stem/progenitor cell behavior and JAK2-mutant clone expansion. In addition to the well-known driver mutations, myeloproliferative neoplasms (MPNs) frequently harbor mutations in genes associated with epigenetic regulation and mRNA splicing. Mutations in TET2, ASXL1, DNMT3A, and EZH2 are commonly found in patients with myeloproliferative neoplasms (MPNs). Loss-of-function mutations in the aforementioned genes led to global epigenetic dysregulation leading to an apparent induction of oncogenes and HSC self-renewal genes [154,159,160,161].

5. Leveraging Gene Regulation for Translational Applications

Understanding how erythroid transcription factors and chromatin landscapes regulate gene expression has not only clarified fundamental principles of erythropoiesis but also revealed a rich repertoire of regulatory elements that can be both targeted and harnessed for therapeutic purposes. Building on these insights, translational research is increasingly focused on leveraging key enhancers, silencers, and elements of chromatin architecture, not only to correct gene expression defects but also to reprogram erythroid gene networks in a disease-specific manner. In this section, we highlight how the therapeutic modulation and strategic use of the erythroid regulatory genome is driving innovation across a spectrum of hematologic diseases.

5.1. Erythroid-Specific Enhancers for Gene Therapy

Precise spatial and temporal control of gene expression is a cornerstone of effective gene therapy. Cis-regulatory elements can be used to fine-tune transgene expression, ensure lineage fidelity, and integrate seamlessly with endogenous regulatory networks.
Erythroid-specific expression of a β-globin transgene has been considered critically important for gene therapy of hemoglobinopathies since the earliest preclinical trials, when the importance of its regulatory machinery was already established [162,163,164,165,166,167]. The incorporation of critical components of the β-globin LCR into gene therapy vectors marked a pivotal advancement in the effort to achieve high-level, erythroid-specific expression of therapeutic transgenes [168,169,170,171], with the first successful design termed TNS9, achieving efficient gene transfer into murine HSPCs and sustaining high-level, erythroid-specific expression in irradiated mice receiving β-thalassemia marrow transplants [169]. Most current gene therapy for hemoglobinopathies relies on a truncated β-globin LCR spanning a ~3 kb in size to ensure both therapeutic and erythroid-specific expression of the respective transgene [172,173,174]. The first clinical trial for β-thalassemia [175] was performed using the Lentiglobin HPV569 vector, which contained a minimal β-globin promoter and the HS2, HS3, and HS4 elements. To avoid positional variegation and insertional genotoxicity, two copies of the chicken β-globin cHS4 insulator were initially inserted into the viral LTRs, but this introduction significantly reduced viral titers [176]. In the most recent version of this vector, Lentiglobin BB305 (created by Bluebird Bio-BBB), the cHS4 element was excluded, resulting in higher vector titers and improved gene transfer efficiency, enhancing its clinical utility [169]. Other insulating elements, of human origin, better efficiency and smaller size have been recently proposed [177,178]. The GLOBE LV, a beta-globin vector of similar design but containing a minimal HS2-HS3-LCR, demonstrated therapeutic efficacy in both murine models and patient-derived CD34+ cells, restoring HbA levels and reducing apoptosis during erythroid differentiation [173,179]. Recently, high-resolution mapping of the beta-globin LCR has allowed for the functional identification of enhancer elements suggesting alternative vector designs are feasible [180]. Mini LCR elements often contain binding sites for GATA1, NF-E2, and EKLF, which promote open chromatin and high-level, position-independent expression of the β-globin genes [181,182]. Some studies have shown that tandem copies of these short enhancer elements can significantly increase transgene expression by promoting an open chromatin structure at the promoter region [182]. However, optimal activity often requires cooperation between enhancer/LCR elements and promoter or intronic elements, highlighting the importance of context and sequence composition [183,184]. LCR free vectors employing alternative beta-globin locus enhancers have also been described. Building on previous work which identified such regions within the globin locus [185,186,187], vectors bearing the Aγ-globin gene have also been described [188,189].
Enhancers and promoters from other erythroid-expressed genes, such as ANK1 (Ankyrin-1), have also been characterized for their robust activity in erythroid cells and incorporated into vector designs [190,191]. Additionally, KLF1-bound enhancers have been explored, leveraging the erythroid-restricted transcription factor KLF1 to drive lineage-specific activation of therapeutic genes [192,193,194]. A recent study developed a clinical-grade lentiviral gene therapy vector for Diamond–Blackfan anemia (DBA) that uses an engineered human GATA1 enhancer (hG1E) to drive regulated expression of the transcription factor GATA1 [195]. The hG1E enhancer was constructed by concatenating three chromatin regions that are selectively accessible during erythroid differentiation, enabling precise expression of GATA1 only after lineage commitment and thus avoiding adverse effects on hematopoietic stem cells. This enhancer-based regulation successfully restored erythropoiesis in DBA models and patient samples, establishing a broadly applicable therapeutic strategy independent of the patient’s specific genetic mutation [195].
Recent research has focused on identifying compact, potent, and lineage-specific enhancer elements capable of recapitulating the activity of the β-globin locus control region (LCR) in a much smaller sequence. High-throughput screening of short (~200 bp) DNA fragments derived from developmentally active DNase I hypersensitive sites (DHSs) during human erythropoiesis has revealed several candidates that can effectively replace the μLCR in therapeutic vectors. Among these, two erythroid-specific enhancers were identified that exhibit minimal activity in non-erythroid lineages and closely mirror the temporal activation dynamics of native erythroid enhancers [196]. One of these enhancers is located within an intron of the PVT1 locus. It is enriched for erythroid transcription factor binding motifs, including GATA1, TAL1, and KLF1, and becomes accessible early during erythroid differentiation, reflecting the activation pattern of its endogenous DHS. Deletion of the enhancer leads to a marked reduction in PVT1 transcription, confirming its regulatory role, despite limited evidence implicating the PVT1 lncRNA in erythropoiesis. The second enhancer originates from an exonic DHS within a gene encoding the Peroxisome Proliferator Activated Receptor Alpha, PPARA. Like the PVT1-derived element, it is enriched for erythroid transcription factor motifs but exhibits peak activity at later stages of erythroid maturation [196]. PPARA itself has been identified as an erythroid-expressed gene, further supporting the relevance of this enhancer to red blood cell development [197,198].
Overall, erythroid-specific enhancers, from the canonical β-globin LCR to novel elements like PPARA or hG1E, have become essential tools in gene therapy, enabling precise, lineage-restricted transgene expression that enhances efficacy while minimizing off-target effects.

5.2. Fetal Hemoglobin Reactivation

As mentioned above, when HPFH is co-inherited with β-hemoglobinopathies, it leads to a milder disease phenotype [120,199,200]. This observation has inspired therapeutic strategies aimed at reactivating γ-globin expression in adult erythroid cells (Figure 4). Such reactivation can compensate for defective or absent β-globin in β-thalassemia and inhibit β-globin polymerization in sickle cell disease (SCD), providing a versatile treatment strategy [201,202].
Multiple preclinical studies have used genome editing tools, including TALENs, CRISPR-Cas9, and base editors, to reactivate the developmentally silenced γ-globin [203] (Figure 4). One set of targets includes the intergenic region between the Aγ- and δ-globin genes, which contains a pseudogene (HBBP1) and a non-coding gene (BGLT3). Deletion of HBBP1 reactivates γ-globin in adult-type human erythroid cells, suggesting a developmental silencing role [204]. BGLT3 transcribes a long non-coding RNA that enhances chromatin looping between the LCR and the γ-globin genes, positively regulating their expression [205]. Several HPFH-associated mutations have been identified within the HBBP1/BGLT3 locus [206,207].
In addition to these distal elements, more proximal regulatory regions within the γ-globin gene promoters have also been targeted to reactivate fetal hemoglobin. Another successful strategy involves direct editing of the γ-globin (HBG) promoters. TALEN-mediated disruption of regulatory elements in these promoters reactivates γ-globin expression ex vivo and in vivo in humanized mouse models [208]. Similarly, introduction of naturally occurring HPFH mutations—such as the Sicilian HPFH variant—using CRISPR-Cas9 leads to a twofold increase in γ-globin mRNA in HSPC-derived erythroid colonies [209].
These promoter-based strategies frequently act by disrupting the binding of BCL11A, a major transcriptional repressor of γ-globin. Notably, the BCL11A binding site overlaps with the region where naturally occurring HPFH mutations such as −115C > T, −114C > T, and −113A > G have been identified in the HBG1 and HBG2 promoters. These point mutations disrupt BCL11A binding, leading to elevated γ-globin expression in individuals with HPFH. Disruption of the BCL11A binding site at the γ-globin promoters via in vitro [210] or in vivo [211] CRISPR-Cas9 editing was shown to reactivate γ-globin expression. In addition to BCL11A, LRF (ZBTB7A) also contributes to γ-globin repression by binding directly to the promoters. Genome editing strategies that delete or mutate the LRF binding motif in the HBG promoters can relieve this repression, further enhancing fetal hemoglobin induction [125].
In addition to disrupting transcription factor binding at the γ-globin promoters, an alternative strategy targets a regulatory element upstream of BCL11A itself. BCL11A expression is controlled by multiple enhancers, one of which, located in intron 2 and consisting of three DHSs (+56, +58, +62), is selectively active in the erythroid lineage. Targeted disruption of a GATA1 binding site [212,213] at the +58 site of this erythroid enhancer of BCL11A leads to reduced BCL11A expression in erythroid cells without affecting its function in other hematopoietic or immune lineages. This selective silencing enables robust reactivation of γ-globin while preserving HSC function, making it particularly attractive for therapeutic genome editing. Importantly, this strategy has demonstrated efficacy in human HSPCs ex vivo and in animal models [213,214,215,216] and has advanced not only to clinical trials [217,218,219] but also to regulatory approval and commercialization [220]. There are two phase 1/2/3 clinical trials funded also by VERTEX-Therapeutics in collaboration with CRISPR-Therapeutics for TDT beta-thalassemia (CLIMB THAL-111, NCT03655678) and sickle cell disease patients (CLIMB SCD-121, NCT03745287). In these trials, the disruption of the erythroid enhancer of BCL11A using the CRISPR-Cas9 system resulted in sustained normal Hb levels 12.5±1.8 g/dl for SCD and 13.1±1.4 g/dl for beta- thalassemia patients After the gene therapy, the 91% of the thalassemic patients were transfusion independent and the 97% of the SCD had reduced vaso-occlusive crisis for a period of 12 months or more [217,218,219].
Building on CRISPR-Cas9 technology, recent advances have led to the development of base editors, which allow the introduction of precise point mutations without creating double-strand breaks [221,222,223]. Using a cytidine base editor (CBE), HPFH-associated mutations at −115 C or −114 C in the HBG promoters were successfully introduced, leading to increased γ-globin expression in HSPCs from both healthy donors and patients with β-thalassemia [224]. Similarly, adenine base editors (ABEs) have been used to introduce the −113 A > G substitution at HBG promoters, resulting in elevated γ-globin expression in β-YAC/CD46tg mice and in HSPCs derived from β-thalassemia and SCD patients [225,226]. Notably, CBEs and ABEs have also been applied to disrupt LRF binding sites or to generate de novo KLF1 binding sites, promoting γ-globin reactivation and preventing sickling both during ex vivo erythroid differentiation and in vivo in humanized SCD mouse models [227]. More recently, Myuranathan et al. used an ABE to introduce a −175 A > G mutation that creates a GATA-TAL1 motif at the HBG promoters. This edit did not compromise the engraftment potential of modified cells and resulted in effective HbF reactivation both ex vivo and in vivo [228].
Beyond DNA editing, epigenome editing technologies have enabled the modulation of gene expression without altering the underlying DNA sequence. A notable example involves targeting the HBG promoters with catalytically inactive Cas9 (dCas9) proteins fused to the p300 histone acetyltransferase domain. The dCas9-p300 complexes, derived from various Cas9 orthologs, successfully increased HBG expression in 293FT cells, highlighting the potential of epigenetic reprogramming in hemoglobinopathies [229].

5.3. Epigenetic Therapeutics in Erythroid Disorders

Epigenetic modifiers represent a versatile class of therapeutic agents with the potential to modulate gene expression without altering the underlying DNA sequence. Classes of such agents, including DNA Methyltransferase Inhibitors (DNMT inhibitors), Histone Deacetylase Inhibitors (HDAC inhibitors), Histone Methyltransferase (HMT) or Demethylase Inhibitors, and Bromodomain and Extra-Terminal Domain (BET) Inhibitors, have been evaluated in various erythroid disorders [230]. A major therapeutic goal has been the reactivation of fetal hemoglobin (HbF) in β-hemoglobinopathies such as sickle cell disease and β-thalassemia. Pharmacologic reactivation of γ-globin expression offers a non-invasive strategy to compensate for defective or absent β-globin chains (Figure 4). DNMT inhibitors such as azacytidine, decitabine, and the DNMT1-selective inhibitor GSK3482364 have demonstrated therapeutic potential in β-hemoglobinopathies by inducing γ-globin expression through promoter hypomethylation, as shown in patients with β-thalassemia and SCD as well as in transgenic SCD mouse models [231,232,233,234]. Moreover, HDAC inhibitors such as butyrate and trichostatin A led to gamma-globin reactivation, through affecting p38 MAPK and STAT-5 signaling pathway [231,232]. The ACY-957, HDAC1/2 inhibitor efficiently reactivates HbF expression in SCD patients’ cells by histone acetylation-induced activation of the GATA2 gene [235]. UNC0638, a Histone Methyltransferase (HMT) Inhibitor, decreases the accumulation of H3K9me2 upstream of gamma-globin, enhancing the loop formation between gamma-globin promoters and LCR locus and promoting gamma-globin reactivation [236,237,238,239] (Figure 4). Notably, the RK-701 inhibitor of the euchromatic histone/lysine N-methyltransferase 2 (EHMT2) reactivates gamma-globin expression by increasing BGLT3 expression [240]. Moreover, lysine-specific demethylase 1 (LSD1) inhibitors, such as tranylcypromine (TCP), RN-1, and ORY-3001 have been widely employed for the pharmacological induction of gamma-globin expression [241]. TCP and RN-1 increase H3K4 di- and tri-methyl lysine at the γ-globin promoter and efficiently reactivating gamma-globin expression in SCD mouse and baboon models [242], while oral administration of ORY-3001 further enhances the effect in vivo [243]. Additionally, peroxisome proliferator–activated receptor γ coactivator-1α, (PGC-1α) triggers gamma-globin expression through its interaction with nuclear receptors testicular receptor 2/4 (TR2/TR4). PGC-1α activation by SR-18292 or the ZLN005 agonist, led to increased percentage of F-cells in normal CD34+ cells, β-YAC transgenic, and SCD mice, reducing splenomegaly and sickling, especially when used in combination with hydroxyurea [244,245].
Beyond β-hemoglobinopathies, epigenetic dysregulation also plays a central role in erythroid malignancies, such as acute erythroid leukemia (AEL). Acute erythroid leukemia is an uncommon and aggressive subtype of AML, representing <1% of adult AML cases. It is typified by predominant erythroblast proliferation, frequent biallelic TP53 mutations, complex cytogenetics, and poor prognosis [145,146,147,148,149,246,247]. Large-scale molecular studies have uncovered recurrent alterations across epigenetic regulators—including DNMT3A, TET2, NSD1, BCOR—and splicing factors, which synergize with TP53 loss to impair erythroid differentiation via widespread DNA methylation and histone modification changes affecting loci such as GATA1, TAL1, KLF1, and EPOR [247]. Mouse models and functional analyses confirm that combined lesions (e.g., TP53 + BCOR + DNMT3A) induce erythroid-skewed acute leukemia, driven by genome-wide hypomethylation at erythroid regulatory elements; such preclinical studies also highlight potential sensitivities to CDK7/9 inhibitors, PARP inhibitors, and menin-bromodomain targeting agents [145,149,246,247] (Table 1).

6. Conclusions and Future Directions

Erythropoiesis is governed by complex, multilayered regulatory mechanisms involving dynamic chromatin remodeling, lineage-defining transcription factors, and stage-specific regulatory elements. This review has outlined how these epigenetic processes coordinate erythroid lineage commitment, differentiation, and globin gene expression and how their disruption contributes to a spectrum of hematologic diseases, from β-hemoglobinopathies to erythroid malignancies.
Recent advances in genomics, gene editing, and epigenetic profiling have not only deepened our understanding of erythroid biology but have also enabled the development of precision therapies, ranging from fetal hemoglobin reactivation to enhancer-driven gene therapy. These therapeutic strategies increasingly reflect the regulatory logic of the erythroid program itself, offering promising avenues for targeted and durable interventions.
While this review focuses on key aspects of epigenetic regulation in steady-state erythropoiesis, it is important to acknowledge that this represents only a fraction of the regulatory complexity underlying red blood cell development. Important areas such as the contribution of non-coding RNAs, post-translational modifications of epigenetic regulators, and the unique regulatory dynamics of stress erythropoiesis remain beyond the scope of this article. Continued integration of these dimensions will be essential to build a more complete and nuanced understanding of erythroid biology in both health and disease.

Author Contributions

G.G. and N.P. conceptualized the review. N.I.V., K.P., C.B., G.G. and N.P. performed the literature search and drafted the initial manuscript. G.G.and N.P. provided critical revisions, and supervised the project. All authors have read and agreed to the published version of the manuscript.

Funding

This project is carried out within the framework of the National Recovery and Resilience Plan Greece 2.0, funded by the European Union—NextGenerationEU (Implementation body: HFRI-OGETHERA-15358) (to N.P.).

Data Availability Statement

No data were generated for this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Baron, M.H.; Isern, J.; Fraser, S.T. The Embryonic Origins of Erythropoiesis in Mammals. Blood 2012, 119, 4828–4837. [Google Scholar] [CrossRef] [PubMed]
  2. Dzierzak, E.; Philipsen, S. Erythropoiesis: Development and Differentiation. Cold Spring Harb. Perspect. Med. 2013, 3, a011601. [Google Scholar] [CrossRef] [PubMed]
  3. Palis, J. Erythropoiesis in the Mammalian Embryo. Exp. Hematol. 2024, 136, 104283. [Google Scholar] [CrossRef] [PubMed]
  4. Palis, J.; Malik, J.; McGrath, K.E.; Kingsley, P.D. Primitive Erythropoiesis in the Mammalian Embryo. Int. J. Dev. Biol. 2010, 54, 1011–1018. [Google Scholar] [CrossRef]
  5. Palis, J. Primitive and Definitive Erythropoiesis in Mammals. Front. Physiol. 2014, 5, 3. [Google Scholar] [CrossRef]
  6. Hattangadi, S.M.; Wong, P.; Zhang, L.; Flygare, J.; Lodish, H.F. From Stem Cell to Red Cell: Regulation of Erythropoiesis at Multiple Levels by Multiple Proteins, RNAs, and Chromatin Modifications. Blood 2011, 118, 6258–6268. [Google Scholar] [CrossRef]
  7. Schulz, V.P.; Yan, H.; Lezon-Geyda, K.; An, X.; Hale, J.; Hillyer, C.D.; Mohandas, N.; Gallagher, P.G. A Unique Epigenomic Landscape Defines Human Erythropoiesis. Cell Rep. 2019, 28, 2996–3009.e7. [Google Scholar] [CrossRef]
  8. Georgolopoulos, G.; Psatha, N.; Iwata, M.; Nishida, A.; Som, T.; Yiangou, M.; Stamatoyannopoulos, J.A.; Vierstra, J. Discrete Regulatory Modules Instruct Hematopoietic Lineage Commitment and Differentiation. Nat. Commun. 2021, 12, 6790. [Google Scholar] [CrossRef]
  9. Martin, E.W.; Rodriguez y Baena, A.; Reggiardo, R.E.; Worthington, A.K.; Mattingly, C.S.; Poscablo, D.M.; Krietsch, J.; McManus, M.T.; Carpenter, S.; Kim, D.H.; et al. Dynamics of Chromatin Accessibility During Hematopoietic Stem Cell Differentiation Into Progressively Lineage-Committed Progeny. Stem Cells 2023, 41, 520–539. [Google Scholar] [CrossRef]
  10. Dixon, J.R.; Jung, I.; Selvaraj, S.; Shen, Y.; Antosiewicz-Bourget, J.E.; Lee, A.Y.; Ye, Z.; Kim, A.; Rajagopal, N.; Xie, W.; et al. Chromatin Architecture Reorganization during Stem Cell Differentiation. Nature 2015, 518, 331–336. [Google Scholar] [CrossRef]
  11. Ludwig, L.S.; Lareau, C.A.; Bao, E.L.; Nandakumar, S.K.; Muus, C.; Ulirsch, J.C.; Chowdhary, K.; Buenrostro, J.D.; Mohandas, N.; An, X.; et al. Transcriptional States and Chromatin Accessibility Underlying Human Erythropoiesis. Cell Rep. 2019, 27, 3228–3240.e7. [Google Scholar] [CrossRef] [PubMed]
  12. Stamatoyannopoulos, G. Human Hemoglobin Switching. Science 1991, 252, 383. [Google Scholar] [CrossRef] [PubMed]
  13. Stamatoyannopoulos, G. Control of Globin Gene Expression during Development and Erythroid Differentiation. Exp. Hematol. 2005, 33, 259–271. [Google Scholar] [CrossRef]
  14. Levings, P.P.; Bungert, J. The Human Β-globin Locus Control Region: A Center of Attraction. Eur. J. Biochem. 2002, 269, 1589–1599. [Google Scholar] [CrossRef]
  15. Li, Q.; Peterson, K.R.; Fang, X.; Stamatoyannopoulos, G. Locus Control Regions. Blood 2002, 100, 3077–3086. [Google Scholar] [CrossRef]
  16. Molete, J.M.; Petrykowska, H.; Sigg, M.; Miller, W.; Hardison, R. Functional and Binding Studies of HS3.2 of the Beta-Globin Locus Control Region. Gene 2002, 283, 185–197. [Google Scholar] [CrossRef]
  17. Bulger, M.; Groudine, M. Functional and Mechanistic Diversity of Distal Transcription Enhancers. Cell 2011, 144, 327–339. [Google Scholar] [CrossRef]
  18. Labbaye, C.; Valtieri, M.; Barberi, T.; Meccia, E.; Masella, B.; Pelosi, E.; Condorelli, G.L.; Testa, U.; Peschle, C. Differential Expression and Functional Role of GATA-2, NF-E2, and GATA-1 in Normal Adult Hematopoiesis. J. Clin. Investig. 1995, 95, 2346–2358. [Google Scholar] [CrossRef]
  19. Carter, D.; Chakalova, L.; Osborne, C.S.; Dai, Y.; Fraser, P. Long-Range Chromatin Regulatory Interactions in Vivo. Nat. Genet. 2002, 32, 623–626. [Google Scholar] [CrossRef]
  20. Palstra, R.-J.; Tolhuis, B.; Splinter, E.; Nijmeijer, R.; Grosveld, F.; Laat, W. de The β-Globin Nuclear Compartment in Development and Erythroid Differentiation. Nat. Genet. 2003, 35, 190–194. [Google Scholar] [CrossRef]
  21. Andrews, N.C.; Erdjument-Bromage, H.; Davidson, M.B.; Tempst, P.; Orkin, S.H. Erythroid Transcription Factor NF-E2 Is a Haematopoietic-Specific Basic–Leucine Zipper Protein. Nature 1993, 362, 722–728. [Google Scholar] [CrossRef] [PubMed]
  22. Fontana, L.; Alahouzou, Z.; Miccio, A.; Antoniou, P. Epigenetic Regulation of β-Globin Genes and the Potential to Treat Hemoglobinopathies through Epigenome Editing. Genes 2023, 14, 577. [Google Scholar] [CrossRef] [PubMed]
  23. Kim, Y.W.; Kim, S.; Kim, C.G.; Kim, A. The Distinctive Roles of Erythroid Specific Activator GATA-1 and NF-E2 in Transcription of the Human Fetal γ-Globin Genes. Nucleic Acids Res. 2011, 39, 6944–6955. [Google Scholar] [CrossRef]
  24. Liu, N.; Xu, S.; Yao, Q.; Zhu, Q.; Kai, Y.; Hsu, J.Y.; Sakon, P.; Pinello, L.; Yuan, G.-C.; Bauer, D.E.; et al. Transcription Factor Competition at the γ-Globin Promoters Controls Hemoglobin Switching. Nat. Genet. 2021, 53, 511–520. [Google Scholar] [CrossRef]
  25. Kim, A.; Dean, A. Chromatin Loop Formation in the β-Globin Locus and Its Role in Globin Gene Transcription. Mol. Cells 2012, 34, 1–6. [Google Scholar] [CrossRef]
  26. Gurumurthy, A.; Yu, D.T.; Stees, J.R.; Chamales, P.; Gavrilova, E.; Wassel, P.; Li, L.; Stribling, D.; Chen, J.; Brackett, M.; et al. Super-Enhancer Mediated Regulation of Adult β-Globin Gene Expression: The Role of ERNA and Integrator. Nucleic Acids Res. 2021, 49, 1383–1396. [Google Scholar] [CrossRef]
  27. Kim, Y.W.; Yun, W.J.; Kim, A. Erythroid Activator NF-E2, TAL1 and KLF1 Play Roles in Forming the LCR HSs in the Human Adult β-Globin Locus. Int. J. Biochem. Cell Biol. 2016, 75, 45–52. [Google Scholar] [CrossRef]
  28. Xu, J.; Sankaran, V.G.; Ni, M.; Menne, T.F.; Puram, R.V.; Kim, W.; Orkin, S.H. Transcriptional Silencing of γ-Globin by BCL11A Involves Long-Range Interactions and Cooperation with SOX6. Genes Dev. 2010, 24, 783–798. [Google Scholar] [CrossRef]
  29. Jawaid, K.; Wahlberg, K.; Thein, S.L.; Best, S. Binding Patterns of BCL11A in the Globin and GATA1 Loci and Characterization of the BCL11A Fetal Hemoglobin Locus. Blood Cells Mol. Dis. 2010, 45, 140–146. [Google Scholar] [CrossRef]
  30. Zheng, G.; Orkin, S.H. Transcriptional Repressor BCL11A in Erythroid Cells. Adv. Exp. Med. Biol. 2024, 1459, 199–215. [Google Scholar] [CrossRef]
  31. Enver, T.; Zhang, J.-W.; Anagnou, N.P.; Stamatoyannopoulos, G.; Papayannopoulou, T. Developmental Programs of Human Erythroleukemia Cells: Globin Gene Expression and Methylation. Mol. Cell. Biol. 1988, 8, 4917–4926. [Google Scholar] [CrossRef] [PubMed]
  32. Shang, S.; Li, X.; Azzo, A.; Truong, T.; Dozmorov, M.; Lyons, C.; Manna, A.K.; Williams, D.C.; Ginder, G.D. MBD2a–NuRD Binds to the Methylated γ-Globin Gene Promoter and Uniquely Forms a Complex Required for Silencing of HbF Expression. Proc. Natl. Acad. Sci. USA 2023, 120, e2302254120. [Google Scholar] [CrossRef] [PubMed]
  33. Enver, T.; Zhang, J.W.; Papayannopoulou, T.; Stamatoyannopoulos, G. DNA Methylation: A Secondary Event in Globin Gene Switching? Genes Dev. 1988, 2, 698–706. [Google Scholar] [CrossRef] [PubMed]
  34. Wang, H.; Maurano, M.T.; Qu, H.; Varley, K.E.; Gertz, J.; Pauli, F.; Lee, K.; Canfield, T.; Weaver, M.; Sandstrom, R.; et al. Widespread Plasticity in CTCF Occupancy Linked to DNA Methylation. Genome Res. 2012, 22, 1680–1688. [Google Scholar] [CrossRef]
  35. Sankaran, V.G.; Menne, T.F.; Xu, J.; Akie, T.E.; Lettre, G.; Handel, B.V.; Mikkola, H.K.A.; Hirschhorn, J.N.; Cantor, A.B.; Orkin, S.H. Human Fetal Hemoglobin Expression Is Regulated by the Developmental Stage-Specific Repressor BCL11A. Science 2008, 322, 1839–1842. [Google Scholar] [CrossRef]
  36. Deng, W.; Lee, J.; Wang, H.; Miller, J.; Reik, A.; Gregory, P.D.; Dean, A.; Blobel, G.A. Controlling Long-Range Genomic Interactions at a Native Locus by Targeted Tethering of a Looping Factor. Cell 2012, 149, 1233–1244. [Google Scholar] [CrossRef]
  37. Song, S.-H.; Kim, A.; Ragoczy, T.; Bender, M.A.; Groudine, M.; Dean, A. Multiple Functions of Ldb1 Required for β-Globin Activation during Erythroid Differentiation. Blood 2010, 116, 2356–2364. [Google Scholar] [CrossRef]
  38. Higgs, D.R.; Garrick, D.; Anguita, E.; Gobbi, M.D.; Hughes, J.; Muers, M.; Vernimmen, D.; Lower, K.; Law, M.; Argentaro, A.; et al. Understanding A-Globin Gene Regulation: Aiming to Improve the Management of Thalassemia. Ann. N. York Acad. Sci. 2005, 1054, 92–102. [Google Scholar] [CrossRef]
  39. Gobbi, M.D.; Anguita, E.; Hughes, J.; Sloane-Stanley, J.A.; Sharpe, J.A.; Koch, C.M.; Dunham, I.; Gibbons, R.J.; Wood, W.G.; Higgs, D.R. Tissue-Specific Histone Modification and Transcription Factor Binding in α Globin Gene Expression. Blood 2007, 110, 4503–4510. [Google Scholar] [CrossRef]
  40. Higgs, D.R.; Wood, W.G. Long-Range Regulation of α Globin Gene Expression during Erythropoiesis. Curr. Opin. Hematol. 2008, 15, 176–183. [Google Scholar] [CrossRef]
  41. Vernimmen, D.; Marques-Kranc, F.; Sharpe, J.A.; Sloane-Stanley, J.A.; Wood, W.G.; Wallace, H.A.C.; Smith, A.J.H.; Higgs, D.R. Chromosome Looping at the Human α-Globin Locus Is Mediated via the Major Upstream Regulatory Element (HS −40). Blood 2009, 114, 4253–4260. [Google Scholar] [CrossRef] [PubMed]
  42. Garrick, D.; Gobbi, M.D.; Samara, V.; Rugless, M.; Holland, M.; Ayyub, H.; Lower, K.; Sloane-Stanley, J.; Gray, N.; Koch, C.; et al. The Role of the Polycomb Complex in Silencing α-Globin Gene Expression in Nonerythroid Cells. Blood 2008, 112, 3889–3899. [Google Scholar] [CrossRef] [PubMed]
  43. Vernimmen, D.; Lynch, M.D.; Gobbi, M.D.; Garrick, D.; Sharpe, J.A.; Sloane-Stanley, J.A.; Smith, A.J.H.; Higgs, D.R. Polycomb Eviction as a New Distant Enhancer Function. Genes Dev. 2011, 25, 1583–1588. [Google Scholar] [CrossRef] [PubMed]
  44. Oudelaar, A.M.; Davies, J.O.J.; Hanssen, L.L.P.; Telenius, J.M.; Schwessinger, R.; Liu, Y.; Brown, J.M.; Downes, D.J.; Chiariello, A.M.; Bianco, S.; et al. Single-Allele Chromatin Interactions Identify Regulatory Hubs in Dynamic Compartmentalized Domains. Nat. Genet. 2018, 50, 1744–1751. [Google Scholar] [CrossRef]
  45. Chiariello, A.M.; Bianco, S.; Oudelaar, A.M.; Esposito, A.; Annunziatella, C.; Fiorillo, L.; Conte, M.; Corrado, A.; Prisco, A.; Larke, M.S.C.; et al. A Dynamic Folded Hairpin Conformation Is Associated with α-Globin Activation in Erythroid Cells. Cell Rep. 2020, 30, 2125–2135.e5. [Google Scholar] [CrossRef]
  46. Oudelaar, A.M.; Beagrie, R.A.; Gosden, M.; de Ornellas, S.; Georgiades, E.; Kerry, J.; Hidalgo, D.; Carrelha, J.; Shivalingam, A.; El-Sagheer, A.H.; et al. Dynamics of the 4D Genome during in Vivo Lineage Specification and Differentiation. Nat. Commun. 2020, 11, 2722. [Google Scholar] [CrossRef]
  47. Hanssen, L.L.P.; Kassouf, M.T.; Oudelaar, A.M.; Biggs, D.; Preece, C.; Downes, D.J.; Gosden, M.; Sharpe, J.A.; Sloane-Stanley, J.A.; Hughes, J.R.; et al. Tissue-Specific CTCF–Cohesin-Mediated Chromatin Architecture Delimits Enhancer Interactions and Function in Vivo. Nat. Cell Biol. 2017, 19, 952–961. [Google Scholar] [CrossRef]
  48. Georgiades, E.; Harrold, C.; Roberts, N.; Kassouf, M.; Riva, S.G.; Sanders, E.; Downes, D.; Francis, H.S.; Blayney, J.; Oudelaar, A.M.; et al. Active Regulatory Elements Recruit Cohesin to Establish Cell Specific Chromatin Domains. Sci. Rep. 2025, 15, 11780. [Google Scholar] [CrossRef]
  49. Hua, P.; Badat, M.; Hanssen, L.L.P.; Hentges, L.D.; Crump, N.; Downes, D.J.; Jeziorska, D.M.; Oudelaar, A.M.; Schwessinger, R.; Taylor, S.; et al. Defining Genome Architecture at Base-Pair Resolution. Nature 2021, 595, 125–129. [Google Scholar] [CrossRef]
  50. Sanborn, A.; Rao, S.; Huang, S.; Durand, N.; Huntley, M.; Jewett, A.; Bochkov, I.; Chinnappan, D.; Cutkosky, A.; Li, J.; et al. Chromatin Extrusion Explains Key Features of Loop and Domain Formation in Wild-type and Engineered Genomes. FASEB J. 2016, 30, 588.1. [Google Scholar] [CrossRef]
  51. Kassouf, M.T.; Francis, H.S.; Gosden, M.; Suciu, M.C.; Downes, D.J.; Harrold, C.; Larke, M.; Oudelaar, M.; Cornell, L.; Blayney, J.; et al. The α-Globin Super-Enhancer Acts in an Orientation-Dependent Manner. Nat. Commun. 2025, 16, 1033. [Google Scholar] [CrossRef] [PubMed]
  52. Tanimoto, K.; Liu, Q.; Bungert, J.; Engel, J.D. Effects of Altered Gene Order or Orientation of the Locus Control Region on Human β-Globin Gene Expression in Mice. Nature 1999, 398, 344–348. [Google Scholar] [CrossRef] [PubMed]
  53. Noordermeer, D.; Branco, M.R.; Splinter, E.; Klous, P.; van IJcken, W.; Swagemakers, S.; Koutsourakis, M.; van der Spek, P.; Pombo, A.; Laat, W. de Transcription and Chromatin Organization of a Housekeeping Gene Cluster Containing an Integrated β-Globin Locus Control Region. PLoS Genet. 2008, 4, e1000016. [Google Scholar] [CrossRef]
  54. Felder, A.-K.; Tjalsma, S.J.D.; Verhagen, H.J.M.P.; Majied, R.; Verstegen, M.J.A.M.; Verheul, T.C.J.; Mohnani, R.; Gremmen, R.; Krijger, P.H.L.; Philipsen, S.; et al. Reactivation of Developmentally Silenced Globin Genes through Genomic Deletions Reveals That Enhancer Distance Matters. bioRxiv 2025. bioRxiv:2025.01.13.632719. [Google Scholar] [CrossRef]
  55. Shearstone, J.R.; Pop, R.; Bock, C.; Boyle, P.; Meissner, A.; Socolovsky, M. Global DNA Demethylation During Mouse Erythropoiesis in Vivo. Science 2011, 334, 799–802. [Google Scholar] [CrossRef]
  56. Bartholdy, B.; Lajugie, J.; Yan, Z.; Zhang, S.; Mukhopadhyay, R.; Greally, J.M.; Suzuki, M.; Bouhassira, E.E. Mechanisms of Establishment and Functional Significance of DNA Demethylation during Erythroid Differentiation. Blood Adv. 2018, 2, 1833–1852. [Google Scholar] [CrossRef]
  57. Lessard, S.; Beaudoin, M.; Benkirane, K.; Lettre, G. Comparison of DNA Methylation Profiles in Human Fetal and Adult Red Blood Cell Progenitors. Genome Med. 2015, 7, 1. [Google Scholar] [CrossRef]
  58. Palii, C.G.; Cheng, Q.; Gillespie, M.A.; Shannon, P.; Mazurczyk, M.; Napolitani, G.; Price, N.D.; Ranish, J.A.; Morrissey, E.; Higgs, D.R.; et al. Single-Cell Proteomics Reveal That Quantitative Changes in Co-Expressed Lineage-Specific Transcription Factors Determine Cell Fate. Cell Stem Cell 2019, 24, 812–820.e5. [Google Scholar] [CrossRef]
  59. Merika, M.; Orkin, S.H. DNA-Binding Specificity of GATA Family Transcription Factors. Mol. Cell. Biol. 1993, 13, 3999–4010. [Google Scholar] [CrossRef]
  60. Kang, H.; Mesquitta, W.-T.; Jung, H.S.; Moskvin, O.V.; Thomson, J.A.; Slukvin, I.I. GATA2 Is Dispensable for Specification of Hemogenic Endothelium but Promotes Endothelial-to-Hematopoietic Transition. Stem Cell Rep. 2018, 11, 197–211. [Google Scholar] [CrossRef]
  61. Peters, I.J.A.; de Pater, E.; Zhang, W. The Role of GATA2 in Adult Hematopoiesis and Cell Fate Determination. Front. Cell Dev. Biol. 2023, 11, 1250827. [Google Scholar] [CrossRef] [PubMed]
  62. Chiba, T.; Ikawa, Y.; Todokoro, K. GATA-1 Transactivates Erythropoietin Receptor Gene, and Erythropoietin Receptor-Mediated Signals Enhance GATA-1 Gene Expression. Nucleic Acids Res. 1991, 19, 3843–3848. [Google Scholar] [CrossRef] [PubMed]
  63. Maeda, T.; Ito, K.; Merghoub, T.; Poliseno, L.; Hobbs, R.M.; Wang, G.; Dong, L.; Maeda, M.; Dore, L.C.; Zelent, A.; et al. LRF Is an Essential Downstream Target of GATA1 in Erythroid Development and Regulates BIM-Dependent Apoptosis. Dev. Cell 2009, 17, 527–540. [Google Scholar] [CrossRef]
  64. Vyas, P.; Ault, K.; Jackson, C.W.; Orkin, S.H.; Shivdasani, R.A. Consequences of GATA-1 Deficiency in Megakaryocytes and Platelets. Blood 1999, 93, 2867–2875. [Google Scholar] [CrossRef]
  65. Shivdasani, R.A.; Fujiwara, Y.; McDevitt, M.A.; Orkin, S.H. A Lineage-selective Knockout Establishes the Critical Role of Transcription Factor GATA-1 in Megakaryocyte Growth and Platelet Development. EMBO J. 1997, 16, 3965–3973. [Google Scholar] [CrossRef]
  66. Fujiwara, Y.; Browne, C.P.; Cunniff, K.; Goff, S.C.; Orkin, S.H. Arrested Development of Embryonic Red Cell Precursors in Mouse Embryos Lacking Transcription Factor GATA-1. Proc. Natl. Acad. Sci. USA 1996, 93, 12355–12358. [Google Scholar] [CrossRef]
  67. Cantor, A.B.; Orkin, S.H. Coregulation of GATA Factors by the Friend of GATA (FOG) Family of Multitype Zinc Finger Proteins. Semin. Cell Dev. Biol. 2005, 16, 117–128. [Google Scholar] [CrossRef]
  68. Stachura, D.L.; Chou, S.T.; Weiss, M.J. Early Block to Erythromegakaryocytic Development Conferred by Loss of Transcription Factor GATA-1. Blood 2006, 107, 87–97. [Google Scholar] [CrossRef]
  69. MORCEAU, F.; SCHNEKENBURGER, M.; DICATO, M.; DIEDERICH, M. GATA-1: Friends, Brothers, and Coworkers. Ann. N. Y. Acad. Sci. 2004, 1030, 537–554. [Google Scholar] [CrossRef]
  70. Lowry, J.A.; Mackay, J.P. GATA-1: One Protein, Many Partners. Int. J. Biochem. Cell Biol. 2006, 38, 6–11. [Google Scholar] [CrossRef]
  71. Cantor, A.B.; Orkin, S.H. Transcriptional Regulation of Erythropoiesis: An Affair Involving Multiple Partners. Oncogene 2002, 21, 3368–3376. [Google Scholar] [CrossRef] [PubMed]
  72. Tsang, A.P.; Visvader, J.E.; Turner, C.A.; Fujiwara, Y.; Yu, C.; Weiss, M.J.; Crossley, M.; Orkin, S.H. FOG, a Multitype Zinc Finger Protein, Acts as a Cofactor for Transcription Factor GATA-1 in Erythroid and Megakaryocytic Differentiation. Cell 1997, 90, 109–119. [Google Scholar] [CrossRef] [PubMed]
  73. Shimizu, R.; Ohneda, K.; Engel, J.D.; Trainor, C.D.; Yamamoto, M. Transgenic Rescue of GATA-1–Deficient Mice with GATA-1 Lacking a FOG-1 Association Site Phenocopies Patients with X-Linked Thrombocytopenia. Blood 2004, 103, 2560–2567. [Google Scholar] [CrossRef] [PubMed]
  74. Letting, D.L.; Chen, Y.-Y.; Rakowski, C.; Reedy, S.; Blobel, G.A. Context-Dependent Regulation of GATA-1 by Friend of GATA-1. Proc. Natl. Acad. Sci. USA 2004, 101, 476–481. [Google Scholar] [CrossRef]
  75. Crispino, J.D.; Lodish, M.B.; MacKay, J.P.; Orkin, S.H. Use of Altered Specificity Mutants to Probe a Specific Protein–Protein Interaction in Differentiation the GATA-1:FOG Complex. Mol. Cell 1999, 3, 219–228. [Google Scholar] [CrossRef]
  76. Chickarmane, V.; Enver, T.; Peterson, C. Computational Modeling of the Hematopoietic Erythroid-Myeloid Switch Reveals Insights into Cooperativity, Priming, and Irreversibility. PLoS Comput. Biol. 2009, 5, e1000268. [Google Scholar] [CrossRef]
  77. Burda, P.; Curik, N.; Kokavec, J.; Pospisil, V.; Skoultchi, A.I.; Zavadil, J.; Stopka, T. Fog1 and Cebpa Are DNA Targets of GATA-1/PU.1 Antagonism during Leukemia Differentiation. Blood 2007, 110, 4121. [Google Scholar] [CrossRef]
  78. Fujiwara, T.; Sasaki, K.; Saito, K.; Hatta, S.; Ichikawa, S.; Kobayashi, M.; Okitsu, Y.; Fukuhara, N.; Onishi, Y.; Harigae, H. Forced FOG1 Expression in Erythroleukemia Cells: Induction of Erythroid Genes and Repression of Myelo-Lymphoid Transcription Factor PU.1. Biochem. Biophys. Res. Commun. 2017, 485, 380–387. [Google Scholar] [CrossRef]
  79. Boyes, J.; Byfield, P.; Nakatani, Y.; Ogryzko, V. Regulation of Activity of the Transcription Factor GATA-1 by Acetylation. Nature 1998, 396, 594–598. [Google Scholar] [CrossRef]
  80. Hung, H.-L.; Lau, J.; Kim, A.Y.; Weiss, M.J.; Blobel, G.A. CREB-Binding Protein Acetylates Hematopoietic Transcription Factor GATA-1 at Functionally Important Sites. Mol. Cell. Biol. 1999, 19, 3496–3505. [Google Scholar] [CrossRef]
  81. Papadopoulos, G.L.; Karkoulia, E.; Tsamardinos, I.; Porcher, C.; Ragoussis, J.; Bungert, J.; Strouboulis, J. GATA-1 Genome-Wide Occupancy Associates with Distinct Epigenetic Profiles in Mouse Fetal Liver Erythropoiesis. Nucleic Acids Res. 2013, 41, 4938–4948. [Google Scholar] [CrossRef] [PubMed]
  82. Gao, Z.; Huang, Z.; Olivey, H.E.; Gurbuxani, S.; Crispino, J.D.; Svensson, E.C. FOG-1-mediated Recruitment of NuRD Is Required for Cell Lineage Re-enforcement during Haematopoiesis. EMBO J. 2010, 29, 457–468. [Google Scholar] [CrossRef] [PubMed]
  83. Gregory, G.D.; Miccio, A.; Bersenev, A.; Wang, Y.; Hong, W.; Zhang, Z.; Poncz, M.; Tong, W.; Blobel, G.A. FOG1 Requires NuRD to Promote Hematopoiesis and Maintain Lineage Fidelity within the Megakaryocytic-Erythroid Compartment. Blood 2010, 115, 2156–2166. [Google Scholar] [CrossRef] [PubMed]
  84. Mancini, E.; Sanjuan-Pla, A.; Luciani, L.; Moore, S.; Grover, A.; Zay, A.; Rasmussen, K.D.; Luc, S.; Bilbao, D.; O’Carroll, D.; et al. FOG-1 and GATA-1 Act Sequentially to Specify Definitive Megakaryocytic and Erythroid Progenitors. EMBO J. 2012, 31, 351–365. [Google Scholar] [CrossRef]
  85. Ferreira, R.; Wai, A.; Shimizu, R.; Gillemans, N.; Rottier, R.; von Lindern, M.; Ohneda, K.; Grosveld, F.; Yamamoto, M.; Philipsen, S. Dynamic Regulation of Gata Factor Levels Is More Important than Their Identity. Blood 2007, 109, 5481–5490. [Google Scholar] [CrossRef]
  86. Robb, L.; Elwood, N.J.; Elefanty, A.G.; Köntgen, F.; Li, R.; Barnett, L.D.; Begley, C.G. The Scl Gene Product Is Required for the Generation of All Hematopoietic Lineages in the Adult Mouse. EMBO J. 1996, 15, 4123–4129. [Google Scholar] [CrossRef]
  87. Porcher, C.; Swat, W.; Rockwell, K.; Fujiwara, Y.; Alt, F.W.; Orkin, S.H. The T Cell Leukemia Oncoprotein SCL/Tal-1 Is Essential for Development of All Hematopoietic Lineages. Cell 1996, 86, 47–57. [Google Scholar] [CrossRef]
  88. Porcher, C.; Chagraoui, H.; Kristiansen, M.S. SCL/TAL1: A Multifaceted Regulator from Blood Development to Disease. Blood 2017, 129, 2051–2060. [Google Scholar] [CrossRef]
  89. Palii, C.G.; Perez-Iratxeta, C.; Yao, Z.; Cao, Y.; Dai, F.; Davison, J.; Atkins, H.; Allan, D.; Dilworth, F.J.; Gentleman, R.; et al. Differential Genomic Targeting of the Transcription Factor TAL1 in Alternate Haematopoietic Lineages. EMBO J. 2011, 30, 494–509. [Google Scholar] [CrossRef]
  90. Finger, L.R.; Kagan, J.; Christopher, G.; Kurtzberg, J.; Hershfield, M.S.; Nowell, P.C.; Croce, C.M. Involvement of the TCL5 Gene on Human Chromosome 1 in T-Cell Leukemia and Melanoma. Proc. Natl. Acad. Sci. USA 1989, 86, 5039–5043. [Google Scholar] [CrossRef]
  91. Sanda, T.; Lawton, L.N.; Barrasa, M.I.; Fan, Z.P.; Kohlhammer, H.; Gutierrez, A.; Ma, W.; Tatarek, J.; Ahn, Y.; Kelliher, M.A.; et al. Core Transcriptional Regulatory Circuit Controlled by the TAL1 Complex in Human T Cell Acute Lymphoblastic Leukemia. Cancer Cell 2012, 22, 209–221. [Google Scholar] [CrossRef] [PubMed]
  92. Anantharaman, A.; Lin, I.-J.; Barrow, J.; Liang, S.Y.; Masannat, J.; Strouboulis, J.; Huang, S.; Bungert, J. Role of Helix-Loop-Helix Proteins during Differentiation of Erythroid Cells. Mol. Cell. Biol. 2011, 31, 1332–1343. [Google Scholar] [CrossRef] [PubMed]
  93. Kassouf, M.T.; Hughes, J.R.; Taylor, S.; McGowan, S.J.; Soneji, S.; Green, A.L.; Vyas, P.; Porcher, C. Genome-Wide Identification of TAL1′s Functional Targets: Insights into Its Mechanisms of Action in Primary Erythroid Cells. Genome Res. 2010, 20, 1064–1083. [Google Scholar] [CrossRef] [PubMed]
  94. Yun, W.J.; Kim, Y.W.; Kang, Y.; Lee, J.; Dean, A.; Kim, A. The Hematopoietic Regulator TAL1 Is Required for Chromatin Looping between the β-Globin LCR and Human γ-Globin Genes to Activate Transcription. Nucleic Acids Res. 2014, 42, 4283–4293. [Google Scholar] [CrossRef] [PubMed]
  95. Malyavantham, K.S.; Bhattacharya, S.; Barbeitos, M.; Mukherjee, L.; Xu, J.; Fackelmayer, F.O.; Berezney, R. Identifying Functional Neighborhoods within the Cell Nucleus: Proximity Analysis of Early S-phase Replicating Chromatin Domains to Sites of Transcription, RNA Polymerase II, HP1γ, Matrin 3 and SAF-A. J. Cell. Biochem. 2008, 105, 391–403. [Google Scholar] [CrossRef]
  96. Porcu, S.; Manchinu, M.F.; Marongiu, M.F.; Sogos, V.; Poddie, D.; Asunis, I.; Porcu, L.; Marini, M.G.; Moi, P.; Cao, A.; et al. Klf1 Affects DNase II-Alpha Expression in the Central Macrophage of a Fetal Liver Erythroblastic Island: A Non-Cell-Autonomous Role in Definitive Erythropoiesis. Mol. Cell. Biol. 2011, 31, 4144–4154. [Google Scholar] [CrossRef]
  97. Xue, L.; Galdass, M.; Gnanapragasam, M.N.; Manwani, D.; Bieker, J.J. Extrinsic and Intrinsic Control by EKLF (KLF1) within a Specialized Erythroid Niche. Development 2014, 141, 2245–2254. [Google Scholar] [CrossRef]
  98. Mateus, J.; Grifoni, A.; Tarke, A.; Sidney, J.; Ramirez, S.I.; Dan, J.M.; Burger, Z.C.; Rawlings, S.A.; Smith, D.M.; Phillips, E.; et al. Selective and Cross-Reactive SARS-CoV-2 T Cell Epitopes in Unexposed Humans. Science 2020, 370, 89–94. [Google Scholar] [CrossRef]
  99. Frontelo, P.; Manwani, D.; Galdass, M.; Karsunky, H.; Lohmann, F.; Gallagher, P.G.; Bieker, J.J. Novel Role for EKLF in Megakaryocyte Lineage Commitment. Blood 2007, 110, 3871–3880. [Google Scholar] [CrossRef]
  100. Pilon, A.M.; Arcasoy, M.O.; Dressman, H.K.; Vayda, S.E.; Maksimova, Y.D.; Sangerman, J.I.; Gallagher, P.G.; Bodine, D.M. Failure of Terminal Erythroid Differentiation in EKLF-Deficient Mice Is Associated with Cell Cycle Perturbation and Reduced Expression of E2F2. Mol. Cell. Biol. 2008, 28, 7394–7401. [Google Scholar] [CrossRef]
  101. Norton, L.J.; Hallal, S.; Stout, E.S.; Funnell, A.P.W.; Pearson, R.C.M.; Crossley, M.; Quinlan, K.G.R. Direct Competition between DNA Binding Factors Highlights the Role of Krüppel-like Factor 1 in the Erythroid/Megakaryocyte Switch. Sci. Rep. 2017, 7, 3137. [Google Scholar] [CrossRef] [PubMed]
  102. Tallack, M.R.; Whitington, T.; Yuen, W.S.; Wainwright, E.N.; Keys, J.R.; Gardiner, B.B.; Nourbakhsh, E.; Cloonan, N.; Grimmond, S.M.; Bailey, T.L.; et al. A Global Role for KLF1 in Erythropoiesis Revealed by ChIP-Seq in Primary Erythroid Cells. Genome Res. 2010, 20, 1052–1063. [Google Scholar] [CrossRef] [PubMed]
  103. Hodge, D.; Coghill, E.; Keys, J.; Maguire, T.; Hartmann, B.; McDowall, A.; Weiss, M.; Grimmond, S.; Perkins, A. A Global Role for EKLF in Definitive and Primitive Erythropoiesis. Blood 2006, 107, 3359–3370. [Google Scholar] [CrossRef]
  104. Drissen, R.; von Lindern, M.; Kolbus, A.; Driegen, S.; Steinlein, P.; Beug, H.; Grosveld, F.; Philipsen, S. The Erythroid Phenotype of EKLF-Null Mice: Defects in Hemoglobin Metabolism and Membrane Stability. Mol. Cell. Biol. 2005, 25, 5205–5214. [Google Scholar] [CrossRef]
  105. Nuez, B.; Michalovich, D.; Bygrave, A.; Ploemacher, R.; Grosveld, F. Defective Haematopoiesis in Fetal Liver Resulting from Inactivation of the EKLF Gene. Nature 1995, 375, 316–318. [Google Scholar] [CrossRef]
  106. Kingsley, P.D.; Malik, J.; Emerson, R.L.; Bushnell, T.P.; McGrath, K.E.; Bloedorn, L.A.; Bulger, M.; Palis, J. “Maturational” Globin Switching in Primary Primitive Erythroid Cells. Blood 2006, 107, 1665–1672. [Google Scholar] [CrossRef]
  107. Sankaran, V.G.; Xu, J.; Orkin, S.H. Advances in the Understanding of Haemoglobin Switching. Br. J. Haematol. 2010, 149, 181–194. [Google Scholar] [CrossRef]
  108. Peschle, C.; Mavilio, F.; Carè, A.; Migliaccio, G.; Migliaccio, A.R.; Salvo, G.; Samoggia, P.; Petti, S.; Guerriero, R.; Marinucci, M.; et al. Haemoglobin Switching in Human Embryos: Asynchrony of ζ → α and ε → γ-Globin Switches in Primitive and Definitive Erythropoietic Lineage. Nature 1985, 313, 235–238. [Google Scholar] [CrossRef]
  109. Parkins, A.C.; Sharpe, A.H.; Orkin, S.H. Lethal β-Thalassaemia in Mice Lacking the Erythroid CACCC-Transcription Factor EKLF. Nature 1995, 375, 318–322. [Google Scholar] [CrossRef]
  110. Zhang, W.; Kadam, S.; Emerson, B.M.; Bieker, J.J. Site-Specific Acetylation by P300 or CREB Binding Protein Regulates Erythroid Krüppel-Like Factor Transcriptional Activity via Its Interaction with the SWI-SNF Complex. Mol. Cell. Biol. 2001, 21, 2413–2422. [Google Scholar] [CrossRef]
  111. Waye, J.S.; Eng, B. Krüppel-like Factor 1: Hematologic Phenotypes Associated with KLF1 Gene Mutations. Int. J. Lab. Hematol. 2015, 37, 78–84. [Google Scholar] [CrossRef] [PubMed]
  112. Tallack, M.R.; Keys, J.R.; Humbert, P.O.; Perkins, A.C. EKLF/KLF1 Controls Cell Cycle Entry via Direct Regulation of E2f2*. J. Biol. Chem. 2009, 284, 20966–20974. [Google Scholar] [CrossRef] [PubMed]
  113. Wells, M.; Steiner, L. Epigenetic and Transcriptional Control of Erythropoiesis. Front. Genet. 2022, 13, 805265. [Google Scholar] [CrossRef] [PubMed]
  114. Jindal, G.A.; Farley, E.K. Enhancer Grammar in Development, Evolution, and Disease: Dependencies and Interplay. Dev. Cell 2021, 56, 575–587. [Google Scholar] [CrossRef]
  115. Snetkova, V.; Ypsilanti, A.R.; Akiyama, J.A.; Mannion, B.J.; Plajzer-Frick, I.; Novak, C.S.; Harrington, A.N.; Pham, Q.T.; Kato, M.; Zhu, Y.; et al. Ultraconserved Enhancer Function Does Not Require Perfect Sequence Conservation. Nat. Genet. 2021, 53, 521–528. [Google Scholar] [CrossRef]
  116. Hindorff, L.A.; Sethupathy, P.; Junkins, H.A.; Ramos, E.M.; Mehta, J.P.; Collins, F.S.; Manolio, T.A. Potential Etiologic and Functional Implications of Genome-Wide Association Loci for Human Diseases and Traits. Proc. Natl. Acad. Sci. USA 2009, 106, 9362–9367. [Google Scholar] [CrossRef]
  117. Huang, J.; Liu, X.; Li, D.; Shao, Z.; Cao, H.; Zhang, Y.; Trompouki, E.; Bowman, T.V.; Zon, L.I.; Yuan, G.-C.; et al. Dynamic Control of Enhancer Repertoires Drives Lineage and Stage-Specific Transcription during Hematopoiesis. Dev. Cell 2016, 36, 9–23. [Google Scholar] [CrossRef]
  118. Wakabayashi, A.; Ulirsch, J.C.; Ludwig, L.S.; Fiorini, C.; Yasuda, M.; Choudhuri, A.; McDonel, P.; Zon, L.I.; Sankaran, V.G. Insight into GATA1 Transcriptional Activity through Interrogation of Cis Elements Disrupted in Human Erythroid Disorders. Proc. Natl. Acad. Sci. USA 2016, 113, 4434–4439. [Google Scholar] [CrossRef]
  119. Jacob, G.F.; Raper, A.B. Hereditary Persistence of Foetal Haemoglobin Production, and Its Interaction with the Sickle-Cell Trait. Br. J. Haematol. 1958, 4, 138–149. [Google Scholar] [CrossRef]
  120. Steinberg, M.H. Fetal Hemoglobin in Sickle Hemoglobinopathies: High HbF Genotypes and Phenotypes. J. Clin. Med. 2020, 9, 3782. [Google Scholar] [CrossRef]
  121. Wonkam, A.; Bitoungui, V.J.N.; Vorster, A.A.; Ramesar, R.; Cooper, R.S.; Tayo, B.; Lettre, G.; Ngogang, J. Association of Variants at BCL11A and HBS1L-MYB with Hemoglobin F and Hospitalization Rates among Sickle Cell Patients in Cameroon. PLoS ONE 2014, 9, e92506. [Google Scholar] [CrossRef] [PubMed]
  122. Wen, T.; Boyden, S.E.; Hocutt, C.M.; Lewis, R.G.; Baldwin, E.E.; Vagher, J.; Andrews, A.; Nicholas, T.J.; Chapin, A.; Fan, E.M.; et al. Identification of 2 Novel Noncoding Variants in Patients with Diamond-Blackfan Anemia Syndrome by Whole Genome Sequencing. Blood Adv. 2025, 9, 2443–2452. [Google Scholar] [CrossRef] [PubMed]
  123. Liu, N.; Hargreaves, V.V.; Zhu, Q.; Kurland, J.V.; Hong, J.; Kim, W.; Sher, F.; Macias-Trevino, C.; Rogers, J.M.; Kurita, R.; et al. Direct Promoter Repression by BCL11A Controls the Fetal to Adult Hemoglobin Switch. Cell 2018, 173, 430–442.e17. [Google Scholar] [CrossRef] [PubMed]
  124. Rahman, S.; Mansour, M.R. The Role of Noncoding Mutations in Blood Cancers. Dis. Model. Mech. 2019, 12, dmm041988. [Google Scholar] [CrossRef]
  125. Weber, L.; Frati, G.; Felix, T.; Hardouin, G.; Casini, A.; Wollenschlaeger, C.; Meneghini, V.; Masson, C.; Cian, A.D.; Chalumeau, A.; et al. Editing a γ-Globin Repressor Binding Site Restores Fetal Hemoglobin Synthesis and Corrects the Sickle Cell Disease Phenotype. Sci. Adv. 2020, 6, eaay9392. [Google Scholar] [CrossRef]
  126. Shaikho, E.M.; Farrell, J.J.; Alsultan, A.; Qutub, H.; Al-Ali, A.K.; Figueiredo, M.S.; Chui, D.H.K.; Farrer, L.A.; Murphy, G.J.; Mostoslavsky, G.; et al. A Phased SNP-Based Classification of Sickle Cell Anemia HBB Haplotypes. BMC Genom. 2017, 18, 608. [Google Scholar] [CrossRef]
  127. Akinbami, A.O.; Campbell, A.D.; Han, Z.J.; Luo, H.-Y.; Chui, D.H.K.; Steinberg, M.H. Hereditary Persistence of Fetal Hemoglobin Caused by Single Nucleotide Promoter Mutations in Sickle Cell Trait and Hb SC Disease. Hemoglobin 2016, 40, 64–65. [Google Scholar] [CrossRef]
  128. Coleman, M.B.; Adams, J.G.; Steinberg, M.H.; Plonczynski, M.W.; Harrell, A.H.; Castro, O.; Winter, W.P. GγAγ(Β+) Hereditary Persistence of Fetal Hemoglobin: The Gγ–158 C → T Mutation in Cis to the − 175 T → C Mutation of the Aγ-globin Gene Results in Increased Gγ-globin Synthesis. Am. J. Hematol. 1993, 42, 186–190. [Google Scholar] [CrossRef]
  129. Choudhuri, A.; Trompouki, E.; Abraham, B.J.; Colli, L.M.; Kock, K.H.; Mallard, W.; Yang, M.-L.; Vinjamur, D.S.; Ghamari, A.; Sporrij, A.; et al. Common Variants in Signaling Transcription-Factor-Binding Sites Drive Phenotypic Variability in Red Blood Cell Traits. Nat. Genet. 2020, 52, 1333–1345. [Google Scholar] [CrossRef]
  130. Menzel, S.; Jiang, J.; Silver, N.; Gallagher, J.; Cunningham, J.; Surdulescu, G.; Lathrop, M.; Farrall, M.; Spector, T.D.; Thein, S.L. The HBS1L-MYB Intergenic Region on Chromosome 6q23.3 Influences Erythrocyte, Platelet, and Monocyte Counts in Humans. Blood 2007, 110, 3624–3626. [Google Scholar] [CrossRef]
  131. Ropero, P.; Peral, M.; Sánchez-Martínez, L.J.; Rochas, S.; Gómez-Álvarez, M.; Nieto, J.M.; González, F.A.; Villegas, A.; Benavente, C. Phenotype of Sickle Cell Disease. Correlation of Haplotypes and Polymorphisms in Cluster β, BCL11A, and HBS1L−MYB. Pilot Study. Front. Med. 2025, 12, 1347026. [Google Scholar] [CrossRef] [PubMed]
  132. Munkongdee, T.; Tongsima, S.; Ngamphiw, C.; Wangkumhang, P.; Peerapittayamongkol, C.; Hashim, H.B.; Fucharoen, S.; Svasti, S. Predictive SNPs for Β0-Thalassemia/HbE Disease Severity. Sci. Rep. 2021, 11, 10352. [Google Scholar] [CrossRef] [PubMed]
  133. Habara, A.H.; Shaikho, E.M.; Steinberg, M.H. Fetal Hemoglobin in Sickle Cell Anemia: The Arab-Indian Haplotype and New Therapeutic Agents. Am. J. Hematol. 2017, 92, 1233–1242. [Google Scholar] [CrossRef] [PubMed]
  134. Kulshrestha, A.; Garrett, M.E.; Telen, M.J.; Ashley-Koch, A.E. Novel Loci Associated with Acute Chest Syndrome in Sickle Cell Disease Patients. Blood 2024, 144, 1110. [Google Scholar] [CrossRef]
  135. Han, W.; Qi, M.; Ye, K.; He, Q.; Yekefenhazi, D.; Xu, D.; Han, F.; Li, W. Genome-Wide Association Study for Growth Traits with 1066 Individuals in Largemouth Bass (Micropterus salmoides). Front. Mol. Biosci. 2024, 11, 1443522. [Google Scholar] [CrossRef]
  136. Galarneau, G.; Coady, S.; Garrett, M.E.; Jeffries, N.; Puggal, M.; Paltoo, D.; Soldano, K.; Guasch, A.; Ashley-Koch, A.E.; Telen, M.J.; et al. Gene-Centric Association Study of Acute Chest Syndrome and Painful Crisis in Sickle Cell Disease Patients. Blood 2013, 122, 434–442. [Google Scholar] [CrossRef]
  137. Papasavva, T.; van IJcken, W.F.J.; Kockx, C.E.M.; van den Hout, M.C.G.N.; Kountouris, P.; Kythreotis, L.; Kalogirou, E.; Grosveld, F.G.; Kleanthous, M. Next Generation Sequencing of SNPs for Non-Invasive Prenatal Diagnosis: Challenges and Feasibility as Illustrated by an Application to β-Thalassaemia. Eur. J. Hum. Genet. 2013, 21, 1403–1410. [Google Scholar] [CrossRef]
  138. Ulirsch, J.C.; Nandakumar, S.K.; Wang, L.; Giani, F.C.; Zhang, X.; Rogov, P.; Melnikov, A.; McDonel, P.; Do, R.; Mikkelsen, T.S.; et al. Systematic Functional Dissection of Common Genetic Variation Affecting Red Blood Cell Traits. Cell 2016, 165, 1530–1545. [Google Scholar] [CrossRef]
  139. Möller, M.; Lee, Y.Q.; Vidovic, K.; Kjellström, S.; Björkman, L.; Storry, J.R.; Olsson, M.L. Disruption of a GATA1-Binding Motif Upstream of XG/PBDX Abolishes Xga Expression and Resolves the Xg Blood Group System. Blood 2018, 132, 334–338. [Google Scholar] [CrossRef]
  140. Fang, F.; Xu, J.; Kang, Y.; Ren, H.; Muyey, D.M.; Chen, X.; Tan, Y.; Xu, Z.; Wang, H. GATA2 Rs2335052 and GATA2 Rs78245253 Single-nucleotide Polymorphisms in Chinese Patients with Acute Myelocytic Leukemia. Int. J. Lab. Hematol. 2021, 43, 1491–1500. [Google Scholar] [CrossRef]
  141. Russo, R.; Andolfo, I.; Gambale, A.; Rosa, G.D.; Manna, F.; Arillo, A.; Wandroo, F.; Bisconte, M.G.; Iolascon, A. GATA1 Erythroid-Specific Regulation of SEC23B Expression and Its Implication in the Pathogenesis of Congenital Dyserythropoietic Anemia Type II. Haematologica 2017, 102, e371–e374. [Google Scholar] [CrossRef] [PubMed]
  142. Kaneko, K.; Furuyama, K.; Fujiwara, T.; Kobayashi, R.; Ishida, H.; Harigae, H.; Shibahara, S. Identification of a Novel Erythroid-Specific Enhancer for the ALAS2 Gene and Its Loss-of-Function Mutation Which Is Associated with Congenital Sideroblastic Anemia. Haematologica 2014, 99, 252–261. [Google Scholar] [CrossRef] [PubMed]
  143. Sankaran, V.G.; Ghazvinian, R.; Do, R.; Thiru, P.; Vergilio, J.-A.; Beggs, A.H.; Sieff, C.A.; Orkin, S.H.; Nathan, D.G.; Lander, E.S.; et al. Exome Sequencing Identifies GATA1 Mutations Resulting in Diamond-Blackfan Anemia. J. Clin. Investig. 2012, 122, 2439–2443. [Google Scholar] [CrossRef] [PubMed]
  144. Hollanda, L.M.; Lima, C.S.P.; Cunha, A.F.; Albuquerque, D.M.; Vassallo, J.; Ozelo, M.C.; Joazeiro, P.P.; Saad, S.T.O.; Costa, F.F. An Inherited Mutation Leading to Production of Only the Short Isoform of GATA-1 Is Associated with Impaired Erythropoiesis. Nat. Genet. 2006, 38, 807–812. [Google Scholar] [CrossRef]
  145. Stachorski, L.; Heckl, D.; Thangapandi, V.R.; Maroz, A.; Reinhardt, D.; Klusmann, J.-H. GATA1-Centered Genetic Network on Chromosome 21 Drives Down Syndrome Acute Megakaryoblastic Leukemia. Blood 2014, 124, 4310. [Google Scholar] [CrossRef]
  146. Kosmidou, A.; Tragiannidis, A.; Gavriilaki, E. Myeloid Leukemia of Down Syndrome. Cancers 2023, 15, 3265. [Google Scholar] [CrossRef]
  147. Chen, C.-C.; Silberman, R.E.; Ma, D.; Perry, J.A.; Khalid, D.; Pikman, Y.; Amon, A.; Hemann, M.T.; Rowe, R.G. Inherent Genome Instability Underlies Trisomy 21-Associated Myeloid Malignancies. Leukemia 2024, 38, 521–529. [Google Scholar] [CrossRef]
  148. Baruchel, A.; Bourquin, J.-P.; Crispino, J.; Cuartero, S.; Hasle, H.; Hitzler, J.; Klusmann, J.-H.; Izraeli, S.; Lane, A.A.; Malinge, S.; et al. Down Syndrome and Leukemia: From Basic Mechanisms to Clinical Advances. Haematologica 2023, 108, 2570–2581. [Google Scholar] [CrossRef]
  149. Malinge, S.; Bliss-Moreau, M.; Kirsammer, G.; Diebold, L.; Chlon, T.; Gurbuxani, S.; Crispino, J.D. Increased Dosage of the Chromosome 21 Ortholog Dyrk1a Promotes Megakaryoblastic Leukemia in a Murine Model of Down Syndrome. J. Clin. Investig. 2012, 122, 948–962. [Google Scholar] [CrossRef]
  150. Dore, L.C.; Amigo, J.D.; dos Santos, C.O.; Zhang, Z.; Gai, X.; Tobias, J.W.; Yu, D.; Klein, A.M.; Dorman, C.; Wu, W.; et al. A GATA-1-Regulated MicroRNA Locus Essential for Erythropoiesis. Proc. Natl. Acad. Sci. USA 2008, 105, 3333–3338. [Google Scholar] [CrossRef]
  151. Pase, L.; Layton, J.E.; Kloosterman, W.P.; Carradice, D.; Waterhouse, P.M.; Lieschke, G.J. miR-451 Regulates Zebrafish Erythroid Maturation in Vivo via Its Target Gata2. Blood 2009, 113, 1794–1804. [Google Scholar] [CrossRef] [PubMed]
  152. Xu, L.; Wu, F.; Yang, L.; Wang, F.; Zhang, T.; Deng, X.; Zhang, X.; Yuan, X.; Yan, Y.; Li, Y.; et al. miR-144/451 Inhibits C-Myc to Promote Erythroid Differentiation. FASEB J. 2020, 34, 13194–13210. [Google Scholar] [CrossRef] [PubMed]
  153. Khurana, E.; Fu, Y.; Chakravarty, D.; Demichelis, F.; Rubin, M.A.; Gerstein, M. Role of Non-Coding Sequence Variants in Cancer. Nat. Rev. Genet. 2016, 17, 93–108. [Google Scholar] [CrossRef] [PubMed]
  154. Zhao, A.; Zhou, H.; Yang, J.; Li, M.; Niu, T. Epigenetic Regulation in Hematopoiesis and Its Implications in the Targeted Therapy of Hematologic Malignancies. Signal Transduct. Target. Ther. 2023, 8, 71. [Google Scholar] [CrossRef]
  155. Sharma, V.; Wright, K.L.; Epling-Burnette, P.K.; Reuther, G.W. Metabolic Vulnerabilities and Epigenetic Dysregulation in Myeloproliferative Neoplasms. Front. Immunol. 2020, 11, 604142. [Google Scholar] [CrossRef]
  156. Mejía-Ochoa, M.; Toro, P.A.A.; Cardona-Arias, J.A. Systematization of Analytical Studies of Polycythemia Vera, Essential Thrombocythemia and Primary Myelofibrosis, and a Meta-Analysis of the Frequency of JAK2, CALR and MPL Mutations: 2000–2018. BMC Cancer 2019, 19, 590. [Google Scholar] [CrossRef]
  157. Grinfeld, J.; Nangalia, J.; Green, A.R. Molecular Determinants of Pathogenesis and Clinical Phenotype in Myeloproliferative Neoplasms. Haematologica 2017, 102, 7–17. [Google Scholar] [CrossRef]
  158. Oddsson, A.; Kristinsson, S.Y.; Helgason, H.; Gudbjartsson, D.F.; Masson, G.; Sigurdsson, A.; Jonasdottir, A.; Jonasdottir, A.; Steingrimsdottir, H.; Vidarsson, B.; et al. The Germline Sequence Variant Rs2736100_C in TERT Associates with Myeloproliferative Neoplasms. Leukemia 2014, 28, 1371–1374. [Google Scholar] [CrossRef]
  159. McKerrell, T.; Park, N.; Moreno, T.; Grove, C.S.; Ponstingl, H.; Stephens, J.; Group, U.S.S.; Crawley, C.; Craig, J.; Scott, M.A.; et al. Leukemia-Associated Somatic Mutations Drive Distinct Patterns of Age-Related Clonal Hemopoiesis. Cell Rep. 2015, 10, 1239–1245. [Google Scholar] [CrossRef]
  160. Genovese, G.; Kähler, A.K.; Handsaker, R.E.; Lindberg, J.; Rose, S.A.; Bakhoum, S.F.; Chambert, K.; Mick, E.; Neale, B.M.; Fromer, M.; et al. Clonal Hematopoiesis and Blood-Cancer Risk Inferred from Blood DNA Sequence. N. Engl. J. Med. 2014, 371, 2477–2487. [Google Scholar] [CrossRef]
  161. Khan, S.N.; Jankowska, A.M.; Mahfouz, R.; Dunbar, A.J.; Sugimoto, Y.; Hosono, N.; Hu, Z.; Cheriyath, V.; Vatolin, S.; Przychodzen, B.; et al. Multiple Mechanisms Deregulate EZH2 and Histone H3 Lysine 27 Epigenetic Changes in Myeloid Malignancies. Leukemia 2013, 27, 1301–1309. [Google Scholar] [CrossRef] [PubMed]
  162. Karlsson, S.; Papayannopoulou, T.; Schweiger, S.G.; Stamatoyannopoulos, G.; Nienhuis, A.W. Retroviral-Mediated Transfer of Genomic Globin Genes Leads to Regulated Production of RNA and Protein. Proc. Natl. Acad. Sci. USA 1987, 84, 2411–2415. [Google Scholar] [CrossRef]
  163. Dzierzak, E.A.; Papayannopoulou, T.; Mulligan, R.C. Lineage-Specific Expression of a Human β-Globin Gene in Murine Bone Marrow Transplant Recipients Reconstituted with Retrovirus-Transduced Stem Cells. Nature 1988, 331, 35–41. [Google Scholar] [CrossRef] [PubMed]
  164. Antoniou, M.; deBoer, E.; Habets, G.; Grosveld, F. The Human Beta-globin Gene Contains Multiple Regulatory Regions: Identification of One Promoter and Two Downstream Enhancers. EMBO J. 1988, 7, 377–384. [Google Scholar] [CrossRef]
  165. Collis, P.; Antoniou, M.; Grosveld, F. Definition of the Minimal Requirements within the Human Beta-globin Gene and the Dominant Control Region for High Level Expression. EMBO J. 1990, 9, 233–240. [Google Scholar] [CrossRef]
  166. Forrester, W.C.; Takegawa, S.; Papayannopoulou, T.; Stamatoyannopoulos, G.; Groudine, M. Evidence for a Locus Activation Region: The Formation of Developmentally Stable Hypersensitive Sites in Globin-Expressing Hybrids. Nucleic Acids Res. 1987, 15, 10159–10177. [Google Scholar] [CrossRef]
  167. Chow, C.-M.; Athanassiadou, A.; Raguz, S.; Psiouri, L.; Harland, L.; Malik, M.; Aitken, M.; Grosveld, F.; Antoniou, M. LCR-Mediated, Long-Term Tissue-Specific Gene Expression within Replicating Episomal Plasmid and Cosmid Vectors. Gene Ther. 2002, 9, 327–336. [Google Scholar] [CrossRef]
  168. Sadelain, M.; Wang, C.H.; Antoniou, M.; Grosveld, F.; Mulligan, R.C. Generation of a High-Titer Retroviral Vector Capable of Expressing High Levels of the Human Beta-Globin Gene. Proc. Natl. Acad. Sci. USA 1995, 92, 6728–6732. [Google Scholar] [CrossRef]
  169. May, C.; Rivella, S.; Callegari, J.; Heller, G.; Gaensler, K.M.L.; Luzzatto, L.; Sadelain, M. Therapeutic Haemoglobin Synthesis in β-Thalassaemic Mice Expressing Lentivirus-Encoded Human β-Globin. Nature 2000, 406, 82–86. [Google Scholar] [CrossRef]
  170. Pawliuk, R.; Westerman, K.A.; Fabry, M.E.; Payen, E.; Tighe, R.; Bouhassira, E.E.; Acharya, S.A.; Ellis, J.; London, I.M.; Eaves, C.J.; et al. Correction of Sickle Cell Disease in Transgenic Mouse Models by Gene Therapy. Science 2001, 294, 2368–2371. [Google Scholar] [CrossRef]
  171. Imren, S.; Payen, E.; Westerman, K.A.; Pawliuk, R.; Fabry, M.E.; Eaves, C.J.; Cavilla, B.; Wadsworth, L.D.; Beuzard, Y.; Bouhassira, E.E.; et al. Permanent and Panerythroid Correction of Murine β Thalassemia by Multiple Lentiviral Integration in Hematopoietic Stem Cells. Proc. Natl. Acad. Sci. USA 2002, 99, 14380–14385. [Google Scholar] [CrossRef] [PubMed]
  172. Negre, O.; Bartholomae, C.; Beuzard, Y.; Cavazzana, M.; Christiansen, L.; Courne, C.; Deichmann, A.; Denaro, M.; de Dreuzy, E.; Finer, M.; et al. Preclinical Evaluation of Efficacy and Safety of an Improved Lentiviral Vector for the Treatment of β-Thalassemia and Sickle Cell Disease. Curr. Gene Ther. 2014, 15, 64–81. [Google Scholar] [CrossRef] [PubMed]
  173. Miccio, A.; Cesari, R.; Lotti, F.; Rossi, C.; Sanvito, F.; Ponzoni, M.; Routledge, S.J.E.; Chow, C.-M.; Antoniou, M.N.; Ferrari, G. In Vivo Selection of Genetically Modified Erythroblastic Progenitors Leads to Long-Term Correction of β-Thalassemia. Proc. Natl. Acad. Sci. USA 2008, 105, 10547–10552. [Google Scholar] [CrossRef] [PubMed]
  174. Brendel, C.; Guda, S.; Renella, R.; Bauer, D.E.; Canver, M.C.; Kim, Y.-J.; Heeney, M.M.; Klatt, D.; Fogel, J.; Milsom, M.D.; et al. Lineage-Specific BCL11A Knockdown Circumvents Toxicities and Reverses Sickle Phenotype. J. Clin. Investig. 2016, 126, 3868–3878. [Google Scholar] [CrossRef]
  175. Cavazzana-Calvo, M.; Payen, E.; Negre, O.; Wang, G.; Hehir, K.; Fusil, F.; Down, J.; Denaro, M.; Brady, T.; Westerman, K.; et al. Transfusion Independence and HMGA2 Activation after Gene Therapy of Human β-Thalassaemia. Nature 2010, 467, 318–322. [Google Scholar] [CrossRef]
  176. Lee, Y.T.; de Vasconcellos, J.F.; Yuan, J.; Byrnes, C.; Noh, S.-J.; Meier, E.R.; Kim, K.S.; Rabel, A.; Kaushal, M.; Muljo, S.A.; et al. LIN28B-Mediated Expression of Fetal Hemoglobin and Production of Fetal-like Erythrocytes from Adult Human Erythroblasts Ex Vivo. Blood 2013, 122, 1034–1041. [Google Scholar] [CrossRef]
  177. Liu, M.; Maurano, M.T.; Wang, H.; Qi, H.; Song, C.-Z.; Navas, P.A.; Emery, D.W.; Stamatoyannopoulos, J.A.; Stamatoyannopoulos, G. Genomic Discovery of Potent Chromatin Insulators for Human Gene Therapy. Nat. Biotechnol. 2015, 33, 198–203. [Google Scholar] [CrossRef]
  178. Papayanni, P.-G.; Psatha, N.; Christofi, P.; Li, X.-G.; Melo, P.; Volpin, M.; Montini, E.; Liu, M.; Kaltsounis, G.; Yiangou, M.; et al. Investigating the Barrier Activity of Novel, Human Enhancer-Blocking Chromatin Insulators for Hematopoietic Stem Cell Gene Therapy. Hum. Gene Ther. 2021, 32, 1186–1199. [Google Scholar] [CrossRef]
  179. Roselli, E.A.; Mezzadra, R.; Frittoli, M.C.; Maruggi, G.; Biral, E.; Mavilio, F.; Mastropietro, F.; Amato, A.; Tonon, G.; Refaldi, C.; et al. Correction of Β-thalassemia Major by Gene Transfer in Haematopoietic Progenitors of Pediatric Patients. EMBO Mol. Med. 2010, 2, 315–328. [Google Scholar] [CrossRef]
  180. Morgan, R.A.; Ma, F.; Unti, M.J.; Brown, D.; Ayoub, P.G.; Tam, C.; Lathrop, L.; Aleshe, B.; Kurita, R.; Nakamura, Y.; et al. Creating New β-Globin-Expressing Lentiviral Vectors by High-Resolution Mapping of Locus Control Region Enhancer Sequences. Mol. Ther. Methods Clin. Dev. 2020, 17, 999–1013. [Google Scholar] [CrossRef]
  181. Nemeth, M.J.; Bodine, D.M.; Garrett, L.J.; Lowrey, C.H. An Erythroid-Specific Chromatin Opening Element Reorganizes β-Globin Promoter Chromatin Structure and Augments Gene Expression. Blood Cells Mol. Dis. 2001, 27, 767–780. [Google Scholar] [CrossRef] [PubMed]
  182. Peslak, S.A.; Demirci, S.; Chandra, V.; Ryu, B.; Bhardwaj, S.K.; Jiang, J.; Rupon, J.W.; Throm, R.E.; Uchida, N.; Leonard, A.; et al. Forced Enhancer-Promoter Rewiring to Alter Gene Expression in Animal Models. Mol. Ther. Nucleic Acids 2023, 31, 452–465. [Google Scholar] [CrossRef] [PubMed]
  183. Reitman, M.; Lee, E.; Westphal, H.; Felsenfeld, G. An Enhancer/Locus Control Region Is Not Sufficient To Open Chromatin. Mol. Cell. Biol. 1993, 13, 3990–3998. [Google Scholar] [CrossRef] [PubMed]
  184. Buzina, A.; Lo, M.Y.M.; Moffett, A.; Hotta, A.; Fussner, E.; Bharadwaj, R.R.; Pasceri, P.; Garcia-Martinez, J.V.; Bazett-Jones, D.P.; Ellis, J. β-Globin LCR and Intron Elements Cooperate and Direct Spatial Reorganization for Gene Therapy. PLoS Genet. 2008, 4, e1000051. [Google Scholar] [CrossRef]
  185. Gazouli, M.; Katsantoni, E.; Kosteas, T.; Anagnou, N.P. Persistent Fetal γ-Globin Expression in Adult Transgenic Mice Following Deletion of Two Silencer Elements Located 3′ to the Human Aγ-Globin Gene. Mol. Med. 2009, 15, 415–424. [Google Scholar] [CrossRef]
  186. Katsantoni, E.Z.; de Krom, M.; Kong-a-San, J.; Imam, A.M.A.; Grosveld, F.; Anagnou, N.P.; Strouboulis, J. An Embryonic-Specific Repressor Element Located 3′ to the Aγ-Globin Gene Influences Transcription of the Human β-Globin Locus in Transgenic Mice. Exp. Hematol. 2004, 32, 224–233. [Google Scholar] [CrossRef]
  187. Katsantoni, E.Z.; Langeveld, A.; Wai, A.W.K.; Drabek, D.; Grosveld, F.; Anagnou, N.P.; Strouboulis, J. Persistent γ-Globin Expression in Adult Transgenic Mice Is Mediated by HPFH-2, HPFH-3, and HPFH-6 Breakpoint Sequences. Blood 2003, 102, 3412–3419. [Google Scholar] [CrossRef]
  188. Drakopoulou, E.; Georgomanoli, M.; Lederer, C.W.; Panetsos, F.; Kleanthous, M.; Voskaridou, E.; Valakos, D.; Papanikolaou, E.; Anagnou, N.P. The Optimized γ-Globin Lentiviral Vector GGHI-MB-3D Leads to Nearly Therapeutic HbF Levels In Vitro in CD34+ Cells from Sickle Cell Disease Patients. Viruses 2022, 14, 2716. [Google Scholar] [CrossRef]
  189. Papanikolaou, E.; Georgomanoli, M.; Stamateris, E.; Panetsos, F.; Karagiorga, M.; Tsaftaridis, P.; Graphakos, S.; Anagnou, N.P. The New Self-Inactivating Lentiviral Vector for Thalassemia Gene Therapy Combining Two HPFH Activating Elements Corrects Human Thalassemic Hematopoietic Stem Cells. Hum. Gene Ther. 2012, 23, 15–31. [Google Scholar] [CrossRef]
  190. Morgan, R.A.; Unti, M.J.; Aleshe, B.; Brown, D.; Osborne, K.S.; Koziol, C.; Ayoub, P.G.; Smith, O.B.; O’Brien, R.; Tam, C.; et al. Improved Titer and Gene Transfer by Lentiviral Vectors Using Novel, Small β-Globin Locus Control Region Elements. Mol. Ther. 2020, 28, 328–340. [Google Scholar] [CrossRef]
  191. Uchida, N.; Ferrara, F.; Drysdale, C.M.; Yapundich, M.; Gamer, J.; Nassehi, T.; DiNicola, J.; Shibata, Y.; Wielgosz, M.; Kim, Y.-S.; et al. Sustained Fetal Hemoglobin Induction in Vivo Is Achieved by BCL11A Interference and Coexpressed Truncated Erythropoietin Receptor. Sci. Transl. Med. 2021, 13, eabb0411. [Google Scholar] [CrossRef] [PubMed]
  192. Tallack, M.R.; Magor, G.W.; Dartigues, B.; Sun, L.; Huang, S.; Fittock, J.M.; Fry, S.V.; Glazov, E.A.; Bailey, T.L.; Perkins, A.C. Novel Roles for KLF1 in Erythropoiesis Revealed by mRNA-Seq. Genome Res. 2012, 22, 2385–2398. [Google Scholar] [CrossRef] [PubMed]
  193. Siatecka, M.; Bieker, J.J. The Multifunctional Role of EKLF/KLF1 during Erythropoiesis. Blood 2011, 118, 2044–2054. [Google Scholar] [CrossRef] [PubMed]
  194. Borg, J.; Papadopoulos, P.; Georgitsi, M.; Gutiérrez, L.; Grech, G.; Fanis, P.; Phylactides, M.; Verkerk, A.J.M.H.; van der Spek, P.J.; Scerri, C.A.; et al. Haploinsufficiency for the Erythroid Transcription Factor KLF1 Causes Hereditary Persistence of Fetal Hemoglobin. Nat. Genet. 2010, 42, 801–805. [Google Scholar] [CrossRef]
  195. Voit, R.A.; Liao, X.; Caulier, A.; Antoszewski, M.; Cohen, B.; Armant, M.; Lu, H.Y.; Fleming, T.J.; Kamal, E.; Wahlster, L.; et al. Regulated GATA1 Expression as a Universal Gene Therapy for Diamond-Blackfan Anemia. Cell Stem Cell 2025, 32, 38–52.e6. [Google Scholar] [CrossRef]
  196. Psatha, N.; Sova, P.; Georgolopoulos, G.; Paschoudi, K.; Iwata, M.; Bloom, J.; Ulyanova, T.; Wang, H.; Kirtsou, A.; Vasiloudis, N.-I.; et al. Large-Scale Discovery of Potent, Compact and Erythroid Specific Enhancers for Gene Therapy Vectors. Nat. Commun. 2025, 16, 4325. [Google Scholar] [CrossRef]
  197. Lodish, H. PPARα Agonists and TGFβ Inhibitors Stimulate Red Blood Cell Production by Enhancing Self-Renewal of BFU-E Erythroid Progenitors. Blood 2016, 128, SCI-48. [Google Scholar] [CrossRef]
  198. Lee, H.-Y.; Gao, X.; Barrasa, M.I.; Li, H.; Elmes, R.R.; Peters, L.L.; Lodish, H.F. PPAR-α and Glucocorticoid Receptor Synergize to Promote Erythroid Progenitor Self-Renewal. Nature 2015, 522, 474–477. [Google Scholar] [CrossRef]
  199. Jane, S.M.; Cunningham, J.M. Understanding Fetal Globin Gene Expression: A Step Towards Effective Hbf Reactivation in Haemoglobinopathies. Br. J. Haematol. 1998, 102, 415–423. [Google Scholar] [CrossRef]
  200. Higgs, D.R.; Engel, J.D.; Stamatoyannopoulos, G. Thalassaemia. Lancet 2012, 379, 373–383. [Google Scholar] [CrossRef]
  201. Psatha, N.; Papayanni, P.-G.; Yannaki, E. A New Era for Hemoglobinopathies: More Than One Curative Option. Curr. Gene Ther. 2018, 17, 364–378. [Google Scholar] [CrossRef] [PubMed]
  202. Paschoudi, K.; Yannaki, E.; Psatha, N. Precision Editing as a Therapeutic Approach for β-Hemoglobinopathies. Int. J. Mol. Sci. 2023, 24, 9527. [Google Scholar] [CrossRef] [PubMed]
  203. Psatha, N.; Paschoudi, K.; Papadopoulou, A.; Yannaki, E. In Vivo Hematopoietic Stem Cell Genome Editing: Perspectives and Limitations. Genes 2022, 13, 2222. [Google Scholar] [CrossRef] [PubMed]
  204. Huang, P.; Keller, C.A.; Giardine, B.; Grevet, J.D.; Davies, J.O.J.; Hughes, J.R.; Kurita, R.; Nakamura, Y.; Hardison, R.C.; Blobel, G.A. Comparative Analysis of Three-Dimensional Chromosomal Architecture Identifies a Novel Fetal Hemoglobin Regulatory Element. Genes Dev. 2017, 31, 1704–1713. [Google Scholar] [CrossRef]
  205. Kiefer, C.M.; Lee, J.; Hou, C.; Dale, R.K.; Lee, Y.T.; Meier, E.R.; Miller, J.L.; Dean, A. Distinct Ldb1/NLI Complexes Orchestrate γ-Globin Repression and Reactivation through ETO2 in Human Adult Erythroid Cells. Blood 2011, 118, 6200–6208. [Google Scholar] [CrossRef]
  206. FORGET, B.G. Molecular Basis of Hereditary Persistence of Fetal Hemoglobin. Ann. N. York Acad. Sci. 1998, 850, 38–44. [Google Scholar] [CrossRef]
  207. Chakalova, L.; Osborne, C.S.; Dai, Y.-F.; Goyenechea, B.; Metaxotou-Mavromati, A.; Kattamis, A.; Kattamis, C.; Fraser, P. The Corfu Δβ Thalassemia Deletion Disrupts γ-Globin Gene Silencing and Reveals Post-Transcriptional Regulation of HbF Expression. Blood 2005, 105, 2154–2160. [Google Scholar] [CrossRef]
  208. Lux, C.T.; Pattabhi, S.; Berger, M.; Nourigat, C.; Flowers, D.A.; Negre, O.; Humbert, O.; Yang, J.G.; Lee, C.; Jacoby, K.; et al. TALEN-Mediated Gene Editing of HBG in Human Hematopoietic Stem Cells Leads to Therapeutic Fetal Hemoglobin Induction. Mol. Ther. Methods Clin. Dev. 2019, 12, 175–183. [Google Scholar] [CrossRef]
  209. Ye, L.; Wang, J.; Tan, Y.; Beyer, A.I.; Xie, F.; Muench, M.O.; Kan, Y.W. Genome Editing Using CRISPR-Cas9 to Create the HPFH Genotype in HSPCs: An Approach for Treating Sickle Cell Disease and β-Thalassemia. Proc. Natl. Acad. Sci. USA 2016, 113, 10661–10665. [Google Scholar] [CrossRef]
  210. Traxler, E.A.; Yao, Y.; Wang, Y.-D.; Woodard, K.J.; Kurita, R.; Nakamura, Y.; Hughes, J.R.; Hardison, R.C.; Blobel, G.A.; Li, C.; et al. A Genome-Editing Strategy to Treat β-Hemoglobinopathies That Recapitulates a Mutation Associated with a Benign Genetic Condition. Nat. Med. 2016, 22, 987–990. [Google Scholar] [CrossRef]
  211. Li, C.; Psatha, N.; Sova, P.; Gil, S.; Wang, H.; Kim, J.; Kulkarni, C.; Valensisi, C.; Hawkins, R.D.; Stamatoyannopoulos, G.; et al. Reactivation of γ-Globin in Adult β-YAC Mice after Ex Vivo and in Vivo Hematopoietic Stem Cell Genome Editing. Blood 2018, 131, 2915–2928. [Google Scholar] [CrossRef] [PubMed]
  212. Canver, M.C.; Smith, E.C.; Sher, F.; Pinello, L.; Sanjana, N.E.; Shalem, O.; Chen, D.D.; Schupp, P.G.; Vinjamur, D.S.; Garcia, S.P.; et al. BCL11A Enhancer Dissection by Cas9-Mediated in Situ Saturating Mutagenesis. Nature 2015, 527, 192–197. [Google Scholar] [CrossRef] [PubMed]
  213. Psatha, N.; Reik, A.; Phelps, S.; Zhou, Y.; Dalas, D.; Yannaki, E.; Levasseur, D.N.; Urnov, F.D.; Holmes, M.C.; Papayannopoulou, T. Disruption of the BCL11A Erythroid Enhancer Reactivates Fetal Hemoglobin in Erythroid Cells of Patients with β-Thalassemia Major. Mol. Ther. Methods Clin. Dev. 2018, 10, 313–326. [Google Scholar] [CrossRef] [PubMed]
  214. Humbert, O.; Peterson, C.W.; Norgaard, Z.K.; Radtke, S.; Kiem, H.-P. A Nonhuman Primate Transplantation Model to Evaluate Hematopoietic Stem Cell Gene Editing Strategies for β-Hemoglobinopathies. Mol. Ther. Methods Clin. Dev. 2018, 8, 75–86. [Google Scholar] [CrossRef]
  215. Wu, Y.; Zeng, J.; Roscoe, B.P.; Liu, P.; Yao, Q.; Lazzarotto, C.R.; Clement, K.; Cole, M.A.; Luk, K.; Baricordi, C.; et al. Highly Efficient Therapeutic Gene Editing of Human Hematopoietic Stem Cells. Nat. Med. 2019, 25, 776–783. [Google Scholar] [CrossRef]
  216. Psatha, N.; Georgakopoulou, A.; Li, C.; Nandakumar, V.; Georgolopoulos, G.; Acosta, R.; Paschoudi, K.; Nelson, J.; Chee, D.; Athanasiadou, A.; et al. Enhanced HbF Reactivation by Multiplex Mutagenesis of Thalassemic CD34+ Cells in Vitro and in Vivo. Blood 2021, 138, 1540–1553. [Google Scholar] [CrossRef]
  217. Locatelli, F.; Lang, P.; Wall, D.; Meisel, R.; Corbacioglu, S.; Li, A.M.; de la Fuente, J.; Shah, A.J.; Carpenter, B.; Kwiatkowski, J.L.; et al. Exagamglogene Autotemcel for Transfusion-Dependent β-Thalassemia. N. Engl. J. Med. 2024, 390, 1663–1676. [Google Scholar] [CrossRef]
  218. Janoudi, T.; Jagdale, M.; Wu, M.; Gorlla, S.; Zhang, P.; Shao, Y.; Li, L.; Bowley, S.R.; Marco, E.; Chang, K.-H. Nonclinical Evaluation of HBG1/2 and BCL11A as Genome-Editing Targets for the Treatment of β-Hemoglobinopathies. Blood Adv. 2025, 9, 808–813. [Google Scholar] [CrossRef]
  219. Frangoul, H.; Locatelli, F.; Sharma, A.; Bhatia, M.; Mapara, M.; Molinari, L.; Wall, D.; Liem, R.I.; Telfer, P.; Shah, A.J.; et al. Exagamglogene Autotemcel for Severe Sickle Cell Disease. N. Engl. J. Med. 2024, 390, 1649–1662. [Google Scholar] [CrossRef]
  220. Singh, A.; Irfan, H.; Fatima, E.; Nazir, Z.; Verma, A.; Akilimali, A. Revolutionary Breakthrough: FDA Approves CASGEVY, the First CRISPR/Cas9 Gene Therapy for Sickle Cell Disease. Ann. Med. Surg. 2024, 86, 4555. [Google Scholar] [CrossRef]
  221. Porto, E.M.; Komor, A.C.; Slaymaker, I.M.; Yeo, G.W. Base Editing: Advances and Therapeutic Opportunities. Nat. Rev. Drug Discov. 2020, 19, 839–859. [Google Scholar] [CrossRef] [PubMed]
  222. Jeong, Y.K.; Song, B.; Bae, S. Current Status and Challenges of DNA Base Editing Tools. Mol. Ther. 2020, 28, 1938–1952. [Google Scholar] [CrossRef] [PubMed]
  223. Gaudelli, N.M.; Lam, D.K.; Rees, H.A.; Solá-Esteves, N.M.; Barrera, L.A.; Born, D.A.; Edwards, A.; Gehrke, J.M.; Lee, S.-J.; Liquori, A.J.; et al. Directed Evolution of Adenine Base Editors with Increased Activity and Therapeutic Application. Nat. Biotechnol. 2020, 38, 892–900. [Google Scholar] [CrossRef]
  224. Wang, L.; Li, L.; Ma, Y.; Hu, H.; Li, Q.; Yang, Y.; Liu, W.; Yin, S.; Li, W.; Fu, B.; et al. Reactivation of γ-Globin Expression through Cas9 or Base Editor to Treat β-Hemoglobinopathies. Cell Res. 2020, 30, 276–278. [Google Scholar] [CrossRef]
  225. Li, C.; Georgakopoulou, A.; Mishra, A.; Gil, S.; Hawkins, R.D.; Yannaki, E.; Lieber, A. In Vivo HSPC Gene Therapy with Base Editors Allows for Efficient Reactivation of Fetal γ-Globin in β-YAC Mice. Blood Adv. 2021, 5, 1122–1135. [Google Scholar] [CrossRef]
  226. Li, C.; Georgakopoulou, A.; Newby, G.A.; Everette, K.A.; Nizamis, E.; Paschoudi, K.; Vlachaki, E.; Gil, S.; Anderson, A.K.; Koob, T.; et al. In Vivo Base Editing by a Single Intravenous Vector Injection for Treatment of Hemoglobinopathies. JCI Insight 2022, 7, e162939. [Google Scholar] [CrossRef]
  227. Antoniou, P.; Hardouin, G.; Martinucci, P.; Frati, G.; Felix, T.; Chalumeau, A.; Fontana, L.; Martin, J.; Masson, C.; Brusson, M.; et al. Base-Editing-Mediated Dissection of a γ-Globin Cis-Regulatory Element for the Therapeutic Reactivation of Fetal Hemoglobin Expression. Nat. Commun. 2022, 13, 6618. [Google Scholar] [CrossRef]
  228. Mayuranathan, T.; Newby, G.A.; Feng, R.; Yao, Y.; Mayberry, K.D.; Lazzarotto, C.R.; Li, Y.; Levine, R.M.; Nimmagadda, N.; Dempsey, E.; et al. Potent and Uniform Fetal Hemoglobin Induction via Base Editing. Nat. Genet. 2023, 55, 1210–1220. [Google Scholar] [CrossRef]
  229. Butterfield, G.L.; Rohm, D.; Roberts, A.; Nethery, M.A.; Rizzo, A.J.; Morone, D.J.; Garnier, L.; Iglesias, N.; Barrangou, R.; Gersbach, C.A. Characterization of Diverse Cas9 Orthologs for Genome and Epigenome Editing. Proc. Natl. Acad. Sci. USA 2025, 122, e2417674122. [Google Scholar] [CrossRef]
  230. Bou-Fakhredin, R.; Franceschi, L.D.; Motta, I.; Cappellini, M.D.; Taher, A.T. Pharmacological Induction of Fetal Hemoglobin in β-Thalassemia and Sickle Cell Disease: An Updated Perspective. Pharmaceuticals 2022, 15, 753. [Google Scholar] [CrossRef]
  231. Chin, J.; Singh, M.; Banzon, V.; Vaitkus, K.; Ibanez, V.; Kouznetsova, T.; Mahmud, N.; DeSimone, J.; Lavelle, D. Transcriptional Activation of the γ-Globin Gene in Baboons Treated with Decitabine and in Cultured Erythroid Progenitor Cells Involves Different Mechanisms. Exp. Hematol. 2009, 37, 1131–1142. [Google Scholar] [CrossRef] [PubMed]
  232. Ley, T.J.; DeSimone, J.; Noguchi, C.T.; Turner, P.H.; Schechter, A.N.; Heller, P.; Nienhuis, A.W. 5-Azacytidine Increases γ-Globin Synthesis and Reduces the Proportion of Dense Cells in Patients With Sickle Cell Anemia. Blood 1983, 62, 370–380. [Google Scholar] [CrossRef] [PubMed]
  233. Koshy, M.; Dorn, L.; Bressler, L.; Molokie, R.; Lavelle, D.; Talischy, N.; Hoffman, R.; van Overveld, W.; DeSimone, J. 2-Deoxy 5-Azacytidine and Fetal Hemoglobin Induction in Sickle Cell Anemia. Blood 2000, 96, 2379–2384. [Google Scholar] [CrossRef] [PubMed]
  234. Gilmartin, A.G.; Groy, A.; Gore, E.R.; Atkins, C.; Long, E.R.; Montoute, M.N.; Wu, Z.; Halsey, W.; McNulty, D.E.; Ennulat, D.; et al. In Vitro and in Vivo Induction of Fetal Hemoglobin with a Reversible and Selective DNMT1 Inhibitor. Haematologica 2020, 106, 1979–1987. [Google Scholar] [CrossRef]
  235. Shearstone, J.R.; Golonzhka, O.; Chonkar, A.; Tamang, D.; van Duzer, J.H.; Jones, S.S.; Jarpe, M.B. Chemical Inhibition of Histone Deacetylases 1 and 2 Induces Fetal Hemoglobin through Activation of GATA2. PLoS ONE 2016, 11, e0153767. [Google Scholar] [CrossRef]
  236. Nualkaew, T.; Khamphikham, P.; Pongpaksupasin, P.; Kaewsakulthong, W.; Songdej, D.; Paiboonsukwong, K.; Sripichai, O.; Engel, J.D.; Hongeng, S.; Fucharoen, S.; et al. UNC0638 Induces High Levels of Fetal Hemoglobin Expression in β-Thalassemia/HbE Erythroid Progenitor Cells. Ann. Hematol. 2020, 99, 2027–2036. [Google Scholar] [CrossRef]
  237. Chen, X.; Skutt-Kakaria, K.; Davison, J.; Ou, Y.-L.; Choi, E.; Malik, P.; Loeb, K.; Wood, B.; Georges, G.; Torok-Storb, B.; et al. G9a/GLP-Dependent Histone H3K9me2 Patterning during Human Hematopoietic Stem Cell Lineage Commitment. Genes Dev. 2012, 26, 2499–2511. [Google Scholar] [CrossRef]
  238. Renneville, A.; Galen, P.V.; Canver, M.C.; McConkey, M.; Krill-Burger, J.M.; Dorfman, D.M.; Holson, E.B.; Bernstein, B.E.; Orkin, S.H.; Bauer, D.E.; et al. EHMT1 and EHMT2 Inhibition Induces Fetal Hemoglobin Expression. Blood 2015, 126, 1930–1939. [Google Scholar] [CrossRef]
  239. Krivega, I.; Byrnes, C.; de Vasconcellos, J.F.; Lee, Y.T.; Kaushal, M.; Dean, A.; Miller, J.L. Inhibition of G9a Methyltransferase Stimulates Fetal Hemoglobin Production by Facilitating LCR/γ-Globin Looping. Blood 2015, 126, 665–672. [Google Scholar] [CrossRef]
  240. Takase, S.; Hiroyama, T.; Shirai, F.; Maemoto, Y.; Nakata, A.; Arata, M.; Matsuoka, S.; Sonoda, T.; Niwa, H.; Sato, S.; et al. A Specific G9a Inhibitor Unveils BGLT3 LncRNA as a Universal Mediator of Chemically Induced Fetal Globin Gene Expression. Nat. Commun. 2023, 14, 23. [Google Scholar] [CrossRef]
  241. Holshouser, S.; Cafiero, R.; Robinson, M.; Kirkpatrick, J.; Casero, R.A.; Hyacinth, H.I.; Woster, P.M. Epigenetic Reexpression of Hemoglobin F Using Reversible LSD1 Inhibitors: Potential Therapies for Sickle Cell Disease. ACS Omega 2020, 5, 14750–14758. [Google Scholar] [CrossRef] [PubMed]
  242. Ibanez, V.; Vaitkus, K.; Ruiz, M.A.; Lei, Z.; Maienschein-Cline, M.; Arbieva, Z.; Lavelle, D. Effect of the LSD1 Inhibitor RN-1 on γ-Globin and Global Gene Expression during Erythroid Differentiation in Baboons (Papio Anubis). PLoS ONE 2023, 18, e0289860. [Google Scholar] [CrossRef] [PubMed]
  243. Rivers, A.; Vaitkus, K.; Jagadeeswaran, R.; Ruiz, M.A.; Ibanez, V.; Ciceri, F.; Cavalcanti, F.; Molokie, R.E.; Saunthararajah, Y.; Engel, J.D.; et al. Oral Administration of the LSD1 Inhibitor ORY-3001 Increases Fetal Hemoglobin in Sickle Cell Mice and Baboons. Exp. Hematol. 2018, 67, 60–64.e2. [Google Scholar] [CrossRef] [PubMed]
  244. Sun, Y.; Habara, A.; Le, C.Q.; Nguyen, N.; Chen, R.; Murphy, G.J.; Chui, D.H.K.; Steinberg, M.H.; Cui, S. Pharmacologic Induction of PGC-1α Stimulates Fetal Haemoglobin Gene Expression. Br. J. Haematol. 2022, 197, 97–109. [Google Scholar] [CrossRef]
  245. Sun, Y.; Benmhammed, H.; Abdullatif, S.A.; Habara, A.; Fu, E.; Brady, J.; Williams, C.; Ilinski, A.; Sharma, A.; Mahdaviani, K.; et al. PGC-1α Agonism Induces Fetal Hemoglobin and Exerts Antisickling Effects in Sickle Cell Disease. Sci. Adv. 2024, 10, eadn8750. [Google Scholar] [CrossRef]
  246. Fernandes, P.; Waldron, N.; Chatzilygeroudi, T.; Naji, N.S.; Karantanos, T. Acute Erythroid Leukemia: From Molecular Biology to Clinical Outcomes. Int. J. Mol. Sci. 2024, 25, 6256. [Google Scholar] [CrossRef]
  247. Fagnan, A.; Piqué-Borràs, M.; Tauchmann, S.; Mercher, T.; Schwaller, J. Molecular Landscapes and Models of Acute Erythroleukemia. HemaSphere 2021, 5, e558. [Google Scholar] [CrossRef]
Figure 1. The human beta-like globin gene cluster along with the 5 hypersensitive sites of the locus control region (LCR) super-enhancer. The developmental control of the genes is mediated by the LCR, which interacts with different genes during embryonic, fetal, and postnatal/adult life. Finally, a scheme of the transcription factor complex which facilitates the interaction between the LCR and the adult globins is shown. Coordinates shown are in GRCh38 genome assembly.
Figure 1. The human beta-like globin gene cluster along with the 5 hypersensitive sites of the locus control region (LCR) super-enhancer. The developmental control of the genes is mediated by the LCR, which interacts with different genes during embryonic, fetal, and postnatal/adult life. Finally, a scheme of the transcription factor complex which facilitates the interaction between the LCR and the adult globins is shown. Coordinates shown are in GRCh38 genome assembly.
Ijms 26 06342 g001
Figure 2. The human alpha-like globin gene cluster along with the identified hypersensitive sites of the locus control region (LCR) super-enhancer harbored in the neighboring NPRL3 gene. Coordinates shown are in GRCh38 genome assembly.
Figure 2. The human alpha-like globin gene cluster along with the identified hypersensitive sites of the locus control region (LCR) super-enhancer harbored in the neighboring NPRL3 gene. Coordinates shown are in GRCh38 genome assembly.
Ijms 26 06342 g002
Figure 3. Schematic representation of human erythroid differentiation and associated markers. Progression from hematopoietic stem cells (HSCs) through erythroid progenitor stages, including burst-forming unit–erythroid (BFU-E), colony-forming unit–erythroid (CFU-E), proerythroblasts (proEB), basophilic erythroblasts (basoEB), polychromatic erythroblasts (polyEB), and orthochromatic erythroblasts (orthoEB), culminating in reticulocytes (RET), pyrenocytes, and mature red blood cells (RBCs). Key transcription factors (GATA1, GATA2), surface markers (CD34, CD45, CD71, CD235a), and hemoglobin expression are indicated at relevant stages.
Figure 3. Schematic representation of human erythroid differentiation and associated markers. Progression from hematopoietic stem cells (HSCs) through erythroid progenitor stages, including burst-forming unit–erythroid (BFU-E), colony-forming unit–erythroid (CFU-E), proerythroblasts (proEB), basophilic erythroblasts (basoEB), polychromatic erythroblasts (polyEB), and orthochromatic erythroblasts (orthoEB), culminating in reticulocytes (RET), pyrenocytes, and mature red blood cells (RBCs). Key transcription factors (GATA1, GATA2), surface markers (CD34, CD45, CD71, CD235a), and hemoglobin expression are indicated at relevant stages.
Ijms 26 06342 g003
Figure 4. Therapeutic strategies for fetal hemoglobin (HbF) induction in β-hemoglobinopathies. Schematic overview of current approaches to reactivate γ-globin expression as a treatment for β-thalassemia and sickle cell disease: (a) Gene addition therapy using lentiviral vectors expressing γ-globin or short hairpin RNAs (shRNAs) targeting the HbF repressor BCL11A; (b) genome editing strategies that disrupt BCL11A expression by targeting its erythroid-specific enhancer or introduce mutations in the HBG promoters to mimic hereditary persistence of fetal hemoglobin (HPFH); (c) pharmacological induction through the use of epigenetic modulators such as DNA methyltransferase (DNMT) inhibitors, histone deacetylase (HDAC) inhibitors, and histone methyltransferase (HMT) inhibitors, which reactivate γ-globin expression via chromatin remodeling.
Figure 4. Therapeutic strategies for fetal hemoglobin (HbF) induction in β-hemoglobinopathies. Schematic overview of current approaches to reactivate γ-globin expression as a treatment for β-thalassemia and sickle cell disease: (a) Gene addition therapy using lentiviral vectors expressing γ-globin or short hairpin RNAs (shRNAs) targeting the HbF repressor BCL11A; (b) genome editing strategies that disrupt BCL11A expression by targeting its erythroid-specific enhancer or introduce mutations in the HBG promoters to mimic hereditary persistence of fetal hemoglobin (HPFH); (c) pharmacological induction through the use of epigenetic modulators such as DNA methyltransferase (DNMT) inhibitors, histone deacetylase (HDAC) inhibitors, and histone methyltransferase (HMT) inhibitors, which reactivate γ-globin expression via chromatin remodeling.
Ijms 26 06342 g004
Table 1. Summary of epigenetic agents, their mechanisms, and effects in erythroid disorders.
Table 1. Summary of epigenetic agents, their mechanisms, and effects in erythroid disorders.
Epigenetic Agent Class Example CompoundsMechanism or TargetTherapeutic Effect
DNMT InhibitorsAzacytidine, Decitabine, GSK3482364Hypomethylation of γ-globin promoters; DNMT1 selective inhibitionReactivation of γ-globin in β-thalassemia and SCD
HDAC InhibitorsButyrate, Trichostatin A, ACY-957Affect p38 MAPK and STAT5 signaling; histone acetylation; GATA2 activationReactivation of γ-globin
in SCD
HMT InhibitorsUNC0638, RK-701Decrease H3K9me2; promote chromatin looping; increase BGLT3 expressionEnhanced γ-globin reactivation
BET InhibitorsApabetalone (RVX-208), JQ1, I-BET762 (Molibresib)Target bromodomain and extra-terminal domainsIncreased γ-globin expression and HbF levels in SCD and β-thalassemia models
LSD1 Inhibitors TCP, RN-1, ORY-3001Target LSD1Reactivation of γ-globin in SCD, β-YAC mice and baboons
PGC-1a modulatorsSR-18292, ZLN005Increase PGC1a expressionIncreased γ-globin expression in human primary cells, SCD, and β-thalassemia mouse models
Abbreviations: DNMT: DNA methyltransferase, HDAC: Histone deacetylase, HMT: Histone methyltransferase, BET: Bromodomain and extra-terminal domain, LSD1: Lysine-specific demethylase 1, PGC-1α: Peroxisome proliferator-activated receptor gamma coactivator 1-alpha, HbF: Fetal hemoglobin, SCD: Sickle cell disease, β-YAC: Beta yeast artificial chromosome, p38 MAPK: p38 mitogen-activated protein kinase, STAT5: Signal transducer and activator of transcription 5, H3K9me2: Histone H3 lysine 9 dimethylation.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Vasiloudis, N.I.; Paschoudi, K.; Beta, C.; Georgolopoulos, G.; Psatha, N. Epigenetic Regulation of Erythropoiesis: From Developmental Programs to Therapeutic Targets. Int. J. Mol. Sci. 2025, 26, 6342. https://doi.org/10.3390/ijms26136342

AMA Style

Vasiloudis NI, Paschoudi K, Beta C, Georgolopoulos G, Psatha N. Epigenetic Regulation of Erythropoiesis: From Developmental Programs to Therapeutic Targets. International Journal of Molecular Sciences. 2025; 26(13):6342. https://doi.org/10.3390/ijms26136342

Chicago/Turabian Style

Vasiloudis, Ninos Ioannis, Kiriaki Paschoudi, Christina Beta, Grigorios Georgolopoulos, and Nikoletta Psatha. 2025. "Epigenetic Regulation of Erythropoiesis: From Developmental Programs to Therapeutic Targets" International Journal of Molecular Sciences 26, no. 13: 6342. https://doi.org/10.3390/ijms26136342

APA Style

Vasiloudis, N. I., Paschoudi, K., Beta, C., Georgolopoulos, G., & Psatha, N. (2025). Epigenetic Regulation of Erythropoiesis: From Developmental Programs to Therapeutic Targets. International Journal of Molecular Sciences, 26(13), 6342. https://doi.org/10.3390/ijms26136342

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop