Skip to main content
Molecular and Cellular Biology logoLink to Molecular and Cellular Biology
. 2014 Oct;34(19):3560–3569. doi: 10.1128/MCB.00714-14

Fetal Globin Gene Repressors as Drug Targets for Molecular Therapies To Treat the β-Globinopathies

Mikiko Suzuki a, Masayuki Yamamoto b,c, James Douglas Engel d,
PMCID: PMC4187729  PMID: 25022757

Abstract

The human β-globin locus is comprised of embryonic, fetal, and adult globin genes that are expressed in a developmental stage-specific manner. Mutations in the globin locus give rise to the β-globinopathies, β-thalassemia and sickle cell disease, which begin to manifest symptoms around the time of birth. Although the fetal globin genes are autonomously silenced in adult-stage erythroid cells, mutations lying both within and outside the locus lead to natural variations in the level of fetal globin gene expression, and some of these significantly ameliorate the clinical symptoms of the β-globinopathies. Multiple reports have now identified several transcription factors that are involved in fetal globin gene repression in definitive (adult)-stage erythroid cells (the TR2/TR4 heterodimer, MYB, KLFs, BCL11A, and SOX6). To carry out their repression functions, chromatin-modifying enzymes (such as DNA methyltransferase, histone deacetylases, and lysine-specific histone demethylase 1) are additionally involved as a consequence of forming large macromolecular complexes with the DNA-binding subunits of these cellular machines. This review focuses on the molecular mechanisms underlying fetal globin gene silencing and possible near-future molecularly targeted therapies for treating the β-globinopathies.

INTRODUCTION

The human β-globin locus is composed of embryonic (ε), fetal (Gγ and Aγ), and adult (δ and β) globin genes (Fig. 1). The ε-globin gene is expressed in the embryonic yolk sac during the first few months of pregnancy, while sometime during the second month of gestation, after definitive erythroid cells first emerge, the ε-globin gene is silenced and fetal γ-globin expression is activated in the fetal liver. Beginning around the time of birth, fetal γ-globin expression is extinguished and the adult β-globin gene is activated. It is in this way that the β-globin genes are each expressed at distinct stages of development through a process termed “hemoglobin switching.” The abundance of all five β-like globin transcripts is regulated by the locus control region (LCR) (Fig. 1), an aggregate enhancer located far 5′ to the locus. The LCR was originally identified as a group of five dispersed, developmentally stable DNase I-hypersensitive sites localized far 5′ to the embryonic ε-globin gene (1, 2). The LCR harbors both enhancer and insulator activities for the β-locus genes (35). The most widely accepted model for how the LCR transcriptionally activates individual genes is through chromatin looping, creating direct contacts between the proteins bound at the LCR and those bound at the individual globin gene promoters (68).

FIG 1.

FIG 1

A current model for embryonic and fetal globin gene silencing. The human β-globin locus is composed of embryonic (ε), fetal (γ), and adult (β) globin genes and the LCR, itself composed of five DNase I-hypersensitive sites. The embryonic and fetal globin genes are autonomously repressed in adult-stage erythroid cells. TR2/TR4 represses embryonic and fetal globin genes directly by binding to DR elements in their promoters. MYB indirectly represses embryonic and fetal globin expression by activating TR2/TR4 and KLF1. KLF1 activates BCL11A, which represses embryonic and fetal globin genes in collaboration with SOX6 and KLF3/8. GATA1 may directly repress the fetal globin genes and additionally contribute to BCL11A-mediated repression.

Proximity to the LCR significantly affects the developmental expression patterns of β-like globin genes. In the human locus, ε (embryonic), Gγ and Aγ (fetal), and δ and β (adult) globin genes are spatially arranged in a proximal-to-distal direction from the LCR and expressed developmentally in the same order. Transgenic mice harboring transgenes in which the LCR was used to direct transcription of the individual γ- or β-globin genes showed misregulated expression of those genes (911). However, transgenic mice harboring transgenes in which both the γ- and β-globin genes were linked to the LCR in their usual order exhibited behavior that reasonably approximated normal switching (1214). Furthermore, by inverting the gene arrangement relative to the LCR in vivo, the order of expression was altered in a manner that was consistent with then-current models (15). These lines of evidence comprise one of the first “rules” of hemoglobin switching: competition for the LCR (enhancer) between the β-like globin genes is generally advantageous for the gene closest to the LCR. However, in addition to competition among the globin promoters, the second “rule” of switching emerged when it was discovered that the embryonic and fetal globin genes were autonomously silenced (became transcriptionally repressed) by local regulatory sequences that acted at the appropriate developmental stages (16, 17). Concordant with this observation, when transgenic mice harboring transgenes in which the LCR was directly linked to both the γ- and β-globin genes in their natural order were examined, γ-globin expression was repressed in adult-stage erythroid cells, while β-globin expression was induced (13). Thus, the second rule of hemoglobin switching posits that developmental stage-specific repression of the embryonic and fetal genes overrides LCR proximity in the determination of which gene in the locus is to be expressed. In summary, the developmental stage-specific expression of the β-like globin genes appears to be governed by two competing principles: the promoters compete for the LCR enhancer activity (68), while autonomous silencing of the embryonic and fetal globin genes takes precedence. Although there are countless cryptic nuances lying under the umbrella that covers this broad conclusion, in the final analysis these two overriding concepts must be considered in any further elaboration of mechanisms that explain the remaining aspects of hemoglobin switching, some of which are still shrouded in mystery.

β-GLOBINOPATHIES

Inherited mutations in the adult β-globin structural gene cause two major β-globinopathies: β-thalassemia (Cooley's anemia, or CA) and sickle cell disease (SCD). β-Thalassemia results from reduced production of β-globin chains, which leads to an imbalanced accumulation of α- and β-globin proteins inside red blood cells (RBCs). SCD is caused by a mutation that results in a substitution of the sixth amino acid of adult β-globin from glutamic acid to valine (18). Hemoglobin tetramers bearing this mutation (HbS) polymerize inside RBCs and distort them into a characteristic distended appearance instead of their normal oblate ellipsoidal shape, which appears to be a biconcave disc. These rigid sickle RBCs are more prone to lysis as they traverse the circulatory system and have been shown to occlude blood flow in the microvasculature, causing pain and organ damage. While the symptoms are not nearly so serious in individuals who are heterozygous for the mutation (called sickle cell trait), patients bearing two sickle alleles commonly develop severe physiological problems.

Persistent expression of fetal globin gene expression in patients ameliorates both the symptoms and at least a subset of the pathophysiologies that occur as a consequence of these β-globinopathies. Musallam et al. observed a negative correlation between fetal globin levels and morbidity due to β-thalassemia intermedia (19). Induction of fetal hemoglobin (HbF) to above 8.6% was found to significantly relieve SCD mortality (20). Fetal γ-globin incorporation inhibits the extension of the long sickle polymers inside sickle cells (21) and thus ameliorates the symptoms and at least some of the pathologies due to SCD.

The only HbF-inducing agent that is currently approved by the U.S. Food and Drug Administration (FDA) for the treatment of SCD is hydroxyurea (HU). HU is an anticancer drug that induces cell cycle arrest by inhibiting ribonucleotide reductase activity for DNA synthesis. HU treatment was one of the very first compounds shown to increase HbF levels and thus improve some pathological effects of SCD (22). However, only about half of HU-treated patients generate sufficient levels of HbF to relieve the symptoms (23), and therefore novel therapeutic strategies must be developed in order to treat patients who are affected by SCD and β-thalassemia. The most direct strategy to induce HbF in adults would be to interfere with the gene-autonomous mechanisms that regulate fetal globin gene silencing. The rest of the review focuses on recent literature that explores the activities of the currently known transcriptional repressors (site-specific DNA-binding proteins) of the fetal globin genes in the temporal order of their original descriptions as participants in the γ-globin silencing process: TR2/TR4, MYB, BCL11A, SOX6, GATA1, and KLFs (Fig. 1).

THE DRED COMPLEX: A DIRECT γ-GLOBIN REPRESSOR

The embryonic and fetal globin genes have direct repeat (DR) elements in their promoters (24), while neither the mouse nor human adult β-globin genes bear such elements. DRs are consensus binding sites for nonsteroidal nuclear receptors. Significantly, a point mutation in either γ-globin promoter DR element leads to elevated, gene-autonomous γ-globin synthesis in adulthood (called HPFH, or hereditary persistence of fetal hemoglobin), indicating that mutation of this element is the underlying cause of aberrantly elevated fetal globin expression in the affected individuals (25, 26). Importantly, numerous point mutations in both the Gγ- and Aγ-globin promoter DR elements have since been identified that each lead to a fetal gene-specific HPFH phenotype. Phenocopying these natural variants, mutation of the DR elements in the ε- or γ-globin promoters of yeast artificial chromosome (YAC) or bacterial artificial chromosome transgenic mice bearing the entire human β-globin locus leads to persistent expression of the closest cis-linked gene into adulthood (27, 28).

Given that the DR cis element in the γ-globin promoters provided initial evidence for the possibility that the nuclear protein that bound to that site was a repressor, significant effort was expended to identify that presumptive nuclear receptor. The first suggestion for its identity as a transcription factor that was likely to be widely expressed came from gel shift analysis of biochemical extracts derived from erythroid cell lines (29, 30), but the identity of the protein(s) was unresolved. Somewhat later, some intriguing experiments identified this DR element-binding protein as a complex between the erythroid cell-specific transcription factor NF-E3 and a member of the COUP transcription factor family, possibly COUP-TFII (31). However, the bona fide identity of NF-E3 has never been firmly established, and although COUP-TFII is abundantly expressed in a human embryonic erythroid cell line (K562) (27, 31), and despite recent evidence suggesting the contrary (32), COUP-TFII is not expressed in definitive adult human erythroid cells (27; M. Sierant and J. D. Engel, unpublished observations).

We biochemically purified a DR element-binding protein complex that we named “direct repeat erythroid-definitive” (DRED) (Fig. 2), well before we had determined the actual identity of the DRED DNA-binding proteins and cofactors (27, 33); no other nuclear receptors were identified in this initial mass spectrometry-based experiment. The DRED complex appears to be composed of a heterodimer of the orphan nuclear receptors TR2 and TR4 (in current nomenclature NR2C1 and NR2C2, respectively) (33). TR2 and TR4 bind to the two DR elements in the ε-globin promoter with somewhat higher affinity than to the single DR element in each of the γ-globin promoters (33, 34). Mice bearing a targeted deletion in the Tr2 gene reportedly display no overt phenotypes (35), while targeted mutation of Tr4 leads to growth retardation and behavioral defects (36). Compound null mutations in both Tr2 and Tr4 leads to embryonic death prior to implantation, with incomplete penetrance (34, 37), and therefore TR2 and TR4 appear to have redundant but not identical activities. Silencing of the embryonic and fetal globin genes was delayed in definitive erythroid cells of Tr2- and Tr4-null mutant mice (34), consistent with a previous report that concluded that TR2 could repress transcription (38). Thus, the DRED complex directly represses embryonic and fetal globin gene expression in definitive erythroid cells.

FIG 2.

FIG 2

Embryonic and fetal globin gene repression by the DRED complex and molecular targets of inhibitory drugs. The DRED complex directly binds to the promoters of the embryonic and fetal globin genes to repress their expression in definitive adult bone marrow-derived erythroid progenitor cells. The DRED core tetrameric complex is composed of TR2/TR4, DNMT1, and LSD1. The core complex also associates with additional corepressors (NuRD and CoREST complexes that contain HDACs 1 and 2 and independently with HDAC3). Inhibitors of DNMT, LSD1, and HDACs, namely, 5-azacytidine/decitabine, tranylcypromine, and butyrate, respectively, induce fetal globin expression, which is at least partially dependent on DRED activity.

In addition to TR2 and TR4, DNA methyltransferase 1 (DNMT1) and lysine-specific histone demethylase 1 (LSD1) participate as cofactors to form a tetrameric core DRED complex (39). DNMT1, previously characterized as the maintenance DNA methylase, is able to recognize and methylate CpG dinucleotides when they are found on the opposite strand of MeCpG, thus maintaining stable DNA methylation patterns in the genome after DNA replication (40). LSD1 removes the methyl groups from mono- and dimethyl-histone H3 lysine 4 (MeH3K4 and Me2H3K4, respectively), which are widely regarded to be epigenetic signatures that mark transcriptionally active genes (41). This tetrameric core repression complex (TR2-TR4-DNMT1-LSD1) can additionally interact with histone deacetylase 1/2 (HDAC1/2), NuRD, CoREST, TRIM28 (also known as TIF1β), and HDAC3 to form an array of even larger multifunctional complexes (39) (Fig. 2).

Curiously (and incongruously), forced overexpression of TR2 and TR4 in erythroid cells leads to the induction of fetal globin expression, thereby reversing SCD phenotypes in mouse models (34, 42). Although the mechanism underlying this seemingly paradoxical phenomenon is unknown, one must consider two possibilities. One possibility is that overly abundant expression of TR2 and TR4 inhibits effective formation of the very large DRED complex because one or more of the enzyme components identified as DRED cofactors may be in limiting abundance, thereby “squelching” inactivation by the repressor (43). The other possibility is that TR2 and TR4, in addition to their repressive role, also function as activators of the fetal globin genes by association with cellular coactivators, such as PGC-1 (44). Several previous reports indeed contended that TR4 can function as a robust transcriptional activator (4551). TR2 and TR4 are very broadly expressed, including in erythroid cells (5254). In nonerythroid cells, TR4 has been shown to transcriptionally activate apolipoprotein E/C-I/C-II (46) and phosphoenolpyruvate carboxykinase in the liver (55), luteinizing hormone receptor in the ovary (56), and CD36 in macrophages (57). In erythroid cells, hundreds of genes are repressed (either directly or indirectly) by TR4, while even more genes are activated (58). Analysis of erythroid differentiation in comparative data sets from chromatin immunoprecipitation-sequencing (ChIP-seq) and RNA-seq experiments indicated that the DR motif is more prevalent in the proximal promoters of TR4 direct target genes that are involved in basic biological functions (e.g., mRNA processing, ribosomal assembly, RNA splicing, and primary metabolic processes [58, 59]). We also previously identified TRIM28 as an interacting protein of TR2/TR4 (39). TRIM28 is dispensable for fetal globin silencing but is required for erythroid differentiation, implying that TRIM28 may function as a coactivator of TR2/TR4 (60). However, it is equally plausible that TRIM28 functions as a corepressor, as it does in other contexts (61), to inhibit the transcription of a currently unidentified activator that is required for erythroid differentiation.

TR2 and TR4 are categorized as orphan nuclear receptors, since no ligand for either of them has been identified. However, structural analyses and biochemical investigations of the TR4 ligand-binding domain by Zhou et al. revealed that retinol and all-trans retinoic acid can act as very weak ligands for TR4 (62), although the concentrations at which they stimulated TR4 activity in transfected cells suggested that they may not be the genuine ligands in vivo. In addition, unsaturated fatty acids and their metabolites have also been reported to promote the transcriptional activity of TR4 (51, 57). In contrast, the amino acid sequence of the closely related TR2 protein suggests that it does not naturally associate with ligands, and thus may exert its activity either as an obligate TR4 partner or by homodimerization, thereby leading to competition for TR4 binding sites or for TR2/TR4 heterodimer binding.

In summary, even though TR2/TR4 was the first γ-globin repressor to be identified, the complexity of its interactions with multiple cofactors, the open question of whether or not either of this pair of orphan nuclear receptors is regulated by small-molecule ligands, and the fact that the TR2 and TR4 subunits are genetically compensatory all complicate potential strategies that can be envisioned by which one might inactivate the repressor and achieve the ultimate goal of increasing γ-globin synthesis. The hypothesis that TR4 could be a typical ligand-regulated nuclear receptor (51, 57, 62) seems promising, since identification of the genuine ligand could be used to subsequently identify antagonists that could prove to be clinically beneficial.

MYB: A CRITICAL HEMATOPOIETIC REGULATOR INVOLVED IN FETAL GLOBIN SILENCING

Thein et al. originally identified an intergenic region lying between the HBS1L and MYB genes on 6q23 as a region controlling fetal globin gene expression in adults based on quantitative trait locus (QTL) analyses (63, 64). Since the abundance of both the HBS1L and MYB genes diminished in the patients with the HBS1L-MYB genotype, the gene responsible for that specific HPFH phenotype was unclear. The answer to this question was resolved via a serendipitous coincidence of similar conclusions from three independent studies.

The first insights regarding the nature of the effects of the HBS1L-MYB intergenic region emerged from a study of newborns bearing trisomy 13 who exhibited elevated HbF levels. Sankaran et al. identified two microRNAs, miR-15a and -16-1, on 13q14 as responsible for the elevated HbF levels in human trisomy 13 patients (65). At least one target of these two microRNAs was the MYB gene, and so inhibition of MYB mRNA function in human adult CD34+ cell-derived erythroid cells resulted in induction of fetal hemoglobin. Therefore, it was concluded that MYB activity was responsible for this particular variant HPFH. In further support of this conclusion, Sankaran et al. recently reported analysis of a patient with a complete loss of function of the HBS1L gene (66). This patient exhibited a normal distribution of hemoglobin subtypes and no hematological abnormalities, but did bear other phenotypic deficiencies. Finally, it was demonstrated that disruption of the Hbs1l-Myb intergenic region by a random transgene insertion in the mouse led to greatly diminished Myb accumulation (Myb expression was significantly more affected than was Hbs1l) and was accompanied by a marked increase in the murine homologues of the human ε- and γ-globin genes (67). These three observations in concert validated the contention that MYB, rather than HBS1L, is responsible for the mutant intergenic region HPFH phenotype.

The Myb gene encodes a transcription factor that is critical for hematopoietic stem cells and progenitors. Myb-null mutant mice die at embryonic day 15 due to a failure in definitive erythropoiesis (68). Analyses of constitutive and conditional Myb mutant mice have revealed the contributions of this vital transcription factor to erythroid cell, T cell, and B cell development (6971). While the Myb gene is highly expressed in hematopoietic stem cells and progenitors, its expression progressively diminishes as erythroid cells mature (72). Since the developmental stages at which the MYB and β-type globins are expressed differ markedly (67), the data can be interpreted to suggest that MYB probably induces fetal globin transcription through an indirect mechanism. In support of this contention, Bianchi et al. and we have found that MYB activates KLF1 and TR2/TR4 in human erythroid cells (67, 73). Since one of the several major functions of KLF1 is to activate the BCL11A gene, itself a robust fetal globin gene repressor (see below), current data support the hypothesis that MYB indirectly regulates γ-globin transcription because of diminished production of either or both of the repressors, BCL11A and TR2/TR4 (Fig. 1).

Importantly, the region on 6q23 that is responsible for elevated HbF levels is located not in the MYB structural gene, but in a putative gene regulatory region. Farrell et al. identified a 3-bp deletion within the HBS1L-MYB intergenic region that was first associated with high HbF levels (74). This tiny indel was located between a TAL1-binding site and a GATA-binding site, suggesting that the deletion may have affected the formation of a precisely structured higher-order TAL1/GATA factor complex, as first described almost 2 decades ago (75). More recently, Stadhouders et al. identified binding sites for the TAL1/GATA1 complexes in the mouse Hbs1l-Myb intergenic region and showed that the binding sites were in close three-dimensional proximity to the Myb gene promoter (76). Indeed, single nucleotide polymorphisms (SNPs) located at these binding sites were more recently shown to influence GATA1 and KLF1 binding and the interaction between those sites and the MYB gene promoter that led to a reduction in MYB gene expression (77). These reports suggested that the cell-type-specific loss of MYB function was achieved by inhibiting the ability of a TAL1/GATA1 complex to form an optimal association with a DNA site that regulates MYB transcription, thereby indirectly leading to an HPFH phenotype.

As described above, we originally identified an anemic mouse that had been created by random insertion of a transgene into the Hbs1l-Myb intergenic region (78). Such mice exhibited elevated levels of the mouse embryonic globins as well as anemic and thrombocytic phenotypes. In these mice, Myb gene expression was muted in megakaryocyte/erythrocyte progenitors (MEPs) but not in T cells, suggesting that the intergenic region contained a regulatory element, possibly an enhancer, that controlled Myb transcription in MEPs. Indeed, human HPFH individuals bearing mutations within the HBS1L-MYB region also exhibited modest (but quantitatively altered) hematopoietic deficiencies, including anemia and thrombocytosis (79). Evolution suggests that mutations or polymorphisms in gene regulatory regions can result in spatiotemporal defects of gene expression, which can in turn give rise to limited (but only rarely beneficial) phenotypic effects; such appears to be the case for the HBS1L-MYB intergenic mutation encountered in humans who additionally bear deleterious mutations in β-globin synthesis (8083).

BCL11A: THE POSTER CHILD FOR GWAS

Genome-wide association studies (GWAS) represent a natural variation-based genetic strategy for interrogating the entire human genome for SNPs that are associated with specific phenotypes (84). GWAS performed on the genomes of sickle cell disease cohorts originally identified three chromosomal positions that influence HbF levels. While two of three represented loci that were already known to differentially modulate fetal and adult β-type globin levels (one within the β-globin locus as well as one in the HBS1L-MYB intergenic region), a third completely novel locus mapped at the position of the BCL11A gene, previously identified as a vital transcription factor for B lymphocyte development (63, 80, 85). In humans, a high level of HbF is associated with low BCL11A expression (86). Bcl11a-null mutant mice and erythroid cell-specific Bcl11a knockout mice both exhibited an HPFH phenotype (87, 88). These reports confirmed that the transcription factor BCL11A is a repressor of the fetal globin genes.

BCL11A forms molecular complexes with GATA1 and FOG1 and additionally with the NuRD chromatin remodeling complex (86). Furthermore, BCL11A physically interacts with transcription factor SOX6 (89). In a previous report, SOX6 was shown to bind to the promoter of the mouse embryonic εY-globin gene and silence its expression in definitive erythroid cells (90). Thus, BCL11A and SOX6 may collaboratively bind to a number of sites within the β-globin locus to repress fetal globin gene expression (89, 91).

Since Bcl11a-null mutant mice die in the perinatal period (85), as a therapeutic strategy a global reduction in BCL11A abundance could prove to be problematic in humans, especially since such a strategy could simultaneously affect unintended functions in other tissues. For example, Bcl11a germ line-null mutant and adult-stage-specific null mutant mice exhibit defects in lymphoid development (85, 92). Additionally, BCL11A plays critical roles in neuronal morphogenesis and sensory circuit formation (93). However, Bauer et al. recently reported the identification of an erythroid cell-specific enhancer of the BCL11A gene, providing a possibly superior, and tissue-specific, alternative target for therapeutic development. As is also the case for the MYB intergenic enhancer, a TAL1/GATA1 complex activates the BCL11A erythroid enhancer (94).

GATA1 AND KLFs: ERYTHROID REGULATORY PROTEINS THAT INFLUENCE FETAL GLOBIN REPRESSION

KLF1, originally designated EKLF (95), has been known to be required for activation of the adult β-globin gene since shortly after its original discovery (96, 97). However, Borg et al. recently revealed a new function of KLF1 as a repressor of the fetal globin gene (98): from the analysis of a Maltese family with prevalent HPFH, the investigators identified a nonsense mutation in the KLF1 gene. Further studies demonstrated that KLF1 represses fetal globin expression indirectly by activation of BCL11A (98, 99), and thus in haploinsufficient individuals a sensitivity to reduced KLF1 abundance in erythroid cells results in diminished production of the BCL11A repressor. In addition to KLF1, other KLF family factors have been shown to be involved in fetal globin silencing. KLF3-deficient (originally characterized as BKLF-deficient) mice exhibited derepression of the embryonic globin genes, and the effect was enhanced in Klf3:Klf8 compound null mutant mice (100). Since the KLF3, KLF8, and BCL11A genes are all activated by KLF1, these factors all participate in fetal globin gene silencing, which is possibly mediated through KLF1.

SNP studies originally suggested that GATA transcription factors may also be directly involved in fetal globin gene silencing. Chen et al. identified an SNP located 567 bp 5′ to the fetal Gγ globin gene (HBG2) in an Iranian-American family that exhibited elevated fetal globin levels (101). This polymorphism disrupted a GATA-binding motif in the Gγ globin far upstream promoter that was also duplicated in the Aγ globin (HBG1) gene (located 566 bp 5′ to the Aγ start site) (102). Disruption of the −566 Aγ GATA site in β-globin YAC mutant mice resulted in persistent expression of the Aγ gene in adult mice (103). These studies suggest a contribution of a GATA transcription factor(s) to fetal globin silencing. As GATA1 also appears to play a significant role in BCL11A transcriptional activity (87), the mechanism of GATA1-related γ-globin repression could be functionally related to the BCL11A pathway.

CHROMATIN MARKS THAT DEFINE FETAL GLOBIN GENE EXPRESSION

As alluded to above, transcription factors that play roles as direct or indirect fetal globin gene repressors have been shown to interact with several enzymes that modify chromatin. While these enzymes are usually referred to as “epigenetic” modifying proteins, most have not been shown to heritably affect gene expression through a causal relationship. However, the marks that these enzymes impart to DNA and histones, and even to transcription factors themselves, are unquestionably affiliated with either activated or repressed transcriptional states, be they causal or simply a reflection of those states. Perhaps, intriguingly, DNMTs, HDACs, and the LSD1 enzymes are all implicated in fetal globin gene silencing by both TR2/TR4 and BCL11A (39, 86). DNA methylation within the locus has for many years been one of the hallmarks of globin gene inactivity through the strong correlation of methylation with repression (104108). Histone acetylation of globin gene promoters also correlates with the expression levels of these genes (109), and HDAC inhibitors such as butyrate have been shown to induce HbF (110).

As mentioned above, methylation of cytosine residues in CpG dinucleotides by DNMTs correlates strongly with transcriptional repression (111). DNA methylation in the embryonic and fetal globin promoters inversely correlates with their expression. Small interfering RNA-mediated reduction of DNMT1 levels in baboon CD34+ bone marrow-derived erythroid progenitors exhibited diminished methylation of the embryonic and fetal globin promoters and elevated gene transcription (112). Furthermore, mice that are genetically deficient for the methyl-CpG-binding protein gene Mbd2 exhibit delayed (human YAC transgene-delivered) fetal globin gene silencing (113). The DNMT inhibitors 5-azacytidine and 5-aza-2′-deoxycytidine (decitabine) have been shown to lead to demethylation of the fetal globin gene promoters and to induce fetal globin gene expression (114117); decitabine, in particular, is currently in clinical trials and is being tested as a possible therapeutic human γ-globin inducer.

In addition to DNMT and HDAC, LSD1 also binds (directly or indirectly) to both TR2/TR4 and BCL11A (39). LSD1 demethylates mono- and dimethyl-histone H3 lysine 4 (H3K4), an activating histone mark (41). We recently showed that inhibition of LSD1 via use of either RNA interference or a small-molecule LSD1 inhibitor (the FDA-approved antidepressant tranylcypromine [TC], trade name Parnate, is a monoamine oxidase inhibitor [MAOI]) robustly induces fetal globin expression in human adult erythroid cells that can be derived in culture from mobilized circulating CD34+ progenitors (118, 119). Furthermore, the combinatorial application of TC and decitabine had synergistic effects on HbF induction in this cell culture model, achieving >50% HbF production in differentiated CD34+ cells ex vivo, while the combination of TC plus HU was much less effective (118). Since it was later shown that higher concentrations of TC inhibited erythroid differentiation (120) and as there exist a number of potentially dangerous drug-metabolite interactions with this compound (121, 122), TC is unlikely to represent a popular medication for treating the β-globinopathies. Perhaps more-specific compounds that target LSD1 activity and that act at lower concentrations for longer periods of time in the bloodstream and have fewer side effects can be identified in the not-too-distant future.

CONCLUDING REMARKS

Development of novel therapies for treating the β-globinopathies by fetal globin gene induction has been contemplated and pursued for decades. Recent studies have identified a number of transcription factor molecules that are directly or indirectly involved in fetal globin gene silencing, and their associations with additional enzyme cofactors that impart distinguishing activating or repressing marks on DNA and histones near the transcription factor binding sites have provided many new potential interaction surfaces and enzymatic mechanisms for interruption of the activity of these proteins and for possible therapeutic development. Due to these achievements, we should be able to find new molecularly targeted drugs or new uses for existent medications to treat these disorders. In this regard, the greatest challenge may be to selectively impair the activity of broadly expressed factors that we might envision as potential drug targets (e.g., TR2, TR4, LSD1, and DNMT1) specifically in erythroid cells or to inactivate molecules that are more highly tissue restricted (e.g., MYB, KLFs, GATA1, and BCL11A) without affecting the essentially universal requirement for MYB, GATA1, and KLF function in hematopoietic cells or the physiological requirement for BCL11A in other tissues. In this regard, the identification of enhancers that control the erythroid cell- or hematopoietic cell-specific transcription of MYB (67, 76) and BCL11A (94) could represent a significant conceptual advance in providing additional, more highly refined, interaction targets that could lead to disruption of regulatory pathways in a tissue-specific manner.

The progression of regenerative medicine utilizing human embryonic stem cells (ESC) and induced pluripotent stem cells (iPSC) in combination with genome editing portends cures for the β-globinopathies after correcting these lesions by essentially autologous transplantation of modified patient-specific iPSC or hematopoietic stem cells. Hanna et al. reported correction of sickle cell disease by transplantation with hematopoietic progenitors derived from genome-targeted iPSC generated from sickle cell model mice (123). In addition, recently developed techniques for genome editing (e.g., TALENs and CRISPR/Cas9) enable rapid and increasingly accurate manipulation of the genome (124, 125). Nevertheless, since those developments as molecular cures likely lie some time away, and since they will have the same restrictions for advanced medical care as any other strategy requiring bone marrow transplantation (and thus will be available to only a small fraction of the affected patient population), the development of safer and more effective drug treatments for the hemoglobinopathies is required. The goals in this field are and should be immediately focused on the development of safe, effective drug therapies that can be accomplished through fetal globin gene induction and at the same time on the prospect of cures through bone marrow transplantation using the promise of genome-editing strategies that will bring a vastly improved quality and quantity of life to patients that suffer from these devastating disorders.

ACKNOWLEDGMENTS

We are grateful to our colleagues Shuaiying Cui and Mary Lee, and to an unknown referee, for careful editing of the manuscript and for providing significant insight and perspective.

This work was supported by the NIH (R21HL114368 and U01HL117658 to J.D.E.) and the JSPS (KAKENHI 22118001 and 24249015 to M.Y. and 26461395 to M.S.).

Footnotes

Published ahead of print 14 July 2014

REFERENCES

  • 1.Tuan D, London IM. 1984. Mapping of DNase I-hypersensitive sites in the upstream DNA of human embryonic e-globin gene in K562 leukemia cells. Proc. Natl. Acad. Sci. U. S. A. 81:2718–2722. 10.1073/pnas.81.9.2718 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Forrester WC, Thompson C, Elder JT, Groudine M. 1986. A developmentally stable chromatin structure in the human beta-globin gene cluster. Proc. Natl. Acad. Sci. U. S. A. 83:1359–1363. 10.1073/pnas.83.5.1359 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Forrester WC, Takegawa S, Papayannopoulou T, Stamatoyannopoulos G, Groudine M. 1987. Evidence for a locus activation region: the formation of developmentally stable hypersensitive sites in globin-expressing hybrids. Nucleic Acids Res. 15:10159–10177. 10.1093/nar/15.24.10159 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Grosveld F, van Assendelft GB, Greaves DR, Kollias G. 1987. Position-independent, high-level expression of the human beta-globin gene in transgenic mice. Cell 51:975–985. 10.1016/0092-8674(87)90584-8 [DOI] [PubMed] [Google Scholar]
  • 5.Li Q, Stamatoyannopoulos G. 1994. Hypersensitive site 5 of the human beta locus control region functions as a chromatin insulator. Blood 84:1399–1401 [PubMed] [Google Scholar]
  • 6.Carter D, Chakalova L, Osborne CS, Dai YF, Fraser P. 2002. Long-range chromatin regulatory interactions in vivo. Nat. Genet. 32:623–626. 10.1038/ng1051 [DOI] [PubMed] [Google Scholar]
  • 7.Choi OR, Engel JD. 1988. Developmental regulation of beta-globin gene switching. Cell 55:17–26. 10.1016/0092-8674(88)90005-0 [DOI] [PubMed] [Google Scholar]
  • 8.Tolhuis B, Palstra RJ, Splinter E, Grosveld F, de Laat W. 2002. Looping and interaction between hypersensitive sites in the active beta-globin locus. Mol. Cell 10:1453–1465. 10.1016/S1097-2765(02)00781-5 [DOI] [PubMed] [Google Scholar]
  • 9.Enver T, Ebens AJ, Forrester WC, Stamatoyannopoulos G. 1989. The human β-globin locus activation region alters the developmental fate of a human fetal globin gene in transgenic mice. Proc. Natl. Acad. Sci. U. S. A. 86:7033–7037. 10.1073/pnas.86.18.7033 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Townes TM, Lingrel JB, Chen HY, Brinster RL, Palmiter RD. 1985. Erythroid-specific expression of human beta-globin genes in transgenic mice. EMBO J. 4:1715–1723 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Kollias G, Wrighton N, Hurst J, Grosveld F. 1986. Regulated expression of human Aγ-, β-, and hybrid γβ-globin genes in transgenic mice: manipulation of the developmental expression patterns. Cell 46:89–94. 10.1016/0092-8674(86)90862-7 [DOI] [PubMed] [Google Scholar]
  • 12.Enver T, Raich N, Ebens AJ, Papayannopoulou T, Costantini F, Stamatoyannopoulos G. 1990. Developmental regulation of human fetal-to-adult globin gene switching in transgenic mice. Nature 344:309–313. 10.1038/344309a0 [DOI] [PubMed] [Google Scholar]
  • 13.Behringer RR, Ryan TM, Palmiter RD, Brinster RL, Townes TM. 1990. Human γ- to β-globin gene switching in transgenic mice. Genes Dev. 4:380–389. 10.1101/gad.4.3.380 [DOI] [PubMed] [Google Scholar]
  • 14.Hanscombe O, Whyatt D, Fraser P, Yannoutsos N, Greaves D, Dillon N, Grosveld F. 1991. Importance of globin gene order for correct developmental expression. Genes Dev. 5:1387–1394. 10.1101/gad.5.8.1387 [DOI] [PubMed] [Google Scholar]
  • 15.Tanimoto K, Liu Q, Bungert J, Engel JD. 1999. Effects of altered gene order or orientation of the locus control region on human beta-globin gene expression in mice. Nature 398:344–348. 10.1038/18698 [DOI] [PubMed] [Google Scholar]
  • 16.Raich N, Enver T, Nakamoto B, Josephson B, Papayannopoulou T, Stamatoyannopoulos G. 1990. Autonomous developmental control of human embryonic globin gene switching in transgenic mice. Science 250:1147–1149. 10.1126/science.2251502 [DOI] [PubMed] [Google Scholar]
  • 17.Dillon N, Grosveld F. 1991. Human gamma-globin genes silenced independently of other genes in the beta-globin locus. Nature 350:252–254. 10.1038/350252a0 [DOI] [PubMed] [Google Scholar]
  • 18.Ingram VM. 1957. Gene mutations in human haemoglobin: the chemical difference between normal and sickle cell haemoglobin. Nature 180:326–328. 10.1038/180326a0 [DOI] [PubMed] [Google Scholar]
  • 19.Musallam KM, Sankaran VG, Cappellini MD, Duca L, Nathan DG, Taher AT. 2012. Fetal hemoglobin levels and morbidity in untransfused patients with beta-thalassemia intermedia. Blood 119:364–367. 10.1182/blood-2011-09-382408 [DOI] [PubMed] [Google Scholar]
  • 20.Platt OS, Brambilla DJ, Rosse WF, Milner PF, Castro O, Steinberg MH, Klug PP. 1994. Mortality in sickle cell disease. Life expectancy and risk factors for early death. N. Engl. J. Med. 330:1639–1644 [DOI] [PubMed] [Google Scholar]
  • 21.Sunshine HR, Hofrichter J, Eaton WA. 1978. Requirement for therapeutic inhibition of sickle haemoglobin gelation. Nature 275:238–240. 10.1038/275238a0 [DOI] [PubMed] [Google Scholar]
  • 22.Veith R, Galanello R, Papayannopoulou T, Stamatoyannopoulos G. 1985. Stimulation of F-cell production in patients with sickle-cell anemia treated with cytarabine or hydroxyurea. N. Engl. J. Med. 313:1571–1575. 10.1056/NEJM198512193132503 [DOI] [PubMed] [Google Scholar]
  • 23.Steinberg MH, Lu ZH, Barton FB, Terrin ML, Charache S, Dover GJ. 1997. Fetal hemoglobin in sickle cell anemia: determinants of response to hydroxyurea. Blood 89:1078–1088 [PubMed] [Google Scholar]
  • 24.Myers RM, Tilly K, Maniatis T. 1986. Fine structure genetic analysis of a beta-globin promoter. Science 232:613–618. 10.1126/science.3457470 [DOI] [PubMed] [Google Scholar]
  • 25.Gelinas R, Endlich B, Pfeiffer C, Yagi M, Stamatoyannopoulos G. 1985. G to A substitution in the distal CCAAT box of the Aγ-globin gene in Greek hereditary persistence of fetal haemoglobin. Nature 313:323–325. 10.1038/313323a0 [DOI] [PubMed] [Google Scholar]
  • 26.Fucharoen S, Shimizu K, Fukumaki Y. 1990. A novel C-T transition within the distal CCAAT motif of the Gγ-globin gene in the Japanese HPFH: implication of factor binding in elevated fetal globin expression. Nucleic Acids Res. 18:5245–5253. 10.1093/nar/18.17.5245 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Tanimoto K, Liu Q, Grosveld F, Bungert J, Engel JD. 2000. Context-dependent EKLF responsiveness defines the developmental specificity of the human epsilon-globin gene in erythroid cells of YAC transgenic mice. Genes Dev. 14:2778–2794. 10.1101/gad.822500 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Omori A, Tanabe O, Engel JD, Fukamizu A, Tanimoto K. 2005. Adult stage gamma-globin silencing is mediated by a promoter direct repeat element. Mol. Cell. Biol. 25:3443–3451. 10.1128/MCB.25.9.3443-3451.2005 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Mantovani R, Malgaretti N, Nicolis S, Giglioni B, Comi P, Cappellini N, Bertero MT, Caligaris-Cappio F, Ottolenghi S. 1988. An erythroid specific nuclear factor binding to the proximal CACCC box of the beta-globin gene promoter. Nucleic Acids Res. 16:4299–4313. 10.1093/nar/16.10.4299 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Mantovani R, Malgaretti N, Nicolis S, Ronchi A, Giglioni B, Ottolenghi S. 1988. The effects of HPFH mutations in the human gamma-globin promoter on binding of ubiquitous and erythroid specific nuclear factors. Nucleic Acids Res. 16:7783–7797. 10.1093/nar/16.16.7783 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Filipe A, Li QL, Deveaux S, Godin I, Romeo PH, Stamatoyannopoulos G, Mignotte V. 1999. Regulation of embryonic/fetal globin genes by nuclear hormone receptors: a novel perspective on hemoglobin switching. EMBO J. 18:687–697. 10.1093/emboj/18.3.687 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Aerbajinai W, Zhu J, Kumkhaek C, Chin K, Rodgers GP. 2009. SCF induces gamma-globin gene expression by regulating downstream transcription factor COUP-TFII. Blood 114:187–194. 10.1182/blood-2008-07-170712 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Tanabe O, Katsuoka F, Campbell AD, Song W, Yamamoto M, Tanimoto K, Engel JD. 2002. An embryonic/fetal beta-type globin gene repressor contains a nuclear receptor TR2/TR4 heterodimer. EMBO J. 21:3434–3442. 10.1093/emboj/cdf340 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Tanabe O, McPhee D, Kobayashi S, Shen Y, Brandt W, Jiang X, Campbell AD, Chen YT, Chang C, Yamamoto M, Tanimoto K, Engel JD. 2007. Embryonic and fetal beta-globin gene repression by the orphan nuclear receptors, TR2 and TR4. EMBO J. 26:2295–2306. 10.1038/sj.emboj.7601676 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Shyr CR, Collins LL, Mu XM, Platt KA, Chang C. 2002. Spermatogenesis and testis development are normal in mice lacking testicular orphan nuclear receptor 2. Mol. Cell. Biol. 22:4661–4666. 10.1128/MCB.22.13.4661-4666.2002 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Collins LL, Lee YF, Heinlein CA, Liu NC, Chen YT, Shyr CR, Meshul CK, Uno H, Platt KA, Chang C. 2004. Growth retardation and abnormal maternal behavior in mice lacking testicular orphan nuclear receptor 4. Proc. Natl. Acad. Sci. U. S. A. 101:15058–15063. 10.1073/pnas.0405700101 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Tanabe O, Shen Y, Liu Q, Campbell AD, Kuroha T, Yamamoto M, Engel JD. 2007. The TR2 and TR4 orphan nuclear receptors repress Gata1 transcription. Genes Dev. 21:2832–2844. 10.1101/gad.1593307 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Lee HJ, Young WJ, Shih CY, Chang C. 1996. Suppression of the human erythropoietin gene expression by the TR2 orphan receptor, a member of the steroid receptor superfamily. J. Biol. Chem. 271:10405–10412. 10.1074/jbc.271.17.10405 [DOI] [PubMed] [Google Scholar]
  • 39.Cui S, Kolodziej KE, Obara N, Amaral-Psarris A, Demmers J, Shi L, Engel JD, Grosveld F, Strouboulis J, Tanabe O. 2011. Nuclear receptors TR2 and TR4 recruit multiple epigenetic transcriptional corepressors that associate specifically with the embryonic beta-type globin promoters in differentiated adult erythroid cells. Mol. Cell. Biol. 31:3298–3311. 10.1128/MCB.05310-11 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Lei H, Oh SP, Okano M, Juttermann R, Goss KA, Jaenisch R, Li E. 1996. De novo DNA cytosine methyltransferase activities in mouse embryonic stem cells. Development 122:3195–3205 [DOI] [PubMed] [Google Scholar]
  • 41.Shi Y, Lan F, Matson C, Mulligan P, Whetstine JR, Cole PA, Casero RA, Shi Y. 2004. Histone demethylation mediated by the nuclear amine oxidase homolog LSD1. Cell 119:941–953. 10.1016/j.cell.2004.12.012 [DOI] [PubMed] [Google Scholar]
  • 42.Campbell AD, Cui S, Shi L, Urbonya R, Mathias A, Bradley K, Bonsu KO, Douglas RR, Halford B, Schmidt L, Harro D, Giacherio D, Tanimoto K, Tanabe O, Engel JD. 2011. Forced TR2/TR4 expression in sickle cell disease mice confers enhanced fetal hemoglobin synthesis and alleviated disease phenotypes. Proc. Natl. Acad. Sci. U. S. A. 108:18808–18813. 10.1073/pnas.1104964108 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Gill G, Ptashne M. 1988. Negative effect of the transcriptional activator GAL4. Nature 334:721–724. 10.1038/334721a0 [DOI] [PubMed] [Google Scholar]
  • 44.Cui S, Tanabe O, Lim KC, Xu HE, Zhou XE, Lin JD, Shi L, Schmidt L, Campbell A, Shimizu R, Yamamoto M, Engel JD. 2014. PGC-1 coactivator activity is required for murine erythropoiesis. Mol. Cell. Biol. 34:1956–1965. 10.1128/MCB.00247-14 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Lee YF, Pan HJ, Burbach JP, Morkin E, Chang C. 1997. Identification of direct repeat 4 as a positive regulatory element for the human TR4 orphan receptor. A modulator for the thyroid hormone target genes. J. Biol. Chem. 272:12215–12220 [DOI] [PubMed] [Google Scholar]
  • 46.Kim E, Xie S, Yeh SD, Lee YF, Collins LL, Hu YC, Shyr CR, Mu XM, Liu NC, Chen YT, Wang PH, Chang C. 2003. Disruption of TR4 orphan nuclear receptor reduces the expression of liver apolipoprotein E/C-I/C-II gene cluster. J. Biol. Chem. 278:46919–46926. 10.1074/jbc.M304088200 [DOI] [PubMed] [Google Scholar]
  • 47.Young WJ, Lee YF, Smith SM, Chang C. 1998. A bidirectional regulation between the TR2/TR4 orphan receptors (TR2/TR4) and the ciliary neurotrophic factor (CNTF) signaling pathway. J. Biol. Chem. 273:20877–20885. 10.1074/jbc.273.33.20877 [DOI] [PubMed] [Google Scholar]
  • 48.Young WJ, Smith SM, Chang C. 1997. Induction of the intronic enhancer of the human ciliary neurotrophic factor receptor (CNTFRα) gene by the TR4 orphan receptor. A member of steroid receptor superfamily. J. Biol. Chem. 272:3109–3116 [DOI] [PubMed] [Google Scholar]
  • 49.Lee CH, Wei LN. 1999. Characterization of an inverted repeat with a zero spacer (IR0)-type retinoic acid response element from the mouse nuclear orphan receptor TR2-11 gene. Biochemistry 38:8820–8825. 10.1021/bi9903547 [DOI] [PubMed] [Google Scholar]
  • 50.Lee CH, Wei LN. 1999. Characterization of receptor-interacting protein 140 in retinoid receptor activities. J. Biol. Chem. 274:31320–31326. 10.1074/jbc.274.44.31320 [DOI] [PubMed] [Google Scholar]
  • 51.Tsai NP, Huq M, Gupta P, Yamamoto K, Kagechika H, Wei LN. 2009. Activation of testicular orphan receptor 4 by fatty acids. Biochim. Biophys. Acta 1789:734–740. 10.1016/j.bbagrm.2009.09.010 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Chang C, Da Silva SL, Ideta R, Lee Y, Yeh S, Burbach JP. 1994. Human and rat TR4 orphan receptors specify a subclass of the steroid receptor superfamily. Proc. Natl. Acad. Sci. U. S. A. 91:6040–6044. 10.1073/pnas.91.13.6040 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Chang C, Kokontis J, Acakpo-Satchivi L, Liao S, Takeda H, Chang Y. 1989. Molecular cloning of new human TR2 receptors: a class of steroid receptor with multiple ligand-binding domains. Biochem. Biophys. Res. Commun. 165:735–741. 10.1016/S0006-291X(89)80028-2 [DOI] [PubMed] [Google Scholar]
  • 54.Lee YF, Lee HJ, Chang C. 2002. Recent advances in the TR2 and TR4 orphan receptors of the nuclear receptor superfamily. J. Steroid Biochem. Mol. Biol. 81:291–308. 10.1016/S0960-0760(02)00118-8 [DOI] [PubMed] [Google Scholar]
  • 55.Liu NC, Lin WJ, Kim E, Collins LL, Lin HY, Yu IC, Sparks JD, Chen LM, Lee YF, Chang C. 2007. Loss of TR4 orphan nuclear receptor reduces phosphoenolpyruvate carboxykinase-mediated gluconeogenesis. Diabetes 56:2901–2909. 10.2337/db07-0359 [DOI] [PubMed] [Google Scholar]
  • 56.Chen LM, Wang RS, Lee YF, Liu NC, Chang YJ, Wu CC, Xie S, Hung YC, Chang C. 2008. Subfertility with defective folliculogenesis in female mice lacking testicular orphan nuclear receptor 4. Mol. Endocrinol. 22:858–867. 10.1210/me.2007-0181 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Xie S, Lee YF, Kim E, Chen LM, Ni J, Fang LY, Liu S, Lin SJ, Abe J, Berk B, Ho FM, Chang C. 2009. TR4 nuclear receptor functions as a fatty acid sensor to modulate CD36 expression and foam cell formation. Proc. Natl. Acad. Sci. U. S. A. 106:13353–13358. 10.1073/pnas.0905724106 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Shi L, Lin YH, Sierant MC, Zhu F, Cui S, Guan Y, Sartor MA, Tanabe O, Lim KC, Engel JD. 2014. Developmental transcriptome analysis of human erythropoiesis. Hum. Mol. Genet.. 10.1093/hmg/ddu167 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Shi L, Sierant MC, Gurdziel K, Zhu F, Cui S, Kolodziej KE, Strouboulis J, Guan Y, Tanabe O, Lim KC, Engel JD. 2014. Biased, non-equivalent gene-proximal and -distal binding motifs of orphan nuclear receptor TR4 in primary human erythroid cells. PLoS Genet. 10:e1004339. 10.1371/journal.pgen.1004339 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Hosoya T, Clifford M, Losson R, Tanabe O, Engel JD. 2013. TRIM28 is essential for erythroblast differentiation in the mouse. Blood 122:3798–3807. 10.1182/blood-2013-04-496166 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Miyagi S, Koide S, Saraya A, Wendt GR, Oshima M, Konuma T, Yamazaki S, Mochizuki-Kashio M, Nakajima-Takagi Y, Wang C, Chiba T, Kitabayashi I, Nakauchi H, Iwama A. 2014. The TIF1β-HP1 system maintains transcriptional integrity of hematopoietic stem cells. Stem Cell Rep. 2:145–152. 10.1016/j.stemcr.2013.12.008 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Zhou XE, Suino-Powell KM, Xu Y, Chan CW, Tanabe O, Kruse SW, Reynolds R, Engel JD, Xu HE. 2011. The orphan nuclear receptor TR4 is a vitamin A-activated nuclear receptor. J. Biol. Chem. 286:2877–2885. 10.1074/jbc.M110.168740 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Menzel S, Garner C, Gut I, Matsuda F, Yamaguchi M, Heath S, Foglio M, Zelenika D, Boland A, Rooks H, Best S, Spector TD, Farrall M, Lathrop M, Thein SL. 2007. A QTL influencing F cell production maps to a gene encoding a zinc-finger protein on chromosome 2p15. Nat. Genet. 39:1197–1199. 10.1038/ng2108 [DOI] [PubMed] [Google Scholar]
  • 64.Craig JE, Rochette J, Fisher CA, Weatherall DJ, Marc S, Lathrop GM, Demenais F, Thein S. 1996. Dissecting the loci controlling fetal haemoglobin production on chromosomes 11p and 6q by the regressive approach. Nat. Genet. 12:58–64. 10.1038/ng0196-58 [DOI] [PubMed] [Google Scholar]
  • 65.Sankaran VG, Menne TF, Scepanovic D, Vergilio JA, Ji P, Kim J, Thiru P, Orkin SH, Lander ES, Lodish HF. 2011. MicroRNA-15a and -16-1 act via MYB to elevate fetal hemoglobin expression in human trisomy 13. Proc. Natl. Acad. Sci. U. S. A. 108:1519–1524. 10.1073/pnas.1018384108 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Sankaran VG, Joshi M, Agrawal A, Schmitz-Abe K, Towne MC, Marinakis N, Markianos K, Berry GT, Agrawal PB. 2013. Rare complete loss of function provides insight into a pleiotropic genome-wide association study locus. Blood 122:3845–3847. 10.1182/blood-2013-09-528315 [DOI] [PubMed] [Google Scholar]
  • 67.Suzuki M, Yamazaki H, Mukai HY, Motohashi H, Shi L, Tanabe O, Engel JD, Yamamoto M. 2013. Disruption of the Hbs1l-Myb locus causes hereditary persistence of fetal hemoglobin in a mouse model. Mol. Cell. Biol. 33:1687–1695. 10.1128/MCB.01617-12 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Mucenski ML, McLain K, Kier AB, Swerdlow SH, Schreiner CM, Miller TA, Pietryga DW, Scott WJ, Jr, Potter SS. 1991. A functional c-myb gene is required for normal murine fetal hepatic hematopoiesis. Cell 65:677–689. 10.1016/0092-8674(91)90099-K [DOI] [PubMed] [Google Scholar]
  • 69.Bender TP, Kremer CS, Kraus M, Buch T, Rajewsky K. 2004. Critical functions for c-Myb at three checkpoints during thymocyte development. Nat. Immunol. 5:721–729. 10.1038/ni1085 [DOI] [PubMed] [Google Scholar]
  • 70.Thomas MD, Kremer CS, Ravichandran KS, Rajewsky K, Bender TP. 2005. c-Myb is critical for B cell development and maintenance of follicular B cells. Immunity 23:275–286. 10.1016/j.immuni.2005.08.005 [DOI] [PubMed] [Google Scholar]
  • 71.Emambokus N, Vegiopoulos A, Harman B, Jenkinson E, Anderson G, Frampton J. 2003. Progression through key stages of haemopoiesis is dependent on distinct threshold levels of c-Myb. EMBO J. 22:4478–4488. 10.1093/emboj/cdg434 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Gonda TJ, Sheiness DK, Bishop JM. 1982. Transcripts from the cellular homologs of retroviral oncogenes: distribution among chicken tissues. Mol. Cell. Biol. 2:617–624 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Bianchi E, Zini R, Salati S, Tenedini E, Norfo R, Tagliafico E, Manfredini R, Ferrari S. 2010. c-myb supports erythropoiesis through the transactivation of KLF1 and LMO2 expression. Blood 116:e99–e110. 10.1182/blood-2009-08-238311 [DOI] [PubMed] [Google Scholar]
  • 74.Farrell JJ, Sherva RM, Chen ZY, Luo HY, Chu BF, Ha SY, Li CK, Lee AC, Li RC, Yuen HL, So JC, Ma ES, Chan LC, Chan V, Sebastiani P, Farrer LA, Baldwin CT, Steinberg MH, Chui DH. 2011. A 3-bp deletion in the HBS1L-MYB intergenic region on chromosome 6q23 is associated with HbF expression. Blood 117:4935–4945. 10.1182/blood-2010-11-317081 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Wadman IA, Osada H, Grutz GG, Agulnick AD, Westphal H, Forster A, Rabbitts TH. 1997. The LIM-only protein Lmo2 is a bridging molecule assembling an erythroid, DNA-binding complex which includes the TAL1, E47, GATA-1 and Ldb1/NLI proteins. EMBO J. 16:3145–3157. 10.1093/emboj/16.11.3145 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Stadhouders R, Thongjuea S, Andrieu-Soler C, Palstra RJ, Bryne JC, van den Heuvel A, Stevens M, de Boer E, Kockx C, van der Sloot A, van den Hout M, van Ijcken W, Eick D, Lenhard B, Grosveld F, Soler E. 2012. Dynamic long-range chromatin interactions control Myb proto-oncogene transcription during erythroid development. EMBO J. 31:986–999. 10.1038/emboj.2011.450 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Stadhouders R, Aktuna S, Thongjuea S, Aghajanirefah A, Pourfarzad F, van Ijcken W, Lenhard B, Rooks H, Best S, Menzel S, Grosveld F, Thein SL, Soler E. 2014. HBS1L-MYB intergenic variants modulate fetal hemoglobin via long-range MYB enhancers. J. Clin. Invest. 124:1699–1710. 10.1172/JCI71520 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Mukai HY, Motohashi H, Ohneda O, Suzuki N, Nagano M, Yamamoto M. 2006. Transgene insertion in proximity to the c-myb gene disrupts erythroid-megakaryocytic lineage bifurcation. Mol. Cell. Biol. 26:7953–7965. 10.1128/MCB.00718-06 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Menzel S, Jiang J, Silver N, Gallagher J, Cunningham J, Surdulescu G, Lathrop M, Farrall M, Spector TD, Thein SL. 2007. The HBS1L-MYB intergenic region on chromosome 6q23.3 influences erythrocyte, platelet, and monocyte counts in humans. Blood 110:3624–3626. 10.1182/blood-2007-05-093419 [DOI] [PubMed] [Google Scholar]
  • 80.Lettre G, Sankaran VG, Bezerra MAC, Araujo AS, Uda M, Sanna S, Cao A, Schlessinger D, Costa FF, Hirschhorn JN, Orkin SH. 2008. DNA polymorphisms at the BCL11A, HBS1L-MYB, and beta-globin loci associate with fetal hemoglobin levels and pain crises in sickle cell disease. P Natl. Acad. Sci. U. S. A. 105:11869–11874. 10.1073/pnas.0804799105 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.van der Harst P, Zhang W, Mateo Leach I, Rendon A, Verweij N, Sehmi J, Paul DS, Elling U, Allayee H, Li X, Radhakrishnan A, Tan ST, Voss K, Weichenberger CX, Albers CA, Al-Hussani A, Asselbergs FW, Ciullo M, Danjou F, Dina C, Esko T, Evans DM, Franke L, Gogele M, Hartiala J, Hersch M, Holm H, Hottenga JJ, Kanoni S, Kleber ME, Lagou V, Langenberg C, Lopez LM, Lyytikainen LP, Melander O, Murgia F, Nolte IM, O'Reilly PF, Padmanabhan S, Parsa A, Pirastu N, Porcu E, Portas L, Prokopenko I, Ried JS, Shin SY, Tang CS, Teumer A, Traglia M, Ulivi S, Westra HJ, Yang J, Zhao JH, Anni F, Abdellaoui A, Attwood A, Balkau B, Bandinelli S, Bastardot F, Benyamin B, Boehm BO, Cookson WO, Das D, de Bakker PI, de Boer RA, de Geus EJ, de Moor MH, Dimitriou M, Domingues FS, Doring A, Engstrom G, Eyjolfsson GI, Ferrucci L, Fischer K, Galanello R, Garner SF, Genser B, Gibson QD, Girotto G, Gudbjartsson DF, Harris SE, Hartikainen AL, Hastie CE, Hedblad B, Illig T, Jolley J, Kahonen M, Kema IP, Kemp JP, Liang L, Lloyd-Jones H, Loos RJ, Meacham S, Medland SE, Meisinger C, Memari Y, Mihailov E, Miller K, Moffatt MF, Nauck M, Novatchkova M, Nutile T, Olafsson I, Onundarson PT, Parracciani D, Penninx BW, Perseu L, Piga A, Pistis G, Pouta A, Puc U, Raitakari O, Ring SM, Robino A, Ruggiero D, Ruokonen A, Saint-Pierre A, Sala C, Salumets A, Sambrook J, Schepers H, Schmidt CO, Sillje HH, Sladek R, Smit JH, Starr JM, Stephens J, Sulem P, Tanaka T, Thorsteinsdottir U, Tragante V, van Gilst WH, van Pelt LJ, van Veldhuisen DJ, Volker U, Whitfield JB, Willemsen G, Winkelmann BR, Wirnsberger G, Algra A, Cucca F, d'Adamo AP, Danesh J, Deary IJ, Dominiczak AF, Elliott P, Fortina P, Froguel P, Gasparini P, Greinacher A, Hazen SL, Jarvelin MR, Khaw KT, Lehtimaki T, Maerz W, Martin NG, Metspalu A, Mitchell BD, Montgomery GW, Moore C, Navis G, Pirastu M, Pramstaller PP, Ramirez-Solis R, Schadt E, Scott J, Shuldiner AR, Smith GD, Smith JG, Snieder H, Sorice R, Spector TD, Stefansson K, Stumvoll M, Tang WH, Toniolo D, Tonjes A, Visscher PM, Vollenweider P, Wareham NJ, Wolffenbuttel BH, Boomsma DI, Beckmann JS, Dedoussis GV, Deloukas P, Ferreira MA, Sanna S, Uda M, Hicks AA, Penninger JM, Gieger C, Kooner JS, Ouwehand WH, Soranzo N, Chambers JC. 2012. Seventy-five genetic loci influencing the human red blood cell. Nature 492:369–375. 10.1038/nature11677 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Nuinoon M, Makarasara W, Mushiroda T, Setianingsih I, Wahidiyat PA, Sripichai O, Kumasaka N, Takahashi A, Svasti S, Munkongdee T, Mahasirimongkol S, Peerapittayamongkol C, Viprakasit V, Kamatani N, Winichagoon P, Kubo M, Nakamura Y, Fucharoen S. 2010. A genome-wide association identified the common genetic variants influence disease severity in β0-thalassemia/hemoglobin E. Hum. Genet. 127:303–314. 10.1007/s00439-009-0770-2 [DOI] [PubMed] [Google Scholar]
  • 83.Galanello R, Sanna S, Perseu L, Sollaino MC, Satta S, Lai ME, Barella S, Uda M, Usala G, Abecasis GR, Cao A. 2009. Amelioration of Sardinian β0 thalassemia by genetic modifiers. Blood 114:3935–3937. 10.1182/blood-2009-04-217901 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.McCarthy MI, Abecasis GR, Cardon LR, Goldstein DB, Little J, Ioannidis JP, Hirschhorn JN. 2008. Genome-wide association studies for complex traits: consensus, uncertainty and challenges. Nat. Rev. Genet. 9:356–369. 10.1038/nrg2344 [DOI] [PubMed] [Google Scholar]
  • 85.Liu P, Keller JR, Ortiz M, Tessarollo L, Rachel RA, Nakamura T, Jenkins NA, Copeland NG. 2003. Bcl11a is essential for normal lymphoid development. Nat. Immunol. 4:525–532. 10.1038/ni925 [DOI] [PubMed] [Google Scholar]
  • 86.Sankaran VG, Menne TF, Xu J, Akie TE, Lettre G, Van Handel B, Mikkola HK, Hirschhorn JN, Cantor AB, Orkin SH. 2008. Human fetal hemoglobin expression is regulated by the developmental stage-specific repressor BCL11A. Science 322:1839–1842. 10.1126/science.1165409 [DOI] [PubMed] [Google Scholar]
  • 87.Sankaran VG, Xu J, Ragoczy T, Ippolito GC, Walkley CR, Maika SD, Fujiwara Y, Ito M, Groudine M, Bender MA, Tucker PW, Orkin SH. 2009. Developmental and species-divergent globin switching are driven by BCL11A. Nature 460:1093–1097. 10.1038/nature08243 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Xu J, Peng C, Sankaran VG, Shao Z, Esrick EB, Chong BG, Ippolito GC, Fujiwara Y, Ebert BL, Tucker PW, Orkin SH. 2011. Correction of sickle cell disease in adult mice by interference with fetal hemoglobin silencing. Science 334:993–996. 10.1126/science.1211053 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Xu J, Sankaran VG, Ni M, Menne TF, Puram RV, Kim W, Orkin SH. 2010. Transcriptional silencing of γ-globin by BCL11A involves long-range interactions and cooperation with SOX6. Genes Dev. 24:783–798. 10.1101/gad.1897310 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Yi Z, Cohen-Barak O, Hagiwara N, Kingsley PD, Fuchs DA, Erickson DT, Epner EM, Palis J, Brilliant MH. 2006. Sox6 directly silences epsilon globin expression in definitive erythropoiesis. PLoS genetics 2:e14. 10.1371/journal.pgen.0020014 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Jawaid K, Wahlberg K, Thein SL, Best S. 2010. Binding patterns of BCL11A in the globin and GATA1 loci and characterization of the BCL11A fetal hemoglobin locus. Blood Cells Mol. Dis. 45:140–146. 10.1016/j.bcmd.2010.05.006 [DOI] [PubMed] [Google Scholar]
  • 92.Yu Y, Wang J, Khaled W, Burke S, Li P, Chen X, Yang W, Jenkins NA, Copeland NG, Zhang S, Liu P. 2012. Bcl11a is essential for lymphoid development and negatively regulates p53. J. Exp. Med. 209:2467–2483. 10.1084/jem.20121846 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.John A, Brylka H, Wiegreffe C, Simon R, Liu P, Juttner R, Crenshaw EB, 3rd, Luyten FP, Jenkins NA, Copeland NG, Birchmeier C, Britsch S. 2012. Bcl11a is required for neuronal morphogenesis and sensory circuit formation in dorsal spinal cord development. Development 139:1831–1841. 10.1242/dev.072850 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Bauer DE, Kamran SC, Lessard S, Xu J, Fujiwara Y, Lin C, Shao Z, Canver MC, Smith EC, Pinello L, Sabo PJ, Vierstra J, Voit RA, Yuan GC, Porteus MH, Stamatoyannopoulos JA, Lettre G, Orkin SH. 2013. An erythroid enhancer of BCL11A subject to genetic variation determines fetal hemoglobin level. Science 342:253–257. 10.1126/science.1242088 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Miller IJ, Bieker JJ. 1993. A novel, erythroid cell-specific murine transcription factor that binds to the CACCC element and is related to the Kruppel family of nuclear proteins. Mol. Cell. Biol. 13:2776–2786 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Nuez B, Michalovich D, Bygrave A, Ploemacher R, Grosveld F. 1995. Defective haematopoiesis in fetal liver resulting from inactivation of the EKLF gene. Nature 375:316–318. 10.1038/375316a0 [DOI] [PubMed] [Google Scholar]
  • 97.Perkins AC, Sharpe AH, Orkin SH. 1995. Lethal beta-thalassaemia in mice lacking the erythroid CACCC-transcription factor EKLF. Nature 375:318–322. 10.1038/375318a0 [DOI] [PubMed] [Google Scholar]
  • 98.Borg J, Papadopoulos P, Georgitsi M, Gutierrez L, Grech G, Fanis P, Phylactides M, Verkerk AJ, van der Spek PJ, Scerri CA, Cassar W, Galdies R, van Ijcken W, Ozgur Z, Gillemans N, Hou J, Bugeja M, Grosveld FG, von Lindern M, Felice AE, Patrinos GP, Philipsen S. 2010. Haploinsufficiency for the erythroid transcription factor KLF1 causes hereditary persistence of fetal hemoglobin. Nat. Genet. 42:801–805. 10.1038/ng.630 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Zhou D, Liu K, Sun CW, Pawlik KM, Townes TM. 2010. KLF1 regulates BCL11A expression and gamma- to beta-globin gene switching. Nat. Genet. 42:742–744. 10.1038/ng.637 [DOI] [PubMed] [Google Scholar]
  • 100.Funnell AP, Mak KS, Twine NA, Pelka GJ, Norton LJ, Radziewic T, Power M, Wilkins MR, Bell-Anderson KS, Fraser ST, Perkins AC, Tam PP, Pearson RC, Crossley M. 2013. Generation of mice deficient in both KLF3/BKLF and KLF8 reveals a genetic interaction and a role for these factors in embryonic globin gene silencing. Mol. Cell. Biol. 33:2976–2987. 10.1128/MCB.00074-13 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Chen Z, Luo HY, Basran RK, Hsu TH, Mang DW, Nuntakarn L, Rosenfield CG, Patrinos GP, Hardison RC, Steinberg MH, Chui DH. 2008. A T-to-G transversion at nucleotide −567 upstream of HBG2 in a GATA-1 binding motif is associated with elevated hemoglobin F. Mol. Cell. Biol. 28:4386–4393. 10.1128/MCB.00071-08 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Harju-Baker S, Costa FC, Fedosyuk H, Neades R, Peterson KR. 2008. Silencing of Aγ-globin gene expression during adult definitive erythropoiesis mediated by GATA-1-FOG-1-Mi2 complex binding at the −566 GATA site. Mol. Cell. Biol. 28:3101–3113. 10.1128/MCB.01858-07 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Costa FC, Fedosyuk H, Chazelle AM, Neades RY, Peterson KR. 2012. Mi2β is required for γ-globin gene silencing: temporal assembly of a GATA-1-FOG-1-Mi2 repressor complex in β-YAC transgenic mice. PLoS Genet. 8:e1003155. 10.1371/journal.pgen.1003155 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.van der Ploeg LH, Flavell RA. 1980. DNA methylation in the human gamma delta beta-globin locus in erythroid and nonerythroid tissues. Cell 19:947–958. 10.1016/0092-8674(80)90086-0 [DOI] [PubMed] [Google Scholar]
  • 105.Groudine M, Weintraub H. 1981. Activation of globin genes during chicken development. Cell 24:393–401. 10.1016/0092-8674(81)90329-9 [DOI] [PubMed] [Google Scholar]
  • 106.Marcaud L, Reynaud CA, Therwath A, Scherrer K. 1981. Modification of the methylation pattern in the vicinity of the chicken globin genes in avian erythroblastosis virus transformed cells. Nucleic Acids Res. 9:1841–1851. 10.1093/nar/9.8.1841 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Haigh LS, Owens BB, Hellewell OS, Ingram VM. 1982. DNA methylation in chicken alpha-globin gene expression. Proc. Natl. Acad. Sci. U. S. A. 79:5332–5336. 10.1073/pnas.79.17.5332 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Busslinger M, Hurst J, Flavell RA. 1983. DNA methylation and the regulation of globin gene expression. Cell 34:197–206. 10.1016/0092-8674(83)90150-2 [DOI] [PubMed] [Google Scholar]
  • 109.Yin W, Barkess G, Fang X, Xiang P, Cao H, Stamatoyannopoulos G, Li Q. 2007. Histone acetylation at the human beta-globin locus changes with developmental age. Blood 110:4101–4107. 10.1182/blood-2007-05-091256 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Fathallah H, Weinberg RS, Galperin Y, Sutton M, Atweh GF. 2007. Role of epigenetic modifications in normal globin gene regulation and butyrate-mediated induction of fetal hemoglobin. Blood 110:3391–3397. 10.1182/blood-2007-02-076091 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Bird A. 2002. DNA methylation patterns and epigenetic memory. Genes Dev. 16:6–21. 10.1101/gad.947102 [DOI] [PubMed] [Google Scholar]
  • 112.Banzon V, Ibanez V, Vaitkus K, Ruiz MA, Peterson K, DeSimone J, Lavelle D. 2011. siDNMT1 increases γ-globin expression in chemical inducer of dimerization (CID)-dependent mouse βYAC bone marrow cells and in baboon erythroid progenitor cell cultures. Exp. Hematol. 39:26–36.e21. 10.1016/j.exphem.2010.10.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Rupon JW, Wang SZ, Gaensler K, Lloyd J, Ginder GD. 2006. Methyl binding domain protein 2 mediates gamma-globin gene silencing in adult human βYAC transgenic mice. Proc. Natl. Acad. Sci. U. S. A. 103:6617–6622. 10.1073/pnas.0509322103 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.DeSimone J, Heller P, Hall L, Zwiers D. 1982. 5-Azacytidine stimulates fetal hemoglobin synthesis in anemic baboons. Proc. Natl. Acad. Sci. U. S. A. 79:4428–4431. 10.1073/pnas.79.14.4428 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Ley TJ, DeSimone J, Anagnou NP, Keller GH, Humphries RK, Turner PH, Young NS, Keller P, Nienhuis AW. 1982. 5-Azacytidine selectively increases gamma-globin synthesis in a patient with beta+ thalassemia. N. Engl. J. Med. 307:1469–1475. 10.1056/NEJM198212093072401 [DOI] [PubMed] [Google Scholar]
  • 116.Ley TJ, DeSimone J, Noguchi CT, Turner PH, Schechter AN, Heller P, Nienhuis AW. 1983. 5-Azacytidine increases gamma-globin synthesis and reduces the proportion of dense cells in patients with sickle cell anemia. Blood 62:370–380 [PubMed] [Google Scholar]
  • 117.Saunthararajah Y, Hillery CA, Lavelle D, Molokie R, Dorn L, Bressler L, Gavazova S, Chen YH, Hoffman R, DeSimone J. 2003. Effects of 5-aza-2′-deoxycytidine on fetal hemoglobin levels, red cell adhesion, and hematopoietic differentiation in patients with sickle cell disease. Blood 102:3865–3870. 10.1182/blood-2003-05-1738 [DOI] [PubMed] [Google Scholar]
  • 118.Shi L, Cui S, Engel JD, Tanabe O. 2013. Lysine-specific demethylase 1 is a therapeutic target for fetal hemoglobin induction. Nat. Med. 19:291–294. 10.1038/nm.3101 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Lee MG, Wynder C, Schmidt DM, McCafferty DG, Shiekhattar R. 2006. Histone H3 lysine 4 demethylation is a target of nonselective antidepressive medications. Chem. Biol. 13:563–567. 10.1016/j.chembiol.2006.05.004 [DOI] [PubMed] [Google Scholar]
  • 120.Xu J, Bauer DE, Kerenyi MA, Vo TD, Hou S, Hsu YJ, Yao H, Trowbridge JJ, Mandel G, Orkin SH. 2013. Corepressor-dependent silencing of fetal hemoglobin expression by BCL11A. Proc. Natl. Acad. Sci. U. S. A. 110:6518–6523. 10.1073/pnas.1303976110 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Cuthill JM, Griffiths AB, Powell DE. 1964. Death associated with tranylcypromine and cheese. Lancet i:1076–1077 [DOI] [PubMed] [Google Scholar]
  • 122.Blackwell B, Marley E, Price J, Taylor D. 1967. Hypertensive interactions between monoamine oxidase inhibitors and foodstuffs. Br. J. Psychiatry 113:349–365. 10.1192/bjp.113.497.349 [DOI] [PubMed] [Google Scholar]
  • 123.Hanna J, Wernig M, Markoulaki S, Sun CW, Meissner A, Cassady JP, Beard C, Brambrink T, Wu LC, Townes TM, Jaenisch R. 2007. Treatment of sickle cell anemia mouse model with iPS cells generated from autologous skin. Science 318:1920–1923. 10.1126/science.1152092 [DOI] [PubMed] [Google Scholar]
  • 124.Jinek M, Chylinski K, Fonfara I, Hauer M, Doudna JA, Charpentier E. 2012. A programmable dual-RNA-guided DNA endonuclease in adaptive bacterial immunity. Science 337:816–821. 10.1126/science.1225829 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Christian M, Cermak T, Doyle EL, Schmidt C, Zhang F, Hummel A, Bogdanove AJ, Voytas DF. 2010. Targeting DNA double-strand breaks with TAL effector nucleases. Genetics 186:757–761. 10.1534/genetics.110.120717 [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Molecular and Cellular Biology are provided here courtesy of Taylor & Francis

RESOURCES