The level of fetal hemoglobin (HbF) modifies the severity of the common β-globin disorders. Knowledge of the normal mechanisms that repress HbF in the adult stage has remained limited until recently despite nearly 3 decades of molecular investigation, in part because of imperfect model systems. Recent studies have provided new insights into the developmental regulation of globin genes and identified specific transcription factors and epigenetic regulators responsible for physiologic silencing of HbF. Most prominent among these regulators is BCL11A, a transcriptional repressor that inhibits adult-stage HbF expression. KLF1 and c-Myb are additional critical HbF-regulating erythroid transcription factors more broadly involved in erythroid gene expression programs. Chromatin modifiers, including histone deacetylases and DNA methyltransferases, also play key roles in orchestrating appropriate globin gene expression. Taken together, these discoveries present novel therapeutic targets for further consideration. Although substantial hurdles remain, opportunities are now rich for the rational design of HbF inducers.

Hemoglobinopathies are among the most common inherited recessive diseases.1  Globin gene mutations occurred numerous times during human history and have been selected to high frequencies in malarial endemic regions. Although many of these globin mutations in the heterozygous state afford modest protection against malaria, the coinheritance of 2 mutant β-globin alleles (in homozygous or compound heterozygous combination) produces the common β-globin disorders. Sickle cell disease (SCD) and the β-thalassemias are chronic diseases with considerable morbidity and mortality. In low-income countries, most affected individuals die in early childhood. Nearly 300 000 infants are estimated born with SCD each year (the majority in Africa), and another 40 000 severely affected with β-thalassemia.2  In the United States alone, the annual medical costs for the approximately 75 000 individuals with SCD are estimated at more than $1 billion.3  Although genetic screening and prenatal diagnosis have reduced the incidence of β-thalassemia in selected countries, such as Sardinia and Cyprus, β-thalassemia remains common in many areas of Asia and the Middle East with limited resources for treatment. Children with β-globin disorders are at particular risk of life-threatening infections. As childhood infectious diseases are brought under better control in the developing world, the β-globin disorders will undoubtedly take on intensified public health significance. Given anticipated population growth, the worldwide prevalence of these diseases is expected to rise dramatically over the next century.4 

Since the initial hematologic descriptions of SCD and β-thalassemia by Herrick and Cooley, respectively, in the early 20th century, studies of the hemoglobinopathies have been at the forefront of human genetics and molecular biology. Sickle cell disease, distinguished by its unique hemoglobin structure because of the characteristic glutamate-to-valine substitution of βS, was heralded as the first “molecular disease.”5page543 The subsequent demonstration of globin chain imbalance as the pathophysiologic underpinning of the thalassemias presaged the molecular biology era in which various thalassemia mutations were dissected, illuminating fundamental aspects of gene regulation.6 

Two gene clusters encode the various globins—the α cluster on chromosome 16 contains the embryonic ζ gene, and adult α1 and α2 genes, and the β cluster on chromosome 11 the embryonic ϵ, the fetal Gγ and Aγ, and adult δ and β genes (Figure 1).7  During early embryonic development, erythropoiesis is yolk-sac–derived, transitioning to the fetal liver midway through the first trimester. Approaching the time of birth, the bone marrow becomes the dominant site of erythropoiesis. Two accompanying switches occur in the expression of genes from the β-globin cluster—a switch from embryonic-to-fetal globins early in gestation, and then from fetal-to-adult globins around the time of birth. Thus HbF (α2γ2) constitutes the major hemoglobin during fetal life, gradually replaced by adult hemoglobin (HbA, α2β2) during infancy. Adults retain low-level expression of HbF (roughly between 0.1% and 1% total hemoglobin), with only a subset of erythrocytes possessing measurable HbF.

Figure 1

The β-globin genes are encoded from a single cluster and under strict developmental control. There are 2 developmental switches in expression from the cluster, from embryonic-to-fetal during the first trimester of conception, and from fetal-to-adult around the time of birth.

Figure 1

The β-globin genes are encoded from a single cluster and under strict developmental control. There are 2 developmental switches in expression from the cluster, from embryonic-to-fetal during the first trimester of conception, and from fetal-to-adult around the time of birth.

Close modal

A preponderance of genetic, biochemical, and clinical observations suggest a salutary role for HbF in the β-globin disorders. Both SCD and β-thalassemia are genetically heterogeneous conditions. The HbS mutation is carried on 5 independent common haplotypes. These haplotypes have identical βS Glu6Val mutations but vary with regard to the level of γ-globin production from the linked gene cluster. Haplotypes associated with higher levels of HbF are associated with milder clinical courses. β-thalassemia may result from diverse point mutations or deletions affecting the β-globin gene. Deletions with differing breakpoints may be associated with relatively low, intermediate, or high rates of γ-globin production from the linked gene cluster (ie, β0-thalassemia, δβ-thalassemia, or hereditary persistence of fetal hemoglobin [HPFH], respectively). Individuals compound heterozygous for β0 and HPFH have a mild clinical course in contrast to the thalassemia major of those homozygous for β0. Similarly, individuals with 1 βS allele and 1 allele of HPFH are protected from the deleterious consequences of SCD, with HbF remaining in the range of 20%-30%, whereas those with 1 βS allele and 1 β-thalassemia allele (βSβ0) have clinical manifestations comparable with those with HbSS.8 

Biochemical studies have demonstrated that the presence of HbF profoundly delays the polymerization and increases the solubility of HbS under deoxygenated conditions. HbF is more potent than HbA in terms of polymer interference.9  The survival advantage of F-cells in SCD has been demonstrated by observing relative enrichment compared with F-reticulocytes.10  γ-globin expression in β-thalassemia mitigates globin chain imbalance, rendering erythropoiesis more effective. In patients with β-hemoglobinopathies who develop mixed chimerism after stem cell transplant, low levels of donor chimerism are sufficient to reverse clinical manifestations of their disease, emphasizing the selective advantage within the bone marrow and peripheral blood of normal erythroid cells.11  Epidemiologic studies have shown that HbF is a major modifier of frequency of painful crises and mortality in SCD.12  Perhaps the most dramatic evidence comes from individual patients—unlike α-thalassemia major which leads to fetal hydrops, β-thalassemia major and SCD only begin to manifest after the first months of life after HbF levels are developmentally silenced.

The sum of this evidence indicates that even modest induction of HbF may be sufficient to ameliorate SCD or β-thalassemia. Importantly, individuals with HPFH as well as their offspring are healthy, suggesting the safety of induction of HbF, even if it were possible to achieve a complete switch from HbA to HbF. Therefore a major goal of hematologists has been to discover targets to reactivate HbF. The supposition has been that the physiologic silencing of HbF during development (known as hemoglobin switching) could be counteracted for therapeutic benefit. In the 1980s, 5-azacytadine, a DNA demethylating agent, was found to increase levels of HbF in baboons and subsequently individuals with β-globin disorders.13,14  Two proposed mechanisms were direct demethylation of the γ-globin promoter and cytotoxicity with induction of HbF inextricably linked to the altered cell-cycle kinetics of stress erythropoiesis.15  Because of concerns related to potential genotoxicity of demethylating agents, studies were conducted with alternative S-phase inhibitors to identify pharmacologics that could induce HbF with more favorable toxicity profiles. This effort led to the discovery of the HbF-inducing potential of hydroxyurea in baboons and then humans.16,17  The mechanism whereby hydroxyurea results in increased HbF remains incompletely understood. Hydroxyurea has been of substantial benefit for many patients with SCD, although it has variable efficacy, requires careful monitoring with dose-limiting myelosuppression, and is of limited utility for β-thalassemia.18,19  A uniformly safe and potent HbF-inducing therapy remains to be discovered. Poor comprehension of the molecular mechanisms operative during the hemoglobin switch has limited the development of novel therapeutics. A recent upsurge of knowledge has reinvigorated the pursuit of rationally designed HbF inducers.

Hemoglobin switching, the result of developmental changes in transcriptional output from the β-globin cluster, has been subject to intensive investigation as a model for the control of gene expression.7  The β-like globin genes are linearly arranged 5′ to 3′ in the order expressed during development. A critical distal regulatory element directing expression within the gene cluster is known as the locus control region (LCR).20  Although transgenic mice carrying LCR–γ-globin constructs display partial developmental silencing indicating some gene-autonomous developmental control, larger transgenes with both the β and γ genes linked to the LCR show more appropriate developmental regulation, suggesting that competition between the globin genes for the LCR is critical for proper developmental regulation.21,22  Inherited deletions of the β-globin cluster that result in considerably elevated HbF levels (ie, deletional HPFH) have pointed to the existence of repressive elements prominently located in the γ-δ intergenic region.23,24 

Novel approaches helped break the logjam in the search for γ-globin repressors to serve as therapeutic targets. Genome-wide association studies (GWAS) identified loci beyond the β-globin cluster important in HbF regulation (Figure 2). Because residual adult-stage HbF level is a heritable quantitative trait, it is particularly suited for genotype–phenotype correlation. GWAS of HbF level have included subjects with and without β-hemoglobinopathies and of diverse ethnic ancestry. The findings across numerous studies have been striking and consistent. In addition to cis-acting variants at the β-globin cluster itself, common variation at 2 additional loci, BCL11A on chromosome 2 and the HBS1L-MYB intergenic interval on chromosome 6, accounts for a substantial fraction (estimated at 15%-20% each) of individual variation in HbF level.25-28  Importantly, variants at each of these loci are also associated with the severity of the β-globin disorders, with high-HbF variants associated with milder disease.29,30 

Figure 2

Genome-wide association studies have revealed 3 loci consistently associated with HbF level and β-globin disorder severity, across various ethnic backgrounds. These include the β-globin cluster itself on chromosome 11, the HBS1L-MYB intergenic interval on chromosome 6, and BCL11A on chromosome 2. A representative Manhattan plot is shown from the CSSCD cohort. Figure courtesy of Guillaume Lettre (Montreal Heart Institute).

Figure 2

Genome-wide association studies have revealed 3 loci consistently associated with HbF level and β-globin disorder severity, across various ethnic backgrounds. These include the β-globin cluster itself on chromosome 11, the HBS1L-MYB intergenic interval on chromosome 6, and BCL11A on chromosome 2. A representative Manhattan plot is shown from the CSSCD cohort. Figure courtesy of Guillaume Lettre (Montreal Heart Institute).

Close modal

Knockdown studies demonstrate that BCL11A, a transcriptional repressor, is required to maintain silencing of HbF expression in primary human adult erythroid progenitors.31  Knockout of BCL11A in mice results in impaired developmental silencing of endogenous murine embryonic globin and transgenic human γ-globin genes.32  BCL11A interacts with GATA1, FOG1, and SOX6, erythroid transcription factors, and with the NuRD nucleosome remodeling and deacetylase complex.31,33  BCL11A occupies critical sites within the β-globin gene cluster, including sequences specifically deleted in HPFH, promoting long-range physical interactions between the LCR and the β promoter at the expense of the γ promoter.24,33  These results support a model in which BCL11A coordinates the hemoglobin switch by participating in multiprotein complexes occupying the β-globin gene cluster.

A recent linkage study mapped a novel form of HPFH to mutations in KLF1.34  Interestingly, the Nan (neonatal anemia) mouse mutant also maps to KLF1, with heterozygotes showing derepressed embryonic globin gene expression.35  KLF1 (also known as EKLF) is a critical erythroid transcription factor required for high-level adult-stage β-globin expression.36-38  It interacts with a CACCC motif at the β promoter at higher affinity than similar elements at the γ promoters.39,40  Knockdown of KLF1 in human erythroid progenitor cells leads to increased HbF expression and murine knockout of KLF1 leads to impaired silencing of transgenic human γ-globin.34,41,42  KLF1 protein occupies the BCL11A promoter and activates BCL11A expression.34,41  Therefore, KLF1 appears to influence hemoglobin switching both by directly activating β-globin in the adult stage as well as promoting the expression of the γ-globin silencer BCL11A. Interestingly, KLF1 mutations have been identified in individuals with a variety of erythroid conditions in addition to HPFH, including the In(Lu) blood antigen phenotype, congenital dyserythropoietic anemia, borderline HbA2, and elevated zinc protoporphyrin.43-46  Heterozygous inactivating mutations of KLF1 do not invariably result in elevated HbF, suggesting that genetic modifiers, including variants at BCL11A, may influence HbF level.43  Moreover, KLF1 has been demonstrated to occupy numerous loci within erythroid precursors and influence a broad array of erythroid gene expression programs.47-50 

Another locus implicated by GWAS is the HBS1L-MYB intergenic interval. Although a contribution for HBS1L cannot be excluded,26  the bulk of evidence implicates MYB in influencing HbF regulation. c-Myb is a hematopoietic transcription factor essential for definitive hematopoiesis.51  Increased expression of c-Myb inhibits γ-globin expression in the K562 cell line.52  Knockdown of c-Myb within primary human erythroid progenitors results in increased HbF.53  c-Myb can influence KLF1 expression.54  A variety of additional factors have been related to hemoglobin switching, including COUP-TF, FOP (Friend of PRMT1), Ikaros, miR-15 and 16, MBD2, NF-E4, NRF2, and TR2/TR4.53,55-61 

The previously mentioned studies implicate specific transcription factors in the developmental control of the globin genes. Each of these factors acts within larger multiprotein complexes with chromatin-modifying activity. The epigenetic state of the globin cluster, including higher-order chromatin structure, histone modifications, and DNA methylation, is correlated with its developmental regulation. Active globin genes are hyperacetylated, whereas inactive globin genes are hypoacetylated.62,63  Moreover, histone deacetylase inhibitors, including short-chain fatty acids, such as butyrate, lead to increased HbF levels.64,65  Specifically, HDAC1 and HDAC2 contribute to HbF repression.66  Chromatin regulators observed to occupy the γ-globin promoter include the arginine methyltransferase PRMT5, lysine methyltransferase SUV4-20h1, serine/threonine kinase CK2α, the NuRD complex, and DNA methyltransferase DNMT3A.67,68  Perhaps the best-characterized epigenetic correlate of HbF level is DNA methylation. Hypermethylation of the β promoter is observed in the embryonic and fetal stages, and hypermethylation of the γ promoter in the adult-stage.69,70 

Many of the critical discoveries regarding mechanisms of the hemoglobin switch derive from human genetics. In part, this is a credit to generations of astute clinical investigators. However, it also speaks to the fact that cellular and animal models of HbF regulation possess significant limitations to their broad applicability.

Cellular models

K562 cells are an extensively studied human cell line with erythroid and megakaryocytic potential isolated from an individual with BCR-ABL+ chronic myeloid leukemia. When erythroid maturation is induced, modest levels of embryonic and fetal globins ζ, ϵ, and γ are synthesized with minimal production of α or β.71  That is, these cells appear to recapitulate a hybrid of embryonic and fetal but not adult-stage erythropoiesis. Likewise these cells express very low levels of BCL11A.31  Therefore their utility in screening and validating targets that participate in adult-stage silencing of γ-globin appears questionable. A variety of other human cell lines with erythroid potential exist (eg, HEL, KU812, etc) but none of these possess an adult-stage globin gene expression pattern.72 

Embryonic stem (ES) and induced pluripotent stem (iPS) cells have been used for erythroid differentiation.73,74  Potential advantages of these systems are the availability from numerous individuals, including those with β-globin disorders. In fact, β-globin disorder patient-derived iPS cells have been generated.75,76  However, current erythroid maturation protocols remain inefficient. Furthermore, erythroid differentiation from pluripotent cells reflects embryonic or fetal stages and does not recapitulate the adult stage.77 

Primary human erythroid cultures are widely used. A variety of 2-phase cell culture systems have been described, with expansion of early hematopoietic progenitors before enforced erythroid maturation. These protocols use either rare circulating progenitors present in peripheral blood, or enriched CD34+ hematopoietic stem/progenitor cells from bone marrow aspirates or G-CSF mobilized peripheral blood.78-80  A clear asset of these systems is their capacity to model aspects of human erythropoiesis, including adult-stage pattern globin gene expression and expression of critical regulators including BCL11A, KLF1, and c-Myb. Disadvantages of the primary human cell systems include limited proliferative capacity and asynchronous differentiation manner, especially with regard to the terminal stages of erythroid maturation and enucleation. These cells are particularly subject to maturational switching so careful monitoring of kinetics is mandatory for studies of γ-globin expression. Furthermore these cells typically have a relatively high background level of HbF and appear unusually permissive for HbF induction, thus rendering them sensitive but not necessarily specific indicators of manipulations inducing HbF.81,82  Extrinsic signals, such as fetal calf serum or specific cytokines, such as stem cell factor, may contribute to elevated HbF levels in these cultures.83  Stress erythropoiesis is a final common pathway associated with elevated HbF production; this is seen in patients recovering from myelosuppression (and may underlie some of the HbF induction by hydroxyurea). Therefore exquisite care must be taken to ensure any perturbations of globin gene expression are out of proportion to impacts on cellular growth and proliferation. Agents that increase HbF in primary human erythroid cultures may not necessarily have equivalent behavior in vivo. Additional studies are required to corroborate effects on HbF induction, total globin synthesis, and erythroid differentiation.

Mouse erythroleukemia (MEL) cells demonstrate adult-pattern globin expression, characterized by high-level expression of the adult globins βmajor and βminor and repression of the embryonic globins ϵy and βh1. Knockdown of BCL11A in these cells recapitulates embryonic globin derepression.33  Recently a MEL line was reported in which a human β-globin cluster transgene was engineered to have fluorescent reporter genes in place of the γ and β-globin coding sequences; however, the relatively small shift in ratio of γ/β-globin reporters on BCL11A knockdown suggests limited robustness for small molecule screening.84  Additional cell lines have been generated from mice carrying the entire human β-globin cluster as a transgene.85  All murine cellular systems face the caveat of species-level developmental differences in physiologic hemoglobin switching.

Animal models

A faithful small animal model would offer enormous advantages for screening and validation. The benefits of such a model could include ability to assess impact of modulation of a target: on globin gene expression throughout the developmental stages of erythropoiesis; on aspects of erythropoiesis in addition to globin gene regulation; on other hematopoietic lineages; and beyond the hematopoietic system. A critical drawback of current small animal models is that human HbF regulation lacks a direct analog in rodents. The emergence of fetal stage γ-globin expression occurred approximately 45 million years ago, before the divergence of Old World monkeys, apes, and humans.86  Other vertebrates lack an HbF equivalent although all have at least 1 developmental switch within the β-globin cluster. For example, in contrast to human erythropoiesis consisting of 2 developmental transitions in globin gene expression (from embryonic-to-fetal and from fetal-to-adult) only 1 switch in globin gene expression (from embryonic-to-adult) takes place in the mouse. Therefore, interventions that interfere with mouse globin switching may not directly apply to the human fetal-to-adult transition although certain critical mechanistic aspects do appear conserved.32  Many studies have taken advantage of mice carrying a human β-globin cluster transgene, consisting of all 5 human β-like globin genes as well as the LCR with preserved spatial organization, to circumvent the absence of an endogenous mouse γ-globin gene. These transgenic mice display high-level, lineage-restricted, developmentally regulated expression of the β-like globin genes. However, silencing of human γ-globin occurs alongside that of human ϵ and mouse embryonic β-like globins ϵy and βh1 during the transition from primitive yolk-sac–derived erythropoiesis to the definitive fetal liver stage.32,87,88  The degree of repression is much more profound for embryonic murine globins and transgenic human γ globin (on the order of 1:10 000 relative to adult β-globin expression) in contrast to the relatively more modest degree of repression of γ-globin observed in humans, in whom residual expression of HbF between 0.1% and 1% persists throughout adulthood.89  Therefore interventions seeking to derepress γ globin expression face a greater quantitative hurdle and potentially unique epigenetic mechanisms in the mouse environment compared with the human. BCL11A plays a critical role in the mouse globin switch. However, even in mice deficient for BCL11A within the erythroid lineage, adult γ-globin production remains at ∼ 10%-15% of total human β-like globin production.89  In contrast, in murine Berkeley sickle mice, with basal γ-globin level orders of magnitude higher (∼ 1% of total β-like globin), BCL11A deficiency results in derepression of γ-globin up to ∼ 30%, enough for reversal of the end-organ damage caused by SCD.89  Of note, this SCD model contains a “mini-LCR” globin gene cassette, which does not possess the full complement of regulatory elements at the β-globin cluster so might not completely mirror regulation of the endogenous human genes.90 

Recent evidence suggests the regulation of transgenic human γ-globin is more similar to murine embryonic βh1 than to ϵy, with expression in early definitive erythroid precursors of the nascent fetal liver.91,92  Intriguingly, βh1 shares an evolutionary origin with human γ-globin whereas murine ϵy is the paralog of human ϵ.93  Moreover, γ-globin is known to undergo maturational switching during normal human erythropoiesis whereby it is expressed at higher levels in less mature erythroid precursors.94  The mechanistic correlates of embryonic-to-fetal switching as well as maturational switching remain incompletely elucidated. However, loss of BCL11A results in derepression of both ϵy and βh1 to a similar degree during adulthood.

The zebrafish was recently shown to undergo 2 developmental switches in globin gene expression, from embryonic-to-larval and larval-to-adult.95  The globin gene clusters share partial synteny and similar distal regulatory elements with the human genes though it remains to be elucidated whether fish and mammalian hemoglobin switches share trans-acting regulatory mechanisms. Nonetheless, the zebrafish model serves as a complementary system for genetic and chemical screening in numerous aspects of hematopoiesis.96 

Large animals have also been used as a model of human hemoglobin switching, as fetal-stage γ-globin expression evolved in a common primate ancestor.86  Studies in baboons have taken advantage of this shared γ-globin regulation to demonstrate that HbF levels increase as a result of stress erythropoiesis or demethylating agents.97  The baboon system has been used to study epigenetic changes at the β-globin cluster.98  Potential limitations with this model include subtle differences in γ-globin regulation, even between various baboon species, and inherent challenges to conducting genetic or even high-throughput chemical studies in such animals.97,99 

The ideal molecular target for therapeutic induction of HbF for the β-globin disorders would meet several criteria including: direct and significant involvement in HbF silencing; limited effects on total hemoglobin synthesis and on erythroid maturation; absence of critical functions in nonerythroid cells; validation in primary human erythroid cultures as well as in animal models; and feasibility of therapeutic modulation. It should be noted that in other “rational therapy” settings, the criterion of cellular specificity is not always stringently applied. For example, kinase inhibitors that are effective in treating hematopoietic malignancies often also inhibit kinases expressed in noncancerous cells. Until pharmacologic agents are available that directly reactivate HbF expression in adult erythroid cells, it is premature to consider effects in other cell types.

Targets

The genetically implicated transcription factors, BCL11A, KLF1, and c-Myb may all be considered prospective targets for therapy. Assets and deficiencies of each are apparent. The ubiquitous expression of c-Myb in HSCs, and progenitors, and requirement for c-Myb in the maintenance and differentiation of HSCs, raise concerns as to whether an adequate therapeutic window exists whereby c-Myb could be modulated without having negative effects on hematopoiesis. KLF1 has the benefit of exclusive expression within the erythroid lineage. However, pleiotropic erythroid phenotypes observed with KLF1 mutations in both mouse and man indicate that it may be difficult to interfere with this factor without broadly influencing red blood cell production and function. Of note, mice that lack KLF1 die in utero from severe anemia, even if lethal thalassemia is averted by enforced expression of globin genes, demonstrating the critical importance of this factor for erythropoiesis.100  BCL11A has several favorable attributes for therapeutic aim. Its potency in HbF silencing has been demonstrated in a variety of systems from cell lines and primary cell cultures to transgenic mice. For example, in mouse models of sickle cell disease, loss of BCL11A alone produces pancellular HbF induction and reverses the characteristic end-organ damage.89  BCL11A influences the expression of very few genes within the erythroid lineage besides the globins, and its loss is not associated with any discernable erythroid phenotype besides embryonic and fetal globin gene derepression.89  However, there are several potential drawbacks to targeting BCL11A.

One concern relates to essential roles of BCL11A outside the erythroid lineage. BCL11A null mice are perinatal lethal for unclear reasons.101  Within the central nervous system BCL11A is expressed both during embryonic development and after birth.102  BCL11A serves important roles in neuronal differentiation and morphogenesis as well as axonal guidance.103,104  Fortunately, potential untoward CNS effects of BCL11A inhibition might be avoided by an intact blood-brain barrier. Genetic variants at the BCL11A locus have been associated with diabetes mellitus, although a direct role for BCL11A in metabolic homeostasis has not been identified.105,106  Within the hematopoietic compartment, BCL11A is essential for proper development of B-lymphocytes during ontogeny, although its role in postnatal B-cell function remains unclear.101  Given the divergent kinetics of B-cell and erythrocyte production, it seems plausible that intermittent blockade of BCL11A could lead to substantial induction of HbF without excessive interference with B-lymphocyte homeostasis. T-cell neoplasms were identified in recipients of BCL11A-deficient fetal-liver transplants, although the effect appeared to be noncell autonomous, and may have been related to the immunodeficient state of the recipients.101 

A more generic concern is that transcription factors have traditionally been considered to be unfavorable drug targets.107  Unlike enzymes, transcription factors lack catalytic domains, and unlike receptors they lack ligands (with the obvious exception of the nuclear hormone receptors). Thus, interference with protein-DNA or protein-protein interactions would appear to be required. By considering the larger chromatin context within which individual transcription factors act, novel targets might materialize. For example, emerging evidence suggests chromatin readers (proteins with recognition modules for specific epigenetic marks) may serve as a susceptible node for small-molecule modulation of transcription. In particular, highly selective small molecules have been designed that interfere with the tandem bromodomain (BET) motif for acetyl-lysine recognition.108,109  These compounds have varied and potent biologic activity in mouse models, preventing LPS-mediated changes in gene expression and subsequent fatal sepsis as well as reversing malignant progression in diverse BET-reliant cancers.108-110  Although the precise contextual determinants of specificity for the chromatin readers remain to be elucidated, the therapeutic potential is striking.

In addition to chromatin “reading,” chromatin “writing” and “erasing” could be important targets in hemoglobin switching. It may be instructive that demethylating agents interfering with DNA methylation were the first compounds to demonstrate HbF induction in patients with β-globin disorders. Next-generation demethylating agents with improved toxicity profiles, such as decitabine, may offer benefit to patients with β-globin disorders.111,112  Although short-chain fatty acid derivatives, many of which appear to exert their action through HDAC inhibition, have shown activity in patients with β-globin disorders, cytotoxicity has limited their clinical applicability.64  Perhaps more specific epigenetic therapies, such as HDAC1 or HDAC2 selective inhibitors, would have an improved therapeutic index.

Purely “agnostic” approaches to genetic or chemical screening may yet yield novel targets for HbF inhibition, although these must tackle the aforementioned challenges of screening platforms and robust validation before moving toward clinical translation. An example of unexpected small molecule modulators is provided by the so-called immunomodulatory compounds lenalidomide and pomalidomide (thalidomide derivatives). These novel agents that have been developed for hematologic malignancies appear to have modest effects in terms of HbF induction.113,114  In this case, the biologic importance of the class of small molecules preceded the identification of the molecular target—although recent studies suggest cereblon, an E3 ubiqutin ligase, may mediate some of its biologic effects.115 

Therapeutic modalities

Even after identification of a suitable target, determination of an appropriate modality to address the target remains a formidable challenge (Figure 3). For enzymes, active-site inhibitors would be an obvious strategy. The recognition modules of chromatin readers are another conceptually straightforward approach for small molecule modulation. For multiprotein complexes, critical interfaces between partner protein subunits may be targeted, for example by stabilized peptides.116  Allosteric inhibitors are plausible for a wide spectrum of proteins. Contemporary strategies of compound discovery, such as small molecule microarrays as well as expanding structural knowledge raise hope that transcription factors may not be as “undruggable” as initially supposed.107 

Figure 3

The BCL11A network as a schematic target for various potential therapeutic modalities. BCL11A is shown occupying sequences within the β-globin cluster distal from the γ-globin genes themselves. It is subject to transcriptional activation by KLF1, which itself may be a target of c-Myb. BCL11A interacts with the NuRD nucleosome remodeling and deacetylase complex, which includes the ATPases CHD3/4 and histone deacetylases HDAC1/2 as well as MBD2. BCL11A also interacts with erythroid transcription factors including GATA1, FOG1, and SOX6. BCL11A could conceptually be therapeutically targeted by various strategies including: (1) decreasing its steady-state level, such as by preventing its activation by KLF1, or by RNA interference; (2) interfering with protein–protein interactions such as between BCL11A and GATA1; (3) interfering directly with BCL11A's protein–DNA interactions; (4) allosteric inhibitors of BCL11A itself or various partners; (5) active-site inhibitors of partner proteins with enzymatic activity such as HDAC1/2 and CHD3/4; (6) blocking associated chromatin reader modules, such as PHD fingers and chromodomains on CHD3/4 and MBD domain on MBD2; (7) direct genome editing, such as of critical BCL11A binding regulatory elements; and (8) as part of combination therapy with additional targets such as low-dose demethylase therapy.

Figure 3

The BCL11A network as a schematic target for various potential therapeutic modalities. BCL11A is shown occupying sequences within the β-globin cluster distal from the γ-globin genes themselves. It is subject to transcriptional activation by KLF1, which itself may be a target of c-Myb. BCL11A interacts with the NuRD nucleosome remodeling and deacetylase complex, which includes the ATPases CHD3/4 and histone deacetylases HDAC1/2 as well as MBD2. BCL11A also interacts with erythroid transcription factors including GATA1, FOG1, and SOX6. BCL11A could conceptually be therapeutically targeted by various strategies including: (1) decreasing its steady-state level, such as by preventing its activation by KLF1, or by RNA interference; (2) interfering with protein–protein interactions such as between BCL11A and GATA1; (3) interfering directly with BCL11A's protein–DNA interactions; (4) allosteric inhibitors of BCL11A itself or various partners; (5) active-site inhibitors of partner proteins with enzymatic activity such as HDAC1/2 and CHD3/4; (6) blocking associated chromatin reader modules, such as PHD fingers and chromodomains on CHD3/4 and MBD domain on MBD2; (7) direct genome editing, such as of critical BCL11A binding regulatory elements; and (8) as part of combination therapy with additional targets such as low-dose demethylase therapy.

Close modal

Interfering with the synthesis of the target is another potential strategy. A direct approach is systemic administration of a compound capable of RNA inhibition, such as modified inhibitory RNA.117  In this approach, effects should be reversible, and potentially titrated to effect. Systemic inhibitory RNA technologies are limited by efficiency of cellular uptake and tissue distribution in vivo. A second approach might involve stable expression of an shRNA for a given target, such as BCL11A, using gene therapy in HSCs ex vivo before transfer back to the patient.118  Expression would need to be persistent for efficacy. Even a limited fraction of transduced HSCs could be of benefit given in vivo selection for HbF-containing erythroid cells. In fact, residual nontransduced HSCs might obviate some possible extraerythroid toxicities. With this strategy, effects of attacking the target outside the hematopoietic lineages would not be relevant. The standard concerns regarding genotoxicity secondary to gene therapy need to be considered. Compared with other gene therapy approaches for SCD, interfering with a target such as BCL11A would be associated with down-regulation of HbS in a reciprocal manner as up-regulation of HbF—an advantage over mere addition of a nonsickling β or γ chain.

Direct genome editing is another possible therapeutic modality.119  Sequence-specific nucleases (such as zinc-finger nucleases or transcription activator-like effector nucleases) are tools that can introduce targeted double-stranded breaks in the genome to produce mutations, such as frameshifts or deletions, or to stimulate homologous recombination. For therapeutic purposes, genome editing would require exquisite specificity to prevent off-target mutagenic events.

Some of these strategies appear more feasible for the relatively short-term horizon; for example, nucleic acid based therapeutics could apply existing technical knowledge to a novel target (eg, BCL11A), and might be suitable for resource-rich environments. Other approaches might require a longer timeframe to materialize as they require the development of novel techniques, the acquisition of structural knowledge, or the identification of novel targets. However, these longer-term approaches might ultimately yield a small-molecule that could be more easily scaled to the developing world where the majority of β-globin disorder patients reside.

Any pharmacologic intervention would be unlikely to achieve complete blockade of its target. For a dose-dependent regulator of HbF, such as BCL11A, incomplete targeting is desirable as it permits a therapeutic window in which adequate HbF induction might occur with minimal toxicity. In addition, combination therapy could allow for various unique strategies to achieve additive or synergistic efficacy. For example, combining genetic BCL11A deficiency with chemical histone deacetylation or DNA demethylation leads to synergistic derepression of transgenic human γ-globin in mice.89  One could envision otherwise modestly effective or toxic therapies playing a valuable role in combination rather than as single agents; for example, low-dose demethylating agents.

After target identification and therapeutic modality design and validation, clinical trials are required to demonstrate safety and efficacy in humans. A challenge for all the model systems is that it is difficult to directly extrapolate the percentage HbF induction expected in humans based on experiments in cells or animals. Fortunately, HbF is an ideal biomarker for clinical studies, inextricably linked to the mechanism of action and compellingly associated with meaningful clinical outcomes in both SCD and β-thalassemia. It may be feasible to define safety and efficacy in healthy individuals with intrapatient dose-escalation before study in patients with β-globin disorders. A greater requirement of γ-globin induction for therapeutic benefit is anticipated for β-thalassemia patients in whom the goal would be to mitigate globin chain imbalance compared with SCD patients in whom the goal is prevention of HbS polymerization.

Targets that directly influence HbF silencing have been identified. Translating this knowledge to the clinics faces hurdles of imperfect model systems and targets. The empiric era of discovery of HbF inducers should now yield to a period of mechanism-based therapeutic design. Exquisitely devised and performed clinical studies will be required to ensure that patients will benefit from safe and efficacious new therapies. It will be necessary to engage interdisciplinary efforts of clinicians, molecular biologists, chemists, and pharmacologists. In addition, as both “orphan” genetic diseases and an emerging public health problem, the β-hemoglobinopathies could benefit enormously from involvement of disease advocates.

The authors apologize to the many authors whose work they were unable to cite because of space constraints.

Contribution: D.E.B. and S.C.K. reviewed the literature; and D.E.B., S.C.K., and S.H.O. wrote the paper.

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

Correspondence: Stuart H. Orkin, Department of Pediatric Oncology, Dana-Farber Cancer Institute, 44 Binney St, Boston, MA 02115; e-mail: orkin@bloodgroup.tch.harvard.edu.

1
Weatherall
 
DJ
The inherited diseases of hemoglobin are an emerging global health burden.
Blood
2010
, vol. 
115
 
22
(pg. 
4331
-
4336
)
2
Modell
 
B
Darlison
 
M
Global epidemiology of haemoglobin disorders and derived service indicators.
Bull World Health Organ
2008
, vol. 
86
 
6
(pg. 
480
-
487
)
3
Kauf
 
TL
Coates
 
TD
Huazhi
 
L
Mody-Patel
 
N
Hartzema
 
AG
The cost of health care for children and adults with sickle cell disease.
Am J Hematol
2009
, vol. 
84
 
6
(pg. 
323
-
327
)
4
World urbanization prospects, the 2011 revision
2012
Accessed May 15, 2012 
5
Pauling
 
L
Itano
 
HA
Sickle cell anemia a molecular disease.
Science
1949
, vol. 
110
 
2865
(pg. 
543
-
548
)
6
Weatherall
 
DJ
Towards molecular medicine; reminiscences of the haemoglobin field, 1960-2000.
Br J Haematol
2001
, vol. 
115
 
4
(pg. 
729
-
738
)
7
Stamatoyannopoulos
 
G
Control of globin gene expression during development and erythroid differentiation.
Exp Hematol
2005
, vol. 
33
 
3
(pg. 
259
-
271
)
8
Bank
 
A
Regulation of human fetal hemoglobin: New players, new complexities.
Blood
2006
, vol. 
107
 
2
(pg. 
435
-
443
)
9
Sunshine
 
HR
Hofrichter
 
J
Eaton
 
WA
Requirement for therapeutic inhibition of sickle haemoglobin gelation.
Nature
1978
, vol. 
275
 
5677
(pg. 
238
-
240
)
10
Dover
 
GJ
Boyer
 
SH
Charache
 
S
Heintzelman
 
K
Individual variation in the production and survival of F cells in sickle-cell disease.
N Engl J Med
1978
, vol. 
299
 
26
(pg. 
1428
-
1435
)
11
Andreani
 
M
Testi
 
M
Gaziev
 
J
, et al. 
Quantitatively different red cell/nucleated cell chimerism in patients with long-term, persistent hematopoietic mixed chimerism after bone marrow transplantation for thalassemia major or sickle cell disease.
Haematologica
2011
, vol. 
96
 
1
(pg. 
128
-
133
)
12
Platt
 
OS
Brambilla
 
DJ
Rosse
 
WF
, et al. 
Mortality in sickle cell disease. life expectancy and risk factors for early death.
N Engl J Med
1994
, vol. 
330
 
23
(pg. 
1639
-
1644
)
13
DeSimone
 
J
Heller
 
P
Hall
 
L
Zwiers
 
D
5-azacytidine stimulates fetal hemoglobin synthesis in anemic baboons.
Proc Natl Acad Sci U S A
1982
, vol. 
79
 
14
(pg. 
4428
-
4431
)
14
Ley
 
TJ
DeSimone
 
J
Anagnou
 
NP
, et al. 
5-azacytidine selectively increases gamma-globin synthesis in a patient with beta+ thalassemia.
N Engl J Med
1982
, vol. 
307
 
24
(pg. 
1469
-
1475
)
15
Ley
 
TJ
Chiang
 
YL
Haidaris
 
D
Anagnou
 
NP
Wilson
 
VL
Anderson
 
WF
DNA methylation and regulation of the human beta-globin-like genes in mouse erythroleukemia cells containing human chromosome 11.
Proc Natl Acad Sci U S A
1984
, vol. 
81
 
21
(pg. 
6618
-
6622
)
16
Letvin
 
NL
Linch
 
DC
Beardsley
 
GP
McIntyre
 
KW
Nathan
 
DG
Augmentation of fetal-hemoglobin production in anemic monkeys by hydroxyurea.
N Engl J Med
1984
, vol. 
310
 
14
(pg. 
869
-
873
)
17
Platt
 
OS
Orkin
 
SH
Dover
 
G
Beardsley
 
GP
Miller
 
B
Nathan
 
DG
Hydroxyurea enhances fetal hemoglobin production in sickle cell anemia.
J Clin Invest
1984
, vol. 
74
 
2
(pg. 
652
-
656
)
18
Platt
 
OS
Hydroxyurea for the treatment of sickle cell anemia.
N Engl J Med
2008
, vol. 
358
 
13
(pg. 
1362
-
1369
)
19
Taher
 
AT
Musallam
 
KM
Karimi
 
M
, et al. 
Overview on practices in thalassemia intermedia management aiming for lowering complication rates across a region of endemicity: The OPTIMAL CARE study.
Blood
2010
, vol. 
115
 
10
(pg. 
1886
-
1892
)
20
Noordermeer
 
D
de Laat
 
W
Joining the loops: beta-globin gene regulation.
IUBMB Life
2008
, vol. 
60
 
12
(pg. 
824
-
833
)
21
Behringer
 
RR
Ryan
 
TM
Palmiter
 
RD
Brinster
 
RL
Townes
 
TM
Human gamma- to beta-globin gene switching in transgenic mice.
Genes Dev
1990
, vol. 
4
 
3
(pg. 
380
-
389
)
22
Enver
 
T
Raich
 
N
Ebens
 
AJ
Papayannopoulou
 
T
Costantini
 
F
Stamatoyannopoulos
 
G
Developmental regulation of human fetal-to-adult globin gene switching in transgenic mice.
Nature
1990
, vol. 
344
 
6264
(pg. 
309
-
313
)
23
Chakalova
 
L
Osborne
 
CS
Dai
 
YF
, et al. 
The corfu deltabeta thalassemia deletion disrupts gamma-globin gene silencing and reveals post-transcriptional regulation of HbF expression.
Blood
2005
, vol. 
105
 
5
(pg. 
2154
-
2160
)
24
Sankaran
 
VG
Xu
 
J
Byron
 
R
, et al. 
A functional element necessary for fetal hemoglobin silencing.
N Engl J Med
2011
, vol. 
365
 
9
(pg. 
807
-
814
)
25
Thein
 
SL
Menzel
 
S
Peng
 
X
, et al. 
Intergenic variants of HBS1L-MYB are responsible for a major quantitative trait locus on chromosome 6q23 influencing fetal hemoglobin levels in adults.
Proc Natl Acad Sci U S A
2007
, vol. 
104
 
27
(pg. 
11346
-
11351
)
26
Menzel
 
S
Garner
 
C
Gut
 
I
, et al. 
A QTL influencing F cell production maps to a gene encoding a zinc-finger protein on chromosome 2p15.
Nat Genet
2007
, vol. 
39
 
10
(pg. 
1197
-
1199
)
27
Uda
 
M
Galanello
 
R
Sanna
 
S
, et al. 
Genome-wide association study shows BCL11A associated with persistent fetal hemoglobin and amelioration of the phenotype of beta-thalassemia.
Proc Natl Acad Sci U S A
2008
, vol. 
105
 
5
(pg. 
1620
-
1625
)
28
Sedgewick
 
AE
Timofeev
 
N
Sebastiani
 
P
, et al. 
BCL11A is a major HbF quantitative trait locus in three different populations with beta-hemoglobinopathies.
Blood Cells Mol Dis
2008
, vol. 
41
 
3
(pg. 
255
-
258
)
29
Lettre
 
G
Sankaran
 
VG
Bezerra
 
MA
, et al. 
DNA polymorphisms at the BCL11A, HBS1L-MYB, and beta-globin loci associate with fetal hemoglobin levels and pain crises in sickle cell disease.
Proc Natl Acad Sci U S A
2008
, vol. 
105
 
33
(pg. 
11869
-
11874
)
30
Nuinoon
 
M
Makarasara
 
W
Mushiroda
 
T
, et al. 
A genome-wide association identified the common genetic variants influence disease severity in beta0-thalassemia/hemoglobin E.
Hum Genet
2010
, vol. 
127
 
3
(pg. 
303
-
314
)
31
Sankaran
 
VG
Menne
 
TF
Xu
 
J
, et al. 
Human fetal hemoglobin expression is regulated by the developmental stage-specific repressor BCL11A.
Science
2008
, vol. 
322
 
5909
(pg. 
1839
-
1842
)
32
Sankaran
 
VG
Xu
 
J
Ragoczy
 
T
, et al. 
Developmental and species-divergent globin switching are driven by BCL11A.
Nature
2009
, vol. 
460
 
7259
(pg. 
1093
-
1097
)
33
Xu
 
J
Sankaran
 
VG
Ni
 
M
, et al. 
Transcriptional silencing of {gamma}-globin by BCL11A involves long-range interactions and cooperation with SOX6.
Genes Dev
2010
, vol. 
24
 
8
(pg. 
783
-
798
)
34
Borg
 
J
Papadopoulos
 
P
Georgitsi
 
M
, et al. 
Haploinsufficiency for the erythroid transcription factor KLF1 causes hereditary persistence of fetal hemoglobin.
Nat Genet
2010
, vol. 
42
 
9
(pg. 
801
-
805
)
35
Siatecka
 
M
Sahr
 
KE
Andersen
 
SG
Mezei
 
M
Bieker
 
JJ
Peters
 
LL
Severe anemia in the nan mutant mouse caused by sequence-selective disruption of erythroid kruppel-like factor.
Proc Natl Acad Sci U S A
2010
, vol. 
107
 
34
(pg. 
15151
-
15156
)
36
Perkins
 
AC
Sharpe
 
AH
Orkin
 
SH
Lethal beta-thalassaemia in mice lacking the erythroid CACCC-transcription factor EKLF.
Nature
1995
, vol. 
375
 
6529
(pg. 
318
-
322
)
37
Nuez
 
B
Michalovich
 
D
Bygrave
 
A
Ploemacher
 
R
Grosveld
 
F
Defective haematopoiesis in fetal liver resulting from inactivation of the EKLF gene.
Nature
1995
, vol. 
375
 
6529
(pg. 
316
-
318
)
38
Guy
 
LG
Mei
 
Q
Perkins
 
AC
Orkin
 
SH
Wall
 
L
Erythroid kruppel-like factor is essential for beta-globin gene expression even in absence of gene competition, but is not sufficient to induce the switch from gamma-globin to beta-globin gene expression.
Blood
1998
, vol. 
91
 
7
(pg. 
2259
-
2263
)
39
Miller
 
IJ
Bieker
 
JJ
A novel, erythroid cell-specific murine transcription factor that binds to the CACCC element and is related to the kruppel family of nuclear proteins.
Mol Cell Biol
1993
, vol. 
13
 
5
(pg. 
2776
-
2786
)
40
Feng
 
WC
Southwood
 
CM
Bieker
 
JJ
Analyses of beta-thalassemia mutant DNA interactions with erythroid kruppel-like factor (EKLF), an erythroid cell-specific transcription factor.
J Biol Chem
1994
, vol. 
269
 
2
(pg. 
1493
-
1500
)
41
Zhou
 
D
Liu
 
K
Sun
 
CW
Pawlik
 
KM
Townes
 
TM
KLF1 regulates BCL11A expression and gamma- to beta-globin gene switching.
Nat Genet
2010
, vol. 
42
 
9
(pg. 
742
-
744
)
42
Perkins
 
AC
Gaensler
 
KM
Orkin
 
SH
Silencing of human fetal globin expression is impaired in the absence of the adult beta-globin gene activator protein EKLF.
Proc Natl Acad Sci U S A
1996
, vol. 
93
 
22
(pg. 
12267
-
12271
)
43
Satta
 
S
Perseu
 
L
Moi
 
P
, et al. 
Compound heterozygosity for KLF1 mutations associated with remarkable increase of fetal hemoglobin and red cell protoporphyrin.
Haematologica
2011
, vol. 
96
 
5
(pg. 
767
-
770
)
44
Singleton
 
BK
Burton
 
NM
Green
 
C
Brady
 
RL
Anstee
 
DJ
Mutations in EKLF/KLF1 form the molecular basis of the rare blood group in(lu) phenotype.
Blood
2008
, vol. 
112
 
5
(pg. 
2081
-
2088
)
45
Arnaud
 
L
Saison
 
C
Helias
 
V
, et al. 
A dominant mutation in the gene encoding the erythroid transcription factor KLF1 causes a congenital dyserythropoietic anemia.
Am J Hum Genet
2010
, vol. 
87
 
5
(pg. 
721
-
727
)
46
Perseu
 
L
Satta
 
S
Moi
 
P
, et al. 
KLF1 gene mutations cause borderline HbA(2).
Blood
2011
, vol. 
118
 
16
(pg. 
4454
-
4458
)
47
Drissen
 
R
von Lindern
 
M
Kolbus
 
A
, et al. 
The erythroid phenotype of EKLF-null mice: Defects in hemoglobin metabolism and membrane stability.
Mol Cell Biol
2005
, vol. 
25
 
12
(pg. 
5205
-
5214
)
48
Hodge
 
D
Coghill
 
E
Keys
 
J
, et al. 
A global role for EKLF in definitive and primitive erythropoiesis.
Blood
2006
, vol. 
107
 
8
(pg. 
3359
-
3370
)
49
Tallack
 
MR
Whitington
 
T
Yuen
 
WS
, et al. 
A global role for KLF1 in erythropoiesis revealed by ChIP-seq in primary erythroid cells.
Genome Res
2010
, vol. 
20
 
8
(pg. 
1052
-
1063
)
50
Pilon
 
AM
Ajay
 
SS
Kumar
 
SA
, et al. 
Genome-wide ChIP-seq reveals a dramatic shift in the binding of the transcription factor erythroid kruppel-like factor during erythrocyte differentiation.
Blood
2011
, vol. 
118
 
17
(pg. 
e139
-
48
)
51
Mucenski
 
ML
McLain
 
K
Kier
 
AB
, et al. 
A functional c-myb gene is required for normal murine fetal hepatic hematopoiesis.
Cell
1991
, vol. 
65
 
4
(pg. 
677
-
689
)
52
Jiang
 
J
Best
 
S
Menzel
 
S
, et al. 
cMYB is involved in the regulation of fetal hemoglobin production in adults.
Blood
2006
, vol. 
108
 
3
(pg. 
1077
-
1083
)
53
Sankaran
 
VG
Menne
 
TF
Scepanovic
 
D
, et al. 
MicroRNA-15a and -16-1 act via MYB to elevate fetal hemoglobin expression in human trisomy 13.
Proc Natl Acad Sci U S A
2011
, vol. 
108
 
4
(pg. 
1519
-
1524
)
54
Bianchi
 
E
Zini
 
R
Salati
 
S
, et al. 
c-myb supports erythropoiesis through the transactivation of KLF1 and LMO2 expression.
Blood
2010
, vol. 
116
 
22
(pg. 
e99
-
e110
)
55
Aerbajinai
 
W
Zhu
 
J
Kumkhaek
 
C
Chin
 
K
Rodgers
 
GP
SCF induces gamma-globin gene expression by regulating downstream transcription factor COUP-TFII.
Blood
2009
, vol. 
114
 
1
(pg. 
187
-
194
)
56
van Dijk
 
TB
Gillemans
 
N
Pourfarzad
 
F
, et al. 
Fetal globin expression is regulated by friend of Prmt1.
Blood
2010
, vol. 
116
 
20
(pg. 
4349
-
4352
)
57
Lopez
 
RA
Schoetz
 
S
DeAngelis
 
K
O'Neill
 
D
Bank
 
A
Multiple hematopoietic defects and delayed globin switching in ikaros null mice.
Proc Natl Acad Sci U S A
2002
, vol. 
99
 
2
(pg. 
602
-
607
)
58
Rupon
 
JW
Wang
 
SZ
Gaensler
 
K
Lloyd
 
J
Ginder
 
GD
Methyl binding domain protein 2 mediates gamma-globin gene silencing in adult human betaYAC transgenic mice.
Proc Natl Acad Sci U S A
2006
, vol. 
103
 
17
(pg. 
6617
-
6622
)
59
Zhou
 
W
Zhao
 
Q
Sutton
 
R
, et al. 
The role of p22 NF-E4 in human globin gene switching.
J Biol Chem
2004
, vol. 
279
 
25
(pg. 
26227
-
26232
)
60
Macari
 
ER
Lowrey
 
CH
Induction of human fetal hemoglobin via the NRF2 antioxidant response signaling pathway.
Blood
2011
, vol. 
117
 
22
(pg. 
5987
-
5997
)
61
Tanabe
 
O
McPhee
 
D
Kobayashi
 
S
, et al. 
Embryonic and fetal beta-globin gene repression by the orphan nuclear receptors, TR2 and TR4.
EMBO J
2007
, vol. 
26
 
9
(pg. 
2295
-
2306
)
62
Forsberg
 
EC
Downs
 
KM
Christensen
 
HM
Im
 
H
Nuzzi
 
PA
Bresnick
 
EH
Developmentally dynamic histone acetylation pattern of a tissue-specific chromatin domain.
Proc Natl Acad Sci U S A
2000
, vol. 
97
 
26
(pg. 
14494
-
14499
)
63
Yin
 
W
Barkess
 
G
Fang
 
X
, et al. 
Histone acetylation at the human beta-globin locus changes with developmental age.
Blood
2007
, vol. 
110
 
12
(pg. 
4101
-
4107
)
64
Perrine
 
SP
Castaneda
 
SA
Chui
 
DH
, et al. 
Fetal globin gene inducers: novel agents and new potential.
Ann N Y Acad Sci
2010
, vol. 
1202
 (pg. 
158
-
164
)
65
Cao
 
H
Stamatoyannopoulos
 
G
Jung
 
M
Induction of human gamma globin gene expression by histone deacetylase inhibitors.
Blood
2004
, vol. 
103
 
2
(pg. 
701
-
709
)
66
Bradner
 
JE
Mak
 
R
Tanguturi
 
SK
, et al. 
Chemical genetic strategy identifies histone deacetylase 1 (HDAC1) and HDAC2 as therapeutic targets in sickle cell disease.
Proc Natl Acad Sci U S A
2010
, vol. 
107
 
28
(pg. 
12617
-
12622
)
67
Zhao
 
Q
Rank
 
G
Tan
 
YT
, et al. 
PRMT5-mediated methylation of histone H4R3 recruits DNMT3A, coupling histone and DNA methylation in gene silencing.
Nat Struct Mol Biol
2009
, vol. 
16
 
3
(pg. 
304
-
311
)
68
Rank
 
G
Cerruti
 
L
Simpson
 
RJ
Moritz
 
RL
Jane
 
SM
Zhao
 
Q
Identification of a PRMT5-dependent repressor complex linked to silencing of human fetal globin gene expression.
Blood
2010
, vol. 
116
 
9
(pg. 
1585
-
1592
)
69
van der Ploeg
 
LH
Flavell
 
RA
DNA methylation in the human gamma delta beta-globin locus in erythroid and nonerythroid tissues.
Cell
1980
, vol. 
19
 
4
(pg. 
947
-
958
)
70
Goren
 
A
Simchen
 
G
Fibach
 
E
, et al. 
Fine tuning of globin gene expression by DNA methylation.
PLoS One
2006
, vol. 
1
 pg. 
e46
 
71
Cioe
 
L
McNab
 
A
Hubbell
 
HR
Meo
 
P
Curtis
 
P
Rovera
 
G
Differential expression of the globin genes in human leukemia K562(S) cells induced to differentiate by hemin or butyric acid.
Cancer Res
1981
, vol. 
41
 
1
(pg. 
237
-
243
)
72
Zein
 
S
Li
 
W
Ramakrishnan
 
V
, et al. 
Identification of fetal hemoglobin-inducing agents using the human leukemia KU812 cell line.
Exp Biol Med (Maywood)
2010
, vol. 
235
 
11
(pg. 
1385
-
1394
)
73
Ma
 
F
Ebihara
 
Y
Umeda
 
K
, et al. 
Generation of functional erythrocytes from human embryonic stem cell-derived definitive hematopoiesis.
Proc Natl Acad Sci U S A
2008
, vol. 
105
 
35
(pg. 
13087
-
13092
)
74
Lapillonne
 
H
Kobari
 
L
Mazurier
 
C
, et al. 
Red blood cell generation from human induced pluripotent stem cells: Perspectives for transfusion medicine.
Haematologica
2010
, vol. 
95
 
10
(pg. 
1651
-
1659
)
75
Zou
 
J
Mali
 
P
Huang
 
X
Dowey
 
SN
Cheng
 
L
Site-specific gene correction of a point mutation in human iPS cells derived from an adult patient with sickle cell disease.
Blood
2011
, vol. 
118
 
17
(pg. 
4599
-
4608
)
76
Papapetrou
 
EP
Lee
 
G
Malani
 
N
, et al. 
Genomic safe harbors permit high beta-globin transgene expression in thalassemia induced pluripotent stem cells.
Nat Biotechnol
2011
, vol. 
29
 
1
(pg. 
73
-
78
)
77
Chang
 
KH
Huang
 
A
Hirata
 
RK
Wang
 
PR
Russell
 
DW
Papayannopoulou
 
T
Globin phenotype of erythroid cells derived from human induced pluripotent stem cells.
Blood
2010
, vol. 
115
 
12
(pg. 
2553
-
2554
)
78
Fibach
 
E
Rachmilewitz
 
EA
The two-step liquid culture: A novel procedure for studying maturation of human normal and pathological erythroid precursors.
Stem Cells
1993
, vol. 
11
 
Suppl 1
(pg. 
36
-
41
)
79
Giarratana
 
MC
Kobari
 
L
Lapillonne
 
H
, et al. 
Ex vivo generation of fully mature human red blood cells from hematopoietic stem cells.
Nat Biotechnol
2005
, vol. 
23
 
1
(pg. 
69
-
74
)
80
van den Akker
 
E
Satchwell
 
TJ
Pellegrin
 
S
Daniels
 
G
Toye
 
AM
The majority of the in vitro erythroid expansion potential resides in CD34(-) cells, outweighing the contribution of CD34(+) cells and significantly increasing the erythroblast yield from peripheral blood samples.
Haematologica
2010
, vol. 
95
 
9
(pg. 
1594
-
1598
)
81
Giarratana
 
MC
Rouard
 
H
Dumont
 
A
, et al. 
Proof of principle for transfusion of in vitro-generated red blood cells.
Blood
2011
, vol. 
118
 
19
(pg. 
5071
-
5079
)
82
Li
 
B
Ding
 
L
Li
 
W
Story
 
MD
Pace
 
BS
Characterization of the transcriptome profiles related to globin gene switching during in vitro erythroid maturation.
BMC Genomics
2012
, vol. 
13
 
1
pg. 
153
 
83
Sripichai
 
O
Kiefer
 
CM
Bhanu
 
NV
, et al. 
Cytokine-mediated increases in fetal hemoglobin are associated with globin gene histone modification and transcription factor reprogramming.
Blood
2009
, vol. 
114
 
11
(pg. 
2299
-
2306
)
84
Chan
 
KS
Xu
 
J
Wardan
 
H
McColl
 
B
Orkin
 
S
Vadolas
 
J
Generation of a genomic reporter assay system for analysis of gamma- and beta-globin gene regulation.
FASEB J
2012
, vol. 
26
 
4
(pg. 
1736
-
1744
)
85
Blau
 
CA
Barbas
 
CF
Bomhoff
 
AL
, et al. 
{Gamma}-globin gene expression in chemical inducer of dimerization (CID)-dependent multipotential cells established from human {beta}-globin locus yeast artificial chromosome ({beta}-YAC) transgenic mice.
J Biol Chem
2005
, vol. 
280
 
44
(pg. 
36642
-
36647
)
86
Hardison
 
R
Hemoglobins from bacteria to man: evolution of different patterns of gene expression.
J Exp Biol
1998
, vol. 
201
 
Pt 8
(pg. 
1099
-
1117
)
87
Peterson
 
KR
Clegg
 
CH
Huxley
 
C
, et al. 
Transgenic mice containing a 248-kb yeast artificial chromosome carrying the human beta-globin locus display proper developmental control of human globin genes.
Proc Natl Acad Sci U S A
1993
, vol. 
90
 
16
(pg. 
7593
-
7597
)
88
Porcu
 
S
Kitamura
 
M
Witkowska
 
E
, et al. 
The human beta globin locus introduced by YAC transfer exhibits a specific and reproducible pattern of developmental regulation in transgenic mice.
Blood
1997
, vol. 
90
 
11
(pg. 
4602
-
4609
)
89
Xu
 
J
Peng
 
C
Sankaran
 
VG
, et al. 
Correction of sickle cell disease in adult mice by interference with fetal hemoglobin silencing.
Science
2011
, vol. 
334
 
6058
(pg. 
993
-
996
)
90
Pászty
 
C
Brion
 
CM
Manci
 
E
, et al. 
Transgenic knockout mice with exclusively human sickle hemoglobin and sickle cell disease.
Science
1997
, vol. 
278
 
5339
(pg. 
876
-
878
)
91
Kingsley
 
PD
Malik
 
J
Emerson
 
RL
, et al. 
“Maturational” globin switching in primary primitive erythroid cells.
Blood
2006
, vol. 
107
 
4
(pg. 
1665
-
1672
)
92
McGrath
 
KE
Frame
 
JM
Fromm
 
GJ
, et al. 
A transient definitive erythroid lineage with unique regulation of the beta-globin locus in the mammalian embryo.
Blood
2011
, vol. 
117
 
17
(pg. 
4600
-
4608
)
93
Chada
 
K
Magram
 
J
Costantini
 
F
An embryonic pattern of expression of a human fetal globin gene in transgenic mice.
Nature
1986
, vol. 
319
 
6055
(pg. 
685
-
689
)
94
Papayannopoulou
 
T
Kalmantis
 
T
Stamatoyannopoulos
 
G
Cellular regulation of hemoglobin switching: Evidence for inverse relationship between fetal hemoglobin synthesis and degree of maturity of human erythroid cells.
Proc Natl Acad Sci U S A
1979
, vol. 
76
 
12
(pg. 
6420
-
6424
)
95
Ganis
 
JJ
Hsia
 
N
Trompouki
 
E
, et al. 
Zebrafish globin switching occurs in two developmental stages and is controlled by the LCR.
Dev Biol
2012
, vol. 
366
 
2
(pg. 
185
-
194
)
96
Jing
 
L
Zon
 
LI
Zebrafish as a model for normal and malignant hematopoiesis.
Dis Model Mech
2011
, vol. 
4
 
4
(pg. 
433
-
438
)
97
Lavelle
 
D
DeSimone
 
J
Heller
 
P
Zwiers
 
D
Hall
 
L
On the mechanism of hb F elevations in the baboon by erythropoietic stress and pharmacologic manipulation.
Blood
1986
, vol. 
67
 
4
(pg. 
1083
-
1089
)
98
Chin
 
J
Singh
 
M
Banzon
 
V
, et al. 
Transcriptional activation of the gamma-globin gene in baboons treated with decitabine and in cultured erythroid progenitor cells involves different mechanisms.
Exp Hematol
2009
, vol. 
37
 
10
(pg. 
1131
-
1142
)
99
Papayannopoulou
 
T
Kurachi
 
S
Brice
 
M
Nakamoto
 
B
Stamatoyannopoulos
 
G
Asynchronous synthesis of HbF and HbA during erythroblast maturation. II. studies of G gamma, A gamma, and beta chain synthesis in individual erythroid clones from neonatal and adult BFU-E cultures.
Blood
1981
, vol. 
57
 
3
(pg. 
531
-
536
)
100
Perkins
 
AC
Peterson
 
KR
Stamatoyannopoulos
 
G
Witkowska
 
HE
Orkin
 
SH
Fetal expression of a human agamma globin transgene rescues globin chain imbalance but not hemolysis in EKLF null mouse embryos.
Blood
2000
, vol. 
95
 
5
(pg. 
1827
-
1833
)
101
Liu
 
P
Keller
 
JR
Ortiz
 
M
, et al. 
Bcl11a is essential for normal lymphoid development.
Nat Immunol
2003
, vol. 
4
 
6
(pg. 
525
-
532
)
102
Leid
 
M
Ishmael
 
JE
Avram
 
D
Shepherd
 
D
Fraulob
 
V
Dolle
 
P
CTIP1 and CTIP2 are differentially expressed during mouse embryogenesis.
Gene Expr Patterns
2004
, vol. 
4
 
6
(pg. 
733
-
739
)
103
Kuo
 
TY
Chen
 
CY
Hsueh
 
YP
Bcl11A/CTIP1 mediates the effect of the glutamate receptor on axon branching and dendrite outgrowth.
J Neurochem
2010
, vol. 
114
 
5
(pg. 
1381
-
1392
)
104
John
 
A
Brylka
 
H
Wiegreffe
 
C
, et al. 
Bcl11a is required for neuronal morphogenesis and sensory circuit formation in dorsal spinal cord development.
Development
2012
, vol. 
139
 
10
(pg. 
1831
-
1841
)
105
Zeggini
 
E
Scott
 
LJ
Saxena
 
R
, et al. 
Meta-analysis of genome-wide association data and large-scale replication identifies additional susceptibility loci for type 2 diabetes.
Nat Genet
2008
, vol. 
40
 
5
(pg. 
638
-
645
)
106
Langberg
 
KA
Ma
 
L
Sharma
 
NK
, et al. 
Single nucleotide polymorphisms in JAZF1 and BCL11A gene are nominally associated with type 2 diabetes in african-american families from the GENNID study.
J Hum Genet
2012
, vol. 
57
 
1
(pg. 
57
-
61
)
107
Koehler
 
AN
A complex task? Direct modulation of transcription factors with small molecules.
Curr Opin Chem Biol
2010
, vol. 
14
 
3
(pg. 
331
-
340
)
108
Filippakopoulos
 
P
Qi
 
J
Picaud
 
S
, et al. 
Selective inhibition of BET bromodomains.
Nature
2010
, vol. 
468
 
7327
(pg. 
1067
-
1073
)
109
Nicodeme
 
E
Jeffrey
 
KL
Schaefer
 
U
, et al. 
Suppression of inflammation by a synthetic histone mimic.
Nature
2010
, vol. 
468
 
7327
(pg. 
1119
-
1123
)
110
Dawson
 
MA
Prinjha
 
RK
Dittmann
 
A
, et al. 
Inhibition of BET recruitment to chromatin as an effective treatment for MLL-fusion leukaemia.
Nature
2011
, vol. 
478
 
7370
(pg. 
529
-
533
)
111
Saunthararajah
 
Y
Hillery
 
CA
Lavelle
 
D
, et al. 
Effects of 5-aza-2′-deoxycytidine on fetal hemoglobin levels, red cell adhesion, and hematopoietic differentiation in patients with sickle cell disease.
Blood
2003
, vol. 
102
 
12
(pg. 
3865
-
3870
)
112
Olivieri
 
NF
Saunthararajah
 
Y
Thayalasuthan
 
V
, et al. 
A pilot study of subcutaneous decitabine in beta-thalassemia intermedia.
Blood
2011
, vol. 
118
 
10
(pg. 
2708
-
2711
)
113
Moutouh-de Parseval
 
LA
Verhelle
 
D
Glezer
 
E
, et al. 
Pomalidomide and lenalidomide regulate erythropoiesis and fetal hemoglobin production in human CD34+ cells.
J Clin Invest
2008
, vol. 
118
 
1
(pg. 
248
-
258
)
114
Meiler
 
SE
Wade
 
M
Kutlar
 
F
, et al. 
Pomalidomide augments fetal hemoglobin production without the myelosuppressive effects of hydroxyurea in transgenic sickle cell mice.
Blood
2011
, vol. 
118
 
4
(pg. 
1109
-
1112
)
115
Ito
 
T
Ando
 
H
Suzuki
 
T
, et al. 
Identification of a primary target of thalidomide teratogenicity.
Science
2010
, vol. 
327
 
5971
(pg. 
1345
-
1350
)
116
Moellering
 
RE
Cornejo
 
M
Davis
 
TN
, et al. 
Direct inhibition of the NOTCH transcription factor complex.
Nature
2009
, vol. 
462
 
7270
(pg. 
182
-
188
)
117
Lanford
 
RE
Hildebrandt-Eriksen
 
ES
Petri
 
A
, et al. 
Therapeutic silencing of microRNA-122 in primates with chronic hepatitis C virus infection.
Science
2010
, vol. 
327
 
5962
(pg. 
198
-
201
)
118
Wilber
 
A
Hargrove
 
PW
Kim
 
YS
, et al. 
Therapeutic levels of fetal hemoglobin in erythroid progeny of beta-thalassemic CD34+ cells after lentiviral vector-mediated gene transfer.
Blood
2011
, vol. 
117
 
10
(pg. 
2817
-
2826
)
119
Cheng
 
LT
Sun
 
LT
Tada
 
T
Genome editing in induced pluripotent stem cells.
Genes Cells
2012
, vol. 
17
 
6
(pg. 
431
-
438
)

Author notes

*

D.E.B. and S.C.K. contributed equally to this work.

Sign in via your Institution