Next Article in Journal
Quercetin-Rich Extracts from Onions (Allium cepa) Play Potent Cytotoxicity on Adrenocortical Carcinoma Cell Lines, and Quercetin Induces Important Anticancer Properties
Next Article in Special Issue
Novel Therapies to Address Unmet Needs in ITP
Previous Article in Journal
Application of In Silico Filtering and Isothermal Titration Calorimetry for the Discovery of Small Molecule Inhibitors of MDM2
Previous Article in Special Issue
Target Therapies for Systemic Mastocytosis: An Update
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Pharmacological Induction of Fetal Hemoglobin in β-Thalassemia and Sickle Cell Disease: An Updated Perspective

by
Rayan Bou-Fakhredin
1,
Lucia De Franceschi
2,
Irene Motta
1,3,
Maria Domenica Cappellini
1,3,* and
Ali T. Taher
4,*
1
Department of Clinical Sciences and Community Health, University of Milan, 20122 Milan, Italy
2
Department of Medicine, University of Verona and Azienda Ospedaliera Universitaria Verona, 37128 Verona, Italy
3
UOC General Medicine, Fondazione IRCCS Ca’ Granda Ospedale Maggiore Policlinico, 20122 Milan, Italy
4
Department of Internal Medicine, Division of Hematology-Oncology, American University of Beirut Medical Center, Beirut 1107 2020, Lebanon
*
Authors to whom correspondence should be addressed.
Pharmaceuticals 2022, 15(6), 753; https://doi.org/10.3390/ph15060753
Submission received: 27 April 2022 / Revised: 12 June 2022 / Accepted: 13 June 2022 / Published: 16 June 2022
(This article belongs to the Special Issue Drug Design and Development for Rare Hematologic Diseases)

Abstract

:
A significant amount of attention has recently been devoted to the mechanisms involved in hemoglobin (Hb) switching, as it has previously been established that the induction of fetal hemoglobin (HbF) production in significant amounts can reduce the severity of the clinical course in diseases such as β-thalassemia and sickle cell disease (SCD). While the induction of HbF using lentiviral and genome-editing strategies has been made possible, they present limitations. Meanwhile, progress in the use of pharmacologic agents for HbF induction and the identification of novel HbF-inducing strategies has been made possible as a result of a better understanding of γ-globin regulation. In this review, we will provide an update on all current pharmacological inducer agents of HbF in β-thalassemia and SCD in addition to the ongoing research into other novel, and potentially therapeutic, HbF-inducing agents.

1. Introduction

Hemoglobinopathies are among the most common inherited monogenic disorders in the world and they exhibit a large range of clinical phenotypes with varying disease severity [1,2,3]. The levels of fetal hemoglobin (HbF) in erythrocytes make up a large part of the clinical heterogeneity that is observed in patients with β-thalassemia and sickle cell disease (SCD).
HbF is the predominant form of Hb expressed throughout gestation and has a crucial role in facilitating transplacental gas exchange [4,5,6]. Composed of two α-globin chains and two γ-globin chains, HbF has a greater affinity for oxygen when compared to hemoglobin A (HbA) [4,7,8]. After birth, the predominant form of Hb produced switches from HbF to HbA, which has a lower affinity for oxygen [4,5,6,8,9,10]. This switch is genetically regulated by transcription factors such as BCL11A, KLF-1 and MYB [4,11] and results in adults typically expressing low levels of HbF [4,12,13]. However, as expression switches from HbF to HbA, individuals who have mutations in the β-globin gene end in either deficiency in globin chain synthesis or production of pathologic Hb such as in β-thalassemia or SCD. Fetal erythrocytes or fetal cells contain nearly 100% of HbF. Following the transcriptional switch in definitive erythroid progenitors from HbF (α2γ2) to adult hemoglobin (α2β2) around birth, very low levels of HbF are found a year after birth [14], suggesting this switch is not complete. In some individuals, higher levels of HbF persist into childhood and adulthood, reflected in a continuous positively skewed distribution of HbF across healthy adults. Adult erythrocytes that contain detectable levels of HbF are called F cells. These cells arise from erythroid precursors that can give rise to both HbF-containing and non-HbF-containing erythrocytes [15]. Hereditary persistence of fetal hemoglobin (HPHF) is defined as the heterogenous group of inherited defects in the switch from fetal to adult hemoglobin [16]. Large deletions or point mutations within the HBB gene cluster that result in higher than normal levels of HbF persisting into adulthood cause the most frequently recognized type of HPFH. Pancellular vs. heterocellular HbF distribution has been used as a defining aspect of HPFH [16]. The three main loci that control HbF levels outside of the locus control region at the 5’ end of the β-globin gene cluster and promoter regions on the β- and γ-globin genes are BCL11A, ZBTB7A and MYB [15].
In recent years, progress on the knowledge of molecular mechanisms involved in Hb switching, the concept of HPHF, combined with epidemiological and clinical observations has provided important evidence on the beneficial role of increasing HbF levels in ameliorating the clinical complications of β-thalassemia and SCD [17]. This can be achieved by either classic drug-modulating approaches or by gene therapy and genome-editing approaches [18,19,20,21,22,23,24]. Although encouraging and promising results have been reported on gene therapy/genome editing and HbF expression, their safety is still to be determined and their expected costs are to be considered as well [25,26,27]. Accordingly, pharmacological induction of HbF production is still an interesting option to decrease the severity of these disorders. In this review, we will provide an update on all current pharmacological inducer agents of HbF in β-thalassemia and SCD in addition to the ongoing research into other novel and potentially curative HbF-inducing agents (Figure 1).

2. Established Pharmacologic Approaches to HbF Induction in Hemoglobinopathies

Different pharmacological agents have been pre-clinically and clinically tested at various stages of clinical trials for their ability to increase γ-globin expression. In this section, we describe these already established and major pharmacological approaches and agents for HbF induction. These are also summarized in Figure 1.

2.1. Chemotherapeutic Agents

2.1.1. Hydroxyurea

The primary mechanism of action of hydroxyurea (HU) in β-thalassemia and SCD is the upregulation of the γ-globin gene expression in erythroid cells [28]. Hydroxyurea enhanced the percentage of F cells in the circulations of patients with hemoglobinopathies, with a heterogenous distribution within red cells, leading to overall improvements in the hematologic phenotype [29,30,31,32,33]. While the clinical efficacy of HU is primarily due to its ability to induce HbF, its exact mechanism of action remains not fully understood [34].
Studies conducted in β-thalassemia patients showed that HU induced a 2-to-9-fold increase in γ-mRNA with a good correlation between γ-mRNA and HbF levels [35,36,37,38]. Responses in these patients were observed at HU doses ranging between 10 and 20 mg/kg/day. In addition to its known effects in stimulating γ-globin production, the use of HU was also associated with a decrease in transfusion requirement and an improvement in the prothrombotic profile, especially in splenectomized patients [39,40]. Collectively, these studies present several limitations such as the small sample size, the definition of transfusion dependence, the heterogeneity of the β-thalassemic patient populations and the lack of control groups. Thus, we might conclude that these data generate a rationale to design new studies to assess the possible beneficial effects of HU in β-thalassemia patients, particularly in those with non-transfusion-dependent thalassemia.
Improvements in HbF levels after HU treatment have been shown in patients with SCD, HbSC disease, HbS-β0 and HbS-β+ thalassemia [40,41,42,43]. In SCD, two different mechanisms exist for HU-induced increases in HbF synthesis: (i) inhibition of ribonucleotide reductase, promoting the selection of high HbF expression erythroid precursors and (ii) the direct selection of HbF cell production by inhibiting soluble guanylate cyclase [44,45,46]. Several studies on SCD have proposed that curative therapies with HU should aim to achieve HbF >30%, F cells >70% and >4–10 pg F/F cell [47,48,49]. Other studies demonstrated that the early initiation of HU using individualized and pharmacokinetics-guided dosing can lead to robust and sustained HbF levels beyond 30–40% in most SCD patients who are adherent to therapy [50,51]. Hydroxyurea has also been associated with improved mortality and morbidity in both adults and children with SCD [52]. Indeed, HU significantly reduced hospitalization rate, acute chest syndrome, VOCs and transfusion requirements [53]. Few studies have shown a possible declining effect of HU on HbF synthesis over time [54]. The beneficial effects of HU are noteworthy and go beyond HbF synthesis in SCD. Hydroxyurea also acts as a multimodal agent targeting neutrophils, modulating inflammatory response and vascular endothelial activation, contributing to nitric oxide biosynthesis [55,56,57,58].

2.1.2. DNA Methyltransferase Inhibitors

DNA methyltransferase (DNMT) inhibitors, such as 5-azacytidine and decitabine, can reactivate γ-globin gene expression via DNA hypomethylation [59]. Initially, 5-azacytidine was studied and used in SCD but it was quickly abandoned due to its toxicity profile and possible carcinogenicity risk [60]. Decitabine, an analogue of 5-azacytidine, is also a potent DNMT1 inhibitor with a more favorable safety profile. A pilot study by Oliviera et al. showed that decitabine administered subcutaneously at a dose of 0.2 mg/kg twice per week for 12 weeks increased total hemoglobin (Hb) from 78.8 to 90.4 g/L, and increased absolute HbF levels from 36.4 to 42.9 g/L in 5 β-TI patients [61]. An improvement in RBC indices was also noted. Treatment was well tolerated overall, with the main adverse event being an elevation in platelet counts [61]. Small studies in SCD have also suggested that decitabine can substantially increase HbF and total Hb levels [62,63,64,65]. However, if taken orally, decitabine is rapidly deaminated and inactivated by cytosine deaminase. To overcome this challenge, one clinical trial used decitabine in combination with tetrahydrouridine (THU), a cytosine deaminase inhibitor, to induce HbF production (NCT01685515). A phase I study showed that the combination of decitabine and THU led to the persistent inhibition of the DNMT1 protein with induction and increase in HbF levels, and more importantly, HbF-enriched red cells (F cells) increased to 80%. While these agents do not have major myelotoxic effects, they might induce thrombocytosis. This might be taken into consideration in patients with SCD that is per se a thrombophilic state [66].
Recently, an innovative and orally bioavailable DNMT1-selective inhibitor known as GSK3482364 has emerged [67]. In contrast with the cytidine analog DNMT1 inhibitors, the inhibitory mechanism of GSK3482364 does not require DNA incorporation and is reversible. In cells, treatment with GSK3482364 caused DNA hypomethylation, and resulted in HbF induction. These effects were approximately equivalent and comparable to decitabine treatment [67]. In an in vitro model of erythropoiesis, GSK3482364 and decitabine led to comparable increases in HbF cells. However, treatment with GSK3482364 resulted in a larger proportion of cells maturing into HbF-expressing reticulocytes [67]. The effects of GSK3482364 on the bone marrow of transgenic SCD mice were also studied and showed a clear induction of F cells that exceeded the corresponding effects of decitabine at tolerated doses over a 12-day period [67]. The compound was well-tolerated by SCD mice and there was no evidence of adverse hematological effects [67]. Notably, and unlike what has been reported and observed with decitabine, the use of GSK3482364 was not associated with significant increases in platelet count. Taken together, these data suggest that GSK3482364 is a promising molecule to further study the role of DNMT1 inhibitors in both in vitro and in vivo models of SCD.

2.2. Histone Deacetylase Inhibitors

Histone deacetylase (HDAC) inhibitors have also been considered as therapeutic targets for HbF induction. This group of regulatory molecules is involved in the epigenetic silencing of the γ-globin genes [34,68]. The earliest HDAC inhibitor investigated and used as a HbF inducer was butyrate. Butyrate increased the transcription rate of the HBG genes, as well as the translation of HBG1/HBG2 mRNA [69]. In clinical trials conducted on patients with β-thalassemia and SCD, intravenous administration of arginine butyrate and oral administration of sodium phenylbutyrate increased HbF levels. However, these oral butyrate compounds were only successful in producing a very mild raise in HbF levels and the overall patient compliance was poor [70]. Panobinostat is a pan-HDAC inhibitor that has been reported to have a greater potency than sodium butyrate [71]. It is currently being tested in a phase I clinical trial on adult SCD patients (NCT01245179).
Vorinostat (also known as suberoylanilide hydroxamic acid or SAHA) is a hydroxamic acid group pan-HDAC inhibitor that has been used in many research studies including cancer treatment [72]. In 2019, Mettananda et al. demonstrated that in human erythroid cells, vorinostat downregulates α-globin expression while inducing γ-globin expression and its use can thus be considered as a potential therapy for β-thalassemia [73]. As a potent inducer of γ-globin and HbF, vorinostat has also been considered as a potential therapy for SCD [74]. Thus, it was postulated that vorinostat may help in the treatment SCD by increasing the amount of HbF in the blood. One 2010 study by Hebbel et al. showed that the HDAC inhibitors trichostatin A and vorinostat were beneficial for the vascular pathobiology of sickle transgenic mice and led to an increase in HbF levels [75]. A phase 2 trial assessing the safety and efficacy of vorinostat (once a day for three consecutive days every week for 12–16 weeks) in adult SCD patients resistant to HU was also initiated (NCT01000155). However, the study was eventually terminated early as recruitment was poor. A novel class I-restricted HDAC inhibitor and largazole analog known as CT-101 has been recently evaluated for its pharmacodynamics, cytotoxicity and targeted epigenetic effects in sickle erythroid precursors [76]. Results demonstrated that CT-101 was successfully able to activate γ-globin transcription selectively and increase F cells and HbF levels [76]. Moreover, the combination of CT-101 with HU produced a better effect on HbF levels. CT-101 produced very limited cell toxicity after 5 days of treatment as shown by slightly reduced sickle erythroid maturation. Despite this, cell numbers continued to increase [76]. CT-101 also increased acetylated histone H3 and H4 levels and conferred an open chromatin conformation in the γ-globin promoter [76]. Thus, CT-101 may be a possible lead candidate to be further developed and transferred to clinical studies as a pharmacologic inducer of HbF.

EHMT1/2 Inhibitors

Other pharmacological agents targeting histone methyl transferase to induce HbF expression include euchromatic histone-lysine-N-methyltransferases 1 and 2 (EHMT1/2) inhibitors. In vitro, UNC0638 has been shown to induce γ-globin mRNA and HbF expression in normal erythroid precursors [77,78,79]. This was associated with a decreased accumulation of H3K9me2 near the γ-globin loci and with increased loop formation between the locus control region and the γ-globin promoters through the recruitment of the LDB1 complex to the γ globin promoters [77,78,79]. In a recent study conducted on erythroid precursors derived from CD34+ cells of β0-thalassemia/HbE patients, the HbF-inducing activity of UNC0638, either alone or in combination with pomalidomide and decitabine, was investigated [80]. UNC0638 was able to increase in γ-globin mRNA levels, HbF levels and F cells. This HbF induction was 25.5 ± 4.2% above baseline levels, with a more pronounced effect when added at an early stage (day 4) of erythropoiesis [80]. It is noteworthy that UNC0638 exhibited an HbF-inducing activity similar to pomalidomide, but higher than decitabine under the same culture conditions [80]. UNC0638 shows a synergic effect on HbF synthesis when used in combination with either pomalidomide or decitabine. Although UNC0638 showed a strong HbF-inducing activity in ex vivo erythroid cell culture systems, its in vivo pharmacokinetic properties were shown to be due to a lack of drug-like properties [81]. In the future, EHMT1/2 inhibitors should undergo further development to achieve a higher potency and improved in vivo pharmacokinetic properties.

2.3. Immunomodulators: Thalidomide and Its Derivatives

Thalidomide is an immunomodulatory compound used in clinical practice for the clinical management of multiple myeloma [82,83]. Thalidomide and its derivate pomalidomide have been considered as new therapeutic options for β-thalassemia. In vitro models of pathologic erythropoiesis have established that thalidomide might induce γ-globin mRNA expression in a dose-dependent manner throughout the modulation of γ-globin mRNA expression targeting BCL11A, SOX6, GATA1 and KLF1 as well as by post-translational modification induced by p38MAPK activation [82,83,84]. Several small case series and few observational studies have reported the efficacy and safety of thalidomide in patients with β-thalassemia [85,86,87,88,89,90,91,92]. Collectively, these studies show that thalidomide (range dosage from 75–100 mg/kg/day to 150–200 mg/kg/day) increases Hb levels by elevating the HbF level and reduces spleen size [93,94,95,96,97]. Furthermore, the efficacy of thalidomide on hematologic phenotype of patients with β-thalassemia was not inferior of HU [93]. Since large doses of thalidomide can be detrimental, the lowest effective dose of 50 mg/day was proposed to improve anemia in β-thalassemic patients [93,94,98,99]. In the future, larger randomized controlled trials are needed to establish the efficacy of thalidomide in patients with β-thalassemia.
Pomalidomide, a third-generation immunomodulatory drug and thalidomide derivative with less adverse events (AEs), has been shown to be an effective and potent HbF inducer in in vitro β-thalassemia/HbE erythropoiesis [94]. Its use has also been documented to increase HbF levels in SCD. Similar to HU, pomalidomide increased the level of HbF production without myelosuppressive effects in a humanized mouse model of SCD [100]. In addition, treatment of human hematopoietic stem cells (HSCs) with pomalidomide and lenalidomide significantly stimulated the proliferation of HSCs and induced HbF by modulating the transcription of HBB and HBG gene through the downregulation of BCL11A IKZF1, KLF1, LSD1 and SOX-6 repression factors [82,101]. A phase 1 trial was conducted to assess the efficacy, safety and maximum tolerated dose of pomalidomide (0.5–4 mg/day for 84 days) in SCD patients (NCT01522547). Two out of four patients treated with pomalidomide at 4 mg/day showed an increase in HbF levels [15]. Lately, pomalidomide–nitric oxide donor derivatives (3a–f) have been synthesized and their suitability as novel and potential HbF inducers has also been evaluated and successfully shown in an early preliminary study [70]. This could pave the way in the future for further investigation to use these pomalidomide derivatives in the treatment of SCD.

2.4. cGMP Modulators: Phosphodiesterase-9 Inhibitor

IMR-687 (Tovinontrine) is a highly selective phosphodiesterase 9 inhibitor that is currently being developed as an orally administered therapy for β-thalassemia and SCD patients. Similar to HU, IMR-687 increases intracellular cGMP levels and its use in preclinical studies showed an increase in HbF expression [102]. In a phase 2a study conducted on adult patients with SCD (N = 93) (NCT03401112), IMR-687 was given at a dose of 50–200 mg once daily (N = 63) for up to 6 months. IMR-687 was generally well-tolerated when used alone or in combination HU and the data showed a mean absolute change of +1.9 and +7.3 in HbF (%) and F cells (%) at 4 months, respectively, with minimal changes in Hb [103]. A phase 2a open-label extension study (NCT04053803) is ongoing to assess the long-term efficacy and safety of IMR-687 in SCD patients for up to 4 years at a dose of 200 mg once daily [104]. Seventeen patients with SCD were treated with IMR-687 monotherapy and seven were treated with a combination of IMR-687 and HU. Preliminary results from the patients that had an evaluable pharmacodynamic data at 8 months showed that seven (47%) patients had a ≥6% absolute increase in F cells and four patients (36%) had a ≥3% absolute increase in HbF [104]. With these encouraging data, a phase 2b study in SCD patients is currently ongoing to further evaluate the efficacy of IMR-687 as a HbF-inducing agent at higher doses (NCT04474314).
A summary of the mechanism of action of these established pharmacologic approaches to HbF induction has been included in Table 1.

3. Novel Experimental Strategies for the Pharmacologic Induction of HbF in Hemoglobinopathies

Progress on the knowledge of molecular mechanisms involved in the modulation of HbF synthesis has paved the way for the identification of new potential therapeutic targets, which are currently under evaluation at the pre-clinical stage to then possibly be transferable to clinical studies in patients with in hemoglobinopathies (Figure 1).

3.1. FTX-6058: Embryonic Ectoderm Development (EED) Inhibitor

FTX-6058 is a small and potent molecule that acts as an inhibitor of embryonic ectoderm development (EED), which is part of the polycomb repressive complex 2 (PRC2) that is involved in the repression of gene transcription by histone 3 methylation. In both in vitro and in vivo models of SCD, FTX-6058 induced the upregulation of HbF expression (up to 40% total Hb) [105]. The safety and efficacy of single- and multiple-ascending doses of FTX-6058 are currently being investigated in a phase-1 trial enrolling healthy adult volunteers (NCT04586985) [106]. FTX-6058 was tested at the dosage of 2 mg to 60 mg both given once daily for 14 days vs. placebo [106]. Interim results showed proportional increases in HbF levels, and increased numbers of F reticulocytes in the dose-ascending groups [106]. After 14 days, the 6 mg dose increased mRNA levels by more than three-fold and the 10 mg dose by more than four-fold. New data from the group of patients receiving a dose of 20 mg and 30 mg showed that by day 14, Hb mRNA levels had risen by a mean of 5.6 times and 6.2 times, respectively [106]. Moreover, 7 to 10 days after dosing, the percentage of F reticulocytes increased by a mean of 1.8 times and 2.4 times in the 20 mg group and 30 mg group, respectively [106]. FTX-6058 was generally well-tolerated without major AEs reported. All the above mentioned results further support the rationale to design clinical studies with FTX-6058 in patients with SCD. Indeed, a phase 1b proof-of-concept study is awaited to evaluate FTX-6058 in people with SCD [106].

3.2. Lysine-Specific Histone Demethylase 1 (LSD1) Inhibitors

With the use of RNA interference strategies and pharmacological inhibitors, lysine-specific histone demethylase 1 (LSD1) has been identified as a therapeutic target for HbF induction [107]. Tranylcypromine (TC) is a selective monoamine oxidase inhibitor of LSD1 that is currently FDA approved and used for the management of major depressive disorders [108]. Shi et al. initially reported that inhibition of LSD1 by TC enhances HbF synthesis in primary human erythroid cells, as well as in β-YAC mice [107]. However, in vitro erythropoiesis, high TC concentrations delayed erythroid maturation with a rapid decrease in total β-like globin mRNA [109].
Additionally, RN-1 is a TC analog that can act as an effective, irreversible LSD1 inhibitor with IC50 lower than TC (0.07 µM vs. 2 μM) [110]. In humanized healthy and SCD mice, LSD1 inactivation by RN-1 was shown to induce γ-globin expression and HbF synthesis [111]. In particular, treatment with RN-1 increased the expression of murine embryonic εy- and βh1-globin genes without any changes in the expression of adult β-globin [111]. When tranylcypromine (TCP) and RN-1 were tested in primate and murine erythroid cell cultures, RN-1 was shown to induce F cells and γ-globin mRNA at much greater levels than either TCP or HU [112]. More recently, the use of RN-1 was investigated in erythroid progenitor cells derived from β0-thalassemia/HbE patients [113]. At a concentration of 0.004 uM, RN-1 significantly increased the expression of γ-globin mRNA and HbF expression without any significant toxicity and regardless of HbF baseline levels [113]. Transcript levels of numerous key γ-globin repressors and co-repressors such as NCOR1, SOX6 and MYB were also modulated by RN-1 treatment [113]. Collectively, these data suggest the consideration of RN-1 for further development as an additional strategy to induce HbF in both SCD and thalassemic syndromes.

3.3. Modulators of Redox-Related Transcriptional Factors: Nrf2 or FOXO3

3.3.1. Dimethyl Fumarate

Dimethyl fumarate (DMF), a small and orally active molecule that acts as a nuclear factor erythroid derived-2-like 2 (Nrf2) agonist, is currently approved for the treatment of relapsing-remitting multiple sclerosis [114,115,116,117]. Being a transcription factor, Nrf2 triggers the cytoprotective and antioxidant pathways in response to oxidation [118,119,120,121]. A study by Krishnamoorthy et al. examined the ability of DMF to activate γ-globin transcription and enhance HbF in SCD-derived erythroid precursors cells, SCD transgenic mice and non-anemic cynomolgus monkeys [122,123]. Across all these settings, DMF (with or without HU) significantly increased the level of γ-globin mRNA, the ratio of γ/β-globin mRNA and the percentage of F cells [122]. The greatest average increase in γ/β-globin mRNA was observed in the group receiving a combination of DMF and HU [122]. This was associated with increased Gγ-globin and Aγ-globin chains synthesis [122]. In humanized SCD, DMF (100 mg/kg) enhanced Aγ-globin expression and Nrf2-dependent genes such as NQO1 and HO-1 with beneficial effects in murine hematologic sickle cell phenotype as well as an improvement in inflammatory vasculopathy [122]. These findings are in agreement with previously published studies where Nrf2 was genetically activated by decreasing expression of one its binding protein Keap1 [124,125]. Thus, DMF might be an interesting new approach to the biocomplexity of SCD by acting as multimodal therapy, not only as an HbF inducer agent but also as a modulator in inflammatory response and vascular dysfunction.

3.3.2. Metformin

Metformin has been shown to activate the redox-related transcriptional factor FOXO3 in both non-erythroid cell lines and hepatocytes [126,127,128]. Gene silencing of FOXO3 reduced γ-globin RNA expression and HbF levels in erythroblasts, whereas overexpression of FOXO3 produced the opposite effect [129]. When primary CD34+ cell-derived erythroid cultures were treated with metformin (0, 50 and 100 mM), dose-related FOXO3-dependent increases in the percentage of HbF as well as in the amounts of HbF-immunostaining cells (F cells) were observed, without any changes in BCL11A, MYB or KLF1 expression [129]. Studies in erythroid precursors from SCD patients treated with metformin alone or in combination with HU show a three-fold reduction in in vitro sickling, comparable to that observed with HU alone. It is noteworthy that the combination of metformin and HU have a synergic effect on the reduction in the percentage of sickled erythroid cells compared with monotherapy. Metformin was also evaluated in 18 patients with SCD with SS, Sβ0, Sβ+ and SC genotypes [130]. The increase in HbF was shown to be minimal in patients with SS, Sβ0 genotype [130]. Although some concerns on the use of metformin in SCD patients was raised due to the possible lactic acidosis which might accelerate HbS polymerization, no data on lactic acidosis were reported in patients with SCD [130]. Although the data on metformin are stimulating, the small number of SCD patients and the absence of data on clinical outcomes limit the conclusion of its use as a HbF inducer in clinical practice [131].

3.4. Agents Involved in Displacement/Suppression of γ-Globin Gene Promoters

This family includes a heterogenous group of molecules, such as benserazide, TN1, nethylpiperazine, acyclovir, tenofovir disoproxil fumarate and cilostazol, which are involved in the displacement/suppression of γ-globin gene promoters such as LSD-1, BCL11A and HDAC3 [132] (Figure 1).

3.4.1. Benserazide

Benserazide is a peripheral dopa decarboxylase inhibitor used in combination with L-DOPA for treatment of Parkinson’s disease. In preclinical models, benserazide was shown to induce HbF production [133,134]. Benserazide was also shown to increase fetal γ-globin gene transcription and the proportions of cells expressing HbF (F reticulocytes and F cells) in erythroid precursors from patients with HbE-β0-thalassemia and SCD [133,135]. An observational study by Santos et al. on a total of 50 individuals was conducted to evaluate the ability of benzeraside (at daily doses that ranged from 100 mg to 700 mg) to increase HbF production and circulating F cells. No correlations were found between the average daily dose of benserazide and HbF levels [136]. Moreover, no hematologic AEs related to benserazide use were recorded, even after up to 22 years of treatment [136]. Another recent study evaluated the efficacy of (R,S)-benserazide in comparison to its enantiomers to identify the best optimal form for clinical development transferable to hemoglobinopathies such as β-thalassemia or SCD [132]. Non-inferiority data on HbF expression between benserazide and its individual enantiomers were reported [132]. In addition, in β-YAC mice, the intermittent treatment with all forms of benserazide significantly increased the proportions of F cells with similar pharmacokinetic profiles in the absence of myelotoxicity [132]. Thus, the use of benserazide either alone or in combination with other HbF-inducing agents, such as HU or decitabine, might be considered another possible strategy to further increase HbF levels and proportions of F cells through complimentary mechanisms.

3.4.2. Purine-Based Fetal Hemoglobin Inducers

An early study by Nam et al. provided the first evidence on TN1 (2,6-diamino-substituted purine), a potent purine-based HbF inducer [137]. In leukemia cell lines, TN1 (100 nM) induced HbF more potently than HU [137]. However, another study conducted on human primary erythroid cells showed that this agent did not significantly increase γ-globin gene expression [138]. These drawbacks limited the development of this compound and its capability of being used in clinical trials in patients with hemoglobinopathies. Lai et al. then reported another potent and orally active purine-based HbF inducer known as Nethylpiperazine, or compound 13a [138]. In vitro assays demonstrated that in primary erythroid cells as well as in HU-resistant primary erythroid cells, compound 13a might efficiently induce γ-globin expression at a non-toxic concentration [138]. In a mouse model for SCD, compound 13a was safe and well tolerated with a dose-dependent beneficial effect on the hematologic SCD phenotype [138]. Thus, compound 13a can be considered an interesting molecule to be further developed as an inducer of HbF. It may be used in combination with HU or in HU low-responder patients with β-thalassemia and SCD.
Acyclovir (ACV) is another cyclic purine nucleotide analog, recently reported to induce HbF. Acyclovir is an FDA-approved antiviral agent against herpes simplex viruses, and its antiviral activity is highly specific [139]. Ali et al. reported the increased expression of γ-globin gene and HbF synthesis in CD34+-derived erythroid precursors. No major effects on either erythroid proliferation and maturation were observed [140]. ACV significantly downregulated the γ-globin repressors BCL11A and SOX6. This was associated with the upregulation of GATA-1 [140]. Data from β-YAC transgenic mice treated with ACX revealed a substantial increase in HbF mRNA expression as well as in the percentage of F cells [140].
Tenofovir disoproxil fumarate (TDF) is an acyclic nucleotide analogue of adenosine used in the treatment of the human immunodeficiency virus and hepatitis B infection. TDF has also been investigated as a pharmacologically active HbF inducer. A study by Khan et al. observed that TDF increased erythroid differentiation, γ- globin gene mRNA transcription and HbF expression in K562 cells [141]. In vivo studies using β-YAC transgenic mice confirmed that the beneficial effects of TDF (at a dose of 200 mg/kg/day for four weeks) was even better than HU on the percentage of HbF-positive RBCs in the absence of myelotoxic and cytotoxic effects [141].

3.4.3. Cilostazol (OPC-13013)

Cilostazol, a quinolinone derivative, is a reversible specific type 3 phosphodiesterase inhibitor. Cilostazol was shown to increase cyclic adenosine monophosphate (cAMP) cellular content, resulting in the inhibition of platelet aggregation and modulation of vascular tone [142]. Ali et al. demonstrated that cilostazol (20, 30, 40, 50 and 100 μM) induces erythroid differentiation of K562 cells in a concentration-dependent manner and significantly increases cell hemoglobinization [143]. This was associated with the upregulation of γ-globin mRNA transcripts as well as the amount of F cells [143]. In β-YAC mice, cilostazol upregulated γ-globin mRNA level and increased the percentage of F cells [143]. No adverse events secondary to the drug were observed. Although these are preliminary data, the multimodal action of cilostazol targeting vascular endothelial cells, platelets and HbF expression make this molecule an ideal candidate to further study in β-thalassemia and SCD.

3.5. Molecules Targeting Post-Translational Modifications Involved in HbF Expression

Recent reports have highlighted the novel role of post-translational modifications such as phosphorylation in the induction of HbF, most likely amplifying signaling pathways activated in response to stress erythropoiesis [144,145] (Figure 1).

3.5.1. Salubrinal (SAL)

Salubrinal is a selective inhibitor of protein phosphatase 1 (PP1) that participates in the recruitment of phosphorylated eukaryotic initiation factor 2α (p-eIF2α) which in turn activates downstream targets such as activating transcription factor 4 (ATF4) [146]. Chen et al. demonstrated that SAL ameliorates anemia in mice genetically lacking the discoidal domain receptor1 (DDRGK1F/F), a model of stress erythropoiesis. [147]. Hahn et al. on the other hand showed that in normal erythroid precursors, SAL might prevent dephosphorylation of p-eIF2α, resulting in increased HbF production by a post-transcriptional mechanism [144]. Lopez et al. investigated whether SAL can induce HbF expression through the stress-signaling pathway by the activation of p-eIF2α and ATF4 trans-activation in the γ-globin gene promoter [148]. In sickle erythroid precursors, SAL (24 μM) increased F cells and significantly reduced oxidative stress, and increased levels of p-eIF2α and ATF4 [148]. In humanized SCD mice, a single intraperitoneal injection of SAL (at a dose of 5 mg/kg for four weeks) mediated a 2.3-fold increase in F cells and reduced the percentage of sickle erythrocytes and erythrocyte ROS production [148]. Although the study did not report an additive effect when SAL was used in combination with HU, it was concluded that SAL was as effective as HU[148].

3.5.2. PGC-1α Agonist: ZLN005

The pharmacologic induction of peroxisome proliferator-activated receptor-γ coactivator 1-α (PGC-1α) has recently been shown to induce HbF gene expression [149]. PGC-1α interacts with different nuclear receptors such as PPARg or TR4, as well as in the recruitment of chromatin complexes. In human CD34+ cell-derived primary erythroid precursors, Sun et al. recently showed that the upregulation of PGC-1α by ZLN005, a small-molecule PGC-1α agonist, induced both γ-globin mRNA expression and HbF and increased F cells without significantly affecting cell proliferation and differentiation [149]. It was also reported that the use of ZLN005 in combination with HU exhibited an additive effect on the expression of γ-globin and the generation of F cells [149]. In addition, ZLN005 induced human γ-globin gene expression in SCD mice [149].
The mechanism of action of the abovementioned novel experimental strategies outlined in this section has also been summarized in Table 2.

4. Conclusions and Future Perspectives

In conclusion, pharmacological agents to induce HbF are currently being used to alleviate the morbidity profile and disease burden that is associated with β-thalassemia and SCD. Understanding the complex regulation of HbF is key and necessary to supporting the generation of newer modalities. Research into novel pharmacologic strategies directed at elevating HbF levels is currently ongoing. The development of such novel curative treatments and approaches for β-thalassemia and SCD is crucial to diminish the clinical disease severity and substantially improve the quality of patients’ lives.

Author Contributions

Conceptualization, R.B.-F. and I.M.; investigation, R.B.-F. and L.D.F.; writing—original draft preparation, R.B.-F.; writing—review and editing, I.M., L.D.F., A.T.T. and M.D.C.; supervision, I.M., A.T.T. and M.D.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

R.B.-F. has nothing to disclose. L.D.F. serves on the advisory board of Novartis, Roche and Vifor. I.M. reports receiving honoraria from Sanofi-Genzyme and Amicus Therapeutics and serves on the advisory boards for Bristol Myers Squibb/Celgene. M.D.C. serves on the advisory boards for Bristol Myers Squibb/Celgene, Sanofi/Genzyme, Agios, Silence Therapeutics, Vertex and Vifor. A.T.T. reports receiving consultancy from Novartis, Bristol Myers Squibb/Celgene, Vifor Pharma, Ionis Pharmaceuticals, Imara and Agios Pharmaceuticals, and research funding from Novartis, Bristol Myers Squibb, Vifor Pharma, Imara and Ionis Pharmaceuticals.

References

  1. Taher, A.T.; Musallam, K.M.; Cappellini, M.D. beta-Thalassemias. N. Engl. J. Med. 2021, 384, 727–743. [Google Scholar] [CrossRef] [PubMed]
  2. Taher, A.T.; Weatherall, D.J.; Cappellini, M.D. Thalassaemia. Lancet 2018, 391, 155–167. [Google Scholar] [CrossRef]
  3. Ware, R.E.; de Montalembert, M.; Tshilolo, L.; Abboud, M.R. Sickle cell disease. Lancet 2017, 390, 311–323. [Google Scholar] [CrossRef]
  4. Sankaran, V.G.; Orkin, S.H. The switch from fetal to adult hemoglobin. Cold Spring Harb. Perspect. Med. 2013, 3, a011643. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Amato, A.; Cappabianca, M.P.; Perri, M.; Zaghis, I.; Grisanti, P.; Ponzini, D.; Di Biagio, P. Interpreting elevated fetal hemoglobin in pathology and health at the basic laboratory level: New and known γ- gene mutations associated with hereditary persistence of fetal hemoglobin. Int. J. Lab. Hematol. 2014, 36, 13–19. [Google Scholar] [CrossRef]
  6. Hampl, V.; Bibova, J.; Stranak, Z.; Wu, X.; Michelakis, E.D.; Hashimoto, K.; Archer, S.L. Hypoxic fetoplacental vasoconstriction in humans is mediated by potassium channel inhibition. Am. J. Physiol. Heart Circ. Physiol. 2002, 283, H2440–H2449. [Google Scholar] [CrossRef] [Green Version]
  7. Murji, A.; Sobel, M.L.; Hasan, L.; McLeod, A.; Waye, J.S.; Sermer, M.; Berger, H. Pregnancy outcomes in women with elevated levels of fetal hemoglobin. J. Matern. Fetal. Neonatal. Med. 2012, 25, 125–129. [Google Scholar] [CrossRef]
  8. Sokolova, A.; Mararenko, A.; Rozin, A.; Podrumar, A.; Gotlieb, V. Hereditary persistence of hemoglobin F is protective against red cell sickling. A case report and brief review. Hematol. Oncol. Stem. Cell Ther. 2019, 12, 215–219. [Google Scholar] [CrossRef]
  9. Alter, B.P.; Rappeport, J.M.; Huisman, T.H.J.; Schroeder, W.A.; Nathan, D.G. Fetal Erythropoiesis Following Bone Marrow Transplantation. Blood 1976, 48, 843–853. [Google Scholar] [CrossRef] [Green Version]
  10. Nachbaur, D.; Kropshofer, G.; Heitger, A.; Lätzer, K.; Glassl, H.; Ludescher, C.; Nussbaumer, W.; Niederwieser, D. Phenotypic and functional lymphocyte recovery after CD34+-enriched versus non-T cell-depleted autologous peripheral blood stem cell transplantation. J. Hematother. Stem Cell Res. 2000, 9, 727–736. [Google Scholar] [CrossRef]
  11. Liu, J.W.; Hong, T.; Qin, X.; Liang, Y.M.; Zhang, P. Recent advance on genome editing for therapy of beta-hemoglobinopathies. Yi Chuan 2018, 40, 95–103. [Google Scholar] [CrossRef] [PubMed]
  12. Liu, N.; Hargreaves, V.V.; Zhu, Q.; Kurland, J.V.; Hong, J.; Kim, W.; Sher, F.; Macias-Trevino, C.; Rogers, J.M.; Kurita, R.; et al. Direct Promoter Repression by BCL11A Controls the Fetal to Adult Hemoglobin Switch. Cell 2018, 173, 430–442.e417. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Thein, S.L.; Craig, J.E. Genetics of Hb F/F cell variance in adults and heterocellular hereditary persistence of fetal hemoglobin. Hemoglobin 1998, 22, 401–414. [Google Scholar] [CrossRef] [PubMed]
  14. Steinberg, M.H.; Forget, B.G.; Higgs, D.R.; Weatherall, D.J. Disorders of Hemoglobin: Genetics, Pathophysiology, and Clinical Management; Cambridge University Press: New York, NY, USA, 2009. [Google Scholar]
  15. Steinberg, M.H. Fetal hemoglobin in sickle cell anemia. Blood 2020, 136, 2392–2400. [Google Scholar] [CrossRef]
  16. Steinberg, M.H. Fetal Hemoglobin in Sickle Hemoglobinopathies: High HbF Genotypes and Phenotypes. J. Clin. Med. 2020, 9, 3782. [Google Scholar] [CrossRef]
  17. Mukherjee, M.; Rahaman, M.; Ray, S.K.; Shukla, P.C.; Dolai, T.K.; Chakravorty, N. Revisiting fetal hemoglobin inducers in beta-hemoglobinopathies: A review of natural products, conventional and combinatorial therapies. Mol. Biol. Rep. 2022, 49, 2359–2373. [Google Scholar] [CrossRef]
  18. Métais, J.-Y.; Doerfler, P.A.; Mayuranathan, T.; Bauer, D.E.; Fowler, S.C.; Hsieh, M.M.; Katta, V.; Keriwala, S.; Lazzarotto, C.R.; Luk, K. Genome editing of HBG1 and HBG2 to induce fetal hemoglobin. Blood Adv. 2019, 3, 3379–3392. [Google Scholar] [CrossRef] [Green Version]
  19. Magrin, E.; Miccio, A.; Cavazzana, M. Lentiviral and genome-editing strategies for the treatment of β-hemoglobinopathies. Blood J. Am. Soc. Hematol. 2019, 134, 1203–1213. [Google Scholar] [CrossRef]
  20. Walters, M.C. Induction of Fetal Hemoglobin by Gene Therapy. Mass Med. Soc. 2021, 384, 284–285. [Google Scholar] [CrossRef]
  21. Frangoul, H.; Altshuler, D.; Cappellini, M.D.; Chen, Y.-S.; Domm, J.; Eustace, B.K.; Foell, J.; de la Fuente, J.; Grupp, S.; Handgretinger, R. CRISPR-Cas9 gene editing for sickle cell disease and β-thalassemia. N. Engl. J. Med. 2021, 384, 252–260. [Google Scholar] [CrossRef]
  22. Samuelson, C.; Radtke, S.; Zhu, H.; Llewellyn, M.; Fields, E.; Cook, S.; Huang, M.-L.W.; Jerome, K.R.; Kiem, H.-P.; Humbert, O. Multiplex CRISPR/Cas9 genome editing in hematopoietic stem cells for fetal hemoglobin reinduction generates chromosomal translocations. Mol. Ther. Methods Clin. Dev. 2021, 23, 507–523. [Google Scholar] [CrossRef] [PubMed]
  23. Breda, L.; Motta, I.; Lourenco, S.; Gemmo, C.; Deng, W.; Rupon, J.W.; Abdulmalik, O.Y.; Manwani, D.; Blobel, G.A.; Rivella, S. Forced chromatin looping raises fetal hemoglobin in adult sickle cells to higher levels than pharmacologic inducers. Blood J. Am. Soc. Hematol. 2016, 128, 1139–1143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Deng, W.; Rupon, J.W.; Krivega, I.; Breda, L.; Motta, I.; Jahn, K.S.; Reik, A.; Gregory, P.D.; Rivella, S.; Dean, A.; et al. Reactivation of developmentally silenced globin genes by forced chromatin looping. Cell 2014, 158, 849–860. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Sherkow, J.S. Focus: Genome Editing: CRISPR, Patents, and the Public Health. Yale J. Biol. Med. 2017, 90, 667. [Google Scholar] [PubMed]
  26. Rigter, T.; Klein, D.; Weinreich, S.S.; Cornel, M.C. Moving somatic gene editing to the clinic: Routes to market access and reimbursement in Europe. Eur. J. Hum. Genet. 2021, 29, 1477–1484. [Google Scholar] [CrossRef]
  27. Cornel, M.C.; Howard, H.C.; Lim, D.; Bonham, V.L.; Wartiovaara, K. Moving towards a cure in genetics: What is needed to bring somatic gene therapy to the clinic? Eur. J. Hum. Genet. 2019, 27, 484–487. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Yasara, N.; Premawardhena, A.; Mettananda, S. A comprehensive review of hydroxyurea for β-haemoglobinopathies: The role revisited during COVID-19 pandemic. Orphanet J. Rare Dis. 2021, 16, 114. [Google Scholar] [CrossRef]
  29. Lemonne, N.; Möckesch, B.; Charlot, K.; Garnier, Y.; Waltz, X.; Lamarre, Y.; Antoine-Jonville, S.; Etienne-Julan, M.; Hardy-Dessources, M.-D.; Romana, M. Effects of hydroxyurea on blood rheology in sickle cell anemia: A two-years follow-up study. Clin. Hemorheol. Microcirc. 2017, 67, 141–148. [Google Scholar] [CrossRef]
  30. Keikhaei, B.; Yousefi, H.; Bahadoram, M. Hydroxyurea: Clinical and hematological effects in patients with sickle cell anemia. Glob. J. Health Sci. 2016, 8, 252. [Google Scholar] [CrossRef]
  31. Ballas, S.K.; McCarthy, W.F.; Guo, N.; Brugnara, C.; Kling, G.; Bauserman, R.L.; Waclawiw, M.A. Early detection of response to hydroxyurea therapy in patients with sickle cell anemia. Hemoglobin 2010, 34, 424–429. [Google Scholar] [CrossRef]
  32. Keikhaei, B.; Yousefi, H.; Bahadoram, M. Clinical and haematological effects of hydroxyurea in β-Thalassemia intermedia patients. J. Clin. Diagn. Res. JCDR 2015, 9, OM01. [Google Scholar] [CrossRef] [PubMed]
  33. Chowdhury, P.K.; Jena, R.; Chowdhury, D. Red Cell Indices as Predictors of Response to Hydroxyurea Therapy in HbE/Beta Thalassaemia Patients; American Society of Hematology: Washington, DC, USA, 2016. [Google Scholar]
  34. Cisneros, G.S.; Thein, S.L. Research in sickle cell disease: From bedside to bench to bedside. Hemasphere 2021, 5, e584. [Google Scholar] [CrossRef] [PubMed]
  35. Italia, K.Y.; Jijina, F.J.; Merchant, R.; Panjwani, S.; Nadkarni, A.H.; Sawant, P.M.; Nair, S.B.; Ghosh, K.; Colah, R.B. Response to hydroxyurea in β thalassemia major and intermedia: Experience in western India. Clin. Chim. Acta 2009, 407, 10–15. [Google Scholar] [CrossRef] [PubMed]
  36. Watanapokasin, R.; Sanmund, D.; Winichagoon, P.; Muta, K.; Fucharoen, S. Hydroxyurea responses and fetal hemoglobin induction in β-thalassemia/HbE patients’ peripheral blood erythroid cell culture. Ann. Hematol. 2006, 85, 164–169. [Google Scholar] [CrossRef]
  37. Watanapokasin, Y.; Chuncharunee, S.; Sanmund, D.; Kongnium, W.; Winichagoon, P.; Rodgers, G.P.; Fucharoen, S. In vivo and in vitro studies of fetal hemoglobin induction by hydroxyurea in β-thalassemia/hemoglobin E patients. Exp. Hematol. 2005, 33, 1486–1492. [Google Scholar] [CrossRef]
  38. Rigano, P.; Pecoraro, A.; Calzolari, R.; Troia, A.; Acuto, S.; Renda, D.; Pantalone, G.R.; Maggio, A.; Marzo, R.D. Desensitization to hydroxycarbamide following long-term treatment of thalassaemia intermedia as observed in vivo and in primary erythroid cultures from treated patients. Br. J. Haematol. 2010, 151, 509–515. [Google Scholar] [CrossRef]
  39. Singer, S.T.; Vichinsky, E.P.; Larkin, S.; Olivieri, N.; Sweeters, N.; Kuypers, F.A. Hydroxycarbamide-induced changes in E/beta thalassemia red blood cells. Am. J. Hematol. 2008, 83, 842–845. [Google Scholar] [CrossRef] [Green Version]
  40. Musallam, K.M.; Taher, A.T.; Cappellini, M.D.; Sankaran, V.G. Clinical experience with fetal hemoglobin induction therapy in patients with β-thalassemia. Blood J. Am. Soc. Hematol. 2013, 121, 2199–2212. [Google Scholar] [CrossRef] [Green Version]
  41. Luchtman-Jones, L.; Pressel, S.; Hilliard, L.; Brown, R.C.; Smith, M.G.; Thompson, A.A.; Lee, M.T.; Rothman, J.; Rogers, Z.R.; Owen, W. Effects of hydroxyurea treatment for patients with hemoglobin SC disease. Am. J. Hematol. 2016, 91, 238–242. [Google Scholar] [CrossRef]
  42. Di Maggio, R.; Hsieh, M.M.; Zhao, X.; Calvaruso, G.; Rigano, P.; Renda, D.; Tisdale, J.F.; Maggio, A. Chronic administration of hydroxyurea (HU) benefits caucasian patients with sickle-beta thalassemia. Int. J. Mol. Sci. 2018, 19, 681. [Google Scholar] [CrossRef] [Green Version]
  43. Voskaridou, E.; Christoulas, D.; Bilalis, A.; Plata, E.; Varvagiannis, K.; Stamatopoulos, G.; Sinopoulou, K.; Balassopoulou, A.; Loukopoulos, D.; Terpos, E. The effect of prolonged administration of hydroxyurea on morbidity and mortality in adult patients with sickle cell syndromes: Results of a 17-year, single-center trial (LaSHS). Blood J. Am. Soc. Hematol. 2010, 115, 2354–2363. [Google Scholar] [CrossRef] [PubMed]
  44. Platt, O.S. Hydroxyurea for the treatment of sickle cell anemia. N. Engl. J. Med. 2008, 358, 1362–1369. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Tang, D.C.; Zhu, J.; Liu, W.; Chin, K.; Sun, J.; Chen, L.; Hanover, J.A.; Rodgers, G.P. The hydroxyurea-induced small GTP-binding protein SAR modulates gamma-globin gene expression in human erythroid cells. Blood 2005, 106, 3256–3263. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Zhu, J.; Chin, K.; Aerbajinai, W.; Kumkhaek, C.; Li, H.; Rodgers, G.P. Hydroxyurea-inducible SAR1 gene acts through the Giα/JNK/Jun pathway to regulate γ-globin expression. Blood J. Am. Soc. Hematol. 2014, 124, 1146–1156. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Hebert, N.; Rakotoson, M.G.; Bodivit, G.; Audureau, E.; Bencheikh, L.; Kiger, L.; Oubaya, N.; Pakdaman, S.; Sakka, M.; Di Liberto, G. Individual red blood cell fetal hemoglobin quantification allows to determine protective thresholds in sickle cell disease. Am. J. Hematol. 2020, 95, 1235–1245. [Google Scholar] [CrossRef]
  48. Buchanan, G.R. “Packaging” of fetal hemoglobin in sickle cell anemia. Blood J. Am. Soc. Hematol. 2014, 123, 464–465. [Google Scholar] [CrossRef]
  49. Steinberg, M.H.; Chui, D.H.; Dover, G.J.; Sebastiani, P.; Alsultan, A. Fetal hemoglobin in sickle cell anemia: A glass half full? Blood J. Am. Soc. Hematol. 2014, 123, 481–485. [Google Scholar] [CrossRef] [Green Version]
  50. Dong, M.; McGann, P.T. Changing the clinical paradigm of hydroxyurea treatment for sickle cell anemia through precision medicine. Clin. Pharmacol. Ther. 2021, 109, 73–81. [Google Scholar] [CrossRef]
  51. McGann, P.T.; Niss, O.; Dong, M.; Marahatta, A.; Howard, T.A.; Mizuno, T.; Lane, A.; Kalfa, T.A.; Malik, P.; Quinn, C.T. Robust clinical and laboratory response to hydroxyurea using pharmacokinetically guided dosing for young children with sickle cell anemia. Am. J. Hematol. 2019, 94, 871–879. [Google Scholar] [CrossRef]
  52. Steinberg, M.H.; McCarthy, W.F.; Castro, O.; Ballas, S.K.; Armstrong, F.D.; Smith, W.; Ataga, K.; Swerdlow, P.; Kutlar, A.; DeCastro, L. The risks and benefits of long-term use of hydroxyurea in sickle cell anemia: A 17.5 year follow-up. Am. J. Hematol. 2010, 85, 403–408. [Google Scholar] [CrossRef] [Green Version]
  53. Thomas, R.; Dulman, R.; Lewis, A.; Notarangelo, B.; Yang, E. Prospective longitudinal follow-up of children with sickle cell disease treated with hydroxyurea since infancy. Pediatric Blood Cancer 2019, 66, e27816. [Google Scholar] [CrossRef] [PubMed]
  54. Green, N.S.; Manwani, D.; Qureshi, M.; Ireland, K.; Sinha, A.; Smaldone, A.M. Decreased fetal hemoglobin over time among youth with sickle cell disease on hydroxyurea is associated with higher urgent hospital use. Pediatric Blood Cancer 2016, 63, 2146–2153. [Google Scholar] [CrossRef] [PubMed]
  55. Matte, A.; Zorzi, F.; Mazzi, F.; Federti, E.; Olivieri, O.; De Franceschi, L. New Therapeutic Options for the Treatment of Sickle Cell Disease. Mediterr. J. Hematol. Infect. Dis. 2019, 11, e2019002. [Google Scholar] [CrossRef] [PubMed]
  56. Matte, A.; Cappellini, M.D.; Iolascon, A.; Enrica, F.; De Franceschi, L. Emerging drugs in randomized controlled trials for sickle cell disease: Are we on the brink of a new era in research and treatment? Expert. Opin. Investig. Drugs 2020, 29, 23–31. [Google Scholar] [CrossRef] [PubMed]
  57. De Franceschi, L.; Cappellini, M.D.; Olivieri, O. Thrombosis and sickle cell disease. Semin. Thromb Hemost. 2011, 37, 226–236. [Google Scholar] [CrossRef] [Green Version]
  58. De Franceschi, L.; Corrocher, R. Established and experimental treatments for sickle cell disease. Haematologica 2004, 89, 348–356. [Google Scholar]
  59. Lavelle, D.; Engel, J.D.; Saunthararajah, Y. Fetal hemoglobin induction by epigenetic drugs. Semin Hematol. 2018, 55, 60–67. [Google Scholar] [CrossRef]
  60. Charache, S.; Dover, G.; Smith, K.; Talbot, C.C.; Moyer, M.; Boyer, S. Treatment of sickle cell anemia with 5-azacytidine results in increased fetal hemoglobin production and is associated with nonrandom hypomethylation of DNA around the gamma-delta-beta-globin gene complex. Proc. Natl. Acad. Sci. USA 1983, 80, 4842–4846. [Google Scholar] [CrossRef] [Green Version]
  61. Olivieri, N.F.; Saunthararajah, Y.; Thayalasuthan, V.; Kwiatkowski, J.; Ware, R.E.; Kuypers, F.A.; Kim, H.-Y.; Trachtenberg, F.L.; Vichinsky, E.P.; Network, T.C.R. A pilot study of subcutaneous decitabine in β-thalassemia intermedia. Blood J. Am. Soc. Hematol. 2011, 118, 2708–2711. [Google Scholar] [CrossRef]
  62. Saunthararajah, Y.; Hillery, C.A.; Lavelle, D.; Molokie, R.; Dorn, L.; Bressler, L.; Gavazova, S.; Chen, Y.-H.; Hoffman, R.; DeSimone, J. Effects of 5-aza-2′-deoxycytidine on fetal hemoglobin levels, red cell adhesion, and hematopoietic differentiation in patients with sickle cell disease. Blood 2003, 102, 3865–3870. [Google Scholar] [CrossRef]
  63. Koshy, M.; Dorn, L.; Bressler, L.; Molokie, R.; Lavelle, D.; Talischy, N.; Hoffman, R.; van Overveld, W.; DeSimone, J. 2-deoxy 5-azacytidine and fetal hemoglobin induction in sickle cell anemia. Blood J. Am. Soc. Hematol. 2000, 96, 2379–2384. [Google Scholar]
  64. DeSimone, J.; Koshy, M.; Dorn, L.; Lavelle, D.; Bressler, L.; Molokie, R.; Talischy, N. Maintenance of elevated fetal hemoglobin levels by decitabine during dose interval treatment of sickle cell anemia. Blood J. Am. Soc. Hematol. 2002, 99, 3905–3908. [Google Scholar] [CrossRef] [PubMed]
  65. Saunthararajah, Y.; Molokie, R.; Saraf, S.; Sidhwani, S.; Gowhari, M.; Vara, S.; Lavelle, D.; DeSimone, J. Clinical effectiveness of decitabine in severe sickle cell disease. Br. J. Haematol. 2008, 141, 126–129. [Google Scholar] [CrossRef] [PubMed]
  66. Molokie, R.; Lavelle, D.; Gowhari, M.; Pacini, M.; Krauz, L.; Hassan, J.; Ibanez, V.; Ruiz, M.A.; Ng, K.P.; Woost, P. Oral tetrahydrouridine and decitabine for non-cytotoxic epigenetic gene regulation in sickle cell disease: A randomized phase 1 study. PLoS Med. 2017, 14, e1002382. [Google Scholar] [CrossRef] [Green Version]
  67. Gilmartin, A.G.; Groy, A.; Gore, E.R.; Atkins, C.; Long, E.R.; Montoute, M.N.; Wu, Z.; Halsey, W.; McNulty, D.E.; Ennulat, D. In vitro and in vivo induction of fetal hemoglobin with a reversible and selective DNMT1 inhibitor. Haematologica 2021, 106, 1979. [Google Scholar] [CrossRef]
  68. Ronzoni, L.; Sonzogni, L.; Fossati, G.; Modena, D.; Trombetta, E.; Porretti, L.; Cappellini, M.D. Modulation of gamma globin genes expression by histone deacetylase inhibitors: An in vitro study. Br. J. Haematol. 2014, 165, 714–721. [Google Scholar] [CrossRef]
  69. Weinberg, R.S.; Ji, X.; Sutton, M.; Perrine, S.; Galperin, Y.; Li, Q.; Liebhaber, S.A.; Stamatoyannopoulos, G.; Atweh, G.F. Butyrate increases the efficiency of translation of gamma-globin mRNA. Blood 2005, 105, 1807–1809. [Google Scholar] [CrossRef] [Green Version]
  70. de Melo, T.R.F.; Dulmovits, B.M.; dos Santos Fernandes, G.F.; de Souza, C.M.; Lanaro, C.; He, M.; Al Abed, Y.; Chung, M.C.; Blanc, L.; Costa, F.F. Synthesis and pharmacological evaluation of pomalidomide derivatives useful for sickle cell disease treatment. Bioorg. Chem. 2021, 114, 105077. [Google Scholar] [CrossRef]
  71. Bradner, J.E.; Mak, R.; Tanguturi, S.K.; Mazitschek, R.; Haggarty, S.J.; Ross, K.; Chang, C.Y.; Bosco, J.; West, N.; Morse, E. Chemical genetic strategy identifies histone deacetylase 1 (HDAC1) and HDAC2 as therapeutic targets in sickle cell disease. Proc. Natl. Acad. Sci. USA 2010, 107, 12617–12622. [Google Scholar] [CrossRef] [Green Version]
  72. Eckschlager, T.; Plch, J.; Stiborova, M.; Hrabeta, J. Histone deacetylase inhibitors as anticancer drugs. Int. J. Mol. Sci. 2017, 18, 1414. [Google Scholar] [CrossRef]
  73. Mettananda, S.; Yasara, N.; Fisher, C.A.; Taylor, S.; Gibbons, R.; Higgs, D. Synergistic silencing of α-globin and induction of γ-globin by histone deacetylase inhibitor, vorinostat as a potential therapy for β-thalassaemia. Sci. Rep. 2019, 9, 11649. [Google Scholar] [CrossRef] [PubMed]
  74. Sankaran, V.G.; Weiss, M.J. Anemia: Progress in molecular mechanisms and therapies. Nat. Med. 2015, 21, 221–230. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Hebbel, R.P.; Vercellotti, G.M.; Pace, B.S.; Solovey, A.N.; Kollander, R.; Abanonu, C.F.; Nguyen, J.; Vineyard, J.V.; Belcher, J.D.; Abdulla, F. The HDAC inhibitors trichostatin A and suberoylanilide hydroxamic acid exhibit multiple modalities of benefit for the vascular pathobiology of sickle transgenic mice. Blood J. Am. Soc. Hematol. 2010, 115, 2483–2490. [Google Scholar] [CrossRef] [PubMed]
  76. Junker, L.H.; Li, B.; Zhu, X.; Koti, S.; Cerbone, R.E.; Hendrick, C.L.; Sangerman, J.; Perrine, S.; Pace, B.S. Novel histone deacetylase inhibitor CT-101 induces γ-globin gene expression in sickle erythroid progenitors with targeted epigenetic effects. Blood Cells Mol. Dis. 2022, 93, 102626. [Google Scholar] [CrossRef]
  77. Krivega, I.; Byrnes, C.; de Vasconcellos, J.F.; Lee, Y.T.; Kaushal, M.; Dean, A.; Miller, J.L. Inhibition of G9a methyltransferase stimulates fetal hemoglobin production by facilitating LCR/γ-globin looping. Blood J. Am. Soc. Hematol. 2015, 126, 665–672. [Google Scholar] [CrossRef] [Green Version]
  78. Renneville, A.; Van Galen, P.; Canver, M.C.; McConkey, M.; Krill-Burger, J.M.; Dorfman, D.M.; Holson, E.B.; Bernstein, B.E.; Orkin, S.H.; Bauer, D.E. EHMT1 and EHMT2 inhibition induces fetal hemoglobin expression. Blood J. Am. Soc. Hematol. 2015, 126, 1930–1939. [Google Scholar] [CrossRef] [Green Version]
  79. Chen, X.; Skutt-Kakaria, K.; Davison, J.; Ou, Y.-L.; Choi, E.; Malik, P.; Loeb, K.; Wood, B.; Georges, G.; Torok-Storb, B. G9a/GLP-dependent histone H3K9me2 patterning during human hematopoietic stem cell lineage commitment. Genes Dev. 2012, 26, 2499–2511. [Google Scholar] [CrossRef] [Green Version]
  80. Nualkaew, T.; Khamphikham, P.; Pongpaksupasin, P.; Kaewsakulthong, W.; Songdej, D.; Paiboonsukwong, K.; Sripichai, O.; Engel, J.D.; Hongeng, S.; Fucharoen, S. UNC0638 induces high levels of fetal hemoglobin expression in β-thalassemia/HbE erythroid progenitor cells. Ann. Hematol. 2020, 99, 2027–2036. [Google Scholar] [CrossRef]
  81. Liu, F.; Barsyte-Lovejoy, D.; Li, F.; Xiong, Y.; Korboukh, V.; Huang, X.-P.; Allali-Hassani, A.; Janzen, W.P.; Roth, B.L.; Frye, S.V. Discovery of an in vivo chemical probe of the lysine methyltransferases G9a and GLP. J. Med. Chem. 2013, 56, 8931–8942. [Google Scholar] [CrossRef] [Green Version]
  82. Dulmovits, B.M.; Appiah-Kubi, A.O.; Papoin, J.; Hale, J.; He, M.; Al-Abed, Y.; Didier, S.; Gould, M.; Husain-Krautter, S.; Singh, S.A. Pomalidomide reverses γ-globin silencing through the transcriptional reprogramming of adult hematopoietic progenitors. Blood J. Am. Soc. Hematol. 2016, 127, 1481–1492. [Google Scholar] [CrossRef]
  83. Khamphikham, P.; Nualkaew, T.; Pongpaksupasin, P.; Kaewsakulthong, W.; Songdej, D.; Paiboonsukwong, K.; Engel, J.D.; Hongeng, S.; Fucharoen, S.; Sripichai, O. High-level induction of fetal haemoglobin by pomalidomide in β-thalassaemia/HbE erythroid progenitor cells. Br. J. Haematol. 2020, 189, e240–e245. [Google Scholar] [CrossRef] [PubMed]
  84. Aerbajinai, W.; Zhu, J.; Gao, Z.; Chin, K.; Rodgers, G.P. Thalidomide induces γ-globin gene expression through increased reactive oxygen species–mediated p38 MAPK signaling and histone H4 acetylation in adult erythropoiesis. Blood J. Am. Soc. Hematol. 2007, 110, 2864–2871. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Jain, M.; Chakrabarti, P.; Dolai, T.K.; Ghosh, P.; Mandal, P.K.; Baul, S.N.; De, R. Comparison of efficacy and safety of thalidomide vs hydroxyurea in patients with Hb E-β thalassemia-a pilot study from a tertiary care Centre of India. Blood Cells Mol. Dis. 2021, 88, 102544. [Google Scholar] [CrossRef] [PubMed]
  86. Li, X.; Hu, S.; Liu, Y.; Huang, J.; Hong, W.; Xu, L.; Xu, H.; Fang, J. Efficacy of thalidomide treatment in children with transfusion dependent β-thalassemia: A retrospective clinical study. Front. Pharmacol. Front Pharmacol. 2021, 12, 722502. [Google Scholar] [CrossRef]
  87. Chen, J.-M.; Zhu, W.-J.; Liu, J.; Wang, G.-Z.; Chen, X.-Q.; Tan, Y.; Xu, W.-W.; Qu, L.-W.; Li, J.-Y.; Yang, H.-J. Safety and efficacy of thalidomide in patients with transfusion-dependent β-thalassemia: A randomized clinical trial. Signal Transduct. Target. Ther. 2021, 6, 405. [Google Scholar] [CrossRef]
  88. Chandra, J.; Parakh, N.; Singh, N.; Sharma, S.; Goel, M.; Pemde, H. Efficacy and safety of thalidomide in patients with transfusion-dependent thalassemia. Indian Pediatrics 2021, 58, 611–616. [Google Scholar] [CrossRef]
  89. Yassin, A.K. Promising Response to Thalidomide in Symptomatic beta-Thalassemia. Indian J. Hematol. Blood Transfus. 2020, 36, 337–341. [Google Scholar] [CrossRef]
  90. Yang, K.; Wu, Y.; Zhou, Y.; Long, B.; Lu, Q.; Zhou, T.; Wang, L.; Geng, Z.; Yin, X. Thalidomide for patients with β-thalassemia: A multicenter experience. Mediterr. J. Hematol. Infect. Dis. 2020, 12, e2020021. [Google Scholar] [CrossRef]
  91. Javed, R.; Radhakrishnan, V.; Basu, S.; Chandy, M. Challenges in transfusion and the role of Thalidomide in E-β-Thalassemia—A case report. Clin. Case Rep. 2020, 8, 2208–2210. [Google Scholar] [CrossRef]
  92. Nag, A.; Radhakrishnan, V.S.; Kumar, J.; Bhave, S.; Mishra, D.K.; Nair, R.; Chandy, M. Thalidomide in patients with transfusion-dependent E-beta thalassemia refractory to hydroxyurea: A single-center experience. Indian J. Hematol. Blood Transfus. 2020, 36, 399–402. [Google Scholar] [CrossRef]
  93. Lu, Y.; Wei, Z.; Yang, G.; Lai, Y.; Liu, R. Investigating the Efficacy and Safety of Thalidomide for Treating Patients With ss-Thalassemia: A Meta-Analysis. Front. Pharmacol. 2021, 12, 814302. [Google Scholar] [CrossRef] [PubMed]
  94. Chen, J.; Zhu, W.; Cai, N.; Bu, S.; Li, J.; Huang, L. Thalidomide induces haematologic responses in patients with beta-thalassaemia. Eur. J. Haematol. 2017, 99, 437–441. [Google Scholar] [CrossRef] [PubMed]
  95. Chen, Y.; Cai, N.; Lai, Y.; Xu, W.; Li, J.; Huang, L.; Huang, Y.; Hu, M.; Yang, H.; Chen, J. Thalidomide for the Treatment of Thrombocytopenia and Hypersplenism in Patients with Cirrhosis or Thalassemia. Front. Pharmacol. 2020, 11, 1137. [Google Scholar] [CrossRef] [PubMed]
  96. Aguilar-Lopez, L.B.; Delgado-Lamas, J.L.; Rubio-Jurado, B.; Perea, F.J.; Ibarra, B. Thalidomide therapy in a patient with thalassemia major. Blood Cells Mol. Dis. 2008, 41, 136–137. [Google Scholar] [CrossRef] [PubMed]
  97. Masera, N.; Tavecchia, L.; Capra, M.; Cazzaniga, G.; Vimercati, C.; Pozzi, L.; Biondi, A.; Masera, G. Optimal response to thalidomide in a patient with thalassaemia major resistant to conventional therapy. Blood Transfus. 2010, 8, 63–65. [Google Scholar] [CrossRef]
  98. Li, Y.; Ren, Q.; Zhou, Y.; Li, P.; Lin, W.; Yin, X. Thalidomide has a significant effect in patients with thalassemia intermedia. Hematology 2018, 23, 50–54. [Google Scholar] [CrossRef] [Green Version]
  99. Ren, Q.; Zhou, Y.L.; Wang, L.; Chen, Y.S.; Ma, Y.N.; Li, P.P.; Yin, X.L. Clinical trial on the effects of thalidomide on hemoglobin synthesis in patients with moderate thalassemia intermedia. Ann. Hematol. 2018, 97, 1933–1939. [Google Scholar] [CrossRef]
  100. Meiler, S.E.; Wade, M.; Kutlar, F.; Yerigenahally, S.D.; Xue, Y.; Moutouh-de Parseval, L.A.; Corral, L.G.; Swerdlow, P.S.; Kutlar, A. Pomalidomide augments fetal hemoglobin production without the myelosuppressive effects of hydroxyurea in transgenic sickle cell mice. Blood J. Am. Soc. Hematol. 2011, 118, 1109–1112. [Google Scholar] [CrossRef] [Green Version]
  101. Moutouh-de Parseval, L.A.; Verhelle, D.; Glezer, E.; Jensen-Pergakes, K.; Ferguson, G.D.; Corral, L.G.; Morris, C.L.; Muller, G.; Brady, H.; Chan, K. Pomalidomide and lenalidomide regulate erythropoiesis and fetal hemoglobin production in human CD34+ cells. J. Clin. Investig. 2008, 118, 248–258. [Google Scholar] [CrossRef]
  102. McArthur, J.G.; Svenstrup, N.; Chen, C.; Fricot, A.; Carvalho, C.; Nguyen, J.; Nguyen, P.; Parachikova, A.; Abdulla, F.; Vercellotti, G.M.; et al. A novel, highly potent and selective phosphodiesterase-9 inhibitor for the treatment of sickle cell disease. Haematologica 2020, 105, 623–631. [Google Scholar] [CrossRef] [Green Version]
  103. Andemariam, B.; Bronte, L.; Gordeuk, V.; Howard, J.; Kanter, J.; Eleftheriou, P.; Pancham, S.; Hagar, R.; Clarke, L.; Blyden, G.; et al. The safety, pharmacokinetics & pharmacodynamic effects of IMR-687, a highly selective PDE9 inhibitor, in adults with sickle cell disease: Phase-2a placebo-controlled & open-label extension studies [abstract]. Hemasphere 2021, 5, 90. [Google Scholar]
  104. Andemariam, B.; Mant, T.; Eleftheriou, P.; Lugthart, S.; Bronté-Hall, L.; Barroso, F.; Blyden, G.; Mason, J.; Barysauskas, C.M.; Yen, J. Treatment with IMR-687, a Highly Selective PDE9 Inhibitor, Increases HbF and Reduces VOCs in Adults with Sickle Cell Disease in a Long-Term, Phase 2a, Open-Label Extension Study. Blood 2021, 138, 2046. [Google Scholar] [CrossRef]
  105. Matson, D.; Xie, K.; Roth, M.; Stuart, B.; Bruno, P.; Efremov, I.; Thompson, L.; Silver, S.; Moxham, C. Ftx-6058 Induces Fetal Hemoglobin Production and Ameliorates Disease Pathology in Sickle Cell Mice. Blood 2021, 138, 2018. [Google Scholar] [CrossRef]
  106. Higher FTX-6058 Doses Raise Hemoglobin in Healthy Adults in Trial. Sickle Cell Disease News. 16 December 2021. Available online: https://sicklecellanemianews.com/2021/12/16/higher-ftx-6058-doses-raise-hemoglobin-healthy-adults-trial/ (accessed on 15 April 2022).
  107. Shi, L.; Cui, S.; Engel, J.D.; Tanabe, O. Lysine-specific demethylase 1 is a therapeutic target for fetal hemoglobin induction. Nat. Med. 2013, 19, 291–294. [Google Scholar] [CrossRef] [Green Version]
  108. Lee, M.G.; Wynder, C.; Schmidt, D.M.; McCafferty, D.G.; Shiekhattar, R. Histone H3 lysine 4 demethylation is a target of nonselective antidepressive medications. Chem. Biol. 2006, 13, 563–567. [Google Scholar] [CrossRef] [Green Version]
  109. Xu, J.; Bauer, D.E.; Kerenyi, M.A.; Vo, T.D.; Hou, S.; Hsu, Y.J.; Yao, H.; Trowbridge, J.J.; Mandel, G.; Orkin, S.H. Corepressor-dependent silencing of fetal hemoglobin expression by BCL11A. Proc. Natl. Acad. Sci. USA 2013, 110, 6518–6523. [Google Scholar] [CrossRef] [Green Version]
  110. Neelamegam, R.; Ricq, E.L.; Malvaez, M.; Patnaik, D.; Norton, S.; Carlin, S.M.; Hill, I.T.; Wood, M.A.; Haggarty, S.J.; Hooker, J.M. Brain-penetrant LSD1 inhibitors can block memory consolidation. ACS Chem. Neurosci. 2012, 3, 120–128. [Google Scholar] [CrossRef]
  111. Cui, S.; Lim, K.C.; Shi, L.; Lee, M.; Jearawiriyapaisarn, N.; Myers, G.; Campbell, A.; Harro, D.; Iwase, S.; Trievel, R.C.; et al. The LSD1 inhibitor RN-1 induces fetal hemoglobin synthesis and reduces disease pathology in sickle cell mice. Blood 2015, 126, 386–396. [Google Scholar] [CrossRef]
  112. Rivers, A.; Vaitkus, K.; Ruiz, M.A.; Ibanez, V.; Jagadeeswaran, R.; Kouznetsova, T.; DeSimone, J.; Lavelle, D. RN-1, a potent and selective lysine-specific demethylase 1 inhibitor, increases γ-globin expression, F reticulocytes, and F cells in a sickle cell disease mouse model. Exp. Hematol. 2015, 43, 546–553.e543. [Google Scholar] [CrossRef]
  113. Kaewsakulthong, W.; Pongpaksupasin, P.; Nualkaew, T.; Hongeng, S.; Fucharoen, S.; Jearawiriyapaisarn, N.; Sripichai, O. Lysine-specific histone demethylase 1 inhibition enhances robust fetal hemoglobin induction in human beta(0)-thalassemia/hemoglobin E erythroid cells. Hematol. Rep. 2021, 13, 9215. [Google Scholar] [CrossRef]
  114. Viglietta, V.; Miller, D.; Bar-Or, A.; Phillips, J.T.; Arnold, D.L.; Selmaj, K.; Kita, M.; Hutchinson, M.; Yang, M.; Zhang, R.; et al. Efficacy of delayed-release dimethyl fumarate in relapsing-remitting multiple sclerosis: Integrated analysis of the phase 3 trials. Ann. Clin. Transl. Neurol. 2015, 2, 103–118. [Google Scholar] [CrossRef] [PubMed]
  115. Gold, R.; Arnold, D.L.; Bar-Or, A.; Fox, R.J.; Kappos, L.; Chen, C.; Parks, B.; Miller, C. Safety and efficacy of delayed-release dimethyl fumarate in patients with relapsing-remitting multiple sclerosis: 9 years’ follow-up of DEFINE, CONFIRM, and ENDORSE. Ther. Adv. Neurol. Disord. 2020, 13, 1756286420915005. [Google Scholar] [CrossRef] [PubMed]
  116. Alroughani, R.; Huppke, P.; Mazurkiewicz-Beldzinska, M.; Blaschek, A.; Valis, M.; Aaen, G.; Pultz, J.; Peng, X.; Beynon, V. Delayed-Release Dimethyl Fumarate Safety and Efficacy in Pediatric Patients with Relapsing-Remitting Multiple Sclerosis. Front. Neurol. 2020, 11, 606418. [Google Scholar] [CrossRef] [PubMed]
  117. Fox, R.J.; Gold, R.; Phillips, J.T.; Okwuokenye, M.; Zhang, A.; Marantz, J.L. Efficacy and Tolerability of Delayed-release Dimethyl Fumarate in Black, Hispanic, and Asian Patients with Relapsing-Remitting Multiple Sclerosis: Post Hoc Integrated Analysis of DEFINE and CONFIRM. Neurol. Ther. 2017, 6, 175–187. [Google Scholar] [CrossRef] [Green Version]
  118. de Franceschi, L.; Turrini, F.; Honczarenko, M.; Ayi, K.; Rivera, A.; Fleming, M.D.; Law, T.; Mannu, F.; Kuypers, F.A.; Bast, A.; et al. In vivo reduction of erythrocyte oxidant stress in a murine model of beta-thalassemia. Haematologica 2004, 89, 1287–1298. [Google Scholar]
  119. De Franceschi, L.; Bertoldi, M.; Matte, A.; Santos Franco, S.; Pantaleo, A.; Ferru, E.; Turrini, F. Oxidative stress and beta-thalassemic erythroid cells behind the molecular defect. Oxid. Med. Cell Longev. 2013, 2013, 985210. [Google Scholar] [CrossRef] [Green Version]
  120. Matte, A.; De Falco, L.; Iolascon, A.; Mohandas, N.; An, X.; Siciliano, A.; Leboeuf, C.; Janin, A.; Bruno, M.; Choi, S.Y. The interplay between peroxiredoxin-2 and nuclear factor-erythroid 2 is important in limiting oxidative mediated dysfunction in β-thalassemic erythropoiesis. Antioxid. Redox Signal. 2015, 23, 1284–1297. [Google Scholar] [CrossRef] [Green Version]
  121. Brugnara, C.; de Franceschi, L. Effect of cell age and phenylhydrazine on the cation transport properties of rabbit erythrocytes. J. Cell Physiol. 1993, 154, 271–280. [Google Scholar] [CrossRef]
  122. Krishnamoorthy, S.; Pace, B.; Gupta, D.; Sturtevant, S.; Li, B.; Makala, L.; Brittain, J.; Moore, N.; Vieira, B.F.; Thullen, T.; et al. Dimethyl fumarate increases fetal hemoglobin, provides heme detoxification, and corrects anemia in sickle cell disease. JCI Insight 2017, 2, e96409. [Google Scholar] [CrossRef] [Green Version]
  123. Krishnamoorthy, S.; Gupta, D.; Sturtevant, S.; Li, B.; Makala, L.C.; Hobbs, W.E.; Light, D.R.; Pace, B. Dimethyl fumarate induces fetal hemoglobin in sickle cell disease. Blood 2015, 126, 410. [Google Scholar] [CrossRef]
  124. Keleku-Lukwete, N.; Suzuki, M.; Otsuki, A.; Tsuchida, K.; Katayama, S.; Hayashi, M.; Naganuma, E.; Moriguchi, T.; Tanabe, O.; Engel, J.D.; et al. Amelioration of inflammation and tissue damage in sickle cell model mice by Nrf2 activation. Proc. Natl. Acad. Sci. USA 2015, 112, 12169–12174. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Belcher, J.D.; Chen, C.; Nguyen, J.; Zhang, P.; Abdulla, F.; Nguyen, P.; Killeen, T.; Xu, P.; O’Sullivan, G.; Nath, K.A.; et al. Control of Oxidative Stress and Inflammation in Sickle Cell Disease with the Nrf2 Activator Dimethyl Fumarate. Antioxid. Redox. Signal. 2017, 26, 748–762. [Google Scholar] [CrossRef] [PubMed]
  126. Hu, T.; Chung, Y.M.; Guan, M.; Ma, M.; Ma, J.; Berek, J.S.; Hu, M.C. Reprogramming ovarian and breast cancer cells into non-cancerous cells by low-dose metformin or SN-38 through FOXO3 activation. Sci. Rep. 2014, 4, 5810. [Google Scholar] [CrossRef] [Green Version]
  127. Takayama, H.; Misu, H.; Iwama, H.; Chikamoto, K.; Saito, Y.; Murao, K.; Teraguchi, A.; Lan, F.; Kikuchi, A.; Saito, R.; et al. Metformin suppresses expression of the selenoprotein P gene via an AMP-activated kinase (AMPK)/FoxO3a pathway in H4IIEC3 hepatocytes. J. Biol. Chem. 2014, 289, 335–345. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Yung, M.M.; Chan, D.W.; Liu, V.W.; Yao, K.M.; Ngan, H.Y. Activation of AMPK inhibits cervical cancer cell growth through AKT/FOXO3a/FOXM1 signaling cascade. BMC Cancer 2013, 13, 327. [Google Scholar] [CrossRef] [Green Version]
  129. Zhang, Y.; Paikari, A.; Sumazin, P.; Ginter Summarell, C.C.; Crosby, J.R.; Boerwinkle, E.; Weiss, M.J.; Sheehan, V.A. Metformin induces FOXO3-dependent fetal hemoglobin production in human primary erythroid cells. Blood 2018, 132, 321–333. [Google Scholar] [CrossRef] [Green Version]
  130. Han, J.; Saraf, S.L.; Molokie, R.E.; Gordeuk, V.R. Use of metformin in patients with sickle cell disease. Am. J. Hematol. 2019, 94, E13–E15. [Google Scholar] [CrossRef] [Green Version]
  131. Badawy, S.M.; Payne, A.B. Association between clinical outcomes and metformin use in adults with sickle cell disease and diabetes mellitus. Blood Adv. 2019, 3, 3297–3306. [Google Scholar] [CrossRef]
  132. Pace, B.S.; Perrine, S.; Li, B.; Makala, L.; Xu, H.; Takezaki, M.; Wolf, R.F.; Wang, A.; Xu, X.; Huang, J.; et al. Benserazide racemate and enantiomers induce fetal globin gene expression in vivo: Studies to guide clinical development for beta thalassemia and sickle cell disease. Blood Cells Mol. Dis. 2021, 89, 102561. [Google Scholar] [CrossRef]
  133. Dai, Y.; Sangerman, J.; Nouraie, M.; Faller, A.D.; Oneal, P.; Rock, A.; Owoyemi, O.; Niu, X.; Nekhai, S.; Maharaj, D.; et al. Effects of hydroxyurea on F-cells in sickle cell disease and potential impact of a second fetal globin inducer. Am. J. Hematol. 2017, 92, E10–E11. [Google Scholar] [CrossRef]
  134. Boosalis, M.S.; Sangerman, J.I.; White, G.L.; Wolf, R.F.; Shen, L.; Dai, Y.; White, E.; Makala, L.H.; Li, B.; Pace, B.S. Novel inducers of fetal globin identified through high throughput screening (HTS) are active in vivo in anemic baboons and transgenic mice. PLoS ONE 2015, 10, e0144660. [Google Scholar] [CrossRef] [PubMed]
  135. Dai, Y.; Sangerman, J.; Luo, H.Y.; Fucharoen, S.; Chui, D.H.; Faller, D.V.; Perrine, S.P. Therapeutic fetal-globin inducers reduce transcriptional repression in hemoglobinopathy erythroid progenitors through distinct mechanisms. Blood Cells Mol. Dis. 2016, 56, 62–69. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Santos, M.E.H.P.; Olops, L.; Vendrame, F.; Tavares, A.H.J.; Leonardo, D.P.; de Azevedo, P.C.; Piovesana, L.G.; Costa, F.F.; Fertrin, K.Y. Benserazide as a potential novel fetal hemoglobin inducer: An observational study in non-carriers of hemoglobin disorders. Blood Cells Mol. Dis. 2021, 87, 102511. [Google Scholar] [CrossRef]
  137. Nam, T.G.; Lee, J.; Walker, J.R.; Brinker, A.; Cho, C.Y.; Schultz, P.G. Identification and characterization of small-molecule inducers of fetal hemoglobin. ChemMedChem 2011, 6, 777–780. [Google Scholar] [CrossRef] [PubMed]
  138. Lai, Z.S.; Yeh, T.K.; Chou, Y.C.; Hsu, T.; Lu, C.T.; Kung, F.C.; Hsieh, M.Y.; Lin, C.H.; Chen, C.T.; James Shen, C.K.; et al. Potent and orally active purine-based fetal hemoglobin inducers for treating beta-thalassemia and sickle cell disease. Eur. J. Med. Chem. 2021, 209, 112938. [Google Scholar] [CrossRef] [PubMed]
  139. Elion, G. Acyclovir: Discovery, mechanism of action, and selectivity. J. Med. Virol. 1993, 41 (Suppl. 1), 2–6. [Google Scholar] [CrossRef]
  140. Ali, H.; Khan, F.; Musharraf, S.G. Acyclovir induces fetal hemoglobin via downregulation of γ-globin repressors, BCL11A and SOX6 trans-acting factors. Biochem. Pharmacol. 2021, 190, 114612. [Google Scholar] [CrossRef]
  141. Khan, F.; Ali, H.; Musharraf, S.G. Tenofovir disoproxil fumarate induces fetal hemoglobin production in K562 cells and beta-YAC transgenic mice: A therapeutic approach for gamma-globin induction. Exp. Cell Res. 2020, 394, 112168. [Google Scholar] [CrossRef]
  142. Liu, Y.; Shakur, Y.; Yoshitake, M.; Kambayashi, J.i. Cilostazol (Pletal®): A dual inhibitor of cyclic nucleotide phosphodiesterase type 3 and adenosine uptake. Cardiovasc. Drug Rev. 2001, 19, 369–386. [Google Scholar] [CrossRef]
  143. Ali, H.; Khan, F.; Musharraf, S.G. Cilostazol-mediated reversion of γ-globin silencing is associated with a high level of HbF production: A potential therapeutic candidate for β-globin disorders. Biomed. Pharmacother. 2021, 142, 112058. [Google Scholar] [CrossRef]
  144. Hahn, C.K.; Lowrey, C.H. Eukaryotic initiation factor 2alpha phosphorylation mediates fetal hemoglobin induction through a post-transcriptional mechanism. Blood 2013, 122, 477–485. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Chen, J.J.; Perrine, S. Stressing HbF synthesis: Role of translation? Blood 2013, 122, 467–468. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Boyce, M.; Bryant, K.F.; Jousse, C.; Long, K.; Harding, H.P.; Scheuner, D.; Kaufman, R.J.; Ma, D.; Coen, D.M.; Ron, D.; et al. A selective inhibitor of eIF2alpha dephosphorylation protects cells from ER stress. Science 2005, 307, 935–939. [Google Scholar] [CrossRef] [PubMed]
  147. Chen, F.; Xing, C.; Zhang, W.; Li, J.; Hu, T.; Li, L.; Li, H.; Cai, Y. Salubrinal, a novel inhibitor of eIF-2alpha dephosphorylation, promotes erythropoiesis at early stage targeted by ufmylation pathway. J. Cell Physiol. 2019, 234, 18560–18570. [Google Scholar] [CrossRef] [PubMed]
  148. Lopez, N.H.; Li, B.; Palani, C.; Siddaramappa, U.; Takezaki, M.; Xu, H.; Zhi, W.; Pace, B.S. Salubrinal induces fetal hemoglobin expression via the stress-signaling pathway in human sickle erythroid progenitors and sickle cell disease mice. PLoS ONE 2022, 17, e0261799. [Google Scholar] [CrossRef]
  149. Sun, Y.; Habara, A.; Le, C.Q.; Nguyen, N.; Chen, R.; Murphy, G.J.; Chui, D.H.K.; Steinberg, M.H.; Cui, S. Pharmacologic induction of PGC-1alpha stimulates fetal haemoglobin gene expression. Br. J. Haematol. 2022, 197, 97–109. [Google Scholar] [CrossRef]
Figure 1. Summary of all established pharmacological approaches and experimental therapeutic strategies for HbF induction in β-thalassemia and SCD.
Figure 1. Summary of all established pharmacological approaches and experimental therapeutic strategies for HbF induction in β-thalassemia and SCD.
Pharmaceuticals 15 00753 g001
Table 1. Mechanism of action of established pharmacologic approaches to HbF induction.
Table 1. Mechanism of action of established pharmacologic approaches to HbF induction.
AgentMechanism of Action
Hydroxyurea
  • Ribonucleotide reductase inhibitor (inihibition of DNA analysis)
DNA methyltransferase inhibitors
  • Hypomethaltion of DNA and post-transcriptional mechanism
HDAC inhibitors
  • Inhibition of HDAC activity
  • Epigenetic silencing of γ-globin genes
Immunomodulators: Thalidomide
and its derivatives
  • Activation of p38 MAPK kinase
  • Histone acetalation at γ-globin gene promoter
cGMP modulators: PDE-9 inhibitor
  • Inhibition of PDE9 and increased levels of cGMP
Abbreviations: HbF: Hemoglobin F; DNA: Deoxyribonucleic acid; HDAC: Histone deacetylase; cGMP: Cyclic guanosine monophosphate; PDE-9: Phosphodiesterase 9.
Table 2. Mechanism of action of novel experimental strategies for HbF induction.
Table 2. Mechanism of action of novel experimental strategies for HbF induction.
AgentMechanism of Action
EED inhibitors
  • Inhibition of the activity of PRC2 (which catalyzes tri methylation of histone H3 at lysine 27) and elevation in gene expression (e.g., HBG1/2)
LSD1 inhibitors
  • Disruption of the DRED complex that controls the expression of γ-globin
Modulators of redox-related transcriptional factors
  • Activation of Nrf2 transcriptional pathway
  • Enhancement of FOXO3 expression
Agents involved in the displacement/ suppression of γ-globin gene promoters
  • Displacement/suppression of γ-globin gene promoters
Modulators of cell signaling
  • Activation of p-eIF2α and ATF4 trans-activation in the γ-globin gene promoter
  • Induction of PGC-1α
Abbreviations: HbF: Hemoglobin F; EED: Embryonic ectoderm development; PRC2: Polycomb repressive complex 2; Nrf2: Nuclear factor erythroid 2-related factor 2; FOXO3: Forkhead box O-3; p-eIF2α: Phosphorylated eukaryotic initiation factor 2α; ATF4: Activating transcription factor 4; PGC-1α: Peroxisome proliferator-activated receptor-γ coactivator 1-α.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bou-Fakhredin, R.; De Franceschi, L.; Motta, I.; Cappellini, M.D.; Taher, A.T. Pharmacological Induction of Fetal Hemoglobin in β-Thalassemia and Sickle Cell Disease: An Updated Perspective. Pharmaceuticals 2022, 15, 753. https://doi.org/10.3390/ph15060753

AMA Style

Bou-Fakhredin R, De Franceschi L, Motta I, Cappellini MD, Taher AT. Pharmacological Induction of Fetal Hemoglobin in β-Thalassemia and Sickle Cell Disease: An Updated Perspective. Pharmaceuticals. 2022; 15(6):753. https://doi.org/10.3390/ph15060753

Chicago/Turabian Style

Bou-Fakhredin, Rayan, Lucia De Franceschi, Irene Motta, Maria Domenica Cappellini, and Ali T. Taher. 2022. "Pharmacological Induction of Fetal Hemoglobin in β-Thalassemia and Sickle Cell Disease: An Updated Perspective" Pharmaceuticals 15, no. 6: 753. https://doi.org/10.3390/ph15060753

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

Article Metrics

Back to TopTop