Next Article in Journal
The Experimental and In Silico-Based Evaluation of NRF2 Modulators, Sulforaphane and Brusatol, on the Transcriptome of Immortalized Bovine Mammary Alveolar Cells
Next Article in Special Issue
Pulmonary Hypertension in Sickle Cell Disease: Novel Findings of Gene Polymorphisms Related to Pathophysiology
Previous Article in Journal
Geniposide and Harpagoside Functionalized Cerium Oxide Nanoparticles as a Potential Neuroprotective
Previous Article in Special Issue
Impact of α-Globin Gene Expression and α-Globin Modifiers on the Phenotype of β-Thalassemia and Other Hemoglobinopathies: Implications for Patient Management
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Pharmacogenomics of Drugs Used in β-Thalassemia and Sickle-Cell Disease: From Basic Research to Clinical Applications

by
Roberto Gambari
1,*,
Aliyu Dahiru Waziri
2,
Hemali Goonasekera
3 and
Emmanuel Peprah
4
1
Center “Chiara Gemmo and Elio Zago” for the Research on Thalassemia, Department of Life Sciences and Biotechnology, Ferrara University, 40124 Ferrara, Italy
2
Department of Hematology and Blood Transfusion, Ahmadu Bello University Teaching Hospital Zaria, Kaduna 810001, Nigeria
3
Department of Anatomy, Genetics and Biomedical Informatics, Faculty of Medicine, University of Colombo, Colombo P.O. Box 271, Sri Lanka
4
Implementing Sustainable Evidence-Based Interventions through Engagement (ISEE) Lab, Department of Global and Environmental Health, School of Global Public Health, New York University, New York, NY 10003, USA
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(8), 4263; https://doi.org/10.3390/ijms25084263
Submission received: 4 March 2024 / Revised: 30 March 2024 / Accepted: 4 April 2024 / Published: 12 April 2024

Abstract

:
In this short review we have presented and discussed studies on pharmacogenomics (also termed pharmacogenetics) of the drugs employed in the treatment of β-thalassemia or Sickle-cell disease (SCD). This field of investigation is relevant, since it is expected to help clinicians select the appropriate drug and the correct dosage for each patient. We first discussed the search for DNA polymorphisms associated with a high expression of γ-globin genes and identified this using GWAS studies and CRISPR-based gene editing approaches. We then presented validated DNA polymorphisms associated with a high HbF production (including, but not limited to the HBG2 XmnI polymorphism and those related to the BCL11A, MYB, KLF-1, and LYAR genes). The expression of microRNAs involved in the regulation of γ-globin genes was also presented in the context of pharmacomiRNomics. Then, the pharmacogenomics of validated fetal hemoglobin inducers (hydroxyurea, butyrate and butyrate analogues, thalidomide, and sirolimus), of iron chelators, and of analgesics in the pain management of SCD patients were considered. Finally, we discuss current clinical trials, as well as international research networks focusing on clinical issues related to pharmacogenomics in hematological diseases.

1. Introduction

β-thalassemias and Sickle-cell disease (SCD) are the most important inherited genetic disorders affecting the hematopoietic system [1,2]. The phenotype of β-thalassemia is primarily characterized by a low or absent production of adult hemoglobin (HbA) due to low or absent production of β-globin, leading to an excess of free α-globin chains with associated ineffective erythropoiesis [1]. This β-thalassemia phenotype is caused by more than 350 mutations of the adult β-globin gene [1,3,4,5]. The phenotype of SCD is characterized by the presence of a mutated β-globin chain with the replacement (due to a mis-sense point mutation of the β-globin gene) of glutamic acid with valine in the sixth position of the protein, originating the so-called Sickle hemoglobin (HbS) [6]. In hypoxic conditions, HbS polymerizes and interacts with the cytoskeleton, causing the abnormal/Sickle-shaped morphology of red blood cells [7]. Relevant to this review, it should be noted that β-thalassemias and SCD usually require lifelong treatments, some of which are similar (for instance, the induction of fetal hemoglobin -HbF) [8]. In this context, it is very important to consider that, in addition to the primary mutation(s) causing β-thalassemias and SCD, they also exhibit single nucleotide polymorphisms (SNPs) in the genome that can predict the response to the high number of therapeutic treatments that these patients need to face during their lifetime [9]. In this respect, the activity of international organizations devoted to facilitate the translation of the use of these genomic SNPs’ testing into actionable prescribing decisions for related drugs should be monitored [10,11,12]. One example of such an organization is the Clinical Pharmacogenetics Implementation Consortium (CPIC) [12]. In Figure 1, the interplay between the therapeutic strategies for β-thalassemia and SCD and these organizations is depicted.
Pharmacogenomics (also termed pharmacogenetics) [13,14,15] and pharmacoepigenomics [16,17] are fields of applied research that study how the genome and the epigenome affect the response of patients (diagnosed with diseases including β-thalassemia and SCD) to pharmacologic treatment [13,18,19]. Pharmacogenomics studies include (but are not limited to) how variation of the human genome affects differential drug responses [16], whereas pharmacoepigenomics studies deal with the influence that epigenetic alterations (including DNA methylation, histone modifications, chromatin remodeling, and altered expression of non-coding RNAs) exert on drug efficacy and safety [17]. Moreover, pharmacogenomics and pharmacoepigenomics studies also include the effects that drug treatments may have on the genome and or the epigenetic machinery [16,17]. This review is mainly focused on pharmacogenomics and only a few issues (for instance, concerning the impact of non-coding RNAs) related to pharmacoepigenomics have been considered.
The aim of pharmacogenomic research is to achieve optimum care for each individual; thus, pharmacogenomics research helps clinicians select the appropriate drug and the correct dosage for each patient [18,19,20]. Pharmacogenomics can play an important role in differentiating, even among large cohorts of patients, those who are responders and non-responders to a given drug, avoiding adverse events, by selecting the most suitable drug dosage (e.g., genotype-specific dosing) [21]. The variability of biomarkers among patients is a factor to be considered to obtain, for single protocols, the highest probability to reach positive effects on the highest number of treated patients [22]. This research activity involves interdisciplinary collaborations between researchers working in genetics, molecular biology, pharmacology, pharmaceutics, and organic chemistry and it is expected to discover useful information on the mechanisms of drug action (e.g., pharmacodynamics and pharmacokinetics), the identification of novel molecular targets, and the design of novel biomedicines [3,4,5]. Furthermore, the clinical application for the prediction of toxicity risk or beneficial effects could be helpful in numerous ways for several pathologies, including hemoglobinopathies [13,20,23,24,25,26,27].
This aspect of the “Pharmacogenomics” endeavor has a clear impact on personalized treatments in precision medicine for patients carrying β-hemoglobinopathies, such as β-thalassemia and SCD [25,26,27,28,29,30]. In fact, with the help of pharmacogenomics studies, the administration of drugs and drug combinations could be optimized on the basis of the genetic characteristics of each individual [25,26,27].
On the other hand, the changes in the “omics” profile (for instance global methylation, global gene expression, and/or microRNA overall expression), following treatment with drugs employed for the therapy of β-thalassemia, might help in developing further protocols when considered together with the genetic and functional background affecting drug response [31,32,33]. In Figure 2, different therapeutic options for β-thalassemia and SCD are considered for which pharmacogenomics information is already available or is expected in the near future.
A simple flowchart summarizing the key points of the pharmacogenomics of HbF inducers is shown in Figure 3.
In this context, two fields of investigation have been highly impactful in understanding the mechanism of action of drugs for β-thalassemia and SCD, as well as the patient-to-patient variation in the response to treatment, as follows: (a) the identification of DNA polymorphisms that predict the response to drug treatment [34] and (b) the post-transcriptional regulation of globin gene expression by non-coding RNAs, including microRNAs [35,36,37]. In vitro, ex vivo, and in vivo studies were deemed to be necessary for a comprehensive understanding of the impact of pharmacogenomics in the management of β-thalassemia patients [38,39,40,41,42,43,44]. In particular, the field of RNA therapeutics (including microRNA therapeutics) should be considered with great attention, because some products from this research field have already reached clinical applications. For instance, inclisiran, a chemically synthesized small interfering RNA targeting PCSK9 messenger RNA [45], has been shown to reduce low-density lipoprotein (LDL) cholesterol levels in clinical trials [46,47]. Interestingly, and of relevance in the context of this review, cardiovascular complications and risk of death are well-known in clinical issues in SCD [48,49]. In this short review, we will first focus on the pharmacogenomics of inducers of fetal hemoglobin (HbF) in erythroid cells from β-thalassemia patients. The increased expression of γ-globin genes and a high HbF induction in β-thalassemia patients are associated with a milder or even asymptomatic disease pattern [50,51,52]. In this case, a transfusion regimen and chelation therapy might not be necessary. In addition to β-thalassemia, the high-level production of HbF could be highly beneficial for patients with SCD [53,54]. Despite the fact that the clinical phenotypes in SCD are considerably heterogeneous, a reduction in pain episodes and avascular necrosis of the femoral head (that are both common clinical features of SCD) are reduced in SCD patients exhibiting high levels of HbF.
In consideration of the role of HbF production in β-thalassemia and SCD, efficient HbF inducers were developed for the treatment of β-thalassemia patients characterized by a low production of HbF [55,56,57,58,59]. Despite the high and growing numbers of HbF inducers, clinical trials are ongoing for a limited number of them, as depicted in Table 1 [60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82].

2. The Search of DNA Polymorphisms Associated with High Expression of γ-Globin Genes: GWAS Studies and CRISPR-Based Gene Editing Approaches

Historically, the most important contribution to the identification of DNA polymorphisms associated with high HbF production came from genome-wide association studies (GWAS) [83,84,85,86,87,88,89]. For instance, GWAS were a key approach to identify BCL11A as a candidate repressor of the expression of γ-globin genes [85]. In addition, GWAS-identified single nucleotide polymorphisms (SNPs) have been mapped within the erythroid enhancer of the BCL11A gene [84]. This was a key contribution to the development of the pharmacogenomics of β-thalassemia, especially after the development of gene editing protocols that were able to demonstrate that the disruption of discrete portions of the BCL11A enhancer are associated with a reduction in the expression of BCL11A and an increase in γ-globin gene expression and HbF production [90,91,92]. More recently, Antoniou et al. have used base editing to generate a large variety of mutations within the −200 region of the γ-globin gene promoters [91]. Most of the mutations were found to deeply alter the binding of transcriptional repressors to the promoter. Editing of hematopoietic stem/precursor cells HSPC isolated from patients was able to induce HbF reactivation, rescuing the pathological phenotype [91]. Therefore, β-thalassemia patients carrying polymorphic DNA sequences causing these clinically relevant effects are expected to respond to some Hb inducers with a high efficiency. This can be studied both ex vivo (using erythroid progenitor cells isolated from β-thalassemia or SCD patients) and in vivo (treating β-thalassemia or SCD patients with the HbF inducer and verifying their response to the therapeutic intervention) (see Figure 3).
The interplay between GWAS, gene editing, and pharmacogenomics is further described in Figure 4, which also underlines that fully integrated procedures might help in the identification of novel targets, in the development of novel drugs, and, consequently, in the identification and characterization of novel DNA polymorphisms.

3. Major DNA Polymorphisms Associated with High Expression of γ-Globin Genes

Several reports concurrently demonstrate that the response to HbF inducers might depend on HbF-associated polymorphisms (e.g., HBG2 XmnI, BCL11A, and MYB polymorphisms) [93,94], and patient stratification based on these genetic characteristics might be useful for clinical management and the choice of the therapeutic protocol. For instance, Roy et al. studied the influence of BCL11A, HBS1L-MYB, and HBBP1 single nucleotide polymorphisms (in addition to the HBG2 XmnI polymorphism) on the levels of endogenous production of HbF and found a significant association of HbF levels with rs2071348 (BCL11A) and rs4895441 (HBS1L-MYB), respectively. The HBG2 XmnI polymorphism showed the strongest association [76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95]. An extensive description of SNPs associated with high production of HbF is outside the objectives of the present review that is mainly focused on the most relevant (in our opinion) DNA polymorphisms considered in published pharmacogenomic studies on β-thalassemia and SCD. However, important studies and review articles on DNA polymorphisms associated with the high expression of γ-globin genes and HbF production are available in the literature, such as (among others) those published by Pace B and Zein [96], Biswas et al. [34], Stadhouders et al. [97], Patrinos et al. [25], Karamperis et al. [26] and Mnika et al. [28].

3.1. The HBG2 XmnI rs7482144 Polymorphism

The HBG2 XmnI polymorphism is among the most studied and characterized genetic variants. Furthermore, this polymorphism was found to be strongly associated with HbF levels in β-thalassemia patients and with the response to HbF inducers [98]. For instance, Tien Nguyen et al. found that in β-thalassemia intermedia patients with various genotypes, the XmnI (G)gamma polymorphism at position −158 of the HBG2 promoter displayed a deep effect on fetal hemoglobin levels [99]. Furthermore, the XmnI rs7482144 polymorphism helps in the genetic prediction for age at onset and life expectancy of β-thalassemia patients [100]. In addition, this polymorphism has been associated with high levels of HbF in Sickle-cell anemia [101,102].

3.2. The LYAR rs368698783 (G>A) Polymorphism: Relevance and Structural Association with HBG2 XmnI

Recently, Ju et al. [103] identified a putative novel regulator of γ-globin gene transcription, LYAR (human homologue of mouse Ly-1 antibody reactive clone). The LYAR DNA-binding motif (5′-GGTTAT-3′) was identified in a DNA region corresponding to the 5′-untranslated region of the Aγ-globin gene. LYAR is a strong repressor of human γ-globin gene transcription, suggesting this is a novel transcription factor essential for γ-globin gene silencing [104]. Interestingly, Bianchi et al. [105] and Breveglieri et al. [106] found that the rs368698783 (G>A) polymorphism is present in β-thalassemia patients [105,106]. Moreover, molecular docking simulations, the electrophoretic mobility shift assay (EMSA), and Biacore analyses demonstrated that the rs368698783 (G>A) polymorphism decreases the LYAR binding efficiency to the Aγ-globin gene, confirming that it might affect cellular responses to HbF inducers [107]. Using rapamycin as an HbF inducer and erythroid precursor cells (ErPCs) isolated from β-thalassemia patients as an experimental model system, Zuccato et al. found that rs368698783 (G>A) is associated with a high induction of HbF [108].

3.3. The BCL11A rs2071348 Polymorphisms

The rs2071348 polymorphisms, together with the HBG2 XmnI rs7482144 polymorphism, help in the genetic prediction for age at onset and life expectancy [79]. Interestingly, Giannopoulou et al., studying a group of β-thalassemia major (β-TM) patients (severe phenotype) and β- thalassemia intermedia (β-TI) patients (mild phenotype), found that the rs2071348 polymorphism was associated with higher HbF levels and a milder β- thalassemia disease phenotype [109].

3.4. The KLF1 rs3817621, rs2072597, and rs2072596 Polymorphisms

The erythroid-specific Krupple-like factor 1(KLF1) is a transcription factor that regulates the beta-like globin gene expression, and variations within the gene can alter the SCD phenotype [110]. KLF1 is the master regulator of the BCL11A gene and also mediates hemoglobin switching [111,112]. Of the three novel SNPs (rs3817621, rs2072597, and rs2072596) identified in 118 SCD patients in central India, no statistically significant association of any of the SNPs identified in the polymorphisms was seen with HbF levels, as well as the SCD-related morbidities [111]. However, rs3817621 was reported in association with fetal hemoglobin in Thai hemoglobin E (HbE)/E β-thalassemia [113] and HbF in β-thalassemia carriers and normal individuals of Portuguese origin [114]. Although rs2072597 was not found to be associated with SCD patients in central India, the same SNP was found to be associated with HbE/β-thalassemia [95]. Finally, rs2072596 is associated with high HbA2 levels in Iranian thalassemia patients [115], but not in patients from Saudi Arabia [116].

3.5. Other Genetic Polymorphisms

Other genetic variants have been studied and their characterization has been included in several reports. For instance, genomic variants of the MAP3K5, NOS2A, and ARG2 genes were associated with hydroxyurea (HU) therapy efficacy [96]. In the same study, it was also reported that the FLT1 and ARG2 genomic variants were associated with a mild phenotype of NTDT patients. The findings reported by Kolliopoulou et al. [117] provide evidence that the MAP3K5, NOS2A, ARG2, and FLT1 genomic variants could be considered as genomic biomarkers to predict HU therapy efficacy in Hb S-β-thalassemia compound heterozygotes and also to describe disease severity in patients with β-type hemoglobinopathies [117].

4. Expression of microRNAs Involved in the Regulation of γ-Globin Genes: Evidence Supporting PharmacomiRNomics

4.1. Non-Coding RNAs and HbF Production

Non-coding RNAs (ncRNAs) are important regulators of gene expression in eukaryotes [118,119] and have been demonstrated to be involved in β-thalassemia [120,121,122] and SCD [123,124]. This is a rapidly evolving field of investigation. The most studied class of non-coding RNAs are microRNAs, a class of short (less than 30 nucleotides in length) ncRNAs with a very important role in the regulation of gene expression at the post-transcriptional level [125,126,127] through sequence-specific interactions with the 3′UTR of target mRNAs. This causes translational repression or mRNA degradation [126]. Several miRNA/mRNA networks are responsible for the control of important biological functions, such as cell growth, apoptosis, and differentiation [127], and their alterations are associated with the onset and/or progression of several human pathologies [128,129,130].
In the context of RNA- and DNA-based therapeutics, including microRNA therapeutics, the issue of delivery should be mentioned, since it is or paramount relevance in bringing laboratory research to the clinic, greatly facilitating the approval of therapeutic protocols by the regulatory agencies, such as the European Medicines Agency (EMA) in Europe and the Food and Drug Administration (FDA) in theUS [131]. This issue has been considered in excellent reviews [131,132,133]. In these review articles, polymeric nanoparticulate delivery systems are extensively described, such as lipid-based nanoparticles, polymeric micelles, dendrimers, and polymer–drug conjugates [131,132,133,134,135,136]. These systems offer stability, protection, and controlled release [131,133]. In addition, they are expected to be ready for clinical applications when important traits are fully developed, such as the consistent bioactivity across batches; the possibility to be re-dosable, without changes in efficacy and safety; the NHP safety at doses greater than those required for efficacy; acceptable on-/off-target delivery profile in NHPs; and the possibility to be manufactured at the human scale, with a scalable and biodegradable chemistry. These features of delivery systems have been fully considered by Paunovskaet al. [131]. An example is constituted by the Phase 1 study of MRX34, a liposomal miR-34a mimic [137]. This is of great interest in the context of our study, as the stable expression of miR-34a was found to mediate HbF induction, possibly involving STAT3 gene silencing [138].

4.2. MicroRNAs and HbF Production

The role of non-coding RNAs (including microRNAs and miRNAs) in HbF regulation in β-thalassemia and SCD has recently been investigated by many research groups [35,36,37,139,140,141,142,143,144,145,146,147,148,149,150,151]. Concerning microRNAs, Kargutkar et al. investigated the possible relationship between the expression of miRNAs and the induction of HbF in subjects (including SCD patients) treated for 3 and 6 months with hydroxyurea [35]. A differential expression of miRNAs during treatment with HU was found and was associated with the HU-mediated increase in HbF. This was particularly evident for the increase in miR-210, miR16-1, and miR-29a expression and the decrease in miR-96. As this study strongly suggests that miR-210, miR-16-1, and miR-29a are positive regulators of γ-globin gene expression and miR-96 is a negative regulator, the pattern of expression of miRNAs should be considered in further studies aimed at verifying the possibility to predict response to HU, based on the miRNome of each patient. A study by Kargutkar et al. confirms a possible role in HbF induction of miR-210, which was proposed by Gasparello et al. to target BCL11A, the major repressor of γ-globin gene expression [143]. Other examples of microRNAs involved in the regulation of γ-globin genes are miR-92a-3p [144] and miR-30a [145]. A partial list of the non-coding RNAs involved in the regulation of γ-globin genes is reported in Table 2 [143,144,145,146,147,148,149,150,151,152,153,154,155,156,157,158].
All the information available supports the concept that non-coding RNAs, including microRNAs (miRNAs), long non-coding RNAs (lncRNAs), and circular RNAs (circRNAs), are deeply involved in the regulation of γ-globin gene expression [139,140,141,142,143,144,145,146,147,148,149,150,151,152,153,154,155,156,157,158,159]. Exploring possible crosstalk between long non-coding RNAs and microRNAs will clarify their possible role as prognostic and therapeutic biomarkers in β-thalassemia [160]. Interestingly, most of the miRNA targets are mRNA coding transcription factors, regulating γ-globin gene expression [143,144,145,146,147,148,149,150,151,152,153,154,155,156,157,158,161]. The impact of non-coding RNAs in the prediction of response to HbF inducers has lead to several observations and/or comments supporting the concept of “PharmacomiRNomics” of HbF inducers (Figure 3), as already proposed by Faraoni et al. (2009) [162] in a study focusing on the variability of miR-155 expression profiles in respect to physiological and pathological processes [162]. Das et al. [163] were able to identify SNPs in pre-miRNA regions and pre-miRNA flanking regions using β-thalassemia as a model system. Most of these SNPs were expected to alter the translation process of miRNA-bound mRNAs, thereby altering HbF production, with an impact on the clinical severity of the disease [163]. Micro-RNAs miR-486 and miR-26b had the highest number of SNPs in upstream and downstream flanking regions, respectively [27]. DNA polymorphisms within the 3′UTR of target mRNAs coding for proteins regulating γ-globin gene expression (for instance BCL11A, MYB, SOX-6, and KLF-1) have also been identified. Furthermore, polymorphisms of the binding sites of microRNAs regulating γ-globin gene expression have also been found [163]. These data further sustain the concept, based on microarray and NGS studies, that a sub-set of microRNAs is associated with the HbF expression levels [159,160,161].
The impact of these data on the possible proposal of “PharmacomiRNomics” (Figure 5) is sustained by two recently published studies. In the first, Hojjati et al. studied β-thalassemia patients treated with HU and grouped them into two groups, as follows: responders and non-responders to the treatment [164]. The results obtained demonstrated that miR-210 and miR-486-3p were expressed at higher levels in the responders than in the non-responders group.
More recently, Kargutkar et al. evaluated the effect of miRNA expression on HbF induction in relation to HU therapy in normal controls, SCD patients [35]. Among the most relevant findings of this study, an increase in miR-210, miR16-1, and miR-29a expression and a decrease in miR-96 expression were strongly associated with the HU-mediated HbF induction [35].
While the results of these studies are promising, further investigations will be necessary to verify the real impact of “PharmacomiRNomics” in the management of the therapeutic options for β-thalassemia and SCD patients, with a particular focus on HbF induction.

5. Pharmacogenomics of Fetal Hemoglobin Inducers in β-Thalassemia and Sickle-Cell Disease: Ex Vivo and In Vivo Studies

5.1. Hydroxyurea

Changes in HbF levels upon HU therapy are likely to be regulated by multiple loci in both β-thalassemia and SCD [165]. The identification and characterization of these SNPs are important, as this improves the clinical management of the patients [25,166]. In the case of HU, several studies indicate a relationship between the HBG2 XmnI polymorphism and the induction of HbF, as reported in different studies [165,166,167,168]; these studies are expected to stimulate further analyses of the degree of response of transfusion-dependent β-thalassemia patients to HU therapy in order to develop a shared diagnostic protocol to identify patients likely to have a significant response [169]. In this respect, Gosh et al. [170]. found that HU therapy conducted in HbE/β-thalassemia-major patients was able to improve the transfusion-free interval and was correlated with the HBG2 XmnI polymorphism [170]. Other SNPs associated with HbF induction should be considered in the development of a final diagnostic protocol for the prediction of HU response in β-thalassemia and SCD. In this respect, Giannopoulou et al. reported that the rs2071348 polymorphism in the HBBP1 gene does not correlate with response to HU treatment [109]. Relevant to this topic is a systematic review focused on seven studies involving SCD patients treated with HU; 50 SNPs of 17 different genes were identified and found to be associated with HbF changes following HU monotherapy (Sales et al., 2021) [166]. The authors confirmed that HU-mediated changes in HbF levels in response to HU therapy are regulated by genetic variations on multiple loci [166]. A second conclusion gathered from the studies reviewed by Sales et al. is that in those studies, there was strong evidence that SNPs located in intron 2 of the BCL11A gene affect HbF changes in SCD patients treated with HU [166]. In conclusion, in spite of the variability of clinical protocols and the heterogeneity of patients studied, the available information strongly supports the concept that pharmacogenomics studies on HbF induction in response to HU are of potential application in improving the clinical management of β-thalassemia and SCD.

5.2. Butyrate and Butyrate Analogues

Butyrates are a class of histone deacetylase (HDAC) inhibitors demonstrated to be able to increase γ-globin gene expression and HbF content in erythroid cells from β-thalassemic and SCD patients [171,172,173]. It has been reported that at least 25% of the patients affected by hemoglobinopathies do not respond to treatment with butyrates [174,175]. However, despite the fact that butyrates have been proposed as good candidate drugs for pharmacogenomic studies [174], few reports have been published on this issue. An interesting study is based on the observation that patients who are unresponsive to HU treatment can be treated with sodium butyrate containing microRNAs to induce HbF synthesis [174,176].

5.3. Thalidomide

A meta-analysis has recently been published on the efficacy of thalidomide in the induction of HbF in thalidomide-treated β-thalassemia patients [177]. This study was unable to detect any statistically significant relationships between the efficacy of thalidomide in transfusion-dependent β-thalassemia patients and the XmnI polymorphism [177]. In another study, Yang et al. considered the relevance of HBG2 (rs7482144), BCL11A (rs11886868, rs4671393, rs766432, and rs1427407), and HBS1L-MYB (rs9399137, rs4895440, and rs4895441) single nucleotide polymorphisms (SNPs) in thalidomide response [178]. This study demonstrated that SNPs in HBG2 and HBS1L-MYB contributed significantly to thalidomide response in Chinese patients with β-thalassemia and that the cumulative number of minor SNP alleles may serve as good predictors of treatment response in this population [178]. In conclusion, pharmacogenomics of HbF induction in response to thalidomide should be considered in further studies on β-thalassemia and SCD patients in order to confirm the contribution of the identified SNPs in HBG2 and HBS1L-MYB in the response to thalidomide and the hypothesis of a lack of correlation between response to thalidomide and the HBG2 XmnI polymorphism.

5.4. Rapamycin (Sirolimus)

Recently, Zuccato et al. reported a study demonstrating that the LYAR rs368698783 (G>A) polymorphism is associated with a high basal and induced production of HbF in β-thalassemia patients [108]. The most striking association was found using rapamycin (sirolimus) as an HbF inducer. The published results are important not only for basic biomedicine, but also in applied translational research for precision medicine in the personalized therapy of β-thalassemia [108]. Accordingly, the rs368698783 polymorphism might be considered among the parameters useful to recruit patients with the highest probability of responding to in vivo rapamycin treatment.

6. Pharmacogenomics of Iron Chelators in β-Thalassemia and Other Hematopoietic Diseases Needing Blood Transfusions

As extensively discussed in several review papers, the management of β-thalassemia patients needing blood transfusion (TDT, transfusion-dependent thalassemia) is based on the use of an iron-chelation therapy in order to eliminate the iron excess associated with blood transfusion. Pharmacogenomics can greatly help clinicians to identify the best iron chelator to be used, considering a number of well-defined genetic polymorphisms, as discussed in the following examples, concerning the two most studied iron chelators, Deferasirox and Deferiprone.

6.1. Deferasirox

Deferasirox (DFX) is an oral iron chelator for the management of chronic iron overload in patients with chronic anemias who require blood transfusions [179]. The available literature suggests the safety and effectiveness of DFX, but a rather heterogenous response in treating β-thalassemia patients [180,181]. The efficacy of DFX in other chronic transfusion-requiring diseases is, at present, extensively studied with promising results. Single nucleotide polymorphisms in genes involved in DFX metabolism (such as UGT1A1, UGT1A3, CYP1A1, CYP1A2, and CYP2D6) and elimination (such as MRP2 and BCRP1) have been investigated, in order to determine whether some of them were found to be associated with DFX plasma concentration and, consequently, the response of the patients to DFX. The most studied polymorphisms are the UGT1A1C>T rs887829, CYP1A1C>A rs2606345, CYP1A2A>C rs762551, CYP1A2C>T rs2470890, and MRP2G>A rs2273697 polymorphisms. The CYP1A1C>A rs2606345 AA and CYP1A2C>T rs2470890 TT genotypes were found to be useful in the prediction of low response of the patients to DFX therapy [182,183].

6.2. Deferiprone

Deferiprone is a bidentate oral iron chelator. Deferiprone is metabolized into deferiprone-3-0-glucuronide (DG) by the enzyme UDP-glucuronosyl transferase (UGT). UGT belongs to four gene families, UGT1, UGT2, UGT3 and UGT8 [184,185]—Among several isoforms of UDPglucuronosyltransferases, UGT1A, UGT2B, and UGT1A9 are responsible for the glucuronidation of deferiprone [184]. Two variants, −2152 C>T and −275 T>A, in the promoter of the UGT1A9 gene are associated with the increased expression of UGT IA9. Therefore, the response to deferiprone-based therapy may be influenced by those genetic variants. It has been known that there are three commonly known polymorphisms in UGT1A6, which are Thr181Ala (541 A>G), Arg 184 Ser (552 A>C), and Ser 7Ala (19 T>G); these are responsible for the efficacy of deferiprone treatment [185].

7. Pharmacogenomics of Analgesics in Pain Management of SCD Patients

Pain is the hallmark of SCD, leading to recurrent admissions to the emergency department. The most common way of combating it is through admissions of analgesics such as NSAIDs and opioids and, in some instances, adjuvants. The cytochrome P450 superfamily, in which CYP2C8 and CYP2C9 belong, plays a major role in their metabolism. Both enzymes have been associated with a decreased enzyme activity alteration in their pharmacokinetics [186,187].
More than 16 alleles and over 60 variants have been characterized. These allelic variants impact their metabolic activity leading to classification into ultra-rapid responders, extensive metabolizers, intermediate metabolizers, and poor metabolizers. Most analgesic dosing and pain plans assume patients to be extensive metabolizers (i.e., having two functional alleles). However, sub-optimal responses and adverse side effects have been demonstrated in individuals having varying polymorphism at CYP2C8 and CYP2C9 [188]. In the context of analgesics, CPIC guidelines and the FDA drug labels are available [189,190].

7.1. Pharmacogenomics of Opioids

Opioids are of interest for the treatment of pain in SCD patients, as they are among the first choices for the management of the patients, with respect to this very important clinical complication. Accordingly, chronic opioid therapy might be, in some cases, considered, especially in the case of SCD patients who are debilitated by chronic pain. The pharmacogenomics of opioids have been considered in a review article by Husain et al. [189], who briefly reviewed the literature on pharmacogenomics in SCD and possible targets for intervention in the context of possible steps on the road to personalized medicine, considering that the SCD patients exhibit a high variability in response to opioids. Polymorphisms in the COMT, OPRM1, and ABCB1 genes can lead to an altered perception of pain and/or a change in response to opioids by the SCD patients. This information can be found in the studies reported by Joly et al. [190] and Jhun et al. [191]. Joly’s study is relevant, since it was conducted in a significant cohort of SCD patients [190]. With respect to pharmacokinetics, important SNPs have been genotyped for the CYP2D6 gene (codeine to morphine conversion) and for three genes involved in morphine elimination (CYP3A, UGT2B7, and ABCB1), but more pharmacokinetic studies on phenotype/genotype correlations are needed to reach definitive conclusions [190]. The pharmacogenomics of morphine have recently been discussed [192]. The role of pharmacogenomics in prescribing opioids has been discussed by Wong et al. [193], discussing the situation concerning cancer patients. In this case, pharmacogenomic evidence guiding opioid prescription is currently available for codeine and tramadol, which relates to the CYP2D6 gene variants. Moderate evidence has been reported for oxycodone (which relates to CYP2D6) and methadone (which relates to CYP2B6 and ABCB1) [193]. The experience in cancer will probably guide the research efforts concerning SCD, as demonstrated by the study by Gammal et al. [194], who discussed the role of pharmacogenomics (in particular focusing on the CYP2D6 genotype for the safe use of codeine in SCD). This study is particularly relevant, considering the adverse effects of postoperative deaths in children who were prescribed codeine [194].

7.2. Inducers of Nitric Oxide in SCD

Nitric oxide (NO) deficiency has been shown to play an important role in the development of vaso-occlusive crises, a characteristic feature of SCD [188]. Accordingly, clinical trials and studies have suggested the importance of artificial NO induction in relation to a reduction in the severity of the disease [195]. In this respect, Weiner et al. [196] reported a preliminary study suggesting that inhaled NO may be beneficial for acute vaso-occlusive crises in SCD [196]. Since endothelial Nitric Oxide Synthase (eNOS) plays a major role in regulating the NO levels in the body, studies have been focused on the polymorphism(s) of the eNOS gene, in order to verify whether they can be predictive of the severity of SCD [197]. In this respect, several SNPs of the eNOS gene were found in SCD patients to be associated with NO levels and the severity of the disease [198,199]. For example, rs2070744 was found in the promoter region and rs1799983 within exon 7 (894 G>T, Glu-298Asp) and the frequency of these polymorphisms was particularly high in patients with severe SCD, associated with lower levels of plasma NO [199]. NO inducers should be considered as therapeutic tools in SCD [200,201]. Interestingly, evidence for the in vivo conversion of hydroxyurea to nitric oxide has been provided [202].

8. Guidelines and Clinical Trials Focusing on Pharmacogenomics/Pharmacogenetics in β-Thalassemia and SCD

Although knowledge of the association of SNPs on pharmacodynamics and pharmacokinetics has increased, as of 2016, the issue of utilizing pharmacogenomics for β-thalassemia and SCD remains contentious. In parallel, several activities are ongoing to help clinicians follow the best possible management of β-thalassemia and SCD patients. For instance, some of the pharmacological interventions that are used to manage SCD have specific guidelines created by the Clinical Pharmacogenetics Implementation Consortium (CPIC) [203]. The CPIC evidence-based guidelines can improve outcomes by utilizing guide dosing [30]. The CPIC publishes genotype-based drug guidelines to help clinicians understand how available genetic test results could be used to optimize drug therapy [203]. Using electronic medical records, Gallaway and colleagues found that CPIC guidelines could have been used during a 16 year (2005 to 2021) period, in which 892 (93%) patients at Indiana University received SCD care [30]. A total of 814 (85.1%) patients received two drugs with CPIC guidelines, 654 (68.3%) received three drugs, 355 (37.1%) received four drugs, 157 (16.4%) received five drugs, 63 (6.6%) received six drugs, 21 (2.2%) received seven drugs, and one (0.1%) received eight drugs [16]. In total, 65 (6.8%) patients received no drugs with CPIC guidelines [8]. Nevertheless, Gallaway and colleagues highlight that SCD patients had a high utilization of drugs with CPIC guidelines [30] and received an average of three unique CPIC drugs/drug classes per person. Table 3 summarizes examples of pharmacogenomics gene–drug association, using information available from the CPIC and FDA concerning drugs employed or proposed for SCA and β-thalassemia therapy [203].
Although many acknowledge that greater progress has been made toward identifying the key genomic variants, mainly in BCL11A, HBS1L-MYB, or SAR1, which contribute to the response to HU treatment [28], some argue that the complete picture of pharmacogenomic/genetic determinants of therapeutic phenotypes remains elusive [28]. Moreover, no large-scale studies have been conducted in sub-Saharan Africa, where the majority of patients with SCD live to elucidate the phenotype/genotype relationships [16]. In contrast, others have suggested tailor-made “theranostics” in SCD as the way forward, able to direct optimal clinical management of the patients [203,204,205,206,207,208].
To assess the current state of clinical trials, we use the US Clinical Trials website. Although this website contains comprehensive information for trials in the US, trials in many countries are not captured in this dataset. After a first search using the ClinicalTrials.gov database (https://www.clinicaltrials.gov, accessed on 25 March 2024), 279 studies were found for β-thalassemia and 972 for SCD. Of these studies, 7–10% (23 studies for β-thalassemia and 72 for SCD) indicated HbF production as an endpoint. When drugs, here reviewed, were considered, most of the studies (15 studies for β-thalassemia and 172 for SCD) were based on HU. No study was primarily focused on “pharmacogenomics in β-thalassemia or SCD”. We can conclude that the design and activation of this field of clinical investigation is still in a preliminary phase.
In the current landscape of trials, 199 trials match the “sickle cell” text search in the ClinicalTrials.gov database that are currently recruiting participants; of those, 44 trials are industry funded, with other trials being a combination of government, academia, and other organizations. The only ongoing clinical trial that matched the search terms “Pharmacogenomic” and “Sickle Cell” in the NIH Clinical Trials database (conducted in June 2023) was the Therapeutic Response Evaluation and Adherence Trial (TREAT). No trials matched the search terms “Pharmacogenetic” and “Sickle Cell”. The TREAT clinical trial is funded by the US NIH and located at Cincinnati Children’s Hospital Medical Center and has the following objectives: (1) to develop and prospectively evaluate a population pharmacokinetic/pharmacodynamics model to predict the maximum tolerated dose (MTD); (2) to identify urine biomarkers of HU adherence using a novel metabolomics approach; (3) to identify pharmacogenomics modifiers of HU MTD; and (4) longitudinal monitoring of the effect of HU upon organ function and quality of life (https://clinicaltrials.gov/ct2/show/NCT02286154, accessed on 25 March 2024) and recruitment is still ongoing. The TREAT recruitment cohort consists of individuals aged from 6 months to 21 years, diagnosed with SCA (HbSS or Hbβ0-thalassemia) and will consist of patients who initiate HU therapy, including patients who are transitioning from chronic transfusions to HU therapy. The enrolled participants will receive HU; however, upon enrollment, participants will be grouped as a “New Cohort” or “Old Cohort”. “New Cohort” participants include those who are not receiving HU therapy upon study entry; for group pharmacodynamics and pharmacokinetics, data will be used to predict the most effective maximum tolerated dose. “Old Cohort” participants include those who are already receiving HU therapy and participants will receive HU escalated to MTD based on local clinical guidelines.
A search of the NIH trials database indicates that currently, 40 clinical trials match the search terms “hydroxyurea” and “sickle cell” with currently 33 recruiting and 7 designated as not yet recruiting. Among clinical trials utilizing a pharmacogenomics/genetics approach, HU is the most studied pharmacological modifier [166]. Sales and colleagues utilized a systematic review and found 50 different SNPs of 17 genes to be associated with HbF changes in patients with SCD treated with hydroxyurea that are VEGF ligand–receptor interactions (R-HSA-194313; R-HSA195399) and the urea cycle(R-HSA-70635) [166]. Further, Sales and colleagues concluded from their systematic review that pharmacogenetic studies in patients with SCD were sparse and very heterogeneous in regard to candidate genes and SNPs examined for the association with HbF changes and outcomes, suggesting further research is needed [166].

9. Conclusions and Future Perspectives

In this short review, we have presented and discussed studies on pharmacogenomics and the response of patients (diagnosed with diseases including β-thalassemia and SCD) to pharmacologic treatment.
The issue of pharmacogenomics in hematological diseases has been considered in several concluded European Union-funded research projects, as well as in ongoing initiatives. For instance, the study of the association between DNA polymorphisms and phenotypes in β-thalassemia was considered a key issue in the ITHANET (Electronic Infrastructure for Thalassemia Research Network, FP6) [209] and THALAMOSS (Thalassemia Modular Stratification System for personalized therapy of β-thalassemia, FP7) [210] EU-funded projects. Furthermore, this field of investigation is central in two ongoing projects related to hemoglobinopathies (such as thalassemia and SCD), the COST-action CA22119–HELIOS and the INHERENT (The International Hemoglobinopathy Research Network) projects.
INHERENT is a recently established network with the aim of investigating the role of genetic modifiers in hemoglobinopathies, through a large-scale, multi-ethnic genome-wide association study (GWAS). INHERENT brings together nine existing international or regional consortia in the field of hemoglobinopathies, namely ITHANET, RADeep [211], ARISE, SPARCO, SADaCC, REDAC, the HVP Global Globin Network [212], the International Health Repository, and the ClinGen Hemoglobinopathy VCEP [213].
In particular, we expect that the large INHERENT research group will be able to sustain research efforts on the discovery/validation of known and new genetic modifiers for the pharmacogenomics of HbF inducers, as well as for other drugs employed in the management of β-thalassemia and SCD patients. In fact, the identification and characterization of a few known genetic modifiers (including those reviewed in the present study) are, as yet, insufficient to guide treatment recommendations or to reliably stratify patients. As proposed in the review article by Kountouris et al. [214], larger multi-ethnic studies are needed to identify and validate further disease modifiers that can be used for patient stratification and personalized treatment. The INHERENT network is expected to greatly help in these research efforts.
In addition, future perspectives include the use of multiple genetic modifiers to verify the response of the patients to drugs to be considered in therapy. Furthermore, we have to comment on the fact that several drugs/molecules for β-thalassemia and SCD have recently reached the market and/or are, at present, being tested in clinical trials. For most of them (for instance, mitapivat, luspatercept, and the gene editing drug CTX001), pharmacogenomic studies are not available yet (see Figure 2). Finally, up to now, most pharmacogenomic-oriented studies have been performed using single-drug treatments. In the future, we expect that combined treatments will be tested in order to reach the highest benefit possible for β-thalassemia and SCD patients. These new protocols will be considered for pharmacogenomic analyses.
A final comment is related to molecular diagnosis and, in particular, to non-invasive prenatal diagnosis (NIPD). In this field of investigation, several approaches have been tested on β-thalassemia, with the aim of identifying the primary mutations [155]. In this respect, the discovery of cell-free fetal DNA (cffDNA) in the peripheral blood of pregnant women has encouraged the research of innovative molecular protocols providing the possibility of NIPD [215]. For instance, the possibility of NIPD has been demonstrated using multiplex PCR/NGS [216], pyrosequencing [217], and droplet digital PCR [218,219,220]. In the near future, we expect that these protocols will be applied to the non-invasive prenatal identification of DNA polymorphisms associated with better outcomes of clinical treatments in β-thalassemia and SCD.

Author Contributions

Conceptualization, R.G.; methodology, E.P.; formal analysis, R.G. and E.P.; writing—original draft preparation, R.G.; writing—review and editing, A.D.W., H.G., E.P. and R.G.; supervision, R.G.; funding acquisition, R.G. All authors have read and agreed to the published version of the manuscript.

Funding

This study was sponsored by the Wellcome Trust (innovator award 208872/Z/17/Z) and AIFA (AIFA-2016-02364887). R.G. also received funding from the UE THALAMOSS Project (Thalassemia Modular Stratification System for Personalized Therapy of Βeta-Thalassemia; no. 306201-FP7-HEALTH-2012-INNOVATION-1).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We thank all colleagues of the INHERENT Network for their continuous support.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Higgs, R.; Engel, J.D.; Stamatoyannopoulo, G. Thalassaemia. Lancet 2012, 37, 373–383. [Google Scholar] [CrossRef] [PubMed]
  2. Weatherall, D.J. Phenotype-genotype relationships in monogenic disease: Lessons from the thalassaemias. Nat. Rev. Genet. 2001, 2, 245–255. [Google Scholar] [CrossRef] [PubMed]
  3. Thein, S.L. Molecular basis of β thalassemia and potential therapeutic targets. Blood Cells Mol. Dis. 2018, 70, 54–65. [Google Scholar] [CrossRef] [PubMed]
  4. Rao, E.; Kumar Chandraker, S.; Misha Singh, M.; Kumar, R. Global distribution of β-thalassemia mutations: An update. Gene 2024, 896, 148022. [Google Scholar] [CrossRef] [PubMed]
  5. Jaing, T.H.; Chang, T.Y.; Chen, S.H.; Lin, C.W.; Wen, Y.C.; Chiu, C.C. Molecular genetics of beta-thalassemia: A narrative review. Medicine 2021, 100, e27522. [Google Scholar] [CrossRef] [PubMed]
  6. Kavanagh, P.L.; Fasipe, T.A.; Wun, T. Sickle Cell Disease: A Review. JAMA 2022, 328, 57–68. [Google Scholar] [CrossRef] [PubMed]
  7. Bunn, H.F.; Noguchi, C.T.; Hofrichter, J.; Schechter, G.P.; Schechter, A.N.; Eaton, W.A. Molecular and cellular pathogenesis of hemoglobin SC disease. Proc. Natl. Acad. Sci. USA 1982, 79, 7527–7531. [Google Scholar] [CrossRef] [PubMed]
  8. May, A.; Choiseul, M. Sickle cell anaemia and thalassaemia: Symptoms, treatment and effects on lifestyle. Health Visit. 1988, 61, 212–215. [Google Scholar] [PubMed]
  9. Thein, S.L. Genetic Basis and Genetic Modifiers of β-Thalassemia and Sickle Cell Disease. Adv. Exp. Med. Biol. 2017, 1013, 27–57. [Google Scholar]
  10. Abdullah-Koolmees, H.; van Keulen, A.M.; Nijenhuis, M.; Deneer, V.H.M. Pharmacogenetics Guidelines: Overview and Comparison of the DPWG, CPIC, CPNDS, and RNPGx Guidelines. Front. Pharmacol. 2021, 11, 595219. [Google Scholar] [CrossRef]
  11. Hulshof, E.C.; Deenen, M.J.; Nijenhuis, M.; Soree, B.; de Boer-Veger, N.J.; Buunk, A.M.; Houwink, E.J.F.; Risselada, A.; Rongen, G.A.P.J.M.; van Schaik, R.H.N.; et al. Dutch pharmacogenetics working group (DPWG) guideline for the gene-drug interaction between UGT1A1 and irinotecan. Eur. J. Hum. Genet. 2023, 31, 982–987. [Google Scholar] [CrossRef] [PubMed]
  12. Bell, G.C.; Caudle, K.E.; Whirl-Carrillo, M.; Gordon, R.J.; Hikino, K.; Prows, C.A.; Gaedigk, A.; Agundez, J.; Sadhasivam, S.; Klein, T.E.; et al. Clinical Pharmacogenetics Implementation Consortium (CPIC) guideline for CYP2D6 genotype and use of ondansetron and tropisetron. Clin. Pharmacol. Ther. 2017, 102, 213–218. [Google Scholar] [CrossRef] [PubMed]
  13. Johnson, J.A. Pharmacogenetics: Potential for individualized drug therapy through genetics. Trends Genet. 2003, 19, 660–666. [Google Scholar] [CrossRef]
  14. Pirmohamed, M. Pharmacogenomics: Current status and future perspectives. Nat. Rev. Genet. 2023, 24, 350–362. [Google Scholar] [CrossRef]
  15. Patrinos, G.P.; Grosveld, F.G. Pharmacogenomics and therapeutics of hemoglobinopathies. Hemoglobin 2008, 32, 229–236. [Google Scholar] [CrossRef]
  16. Peedicayil, J. Pharmacoepigenetics and Pharmacoepigenomics: An Overview. Curr. Drug Discov. Technol. 2019, 16, 392–399. [Google Scholar] [CrossRef]
  17. López-Camarillo, C.; Gallardo-Rincón, D.; Álvarez-Sánchez, M.E.; Marchat, L.A. Pharmaco-epigenomics: On the Road of Translation Medicine. Adv. Exp. Med. Biol. 2019, 1168, 31–42. [Google Scholar] [PubMed]
  18. Patrinos, G.P.; Innocenti, F. Pharmacogenomics: Paving the path to personalized medicine. Pharmacogenomics 2010, 11, 141–146. [Google Scholar] [CrossRef]
  19. Udagawa, C.; Zembutsu, H. Pharmacogenetics for severe adverse drug reactions induced by molecular-targeted therapy. Cancer Sci. 2020, 111, 3445–3457. [Google Scholar] [CrossRef]
  20. Eichelbaum, M.; Ingelman-Sundberg, M.; Evans, W.E. Pharmacogenomics and individualized drug therapy. Annu. Rev. Med. 2006, 57, 119–137. [Google Scholar] [CrossRef]
  21. Ahangari, N.; Doosti, M.; Ghayour Mobarhan, M.; Sahebkar, A.; Ferns, G.A.; Pasdar, A. Personalised medicine in hypercholesterolaemia: The role of pharmacogenetics in statin therapy. Ann. Med. 2020, 52, 462–470. [Google Scholar] [CrossRef]
  22. Wagner, J.B. Children Are Not Small Adults: Specific Findings in Statin Exposure and Response in a Growing Population. Clin. Pharmacol. Ther. 2019, 106, 278–280. [Google Scholar] [CrossRef] [PubMed]
  23. Gummadi, A.C.; Guddati, A.K. Genetic Polymorphisms in Pharmaceuticals and Chemotherapy. World J. Oncol. 2021, 12, 149–154. [Google Scholar] [CrossRef]
  24. Del Re, M.; Cucchiara, F.; Rofi, E.; Fontanelli, L.; Petrini, I.; Gri, N.; Pasquini, G.; Rizzo, M.; Gabelloni, M.; Belluomini, L.; et al. A multiparametric approach to improve the prediction of response to immunotherapy in patients with metastatic NSCLC. Cancer Immunol. Immunother. 2021, 70, 1667–1678. [Google Scholar] [CrossRef]
  25. Patrinos, G.P.; Chui, D.H.K.; Hardison, R.C.; Steinberg, M.H. Strategies to improve pharmacogenomic-guided treatment options for patients with β-hemoglobinopathies. Expert Rev. Hematol. 2021, 14, 883–885. [Google Scholar] [CrossRef]
  26. Karamperis, K.; Tsoumpeli, M.T.; Kounelis, F.; Koromina, M.; Mitropoulou, C.; Moutinho, C.; Patrinos, G.P. Genome-based therapeutic interventions for β-type hemoglobinopathies. Hum. Genom. 2021, 15, 32. [Google Scholar] [CrossRef] [PubMed]
  27. Gravia, A.; Chondrou, V.; Sgourou, A.; Papantoni, I.; Borg, J.; Katsila, T.; Papachatzopoulou, A.; Patrinos, G.P. Individualizing fetal hemoglobin augmenting therapy for β-type hemoglobinopathies patients. Pharmacogenomics 2014, 15, 1355–1364. [Google Scholar] [CrossRef]
  28. Mnika, K.; Pule, G.D.; Dandara, C.; Wonkam, A. An Expert Review of Pharmacogenomics of Sickle Cell Disease Therapeutics: Not Yet Ready for Global Precision Medicine. OMICS 2016, 20, 565–574. [Google Scholar] [CrossRef] [PubMed]
  29. Iolascon, A.; Andolfo, I.; Russo, R. Red cells in post-genomic era: Impact of personalized medicine in the treatment of anemias. Haematologica 2015, 100, 3–6. [Google Scholar]
  30. Gallaway, K.A.; Sakon, C.; Ongeri, J.; Patel, K.S.; Oliver, J.; Patacca, H.; O’Brien, A.R.; Skaar, T.C.; Tillman, E.M. Opportunity for pharmacogenetics testing in patients with sickle cell anemia. Pharmacogenomics 2022, 23, 925–931. [Google Scholar] [CrossRef]
  31. Brewin, J.; Tewari, S.; Menzel, S.; Kirkham, F.; Inusa, B.; Renney, G.; Ward, M.; Rees, D.C. The effects of hydroxycarbamide on the plasma proteome of children with sickle cell anaemia. Br. J. Haematol. 2019, 186, 879–886. [Google Scholar] [CrossRef]
  32. Zohaib, M.; Ansari, S.H.; Shamsi, T.S.; Zubarev, R.A.; Zarina, S. Pharmacoproteomics Profiling of Plasma from β-Thalassemia Patients in Response to Hydroxyurea Treatment. J. Clin. Pharmacol. 2019, 59, 98–106. [Google Scholar] [CrossRef] [PubMed]
  33. Lance, E.I.; Faulcon, L.M.; Fu, Z.; Yang, J.; Whyte-Stewart, D.; Strouse, J.J.; Barron-Casella, E.; Jones, K.; Van Eyk, J.E.; Casella, J.F.; et al. Proteomic discovery in sickle cell disease: Elevated neurogranin levels in children with sickle cell disease. Proteom. Clin. Appl. 2021, 15, e2100003. [Google Scholar] [CrossRef] [PubMed]
  34. Biswas, S.; Nag, A.; Ghosh, K.; Ray, R.; Roy, K.; Bandyopadhyay, A.; Bhattacharyya, M. Genetic determinants related to pharmacological induction of foetal haemoglobin in transfusion-dependent HbE-β thalassaemia. Ann. Hematol. 2019, 98, 289–299. [Google Scholar] [CrossRef] [PubMed]
  35. Kargutkar, N.; Sawant-Mulay, M.; Hariharan, P.; Chandrakala, S.; Nadkarni, A. Role of microRNA in hydroxyurea mediated HbF induction in sickle cell anaemia patients. Sci. Rep. 2023, 13, 369. [Google Scholar] [CrossRef] [PubMed]
  36. Starlard-Davenport, A.; Gu, Q.; Pace, B.S. Targeting Genetic Modifiers of HBG Gene Expression in Sickle Cell Disease: The miRNA Option. Mol. Diagn. Ther. 2022, 26, 497–509. [Google Scholar] [CrossRef] [PubMed]
  37. Okeke, C.; Silas, U.; Nnodu, O.; Clementina, O. HSC and miRNA Regulation with Implication for Foetal Haemoglobin Induction in Beta Haemoglobinopathies. Curr. Stem Cell Res. Ther. 2022, 17, 339–347. [Google Scholar] [CrossRef] [PubMed]
  38. Muralidhar, S.A.; Ramakrishnan, V.; Kalra, I.S.; Li, W.; Pace, B.S. Histone deacetylase 9 activates gamma-globin gene expression in primary erythroid cells. J. Biol. Chem. 2011, 286, 2343–2353. [Google Scholar] [CrossRef]
  39. Zein, S.; Li, W.; Ramakrishnan, V.; Lou, T.F.; Sivanand, S.; Mackie, A.; Pace, B. Identification of fetal hemoglobin-inducing agents using the human leukemia KU812 cell line. Exp. Biol. Med. 2010, 235, 1385–1394. [Google Scholar] [CrossRef] [PubMed]
  40. 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.; et al. 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]
  41. Perrine, S.P.; Mankidy, R.; Boosalis, M.S.; Bieker, J.J.; Faller, D.V. Erythroid Kruppel-like factor (EKLF) is recruited to the gamma-globin gene promoter as a co-activator and is required for gamma-globin gene induction by short-chain fatty acid derivatives. Eur. J. Haematol. 2009, 82, 466–476. [Google Scholar] [CrossRef]
  42. Zhou, G.; Zhang, H.; Lin, A.; Wu, Z.; Li, T.; Zhang, X.; Chen, H.; Lu, D. Multi-Omics Analysis in β-Thalassemia Using an HBB Gene-Knockout Human Erythroid Progenitor Cell Model. Int. J. Mol. Sci. 2022, 23, 2807. [Google Scholar] [CrossRef] [PubMed]
  43. Gotardo, É.M.F.; Brito, P.L.; Gushiken, L.F.S.; Chweih, H.; Leonardo, F.C.; Costa, F.F.; Conran, N. Molecular and cellular effects of in vivo chronic intravascular hemolysis and anti-inflammatory therapeutic approaches. Vasc. Pharmacol. 2023, 150, 107176. [Google Scholar] [CrossRef] [PubMed]
  44. Pradhan-Sundd, T.; Kato, G.J.; Novelli, E.M. Molecular mechanisms of hepatic dysfunction in sickle cell disease: Lessons from Townes mouse model. Am. J. Physiol. Cell Physiol. 2022, 323, C494–C504. [Google Scholar] [CrossRef] [PubMed]
  45. Fitzgerald, K.; White, S.; Borodovsky, A.; Bettencourt, B.R.; Strahs, A.; Clausen, V.; Wijngaard, P.; Horton, J.D.; Taubel, J.; Brooks, A.; et al. A Highly Durable RNAi Therapeutic Inhibitor of PCSK9. N. Engl. J. Med. 2017, 376, 41–51. [Google Scholar] [CrossRef] [PubMed]
  46. Ray, K.K.; Landmesser, U.; Leiter, L.A.; Kallend, D.; Dufour, R.; Karakas, M.; Hall, T.; Troquay, R.P.; Turner, T.; Visseren, F.L.; et al. Inclisiran in Patients at High Cardiovascular Risk with Elevated LDL Cholesterol. N. Engl. J. Med. 2017, 376, 1430–1440. [Google Scholar] [CrossRef] [PubMed]
  47. Wright, R.S.; Ray, K.K.; Raal, F.J.; Kallend, D.G.; Jaros, M.; Koenig, W.; Leiter, L.A.; Landmesser, U.; Schwartz, G.G.; Friedman, A.; et al. Pooled Patient-Level Analysis of Inclisiran Trials in Patients with Familial Hypercholesterolemia or Atherosclerosis. J. Am. Coll. Cardiol. 2021, 77, 1182–1193. [Google Scholar] [CrossRef] [PubMed]
  48. Gladwin, M.T. Cardiovascular complications and risk of death in sickle-cell disease. Lancet 2016, 387, 2565–2574. [Google Scholar] [CrossRef] [PubMed]
  49. Sachdev, V.; Rosing, D.R.; Thein, S.L. Cardiovascular complications of sickle cell disease. Trends Cardiovasc. Med. 2021, 31, 187–193. [Google Scholar] [CrossRef]
  50. Sripichai, O.; Fucharoen, S. Fetal hemoglobin regulation in β-thalassemia: Heterogeneity, modifiers and therapeutic approaches. Expert Rev. Hematol. 2016, 9, 1129–1137. [Google Scholar] [CrossRef]
  51. Musallam, K.M.; Sankaran, V.G.; Cappellini, M.D.; Duca, L.; Nathan, D.G.; Taher, A.T. Fetal hemoglobin levels and morbidity in untransfused patients with β-thalassemia intermedia. Blood 2012, 119, 364–367. [Google Scholar] [CrossRef]
  52. Danjou, F.; Anni, F.; Perseu, L.; Satta, S.; Dessì, C.; Lai, M.E.; Fortina, P.; Devoto, M.; Galanello, R. Genetic modifiers of β-thalassemia and clinical severity as assessed by age at first transfusion. Haematologica 2012, 97, 989–993. [Google Scholar] [CrossRef] [PubMed]
  53. 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] [PubMed]
  54. Adekile, A. The Genetic and Clinical Significance of Fetal Hemoglobin Expression in Sickle Cell Disease. Med. Princ. Pract. 2021, 30, 201–211. [Google Scholar] [CrossRef]
  55. Gambari, R.; Fibach, E. Medicinal chemistry of fetal hemoglobin inducers for treatment of β-thalassemia. Curr. Med. Chem. 2007, 14, 199–212. [Google Scholar] [CrossRef] [PubMed]
  56. Lavelle, D.; Engel, J.D.; Saunthararajah, Y. Fetal Hemoglobin Induction by Epigenetic Drugs. Semin. Hematol. 2018, 55, 60–67. [Google Scholar] [CrossRef]
  57. Prosdocimi, M.; Zuccato, C.; Cosenza, L.C.; Borgatti, M.; Lampronti, I.; Finotti, A.; Gambari, R. A Rational Approach to Drug Repositioning in β-thalassemia: Induction of Fetal Hemoglobin by Established Drugs. Wellcome Open Res. 2022, 7, 150. [Google Scholar] [CrossRef]
  58. Pavan, A.R.; Lopes, J.R.; Dos Santos, J.L. The state of the art of fetal hemoglobin-inducing agents. Expert Opin. Drug Discov. 2022, 17, 1279–1293. [Google Scholar] [CrossRef]
  59. Lavelle, D.; Ibanez, V.; Vaitkus, K.; Zhang, X.; Ramasamy, J.; Rivers, A.E.; Saunthararajah, Y.; Molokie, R. Combinatorial targeting of epigenome-modifying enzymes with small molecule drugs synergistically increases HbF. Blood Adv. 2023, 7, 3891–3902. [Google Scholar]
  60. Kalantri, S.A.; Ray, R.; Chattopadhyay, A.; Bhattacharjee, S.; Biswas, A.; Bhattacharyya, M. Efficacy of decitabine as hemoglobin F inducer in HbE/β-thalassemia. Ann. Hematol. 2018, 97, 1689–1694. [Google Scholar] [CrossRef]
  61. Fucharoen, S.; Siritanaratkul, N.; Winichagoon, P.; Chowthaworn, J.; Siriboon, W.; Muangsup, W.; Chaicharoen, S.; Poolsup, N.; Chindavijak, B.; Pootrakul, P.; et al. Hydroxyurea increases hemoglobin F levels and improves the effectiveness of erythropoiesis in beta-thalassemia/hemoglobin E disease. Blood 1996, 87, 887–892. [Google Scholar] [CrossRef]
  62. Bianchi, N.; Chiarabelli, C.; Zuccato, C.; Lampronti, I.; Borgatti, M.; Amari, G.; Delcanale, M.; Chiavilli, F.; Prus, E.; Fibach, E.; et al. Erythroid differentiation ability of butyric acid analogues: Identification of basal chemical structures of new inducers of foetal haemoglobin. Eur. J. Pharmacol. 2015, 752, 84–91. [Google Scholar] [CrossRef] [PubMed]
  63. Reid, M.E.; El Beshlawy, A.; Inati, A.; Kutlar, A.; Abboud, M.R.; Haynes, J., Jr.; Ward, R.; Sharon, B.; Taher, A.T.; Smith, W.; et al. A double-blind, placebo-controlled phase II study of the efficacy and safety of 2,2-dimethylbutyrate (HQK-1001), an oral fetal globin inducer, in sickle cell disease. Am. J. Hematol. 2014, 89, 709–713. [Google Scholar] [CrossRef]
  64. Rönndahl, G.; Mönkemeyer, S.; Schulze, S.; Pekrun, A.; Eikel, D.; Nau, H.; Witt, O. Novel valproic acid derivatives with hemoglobin F inducing activity. Am. J. Hematol. 2006, 81, 374–376. [Google Scholar] [CrossRef]
  65. Aerbajinai, W.; Zhu, J.; Gao, Z.; Chin, K.; Rodgers, G.P. Thalidomide induces gamma-globin gene expression through increased reactive oxygen species-mediated p38 MAPK signaling and histone H4 acetylation in adult erythropoiesis. Blood 2007, 110, 2864–2871. [Google Scholar] [CrossRef] [PubMed]
  66. 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] [PubMed]
  67. Marianna, P.; Kollia, P.; Akel, S.; Papassotiriou, Y.; Stamoulakatou, A.; Loukopoulos, D. Valproic acid, trichostatin and their combination with hemin preferentially enhance gamma-globin gene expression in human erythroid liquid cultures. Haematologica 2001, 86, 700–705. [Google Scholar]
  68. Guo, L.; Chen, J.; Wang, Q.; Zhang, J.; Huang, W. Oridonin enhances γ-globin expression in erythroid precursors from patients with β-thalassemia via activation of p38 MAPK signaling. Mol. Med. Rep. 2020, 21, 909–917. [Google Scholar] [CrossRef]
  69. Fibach, E.; Bianchi, N.; Borgatti, M.; Zuccato, C.; Finotti, A.; Lampronti, I.; Prus, E.; Mischiati, C.; Gambari, R. Effects of rapamycin on accumulation of alpha-, beta- and gamma-globin mRNAs in erythroid precursor cells from beta-thalassaemia patients. Eur. J. Haematol. 2006, 77, 437–441. [Google Scholar] [CrossRef]
  70. Zuccato, C.; Cosenza, L.C.; Zurlo, M.; Gasparello, J.; Papi, C.; D’Aversa, E.; Breveglieri, G.; Lampronti, I.; Finotti, A.; Borgatti, M.; et al. Expression of γ-globin genes in β-thalassemia patients treated with sirolimus: Results from a pilot clinical trial (Sirthalaclin). Ther. Adv. Hematol. 2022, 13, 20406207221100648. [Google Scholar] [CrossRef]
  71. Zuccato, C.; Bianchi, N.; Borgatti, M.; Lampronti, I.; Massei, F.; Favre, C.; Gambari, R. Everolimus is a potent inducer of erythroid differentiation and gamma-globin gene expression in human erythroid cells. Acta Haematol. 2007, 117, 168–176. [Google Scholar] [CrossRef]
  72. Fibach, E.; Bianchi, N.; Borgatti, M.; Prus, E.; Gambari, R. Mithramycin induces fetal hemoglobin production in normal and thalassemic human erythroid precursor cells. Blood 2003, 102, 1276–1281. [Google Scholar] [CrossRef] [PubMed]
  73. Fibach, E.; Prus, E.; Bianchi, N.; Zuccato, C.; Breveglieri, G.; Salvatori, F.; Finotti, A.; Lipucci di Paola, M.; Brognara, E.; Lampronti, I.; et al. Resveratrol: Antioxidant activity and induction of fetal hemoglobin in erythroid cells from normal donors and β-thalassemia patients. Int. J. Mol. Med. 2012, 29, 974–982. [Google Scholar]
  74. Lampronti, I.; Bianchi, N.; Borgatti, M.; Fibach, E.; Prus, E.; Gambari, R. Accumulation of gamma-globin mRNA in human erythroid cells treated with angelicin. Eur. J. Haematol. 2003, 71, 189–195. [Google Scholar] [CrossRef]
  75. Iftikhar, F.; Ali, H.; Musharraf, S.G. Cinchona alkaloids as natural fetal hemoglobin inducing agents in human erythroleukemia cells. RSC Adv. 2019, 9, 17551–17559. [Google Scholar] [CrossRef] [PubMed]
  76. Zuccato, C.; Cosenza, L.C.; Zurlo, M.; Lampronti, I.; Borgatti, M.; Scapoli, C.; Gambari, R.; Finotti, A. Treatment of Erythroid Precursor Cells from β-Thalassemia Patients with Cinchona Alkaloids: Induction of Fetal Hemoglobin Production. Int. J. Mol. Sci. 2021, 22, 13433. [Google Scholar] [CrossRef]
  77. 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]
  78. 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]
  79. Iftikhar, F.; Rahman, S.; Khan, M.B.N.; Khan, K.; Khan, M.N.; Uddin, R.; Musharraf, S.G. In Vitro and In Vivo Studies for the Investigation of γ-Globin Gene Induction by Adhatoda vasica: A Pre-Clinical Study of HbF Inducers for β-Thalassemia. Front. Pharmacol. 2022, 13, 797853. [Google Scholar] [CrossRef]
  80. Yasara, N.; Wickramarathne, N.; Mettananda, C.; Silva, I.; Hameed, N.; Attanayaka, K.; Rodrigo, R.; Wickramasinghe, N.; Perera, L.; Manamperi, A.; et al. A randomised double-blind placebo-controlled clinical trial of oral hydroxyurea for transfusion-dependent β-thalassaemia. Sci. Rep. 2022, 12, 2752. [Google Scholar] [CrossRef]
  81. Ansari, S.H.; Ansari, I.; Wasim, M.; Sattar, A.; Khawaja, S.; Zohaib, M.; Hussain, Z.; Adil, S.O.; Ansari, A.H.; Ansari, U.H.; et al. Evaluation of the combination therapy of hydroxyurea and thalidomide in β-thalassemia. Blood Adv. 2022, 6, 6162–6168. [Google Scholar] [CrossRef]
  82. Gamberini, M.R.; Prosdocimi, M.; Gambari, R. Sirolimus for treatment of β-thalassemia: From pre-clinical studies to the design of clinical trials. Health Educ. Public Health 2021, 4, 425–435. [Google Scholar]
  83. Smith, E.C.; Orkin, S.H. Hemoglobin genetics: Recent contributions of GWAS and gene editing. Hum. Mol. Genet. 2016, 25, R99–R105. [Google Scholar] [CrossRef] [PubMed]
  84. Uda, M.; Galanello, R.; Sanna, S.; Lettre, G.; Sankaran, V.G.; Chen, W.; Usala, G.; Busonero, F.; Maschio, A.; Albai, G.; et al. Genome wide association study shows BCL11A associated with persistent fetal hemoglobin and amelioration of the phenotype of β-thalassemia. Proc. Natl. Acad. Sci. USA 2008, 105, 1620–1625. [Google Scholar] [CrossRef] [PubMed]
  85. Nuinoon, M.; Makarasara, W.; Mushiroda, T.; Setianingsih, I.; Wahidiyat, P.A.; Sripichai, O.; Kumasaka, N.; Takahashi, A.; Svasti, S.; Munkongdee, T.; et al. A genome-wide association identified the common genetic variants influence disease severity in beta0-thalassemia/hemoglobin E. Hum. Genet. 2010, 127, 303–314. [Google Scholar] [CrossRef] [PubMed]
  86. Sherva, R.; Sripichai, O.; Abel, K.; Ma, Q.; Whitacre, J.; Angkachatchai, V.; Makarasara, W.; Winichagoon, P.; Svasti, S.; Fucharoen, S.; et al. Genetic modifiers of Hb E/beta0 thalassemia identified by a two-stage genome-wide association study. BMC Med. Genet. 2010, 11, 51. [Google Scholar] [CrossRef] [PubMed]
  87. Habara, A.; Steinberg, M.H. Minireview: Genetic basis of heterogeneity and severity in sickle cell disease. Exp. Biol. Med. 2016, 241, 689–696. [Google Scholar] [CrossRef] [PubMed]
  88. Solovieff, N.; Milton, J.N.; Hartley, S.W.; Sherva, R.; Sebastiani, P.; Dworkis, D.A.; Klings, E.S.; Farrer, L.A.; Garrett, M.E.; Ashley-Koch, A.; et al. Fetal hemoglobin in sickle cell anemia: Genome-wide association studies suggest a regulatory region in the 5′ olfactory receptor gene cluster. Blood 2010, 115, 1815–1822. [Google Scholar] [CrossRef] [PubMed]
  89. Ravi, N.S.; Wienert, B.; Wyman, S.K.; Bell, H.W.; George, A.; Mahalingam, G.; Vu, J.T.; Prasad, K.; Bandlamudi, B.P.; Devaraju, N.; et al. Identification of novel HPFH-like mutations by CRISPR base editing that elevate the expression of fetal hemoglobin. Elife 2022, 11, e65421. [Google Scholar] [CrossRef]
  90. Venkatesan, V.; Christopher, A.C.; Rhiel, M.; Azhagiri, M.K.K.; Babu, P.; Walavalkar, K.; Saravanan, B.; Andrieux, G.; Rangaraj, S.; Srinivasan, S.; et al. Editing the core region in HPFH deletions alters fetal and adult globin expression for treatment of β-hemoglobinopathies. Mol. Ther. Nucleic Acids 2023, 32, 671–688. [Google Scholar] [CrossRef]
  91. Antoniou, P.; Hardouin, G.; Martinucci, P.; Frati, G.; Felix, T.; Chalumeau, A.; Fontana, L.; Martin, J.; Masson, C.; Brusson, M.; et al. Base-editing-mediated dissection of a γ-globin cis-regulatory element for the therapeutic reactivation of fetal hemoglobin expression. Nat. Commun. 2022, 13, 6618. [Google Scholar] [CrossRef]
  92. Quagliano, A.; Acevedo, D.; Hardigan, P.; Prasad, S. Using Clustered Regularly Interspaced Short Palindromic Repeats gene editing to induce permanent expression of fetal hemoglobin in β-thalassemia and sickle cell disease: A comparative meta-analysis. Front. Med. 2022, 9, 943631. [Google Scholar] [CrossRef] [PubMed]
  93. Manco, L.; Santos, R.; Rocha, C.; Relvas, L.; Bento, C.; Maia, T.; Gomes, V.; Amorim, A.; Prata, M.J. Hb F Levels in β-Thalassemia Carriers and Normal Individuals: Known and Unknown Quantitative Trait Loci in the β-Globin Gene Cluster. Hemoglobin 2022, 46, 168–175. [Google Scholar] [CrossRef] [PubMed]
  94. Bashir, S.; Mahmood, S.; Mohsin, S.; Tabassum, I.; Ghafoor, M.; Sajjad, O. Modulatory effect of single nucleotide polymorphism in Xmn1, BCL11A and HBS1L-MYB loci on foetal haemoglobin levels in β-thalassemia major and Intermedia patients. J. Pak. Med. Assoc. 2021, 71, 1394–1398. [Google Scholar] [PubMed]
  95. Roy, P.; Bhattacharya, G.; Mandal, A.; Dasgupta, U.B.; Banerjee, D.; Chandra, S.; Das, M. Influence of BCL11A, HBS1L-MYB, HBBP1 single nucleotide polymorphisms and the HBG2 XmnI polymorphism on Hb F levels. Hemoglobin 2012, 36, 592–599. [Google Scholar] [CrossRef] [PubMed]
  96. Pace, B.; Zein, S. Understanding mechanisms of gamma-globin gene regulation to develop strategies for pharmacological fetal hemoglobin induction. Dev. Dyn. 2006, 235, 1727–1737. [Google Scholar] [CrossRef] [PubMed]
  97. Stadhouders, R.; Aktuna, S.; Thongjuea, S.; Aghajanirefah, A.; Pourfarzad, F.; van Ijcken, W.; Lenhard, B.; Rooks, H.; Best, S.; Menzel, S.; et al. HBS1L-MYB intergenic variants modulate fetal hemoglobin via long-range MYB enhancers. J. Clin. Investig. 2014, 124, 1699–1710. [Google Scholar] [CrossRef] [PubMed]
  98. Tepakhan, W.; Kanjanaopas, S.; Srewaradachpisal, K. Association Between Genetic Polymorphisms and Hb F Levels in Heterozygous β-Thalassemia 3.5 kb Deletions. Hemoglobin 2020, 44, 338–343. [Google Scholar] [CrossRef] [PubMed]
  99. Nguyen, T.K.; Joly, P.; Bardel, C.; Moulsma, M.; Bonello-Palot, N.; Francina, A. The XmnI (G)gamma polymorphism influences hemoglobin F synthesis contrary to BCL11A and HBS1L-MYB SNPs in a cohort of 57 beta-thalassemia intermedia patients. Blood Cells Mol. Dis. 2010, 45, 124–127. [Google Scholar] [CrossRef] [PubMed]
  100. Nuinoon, M.; Rattanaporn, P.; Benjchareonwong, T.; Choowet, A.; Suwanno, K.; Saekoo, N.; Lekpetch, K.; Thipthara, O.; Svasti, S.; Fucharoen, S. Genetic predictions of life expectancy in southern Thai patients with β0-thalassemia/Hb E. Biomed. Rep. 2022, 16, 52. [Google Scholar] [CrossRef]
  101. Bhagat, S.; Patra, P.K.; Thakur, A.S. Association between XmnI Polymorphism and HbF Level in Sickle Cell Disease Patients from Chhattisgarh. Int. J. Biomed. Sci. 2012, 8, 36–39. [Google Scholar] [CrossRef]
  102. Menzel, S.; Thein, S.L. Genetic Modifiers of Fetal Haemoglobin in Sickle Cell Disease. Mol. Diagn. Ther. 2019, 23, 235–244. [Google Scholar] [CrossRef]
  103. Ju, J.; Wang, Y.; Liu, R.; Zhang, Y.; Xu, Z.; Wang, Y.; Wu, Y.; Liu, M.; Cerruti, L.; Zou, F.; et al. Human fetal globin gene expression is regulated by LYAR. Nucleic Acids Res. 2014, 42, 9740–9752. [Google Scholar] [CrossRef]
  104. Chen, D.; Zuo, Y.; Zhang, X.; Ye, Y.; Bao, X.; Huang, H.; Tepakhan, W.; Wang, L.; Ju, J.; Chen, G.; et al. A Genetic Variant Ameliorates β-Thalassemia Severity by Epigenetic-Mediated Elevation of Human Fetal Hemoglobin Expression. Am. J. Hum. Genet. 2017, 101, 130–138. [Google Scholar] [CrossRef] [PubMed]
  105. Bianchi, N.; Cosenza, L.C.; Lampronti, I.; Finotti, A.; Breveglieri, G.; Zuccato, C.; Fabbri, E.; Marzaro, G.; Chilin, A.; De Angelis, G.; et al. Structural and Functional Insights on an Uncharacterized Aγ-Globin-Gene Polymorphism Present in Four β0-Thalassemia Families with High Fetal Hemoglobin Levels. Mol. Diagn. Ther. 2016, 20, 161–173. [Google Scholar] [CrossRef]
  106. Breveglieri, G.; Bianchi, N.; Cosenza, L.C.; Gamberini, M.R.; Chiavilli, F.; Zuccato, C.; Montagner, G.; Borgatti, M.; Lampronti, I.; Finotti, A.; et al. An Aγ-globin G->A gene polymorphism associated with β039 thalassemia globin gene and high fetal hemoglobin production. BMC Med. Genet. 2017, 18, 93. [Google Scholar] [CrossRef]
  107. Gemmo, C.; Breveglieri, G.; Marzaro, G.; Lampronti, I.; Cosenza, L.C.; Gasparello, J.; Zuccato, C.; Fabbri, E.; Borgatti, M.; Chilin, A.; et al. Surface plasmon resonance based analysis of the binding of LYAR protein to the rs368698783 (G>A) polymorphic Aγ-globin gene sequences mutated in β-thalassemia. Anal. Bioanal. Chem. 2019, 411, 7699–7707. [Google Scholar] [CrossRef] [PubMed]
  108. Zuccato, C.; Cosenza, L.C.; Zurlo, M.; Breveglieri, G.; Bianchi, N.; Lampronti, I.; Gasparello, J.; Scapoli, C.; Borgatti, M.; Finotti, A.; et al. The rs368698783 (G>A) Polymorphism Affecting LYAR Binding to the Aγ-Globin Gene Is Associated with High Fetal Hemoglobin (HbF) in β-Thalassemia Erythroid Precursor Cells Treated with HbF Inducers. Int. J. Mol. Sci. 2023, 24, 776. [Google Scholar] [CrossRef]
  109. Giannopoulou, E.; Bartsakoulia, M.; Tafrali, C.; Kourakli, A.; Poulas, K.; Stavrou, E.F.; Papachatzopoulou, A.; Georgitsi, M.; Patrinos, G.P. A single nucleotide polymorphism in the HBBP1 gene in the human β-globin locus is associated with a mild β-thalassemia disease phenotype. Hemoglobin 2012, 36, 433–445. [Google Scholar] [CrossRef] [PubMed]
  110. Perkins, A.; Xu, X.; Higgs, D.R.; Patrinos, G.P.; Arnaud, L.; Bieker, J.J.; Philipsen, S.; KLF1 Consensus Workgroup. Krüppeling erythropoiesis: An unexpected broad spectrum of human red blood cell disorders due to KLF1 variants. Blood 2016, 127, 1856–1862. [Google Scholar] [CrossRef]
  111. Kumar, R.; Yadav, R.; Mishra, S.; Singh, M.P.S.S.; Gwal, A.; Bharti, P.K.; Rajasubramaniam, S. Krüppel-like factor 1 (KLF1) gene single nucleotide polymorphisms in sickle cell disease and its association with disease-related morbidities. Ann. Hematol. 2021, 100, 365–373. [Google Scholar] [CrossRef]
  112. Zhou, D.; Liu, K.; Sun, C.W.; Pawlik, K.M.; Townes, T.M. KLF1 regulates BCL11A expression and gamma- to beta-globin gene switching. Nat. Genet. 2010, 42, 742–744. [Google Scholar] [CrossRef] [PubMed]
  113. Khamphikham, P.; Sripichai, O.; Munkongdee, T.; Fucharoen, S.; Tongsima, S.; Smith, D.R. Genetic variation of Krüppel-like factor 1 (KLF1) and fetal hemoglobin (HbF) levels in β0-thalassemia/HbE disease. Int. J. Hematol. 2018, 107, 297–310. [Google Scholar] [CrossRef]
  114. Pereira, C.; Relvas, L.; Bento, C.; Abade, A.; Ribeiro, M.L.; Manco, L. Polymorphic variations influencing fetal hemoglobin levels: Association study in beta-thalassemia carriers and in normal individuals of Portuguese origin. Blood Cells Mol. Dis. 2015, 54, 315–320. [Google Scholar] [CrossRef] [PubMed]
  115. Zaker-Kandjani, B.; Namdar-Aligoodarzi, P.; Azarkeivan, A.; Najmabadi, H.; Banan, M. Mutation screening of the Krüppel-like factor 1 gene using single-strand conformational polymorphism in a cohort of Iranian β-thalassemia patients. Hemoglobin 2015, 39, 24–29. [Google Scholar] [CrossRef] [PubMed]
  116. Borgio, J.F.; AbdulAzeez, S.; Al-Muslami, A.M.; Naserullah, Z.A.; Al-Jarrash, S.; Al-Suliman, A.M.; Al-Madan, M.S.; Al-Ali, A.K. KLF1 gene and borderline hemoglobin A2 in Saudi population. Arch. Med. Sci. 2018, 14, 230–236. [Google Scholar] [CrossRef] [PubMed]
  117. Kolliopoulou, A.; Siamoglou, S.; John, A.; Sgourou, A.; Kourakli, A.; Symeonidis, A.; Vlachaki, E.; Chalkia, P.; Theodoridou, S.; Ali, B.R.; et al. Role of Genomic Biomarkers in Increasing Fetal Hemoglobin Levels Upon Hydroxyurea Therapy and in β-Thalassemia Intermedia: A Validation Cohort Study. Hemoglobin 2019, 43, 27–33. [Google Scholar] [CrossRef]
  118. Statello, L.; Guo, C.J.; Chen, L.L.; Huarte, M. Gene regulation by long non-coding RNAs and its biological functions. Nat. Rev. Mol. Cell Biol. 2021, 22, 96–118. [Google Scholar] [CrossRef] [PubMed]
  119. Schmitz, S.U.; Grote, P.; Herrmann, B.G. Mechanisms of long noncoding RNA function in development and disease. Cell. Mol. Life Sci. 2016, 73, 2491–2509. [Google Scholar] [CrossRef] [PubMed]
  120. Jia, W.; Jia, S.; Chen, P.; He, Y. Construction and Analysis of a Long Non-Coding RNA (lncRNA)-Associated ceRNA Network in β-Thalassemia and Hereditary Persistence of Fetal Hemoglobin. Med. Sci. Monit. 2019, 25, 7079–7086. [Google Scholar] [CrossRef]
  121. Yang, F.; Ruan, H.; Li, S.; Hou, W.; Qiu, Y.; Deng, L.; Su, S.; Chen, P.; Pang, L.; Lai, K. Analysis of circRNAs and circRNA-associated competing endogenous RNA networks in β-thalassemia. Sci. Rep. 2022, 12, 8071. [Google Scholar] [CrossRef]
  122. Fakhr-Eldeen, A.; Toraih, E.A.; Fawzy, M.S. Long non-coding RNAs MALAT1, MIAT and ANRIL gene expression profiles in beta-thalassemia patients: A cross-sectional analysis. Hematology 2019, 24, 308–317. [Google Scholar] [CrossRef] [PubMed]
  123. Finotti, A.; Fabbri, E.; Lampronti, I.; Gasparello, J.; Borgatti, M.; Gambari, R. MicroRNAs and Long Non-coding RNAs in Genetic Diseases. Mol. Diagn. Ther. 2019, 23, 155–171. [Google Scholar] [CrossRef] [PubMed]
  124. Wonkam, A. The future of sickle cell disease therapeutics rests in genomics. Dis. Model. Mech. 2023, 16, dmm049765. [Google Scholar] [CrossRef] [PubMed]
  125. He, L.; Hannon, G.J. MicroRNAs: Small RNAs with a big role in gene regulation. Nat. Rev. Genet. 2004, 5, 522–531. [Google Scholar] [CrossRef] [PubMed]
  126. Sontheimer, E.J.; Carthew, R.W. Silence from within: Endogenous siRNAs and miRNAs. Cell 2005, 122, 9–12. [Google Scholar] [CrossRef] [PubMed]
  127. Alvarez-Garcia, I.; Miska, E.A. MicroRNA functions in animal development and human disease. Development 2005, 132, 4653–4662. [Google Scholar] [CrossRef] [PubMed]
  128. Li, Y.; Kowdley, K.V. MicroRNAs in common human diseases. Genom. Proteom. Bioinform. 2012, 10, 246–253. [Google Scholar] [CrossRef]
  129. Lu, M.; Zhang, Q.; Deng, M.; Miao, J.; Guo, Y.; Gao, W.; Cui, Q. An analysis of human microRNA and disease associations. PLoS ONE 2008, 3, e3420. [Google Scholar] [CrossRef]
  130. Das, J.; Podder, S.; Ghosh, T.C. Insights into the miRNA regulations in human disease genes. BMC Genom. 2014, 15, 1010. [Google Scholar] [CrossRef]
  131. Paunovska, K.; Loughrey, K.; Dahlman, J.E. Drug delivery systems for RNA therapeutics. Nat. Rev. Genet. 2022, 23, 265–280. [Google Scholar] [CrossRef]
  132. Buck, J.; Grossen, P.; Cullis, P.R.; Huwyler, J.; Witzigmann, D. Lipid-based DNA therapeutics: Hallmarks of non-viral gene delivery. ACS Nano 2019, 13, 3754–3782. [Google Scholar] [CrossRef] [PubMed]
  133. Sristi Almalki, W.H.; Karwasra, R.; Gupta, G.; Singh, S.; Sharma, A.; Sahebkar, A.; Kesharwani, P. Advances in the polymeric nanoparticulate delivery systems for RNA therapeutics. Prog. Mol. Biol. Transl. Sci. 2024, 204, 219–248. [Google Scholar]
  134. Roberts, T.C.; Langer, R.; Wood MJ, A. Advances in oligonucleotide drug delivery. Nat. Rev. Drug Discov. 2020, 19, 673–694. [Google Scholar] [CrossRef] [PubMed]
  135. Yamamoto, T.; Mukai, Y.; Wada, F.; Terada, C.; Kayaba, Y.; Oh, K.; Yamayoshi, A.; Obika, S.; Harada-Shiba, M. Highly Potent GalNAc-Conjugated Tiny LNA Anti-miRNA-122 Antisense Oligonucleotides. Pharmaceutics 2021, 13, 817. [Google Scholar] [CrossRef] [PubMed]
  136. Semple, S.C.; Akinc, A.; Chen, J.; Sandhu, A.P.; Mui, B.L.; Cho, C.K.; Sah, D.W.; Stebbing, D.; Crosley, E.J.; Yaworski, E.; et al. Rational design of cationic lipids for siRNA delivery. Nat. Biotechnol. 2010, 28, 172–176. [Google Scholar] [CrossRef] [PubMed]
  137. Hong, D.S.; Kang, Y.K.; Borad, M.; Sachdev, J.; Ejadi, S.; Lim, H.Y.; Brenner, A.J.; Park, K.; Lee, J.L.; Kim, T.Y.; et al. Phase 1 study of MRX34, a liposomal miR-34a mimic, in patients with advanced solid tumours. Br. J. Cancer 2020, 122, 1630–1637. [Google Scholar] [CrossRef] [PubMed]
  138. Ward, C.M.; Li, B.; Pace, B.S. Original Research: Stable expression of miR-34a mediates fetal hemoglobin induction in K562 cells. Exp. Biol. Med. 2016, 241, 719–729. [Google Scholar] [CrossRef] [PubMed]
  139. Wang, F.; Ling, L.; Yu, D. MicroRNAs in β-thalassemia. Am. J. Med. Sci. 2021, 362, 5–12. [Google Scholar] [CrossRef] [PubMed]
  140. Saki, N.; Abroun, S.; Soleimani, M.; Kavianpour, M.; Shahjahani, M.; Mohammadi-Asl, J.; Hajizamani, S. MicroRNA Expression in β-Thalassemia and Sickle Cell Disease: A Role in the Induction of Fetal Hemoglobin. Cell J. 2016, 17, 583–592. [Google Scholar]
  141. Starlard-Davenport, A.; Fitzgerald, A.; Pace, B.S. Exploring epigenetic and microRNA approaches for γ-globin gene regulation. Exp. Biol. Med. 2021, 246, 2347–2357. [Google Scholar] [CrossRef]
  142. Das, S.S.; Das, S.; Byram, P.K.; Rahaman, M.; Dolai, T.K.; Chatterjee, A.; Chakravorty, N. MicroRNA expression patterns in HbE/β-thalassemia patients: The passwords to unlock fetal hemoglobin expression in β-hemoglobinopathies. Blood Cells Mol. Dis. 2021, 87, 102523. [Google Scholar] [CrossRef] [PubMed]
  143. Gasparello, J.; Fabbri, E.; Bianchi, N.; Breveglieri, G.; Zuccato, C.; Borgatti, M.; Gambari, R.; Finotti, A. BCL11A mRNA Targeting by miR-210, A Possible Network Regulating γ-Globin Gene Expression. Int. J. Mol. Sci. 2017, 18, 2530. [Google Scholar] [CrossRef] [PubMed]
  144. Li, H.; Lin, R.; Li, H.; Ou, R.; Wang, K.; Lin, J.; Li, C. MicroRNA-92a-3p-mediated inhibition of BCL11A upregulates γ-globin expression and inhibits oxidative stress and apoptosis in erythroid precursor cells. Hematology 2022, 27, 1152–1162. [Google Scholar] [CrossRef] [PubMed]
  145. Gholampour, M.A.; Asadi, M.; Naderi, M.; Azarkeivan, A.; Soleimani, M.; Atashi, A. miR-30a regulates γ-globin expression in erythoid precursors of intermedia thalassemia through targeting, BCL11A. Mol. Biol. Rep. 2020, 47, 3909–3918. [Google Scholar] [CrossRef] [PubMed]
  146. Cheng, Y.; Shang, X.; Chen, D.; Pang, D.; Zhao, C.; Xu, X. MicroRNA-2355-5p regulates γ-globin expression in human erythroid cells by inhibiting KLF6. Br. J. Haematol. 2021, 193, 401–405. [Google Scholar] [CrossRef] [PubMed]
  147. Srinoun, K.; Sathirapongsasuti, N.; Paiboonsukwong, K.; Sretrirutchai, S.; Wongchanchailert, M.; Fucharoen, S. miR-144 regulates oxidative stress tolerance of thalassemic erythroid cell via targeting NRF2. Ann. Hematol. 2019, 98, 2045–2052. [Google Scholar] [CrossRef] [PubMed]
  148. Li, Y.; Liu, D.; Zhang, X.; Li, Z.; Ye, Y.; Liu, Q.; Shen, J.; Chen, Z.; Huang, H.; Liang, Y.; et al. miR-326 regulates HbF synthesis by targeting EKLF in human erythroid cells. Exp. Hematol. 2018, 63, 33–40.e2. [Google Scholar] [CrossRef]
  149. Srinoun, K.; Nopparatana, C.; Wongchanchailert, M.; Fucharoen, S. MiR-155 enhances phagocytic activity of β-thalassemia/HbE monocytes via targeting of BACH1. Int. J. Hematol. 2017, 106, 638–647. [Google Scholar] [CrossRef]
  150. Pule, G.D.; Mowla, S.; Novitzky, N.; Wonkam, A. Hydroxyurea down-regulates BCL11A, KLF-1 and MYB through miRNA-mediated actions to induce γ-globin expression: Implications for new therapeutic approaches of sickle cell disease. Clin. Transl. Med. 2016, 5, 15. [Google Scholar] [CrossRef]
  151. Starlard-Davenport, A.; Smith, A.; Vu, L.; Li, B.; Pace, B.S. MIR29B mediates epigenetic mechanisms of HBG gene activation. Br. J. Haematol. 2019, 186, 91–100. [Google Scholar] [CrossRef]
  152. Lulli, V.; Romania, P.; Morsilli, O.; Cianciulli, P.; Gabbianelli, M.; Testa, U.; Giuliani, A.; Marziali, G. MicroRNA-486-3p regulates γ-globin expression in human erythroid cells by directly modulating BCL11A. PLoS ONE 2013, 8, e60436. [Google Scholar] [CrossRef] [PubMed]
  153. Sankaran, V.G.; Menne, T.F.; Šćepanović, D.; Vergilio, J.A.; Ji, P.; Kim, J.; Thiru, P.; Orkin, S.H.; Lander, E.S.; Lodish, H.F. MicroRNA-15a and -16-1 act via MYB to elevate fetal hemoglobin expression in human trisomy 13. Proc. Natl. Acad. Sci. USA 2011, 108, 1519–1524. [Google Scholar] [CrossRef] [PubMed]
  154. Chen, M.; Wang, X.; Wang, H.; Zhang, M.; Chen, L.; Chen, H.; Pan, Y.; Zhang, Y.; Xu, L.; Huang, H. The clinical value of hsa-miR-190b-5p in peripheral blood of pediatric beta-thalassemia and its regulation on BCL11A expression. PLoS ONE 2023, 18, e0292031. [Google Scholar]
  155. Chen, M.; Lv, A.; Zhang, S.; Zheng, J.; Lin, N.; Xu, L.; Huang, H. Peripheral blood circular RNA circ-0008102 may serve as a novel clinical biomarker in beta-thalassemia patients. Eur. J. Pediatr. 2024, 183, 1367–1379. [Google Scholar] [CrossRef] [PubMed]
  156. Xie, D.; Han, Y.; Zhang, W.; Wu, J.; An, B.; Huang, S.; Sun, F. Long Non-Coding RNA H19 Leads to Upregulation of gamma-Globin Gene Expression during Erythroid Differentiation. Hemoglobin 2024, 1–11. [Google Scholar] [CrossRef] [PubMed]
  157. Ma, S.P.; Gao, X.X.; Zhou, G.Q.; Zhang, H.K.; Yang, J.M.; Wang, W.J.; Song, X.M.; Chen, H.Y.; Lu, D.R. Reactivation of gamma-globin expression using a minicircle DNA system to treat beta-thalassemia. Gene 2022, 820, 146289. [Google Scholar] [CrossRef] [PubMed]
  158. Morrison, T.A.; Wilcox, I.; Luo, H.Y.; Farrell, J.J.; Kurita, R.; Nakamura, Y.; Murphy, G.J.; Cui, S.; Steinberg, M.H.; Chui, D.H.K. A long noncoding RNA from the HBS1L-MYB intergenic region on chr6q23 regulates human fetal hemoglobin expression. Blood Cells Mol. Dis. 2018, 69, 1–9. [Google Scholar] [CrossRef] [PubMed]
  159. Papasavva, P.L.; Papaioannou, N.Y.; Patsali, P.; Kurita, R.; Nakamura, Y.; Sitarou, M.; Christou, S.; Kleanthous, M.; Lederer, C.W. Distinct miRNA Signatures and Networks Discern Fetal from Adult Erythroid Differentiation and Primary from Immortalized Erythroid Cells. Int. J. Mol. Sci. 2021, 22, 3626. [Google Scholar] [CrossRef]
  160. Rahaman, M.; Mukherjee, M.; Bhattacharya, S.; Mukherjee, B.; Shukla, P.C.; Dolai, T.K.; Chakravorty, N. Exploring the crosstalk between long non-coding RNAs and microRNAs to unravel potential prognostic and therapeutic biomarkers in β-thalassemia. Mol. Biol. Rep. 2022, 49, 7057–7068. [Google Scholar] [CrossRef]
  161. Kalaigar, S.S.; Rajashekar, R.B.; Nataraj, S.M.; Vishwanath, P.; Prashant, A. Bioinformatic Tools for the Identification of MicroRNAs Regulating the Transcription Factors in Patients with β-Thalassemia. Bioinform. Biol. Insights 2022, 16, 11779322221115536. [Google Scholar] [CrossRef]
  162. Faraoni, I.; Antonetti, F.R.; Cardone, J.; Bonmassar, E. miR-155 gene: A typical multifunctional, microRNA. Biochim. Biophys. Acta 2009, 1792, 497–505. [Google Scholar] [CrossRef] [PubMed]
  163. Das, S.S.; Mitra, A.; Chakravorty, N. Diseases and their clinical heterogeneity—Are we ignoring the SNiPers and micRomaNAgers? An illustration using Beta-thalassemia clinical spectrum and fetal hemoglobin levels. Genomics 2019, 111, 67–75. [Google Scholar] [CrossRef] [PubMed]
  164. Hojjati, M.T.; Azarkeivan, A.; Pourfathollah, A.A.; Amirizadeh, N. Comparison of MicroRNAs Mediated in Reactivation of the γ-Globin in β-Thalassemia Patients, Responders and Non-Responders to Hydroxyurea. Hemoglobin 2017, 41, 110–115. [Google Scholar] [CrossRef] [PubMed]
  165. Sales, R.R.; Nogueira, B.L.; Luizon, M.R. Pharmacogenomics of hydroxyurea therapy and fetal hemoglobin levels in sickle cell anemia. Pharmacogenomics 2022, 23, 393–396. [Google Scholar] [CrossRef] [PubMed]
  166. Sales, R.R.; Nogueira, B.L.; Tosatti, J.A.G.; Gomes, K.B.; Luizon, M.R. Do Genetic Polymorphisms Affect Fetal Hemoglobin (HbF) Levels in Patients with Sickle Cell Anemia Treated with Hydroxyurea? A Systematic Review and Pathway Analysis. Front. Pharmacol. 2022, 12, 779497. [Google Scholar] [CrossRef] [PubMed]
  167. Ali, N.; Ayyub, M.; Khan, S.A.; Ahmed, S.; Abbas, K.; Malik, H.S.; Tashfeen, S. Frequency of Ggamma-globin promoter −158 (C>T) XmnI polymorphism in patients with homozygous/compound heterozygous beta thalassaemia. Hematol. Oncol. Stem Cell Ther. 2015, 8, 10–15. [Google Scholar] [CrossRef] [PubMed]
  168. Banan, M.; Bayat, H.; Azarkeivan, A.; Mohammadparast, S.; Kamali, K.; Farashi, S.; Bayat, N.; Khani, M.H.; Neishabury, M.; Najmabadi, H. The XmnI and BCL11A single nucleotide polymorphisms may help predict hydroxyurea response in Iranian beta-thalassemia patients. Hemoglobin 2012, 36, 371–380. [Google Scholar] [CrossRef] [PubMed]
  169. Koren, A.; Levin, C.; Dgany, O.; Kransnov, T.; Elhasid, R.; Zalman, L.; Palmor, H.; Tamary, H. Response to hydroxyurea therapy in beta-thalassemia. Am. J. Hematol. 2008, 83, 366–370. [Google Scholar] [CrossRef] [PubMed]
  170. Ghosh, D.; Panja, A.; Saha, D.; Banerjee, U.; Datta, A.K.; Basu, A. Drug Repurposing: Hydroxyurea Therapy Improves the Transfusion-Free Interval in HbE/Beta-Thalassemia-Major Patients with the XmnI Polymorphism. Genet. Test. Mol. Biomark. 2021, 25, 563–570. [Google Scholar] [CrossRef]
  171. Perrine, S.P.; Miller, B.A.; Faller, D.V.; Cohen, R.A.; Vichinsky, E.P.; Hurst, D.; Lubin, B.H.; Papayannopoulou, T. Sodium butyrate enhances fetal globin gene expression in erythroid progenitors of patients with Hb SS and beta thalassemia. Blood 1989, 74, 454–459. [Google Scholar] [CrossRef]
  172. Perrine, S.P.; Ginder, G.D.; Faller, D.V.; Dover, G.H.; Ikuta, T.; Witkowska, H.E.; Cai, S.P.; Vichinsky, E.P.; Olivieri, N.F. A short-term trial of butyrate to stimulate fetal-globin-gene expression in the beta-globin disorders. N. Engl. J. Med. 1993, 328, 81–86. [Google Scholar] [CrossRef]
  173. Faller, D.V.; Perrine, S.P. Butyrate in the treatment of sickle cell disease and beta-thalassemia. Curr. Opin. Hematol. 1995, 2, 109–117. [Google Scholar] [CrossRef] [PubMed]
  174. Panja, A.; Basu, A. Pharmacogenomics of the Drugs used for the Treatment of Thalassemia. J. Cytol. Histol. 2015, 6, 5. [Google Scholar]
  175. Testa, U. Fetal hemoglobin chemical inducers for treatment of hemoglobinopathies. Ann. Hematol. 2009, 88, 505–528. [Google Scholar] [CrossRef]
  176. Tayebi, B.; Abrishami, F.; Alizadeh, S.; Minayi, N.; Mohammadian, M.; Soleimani, M.; Dehghanifard, A.; Atwan, H.; Ajami, M.; Ajami, M. Modulation of microRNAs expression in hematopoietic stem cells treated with sodium butyrate in inducing fetal hemoglobin expression. Artif. Cells Nanomed. Biotechnol. 2016, 45, 146–156. [Google Scholar] [CrossRef] [PubMed]
  177. 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. 2022, 12, 814302. [Google Scholar] [CrossRef]
  178. Yang, K.; Wu, Y.; Ma, Y.; Xiao, J.; Zhou, Y.; Yin, X. The association of HBG2, BCL11A, and HBS1L-MYB polymorphisms to thalidomide response in Chinese β-thalassemia patients. Blood Cells Mol. Dis. 2020, 84, 102442. [Google Scholar] [CrossRef]
  179. Taher, A.; Cappellini, M.D. Update on the use of deferasirox in the management of iron overload. Ther. Clin. Risk Manag. 2009, 5, 857–868. [Google Scholar] [CrossRef]
  180. Lee, J.W.; Kang, H.J.; Choi, J.Y.; Kim, N.H.; Jang, M.K.; Yeo, C.W.; Lee, S.S.; Kim, H.; Park, J.D.; Park, K.D.; et al. Pharmacogenetic study of deferasirox, an iron chelating agent. PLoS ONE 2013, 8, e64114. [Google Scholar] [CrossRef]
  181. Ricchi, P.; Marsella, M. Profile of deferasirox for the treatment of patients with non-transfusion-dependent thalassemia syndromes. Drug Des. Dev. Ther. 2015, 9, 6475–6482. [Google Scholar] [CrossRef]
  182. Cusato, J.; Allegra, S.; De Francia, S.; Massano, D.; Piga, A.; D’Avolio, A. Role of pharmacogenetics on deferasirox AUC and efficacy. Pharmacogenomics 2016, 17, 561–572. [Google Scholar] [CrossRef] [PubMed]
  183. Cusato, J.; Allegra, S.; Massano, D.; De Francia, S.; Piga, A.; D’Avolio, A. Influence of single-nucleotide polymorphisms on deferasirox C trough levels and effectiveness. Pharmacogenomics J. 2015, 15, 263–271. [Google Scholar] [CrossRef]
  184. Biancamano, B.M.; Connelly, J.; Villeneuve, L.; Caron, P.; Guillemette, C. Deferiprone Glucuronidation by Human Tissues and Recombinant UDP Glucuronosyltransferase1A6, An In Vitro Investigation of Genetic and Splice Variants. Am. Soc. Pharmacol. Exp. Ther. 1997, 37, 322–329. [Google Scholar]
  185. Dadheech, S.; Rao, A.V.; Shaheen, U.; Hussien, M.D.; Jain, S.; Jyothy, A.; Munshi, A. Three most common nonsynonymous UGT1A6*2 polymorphisms (Thr181Ala, Arg184Ser and Ser7Ala) and therapeutic response to deferiprone in β-thalassemia major patients. Gene 2013, 531, 301–305. [Google Scholar] [CrossRef] [PubMed]
  186. Wang, B.; Wang, J.; Huang, S.Q.; Su, H.H.; Zhou, S.F. Genetic polymorphism of the human cytochrome P450 2C9 gene and its clinical significance. Curr. Drug Metab. 2009, 10, 781–834. [Google Scholar] [CrossRef]
  187. Zobdeh, F.; Eremenko, I.I.; Akan, M.A.; Tarasov, V.V.; Chubarev, V.N.; Schiöth, H.B.; Mwinyi, J. Pharmacogenetics and PainTreatment with a Focus on NonSteroidal Anti-Inflammatory Drugs (NSAIDs) and Antidepressants: A Systematic Review. Pharmaceutics 2022, 14, 1190. [Google Scholar] [CrossRef] [PubMed]
  188. Agúndez, J.A.; García-Martín, E.; Martínez, C. Genetically based impairment in CYP2C8- and CYP2C9-dependent NSAID metabolism as a risk factor for gastrointestinal bleeding: Is a combination of pharmacogenomics and metabolomics required to improve personalized medicine? Expert Opin. Drug Metab. Toxicol. 2009, 5, 607–620. [Google Scholar] [CrossRef] [PubMed]
  189. Husain, M.; Hartman, A.D.; Desai, P. Pharmacogenomics of sickle cell disease: Steps toward personalized medicine. Pharmgenomics Pers. Med. 2017, 10, 261–265. [Google Scholar] [CrossRef]
  190. Joly, P.; Gagnieu, M.C.; Bardel, C.; Francina, A.; Pondarre, C.; Martin, C. Genotypic screening of the main opiate-related polymorphisms in a cohort of 139 sickle cell disease patients. Am. J. Hematol. 2012, 87, 534–536. [Google Scholar] [CrossRef]
  191. Jhun, E.H.; Yao, Y.; He, Y.; Mack, A.K.; Wilkie, D.J.; Molokie, R.E.; Wang, Z.J. Prevalence of pain-related single nucleotide polymorphisms in patients of African origin with sickle cell disease. Pharmacogenomics 2015, 16, 1795–1806. [Google Scholar] [CrossRef]
  192. Ofoegbu, A.; BEttienne, E. Pharmacogenomics and Morphine. J. Clin. Pharmacol. 2021, 61, 1149–1155. [Google Scholar] [CrossRef] [PubMed]
  193. Wong, A.K.; Somogyi, A.A.; Rubio, J.; Philip, J. The Role of Pharmacogenomics in Opioid Prescribing. Curr. Treat. Options Oncol. 2022, 23, 1353–1369. [Google Scholar] [CrossRef]
  194. Gammal, R.S.; Crews, K.R.; Haidar, C.E.; Hoffman, J.M.; Baker, D.K.; Barker, P.J.; Estepp, J.H.; Pei, D.; Broeckel, U.; Wang, W.; et al. Pharmacogenetics for Safe Codeine Use in Sickle Cell Disease. Pediatrics 2016, 138, e20153479. [Google Scholar] [CrossRef] [PubMed]
  195. Gupta, P.; Kumar, R. Nitric oxide: A potential etiological agent for vaso-occlusive crises in sickle cell disease. Nitric Oxide 2024, 144, 40–46. [Google Scholar] [CrossRef] [PubMed]
  196. Weiner, D.L.; Hibberd, P.L.; Betit, P.; Cooper, A.B.; Botelho, C.A.; Brugnara, C. Preliminary assessment of inhaled nitric oxide for acute vaso-occlusive crisis in pediatric patients with sickle cell disease. JAMA 2003, 289, 1136–1142. [Google Scholar] [CrossRef] [PubMed]
  197. Tsukada, T.; Yokoyama, K.; Arai, T.; Takemoto, F.; Hara, S.; Yamada, A.; Kawaguchi, Y.; Hosoya, T.; Igari, J. Evidence of association of the ecNOS gene polymorphism with plasma NO metabolite levels in humans. Biochem. Biophys. Res. Commun. 1998, 245, 190–193. [Google Scholar] [CrossRef] [PubMed]
  198. Yousry, S.M.; Ellithy, H.N.; Shahin, G.H. Endothelial nitric oxide synthase gene polymorphisms and the risk of vasculopathy in sickle cell disease. Hematology 2016, 21, 359–367. [Google Scholar] [CrossRef] [PubMed]
  199. Nishank, S.S.; Singh, M.P.; Yadav, R.; Gupta, R.B.; Gadge, V.S.; Gwal, A. Endothelial nitric oxide synthase gene polymorphism is associated with sickle cell disease patients in India. J. Hum. Genet. 2013, 58, 775–779. [Google Scholar] [CrossRef] [PubMed]
  200. Hebbel, R.P.; Vercellotti, G.M. Multiple inducers of endothelial NOS (eNOS) dysfunction in sickle cell disease. Am. J. Hematol. 2021, 96, 1505–1517. [Google Scholar] [CrossRef]
  201. Gladwin, M.T.; Schechter, A.N. Nitric oxide therapy in sickle cell disease. Semin. Hematol. 2001, 38, 333–342. [Google Scholar] [CrossRef]
  202. King, S.B. Nitric oxide production from hydroxyurea. Free Radic. Biol. Med. 2004, 37, 737–744. [Google Scholar] [CrossRef]
  203. Caudle, K.E.; Klein, T.E.; Hoffman, J.M.; Muller, D.J.; Whirl-Carrillo, M.; Gong, L.; McDonagh, E.M.; Sangkuhl, K.; Thorn, C.F.; Schwab, M.; et al. Incorporation of pharmacogenomics into routine clinical practice: The Clinical Pharmacogenetics Implementation Consortium (CPIC) guideline development process. Curr. Drug Metab. 2014, 15, 209–217. [Google Scholar] [CrossRef] [PubMed]
  204. Cho, G.; Anie, K.A.; Buckton, J.; Kiilu, P.; Layton, M.; Alexander, L.; Hemmaway, C.; Sutton, D.; Amos, C.; Doré, C.J.; et al. SWIM (sickle with ibuprofen and morphine) randomised controlled trial fails to recruit: Lessons learnt. BMJ Open 2016, 6, e011276. [Google Scholar] [CrossRef]
  205. Adam, S.S.; Hoppe, C. Potential role for statins in sickle cell disease. Pediatr. Blood Cancer 2013, 60, 550–557. [Google Scholar] [CrossRef] [PubMed]
  206. Selby, R.; Nisbet-Brown, E.; Basran, R.K.; Chang, L.; Olivieri, N.F. Valproic acid and augmentation of fetal hemoglobin in individuals with and without sickle cell disease. Blood 1997, 90, 891–893. [Google Scholar] [CrossRef] [PubMed]
  207. Lampronti, I.; Bianchi, N.; Zuccato, C.; Medici, A.; Bergamini, P.; Gambari, R. Effects on erythroid differentiation of platinum(II) complexes of synthetic bile acid derivatives. Bioorg. Med. Chem. 2006, 14, 5204–5210. [Google Scholar] [CrossRef]
  208. Kolliopoulou, A.; Stratopoulos, A.; Siamoglou, S.; Sgourou, A.; Ali, B.R.; Papachatzopoulou, A.; Katsila, T.; Patrinos, G.P. Key Pharmacogenomic Considerations for Sickle Cell Disease Patients. OMICS 2017, 21, 314–322. [Google Scholar] [CrossRef]
  209. Kountouris, P.; Lederer, C.W.; Fanis, P.; Feleki, X.; Old, J.; Kleanthous, M. IthaGenes: An interactive database for haemoglobin variations and epidemiology. PLoS ONE 2014, 9, e103020. [Google Scholar] [CrossRef]
  210. Gambari, R. The Role of OMICS Research in Understanding Phenotype Variation in Thalassaemia: The THALAMOSS Project. Thalass. Rep. 2014, 4, 4877. [Google Scholar] [CrossRef]
  211. Mañú Pereira, M.D.M.; Colombatti, R.; Alvarez, F.; Bartolucci, P.; Bento, C.; Brunetta, A.L.; Cela, E.; Christou, S.; Collado, A.; de Montalembert, M.; et al. Sickle cell disease landscape and challenges in the EU: The ERN-EuroBloodNet perspective. Lancet Haematol. 2023, 10, e687–e694. [Google Scholar] [CrossRef]
  212. Halim-Fikri, B.H.; Lederer, C.W.; Baig, A.A.; Mat-Ghani, S.N.A.; Syed-Hassan, S.R.; Yusof, W.; Abdul Rashid, D.; Azman, N.F.; Fucharoen, S.; Panigoro, R.; et al. Global Globin Network Consensus Paper: Classification and Stratified Roadmaps for Improved Thalassaemia Care and Prevention in 32 Countries. J. Pers. Med. 2022, 12, 552. [Google Scholar] [CrossRef] [PubMed]
  213. Kountouris, P.; Stephanou, C.; Lederer, C.W.; Traeger-Synodinos, J.; Bento, C.; Harteveld, C.L.; Fylaktou, E.; Koopmann, T.T.; Halim-Fikri, H.; Michailidou, K.; et al. ClinGen Hemoglobinopathy Variant Curation Expert Panel. Adapting the ACMG/AMP variant classification framework: A perspective from the ClinGen Hemoglobinopathy Variant Curation Expert Panel. Hum. Mutat. 2022, 43, 1089–1096. [Google Scholar] [CrossRef] [PubMed]
  214. Kountouris, P.; Stephanou, C.; Archer, N.; Bonifazi, F.; Giannuzzi, V.; Kuo, K.H.M.; Maggio, A.; Makani, J.; Mañú-Pereira, M.D.M.; Michailidou, K.; et al. The International Hemoglobinopathy Research Network (INHERENT): An international initiative to study the role of genetic modifiers in hemoglobinopathies. Am. J. Hematol. 2021, 96, E416–E420. [Google Scholar] [CrossRef] [PubMed]
  215. Liu, D.; Nong, C.; Lai, F.; Tang, Y.; Wang, T. Research Progress of Cell-Free Fetal DNA in Non-Invasive Prenatal Diagnosis of Thalassemia. Hemoglobin 2023, 47, 80–84. [Google Scholar] [CrossRef] [PubMed]
  216. Yang, X.; Ye, Y.; Fan, D.; Lin, S.; Li, M.; Hou, H.; Zhang, J.; Yang, X. Non-invasive prenatal diagnosis of thalassemia through multiplex PCR, target capture and next-generation sequencing. Mol. Med. Rep. 2020, 22, 1547–1557. [Google Scholar] [CrossRef] [PubMed]
  217. Maskoen, A.M.; Rahayu, N.S.; Laksono, B.; Fibriani, A.; Soewondo, W.; Mose, J.C.; Sahiratmadja, E.; Panigoro, R. Cell-free fetal DNA as a non-invasive method using pyrosequencing in detecting beta-globin gene mutation: A pilot study from area with limited facilities in Indonesia. Front. Pediatr. 2022, 10, 902879. [Google Scholar] [CrossRef]
  218. Sawakwongpra, K.; Tangmansakulchai, K.; Ngonsawan, W.; Promwan, S.; Chanchamroen, S.; Quangkananurug, W.; Sriswasdi, S.; Jantarasaengaram, S.; Ponnikorn, S. Droplet-based digital PCR for non-invasive prenatal genetic diagnosis of α and β-thalassemia. Biomed. Rep. 2021, 15, 82. [Google Scholar] [CrossRef] [PubMed]
  219. D’Aversa, E.; Breveglieri, G.; Boutou, E.; Balassopoulou, A.; Voskaridou, E.; Pellegatti, P.; Guerra, G.; Scapoli, C.; Gambari, R.; Borgatti, M. Droplet Digital PCR for Non-Invasive Prenatal Detection of Fetal Single-Gene Point Mutations in Maternal Plasma. Int. J. Mol. Sci. 2022, 23, 2819. [Google Scholar] [CrossRef]
  220. Constantinou, C.G.; Karitzi, E.; Byrou, S.; Stephanou, C.; Michailidou, K.; Makariou, C.; Hadjilambi, G.; Christofides, A.; Kleanthous, M.; Papasavva, T. Optimized Droplet Digital PCR Assay on Cell-Free DNA Samples for Non-Invasive Prenatal Diagnosis: Application to Beta-Thalassemia. Clin. Chem. 2022, 68, 1053–1063. [Google Scholar] [CrossRef]
Figure 1. Possible interactions between strategies commonly used and/or proposed for the therapeutic intervention on β-thalassemia and SCD and organizations/networks aiming to facilitate the translation of pharmacogenomics-based information into actionable prescribing decisions for the drugs/agents employed in their treatment. (*) Luspatercept and Mitapivat are under investigation in SCD.
Figure 1. Possible interactions between strategies commonly used and/or proposed for the therapeutic intervention on β-thalassemia and SCD and organizations/networks aiming to facilitate the translation of pharmacogenomics-based information into actionable prescribing decisions for the drugs/agents employed in their treatment. (*) Luspatercept and Mitapivat are under investigation in SCD.
Ijms 25 04263 g001
Figure 2. Different therapeutic options for β-thalassemia and SCD in the context of available pharmacogenomics-based studies (most of which are considered in the present review). Red arrows and boxes indicate therapeutic options for which pharmacogenomics information is already available. Blue arrows and boxes indicate therapeutic options for which pharmacogenomics information is expected in the near future.
Figure 2. Different therapeutic options for β-thalassemia and SCD in the context of available pharmacogenomics-based studies (most of which are considered in the present review). Red arrows and boxes indicate therapeutic options for which pharmacogenomics information is already available. Blue arrows and boxes indicate therapeutic options for which pharmacogenomics information is expected in the near future.
Ijms 25 04263 g002
Figure 3. Flow chart of the evaluation of the effects of DNA polymorphism(s) on the response of β-thalassemia patients to an HbF inducer. In the ex vivo protocols, the objective is to verify whether cells isolated from the patient(s) respond to in vitro induction (i.e., increased γ-globin gene expression and HbF production following treatment). The in vivo protocols are based on the treatment of β-thalassemia patients with the HbF inducer (endpoint: reduction, or elimination, of red blood transfusions). In the case of a correlation existing between ex vivo and in vivo protocols, ex vivo assays are predictive of in vivo response.
Figure 3. Flow chart of the evaluation of the effects of DNA polymorphism(s) on the response of β-thalassemia patients to an HbF inducer. In the ex vivo protocols, the objective is to verify whether cells isolated from the patient(s) respond to in vitro induction (i.e., increased γ-globin gene expression and HbF production following treatment). The in vivo protocols are based on the treatment of β-thalassemia patients with the HbF inducer (endpoint: reduction, or elimination, of red blood transfusions). In the case of a correlation existing between ex vivo and in vivo protocols, ex vivo assays are predictive of in vivo response.
Ijms 25 04263 g003
Figure 4. The interplay between GWAS approaches, gene editing, and pharmacogenomics. These combined approaches can ultimately help in identifying novel targets to be modulated with novel drugs to be further studied with pharmacogenomic protocols. (HPFH: hereditary persistence of fetal haemoglobin; HbF: fetal hemoglobin; CRISPR: Clustered regularly interspaced palindromic repeats.
Figure 4. The interplay between GWAS approaches, gene editing, and pharmacogenomics. These combined approaches can ultimately help in identifying novel targets to be modulated with novel drugs to be further studied with pharmacogenomic protocols. (HPFH: hereditary persistence of fetal haemoglobin; HbF: fetal hemoglobin; CRISPR: Clustered regularly interspaced palindromic repeats.
Ijms 25 04263 g004
Figure 5. The basic steps of “PharmacomiRNomics” of HbF inducers in β-thalassemia and SCD.
Figure 5. The basic steps of “PharmacomiRNomics” of HbF inducers in β-thalassemia and SCD.
Ijms 25 04263 g005
Table 1. Fetal hemoglobin inducers and ongoing clinical trials.
Table 1. Fetal hemoglobin inducers and ongoing clinical trials.
AgentCommentsReferences
5-AzacytidineDNA hypomethylationKalantri et al., 2018 [60]
HydroxyureaDNA hypomethylationFucharoen et al., 1996 [61]
ButyrateHDAC activity inhibitionBianchi et al., 2015 [62]
2,2-dimethylbutyrate
(HQK-1001)
A phase II study of efficacy and safety of the oral fetal globin inducer, HQK-1001 Reid et al., 2014 [63]
ValproateActivation of p38 MAPK pathway, HDAC inhibitionRönndahl et al., 2006 [64]
Thalidomide derivatives Activation of p38 MAPK pathwayAerbajinai et al., 2007 [65]
Moutouh-De Parseval et al., 2008 [66]
Trichostatin AHDAC inhibitionMarianna et al., 2001 [67]
OridoninActivation of p38 MAPK signalingGuo L et al., 2020 [68]
Rapamycin (sirolimus) mTOR inhibitorFibach et al., 2006 [69]
Zuccato et al., 2022 [70]
EverolimusmTOR inhibitorZuccato et al., 2007 [71]
MithramycinInhibition of Sp1/DNA interactionsFibach et al., 2003 [72]
ResveratrolActivation of p38 MAPK signalingFibach et al., 2012 [73]
AngelicinInduction of γ-globin expression via NRF2/ARE stress response pathwayLampronti et al., 2003 [74]
Cinchonidine, Quinidine and CinchonineCinchona alkaloids are potent inducers of the expression of γ-globin genes in erythroid cellsIftikhar et al., 2019 [75]
Zuccato et al., 2021 [76]
TranylcypromineLysine-specific demethylase 1 (LSD1) inhibitionShi et al., 2013 [77]
VorinostatHDAC inhibition reduces the expression of α-globin, induces γ-globin expressionMettananda et al., 2019 [78]
Vasicinol and Vasicine Induction of γ-globin genes in a pre-clinical study of HbF inducers isolated from Adhatoda vasicaIftikhar et al., 2022 [79]
HydroxyureaClinical trials NCT00001958 and NCT03183375Yasara et al., 2022 [80]
HQK-1001Clinical trial NCT00790127Reid et al. [63]
ThalidomideClinical trial NCT05132270Ansari et al. [81]
SirolimusClinical trials NCT03877809 and NCT04247750Gamberini et al., 2021 [82]
Table 2. Examples of non-coding RNAs regulating γ-globin gene expression.
Table 2. Examples of non-coding RNAs regulating γ-globin gene expression.
Non-Coding RNATarget mRNACommentsReference
miR-34aSTAT3Stable expression mediates fetal hemoglobin induction Ward et al., 2016 [138]
miR-210-3pBCL11AHigh expression associated with high γ-globin mRNA content; down-regulated BCL11A in erythroid precursor cells induced with mithramycinGasparello et al., 2017 [143]
miR-92a-3pBCL11AUpregulates γ-globin expression and inhibits oxidative stress and apoptosis in erythroid precursor cellsLi H et al., 2022 [144]
miR-30aBCL11ARegulation of γ-globin gene expression through targeting BCL11AGholampour et al., 2020 [145]
miR-2355-5pKLF6Upregulates γ-globin gene expressionCheng et al., 2021 [146]
miR-144NRF2Regulates oxidative stress tolerance of thalassemic erythroid cellsSrinoun et al., 2019 [147]
miR-326EKLFRegulates HbF synthesis by targeting EKLF in human erythroid cellsLi et al., 2018 [148]
miR-155BACH1Enhances phagocytic activity of β-thalassemia/HbE monocytes Srinoun et al., 2017 [149]
miR-26bMYB 3′UTRUpregulates HbF expression in erythroleukemic K-562 cell line.Pule et al., 2016 [150]
miR-29b-3pSp1Mediates epigenetic mechanisms of HBG gene activationStarlard-Davenport A 2019 [151]
miR-486-3pBCL11ARegulation of γ-globin gene expression through direct modulation of BCL11ALulli et al., 2013 [152]
miR-15a and miR-16-1MYBElevates fetal hemoglobin expressionSankaran et al., 2011 [153]
miR-190b-5pBCL11ARegulates BCL11A expressionChen et al., 2023 [154]
circ-0008102tbd (*)May serve as a novel clinical biomarker in beta-thalassemiaChen et al., 2024 [155]
lncRNA H19tbd (*)Leads to upregulation of γ-globin gene expressionXie et al., 2024 [156]
lncRNA HBBP1ELK1Leads to upregulation of γ-globin gene expressionA et al., 2021 [157]
lncRNA HMItbd(*)Regulates human fetal hemoglobin expressionMorrison et al., 2018 [158]
(*) tbd = To be determined.
Table 3. Examples of pharmacogenomics gene–drug association: drugs employed or proposed for SCA therapy.
Table 3. Examples of pharmacogenomics gene–drug association: drugs employed or proposed for SCA therapy.
Gene/
Biomarker
DrugDrug
RxNorm
ID
Medical Intervention
(Established or Proposed)
FDA Labeling SectionsCPIC Guideline Classification of Recommendation Reference from the Literature
CYP2D6Codeine2670Pain managementBoxed Warning, Warnings and Precautions, Use in Specific Populations, Patient Counseling Informationhttps://cpicpgx.org/guidelines/guideline-for-codeine-and-cyp2d6/
(accessed on 25 March 2024)
Gammal et al. (2016) [194]
CYP2C9Ibuprofen5640Anti-inflammatory analgesicsn.a. (*)https://cpicpgx.org/guidelines/cpic-guideline-for-nsaids-based-on-cyp2c9-genotype/
(accessed on 25 March 2024)
Cho et al. (2016) [204]
CYP3A4Rosuvastatin301542Reduction in serum cholesterol and cardiovascular complicationsClinical Pharmacologyhttps://cpicpgx.org/guidelines/cpic-guideline-for-statins/
(accessed on 25 March 2024)
Adam et al. (2013) [205]
OTC, POLGValproic acid11118Fetal hemoglobin inductionBoxed Warning, Contraindications, Warnings and Precautions; Testing requiredn.a. (*)Selby et al. (1997) [206]
CYP3A5Sirolimus35302Fetal hemoglobin inductionn.a. (*)Provisional (accessed on 25 March 2024).Zuccato et al. (2022) [70]
ACTP2,
ERCC1,
TPMT
Cisplatin2555Fetal hemoglobin induction, Bone marrow transplantationAdverse ReactionsProvisional (accessed on 25 March 2024).Lampronti et al. (2006) [207]
CYP2D6Quinidine9068Fetal hemoglobin inductionWarnings and Precautions, Clinical PharmacologyProvisional (accessed on 25 March 2024).Zuccato et al. (2021) [76]
CYP2D6Quinine9071Fetal hemoglobin inductionDrug Interactions, Warnings and PrecautionsProvisional (accessed on 25 March 2024).Zuccato et al. (2021) [76]
(*) n.a. = not available.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gambari, R.; Waziri, A.D.; Goonasekera, H.; Peprah, E. Pharmacogenomics of Drugs Used in β-Thalassemia and Sickle-Cell Disease: From Basic Research to Clinical Applications. Int. J. Mol. Sci. 2024, 25, 4263. https://doi.org/10.3390/ijms25084263

AMA Style

Gambari R, Waziri AD, Goonasekera H, Peprah E. Pharmacogenomics of Drugs Used in β-Thalassemia and Sickle-Cell Disease: From Basic Research to Clinical Applications. International Journal of Molecular Sciences. 2024; 25(8):4263. https://doi.org/10.3390/ijms25084263

Chicago/Turabian Style

Gambari, Roberto, Aliyu Dahiru Waziri, Hemali Goonasekera, and Emmanuel Peprah. 2024. "Pharmacogenomics of Drugs Used in β-Thalassemia and Sickle-Cell Disease: From Basic Research to Clinical Applications" International Journal of Molecular Sciences 25, no. 8: 4263. https://doi.org/10.3390/ijms25084263

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