Next Article in Journal
6-Bromoindole- and 6-Bromoindazole-Based Inhibitors of Bacterial Cystathionine γ-Lyase Containing 3-Aminothiophene-2-Carboxylate Moiety
Previous Article in Journal
Differences in Nanoplastic Formation Behavior Between High-Density Polyethylene and Low-Density Polyethylene
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Hemoglobin Variants as Targets for Stabilizing Drugs

by
Miroslava Žoldáková
1,
Michal Novotný
2,
Krishna P. Khakurel
3,* and
Gabriel Žoldák
1,*
1
Faculty of Science, Pavol Jozef Šafárik University in Košice, Park Angelinum 19, 040 01 Košice, Slovakia
2
AURORA R&D s.r.o., Mojmírova 12, 040 01 Košice, Slovakia
3
Extreme Light Infrastructure ERIC, Za Radnici 835, 25241 Dolni Brezany, Czech Republic
*
Authors to whom correspondence should be addressed.
Molecules 2025, 30(2), 385; https://doi.org/10.3390/molecules30020385
Submission received: 12 November 2024 / Revised: 27 December 2024 / Accepted: 8 January 2025 / Published: 17 January 2025

Abstract

:
Hemoglobin is an oxygen-transport protein in red blood cells that interacts with multiple ligands, e.g., oxygen, carbon dioxide, carbon monoxide, and nitric oxide. Genetic variations in hemoglobin chains, such as those underlying sickle cell disease and thalassemias, present substantial clinical challenges. Here, we review the progress in research, including the use of allosteric modulators, pharmacological chaperones, and antioxidant treatments, which has begun to improve hemoglobin stability and oxygen affinity. According to UniProt (as of 7 August 2024), 819 variants of the α-hemoglobin subunit and 771 variants of the β-hemoglobin subunit have been documented, with over 116 classified as unstable. These data demonstrate the urgent need to develop variant-specific stabilizing options. Beyond small-molecule drugs/binders, novel protein-based strategies—such as engineered hemoglobin-binding proteins (including falcilysin, llama-derived nanobodies, and α-hemoglobin-stabilizing proteins)—offer promising new options. As our understanding of hemoglobin’s structural and functional diversity grows, so does the potential for genotype-driven approaches. Continued research into hemoglobin stabilization and ligand-binding modification may yield more precise, effective treatments and pave the way toward effective strategies for hemoglobinopathies.

1. Introduction

Hemoglobin (Hb) is a crucial protein for oxygen transport in mammals, with a complex structure and function. The cooperative binding of oxygen to Hb involves an equilibrium between two quaternary structures with different oxygen affinities [1,2]. This process is influenced by various factors, including pH (the Bohr effect), organic phosphates, and other allosteric effectors [3,4]. The two-state model proposed by Monod, Wyman, and Changeux has been widely used to explain Hb cooperativity, although some discrepancies have been observed [2].
Hb exhibits complex allosteric regulation, crucial for its oxygen-binding function. The interplay between tense and relaxed states governs Hb’s behavior [5,6]. Allosteric effectors, both endogenous and synthetic, modulate Hb’s oxygen affinity by binding to specific sites, altering its structure and function [7,8]. This allosteric control involves both homotropic and heterotropic effects, where ligand binding at one site influences binding at other sites [9]. Protein engineering strategies have explored mutations to optimize Hb’s functional properties, revealing additive effects of interfacial and distal-heme pocket mutations on structure and function [10]. The evolution of Hb has resulted in diverse variants with altered oxygen-binding properties, while maintaining fundamental structural features [11]. Understanding these mechanisms is crucial for developing Hb-based oxygen carriers and treating blood-related disorders [12]. Hemoglobin oxygen affinity is influenced by several physiological ligands (Figure 1): (1) pH aka Bohr effect—a decrease in pH reduces hemoglobin’s oxygen affinity, promoting oxygen release in tissues [3]; (2) CO2 levels—increased CO2 levels form carbaminohemoglobin, stabilizing the T state and reducing oxygen affinity [13]; (3) NO, which can bind to Hb’s heme groups, forming nitrosylhemoglobin, or to cysteine residues, creating S-nitrosohemoglobin (SNO-Hb) [14,15] and (4) 2,3-Bisphosphoglycerate (2,3-BPG)—this metabolite binds to deoxygenated hemoglobin, stabilizing the T state and promoting oxygen release in tissues. 2,3-Bisphosphoglycerate (2,3-BPG) is a crucial metabolite that binds to hemoglobin, regulating oxygen affinity and delivery to tissues. It binds more strongly to deoxyhemoglobin than oxyhemoglobin, with one binding site per tetramer [16,17,18]. 2,3-BPG levels increase during hypoxia, enhancing oxygen delivery [19]. Its concentration is regulated by complex interactions between glycolysis, the 2,3-BPG shunt, and the pentose phosphate pathway [20]. Feedback inhibition of hexokinase and phosphofructokinase by 2,3-BPG, along with product inhibition of 2,3-BPG synthase, control its steady-state concentration. H+ and oxygen effectively regulate 2,3-BPG levels, primarily through hexokinase and phosphofructokinase. The 2,3-BPG shunt also plays a role in stabilizing ATP levels.
Hemoglobin (Hb) is an essential protein for oxygen transport in mammals, with a complex allosteric regulation that enables efficient oxygen delivery to tissues. Genetic mutations affecting hemoglobin structure and function lead to a wide range of disorders, including hemoglobinopathies such as sickle cell disease and thalassemias. These conditions represent significant health challenges worldwide, driving the need for targeted and effective therapeutic strategies.
The objective of this review is to synthesize current knowledge on the physiological roles of hemoglobin and the clinical implications of its variants. We focus on exploring the mechanisms of hemoglobin stability and instability, the impact of genetic diversity, and the therapeutic potential of small molecules and protein engineering. By addressing these aspects, the review aims to provide insights into the development of personalized therapies for hemoglobinopathies. This work also emphasizes the importance of understanding structural dynamics and variant-specific responses.

2. Hemoglobin Function and Clinical Variants

The α and β globin genes are located on different chromosomes, with α globin genes on chromosome 16 and β globin genes on chromosome 11. The arrangement and expression of these genes are tightly regulated to ensure proper hemoglobin formation and function. Two main natural variants exist: fetal hemoglobin (HbF, α2γ2) and adult hemoglobin (HbA, α2β2) [22]. HbF has higher oxygen affinity than HbA, facilitating oxygen transfer from mother to fetus [23]. During development, HbF levels decrease from 70% at birth to <1% in adults, while HbA becomes predominant [24,25]. HbA2 (α2δ2) comprises 1.5–3.5% of adult hemoglobin [26]. Genetic factors, including mutations and polymorphisms, can influence HbF production in adulthood [27]. Structural hemoglobinopathies and thalassemias result from mutations affecting globin genes, leading to abnormal or decreased hemoglobin production [28]. These variants can interfere with measurements of glycated hemoglobin, HbA1c, impacting diabetes monitoring [29].
The α- and β-globin gene clusters are characterized by sophisticated genetic regulatory mechanisms designed for expression control during various developmental stages. The regulation process of globin genes was covered in a recent review [30]. The α-globin cluster is located on chromosome 16 at position 16p13.3. The α-globin cluster encompasses key genes including two nearly identical α-globin genes (HBA1 and HBA2) and pseudogenes. These genes are primarily expressed in postnatal and adult life. Additionally, the embryonic ζ-globin genes (HBZ), expressed during the embryonic stage, are also part of this cluster. The entire α-globin cluster spans approximately 30 kilobases and includes regulatory elements crucial for gene expression. The β-globin cluster is situated on chromosome 11 at position 11p15.5, the β-globin cluster extends over about 60 kilobases. It contains several genes essential for hemoglobin production through different stages of development: ε (HBE), expressed in the embryonic stage; two fetal γ-globin genes (HBG1, HBG2); and the adult δ (HBD) and β (HBB) globin genes. The locus control region (LCR), located upstream of these genes, plays a pivotal role in regulating their expression through chromatin looping mechanisms, ensuring that each gene is expressed at the appropriate developmental stage. The regulation of these gene clusters is complex and involves multiple mechanisms. LCRs are key regulatory DNA sequences that interact with globin genes to enhance their transcription. The LCRs achieve this by forming chromatin loops that bring the LCRs into proximity with the genes they regulate, facilitating a high level of gene expression during specific developmental stages. In addition to that several transcription factors and chromatin modifications are involved. Transcription factors, along with various chromatin modifications such as DNA methylation and histone acetylation or methylation, modulate the transcriptional activity of these genes. For instance, during the transition from fetal to adult hemoglobin, specific transcription factors such as BCL11A play a role in repressing γ-globin gene expression while enhancing β-globin gene expression.
Hence, during the embryonic stage, the ε and ζ genes are primarily active, producing embryonic variants of hemoglobin. Next, during the fetal stage, γ-globin genes dominate, producing fetal hemoglobin (HbF), which has a higher affinity for oxygen than adult hemoglobin. During the adult stage, the expression shifts towards the δ and β genes, which contribute to the formation of adult hemoglobin (HbA). The transition from fetal (HbF) to adult hemoglobin (HbA) involves a decrease in the expression of γ-globin coupled with an increase in β-globin production. This switch is critical for adapting the oxygen-carrying capacity of hemoglobin to the lower oxygen environment outside the womb. The regulatory mechanisms behind this switch are crucial for understanding diseases like sickle cell disease and β-thalassemia, where therapeutic strategies often involve reactivating fetal hemoglobin production to counteract the effects of defective β-globin.
In summary, the regulation of α-globin and β-globin gene expression is a complex process involving multiple transcription factors, epigenetic modifications, and chromatin remodeling (see [30] for more details).

2.1. Diversity of Genetic Variants of Hemoglobin

Hemoglobin variants are naturally occurring mutations in the hemoglobin molecule that can have significant implications for clinical diagnosis and treatment. According to the World Health Organization, approximately 7% of the global population is affected by hemoglobin variants [31]. Hemoglobin variants can lead to various hematological disorders, such as sickle cell disease and β-thalassemia, which require effective therapeutic strategies. Several approaches have been explored to stabilize hemoglobin variants and mitigate the associated clinical complications. One promising approach is the induction of fetal hemoglobin (HbF) expression. Stabilizing the structure and function of hemoglobin variants is also crucial. The discovery of the alpha-hemoglobin stabilizing protein (AHSP), a chaperone for free alpha-hemoglobin, has provided insights into the importance of balanced globin levels for effective [32]. Targeting the interaction between AHSP and abnormal hemoglobin variants may help stabilize these variants and mitigate the associated pathologies [33]. Another viable option is to explore already-known small-molecule binders of hemoglobin (see Section 2.4). Furthermore, gene therapy approaches, such as CRISPR-Cas9-mediated knockin of the HBB gene, have shown promise in correcting hemoglobin disorders [34]. Lentiviral vector-mediated gene therapy has also demonstrated long-term safety and stability in preclinical models [35,36].
Around 10 years ago, over 1000 naturally occurring hemoglobin variants were identified; these variants show various effects on hemoglobin structure and biochemical properties with varying physiological effects [37]. Since then, several new variants have been found, and their sequences can be retrieved from the UniProt database. UniProt entries for human hemoglobin α and β subunits are
In the UniProt database, there are 819 variants of the hemoglobin subunit α and 771 variants of the hemoglobin subunit β (as of 7 August 2024, Figure 2). Assuming homozygotes, there are 631,749 potential combinations of hemoglobin. However, the genetic diversity increases exponentially when considering heterozygotes, resulting in over 1011 possible combinations. This number surpasses the current human population on Earth, illustrating the astonishing potential variations in known combinations of hemoglobin variants. The actual diversity is likely much lower because Hb variants are often present in specific geographically constrained sub-populations (see Table 1). For example, HbS is prevalent in Africa and India, thalassemias are common in the Mediterranean, Middle East, and Southeast Asia, HbC is common in West Africa, HbD-Punjab has been found in the Indian subcontinent, HbE is highly prevalent in Southeast Asia, Hb Lepore is prevalent in Southern Italy and Mediterranean populations, Hb Constant Spring (Hb CS) is common in Southeast Asia, Hb Bart’s (γ4) has been found in individuals with α-thalassemia, prevalent in Saudi Arabia and Thailand. Summarizing details can be found in Table 1 and in the main text.
The classification of genetic variants, particularly in hemoglobinopathies, is crucial for accurate diagnosis and clinical management. The American College of Medical Genetics and Genomics (ACMG) guidelines provide a five-category system: pathogenic, likely pathogenic, uncertain significance, likely benign, and benign [38,39]. However, this system has limitations in addressing the spectrum of variant effects. An expanded seven-category system, including “predisposing” and “likely predisposing”, has been proposed to better classify variants in disease-causing genes [38,40]. Variant reclassification, driven by new evidence, can significantly impact patient care, with reclassification rates ranging from 3.6% to 58.8% [41]. For hemoglobinopathies, molecular genetic testing is essential for accurate diagnosis and genetic counseling [42,43]. The clinical implications of variant classification are profound, affecting organ surveillance, prophylactic surgery, and cascade testing [44].
Hemoglobinopathies are a group of disorders affecting the hemoglobin molecule. These disorders typically result from genetic mutations that lead to abnormal structure or production of hemoglobin, causing various health problems ranging from mild anemia to severe complications (Table 1). For example, HbM variants lead to methemoglobinemia, where iron in the heme group is oxidized to the ferric form, impairing oxygen delivery. Methemoglobinemia is a blood disorder characterized by elevated levels of methemoglobin, which impairs oxygen delivery to tissues [45]. It can be inherited or acquired, with the latter being more common and often caused by medications or chemicals [46]. Symptoms include cyanosis unresponsive to oxygen therapy and chocolate-colored blood [47].
Diagnosis is confirmed through co-oximetry, arterial blood gas analysis, and serum methemoglobin levels [48]. Treatment involves removing the offending agent, administering oxygen, and using methylene blue as an antidote in severe cases [49]. Dapsone is a common cause of acquired methemoglobinemia, while benzocaine spray has been associated with severe cases [50]. In pediatric patients, underlying hematologic diseases and G6PD deficiency can predispose to methemoglobinemia [51]. Increased awareness and primary prevention efforts are crucial to reduce morbidity and mortality associated with this condition [50].
Table 1. Summary of selected hemoglobin variants, their genetic background, clinical presentations, and diagnostic/therapeutic considerations.
Table 1. Summary of selected hemoglobin variants, their genetic background, clinical presentations, and diagnostic/therapeutic considerations.
Hb VariantMutationGeo. DistributionClinical FeaturesDiag. MethodNotes
HbCβ-globin gene mutation (Glu → Lys at position 6) [52]Common in West AfricaMild chronic hemolysis, splenomegaly, jaundice in homozygotes; compound heterozygosity with HbS can worsen severity [52,53]Hemoglobin electrophoresis, visualization of HbC crystals [54,55]Affects RBC rigidity and may influence acquired immunity to malaria [56]
HbD-Punjabβ-globin gene mutation at codon 121 (Glu → Gln) [57]Relatively low (~0.06% in one study) [57]; found in Indian subcontinent and other regionsGenerally asymptomatic in heterozygotes; homozygotes may have mild hemolytic anemia and splenomegaly; severity increases when co-inherited with SCD or thalassemias [31,57,58,59,60,61,62]Hemoglobin electrophoresis, molecular studies [60]Treatment may include transfusions, hydroxyurea; HbSD-Punjab can mimic sickle cell disease [59]
HbEβ-globin gene mutation (Glu → Lys at position 26) [63]Common in Southeast Asia; frequencies up to 0.433 in some Lao populations [64]Variable severity from asymptomatic to severe anemia when co-inherited with β-thalassemia; altered redox properties and susceptibility to oxidative damage [65,66,67,68]Hemoglobin electrophoresis, molecular analysisOrigin linked to malarial selection; minimal allosteric changes but altered redox properties [63,65]
Hb Bart’s (γ4)α-globin gene deletions leading to excess γ chains (γ4 tetramers) [69]High rates in regions with α-thalassemia (e.g., Saudi Arabia, Thailand) [70,71]Indicates α-thalassemia severity; elevated levels correspond to number of α-globin gene deletions; high oxygen affinity but no cooperativity [69,72,73,74]Quantification in cord blood; spectroscopic studies [74,75]In some cases, elevated Hb Bart’s may be due to developmental asynchrony rather than true α-thalassemia [76]
Hb Chesapeakeα92 Arg → Leu mutation [77,78]Reported in certain families; not widely prevalentHigh oxygen affinity variant causing mild erythrocytosis; reduces tissue oxygen delivery, triggering increased RBC production [77,78,79,80]Hemoglobin electrophoresis, oxygen dissociation studies, molecular analysisSimilar to other high-affinity variants (e.g., Hb Kempsey); fetal form also has increased O2 affinity [77,78]
Hb LeporeFusion of δ and β globin genes due to unequal crossover [81,82]Prevalent in Southern Italy, found globally due to migration [83]β-thalassemia minor phenotype in carriers; severe anemia in homozygotes or with co-inherited β-thalassemia [83,84,85,86]Hematological, biochemical, and molecular analyses [82]Variants include Lepore-Boston, Lepore-Hollandia, and Lepore-Washington-Boston; new variant Hb Lepore-Hong Kong identified [87]
Hb Constant Spring (CS)Mutation in α2-globin termination codon, elongating α-chain [88]Common in Southeast Asia; gene frequencies vary from 0.008 in Thailand to 0.143 in Vietnam [89,90]Can lead to thalassemia intermedia when combined with α-thalassemia; rare homozygous cases cause fetal anemia and hydrops [91,92]Selective enzymatic amplification of α2-globin DNA; hemoglobin electrophoresis [93]Interferes with glycated hemoglobin measurements; important for genetic counseling [94]
Hb O-Arabβ121 Glu → Lys mutation [95,96]Found in populations from the Middle East, North Africa, African Americans, West Africans [95,97,98]Generally mild in homozygous form, but compound heterozygosity (e.g., Hb S/O Arab) can cause severe disease [95,99]Hemoglobin electrophoresis with specialized techniques [95]Management may involve transfusions and splenectomy [100]
Hb Seal RockExtended α-chain variant [101]Rare, limited reportsAssociated with mild Hb H disease and α-thalassemia-2 trait [101]Hemoglobin electrophoresis, molecular studiesImpact depends on mutation location within the gene [102]
Hb IndianapolisRare, unstable β-globin variant [103]Reported in a Brazilian patient [103]Moderate hemolytic anemia and renal damage [103]Routine DNA sequencing of globin genes [104]Highlights the growing diversity of novel β-chain variants
Unstable Variants (e.g., Hb Madrid, Hb Showa-Yakushiji, Hb Santander, Hb Yokohama, Hb Seattle, Hb Miami, Hb Hershey, Hb Abington)Various mutations in β-globin affecting amino acid positions [105,106,107,108,109,110,111]Reported in disparate geographic locations (Spain, Republic of Korea, India, Japan, etc.) [105,106,107,108,109,110]Mild to moderate hemolytic anemia; severity often increases with co-inherited thalassemia mutations [105,106,107,108,109,110,111]Hemoglobin electrophoresis, DNA sequencing, RBC morphology analysisDemonstrate clinical severity range and genetic complexity of unstable hemoglobin variants [104,105,106,107,108,109,110,111]
Sickle cell disease (SCD) is a genetic disorder caused by a single nucleotide mutation in the β-globin gene, resulting in the production of abnormal hemoglobin S (HbS) [112,113]. This E6V mutation leads to HbS polymerization under low oxygen conditions, causing red blood cell (RBC) sickling and hemolysis [112,114]. SCD is characterized by chronic inflammation, oxidative stress, and vascular dysfunction, contributing to various complications such as stroke, pulmonary hypertension, and renal disorders [113,115]. The disease affects millions worldwide, with higher prevalence in Africa and India [114]. Recent advancements in care, including newborn screening, penicillin prophylaxis, and hydroxyurea treatment, have improved patient outcomes [116,117]. Ongoing research explores gene editing techniques like CRISPR/Cas9 to target the β-globin gene in hematopoietic stem cells as a potential therapeutic approach [118]. Gene therapy approaches using βT87Q-globin have shown anti-sickling effects by reducing endogenous βS-globin expression [119]. These advancements in antisickling agents and therapies offer new possibilities for treating sickle cell disease.
Thalassemias are inherited disorders of hemoglobin synthesis, characterized by deficient production of globin chains [120,121]. They are prevalent in the Mediterranean, Middle Eastern, and Southeast Asian regions [122,123]. Alpha and beta thalassemias are the most common types, resulting from mutations in α and β globin genes, respectively [124]. Symptoms range from asymptomatic in carriers to severe anemia requiring regular blood transfusions in major forms [125]. Complications include iron overload, cardiac issues, and endocrine disorders [126]. Treatment options include blood transfusions, iron chelation therapy, and bone marrow transplantation [122]. Gene therapy and editing strategies are emerging as potential curative treatments [120]. Prevention through premarital screening and prenatal diagnosis is crucial in reducing the prevalence of thalassemia [126]. Despite its significant impact, thalassemia is often underrecognized in global health assessments [121].
Hemoglobin C (HbC) is a common structural variant of normal hemoglobin, resulting from a mutation in the beta-globin gene [52]. HbC is less soluble than normal hemoglobin, leading to increased rigidity of red blood cells [53]. This can cause mild chronic hemolysis, splenomegaly, and jaundice in homozygous individuals [52]. HbC interacts more strongly with erythrocyte membranes than normal hemoglobin [127]. While HbC disease is generally mild, its inheritance with other hemoglobinopathies like HbS can have serious consequences [52]. Both HbC and HbS may affect the development of acquired immunity against malaria [56]. Efforts have been made to develop simple, inexpensive methods for detecting HbC in resource-limited settings, including microscopic visualization of HbC crystals [54,55]. These methods could potentially determine zygosity and aid in diagnosis in underdeveloped countries where HbC is prevalent.
Hemoglobin D-Punjab (HbD-Punjab) is a variant resulting from a mutation in codon 121 of the β-globin gene [57]. While heterozygous HbD-Punjab is generally asymptomatic, homozygous cases can cause mild hemolytic anemia and splenomegaly [57,128]. The clinical presentation of HbD-Punjab can vary, especially when coinherited with other hemoglobinopathies or thalassemias [58,60]. HbSD-Punjab, a compound heterozygous condition, can lead to severe symptoms resembling sickle cell disease [59]. HbD-Punjab/β-thalassemia combinations may result in moderate to severe anemia [61,62,128]. Diagnosis typically involves hemoglobin electrophoresis and molecular studies [60]. Treatment options include blood transfusions and hydroxyurea, which have shown promise in managing symptoms [59]. The prevalence of HbD-Punjab is relatively low, estimated at 0.06% in one study [57].
Hemoglobin E (HbE), a common β-globin variant in Southeast Asia, results from a Glu26Lys mutation [63,64]. This variant exhibits extended linkage disequilibrium and likely arose 1240–4440 years ago due to malarial selection [63]. HbE frequencies reach up to 0.433 in some Lao populations [64]. While HbE shows minimal allosteric changes, it demonstrates altered redox properties, including decreased nitrite reductase activity and accelerated reduction by cysteine [65]. HbE is more susceptible to oxidative damage, especially in the presence of free α subunits [66]. When co-inherited with β-thalassemia, HbE can cause severe clinical manifestations [67]. However, its association with protection against cerebral malaria remains inconclusive [68]. HbE disorders present a wide range of clinical severity, from asymptomatic carriers to severe anemia [129].
Hemoglobin [130] Bart’s (γ4) is a homotetrameric hemoglobin found in individuals with α-thalassemia [69]. It lacks cooperativity and has a higher oxygen affinity than adult or fetal hemoglobin. Crystal structures reveal that Hb Bart’s resembles the R state of adult hemoglobin [69,72]. Quantification of Hb Bart’s in cord blood accurately indicates α-thalassemia status, with levels correlating to the number of deleted or inactivated α-globin genes [73,74]. The prevalence of Hb Bart’s varies among populations, with high rates reported in Saudi Arabia and Thailand [70,71]. Spectroscopic studies show that Hb Bart’s is structurally similar to Hb H (β4) [75]. In some populations, elevated Hb Bart’s may result from developmental asynchrony rather than α-thalassemia [76].
Hemoglobin Chesapeake (α92 Arg → Leu) is a high-oxygen-affinity hemoglobin variant associated with erythrocytosis [77,78]. This mutation occurs at the α1β2 interface, similar to other high-affinity variants like Hb Wood and Hb Malmö [131]. The increased oxygen affinity results from an early conversion from the T to R state during ligand binding [77]. Hb Chesapeake causes mild erythrocytosis in carriers due to reduced oxygen delivery to tissues, necessitating increased red cell production [79]. This mechanism is also observed in other high-affinity variants like Hb Kempsey [132]. The fetal form of Hb Chesapeake (α2Chesγ2) also exhibits increased oxygen affinity, suggesting similar conformational changes in the γ chain [78]. Generally, high-affinity hemoglobin variants impair the formation of a stable T state or modify the heme environment, leading to compensatory erythrocytosis [80].
Hemoglobin Lepore is a structural variant formed by the fusion of δ and β globin genes, resulting from unequal crossover events [81,82]. It is associated with a β-thalassemia minor phenotype in carriers and can lead to severe anemia in homozygotes or compound heterozygotes with β-thalassemia [83,84]. Several variants exist, including Lepore-Boston, Lepore-Hollandia, and Lepore-Washington-Boston [85,86]. The condition is prevalent in Southern Italy and has spread globally due to migration [83]. Diagnosis requires hematological, biochemical, and molecular analyses [82]. Hb Lepore can interact with other hemoglobinopathies, resulting in various clinical phenotypes [86]. Recently, a novel variant, Hb Lepore-Hong Kong, was identified in a Chinese family [87]. Accurate diagnosis is crucial for genetic counseling and prenatal screening in affected populations.
Hemoglobin Constant Spring (Hb CS) is a variant caused by a mutation in the α2-globin gene termination codon, resulting in an elongated α-chain [88]. It is prevalent in Southeast Asian populations, with gene frequencies ranging from 0.008 in Thailand to 0.143 in Vietnam [89,90]. Hb CS can lead to thalassemia intermedia when combined with α-thalassemia [91]. Diagnosis is challenging due to low amounts of the mutant hemoglobin, but selective enzymatic amplification of α2-globin DNA allows for accurate detection [93]. Hb CS can interfere with glycated hemoglobin measurements using boronate affinity chromatography [94]. In rare cases, homozygous Hb CS can cause fetal anemia and hydrops [92]. Understanding Hb CS is crucial for genetic counseling and prenatal diagnosis in affected populations. Hb O-Arab (Glu121Lys) causes mild hemolytic anemia in populations from the Middle East and North Africa.
Hemoglobin O Arab is a rare abnormal hemoglobin variant characterized by a β121Glu → Lys mutation [95,96]. It can occur in homozygous form or in combination with other hemoglobinopathies. The homozygous form is generally well-tolerated [97,100], while compound heterozygous forms, particularly Hb S/O Arab, can result in severe clinical manifestations like sickle cell anemia [95,99]. Hb S/O Arab patients may experience acute chest syndrome, vasoocclusive events, and other complications [95]. Diagnosis requires specific electrophoresis techniques due to Hb O Arab’s migration patterns [95]. The condition has been reported in various populations, including African Americans, West Africans, and Arabs [95,97,98]. Treatment may involve transfusions and splenectomy in cases of hypersplenism [100].
Hb Seal Rock, an extended α-chain variant, is associated with mild Hb H disease and α-thalassemia-2 trait [101]. Some α-chain variants can lead to chronic hemolytic anemia, with the mutation’s location within the gene sequence playing a crucial role in determining its effect [102]. Hb Indianapolis, a rare and slightly unstable β-globin variant, was reported to cause moderate hemolytic anemia and renal damage in a Brazilian patient [103]. Routine DNA sequencing of α- and β-globin genes has revealed numerous novel mutations, including 11 new β-chain variants, 15 α-chain variants, 19 β-thalassemia mutations, and 15 α+-thalassemia mutations, highlighting the genetic diversity in hemoglobinopathies [104].
Several unstable hemoglobin variants have been identified that cause mild to moderate hemolytic anemia. Hb Madrid (β115Ala → Pro) was found in a Spanish boy and a Korean family, resulting in moderately severe hemolytic anemia [105,106]. Hb Showa-Yakushiji (β110Leu → Pro) was observed in four unrelated Indian patients [107]. Other variants include Hb Santander (β34Val → Asp) in a Spanish patient [108], Hb Yokohama (β31Leu → Pro) in a Japanese family [109], and Hb Seattle (β76Ala → Glu) in a mother and her two sons [110]. Hb Miami (β116His → Pro) and Hb Hershey (β70Ala → Gly) were found in association with beta-thalassemia mutations, while Hb Abington (β70Ala → Pro) was another unstable variant [111]. These variants demonstrate the range of clinical severity observed with unstable hemoglobin mutations, particularly when combined with thalassemic mutations.

2.2. Hemoglobin Instability and Heinz Bodies

Heinz bodies are aggregates of denatured hemoglobin that form within red blood cells (RBCs) as a result of oxidative stress or intrinsic instability of hemoglobin. Their presence is indicative of various hemolytic conditions and can be associated with unstable hemoglobin variants, such as those seen in hemoglobinopathies. The formation of Heinz bodies is a complex process influenced by multiple factors, including oxidative damage, genetic mutations, and environmental toxins. This synthesis aims to explore the stability of hemoglobin in relation to Heinz body formation, drawing on a range of studies that elucidate the mechanisms and implications of this phenomenon.
The historical context of Heinz body formation dates back to the late 19th century when Robert Heinz first described these inclusions in cells subjected to oxidative stress. Since then, research has established that Heinz bodies are not exclusive to unstable hemoglobinopathies; they can also arise in conditions such as glucose-6-phosphate dehydrogenase (G6PD) deficiency, where oxidative stress leads to hemolytic anemia [133]. In cases of unstable hemoglobin variants, such as those documented in hemoglobin Montreal II, the instability of the hemoglobin molecule can precipitate under physiological conditions, resulting in Heinz body formation. This instability is often exacerbated by factors such as viral infections, which can further complicate the clinical picture by increasing hemolysis [133].
The biochemical basis for Heinz body formation involves the denaturation of hemoglobin, which can occur due to oxidative damage from various sources, including environmental toxins. For instance, exposure to polycyclic aromatic hydrocarbons (PAHs) has been shown to induce oxidative injury to hemoglobin, leading to the aggregation of denatured hemoglobin into Heinz bodies [134]. This oxidative damage is not limited to specific species; studies have documented similar phenomena in avian species, where exposure to oil pollutants resulted in the formation of Heinz bodies and subsequent hemolytic anemia [135]. The presence of Heinz bodies in these contexts serves as a biomarker for oxidative stress and cellular damage.
In addition to environmental factors, genetic mutations play a crucial role in the stability of hemoglobin. Variants such as Hb Volga exhibit decreased stability, leading to chronic hemolytic anemia characterized by the formation of Heinz bodies composed of precipitated hemichromes and heme-free globins [136]. More than 100 hemoglobin variants have been identified with reduced stability, highlighting the genetic underpinnings of this condition [136]. Furthermore, the role of molecular chaperones, such as the alpha hemoglobin stabilizing protein (AHSP), has been implicated in maintaining hemoglobin stability. AHSP assists in the proper folding of α-globin chains and prevents their aggregation, thereby reducing the likelihood of Heinz body formation [137].
The physiological implications of Heinz body formation are significant, as these inclusions can compromise the functionality of RBCs. The presence of Heinz bodies can lead to increased erythrocyte fragility and a shortened lifespan, as cells containing these aggregates are often removed from circulation by the spleen [138]. This process can result in hemolytic anemia, characterized by fatigue and decreased oxygen-carrying capacity [139]. The clinical manifestations of Heinz body-related hemolytic anemia can vary, with some individuals exhibiting mild symptoms while others may experience severe anemia requiring medical intervention.
Diagnostic approaches to identify Heinz bodies typically involve supravital staining techniques, which can reveal these inclusions in RBCs. For instance, the use of brilliant cresyl blue staining has been shown to enhance the visibility of Heinz bodies, particularly in cases where oxidative damage has occurred [140]. This staining method can be particularly useful in neonatal populations, where Heinz bodies may be present but not readily detected without appropriate staining protocols [140]. The identification of Heinz bodies in clinical samples can provide valuable insights into the underlying etiology of hemolytic anemia and guide further diagnostic testing, such as hemoglobin electrophoresis or high-performance liquid chromatography (HPLC) to characterize abnormal hemoglobin variants [133].
The formation of Heinz bodies is also influenced by dietary factors and exposure to certain plants known to induce oxidative stress. For example, garlic extracts have been shown to deplete thiol levels in erythrocytes, leading to increased susceptibility to oxidative damage and subsequent Heinz body formation [141]. Similarly, the ingestion of Allium species, such as onions, has been documented to cause oxidative damage to hemoglobin, resulting in the precipitation of denatured hemoglobin and the formation of Heinz bodies in affected animals [142]. These findings underscore the importance of environmental and dietary factors in the pathogenesis of hemolytic anemia associated with Heinz body formation.
In summary, the stability of hemoglobin is a critical factor in the formation of Heinz bodies, with both genetic and environmental influences playing significant roles. Unstable hemoglobin variants, oxidative stress from environmental toxins, and dietary factors can all contribute to the denaturation of hemoglobin and the subsequent aggregation into Heinz bodies. The presence of these inclusions serves as a marker for hemolytic anemia and highlights the need for comprehensive diagnostic approaches to identify underlying causes and guide appropriate management strategies. Future research should continue to explore the molecular mechanisms underlying hemoglobin stability and the factors that contribute to Heinz body formation in various populations.
To validate destabilizing mutations in hemoglobin, further genetic testing is needed. Additionally, it can provide some information for treatment choices by recommending the possible use of specific stabilizers or antioxidants to shield red blood cells, lessen oxidative damage, and enhance patient outcomes [143,144,145]. Of course, hemoglobin genetic variants can differ in their susceptibility to oxidation. Certain unstable variants are more likely to undergo oxidation or lose the heme. Losing the heme group, the heme-free globin chains can eventually result in the production of Heinz bodies. Since they exhibit a variety of inclinations to form these inclusions under oxidative stress, approximately 3.8% of α-globin and 11% of β-globin variations are regarded as unstable [143,146]. This link emphasizes how Heinz bodies reflect the distinct molecular features of particular hemoglobin types. In this review, we highlight Heinz bodies in connection with hemoglobin structural and genetic variations. It might be the case that Heinz bodies themselves might not be the main targets of treatment.
However, Heinz bodies highlight their role as clinical indicators. Their function serves to emphasize how crucial it is to take structural weaknesses and oxidative stress into account when creating management or treatment plans for hemoglobin-related illnesses. In general, hemoglobin structural instability can lead to denaturation and precipitation, resulting in the formation of Heinz bodies, which are inclusions of denatured hemoglobin in red blood cells. This instability is often due to amino acid substitutions or deletions that disrupt the hemoglobin structure, enhancing oxidation to methemoglobin and subsequent conversion to hemichrome [143]. The loss of heme, particularly from beta chains, is a critical factor in Heinz body formation [146,147]. The presence of Heinz bodies can cause hemolytic anemia due to mechanical obstruction in microcirculation and oxidative damage to red cell membranes [130,148]. Additionally, factors such as temperature and oxidative stress can exacerbate the formation of Heinz bodies [144]. Oxidative stress occurs when there is an imbalance between reactive oxygen species (ROS) and the body’s antioxidant defenses. Hemoglobin is particularly vulnerable to oxidative damage, which can lead to hemolysis. This oxidative damage is exacerbated in conditions such as thalassemia and sickle cell disease. Mechanisms leading to Heinz body formation include oxidative damage, genetic mutations, chemical exposure, and enzyme deficiencies. Reactive oxygen species (ROS) can oxidize hemoglobin, causing it to denature and form Heinz bodies. This oxidative stress is common in conditions with an elevated level of ROS, such as inflammation or chronic diseases.
In summary, Heinz bodies are microscopic clusters that form when hemoglobin oxidizes and becomes unstable. They frequently serve as a precursor to hemolytic disorders. Their presence in red blood cells may indicate certain underlying problems, such as hemolysis caused by genetic hemoglobin variations that induce instability.
As mentioned above, genetic factors significantly influence the stability of hemoglobin. Mutations in the genes encoding hemoglobin can lead to structurally unstable hemoglobin variants, making them more susceptible to denaturation and resulting in disorders such as thalassemias and hemoglobinopathies. These genetic mutations can affect the function and stability of hemoglobin molecules, leading to various clinical manifestations. Out of the 819 known variants of the hemoglobin subunit α, 31 are classified as unstable. This represents approximately 3.8% of all α subunit variants. In contrast, 85 out of 771 known variants of the hemoglobin subunit β are considered unstable, accounting for about 11% of the β subunit variants. This comparison highlights a significant difference in the frequency of unstable variants between the two subunits. The frequency of unstable variants in the β subunit is nearly three times higher than that in the α subunit, suggesting that the β-globin gene may be more susceptible to mutations that result in instability or that the structural constraints of the β subunit allow for more potentially unstable configurations.
The human α-globin gene cluster, including HBA1 and HBA2, is located on chromosome 16 at position 16p13.2-pter [149,150]. This region is extremely gene-rich, containing 100 confirmed and 20 predicted genes [151]. The α-globin locus is highly polymorphic, with 15 dimorphic and 2 multiallelic genetic markers identified [152]. Upstream regulatory elements, particularly HS-40, play crucial roles in α-globin gene expression [153]. α-Thalassemia, a disorder of reduced α-chain synthesis, is often caused by large deletions within the gene cluster [154]. Interestingly, a 62 kb deletion upstream of the α-globin genes can also cause α-thalassemia without directly affecting the genes themselves [155]. The α-globin locus serves as an ideal genetic marker for constructing human linkage maps and studying the molecular genetics of the cluster [152].
The β-globin gene cluster, located on chromosome 11 at position 11p15.5, includes the HBB gene, which encodes the β-globin chain of adult hemoglobin [156,157]. This cluster comprises several genes, including HBD, which encodes the δ-globin chain, and undergoes critical expression switches during development [158]. The precise localization of the β-globin gene has been confirmed through various studies, indicating its location within the 11p15.4 to 11p15.5 region [159,160,161]. The regulation of these genes is crucial for understanding hemoglobinopathies, with recent findings highlighting the role of factors like BCL11A in silencing fetal hemoglobin [162]. Overall, the β-globin gene cluster is essential for normal hemoglobin function and is implicated in disorders such as sickle cell anemia and β-thalassemia [157].
The observed difference in the frequency of unstable variants between the α and β subunits, 3.8% vs. 11%, may be due to several factors. One would expect that the α-globin gene cluster location on chromosome 16 and its gene duplication may confer additional stability, resulting in a higher chance of accepting unstable variants compared to the β-globin gene cluster on chromosome 11. The presence of two nearly identical genes (HBA1 and HBA2) in the α-globin cluster means there is a redundancy that can potentially compensate for potential mutations. If one gene is mutated, the other can often continue to produce functional α-globin chains, reducing the impact of any single mutation and enhancing overall stability. This redundancy might act as a genetic buffer, making the α-globin production more resilient to genetic mutations and environmental pressures. However, we found that less unstable variants of α-chain are nearly three-fold less indicating that other factors are responsible for a higher number of unstable β-chain variants. Another factor would be structural differences between the α and β subunits might inherently affect their susceptibility to mutations that lead to instability. While both subunits share a high sequence identity of 43% and 50–60% sequence similarity, they may have distinct tolerance toward mutations constrained by their differences in function.
The β-globin’s versus α-globin’s specific role and interactions in the hemoglobin tetramer may be the reason for higher counts of unstable mutations. In fact, the β-globin plays a crucial role in hemoglobin structure and function. It forms strong interactions with α-globin to create stable α1β1 dimers, which then assemble into α2β2 tetramers [163]. The β-chains significantly influence oxygen binding characteristics, particularly through alterations in the T state properties [164]. The β112 Cys residue is critical for both homo- (β4) and hetero-tetramer (α2β2) formation, with mutations at this position affecting assembly and stability [165]. The β4 tetramer structure closely resembles the R state of liganded α2β2 hemoglobin, but with some unique interface interactions [166]. The β chains also interact with the cytoplasmic domain of band 3 in erythrocyte membranes, preferentially binding to deoxyhemoglobin [167]. On the other hand, α-globin plays a crucial role in hemoglobin formation and function. The α-Hemoglobin Stabilizing Protein (AHSP) is essential for maintaining α-globin stability and preventing its precipitation [168]. AHSP binds specifically to the G and H helices of α-globin, with the N-terminal part of the H helix being critical for this interaction [169,170]. AHSP acts as a molecular chaperone, facilitating α-globin folding, refolding after denaturation, and incorporation into hemoglobin A [171]. It also protects against oxidative damage and helps maintain the correct ratios of α and β globins during hemoglobin biosynthesis [172]. Mutations in α-globin that impair AHSP binding can lead to protein instability and anemia [173]. The complex interplay between α-globin expression, AHSP, and hemoglobin assembly is crucial for normal erythropoiesis [174]. Hence, we can speculate that these differences between globins may result in observed differences in frequencies of unstable variants that α-globin mutations produce a significant disadvantage, so they are much less tolerated than β-globin unstable variants. In addition, a possible correlation can be found between amino acid residues contacting heme and unstable variants (Figure 2d).

2.3. Pharmacological Approaches to Hemoglobin Stabilization and Oxidative Stress in Hemolytic Disorders

Certain molecules and toxins can induce oxidative stress or directly interact with hemoglobin, leading to its denaturation and the formation of so-called Heinz bodies. Even though the below-mentioned small molecules do not directly interact with Hb, they can influence the metabolism of erythrocytes, reduce oxidative stress, or induce other hemoglobin chain expressions, and by doing this, they can provide stability for hemoglobin or prevent sickling.
Glucose-6-phosphate dehydrogenase (G6PD) deficiency, affecting approximately 400 million people worldwide, is a crucial enzyme in protecting red blood cells from oxidative stress [145]. G6PD deficiency can lead to hemolytic anemia, neonatal jaundice, and increased susceptibility to oxidative damage [175]. The deficiency results in reduced NADPH production, compromising the cell’s ability to maintain glutathione in its reduced state [176]. This leads to increased hemoglobin denaturation and ferriheme release, potentially causing hemolysis [177]. Contrary to previous beliefs, NADPH status, rather than glutathione levels, modulates oxidant sensitivity in G6PD-deficient erythrocytes [178]. G6PD deficiency also has broader implications, including increased risk of prenatal and postnatal death, infertility, and teratogenesis [179]. In transfusion medicine, G6PD-deficient blood may pose risks to certain patient populations [180].
Oxidative stress plays a significant role in hemolytic anemias, causing damage to red blood cells and exacerbating symptoms [181]. Antioxidants, such as N-acetylcysteine (NAC) and Vitamin E, have shown promise in reducing oxidative stress and improving hemoglobin levels, particularly in children with transfusion-dependent thalassemia [182]. These antioxidants can inhibit cation pathways responsible for red blood cell dehydration and reduce phosphatidylserine exposure, potentially improving cell rheology and reducing vascular adhesion [183]. Supplementation with vitamin C and NAC has been found to preserve glutathione homeostasis and reduce hemolysis in stored red blood cells [184]. However, high-dose vitamin C and E supplementation may worsen hemolysis in some cases [185]. Combining antioxidants with iron chelators may provide substantial improvements in the pathophysiology of hemolytic anemias, especially thalassemia [186].
Hemin is a pharmacological agent used in treating acute porphyrias, which are metabolic disorders characterized by heme biosynthesis defects. It functions primarily by inhibiting delta-aminolevulinic acid synthase (ALAS-1), the rate-limiting enzyme in heme production, thereby reducing the accumulation of toxic heme precursors like delta-aminolevulinic acid [187,188]. Hemin administration has been shown to alleviate symptoms during acute attacks by restoring heme levels and downregulating ALAS-1 activity [189,190]. Additionally, heme’s feedback inhibition on ALAS-1 is mediated through various mechanisms, including proteolytic degradation of the enzyme [191]. The clinical efficacy of hemin underscores its role in managing acute porphyrias, as it directly addresses the biochemical disturbances underlying these conditions [192,193].
Hydroxyurea is a promising treatment for sickle cell disease, primarily due to its ability to induce fetal hemoglobin (HbF) production [194,195,196]. HbF inhibits the polymerization of sickle hemoglobin (HbS), reducing disease severity [195]. Hydroxyurea works by perturbing erythroid precursor maturation and potentially through nitric oxide-dependent mechanisms [197]. Clinical studies have shown that hydroxyurea treatment increases HbF levels, reduces hemolysis, and decreases the frequency of vaso-occlusive crises and hospitalizations [198,199]. The drug also improves red cell survival and reduces sickling at partial oxygen saturation [200]. While hydroxyurea’s effects can be inconsistent, it has been shown to reduce morbidity and mortality in adults with sickle cell anemia [194]. Ongoing research is exploring its use in children and in combination with other HbF-inducing agents [195,201].
Mitapivat, an oral pyruvate kinase activator, has shown promising results in treating various hemoglobinopathies. In pyruvate kinase deficiency, it increased hemoglobin levels and reduced transfusion burden, with sustained long-term effects [202,203]. For non-transfusion-dependent α- and β-thalassemia, mitapivat improved hemoglobin levels, markers of erythropoiesis, and hemolysis [204,205,206]. In sickle cell disease, it demonstrated improvements in anemia, hemolysis, and hemoglobin S polymerization kinetics [207]. Mitapivat’s mechanism of action involves increasing ATP production in red blood cells, which may help stabilize hemoglobin and improve cell integrity [208]. The drug has shown a favorable safety profile across studies, with common adverse events being generally mild and transient [203,204]. These findings suggest that mitapivat may represent a novel therapeutic approach for various hereditary anemias [209].

2.4. Small-Molecule Binders to Hemoglobin

Hemoglobin is also a direct drug target and, hence, several classes of molecules were found to bind and stabilize hemoglobin, which should improve oxygen delivery in patients with hemoglobinopathies such as sickle cell disease and β-thalassemia. Pharmacological chaperones are drugs that stabilize hemoglobin by promoting proper folding and preventing denaturation. These agents help maintain hemoglobin’s functional state under stress conditions, reducing the likelihood of hemolysis (see above). Most of the development of drugs targeting hemoglobin focus on sickle cell diseases, which is caused by a single mutation that results in the substitution of valine for glutamic acid in the sixth position of the β-globin chain of hemoglobin, leading to the polymerization of hemoglobin S upon deoxygenation. Peptides and amino acids stabilizing hemoglobin have been summarized in detail in a recent review [210]. Recent advances in SCD treatment have expanded beyond the traditional therapies of hydroxyurea and L-glutamine. New approaches target various aspects of SCD pathophysiology, including HbS polymerization, cellular adhesion, inflammation, and oxidative stress [211]. The FDA has approved three new drugs: crizanlizumab, voxelotor, and L-glutamine, which have shown efficacy in reducing pain crises and improving hemoglobin levels [212]. Emerging therapies include peptide-based inhibitors of HbS polymerization, which represent a novel approach to SCD treatment [210]. Additionally, researchers are exploring combination therapies to address the complex nature of SCD manifestations [213,214]. While gene therapy holds promise for a cure, its widespread application remains limited due to technical and cost challenges [215]. These developments offer hope for improved quality of life and survival for SCD patients.
Hemoglobin exhibits remarkable versatility in binding a wide variety of small molecules, reflecting its functional and structural adaptability (Table 2, Figure 3). Figure 3 showcases examples of these interactions, supported by three-dimensional structures available in the Protein Data Bank (PDB). Representative hemoglobin-small molecule complexes include toluene and 2,3-dihydro-1,4-benzodioxin-2-yl)-4H-1,2,4-triazol-3-yl)disulfide (TD1, PDB: 4NI0), 1H-1,2,3-triazole-5-thiol (TD3, PDB: 6BWU), 2-[(2-methoxy-5-methylphenoxy)methyl]pyridine (INN-298, PDB: 3IC0), 5-hydroxymethyl-furfural (5HMF, PDB: 1QXE), inositol hexakisphosphate (IHP, PDB: 3HXN), 1,4,7,10,13,16-hexaoxacyclooctadecane (18-crown-6, PDB: 3WHM), 2-[(4-methoxy-2-methylphenoxy)methyl]pyridine (vanillin derivative screening, INN-310, PDB: 6BNR), 2-hydroxy-6-((6-(hydroxymethyl)pyridin-2-yl)methoxy)benzaldehyde (VZHE-039, PDB: 6XD6), 2-[4-({[(3,5-dichlorophenyl)amino]carbonyl}amino)phenoxy]-2-methylpropanoic acid (L35, PDB: 2D5Z), (2R,4S)-2-hydroxy-1-[2-hydroxy-3-(trifluoromethyl)phenyl]-4-[4-(trifluoromethyl)phenyl]pyrrolidine-3,5-dione (GBT440, Voxeltor, PDB: 5E83) and 6-{(1S)-1-[(2-amino-6-fluoroquinolin-3-yl)oxy]ethyl}-5-(1H-pyrazol-1-yl)pyridin-2(1H)-one (compound 23, PF-07059013, PDB: 7JY3). While these examples illustrate the chemical diversity of molecules that bind to hemoglobin, they represent only a subset of the available structural data. Not all hemoglobin–ligand structures have been published, and the presented examples are chosen to demonstrate the range of chemical scaffolds and binding interactions observed. Importantly, these small molecules interact with specific residues in hemoglobin chains, underscoring the molecular specificity of binding. Some molecules, such as TD1 (PDB: 4NI0), interact with residues in both α and β chains of hemoglobin, while others bind selectively to residues within a single chain. This variability highlights the dynamic nature of hemoglobin’s interaction with ligands, as well as its potential for modulating functional states through small-molecule binding. The provided examples demonstrate hemoglobin’s capacity to bind chemically distinct compounds, offering valuable insights for designing therapeutics targeting hemoglobin-related disorders or for understanding its physiological roles in molecular detail.
Compounds like TD7, VZHE039, and their nitric oxide-releasing derivatives have shown promising antisickling properties by increasing hemoglobin’s oxygen affinity and disrupting HbS polymerization [216]. TD-1, a small molecule identified through screening, demonstrated dual allosteric and antioxidant effects by binding to βCys93 and βCys112, stabilizing hemoglobin’s relaxed state and reducing oxidative changes [217,218]. Other compounds like TD-3 and hydroxyurea also target βCys93, providing both antioxidant and antisickling benefits [219]. These studies have elucidated the mechanisms of action for various antisickling agents, including their effects on oxygen affinity, HbS polymerization, and oxidative stress reduction. Further research on antisickling agents targeting hemoglobin’s β-Cys93 has shown promising results for treating sickle cell disease. Other thiol reagents reacting with β-Cys93 have shown varying antisickling effects [220,221]. Overall, increasing hemoglobin’s oxygen affinity represents a promising therapeutic approach for sickle cell disease.
5-hydroxymethyl-2-furfural (5HMF) is a promising compound for treating sickle cell disease. It forms a Schiff-base adduct with hemoglobin, increasing oxygen affinity and inhibiting red blood cell sickling [222,223]. 5HMF demonstrates dose-dependent effects on hemoglobin oxygen affinity in phase 1 clinical trials [224]. It improves cardiac function during severe hypoxia [225] and reduces deoxygenation-induced dehydration of red blood cells by inhibiting cation conductance and Ca2+-activated K+ channels [226]. Researchers have developed ester and ether derivatives of 5HMF with enhanced antisickling properties [223] and explored nitric oxide-releasing prodrugs [227]. The mechanism of action involves stabilizing the R state of hemoglobin, which has higher oxygen affinity and slower polymerization kinetics [228,229]. These findings support 5HMF as a potential therapeutic agent for sickle cell disease.
Hb plays a crucial role in nitric oxide metabolism and oxygen transport. Hb can form various NO-related compounds, including nitrosylated Hb (HbNO) and S-nitrosohemoglobin (SNO-Hb) [230,231]. These compounds are involved in regulating vascular function and oxygen delivery [15,232]. The interaction between Hb and NO is complex, involving redox reactions and allosteric effects [230,233]. SNO-Hb has been proposed as a key mediator of NO transport and bioactivity, with its formation and release coupled with oxygen saturation [15,234]. However, the precise mechanisms of NO activity preservation and regulation by Hb remain debated [235,236]. Understanding these interactions is crucial for developing transfusion therapeutics and treating various cardiovascular diseases [232].
Recent research on SCD has also focused on developing aromatic aldehydes as potential treatments by increasing hemoglobin oxygen affinity and inhibiting red blood cell sickling. Novel compounds, including pyridyl derivatives of vanillin and azolylacryloyl derivatives, have shown improved potency, metabolic stability, and sustained antisickling effects in vitro [237,238]. Some compounds exhibit both oxygen-dependent and oxygen-independent antisickling mechanisms, potentially disrupting key intermolecular contacts necessary for hemoglobin S polymer formation. Crystal structure studies have provided insights into the binding modes and mechanisms of action of these compounds [238]. The development of hemoglobin modulators represents a promising approach to treating SCD, with ongoing research and growing intellectual property in this field [239]. These advancements complement other therapeutic strategies, including targeting cellular adhesion, inflammation, and oxidant injury.
PF-07059013 (compound 23) is a noncovalent hemoglobin modulator that has advanced to phase 1 clinical trials for sickle cell disease (SCD) treatment. It demonstrated a 37.8% reduction in red blood cell sickling in a mouse model [240]. This compound represents a new approach to SCD treatment, targeting the root cause of the disease: hemoglobin S polymerization [241]. Unlike earlier covalent modifiers with potential safety concerns, PF-07059013 binds noncovalently to hemoglobin [242]. Similar approaches, such as GBT440, have shown promise in preclinical testing [243,244]. These developments are part of a broader effort to create new SCD therapies, including fetal hemoglobin inducers and agents targeting cellular adhesion, inflammation, and vascular tone [245]. Such advancements offer hope for improved SCD management beyond current treatments like hydroxyurea and blood transfusions [246].
Vanillin derivatives have shown promise as potential treatments for sickle cell disease and other conditions. Novel compounds like SAJ-009, SAJ-310, and SAJ-270 demonstrated improved binding and pharmacokinetic properties compared to vanillin, with enhanced allosteric and antisickling effects [238]. SAJ-310 exhibited sustained antisickling activity and potential metabolic stability [247]. Vanillin derivatives also showed antioxidant properties and inhibitory activity against acetylcholinesterase and β-amyloid aggregation, suggesting potential applications in Alzheimer’s disease treatment [248]. Additionally, vanillin demonstrated analgesic effects on mechanical allodynia in a neuropathic pain model [249] and inhibitory activity against mushroom tyrosinase [250]. The compound’s ability to react covalently with sickle hemoglobin and inhibit cell sickling supports its potential as a treatment for sickle cell anemia [251]. Vanillin’s diverse pharmacological activities and safety profile make it a promising candidate for further research and development.
VZHE-039 is a novel antisickling agent that exhibits both oxygen-dependent and oxygen-independent mechanisms of action [252]. In this study, VZHE-039 demonstrates enhanced pharmacologic activities and metabolic stability. Its oxygen-independent activity is attributed to interactions with the αF-helix of hemoglobin, potentially destabilizing the sickle hemoglobin polymer [253]. Other antisickling agents, such as voxelotor and FT-4202, also show promise in treating sickle cell disease by increasing hemoglobin oxygen affinity.
Crown ethers, cyclic polyethers with oxygen atoms, have shown potential in modulating protein surface properties and interactions [254]. They can bind to metal ions and positively charged amino acid side chains, affecting protein behavior such as oligomerization and crystallization. In hemoglobin research, crown ethers and related compounds have been explored for their ability to modify hemoglobin’s oxygen-binding properties. Acylation of hemoglobin with polyoxyethylene derivatives has been shown to alter its oxygen affinity [255]. Other synthetic allosteric effectors, such as RSR-4 and RSR-13, have demonstrated the ability to shift hemoglobin’s allosteric equilibrium towards the low-affinity T-state [256]. These studies highlight the potential of crown ethers and related compounds in modulating hemoglobin function and developing novel therapeutic approaches.
Complex interactions between hemoglobin (Hb) can occur also with natural molecules such as sphingosine 1-phosphate (S1P) and lipopolysaccharide (LPS). S1P is a bioactive lipid involved in immune responses, cell growth, and survival. It has been found to bind specifically to deoxygenated sickle hemoglobin (HbS), the form of hemoglobin that promotes the sickling of red blood cells in sickle cell disease. When S1P binds to HbS, it alters glucose metabolism and increases oxidative stress, both of which are significant contributors to the pathology of sickle cell disease. Altered glucose metabolism disrupts energy production in red blood cells, while increased oxidative stress leads to cellular damage and inflammation, further exacerbating the disease’s symptoms [257]. Hb also interacts with lipopolysaccharide (LPS), a component of the outer membrane of Gram-negative bacteria, enhancing its biological activity and potentially contributing to the immune defense against bacterial infections. This interaction suggests that hemoglobin may play a role beyond oxygen transport, acting as a modulator of the immune response [258,259]. Multiple binding sites for LPS have been identified on both the α and β-subunits of hemoglobin, highlighting the molecule’s multifunctional nature and its potential importance in pathogen defense [259,260]. Hemoglobin also binds to erythrocyte membranes, an interaction crucial for the stability and integrity of red blood cells. This binding is driven by electrostatic forces influenced by factors such as pH and ionic strength, which may vary under different physiological conditions, such as during oxygenation/deoxygenation cycles [261,262]. A critical protein involved in this interaction is band 3 protein, a major structural component of the erythrocyte membrane. The acidic N-terminal segment of the cytoplasmic domain of band 3 serves as a primary binding site for hemoglobin [167]. This interaction is essential for maintaining the structural integrity of red blood cells and regulating their lifespan. Disruptions in this binding can lead to membrane instability and hemolysis, a hallmark of hemolytic anemias such as sickle cell disease. In pathological conditions like sickle cell disease, these hemoglobin interactions are further altered, contributing to disease progression. The S1P-Hb interaction exacerbates oxidative stress, worsening the symptoms of sickle cell disease. The LPS-Hb interaction could play a dual role in immune defense and the potential triggering of exaggerated immune responses.
Finally, hemoglobin’s binding to the erythrocyte membrane via band 3 protein is critical for cell stability, and disruptions in this interaction could contribute to the shortened lifespan of red blood cells seen in sickle cell disease and other hemoglobinopathies. In summary, these complex interactions between hemoglobin, S1P, LPS, and the erythrocyte membrane are critical for both normal physiological processes and pathological conditions, especially sickle cell disease.
Voxelotor is a novel oral treatment for SCD that binds to hemoglobin, increasing its oxygen affinity and preventing the polymerization of sickle hemoglobin (HbS) [263,264]. By forming a reversible covalent bond with hemoglobin, voxelotor reduces the sickling of red blood cells and interrupts the molecular pathogenesis of SCD. Clinical trials have demonstrated that voxelotor increases hemoglobin concentration, reduces hemolysis, and improves hematological parameters in SCD patients [265,266]. The drug has shown a dose-dependent pharmacokinetic and pharmacodynamic response and is well-tolerated [267]. Voxelotor represents a new disease-modifying approach to SCD treatment, targeting the root cause of the disease [268].

3. Search for Potentially New Hb Stabilisers: Hemoglobin Binding Proteins

Innovative approaches to stabilizing hemoglobin leverage specialized molecular chaperones, engineered enzymes, and highly specific binding agents to counter the structural vulnerabilities of various hemoglobin variants. These methods go beyond traditional therapies by focusing directly on the molecular factors that lead to hemoglobin instability, aggregation, or ineffective oxygen transport. For example, alpha-hemoglobin-stabilizing protein (AHSP) is an erythroid-specific chaperone that binds free α-globin chains and prevents their precipitation [168,170,269,270,271,272]. In conditions where the balance of α- and β-globin is disrupted (e.g., α-thalassemia or other unstable hemoglobin variants), AHSP can mitigate the harmful accumulation of unpaired α-globin, which would otherwise form inclusion bodies, generate oxidative stress, and shorten red blood cell lifespan. By enhancing AHSP expression or designing small molecules that mimic its stabilizing interface, one could restore proper globin chain stoichiometry and improve red blood cell survival. For instance, future therapies might involve drugs that upregulate AHSP in erythroid progenitors or engineered AHSP analogs that bind mutated α-globins more efficiently, thus preventing hemoglobin instability at its source.
Enzymes like falcilysin, a metalloprotease from the malaria parasite Plasmodium falciparum, also inspire innovative strategies. Although falcilysin’s natural function is to degrade hemoglobin within the parasite’s food vacuole, its specificity and mechanism highlight how selective proteolysis could be harnessed or inhibited to influence hemoglobin stability [273,274,275,276,277]. Inhibiting falcilysin is already a potential antimalarial strategy, but repurposing or engineering similar proteolytic activities could help clear unstable hemoglobin fragments in other contexts. For example, engineered human proteases or inhibitors could be designed to selectively remove or neutralize harmful, partially denatured hemoglobin species in patients with unstable variants. This would prevent the formation of damaging aggregates, such as Heinz bodies, and reduce the oxidative burden on red blood cells.
Nanobodies, small antibody-derived fragments with high specificity and affinity, represent another frontier. By targeting key regulators of hemoglobin gene expression, such as BCL11A, nanobodies can potentially shift the balance toward fetal hemoglobin (HbF) production, alleviating symptoms in disorders like sickle cell disease and β-thalassemia [278,279]. Beyond gene regulation, nanobodies like NbE11 that precisely bind human hemoglobin have immediate applications in diagnostics. Their exceptional specificity could be adapted for therapeutic delivery systems that “escort” stabilizing compounds directly to unstable hemoglobin variants within red blood cells, thus minimizing off-target effects. For instance, a nanobody could be fused to a small molecule or peptide stabilizer, ensuring the agent reaches and binds specifically to the defective hemoglobin subunits, improving stability and function without altering healthy hemoglobin.
In addition, molecularly imprinted polymeric nanoparticles provide a synthetic approach to selectively bind and purify human hemoglobin [280]. These “plastic antibodies” could be tailored to recognize and capture unstable hemoglobin variants from patient samples, aiding in precise diagnostics and screening assays. By modifying the imprinting process, one might even create nanoparticles that deliver stabilizing agents to red blood cells or sequester oxidized hemoglobin species, improving red cell viability.
Collectively, these strategies—enhancing AHSP function, leveraging proteases like falcilysin for targeted intervention, employing nanobody-based precision tools, and using molecular imprinting technologies—offer a sophisticated toolkit for addressing hemoglobin instability at a molecular level. By focusing on the structural and mechanistic aspects of hemoglobin, these innovative approaches have the potential to yield more effective, variant-specific treatments and diagnostic platforms, ultimately improving patient outcomes in hemoglobinopathies.
Innovative approaches targeting hemoglobin stability and function may enhance protein platforms for protein engineering by the use of protein chaperones, enzymes like falcilysin, and nanobodies.

4. Conclusions

Hemoglobinopathies such as sickle cell disease and thalassemias are driven by diverse hemoglobin variants, posing challenges for universal therapies. This review highlights promising approaches like small-molecule binders (e.g., voxelotor) and peptides that enhance hemoglobin stability, increase oxygen affinity, and reduce oxidative stress [210,219,263,264,265]. Combining these therapies with antioxidants offers additional potential [181,182,183,184]. Future research should focus on discovering new hemoglobin-targeting molecules, utilizing advanced techniques like time-resolved X-ray diffraction [281] and in situ crystallization [282], and refining gene-editing tools such as CRISPR/Cas9 for curative treatments [118,119]. By integrating these strategies, tailored and effective therapies can significantly improve outcomes for hemoglobinopathy patients.

Author Contributions

Conceptualization, M.Ž., G.Ž. and K.P.K.; methodology, M.Ž.; software, G.Ž.; validation, M.N., K.P.K. and M.Ž.; formal analysis, G.Ž.; investigation, K.P.K. and G.Ž.; resources, G.Ž. and M.N.; data curation, G.Ž.; writing—original draft preparation, G.Ž.; writing—review and editing, G.Ž. and M.Ž.; visualization, M.Ž. and G.Ž. All authors have read and agreed to the published version of the manuscript.

Funding

The research was supported by the Slovak Research and Development Agency under the Contract no. APVV-23-0212 and Visegrad Grants from the International Visegrad Fund and Fellowship (No.#62410140). G.Ž. is partly funded by the European Union’s Next Generation EU through the Recovery and Resilience Plan for Slovakia under the project No. 09I03-03-V03-00008. The Extreme Light Infrastructure ERIC funded part of this research.

Acknowledgments

Róbert Onufrák from SOLUTION s.r.o., Werferova 1, 040 11 Košice, Slovakia.

Conflicts of Interest

Author Michal Novotný was employed by the company AURORA R&D s.r.o. and Krishna P. Khakurel was employed by the company Extreme Light Infrastructure ERIC. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
HbHemoglobin
SNOS-nitrosohemoglobin
BPGBisphosphoglycerate
HbFFetal Hemoglobin
HbAAdult Hemoglobin
HbCHemoglobin C
HbDHemoglobin D
HbEHemoglobin E
Hb CSHb Constant Spring
ACMGAmerican College of Medical Genetics and Genomics
SCDSickle Cell Disease
RBCRed Blood Cell
ROSReactive Oxygen Species
AHSPα-Hemoglobin Stabilizing Protein
G6PDGlucose-6-phosphate Dehydrogenase
NACN-acetylcysteine
ALAS-1Aminolevulinic Acid Synthase
S1PSphingosine 1-phosphate
LPSLipopolysaccharide

References

  1. Perutz, M.F. Stereochemical mechanism of oxygen transport by haemoglobin. Proc. R. Soc. Lond. Ser. B Biol. Sci. 1980, 208, 135–162. [Google Scholar]
  2. Bellelli, A. Hemoglobin and Cooperativity: Experiments and Theories. Curr. Protein Pept. Sci. 2010, 11, 2–36. [Google Scholar] [CrossRef] [PubMed]
  3. Riggs, A.F. The Bohr Effect. Annu. Rev. Physiol. 1988, 50, 181–204. [Google Scholar] [CrossRef] [PubMed]
  4. Tyuma, I.; Shimizu, K. Different response to organic phosphates of human fetal and adult hemoglobins. Arch. Biochem. Biophys. 1969, 129, 404–405. [Google Scholar] [CrossRef]
  5. Safo, M.K.; Ahmed, M.H.; Ghatge, M.S.; Boyiri, T. Hemoglobin–ligand binding: Understanding Hb function and allostery on atomic level. Biochim. Biophys. Acta (BBA) Proteins Proteom. 2011, 1814, 797–809. [Google Scholar] [CrossRef]
  6. Ahmed, M.H.; Ghatge, M.S.; Safo, M.K. Hemoglobin: Structure, Function and Allostery. Subcell. Biochem. 2020, 94, 345–382. [Google Scholar]
  7. Randad, R.S.; Mahran, M.A.; Mehanna, A.S.; Abraham, D.J. Allosteric modifiers of hemoglobin. 1. Design, synthesis, testing, and structure-allosteric activity relationship of novel hemoglobin oxygen affinity decreasing agents. J. Med. Chem. 1991, 34, 752–757. [Google Scholar] [CrossRef]
  8. Ronda, L.; Bruno, S.; Abbruzzetti, S.; Viappiani, C.; Bettati, S. Ligand reactivity and allosteric regulation of hemoglobin-based oxygen carriers. Biochim. Biophys. Acta (BBA) Proteins Proteom. 2008, 1784, 1365–1377. [Google Scholar] [CrossRef]
  9. Ciaccio, C.; Coletta, A.; De Sanctis, G.; Marini, S.; Coletta, M. Cooperativity and allostery in haemoglobin function. IUBMB Life 2008, 60, 112–123. [Google Scholar] [CrossRef]
  10. Maillett, D.H.; Simplaceanu, V.; Shen, T.-J.; Ho, N.T.; Olson, J.S.; Ho, C. Interfacial and Distal-Heme Pocket Mutations Exhibit Additive Effects on the Structure and Function of Hemoglobin. Biochemistry 2008, 47, 10551–10563. [Google Scholar] [CrossRef]
  11. Brunori, M.; Miele, A.E. Modulation of Allosteric Control and Evolution of Hemoglobin. Biomolecules 2023, 13, 572. [Google Scholar] [CrossRef] [PubMed]
  12. Storz, J.F. Hemoglobin structure and allosteric mechanism. In Hemoglobin; Oxford University Press: Oxford, UK, 2018; pp. 58–93. [Google Scholar]
  13. Ferguson, J.K.W.; Roughton, F.J.W. The chemical relationships and physiological importance of carbamino compounds of CO2 with hæmoglobin. J. Physiol. 1934, 83, 87–102. [Google Scholar] [CrossRef] [PubMed]
  14. Chan, N.-L.; Kavanaugh, J.S.; Rogers, P.H.; Arnone, A. Crystallographic Analysis of the Interaction of Nitric Oxide with Quaternary-T Human Hemoglobin. Biochemistry 2003, 43, 118–132. [Google Scholar] [CrossRef] [PubMed]
  15. Allen, B.W.; Stamler, J.S.; Piantadosi, C.A. Hemoglobin, nitric oxide and molecular mechanisms of hypoxic vasodilation. Trends Mol. Med. 2009, 15, 452–460. [Google Scholar] [CrossRef] [PubMed]
  16. Hamasaki, N.; Rose, Z.B. The Binding of Phosphorylated Red Cell Metabolites to Human Hemoglobin A. J. Biol. Chem. 1974, 249, 7896–7901. [Google Scholar] [CrossRef]
  17. Garby, L.; De Verdier, C.-H. Affinity of Human Hemoglobin a to 2,3—Diphosphoglycerate. Effect of Hemoglobin Concentration and of pH. Scand. J. Clin. Lab. Investig. 1971, 27, 345–350. [Google Scholar] [CrossRef]
  18. Beek, G.G.M.; Bruin, S.H. The pH Dependence of the Binding of d-Glycerate 2,3-Bisphosphate to Deoxyhemoglobin and Oxyhemoglobin. Determination of the Number of Binding Sites in Oxyhemoglobin. Eur. J. Biochem. 1979, 100, 497–502. [Google Scholar] [CrossRef]
  19. Isaacks, R.E. Can Metabolites Contribute in Regulating Blood Oxygen Affinity? In Advances in Experimental Medicine and Biology; Springer: New York, NY, USA, 1988; pp. 137–143. [Google Scholar]
  20. Mulquiney, P.J.; Kuchel, P.W. Model of 2,3-bisphosphoglycerate metabolism in the human erythrocyte based on detailed enzyme kinetic equations: Equations and parameter refinement. Biochem. J. 1999, 342, 581–596. [Google Scholar] [CrossRef]
  21. Kayikci, M.; Venkatakrishnan, A.J.; Scott-Brown, J.; Ravarani, C.N.J.; Flock, T.; Babu, M.M. Visualization and analysis of non-covalent contacts using the Protein Contacts Atlas. Nat. Struct. Mol. Biol. 2018, 25, 185–194. [Google Scholar] [CrossRef]
  22. López, F.J.B.; Centurion, I.F.; Díaz, F.J.P.; Pacheco, C.M.; Cortés, B.I.; Mendoza, B.M.T.; Flores-Jimenez, J.A.; Torre, L.D.C.R.-D.L. Structural Effect of Gγ and Aγ Globin Chains in Fetal Hemoglobin Tetramer. Blood 2023, 142 (Suppl. S1), 2291. [Google Scholar] [CrossRef]
  23. Sharma, S.K.; Lechner, R.B. Hematologic and coagulation disorders. In Obstetric Anesthesia: Principles and Practice, 2nd ed.; Chestnut, D.H., Ed.; Mosby Inc.: St. Louis, MO, USA, 1999; ISBN 0-3230-0383-4. [Google Scholar]
  24. Manca, L.; Masala, B. Disorders of the synthesis of human fetal hemoglobin. IUBMB Life 2008, 60, 94–111. [Google Scholar] [CrossRef] [PubMed]
  25. He, L.; Rockwood, A.L.; Agarwal, A.M.; Anderson, L.C.; Weisbrod, C.R.; Hendrickson, C.L.; Marshall, A.G. Diagnosis of Hemoglobinopathy and β-Thalassemia by 21 Tesla Fourier Transform Ion Cyclotron Resonance Mass Spectrometry and Tandem Mass Spectrometry of Hemoglobin from Blood. Clin. Chem. 2019, 65, 986–994. [Google Scholar] [CrossRef] [PubMed]
  26. Kim, J.-S.; Kim, H.S. Diagnosis of hemoglobinopathy and β-thalassemia by 21-Tesla Fourier transform ion cyclotron resonance mass spectrometry. Ann. Transl. Med. 2019, 7 (Suppl. S6), S239. [Google Scholar] [CrossRef] [PubMed]
  27. Carrocini, G.C.d.S.; Zamaro, P.J.A.; Bonini-Domingos, C.R. What influences Hb fetal production in adulthood? Rev. Bras. Hematol. Hemoter. 2011, 33, 231–236. [Google Scholar] [CrossRef]
  28. Hempe, J.M.; Craver, R.D. Laboratory Diagnosis of Structural Hemoglobinopathies and Thalassemias by Capillary Isoelectric Focusing. In Clinical Applications of Capillary Electrophoresis; Humana Press: Totowa, NJ, USA, 1999; pp. 81–98. [Google Scholar]
  29. Little, R.R.; Roberts, W.L. A Review of Variant Hemoglobins Interfering with Hemoglobin A1c Measurement. J. Diabetes Sci. Technol. 2009, 3, 446–451. [Google Scholar] [CrossRef]
  30. Fontana, L.; Alahouzou, Z.; Miccio, A.; Antoniou, P. Epigenetic Regulation of β-Globin Genes and the Potential to Treat Hemoglobinopathies through Epigenome Editing. Genes 2023, 14, 577. [Google Scholar] [CrossRef]
  31. Modell, B. Global epidemiology of haemoglobin disorders and derived service indicators. Bull. World Health Organ. 2008, 86, 480–487. [Google Scholar] [CrossRef]
  32. Turbpaiboon, C.; Wilairat, P. Alpha-hemoglobin stabilizing protein: Molecular function and clinical correlation. Front. Biosci. 2010, 15, 1–11. [Google Scholar] [CrossRef]
  33. Cardiero, G.; Musollino, G.; Prezioso, R.; Lacerra, G. mRNA Analysis of Frameshift Mutations with Stop Codon in the Last Exon: The Case of Hemoglobins Campania [α1 cod95 (−C)] and Sciacca [α1 cod109 (−C)]. Biomedicines 2021, 9, 1390. [Google Scholar] [CrossRef]
  34. Yang, P.; Chou, S.-J.; Li, J.; Hui, W.; Liu, W.; Sun, N.; Zhang, R.Y.; Zhu, Y.; Tsai, M.-L.; Lai, H.I.; et al. Supramolecular nanosubstrate–mediated delivery system enables CRISPR-Cas9 knockin of hemoglobin beta gene for hemoglobinopathies. Sci. Adv. 2020, 6, eabb7107. [Google Scholar] [CrossRef]
  35. Nualkaew, T.; Sii-Felice, K.; Giorgi, M.; McColl, B.; Gouzil, J.; Glaser, A.; Voon, H.P.; Tee, H.Y.; Grigoriadis, G.; Svasti, S.; et al. Coordinated β-globin expression and α2-globin reduction in a multiplex lentiviral gene therapy vector for β-thalassemia. Mol. Ther. 2021, 29, 2841–2853. [Google Scholar] [CrossRef] [PubMed]
  36. Kiem, H.-P.; I Arumugam, P.; Burtner, C.R.; Fox, C.F.; Beard, B.C.; Dexheimer, P.; E Adair, J.; Malik, P. Pigtailed macaques as a model to study long-term safety of lentivirus vector-mediated gene therapy for hemoglobinopathies. Mol. Ther. Methods Clin. Dev. 2014, 1, 14055. [Google Scholar] [CrossRef] [PubMed]
  37. Thom, C.S.; Dickson, C.F.; Gell, D.A.; Weiss, M.J. Hemoglobin Variants: Biochemical Properties and Clinical Correlates. Cold Spring Harb. Perspect. Med. 2013, 3, a011858. [Google Scholar] [CrossRef] [PubMed]
  38. Masson, E.; Zou, W.-B.; Génin, E.; Cooper, D.N.; Le Gac, G.; Fichou, Y.; Pu, N.; Rebours, V.; Férec, C.; Liao, Z.; et al. Expanding ACMG variant classification guidelines into a general framework. Hum. Genom. 2022, 16, 31. [Google Scholar] [CrossRef]
  39. Carss, K.; Goldstein, D.; Aggarwal, V.; Petrovski, S. Variant Interpretation and Genomic Medicine. In Handbook of Statistical Genomics; Balding, D., Moltke, I., Marioni, J., Eds.; John Wiley & Sons Ltd.: Hoboken, NJ, USA, 2019. [Google Scholar] [CrossRef]
  40. Chen, J.-M.; Masson, E.; Zou, W.-B.; Liao, Z.; Génin, E.; Cooper, D.N.; Férec, C. Validation of the ACMG/AMP guidelines-based seven-category variant classification system. medRxiv 2023. [Google Scholar] [CrossRef]
  41. Walsh, N.; Cooper, A.; Dockery, A.; O’Byrne, J.J. Variant reclassification and clinical implications. J. Med. Genet. 2024, 61, 207–211. [Google Scholar] [CrossRef]
  42. Sabath, D.E. Molecular Diagnosis of Thalassemias and Hemoglobinopathies. Am. J. Clin. Pathol. 2017, 148, 6–15. [Google Scholar] [CrossRef]
  43. 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. Adapting the ACMG/AMP variant classification framework: A perspective from the ClinGen Hemoglobinopathy Variant Curation Expert Panel. Hum. Mutat. 2021, 43, 1089–1096. [Google Scholar] [CrossRef]
  44. Turner, S.A.; Rao, S.K.; Morgan, R.H.; Vnencak-Jones, C.L.; Wiesner, G.L. The impact of variant classification on the clinical management of hereditary cancer syndromes. Anesth. Analg. 2019, 21, 426–430. [Google Scholar] [CrossRef]
  45. David, S.; Nora Syahirah, S.; Muhammad Nur Salam Bin, H.; Rajan, R. The Blood Blues: A Review on Methemoglobinemia. J. Pharmacol. Pharmacother. 2018, 9, 5. [Google Scholar]
  46. Boylston, M.; Beer, D. Methemoglobinemia: A case study. Crit. Care Nurse 2002, 22, 50–55. [Google Scholar] [CrossRef] [PubMed]
  47. Ashurst, J.; Wasson, M. Methemoglobinemia: A systematic review of the pathophysiology, detection, and treatment. Delw. Med. J. 2011, 83, 203–208. [Google Scholar] [PubMed]
  48. Nascimento, T.S.D.; Pereira, R.O.L.; de Mello, H.L.D.; Costa, J. Metemoglobinemia: Do diagnóstico ao tratamento. Braz. J. Anesthesiol. 2008, 58, 651–664. [Google Scholar] [CrossRef]
  49. Iolascon, A.; Bianchi, P.; Andolfo, I.; Russo, R.; Barcellini, W.; Fermo, E.; Toldi, G.; Ghirardello, S.; Rees, D.; Van Wijk, R.; et al. Recommendations for diagnosis and treatment of methemoglobinemia. Am. J. Hematol. 2021, 96, 1666–1678. [Google Scholar] [CrossRef] [PubMed]
  50. Ash-Bernal, R.; Wise, R.; Wright, S.M. Acquired Methemoglobinemia: A retrospective series of 138 cases at 2 teaching hospitals. Medicine 2004, 83, 265–273. [Google Scholar] [CrossRef]
  51. Friedman, N.; Scolnik, D.; McMurray, L.; Bryan, J. Acquired methemoglobinemia presenting to the pediatric emergency department: A clinical challenge. CJEM 2020, 22, 673–677. [Google Scholar] [CrossRef] [PubMed]
  52. Karna, B.; Jha, S.K.; Al Zaabi, E. Hemoglobin C Disease. 2023. Available online: https://www.ncbi.nlm.nih.gov/books/NBK559043/ (accessed on 12 August 2024).
  53. Charache, S.; Conley, C.L.; Waugh, D.F.; Ugoretz, R.J.; Spurrell, J.R. Pathogenesis of Hemolytic Anemia in Homozygous Hemoglobin C Disease. J. Clin. Investig. 1967, 46, 1795–1811. [Google Scholar] [CrossRef] [PubMed]
  54. Schmidt, K.L.; Randolph, T.R. Development of a Microscopic Method to Identify Hemoglobin C Conditions for Use in Developing Countries. Am. Soc. Clin. Lab. Sci. 2019, 32, 61–66. [Google Scholar] [CrossRef]
  55. Ohiri, C.D.; Randolph, T.R. Development of an easy, inexpensive, and precise method to identify Hemoglobin C for use in underdeveloped countries. FASEB J. 2016, 30, 61–66. [Google Scholar] [CrossRef]
  56. Verra, F.; Simpore, J.; Warimwe, G.M.; Tetteh, K.K.; Howard, T.; Osier, F.H.A.; Bancone, G.; Avellino, P.; Blot, I.; Fegan, G.; et al. Haemoglobin C and S Role in Acquired Immunity against Plasmodium falciparum Malaria. PLoS ONE 2007, 2, e978. [Google Scholar] [CrossRef]
  57. Ghosh, A.; Basak, J.; Mukhopadhyay, A. Coexistence of rare variant HbD Punjab [α2β2121(Glu→Gln)] and alpha 3.7 kb deletion in a young boy of Hindu family in West Bengal, India. Cell. Mol. Biol. Lett. 2015, 20, 736–742. [Google Scholar] [CrossRef] [PubMed]
  58. Italia, K.; Upadhye, D.; Dabke, P.; Kangane, H.; Colaco, S.; Sawant, P.; Nadkarni, A.; Gorakshakar, A.; Jain, D.; Italia, Y.; et al. Clinical and hematological presentation among Indian patients with common hemoglobin variants. Clin. Chim. Acta 2014, 431, 46–51. [Google Scholar] [CrossRef] [PubMed]
  59. Oberoi, S.M.; Das, R.; Trehan, A.; Ahluwalia, J.; Bansal, D.M.; Malhotra, P.; Marwaha, R.K.M. HbSD-Punjab. J. Pediatr. Hematol. 2014, 36, e140–e144. [Google Scholar] [CrossRef] [PubMed]
  60. Pandey, S.; Mishra, R.M.; Pandey, S.; Shah, V.; Saxena, R. Molecular characterization of hemoglobin D Punjab traits and clinical-hematological profile of the patients. Sao Paulo Med. J. 2012, 130, 248–251. [Google Scholar] [CrossRef] [PubMed]
  61. A Petrenko, A.; Pivnik, A.V.; Kim, P.P.; Demidova, E.Y.; Surin, V.L.; O Abdullaev, A.; Sudarikov, A.B.; A Petrova, N.; A Maryina, S. Coinheritance of HbD-Punjab/β+-thalassemia (IVSI+5 G-C) in patient with Gilbert’s syndrome. Ter. Arkh. 2018, 90, 105–109. [Google Scholar] [CrossRef]
  62. Rieder, R. Globin chain synthesis in HbD (Punjab)-beta-thalassemia. Blood 1976, 47, 113–120. [Google Scholar] [CrossRef]
  63. Ohashi, J.; Naka, I.; Patarapotikul, J.; Hananantachai, H.; Brittenham, G.; Looareesuwan, S.; Clark, A.G.; Tokunaga, K. Extended Linkage Disequilibrium Surrounding the Hemoglobin E Variant Due to Malarial Selection. Am. J. Hum. Genet. 2004, 74, 1198–1208. [Google Scholar] [CrossRef]
  64. Flatz, G.; Sanguansermsri, T.; Sengchanh, S.; Horst, D.; Horst, J. The ‘Hot Spot’ of Hb E [β26(B8)Glu→Lys] in Southeast Asia: β-Globin Anomalies in the Lao Theung Population of Southern Laos. Hemoglobin 2004, 28, 197–204. [Google Scholar] [CrossRef]
  65. Roche, C.J.; Malashkevich, V.; Balazs, T.C.; Dantsker, D.; Chen, Q.; Moreira, J.; Almo, S.C.; Friedman, J.M.; Hirsch, R.E. Structural and Functional Studies Indicating Altered Redox Properties of Hemoglobin E. J. Biol. Chem. 2011, 286, 23452–23466. [Google Scholar] [CrossRef]
  66. Strader, M.B.; Kassa, T.; Meng, F.; Wood, F.B.; Hirsch, R.E.; Friedman, J.M.; Alayash, A.I. Oxidative instability of hemoglobin E (β26 Glu→Lys) is increased in the presence of free α subunits and reversed by α-hemoglobin stabilizing protein (AHSP): Relevance to HbE/β-thalassemia. Redox Biol. 2016, 8, 363–374. [Google Scholar] [CrossRef]
  67. Jamsai, D.; Zaibak, F.; Vadolas, J.; Voullaire, L.; Fowler, K.J.; Gazeas, S.; Peters, H.; Fucharoen, S.; Williamson, R.; Ioannou, P.A. A humanized BAC transgenic/knockout mouse model for HbE/β-thalassemia. Genomics 2006, 88, 309–315. [Google Scholar] [CrossRef] [PubMed]
  68. Naka, I.; Ohashi, J.; Nuchnoi, P.; Hananantachai, H.; Looareesuwan, S.; Tokunaga, K.; Patarapotikul, J. Lack of Association of the HbE Variant with Protection from Cerebral Malaria in Thailand. Biochem. Genet. 2008, 46, 708–711. [Google Scholar] [CrossRef] [PubMed]
  69. Kidd, R.D.; Baker, H.M.; Mathews, A.J.; Brittain, T.; Baker, E.N. Oligomerization and ligand binding in a homotetrameric hemoglobin: Two high-resolution crystal structures of hemoglobin Bart’s (γ4), a marker for α-thalassemia. Protein Sci. 2001, 10, 1739–1749. [Google Scholar] [CrossRef] [PubMed]
  70. Pootrakul, S.; Wasi, P.; Na-Nakorn, S. Haemoglobin Bart’s hydrops foetalis in Thailand. Ann. Hum. Genet. 1967, 30, 293–311. [Google Scholar] [CrossRef] [PubMed]
  71. Pembrey, M.E.; Weatherall, D.J.; Clegg, J.B.; Bunch, C.; Perrine, R.P. Haemoglobin Bart’s in Saudi Arabia. Br. J. Haematol. 1975, 29, 221–234. [Google Scholar] [CrossRef]
  72. Czerwinski, E.; Czerwinski, E.; Risk, M.; Risk, M.; Matustik, M.; Matustik, M. Crystallization and preliminary X-ray diffraction studies of methemoglobin Bart’s. J. Biol. Chem. 1981, 256, 13128–13129. [Google Scholar] [CrossRef]
  73. Papassotiriou, I.; Traeger-Synodinos, J.; Vlachou, C.; Karagiorga, M.; Metaxotou, A.; Kanavakis, E.; Stamoulakatou, A. Rapid and Accurate Quantitation of Hb Bart’s and Hb H Using Weak Cation Exchange High Performance Liquid Chromatography: Correlation with the α-Thalassemia Genotype. Hemoglobin 1999, 23, 203–211. [Google Scholar] [CrossRef]
  74. Rugless, M.J.; Fisher, C.A.; Stephens, A.D.; Amos, R.J.; Mohammed, T.; Old, J.M. Hb Bart’s in Cord Blood: An Accurate Indicator of α-Thalassemia. Hemoglobin 2006, 30, 57–62. [Google Scholar] [CrossRef]
  75. Sasazuki, T.; Isomoto, A.; Nakajima, H. Circular dichroism and absorption spectra of haemoglobin Bart’s. J. Mol. Biol. 1972, 65, 365–369. [Google Scholar] [CrossRef]
  76. Esan, G.J.F. Haemoglobin Bart’s in Newborn Nigerians. Br. J. Haematol. 1972, 22, 73–86. [Google Scholar] [CrossRef]
  77. Gibson, Q.H.; Nagel, R.L. Allosteric Transition and Ligand Binding in Hemoglobin Chesapeake. J. Biol. Chem. 1974, 249, 7255–7259. [Google Scholar] [CrossRef] [PubMed]
  78. Jones, C.M.; Charache, S.; Hathaway, P. The effect of hemoglobin F-Chesapeake (α292 Arg.→Leuγ2) on fetal oxygen affinity and erythropoiesis. Pediatr. Res. 1979, 13, 851–853. [Google Scholar] [CrossRef] [PubMed]
  79. Charache, S. A manifestation of abnormal hemoglobins of man: Altered oxygen affinity. Hemoglobin Chesapeake: From the clinic to the laboratory, and back again. Ann. N. Y. Acad. Sci. 1974, 241, 449–455. [Google Scholar] [CrossRef] [PubMed]
  80. Wajcman, H.; Galactéros, F. Hemoglobins with High Oxygen Affinity Leading to Erythrocytosis. New Variants and New Concepts. Hemoglobin 2005, 29, 91–106. [Google Scholar] [CrossRef] [PubMed]
  81. Harteveld, C.L.; Wijermans, P.W.; Arkesteijn, S.G.; Van Delft, P.; Kerkhoffs, J.-L.; Giordano, P.C. Hb Lepore-Leiden: A New δ/β Rearrangement Associated with a β-Thalassemia Minor Phenotype. Hemoglobin 2008, 32, 446–453. [Google Scholar] [CrossRef]
  82. Pirastru, M.; Manca, L.; Trova, S.; Mereu, P. Biochemical and Molecular Analysis of the Hb Lepore Boston Washington in a Syrian Homozygous Child. BioMed Res. Int. 2017, 2017, 1261972. [Google Scholar] [CrossRef]
  83. Mirabile, E.; Testa, R.; Consalvo, C.; Dickerhoff, R.; Schilirò, G. Association of Hb S/Hb lepore and δβ-thalassemia/Hb lepore in Sicilian patients: Review of the presence of Hb lepore in Sicily. Eur. J. Haematol. 1995, 55, 126–130. [Google Scholar] [CrossRef]
  84. Seward, D.P.; Ware, R.E.; Kinney, T.R. Hemoglobin sickle-lepore: Report of two siblings and review of the literature. Am. J. Hematol. 1993, 44, 192–195. [Google Scholar] [CrossRef]
  85. Efremov, G.D. Hemoglobins Lepore and Anti-Lepore. Hemoglobin 1978, 2, 197–233. [Google Scholar] [CrossRef]
  86. Chaibunruang, A.; Srivorakun, H.; Fucharoen, S.; Fucharoen, G.; Sae-Ung, N.; Sanchaisuriya, K. Interactions of hemoglobin Lepore (deltabeta hybrid hemoglobin) with v arious hemoglobinopathies: A molecular and hematological characteristi cs and differential diagnosis. Blood Cells Mol. Dis. 2010, 44, 140–145. [Google Scholar] [CrossRef]
  87. Jiang, F.; Tang, X.-W.; Li, J.; Zhou, J.-Y.; Zuo, L.-D.; Li, D.-Z. Hb Lepore-Hong Kong: First Report of a Novel δ/β-Globin Gene Fusion in a Chinese Family. Hemoglobin 2021, 45, 220–224. [Google Scholar] [CrossRef] [PubMed]
  88. Hunt, D.M.; Higgs, D.R.; Winichagoon, P.; Clegg, J.B.; Weatherall, D.J. Haemoglobin Constant Spring has an unstable α chain messenger RNA. Br. J. Haematol. 1982, 51, 405–413. [Google Scholar] [CrossRef] [PubMed]
  89. Thonglairoam, V.; Fucharoen, S.; Tanphaichitr, V.S.; Pung-Amritt, P.; Embury, S.H.; Winichagoon, P.; Wasi, P. Hemoglobin constant spring in bangkok: Molecular screening by selective enzymatic amplification of the α2-globin gene. Am. J. Hematol. 1991, 38, 277–280. [Google Scholar] [CrossRef] [PubMed]
  90. Nguyen, V.H.; Sanchaisuriya, K.; Wongprachum, K.; Nguyen, M.D.; Phan, T.T.H.; Vo, V.T.; Sanchaisuriya, P.; Fucharoen, S.; Schelp, F.P. Hemoglobin Constant Spring is markedly high in women of an ethnic minority group in Vietnam: A community-based survey and hematologic features. Blood Cells Mol. Dis. 2014, 52, 161–165. [Google Scholar] [CrossRef]
  91. Jomoui, W.; Fucharoen, G.; Sanchaisuriya, K.; Nguyen, V.H.; Fucharoen, S. Hemoglobin Constant Spring among Southeast Asian Populations: Haplotypic Heterogeneities and Phylogenetic Analysis. PLoS ONE 2015, 10, e0145230. [Google Scholar] [CrossRef]
  92. Charoenkwan, P.; Sirichotiyakul, S.; Chanprapaph, P.; Tongprasert, F.; Taweephol, R.; Sae-Tung, R.; Sanguansermsri, T. Anemia and Hydrops in a Fetus with Homozygous Hemoglobin Constant Spring. J. Pediatr. Hematol. 2006, 28, 827–830. [Google Scholar] [CrossRef]
  93. Kropp, G.; Fucharoen, S.; Embury, S. Selective enzymatic amplification of alpha 2-globin DNA for detection of the hemoglobin Constant Spring mutation. Blood 1989, 73, 1987–1992. [Google Scholar] [CrossRef]
  94. Roberts, W.L. Hemoglobin Constant Spring Can Interfere with Glycated Hemoglobin Measurements by Boronate Affinity Chromatography. Clin. Chem. 2007, 53, 142–143. [Google Scholar] [CrossRef]
  95. Zimmerman, S.A.; O’Branski, E.E.; Rosse, W.F.; Ware, R.E. Hemoglobin S/O(Arab): Thirteen new cases and review of the lit-erature. Am. J. Hematol. 1999, 60, 279–284. [Google Scholar] [CrossRef]
  96. Dror, S. Clinical and hematological features of homozygous hemoglobin O-Arab [beta 121 Glu → Lys]. Pediatr. Blood Cancer 2012, 60, 506–507. [Google Scholar] [CrossRef]
  97. Sangaré, A.; Sanogo, I.; Meité, M.; Ambofo, Y.; Abesopie, V.; Ségbéna, A.; Tolo, A. Hemoglobin O Arab in Ivory Coast and western Africa. Med. Trop. (Mars) 1992, 52, 163–167. [Google Scholar] [PubMed]
  98. El-Hazmi, M.; Lehmann, H. Human Haemoglobins and Haemoglobinopathies in Arabia: Hb O Arab in Saudi Arabia. Acta Haematol. 1980, 63, 268–273. [Google Scholar] [CrossRef] [PubMed]
  99. Milner, P.F.; Miller, C.; Grey, R.; Seakins, M.; DeJong, W.W.; Went, L.N. Hemoglobin O Arab in Four Negro Families and Its Interaction with Hemoglobin S and Hemoglobin C. N. Engl. J. Med. 1970, 283, 1417–1425. [Google Scholar] [CrossRef] [PubMed]
  100. Hafsia, R.; Gouider, E.; Ben Moussa, S.; Ben Salah, N.; Elborji, W.; Hafsia, A. Hemoglobin O Arab: About 20 cases. Tunis. Med. 2007, 85, 637–640. [Google Scholar] [PubMed]
  101. Merritt, D.; Jones, R.T.; Head, C.; Thibodeau, S.N.; Fairbanks, V.F.; Steinberg, M.H.; Coleman, M.B.; Rodgers, G.P. HB Seal Rock [(α2)142 Term→glu, Codon 142 TAA→GAA]: An Extended α Chain Variant Associated with Anemia, Microcytosis, and α-Thalassemia-2 (-3.7 KB). Hemoglobin 1997, 21, 331–344. [Google Scholar] [CrossRef]
  102. Préhu, C.; Moradkhani, K.; Riou, J.; Bahuau, M.; Launay, P.; Martin, N.; Wajcman, H.; Goossens, M.; Galactéros, F. Chronic hemolytic anemia due to novel -globin chain variants: Critical location of the mutation within the gene sequence for a dominant effect. Haematologica 2009, 94, 1624–1625. [Google Scholar] [CrossRef]
  103. Fattori, A.; Kimura, E.; Albuquerque, D.; Oliveira, D.; Costa, F.; Sonati, M. Hb Indianapolis [β112 (G14) Cys→Arg] as the probable cause of moderate hemolytic anemia and renal damage in a Brazilian patient. Am. J. Hematol. 2007, 82, 672–675. [Google Scholar] [CrossRef]
  104. Henderson, S.J.; Timbs, A.T.; McCarthy, J.; Gallienne, A.E.; Proven, M.; Rugless, M.J.; Lopez, H.; Eglinton, J.; Dziedzic, D.; Beardsall, M.; et al. Ten Years of Routineα- andβ-Globin Gene Sequencing in UK Hemoglobinopathy Referrals Reveals 60 Novel Mutations. Hemoglobin 2016, 40, 75–84. [Google Scholar] [CrossRef]
  105. Outeirino, J.; Casey, R.; White, J.; Lehmann, H. Haemoglobin Madrid β115 (G17) Alanine→Proline: An Unstable Variant Associated with Haemolytic Anaemia. Acta Haematol. 1974, 52, 53–60. [Google Scholar] [CrossRef]
  106. Kim, J.Y.; Park, S.S.; Jung, H.L.; Keum, D.H.; Park, H.; Chang, Y.H.; Lee, Y.J.; Cho, H.I. Hb Madrid [β115(G17)Ala→Pro] in a Korean Family with Chronic Hemolytic Anemia. Hemoglobin 2000, 24, 133–138. [Google Scholar] [CrossRef]
  107. Edison, E.S.; Shaji, R.V.; Devi, S.G.; Kumar, S.S.; Srivastava, A.; Chandy, M. Hb Showa-Yakushiji [beta110(G12)Leu-->Pro] in four unrelated patients from west Bengal. Hemoglobin 2005, 29, 19–25. [Google Scholar] [CrossRef] [PubMed]
  108. Villegas, A.; Ropero, P.; Nogales, A.; González, F.A.; Mateo, M.; Mazo, E.; Rodrigo, E.; Arias, M. Hb Santander [β34(B16)Val→Asp (GTC → GAC)]: A New Unstable Variant Found as a De Novo Mutation in a Spanish Patient. Hemoglobin 2003, 27, 31–35. [Google Scholar] [CrossRef] [PubMed]
  109. Nakatsuji, T.; Miwa, S.; Ohba, Y.; Hattori, Y.; Miyaji, T.; Hino, S.; Matsumoto, N. A New Unstable Hemoglobin, Hb Yokohama β31(B13)LEU → Pro, Causing Hemolytic Anemia. Hemoglobin 1981, 5, 667–678. [Google Scholar] [CrossRef] [PubMed]
  110. Huehns, E.R.; Hecht, F.; Yoshida, A.; Stamatoyannopoulos, G.; Hartman, J.; Motulsky, A.G. Hemoglobin-Seattle (α2Aβ276 Glu): An Unstable Hemoglobin Causing Chronic Hemolytic Anemia. Blood 1970, 36, 209–218. [Google Scholar] [CrossRef]
  111. Hoyer, J.D.; Baxter, J.K.; Moran, A.M.; Kubic, K.S.; Ehmann, W.C. Two Unstable β Chain Variants Associated with β-Thalassemia: Hb Miami [β116(G18)His→Pro], and Hb Hershey [β70(E14)Ala→Gly], and a Second Unstable Hb Variant at β70: Hb Abington [β70(E14)Ala→Pro]. Hemoglobin 2005, 29, 241–248. [Google Scholar] [CrossRef] [PubMed]
  112. Li, W. Biophysical Basis of Hb-S Polymerization in Red Blood Cell Sickling. bioRxiv 2019. [Google Scholar] [CrossRef]
  113. Karen, C. Sickle Cell Disease: A Genetic Disorder of Beta-Globin. Thalassemia and Other Hemolytic Anemias; IntechOpen: Rijeka, Croatia, 2018. [Google Scholar]
  114. Ellsworth, P.; Sparkenbaugh, E.M. Targeting the von Willebrand Factor–ADAMTS-13 axis in sickle cell disease. J. Thromb. Haemost. 2023, 21, 2–6. [Google Scholar] [CrossRef]
  115. Nader, E.; Romana, M.; Connes, P. The Red Blood Cell—Inflammation Vicious Circle in Sickle Cell Disease. Front. Immunol. 2020, 11, 454. [Google Scholar] [CrossRef]
  116. Ashley-Koch, A.; Yang, Q.; Olney, R.S. Sickle Hemoglobin (Hb S) Allele and Sickle Cell Disease: A HuGE Review. Am. J. Epidemiol. 2000, 151, 839–845. [Google Scholar] [CrossRef]
  117. Al-Fatlawi, A.C.Y. A Review on Sickle Cells Disease. Sci. J. Med. Res. 2019, 3, 146–148. [Google Scholar] [CrossRef]
  118. Poli, M.C.; Orange, J. CRISPR/Cas9β-globin Gene Targeting in Human Haematopoietic Stem Cells. Nature 2017, 140 (Suppl. S3), S226–S227. [Google Scholar]
  119. Demirci, S.; Gudmundsdottir, B.; Li, Q.; Haro-Mora, J.J.; Nassehi, T.; Drysdale, C.; Yapundich, M.; Gamer, J.; Seifuddin, F.; Tisdale, J.F.; et al. βT87Q-Globin Gene Therapy Reduces Sickle Hemoglobin Production, Allowing for Ex Vivo Anti-sickling Activity in Human Erythroid Cells. Mol. Ther. Methods Clin. Dev. 2020, 17, 912–921. [Google Scholar] [CrossRef] [PubMed]
  120. De Simone, G.; Quattrocchi, A.; Mancini, B.; di Masi, A.; Nervi, C.; Ascenzi, P. Thalassemias: From gene to therapy. Mol. Asp. Med. 2022, 84, 101028. [Google Scholar] [CrossRef]
  121. Angastiniotis, M.; Lobitz, S. Thalassemias: An Overview. Int. J. Neonatal Screen. 2019, 5, 16. [Google Scholar] [CrossRef]
  122. Weatherall, D.J. Fortnightly review: The thalassaemias. BMJ 1997, 314, 1675–1678. [Google Scholar] [CrossRef] [PubMed]
  123. Muncie, H.; James, C. Alpha and beta thalassemia. Am. Fam. Physician 2009, 80, 339–344. [Google Scholar]
  124. Vernimmen, D. Globins, from Genes to Physiology and Diseases. Blood Cells Mol. Dis. 2018, 70, 1. [Google Scholar] [CrossRef]
  125. Aksu, T.; Unal, S. Thalassemia. Trends Pediatr. 2021, 2, 1–7. [Google Scholar] [CrossRef]
  126. Meri, M.A.; Al-Hakeem, A.H.; Al-Abeadi, R.S. Overview on thalassemia: A review article. Med. Sci. J. Adv. Res. 2022, 3, 26–32. [Google Scholar] [CrossRef]
  127. Reiss, G.H.; Ranney, H.M.; Shaklai, N. Association of hemoglobin C with erythrocyte ghosts. J. Clin. Investig. 1982, 70, 946–952. [Google Scholar] [CrossRef]
  128. Singh, N.; Seth, T.; Tyagi, S. Review of Clinical and Hematological Profile of Hemoglobin D Cases in a Single Centre. J. Mar. Med. Soc. 2023, 25 (Suppl. S1), S74–S79. [Google Scholar] [CrossRef]
  129. Mukherjee, S.; Das, M.; Basu, K.; Sengupta, M.; Karmakar, S.; Jha, A.K.; Bandopadhyay, M. HbE Variants: An Experience from Tertiary Care Centre of Eastern India. Ann. Pathol. Lab. Med. 2020, 7, A570–A575. [Google Scholar] [CrossRef]
  130. Winterbourn, C.C. Oxidative denaturation in congenital hemolytic anemias: The unstable hemoglobins. Semin. Hematol. 1990, 27, 41–50. [Google Scholar] [PubMed]
  131. Taketa, F.; Huang, Y.; Libnoch, J.; Dessel, B. Hemoglobin wood β97(FG4) His → Leu: A new high-oxygen-affinity hemoglobin associated with familial erythrocytosis. Biochim. Biophys. Acta (BBA) Protein Struct. 1975, 400, 348–353. [Google Scholar] [CrossRef]
  132. Reed, C.S.; Hampson, R.; Gordon, S.; Jones, R.T.; Novy, M.J.; Brimhall, B.; Edwards, M.J.; Koler, R.D. Erythrocytosis Secondary to Increased Oxygen Affinity of a Mutant Hemoglobin, Hemoglobin Kempsey. Blood 1968, 31, 623–632. [Google Scholar] [CrossRef]
  133. Medri, C.; Méndez, A.; Hammerer-Lercher, A.; Rovó, A.; Angelillo-Scherrer, A. Unstable hemoglobin Montreal II uncovered in an adult with unexplained hemolysis exacerbated by a presumed viral infection: A case report. J. Med. Case Rep. 2022, 16, 145. [Google Scholar] [CrossRef]
  134. Fallon, J.A.; Smith, E.P.; Schoch, N.; Paruk, J.D.; Adams, E.A.; Evers, D.C.; Jodice, P.G.; Perkins, C.; Schulte, S.; Hopkins, W.A. Hematological indices of injury to lightly oiled birds from the Deepwater Horizon oil spill. Environ. Toxicol. Chem. 2018, 37, 451–461. [Google Scholar] [CrossRef]
  135. Harr, K.E.; Cunningham, F.L.; Pritsos, C.A.; Pritsos, K.L.; Muthumalage, T.; Dorr, B.S.; Horak, K.E.; Hanson-Dorr, K.C.; Dean, K.M.; Cacela, D.; et al. Weathered MC252 crude oil-induced anemia and abnormal erythroid morphology in double-crested cormorants (Phalacrocorax auritus) with light microscopic and ultrastructural description of Heinz bodies. Ecotoxicol. Environ. Saf. 2017, 146, 29–39. [Google Scholar] [CrossRef]
  136. Sözen, M.; Karaaslan, C.; Öner, R.; Gümrük, F.; Özdemir, M.; Altay, C.; Gürgey, A.; Öner, C. Severe hemolytic anemia associated with Hb Volga [β27(B9)Ala→Asp]: GCC→GAC at codon 27 in a Turkish family. Am. J. Hematol. 2004, 76, 378–382. [Google Scholar] [CrossRef]
  137. Wali, Y.; Al Zadjali, S.; Elshinawy, M.; Beshlawi, I.; Fawaz, N.; AlKindi, S.; Rawas, A.; Alsinani, S.; Daar, S.; Krishnamoorthy, R. Severity ranking of non-deletional alpha thalassemic alleles: Insights from an Omani family study. Eur. J. Haematol. 2011, 86, 507–511. [Google Scholar] [CrossRef]
  138. Gallagher, P.G. Diagnosis and management of rare congenital nonimmune hemolytic disease. Hematol. Am. Soc. Hematol. Educ. Program 2015, 2015, 392–399. [Google Scholar] [CrossRef] [PubMed]
  139. Fallon, J.A.; Hopkins, W.A.; Fox, L. A practical quantification method for Heinz bodies in birds applicable to rapid response field scenarios. Environ. Toxicol. Chem. 2013, 32, 401–405. [Google Scholar] [CrossRef] [PubMed]
  140. Drouilly, M.; Jourdan, L.; Gérard, D.; Russello, J.; Bobée, V.; Audouy, A.; Phulpin, A.; Perrin, J. Infantile pyknocytosis, a neonatal hemolytic anemia with Heinz bodies: A cohort study. Pediatr. Blood Cancer 2024, 71, e31078. [Google Scholar] [CrossRef]
  141. Furdak, P.; Bartosz, G.; Stefaniuk, I.; Cieniek, B.; Bieszczad-Bedrejczuk, E.; Soszyński, M.; Sadowska-Bartosz, I. Effect of Garlic Extract on the Erythrocyte as a Simple Model Cell. Int. J. Mol. Sci. 2024, 25, 5115. [Google Scholar] [CrossRef] [PubMed]
  142. Salgado, B.; Monteiro, L.; Rocha, N. Allium species poisoning in dogs and cats. J. Venom. Anim. Toxins Incl. Trop. Dis. 2011, 17, 4–11. [Google Scholar] [CrossRef]
  143. Ideguchi, H. Effects of abnormal Hb on red cell membranes. Rinsho Byori 1999, 47, 232–237. [Google Scholar]
  144. Sugawara, Y.; Hayashi, Y.; Shigemasa, Y.; Abe, Y.; Ohgushi, I.; Ueno, E.; Shimamoto, F. Molecular Biosensing Mechanisms in the Spleen for the Removal of Aged and Damaged Red Cells from the Blood Circulation. Sensors 2010, 10, 7099–7121. [Google Scholar] [CrossRef]
  145. Ronquist, G.; Theodorsson, E. Inherited, non-spherocytic haemolysis due to deficiency of glucose-6-phosphate dehydrogenase. Scand. J. Clin. Lab. Investig. 2007, 67, 105–111. [Google Scholar] [CrossRef]
  146. Jacob, H.S.; Winterhalter, K.H. The role of hemoglobin heme loss in Heinz body formation: Studies with a partially heme-deficient hemoglobin and with genetically unstable hemoglobins. J. Clin. Investig. 1970, 49, 2008–2016. [Google Scholar] [CrossRef]
  147. Jacob, H.; Winterhalter, K. Unstable Hemoglobins: The Role of Heme Loss in Heinz Body Formation. Proc. Natl. Acad. Sci. USA 1970, 65, 697–701. [Google Scholar] [CrossRef]
  148. Winterbourn, C.C.; Carrell, R.W. Studies of Hemoglobin Denaturation and Heinz Body Formation in the Unstable Hemoglobins. J. Clin. Investig. 1974, 54, 678–689. [Google Scholar] [CrossRef] [PubMed]
  149. Simmers, R.N.; Mulley, J.C.; Hyland, V.J.; Callen, D.F.; Sutherland, G.R. Mapping the human alpha globin gene complex to 16p13.2—Pter. J. Med. Genet. 1987, 24, 761–766. [Google Scholar] [CrossRef] [PubMed]
  150. Nicholls, R.D.; A Jonasson, J.; O McGee, J.; Patil, S.; Ionasescu, V.V.; Weatherall, D.J.; Higgs, D.R. High resolution gene mapping of the human alpha globin locus. J. Med. Genet. 1987, 24, 39–46. [Google Scholar] [CrossRef] [PubMed]
  151. Daniels, R.J.; Peden, J.F.; Lloyd, C.; Horsley, S.W.; Clark, K.; Tufarelli, C.; Kearney, L.; Buckle, V.J.; Doggett, N.A.; Flint, J.; et al. Sequence, structure and pathology of the fully annotated terminal 2 Mb of the short arm of human chromosome 16. Hum. Mol. Genet. 2001, 10, 339–352. [Google Scholar] [CrossRef]
  152. Higgs, D.R.; Wainscoat, J.S.; Flint, J.; Hill, A.V.; Thein, S.L.; Nicholls, R.D.; Teal, H.; Ayyub, H.; E Peto, T.; Falusi, A.G. Analysis of the human alpha-globin gene cluster reveals a highly informative genetic locus. Proc. Natl. Acad. Sci. USA 1986, 83, 5165–5169. [Google Scholar] [CrossRef]
  153. Coelho, A.; Picanço, I.; Seuanes, F.; Seixas, M.T.; Faustino, P. Novel large deletions in the human α-globin gene cluster: Clarifying the HS-40 long-range regulatory role in the native chromosome environment. Blood Cells Mol. Dis. 2010, 45, 147–153. [Google Scholar] [CrossRef]
  154. Higgs, D.R.; Hill, A.V.S.; Nicholls, R.; Goodbourn, S.E.Y.; Ayyub, H.; Teal, H.; Clegg, J.B.; Weatherall, D.J. Molecular Rearrangements of the Human α-Globin Gene Cluster. Ann. N. Y. Acad. Sci. 1985, 445, 45–56. [Google Scholar] [CrossRef]
  155. Hatton, C.S.; Wilkie, A.O.; Drysdale, H.C.; Wood, W.G.; Vickers, M.A.; Sharpe, J.; Ayyub, H.; Pretorius, I.M.; Buckle, V.J.; Higgs, D.R. Alpha-thalassemia caused by a large (62 kb) deletion upstream of the human alpha globin gene cluster. Blood 1990, 76, 221–227. [Google Scholar] [CrossRef]
  156. Steinberg, M.H. A New Trans-Acting Modulator of Fetal Hemoglobin? Acta Haematol. 2018, 140, 112–113. [Google Scholar] [CrossRef]
  157. Cao, A.; Galanello, R. Beta-thalassemia. Anesth. Analg. 2010, 12, 61–76. [Google Scholar] [CrossRef]
  158. Moleirinho, A.; Seixas, S.; Lopes, A.M.; Bento, C.; Prata, M.J.; Amorim, A. Evolutionary Constraints in the β-Globin Cluster: The Signature of Purifying Selection at the δ-Globin (HBD) Locus and Its Role in Developmental Gene Regulation. Genome Biol. Evol. 2013, 5, 559–571. [Google Scholar] [CrossRef] [PubMed]
  159. Lin, C.; Draper, P.; De Braekeleer, M. High-resolution chromosomal localization of the β-gene of the human β-globin gene complex by in situ hybridization. Cytogenet. Genome Res. 1985, 39, 269–274. [Google Scholar] [CrossRef] [PubMed]
  160. Gusella, J.; Varsanyi-Breiner, A.; Kao, F.T.; Jones, C.; Puck, T.T.; Keys, C.; Orkin, S.; Housman, D. Precise localization of human beta-globin gene complex on chromosome 11. Proc. Natl. Acad. Sci. USA 1979, 76, 5239–5242. [Google Scholar] [CrossRef] [PubMed]
  161. Deisseroth, A.; Nienhuis, A.; Lawrence, J.; Giles, R.; Turner, P.; Ruddle, F.H. Chromosomal localization of human β globin gene on human chromosome 11 in somatic cell hybrids. Proc. Natl. Acad. Sci. USA 1978, 75, 1456–1460. [Google Scholar] [CrossRef]
  162. E Bauer, D.; Orkin, S.H. Update on fetal hemoglobin gene regulation in hemoglobinopathies. Curr. Opin. Pediatr. 2011, 23, 1–8. [Google Scholar] [CrossRef]
  163. Manning, J.M.; Dumoulin, A.; Li, X.; Manning, L.R. Normal and Abnormal Protein Subunit Interactions in Hemoglobins. J. Biol. Chem. 1998, 273, 19359–19362. [Google Scholar] [CrossRef]
  164. Kidd, R.D.; Russell, J.E.; Watmough, N.J.; Baker, E.N.; Brittain, T. The Role of β Chains in the Control of the Hemoglobin Oxygen Binding Function: Chimeric Human/Mouse Proteins, Structure, and Function. Biochemistry 2001, 40, 15669–15675. [Google Scholar] [CrossRef]
  165. Yamaguchi, T.; Pang, J.; Reddy, K.S.; Surrey, S.; Adachi, K. Role of β112 Cys (G14) in Homo-(β4) and Hetero-(α2β2) Tetramer Hemoglobin Formation. J. Biol. Chem. 1998, 273, 14179–14185. [Google Scholar] [CrossRef]
  166. Borgstahl, G.E.; Rogers, P.H.; Arnone, A. The 1·8 Å Structure of Carbonmonoxy-β4 Hemoglobin: Analysis of a Homotetramer with the R Quaternary Structure of Liganded α2β2 Hemoglobin. J. Mol. Biol. 1994, 236, 817–830. [Google Scholar] [CrossRef]
  167. A Walder, J.; Chatterjee, R.; Steck, T.L.; Low, P.S.; Musso, G.F.; Kaiser, E.T.; Rogers, P.H.; Arnone, A. The interaction of hemoglobin with the cytoplasmic domain of band 3 of the human erythrocyte membrane. J. Biol. Chem. 1984, 259, 10238–10246. [Google Scholar] [CrossRef]
  168. Gell, D.; Kong, Y.; Eaton, S.A.; Weiss, M.J.; Mackay, J.P. Biophysical Characterization of the α-Globin Binding Protein α-Hemoglobin Stabilizing Protein. J. Biol. Chem. 2002, 277, 40602–40609. [Google Scholar] [CrossRef] [PubMed]
  169. Domingues-Hamdi, E.; Vasseur, C.; Fournier, J.-B.; Marden, M.C.; Wajcman, H.; Baudin-Creuza, V. Role of α-Globin H Helix in the Building of Tetrameric Human Hemoglobin: Interaction with α-Hemoglobin Stabilizing Protein (AHSP) and Heme Molecule. PLoS ONE 2014, 9, e111395. [Google Scholar] [CrossRef] [PubMed]
  170. Feng, L.; Gell, D.A.; Zhou, S.; Gu, L.; Kong, Y.; Li, J.; Hu, M.; Yan, N.; Lee, C.; Rich, A.M.; et al. Molecular Mechanism of AHSP-Mediated Stabilization of α-Hemoglobin. Cell 2004, 119, 629–640. [Google Scholar] [CrossRef] [PubMed]
  171. Yu, X.; Kong, Y.; Dore, L.C.; Abdulmalik, O.; Katein, A.M.; Zhou, S.; Choi, J.K.; Gell, D.; Mackay, J.P.; Gow, A.J.; et al. An erythroid chaperone that facilitates folding of α-globin subunits for hemoglobin synthesis. J. Clin. Investig. 2007, 117, 1856–1865. [Google Scholar] [CrossRef]
  172. Mollan, T.L.; Yu, X.; Weiss, M.J.; Olson, J.S. The Role of Alpha-Hemoglobin Stabilizing Protein in Redox Chemistry, Denaturation, and Hemoglobin Assembly. Antioxid. Redox Signal. 2010, 12, 219–231. [Google Scholar] [CrossRef]
  173. Yu, X.; Mollan, T.L.; Butler, A.; Gow, A.J.; Olson, J.S.; Weiss, M.J. Analysis of human α globin gene mutations that impair binding to the α hemoglobin stabilizing protein. Blood 2009, 113, 5961–5969. [Google Scholar] [CrossRef]
  174. Voon, H.P.J.; Vadolas, J. Controlling-globin: A review of-globin expression and its impact on -thalassemia. Haematologica 2008, 93, 1868–1876. [Google Scholar] [CrossRef]
  175. Ho, H.-Y.; Cheng, M.-L.; Chiu, D.T.-Y. Glucose-6-phosphate dehydrogenase—From oxidative stress to cellular functions and degenerative diseases. Redox Rep. 2007, 12, 109–118. [Google Scholar] [CrossRef]
  176. Arese, P.; Gallo, V.; Pantaleo, A.; Turrini, F. Life and Death of Glucose-6-Phosphate Dehydrogenase (G6PD) Deficient Erythrocytes—Role of Redox Stress and Band 3 Modifications. Transfus. Med. Hemotherapy 2012, 39, 328–334. [Google Scholar] [CrossRef]
  177. Janney, S.; Joist, J.; Fitch, C. Excess release of ferriheme in G6PD-deficient erythrocytes: Possible cause of hemolysis and resistance to malaria. Blood 1986, 67, 331–333. [Google Scholar] [CrossRef]
  178. Scott, M.D.; Zuo, L.; Lubin, B.H.; Chiu, D.T. NADPH, not glutathione, status modulates oxidant sensitivity in normal and glucose-6-phosphate dehydrogenase-deficient erythrocytes. Blood 1991, 77, 2059–2064. [Google Scholar] [CrossRef]
  179. Nicol, C.J.; Zielenski, J.; Tsui, L.; Wells, P.G. An embryoprotective role for glucose-6-phosphate dehydrogenase in developmental oxidative stress and chemical teratogenesis. FASEB J. 2000, 14, 111–127. [Google Scholar] [CrossRef] [PubMed]
  180. Francis, R.O.; Jhang, J.S.; Pham, H.P.; Hod, E.A.; Zimring, J.C.; Spitalnik, S.L. Glucose-6-phosphate dehydrogenase deficiency in transfusion medicine: The unknown risks. Vox Sang. 2013, 105, 271–282. [Google Scholar] [CrossRef] [PubMed]
  181. Fibach, E.; Rachmilewitz, E. The Role of Oxidative Stress in Hemolytic Anemia. Curr. Mol. Med. 2008, 8, 609–619. [Google Scholar] [CrossRef] [PubMed]
  182. Haghpanah, S.; Hosseini-Bensenjan, M.; Zekavat, O.R.; Bordbar, M.; Karimi, M.; Ramzi, M.; Asmarian, N. The Effect of N-Acetyl Cysteine and Vitamin E on Oxidative Status and Hemoglobin Level in Transfusion-Dependent Thalassemia Patients: A Systematic Review and Meta-Analysis. Iran. J. Blood Cancer 2021, 15, 22–35. [Google Scholar] [CrossRef]
  183. Halima, W.M.A.B.; Hannemann, A.; Rees, D.; Brewin, J.; Gibson, J. The Effect of Antioxidants on the Properties of Red Blood Cells from P atients with Sickle Cell Anemia. Front. Physiol. 2019, 10, 976. [Google Scholar]
  184. Pallotta, V.; Gevi, F.; D’Alessandro, A.; Zolla, L. Storing red blood cells with vitamin C and N-acetylcysteine prevents oxidative stress-related lesions: A metabolomics overview. Blood Transfus. 2014, 12, 376–387. [Google Scholar]
  185. Delesderrier, E.; Curioni, C.; Omena, J.; Macedo, C.R.; Cople-Rodrigues, C.; Citelli, M. Antioxidant nutrients and hemolysis in sickle cell disease. Clin. Chim. Acta 2020, 510, 381–390. [Google Scholar] [CrossRef]
  186. Fibach, E.; Rachmilewitz, E.A. The role of antioxidants and iron chelators in the treatment of oxidative stress in thalassemia. Ann. N. Y. Acad. Sci. 2010, 1202, 10–16. [Google Scholar] [CrossRef]
  187. Bonkovsky, H.L.; Guo, J.T.; Hou, W.; Li, T.; Narang, T.; Thapar, M. Porphyrin and heme metabolism and the porphyrias. Compr. Physiol. 2013, 3, 365–401. [Google Scholar]
  188. Bissell, D.M.; Lai, J.C.; Meister, R.K.; Blanc, P.D. Role of Delta-aminolevulinic Acid in the Symptoms of Acute Porphyria. Am. J. Med. 2014, 128, 313–317. [Google Scholar] [CrossRef] [PubMed]
  189. Watson, C. Hematin and Porphyria. N. Engl. J. Med. 1975, 293, 605–607. [Google Scholar] [CrossRef] [PubMed]
  190. Tenhunen, R.; Mustajoki, P. Acute Porphyria: Treatment with Heme. Semin. Liver Dis. 1998, 18, 53–55. [Google Scholar] [CrossRef] [PubMed]
  191. Tian, Q.; Li, T.; Hou, W.; Zheng, J.; Schrum, L.W.; Bonkovsky, H.L. Lon Peptidase 1 (LONP1)-dependent Breakdown of Mitochondrial 5-Aminolevulinic Acid Synthase Protein by Heme in Human Liver Cells. J. Biol. Chem. 2011, 286, 26424–26430. [Google Scholar] [CrossRef]
  192. Hift, R.J.; Thunell, S.; Brun, A. Drugs in porphyria: From observation to a modern algorithm-based system for the prediction of porphyrogenicity. Pharmacol. Ther. 2011, 132, 158–169. [Google Scholar] [CrossRef]
  193. Gardner, L.; Smith, S.; Cox, T. Biosynthesis of delta-aminolevulinic acid and the regulation of heme formation by immature erythroid cells in man. J. Biol. Chem. 1991, 266, 22010–22018. [Google Scholar] [CrossRef]
  194. Steinberg, M.H. Hydroxyurea Treatment for Sickle Cell Disease. Sci. World J. 2002, 2, 1706–1728. [Google Scholar] [CrossRef]
  195. Steinberg, M.H. Therapies to increase fetal hemoglobin in sickle cell disease. Curr. Hematol. Rep. 2003, 2, 95–101. [Google Scholar]
  196. Fathallah, H.; Atweh, G.F. Induction of Fetal Hemoglobin in the Treatment of Sickle Cell Disease. Hematol. Am. Soc. Hematol. Educ. Program 2006, 2006, 58–62. [Google Scholar] [CrossRef]
  197. Cokic, V.P.; Smith, R.D.; Beleslin-Cokic, B.B.; Njoroge, J.M.; Miller, J.L.; Gladwin, M.T.; Schechter, A.N. Hydroxyurea induces fetal hemoglobin by the nitric oxide–dependent activation of soluble guanylyl cyclase. J. Clin. Investig. 2003, 111, 231–239. [Google Scholar] [CrossRef]
  198. Rodgers, G.P.; Dover, G.J.; Noguchi, C.T.; Schechter, A.N.; Nienhuis, A.W. Hematologic Responses of Patients with Sickle Cell Disease to Treatment with Hydroxyurea. N. Engl. J. Med. 1990, 322, 1037–1045. [Google Scholar] [CrossRef] [PubMed]
  199. Anderson, N. Hydroxyurea therapy: Improving the lives of patients with sickle cell disease. Pediatr. Nurs. 2006, 32, 541–543. [Google Scholar]
  200. Goldberg, M.A.; Brugnara, C.; Dover, G.J.; Schapira, L.; Charache, S.; Bunn, H.F. Treatment of Sickle Cell Anemia with Hydroxyurea and Erythropoietin. N. Engl. J. Med. 1990, 323, 366–372. [Google Scholar] [CrossRef] [PubMed]
  201. Dover, G.; Samuel, C. Hydroxyurea induction of fetal hemoglobin synthesis in sickle-cell disease. Semin. Oncol. 1992, 19, 5. [Google Scholar]
  202. Grace, R.F.; Glenthøj, A.; Barcellini, W.; Verhovsek, M.; Rothman, J.A.; Morado, M.; Layton, D.M.; Andres, O.; Galactéros, F.; van Beers, E.J.; et al. Long-Term Hemoglobin Response and Reduction in Transfusion Burden Are Maintained in Patients with Pyruvate Kinase Deficiency Treated with Mitapivat. Blood 2022, 140, 5313–5315. [Google Scholar] [CrossRef]
  203. Grace, R.F.; Rose, C.; Layton, D.M.; Galactéros, F.; Barcellini, W.; Morton, D.H.; van Beers, E.J.; Yaish, H.; Ravindranath, Y.; Kuo, K.H.; et al. Safety and Efficacy of Mitapivat in Pyruvate Kinase Deficiency. N. Engl. J. Med. 2019, 381, 933–944. [Google Scholar] [CrossRef]
  204. Kuo, K.; Layton, D.; Lal, A.; Al-Samkari, H.; Bhatia, J.; Tong, B.; Lynch, M.; Uhlig, K.; Vichinsky, E. Results from a phase 2 study of mitapivat in adults with non–transfusion-dependent alpha- or beta-thalassemia. Hematol. Transfus. Cell Ther. 2021, 43, S28. [Google Scholar] [CrossRef]
  205. Kuo, K.H.; Layton, D.M.; Lal, A.; Al-Samkari, H.; Kosinski, P.A.; Tong, B.; Estepp, J.H.; Uhlig, K.; Vichinsky, E.P. Mitapivat Improves Markers of Erythropoietic Activity in Long-Term Study of Adults with Alpha- or Beta-Non-Transfusion-Dependent Thalassemia. Blood 2022, 140 (Suppl. S1), 2479–2480. [Google Scholar] [CrossRef]
  206. Kuo, K.; Layton, D.; Lal, A.; Al-Samkari, H.; Bhatia, J.; Kosinski, P.; Tong, B.; Lynch, M.; Uhlig, K.; Vichinsky, E. S116: Long-term efficacy and safety of the oral pyruvate kinase activator mitapivat in adults with non—Transfusion-dependent alpha- or beta-thalassemia. HemaSphere 2022, 6, 8–9. [Google Scholar] [CrossRef]
  207. Xu, J.Z.; Conrey, A.; Frey, I.; Gwaabe, E.; A Menapace, L.; Tumburu, L.; Lundt, M.; Li, Q.; Glass, K.; Iyer, V.; et al. Mitapivat (AG-348) Demonstrates Safety, Tolerability, and Improvements in Anemia, Hemolysis, Oxygen Affinity, and Hemoglobin S Polymerization Kinetics in Adults with Sickle Cell Disease: A Phase 1 Dose Escalation Study. Blood 2021, 138 (Suppl. S1), 10. [Google Scholar] [CrossRef]
  208. Pilo, F.; Angelucci, E. Mitapivat for sickle cell disease and thalassemia. Drugs Today 2023, 59, 125–134. [Google Scholar] [CrossRef] [PubMed]
  209. Musallam, K.M.; Taher, A.T.; Cappellini, M.D. Right in time: Mitapivat for the treatment of anemia in α- and β-thalassemia. Cell Rep. Med. 2022, 3, 100790. [Google Scholar] [CrossRef] [PubMed]
  210. Olubiyi, O.O.; Olagunju, M.O.; Strodel, B. Rational Drug Design of Peptide-Based Therapies for Sickle Cell Disease. Molecules 2019, 24, 4551. [Google Scholar] [CrossRef] [PubMed]
  211. Lee, M.T.; Ogu, U.O. Sickle cell disease in the new era: Advances in drug treatment. Transfus. Apher. Sci. 2022, 61, 103555. [Google Scholar] [CrossRef]
  212. Ali, M.A.; Ahmad, A.; Chaudry, H.; Aiman, W.; Aamir, S.; Anwar, M.Y.; Khan, A. Efficacy and safety of recently approved drugs for sickle cell disease: A review of clinical trials. Exp. Hematol. 2020, 92, 11–18.e1. [Google Scholar] [CrossRef]
  213. Ataga, K.I.; Desai, P.C. Advances in new drug therapies for the management of sickle cell disease. Expert Opin. Orphan Drugs 2018, 6, 329–343. [Google Scholar] [CrossRef]
  214. Torres, L.; Conran, N. Emerging pharmacotherapeutic approaches for the management of sickle cell disease. Expert Opin. Pharmacother. 2018, 20, 173–186. [Google Scholar] [CrossRef]
  215. Carden, M.A.; Little, J. Emerging disease-modifying therapies for sickle cell disease. Haematologica 2019, 104, 1710–1719. [Google Scholar] [CrossRef]
  216. Huang, B.; Ghatge, M.S.; Donkor, A.K.; Musayev, F.N.; Deshpande, T.M.; Al-Awadh, M.; Alhashimi, R.T.; Zhu, H.; Omar, A.M.; Telen, M.J.; et al. Design, Synthesis, and Investigation of Novel Nitric Oxide (NO)-Releasing Aromatic Aldehydes as Drug Candidates for the Treatment of Sickle Cell Disease. Molecules 2022, 27, 6835. [Google Scholar] [CrossRef]
  217. Nakagawa, A.; Lui, F.E.; Wassaf, D.; Yefidoff-Freedman, R.; Casalena, D.; Palmer, M.A.; Meadows, J.; Mozzarelli, A.; Ronda, L.; Abdulmalik, O.; et al. Identification of a Small Molecule that Increases Hemoglobin Oxygen Affinity and Reduces SS Erythrocyte Sickling. ACS Chem. Biol. 2014, 9, 2318–2325. [Google Scholar] [CrossRef]
  218. Kassa, T.; Strader, M.B.; Nakagawa, A.; Zapol, W.M.; Alayash, A.I. Targeting βCys93 in hemoglobin S with an antisickling agent possessing dual allosteric and antioxidant effects. Metallomics 2017, 9, 1260–1270. [Google Scholar] [CrossRef] [PubMed]
  219. Kassa, T.; Wood, F.; Strader, M.B.; Alayash, A.I. Antisickling Drugs Targeting βCys93 Reduce Iron Oxidation and Oxidative Changes in Sickle Cell Hemoglobin. Front. Physiol. 2019, 10, 931. [Google Scholar] [CrossRef] [PubMed]
  220. Garel, M.C.; Domenget, C.; Caburi-Martin, J.; Prehu, C.; Galacteros, F.; Beuzard, Y. Covalent binding of glutathione to hemoglobin. I. Inhibition of hemoglobin S polymerization. J. Biol. Chem. 1986, 261, 14704–14709. [Google Scholar] [CrossRef] [PubMed]
  221. Garel, M.; Domenget, C.; Galacteros, F.; Martincaburi, J.; Beuzard, Y. Inhibition of erythrocyte sickling by thiol reagents. Mol. Pharmacol. 1984, 26, 559–565. [Google Scholar]
  222. Abdulmalik, O.; Safo, M.K.; Chen, Q.; Yang, J.; Brugnara, C.; Ohene-Frempong, K.; Abraham, D.J.; Asakura, T. 5-hydroxymethyl-2-furfural modifies intracellular sickle haemoglobin and inhibits sickling of red blood cells†,‡. Br. J. Haematol. 2005, 128, 552–561. [Google Scholar] [CrossRef]
  223. Xu, G.G.; Pagare, P.P.; Ghatge, M.S.; Safo, R.P.; Gazi, A.; Chen, Q.; David, T.; Alabbas, A.B.; Musayev, F.N.; Venitz, J.; et al. Design, Synthesis, and Biological Evaluation of Ester and Ether Derivatives of Antisickling Agent 5-HMF for the Treatment of Sickle Cell Disease. Mol. Pharm. 2017, 14, 3499–3511. [Google Scholar] [CrossRef]
  224. Stern, W.; Mathews, D.; McKew, J.; Shen, X.; Kato, G.J. A Phase 1, First-in-Man, Dose-Response Study of Aes-103 (5-HMF), an Anti-Sickling, Allosteric Modifier of Hemoglobin Oxygen Affinity in Healthy Norman Volunteers. Blood 2012, 120, 3210. [Google Scholar] [CrossRef]
  225. Lucas, A.; Ao-Ieong, E.S.Y.; Williams, A.T.; Jani, V.P.; Muller, C.R.; Yalcin, O.; Cabrales, P. Increased Hemoglobin Oxygen Affinity With 5-Hydroxymethylfurfural Supports Cardiac Function During Severe Hypoxia. Front. Physiol. 2019, 10, 1350. [Google Scholar] [CrossRef]
  226. Hannemann, A.; Cytlak, U.M.; Rees, D.C.; Tewari, S.; Gibson, J.S. Effects of 5-hydroxymethyl-2-furfural on the volume and membrane permeability of red blood cells from patients with sickle cell disease. J. Physiol. 2014, 592, 4039–4049. [Google Scholar] [CrossRef]
  227. Alhashimi, R.T.; Ghatge, M.S.; Donkor, A.K.; Deshpande, T.M.; Anabaraonye, N.; Alramadhani, D.; Danso-Danquah, R.; Huang, B.; Zhang, Y.; Musayev, F.N.; et al. Design, Synthesis, and Antisickling Investigation of a Nitric Oxide-Releasing Prodrug of 5HMF for the Treatment of Sickle Cell Disease. Biomolecules 2022, 12, 696. [Google Scholar] [CrossRef]
  228. Safo, M.K.; Abdulmalik, O.; Danso-Danquah, R.; Burnett, J.C.; Nokuri, S.; Joshi, G.S.; Musayev, F.N.; Asakura, T.; Abraham, D.J. Structural Basis for the Potent Antisickling Effect of a Novel Class of Five-Membered Heterocyclic Aldehydic Compounds. J. Med. Chem. 2004, 47, 4665–4676. [Google Scholar] [CrossRef] [PubMed]
  229. Safo, M.K.; Kato, G.J. Therapeutic Strategies to Alter the Oxygen Affinity of Sickle Hemoglobin. Hematol. Clin. N. Am. 2014, 28, 217–231. [Google Scholar] [CrossRef] [PubMed]
  230. Bonaventura, C.; Fago, A.; Henkens, R.; Crumbliss, A.L. Critical Redox and Allosteric Aspects of Nitric Oxide Interactions with Hemoglobin. Antioxid. Redox Signal. 2004, 6, 979–991. [Google Scholar]
  231. Gladwin, M.T.; Ognibene, F.P.; Pannell, L.K.; Nichols, J.S.; Pease-Fye, M.E.; Shelhamer, J.H.; Schechter, A.N. Relative role of heme nitrosylation and β-cysteine 93 nitrosation in the transport and metabolism of nitric oxide by hemoglobin in the human circulation. Proc. Natl. Acad. Sci. USA 2000, 97, 9943–9948. [Google Scholar] [CrossRef] [PubMed]
  232. Sonveaux, P.; Lobysheva, I.I.; Feron, O.; McMahon, T.J. Transport and Peripheral Bioactivities of Nitrogen Oxides Carried by Red Blood Cell Hemoglobin: Role in Oxygen Delivery. Physiology 2007, 22, 97–112. [Google Scholar] [CrossRef]
  233. Stamler, J.S.; Jia, L.; Eu, J.P.; McMahon, T.J.; Demchenko, I.T.; Bonaventura, J.; Gernert, K.; Piantadosi, C.A. Blood Flow Regulation by S-Nitrosohemoglobin in the Physiological Oxygen Gradient. Science 1997, 276, 2034–2037. [Google Scholar] [CrossRef]
  234. Frehm, E.; Bonaventura, J.; Gow, A. Serial Review: Biomedical Implications for Hemoglobin Interactions with Nitric Oxide Serial Review Editors: Mark T. Gladwin and Rakesh Patel S-nitrosohemoglobin: An allosteric mediator of no group function in m ammalian vasculature. Free Radic. Biol. Med. 2004, 37, 11. [Google Scholar]
  235. Su, H.; Liu, X.; Du, J.; Deng, X.; Fan, Y. The role of hemoglobin in nitric oxide transport in vascular system. Med. Nov. Technol. Devices 2020, 5, 100034. [Google Scholar] [CrossRef]
  236. Lancaster, J.; Hutchings, A.; Kerby, J.D.; Patel, R.P. The hemoglobin-nitric oxide axis: Implications for transfusion therapeutics. Transfus. Altern. Transfus. Med. 2007, 9, 273–280. [Google Scholar] [CrossRef]
  237. Omar, A.M.; Abdulmalik, O.; Ghatge, M.S.; Muhammad, Y.A.; Paredes, S.D.; El-Araby, M.E.; Safo, M.K. An Investigation of Structure-Activity Relationships of Azolylacryloyl Derivatives Yielded Potent and Long-Acting Hemoglobin Modulators for Reversing Erythrocyte Sickling. Biomolecules 2020, 10, 1508. [Google Scholar] [CrossRef]
  238. Pagare, P.P.; Ghatge, M.S.; Musayev, F.N.; Deshpande, T.M.; Chen, Q.; Braxton, C.; Kim, S.; Venitz, J.; Zhang, Y.; Abdulmalik, O.; et al. Rational design of pyridyl derivatives of vanillin for the treatment of sickle cell disease. Bioorg. Med. Chem. 2018, 26, 2530–2538. [Google Scholar] [CrossRef] [PubMed]
  239. Pagare, P.P.; Rastegar, A.; Abdulmalik, O.; Omar, A.M.; Zhang, Y.; Fleischman, A.; Safo, M.K. Modulating hemoglobin allostery for treatment of sickle cell disease: Current progress and intellectual property. Expert Opin. Ther. Patents 2021, 32, 115–130. [Google Scholar] [CrossRef] [PubMed]
  240. Gopalsamy, A.; Aulabaugh, A.E.; Barakat, A.; Beaumont, K.C.; Cabral, S.; Canterbury, D.P.; Casimiro-Garcia, A.; Chang, J.S.; Chen, M.Z.; Choi, C.; et al. PF-07059013: A Noncovalent Modulator of Hemoglobin for Treatment of Sickle Cell Disease. J. Med. Chem. 2020, 64, 326–342. [Google Scholar] [CrossRef] [PubMed]
  241. Saunthararajah, Y. Targeting sickle cell disease root-cause pathophysiology with small molecules. Haematologica 2019, 104, 1720–1730. [Google Scholar] [CrossRef] [PubMed]
  242. Knee, K.M.; Jasuja, R.; Barakat, A.; Rao, D.; Wenzel, Z.; Sahasrabudhe, P.; Narula, J.; Jasti, J.; Chang, J.S.; Beaumont, K.; et al. A Novel Non-Covalent Modulator of Hemoglobin Improves Anemia and Reduces Sickling in a Mouse Model of Sickle Cell Disease. Blood 2019, 134 (Suppl. S1), 207. [Google Scholar] [CrossRef]
  243. Oder, E.; Safo, M.K.; Abdulmalik, O.; Kato, G.J. New developments in anti-sickling agents: Can drugs directly prevent the polymerization of sickle haemoglobin in vivo? Br. J. Haematol. 2016, 175, 24–30. [Google Scholar] [CrossRef]
  244. Metcalf, B.; Chuang, C.; Dufu, K.; Patel, M.P.; Silva-Garcia, A.; Johnson, C.; Lu, Q.; Partridge, J.R.; Patskovska, L.; Patskovsky, Y.; et al. Discovery of GBT440, an Orally Bioavailable R-State Stabilizer of Sickle Cell Hemoglobin. ACS Med. Chem. Lett. 2017, 8, 321–326. [Google Scholar] [CrossRef]
  245. Kapoor, S.; Little, J.A.; Pecker, L.H. Advances in the Treatment of Sickle Cell Disease. Mayo Clin. Proc. 2018, 93, 1810–1824. [Google Scholar] [CrossRef]
  246. Archer, N.; Galacteros, F.; Brugnara, C. 2015 Clinical trials update in sickle cell anemia. Am. J. Hematol. 2015, 90, 934–950. [Google Scholar] [CrossRef]
  247. Deshpande, T.M.; Pagare, P.P.; Ghatge, M.S.; Chen, Q.; Musayev, F.N.; Venitz, J.; Zhang, Y.; Abdulmalik, O.; Safo, M.K. Rational modification of vanillin derivatives to stereospecifically destabilize sickle hemoglobin polymer formation. Acta Crystallogr. Sect. D Struct. Biol. 2018, 74, 956–964. [Google Scholar] [CrossRef]
  248. Scipioni, M.; Kay, G.; Megson, I.L.; Lin, P.K.T. Synthesis of novel vanillin derivatives: Novel multi-targeted scaffold ligands against Alzheimer’s disease. MedChemComm 2019, 10, 764–777. [Google Scholar] [CrossRef] [PubMed]
  249. Beaudry, F.; Ross, A.; Lema, P.P.; Vachon, P. Pharmacokinetics of vanillin and its effects on mechanical hypersensitivity in a rat model of neuropathic pain. Phytother. Res. 2009, 24, 525–530. [Google Scholar] [CrossRef] [PubMed]
  250. Ashraf, Z.; Rafiq, M.; Seo, S.-Y.; Babar, M.M.; Zaidi, N.-U.S. Synthesis, kinetic mechanism and docking studies of vanillin derivatives as inhibitors of mushroom tyrosinase. Bioorg. Med. Chem. 2015, 23, 5870–5880. [Google Scholar] [CrossRef] [PubMed]
  251. Abraham, D.; Mehanna, A.; Wireko, F.; Whitney, J.; Thomas, R.; Orringer, E. Vanillin, a potential agent for the treatment of sickle cell anemia. Blood 1991, 77, 1334–1341. [Google Scholar] [CrossRef]
  252. Abdulmalik, O.; Pagare, P.P.; Huang, B.; Xu, G.G.; Ghatge, M.S.; Xu, X.; Chen, Q.; Anabaraonye, N.; Musayev, F.N.; Omar, A.M.; et al. VZHE-039, a novel antisickling agent that prevents erythrocyte sickling under both hypoxic and anoxic conditions. Sci. Rep. 2020, 10, 20277. [Google Scholar] [CrossRef]
  253. Pagare, P.P.; Ghatge, M.S.; Chen, Q.; Musayev, F.N.; Venitz, J.; Abdulmalik, O.; Zhang, Y.; Safo, M.K. Exploration of Structure-Activity Relationship of Aromatic Aldehydes Bearing Pyridinylmethoxy-Methyl Esters as Novel Antisickling Agents. J. Med. Chem. 2020, 63, 14724–14739. [Google Scholar] [CrossRef]
  254. Lee, C.; Maestre-Reyna, M.; Hsu, K.; Wang, H.; Liu, C.; Jeng, W.; Lin, L.; Wood, R.; Chou, C.; Yang, J.; et al. Crowning Proteins: Modulating the Protein Surface Properties using Crown Ethers. Angew. Chem. Int. Ed. Engl. 2014, 53, 13054–13058. [Google Scholar] [CrossRef]
  255. Leonard, M.; Dellacherie, E. Acylation of human hemoglobin with polyoxyethylene derivatives. Biochim. Biophys. Acta (BBA) Protein Struct. Mol. Enzymol. 1984, 791, 219–225. [Google Scholar] [CrossRef]
  256. Abraham, D.J.; Wireko, F.C.; Randad, R.S.; Poyart, C.; Kister, J.; Bohn, B.; Liard, J.F.; Kunert, M.P. Allosteric modifiers of hemoglobin: 2-[4-[[(3,5-disubstituted anilino)carbonyl]methyl]phenoxy]-2-methylpropionic acid derivatives that lower the oxygen affinity of hemoglobin in red cell suspensions, in whole blood, and in vivo in rats. Biochemistry 1992, 31, 9141–9149. [Google Scholar] [CrossRef]
  257. Sun, K.; D’alessandro, A.; Ahmed, M.H.; Zhang, Y.; Song, A.; Ko, T.-P.; Nemkov, T.; Reisz, J.A.; Wu, H.; Adebiyi, M.; et al. Structural and Functional Insight of Sphingosine 1-Phosphate-Mediated Pathogenic Metabolic Reprogramming in Sickle Cell Disease. Sci. Rep. 2017, 7, 15281. [Google Scholar] [CrossRef]
  258. Kaca, W.; Roth, R.; Levin, J. Hemoglobin, a newly recognized lipopolysaccharide (LPS)-binding protein that enhances LPS biological activity. J. Biol. Chem. 1994, 269, 25078–25084. [Google Scholar] [CrossRef] [PubMed]
  259. Bahl, N.; Du, R.; Winarsih, I.; Ho, B.; Tucker-Kellogg, L.; Tidor, B.; Ding, J.L. Delineation of Lipopolysaccharide (LPS)-binding Sites on Hemoglobin: From in silico predictions to biophysical characterization. J. Biol. Chem. 2011, 286, 37793–37803. [Google Scholar] [CrossRef] [PubMed]
  260. Lechuga, G.C.; Souza-Silva, F.; Sacramento, C.Q.; Trugilho, M.R.O.; Valente, R.H.; Napoleão-Pêgo, P.; Dias, S.S.G.; Fintelman-Rodrigues, N.; Temerozo, J.R.; Carels, N.; et al. SARS-CoV-2 Proteins Bind to Hemoglobin and Its Metabolites. Int. J. Mol. Sci. 2021, 22, 9035. [Google Scholar] [CrossRef] [PubMed]
  261. Fischer, S.; Nagel, R.L.; Bookchin, R.M.; Roth, E.F.; Tellez-Nagel, I. The binding of hemoglobin to membranes of normal and sickle erythrocytes. Biochim. Biophys. Acta (BBA) Biomembr. 1975, 375, 422–433. [Google Scholar] [CrossRef]
  262. Shaklai, N.; Yguerabide, J.; Ranney, H.M. Interaction of hemoglobin with red blood cell membranes as shown by a fluorescent chromophore. Biochemistry 1977, 16, 5585–5592. [Google Scholar] [CrossRef]
  263. Yenamandra, A.; Marjoncu, D. Voxelotor: A Hemoglobin S Polymerization Inhibitor for the Treatment of Sickle Cell Disease. J. Adv. Pract. Oncol. 2020, 11, 873–877. [Google Scholar] [CrossRef]
  264. Singh, J.; Maggo, S.; Sadananden, U.K. Voxelotor: Novel drug for sickle cell disease. Int. J. Basic Clin. Pharmacol. 2020, 9, 513–517. [Google Scholar] [CrossRef]
  265. Han, J.; Saraf, S.L.; Gordeuk, V.R. Systematic Review of Voxelotor: A First-in-Class Sickle Hemoglobin Polymerization Inhibitor for Management of Sickle Cell Disease. Pharmacother. J. Hum. Pharmacol. Drug Ther. 2020, 40, 525–534. [Google Scholar] [CrossRef]
  266. Howard, J.; Hemmaway, C.J.; Telfer, P.; Layton, D.M.; Porter, J.; Awogbade, M.; Mant, T.; Gretler, D.D.; Dufu, K.; Hutchaleelaha, A.; et al. A phase 1/2 ascending dose study and open-label extension study of voxelotor in patients with sickle cell disease. Blood 2019, 133, 1865–1875. [Google Scholar] [CrossRef]
  267. Hutchaleelaha, A.; Patel, M.; Washington, C.; Siu, V.; Allen, E.; Oksenberg, D.; Gretler, D.D.; Mant, T.; Lehrer-Graiwer, J. Pharmacokinetics and pharmacodynamics of voxelotor (GBT440) in healthy adults and patients with sickle cell disease. Br. J. Clin. Pharmacol. 2019, 85, 1290–1302. [Google Scholar] [CrossRef]
  268. Vissa, M.; Vichinsky, E. Voxelotor for the treatment of sickle cell disease. Expert Rev. Hematol. 2021, 14, 253–262. [Google Scholar] [CrossRef] [PubMed]
  269. Vasseur, C.; Baudin-Creuza, V. Role of alpha-hemoglobin molecular chaperone in the hemoglobin formation and clinical expression of some hemoglobinopathies. Transfus. Clin. Biol. 2015, 22, 49–57. [Google Scholar] [CrossRef] [PubMed]
  270. Favero, M.E.; Costa, F.F. Alpha-Hemoglobin-Stabilizing Protein: An Erythroid Molecular Chaperone. Biochem. Res. Int. 2011, 2011, 373859. [Google Scholar] [CrossRef] [PubMed]
  271. Weiss, M.J.; Zhou, S.; Feng, L.; Gell, D.A.; Mackay, J.P.; Shi, Y.; Gow, A.J. Role of Alpha Hemoglobin-Stabilizing Protein in Normal Erythropoiesis and β-Thalassemia. Ann. N. Y. Acad. Sci. 2005, 1054, 103–117. [Google Scholar] [CrossRef]
  272. Khandros, E.; Mollan, T.L.; Yu, X.; Wang, X.; Yao, Y.; D’Souza, J.; Gell, D.A.; Olson, J.S.; Weiss, M.J. Insights into Hemoglobin Assembly through in Vivo Mutagenesis of α-Hemoglobin Stabilizing Protein. J. Biol. Chem. 2012, 287, 11325–11337. [Google Scholar] [CrossRef]
  273. Eggleson, K.K.; Duffin, K.L.; Goldberg, D.E. Identification and Characterization of Falcilysin, a Metallopeptidase Involved in Hemoglobin Catabolism within the Malaria Parasite Plasmodium falciparum. J. Biol. Chem. 1999, 274, 32411–32417. [Google Scholar] [CrossRef]
  274. Christina, E.M.; Goldberg, D. Plasmodium falciparum falcilysin: A metalloprotease with dual specificity. J. Biol. Chem. 2003, 278, 38022–38028. [Google Scholar]
  275. Ralph, S.A. Subcellular multitasking—Multiple destinations and roles for the Plasmodium falcilysin protease. Mol. Microbiol. 2007, 63, 309–313. [Google Scholar] [CrossRef]
  276. Ponpuak, M.; Klemba, M.; Park, M.; Gluzman, I.Y.; Lamppa, G.K.; Goldberg, D.E. A role for falcilysin in transit peptide degradation in the Plasmodium falciparum apicoplast. Mol. Microbiol. 2006, 63, 314–334. [Google Scholar] [CrossRef]
  277. Chance, J.; Hannah, F.; Hernandez, O.; Istvan, E.; Armann, A.; Maslov, N.; Ruby, A.; Teodulo, C.; Huyen, N.; Brian, V.; et al. Development of piperazine-based hydroxamic acid inhibitors against fal cilysin, an essential malarial protease. Bioorg. Med. Chem. Lett. 2018, 28, 1846–1848. [Google Scholar] [CrossRef]
  278. Shen, F.; Zheng, G.; Setegne, M.; Tenglin, K.; Izada, M.; Xie, H.; Zhai, L.; Orkin, S.H.; Dassama, L.M.K. A cell-permeant nano-body-based degrader that induces fetal hemoglobin. ACS Cent. Sci. 2022, 8, 1695–1703. [Google Scholar] [CrossRef] [PubMed]
  279. Maolu, Y.; Manizheh, I.; Karin, T.; Thibault, V.; Liting, Z.; Ge, Z.; Arthanari, H.; Laura, M.K.D.; Orkin, S. Evolution of nanobodies specific for BCL11A. Proc. Natl. Acad. Sci. USA 2022, 120, e2218959120. [Google Scholar]
  280. Gülgün, A.; Andaç, M.; Denizli, A.; Duman, M. Recognition of human hemoglobin with macromolecularly imprinted polyme ric nanoparticles using non-covalent interactions. J. Mol. Recognit. 2021, 34, e2935. [Google Scholar]
  281. Khakurel, K.P.; Žoldák, G.; Angelov, B.; Andreasson, J. On the feasibility of time-resolved X-ray powder diffraction of macromolecules using laser-driven ultrafast X-ray sources. J. Appl. Crystallogr. 2024, 57 Pt 4, 1205–1211. [Google Scholar] [CrossRef]
  282. Khakurel, K.P.; Nemergut, M.; Džupponová, V.; Kropielnicki, K.; Savko, M.; Žoldák, G.; Andreasson, J. Design and fabrication of 3D-printed in situ crystallization plates for probing microcrystals in an external electric field. J. Appl. Crystallogr. 2024, 57 Pt 4, 842–847. [Google Scholar] [CrossRef]
Figure 1. (a) Structure of hemoglobin (α2β2): α-subunits are represented in light and dark green, while β-subunits are shown in blue and turquoise. Hemoglobin can interact with natural ligands such as oxygen, carbon dioxide, carbon monoxide, and nitric oxide. (b) Heme–protein interactions in hemoglobin, illustrating the key residues in the first layer (L83, L86, L66, A65, V62, K61, H58, F46, H45, F43, Y42, T39, L136, M32, V132, L101, F98, N97, L91, H87, V93) involved in stabilizing the heme group within the protein. Asteroid plot was calculated using web server Protein Contacts Atlas (https://pca.mbgroup.bio/index.html, accessed on 12 August 2024) [21]. Created with www.BioRender.com.
Figure 1. (a) Structure of hemoglobin (α2β2): α-subunits are represented in light and dark green, while β-subunits are shown in blue and turquoise. Hemoglobin can interact with natural ligands such as oxygen, carbon dioxide, carbon monoxide, and nitric oxide. (b) Heme–protein interactions in hemoglobin, illustrating the key residues in the first layer (L83, L86, L66, A65, V62, K61, H58, F46, H45, F43, Y42, T39, L136, M32, V132, L101, F98, N97, L91, H87, V93) involved in stabilizing the heme group within the protein. Asteroid plot was calculated using web server Protein Contacts Atlas (https://pca.mbgroup.bio/index.html, accessed on 12 August 2024) [21]. Created with www.BioRender.com.
Molecules 30 00385 g001
Figure 2. (a) Depiction of hemoglobin (Hb) instability leading to the formation of Heinz bodies. Structure of hemoglobin (α2β2): α-subunits are represented in light and dark green, while β-subunits are shown in blue and turquoise. (b) Distribution of 819 unstable α-globin variants, with 3.8% classified as unstable, mapped on chromosome 16 (16p13.3) and divided into clinical significance categories: benign, pathogenic, and uncertain significance. (c) Distribution of 771 unstable β-globin variants, with 11% classified as unstable, mapped on chromosome 11 (11p15.5), similarly categorized. (d) Graph showing the number of unstable variants along the amino acid sequence for both α- and β-globin chains, highlighting regions where residues contact the heme group (grey-filled circles). Created with www.BioRender.com.
Figure 2. (a) Depiction of hemoglobin (Hb) instability leading to the formation of Heinz bodies. Structure of hemoglobin (α2β2): α-subunits are represented in light and dark green, while β-subunits are shown in blue and turquoise. (b) Distribution of 819 unstable α-globin variants, with 3.8% classified as unstable, mapped on chromosome 16 (16p13.3) and divided into clinical significance categories: benign, pathogenic, and uncertain significance. (c) Distribution of 771 unstable β-globin variants, with 11% classified as unstable, mapped on chromosome 11 (11p15.5), similarly categorized. (d) Graph showing the number of unstable variants along the amino acid sequence for both α- and β-globin chains, highlighting regions where residues contact the heme group (grey-filled circles). Created with www.BioRender.com.
Molecules 30 00385 g002
Figure 3. Examples of various small molecules binding to hemoglobin and interacting with specific hemoglobin residues. The figure includes the chemical structures of the ligands, such as toluene, 1H-1,2,3-triazole-5-thiol, inositol hexakisphosphate, and others. Additionally, it lists the specific residues in one of the globin chains that interact with these small molecules. The corresponding PDB codes for the hemoglobin-small molecule complexes are also provided for reference.
Figure 3. Examples of various small molecules binding to hemoglobin and interacting with specific hemoglobin residues. The figure includes the chemical structures of the ligands, such as toluene, 1H-1,2,3-triazole-5-thiol, inositol hexakisphosphate, and others. Additionally, it lists the specific residues in one of the globin chains that interact with these small molecules. The corresponding PDB codes for the hemoglobin-small molecule complexes are also provided for reference.
Molecules 30 00385 g003
Table 2. Overview of small molecules and binders targeting hemoglobin: interactions and therapeutic or research potential. The exact appearance times of these compounds may vary, as they depend on how “appearance” is defined—whether it refers to their initial discovery, release date, or first mention in research publications. For this overview, we assume the timeline is based on the release date and/or first mention in relevant research papers.
Table 2. Overview of small molecules and binders targeting hemoglobin: interactions and therapeutic or research potential. The exact appearance times of these compounds may vary, as they depend on how “appearance” is defined—whether it refers to their initial discovery, release date, or first mention in research publications. For this overview, we assume the timeline is based on the release date and/or first mention in relevant research papers.
Small Molecule/BinderAppearance TimeMain Interaction with HemoglobinTherapeutic or Research Impact
VZHE-039
Molecules 30 00385 i001
2020Enhances oxygen transport efficiencyPotential therapeutic agent to improve oxygen delivery
Compound 23 (PF-07059013)
Molecules 30 00385 i002
2020Noncovalent binder that enhances hemoglobin stabilityExplored for reducing sickling in sickle cell disease
INN-310
Molecules 30 00385 i003
2017Vanillin derivative that influences hemoglobin stabilityExplored for impacts on oxygen affinity and hemoglobin stability
TD3
Molecules 30 00385 i004
2017Affects oxygen binding dynamicsPotential for therapeutic use in modifying hemoglobin function
GBT440 (Voxeltor)
Molecules 30 00385 i005
2015Increases oxygen affinity to prevent sickle hemoglobin polymerizationUsed in treating sickle cell disease
18-crown-6
Molecules 30 00385 i006
2013Alters oxygen-binding propertiesUsed in studies for potential modulation of hemoglobin function
Toluene
Molecules 30 00385 i007
2013Solvent in structural studiesUsed to understand hemoglobin structure
TD1
Molecules 30 00385 i008
2013Alters hemoglobin’s structural stabilityStudied for potential therapeutic impacts on hemoglobin function
INN-298
Molecules 30 00385 i009
2005Modifies hemoglobin functionInvestigated for its effects on hemoglobin and potential treatments
L35
Molecules 30 00385 i010
2005Modifies oxygen affinity and functionInvestigated for its potential to treat hemoglobinopathies
5-Hydroxymethylfurfural (5HMF)
Molecules 30 00385 i011
2003Increases oxygen affinity, reducing sicklingExplored for treatment of sickle cell disease
Inositol Hexakisphosphate (IHP)
Molecules 30 00385 i012
1974Stabilizes deoxyhemoglobin, reduces oxygen affinityResearch tool for studying oxygen release mechanics
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

Žoldáková, M.; Novotný, M.; Khakurel, K.P.; Žoldák, G. Hemoglobin Variants as Targets for Stabilizing Drugs. Molecules 2025, 30, 385. https://doi.org/10.3390/molecules30020385

AMA Style

Žoldáková M, Novotný M, Khakurel KP, Žoldák G. Hemoglobin Variants as Targets for Stabilizing Drugs. Molecules. 2025; 30(2):385. https://doi.org/10.3390/molecules30020385

Chicago/Turabian Style

Žoldáková, Miroslava, Michal Novotný, Krishna P. Khakurel, and Gabriel Žoldák. 2025. "Hemoglobin Variants as Targets for Stabilizing Drugs" Molecules 30, no. 2: 385. https://doi.org/10.3390/molecules30020385

APA Style

Žoldáková, M., Novotný, M., Khakurel, K. P., & Žoldák, G. (2025). Hemoglobin Variants as Targets for Stabilizing Drugs. Molecules, 30(2), 385. https://doi.org/10.3390/molecules30020385

Article Metrics

Back to TopTop