6,879
Views
55
CrossRef citations to date
0
Altmetric
Review

Genomic polymorphisms in sickle cell disease: implications for clinical diversity and treatment

&
Pages 443-458 | Published online: 10 Jan 2014

Abstract

Sickle cell disease (SCD) is one of the best characterized human monogenic disorders. The development of molecular biology allowed the identification of several genomic polymorphisms responsible for its clinical diversity. Research on the first genetic modulators of SCD, such as coinheritance of α-thalassemia and haplotypes in the β-globin gene cluster, have been followed by studies associating single nucleotide polymorphisms (SNPs) with variable risks for stroke, leg ulceration, pulmonary hypertension, priapism and osteonecrosis, with differences in the response to hydroxyurea, and with variability in the management of pain. Furthermore, multigenic analyses based on genome-wide association studies have shed light on the importance of the TGF-β superfamily and oxidative stress to the pathogenesis of complex traits in SCD, and may guide future therapeutic interventions on a genetically oriented basis.

Figure 1. Timeline of major scientific achievements relevant to the study of sickle cell disease.

HbF: Fetal hemoglobin; SCD: Sickle cell disease; SNP: Single nucelotide polymorphism.

Figure 1. Timeline of major scientific achievements relevant to the study of sickle cell disease.HbF: Fetal hemoglobin; SCD: Sickle cell disease; SNP: Single nucelotide polymorphism.

More than 50 years after the first description of the chemical difference between normal and sickle cell hemoglobins Citation[1], sickle cell disease (SCD) has become one of the best characterized monogenic human disorders. However, while its clinical heterogeneity has been long recognized, scientific research over the last 20 years has tried to elucidate the role of several factors responsible for its clinical variability. The first single nucleotide polymorphism (SNP) was described in SCD. Since the first identification of haplotypes in the β-globin gene cluster to the present day Citation[2,3], over 19 million SNPs have been described in the human genome Citation[201], and most of these are still lacking a possible clinical correlation or function description. In SCD, previous reviews have reported that several of these polymorphisms are associated not only with the degree of anemia, but also with pain rate, prevalence of stroke, leg ulcers, pulmonary hypertension, osteonecrosis, hepatobiliary complications and priapism, among other several clinical aspects Citation[4–6]. Moreover, treatment with hydroxyurea (HU), the main pharmacological agent known to improve SCD complications, has variable rates of effectiveness and many studies have aimed to investigate genetic variations that could explain why only some patients tolerate and respond to this treatment, while the rest still need to be treated with blood transfusion-based strategies Citation[7]. This article aims to summarize the current knowledge on genetic polymorphisms that are related to the modulation factors of SCD clinical diversity and influence its treatment.

Study types in genomic research

Genomic research has rapidly evolved during the past decades, and the scientific approach to genetic-based understanding of human disease has changed along with technological development. shows a timeline of scientific discoveries important for the study of SCD, and it has become evident that knowledge in this field has been growing faster in time and in quantity alike Citation[8–13]. summarizes the basic characteristics of the main approaches used in genomic research, their advantages and some of their limitations Citation[14–21]. On the one hand, a reasonably great amount of what is known about genetic modulation leading to phenotype variability in SCD stills derives from genetic epidemiological- and candidate gene-based studies (see later). On the other hand, recent technological developments strive to accomplish more extensive and refined, yet progressively less time-consuming, strategies in the study of the human genome. As a matter of fact, the importance of the first, smaller family-based studies that managed to characterize SCD and its Mendelian heritability, and of subsequent studies (e.g., on haplotypes and interaction with α-thalassemia and candidate gene SNPs) should not be underestimated. The completion of the human genome sequencing led to growing knowledge of human genetic variability, enabling the construction of whole electronic libraries containing the SNPs data, such as the International HapMap Project, launched in 2003 Citation[11]. After this, the recognition of the inherent variation in the human genome as a possible explanation to diverse phenotypes, along with the feasibility of DNA sequencing in a large scale, initiated what could be called the ‘genome-wide study era’. Genome-wide association studies (GWASs) have become a new standard in genomic research in several medical fields for the search for genetic risk factors for common diseases (e.g., studies by the Wellcome Trust Case Control Consortium), focusing on coronary artery disease, hypertension, bipolar disorder, breast cancer, rheumatoid arthritis, Crohn’s disease and diabetes Citation[22]. GWASs use microarrays currently able to identify millions of SNPs in a single run, and the further discovery that copy number variants (CNVs; DNA sequences undergoing deletions, duplications, inversions and translocations) may be more frequent in the human genome than previously expected, has been followed by the development of detection and sequencing technologies enabling scientists to look into these variations and their potential biological importance, although a recent study suggested that CNV genome-wide detection not necessarily enriches previous SNP-based GWASs Citation[23]. This warrants further investigation before drawing definitive conclusions on the use of each approach. Deep DNA sequencing should play a major role in confirming genetic variants in a large number of samples from different populations, and should further extend our understanding of differential expression of genes in normal and pathological states. Regarding hematological disorders, GWASs have recently started a new chapter in genetic characterization of SCD, and these first published results have been included in this review. These newer data might give researchers a glimpse of what future steps should be taken in SCD genomics.

The following sections present brief summaries of what is already known about genetic modulation in SCD according to common subphenotypes encountered by the clinician in the care for these patients.

Anemia & hemolysis

Hemoglobin (Hb) and hematocrit levels are highly variable among patients with SCD, usually being lower in homozygous sickle cell anemia and progressively higher in HbS-β thalassemia and HbSC disease Citation[24]. As variability is broad, even among homozygous sickle cell patients alone, one of the first identified genetic modulators was the coinheritance of α-thalassemia (Box 1). Approximately 30–40% of African–American and African–Brazilian patients with sickle cell anemia are heterozygous, and up to 3% are homozygous for the most common deletion-type α-thalassemia mutation – α3,7kbCitation[25–31]. Patients with one or two deleted α-globin genes have reduced hemolysis, lower reticulocyte counts and lower plasmatic lactate dehydrogenase (LDH) levels, owing to a lower mean corpuscular hemoglobin content (MCHC), less deoxygenated HbS polymerization, fewer denser dehydrated cells and fewer irreversibly sickled cells Citation[32,33]. The averages of some studies report that, comparing genotypes αα/αα, α-/αα and α-/α-, mean hemoglobin levels are 8.1, 8.6 and 9.2 g/dl, respectively; mean corpuscular volume (MCV) is 92, 83 and 72 fl, respectively, and reticulocyte percentage is 11.4, 9 and 6.7%, respectively Citation[33–36]. As far as the reduction of denser cells is concerned, studies in the future may reveal whether noncarriers of α-thalassemia have a better response to treatment with newly developed drugs aiming at reversing hemolysis, such as senicapoc, a Gardos channel blocker that preserves sickle cell hydration and increases red blood cell survival Citation[37]. More recently, the coinheritance of the sickle cell gene and α-thalassemia genes has been studied at the populational level, demonstrating the negative epistatic effect between these genes based on how each of them affects intracellular globin production and on protection against malaria. This study shed light on why high frequencies of α-thalassemia are found among sickle cell patients in certain areas, and presented a possible explanation on the exclusion of the sickle cell alelle from thalassemia-containing populations, thus accounting for the different distributions of these mutations across Africa and the Mediterranean Citation[38], and giving a geographic perspective of the SCD/α-thalassemia phenotype.

Fetal Hb & response to HU

The first study on genetic differences accounting for clinical variability in SCD reported the discovery of a site in linkage disequilibrium with the βS gene (HBB Glu6Val) by use of restriction endonucleases Citation[13]. Subsequent studies characterized the presence of at least five different haplotypes of the β-like-globin gene cluster, suggesting distinct geographic origins of the same βS gene (Senegal, Benin, Bantu, Arab–Indian and Cameroon) and documented that haplotypes differed in fetal Hb (HbF) levels Citation[2,3]. HbF (α2γ2) has the ability to inhibit the polymerization of deoxygenated HbS, making its levels an important modulator of the disease Citation[39–42]. Senegal and Arab–Indian haplotypes are strongly associated with higher HbF levels, especially in carriers of the SNP C-T 158 bp 5´ to the γ-globin gene HBG2 (rs7482144), which creates a restriction site for the enzyme XmnI Citation[43–45]. The XmnI-HBG2 polymorphism was the first quantitative-trait locus (QTL) described in HbF heritability, and elevates the relative production of the Gγ-globin. A high Gγ:Aγ ratio is also found in the Cameroon haplotype, but lacks the XmnI-HBG2 SNP Citation[46]. Instead, this haplotype is characterized by an (AT)8(T)5 sequence in the -500 region of the HBB gene, and is associated with an AGCA deletion 5´ to the β-globin gene linked to the AγT allele, possibly increasing the expression of γ-globin Citation[47]. Absence of a Bantu haplotype is apparently associated with higher responses to HU Citation[48]. HU is an antineoplastic agent known to increase HbF levels and is currently a useful pharmacological treatment for SCD. Nonetheless, not all patients in one specific haplotype demonstrate enhanced HbF production with HU, reinforcing the multigenic character of the HbF expression. Although there is evidence that the 5´ hypersensitive site (5´ HS) is highly polymorphic among sickle cell anemia patients Citation[49,50], definite polymorphisms responsible for modulating HbF levels remain to be described.

Since the discovery that haplotypes with higher HbF levels and the association of SCD with hereditary persistence of HbF are associated with decreased comorbidity, there is an increasing interest in unraveling how HbF production is genetically determined, providing sites for modulation and treatment in SCD. Cis-acting regulatory elements are estimated to account for less than 25% of the variability, which has led to new insights into HbF regulation by trans-acting elements. Four trans-acting QTLs have already been linked to F cell number. The chromosomic region Xp22 maps to the FCP locus Citation[51,52], but has not been validated in subsequent studies. At 6q22.3–24, a QTL was described in an Asian–Indian subject Citation[53–55], and higher resolution studies identified the HBS1L-MYB intergenic polymorphism (HMIP) locus blocks 1–3, with HMIP-2 responsible for the most important effect Citation[56], accounting for 3–7% of HbF levels variance. An interaction between a QTL at 8q and the C–T 158 bp SNP appeared to occur in this same family, and such findings were replicated by a study with 874 twin pairs Citation[57]. Six SNPs in a gene in chromosome 8 called TOX have been described to be associated with HbF levels in more than 1500 SCD patients Citation[58]. At 2p16.1, a region mapping to the BCL11A gene was defined by a GWAS in two different cohorts, African–American and African–Brazilian patients Citation[59]. The QTL in intron 2 of this gene, formerly described as an oncogene, accounted for 15% of F cell number variance, and simultaneous analysis of BCL11A rs11886868, XmnI-HBG2 rs7482144 and HMIP loci yielded more than 20% of HbF variation in SCD patients. More recently, another GWAS described a novel region on chromosome 11 containing the ORF gene cluster, which might be important in the regulation of g-globin gene’ expression and HbF production, but it remains to be determined if BCL11A and ORF are involved in the response to HU Citation[60]. SNPs in genes within the 6q22.3–23.2 and 8q11–q12 linkage peaks, and also the ARG2, FLT1, HAO2 and NOS1 genes, have been associated with the response of HbF to HU Citation[61]. SAR1A is a gene encoding the small GTP-binding protein, secretion-associated and RAS-related (SAR) protein. SNPs have been identified in the upstream 5´ untranslated region (-809 C>T, -502 G>T and -385 C>A) of this gene, possibly associated with the HbF response to HU Citation[62]. GWASs and CNV studies should also yield further knowledge about other HbF-related genes and provide in-depth understanding of the molecular mechanism involved in globin switching, offering the possibility of targeted activation of HbF production in the treatment of SCD Citation[63–67].

Vaso-occlusive events & pain treatment

Either in the form of painful episodes, or presenting as a chronic disorder, pain remains an absolute hallmark of SCD, and clinical heterogeneity has always been a challenge to physicians in charge of the care for sickle cell patients. The first genetic trait described to influence the clinical severity of sickle cell painful episodes was HbF level, since high levels of HbF were recognized to be associated with a lower incidence of crises and, therefore, lower mortality Citation[42]. Sickle cell anemia α-thalassemia is not associated with higher HbF levels and by raising the hemoglobin concentration, consequently elevating the blood viscosity, there seems to be an increase in the likelihood of some types of vaso-occlusive events, although there are conflicting results among studies Citation[68–70]. Pharmacogenomics of opioid metabolism has tried to shed some light on the challenging management of sickle cell pain. Some polymorphisms of the CYP2D6 gene cause decreased metabolization of codeine to morphine, having been associated with failure of codeine treatment for pain crisis in children on HU Citation[71], and possibly with a higher likelihood of adults being admitted to a hospital Citation[72]. Differential morphine metabolism by UDP-glucuronosyltransferase 2B7 (UGT2B7) is associated with the SNP -840G→A, contributing to the variability of hepatic clearance of opioids, and represents a suitable candidate for future genotype–phenotype studies in pain management Citation[73]. Tetrahydrobiopterin is a cofactor required for nitric oxide production, and its synthesis rate-limiting enzyme is GTP cyclohydrolase (GCH1). A GCH1 haplotype defined by three linked SNPs, for the first time, has been described to be possibly associated with patients with more frequent pain. Its relevance needs to be confirmed in other samples, as it might be relevant to future studies in pain management Citation[74]. More recent work has correlated polymorphisms in the MBL2 gene to vaso-occlusive events in children with sickle cell anemia. Genotypes related to lower/intermediate levels of serum mannose-binding lectin (MBL) have been associated with a higher frequency of painful episodes, but how MBL variation and the activation of the complement system by the lectin pathway influence the sickling process remains to be better characterized Citation[75,76]. Acute chest syndrome (ACS) is a type of severe vaso-occlusive episode, also thought to be multifactorial. Preliminary data showed that polymorphisms in genes, such as TGFBR3 and SMAD7, are associated with its occurrence, and corroborate the importance of the TGF-β pathway in the modulation of sickle cell anemia Citation[77]. Pediatric ACS was also associated with SNPs in the gene encoding a member of the PI3K/PI4K family involved in cell–cell adhesion, PIK3CG. Older children and adults presenting with ACS were described, showing association with polymorphisms in other genes, such as SMAD1, NRCAM, SMAD3, klotho (KL) and STARD13 (near KL). The klotho protein is encoded by the KL gene in the chromosomal region 13q12. This β-glycosidase-like protein can be membrane bound and secreted, and it may have a role in nitric oxide metabolism, by stimulating nitric oxide production by endothelial cells. Polymorphisms affecting its production could modulate the intensity of oxidative stress in the microcirculation, alter endothelial-dependent vasodilation and, therefore, affect how the vaso-occlusive process takes place in SCD patients Citation[78]. One study has shown the association of the T-786C endothelial nitric oxide synthase (eNOS) gene polymorphism with increased susceptibility to ACS in female patients Citation[79].

Femoral & humeral avascular necrosis

The pathogenesis of femoral and humeral head osteonecrosis in sickle cell patients is not completely understood, but the impression that the coagulation mechanisms play some role in idiopathic osteonecrosis has led to several studies looking at polymorphisms involved in inherited disorders of thrombosis that would justify a putative hypercoagulable state. Controversial published results have failed to clearly demonstrate that known thrombophilic mutations are implicated in SCD osteonecrosis, even though a potential increased risk for avascular necrosis may involve the 5,10-methylene tetrahydrofolate reductase (MTHFR) C677T mutation Citation[80–84]. Platelet adhesion may also play an important role in sickle cell vaso-occlusion. Inherited polymorphisms of platelet glycoproteins have been associated with occlusive arterial disease. One study described the association of the HPA-5b allele with vascular complications in SCD, such as osteonecrosis, stroke, priapism and ACS Citation[85]. Another study failed to confirm this association in a multivariate analysis, but correlated more frequent and more severe vaso-occlusive crises to the HPA-3b/3b genotype Citation[86]. Deletion of one or more α-globin genes remains a well-known genetic risk factor for osteonecrosis in sickle cell anemia Citation[87], as do other forms of SCD with higher hematocrits, such as HbS-β+ thalassemia and HbSC disease, in which avascular necrosis is more frequent and appears earlier Citation[88]. One study described the association of osteonecrosis with SNPs in several genes related to bone metabolism Citation[89]. Ten SNPs in the KL gene could relate to a deficiency of klotho protein function, generating a dysregulation of the vitamin D metabolism, which may trigger cellular damage by the toxic action of increased levels of calcium, phosphorus and 1,25(OH)2D3Citation[90]. Six SNPs in the ANXA2 gene encoding annexin A2, which is involved in osteoblastic mineralization, were also associated with osteonecrosis. Genes encoding proteins in the TGF-β/SMAD/BMP pathway have also been studied. SNPs in the TGF-β receptors type 2 and type 3 genes (TGFBR2 and TGFBR3) have been associated with a higher risk for avascular necrosis. Five SNPs in the BMP6 may be related to the role of BMP6 in bone formation and in inflammation, and the association with BMP6-3 SNP (rs3812163) was apparently confirmed by another study Citation[91,92].

Infection & bacteremia

Infections are one of the leading causes of morbidity among SCD patients, and much effort has been made at discovering genetic predispositions to certain organisms. Children bearing the H/H131 Fcγ RIIA genotype have more frequently shown a positive history of Haemophilus influenzae type B infection Citation[93]. Several studies have investigated the HLA locus and, if on one hand HLA class II DRB1*15 has a potential protective effect, on the other hand, HLA class II DQB1*03 and HLA-E*0101, in homozygosity, seem to increase infection susceptibility Citation[94,95]. HLA DRB1*100101 was more frequently found in children with osteomyelitis Citation[96], and the HLA-G +3142 polymorphism was associated with a possible protective effect against hepatitis C virus (HCV) Citation[97]. Children with SCD were less likely to have infections if bearing genetic variants in the promoter region of the MBL2 gene, which encodes the mannose-binding protein (MBP), although the relationship between protection against infection and MBP low-producing variants has not yet been explained Citation[98]. In sickle cell anemia patients, the G463A polymorphism in the gene encoding myeloperoxidase (MPO) may be a significant genetic modulator that renders patients more susceptible to infection Citation[99]. Furthermore, a case–control study on bacteremia in SCD showed an association of several SNPs in IGF1R, as well as genes of the TGF-β/SMAD/BMP pathway (BMP6, TGFBR3, BMPR1A, SMAD3 and SMAD6), suggesting the importance of this pathway in the immune function Citation[100]. The gene encoding RANTES (CCL5), a chemokine responsible for bringing immune cells to areas of infection or injury, has a genetic variant associated with protection against infection in sickle cell patients Citation[101], while a study on polymorphisms in the gene encoding its receptor, CCR5, has suggested this gene as a potential candidate gene influencing clinical severity in SCD patients, although there was no statistical significance in this study Citation[102].

Stroke

Sickle cell disease is the most common cause of stroke in childhood, and as one of the most important and potentially devastating complications in SCD patients, its genetic susceptibility has been focused on a large number of polymorphism-association studies Citation[103]. Classic modulators of SCD severity, such as βS cluster haplotype and α-thalassemia, have controversial effects. Decreased hemolysis and less dense cells seem to protect carriers of α-thalassemia Citation[24,104–106], but this is still not a scientific consensus Citation[107–109]. The first studies on HLA genes showed that apparently both HLA class I and II are involved in genetic modulation of the stroke phenotype. HLA B*5301, DQB1*0201, DRB1*0301 and *0302 increased the risk for stroke, while HLA DRB1*1501 protected against it Citation[110]. Further research showed that the genetic contribution to stroke is distinct between large-vessel and small-vessel disease, suggesting different pathological mechanisms. With regard to small-vessel disease, adhesion molecules, such as VCAM-1, have been a preferential target of studies. The G1238C allele of the VCAM1 gene was associated with protection against small-vessel stroke, while carriers of the T-1594C allele were reported as more predisposed to cerebrovascular disease Citation[111,112]. A SNP in the low-density lipoprotein receptor, LDLR, has also been shown to be protective against this type of stroke in children Citation[112]. In addition, HLA DPB1*0401 was associated with small-vessel stroke, while HLA DPB1*1701 had a small protective effect against it Citation[113]. Large-vessel stroke seems to have a different genetic predisposition and SNPs in the IL4R and TNFA genes seem to increase risk for this type of stroke more than fivefold Citation[112,114]. A recently published abstract reported some TNFA polymorphisms associated with CNS events other than stroke, such as seizures and transient ischemic attacks Citation[115]. A genetic variant that upregulates transcription of the gene encoding leukotriene C4 synthase, LTC4S (-444)C, is associated with protection against large-vessel disease, even though it increases proinflammatory mediators Citation[114]. HLA A*1012, A*2612 and A*3301 were described to confer susceptibility to large-vesel stroke Citation[113]. A Bayesian network has been developed to analyze SNPs and their complex interaction in the modulation of risk for stroke in SCD. This kind of approach seems to have a better accuracy in predicting multigenic traits, and the findings showed involvement of proteins from the TGF-β pathway, such as BMP6, TGFBR2 and TGFBR3, and P-selectin. GWASs should also lead to functional research of genes involved in angiogenesis in the modulation of stroke risk Citation[116–118].

Gallstones & bilirubin levels

Biliary complications with higher incidence of gallstones are frequent in hemolytic anemias Citation[119,120]. In SCD, studies have shown the influence of a polymorphism in the promoter region of the uridine diphosphate glucuronosyltransferase 1A (UGT1A) gene, which codifies the enzyme uridine diphosphate glucuronosyltransferase 1A, responsible for bilirubin conjugation. The promoter region contains a sequence of TA repetitions, with the wild type containing six, and polymorphic alleles containing five, seven or eight repetitions. Longer repetitions lead to a decrease in the expression of the gene, thus decreasing effectiveness of bilirubin glucuronidation Citation[121], and a reduction in UGT1A transcription associated with the 7/7 genotype accounts for the genetic basis of Gilbert syndrome. In sickle cell patients, the number of TA repetitions has been shown to exert an important influence on serum bilirubin levels Citation[122]. Homozygosity for the 7 alelle has been associated with earlier and increased incidence of gallstones in children, and a higher need for cholecystectomy in adults with symptomatic biliary disease Citation[123–127]. Coinheritance of α-thalassemia decreases hemolysis and, therefore, can decrease bilirubin levels, but data are contradictory in defining whether it is sufficient to compensate for the effect of the 7/7 genotype Citation[128,129].

Renal impairment

Many renal complications are observed in patients with SCD and kidney failure has emerged as an important comorbidity, as survival of these patients has improved over the last decades but, still, little is known about the genetic modulation of renal function in these patients Citation[130–132]. There are insufficient data on the influence of α-thalassemia on sickle cell renal disease, except in sickle cell-trait bearers, where coinheritance seems to have a protective effect against loss of urinary-concentrating ability Citation[133]. The inheritance of the Bantu haplotype has been associated with worse renal function, supposedly because of the lower HbF levels found in this haplotype Citation[134]. With growing evidence of the importance of the TGF-β/BMP pathway in sickle cell-associated complications, one study managed to demonstrate the association of haplotypes in the TGFBR1 gene harboring three SNPs with glomerular filtration rate. The common haplotype I A-A-G was inversely associated with renal function, while higher filtration rates were found in association with the haplotype VI G-G-A Citation[135]. A better knowledge of how genetic modulation may account for renal impairment should develop increasing importance, as late complications, such as heart failure and transfusion-related iron overload, invariably indicate the use of drugs with potential renal adverse effects (e.g., hyperpotassemia with use of angiotensin-converting enzyme inhibitors or proteinuria with use of deferasirox).

Leg ulceration

Leg ulceration is a complication that varies widely among SCD patients, although it is also common to other hemolytic anemias. They affect between 2 and 40% of patients, depending on the origin of the cohort analyzed. Furthermore, this symptom seems to be closely related to lower Hb levels, lower socioeconomical status, and is more frequent in males and virtually inexistent in children. From the molecular point of view, several genes have been linked recently to the development of chronic leg ulcers in SCD. Individuals with coinheritance of α-thalassemia, as well as compound genotypes in SCD (e.g., HbSC and HbS-β) present with a lower degree of hemolysis, higher Hb levels and, consequently, have been shown to protect against the development of chronic cutaneous ulceration Citation[136]. The polymorphisms rs685417 and rs516306 in the KL gene have been associated with leg ulceration, but they seem important in other SCD complications as well, such as ACS, stroke, osteonecrosis and priapism Citation[89,117,137]. TEK tyrosine kinase (or TIE2) is a receptor for angiopoietins implicated in hypoxia-induced and normal angiogenesis Citation[138]. This could account for the importance of the SNPs rs603085 and rs671084 in impairing the healing process of leg ulcers. Other polymorphic genes possibly implicated in the pathogenesis of leg ulcers belong to the TGF-β/BMP pathway. Members of the TGF-β superfamily activate MAPKs through SMAD signaling, affecting cell proliferation, inflammation, immune regulation and response to tissue injury, among others. SNPs increasing the risk for leg ulcers were described in several genes, such as those encoding receptors TGFBR2, TGFBR3 and BMPR1B, BMP6, signaling proteins SMAD7 and SMAD9, SMURF1 (a ligase specific for SMAD), and downstream enzymes, such as MAP2K1 and MAP3K7Citation[137].

Pulmonary hypertension

Pulmonary hypertension (PH) has increasingly become a major issue in SCD care since it was recognized as a relatively frequent complication and strongly related to early mortality, as predicted by the measurement of tricuspid regurgitant jet velocities Citation[139]. Compound heterozygotes, such as HbSC disease and HbS-β-thalassemia, present with a lower risk for PH, probably owing to a lower rate of hemolysis Citation[140]. Thus, genetic susceptibility to the development of PH seems to be related to factors affecting the hemolytic process itself. Once again, the possible association of polymorphisms has been described in KL, TEK and in several genes from the TGF-β/BMP signaling pathway (ACVRL1 and BMPR2 – previously associated with primary idiopathic PH Citation[141,142], along with BMP6 and ADRB1), and may influence how impaired nitric oxide metabolism secondary to hemolysis leads to PH Citation[143]. Preliminary results showed that SNPs in intron 1 of the NEDD4L gene may be associated with an elevated serum level of N-terminal pro-brain natriuretic peptide (NT-pro-BNP) in sickle cell anemia patients enrolled in a GWAS. Since high levels of NT-pro-BNP are related to high mortality and PH, NEDD4L should be a candidate gene in PH pathogenesis Citation[144]. Another abstract reported a protective effect against pulmonary hypertension attributed to CYBR5 T116S polymorphism, which was associated with lower tricuspid regurgitation velocities, presumably due to lower hemolysis rates in carriers of the SNP Citation[145].

Priapism

Priapism is classically associated with sickle cell hemoglobinopathies Citation[146,147] and, as up to 30% of male patients experience this manifestation, some studies have identified genetic polymorphisms associated with priapism as a dominant feature. An association between SNPs in the KL gene and priapism has been established, corroborating a common phenotype in which severe hemolysis, pulmonary hypertension, leg ulceration and priapism are associated with each other Citation[148]. Carrying a TT genotype for rs2249358 yielded an odds ratio of 2.6 compared to a C_ genotype. The odds ratio for the G–G haplotype associating two other SNPs (rs211239 and rs211234) was 2.3 Citation[149]. Other genes whose polymorphisms have been identified as associated with priapism are TEK, TGFBR3, ITGAV, AQP1 and F13A1. The TGFBR3 gene has already been implicated in stroke and avascular necrosis. The type III receptor for TGF-β is strongly expressed by endothelial cells and is required for endothelial cell transformation Citation[150]. AQP1 encodes acquaporin-1, a water channel expressed by both erythrocytes and endothelial cells. AQP1 may influence the sickling process because it regulates cell volume and controls carbon dioxide permeability through the erythrocyte membrane Citation[151,152]. ITGAV encodes the αv subunit of an integrin expressed by the endothelium, and with a key role in angiogenesis Citation[153]. Some polymorphisms in the α-subunit of Factor XIII (F13A1) have been correlated to priapism, and there is evidence of its importance in coagulation, based on a higher risk of ischemic stroke in the general population involving this same gene Citation[154]. Priapism has also been associated with polymorphism of the human platelet antigen-5 gene (ITGA2) Citation[85].

Transfusion therapy in SCD patients

Patients with SCD may need several blood transfusions throughout their lifetime and may be placed on a chronic transfusion program, with increased risks for alloimmunization and iron overload. Polymorphic mutations of the human hemocromatosis HFE gene have been studied in SCD, but no association with higher degrees of iron overload has been established Citation[155]. Alloantibody development after blood transfusion is relatively common in SCD, reaching up to 47% of patients, depending on the cohort Citation[156]. The presence of a polymorphism in the TRIM21 gene has been associated with the rate and time until alloimmunization in children with SCD. Carriers of the T/T genotype were possibly more tolerant to allogenic red blood cells, causing lower rates of alloimmunization and later development of alloantibodies Citation[157].

Clinical severity of SCD

A fine GWAS applied to SCD was conducted by Sebastiani et al.Citation[158], which managed to correlate clinical severity of SCD based on a network model (composed of 25 different clinical parameters) to 40 different SNPs, including genes not previously identified as pathogenetically important in SCD. A microarray containing probes able to identify more than 600,000 SNPs was used, and the need for better statistical analysis was supplied by the development of an additional subanalysis method called SNP set enrichment analysis (SSEA), designed to further evaluate for sets of SNPs that could define regions of the genome where these genetic variations occur with statistical significance; therefore, identifying more candidate genes associated with a certain subphenotype. New genes identified as probably important to the pathogenesis of SCD included KCNK6 (a potassium channel protein) and TNKS (which encodes a protein called tankyrase1, a polymerase that may influence telomere shortening), and also several other genes that are not directly related to SCD pathophysiology or have unknown functions.

This first study of SCD heterogeneity through a GWAS still needs additional validation and functional studies, but it represents a major scientific breakthrough regarding the possibilities to identify novel genes and signaling pathways possibly involved in the pathophysiology of different SCD complications. Furthermore, GWAS studies should yield more data about non-protein-encoding regions of the human genome, whose functions remain obscure. It is also an example of how important developments in the statistics field have become to the study in populational and genome-wide genetic studies.

Expert commentary

The unexpected clinical diversity in a monogenic disease such as SCD has led to countless genetic studies and current knowledge has evolved, together with technological development in molecular biology. Evolution from the basic identification of polymorphic sites with restriction endonucleases to robust GWASs has provided the tools to discover the genetic complexity that affects genotype–phenotype correlation. The main associations between polymorphic genes, their importance and subphenotypes in SCD are summarized in , with special attention to the KL gene and genes involved in the TGF-β/SMAD/BMP pathway in .

The number of SNPs and genes possibly associated with clinical variability of SCD are increasing very rapidly, but great amounts of data need to be confirmed in larger multicentric studies before they may be used as an effective tool in the management of the patients. Possibly, in the near future, a number of genetic predictors of the variability will be relevant for clinical purposes.

The possibility of working with a whole-genome perspective has turned sample size into an important issue, since the pursuit for genetic variants with smaller effects meant a larger number of samples, which has led to an increasing risk of discovering ‘false-positive’ associations. This, along with the difficulty to validate and replicate results, might be the main challenges researchers will have to face in GWASs, so multicentric studies and multiprofessional work (particularly including bioinformatic experts) should progressively become a standard in this field.

Five-year view

In 5 years, GWASs, Bayesian networks and CNV studies will have yielded much more information about new genes, polymorphisms and their correlation with most subphenotypes in SCD. Individual studies of each gene shall lead to a better interpretation of associations with SNPs lacking an obvious pathophysiological link to clinical diversity, as well as a better understanding of the underlying mechanisms of gene expression and regulation and, possibly, lead to the characterization of new signaling pathways that could provide novel therapeutics aims. The probable evolution of GWASs should be genome-wide gene-interaction studies, in which characterization of epistatic phenomena might uncover how genetic variations in noncoding regions could be responsible for transcriptional, translational or post-translational regulation, and could affect phenotype and gene–environment interactions as well. Further studies should also strive for a more genetically oriented therapeutic approach to this heterogenous disease, with special interest in the influence of pharmacogenomics on pain management and HbF induction, and the development of new drugs based on updated knowledge on vascular biology in the sickle cell vaso-occlusive process.

Table 1. Comparison among different approaches to genomic study in human diseases.

Table 2. Main associations between genes and complications in sickle cell disease.

Table 3. Polymorphisms in genes involved in the TGF-β/SMAD/BMP pathway and in the klotho gene, and genotypes associated with subphenotypes in sickle cell disease.

Box 1. Effects of α-thalassemia on sickle cell disease.

  • • Lower mean corpuscular hemoglobin content

  • • Fewer denser dehydrated cells and irreversibly sickled cells

  • • Reduced hemolysis

  • • Lower reticulocyte counts

  • • Lower plasmatic lactate dehydrogenase levels

  • • Increased frequency of vaso-occlusive events

  • • Increased risk for osteonecrosis

  • • Decreased risk for stroke

  • • Decreased risk for leg ulceration

Key issues

  • • Sickle cell disease is a monogenic, yet clinically heterogenous entity, and several genetic polymorphisms are markers of genes that may account for this variation.

  • • Higher fetal hemoglobin (HbF) levels associated with the XmnI-HBG2 single nucleotide polymorphism (SNP) and coinheritance of α-thalassemia were among the first genetic modulators to be described.

  • • Current knowledge on BCL11A and other genes may explain how HbF levels are regulated, and may lead to novel therapeutic approaches in HbF induction in sickle cell disease.

  • • Further functional studies of genetic polymorphisms involving KL, TEK and genes in the TGF-β/SMAD/BMP pathway may explain different clinical presentations of several complications, such as osteonecrosis, pulmonary hypertension, priapism, leg ulcers and renal failure, and may allow the development of novel therapies for these complications.

  • • Genome-wide association studies and deep DNA sequencing studies may lead to Bayesian network models able to explain multigenic traits in sickle cell disease, identify novel genes and subsequent variations in signaling pathways involved in its pathophysiology, and guide future therapeutic interventions on a genetically oriented basis.

  • • Further multicentric research and development of more complex bioinformatic analysis are warranted to characterize genetic variants in sickle cell disease among different populations across the world.

Acknowledgements

We would like to thank Nicola Conran for the English review of this manuscript.

Financial & competing interests disclosure

The authors are supported by FAPESP and CNPq, Brazil. The Hematology and Hemotherapy Center, UNICAMP, forms part of the National Institute of Science and Technology of Blood (INCT do Sangue - CNPq/MCT). The authors have no other relevant affiliations or financial involvement with any organization or entity with a financial interest in or financial conflict with the subject matter or materials discussed in the manuscript apart from those disclosed.

No writing assistance was utilized in the production of this manuscript.

Notes

Indicates studies with controversial results.

References

  • Ingram VM. Gene mutations in human haemoglobin: the chemical difference between normal and sickle cell haemoglobin. Nature180(4581), 326–328 (1957).
  • Nagel RL, Labie D. DNA haplotypes and the bS globin gene. Prog. Clin. Biol. Res.316B, 371–393 (1989).
  • Nagel RL, Steinberg MH. Role of epistatic (modifier) genes in the modulation of the phenotypic diversity of sickle cell anemia. Pediatr. Pathol. Mol. Med.20, 123–136 (2001).
  • Steinberg MH, Adewoye AH. Modifier genes and sickle cell anemia. Curr. Opin. Hematol.13, 131–136 (2006).
  • Steinberg MH. Predicting clinical severity in sickle cell anaemia. Br. J. Haematol.129, 465–481 (2005).
  • Steinberg MH. Genetic etiologies for phenotypic diversity in sickle cell anemia. ScientificWorldJournal18, 46–67 (2009).
  • Steinberg MH. Sickle cell anemia, the first molecular disease: overview of molecular etiology, pathophysiology, and therapeutic approaches. ScientificWorldJournal8, 1295–1324 (2008).
  • Pauling L, Itano HA, Singer SJ, Wells IC. Sickle cell anemia, a molecular disease. Science109, 443 (1949).
  • Platt OS, Orkin SH, Dover G, Beardsley GP, Miller B, Nathan DG. Hydroxyurea enhances fetal hemoglobin production in sickle cell anemia. J. Clin. Invest.74, 652–656 (1984).
  • Charache S, Terrin ML, Moore RD et al. Effect of hydroxyurea on the frequency of painful crises in sickle cell anemia. Investigators of the Multicenter Study of Hydroxyurea in Sickle Cell Anemia. N. Engl. J. Med.332, 1317–1322 (1995).
  • International HapMap Consortium. The International HapMap project. Nature426, 789–796 (2003).
  • Sedgewick AE, Timofeev N, Sebastiani P et al. BCL11A is a major HbF quantitative trait locus in three different populations with β-hemoglobinopathies. Blood Cells Mol. Dis.41, 255–258 (2008).
  • Kan YW, Dozy AM. Polymorphism of DNA sequence adjacent to human β-globin structural gene: relationship to sickle mutation. Proc. Natl Acad. Sci. USA75(11), 5631–5635 (1978).
  • Dekker MCJ, van Duijn CM. Prospects of genetic epidemiology in the 21st Century. Eur. J. Epidemiol.18, 607–616 (2003).
  • Freimer N, Sabatti C. The use of pedigree, sib-pair and association studies of common diseases for genetic mapping and epidemiology. Nat. Genet.36, 1045–1051 (2004).
  • Cordell HJ. Detecting gene–gene interactions that underlie human diseases. Nat. Rev. Genet.10, 392–404 (2009).
  • Manolio TA, Collins FS, Cox NJ et al. Finding the missing heritability of complex diseases. Nature461, 747–753 (2009).
  • Teo YY, Small KS, Kwiatkowski DP. Methodological challenges of genome-wide association analysis in Africa. Nat. Rev. Genet.11, 149–160 (2010)
  • Pepke S, Wold B, Mortazavi A. Computation for ChIP-seq and RNA-seq studies. Nat. Methods6(11 Suppl.), S22–S32 (2009).
  • Park PJ. ChIP-seq: advantages and challenges of a maturing technology. Nat. Rev. Genet.10, 669–680 (2009).
  • Gräslund T, Li X, Magnenat L, Popkov M, Barbas CF 3rd. Exploring strategies for the design of artificial transcription factors: targeting sites proximal to known regulatory regions for the induction of γ-globin expression and the treatment of sickle cell disease. J. Biol. Chem.280, 3707–3714 (2005).
  • Wellcome Trust Case Control Consortium. Genome-wide association study of 14,000 cases of seven common diseases and 3,000 shared controls. Nature447, 661–678 (2007).
  • Wellcome Trust Case Control Consortium. Genome-wide association study of CNVs in 16,000 cases of eight common diseases and 3,000 shared controls. Nature464, 713–720 (2010).
  • Steinberg MH. Compound heterozygous and other hemoglobinopathies. In: Disorders of Hemoglobin: Genetics, Pathophysiology, and Clinical Management. Steinberg MH, Forget BG, Higgs DR, Nagel RL (Eds). Cambridge University Press, Cambridge, UK 786–810 (2001).
  • Figueiredo MS, Kerbauy J, Gonçalves MS et al. Effect of α-thalassemia and β-globin gene cluster haplotypes on the hematological and clinical features of sickle-cell anemia in Brazil. Am. J. Hematol.53, 72–76 (1996).
  • Kéclard L, Ollendorf V, Berchel C et al. βS haplotypes, α-globin gene status, and hematological data of sickle cell disease patients in Guadeloupe. Hemoglobin20, 63–74 (1996).
  • Steinberg MH, Embury SH. α-thalassemia in blacks: genetic and clinical aspects and interactions with the sickle hemoglobin gene. Blood68, 985–990 (1986).
  • Schroeder WA, Powars DR, Kay LM et al. β-cluster haplotypes, α-gene status, and hematological data from SS, SC, and S-β-thalassemia patients in southern California. Hemoglobin13(4), 325–353 (1989).
  • Mukherjee MB, Colah RB, Ghosh K et al. Milder clinical course of sickle cell disease in patients with α-thalassemia in the Indian subcontinent. Blood89, 732 (1997).
  • Costa FF, Tavella MH, Zago MA. Deletion type α-thalassemia among Brazilian patients with sickle cell anemia. Brazil. J. Genetics12(3), 605–611 (1989).
  • Sonati MF, Farah SB, Ramalho AS, Costa FF. High prevalence of α-thalassemia in a black population of Brazil. Hemoglobin15(4), 309–311 (1991).
  • Noguchi CT, Dover GJ, Rodgers GP et al. α thalassemia changes erythrocyte heterogeneity in sickle cell disease. J. Clin. Invest.75(5), 1632–1637 (1985).
  • Embury SH, Dozy AM, Miller J et al. Concurrent sickle-cell anemia and α-thalassemia. Effect on severity of anemia. N. Engl. J. Med.306, 270 (1982).
  • de Ceulaer D, Higgs DR. Weatherall DJ, Hayes RJ, Serjeant BE, Serjeant GR. α-thalassemia reduces the hemolytic rate in homozygous sickle cell disease. N. Engl. J. Med.309, 189 (1983).
  • Higgs DR, Aldridge BE, Lamb J et al. The interaction of α-thalassemia and homozygous sickle cell disease. N. Engl. Med.306, 1441 (1982).
  • Steinberg MH, Rosenstock W, Coleman MB et al. Effects of thalassemia and microcytosis upon the hematological and vaso-occlusive severity of sickle cell anemia. Blood63, 1353 (1984).
  • Ataga KI, Smith WR, De Castro LM et al. Efficacy and safety of the Gardos channel blocker, senicapoc (ICA-17043), in patients with sickle cell anemia. Blood111(8), 3991–3997 (2008).
  • Penman BS, Pybus OG, Weatherall DJ, Gupta S. Epistatic interactions between genetic disorders of hemoglobin can explain why the sickle-cell gene is uncommon in the Mediterranean. Proc. Natl Acad. Sci. USA106, 21242–21246. (2009).
  • Goldberg MA, Husson MA, Bunn HF. Participation of hemoglobins A and F in polymerization of sickle hemoglobin. J. Biol. Chem.252, 3414–3421 (1977).
  • Sunshine HR, Hofrichter J, Eaton WA. Gelation of sickle cell hemoglobin in mixtures with normal adult and fetal hemoglobins. J. Mol. Biol.133, 435–467 (1979).
  • Charache S. Fetal hemoglobin, sickling, and sickle cell disease. Adv. Pediatr.37, 1–31 (1990).
  • Benesch RE, Edalji R, Benesch R, Kwong S. Solubilization of hemoglobin S by other hemoglobins. Proc. Natl Acad. Sci. USA77, 5130–5134 (1980).
  • Gonçalves MS, Nechtman JF, Figueiredo MS et al. Sickle cell disease in a Brazilian population from São Paulo: a study of the βS haplotypes. Hum. Hered.44, 322–327 (1994).
  • Ballas SK, Talacki CA, Adachi K, Schwartz E, Surrey S, Rappaport E. The Xmn I site (-158, C–T) 5´ to the G γ gene: correlation with the Senegalese haplotype and G γ globin expression. Hemoglobin15(5), 393–405 (1991).
  • el-Hazmi MA, Bahakim HM, Warsy AS. DNA polymorphism in the β-globin gene cluster in Saudi Arabs: relation to severity of sickle cell anaemia. Acta Haematol.88(2–3), 61–66 (1992).
  • Lapoumeroulie C, Dunda O, Ducrocq R et al. A novel sickle gene of yet another origin in Africa: the Cameroon type. Hum. Genet.89, 333–337 (1989).
  • Gilman JG, Huisman THJ. DNA sequence variation associated with elevated foetal Gγ globin production. Blood66, 783–787 (1985).
  • Steinberg MH, Lu ZH, Barton FB, Terrin ML, Charache S, Dover GJ. Fetal hemoglobin in sickle cell anemia: determinants of response to hydroxyurea. Multicenter Study of Hydroxyurea. Blood89, 1078–1088 (1997).
  • Lu ZH, Steinberg MH. Fetal hemoglobin in sickle cell anemia: relation to regulatory sequences cis to the β-globin gene. Blood87, 1604–1611 (1996).
  • Ofori-Acquah SF, Lalloz MRA, Layton DM. Localisation of cis-active determinants of fetal hemoglobin level in sickle cell anemia. Blood88, 493a (1996).
  • Chang YPC, Maier-Redelsperger M, Smith KD et al. The relative importance of the X-linked FCP locus and β-globin haplotypes in determining hemoglobin F levels: a study of SS patients homozygous for βS haplotypes. Br. J. Haematol.96, 806–814 (1997).
  • Dover GJ, Smith KD, Chang YC et al. Fetal hemoglobin levels in sickle cell disease and normal individuals are partially controlled by an X-linked gene located at Xp22.2. Blood80, 816–824 (1992).
  • Thein SL, Weatherall DJ. A non-deletion hereditary persistence of foetal hemoglobin (HPFH) determinant not linked to the β-globin gene complex. In: Hemoglobin Switching, Part B: Cellular and Molecular Mechanisms. Stamatoyannopoulos G, Nienhuis AW (Eds). Alan R Liss Inc., NY, USA 97–111 (1989).
  • Garner C, Mitchell J, Hatzis T, Reittie J, Farrall M, Thein SL. Haplotype mapping of a major quantitative-trait locus for fetal hemoglobin production, on chromosome 6q23. Am. J. Hum. Genet.62, 1468–1474 (1998).
  • Craig JE, Rochette J, Fisher CA et al. Dissecting the loci controlling fetal haemoglobin production on chromosomes 11p and 6q by the regressive approach. Nat. Genet.12, 58–64 (1996).
  • Creary LE, Ulug P, Menzel S et al. Genetic variation on chromosome 6 influences F cell levels in healthy individuals of African descent and HbF levels in sickle cell patients. PLoS One4, e4218 (2009).
  • Garner C, Menzel S, Martin C et al. Interaction between two quantitative trait loci affects fetal haemoglobin expression. Ann. Hum. Genet.69,707–714 (2005).
  • Sebastiani P, Wang L, Nolan VG et al. Fetal hemoglobin in sickle cell anemia: Bayesian modeling of genetic associations. Am. J. Hematol.83,189–195 (2008)
  • Lettre G, Sankaran VG, Bezerra MA et al. DNA polymorphisms at the BCL11A, HBS1L-MYB, and β-globin loci associate with fetal hemoglobin levels and pain crises in sickle cell disease. Proc. Natl Acad. Sci. USA105(33), 11869–11874 (2008).
  • Solovieff N, Milton JN, Hartley SW et al. Fetal hemoglobin in sickle cell anemia: genome-wide association studies suggest a regulatory region in the 5´ olfactory receptor gene cluster. Blood115(9), 1815–1822 (2010).
  • Ma Q, Wyszynski DF, Farrell JJ et al. Fetal hemoglobin in sickle cell anemia: genetic determinants of response to hydroxyurea. Pharmacogenomics J.7, 386–394 (2007).
  • Kumkhaek C, Zhu J, Taylor JG et al. Variation in the small guanosine triphosphate (GTP)-binding protein, secretion-associated and RAS-related (SAR1A) gene and response to hydroxyurea treatment in sickle cell disease. Blood110, 3392 (2007).
  • Dworkis D, Sebastiani P, Melista E et al. Fetal hemoglobin in sickle cell anemia: a novel method for high-resolution discovery of associated genomic copy number variations. Blood112, 2491 (2008).
  • Timofeev N, Sebastiani P, Hartley SH et al. Fetal hemoglobin in sickle cell anemia: a genome-wide association study of the response to hydroxyurea. Blood112, 2471 (2008).
  • McDade J, Flanagan JM, Mortier N et al. Genetic predictors of hydroxyurea response in children with sickle cell disease. Blood114, 820 (2009).
  • Timofeev N, Milton JN, Hartley SW et al. Genome-wide studies in sickle cell anemia show associations between SNPs in the olfactory receptor gene cluster and fetal hemoglobin concentration. Blood114, 821 (2009).
  • Xiu J, Sankaran VG, Fujiwara Y et al. Control of hemoglobin switching by BCL11A. Blood114, 5 (2009).
  • Platt OS, Thorington BD, Brambilla DJ et al. Pain in sickle cell disease. Rates and risk factors. N. Engl. J. Med.325(1), 11–16 (1991).
  • Ballas SK, Larner J, Smith ED, Surrey S, Schwartz E, Rappaport EF. Rheologic predictors of the severity of the painful sickle cell crisis. Blood72(4), 1216–1223 (1988).
  • Billett HH, Nagel RL, Fabry ME. Paradoxical increase of painful crises in sickle cell patients with α-thalassemia. Blood86(11), 4382 (1995).
  • Brousseau DC, McCarver DG, Drendel AL, Divakaran K, Panepinto JA. The effect of CYP2D6 polymorphisms on the response to pain treatment for pediatric sickle cell pain crisis . J. Pediatr.150(6), 623–626 (2007).
  • Shord SS, Cavallari LH, Gao W et al. The pharmacokinetics of codeine and its metabolites in Blacks with sickle cell disease. Eur. J. Clin. Pharmacol.65, 651–658 (2009).
  • Darbari DS, van Schaik RHN, Capparelli EV, Rana S, McCarter R, van den Anker J. UGT2B7 promoter variant -840G>A contributes to the variability in hepatic clearance of morphine in patients with sickle cell disease. Am. J. Hematol.83, 200–202 (2008).
  • Taylor JG, Belfer I, Desai K et al. A GCH1 haplotype associated with susceptibility to vasoocclusive pain and impaired vascular function in sickle cell anemia. Blood114, 575 (2009).
  • Oliveira MC, Mendonça TF, Vasconcelos LR et al. Association of the MBL2 gene EXON1 polymorphism and vasoocclusive crisis in patients with sickle cell anemia. Acta Haematol.121(4), 212–215 (2009).
  • Mendonça TF, Oliveira MC, Vasconcelos LR et al. Association of variant alleles of MBL2 gene with vasoocclusive crisis in children with sickle cell anemia. Blood Cells Mol. Dis. DOI: 10.1016/j.bcmd.2010.02.004 (2010) (Epub ahead of print).
  • Martinez-Castaldi C, Nolan VG, Baldwin CT et al. Association of genetic polymorphisms in the TGF-β pathway with the acute chest syndrome of sickle cell anemia. Blood118, 666a (2007).
  • Saito Y, Yamagishi T, Nakamura T et al. Klotho protein protects against endothelial dysfunction. Biochem. Biophys. Res. Commun.248, 324–329 (1998).
  • Sharan K, Surrey S, Ballas S et al. Association of T-786C eNOS gene polymorphism with increased susceptibility to acute chest syndrome in females with sickle cell disease. Br. J. Haematol.124(2), 240–243 (2004).
  • Kutlar A, Kutlar F, Turker I, Tural C. The methylene tetrahydrofolate reductase (C677T) mutation as a potential risk factor for avascular necrosis in sickle cell disease. Hemoglobin25(2), 213–217 (2001).
  • Andrade FL, Annichino-Bizzacchi JM, Saad ST, Costa FF, Arruda VR. Prothrombin mutant, Factor V Leiden, and thermolabile variant of methylenetetrahydrofolate reductase among patients with sickle cell disease in Brazil. Am. J. Hematol.59(1), 46–50 (1998).
  • Moreira Neto F, Lourenço DM, Noguti MAE et al. The clinical impact of MTHFR polymorphism on the vascular complications of sickle cell disease. Braz. J. Med. Biol. Res.39, 1291–1295 (2006).
  • DeCastro L, Rinder HM, Howe JG, Smith BR. Thrombophilic genotypes do not adversely affect the course of sickle cell disease (SCD). Blood92, 161a (1998).
  • Zimmerman SA, Ware RE. Inherited DNA mutations contributing to thrombotic complications in patients with sickle cell disease. Am. J. Hematol.59(4), 267–272 (1998).
  • Castro V, Alberto FL, Costa RN et al. Polymorphism of the human platelet antigen-5 system is a risk factor for occlusive vascular complications in patients with sickle cell anemia. Vox Sanguinis87, 118–123 (2004).
  • Al-Subaie AM, Fawaz NA, Mahdi N et al. Human platelet alloantigens (HPA) 1, HPA2, HPA3, HPA4, and HPA5 polymorphisms in sickle cell anemia patients with vaso-occlusive crisis. Eur. J. Haematol.83(6), 579–585 (2009).
  • Ballas SK, Talacki CA, Rao VM, Steiner RM. The prevalence of avascular necrosis in sickle cell anemia: correlation with α-thalassemia. Hemoglobin13(7–8), 649–655 (1989).
  • Milner PF, Kraus AP, Sebes JI et al. Sickle cell disease as a cause of osteonecrosis of the femoral head. N. Engl. J. Med.325(21), 1476–1481 (1991).
  • Baldwin C, Nolan VG, Wyszynski DF et al. Association of klotho, bone morphogenic protein 6 and annexin A2 polymorphisms with sickle cell osteonecrosis. Blood106(1), 372–375 (2005).
  • Tsujikawa H, Kurotaki Y, Fujimori T, Fukuda K, Nabeshima Y. Klotho, a gene related to a syndrome resembling human premature aging, functions in a negative regulatory circuit of vitamin D endocrine system. Mol. Endocrinol.17, 2393–2403 (2003).
  • Ulug P, Vasavda N, Awogbade M, Cunningham J, Menzel S, Thein SL. Association of sickle avascular necrosis with bone morphogenic protein 6. Ann. Hematol.88, 803–805 (2009).
  • Steinberg MH, Nolan VG, Baldwin CT et al. Association of polymorphisms in genes of the transforming growth factor-β pathway with sickle cell osteonecrosis. Blood102, 262a–263a (2003).
  • Norris CF, Surrey S, Bunin GR, Schwartz E, Buchanan GR, McKenzie SE. Relationship between Fc receptor IIA polymorphism and infection in children with sickle cell disease. J. Pediatr.128(6), 813–819 (1996).
  • Tamouza R, Neonato M, Busson M et al. Infectious complications in sickle cell disease are influenced by HLA class II alleles. Hum. Immunol.63, 194–199 (2002).
  • Tamouza R, Busson M, Fortier C et al. HLA-E*0101 allele in homozygous state favors severe bacterial infections in sickle cell anemia. Hum. Immunol.68, 849–853 (2007).
  • Al-Ola K, Mahdi N, Al-Subaie AM, Ali ME, Al-Absi IK, Almawi WY. Evidence for HLA class II susceptible and protective haplotypes for osteomyelitis in pediatric patients with sickle cell anemia. Tissue Antigens71(5), 453–457 (2008).
  • Cordero EA, Veit TD, da Silva MA, Jacques SM, Silla LM, Chies JÁ. HLA-G polymorphism influences the susceptibility to HCV infection in sickle cell disease patients. Tissue Antigens74(4), 308–313 (2009).
  • Neonato MG, Lu CY, Guilloud-Bataille M et al. Genetic polymorphism of the mannose-binding protein gene in children with sickle cell disease: identification of three new variant alleles and relationship to infections. Eur. J. Hum. Genet.7, 679–686 (1999).
  • Costa RNP, Conran N, Albuquerque DM, Soares PH, Saad STO, Costa FF. Association of the G-463A myeloperoxidase polymorphism with infection in sickle cell anemia. Haematologica90, 977–979 (2005).
  • Adewoye AH, Nolan VG, Ma Q et al. Association of polymorphisms of IGF1R and genes in the transforming growth factor-β/bone morphogenetic protein pathway with bacteremia in sickle cell anemia. Clin. Infect. Dis.43, 593–598 (2006).
  • Dossou-Yovo OP, Zaccaria I, Benkerrou M et al. Effects of RANTES and MBL2 gene polymorphisms in sickle cell disease clinical outcomes: association of the g.In1.1T>C RANTES variant with protection against infections. Am. J. Hematol.84(6), 378–380 (2009).
  • Vargas AE, da Silva MA, Silla L, Chies JA. Polymorphisms of chemokine receptors and eNOS in Brazilian patients with sickle cell disease. Tissue Antigens66, 683–690 (2005).
  • Hoppe C. Defining stroke risk in children with sickle cell anaemia. Br. J. Haematol.128(6), 751–766 (2005).
  • Adams GT, Snieder H, McKie VC et al. Genetic risk factors for cerebrovascular disease in children with sickle cell disease: design of a case–control association study and genomewide screen. BMC Medical Genet.4, 6 (2003).
  • Neonato MG, Guilloud-Bataille M, Beauvais P et al. Acute clinical events in 299 homozygous sickle cell patients living in France. French Study Group on Sickle Cell Disease. Eur. J. Haematol.65, 155–164 (2000).
  • Hsu LL, Miller ST, Wright E et al. α-thalassemia is associated with decreased risk of abnormal transcranial Doppler ultrasonography in children with sickle cell anemia. J. Ped. Hematol/Oncol.25, 622–628 (2003).
  • Ohene-Frempong K, Weiner SJ, Sleeper LA et al. Cerebrovascular accidents in sickle cell disease: rates and risk factors. Blood91, 288–294 (1998).
  • Miller ST, Macklin EA, Pegelow CH et al. Silent infarction as a risk factor for overt stroke in children with sickle cell anemia: a report from the Cooperative Study of Sickle Cell Disease. J. Pediatr.139, 385–390 (2001).
  • Balkaran B, Char G, Morris J, Thomas P, Serjeant B, Serjeant G. Stroke in a cohort of patients with homozygous sickle cell disease. J. Pediatr.120, 360–366 (1992).
  • Styles LA, Hoppe C, Klitz W et al. Evidence for HLA-related susceptibility for stroke in children with sickle cell disease. Blood95(11), 3562–3567 (2000).
  • Taylor JG, Tang DC, Savage SA et al. Variants in the VCAM1 gene and risk for symptomatic stroke in sickle cell disease. Blood100(13), 4303–4309 (2002).
  • Hoppe C, Klitz W, Cheng S et al. Gene interactions and stroke risk in children with sickle cell anemia. Blood103(6), 2391–2396 (2004).
  • Hoppe C, Klitz W, Noble J et al. Distinct HLA associations by stroke subtype in children with sickle cell anemia. Blood101(7), 2865–2869 (2003).
  • Hoppe C, Klitz W, D’Harlingue K et al. Confirmation of an association between the TNF(-308) promoter polymorphism and stroke risk in children with sickle cell anemia. Stroke38, 2241–2246 (2007).
  • Barber LA, Ashley-Koch AE, Garrett ME et al. Polymorphisms in TNFα are associated with cerebrovascular events in sickle cell disease. Blood114, 1540 (2009).
  • Sebastiani P, Ramoni MF, Nolan V, Baldwin CT, Steinberg MH. Genetic dissection and prognostic modeling of overt stroke in sickle cell anemia. Nat. Genet.37, 435–440 (2005).
  • Steinberg MH, Baldwin CT, Wyszynski DF et al. Stroke in sickle cell anemia: association with single nucleotide polymorphisms in genes affecting vascular function. Blood102, 926 (2003).
  • Sebastiani P, Milton JN, Timofeev N et al. Genome-wide association study of stroke in sickle cell anemia. Blood114, 1528 (2009).
  • Barrett-Connor E. Cholelithiasis in sickle cell anemia. Am. J. Med.45, 889–898 (1968).
  • del Giudice EM, Perrotta S, Nobili B, Specchia C, d’Urzo G, Iolascon A. Coinheritance of Gilbert syndrome increases the risk for developing gallstones in patients with hereditary spherocytosis. Blood94, 2259–2262 (1999).
  • Bosma PJ, Chowdhury JR, Bakker C et al. The genetic basis of the reduced expression of bilirubin UDP-glucuronosultransferase 1 in Gilbert’s syndrome. N. Engl. J. Med.333, 1171–1175 (1995).
  • Carpenter SL, Lieff S, Howard TA, Eggleston B, Ware RE. UGT1A1 promoter polymorphisms and the development of hyperbilirubinemia and gallbladder disease in children with sickle cell anemia. Am. J. Hematol.83(10), 800–803 (2008).
  • Chaar V, Keclard L, Diara JP et al. Association of UGT1A1 polymorphism with prevalence and age at onset of cholelithiasis in sickle cell anemia. Haematologica90, 188–199 (2005).
  • Haverfield EV, McKenzie CA, Forrester T et al. UGT1A1 variation and gallstone formation in sickle cell disease. Blood105, 968–972 (2004).
  • Fertrin KY, Melo MB, Assis AM, Saad ST, Costa FF. UDP-glucuronosyltransferase 1 gene promoter polymorphism is associated with increased serum bilirubin levels and cholecystectomy in patients with sickle cell anemia. Clin. Genet.64, 160–162 (2003).
  • Passon RG, Howard TA, Zimmerman SA, Schultz WH, Ware RE. Influence of bilirubin uridine diphosphate-glucuronosyltransferase 1A promoter polymorphisms on serum bilirubin levels and cholelithiasis in children with sickle cell anemia. J. Pediatr. Hematol. Oncol.23, 448–451 (2001).
  • Martins R, Morais A, Dias A et al. Early modification of sickle cell disease clinical course by UDP-glucuronosyltransferase 1A1 gene promoter polymorphism. Hum. Genet.53, 524–528 (2008).
  • Chaar V, Keclard L, Etienne-Julan M et al. UGT1A1 polymorphism outweighs the modest effect of deletional (-α 3.7kb) α-thalassemia on cholelithogenesis in sickle cell anemia. Am. J. Hematol.81, 377–379 (2006).
  • Vasavda N, Menzel S, Kondaveeti S et al. The linear effects of α-thalassaemia, the UGT1A1 and HMOX1 polymorphisms on cholelithiasis in sickle cell disease. Br. J. Haematol.138, 263–270 (2007).
  • Serjeant GR. Sickle Cell Disease. Oxford University Press, Oxford, UK 261–281 (1992).
  • Pham PT, Pham PC, Wilkinson AH, Lew SQ. Renal abnormalities in sickle cell disease. Kidney Int.57(1), 1–8 (2000).
  • Ataga KI, Orringer EP. Renal abnormalities in sickle cell disease. Am. J. Hematol.63, 205–211 (2000).
  • Gupta AK, Kirchner KA, Nicholson R et al. Effects of α-thalassemia and sickle polymerization tendency on the urine-concentrating defect of individuals with sickle cell trait. J. Clin. Invest.88, 1963–1968 (1991).
  • Powars DR, Elliott M, Chan L. Chronic renal failure in sickle cell disease: risk factors, clinical course and mortality. Ann. Int. Med.115, 614–620 (1991).
  • Nolan VG, Ma Q, Cohen HT et al. Estimated glomerular filtration rate in sickle cell anemia is associated with polymorphisms of bone morphogenetic protein receptor 1B. Am. J. Hematol.82, 179–184 (2007).
  • Koshy M, Entsuah R, Koranda A et al. Leg ulcers in patients with sickle cell disease. Blood74, 1403–1408 (1989).
  • Nolan VG, Adewoye A, Baldwin C et al. Sickle cell leg ulcers: associations with haemolysis and SNPs in Klotho, TEK and genes of the TGF-β/BMP pathway. Br. J. Haematol.133, 570–578 (2006).
  • Peters KG, Kontos CD, Lin PC et al. Functional significance of Tie2 signaling in the adult vasculature. Recent Prog. Horm. Res.59, 51–71 (2004).
  • Gladwin MT, Sachdev V, Jison ML et al. Pulmonary hypertension as a risk factor for death in patients with sickle cell disease. N. Engl. J. Med.350(9), 886–895 (2004).
  • Taylor JG 6th, Ackah D, Cobb C et al. Mutations and polymorphisms in hemoglobin genes and the risk of pulmonary hypertension and death in sickle cell disease. Am. J. Hematol.83(1), 6–14 (2008).
  • Roberts KE, McElroy JJ, Wong WP et al. BMPR2 mutations in pulmonary arterial hypertension with congenital heart disease. Eur. Respir. J.24, 371–374 (2004).
  • Trembath RC, Thomson JR, Machado RD et al. Clinical and molecular genetic features of pulmonary hypertension in patients with hereditary hemorrhagic telangiectasia. N. Engl. J. Med.345, 325–334 (2001).
  • Ashley-Koch AE, Elliott L, Kail ME et al. Identification of genetic polymorphisms associated with risk for pulmonary hypertension in sickle cell disease. Blood111(12), 5721–5726 (2008).
  • Klings ES, Dworkis DA, Sedgewick A et al. Genetic polymorphisms in NEDD4L are associated with pulmonary hypertension of sickle cell anemia. Blood114, 2562 (2009).
  • Nouraie M, Reading NS, Campbell A et al. Cytochrome b5 reductase T116S mutation and hemolysis in sickle cell disease. Blood114, 903 (2009)
  • Fowler Jr JE, Koshy M, Strub M, Chinn SK. Priapism associated with the sickle cell hemoglobinopathies: prevalence, natural history and sequelae. J. Urol.145, 65–68 (1991).
  • Hamre MR, Harmon EP, Kirkpatrick DV, Stern MJ, Humbert JR. Priapism as a complication of sickle cell disease. J. Urol.145, 1–5 (1991).
  • Elliott L, Ashley-Koch AE, De Castro L et al. Genetic polymorphisms associated with priapism in sickle cell disease. Br. J. Haematol.137(3), 262–267 (2007).
  • Nolan VG, Baldwin C, Ma Q et al. Association of single nucleotide polymorphisms in klotho with priapism in sickle cell anaemia. Br. J. Haematol.128(2), 266–272 (2005).
  • Brown CB, Boyer AS, Runyan RB, Barnett JV. Requirement of type III TGF-β receptor for endocardial cell transformation in the heart. Science283, 2080–2082 (1999).
  • Blanc L, Liu J, Vidal M, Chasis JA, An X, Mohandas N. The water channel aquaporin-1 partitions into exosomes during reticulocyte maturation: implication for the regulation of cell volume. Blood114(18), 3928–3934 (2009).
  • Endeward V, Musa-Aziz R, Cooper GJ et al. Evidence that aquaporin 1 is a major pathway for CO2 transport across the human erythrocyte membrane. FASEB J.20(12), 1974–1981 (2006).
  • Friedlander M, Brooks PC, Shaffer RW, Kincaid CM, Varner JA, Cheresh DA. Definition of two angiogenic pathways by distinct αv integrins. Science270(5241), 1500–1502 (1995).
  • Pruissen DM, Slooter AJ, Rosendaal FR, van der Graaf Y, Algra A. Coagulation factor XIII gene variation, oral contraceptives, and risk of ischemic stroke. Blood111(3), 1282–1286 (2008).
  • Jeng MR, Adams-Graves P, Howard TA, Whorton MR, Li CS, Ware RE. Identification of hemochromatosis gene polymorphisms in chronically transfused patients with sickle cell disease. Am. J. Hematol.74(4), 243–248 (2003).
  • Aygun B, Padmanabhan S, Paley C, Chandrasekaran V. Clinical significance of RBC alloantibodies and autoantibodies in sickle cell patients who received transfusions. Transfusion42(1), 37–43 (2002).
  • Tatari-Calderone Z, Minniti CP, Kratovil T et al. rs660 polymorphism in Ro52 (SSA1; TRIM21) is a marker for age-dependent tolerance induction and efficiency of alloimmunization in sickle cell disease. Mol. Immunol.47(1), 64–70 (2009).
  • Sebastiani P, Solovieff N, Hartley SW et al. Genetic modifiers of the severity of sickle cell anemia identified through a genome-wide association study. Am. J. Hematol.85(1), 29–35 (2010).

Website

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.