Skip to main content
Advertisement
  • Loading metrics

Meta-Analysis of 28,141 Individuals Identifies Common Variants within Five New Loci That Influence Uric Acid Concentrations

  • Melanie Kolz ,

    Contributed equally to this work with: Melanie Kolz, Toby Johnson

    Affiliation Institute of Epidemiology, Helmholtz Zentrum München, National Research Center for Environment and Health, Neuherberg, Germany

  • Toby Johnson ,

    Contributed equally to this work with: Melanie Kolz, Toby Johnson

    Affiliations Department of Medical Genetics, University of Lausanne, Lausanne, Switzerland, Swiss Institute of Bioinformatics, Lausanne, Switzerland, University Institute of Social and Preventive Medicine, Centre Hospitalier Universitaire Vaudois and University of Lausanne, Switzerland

  • Serena Sanna ,

    These authors also contributed equally to this work.

    Affiliation Istituto di Neurogenetica e Neurofarmacologia, Cagliari, Italy

  • Alexander Teumer ,

    These authors also contributed equally to this work.

    Affiliation Interfaculty Institute for Genetics and Functional Genomics, Ernst-Moritz-Arndt-Universität Greifswald, Greifswald, Germany

  • Veronique Vitart ,

    These authors also contributed equally to this work.

    Affiliation MRC Human Genetics Unit, IGMM, Western General Hospital, Crewe Road, Edinburgh, United Kingdom

  • Markus Perola ,

    These authors also contributed equally to this work.

    Affiliations Department of Chronic Disease Prevention, Institute of Health and Welfare, Helsinki, Finland, FIMM, Institute of Molecular Medicine, Helsinki, Finland

  • Massimo Mangino ,

    These authors also contributed equally to this work.

    Affiliation DTR Department of Twin Research and Genetic Epidemiology, King's College London, London, United Kingdom

  • Eva Albrecht ,

    These authors also contributed equally to this work.

    Affiliation Institute of Epidemiology, Helmholtz Zentrum München, National Research Center for Environment and Health, Neuherberg, Germany

  • Chris Wallace ,

    These authors also contributed equally to this work.

    Affiliation The Diabetes Inflammation Laboratory, Cambridge Institute of Medical Research, Cambridge University, Cambridge, United Kingdom

  • Martin Farrall ,

    These authors also contributed equally to this work.

    Affiliations Department of Cardiovascular Medicine, University of Oxford, Oxford, United Kingdom, The Wellcome Trust Centre for Human Genetics, University of Oxford, Oxford, United Kingdom

  • Åsa Johansson ,

    These authors also contributed equally to this work.

    Affiliation Department of Genetics and Pathology, Rudbeck Laboratory, Uppsala University, Uppsala, Sweden

  • Dale R. Nyholt ,

    These authors also contributed equally to this work.

    Affiliation Genetic Epidemiology Unit, Queensland Institute of Medical Research, Brisbane, Australia

  • Yurii Aulchenko,

    Affiliation Department of Epidemiology, Erasmus MC Rotterdam, Rotterdam, The Netherlands

  • Jacques S. Beckmann,

    Affiliations Department of Medical Genetics, University of Lausanne, Lausanne, Switzerland, Service of Medical Genetics, Centre Hospitalier Universitaire Vaudois, Lausanne, Switzerland

  • Sven Bergmann,

    Affiliations Department of Medical Genetics, University of Lausanne, Lausanne, Switzerland, Swiss Institute of Bioinformatics, Lausanne, Switzerland

  • Murielle Bochud,

    Affiliation University Institute of Social and Preventive Medicine, Centre Hospitalier Universitaire Vaudois and University of Lausanne, Switzerland

  • Morris Brown,

    Affiliation Clinical Pharmacology Unit, Addenbrookes Hospital, University of Cambridge, Cambridge, United Kingdom

  • Harry Campbell,

    Affiliation Centre for Population Health Sciences, University of Edinburgh, Edinburgh, United Kingdom

  • for the EUROSPAN Consortium,
  • John Connell,

    Affiliation Glasgow Cardiovascular Research Centre, University of Glasgow, Glasgow, United Kingdom

  • Anna Dominiczak,

    Affiliation Glasgow Cardiovascular Research Centre, University of Glasgow, Glasgow, United Kingdom

  • Georg Homuth,

    Affiliation Interfaculty Institute for Genetics and Functional Genomics, Ernst-Moritz-Arndt-Universität Greifswald, Greifswald, Germany

  • Claudia Lamina,

    Affiliations Institute of Epidemiology, Helmholtz Zentrum München, National Research Center for Environment and Health, Neuherberg, Germany, Division of Genetic Epidemiology, Department of Medical Genetics, Molecular and Clinical Pharmacology, Innsbruck Medical University, Innsbruck, Austria

  • Mark I. McCarthy,

    Affiliations The Wellcome Trust Centre for Human Genetics, University of Oxford, Oxford, United Kingdom, Oxford Centre for Diabetes, Endocrinology and Metabolism, University of Oxford, Churchill Hospital, Oxford, United Kingdom, National Institute for Health Research, Oxford Biomedical Research Centre, University of Oxford, Headington, Oxford, United Kingdom

  • for the ENGAGE Consortium,
  • Thomas Meitinger,

    Affiliation Institute of Human Genetics, Helmholtz Zentrum München, National Research Center for Environment and Health, Neuherberg, Germany

  • Vincent Mooser,

    Affiliation Genetics Division, GlaxoSmithKline, King of Prussia, Pennsylvania, United States of America

  • Patricia Munroe,

    Affiliation The William Harvey Research Institute, Barts and The London School of Medicine and Dentistry, Queen Mary University of London, London, England

  • Matthias Nauck,

    Affiliation Institut für Klinische Chemie und Laboratoriumsmedizin, Ernst-Moritz-Arndt-Universität Greifswald, Greifswald, Germany

  • John Peden,

    Affiliations Department of Cardiovascular Medicine, University of Oxford, Oxford, United Kingdom, The Wellcome Trust Centre for Human Genetics, University of Oxford, Oxford, United Kingdom

  • Holger Prokisch,

    Affiliation Institute of Human Genetics, Helmholtz Zentrum München, National Research Center for Environment and Health, Neuherberg, Germany

  • Perttu Salo,

    Affiliations Department of Chronic Disease Prevention, Institute of Health and Welfare, Helsinki, Finland, FIMM, Institute of Molecular Medicine, Helsinki, Finland

  • Veikko Salomaa,

    Affiliation Department of Chronic Disease Prevention, Institute of Health and Welfare, Helsinki, Finland

  • Nilesh J. Samani,

    Affiliation Cardiovascular Sciences, University of Leicester, Glenfield Hospital, Leicester, United Kingdom

  • David Schlessinger,

    Affiliation Laboratory of Genetics, Intramural Research Program, National Institute on Aging, National Institutes of Health, Baltimore, Maryland, United States of America

  • Manuela Uda,

    Affiliation Istituto di Neurogenetica e Neurofarmacologia, Cagliari, Italy

  • Uwe Völker,

    Affiliation Interfaculty Institute for Genetics and Functional Genomics, Ernst-Moritz-Arndt-Universität Greifswald, Greifswald, Germany

  • Gérard Waeber,

    Affiliation Department of Medicine and Internal Medicine, Centre Hospitalier Universitaire Vaudois, Lausanne, Switzerland

  • Dawn Waterworth,

    Affiliation Genetics Division, GlaxoSmithKline, King of Prussia, Pennsylvania, United States of America

  • Rui Wang-Sattler,

    Affiliation Institute of Epidemiology, Helmholtz Zentrum München, National Research Center for Environment and Health, Neuherberg, Germany

  • Alan F. Wright,

    Affiliation MRC Human Genetics Unit, IGMM, Western General Hospital, Crewe Road, Edinburgh, United Kingdom

  • Jerzy Adamski,

    Affiliation Genome Analysis Centre, Institute for Experimental Genetics, Helmholtz Zentrum München, National Research Center for Environment and Health, Neuherberg, Germany

  • John B. Whitfield,

    Affiliation Genetic Epidemiology Unit, Queensland Institute of Medical Research, Brisbane, Australia

  • Ulf Gyllensten,

    Affiliation Department of Genetics and Pathology, Rudbeck Laboratory, Uppsala University, Uppsala, Sweden

  • James F. Wilson,

    Affiliation Centre for Population Health Sciences, University of Edinburgh, Edinburgh, United Kingdom

  • Igor Rudan,

    Affiliations Centre for Population Health Sciences, University of Edinburgh, Edinburgh, United Kingdom, Croatian Centre for Global Health, Faculty of Medicine, University of Split, Split, Croatia, Gen-Info, Zagreb, Croatia

  • Peter Pramstaller,

    Affiliations Institute of Genetic Medicine, European Academy Bozen/Bolzano (EURAC), Bolzano, Italy, Affiliated Institute of the University Lübeck, Department of Neurology, Central Hospital, Bolzano, Italy

  • Hugh Watkins,

    Affiliations Department of Cardiovascular Medicine, University of Oxford, Oxford, United Kingdom, The Wellcome Trust Centre for Human Genetics, University of Oxford, Oxford, United Kingdom

  • for the PROCARDIS Consortium,
  • Angela Doering,

    Affiliation Institute of Epidemiology, Helmholtz Zentrum München, National Research Center for Environment and Health, Neuherberg, Germany

  • H.-Erich Wichmann,

    Affiliations Institute of Epidemiology, Helmholtz Zentrum München, National Research Center for Environment and Health, Neuherberg, Germany, IBE, Chair of Epidemiology, University of Munich, Germany

  • for the KORA Study,
  • Tim D. Spector,

    Affiliation DTR Department of Twin Research and Genetic Epidemiology, King's College London, London, United Kingdom

  • Leena Peltonen,

    Affiliations FIMM, Institute of Molecular Medicine, Helsinki, Finland, The Wellcome Trust Sanger Institute, Wellcome Trust Genome Campus, Hinxton, Cambridge, United Kingdom

  • Henry Völzke,

    Affiliation Institute for Community Medicine, Ernst-Moritz-Arndt-Universität Greifswald, Greifswald, Germany

  • Ramaiah Nagaraja,

    Affiliation Laboratory of Genetics, Intramural Research Program, National Institute on Aging, National Institutes of Health, Baltimore, Maryland, United States of America

  • Peter Vollenweider,

    Affiliation Department of Medicine and Internal Medicine, Centre Hospitalier Universitaire Vaudois, Lausanne, Switzerland

  • Mark Caulfield,

    Affiliation The William Harvey Research Institute, Barts and The London School of Medicine and Dentistry, Queen Mary University of London, London, England

  • for the WTCCC,
  • Thomas Illig,

    Affiliation Institute of Epidemiology, Helmholtz Zentrum München, National Research Center for Environment and Health, Neuherberg, Germany

  •  [ ... ],
  • Christian Gieger

    christian.gieger@helmholtz-muenchen.de

    Affiliation Institute of Epidemiology, Helmholtz Zentrum München, National Research Center for Environment and Health, Neuherberg, Germany

  • [ view all ]
  • [ view less ]

Abstract

Elevated serum uric acid levels cause gout and are a risk factor for cardiovascular disease and diabetes. To investigate the polygenetic basis of serum uric acid levels, we conducted a meta-analysis of genome-wide association scans from 14 studies totalling 28,141 participants of European descent, resulting in identification of 954 SNPs distributed across nine loci that exceeded the threshold of genome-wide significance, five of which are novel. Overall, the common variants associated with serum uric acid levels fall in the following nine regions: SLC2A9 (p = 5.2×10−201), ABCG2 (p = 3.1×10−26), SLC17A1 (p = 3.0×10−14), SLC22A11 (p = 6.7×10−14), SLC22A12 (p = 2.0×10−9), SLC16A9 (p = 1.1×10−8), GCKR (p = 1.4×10−9), LRRC16A (p = 8.5×10−9), and near PDZK1 (p = 2.7×10−9). Identified variants were analyzed for gender differences. We found that the minor allele for rs734553 in SLC2A9 has greater influence in lowering uric acid levels in women and the minor allele of rs2231142 in ABCG2 elevates uric acid levels more strongly in men compared to women. To further characterize the identified variants, we analyzed their association with a panel of metabolites. rs12356193 within SLC16A9 was associated with DL-carnitine (p = 4.0×10−26) and propionyl-L-carnitine (p = 5.0×10−8) concentrations, which in turn were associated with serum UA levels (p = 1.4×10−57 and p = 8.1×10−54, respectively), forming a triangle between SNP, metabolites, and UA levels. Taken together, these associations highlight additional pathways that are important in the regulation of serum uric acid levels and point toward novel potential targets for pharmacological intervention to prevent or treat hyperuricemia. In addition, these findings strongly support the hypothesis that transport proteins are key in regulating serum uric acid levels.

Author Summary

Elevated serum uric acid levels cause gout and are a risk factor for cardiovascular disease and diabetes. The regulation of serum uric acid levels is under a strong genetic control. This study describes the first meta-analysis of genome-wide association scans from 14 studies totalling 28,141 participants of European descent. We show that common DNA variants at nine different loci are associated with uric acid concentrations, five of which are novel. These variants are located within the genes coding for organic anion transporter 4 (SLC22A11), monocarboxylic acid transporter 9 (SLC16A9), glucokinase regulatory protein (GCKR), Carmil (LRRC16A), and near PDZ domain-containing 1 (PDZK1). Gender-specific effects are shown for variants within the recently identified genes coding for glucose transporter 9 (SLC2A9) and the ATP-binding cassette transporter (ABCG2). Based on screening of 163 metabolites, we show an association of one of the identified variants within SLC16A9 with DL-carnitine and propionyl-L-carnitine. Moreover, DL-carnitine and propionyl-L-carnitine were strongly correlated with serum UA levels, forming a triangle between SNP, metabolites and UA levels. Taken together, these associations highlight pathways that are important in the regulation of serum uric acid levels and point toward novel potential targets for pharmacological intervention to prevent or treat hyperuricemia.

Introduction

Uric acid (UA) is the final catabolic, heterocyclic purine derivative resulting from the oxidation of purines in humans. Due to the loss of hepatic uricase activity during human evolution, UA is excreted as such and is not further metabolized into carbon dioxide and ammonia. A major mechanism underlying hyperuricemia is impaired renal excretion of urate. Most notably, UA is causally involved in the pathogenesis of gouty arthritis that results from deposition of monosodium urate crystals in the joints [1]. Increased UA concentrations have been implicated in cardiovascular disease for more than five decades [2]. In addition, elevated urate is associated with obesity, blood pressure and insulin resistance, and consequently with the metabolic syndrome and type 2 diabetes [2],[3]. However, UA also has a positive role as an antioxidant, and is correlated with longevity in mammals [4]. Thus, human physiology is especially sensitive to the precise range of UA levels.

Besides environmental factors, there is evidence for a strong genetic influence upon serum UA concentrations, with heritability estimates of up to 73% [5]. Recently, genome-wide association (GWA) studies have identified single nucleotide polymorphisms (SNPs) in the SLC2A9 gene (solute carrier family 2, member 9 gene), a putative glucose transporter, which are strongly associated with serum UA concentrations and gout [6][9]. This novel gene locus functions as a high-capacity urate transporter in humans [8],[10]. This emphasises the power of GWA studies in expanding our understanding at the molecular level of disease mechanisms and in pointing to innovative therapeutic strategies.

The power of GWA studies to detect common variants with modest effects directly depends on the size of the study group. Therefore, the present study sought to detect novel genetic variants related to serum UA levels by conducting a meta-analysis of GWA findings from 14 studies (BRIGHT, CoLaus, CROATIA, Health 2000, KORA F3, KORA S4, ORCADES, PROCARDIS, NSPHS, SardiNIA, SHIP, SSAGA, MICROS, and TwinsUK) totalling 28,141 participants. In addition, the meta-analysis was performed independently on sex specific GWA results to address the pronounced gender differences in the regulation of UA concentrations that have previously been reported [1],[6]. Identified variants were further analyzed for association with metabolite profiles.

Results

The sample size and participant characteristics for each participating study are shown in Table S1. Meta-analysis of GWA data of 28,141 individuals of European ancestry yielded 954 SNPs (full list is provided in Table S4) that exceeded the genome-wide significance threshold of 5×10−8 (Figure 1A).

thumbnail
Figure 1. Genome-wide association results.

Manhattan plots showing significance of association of all SNPs in the meta-analysis for (A) men and women combined, (B) men only, and (C) women only. SNPs are plotted on the x-axis according to their position on each chromosome against association with uric acid concentrations on the y-axis (shown as −log10 p-value).

https://doi.org/10.1371/journal.pgen.1000504.g001

Those SNPs cluster around nine loci (Table 1), four of which are well known regulators of serum UA levels: SLC2A9 (p = 5.2×10−201), ABCG2 (p = 3.1×10−26), SLC17A1 (p = 3.0×10−14), and SLC22A12 (p = 2.0×10−9). The first, SLC2A9, was identified in previous GWA scans (Figure 2C) [6][9]. A total of 788 SNPs reached the genome-wide significance threshold at the SLC2A9 locus. The strongest associated marker was the intronic SNP rs734553 (p = 5.2×10−201, Table 1), which is in high linkage disequilibrium (r2 = 0.88) with the missense mutation rs16890979 previously described [11]. The second locus was on chromosome 4q22, harbouring the ABCG2 gene (Figure 2D). In accordance with previous results, the strongest observed association was at rs2231142 (p = 3.1×10−26, Table 1), a coding SNP leading to a glutamine-to-lysine amino acid change at position 141 [11]. The third previously implicated locus influencing UA levels was on chromosome 6p23-p21.3, which contains the SLC17A3 gene (Figure 2F) [11]. The top associated marker was SNP rs1183201 (p = 3.0×10−14, Table 1), intronic of SLC17A1, but the association signal encompassed a larger region including the SLC17A1, SLC17A3, SLC17A4 genes and downstream to HIST1H4C, in agreement with the linkage disequilibrium at this locus. SNP rs1183201 is in high linkage disequilibrium (r2 = 0.97) with rs1165205, a SNP intronic of SLC17A3 gene identified by a previous GWA scan [11].

thumbnail
Figure 2. Regional association plots of nine urate loci.

P-value plots showing the association signals in the region of (A) PDZK1 on chromosome 1, (B) GCKR on chromosome 2, (C) SLC2A9 on chromosome 4, (D) ABCG2 on chromosome 4, (E) LRRC16A on chromosome 6, (F) SLC17A1 on chromosome 6, (G) SLC16A9 on chromosome 10, (H) SLC22A11 on chromosome 11, and (I) SLC22A12 on chromosome 11. −log10 p-values are plotted as a function of genomic position (NCBI Build 36). Large diamonds in red indicate the most significant SNP in the region while other SNPs in the region are given as colour-coded smaller diamonds. Red diamonds indicate high correlation with the lead SNP (r2>0.8), orange diamonds indicate moderate correlation with the most significant SNP (0.5<r2<0.8), yellow indicates markers in weak correlation with the most significant SNP (0.2<r2<0.5), white indicates no correlation with the most significant SNP (r2<0.2). Estimated recombination rates (HapMap Phase II) are given in light blue, genes as well as the direction of transcription (NCBI) are displayed by green bars.

https://doi.org/10.1371/journal.pgen.1000504.g002

thumbnail
Table 1. Nine loci associated with uric acid concentrations.

https://doi.org/10.1371/journal.pgen.1000504.t001

Among the novel loci identified, the strongest was on chromosome 11q13. One locus was localized upstream and within the SLC22A11 gene, and represented by SNP rs17300741 (p = 6.7×10−14, Table 1, Figure 2H). The second signal was SNP rs505802 (p = 2.0×10−9), representative of all associated markers falling within and downstream the extensively studied SLC22A12 gene coding for URAT1 (Figure 2I). The p-value plot as well as the LD block structure (r2 = 0.09) suggested two nearby but independently associated regions, which was verified in multiple regression analysis (Table S5).

The second novel region was on chromosome 2p23.3-p23.2 (Figure 2B). The most significant SNP in this region was SNP rs780094 (p = 1.4×10−9), intronic of GCKR, a glucokinase regulator protein recently associated with several quantitative traits including the regulation of triglycerides levels [12]. We also identified genome-wide significant association on chromosome 1q21 (Figure 2A). The top ranking SNP in this region was rs12129861 (p = 2.7×10−9, Table 1), located 2 kb upstream of PDZK1 coding for PDZ domain-containing 1 reported to interact with URAT1 [13]. The fourth newly detected region was found on chromosome 6p22.2 (Figure 2E), where the association signal spans two genes, LRRC16A and SCGN, within one highly preserved LD block. The strongest p-value was observed for the SNP rs742132, located within an intron of LRRC16A (p = 8.5×10−9, Table 1). Independence of the LRRC16A and the SLC17A1 loci (r2 = 0.07) was verified in multiple regression analysis. P-values and effect estimates only slightly changed between single SNP analysis and multiple regression analysis (Table S5). Finally, we also observed some evidence of association on chromosome 10q21.3 (Figure 2G). One SNP within SLC16A9, rs12356193, reached genome-wide significance (p = 1.1×10−8). However, there were several additional SNPs within this gene with borderline significance, supporting the hypothesis that this locus may be a true signal rather than a false positive result.

Sex-Stratified Meta-Analysis Identifies Male and Female Specific Variants

We have also performed a meta-analysis of sex specific GWA results using all 14 studies (12,328 males, 15,813 females). Although the results did not show any additional genome-wide significant locus (Figure 1B and 1C), we were able to query which of the aforementioned SNPs have sex-specific effects on serum UA levels (Table 2). For the SLC2A9 gene, we found that in males the top ranking SNP was still rs734553 (p = 1.1×10−41), while for women it was the nearby intronic SNP rs12498742 (p = 2.4×10−196). Supporting previously reported results, we found for both markers that the effect size of the minor allele observed in women was twice the effect size observed in men (p<3.8×10−17, Table 2) [6]. The minor allele of rs2231142 in the ABCG2 gene showed a greater effect size in men compared to women (p = 0.01, Table 2). Similar, the effect size of the most significant SNP for males in the ABCG2 gene locus, rs2199936, was greater in men compared to women (p = 0.008, Table 2). The effect sizes of the other SNPs were comparable in men and women (Table 2).

thumbnail
Table 2. Gender specific association results at the nine loci.

https://doi.org/10.1371/journal.pgen.1000504.t002

Association of Identified Variants with Metabolite Profiles

To further characterize the identified variants, we analyzed their association with a panel of 163 metabolites measured in the KORA F4 survey. After correction for multiple testing, one SNP within SLC16A9, rs12356193, was associated with DL-carnitine concentrations (β = −3.58, p = 4.0×10−26), which in turn were associated with serum UA levels (β = 0.06, p = 1.4×10−57). In addition, this SNP was associated with propionyl-L-carnitine (β = −0.06, p = 5.0×10−8). Similar to DL-carnitine, propionyl-L-carnitine concentrations were also strongly associated with serum UA levels (β = 1.78, p = 8.1×10−54), forming a triangle between SNP, metabolites and UA levels. None of the other SNPs were significantly associated with the measured metabolites.

Discussion

Based on meta-analysis of GWA studies including 28,141 individuals, we have mapped 5 novel loci and confirmed 4 previously implicated loci that influence serum UA levels. Altogether, these associations highlight biological pathways that are important in the regulation of urate concentrations and may point to novel targets for pharmacological interventions to prevent or treat hyperuricemia.

A genome-wide significant p-value was observed for one SNP within SLC16A9 gene locus, coding for monocarboxylic acid transporter 9 (MCT9). This is a member of the monocarboxylate co-transporter family that has been demonstrated to catalyze transport of monocarboxylic acids across cell membranes [14]. MCT9 is expressed in various tissues including the kidney [15]. As other sodium monocarboxylate transporters have been found to influence urate in knockout models this MCT9 isoform might be a sodium-dependent transporter in the kidney. The second newly identified locus was GCKR (glucokinase regulatory protein) a regulator of glucokinase, the first glycolytic enzyme which serves as a glucose sensor, responsible for glucose phosphorylation in the liver. Recently, GWA studies for type 2 diabetes identified the same rs780094 SNP as a potential marker for modulation of triglyceride levels [16]. Meanwhile, GCKR polymorphisms were also shown to be associated with metabolic traits like fasting glucose and, modestly, type 2 diabetes [12],[17],[18]. Several potential mechanisms have been proposed to link serum UA concentrations with metabolic traits. Exogenous insulin decreases renal sodium and urate excretion [19]. Furthermore, renal clearance of UA is inversely related to the degree of insulin resistance [20]. Finally, insulin resistance is thought to be accompanied by impaired oxidative phosphorylation in hepatic mitochondria, leading to increased concentrations of co-enzyme A esters and thus to increased systemic adenosine concentrations [21]. Increased adenosine, in turn, may result in renal retention of sodium, urate, and water [21],[22]. This provides a putative mechanism for hyperuricaemia via both the break down of adenosine to urate and increased renal urate retention [21],[22].

We also found evidence for association in a region containing two genes, LRRC16A and SCGN. The strongest association was located within LRRC16A coding for CARMIL. This large protein is most abundant in kidney and epithelial tissues and serves as an inhibitor of the heterodimeric actin capping protein (CP), an essential element of the actin cytoskeleton which binds to the barbed ends of actin filaments and regulates their polymerization [23]. The multiple biochemical functions associated with CARMIL raise many possibilities for its mechanism of action in cells, but the relation of CARMIL to UA concentration is thus far unclear. The nearby SCGN is coding for Secretagogin, a calcium-binding protein selectively expressed in neuroendocrine tissue and pancreatic beta-cells. The function of Secretagogin is unknown, but it has been suggested to influence calcium influx and insulin secretion [24].

We also demonstrated association of SNPs in SLC22A11 and SLC22A12 with UA concentrations. SLC22A12 encodes the extensively studied URAT1, a member of the organic anion transporter (OAT) family [25]. URAT1, a well known candidate gene for UA accumulation/transport, mediates the non-voltage-dependent exchange of urate for several organic anions [1]. SLC22A11 codes for OAT4, an OAT isoform which, like URAT1, is localized at the apical membrane of the proximal tubules. OAT4 serves as an organic anion–dicarboxylate exchanger, which mediates urate transport across the apical membrane of kidney [26],[27]. In combination with these findings, we also identified genome-wide significant association of SNPs in and upstream of PDZK1, coding for PDZ domain containing 1, a scaffolding protein reported to interact with OAT4, URAT1 and NTP1 (SLC17A1) via their C-terminal PDZ motifs [13],[28]. It has been proposed that the PDZ scaffold may form a bidirectional transport system by linking URAT1 (reabsorption) and NPT1 (secretion) leading to a functional complex responsible for the balanced urate transport regulation at the apical membrane of renal proximal tubules [1],[28].

In accordance with previous genome-wide studies, the strongest effect on serum UA concentrations was detected for SLC2A9, [6][9] coding for GLUT9, which has been shown to be strongly associated with hyperuricemia and gout and to serve as a high-capacity urate transporter in humans [8],[10]. Additional confirmed loci include ABCG2 and SLC17A1 [11]. ABCG2 is a member of the ATP-binding cassette (ABC) superfamily of membrane transporters, while the SLC17A1 locus, located directly downstream of the recently identified SLC17A3 locus (NPT4), encodes NPT1 (renal sodium phosphate transport protein 1). The human NPT1 is localized at the apical membrane of renal proximal tubules and serves as a voltage-driven UA transporter in model systems [28].

Although several of the SNPs associated with uric acid concentrations in this meta-analysis are located within genes that are plausible candidates for influencing uric acid concentrations, our association approach is not able to identify underlying genes or mechanisms in the regions of association signals. Therefore, other genes might be responsible for the observed associations and functional studies are warranted to identify the causal variants and provide insights in the underlying biological mechanisms.

Pronounced gender differences in the regulation of serum UA concentrations have been reported for both humans and animals [1],[6]. In line with our previous findings, [6] the strongest gender-specific effect was observed for the minor allele of rs734553 (SLC2A9), resulting in a 2-fold larger effect size on serum UA concentrations in women compared to men. For ABCG2, the effect of the minor allele of rs2231142 demonstrated a larger effect on UA concentrations in men compared to women. For the other loci, effect sizes did not significantly differ by gender.

The rapidly evolving field of metabolomics aims at a comprehensive measurement of endogenous metabolites in a cell or body fluid [29]. Based on screening of 163 metabolites, we have observed an association of one of the identified variants, rs12356193 within SLC16A9, with DL-carnitine and propionyl-L-carnitine. Moreover, DL-carnitine and propionyl-L-carnitine were strongly correlated with serum UA levels, forming a triangle between SNP, metabolites and UA levels. Carnitine is acquired from diet and endogenous biosynthesis. Its primary function is in the transport of long chain fatty acids. After strenuous physical exercise, both acylcarnitine and UA levels increase in the serum of healthy humans [30]. In spontaneously hypertensive rats, L-carnitine decreases serum UA levels and the age-dependent rise in serum UA [31],[32]. Kidneys absorb 95% of carnitine from the glomerular filtrate via an active Na+-dependent transport mechanism [33]. Impairment of this reabsorptive function can lead to carnitine deficiency, in which hyperuricemia may be present because carnitine competes for renal tubular excretion [34]. Although experimental data are few, currently available data suggest that urinary acylcarnitine, which reflects the balance between dietary intake of carnitine and renal excretion, may be linked to serum UA via oxidative stress pathways [35]. Given that palmitoyl carnitine inhibits binding of Ca2+ channel ligands to rat brain cortical membranes and to inhibit voltage-activated Ca2+ channel currents, acylcarnitines may also have direct influences on MCT9 [36].

Overall, serum UA concentrations are mainly determined by the balance between urate production and renal excretion. We have identified nine loci that are associated with serum UA levels and six of them harbor genes that code for renal transport proteins. Most notably, five of these transport proteins belong to the family–and moreover, to one phylogenetic cluster within this family [37]. These findings strongly support the hypothesis that genetic variation in urate transport proteins are the key influences upon regulation of serum UA levels in humans.

Materials and Methods

Study Participants

The present meta-analysis combined data from 14 GWA scans: British Genetics of Hypertension (BRIGHT), Cohorte Lausannoise (CoLaus), Vis Island Isolate Study (CROATIA), Health 2000 cohort (Health 2000), two surveys of the Cooperative Health Research in the Region of Augsburg (KORA F3, KORA S4), Orkney Complex Disease Study (ORCADES), Precocious Coronary Artery Disease (PROCARDIS), Northern Swedish Population Health Study (NSPHS), SardiNIA Study of Aging (SardiNIA), Study of Health in Pomerania (SHIP), Semi-Structured Assessment for Genetics of Alcoholism (SSAGA), Microisolates in South Tyrol (MICROS), and UK Adult Twin Register (TwinsUK). Altogether, the meta-analysis comprises 28,141 individuals (12,328 males, 15,813 females) of European ancestry with measured serum UA concentrations (Table S1). Approval was obtained by local ethic committees for all studies and informed consent was given from the study participants. A detailed individual description of study designs is provided in Text S1.

Genome-Wide Genotyping and Imputation

Six different platforms/arrays were used for genotyping: the Affymetrix 500 K GeneChip array (4 cohorts, n = 13,103), the Affymetrix 6.0 GeneChip array (2 cohorts, n = 5,901), Illumina HumanHap 300 (5 cohorts, n = 3,609), Illumina Human 610 K Beadchip (1 cohort, n = 2,212), Illumina HumanHap 300-Duo (1 cohort, n = 2,113), and Illumina Human 1 M beadchip (1 cohort, n = 1,203). Imputation of allele dosage of SNPs typed in the HapMap CEU population was performed using either MACH [38] or IMPUTE [39] with parameters and pre-imputation filters as specified in Table S2. All SNPs with a minor allele frequency <0.01 were excluded from analysis. SNPs were also excluded if the cohort-specific imputation quality as assessed by r2.hat (MACH) or .info (IMPUTE) metrics was <0.30 or <0.40, respectively. In total, up to 2,493,963 genotyped or imputed autosomal SNPs were analyzed.

Uric Acid Measurements

Non-fasting blood samples were obtained from study participants of BRIGHT, KORA, NSPHS, SardiNIA, SHIP and SSAGA and fasting samples from those of CoLaus, PROCARDIS, CROATIA, Health 2000, MICROS, ORCADES and TwinsUK. UA analyses were carried out on fresh samples in all studies except from BRIGHT, NSPHS, CROATIA, MICROS and SSAGA, where frozen serum was used that was stored at −20°C (BRIGHT) or −70°C (NSPHS, SSAGA, CROATIA, MICROS). UA concentrations were measured using an uricase/peroxidase method (CROATIA, MICROS, NSPHS and ORCADES: DVIA1650-Autoanalyzer, Siemens Healthcare Diagnostics) or an uricase method (BRIGHT: Hitachi, Roche Diagnostics; CoLaus: uricase PGP, Roche Diagnostics; Health 2000: Thermo Fisher Scientific, Vantaa; KORA F3: URCA Flex, Dade Behring; KORA S4: UA Plus, Roche; PROCARDIS: Hitachi 917, Roche Diagnostics; SardiNIA: Bayer; SHIP: UA PAP, Boehringer; SSAGA: Hitachi 747, Boehringer; TwinsUK: Ektachem/Vitros system, Johnson & Johnson Clinical Diagnostics).

Metabolite Measurements

Metabolomic analyses were conducted in 2020 randomly selected participants (ages 32–81 years) of the KORA F4 survey, a follow-up survey of KORA S4. Genotype information was available for 1814 of these participants. Fasting blood samples were collected in 2006–2008. Blood was drawn into serum gel tube in the morning between 8 and 10 am. The tube was gently inverted two times, followed by 30 minutes resting at room temperature to obtain complete coagulation, and finally centrifugation of blood was performed at 2750 g, 15°C for 10 minutes for serum collection. Serum was aliquoted and kept at 4°C for a maximum of 6 hours, after which it was frozen at −80°C until analyses. Liquid handling of serum samples (10 µl) was performed with Hamilton Star (Hamilton Bonaduz AG, Bonaduz, Switzerland) robot and prepared for quantification with AbsoluteIDQ kit (BIOCRATES Life Sciences AG, Innsbruck, Austria). Sample analyses were done on API4000 Q TRAP LC/MS/MS System (Applied Biosystems, Darmstadt, Germany) equipped with Schimadzu Prominence LC20AD pump and SIL-20AC auto sampler. The complete analytical process (e.g. the targeted metabolite concentration) was performed using the MetIQ software package, which is an integral part of the AbsoluteIDQ Kit. A total of 163 metabolites were measured. The metabolomics dataset contains 14 amino acids, one sugar, 41 acylcarnitines, 15 sphingolipids, and 92 glycerophospholipids.

Statistical Analysis

GWA scans were made using an additive genetic model on Z-scores, calculated by adjusting serum UA levels for age and sex using linear regression and standardizing residuals. In sex-specific association testing Z-scores were calculated in each stratum separately. Study-specific results of the most significant SNP at each locus are presented in Table S3. The results from all 14 GWA scans were combined into a fixed-effects meta-analysis with inverse variance weighting, using the METAL package (www.sph.umich.edu/csg/abecasis/metal). The individual studies were corrected for residual inflation of the test statistic using genomic control methods for genotyped and imputed SNPs combined [40]. For the overall meta-analysis, the inflation factor was 1.028, no further correction was applied. Quantile-quantile plots of the association results are shown in Figure S1, study-specific quantile-quantile plots are illustrated in Figure S2 and S3. Associations were considered genome-wide significant below p = 5×10−8, which corresponds to a Bonferroni correction for the estimated one million independent common variant tests in the human genome of European individuals [41]. We also tested whether the effect estimate resulting from the gender-specific fixed effect meta-analysis differed significantly between men and women by applying a t-test comparing effect and standard error estimates in men with the effect and standard error estimates in women. Genome-wide significant SNPs were tested for independent associations, by including all SNPs in a multiple regression model, and then performing inverse variance weighted meta-analysis, across all cohorts (except for Health 2000), of the coefficient for each SNP. The analysis of metabolites was performed using the same linear regression adjusted by sex and gender as in the genome-wide scan. To specify the dependency of uric acid on metabolite concentration, a univariate regression model without further transformation or adjustment was used. The multiple regression and metabolite analysis were performed using either posterior expected allele dosages, or on best-guess imputed genotypes, with the statistical analysis software R.

Accession Numbers

The OMIM (http://www.ncbi.nlm.nih.gov/OMIM) accession numbers for genes mentioned in this article are PDZK1 (603831), GCKR (600842), SLC2A9 (606142), ABCG2 (603756), LRRC16A (609593), SLC17A1 (182308), SLC22A11 (607097), and SLC22A12 (607096). The HGNC (http://www.gene.ucl.ac.uk) accession number for SLC16A9 is 23520.

Supporting Information

Figure S1.

Quantile-quantile plots of association results. Meta-analysis was performed using sample-size weighted Z-scores after cohort-specific genomic control. Shown are expected p-values plotted against observed p-values resulting from meta-analysis based on all subjects (1st row), only males (2nd row) and only females (3rd row) for all analysed SNPs (left column) and for all analysed SNPs excluding the SLC2A9 region (GLUT9, right column).

https://doi.org/10.1371/journal.pgen.1000504.s001

(1.25 MB TIF)

Figure S2.

Study-specific quantile-quantile plots. Shown are expected p-values plotted against observed p-values resulting from each single study before (black dots) and after (blue dots) genomic control correction. The study-specific λ-values were λ = 1.007 (BRIGHT), λ = 1.025 (CoLaus), λ = 1.013 (CROATIA), λ = 1.024 (Health 2000), λ = 1.006 (KORA F3), λ = 1.016 (KORA S4), λ = 1.246 (MICROS), λ = 1.253 (NSPHS), λ = 1.182 (ORCADES), λ = 1.022 (PROCARDIS), λ = 1.090 (SardiNIA), λ = 1.031 (SHIP), λ = 1.022 (SSAGA) and λ = 1.122 (TwinsUK). For the overall meta-analysis, the inflation factor was 1.028.

https://doi.org/10.1371/journal.pgen.1000504.s002

(1.95 MB TIF)

Figure S3.

Study-specific quantile-quantile plots excluding GLUT9. Shown are expected p-values plotted against observed p-values resulting from each single study before (black dots) and after (blue dots) genomic control correction, excluding SNPs in the SLC2A9 (GLUT9) region on chromosome 4 (positions 9194245 to 10270832).

https://doi.org/10.1371/journal.pgen.1000504.s003

(1.91 MB TIF)

Table S1.

Study sample characteristics. Characteristics are shown by study for British Genetics of Hypertension (BRIGHT), Cohorte Lausannoise (CoLaus), Vis island isolate study (CROATIA), Health 2000 cohort (Health 2000), two surveys of the Cooperative Health Research in the Region of Augsburg (KORA F3, KORA S4), Orkney Complex Disease Study (ORCADES), Precocious Coronary Artery Disease (PROCARDIS), Northern Swedish Population Health Study (NSPHS), SardiNIA Study of Aging (SardiNIA), Study of Health in Pomerania (SHIP), Semi-Structured Assessment for Genetics of Alcoholism (SSAGA), Microisolates in South Tyrol (MICROS) and UK Adult Twin Register (TwinsUK). Age is given as mean and range in brackets. Uric acid concentrations (UA) are given as mean and appropriate standard deviation (SD). NA indicates not applicable.

https://doi.org/10.1371/journal.pgen.1000504.s004

(0.06 MB DOC)

Table S2.

Genotyping, imputation and analysis procedures by study. Shown are the genotyping platforms, quality control (QC) filters applied before imputation, imputation software, number of SNPs and genotype-phenotype association software.

https://doi.org/10.1371/journal.pgen.1000504.s005

(0.07 MB DOC)

Table S3.

Study-specific results. Shown are study-specific results of the most significant SNP at each locus.

https://doi.org/10.1371/journal.pgen.1000504.s006

(0.23 MB DOC)

Table S4.

Full list of genome-wide significant SNPs. Shown is a full list of SNPs that exceeded the threshold of genome-wide significance (p<5×10−8). Position is given for NCBI Build 36. Effect estimates result from additive linear regression on Z-scores of uric acid concentrations. P-values have been calculated using weighting by the inverse variance. The effect allele is the allele to which the beta (effect) estimate refers.

https://doi.org/10.1371/journal.pgen.1000504.s007

(2.01 MB DOC)

Table S5.

Multiple regression analysis. Genome-wide significant SNPs were tested for independent associations, by including all nine SNPs in a multiple regression model, and then performing inverse variance weighted meta-analysis, across participating cohorts (except for Health2000), of the coefficient for each SNP.

https://doi.org/10.1371/journal.pgen.1000504.s008

(0.04 MB DOC)

Text S1.

Study design. This section describes additional study specific characteristics.

https://doi.org/10.1371/journal.pgen.1000504.s009

(0.14 MB DOC)

Acknowledgments

The authors would like to acknowledge those who agreed to participate in the contributing studies. The computations for CoLaus imputation were performed in part at the Vital-IT center for high performance computing of the Swiss Institute of Bioinformatics. The CoLaus authors thank Yolande Barreau, Mathieu Firmann, Vladimir Mayor, Anne-Lise Bastian, Binasa Ramic, Martine Moranville, Martine Baumer, Marcy Sagette, Jeanne Ecoffey, and Sylvie Mermoud for data collection. The CROATIA authors are grateful to Professor Pavao Rudan and the staff of the Institute for Anthropological Research in Zagreb, Croatia (organization of the field work, anthropometric and physiological measurements, and DNA extraction); Professor Ariana Vorko-Jovic and the staff and medical students of the Andrija Stampar School of Public Health of the Faculty of Medicine, University of Zagreb, Croatia (questionnaires, genealogial reconstruction and data entry). The CROATIA authors acknowledge the Wellcome Trust Clinical Research Facility (Edinburgh) for performing the genotyping. The KORA authors acknowledge the contribution of P. Lichtner, G. Eckstein, G. Fischer, L. Geistlinger, and N. Klopp and all other members of the Helmholtz Zentrum München genotyping staff for generating the SNP data. The KORA authors also want to thank C. Prehn and W. Römisch-Margl for support and advice, and A. Sabunchi, H. Chavez, B. Hochstrat and T. Halex for technical assistance during the metabolite analyses as well as all members of field staffs who were involved in the planning and conduct of the KORA Augsburg studies. The MICROS authors are grateful to the primary care practitioners Raffaela Stocker, Stefan Waldner, Toni Pizzecco, Josef Plangger, Ugo Marcadent, and the personnel of the Hospital of Silandro (Department of Laboratory Medicine) for their participation and collaboration in the research project. The NSPHS authors are grateful to the Uppsala Genome Center for performing the genotyping, the contribution of samples from the Medical Biobank in Umeå and the contribution of the district nurse Svea Hennix in the Karesuando study. DNA extractions for ORCADES were performed at the Wellcome Trust Clinical Research Facility in Edinburgh. The ORCADES authors would like to acknowledge the invaluable contributions of Lorraine Anderson and the research nurses in Orkney, the administrative team in Edinburgh and the people of Orkney. The PROCARDIS authors acknowledge the technical contributions of Anuj Goel, Simon Heath, Ivo Gut, and Ann-Christine Syvänen. The SardiNIA authors thank the team of the physician, the nurses and the genotyping team of biologists: Gianluca Usala, Fabio Busonero, Andrea Maschio, Sandra Lai, Mariano Dei, and Antonella Mulas. The SHIP authors are grateful to Stefan Funke for the opportunity to use his Server Cluster for SNP Imputation, and to the contribution of Florian Ernst, Anja Hoffmann, and Astrid Petersmann in generating the SNP data. The SSAGA authors thank Professors Nicholas Martin and Grant Montgomery and all members of the Genetic Epidemiology and Molecular Epidemiology Laboratories who were involved in the planning and conduct of the SSAGA studies. The TwinsUK authors thank the staff from the TwinsUK, the DNA Collections and Genotyping Facilities at the Wellcome Trust Sanger Institute for sample preparation; Le Centre National de Génotypage, France, led by Mark Lathrop, for genotyping; Duke University, North Carolina, United States of America, led by David Goldstein, for genotyping; and the Finnish Institute of Molecular Medicine, Finnish Genome Center, University of Helsinki, led by Aarno Palotie. The UA measurements (ORCADES, MICROS, CROATIA) were performed in Regensburg in the laboratory of Professor Schmitz. The details and full memberships of the consortia are listed at http://www.procardis.org for the PROCARDIS consortium, http://www.helmholtz-muenchen.de/kora/ for the KORA study, http://www.euengage.org for the ENGAGE consortium and http://homepages.ed.ac.uk/s0565445/ for the EUROSPAN consortium. See Text S1 for membership and affiliations of the Wellcome Trust Case Control Consortium.

Author Contributions

Conceived and designed the experiments: AFW JA JBW UG JFW IR PP HW AD HEW TDS LP HV RN PV MC TI CG. Performed the experiments: HP RWS. Analyzed the data: MK TJ SS AT VV MP MM EA CW MF ÅJ DRN YA CL PS RWS CG. Contributed reagents/materials/analysis tools: JSB SB MB MB HC JC AD GH MIM TM VM PM MN JP VS NJS DS MU UV GW DW TI. Wrote the paper: MK TJ CG. Conducted meta-analysis: MK TJ EA CG.

References

  1. 1. Taniguchi A, Kamatani N (2008) Control of renal uric acid excretion and gout. Curr Opin Rheumatol 20: 192–197.
  2. 2. Koenig W, Meisinger C (2008) Uric acid, type 2 diabetes, and cardiovascular diseases: fueling the common soil hypothesis? Clin Chem 54: 231–233.
  3. 3. Hayden MR, Tyagi SC (2004) Uric acid: A new look at an old risk marker for cardiovascular disease, metabolic syndrome, and type 2 diabetes mellitus: The urate redox shuttle. Nutr Metab (Lond) 1: 10.
  4. 4. Cutler RG (1984) Urate and ascorbate: their possible roles as antioxidants in determining longevity of mammalian species. Arch Gerontol Geriatr 3: 321–348.
  5. 5. Whitfield JB, Martin NG (1983) Inheritance and alcohol as factors influencing plasma uric acid levels. Acta Genet Med Gemellol (Roma ) 32: 117–126.
  6. 6. Doring A, Gieger C, Mehta D, Gohlke H, Prokisch H, et al. (2008) SLC2A9 influences uric acid concentrations with pronounced sex-specific effects. Nat Genet 40: 430–436.
  7. 7. Li S, Sanna S, Maschio A, Busonero F, Usala G, et al. (2007) The GLUT9 gene is associated with serum uric acid levels in Sardinia and Chianti cohorts. PLoS Genet 3: e194.
  8. 8. Vitart V, Rudan I, Hayward C, Gray NK, Floyd J, et al. (2008) SLC2A9 is a newly identified urate transporter influencing serum urate concentration, urate excretion and gout. Nat Genet 40: 437–442.
  9. 9. Wallace C, Newhouse SJ, Braund P, Zhang F, Tobin M, et al. (2008) Genome-wide association study identifies genes for biomarkers of cardiovascular disease: serum urate and dyslipidemia. Am J Hum Genet 82: 139–149.
  10. 10. Caulfield MJ, Munroe PB, O'Neill D, Witkowska K, Charchar FJ, et al. (2008) SLC2A9 is a high-capacity urate transporter in humans. PLoS Med 5: e197.
  11. 11. Dehghan A, Kottgen A, Yang Q, Hwang SJ, Kao WL, et al. (2008) Association of three genetic loci with uric acid concentration and risk of gout: a genome-wide association study. Lancet 372: 1953–1961.
  12. 12. Sparso T, Andersen G, Nielsen T, Burgdorf KS, Gjesing AP, et al. (2008) The GCKR rs780094 polymorphism is associated with elevated fasting serum triacylglycerol, reduced fasting and OGTT-related insulinaemia, and reduced risk of type 2 diabetes. Diabetologia 51: 70–75.
  13. 13. Anzai N, Miyazaki H, Noshiro R, Khamdang S, Chairoungdua A, et al. (2004) The multivalent PDZ domain-containing protein PDZK1 regulates transport activity of renal urate-anion exchanger URAT1 via its C terminus. J Biol Chem 279: 45942–45950.
  14. 14. Halestrap AP, Price NT (1999) The proton-linked monocarboxylate transporter (MCT) family: structure, function and regulation. Biochem J 343 Pt 2: 281–299.
  15. 15. Halestrap AP, Meredith D (2004) The SLC16 gene family-from monocarboxylate transporters (MCTs) to aromatic amino acid transporters and beyond. Pflugers Arch 447: 619–628.
  16. 16. Saxena R, Voight BF, Lyssenko V, Burtt NP, de Bakker PI, et al. (2007) Genome-wide association analysis identifies loci for type 2 diabetes and triglyceride levels. Science 316: 1331–1336.
  17. 17. Vaxillaire M, Cavalcanti-Proenca C, Dechaume A, Tichet J, Marre M, et al. (2008) The common P446L polymorphism in GCKR inversely modulates fasting glucose and triglyceride levels and reduces type 2 diabetes risk in the DESIR prospective general French population. Diabetes 57: 2253–2257.
  18. 18. Orho-Melander M, Melander O, Guiducci C, Perez-Martinez P, Corella D, et al. (2008) Common missense variant in the glucokinase regulatory protein gene is associated with increased plasma triglyceride and C-reactive protein but lower fasting glucose concentrations. Diabetes 57: 3112–3121.
  19. 19. Ter Maaten JC, Voorburg A, Heine RJ, Ter Wee PM, Donker AJ, et al. (1997) Renal handling of urate and sodium during acute physiological hyperinsulinaemia in healthy subjects. Clin Sci (Lond) 92: 51–58.
  20. 20. Facchini F, Chen YD, Hollenbeck CB, Reaven GM (1991) Relationship between resistance to insulin-mediated glucose uptake, urinary uric acid clearance, and plasma uric acid concentration. JAMA 266: 3008–3011.
  21. 21. Choi HK, Mount DB, Reginato AM (2005) Pathogenesis of gout. Ann Intern Med 143: 499–516.
  22. 22. Bakker SJ, Gans RO, Ter Maaten JC, Teerlink T, Westerhoff HV, et al. (2001) The potential role of adenosine in the pathophysiology of the insulin resistance syndrome. Atherosclerosis 155: 283–290.
  23. 23. Yang C, Pring M, Wear MA, Huang M, Cooper JA, et al. (2005) Mammalian CARMIL inhibits actin filament capping by capping protein. Dev Cell 9: 209–221.
  24. 24. Skovhus KV, Bergholdt R, Erichsen C, Sparre T, Nerup J, et al. (2006) Identification and characterization of secretagogin promoter activity. Scand J Immunol 64: 639–645.
  25. 25. Enomoto A, Kimura H, Chairoungdua A, Shigeta Y, Jutabha P, et al. (2002) Molecular identification of a renal urate anion exchanger that regulates blood urate levels. Nature 417: 447–452.
  26. 26. Anzai N, Enomoto A, Endou H (2005) Renal urate handling: clinical relevance of recent advances. Curr Rheumatol Rep 7: 227–234.
  27. 27. Ekaratanawong S, Anzai N, Jutabha P, Miyazaki H, Noshiro R, et al. (2004) Human organic anion transporter 4 is a renal apical organic anion/dicarboxylate exchanger in the proximal tubules. J Pharmacol Sci 94: 297–304.
  28. 28. Anzai N, Kanai Y, Endou H (2007) New insights into renal transport of urate. Curr Opin Rheumatol 19: 151–157.
  29. 29. Gieger C, Geistlinger L, Altmaier E, Hrabe dA, Kronenberg F, et al. (2008) Genetics meets metabolomics: a genome-wide association study of metabolite profiles in human serum. PLoS Genet 4: e1000282.
  30. 30. Nozaki S, Tanaka M, Mizuno K, Ataka S, Mizuma H, et al. (2009) Mental and physical fatigue-related biochemical alterations. Nutrition 25: 51–57.
  31. 31. Nakazono K, Watanabe N, Matsuno K, Sasaki J, Sato T, et al. (1991) Does superoxide underlie the pathogenesis of hypertension? Proc Natl Acad Sci U S A 88: 10045–10048.
  32. 32. Rauchova H, Dobesova Z, Drahota Z, Zicha J, Kunes J (1998) The effect of chronic L-carnitine treatment on blood pressure and plasma lipids in spontaneously hypertensive rats. Eur J Pharmacol 342: 235–239.
  33. 33. Huang W, Shaikh SN, Ganapathy ME, Hopfer U, Leibach FH, et al. (1999) Carnitine transport and its inhibition by sulfonylureas in human kidney proximal tubular epithelial cells. Biochem Pharmacol 58: 1361–1370.
  34. 34. Roschinger W, Muntau AC, Duran M, Dorland L, IJlst L, et al. (2000) Carnitine-acylcarnitine translocase deficiency: metabolic consequences of an impaired mitochondrial carnitine cycle. Clin Chim Acta 298: 55–68.
  35. 35. Loots DT, Mienie LJ, Bergh JJ, Van der Schyf CJ (2004) Acetyl-L-carnitine prevents total body hydroxyl free radical and uric acid production induced by 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) in the rat. Life Sci 75: 1243–1253.
  36. 36. Stapleton SR, Currie KP, Scott RH, Bell BA (1992) Palmitoyl-DL-carnitine has calcium-dependent effects on cultured neurones from rat dorsal root ganglia. Br J Pharmacol 107: 1192–1197.
  37. 37. Fredriksson R, Nordstrom KJ, Stephansson O, Hagglund MG, Schioth HB (2008) The solute carrier (SLC) complement of the human genome: phylogenetic classification reveals four major families. FEBS Lett 582: 3811–3816.
  38. 38. Li Y, Abecasis GR (2006) Mach 1.0: rapid haplotype reconstruction and missing genotype inference. Am J Hum Genet S79: 2290.
  39. 39. Marchini J, Howie B, Myers S, McVean G, Donnelly P (2007) A new multipoint method for genome-wide association studies by imputation of genotypes. Nat Genet 39: 906–913.
  40. 40. Bacanu SA, Devlin B, Roeder K (2000) The power of genomic control. Am J Hum Genet 66: 1933–1944.
  41. 41. Dudbridge F, Gusnanto A (2008) Estimation of significance thresholds for genomewide association scans. Genet Epidemiol 32: 227–234.