Physiol Genomics 47: 365–375, 2015. First published June 23, 2015; doi:10.1152/physiolgenomics.00004.2015.

Review

Genome-wide association studies and contribution to cardiovascular physiology Patricia B. Munroe and Andrew Tinker Clinical Pharmacology and The Heart Centre, William Harvey Research Institute, Barts and the London School of Medicine and Dentistry, London, United Kingdom Submitted 9 January 2015; accepted in final form 11 June 2015

Munroe PB, Tinker A. Genome-wide association studies and contribution to cardiovascular physiology. Physiol Genomics 47: 365–375, 2015. First published June 23, 2015; doi:10.1152/physiolgenomics.00004.2015.—The study of family pedigrees with rare monogenic cardiovascular disorders has revealed new molecular players in physiological processes. Genome-wide association studies of complex traits with a heritable component may afford a similar and potentially intellectually richer opportunity. In this review we focus on the interpretation of genetic associations and the issue of causality in relation to known and potentially new physiology. We mainly discuss cardiometabolic traits as it reflects our personal interests, but the issues pertain broadly in many other disciplines. We also describe some of the resources that are now available that may expedite follow up of genetic association signals into observations on causal mechanisms and pathophysiology. genome-wide association studies; cardiovascular physiology; cardiometabolic traits

traits with a complex but heritable component have been subjected to modern genomic techniques in an effort to understand their architecture. These analyses include traits that might be considered physiological and are our main focus here, such as heart rate, electrocardiogram (ECG) parameters, blood pressure, and those that are directly related to cardiovascular pathology such as coronary artery disease. This wealth of information has been only modestly exploited to date. In this review, we discuss the general issues that are raised in such studies and describe resources that will facilitate their investigation. Furthermore, we give exemplars illustrating specific physiological problems illuminated by the approach; however, we contend that the majority of the genetic data sets have been little utilized for physiological understanding.

A LARGE NUMBER OF CARDIOMETABOLIC

Design of Genome-wide Association Studies Before examining some specific studies, it is important to understand the general design of genetic association studies and what an association means. Throughout the human genome there is a degree of genetic variation: most commonly thought of as variations in a single base (single nucleotide polymorphisms, SNPs) but also includes insertions, deletions, and variations in copies of repeating sequence. There are currently over 78 million SNPs reported (Box 1), and these may be common or rare, present in introns or exons and in the latter case may affect the protein coding sequence or not (nonsynonymous/synonymous). Genetic variants occur on average every 1 Kb across the genome, and at some loci these alleles tend Address for reprint requests and other correspondence: A. Tinker, Clinical Pharmacology and The Heart Centre, William Harvey Research Inst., Barts and the London School of Medicine and Dentistry, Charterhouse Sq., London, EC1M6BQ, UK (e-mail: [email protected]).

to cosegregate together in blocks of genomic DNA (haplotype blocks); this is a process known as linkage disequilibrium (LD). This is advantageous for mapping traits to a chromosomal location as a single SNP (tagSNP) can be used to impute the identity of other SNPs across a larger genomic region. TagSNPs are included on high-density SNP arrays, and these are used for genotyping of the whole genome. The genetic background of a trait can be analyzed by performing a genetic association study and these analyses can be done for either binary (case-control) or continuous traits. Typically a genome-wide association study (GWAS) is conducted on hundreds to thousands of unrelated individuals, and these individuals are genotyped using commercial arrays with up to two million SNPs (mostly common SNPs with minor allele frequencies ⬎5%). Then for each SNP (genotyped or imputed) the allele (a version of the variant) present is tested for association with the trait of interest. The SNP associations that are significant or suggestive after a multiple testing correction (genome-wide significance is usually defined as P ⬍ 5 ⫻ 10⫺8) are then usually followed up in independent samples to confirm these findings. It is worth noting that prior to 2005 and the publication of the first GWAS, investigators generally chose a candidate gene based on known physiology for example and genotyped one or more SNPs within it and then tested for association with the trait of interest. If SNPs were in the promoter or coding region and nonsynonymous, then this was combined with an appropriate functional correlate. The GWAS approach identifies a genomic region, and within this region there may be several genes, all being considered initially as a potential “candidate” gene. Furthermore, the most statistically significant associated SNP may be in high LD (normally defined by r2 ⬎ 0.8) with a large number of other SNPs, and any of these may be the actual variant that leads to

1094-8341/15 Copyright © 2015 the American Physiological Society

365

Review 366

CARDIOVASCULAR PHYSIOLOGY AND GWAS

the biological effect. The individual contribution of genetic variants found by GWAS is actually quite small, and this is a feature of many cardiometabolic traits. Missing Heritability A large number of loci have been identified for cardiometabolic traits using GWAS; however, the identified loci explain very little of the heritability of the trait. For example blood pressure has a heritability of ⬃30%. The loci discovered thus far collectively explain only a small proportion of the heritability (⬃2–3%), and the associated variants are mostly common with small effect sizes (mostly ⬍1 mmHg systolic blood pressure and ⬍0.5 mmHg diastolic blood pressure) (8). In contrast the heritability of Type 2 diabetes is estimated to be 64%; here the loci are estimated to explain ⬃10% of the heritability (47). The explanation for missing heritability for complex traits is a much-debated issue with more recent discussions suggesting that the heritability for some traits may be overestimated (43, 74). One approach to try and find missing heritability is the study of low-frequency and rare genetic variants (allele frequencies ⬍1%) as these may have larger effect sizes. There are bespoke genotyping arrays (for example the exome chip) whose content is primarily lowfrequency and rare nonsynonymous SNPs discovered by sequencing of 12,000 whole exomes or whole genomes (http:// genome.sph.umich.edu/wiki/Exome_Chip_Design). These arrays have had some success in discovering new loci and have identified low-frequency variants with larger effect sizes at known GWAS loci (41). Resequencing studies of GWAS loci across cases and controls and exome sequencing of individuals with extreme phenotypes have also indicated there are some rare genetic variants with large effect sizes at GWAS loci, e.g., LDLR and APOA5 for LDL-C and ABCA1 for HDL-C (16, 61). These studies have marginally increased the percentage of heritability explained for some traits. Missing heritability may also be explained by contributions from structural variation and epigenetic effects, and these are hypotheses that are currently under exploration. Annotation of GWAS Loci for Identification of Candidate Genes and Causal SNPs Following a GWAS, fine mapping can be deployed to refine the size of a large association interval and narrow the list of potentially causal SNPs. This is usually achieved by genotyping additional SNPs at a locus using bespoke genotyping arrays with fine mapping content such as the CardioMetabochip (71), imputing existing datasets to a 1000 Genomes reference panel (Box 1) or genotyping individuals of non-European ancestry. The LD patterns vary according to ancestry, and individuals of African ancestry have higher diversity and lower LD, permitting refinement of association signals for some loci, and genotyping of exomic variants that have functional annotation (Box 1) (7). To determine candidate genes at a GWAS locus you can generate a list of SNPs in LD (r2 ⬎ 0.5) with the most associated SNP and then map these SNPs to genomic location and corresponding genes. Genes that map up to 500 Kb either side of the most associated SNP should also be considered. A literature review of these genes may provide some information on the known function. To delineate the SNP(s) you would take forward into functional testing you can check if the most

associated SNP is a nonsynonymous SNP or if it is in high LD with one. If this is the case, it can be postulated that the nonsynonymous SNP might affect protein function in a modest fashion, and an experimental plan can be devised to test this hypothesis. Researchers can also assess if the most associated SNP or a SNP in high LD is associated with expression levels of nearby gene(s) by performing a cis-expression quantitative trait locus (eQTL) analysis. There are online bioinformatics resources for performing these preliminary analyses in a range of different tissue and cell types (Box 1). Currently the numbers of samples from cardiovascular relevant tissues is limited. The number of GWAS SNPs that are annotated as being nonsynonymous is small, and the majority (93%) are located in introns and intergenic regions (44). It is likely that several of these variants are involved in the regulation of transcriptional activity. The ENCODE and Roadmap Epigenomics projects have revolutionized our understanding of the regulatory landscape, providing detailed information on promoter and enhancer DNA sequences across the human genome across hundreds of different cell types and tissues (Box 1). We can utilize a range of bioinformatics tools to check if GWAS SNPs or their LD proxies are located in regulatory sequences in different tissues and if the SNP(s) are predicted to affect transcription binding sites (Box 1). These online tools provide a detailed annotation of all currently known SNPs at a GWAS locus. We also need to consider if SNP(s) found in enhancer sequences are interacting with a promoter of a nearby gene or if there are long-range chromosomal interactions? The experimental methods of 4C-Seq and HiC can be utilized for this purpose (13). Figure 1 illustrates examples of follow-up analyses for a heart rate locus on chromosome 3 (14). Some Case Studies in Cardiometabolic Traits Illustrating Some of the Issues in Interpretation As detailed above the mapping and identification of a gene from a GWAS is a challenging process and is much less certain than in hereditary monogenic disorders. The commonest justification given for the substantial investment in GWAS is identifying new genes and new biology, and processes that might lead to new targets for therapeutic innovation. How well does this stack up to closer scrutiny? We consider three potential scenarios. Pointing to new physiology. Circulating blood lipid levels are important risk factors for cardiovascular disease, and the GWAS approach has yielded numerous variants associated with lipid traits and coronary artery disease (34). A recent study has reported genotyping of the exome array (Box 1) with lipid traits (24). This analysis yielded genome-wide significant associations with coding variants at 10 known lipid loci, nine of which are in genes with recognized roles in lipid metabolism. Of particular interest to the wider community interested in fine mapping of GWAS loci was the observation of significant association with rs58542926, a Glu167Lys variant located in TM6SF2. This variant is in high LD with the previously reported GWAS variant in the region. There are 21 candidate genes at this locus; further analyses of TM6SF2 indicated high levels of expression in the liver, and “in vivo” experiments of transient overexpression of TM6SF2 using a liver-specific adenovirus in mice led to increased total cholesterol levels and triglycerides in the mice, and transient knock-

Physiol Genomics • doi:10.1152/physiolgenomics.00004.2015 • www.physiolgenomics.org

Review 367

CARDIOVASCULAR PHYSIOLOGY AND GWAS

Plotted SNPs 10

r2

100

rs9647379

0.6

− log10(p−value)

8

80

0.4 0.2

6

60

4

40

2

20

Recombination rate (cM/Mb)

0.8

0

0 PLD1

FNDC3B TMEM212

173

GHSR

173.2

173.4

173.6

Position on chr3 (Mb)

Fig. 1. The pathway from genome-wide association study (GWAS) signal to functional analyses. A regional association plot of a GWAS locus for heart rate is shown (12). The region extending to within 500 Kb of rs9647379 [the most associated single nucleotide polymorphism (SNP)] is shown. The statistical significance of SNPs in the region are illustrated on the ⫺log10(P) scale as a function of chromosomal position (NCBI Build 37). The most associated SNP is indicated with the purple diamond, and the combined discovery and replication P value is shown. The correlation of rs9647379 to other SNPs at the locus is shown on a scale from minimal (gray to blue) to maximal (red). The fine-scale recombination rate is indicated in aqua blue. The genes in the region are indicated along the x-axis. This locus has 4 candidate genes: fibronectin type III domain 3B (FNDC3B), growth hormone secretagogue receptor (GHSR), placenta specific 1-like (TMEM122), and phospholipase D1, phosphatidylcholine-specific (PLD1). Bioinformatics analysis indicates PLD1 and FNDC3B are expressed in heart tissue; there are no nonsynonymous (ns)SNPs in high linkage disequilibrium (LD) and no evidence from public databases of cis-express quantitative trait loci (eQTLs). Functional experimentation in Drosophila melanogaster and Dario rerio using RNA interference (RNAi) to knock down the expression levels of candidate genes where the orthologs were available found lower resting heart rates in D. melongaster pupae for GHSR and PLD1 and reduced heart rates and unlooped hearts in some of the D. rerio embryos for PLD1 (12). The functional analyses indicate GSHR and PLD1 as 2 candidate genes for further studies.

down of TM6SF2 led to protein levels being significantly decreased in the liver. These data are concordant with carriers of the lysine amino acid at position 167 in TMS6SF2 having decreased lipid levels. The variant was also shown to be associated with a decreased risk of myocardial infarction. The function of TM6SF2 (transmembrane 6 superfamily member 2) was previously not known; these analyses pinpoint this as the likely candidate gene at this GWAS locus, paving the way for further functional analyses. The long QT syndrome refers to a characteristic ECG abnormality that is associated with torsade de pointes, a form of ventricular tachycardia, and sudden death. There are a number of hereditary monogenic diseases in which potassium channels (KCNQ1, KCNH2), the cardiac sodium channel (SCN5A), and interacting proteins are implicated as the cause (60). The QT interval is also heritable and this can be analyzed in GWAS (2, 3, 56, 74). Reassuringly, a number of genes such as SCN5A and KCNQ1 overlapped between the monogenic diseases and the GWAS studies. However, one of the strongest associations was with a locus that included the nitric oxide synthase 1 adaptor protein (NOS1AP) gene. This prompted an investigation into NOS1AP function in ventricular cardiomyocytes (9). The protein did indeed interact with NOS1 in the heart and overexpression inhibited L-type Ca2⫹ channel function. In a further twist, polymorphisms in NOS1AP in monogenic long QT syndromes may act as disease-modifying genes

(11) and could underlie some of the predisposition to the more common drug-induced long QT syndrome (27). Finally, subsequent studies have identified the causative SNP as being within an enhancer element that affects NOS1AP expression. Overexpression of NOS1AP led to a shortened action potential duration and an increased conduction velocity in neonatal rate ventricular myocytes. Furthermore, NOS1AP localized to the intercalated disc and a panel of genes expressed at this site were enriched and associated with the QT interval (31). Uromodulin is a protein secreted in the urine from the kidney and was originally identified as the Tamm-Horsfall protein (54). Recently, GWAS have discovered variants in the UMOD gene to be associated with renal function, chronic kidney disease, and hypertension (36, 37, 53). There was some prior work suggesting UMOD as a candidate gene for hypertension, but there were limited genetic analyses (26, 69). Uromodulin is expressed solely in the thick ascending limb of the loop of Henle and is inserted as a glycosylphosphatidylinositol anchored protein in the apical membrane of the tubular epithelial cells. It is cleaved and released into the urine from there (52). This has prompted studies examining the interaction of uromodulin with key transport proteins in the thick ascending limb namely the Na⫹/K⫹/2Cl⫺ cotransporter and the renal outer medullary potassium channel (ROMK2). Experimental work revealed that uromodulin promoted the surface expression of both proteins and loss of function would potentially

Physiol Genomics • doi:10.1152/physiolgenomics.00004.2015 • www.physiolgenomics.org

Review 368

CARDIOVASCULAR PHYSIOLOGY AND GWAS

promote salt loss, while increased expression might lead to salt retention and hypertension (48, 57). This hypothesis was supported by subsequent work showing common variants in the promoter of the UMOD gene led to increased expression of the protein and that overexpression of the protein in mice led to hypertension (70). There are no direct studies in humans indicating increased blood pressure. However, in a small observational study of “never treated” hypertensive individuals, those with the “at risk” genotype had higher blood pressures, and in a small subset of these that had treatment with a loop diuretic that blocks Na⫹/K⫹/2Cl⫺ cotransporter there was a greater decrease in blood pressure in individuals with the at-risk genotype (n ⫽ 118) compared with individuals heterozygous or homozygous for the other genotypes (n ⫽ 47). Contribution to known physiological pathways. Blood pressure homeostasis is maintained by several systems including the central and sympathetic nervous systems, the vasculature, and the kidney, together with various hormonal regulators. Recent large meta-analyses of GWAS have led to the discovery of over 63 loci influencing blood pressure (8). A review of the SNP associations finds many that are in or near genes that are known to affect blood pressure based on prior functional and physiological experiments. GWAS have indicated several SNP associations in vasodilator and vasoconstrictor stimulators of vascular smooth muscle cells (Fig. 2). These include a SNP association upstream of endothelial nitric oxide synthase (30,

59), three SNP associations near the natriuretic peptides A and B on chromosome 1 (50, 68), a SNP near the guanylate cyclase ␣- and ␤-subunits, and two SNPs near the C-type natriuretic peptide receptor (17, 33). In the vasoconstrictor pathways there was a SNP association in intron 1 of the angiotensinogen gene, a precursor of angiotensin II (28 –30), and a SNP association in the adrenergic beta 1 receptor, a known target for beta-blockers (33). There has been limited functional work understanding the mechanism of GWAS SNP associations. The rs5068 (A/G) variant is associated with blood pressure and is located in the 3=-untranslated region close to the NPPA gene that encodes the atrial natriuretic peptide. Carriers of the G allele have increased atrial natriuretic peptide levels and a decreased risk of hypertension (51). In a recent study, the response of different genotypes to an intravenous salt challenge was compared after a week-long low- or high-salt diet (4). In both cases atrial natriuretic peptide levels were significantly higher in AG individuals before, during, and after the salt challenge. Furthermore, the authors showed that the microRNA miR-425 could influence the expression of luciferase reporter constructs containing the AA but not AG allele. Finally, the transfection of miR-425 into cardiomyocytes obtained from induced pluripotent stem cells (iPSCs) reduced NPPA mRNA levels and propeptide release of the precursor of atrial natriuretic peptide (4). These analyses make a compelling case for the rs5068

Fig. 2. Key mechanisms regulating vascular smooth muscle tone. Signaling events that promote vasoconstriction are indicated in red, and those that promote vasodilation are indicated in blue. The phosphophorylation (P) of myosin is a critical step in contraction of vascular smooth muscle. Vasoconstrictors such as angiotensin II and norepinephrine promote increased cystolic calcium concentration (Ca2⫹); this in turn activates myosin kinase leading to myosin-actin interactions leading to vasoconstriction of the vessel. SNP associations are observed in the genes indicated in the red text (AGT, angiotensinogen; ADRB1, beta 1 adrenergic receptor). Vasodilators such as atrial natriuretic peptide (ANP) and nitric oxide via cyclic guanosine monophosphate (cGMP) activate myosin phosphatase, which leads to dephosphorylation of myosin and relaxation. Further SNP associations have been found in this pathway, also indicated in red text [endothelial nitric oxide synthase (eNOS); ANP; guanylate cyclase ␣- and ␤-subunits (GUCYA and GUCY B, respectively)]. Genetic associations are also observed in the natriuretic peptide receptor-C (NPR-C) gene that, via a separate pathway, leads to vasodilation in vascular smooth muscle. Other abbreviations: AT1-R, angiotensin 1 receptor; CNP, c type natriuretic peptide; ANP-R, atrial natriuretic peptide receptor. Figure is adapted from Ref. 39. Physiol Genomics • doi:10.1152/physiolgenomics.00004.2015 • www.physiolgenomics.org

Review CARDIOVASCULAR PHYSIOLOGY AND GWAS

variant having differential effects on atrial natriuretic peptide via microRNA regulation. However, in the relatively small numbers of healthy young volunteers they studied they were not able to show a significant effect of the variant on blood pressure. It will be important to show robust changes in blood pressure according to genotype, as well as association between the genetic marker and BNP levels to move the field forward. A more complex scenario. One of the first GWAS in 2007 reported association between SNPs at the FTO (fat mass and obesity associated) locus and obesity (21). Individuals with two copies of the risk allele had a 1.67-fold higher rate of obesity compared with individuals with no copies. Follow-up studies of FTO gene included the development of murine models to investigate the function of fto; mice overexpressing fto were found to be overweight, whereas null allele fto mice were lean (10, 19). These experiments suggested FTO as the likely candidate gene in the associated interval. The most significantly associated SNP (rs9939609) is located in intron 1 of the FTO gene. An eQTL analysis in a range of different tissues including adipose tissue, however, found no association between FTO SNPs and FTO expression levels (23, 40). This lack of association between the associated variant and gene expression has perplexed researchers searching for causal mechanism(s) based on the human GWAS results. The FTO locus is a 47 Kb region of high LD, and the obesity-associated SNPs are located in introns 1 and 2 of the FTO gene. Recently, Smemo et al. (2014) in a series of elegant experiments demonstrated that DNA sequence containing the FTO variants is in contact with the promoter sequence of IRX3, a gene located 0.5 Mb away (64). The IRX3 gene encodes a transcription factor, and currently very little is known about its function. This physical DNA interaction occurs in humans, mice, and zebrafish, and the authors showed an association between FTO genetic variants and expression levels of IRX3 in the brain. These analyses and further follow-up studies in murine models indicate that the associated variants at the FTO locus might be influencing body mass via IRX3 and not FTO. Experimental Resources to Follow up GWAS Findings. We have proposed that the commonest scenario is for a GWAS study to point to one or more genes whose function will need to be explored, though the specifics of this might be more complex than anticipated. In this section we highlight some recent initiatives that might facilitate this endeavor. Murine resources. One key strategy is the use of animal models, in particular the mouse, rat, and zebrafish. Importantly the genomes of each of these can be readily manipulated and large-scale international mutagenesis projects are being undertaken (Box 1). The International Mouse Phenotyping Consortium aims to generate and phenotype knockout mouse models for every gene in the genome. Each of these lines will be subjected to a broad phenotyping pipeline that includes at some centers measurement of the ECG and cardiac function by echocardiography. In an interesting early analysis of 250 lines passing through one of these pipelines, a large number of mice had some phenotype, and this was often unexpected (72). Genes without an obvious paralog (i.e., a gene that was duplicated and homologous in the same species) were more likely to be essential. Secondly, previously unstudied genes were as likely to give a phenotype as those with a prior literature. As the authors point out there was a large bias to investigation of known genes, and a majority of genes in the

369

human genome remain completely unstudied. Furthermore, if a homozygotic knockout was lethal then heterozygotes were studied, and a potential phenotype in these was surprisingly common, suggesting haploinsufficiency is important, i.e., loss of expression of the protein from a single one of the two alleles was enough to lead to functional deficits. In many of these mice the targeting strategy enables the generation of an allele in which a critical exon is flanked by loxP sites suitable for conditional deletion in a particular tissue and/or time after interbreeding with cre recombinase expressing murine lines (63). Knockout mouse models permit the direct testing of a candidate gene for a functional role in the disease without the requirement of understanding the role of the functional polymorphisms levered from genetic studies. Mouse models overexpressing a gene of interest and models with inclusion of nonsynonymous variants that are present at some GWAS loci will also be important going forward. In addition to the increasing availability of genetically modified mice, in vivo phenotyping capabilities have significantly advanced, and these can complement traditional ex vivo, biochemical, and single-cell studies. Many of the traits such as blood pressure represent integrative physiological processes dependent on pathways intrinsic to vascular smooth muscle and endothelium but also extrinsic events for example cardiovascular control centers in the brain stem (65). In particular microfabrication techniques and bespoke solutions for small mammals have allowed the real-time monitoring of ECG and blood pressure with radiotelemetry systems in conscious living animals, complex imaging using echocardiography and MRI, and measurement of in vivo electrophysiological and cardiac function (e.g., pressure-volume loops) parameters using catheters and arrays. Murine models of cardiovascular (patho)physiology have contributed greatly to the field although there are species differences. For example, the murine cardiac action potential is much shorter than that of large mammals and is determined by a different set of potassium currents (12). Rat resources. The rat is a long-standing model system for studying cardiovascular disease (hypertension and stroke) and metabolic traits (diabetes). Monitoring the physiology of disease is thought to be easier, and for some traits (e.g., diabetes) the rat model is thought to resemble the human disease better (1). The Rat Genome Database (RGD; Box 1) provides a central resource for information on the different rat strains and includes genetic, genomic, and phenotypic data on the different models and bioinformatics tools (62). Until recently, the production of rat models with specific mutations or the generation of knockout or conditional strains was technically challenging because of issues with embryonic stem cell manipulation. These issues are now resolved, and as of Sept 2014 there were 2,998 rat strains in the RGD, and 1,977 of these were created by genome-modifying techniques. Rat models have been used to ascertain likely candidate genes at GWAS loci. Using zinc finger nuclease mutagenesis-damaging alleles were introduced into six hypertension candidate genes (agtrap, mthfr, clcn6, anp, bnp, and plod1) in the SS rat model (20). The MTHFRNPPB locus has multiple SNPs associated with blood pressure traits (17, 50, 51). Analysis of blood pressure and renal phenotypes in each of these rat models implicated five genes to be associated with hypertension, indicating complex interactions of alleles and genes at this locus. The gene-editing approach has been also been used to assess other candidate

Physiol Genomics • doi:10.1152/physiolgenomics.00004.2015 • www.physiolgenomics.org

Review 370

CARDIOVASCULAR PHYSIOLOGY AND GWAS

genes at established GWAS loci (e.g., SH2B3, PLEKHA7) (18, 58) and NR2F2, a candidate from GWAS and linkage analyses; these studies provided supportive data for each of the genes tested (38). Zebrafish (Dario rerio). The argument has been made that zebrafish cardiovascular physiology is surprisingly close to that of mammals and man (15). The small size, ease of breeding, and conservation of functional domains and proteins lend themselves to genomic and chemical screens. In relation to GWAS, the zebrafish may be particularly useful when a large number of candidate genes need to be screened at a particular locus as was recently demonstrated in the study of genes implicated in a GWAS in heart rate (14). Knockdown of genes in the zebrafish is readily achievable using morpholino knockdown, but the investigation of cardiovascular function is quite specialized. However, there are some differences: the zebrafish does not have a secondary heart field and the heart has a high regenerative capacity. Furthermore, there are efforts using chemical mutagenesis to systematically disrupt every zebrafish gene as part of the Zebrafish Mutation Project (Box 1) (35). Human iPSCs and disease in a dish. Animal models are much used, but direct approaches in humans and associated tissues and cells would circumvent species differences. Another technological advance is the use of human iPSCs. Adult cells, usually fibroblasts, obtained from a skin biopsy from a patient are reprogrammed to iPSCs by expressing a number of transcription factors (66). These can then be reprogrammed to a variety of tissues including cardiac myocytes (iPSC-CMs) and smooth muscle and then used as a cellular model for the relevant pathology (25). This is intellectually appealing and potentially overcomes some of the limitations of animal models. For example, it has been used by a number of groups to phenocopy the monogenic long QT syndromes. Mutations in KCNQ1 in LQT1 lead to action potential prolongation in the iPSC-CMs, and the channel complex fails to traffic to the cell membrane as we found in heterologous expression systems (46, 73). In principle such approaches could be used to investigate cardiovascular complex traits. However, the effect sizes of SNPs are much smaller than in monogenic diseases, and the effects are likely to be quite subtle. Secondly during the generation of iPSCs the cells accumulate genetic changes, particularly copy number variations (55). Finally, the iPSCCMs have a mixed electrophysiological signature with pacemaker, atrial, and ventricular-like action potential phenotypes with a relatively depolarized resting membrane potential (6, 46, 67). The iPSC-CMs do not have a typical adult ventricular morphology and lack well developed t-tubular networks. Genome editing. Recent advances in genome editing may have wide-ranging ramifications. The initial tools, zinc finger nucleases and transcription activator-like effector nucleases, were labor intensive to construct and required considerable optimization. However, the advent of CRISPR/Cas9 technology has meant the approach is technically simpler. The comparisons with other approaches and details are given in several recent articles (22, 42), and there are tools for designing appropriate strategies (see Box 1 and Ref. 45). Bespoke mouse models that mimic human mutations, for example, nonsynonymous, splice site and at transcription factor binding sites are all now technically possible and at a reduced cost compared with manipulating embryonic stem cells. There is also progress in developing the technology for editing of promoter and enhanc-

ers and introducing risk haplotypes; this will permit testing of SNPs from GWAS studies and the creation of potentially more relevant animal models. Another consequence is that it is possible to engineer genomes other than those of the traditional model organisms. This may open the door to large mammal models of cardiac disease overcoming the limitations of rodent models. Secondly, in the initial disease modeling using iPSC approaches the controls were family members without the disease. The problem is not only does the reprogramming itself introduce genetic changes but the individuals, even if siblings, are likely to have a number of genetic differences. Genome editing allows true isogenic controls by correcting for the causative mutation and is also a powerful proof of causation. This has been demonstrated in LQT2 (5). Furthermore, you needn’t necessarily use iPSCs but could genome-edit human embryonic stem cells instead to generate the relevant models. Studies in humans. The above approaches are potentially highly revealing, but they do not actually directly address the link in humans between the genomic variation and the physiological trait. We have already mentioned that one of the major mechanisms is likely to be association of the causative SNP with small variations in expression through changes in regulatory networks that are significant at a population level. This can be approached with bioinformatics databases and tools (Box 1), novel tissue collections, and subsequent eQTL analyses (see above) (4, 32). An especially useful resource in the future is likely to be the UK Biobank. This is an open resource in which 500,000 individuals aged 40 – 60 yr of age have a core assessment measuring a number of physiological and clinical parameters. Additionally, the individuals have samples collected from them for archiving including genotyping, and their health will be followed prospectively. Furthermore, various substudies have been completed or are underway in groups of 100,000 with more specialized investigations such as exercise testing, MRI, etc. A second initiative is the electronic mining of medical records, and this can generate very large epidemiological cohorts for understanding the progression of disease. The Farr Institute interrogating medical records within the British National Health Service is an example of such an enabling resource (Box 1). Conclusion Physiology has underpinned the rational development of pharmacological agents. The pharmaceutical industry is experiencing difficulties in drug development not only in finding new targets but also in the late failure of agents in phase III clinical trials. Despite the complexity in teasing out the causative gene and mechanism from GWAS we feel that modern genomics may offer another route to target identification and validation. It is telling that a large number of proteins remain completely unstudied in animal models and human disease. Furthermore, the GWAS approach might be tailored to measure the response of populations of individuals to specific physiological challenges such as is exemplified in many of the studies underway in UK Biobank. APPENDIX: LIST OF USEFUL URLs

1000 Genomes project, http://www.1000genomes.org SNP Annotation and Proxy Search - SNAP, http://www.broadinstitute. org/mpg/snap/

Physiol Genomics • doi:10.1152/physiolgenomics.00004.2015 • www.physiolgenomics.org

Review CARDIOVASCULAR PHYSIOLOGY AND GWAS

eQTL browser at NCBI, http://www.ncbi.nlm.nih.gov/projects/ gap/eqtl/index.cgi GenVar at the Sanger Institute, http://www.sanger.ac.uk/resources/ software/genevar/ Scan DB, http://www.scandb.org Broad Institute GTex Browser, http://www.gtexportal.org eQTL resources at the Prichard Lab, http://eqtl.uchicago.edu Encylopedia of DNA Elements - ENCODE, http://genome.ucsc. edu/ENCODE/ Roadmap Epigenomics Project, http://www.roadmapepigenomics. org HAPLOREG, http://www.broadinstitute.org/mammals/haploreg RegulomeDB, http://regulome.stanford.edu Exome chip, http://genome.sph.umich.edu/wiki/Exome_Chip_ Design International Mouse Phenotyping Consortium http://www. mousephenotype.org/ Rat Genome Database, http://rgd.mcw.edu Zebrafish Mutation Project, http://www.sanger.ac.uk/Projects/D_ rerio/zmp/ CHOPCHOP, https://chopchop.rc.fas.harvard.edu UK Biobank, http://www.ukbiobank.ac.uk/ Farr Institute, http://www.farrinstitute.org/ GRANTS The work in our laboratories is supported by the British Heart Foundation, Medical Research Council, and The National Institute for Health Research Barts Cardiovascular Biomedical Research Unit. DISCLOSURES No conflicts of interest, financial or otherwise, are declared by the author(s). AUTHOR CONTRIBUTIONS Author contributions: P.B.M. and A.T. prepared figures; P.B.M. and A.T. drafted manuscript; P.B.M. and A.T. edited and revised manuscript; P.B.M. and A.T. approved final version of manuscript.

4.

REFERENCES 1. Aitman TJ, Critser JK, Cuppen E, Dominiczak A, Fernandez-Suarez XM, Flint J, Gauguier D, Geurts AM, Gould M, Harris PC, Holmdahl R, Hubner N, Izsvak Z, Jacob HJ, Kuramoto T, Kwitek AE, Marrone A, Mashimo T, Moreno C, Mullins J, Mullins L, Olsson T, Pravenec M, Riley L, Saar K, Serikawa T, Shull JD, Szpirer C, Twigger SN, Voigt B, Worley K. Progress and prospects in rat genetics: a community view. Nat Genet 40: 516 –522, 2008. 2. Arking DE, Junttila MJ, Goyette P, Huertas-Vazquez A, Eijgelsheim M, Blom MT, Newton-Cheh C, Reinier K, Teodorescu C, Uy-Evanado A, Carter-Monroe N, Kaikkonen KS, Kortelainen ML, Boucher G, Lagace C, Moes A, Zhao X, Kolodgie F, Rivadeneira F, Hofman A, Witteman JC, Uitterlinden AG, Marsman RF, Pazoki R, Bardai A, Koster RW, Dehghan A, Hwang SJ, Bhatnagar P, Post W, Hilton G, Prineas RJ, Li M, Kottgen A, Ehret G, Boerwinkle E, Coresh J, Kao WH, Psaty BM, Tomaselli GF, Sotoodehnia N, Siscovick DS, Burke GL, Marban E, Spooner PM, Cupples LA, Jui J, Gunson K, Kesaniemi YA, Wilde AA, Tardif JC, O’Donnell CJ, Bezzina CR, Virmani R, Stricker BH, Tan HL, Albert CM, Chakravarti A, Rioux JD, Huikuri HV, Chugh SS. Identification of a sudden cardiac death susceptibility locus at 2q24.2 through genome-wide association in European ancestry individuals. PLoS Genet 7: e1002158, 2011. 3. Arking DE, Pulit SL, Crotti L, van der Harst P, Munroe PB, Koopmann TT, Sotoodehnia N, Rossin EJ, Morley M, Wang X, Johnson AD, Lundby A, Gudbjartsson DF, Noseworthy PA, Eijgelsheim M, Bradford Y, Tarasov KV, Dorr M, Muller-Nurasyid M, Lahtinen AM, Nolte IM, Smith AV, Bis JC, Isaacs A, Newhouse SJ, Evans DS, Post WS, Waggott D, Lyytikainen LP, Hicks AA, Eisele L, Ellinghaus D, Hayward C, Navarro P, Ulivi S, Tanaka T, Tester DJ, Chatel S, Gustafsson S, Kumari M, Morris RW, Naluai AT, Padmanabhan S, Kluttig A, Strohmer B, Panayiotou AG, Torres M, Knoflach M, Hubacek JA, Slowikowski K, Raychaudhuri S, Kumar RD, Harris

5.

6.

7.

8.

9.

10.

371

TB, Launer LJ, Shuldiner AR, Alonso A, Bader JS, Ehret G, Huang H, Kao WH, Strait JB, Macfarlane PW, Brown M, Caulfield MJ, Samani NJ, Kronenberg F, Willeit J, CARe Consortium, COGENT Consortium, Smith JG, Greiser KH, Meyer Zu Schwabedissen H, Werdan K, Carella M, Zelante L, Heckbert SR, Psaty BM, Rotter JI, Kolcic I, Polasek O, Wright AF, Griffin M, Daly MJ, Dcct/Edic, Arnar DO, Holm H, Thorsteinsdottir U, eMERGE Consortium, Denny JC, Roden DM, Zuvich RL, Emilsson V, Plump AS, Larson MG, O’Donnell CJ, Yin X, Bobbo M, D’Adamo AP, Iorio A, Sinagra G, Carracedo A, Cummings SR, Nalls MA, Jula A, Kontula KK, Marjamaa A, Oikarinen L, Perola M, Porthan K, Erbel R, Hoffmann P, Jockel KH, Kalsch H, Nothen MM, Consortium H, den Hoed M, Loos RJ, Thelle DS, Gieger C, Meitinger T, Perz S, Peters A, Prucha H, Sinner MF, Waldenberger M, de Boer RA, Franke L, van der Vleuten PA, Beckmann BM, Martens E, Bardai A, Hofman N, Wilde AA, Behr ER, Dalageorgou C, Giudicessi JR, Medeiros-Domingo A, Barc J, Kyndt F, Probst V, Ghidoni A, Insolia R, Hamilton RM, Scherer SW, Brandimarto J, Margulies K, Moravec CE, Greco MF, Fuchsberger C, O’Connell JR, Lee WK, Watt GC, Campbell H, Wild SH, El Mokhtari NE, Frey N, Asselbergs FW, Mateo Leach I, Navis G, van den Berg MP, van Veldhuisen DJ, Kellis M, Krijthe BP, Franco OH, Hofman A, Kors JA, Uitterlinden AG, Witteman JC, Kedenko L, Lamina C, Oostra BA, Abecasis GR, Lakatta EG, Mulas A, Orru M, Schlessinger D, Uda M, Markus MR, Volker U, Snieder H, Spector TD, Arnlov J, Lind L, Sundstrom J, Syvanen AC, Kivimaki M, Kahonen M, Mononen N, Raitakari OT, Viikari JS, Adamkova V, Kiechl S, Brion M, Nicolaides AN, Paulweber B, Haerting J, Dominiczak AF, Nyberg F, Whincup PH, Hingorani AD, Schott JJ, Bezzina CR, Ingelsson E, Ferrucci L, Gasparini P, Wilson JF, Rudan I, Franke A, Muhleisen TW, Pramstaller PP, Lehtimaki TJ, Paterson AD, Parsa A, Liu Y, van Duijn CM, Siscovick DS, Gudnason V, Jamshidi Y, Salomaa V, Felix SB, Sanna S, Ritchie MD, Stricker BH, Stefansson K, Boyer LA, Cappola TP, Olsen JV, Lage K, Schwartz PJ, Kaab S, Chakravarti A, Ackerman MJ, Pfeufer A, de Bakker PI, NewtonCheh C. Genetic association study of QT interval highlights role for calcium signaling pathways in myocardial repolarization. Nat Genet 46: 826 –836, 2014. Arora P, Wu C, Khan AM, Bloch DB, Davis-Dusenbery BN, Ghorbani A, Spagnolli E, Martinez A, Ryan A, Tainsh LT, Kim S, Rong J, Huan T, Freedman JE, Levy D, Miller KK, Hata A, Del Monte F, Vandenwijngaert S, Swinnen M, Janssens S, Holmes TM, Buys ES, Bloch KD, Newton-Cheh C, Wang TJ. Atrial natriuretic peptide is negatively regulated by microRNA-425. J Clin Invest 123: 3378 –3382, 2013. Bellin M, Casini S, Davis RP, D’Aniello C, Haas J, Ward-van Oostwaard D, Tertoolen LG, Jung CB, Elliott DA, Welling A, Laugwitz KL, Moretti A, Mummery CL. Isogenic human pluripotent stem cell pairs reveal the role of a KCNH2 mutation in long-QT syndrome. EMBO J 32: 3161–3175, 2013. Braam SR, Tertoolen L, Casini S, Matsa E, Lu HR, Teisman A, Passier R, Denning C, Gallacher DJ, Towart R, Mummery CL. Repolarization reserve determines drug responses in human pluripotent stem cell derived cardiomyocytes. Stem Cell Res 10: 48 –56, 2013. Buyske S, Wu Y, Carty CL, Cheng I, Assimes TL, Dumitrescu L, Hindorff LA, Mitchell S, Ambite JL, Boerwinkle E, Buzkova P, Carlson CS, Cochran B, Duggan D, Eaton CB, Fesinmeyer MD, Franceschini N, Haessler J, Jenny N, Kang HM, Kooperberg C, Lin Y, Le Marchand L, Matise TC, Robinson JG, Rodriguez C, Schumacher FR, Voight BF, Young A, Manolio TA, Mohlke KL, Haiman CA, Peters U, Crawford DC, North KE. Evaluation of the metabochip genotyping array in African Americans and implications for fine mapping of GWAS-identified loci: the PAGE study. PLoS One 7: e35651, 2012. Cabrera CP, Ng FL, Warren HR, Barnes MR, Munroe PB, Caulfield MJ. Exploring hypertension genome-wide association studies findings and impact on pathophysiology, pathways, and pharmacogenetics. Wiley Interdiscip Rev Syst Biol Med 7: 73–90, 2015. Chang KC, Barth AS, Sasano T, Kizana E, Kashiwakura Y, Zhang Y, Foster DB, Marban E. CAPON modulates cardiac repolarization via neuronal nitric oxide synthase signaling in the heart. Proc Natl Acad Sci USA 105: 4477–4482, 2008. Church C, Lee S, Bagg EA, McTaggart JS, Deacon R, Gerken T, Lee A, Moir L, Mecinovic J, Quwailid MM, Schofield CJ, Ashcroft FM, Cox RD. A mouse model for the metabolic effects of the human fat mass and obesity associated FTO gene. PLoS Genet 5: e1000599, 2009.

Physiol Genomics • doi:10.1152/physiolgenomics.00004.2015 • www.physiolgenomics.org

Review 372

CARDIOVASCULAR PHYSIOLOGY AND GWAS

11. Crotti L, Monti MC, Insolia R, Peljto A, Goosen A, Brink PA, Greenberg DA, Schwartz PJ, George AL Jr. NOS1AP is a genetic modifier of the long-QT syndrome. Circulation 120: 1657–1663, 2009. 12. Davis RP, van den Berg CW, Casini S, Braam SR, Mummery CL. Pluripotent stem cell models of cardiac disease and their implication for drug discovery and development. Trends Mol Med 17: 475–484, 2011. 13. de Wit E, de Laat W. A decade of 3C technologies: insights into nuclear organization. Genes Dev 26: 11–24, 2012. 14. den Hoed M, Eijgelsheim M, Esko T, Brundel BJ, Peal DS, Evans DM, Nolte IM, Segre AV, Holm H, Handsaker RE, Westra HJ, Johnson T, Isaacs A, Yang J, Lundby A, Zhao JH, Kim YJ, Go MJ, Almgren P, Bochud M, Boucher G, Cornelis MC, Gudbjartsson D, Hadley D, van der Harst P, Hayward C, den Heijer M, Igl W, Jackson AU, Kutalik Z, Luan J, Kemp JP, Kristiansson K, Ladenvall C, Lorentzon M, Montasser ME, Njajou OT, O’Reilly PF, Padmanabhan S, St Pourcain B, Rankinen T, Salo P, Tanaka T, Timpson NJ, Vitart V, Waite L, Wheeler W, Zhang W, Draisma HH, Feitosa MF, Kerr KF, Lind PA, Mihailov E, Onland-Moret NC, Song C, Weedon MN, Xie W, Yengo L, Absher D, Albert CM, Alonso A, Arking DE, de Bakker PI, Balkau B, Barlassina C, Benaglio P, Bis JC, Bouatia-Naji N, Brage S, Chanock SJ, Chines PS, Chung M, Darbar D, Dina C, Dorr M, Elliott P, Felix SB, Fischer K, Fuchsberger C, de Geus EJ, Goyette P, Gudnason V, Harris TB, Hartikainen AL, Havulinna AS, Heckbert SR, Hicks AA, Hofman A, Holewijn S, Hoogstra-Berends F, Hottenga JJ, Jensen MK, Johansson A, Junttila J, Kaab S, Kanon B, Ketkar S, Khaw KT, Knowles JW, Kooner AS, Kors JA, Kumari M, Milani L, Laiho P, Lakatta EG, Langenberg C, Leusink M, Liu Y, Luben RN, Lunetta KL, Lynch SN, Markus MR, Marques-Vidal P, Mateo Leach I, McArdle WL, McCarroll SA, Medland SE, Miller KA, Montgomery GW, Morrison AC, Muller-Nurasyid M, Navarro P, Nelis M, O’Connell JR, O’Donnell CJ, Ong KK, Newman AB, Peters A, Polasek O, Pouta A, Pramstaller PP, Psaty BM, Rao DC, Ring SM, Rossin EJ, Rudan D, Sanna S, Scott RA, Sehmi JS, Sharp S, Shin JT, Singleton AB, Smith AV, Soranzo N, Spector TD, Stewart C, Stringham HM, Tarasov KV, Uitterlinden AG, Vandenput L, Hwang SJ, Whitfield JB, Wijmenga C, Wild SH, Willemsen G, Wilson JF, Witteman JC, Wong A, Wong Q, Jamshidi Y, Zitting P, Boer JM, Boomsma DI, Borecki IB, van Duijn CM, Ekelund U, Forouhi NG, Froguel P, Hingorani A, Ingelsson E, Kivimaki M, Kronmal RA, Kuh D, Lind L, Martin NG, Oostra BA, Pedersen NL, Quertermous T, Rotter JI, van der Schouw YT, Verschuren WM, Walker M, Albanes D, Arnar DO, Assimes TL, Bandinelli S, Boehnke M, de Boer RA, Bouchard C, Caulfield WL, Chambers JC, Curhan G, Cusi D, Eriksson J, Ferrucci L, van Gilst WH, Glorioso N, de Graaf J, Groop L, Gyllensten U, Hsueh WC, Hu FB, Huikuri HV, Hunter DJ, Iribarren C, Isomaa B, Jarvelin MR, Jula A, Kahonen M, Kiemeney LA, van der Klauw MM, Kooner JS, Kraft P, Iacoviello L, Lehtimaki T, Lokki ML, Mitchell BD, Navis G, Nieminen MS, Ohlsson C, Poulter NR, Qi L, Raitakari OT, Rimm EB, Rioux JD, Rizzi F, Rudan I, Salomaa V, Sever PS, Shields DC, Shuldiner AR, Sinisalo J, Stanton AV, Stolk RP, Strachan DP, Tardif JC, Thorsteinsdottir U, Tuomilehto J, van Veldhuisen DJ, Virtamo J, Viikari J, Vollenweider P, Waeber G, Widen E, Cho YS, Olsen JV, Visscher PM, Willer C, Franke L, Global BPgen Consortium, CARDIoGRAM Consortium, Erdmann J, Thompson JR, PR GWAS Consortium, Pfeufer A, QRS GWAS Consortium, Sotoodehnia N, QT-IGC Consortium, Newton-Cheh C, CHARGE-AF Consortium, Ellinor PT, Stricker BH, Metspalu A, Perola M, Beckmann JS, Smith GD, Stefansson K, Wareham NJ, Munroe PB, Sibon OC, Milan DJ, Snieder H, Samani NJ, Loos RJ. Identification of heart rate-associated loci and their effects on cardiac conduction and rhythm disorders. Nat Genet 45: 621–631, 2013. 15. Deo RC, MacRae CA. The zebrafish: scalable in vivo modeling for systems biology. Wiley Interdiscip Rev Syst Biol Med 3: 335–346, 2011. 16. Do R, Stitziel NO, Won H, Jorgensen AB, Duga S, Angelica MP, Kiezun A, Farrall M, Goel A, Zuk O, Guella I, Asselta R, Lange LA, Peloso GM, Auer PL, Girelli D, Martinelli N, Farlow DN, DePristo MA, Roberts R, Stewart AF, Saleheen D, Danesh J, Epstein SE, Sivapalaratnam S, Kees HG, Kastelein JJ, Samani NJ, Schunkert H, Erdmann J, Shah SH, Kraus WE, Davies R, Nikpay M, Johansen CT, Wang J, Hegele RA, Hechter E, Marz W, Kleber ME, Huang J, Johnson AD, Li M, Burke GL, Gross M, Liu Y, Assimes TL, Heiss G, Lange EM, Folsom AR, Taylor HA, Olivieri O, Hamsten A, Clarke R, Reilly DF, Yin W, Rivas MA, Donnelly P, Rossouw JE, Psaty BM, Herrington DM, Wilson JG, Rich SS, Bamshad MJ, Tracy RP,

17.

18.

19. 20.

21.

Adrienne CL, Rader DJ, Reilly MP, Spertus JA, Cresci S, Hartiala J, Wilson Tang WH, Hazen SL, Allayee H, Reiner AP, Carlson CS, Kooperberg C, Jackson RD, Boerwinkle E, Lander ES, Schwartz SM, Siscovick DS, McPherson R, Tybjaerg-Hansen A, Abecasis GR, Watkins H, Nickerson DA, Ardissino D, Sunyaev SR, O’Donnell CJ, Altshuler D, Gabriel S, Kathiresan S. Exome sequencing identifies rare LDLR and APOA5 alleles conferring risk for myocardial infarction. Nature 518: 102–106, 2015. Ehret GB, Munroe PB, Rice KM, Bochud M, Johnson AD, Chasman DI, Smith AV, Tobin MD, Verwoert GC, Hwang SJ, Pihur V, Vollenweider P, O’Reilly PF, Amin N, Bragg-Gresham JL, Teumer A, Glazer NL, Launer L, Zhao JH, Aulchenko Y, Heath S, Sober S, Parsa A, Luan J, Arora P, Dehghan A, Zhang F, Lucas G, Hicks AA, Jackson AU, Peden JF, Tanaka T, Wild SH, Rudan I, Igl W, Milaneschi Y, Parker AN, Fava C, Chambers JC, Fox ER, Kumari M, Go MJ, van der Harst P, Kao WH, Sjogren M, Vinay DG, Alexander M, Tabara Y, Shaw-Hawkins S, Whincup PH, Liu Y, Shi G, Kuusisto J, Tayo B, Seielstad M, Sim X, Nguyen KD, Lehtimaki T, Matullo G, Wu Y, Gaunt TR, Onland-Moret NC, Cooper MN, Platou CG, Org E, Hardy R, Dahgam S, Palmen J, Vitart V, Braund PS, Kuznetsova T, Uiterwaal CS, Adeyemo A, Palmas W, Campbell H, Ludwig B, Tomaszewski M, Tzoulaki I, Palmer ND, Aspelund T, Garcia M, Chang YP, O’Connell JR, Steinle NI, Grobbee DE, Arking DE, Kardia SL, Morrison AC, Hernandez D, Najjar S, McArdle WL, Hadley D, Brown MJ, Connell JM, Hingorani AD, Day IN, Lawlor DA, Beilby JP, Lawrence RW, Clarke R, Hopewell JC, Ongen H, Dreisbach AW, Li Y, Young JH, Bis JC, Kahonen M, Viikari J, Adair LS, Lee NR, Chen MH, Olden M, Pattaro C, Bolton JA, Kottgen A, Bergmann S, Mooser V, Chaturvedi N, Frayling TM, Islam M, Jafar TH, Erdmann J, Kulkarni SR, Bornstein SR, Grassler J, Groop L, Voight BF, Kettunen J, Howard P, Taylor A, Guarrera S, Ricceri F, Emilsson V, Plump A, Barroso I, Khaw KT, Weder AB, Hunt SC, Sun YV, Bergman RN, Collins FS, Bonnycastle LL, Scott LJ, Stringham HM, Peltonen L, Perola M, Vartiainen E, Brand SM, Staessen JA, Wang TJ, Burton PR, Soler AM, Dong Y, Snieder H, Wang X, Zhu H, Lohman KK, Rudock ME, Heckbert SR, Smith NL, Wiggins KL, Doumatey A, Shriner D, Veldre G, Viigimaa M, Kinra S, Prabhakaran D, Tripathy V, Langefeld CD, Rosengren A, Thelle DS, Corsi AM, Singleton A, Forrester T, Hilton G, McKenzie CA, Salako T, Iwai N, Kita Y, Ogihara T, Ohkubo T, Okamura T, Ueshima H, Umemura S, Eyheramendy S, Meitinger T, Wichmann HE, Cho YS, Kim HL, Lee JY, Scott J, Sehmi JS, Zhang W, Hedblad B, Nilsson P, Smith GD, Wong A, Narisu N, Stancakova A, Raffel LJ, Yao J, Kathiresan S, O’Donnell CJ, Schwartz SM, Ikram MA, Longstreth WT Jr, Mosley TH, Seshadri S, Shrine NR, Wain LV, Morken MA, Swift AJ, Laitinen J, Prokopenko I, Zitting P, Cooper JA, Humphries SE, Danesh J, Rasheed A, Goel A, Hamsten A, Watkins H, Bakker SJ, van Gilst WH, Janipalli CS, Mani KR, Yajnik CS, Hofman A, MattaceRaso FU, Oostra BA, Demirkan A, Isaacs A, Rivadeneira F, Lakatta EG, Orru M, Scuteri A, Ala-Korpela M, Kangas AJ, Lyytikainen LP, Soininen P, Tukiainen T, Wurtz P, Ong RT, Dorr M, Kroemer HK, Volker U, Volzke H, Galan P, Hercberg S, Lathrop M, Zelenika D, Deloukas P, Mangino M, Spector TD, Zhai G. Genetic variants in novel pathways influence blood pressure and cardiovascular disease risk. Nature 478: 103–109, 2011. Endres BT, Priestley JR, Palygin O, Flister MJ, Hoffman MJ, Weinberg BD, Grzybowski M, Lombard JH, Staruschenko A, Moreno C, Jacob HJ, Geurts AM. Mutation of Plekha7 attenuates salt-sensitive hypertension in the rat. Proc Natl Acad Sci USA 111: 12817–12822, 2014. Fischer J, Koch L, Emmerling C, Vierkotten J, Peters T, Bruning JC, Ruther U. Inactivation of the Fto gene protects from obesity. Nature 458: 894 –898, 2009. Flister MJ, Tsaih SW, O’Meara CC, Endres B, Hoffman MJ, Geurts AM, Dwinell MR, Lazar J, Jacob HJ, Moreno C. Identifying multiple causative genes at a single GWAS locus. Genome Res 23: 1996 –2002, 2013. Frayling TM, Timpson NJ, Weedon MN, Zeggini E, Freathy RM, Lindgren CM, Perry JR, Elliott KS, Lango H, Rayner NW, Shields B, Harries LW, Barrett JC, Ellard S, Groves CJ, Knight B, Patch AM, Ness AR, Ebrahim S, Lawlor DA, Ring SM, Ben-Shlomo Y, Jarvelin MR, Sovio U, Bennett AJ, Melzer D, Ferrucci L, Loos RJ, Barroso I, Wareham NJ, Karpe F, Owen KR, Cardon LR, Walker M, Hitman GA, Palmer CN, Doney AS, Morris AD, Smith GD, Hattersley AT, McCarthy MI. A common variant in the FTO gene is associated with

Physiol Genomics • doi:10.1152/physiolgenomics.00004.2015 • www.physiolgenomics.org

Review CARDIOVASCULAR PHYSIOLOGY AND GWAS

22. 23.

24.

25. 26. 27.

28.

29.

30.

31.

32.

33.

body mass index and predisposes to childhood and adult obesity. Science 316: 889 –894, 2007. Gaj T, Gersbach CA, Barbas CF 3rd. ZFN, TALEN, and CRISPR/Casbased methods for genome engineering. Trends Biotechnol 31: 397–405, 2013. Grunnet LG, Nilsson E, Ling C, Hansen T, Pedersen O, Groop L, Vaag A, Poulsen P. Regulation and function of FTO mRNA expression in human skeletal muscle and subcutaneous adipose tissue. Diabetes 58: 2402–2408, 2009. Holmen OL, Zhang H, Fan Y, Hovelson DH, Schmidt EM, Zhou W, Guo Y, Zhang J, Langhammer A, Lochen ML, Ganesh SK, Vatten L, Skorpen F, Dalen H, Zhang J, Pennathur S, Chen J, Platou C, Mathiesen EB, Wilsgaard T, Njolstad I, Boehnke M, Chen YE, Abecasis GR, Hveem K, Willer CJ. Systematic evaluation of coding variation identifies a candidate causal variant in TM6SF2 influencing total cholesterol and myocardial infarction risk. Nat Genet 46: 345–351, 2014. Inoue H, Nagata N, Kurokawa H, Yamanaka S. iPS cells: a game changer for future medicine. EMBO J 33: 409 –417, 2014. Iwai N, Kajimoto K, Kokubo Y, Tomoike H. Extensive genetic analysis of 10 candidate genes for hypertension in Japanese. Hypertension 48: 901–907, 2006. Jamshidi Y, Nolte IM, Dalageorgou C, Zheng D, Johnson T, Bastiaenen R, Ruddy S, Talbott D, Norris KJ, Snieder H, George AL, Marshall V, Shakir S, Kannankeril PJ, Munroe PB, Camm AJ, Jeffery S, Roden DM, Behr ER. Common variation in the NOS1AP gene is associated with drug-induced QT prolongation and ventricular arrhythmia. J Am Coll Cardiol 60: 841–850, 2012. Johnson A, Newton-Cheh C, Chasman D, Ehret G, Johnson T, Rose L, Rice K, Verwoert G, Launer L, Gudnason V. Association of hypertension drug target genes with blood pressure and hypertension in 86,588 individuals. Hypertension 57: 903–910, 2011. Johnson AD, Newton-Cheh C, Chasman DI, Ehret GB, Johnson T, Rose L, Rice K, Verwoert GC, Launer LJ, Gudnason V, Larson MG, Chakravarti A, Psaty BM, Caulfield M, van Duijn CM, Ridker PM, Munroe PB, Levy D. Association of hypertension drug target genes with blood pressure and hypertension in 86,588 individuals. Hypertension 57: 903–910, 2011. Johnson T, Gaunt TR, Newhouse SJ, Padmanabhan S, Tomaszewski M, Kumari M, Morris RW, Tzoulaki I, O’Brien ET, Poulter NR, Sever P, Shields DC, Thom S, Wannamethee SG, Whincup PH, Brown MJ, Connell JM, Dobson RJ, Howard PJ, Mein CA, Onipinla A, Shaw-Hawkins S, Zhang Y, Davey Smith G, Day IN, Lawlor DA, Goodall AH, Fowkes FG, Abecasis GR, Elliott P, Gateva V, Braund PS, Burton PR, Nelson CP, Tobin MD, van der Harst P, Glorioso N, Neuvrith H, Salvi E, Staessen JA, Stucchi A, Devos N, Jeunemaitre X, Plouin PF, Tichet J, Juhanson P, Org E, Putku M, Sober S, Veldre G, Viigimaa M, Levinsson A, Rosengren A, Thelle DS, Hastie CE, Hedner T, Lee WK, Melander O, Wahlstrand B, Hardy R, Wong A, Cooper JA, Palmen J, Chen L, Stewart AF, Wells GA, Westra HJ, Wolfs MG, Clarke R, Franzosi MG, Goel A, Hamsten A, Lathrop M, Peden JF, Seedorf U, Watkins H, Ouwehand WH, Sambrook J, Stephens J, Casas JP, Drenos F, Holmes MV, Kivimaki M, Shah S, Shah T, Talmud PJ, Whittaker J, Wallace C, Delles C, Laan M, Kuh D, Humphries SE, Nyberg F, Cusi D, Roberts R, Newton-Cheh C, Franke L, Stanton AV, Dominiczak AF, Farrall M, Hingorani AD, Samani NJ, Caulfield MJ, Munroe PB. Blood pressure loci identified with a gene-centric array. Am J Hum Genet 89: 688 –700, 2011. Kapoor A, Sekar RB, Hansen NF, Fox-Talbot K, Morley M, Pihur V, Chatterjee S, Brandimarto J, Moravec CS, Pulit SL, Pfeufer A, Mullikin J, Ross M, Green ED, Bentley D, Newton-Cheh C, Boerwinkle E, Tomaselli GF, Cappola TP, Arking DE, Halushka MK, Chakravarti A. An enhancer polymorphism at the cardiomyocyte intercalated disc protein NOS1AP locus is a major regulator of the QT interval. Am J Hum Genet 94: 854 –869, 2014. Karra E, O’Daly OG, Choudhury AI, Yousseif A, Millership S, Neary MT, Scott WR, Chandarana K, Manning S, Hess ME, Iwakura H, Akamizu T, Millet Q, Gelegen C, Drew ME, Rahman S, Emmanuel JJ, Williams SC, Ruther UU, Bruning JC, Withers DJ, Zelaya FO, Batterham RL. A link between FTO, ghrelin, and impaired brain foodcue responsivity. J Clin Invest 123: 3539 –3551, 2013. Kato N, Takeuchi F, Tabara Y, Kelly TN, Go MJ, Sim X, Tay WT, Chen CH, Zhang Y, Yamamoto K, Katsuya T, Yokota M, Kim YJ, Ong RT, Nabika T, Gu D, Chang LC, Kokubo Y, Huang W, Ohnaka K, Yamori Y, Nakashima E, Jaquish CE, Lee JY, Seielstad M, Isono

34. 35.

36.

37.

38. 39. 40.

41.

42. 43.

44.

373

M, Hixson JE, Chen YT, Miki T, Zhou X, Sugiyama T, Jeon JP, Liu JJ, Takayanagi R, Kim SS, Aung T, Sung YJ, Zhang X, Wong TY, Han BG, Kobayashi S, Ogihara T, Zhu D, Iwai N, Wu JY, Teo YY, Tai ES, Cho YS, He J. Meta-analysis of genome-wide association studies identifies common variants associated with blood pressure variation in east Asians. Nat Genet 43: 531–538, 2011. Keenan TE, Rader DJ. Genetics of lipid traits and relationship to coronary artery disease. Curr Cardiol Rep 15: 396, 2013. Kettleborough RN, Busch-Nentwich EM, Harvey SA, Dooley CM, de BE, van EF, Sealy I, White RJ, Herd C, Nijman IJ, Fenyes F, Mehroke S, Scahill C, Gibbons R, Wali N, Carruthers S, Hall A, Yen J, Cuppen E, Stemple DL. A systematic genome-wide analysis of zebrafish protein-coding gene function. Nature 496: 494 –497, 2013. Kottgen A, Glazer NL, Dehghan A, Hwang SJ, Katz R, Li M, Yang Q, Gudnason V, Launer LJ, Harris TB, Smith AV, Arking DE, Astor BC, Boerwinkle E, Ehret GB, Ruczinski I, Scharpf RB, Chen YD, de Boer IH, Haritunians T, Lumley T, Sarnak M, Siscovick D, Benjamin EJ, Levy D, Upadhyay A, Aulchenko YS, Hofman A, Rivadeneira F, Uitterlinden AG, van Duijn CM, Chasman DI, Pare G, Ridker PM, Kao WH, Witteman JC, Coresh J, Shlipak MG, Fox CS. Multiple loci associated with indices of renal function and chronic kidney disease. Nat Genet 41: 712–717, 2009. Kottgen A, Hwang SJ, Larson MG, Van Eyk JE, Fu Q, Benjamin EJ, Dehghan A, Glazer NL, Kao WH, Harris TB, Gudnason V, Shlipak MG, Yang Q, Coresh J, Levy D, Fox CS. Uromodulin levels associate with a common UMOD variant and risk for incident CKD. J Am Soc Nephrol 21: 337–344, 2010. Kumarasamy S, Waghulde H, Gopalakrishnan K, Mell B, Morgan E, Joe B. Mutation within the hinge region of the transcription factor Nr2f2 attenuates salt-sensitive hypertension. Nat Commun 6: 6252, 2015. Landry DW, Oliver JA. The pathogenesis of vasodilatory shock. N Engl J Med 345: 588 –595, 2001. Lappalainen T, Kolehmainen M, Schwab U, Pulkkinen L, de Mello VD, Vaittinen M, Laaksonen DE, Poutanen K, Uusitupa M, Gylling H. Gene expression of FTO in human subcutaneous adipose tissue, peripheral blood mononuclear cells and adipocyte cell line. J Nutrigenet Nutrigenom 3: 37–45, 2010. Mahajan A, Sim X, Ng HJ, Manning A, Rivas MA, Highland HM, Locke AE, Grarup N, Im HK, Cingolani P, Flannick J, Fontanillas P, Fuchsberger C, Gaulton KJ, Teslovich TM, Rayner NW, Robertson NR, Beer NL, Rundle JK, Bork-Jensen J, Ladenvall C, Blancher C, Buck D, Buck G, Burtt NP, Gabriel S, Gjesing AP, Groves CJ, Hollensted M, Huyghe JR, Jackson AU, Jun G, Justesen JM, Mangino M, Murphy J, Neville M, Onofrio R, Small KS, Stringham HM, Syvanen AC, Trakalo J, Abecasis G, Bell GI, Blangero J, Cox NJ, Duggirala R, Hanis CL, Seielstad M, Wilson JG, Christensen C, Brandslund I, Rauramaa R, Surdulescu GL, Doney AS, Lannfelt L, Linneberg A, Isomaa B, Tuomi T, Jorgensen ME, Jorgensen T, Kuusisto J, Uusitupa M, Salomaa V, Spector TD, Morris AD, Palmer CN, Collins FS, Mohlke KL, Bergman RN, Ingelsson E, Lind L, Tuomilehto J, Hansen T, Watanabe RM, Prokopenko I, Dupuis J, Karpe F, Groop L, Laakso M, Pedersen O, Florez JC, Morris AP, Altshuler D, Meigs JB, Boehnke M, McCarthy MI, Lindgren CM, Gloyn AL, T2D-GENES Consortium, GoT2D Consortium. Identification and functional characterization of G6PC2 coding variants influencing glycemic traits define an effector transcript at the G6PC2-ABCB11 locus. PLoS Genet 11: e1004876, 2015. Mali P, Esvelt KM, Church GM. Cas9 as a versatile tool for engineering biology. Nat Meth 10: 957–963, 2013. Manolio TA, Collins FS, Cox NJ, Goldstein DB, Hindorff LA, Hunter DJ, McCarthy MI, Ramos EM, Cardon LR, Chakravarti A, Cho JH, Guttmacher AE, Kong A, Kruglyak L, Mardis E, Rotimi CN, Slatkin M, Valle D, Whittemore AS, Boehnke M, Clark AG, Eichler EE, Gibson G, Haines JL, Mackay TF, McCarroll SA, Visscher PM. Finding the missing heritability of complex diseases. Nature 461: 747– 753, 2009. Maurano MT, Humbert R, Rynes E, Thurman RE, Haugen E, Wang H, Reynolds AP, Sandstrom R, Qu H, Brody J, Shafer A, Neri F, Lee K, Kutyavin T, Stehling-Sun S, Johnson AK, Canfield TK, Giste E, Diegel M, Bates D, Hansen RS, Neph S, Sabo PJ, Heimfeld S, Raubitschek A, Ziegler S, Cotsapas C, Sotoodehnia N, Glass I, Sunyaev SR, Kaul R, Stamatoyannopoulos JA. Systematic localization of common disease-associated variation in regulatory DNA. Science 337: 1190 –1195, 2012.

Physiol Genomics • doi:10.1152/physiolgenomics.00004.2015 • www.physiolgenomics.org

Review 374

CARDIOVASCULAR PHYSIOLOGY AND GWAS

45. Montague TG, Cruz JM, Gagnon JA, Church GM, Valen E. CHOPCHOP: a CRISPR/Cas9 and TALEN web tool for genome editing. Nucleic Acids Res 42: W401–W407, 2014. 46. Moretti A, Bellin M, Welling A, Jung CB, Lam JT, Bott-Flugel L, Dorn T, Goedel A, Hohnke C, Hofmann F, Seyfarth M, Sinnecker D, Schomig A, Laugwitz KL. Patient-specific induced pluripotent stem-cell models for long-QT syndrome. N Engl J Med 363: 1397–1409, 2010. 47. Morris AP, Voight BF, Teslovich TM, Ferreira T, Segre AV, Steinthorsdottir V, Strawbridge RJ, Khan H, Grallert H, Mahajan A, Prokopenko I, Kang HM, Dina C, Esko T, Fraser RM, Kanoni S, Kumar A, Lagou V, Langenberg C, Luan J, Lindgren CM, MullerNurasyid M, Pechlivanis S, Rayner NW, Scott LJ, Wiltshire S, Yengo L, Kinnunen L, Rossin EJ, Raychaudhuri S, Johnson AD, Dimas AS, Loos RJ, Vedantam S, Chen H, Florez JC, Fox C, Liu CT, Rybin D, Couper DJ, Kao WH, Li M, Cornelis MC, Kraft P, Sun Q, van Dam RM, Stringham HM, Chines PS, Fischer K, Fontanillas P, Holmen OL, Hunt SE, Jackson AU, Kong A, Lawrence R, Meyer J, Perry JR, Platou CG, Potter S, Rehnberg E, Robertson N, Sivapalaratnam S, Stancakova A, Stirrups K, Thorleifsson G, Tikkanen E, Wood AR, Almgren P, Atalay M, Benediktsson R, Bonnycastle LL, Burtt N, Carey J, Charpentier G, Crenshaw AT, Doney AS, Dorkhan M, Edkins S, Emilsson V, Eury E, Forsen T, Gertow K, Gigante B, Grant GB, Groves CJ, Guiducci C, Herder C, Hreidarsson AB, Hui J, James A, Jonsson A, Rathmann W, Klopp N, Kravic J, Krjutskov K, Langford C, Leander K, Lindholm E, Lobbens S, Mannisto S, Mirza G, Muhleisen TW, Musk B, Parkin M, Rallidis L, Saramies J, Sennblad B, Shah S, Sigurethsson G, Silveira A, Steinbach G, Thorand B, Trakalo J, Veglia F, Wennauer R, Winckler W, Zabaneh D, Campbell H, van Duijn C, Uitterlinden AG, Hofman A, Sijbrands E, Abecasis GR, Owen KR, Zeggini E, Trip MD, Forouhi NG, Syvanen AC, Eriksson JG, Peltonen L, Nothen MM, Balkau B, Palmer CN, Lyssenko V, Tuomi T, Isomaa B, Hunter DJ, Qi L, Shuldiner AR, Roden M, Barroso I, Wilsgaard T, Beilby J, Hovingh K, Price JF, Wilson JF, Rauramaa R, Lakka TA, Lind L, Dedoussis G, Njolstad I, Pedersen NL, Khaw KT, Wareham NJ, Keinanen-Kiukaanniemi SM, Saaristo TE, Korpi-Hyovalti E, Saltevo J, Laakso M, Kuusisto J, Metspalu A, Collins FS, Mohlke KL, Bergman RN, Tuomilehto J, Boehm BO, Gieger C, Hveem K, Cauchi S, Froguel P, Baldassarre D, Tremoli E, Humphries SE, Saleheen D, Danesh J, Ingelsson E, Ripatti S, Salomaa V, Erbel R, Jockel KH, Moebus S, Peters A, Illig T, de Faire U, Hamsten A, Morris AD, Donnelly PJ, Frayling TM, Hattersley AT, Boerwinkle E, Melander O, Kathiresan S, Nilsson PM, Deloukas P, Thorsteinsdottir U, Groop LC, Stefansson K, Hu F, Pankow JS, Dupuis J, Meigs JB, Altshuler D, Boehnke M, McCarthy MI. Large-scale association analysis provides insights into the genetic architecture and pathophysiology of type 2 diabetes. Nat Genet 44: 981–990, 2012. 48. Mutig K, Kahl T, Saritas T, Godes M, Persson P, Bates J, Raffi H, Rampoldi L, Uchida S, Hille C, Dosche C, Kumar S, Castaneda-Bueno M, Gamba G, Bachmann S. Activation of the bumetanide-sensitive Na⫹,K⫹,2Cl- cotransporter (NKCC2) is facilitated by Tamm-Horsfall protein in a chloride-sensitive manner. J Biol Chem 286: 30200 –30210, 2011. 50. Newton-Cheh C, Johnson T, Gateva V, Tobin MD, Bochud M, Coin L, Najjar SS, Zhao JH, Heath SC, Eyheramendy S, Papadakis K, Voight BF, Scott LJ, Zhang F, Farrall M, Tanaka T, Wallace C, Chambers JC, Khaw KT, Nilsson P, van der Harst P, Polidoro S, Grobbee DE, Onland-Moret NC, Bots ML, Wain LV, Elliott KS, Teumer A, Luan J, Lucas G, Kuusisto J, Burton PR, Hadley D, McArdle WL, Brown M, Dominiczak A, Newhouse SJ, Samani NJ, Webster J, Zeggini E, Beckmann JS, Bergmann S, Lim N, Song K, Vollenweider P, Waeber G, Waterworth DM, Yuan X, Groop L, Orho-Melander M, Allione A, Di GA, Guarrera S, Panico S, Ricceri F, Romanazzi V, Sacerdote C, Vineis P, Barroso I, Sandhu MS, Luben RN, Crawford GJ, Jousilahti P, Perola M, Boehnke M, Bonnycastle LL, Collins FS, Jackson AU, Mohlke KL, Stringham HM, Valle TT, Willer CJ, Bergman RN, Morken MA, Doring A, Gieger C, Illig T, Meitinger T, Org E, Pfeufer A, Wichmann HE, Kathiresan S, Marrugat J, O’Donnell CJ, Schwartz SM, Siscovick DS, Subirana I, Freimer NB, Hartikainen AL, McCarthy MI, O’Reilly PF, Peltonen L, Pouta A, de Jong PE, Snieder H, van Gilst WH, Clarke R, Goel A, Hamsten A, Peden JF, Seedorf U, Syvanen AC, Tognoni G, Lakatta EG, Sanna S, Scheet P, Schlessinger D, Scuteri A, Dorr M, Ernst F, Felix SB, Homuth G, Lorbeer R, Reffelmann T, Rettig R, Volker U, Galan P, Gut IG,

51.

52. 53.

54. 55. 56.

57.

58.

59.

60. 61. 62.

63.

Hercberg S, Lathrop GM, Zelenika D, Deloukas P, Soranzo N, Williams FM, Zhai G, Salomaa V, Laakso M, Elosua R, Forouhi NG, Volzke H, Uiterwaal CS, van der Schouw YT, Numans ME, Matullo G, Navis G, Berglund G, Bingham SA, Kooner JS, Connell JM, Bandinelli S, Ferrucci L, Watkins H, Spector TD, Tuomilehto J, Altshuler D, Strachan DP, Laan M, Meneton P, Wareham NJ, Uda M, Jarvelin MR, Mooser V, Melander O, Loos RJ, Elliott P, Abecasis GR, Caulfield M, Munroe PB. Genome-wide association study identifies eight loci associated with blood pressure. Nat Genet 41: 666 –676, 2009. Newton-Cheh C, Larson MG, Vasan RS, Levy D, Bloch KD, Surti A, Guiducci C, Kathiresan S, Benjamin EJ, Struck J. Association of common variants in NPPA and NPPB with circulating natriuretic peptides and blood pressure. Nat Genet 41: 348 –353, 2009. Padmanabhan S, Graham L, Ferreri NR, Graham D, McBride M, Dominiczak AF. Uromodulin, an emerging novel pathway for blood pressure regulation and hypertension. Hypertension 64: 918 –923, 2014. Padmanabhan S, Melander O, Johnson T, Di Blasio AM, Lee WK, Gentilini D, Hastie CE, Menni C, Monti MC, Delles C, Laing S, Corso B, Navis G, Kwakernaak AJ, van der Harst P, Bochud M, Maillard M, Burnier M, Hedner T, Kjeldsen S, Wahlstrand B, Sjogren M, Fava C, Montagnana M, Danese E, Torffvit O, Hedblad B, Snieder H, Connell JM, Brown M, Samani NJ, Farrall M, Cesana G, Mancia G, Signorini S, Grassi G, Eyheramendy S, Wichmann HE, Laan M, Strachan DP, Sever P, Shields DC, Stanton A, Vollenweider P, Teumer A, Volzke H, Rettig R, Newton-Cheh C, Arora P, Zhang F, Soranzo N, Spector TD, Lucas G, Kathiresan S, Siscovick DS, Luan J, Loos RJ, Wareham NJ, Penninx BW, Nolte IM, McBride M, Miller WH, Nicklin SA, Baker AH, Graham D, McDonald RA, Pell JP, Sattar N, Welsh P, Munroe P, Caulfield MJ, Zanchetti A, Dominiczak AF. Genome-wide association study of blood pressure extremes identifies variant near UMOD associated with hypertension. PLoS Genet 6: e1001177, 2010. Pennica D, Kohr WJ, Kuang WJ, Glaister D, Aggarwal BB, Chen EY, Goeddel DV. Identification of human uromodulin as the Tamm-Horsfall urinary glycoprotein. Science 236: 83–88, 1987. Pera MF. Stem cells: the dark side of induced pluripotency. Nature 471: 46 –47, 2011. Pfeufer A, Sanna S, Arking DE, Muller M, Gateva V, Fuchsberger C, Ehret GB, Orru M, Pattaro C, Kottgen A, Perz S, Usala G, Barbalic M, Li M, Putz B, Scuteri A, Prineas RJ, Sinner MF, Gieger C, Najjar SS, Kao WH, Muhleisen TW, Dei M, Happle C, Mohlenkamp S, Crisponi L, Erbel R, Jockel KH, Naitza S, Steinbeck G, Marroni F, Hicks AA, Lakatta E, Muller-Myhsok B, Pramstaller PP, Wichmann HE, Schlessinger D, Boerwinkle E, Meitinger T, Uda M, Coresh J, Kaab S, Abecasis GR, Chakravarti A. Common variants at ten loci modulate the QT interval duration in the QTSCD Study. Nat Genet 41: 407–414, 2009. Renigunta A, Renigunta V, Saritas T, Decher N, Mutig K, Waldegger S. Tamm-Horsfall glycoprotein interacts with renal outer medullary potassium channel ROMK2 and regulates its function. J Biol Chem 286: 2224 –2235, 2011. Rudemiller NP, Lund H, Priestley JR, Endres BT, Prokop JW, Jacob HJ, Geurts AM, Cohen EP, Mattson DL. Mutation of SH2B3 (LNK), a genome-wide association study candidate for hypertension, attenuates Dahl salt-sensitive hypertension via inflammatory modulation. Hypertension 65: 1111–1117, 2015. Salvi E, Kutalik Z, Glorioso N, Benaglio P, Frau F, Kuznetsova T, Arima H, Hoggart C, Tichet J, Nikitin YP. Genomewide association study using a high-density single nucleotide polymorphism array and case-control design identifies a novel essential hypertension susceptibility locus in the promoter region of endothelial NO synthase. Hypertension 59: 248 –255, 2012. Schwartz PJ, Crotti L, Insolia R. Long-QT syndrome: from genetics to management. Circ Arrhythm Electrophysiol 5: 868 –877, 2012. Service S, Sabatti C, Freimer N. Tag SNPs chosen from HapMap perform well in several population isolates. Genet Epidemiol 31: 189 –194, 2007. Shimoyama M, De Pons J, Hayman GT, Laulederkind SJ, Liu W, Nigam R, Petri V, Smith JR, Tutaj M, Wang SJ, Worthey E, Dwinell M, Jacob H. The Rat Genome Database 2015: genomic, phenotypic and environmental variations and disease. Nucleic Acids Res 43: D743–750, 2015. Skarnes WC, Rosen B, West AP, Koutsourakis M, Bushell W, Iyer V, Mujica AO, Thomas M, Harrow J, Cox T, Jackson D, Severin J, Biggs P, Fu J, Nefedov M, de Jong PJ, Stewart AF, Bradley A. A conditional

Physiol Genomics • doi:10.1152/physiolgenomics.00004.2015 • www.physiolgenomics.org

Review CARDIOVASCULAR PHYSIOLOGY AND GWAS

64.

65. 66. 67.

68.

69. 70.

knockout resource for the genome-wide study of mouse gene function. Nature 474: 337–342, 2011. Smemo S, Tena JJ, Kim KH, Gamazon ER, Sakabe NJ, Gomez-Marin C, Aneas I, Credidio FL, Sobreira DR, Wasserman NF, Lee JH, Puviindran V, Tam D, Shen M, Son JE, Vakili NA, Sung HK, Naranjo S, Acemel RD, Manzanares M, Nagy A, Cox NJ, Hui CC, GomezSkarmeta JL, Nobrega MA. Obesity-associated variants within FTO form long-range functional connections with IRX3. Nature 507: 371–375, 2014. Spyer KM. Annual review prize lecture. Central nervous mechanisms contributing to cardiovascular control. J Physiol 474: 1–19, 1994. Takahashi K, Tanabe K, Ohnuki M, Narita M, Ichisaka T, Tomoda K, Yamanaka S. Induction of pluripotent stem cells from adult human fibroblasts by defined factors. Cell 131: 861–872, 2007. Terrenoire C, Wang K, Tung KW, Chung WK, Pass RH, Lu JT, Jean JC, Omari A, Sampson KJ, Kotton DN, Keller G, Kass RS. Induced pluripotent stem cells used to reveal drug actions in a long QT syndrome family with complex genetics. J Gen Physiol 141: 61–72, 2013. Tomaszewski M, Debiec R, Braund PS, Nelson CP, Hardwick R, Christofidou P, Denniff M, Codd V, Rafelt S, van der Harst P, Waterworth D, Song K, Vollenweider P, Waeber G, ZukowskaSzczechowska E, Burton PR, Mooser V, Charchar FJ, Thompson JR, Tobin MD, Samani NJ. Genetic architecture of ambulatory blood pressure in the general population: insights from cardiovascular gene-centric array. Hypertension 56: 1069 –1076, 2010. Torffvit O, Melander O, Hulten UL. Urinary excretion rate of TammHorsfall protein is related to salt intake in humans. Nephron Physiol 97: p31–36, 2004. Trudu M, Janas S, Lanzani C, Debaix H, Schaeffer C, Ikehata M, Citterio L, Demaretz S, Trevisani F, Ristagno G, Glaudemans B,

71.

72.

73.

74.

375

Laghmani K, Dell’Antonio G, Loffing J, Rastaldi MP, Manunta P, Devuyst O, Rampoldi L. Common noncoding UMOD gene variants induce salt-sensitive hypertension and kidney damage by increasing uromodulin expression. Nat Med 19: 1655–1660, 2013. Voight BF, Kang HM, Ding J, Palmer CD, Sidore C, Chines PS, Burtt NP, Fuchsberger C, Li Y, Erdmann J, Frayling TM, Heid IM, Jackson AU, Johnson T, Kilpelainen TO, Lindgren CM, Morris AP, Prokopenko I, Randall JC, Saxena R, Soranzo N, Speliotes EK, Teslovich TM, Wheeler E, Maguire J, Parkin M, Potter S, Rayner NW, Robertson N, Stirrups K, Winckler W, Sanna S, Mulas A, Nagaraja R, Cucca F, Barroso I, Deloukas P, Loos RJ, Kathiresan S, Munroe PB, NewtonCheh C, Pfeufer A, Samani NJ, Schunkert H, Hirschhorn JN, Altshuler D, McCarthy MI, Abecasis GR, Boehnke M. The metabochip, a custom genotyping array for genetic studies of metabolic, cardiovascular, and anthropometric traits. PLoS Genet 8: e1002793, 2012. White JK, Gerdin AK, Karp NA, Ryder E, Buljan M, Bussell JN, Salisbury J, Clare S, Ingham NJ, Podrini C, Houghton R, Estabel J, Bottomley JR, Melvin DG, Sunter D, Adams NC, Tannahill D, Logan DW, Macarthur DG, Flint J, Mahajan VB, Tsang SH, Smyth I, Watt FM, Skarnes WC, Dougan G, Adams DJ, RamirezSolis R, Bradley A, Steel KP. Genome-wide generation and systematic phenotyping of knockout mice reveals new roles for many genes. Cell 154: 452–464, 2013. Wilson AJ, Quinn KV, Graves FM, Bitner-Glindzicz M, Tinker A. Abnormal KCNQ1 trafficking influences disease pathogenesis in hereditary long QT syndromes (LQT1). Cardiovasc Res 67: 476 –486, 2005. Zuk O, Hechter E, Sunyaev SR, Lander ES. The mystery of missing heritability: Genetic interactions create phantom heritability. Proc Natl Acad Sci USA 109: 1193–1198, 2012.

Physiol Genomics • doi:10.1152/physiolgenomics.00004.2015 • www.physiolgenomics.org

Genome-wide association studies and contribution to cardiovascular physiology.

The study of family pedigrees with rare monogenic cardiovascular disorders has revealed new molecular players in physiological processes. Genome-wide ...
630KB Sizes 0 Downloads 10 Views