Skip to main content
Research Article
Originally Published 19 February 2016
Free Access

Genetics of Coronary Artery Disease

Abstract

Genetic factors contribute importantly to the risk of coronary artery disease (CAD), and in the past decade, there has been major progress in this area. The tools applied include genome-wide association studies encompassing >200 000 individuals complemented by bioinformatic approaches, including 1000 Genomes imputation, expression quantitative trait locus analyses, and interrogation of Encyclopedia of DNA Elements, Roadmap, and other data sets. close to 60 common SNPs (minor allele frequency>0.05) associated with CAD risk and reaching genome-wide significance (P<5×10−8) have been identified. Furthermore, a total of 202 independent signals in 109 loci have achieved a false discovery rate (q<0.05) and together explain 28% of the estimated heritability of CAD. These data have been used successfully to create genetic risk scores that can improve risk prediction beyond conventional risk factors and identify those individuals who will benefit most from statin therapy. Such information also has important applications in clinical medicine and drug discovery by using a Mendelian randomization approach to interrogate the causal nature of many factors found to associate with CAD risk in epidemiological studies. In contrast to genome-wide association studies, whole-exome sequencing has provided valuable information directly relevant to genes with known roles in plasma lipoprotein metabolism but has, thus far, failed to identify other rare coding variants linked to CAD. Overall, recent studies have led to a broader understanding of the genetic architecture of CAD and demonstrate that it largely derives from the cumulative effect of multiple common risk alleles individually of small effect size rather than rare variants with large effects on CAD risk. Despite this success, there has been limited progress in understanding the function of the novel loci; the majority of which are in noncoding regions of the genome.

Introduction

Coronary artery disease (CAD) has important genetic underpinnings considered equivalent to that of environmental factors. The heritability of CAD has been estimated between 40% and 60%, on the basis of family and twin studies, a method that yields high precision despite potential bias (Vinkhuyzen et al1). In the Framingham Offspring Study, the age-specific incidence of CAD increased by >2-fold after adjustment for conventional CAD risk factors in participants with a family history of premature disease.2 The Swedish Twin registry reported on close to 21 000 subjects followed up for >35 years and calculated the heritability of fatal CAD events to be 0.57 and 0.38, for men and women, respectively. Of note, heritable effects are most manifest in younger individuals.3 This accords with other data, indicating that the genetic influence is the greatest for early-onset CAD events.4
Unraveling the genetic basis of CAD has evolved slowly during the past half-century. Candidate genes coding for proteins of known biological significance in a disease process seemed to provide a logical first step in understanding the genetics of common disease states. Populations of affected and unaffected individuals could be studied by genotyping common single-nucleotide polymorphisms (SNPs) within a gene and its regulatory sequences. Although economically attractive, this approach had many limitations. By definition, studies are limited to genes with a known or suspected role in defining a given phenotype and do not provide new insight into biological pathways leading to disease. Furthermore, candidate gene associations generally failed to replicate for multiple reasons, including statistical power related to inadequate, sample size, heterogeneity of causality, and population stratification.5 With the exception of rare monogenic disorders of lipid metabolism, such as familial hypercholesterolemia, there is as yet little support for a role of single-gene disorders in coronary atherosclerosis or plaque rupture as discussed below.

Genome-Wide Association Studies

Recent progress in understanding the genetics of CAD and other complex diseases has been driven by technological advances, including high-throughput DNA microarray technology using chips containing up to a million DNA markers consisting of single-nucleotide polymorphisms (SNPs). In the commercial arrays used for genome-wide association studies (GWASs), SNPs (generally 0.5–1 mol/L) are used to tag common variation (SNPs present in ≥5% of the population) across the human genome. Importantly, these are tag SNPs that point to a causative locus but are rarely functional variants. This approach makes use of linkage disequilibrium (LD) that is the nonrandom coinheritance of genetic variants across the human genome. Comparison of the allele frequency of each SNP in cases and controls as part of a GWAS provides an agnostic approach that involves no underlying assumption on candidate genes or pathways. However, the statistical bar for association is high; comparison of a million markers implies a P value of ≤5×10−8 for GWAS significance. Thus, success has been contingent on the allele frequency and effect size of a given variant and the recruitment of thousands of carefully phenotyped individuals with or without clinical evidence of CAD and collaboration among many groups across the world.
The first robust association with CAD identified by the GWAS approach, a 53-kb LD block containing multiple highly correlated SNPs at 9p21.3 locus, was identified by 3 independent groups in 2007.68 The early discovery of this risk locus was facilitated by its large effect size (>1.3) and risk allele frequency (≈0.48). Approximately 25% of Europeans carry 2 copies of the risk allele and have a 40% increased risk of CAD in general and a 2-fold risk of premature CAD. Here, it is notable that in the Ottawa Heart Study,6 an extreme phenotype approach was used comparing young patients with multivessel CAD to elderly, physically active, asymptomatic subjects. Although the discovery population consisted of only 322 cases and 311 controls, the lead SNP rs10757274 achieved a P value of 3.7×10−6. Consistently, in the large Coronary Artery Disease Genome-wide Replication and Meta-Analysis (CARDIoGRAM) study, the allele-specific odds ratio (OR) for CAD in subjects with onset before 50 years of age was 1.41 (confidence interval, 1.34–1.48) versus 1.24 (confidence interval, 1.20–1.28)4 for older individuals.4 The 9p21 locus is also associated with the extent and severity of atherosclerosis9,10 and with a higher risk allele frequency in subjects with multivessel disease.9 It has been consistently shown that the risk conferred by this locus is independent of known risk factors, including plasma lipids, blood pressure, diabetes mellitus, obesity, markers of inflammation, age, and sex. Several other vascular phenotypes are associated with the 9p21 risk alleles, including carotid atherosclerosis,11 stroke,12,13 peripheral arterial disease,14 and abdominal aortic aneurysm15 consistent with an effect on atherosclerosis. However, there is an important association with intracranial aneurysms,16 thus highlighting the possible effects on vascular wall integrity. Distinct haplotype blocks are associated with platelet reactivity17 and myocardial infarction versus atherosclerosis.18 There is also a surprising association with periodontitis.19,20
Since the 9p21 discovery in 2007, large meta-analyses of additional GWASs, the majority of which have been conducted in individuals of European descent, have identified additional loci of smaller effect size but with genome-wide (P<5×10−8) significance. This success has been built on large collaborative efforts, including the Myocardial Infarction Genomics Consortium,21 the CARDIoGRAM consortium,4 the Coronary Artery Disease (C4D) Genetics Consortium,22 CARDIoGRAMplusC4D,23 and others. The risk loci identified include genes known to function in lipoprotein metabolism, hypertension, and other CAD-associated phenotypes but, importantly, include several novel regions harboring genes of unknown relevance to atherosclerosis and plaque rupture, thus highlighting the discovery potential of the GWAS approach. Several risk regions exhibit pleiotropic effects; for example, ABO and SH2B3 associate with multiple CAD and non–CAD-related phenotypes. With the exception of the LPA gene24 encoding lipoprotein(a) and the 9p21.3 locus, the allele-specific ORs for CAD of replicated loci are <1.15, as might be expected for common variants affecting a complex trait.
GWASs of CAD have been designed to test the common disease–common variant hypothesis, and most have applied imputation from HapMap training sets interrogating mainly SNPs present in >5% of the population (minor allele frequency [MAF]>0.05). In contrast to HapMap, the 1000 Genomes Project phase 1, version 3 data set includes >38 million variants; half of which are present in <1 in 200 individuals (MAF<0.005), as well as insertions and deletions (indels). These data were leveraged in a recent meta-analysis led by Martin Farrall that included >185 000 individuals (60 801 CAD cases and 123 504 controls) from 48 studies providing imputed data on 9.4 million SNPs, including 2.7 million low-frequency SNPs (MAF, 0.005–0.05). The summary Manhattan plot is shown in Figure 1. By using both a recessive and an additive model of inheritance, 2 novel loci were identified that would have been missed by conventional analysis. In total, this analysis identified 10 novel CAD loci at GWAS significance, bringing the total number of replicated loci in European populations to 58 (Table 1). Here, it is notable that <25% of the significant loci are related to known CAD risk factors, highlighting the discovery potential of the GWAS approach. Joint association analysis using genome-wide complex trait analysis software revealed no evidence that the signal conferred by common SNPs was because of synthetic associations, that is, linkage with a rare variant(s) of high effect size.25
Table 1. Replicated Genome-Wide Significant Loci for CAD
ChrClosest Gene(s)Putative Functions of Possible Relevance to CADLead SNPEAFOR
1PPAP2BRegulation of cell–cell interactionsrs171140360.921.13
1PCSK9Regulation of LDL receptor recyclingrs112065100.851.08
1SORT1Regulate apoB secretion and LDL catabolismrs5998390.781.11
1IL6RIL-6 receptor, immune responsers48456250.461.05
1MIA3Collagen secretionrs174656370.731.08
2LINC00954LncRNA of unknown functionrs169869530.071.09
2APOBMajor apolipoprotein of LDLrs5151350.781.07
2ABCG5/G8Cholesterol absorption and secretionrs65447130.291.05
2VAMP5/8-GGCXIntracellular vesicle traffickingrs15611980.471.06
2ZEB2-AC074093.1ZEB2-transcriptional repressorrs22526410.441.03
2WDR12Component of nucleolar protein complexrs67258870.141.14
3MRASCell growth and differentiationrs98188700.151.07
4EDNRAReceptor for endothelin—vasoconstrictionrs18784060.161.06
4GUCY1A3Nitric oxide signalingrs76923870.801.07
4REST-NOA1REST maintains VSMCs in a quiescent staters170873350.211.06
5SLC22A4/A5Organic cation transporterrs2739090.141.06
6ANKS1AMay inhibit PDGF-induced mitogenesisrs176099400.821.03
6PHACTR1Regulates protein phosphatase 1 activityrs125264530.711.10
6KCNK5Potassium channel proteinrs109477890.781.05
6TCF21Transcriptional regulatorrs121902870.641.06
6SLC22A3-LPAL2-LPALipoprotein(a)rs20483270.351.06
   rs37892200.021.42
6PLGFibrinolysisrs42521200.741.03
7NOS3Production of nitric oxiders39182260.061.14
7HDAC9Represses MEF2 activity/beige adipogenesisrs20239380.101.08
7ZC3HC1Encodes NIPA, regulator of cell proliferationrs115569240.691.08
8LPLLipolysis of TG-rich lipoproteinsrs2640.851.06
8TRIB1TG, MAPK signaling, SMC proliferationrs29540690.551.04
9CDKN2BASCellular proliferation, platelet functionrs107572740.481.21
9ABOIL-6, E-selectin, LDL-C levelsrs5794590.211.08
10KIAA1462Component of endothelial cell–cell junctionsrs25050830.401.07
10CXCL12Endothelial regeneration; neutrophil migrationrs5011200.811.08
   rs20470090.481.06
10LIPAIntracellular hydrolysis of cholesteryl estersrs14124440.371.07
10  rs112030420.451.04
10CYP17A1-CNNM2-NT5C2CYP17A1: steroidogenic pathwayrs124134090.891.08
11PDGFDRole in SMC proliferationrs9748190.331.07
11SWAP70Leukocyte and VSMC migration and adhesionrs108402930.551.06
11ZNF259 APOA5 APOC3TG-rich lipoprotein metabolismrs9641840.181.05
12SH2B3Negative regulator of cytokine signalingrs31845040.421.07
12ATP2B1Intracellular calcium homeostasisRs71362590.431.04
12KSR2Suppressor of Ras2–cell proliferation; obesityrs118301570.361.12*
13FLT1VEGFR family; angiogenesisRs93194280.311.04
13COL4A1/A2Type IV collagen chain of basement membraners47731440.431.05
   rs95152030.761.07
14HHIPL1Unknownrs28958110.411.04
15ADAMTS7Proliferative response to vascular injuryrs71737430.561.08
15SMAD3Downstream mediator of TGF-β signalingrs560621350.791.07
15MFGE8-ABHD2MFGE8: lactadherin–VEGF neovascularizationrs80422710.901.10
15FURINEndoprotease—TGF-β1 precursor and type I MMPrs175148460.441.05
17BCAS3Rudhira—EC polarity and angiogenesisrs72127980.151.08
17RAI1-PEMT-RASD1PEMT encoded protein converts PE to PCrs129365870.611.03
17SMG6Role in nonsense mediated RNA decayrs2161720.351.05
17UBE2ZProtein ubiquination; apoptosisrs465220.511.04
18PMAIP1-MC4RPMAIP1: HIF1A-induced proapoptotic gene; MC4R: leptin signaling—obesityrs6631290.261.06
19LDLRLDL clearancers11226080.771.08
19APOELDL and VLDL clearancers44206380.171.10
19ZNF507Unknownrs129764110.090.67*
21KCNE2Maintains cardiac electric stabilityrs99826010.131.12
22POM121L9P-ADORA2AAdenosine A2a receptor: infarct-sparing effectsrs1808030.971.20
CAD indicates coronary artery disease; Chr, chromosome; EAF, effect allele frequency; EC, endothelial cells; HIF1A, hypoxia inducible factor 1A; IL, interleukin; LDL, low-density lipoprotein; LDL-C, low-density lipoprotein cholesterol; LncRNA, long noncoding RNA; MAPK, mitogen-activated protein kinase; MEF2, myocyte enhancer factor 2; MMP, matrix metalloproteinase; NIPA, nuclear interacting partner of anaplastic lymphoma kinase; OR, odds ratio; PC, phosphatidylcholine; PDGF, platelet-derived growth factor; PE, phosphatidylethanolamine; REST, RE-1 silencing transcription factor; SMC, smooth muscle cell; SNP, single-nucleotide polymorphism; TG, triglyceride; TGF-β, transforming growth factor-β; VEGFR, vascular endothelial growth factor receptor; VLDL, very low-density lipoprotein; and VSMC, vascular smooth muscle cells.
*
By recessive model.
Figure 1. Manhattan plot of genome-wide association study (GWAS) meta-analysis that included ≈185 000 coronary artery disease (CAD) cases and controls from 48 studies providing 1000 Genomes–imputed data on 9.4 million single-nucleotide polymorphisms (SNPs), including 2.7 million low-frequency SNPs (minor allele frequency, 0.005–0.05). Two novel loci were identified that would have been missed by conventional analysis. The meta-analysis statistics have been adjusted for overdispersion (genomic control parameter, 1.18) and have been capped to P=1×10−20. The genome-wide significance threshold is shown as a horizontal blue line at P<5×10−8. Novel CAD loci are presented with red stacks and gene names. Previously reported loci showing GWAS significance are shown in brown, and those showing nominal significance (P<0.05) in this meta-analysis are shown in blue. Reprinted from Nikpay et al25 with permission of the publisher. Copyright ©2015, Nature Publishing Group.
The 48 GWAS significant loci previously identified explained only ≈11% of the estimated heritability of CAD. Here, beyond considering only GWAS significant loci (P≤5×10−8), an approximate joint association analysis was performed using genome-wide complex trait analysis software26,27 to identify 202 false discovery rate variants (q<0.05) in 129 loci. Thus, several loci contained multiple independent signals for CAD association. Together, the 202 independent variants explained ≈28% of predicted CAD heritability.25 Of these, 15 were of low frequency (MAF<0.05) and explained only ≈2% of CAD heritability. Here, it is important to acknowledge that GWAS analysis based on SNP array data has limited power to resolve genes with rare mutation burdens.25 Overall, this analysis not only provided new information on the genetic architecture of CAD but also strongly supports the conclusion that genetic susceptibility to CAD is largely derived from the effect of multiple common SNPs of small effect size. The 202 independent variants reaching a false discovery rate of q<0.05 showed independent enrichment across 11 cell types for histone/chromatin modifications, DNase I hypersensitive sites consistent with other data that GWAS signals are enriched in regulatory regions of the genome.28,29

Studies in Non-European Populations

The majority of GWASs for CAD have been carried out in populations of European white ancestry with smaller but important studies in East Asian, South Asian, and black populations. Haplotype blocks are segments of DNA that are known and shared within ancestral groups. Because of the recent arrival of humans into Europe, individuals of European ancestry are more similar to each other and have more correlated SNPs and longer haplotype blocks when compared with the more ancient and genetically diverse populations of African ancestry.3032 The average haplotype block is ≈20.7 kb in European whites, ≈8.8 kb in blacks, and ≈ 25.2 kb in Han Chinese. Thus, fewer tag SNPs are required for genotyping a population of European or East Asian versus African ancestry. With the supposition that causal variants are shared among different populations, transethnic mapping by taking advantage of differences in LD and allele frequency might be expected to facilitate identification of causal variants. Here, the significantly shorter haplotype blocks in black populations could help to both fine-map association signals identified in European populations and identify novel ethnic-specific signals.33
Most of the disease-related GWAS loci discovered in Europeans have been replicated in populations of East Asian ancestry. A strong and significant correlation of ORs of specific SNPs for CAD and 27 other diseases across European and East Asian samples has recently been demonstrated, indicating that causal variants, in general, are shared between these 2 populations.34 As would be expected, the SNPs that failed replication in East Asian populations mapped to genomic regions with different patterns of LD. The first published GWAS for CAD, including a substantial number of South Asians, was by the C4D Consortium. Their discovery data set included ≈15 000 CAD cases; half of whom were of South Asian descent.22 Of the 11 previously reported common variants for CAD confirmed in this study, the effect was directionally consistent in European and South Asian populations. However, the OR for several of these, including 9p21.3 and SORT1 (encoding sortilin-1 and related to lipid metabolism), was somewhat lower in the South Asian studies. The recent 1000 Genomes–based meta-analysis included a small number of South Asian (13%) and East Asian (7%) subjects but did not report on ethnic-specific effects.25 In summary, despite smaller data sets, the directional effects of CAD risk alleles identified in European populations on CAD risk in South Asians are generally concordant. In contrast to a similar effect size of CAD risk alleles in East Asian and European populations, for many, the effect size is apparently attenuated in South Asians, possibly not only because of the interaction with unknown genetic or environmental factors but also because the causal variant(s) may be inadequately tagged by markers present on available genotyping arrays, resulting in blunted genetic effects.
In contrast to the findings in East and South Asians, the majority of CAD-associated loci (an exception being rs599839 at the SORT1 locus)35 identified in European populations have failed to replicate or shown considerably reduced effect size in black populations.34 This is despite the high prevalence of CAD among blacks and the potential advantage of interrogating their shorter LD blocks for fine mapping of previously identified CAD loci. Two signals previously reported for CAD at the 9p21.3 locus36,37 did not achieve replication in independent black data sets, and other signals for CAD have not been identified. In the Population Architecture Using Genomics and Epidemiology (PAGE) study, a consortium of multiancestry, population-based studies, 25% of tag SNPs identified in European GWAS had significantly different effect sizes in black cohorts.38 This might imply limited sharing of causal variants between Europeans and Africans when compared with other ethnicities. Given the lower level of LD in African populations, the index SNP identified in European studies may simply fail to tag potentially shared casual variants.34 However, replication has been achieved for signals associated with discrete CAD risk factors.37 It is possible that the genetic risk of CAD in blacks is related more strongly to the genetic contribution to prevalent risk factors, including hypertension and obesity.39 Given that multiple genetic variants of small effect size are believed to account for much of the heritability of CAD in Europeans, these are likely to be greater in number and more difficult to detect in populations of African descent. Here, it should be noted that the sample sizes of these studies in blacks are of smaller magnitudes than the sample sizes of the studies involving individuals of other ancestries, thus limiting statistical power.

Mendelian Randomization

Epidemiological studies have identified a large number of biomarkers that are associated with CAD. However, 2 of the most important problems of observational epidemiology are confounding and reverse causation. These factors limit the ability of these types of studies to identify causality of a biomarker or a risk factor. To circumvent confounding and reverse causation, a growing number of epidemiologists are turning to the Mendelian randomization (MR) method, a method that can be likened to the randomized controlled trial (RCT). This method that combines genetic and observational data takes advantage of the random assortment of genetic variants (=alleles) at conception; therefore, by studying common genetic variants that associate with the exposure variable of interest (eg, low-density lipoprotein [LDL] cholesterol [LDL-C]), reverse causation and most confounding can be avoided. The methodology is based on a simple tenet—if a biomarker has a causal association with a disease, the genetic determinants of the biomarker will also associate with disease risk.40 MR has limitations that are shared with RCTs, and one of the most important of these is the lack of pleiotropy, that is, the genetic variant(s) in question must influence only the biomarker of interest. This is not always the case for plasma lipid traits, where pleiotropic effects are evident for several genes, including CETP, LPL, and APOA5. Using MR, major advances have been made in determining the causal associations between plasma levels of lipoprotein(a), LDL-C, and triglycerides (as a marker of remnant cholesterol) and risk of CAD, whereas C-reactive protein (CRP) and high-density lipoprotein (HDL) cholesterol, despite both being robust risk markers, have not been shown to be causal (Figure 2).
Figure 2. Methodology is based on the tenet that if the biomarker has a causal association with disease, the genetic determinants of the biomarker will also associate with disease risk. A, If evidence #1 to #3 are all documented robustly, the interpretation is that the data are compatible with a causal relationship. B, If evidence #1 and #2 are documented robustly, but the genetic determinants of the biomarker do not associate with disease risk, the interpretation is that the association is noncausal. Using Mendelian randomization, major advances have been made in determining the causal associations between plasma levels of lipoprotein(a), low-density lipoprotein (LDL) cholesterol, triglycerides (TG; as a marker of remnant cholesterol), and body mass index (as a surrogate for obesity) with risk of coronary artery disease (CAD). In contrast, C-reactive protein, Lp-PLA2, high-density lipoprotein (HDL) cholesterol, fibrinogen, and homocysteine, despite being robust risk markers, have not been shown to be causal.
Elevated levels of lipoprotein(a) are associated with increased risk of CAD, and common copy number variants and SNPs in the LPA gene, which associate with increased lipoprotein(a) levels, cause an increased risk of CAD4143 and of aortic valve stenosis.44,45 Although the exact mechanisms are not clear,46 these data suggest that apolipoprotein(a) is directly causal and therefore a potential drug target.
That elevated levels of LDL-C are causally associated with an increased risk of CAD is evident from familial hypercholesterolemia because of mutations in the LDLR. Genetic variants at the PCSK9, NPC1L1, and HMGCR loci uniquely associate with plasma concentrations of LDL-C and are therefore also predictive of CAD risk. As originally reported by Abifadel et al,47 gain-of-function mutations in PCSK9 cause autosomal dominant familial hypercholesterolemia. Cohen et al48 later reported that loss-of-function mutations in PCSK9 were associated with reduced plasma concentrations of LDL-C and lower risk of CAD, establishing PCSK9 as a relevant new drug target for lowering LDL-C levels. Ezetimibe reduces LDL-C levels by inhibiting Niemann–Pick C1-like protein1 (NPC1L1), a transporter that in humans is responsible for cholesterol uptake from the intestine into enterocytes and from bile into hepatocytes. Despite this, ezetimibe has only recently been shown to reduce cardiovascular risk.49 The Myocardial Infarction Genetics Consortium investigators sequenced NPC1L1 in 7364 CAD cases and 14 728 controls and identified 34 loss-of-function mutations. One of these, p.Arg406X, in a much larger replication study was found to be associated with a 10% lower LDL-C and a 50% decrease in CAD risk.50 Similarly, Lauridsen et al51, in a single-center study of the general population, included 67 385 individuals; of them, 5255 and 3886 developed incident ischemic vascular disease and symptomatic gallstone disease, respectively. Using a genetic score of common variants in NPC1L1, mimicking the effect of ezetimibe, they showed that these variants were associated with lifelong, stepwise reductions in LDL-C levels of ≤3.5%, with a corresponding 18% reduction in risk of CVD and a 22% increase in risk of symptomatic gallstone disease. The study, therefore, suggested that a biologically plausible long-term side effect of ezetimibe treatment might be an increased risk of symptomatic gallstone disease. Finally, Ference et al52 in a study of 108 376 subjects in 14 clinical trials reported that common genetic variants associated with each of NPC1L1 and HMGCR conferred a reduction in LDL-C of 2.4 and 2.9 mg/dL and a respective 4.8% and 5.3% lower CAD risk. Because PCSK9, NPC1L1, and HMGCR regulate LDL but do not exhibit effects on other lipid and nonlipid parameters, such studies strongly support a causative relationship between plasma concentrations of LDL-C and CAD. Importantly, multiple variants associated with LDL-C concentrations, that is, in PCSK948,53 and NPC1L150,51 genes, are more strongly associated with CAD than are associated differences in measured LDL-C or the effect of LDL-C reduction in the statin (RCTs). This is presumably because genetic effects encompass the lifetime exposure to LDL-C (Table 2).52,54
Table 2. Common Genetic Terminology
TermDefinition
Complex traitsTraits that are influenced by both the environment and a combination of variants in at least several genes; each of which may have a variable effect
Copy number variantsSegments of DNA that are duplicated or deleted
ExomeThe protein-coding region of the genome
Fine mappingThe use of dense genotyping data around an associated allele in an attempt to identify the causal allele(s)
Genotype imputationA statistical technique that exploits LD (see below) between genotyped and ungenotyped variants to infer missing genotypes in a set of individuals, using a reference panel of genotyped individuals, such as the HapMap or 1000 Genomes data sets
GWASA study that searches, in an agnostic fashion, for allelic association with a particular phenotype by genotyping tag single-nucleotide polymorphisms across the entire genome
Genome-wide significanceA level of statistical significance required to establish association for a common variant in GWASs (P=5×10−8), assuming Bonferroni correction for ≈1 000 000 independent genomic markers
HaplotypeA combination of alleles transmitted together on a single chromosome
HeritabilityThe proportion of the total phenotypic variation that can be attributed to additive genetic effects
LDThe nonrandom association of alleles at ≥2 loci because of infrequent recombination events between them
Purifying selectionSelection acting to remove new deleterious mutations that impair the reproductive fitness of an individual
Whole-exome sequencingThe protein-coding sequences of all known genes (≈30-Mb pairs) are targeted, captured, and sequenced
Whole-genome sequencingThe entire genome (≈3-Gb pairs) is sequenced, and the sequencing data mapped to a reference genome sequence
GWAS indicates genome-wide association study; and LD, Linkage disequilibrium.
The role of plasma triglycerides as an independent risk factor of CAD was debated for many years. However, genetic studies have lent credence to the hypothesis that triglyceride-rich lipoproteins or their remnants have a causative role in atherosclerosis.55 The major determinant of plasma triglyceride metabolism is lipoprotein lipase, and genetic variants that decrease LPL function confer increased cardiovascular risk54,5658 and all-cause mortality.54,59 LPL activity is increased by ApoA5 and inhibited by ApoC3 and the angiopoietin-like proteins 3 and 4. Loss-of-function mutations in APOC360,61 are associated with lower plasma triglycerides and CAD risk, whereas loss-of-function mutations in APOA5 have opposite effects.62 In a collaborative analysis, including 20 842 individuals with CAD and 35 206 controls, Sarwar et al63 reported that a common promoter polymorphism (–1131T>C; rs662799) in APOA5 was associated with CAD with an allele-specific OR of 1.18 directionally similar but quantitatively greater than the effects on plasma triglycerides. Similarly, Jørgensen et al64 found that genotype combinations of 3 common variants in APOA5 associated with increases in nonfasting triglycerides and calculated remnant cholesterol of 1.10 and 0.40 mmol/L, respectively and with corresponding ORs for myocardial infarction of 1.87 (1.25–2.81).
Essential for successful MR studies is the selection of genetic variants without pleiotropic effects. The major difficulty in studying raised concentrations of triglycerides or remnant cholesterol is the inverse association with HDL cholesterol concentrations.55 However, a MR study with genetic variants in several candidate genes that affect the concentrations of remnant cholesterol or HDL cholesterol or both showed that an increase of 1 mmol/L in remnant cholesterol was associated with a 2.8× increased risk of ischemic heart disease that was not attributable to lower HDL cholesterol concentrations.65 A recent GWAS also supports that variants associated with high concentrations of triglycerides are causally associated with CAD, after adjustment for HDL cholesterol.66 These 2 studies62,65,66 also showed that genetically low HDL cholesterol was unrelated to cardiovascular risk, in agreement with several previous studies essentially using an MR approach to test causality for HDL cholesterol.6065 Because triglycerides can be degraded by most cells but cholesterol cannot, the cholesterol content of triglyceride-rich remnant lipoproteins (remnant cholesterol) is more likely to be the cause of atherosclerosis and cardiovascular disease rather than raised triglycerides. Remnant lipoproteins can, like LDL, enter and be trapped in the arterial intima simply because of their size and possibly via attachment to extracellular proteoglycans.55 Taken together, genetic studies strongly support that triglyceride-rich lipoproteins and remnant cholesterol are causal risk factors of CAD and all-cause mortality.
As noted above, MR studies have not supported a direct protective function of HDL cholesterol levels against CAD. Apolipoprotein A-I is the major protein on HDL, but common variants in APOA1 associated with low HDL cholesterol are not associated with an increased risk of CAD.67 For the effect of rare variants on CAD risk, the picture is complicated by the fact that some rare variants in APOA1 that associate with low HDL cholesterol, cause amyloidosis, and this, in addition to other symptoms, may manifest as CAD due to small vessel disease, increased intima media thickness, or in later stages as cardiomyopathy. Haase et al68,69 resequenced the APOA1 gene in >10 000 individuals, genotyped the identified variants in >55 000 individuals from the general population, and reported that the apparent increased risk associated with nonsynonymous mutations in APOA1 became insignificant when variants previously associated with amyloidosis were removed from the analysis. In support, mutations in APOA1 reported to associate with endothelial dysfunction, increased arterial wall thickness, and premature coronary heart disease70 have more recently been shown to cause systemic amyloidosis.71
For other biomarkers, MR studies have provided important insights relevant to drug targets, drug discovery, and potential long-term side effects of drugs. Although high-sensitivity CRP is a robust biomarker for CAD risk and response to stain therapy, neither common genetic variants in the CRP gene associated with substantial increases in CRP levels in the general population72 nor rare exonic variants associated with CRP levels73 associate with CAD risk, demonstrating that CRP is not a causal risk factor and hence not a relevant drug target. Similarly, although data from epidemiological and clinical studies supported a potentially important role of lipoprotein-associated phospholipase A2 in atherosclerosis and its sequelae,74,75 the genetic variants near PLA2G7 and CETP associated with LP-PLA2 levels and activity did not confer CAD risk.76,77 In accord with these findings, the results of phase III randomized controlled trials of inhibitors of these enzymes (varespladib78 and darapladib79) were also negative. A closer inspection of MR data before drug design and initiation of clinical trials could, thus, help to mitigate the high costs of drug development. Conversely, 3 of 4 CETP inhibitors either have failed because of adverse events (torcetrapib) or have been stopped because of futility (dalcetrapib and evacetrapib). Nevertheless, common genetic variants in CETP associated with reduced mass and activity are associated with increased HDL cholesterol, smaller decreases in LDL-C and triglycerides and, in the Copenhagen Heart Study, corresponding stepwise reductions in risk of all ischemic end points and all-cause mortality80 consistent with other reports.8183 However, the effects of small molecule inhibition of CETP are not analogous to the small changes in CETP concentrations because of common genetic variants. Statin therapy lowers CETP by ≈20%,84 and as reported by Kuivenhoven et al,85 individuals with a common CETP polymorphism associated with higher CETP levels showed reduced progression of atherosclerosis when treated with pravastatin, whereas those with genetically low CETP concentrations did not, suggesting that there may be an optimal window of CETP activity.86

Utility of a Genetic Risk Score for Risk Assessment and Treatment Decisions

Although the effect of the identified susceptibility variants of CAD is individually small, their effects are independent and additive. These can be incorporated into a genetic risk score (GRS) consisting of the number of risk alleles adjusted for their individual effect size. In an analysis performed in 2 prospective cohorts, Whitehall II (WHII; n=5059) and the British Women’s Heart and Health Study (BWHHS; n=3414), individuals in the top versus bottom quintile of an LDL-C GRS consisting of 23 SNPs had a higher risk of CAD (WHII: OR=1.43; BWHHS: OR=1.31).87 After adjusting for LDL-C levels, this association was completely attenuated in WHII but not in BWHHS. The value of a GRS for CAD risk prediction beyond conventional risk factors has improved beyond previous studies considering a small number of risk variants88,89 beyond those related to plasma lipid traits to more recent analyses incorporating a more comprehensive list of recently identified CAD loci.9092
Tikkanen et al92 constructed a GRS consisting of 28 CAD genetic variants and in 24 124 participants, in 4 population-based, prospective cohorts, determined its association with incident cardiovascular disease events for a mean 12-year follow-up period. Compared with conventional risk factors and family history alone, the GRS improved CAD risk discrimination (C index, 0.856 versus 0.851; P=0.0002). More recently, Mega et al93 evaluated the association of a GRS based on 27 genetic variants with incident or recurrent coronary heart disease, adjusting for traditional clinical risk factors in a community-based cohort study (48 421 individuals in the Malmo Diet and Cancer Study) and 4 RCTs of statin therapy, in primary (Justification for the Use of Statins in Prevention: an Intervention Trial Evaluating Rosuvastatin [JUPITER] and Anglo-Scandinavian Cardiac Outcomes Trial [ASCOT]) and secondary prevention (Cholesterol and Recurrent Events [CARE] and Pravastatin or Atorvastatin Evaluation and Infection Therapy-Thrombolysis in Myocardial Infarction 22 Trials [PROVE IT-TIMI 22]) populations. Their major findings were that the multivariable adjusted hazard ratio for CAD was directly related to the GRS (hazard ratio, 1.72 for individuals in the highest versus the lowest GRS quintile) and that in the statin RCTs, both relative and absolute risk reductions were 3-fold greater in individuals in the highest genetic risk category, with highly significant effects on the number needed to treat. Thus, a GRS can provide important information on future CAD risk beyond traditional risk factors and may have both clinical and economic values in identifying those individuals who will benefit most from the early application of various disease-deferring therapies, including statin treatment.

Interrogation of Rare Coding Variants Contributing to CAD

Overall, the most recent GWAS analysis strongly supports the conclusion that genetic susceptibility to CAD is largely derived from the effect of multiple common SNPs of small effect size.25 However, this analysis was not able to interrogate rare variants present in >1 in 200 individuals, including nonsynonymous mutations in genes of relevance to CAD. Earlier studies that entailed only limited resequencing of candidate regions of the genome in CAD kindreds reported various rare mutations or copy number variants associated with CAD but for most replication has not been achieved.94 A 21-bp (7 amino acids) deletion in MEF2A was associated with apparent autosomal dominant CAD in 1 large kindred,95 but this was not confirmed in other studies, including the one involving 2 extended families bearing the same copy number variant.96
A comprehensive approach to identify and classify coding variants linked to human disease is whole-exome sequencing. This has yielded important findings for Mendelian disease particularly when family members were also available.97 It has also shown utility for the genetic diagnosis of various cardiomyopathies and cardiac conduction disorders.98 With reference to CAD, the Myocardial Infarction Genetics Consortium Exome Sequencing Project reported on whole-exome sequencing efforts in ≈11 000 individuals. Of note, the identified coding variants were limited to genes relevant to lipoprotein metabolism, including LDLR, APOA5, APOC3, and NPC1L150,60,62 and for which there was a large body of pre-existing information. This finding was somewhat disappointing but did provide further credence to the relevance of triglyceride metabolism to atherosclerosis55,61,6365 and the validity of NPC1L1 as a drug target.49,51,99 Given the perceived need for large sample sizes and subsequent replication, the sample included CAD cases with limited phenotypes, such as acute myocardial infarction without angiographic documentation of disease burden, and controls who may have harbored a burden of nonobstructive atherosclerosis. The inability to follow up on rare variants of high effect size in family members with and without CAD may also have limited the ability to detect novel mutations linked to atherosclerosis. Importantly, studies to date have been significantly underpowered. Recent stimulation analyses indicate that tens of thousands of individuals will be required to confidently detect or exclude rare variant signals at complex disease loci, such as CAD.100

Beyond the Single SNP

Despite the recent success of GWASs, close to 80% of the estimated heritability of CAD remains unknown. Some of this missing heritability may be because of gene environment interaction101 where the effect of a given genetic variant is only manifest in the presence of a modifier, such as obesity or cigarette smoking. For example, the effect of common risk alleles for elevated plasma triglycerides was markedly attenuated in lean versus obese individuals.102 Epistasis is a statistical interaction between ≥2 genetic loci where the effects are nonadditive and has been hypothesized to play an important role in the genetic determination of complex diseases, such as CAD. Epistasis on a genome-wide level has proven challenging because of the requirement for individual genotype data, computational power required for a large number of pair-wise or high-order tests, and the need to correct for multiple testing. However, filtering to consider only those regions with potential biological interaction reduces the necessity for high computational power and the burden of multiple testing correction. We recently reported that SMAD3 is a necessary factor for transforming growth factor-β–mediated stimulation of mRNA and protein expression of type IV collagen genes in human vascular smooth muscle cells and demonstrated a statistical interaction between the COL4A1/COL4A2 locus and the SMAD3 locus by performing pair-wise tests between SNPs at the 2 loci.103
It is also likely that many more common variants are linked to CAD but have not achieved genome-wide significance in GWAS because of small effect size or low allele frequency and insufficient sample size. However, on the basis of the premise that clinically informative polymorphisms related to complex disease occur in systems of closely interacting genes,104 even weakly associated variants may provide important information on the biological basis of disease when such variants cluster within a common functional module or a pathway.105 Thus, novel insights into the genetic architecture of CAD can be obtained by interpreting genetic findings in the context of biological processes and functional interactions among genes.
A modified version of gene-set enrichment analysis, developed for the interrogation of gene expression data,106 was designed by Wang et al107 for the analysis of genome-wide SNP associations, and other gene-set enrichment analysis methods have been more recently applied to GWAS data.108112 These analytic algorithms seek to identify a sets of genes whose variants collectively demonstrate strong association with a trait of interest even if the component SNPs individually exhibit relatively modest or nonsignificant association.105 To identify novel associations between established biological mechanisms and CAD, we recently performed a 2-stage pathway–based gene-set enrichment analysis of 16 GWAS data sets for CAD that included >25 000 subjects with CAD and >66 000 controls105 using the i-GSEA4GWAS tool112 and the Reactome pathway database.113 A total of 32 of 639 Reactome pathways tested demonstrated convincing association with CAD. These resided in core biological processes and included pathways relevant to extracellular matrix integrity, innate immunity, axon guidance, and signaling by platelet-derived growth factor, NOTCH, and the transforming growth factor-β–SMAD receptor complex. Importantly, many of these pathways were shown to have strengths of association comparable with those observed in lipid transport pathways.105

Systems Genetics Approach to Understanding the Genetic Basis of CAD

A further extension of the pathway approach is the interrogation of molecular networks that underlie the complex architecture of complex disease.114 A molecular network is based on interactions among diverse molecules, including genes, proteins, and metabolites. These can include interactions between proteins, effects on gene regulation, coexpression, and various functional interactions. In general, genes associated with the same or similar disorders tend to occupy similar molecular networks through physical or functional modules.115,116 Furthermore, this approach indicates that disease-related genes exhibit network connectivity and network centrality properties that are distinct from other genes.116
Thus, interrogation of molecular networks provides additional information on interactions among gene subsets within a given pathway and highlights potentially important interactions between components of different biological pathways. In the study of Ghosh et al,105 network analysis of unique genes within the replicated pathways not only further confirmed the known processes, such as lipid metabolism, but also revealed many interconnected functional and topologically interacting modules representing novel associations, for example, the semaphorin-regulated axonal guidance pathway. The axon guidance pathways modulate diverse biological phenomena, including cellular adhesion, migration, proliferation, differentiation, survival, and synaptic plasticity, through the participation of highly conserved families of guidance molecules, including netrins, slits, semaphorins, and ephrins, and their cognate receptors117 and members of these pathways have been highlighted in many recent reports related to macrophage migration,118 expression of inflammatory markers and M2 signals in atherosclerotic plaque,119 and chemokine-directed migration of human monocytes.120,121 Network centrality analysis further identified genes (eg, NCAM1, FYN, and FURIN) likely to play critical roles in the maintenance and functioning of several of the replicated pathways.

From Locus to Function

Despite the success of recent GWASs, there has been limited progress in understanding the function of the multiple risk loci identified. Notably, the majority of signals for CAD identified by the GWAS approach are in noncoding regions of the genome. This is not surprising, given that 99% of the genome is nonprotein coding. Nonetheless, ≈10% of this region is under purifying selection,122 implying important functional effects. These regions include noncoding RNAs that may influence gene transcription through multiple mechanisms,123 active promoter and enhancers, regions affecting histone acetylation and deacetylation and chromatin remodeling,124 susceptibility to DNA methylation, and micro-RNAs and micro-RNA binding sites.
The GWAS-identified polymorphisms are themselves rarely causal but rather in LD with a neighboring or even distal causal polymorphism. Most of the CAD-associated loci identified by GWASs span several kilobases, and some of these may encompass multiple causative variants in ≥1 genes. For common variants, fine-mapping approaches can be limited by large blocks of LD. As an example, the 9p21.3 risk locus encompasses multiple SNPs in tight LD. Dense resequencing failed to identify less common genetic variants with a larger effect size than the original GWAS SNPs.125 As reported by the 1000 Genomes Consortium, several CAD loci contain >1 independent signal.25 For example, 7 independent false discovery rate (q<0.05) SNPs were identified at the COL4A1/COL4A2 locus and 5 in the CDKN2BAS (9p21.3) region with 1 uncommon SNP, rs7855162 (MAF, 0.03), exhibiting a higher effect size than the index SNP.25
It is often not fully appreciated that although the association is at the level of genomic DNA, the relevance may be restricted to a particular tissue or organ. An expression quantitative trait locus (eQTL) effect, that is, an effect on mRNA expression on a nearby (cis-eQTL) or distant (trans-eQTL) gene, provides strong evidence for a functional effect of a given SNP. However, recent studies have shown that some eQTLs are shared among tissues, whereas others are tissue specific. Hence, interrogating eQTL databases, for example,126 Genotype Tissue expression and 127 SNP and CNV Annotation database, requires careful consideration of the cell type or tissue of interest.128 Finally, it must be acknowledged that the direct contribution of any given locus could be temporally restricted, for instance, early transient activation followed by sustained epigenetic effects.
The majority of common CAD-associated variants exhibit small individual effects on disease risk, and many are likely to exert their effects by altering promoter or enhancer activity. As a complementary approach to eQTL analysis, public databases, generated by the Encyclopedia of DNA Elements129 and Roadmap Epigenomics projects,130 genome-wide chromatin profiles of histone modifications, data on transcription factor–binding sites, and chromatin immunoprecipitation sequencing data131 can be used to identify specific functional regulatory elements.132 H3K4me3 denotes active promoters, H3K4me1 denotes enhancers, and H3K4me2, as well as most histone deacetylations, denotes both promoter and enhancer regions.133 DNase I hypersensitive sites are also a mark of open chromatin regions containing promoters and enhancers.134 Assay for transposase-accessible chromatin with high-throughput sequencing is a useful method for measuring chromatin accessibility genome wide.135 An important caveat is that these analyses require genome-wide chromatin data from a relevant cell type.136 These approaches to the identification of regulatory sequences can be followed up by more specific methodologies appropriate to the characteristics of the risk locus in the cell or tissue of interest. Thus, it can first be determined whether a risk locus is likely to harbor functional cis-acting regulatory modules whose activity is altered by an SNP in close LD with a particular risk variant before standard molecular biology approaches in the laboratory (Figure 3).
Figure 3. A, Typical Manhattan plot of a significant locus for coronary artery disease (CAD) identified in a meta-analysis of genome-wide association study (GWAS) for myocardial infarction (MI)/CAD. B, Visualization of the various linkage disequilibrium (LD) blocks in the closest gene using the Haploview software and linkage data taken from the 1000 Genomes Browser. C, UCSC genome browser annotation and Encyclopedia of DNA Elements (ENCODE) project data for the region bearing the lead single-nucleotide polymorphisms (SNPs), including chromatin regulatory features, such as DNase hypersensitivity sites, histone modifications (H3K4me1, H3K4me3, and H3K27ac), and predicted transcription factor binding from chromatin immunoprecipitation sequencing (ChIP-Seq) data. D, Interrogation of publically available databases, such HaploReg (motif analysis), RegulomeDB (a database that annotates SNPs with known and predicted regulatory elements in noncoding regions of the genomes), SCAN (SNP and CNV Annotation database; a database of genetic and genomic data with online methodology for mining these data), Genotype Tissue expression (GTex), and other expression quantitative trait locus (eQTL) databases, can provide further information relevant to function. E, Finally, extensive laboratory analyses, including enhancer and promoter assays, allele-specific ChIP, allele-specific expression analyses, and eQTL analyses in cell or tissue of relevance to CAD, is necessary to identify and molecularly characterize the functional variant underlying the GWAS signal.
The poster child for this approach was the identification of a previously unknown role for sortilin-1 in lipoprotein metabolism. Four common GWAS-identified SNPs associated with LDL-C and CAD lie in a noncoding DNA region on chromosome 1p31 between 2 genes, CELSR2 and PSRC1, and downstream of SORT1. In a series of elegant studies, including primary human hepatocytes, Musunuru et al137 demonstrated that these SNPs have the strongest hepatocyte eQTL effect for SORT1 and that one of these, rs12740374, creates a C/EBP (CCAAT/enhancer binding protein) transcription factor–binding site and alters the hepatic expression of SORT1 gene, encoding sortilin-1. In further studies in the mouse, genetically increased hepatic sortilin expression was shown to both reduce hepatic APOB secretion and increase LDL catabolism, providing dual mechanisms for the strong association between increased hepatic sortilin-1 expression and reduced plasma LDL-C levels in humans.138
In another recent investigation of GWAS significant loci, Sazonova et al139 used a complex series of bioinformatic, molecular, cell biology, and system genetic approaches to show that the transcription factor, TCF21, regulates a transcriptional network linking multiple independent CAD loci. The same group demonstrated that TCF21 regulates the development of the epicardial progenitor cells that give risk to smooth muscle cells that contribute to the fibrous cap.140 Similar approaches have provided new insight into the roles of PHACTR1141 and ADAMTS7142 in atherosclerosis. ADAMTS7 plays a role in the regulation of vascular smooth muscle cell migration, and Bauer et al143 recently demonstrated that Adamsts7−/− mice were protected from atherosclerosis on an Ldlr−/− or Apoe−/− background and showed reduced neointimal formation after femoral wire injury and that Adamsts7−/− vascular smooth muscle cells showed reduced migration in the setting of tumor necrosis factor-α stimulation, consistent with a proatherogenic effect of ADAMTS7.

Summary

The past decade of research has provided a broader understanding of the genetic architecture of CAD and demonstrates that the genetic basis of CAD largely derives from the cumulative effect of multiple common risk alleles individually of small effect size rather than rare variants with large effects on CAD risk. Although traditional risk factors remain important, application of these data using a systems genetics approach has pointed to substantial roles for genes and pathways relevant to vessel wall biology and immune function. A major priority is to apply high-throughput methodology to understand at the molecular and cellular levels the function of each of the novel loci; the majority of which are in noncoding regions of the genome. Beyond functional insight into disease mechanisms, these data have proven clinical utility for interrogation of biomarker causality and drug discovery, through MR, creation of a GRS with predictive power to better identify those individuals who will benefit most from statin therapy.

Footnotes

Circulation Research Compendium on Atherosclerosis
Atherosclerosis: Successes, Surprises, and Future Challenges
Epidemiology of Atherosclerosis and the Potential to Reduce the Global Burden of Atherothrombotic Disease
Triglyceride-Rich Lipoproteins and Atherosclerotic Cardiovascular Disease: New From Epidemiology, Genetics, and Biology
Genetics of Coronary Artery Disease
Surprises From Genetic Analyses of Lipid Risk Factors for Atherosclerosis
From Loci to Biology: Functional Genomics of Genome-Wide Association for Coronary Disease
Are Genetic Tests for Atherosclerosis Ready for Routine Clinical Use?
Endothelial Cell Dysfunction and the Pathobiology of Atherosclerosis
Macrophages and Dendritic Cells: Partners in Atherogenesis
Macrophage Phenotype and Function in Different Stages of Atherosclerosis
Adaptive Response of T and B Cells in Atherosclerosis
Microdomains, Inflammation, and Atherosclerosis
Vascular Smooth Muscle Cells in Atherosclerosis
MicroRNA Regulation of Atherosclerosis
The Success Story of LDL Cholesterol Lowering
From Lipids to Inflammation: New Approaches to Reducing Atherosclerotic Risk
Imaging Atherosclerosis
Peter Libby, Karin E. Bornfeldt, and Alan R. Tall, Editors

Nonstandard Abbreviations and Acronyms

BWHHS
British Women’s Heart and Health Study
CAD
coronary artery disease
CARDIoGRAM
Coronary Artery Disease Genome-Wide Replication and Meta-Analysis
CRP
C-reactive protein
eQTL
expression quantitative trait locus
GRS
genetic risk score
GWAS
genome-wide association study
HDL
high-density lipoprotein
LD
linkage disequilibrium
LDL
low-density lipoprotein
LDL-C
low-density lipoprotein cholesterol
MAF
minor allele frequency
MR
Mendelian randomization
NPC1L1
Niemann–Pick C1-like protein1
OR
odds ratio
PAGE
Population Architecture Using Genomics and Epidemiology
RCT
randomized controlled trial
SNP
single-nucleotide polymorphism
WHII
Whitehall II

References

1.
Vinkhuyzen AA, Wray NR, Yang J, Goddard ME, Visscher PM. Estimation and partition of heritability in human populations using whole-genome analysis methods. Annu Rev Genet. 2013;47:75–95. doi: 10.1146/annurev-genet-111212-133258.
2.
Lloyd-Jones DM, Nam BH, D’Agostino RB, Levy D, Murabito JM, Wang TJ, Wilson PW, O’Donnell CJ. Parental cardiovascular disease as a risk factor for cardiovascular disease in middle-aged adults: a prospective study of parents and offspring. JAMA. 2004;291:2204–2211. doi: 10.1001/jama.291.18.2204.
3.
Zdravkovic S, Wienke A, Pedersen NL, Marenberg ME, Yashin AI, De Faire U. Heritability of death from coronary heart disease: a 36-year follow-up of 20 966 Swedish twins. J Intern Med. 2002;252:247–254.
4.
Schunkert H, König IR, Kathiresan S, et al.; Cardiogenics; CARDIoGRAM Consortium. Large-scale association analysis identifies 13 new susceptibility loci for coronary artery disease. Nat Genet. 2011;43:333–338. doi: 10.1038/ng.784.
5.
McPherson R. A gene-centric approach to elucidating cardiovascular risk. Circ Cardiovasc Genet. 2009;2:3–6. doi: 10.1161/CIRCGENETICS.109.848986.
6.
McPherson R, Pertsemlidis A, Kavaslar N, Stewart A, Roberts R, Cox DR, Hinds DA, Pennacchio LA, Tybjaerg-Hansen A, Folsom AR, Boerwinkle E, Hobbs HH, Cohen JC. A common allele on chromosome 9 associated with coronary heart disease. Science. 2007;316:1488–1491. doi: 10.1126/science.1142447.
7.
Helgadottir A, Thorleifsson G, Manolescu A, et al. A common variant on chromosome 9p21 affects the risk of myocardial infarction. Science. 2007;316:1491–1493. doi: 10.1126/science.1142842.
8.
Samani NJ, Erdmann J, Hall AS, et al.; WTCCC and the Cardiogenics Consortium. Genomewide association analysis of coronary artery disease. N Engl J Med. 2007;357:443–453. doi: 10.1056/NEJMoa072366.
9.
Dandona S, Stewart AF, Chen L, Williams K, So D, O’Brien E, Glover C, Lemay M, Assogba O, Vo L, Wang YQ, Labinaz M, Wells GA, McPherson R, Roberts R. Gene dosage of the common variant 9p21 predicts severity of coronary artery disease. J Am Coll Cardiol. 2010;56:479–486. doi: 10.1016/j.jacc.2009.10.092.
10.
Chan K, Patel RS, Newcombe P, et al. Association between the chromosome 9p21 locus and angiographic coronary artery disease burden: a collaborative meta-analysis. J Am Coll Cardiol. 2013;61:957–970. doi: 10.1016/j.jacc.2012.10.051.
11.
Ye S, Willeit J, Kronenberg F, Xu Q, Kiechl S. Association of genetic variation on chromosome 9p21 with susceptibility and progression of atherosclerosis: a population-based, prospective study. J Am Coll Cardiol. 2008;52:378–384. doi: 10.1016/j.jacc.2007.11.087.
12.
Smith JG, Melander O, Lövkvist H, Hedblad B, Engström G, Nilsson P, Carlson J, Berglund G, Norrving B, Lindgren A. Common genetic variants on chromosome 9p21 confers risk of ischemic stroke: a large-scale genetic association study. Circ Cardiovasc Genet. 2009;2:159–164. doi: 10.1161/CIRCGENETICS.108.835173.
13.
Anderson CD, Biffi A, Rost NS, Cortellini L, Furie KL, Rosand J. Chromosome 9p21 in ischemic stroke: population structure and meta-analysis. Stroke. 2010;41:1123–1131. doi: 10.1161/STROKEAHA.110.580589.
14.
Cluett C, McDermott MM, Guralnik J, Ferrucci L, Bandinelli S, Miljkovic I, Zmuda JM, Li R, Tranah G, Harris T, Rice N, Henley W, Frayling TM, Murray A, Melzer D. The 9p21 myocardial infarction risk allele increases risk of peripheral artery disease in older people. Circ Cardiovasc Genet. 2009;2:347–353. doi: 10.1161/CIRCGENETICS.108.825935.
15.
Bown MJ, Braund PS, Thompson J, London NJ, Samani NJ, Sayers RD. Association between the coronary artery disease risk locus on chromosome 9p21.3 and abdominal aortic aneurysm. Circ Cardiovasc Genet. 2008;1:39–42. doi: 10.1161/CIRCGENETICS.108.789727.
16.
Helgadottir A, Thorleifsson G, Magnusson KP, et al. The same sequence variant on 9p21 associates with myocardial infarction, abdominal aortic aneurysm and intracranial aneurysm. Nat Genet. 2008;40:217–224. doi: 10.1038/ng.72.
17.
Musunuru K, Post WS, Herzog W, et al. Association of single nucleotide polymorphisms on chromosome 9p21.3 with platelet reactivity: a potential mechanism for increased vascular disease. Circ Cardiovasc Genet. 2010;3:445–453. doi: 10.1161/CIRCGENETICS.109.923508.
18.
Fan M, Dandona S, McPherson R, Allayee H, Hazen SL, Wells GA, Roberts R, Stewart AF. Two chromosome 9p21 haplotype blocks distinguish between coronary artery disease and myocardial infarction risk. Circ Cardiovasc Genet. 2013;6:372–380. doi: 10.1161/CIRCGENETICS.113.000104.
19.
Ernst FD, Uhr K, Teumer A, Fanghänel J, Schulz S, Noack B, Gonzales J, Reichert S, Eickholz P, Holtfreter B, Meisel P, Linden GJ, Homuth G, Kocher T. Replication of the association of chromosomal region 9p21.3 with generalized aggressive periodontitis (gAgP) using an independent case-control cohort. BMC Med Genet. 2010;11:119. doi: 10.1186/1471-2350-11-119.
20.
Schaefer AS, Richter GM, Groessner-Schreiber B, Noack B, Nothnagel M, El Mokhtari NE, Loos BG, Jepsen S, Schreiber S. Identification of a shared genetic susceptibility locus for coronary heart disease and periodontitis. PLoS Genet. 2009;5:e1000378. doi: 10.1371/journal.pgen.1000378.
21.
Kathiresan S, Voight BF, Purcell S, et al. Genome-wide association of early-onset myocardial infarction with single nucleotide polymorphisms and copy number variants. Nat Genet. 2009;41:334–41. doi: 10.1038/ng.327.
22.
Peden JF, Hopewell JC, Saleheen D, et al. A genome-wide association study in Europeans and South Asians identifies five new loci for coronary artery disease. Nat Genet. 2011;43:339–44. doi: 10.1038/ng.782.
23.
Deloukas P, Kanoni S, Willenborg C, et al. Large-scale association analysis identifies new risk loci for coronary artery disease. Nat Genet. 2013;45:25–33. doi: 10.1038/ng.2480.
24.
Trégouët DA, König IR, Erdmann J, et al.; Wellcome Trust Case Control Consortium; Cardiogenics Consortium. Genome-wide haplotype association study identifies the SLC22A3-LPAL2-LPA gene cluster as a risk locus for coronary artery disease. Nat Genet. 2009;41:283–285. doi: 10.1038/ng.314.
25.
Nikpay M, Goel A, Won HH, et al.; CARDIoGRAMplusC4D Consortium. A comprehensive 1,000 genomes-based genome-wide association meta-analysis of coronary artery disease. Nat Genet. 2015;47:1121–1130. doi: 10.1038/ng.3396.
26.
Yang J, Lee SH, Goddard ME, Visscher PM. GCTA: a tool for genome-wide complex trait analysis. Am J Hum Genet. 2011;88:76–82. doi: 10.1016/j.ajhg.2010.11.011.
27.
Yang J, Ferreira T, Morris AP, Medland SE, Madden PA, Heath AC, Martin NG, Montgomery GW, Weedon MN, Loos RJ, Frayling TM, McCarthy MI, Hirschhorn JN, Goddard ME, Visscher PM; Genetic Investigation of ANthropometric Traits (GIANT) Consortium; DIAbetes Genetics Replication And Meta-analysis (DIAGRAM) Consortium. Conditional and joint multiple-SNP analysis of GWAS summary statistics identifies additional variants influencing complex traits. Nat Genet. 2012;44:369–375, S1–S3. doi: 10.1038/ng.2213.
28.
Maurano MT, Humbert R, Rynes E, et al. Systematic localization of common disease-associated variation in regulatory DNA. Science. 2012;337:1190–1195. doi: 10.1126/science.1222794.
29.
Won HH, Natarajan P, Dobbyn A, Jordan DM, Roussos P, Lage K, Raychaudhuri S, Stahl E, Do R. Disproportionate contributions of select genomic compartments and cell types to genetic risk for coronary artery disease. PLoS Genet. 2015;11:e1005622. doi: 10.1371/journal.pgen.1005622.
30.
Lonjou C, Zhang W, Collins A, Tapper WJ, Elahi E, Maniatis N, Morton NE. Linkage disequilibrium in human populations. Proc Natl Acad Sci U S A. 2003;100:6069–6074. doi: 10.1073/pnas.1031521100.
31.
Campbell MC, Tishkoff SA. African genetic diversity: implications for human demographic history, modern human origins, and complex disease mapping. Annu Rev Genomics Hum Genet. 2008;9:403–433. doi: 10.1146/annurev.genom.9.081307.164258.
32.
Hinds DA, Stuve LL, Nilsen GB, Halperin E, Eskin E, Ballinger DG, Frazer KA, Cox DR. Whole-genome patterns of common DNA variation in three human populations. Science. 2005;307:1072–1079. doi: 10.1126/science.1105436.
33.
Edwards SL, Beesley J, French JD, Dunning AM. Beyond GWASs: illuminating the dark road from association to function. Am J Hum Genet. 2013;93:779–797. doi: 10.1016/j.ajhg.2013.10.012.
34.
Marigorta UM, Navarro A. High trans-ethnic replicability of GWAS results implies common causal variants. PLoS Genet. 2013;9:e1003566. doi: 10.1371/journal.pgen.1003566.
35.
Franceschini N, Carty C, Bůzková P, et al. Association of genetic variants and incident coronary heart disease in multiethnic cohorts: the PAGE study. Circ Cardiovasc Genet. 2011;4:661–672. doi: 10.1161/CIRCGENETICS.111.960096.
36.
Kral BG, Mathias RA, Suktitipat B, Ruczinski I, Vaidya D, Yanek LR, Quyyumi AA, Patel RS, Zafari AM, Vaccarino V, Hauser ER, Kraus WE, Becker LC, Becker DM. A common variant in the CDKN2B gene on chromosome 9p21 protects against coronary artery disease in Americans of African ancestry. J Hum Genet. 2011;56:224–229. doi: 10.1038/jhg.2010.171.
37.
Lettre G, Palmer CD, Young T, et al. Genome-wide association study of coronary heart disease and its risk factors in 8,090 African Americans: the NHLBI CARe Project. PLoS Genet. 2011;7:e1001300. doi: 10.1371/journal.pgen.1001300.
38.
Carlson CS, Matise TC, North KE, et al.; PAGE Consortium. Generalization and dilution of association results from European GWAS in populations of non-European ancestry: the PAGE study. PLoS Biol. 2013;11:e1001661. doi: 10.1371/journal.pbio.1001661.
39.
Wojczynski MK, Li M, Bielak LF, et al. Genetics of coronary artery calcification among African Americans, a meta-analysis. BMC Med Genet. 2013;14:75. doi: 10.1186/1471-2350-14-75.
40.
Evans DM, Davey Smith G. Mendelian randomization: new applications in the coming age of hypothesis-free causality. Annu Rev Genomics Hum Genet. 2015;16:327–350. doi: 10.1146/annurev-genom-090314-050016.
41.
Kamstrup PR, Tybjaerg-Hansen A, Steffensen R, Nordestgaard BG. Genetically elevated lipoprotein(a) and increased risk of myocardial infarction. JAMA. 2009;301:2331–2339. doi: 10.1001/jama.2009.801.
42.
Erqou S, Kaptoge S, Perry PL, Di AE, Thompson A, White IR, Marcovina SM, Collins R, Thompson SG, Danesh J. Lipoprotein(a) concentration and the risk of coronary heart disease, stroke, and nonvascular mortality. JAMA. 2009;302:412–423. doi: 10.1001/jama.2009.1063.
43.
Clarke R, Peden JF, Hopewell JC, et al.; PROCARDIS Consortium. Genetic variants associated with Lp(a) lipoprotein level and coronary disease. N Engl J Med. 2009;361:2518–2528. doi: 10.1056/NEJMoa0902604.
44.
Thanassoulis G, Campbell CY, Owens DS, et al.; CHARGE Extracoronary Calcium Working Group. Genetic associations with valvular calcification and aortic stenosis. N Engl J Med. 2013;368:503–512. doi: 10.1056/NEJMoa1109034.
45.
Kamstrup PR, Tybjærg-Hansen A, Nordestgaard BG. Elevated lipoprotein(a) and risk of aortic valve stenosis in the general population. J Am Coll Cardiol. 2014;63:470–477. doi: 10.1016/j.jacc.2013.09.038.
46.
Koschinsky ML, Boffa MB. Lipoprotein(a): an important cardiovascular risk factor and a clinical conundrum. Endocrinol Metab Clin North Am. 2014;43:949–962. doi: 10.1016/j.ecl.2014.08.002.
47.
Abifadel M, Varret M, Rabès JP, et al. Mutations in PCSK9 cause autosomal dominant hypercholesterolemia. Nat Genet. 2003;34:154–156. doi: 10.1038/ng1161.
48.
Cohen JC, Boerwinkle E, Mosley TH, Hobbs HH. Sequence variations in PCSK9, low LDL, and protection against coronary heart disease. N Engl J Med. 2006;354:1264–1272. doi: 10.1056/NEJMoa054013.
49.
Cannon CP, Blazing MA, Giugliano RP, et al.; IMPROVE-IT Investigators. Ezetimibe added to statin therapy after acute coronary syndromes. N Engl J Med. 2015;372:2387–2397. doi: 10.1056/NEJMoa1410489.
50.
Stitziel NO, Won HH, Morrison AC, et al. Inactivating mutations in NPC1L1 and protection from coronary heart disease. N Engl J Med. 2014;371:2072–2082. doi: 10.1056/NEJMoa1405386.
51.
Lauridsen BK, Stender S, Frikke-Schmidt R, Nordestgaard BG, Tybjærg-Hansen A. Genetic variation in the cholesterol transporter NPC1L1, ischaemic vascular disease, and gallstone disease. Eur Heart J. 2015;36:1601–1608. doi: 10.1093/eurheartj/ehv108.
52.
Ference BA, Majeed F, Penumetcha R, Flack JM, Brook RD. Effect of naturally random allocation to lower low-density lipoprotein cholesterol on the risk of coronary heart disease mediated by polymorphisms in NPC1L1, HMGCR, or both: a 2 × 2 factorial Mendelian randomization study. J Am Coll Cardiol. 2015;65:1552–1561. doi: 10.1016/j.jacc.2015.02.020.
53.
Benn M, Nordestgaard BG, Grande P, Schnohr P, Tybjaerg-Hansen A. PCSK9 R46L, low-density lipoprotein cholesterol levels, and risk of ischemic heart disease: 3 independent studies and meta-analyses. J Am Coll Cardiol. 2010;55:2833–2842. doi: 10.1016/j.jacc.2010.02.044.
54.
Ference BA, Yoo W, Alesh I, Mahajan N, Mirowska KK, Mewada A, Kahn J, Afonso L, Williams KA, Flack JM. Effect of long-term exposure to lower low-density lipoprotein cholesterol beginning early in life on the risk of coronary heart disease: a Mendelian randomization analysis. J Am Coll Cardiol. 2012;60:2631–2639. doi: 10.1016/j.jacc.2012.09.017.
55.
Nordestgaard BG, Varbo A. Triglycerides and cardiovascular disease. Lancet. 2014;384:626–635. doi: 10.1016/S0140-6736(14)61177-6.
56.
Nordestgaard BG, Abildgaard S, Wittrup HH, Steffensen R, Jensen G, Tybjaerg-Hansen A. Heterozygous lipoprotein lipase deficiency: frequency in the general population, effect on plasma lipid levels, and risk of ischemic heart disease. Circulation. 1997;96:1737–1744.
57.
Wittrup HH, Tybjaerg-Hansen A, Abildgaard S, Steffensen R, Schnohr P, Nordestgaard BG. A common substitution (Asn291Ser) in lipoprotein lipase is associated with increased risk of ischemic heart disease. J Clin Invest. 1997;99:1606–1613. doi: 10.1172/JCI119323.
58.
Wittrup HH, Tybjaerg-Hansen A, Nordestgaard BG. Lipoprotein lipase mutations, plasma lipids and lipoproteins, and risk of ischemic heart disease. A meta-analysis. Circulation. 1999;99:2901–2907.
59.
Thomsen M, Varbo A, Tybjærg-Hansen A, Nordestgaard BG. Low nonfasting triglycerides and reduced all-cause mortality: a Mendelian randomization study. Clin Chem. 2014;60:737–746. doi: 10.1373/clinchem.2013.219881.
60.
Crosby J, Peloso GM, Auer PL, et al.; TG and HDL Working Group of the Exome Sequencing Project; National Heart, Lung, and Blood Institute. Loss-of-function mutations in APOC3, triglycerides, and coronary disease. N Engl J Med. 2014;371:22–231. doi: 10.1056/NEJMoa1307095.
61.
Jørgensen AB, Frikke-Schmidt R, Nordestgaard BG, Tybjærg-Hansen A. Loss-of-function mutations in APOC3 and risk of ischemic vascular disease. N Engl J Med. 2014;371:32–41. doi: 10.1056/NEJMoa1308027.
62.
Do R, Stitziel NO, Won HH, et al.; NHLBI Exome Sequencing Project. Exome sequencing identifies rare LDLR and APOA5 alleles conferring risk for myocardial infarction. Nature. 2015;518:102–106. doi: 10.1038/nature13917.
63.
Sarwar N, Sandhu MS, Ricketts SL, et al. Triglyceride-mediated pathways and coronary disease: collaborative analysis of 101 studies. Lancet. 2010;375:1634–1639. doi: 10.1016/S0140-6736(10)60545-4.
64.
Jørgensen AB, Frikke-Schmidt R, West AS, Grande P, Nordestgaard BG, Tybjærg-Hansen A. Genetically elevated non-fasting triglycerides and calculated remnant cholesterol as causal risk factors for myocardial infarction. Eur Heart J. 2013;34:1826–1833. doi: 10.1093/eurheartj/ehs431.
65.
Varbo A, Benn M, Tybjærg-Hansen A, Jørgensen AB, Frikke-Schmidt R, Nordestgaard BG. Remnant cholesterol as a causal risk factor for ischemic heart disease. J Am Coll Cardiol. 2013;61:427–436. doi: 10.1016/j.jacc.2012.08.1026.
66.
Do R, Willer CJ, Schmidt EM, et al. Common variants associated with plasma triglycerides and risk for coronary artery disease. Nat Genet. 2013;45:1345–1352. doi: 10.1038/ng.2795.
67.
Haase CL, Tybjærg-Hansen A, Grande P, Frikke-Schmidt R. Genetically elevated apolipoprotein A-I, high-density lipoprotein cholesterol levels, and risk of ischemic heart disease. J Clin Endocrinol Metab. 2010;95:E500–E510. doi: 10.1210/jc.2010-0450.
68.
Haase CL, Frikke-Schmidt R, Nordestgaard BG, Kateifides AK, Kardassis D, Nielsen LB, Andersen CB, Køber L, Johnsen AH, Grande P, Zannis VI, Tybjaerg-Hansen A. Mutation in APOA1 predicts increased risk of ischaemic heart disease and total mortality without low HDL cholesterol levels. J Intern Med. 2011;270:136–146. doi: 10.1111/j.1365-2796.2011.02381.x.
69.
Haase CL, Frikke-Schmidt R, Nordestgaard BG, Tybjærg-Hansen A. Population-based resequencing of APOA1 in 10,330 individuals: spectrum of genetic variation, phenotype, and comparison with extreme phenotype approach. PLoS Genet. 2012;8:e1003063. doi: 10.1371/journal.pgen.1003063.
70.
Hovingh GK, Brownlie A, Bisoendial RJ, Dube MP, Levels JH, Petersen W, Dullaart RP, Stroes ES, Zwinderman AH, de Groot E, Hayden MR, Kuivenhoven JA, Kastelein JJ. A novel apoA-I mutation (L178P) leads to endothelial dysfunction, increased arterial wall thickness, and premature coronary artery disease. J Am Coll Cardiol. 2004;44:1429–1435. doi: 10.1016/j.jacc.2004.06.070.
71.
Hazenberg AJ, Dikkers FG, Hawkins PN, Bijzet J, Rowczenio D, Gilbertson J, Posthumus MD, Leijsma MK, Hazenberg BP. Laryngeal presentation of systemic apolipoprotein A-I-derived amyloidosis. Laryngoscope. 2009;119:608–615. doi: 10.1002/lary.20106.
72.
Zacho J, Tybjaerg-Hansen A, Jensen JS, Grande P, Sillesen H, Nordestgaard BG. Genetically elevated C-reactive protein and ischemic vascular disease. N Engl J Med. 2008;359:1897–1908. doi: 10.1056/NEJMoa0707402.
73.
Schick UM, Auer PL, Bis JC, et al.; Cohorts for Heart and Aging Research in Genomic Epidemiology; National Heart, Lung, and Blood Institute GO Exome Sequencing Project. Association of exome sequences with plasma C-reactive protein levels in >9000 participants. Hum Mol Genet. 2015;24:559–571. doi: 10.1093/hmg/ddu450.
74.
Mallat Z, Lambeau G, Tedgui A. Lipoprotein-associated and secreted phospholipases A2 in cardiovascular disease: roles as biological effectors and biomarkers. Circulation. 2010;122:2183–2200. doi: 10.1161/CIRCULATIONAHA.110.936393.
75.
Thompson A, Gao P, Orfei L, et al. Lipoprotein-associated phospholipase A(2) and risk of coronary disease, stroke, and mortality: collaborative analysis of 32 prospective studies. Lancet. 2010;375:1536–1544. doi: 10.1016/S0140-6736(10)60319-4.
76.
Casas JP, Ninio E, Panayiotou A, et al. PLA2G7 genotype, lipoprotein-associated phospholipase A2 activity, and coronary heart disease risk in 10 494 cases and 15 624 controls of European Ancestry. Circulation. 2010;121:2284–2293. doi: 10.1161/CIRCULATIONAHA.109.923383.
77.
Holmes MV, Simon T, Exeter HJ, et al. Secretory phospholipase A(2)-IIA and cardiovascular disease: a Mendelian randomization study. J Am Coll Cardiol. 2013;62:1966–1976. doi: 10.1016/j.jacc.2013.06.044.
78.
Nicholls SJ, Kastelein JJ, Schwartz GG, Bash D, Rosenson RS, Cavender MA, Brennan DM, Koenig W, Jukema JW, Nambi V, Wright RS, Menon V, Lincoff AM, Nissen SE; VISTA-16 Investigators. Varespladib and cardiovascular events in patients with an acute coronary syndrome: the VISTA-16 randomized clinical trial. JAMA. 2014;311:252–262. doi: 10.1001/jama.2013.282836.
79.
White HD, Held C, Stewart R, et al. Darapladib for preventing ischemic events in stable coronary heart disease. N Engl J Med. 2014;370:1702–1711. doi: 10.1056/NEJMoa1315878.
80.
Johannsen TH, Frikke-Schmidt R, Schou J, Nordestgaard BG, Tybjærg-Hansen A. Genetic inhibition of CETP, ischemic vascular disease and mortality, and possible adverse effects. J Am Coll Cardiol. 2012;60:2041–2048. doi: 10.1016/j.jacc.2012.07.045.
81.
Thompson A, Di Angelantonio E, Sarwar N, Erqou S, Saleheen D, Dullaart RP, Keavney B, Ye Z, Danesh J. Association of cholesteryl ester transfer protein genotypes with CETP mass and activity, lipid levels, and coronary risk. JAMA. 2008;299:2777–2788. doi: 10.1001/jama.299.23.2777.
82.
Ridker PM, Paré G, Parker AN, Zee RY, Miletich JP, Chasman DI. Polymorphism in the CETP gene region, HDL cholesterol, and risk of future myocardial infarction: genomewide analysis among 18 245 initially healthy women from the Women’s Genome Health Study. Circ Cardiovasc Genet. 2009;2:26–33. doi: 10.1161/CIRCGENETICS.108.817304.
83.
Voight BF, Peloso GM, Orho-Melander M, et al. Plasma HDL cholesterol and risk of myocardial infarction: a Mendelian randomisation study. Lancet. 2012;380:572–580. doi: 10.1016/S0140-6736(12)60312-2.
84.
Inazu A, Koizumi J, Kajinami K, Kiyohar T, Chichibu K, Mabuchi H. Opposite effects on serum cholesteryl ester transfer protein levels between long-term treatments with pravastatin and probucol in patients with primary hypercholesterolemia and xanthoma. Atherosclerosis. 1999;145:405–413.
85.
Kuivenhoven JA, Jukema JW, Zwinderman AH, de Knijff P, McPherson R, Bruschke AV, Lie KI, Kastelein JJ. The role of a common variant of the cholesteryl ester transfer protein gene in the progression of coronary atherosclerosis. The Regression Growth Evaluation Statin Study Group. N Engl J Med. 1998;338:86–93. doi: 10.1056/NEJM199801083380203.
86.
Regieli JJ, Jukema JW, Grobbee DE, Kastelein JJ, Kuivenhoven JA, Zwinderman AH, van der Graaf Y, Bots ML, Doevendans PA. CETP genotype predicts increased mortality in statin-treated men with proven cardiovascular disease: an adverse pharmacogenetic interaction. Eur Heart J. 2008;29:2792–2799. doi: 10.1093/eurheartj/ehn465.
87.
Shah S, Casas JP, Drenos F, et al. Causal relevance of blood lipid fractions in the development of carotid atherosclerosis: Mendelian randomization analysis. Circ Cardiovasc Genet. 2013;6:63–72. doi: 10.1161/CIRCGENETICS.112.963140.
88.
Brautbar A, Ballantyne CM, Lawson K, Nambi V, Chambless L, Folsom AR, Willerson JT, Boerwinkle E. Impact of adding a single allele in the 9p21 locus to traditional risk factors on reclassification of coronary heart disease risk and implications for lipid-modifying therapy in the Atherosclerosis Risk in Communities study. Circ Cardiovasc Genet. 2009;2:279–285. doi: 10.1161/CIRCGENETICS.108.817338.
89.
Davies RW, Dandona S, Stewart AF, Chen L, Ellis SG, Tang WH, Hazen SL, Roberts R, McPherson R, Wells GA. Improved prediction of cardiovascular disease based on a panel of single nucleotide polymorphisms identified through genome-wide association studies. Circ Cardiovasc Genet. 2010;3:468–474. doi: 10.1161/CIRCGENETICS.110.946269.
90.
Ripatti S, Tikkanen E, Orho-Melander M, et al. A multilocus genetic risk score for coronary heart disease: case-control and prospective cohort analyses. Lancet. 2010;376:1393–1400. doi: 10.1016/S0140-6736(10)61267-6.
91.
Ganna A, Magnusson PK, Pedersen NL, de Faire U, Reilly M, Arnlöv J, Sundström J, Hamsten A, Ingelsson E. Multilocus genetic risk scores for coronary heart disease prediction. Arterioscler Thromb Vasc Biol. 2013;33:2267–2272. doi: 10.1161/ATVBAHA.113.301218.
92.
Tikkanen E, Havulinna AS, Palotie A, Salomaa V, Ripatti S. Genetic risk prediction and a 2-stage risk screening strategy for coronary heart disease. Arterioscler Thromb Vasc Biol. 2013;33:2261–2266. doi: 10.1161/ATVBAHA.112.301120.
93.
Mega JL, Stitziel NO, Smith JG, et al. Genetic risk, coronary heart disease events, and the clinical benefit of statin therapy: an analysis of primary and secondary prevention trials. Lancet. 2015;385:2264–2271. doi: 10.1016/S0140-6736(14)61730-X.
94.
Casas JP, Cooper J, Miller GJ, Hingorani AD, Humphries SE. Investigating the genetic determinants of cardiovascular disease using candidate genes and meta-analysis of association studies. Ann Hum Genet. 2006;70:145–169. doi: 10.1111/j.1469-1809.2005.00241.x.
95.
Wang L, Fan C, Topol SE, Topol EJ, Wang Q. Mutation of MEF2A in an inherited disorder with features of coronary artery disease. Science. 2003;302:1578–1581. doi: 10.1126/science.1088477.
96.
Weng L, Kavaslar N, Ustaszewska A, Doelle H, Schackwitz W, Hébert S, Cohen JC, McPherson R, Pennacchio LA. Lack of MEF2A mutations in coronary artery disease. J Clin Invest. 2005;115:1016–1020. doi: 10.1172/JCI24186.
97.
Chong JX, Buckingham KJ, Jhangiani SN, et al.; Centers for Mendelian Genomics. The genetic basis of Mendelian phenotypes: discoveries, challenges, and opportunities. Am J Hum Genet. 2015;97:199–215. doi: 10.1016/j.ajhg.2015.06.009.
98.
Arndt AK, MacRae CA. Genetic testing in cardiovascular diseases. Curr Opin Cardiol. 2014;29:235–240. doi: 10.1097/HCO.0000000000000055.
99.
McPherson R, Hegele RA. Ezetimibe: rescued by randomization (clinical and Mendelian). Arterioscler Thromb Vasc Biol. 2015;35:e13–e15. doi: 10.1161/ATVBAHA.114.305012.
100.
Moutsianas L, Agarwala V, Fuchsberger C, Flannick J, Rivas MA, Gaulton KJ, Albers PK, McVean G, Boehnke M, Altshuler D, McCarthy MI; GoT2D Consortium. The power of gene-based rare variant methods to detect disease-associated variation and test hypotheses about complex disease. PLoS Genet. 2015;11:e1005165. doi: 10.1371/journal.pgen.1005165.
101.
Cole CB, Nikpay M, McPherson R. Gene-environment interaction in dyslipidemia. Curr Opin Lipidol. 2015;26:133–138. doi: 10.1097/MOL.0000000000000160.
102.
Cole CB, Nikpay M, Lau P, Stewart AF, Davies RW, Wells GA, Dent R, McPherson R. Adiposity significantly modifies genetic risk for dyslipidemia. J Lipid Res. 2014;55:2416–2422. doi: 10.1194/jlr.P052522.
103.
Turner AW, Nikpay M, Silva A, Lau P, Martinuk A, Linseman TA, Soubeyrand S, McPherson R. Functional interaction between COL4A1/COL4A2 and SMAD3 risk loci for coronary artery disease. Atherosclerosis. 2015;242:543–552. doi: 10.1016/j.atherosclerosis.2015.08.008.
104.
Jia P, Wang L, Fanous AH, Pato CN, Edwards TL, Zhao Z; International Schizophrenia Consortium. Network-assisted investigation of combined causal signals from genome-wide association studies in schizophrenia. PLoS Comput Biol. 2012;8:e1002587. doi: 10.1371/journal.pcbi.1002587.
105.
Ghosh S, Vivar J, Nelson CP, et al. Systems genetics analysis of genome-wide association study reveals novel associations between key biological processes and coronary artery disease. Arterioscler Thromb Vasc Biol. 2015;35:1712–1722. doi: 10.1161/ATVBAHA.115.305513.
106.
Subramanian A, Tamayo P, Mootha VK, Mukherjee S, Ebert BL, Gillette MA, Paulovich A, Pomeroy SL, Golub TR, Lander ES, Mesirov JP. Gene set enrichment analysis: a knowledge-based approach for interpreting genome-wide expression profiles. Proc Natl Acad Sci U S A. 2005;102:15545–15550. doi: 10.1073/pnas.0506580102.
107.
Wang K, Li M, Bucan M. Pathway-based approaches for analysis of genomewide association studies. Am J Hum Genet. 2007;81:1278–1283. doi: 10.1086/522374.
108.
Wang K, Li M, Hakonarson H. Analysing biological pathways in genome-wide association studies. Nat Rev Genet. 2010;11:843–854. doi: 10.1038/nrg2884.
109.
Wang L, Jia P, Wolfinger RD, Chen X, Zhao Z. Gene set analysis of genome-wide association studies: methodological issues and perspectives. Genomics. 2011;98:1–8. doi: 10.1016/j.ygeno.2011.04.006.
110.
Segrè AV, Groop L, Mootha VK, Daly MJ, Altshuler D; DIAGRAM Consortium; MAGIC Investigators. Common inherited variation in mitochondrial genes is not enriched for associations with type 2 diabetes or related glycemic traits. PLoS Genet. 2010;6. doi: 10.1371/journal.pgen.1001058.
111.
Nam D, Kim J, Kim SY, Kim S. GSA-SNP: a general approach for gene set analysis of polymorphisms. Nucleic Acids Res. 2010;38:W749–W754. doi: 10.1093/nar/gkq428.
112.
Zhang K, Cui S, Chang S, Zhang L, Wang J. i-GSEA4GWAS: a web server for identification of pathways/gene sets associated with traits by applying an improved gene set enrichment analysis to genome-wide association study. Nucleic Acids Res. 2010;38:W90–W95. doi: 10.1093/nar/gkq324.
113.
Matthews L, Gopinath G, Gillespie M, et al. Reactome knowledgebase of human biological pathways and processes. Nucleic Acids Res. 2009;37:D619–D622. doi: 10.1093/nar/gkn863.
114.
Björkegren JL, Kovacic JC, Dudley JT, Schadt EE. Genome-wide significant loci: how important are they? Systems genetics to understand heritability of coronary artery disease and other common complex disorders. J Am Coll Cardiol. 2015;65:830–845. doi: 10.1016/j.jacc.2014.12.033.
115.
Oti M, Brunner HG. The modular nature of genetic diseases. Clin Genet. 2007;71:1–11. doi: 10.1111/j.1399-0004.2006.00708.x.
116.
Feldman I, Rzhetsky A, Vitkup D. Network properties of genes harboring inherited disease mutations. Proc Natl Acad Sci U S A. 2008;105:4323–4328. doi: 10.1073/pnas.0701722105.
117.
Schmidt EF, Strittmatter SM. The CRMP family of proteins and their role in Sema3A signaling. Adv Exp Med Biol. 2007;600:1–11. doi: 10.1007/978-0-387-70956-7_1.
118.
van Gils JM, Derby MC, Fernandes LR, et al. The neuroimmune guidance cue netrin-1 promotes atherosclerosis by inhibiting the emigration of macrophages from plaques. Nat Immunol. 2012;13:136–143. doi: 10.1038/ni.2205.
119.
Oksala N, Pärssinen J, Seppälä I, Raitoharju E, Kholova I, Ivana K, Hernesniemi J, Lyytikäinen LP, Levula M, Mäkelä KM, Sioris T, Kähönen M, Laaksonen R, Hytönen V, Lehtimäki T. Association of neuroimmune guidance cue netrin-1 and its chemorepulsive receptor UNC5B with atherosclerotic plaque expression signatures and stability in human(s): Tampere Vascular Study (TVS). Circ Cardiovasc Genet. 2013;6:579–587. doi: 10.1161/CIRCGENETICS.113.000141.
120.
Wanschel A, Seibert T, Hewing B, Ramkhelawon B, Ray TD, van Gils JM, Rayner KJ, Feig JE, O’Brien ER, Fisher EA, Moore KJ. Neuroimmune guidance cue Semaphorin 3E is expressed in atherosclerotic plaques and regulates macrophage retention. Arterioscler Thromb Vasc Biol. 2013;33:886–893. doi: 10.1161/ATVBAHA.112.300941.
121.
van Gils JM, Ramkhelawon B, Fernandes L, Stewart MC, Guo L, Seibert T, Menezes GB, Cara DC, Chow C, Kinane TB, Fisher EA, Balcells M, Alvarez-Leite J, Lacy-Hulbert A, Moore KJ. Endothelial expression of guidance cues in vessel wall homeostasis dysregulation under proatherosclerotic conditions. Arterioscler Thromb Vasc Biol. 2013;33:911–919. doi: 10.1161/ATVBAHA.112.301155.
122.
Davydov EV, Goode DL, Sirota M, Cooper GM, Sidow A, Batzoglou S. Identifying a high fraction of the human genome to be under selective constraint using GERP++. PLoS Comput Biol. 2010;6:e1001025. doi: 10.1371/journal.pcbi.1001025.
123.
Esteller M. Non-coding RNAs in human disease. Nat Rev Genet. 2011;12:861–874. doi: 10.1038/nrg3074.
124.
Gillette TG, Hill JA. Readers, writers, and erasers: chromatin as the whiteboard of heart disease. Circ Res. 2015;116:1245–1253. doi: 10.1161/CIRCRESAHA.116.303630.
125.
Shea J, Agarwala V, Philippakis AA, Maguire J, Banks E, Depristo M, Thomson B, Guiducci C, Onofrio RC, Kathiresan S, Gabriel S, Burtt NP, Daly MJ, Groop L, Altshuler D; Myocardial Infarction Genetics Consortium. Comparing strategies to fine-map the association of common SNPs at chromosome 9p21 with type 2 diabetes and myocardial infarction. Nat Genet. 2011;43:801–805. doi: 10.1038/ng.871.
126.
Yang J, Huang T, Petralia F, Long Q, Zhang B, Argmann C, Zhao Y, Mobbs CV, Schadt EE, Zhu J, Tu Z; GTEx Consortium. Synchronized age-related gene expression changes across multiple tissues in human and the link to complex diseases. Sci Rep. 2015;5:15145. doi: 10.1038/srep15145.
127.
Zhang W, Gamazon ER, Zhang X, Konkashbaev A, Liu C, Szilágyi KL, Dolan ME, Cox NJ. SCAN database: facilitating integrative analyses of cytosine modification and expression QTL. Database (Oxford). 2015, 1–7. doi: 10.1093/database/bav025.
128.
Brown CD, Mangravite LM, Engelhardt BE. Integrative modeling of eQTLs and cis-regulatory elements suggests mechanisms underlying cell type specificity of eQTLs. PLoS Genet. 2013;9:e1003649. doi: 10.1371/journal.pgen.1003649.
129.
ENCODE Project Consortium. A user’s guide to the Encyclopedia of DNA Elements (ENCODE). PLoS Biol. 2011;9:e1001046. doi: 10.1371/journal.pbio.1001046.
130.
Chadwick LH. The NIH Roadmap Epigenomics Program data resource. Epigenomics. 2012;4:317–324. doi: 10.2217/epi.12.18.
131.
Park PJ. ChIP-seq: advantages and challenges of a maturing technology. Nat Rev Genet. 2009;10:669–680. doi: 10.1038/nrg2641.
132.
Raychaudhuri S. Mapping rare and common causal alleles for complex human diseases. Cell. 2011;147:57–69. doi: 10.1016/j.cell.2011.09.011.
133.
Wang Z, Zang C, Rosenfeld JA, Schones DE, Barski A, Cuddapah S, Cui K, Roh TY, Peng W, Zhang MQ, Zhao K. Combinatorial patterns of histone acetylations and methylations in the human genome. Nat Genet. 2008;40:897–903. doi: 10.1038/ng.154.
134.
Mercer TR, Edwards SL, Clark MB, Neph SJ, Wang H, Stergachis AB, John S, Sandstrom R, Li G, Sandhu KS, Ruan Y, Nielsen LK, Mattick JS, Stamatoyannopoulos JA. DNase I-hypersensitive exons colocalize with promoters and distal regulatory elements. Nat Genet. 2013;45:852–859. doi: 10.1038/ng.2677.
135.
Buenrostro JD, Wu B, Chang HY, Greenleaf WJ. ATAC-seq: a method for assaying chromatin accessibility genome-wide. Curr Protoc Mol Biol. 2015;109:21.29.1–21.29.9. doi: 10.1002/0471142727.mb2129s109.
136.
Boyle AP, Song L, Lee BK, London D, Keefe D, Birney E, Iyer VR, Crawford GE, Furey TS. High-resolution genome-wide in vivo footprinting of diverse transcription factors in human cells. Genome Res. 2011;21:456–464. doi: 10.1101/gr.112656.110.
137.
Musunuru K, Strong A, Frank-Kamenetsky M, et al. From noncoding variant to phenotype via SORT1 at the 1p13 cholesterol locus. Nature. 2010;466:714–719. doi: 10.1038/nature09266.
138.
Strong A, Ding Q, Edmondson AC, et al. Hepatic sortilin regulates both apolipoprotein B secretion and LDL catabolism. J Clin Invest. 2012;122:2807–2816. doi: 10.1172/JCI63563.
139.
Sazonova O, Zhao Y, Nürnberg S, Miller C, Pjanic M, Castano VG, Kim JB, Salfati EL, Kundaje AB, Bejerano G, Assimes T, Yang X, Quertermous T. Characterization of TCF21 downstream target regions identifies a transcriptional network linking multiple independent coronary artery disease loci. PLoS Genet. 2015;11:e1005202. doi: 10.1371/journal.pgen.1005202.
140.
Nurnberg ST, Cheng K, Raiesdana A, et al. Coronary artery disease associated transcription factor TCF21 regulates smooth muscle precursor cells that contribute to the fibrous cap. PLoS Genet. 2015;11:e1005155. doi: 10.1371/journal.pgen.1005155.
141.
Beaudoin M, Gupta RM, Won HH, Lo KS, Do R, Henderson CA, Lavoie-St-Amour C, Langlois S, Rivas D, Lehoux S, Kathiresan S, Tardif JC, Musunuru K, Lettre G. Myocardial infarction-associated SNP at 6p24 interferes with MEF2 binding and associates with PHACTR1 expression levels in human coronary arteries. Arterioscler Thromb Vasc Biol. 2015;35:1472–1479. doi: 10.1161/ATVBAHA.115.305534.
142.
Pu X, Xiao Q, Kiechl S, et al. ADAMTS7 cleavage and vascular smooth muscle cell migration is affected by a coronary-artery-disease-associated variant. Am J Hum Genet. 2013;92:366–374. doi: 10.1016/j.ajhg.2013.01.012.
143.
Bauer RC, Tohyama J, Cui J, Cheng L, Yang J, Zhang X, Ou K, Paschos GK, Zheng XL, Parmacek MS, Rader DJ, Reilly MP. Knockout of Adamts7, a novel coronary artery disease locus in humans, reduces atherosclerosis in mice. Circulation. 2015;131:1202–1213. doi: 10.1161/CIRCULATIONAHA.114.012669.

eLetters(0)

eLetters should relate to an article recently published in the journal and are not a forum for providing unpublished data. Comments are reviewed for appropriate use of tone and language. Comments are not peer-reviewed. Acceptable comments are posted to the journal website only. Comments are not published in an issue and are not indexed in PubMed. Comments should be no longer than 500 words and will only be posted online. References are limited to 10. Authors of the article cited in the comment will be invited to reply, as appropriate.

Comments and feedback on AHA/ASA Scientific Statements and Guidelines should be directed to the AHA/ASA Manuscript Oversight Committee via its Correspondence page.

Information & Authors

Information

Published In

Go to Circulation Research
Go to Circulation Research

This target diagram depicts some candidate “triggers” for the atherogenic process in the bull’s-eye (reddish orange). These inciting factors include low-density lipoprotein (LDL), angiotensin II (Ang II), triglyceride-rich lipoproteins (TGRL), and putative antigens that may drive the adaptive immune response. The second circle (orange) portrays possible mediators that link the potential triggers to altered behavior of cells intrinsic to the arterial wall (endothelial cells and smooth muscle cells, SMC) as well as of leukocytes of various classes that reside in or enter the artery wall during atherogenesis. These mediators include reactive oxygen species (ROS) and reactive nitrogen species (RNS). The next ring (light orange) depicts cells whose functions can promote atherogenesis when encountering the mediators such as those shown in the orange ring. “Th” refers to helper T cells. “B” refers to B lymphocytes. The outermost ring (blue) shows cells and mediators implicated in modulating atherogenesis. These include regulatory T cells (Treg) that secrete transforming growth factor beta (TGF-β) and reparative macrophages. This pictorial representation that depicts reparative, intermediate, and classical macrophages intends to convey a fuller palette of mononuclear phagocytesubtypes than captured in the simplistic concept of an “M1/M2” dichotomy. Ly6C refers to a surface marker found on mouse monocytes. In humans, a CD14+ CD16low monocyte subset corresponds roughly to the Ly6Chigh monocyte population, while CD14low/CD16high monocytes may resemble the Ly6Clow population in mice. We thank Richard N. Mitchell, MD, PhD, Brigham and Women’s Hospital, for the photomicrograph of a cholesterol crystal-rich human coronary artery atherosclerotic plaque section that provides the background for this figure. See related article, page 531.

Circulation Research
Pages: 564 - 578
PubMed: 26892958

History

Received: 3 November 2015
Revision received: 28 December 2015
Accepted: 6 January 2016
Published online: 19 February 2016
Published in print: 19 February 2016

Permissions

Request permissions for this article.

Keywords

  1. coronary artery disease
  2. genetics
  3. genome-wide association study
  4. Mendelian randomization analysis
  5. polymorphism, single nucleotide

Subjects

Authors

Affiliations

Ruth McPherson
From the Department of Medicine, Atherogenomics Laboratory, Division of Cardiology, Ruddy Canadian Cardiovascular Genetics Center, University of Ottawa Heart Institute, Ottawa, Ontario, Canada (R.M.); and Department of Clinical Biochemistry, Rigshospitalet, Copenhagen University Hospital, and Faculty of Health and Medical Sciences, University of Copenhagen, Copenhagen, Denmark (A.T.-H.).
Anne Tybjaerg-Hansen
From the Department of Medicine, Atherogenomics Laboratory, Division of Cardiology, Ruddy Canadian Cardiovascular Genetics Center, University of Ottawa Heart Institute, Ottawa, Ontario, Canada (R.M.); and Department of Clinical Biochemistry, Rigshospitalet, Copenhagen University Hospital, and Faculty of Health and Medical Sciences, University of Copenhagen, Copenhagen, Denmark (A.T.-H.).

Notes

Correspondence to Ruth McPherson, MD, PhD, University of Ottawa Heart Institute, 40 Ruskin St, H4203, Ottawa, Canada K1Y 4W7. E-mail [email protected]

Disclosures

R. McPherson has received funding from Merck, Pfizer, Sanofi, and Amgen. A. Tybjaerg-Hansen has received honoraria from Eli Lilly and LGC Genomics.

Sources of Funding

This study was supported by Canadian Institutes of Health Research MOP-136936 and Heart and Stroke Foundation of Canada BR-7519 and T-7268 (to R. McPherson).

Metrics & Citations

Metrics

Citations

Download Citations

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Select your manager software from the list below and click Download.

  1. Genetic Insights Into Coronary Microvascular Disease, Microcirculation, 32, 1, (2025).https://doi.org/10.1111/micc.12896
    Crossref
  2. Exploring the Factors Influencing Coronary Heart Disease Prevalence in the US Population: A Retrospective Observational Study, Cureus, (2024).https://doi.org/10.7759/cureus.62741
    Crossref
  3. Single-cell ‘omic profiles of human aortic endothelial cells in vitro and human atherosclerotic lesions ex vivo reveal heterogeneity of endothelial subtype and response to activating perturbations, eLife, 12, (2024).https://doi.org/10.7554/eLife.91729.3
    Crossref
  4. Single-cell ‘omic profiles of human aortic endothelial cells in vitro and human atherosclerotic lesions ex vivo reveal heterogeneity of endothelial subtype and response to activating perturbations, eLife, 12, (2024).https://doi.org/10.7554/eLife.91729
    Crossref
  5. Risk Factors Associated with Cardiovascular Disorders, Pakistan BioMedical Journal, (03-10), (2024).https://doi.org/10.54393/pbmj.v7i02.1034
    Crossref
  6. The Benefits of Continuous Health Data Monitoring in Cardiovascular Diseases and Dementia, Encyclopedia of Information Science and Technology, Sixth Edition, (1-22), (2024).https://doi.org/10.4018/978-1-6684-7366-5.ch014
    Crossref
  7. Genetic Variants Linked to Myocardial Infarction in Individuals with Non-Alcoholic Fatty Liver Disease and Their Potential Interaction with Dietary Patterns, Nutrients, 16, 5, (602), (2024).https://doi.org/10.3390/nu16050602
    Crossref
  8. Beyond Natriuretic Peptides: Unveiling the Power of Emerging Biomarkers in Heart Failure, Biomolecules, 14, 3, (309), (2024).https://doi.org/10.3390/biom14030309
    Crossref
  9. Identification of patients with unstable angina based on coronary CT angiography: the application of pericoronary adipose tissue radiomics, Frontiers in Cardiovascular Medicine, 11, (2024).https://doi.org/10.3389/fcvm.2024.1462566
    Crossref
  10. Polygenic risk score predicts all-cause death in East Asian patients with prior coronary artery disease, Frontiers in Cardiovascular Medicine, 11, (2024).https://doi.org/10.3389/fcvm.2024.1296415
    Crossref
  11. See more
Loading...

View Options

View options

PDF and All Supplements

Download PDF and All Supplements

PDF/EPUB

View PDF/EPUB
Login options

Check if you have access through your login credentials or your institution to get full access on this article.

Personal login Institutional Login
Purchase Options

Purchase this article to access the full text.

Purchase access to this article for 24 hours

Genetics of Coronary Artery Disease
Circulation Research
  • Vol. 118
  • No. 4

Purchase access to this journal for 24 hours

Circulation Research
  • Vol. 118
  • No. 4
Restore your content access

Enter your email address to restore your content access:

Note: This functionality works only for purchases done as a guest. If you already have an account, log in to access the content to which you are entitled.

Figures

Tables

Media

Share

Share

Share article link

Share

Comment Response