Genetic and environmental factors contribute to a substantial variation in platelet function seen among normal persons. Candidate gene association studies represent a valiant effort to define the genetic component in an era where genetic tools were limited, but the single nucleotide polymorphisms identified in those studies need to be validated by more objective, comprehensive approaches, such as genome-wide association studies (GWASs) of quantitative functional traits in much larger cohorts of more carefully selected normal subjects. During the past year, platelet count and mean platelet volume, which indirectly affect platelet function, were the subjects of GWAS. The majority of the GWAS signals were located to noncoding regions, a consistent outcome of all GWAS to date, suggesting a major role for mechanisms that alter phenotype at the level of transcription or posttranscriptional modifications. Of 15 quantitative trait loci associated with mean platelet volume and platelet count, one located at 12q24 is also a risk locus for coronary artery disease. In most cases, the effect sizes of individual quantitative trait loci are admittedly small, but the results of these studies have led to new insight into regulators of hematopoiesis and megakaryopoiesis that would otherwise be unapparent and difficult to define.

Interindividual platelet responsiveness to a variety of agonists is highly variable, as documented in several studies of large cohorts of normal persons.1-12  At the same time, these and other studies of siblings, twins, and families with a history of coronary artery disease (CAD) have documented that intraindividual responsiveness is highly reproducible over time, regardless of the agonist tested or the chosen method of assessment. These findings strongly suggest that there is a high level of heritability of platelet function, and this has prompted numerous attempts to define the genetic basis for platelet function variability.

To fully appreciate the gene association studies, it is necessary to briefly summarize key aspects of platelet function and establish a molecular and physiologic basis for the selection of genes to study.

When a blood vessel is damaged, circulating platelets interact with components of the extracellular matrix, particularly collagen, and a complex series of receptor-ligand interactions ensues that ultimately leads to the formation of a stable platelet plug or thrombus. This process is a continuum of at least 3 phases that we can describe as initiation, extension, and consolidation, each of which entails the cooperation of a different group of receptors.

In the initiation phase, plasma von Willebrand factor (VWF) binds to collagen via its A3 domain and becomes structurally altered such that its A1 domain then binds to the platelet membrane receptor glycoprotein Ib-IX-V complex (GPIb complex). It is the GPIbα or larger subunit of GPIb that makes direct contact with VWF. This association is a requisite step in the adhesion of platelets to exposed thrombogenic surfaces at sites of vessel wall injury or in regions of atherosclerotic plaque rupture. Concurrently, a more stable platelet monolayer is formed on the collagen surface mediated predominantly by the platelet-specific receptor glycoprotein VI (GPVI) and platelet integrin α2β1.

The engagement of these receptors enhances platelet activation leading to the extension phase, mediated largely by the conversion of prothrombin to thrombin at the activated platelet surface and the secretion of active compounds from platelet granules (α-granules and δ-granules), which can further stimulate platelets. One of these, adenosine diphosphate (ADP), plays a particularly important role in the platelet response, binding to its cognate platelet receptors, the purinogenic receptors P2Y1 or P2Y12, to augment platelet activation. The activated platelet also produces and/or releases additional agonists, including the agonist thromboxane A2 (TXA2), which then binds to the platelet TXA2 receptor. Most of the receptors involved in the events of the extension phase are members of the G- protein–coupled receptor family.

Other agonists are at work during the extension phase, and each has the potential to initiate the involvement of additional platelets. Two prominent agonists are thrombin and epinephrine (EPI); thrombin (also known as factor II) binds to several receptors, but of particular interest to this discussion, thrombin binds the factor II receptor (F2R) that is also known as protease-activated receptor-1. EPI binds to the platelet α-adrenergic receptor 2A (A2AR). EPI can contribute to the initiation phase and, in low doses, is thought to prime platelets for enhanced activation by other agonists. The A2AR is a G-protein–coupled receptor and thus activates heterotrimeric G-proteins, including those containing the β3 subunit (GNB3).13  Alternative splicing variants of GNB3 appear to enhance G-protein signal transduction14,15  and could thus alter the effect of EPI on platelets. Thrombin makes a particularly relevant contribution to the augmentation of platelet activation during the extension phase because the activated platelet surface is a nidus for prothrombin conversion. Nonetheless, both thrombin and EPI, as well as ADP, are capable of binding to and activating the naive platelet in an alternative initiation phase.

In the consolidation phase, platelet-platelet cohesion (aggregation), mediated by the binding of fibrinogen and/or VWF to the activated platelet integrin αIIbβ3 (also known as GPIIb-IIIa), together with the assembly of a fibrin network, results in the generation of platelet-rich aggregates or thrombi. Further complexity is inherent in this final phase of platelet plug formation, and current research indicates an essential role for outside-in signaling through integrins and via receptor tyrosine kinases, including members of the Eph kinase family.

To this point in time, most attempts to identify genetic differences that influence platelet function in normal persons relied on a candidate gene approach. Target genes were rationally selected as candidates (Table 1) based on their established or logical involvement in any of the important molecular events that contribute to normal platelet function.

Table 1

Candidate gene SNPs used in association studies

GeneProtein identificationdbSNP or chromosomal location*CytobandMAF
ADRA2A α-2A-adrenergic receptor chr10: 112 839 580 10q24-q26 0.41 
F2R Coagulation factor II (thrombin) receptor; proteinase-activated receptor 1 (PAR-1) rs168753 5q13 0.14 
FCGR2A IgG Fc receptor type IIa chr1: 161 479 745 1q23 0.44 
GNB3 Guanine nucleotide-binding protein beta-3 subunit variant rs5443 12p13 0.45 
GP1BA Platelet glycoprotein Ib, α subunit rs6065 17pter-p12 0.10 
GP6 Platelet glycoprotein VI rs1613662 19q13.4 0.16 
ITGA2 Integrin subunit-α2 rs1126643 rs28095 rs1801106 5q11.2 0.38 0.36 0.08 
ITGA2B Integrin subunit-αIIb rs5911 17q21.32 0.41 
ITGB3 Integrin subunit-β3 rs5918 17q21.32 0.17 
P2RY1 Purinogenic receptor P2Y1 rs1065776 3q25.2 0.05 
P2RY12 Purinogenic receptor P2Y12 rs6809699 3q24-q25 0.14 
PTGS1 Prostaglandin-endoperoxide synthase 1; cyclooxygenase-1 (COX-1) rs3842787 9q32-q33.3 0.06 
TBXA2R Thromboxane A2 receptor rs1131882 rs4523 rs5758 19p13.3 0.13 0.30 0.45 
GeneProtein identificationdbSNP or chromosomal location*CytobandMAF
ADRA2A α-2A-adrenergic receptor chr10: 112 839 580 10q24-q26 0.41 
F2R Coagulation factor II (thrombin) receptor; proteinase-activated receptor 1 (PAR-1) rs168753 5q13 0.14 
FCGR2A IgG Fc receptor type IIa chr1: 161 479 745 1q23 0.44 
GNB3 Guanine nucleotide-binding protein beta-3 subunit variant rs5443 12p13 0.45 
GP1BA Platelet glycoprotein Ib, α subunit rs6065 17pter-p12 0.10 
GP6 Platelet glycoprotein VI rs1613662 19q13.4 0.16 
ITGA2 Integrin subunit-α2 rs1126643 rs28095 rs1801106 5q11.2 0.38 0.36 0.08 
ITGA2B Integrin subunit-αIIb rs5911 17q21.32 0.41 
ITGB3 Integrin subunit-β3 rs5918 17q21.32 0.17 
P2RY1 Purinogenic receptor P2Y1 rs1065776 3q25.2 0.05 
P2RY12 Purinogenic receptor P2Y12 rs6809699 3q24-q25 0.14 
PTGS1 Prostaglandin-endoperoxide synthase 1; cyclooxygenase-1 (COX-1) rs3842787 9q32-q33.3 0.06 
TBXA2R Thromboxane A2 receptor rs1131882 rs4523 rs5758 19p13.3 0.13 0.30 0.45 
*

NCBI SNP database (dbSNP) identification number (Build 131) or flanking nucleotide sequence (when SNP has not been assigned a dbSNP number).

CATTCCCAACTCTCTCTCTCTTTTT(G/A)AAGAAAAATGCTAAGGGCAGCCCTG.

AATGGAAAATCCCAGAAATTCTCCC(A/G)TTTGGATCCCACCTTCTCCATCCCA.

Most of the candidate gene association studies have focused on platelet surface receptors, for example, those involved in (1) fibrinogen-mediated platelet aggregation via integrin αIIbβ316-20 ; (2) collagen-induced platelet adhesion and signal transduction via integrin α2β121,22  or GPVI23,24 ; (3) VWF-mediated platelet adhesion via GPIbα25-28 ; or (4) platelet signaling induced by various physiologic agonists.3-5,9,29 

Many of these gene association studies were limited by small cohort sizes and/or a poor selection of single nucleotide polymorphisms (SNPs). The most appropriate SNP choices, or tagSNPs, are those that take into consideration linkage and reflect haplotype definitions.30  However, only a few of the foregoing studies used well-defined tagSNPs. Consequently, although the reported associations for ITGA2 C807T (rs1126643),22 GP1BA T-5C,27,31  and the GP6 haplotype b23  have been replicated, others have not been as reproducible.32,33 

Genes for the following proteins and their representative SNPs have been included in various gene association studies to assess variation in platelet function among normal persons (Table 1):

  1. α-Adrenergic receptor 2A (ADRA2A G1780A)29 ;

  2. F2R (A15992T)3 ;

  3. Immunoglobulin-γ Fc receptor IIA (FCRG2A H131R)34 ;

  4. GNB3 (C825T)6 ;

  5. GPIbα (GP1BA T-5C)27,35 ;

  6. GPVI (GP6 A13039G)23 ;

  7. Integrin subunit α2 (platelet GPIa) (ITGA2 C807T,36,37  C-52T,38  and G1648A)39 ;

  8. Integrin subunit αIIb (platelet GPIIb) (ITGA2B I843S)40 ;

  9. Integrin subunit β3 (platelet GPIIIa) (ITGB3 T196C)41 ;

  10. ADP receptor P2Y1 (P2RY1 C893T)42 ;

  11. ADP receptor P2Y12 (P2RY12 G52T)4 ;

  12. Prostaglandin G/H synthase 1 (cyclooxygenase-1) (PTGS1 C50T)43 ; and

  13. TXA2 receptor (TBXA2R C795T or C924T or G1686A).9 

The accuracy of an association study will depend on several factors, and 4 that are considered most relevant are selected here for discussion.

First, the selection of an accurate, reproducible, and sufficiently vetted measure of platelet function is probably the most critical decision. Several assays have been selected that fulfill these criteria, including light transmission aggregometry (LTA),6  the Platelet Function Analyzer-100 (PFA-100; Siemens Diagnostics),44  and the measurement of platelet activation events by flow cytometry,2  but each is used in association studies under carefully considered conditions that may not be identical to those used in the routine clinical laboratory.

Second, cohort sizes must be sufficiently large to provide adequate power to the statistical analyses. Unless adequate cohort sizes are studied, the chance of false-positive association results will increase, particularly when an SNP with a low minor allele frequency (MAF; eg, < 0.2) is studied. Two examples are the SNPs rs5918 and rs6065, which underlie the platelet alloantigen systems HPA-1 and HPA-2, respectively. Initial studies using inadequate cohort sizes concluded that these were novel risk factors for CAD, but a subsequent meta-analysis, with combined cohort sizes, has not confirmed these conclusions.45 

A third factor that has been largely ignored until recent studies is the ethnic homogeneity of the cohorts. It is now well documented that the presence of “ethnic outliers” between genotype groups can produce false positives if the MAF differs substantially between ethnic groups.46 

A fourth factor is appropriate SNP selection. Often, a single SNP does not represent the underlying sequence variation within a locus that may result from several haplotype blocks. The careful selection of an adequate number of tagSNPs, as well as a certain degree of resequencing, can ensure that this underlying sequence variation is adequately represented.2 

Yee et al7  have provided a useful paradigm whereby platelet hyperreactivity could be assessed using submaximal concentrations of EPI (0.4μM) or crosslinked collagen-related peptide (CRP-XL; 0.005 μg/mL) as agonists. Persons hyperreactive to one agonist are generally hyperreactive to others, and in the case of EPI-induced aggregation, enhanced reactivity is associated with female sex or higher plasma fibrinogen level. Day-to-day reproducibility of hyperreactivity to EPI under these conditions was very high, probably because of the use of such low agonist concentrations, and could be replicated over an extended period of time (> 2 years).

A substantial minority (14%) exhibited the hyperreactive phenotype (ie, ≥ 60% platelet aggregation induced by 0.04μM EPI). The levels of GPIbα, αIIbβ3, α2β1, or FcγRIIa were not different between the normal and hyperreactive cohorts, after adjustment for sex, fibrinogen level, platelet count, and mean platelet volume (MPV). A2AR mediates platelet activation by EPI,47,48  yet no association (P = .654) was observed between hyperreactivity and the ADRA2A G1838A polymorphism, previously linked to variation in EPI-mediated platelet aggregation in blacks.49  An explanation for the negative findings of Yee et al6  may be that the study did not take into account ethnic heterogeneity. On the other hand, hyperreactivity was associated with C825T of GNB3 (rs5443; MAF = 0.45), the gene for the β3 subunit of G-proteins, with a modest degree of statistical significance (P = .03).

The major weakness of platelet function testing in previous studies has always been that there is not one ex vivo assay that accurately predicts or reflects the functionality of the same platelet sample in vivo. Thus, a noteworthy finding of Yee et al6  is that, with the low-dose, EPI defined hyperreactivity by LTA identified persons with a generalized hyperreactivity that was reflected by diverse forms of platelet stimulation and different aspects of platelet function. Thus, hyperreactivity was manifested in LTA using different agonists, in the PFA-100 apparatus, wherein thrombus formation under high shear is induced with collagen plus EPI (CEPI) or collagen plus ADP, and by direct assay of activation events (P-selectin expression or binding of the activation-dependent, αIIbβ3-specific, murine monoclonal antibody PAC-1) using flow cytometry. Hyperreactivity was also consistently observed either in whole blood or platelet-rich plasma and using different anticoagulants, such as sodium citrate or D-phenylalanyl-L-prolyl-arginine chloromethyl ketone.

Yabe et al29  subsequently established that the ADRA2A SNP A1780G (MAF = 0.41) was associated with CEPI CT in the PFA-100, within a cohort of 211 Japanese males. The 1780AA genotype was associated with a shorter CT (ie, platelet hyper-reactivity). Even though the SNP designations are different in Yabe et el29  and Yee et al,6  it is probable that these represent the same SNPs because the flanking sequence of the major allele in each case (TTTAAA) represents a DraI restriction site. According to the current dbSNP database (Build 131), this polymorphism is located at chr10:112,839,580 but has not been named.

Kondkar et al50  used an unbiased RNA expression approach to identify genes that are up-regulated or down-regulated in platelets exhibiting hyperresponsiveness to low-dose EPI, as defined by LTA. A total of 290 genes were differentially expressed between subjects of differing platelet reactivity, with generally moderate or small effect sizes, suggesting that several genes contribute to this phenotype. One of the mRNA species that was differentially up-regulated in hyperresponsive platelets is that coding for vesicle-associated membrane protein 8/endobrevin, a member of a group of proteins known as v-SNAREs. This made perfect sense because platelet aggregation in response to EPI is absolutely dependent on the release of granule contents, and release is facilitated by the interaction of integral membrane proteins of the platelet plasma membrane and granule membrane, collectively known as t-SNAREs and v-SNAREs, respectively.51  Previous studies by others have shown that modulation of platelet SNAREs affects platelet function in vitro.52 

The PFA-100 measures platelet function initiated by exposure to collagen and ADP or EPI under high shear stress in citrated whole blood and is uniquely sensitive to genetic differences that modulate the initial stages of platelet activation, namely, those that are involved in the VWF-dependent adhesion of platelets to collagen and the enhanced adhesion mediated by GPVI and integrin α2β1.53,54  Harrison et al44  were the first to show that the PFA-100 is capable of detecting platelet hyperfunction, which is influenced by platelet count, hematocrit, and most significantly plasma VWF antigen level.55-57 

In a cohort of 123 healthy subjects, we measured the combined effects of plasma VWF level, platelet count, hematocrit, sex, and 7 candidate gene polymorphisms on baseline platelet reactivity in the PFA-100.58  This study was, at the time, the largest gene association study of PFA-100 reactivity in normal donors and confirmed that the PFA-100 can be used to compare normal donors based on their responsiveness to collagen combined with EPI or ADP. We compared the alleles of 7 candidate genes: ITGA2, ITGB3, GP1BA, GP6, P2RY1, P2RY12, and PTGS1 (Table 1).58  Based on linear and logistic regression models, we found an inverse correlation between baseline CEPI CT and plasma VWF antigen level, ITGA2 807T and P2RY1 893C. On the other hand, we observed an inverse correlation between baseline collagen plus ADP CT and P2RY1 893C or GP1BA-5C.

Whereas the preceding studies focused on surface receptor genes, others encoding relevant proteins in signaling cascades downstream of these receptors were not studied. The availability of high- throughput, cost-effective genomic sequencing enabled a more objective study of larger cohort sizes and the simultaneous evaluation of many more SNPs.

Through the comparison of a relatively large number of SNPs (n = 1327) in a selected group of several candidate genes (n = 97), Jones et al2  recently identified 17 independently associated SNPs that account for 48% of the variation in platelet reactivity induced by either ADP or CRP-XL. The success of this study reflects a careful attention to study design that establishes a paradigm for future genomic studies:

  1. The first consideration was cohort selection. The Platelet Function Cohort was established as a group of 500 healthy persons with a similar Northern European ancestry.8 

  2. Second, markers of platelet function were selected. Platelet function was compared by measuring fibrinogen binding to activated αIIbβ3, the requisite step for aggregation, and the expression of P-selectin, a marker of platelet degranulation, after induction with either the GPVI agonist, CRP-XL, or ADP. These agonists represent 2 distinct signaling pathways. GPVI signaling is an early event that acts via an FcRγ/ITAM pathway, and ADP amplifies platelet activation in later stages through the G protein-coupled receptors, P2Y1 and P2Y12.

  3. Marker selection. Based on a priori knowledge of the protein components of each of the signaling cascades, 97 candidate genes were selected. The exons of these genes were resequenced in 48 reference DNA samples from the Center d'Etude du Polymorphism Humain European Cohort to maximize the database of sequence variation, leading to the identification of 1327 SNPs that were then genotyped in the Platelet Function Cohort samples.

This association study confirmed that GP6 is a strong quantitative trait locus (QTL) for platelet response to CRP-XL, but the success of this study lay in the analysis of a large number of genes each exerting a small effect. Using a P value less than .005 as a cut-off, to overcome the errors inherent in multiple testing, 68 SNPs encompassed within 15 genes showed an association with the platelet response to CRP-XL, ADP, or both (Table 2).

Table 2

Loci and genes associated with platelet hyperreactivity as determined by flow cytometric measurement of platelet function and a functional genomics approach

GeneProtein functionSNP*CytobandFunctional trait
PAFAPCFC
GP6 Platelet glycoprotein VI rs1613662 19q13.4 ✓ — ✓ ✓ 
FCER1G Adapter protein for agonist receptor rs3557 1q23.3 — — ✓ ✓ 
PEAR1 Platelet surface receptor PEAR1 rs3737224 1q23.1 — — ✓ ✓ 
  rs11264579  ✓ ✓ — — 
RAF1 Protein kinase rs3729931 3p25.1  ✓ ✓ ✓ 
JAK2 Protein kinase rs10429491 9p24.1 ✓ ✓ — — 
P2RY12 Purinogenic (ADP) receptor P2Y12 rs1472122 3q25.1 ✓ ✓ ✓ — 
GNAZ Adapter protein for agonist receptor rs3788337 22q11.22 ✓ — — — 
VAV3 Intracellular signaling molecule rs17229705 1p13.3 ✓ ✓ ✓ — 
CD36 Platelet surface receptor rs1049654 7q21.11 — — ✓ ✓ 
MAP2K2 Protein kinase rs350916 19p13.3 — ✓ — — 
ITGA2 Integrin subunit-α2 rs41305896 5q11.2 ✓ ✓ — — 
AKT2 Protein kinase rs41275750 19q13.2   ✓ ✓ 
MAPK14 Protein kinase rs851007 6p21.31 ✓ ✓ ✓ — 
MAP2K4 Protein kinase rs41307923 17p12 — — ✓ ✓ 
ITPR1 Intracellular signaling molecule rs17786144 3p26.2 ✓ ✓ — — 
GeneProtein functionSNP*CytobandFunctional trait
PAFAPCFC
GP6 Platelet glycoprotein VI rs1613662 19q13.4 ✓ — ✓ ✓ 
FCER1G Adapter protein for agonist receptor rs3557 1q23.3 — — ✓ ✓ 
PEAR1 Platelet surface receptor PEAR1 rs3737224 1q23.1 — — ✓ ✓ 
  rs11264579  ✓ ✓ — — 
RAF1 Protein kinase rs3729931 3p25.1  ✓ ✓ ✓ 
JAK2 Protein kinase rs10429491 9p24.1 ✓ ✓ — — 
P2RY12 Purinogenic (ADP) receptor P2Y12 rs1472122 3q25.1 ✓ ✓ ✓ — 
GNAZ Adapter protein for agonist receptor rs3788337 22q11.22 ✓ — — — 
VAV3 Intracellular signaling molecule rs17229705 1p13.3 ✓ ✓ ✓ — 
CD36 Platelet surface receptor rs1049654 7q21.11 — — ✓ ✓ 
MAP2K2 Protein kinase rs350916 19p13.3 — ✓ — — 
ITGA2 Integrin subunit-α2 rs41305896 5q11.2 ✓ ✓ — — 
AKT2 Protein kinase rs41275750 19q13.2   ✓ ✓ 
MAPK14 Protein kinase rs851007 6p21.31 ✓ ✓ ✓ — 
MAP2K4 Protein kinase rs41307923 17p12 — — ✓ ✓ 
ITPR1 Intracellular signaling molecule rs17786144 3p26.2 ✓ ✓ — — 

PA indicates P-selectin expression in response to ADP; FA, fibrinogen binding in response to ADP; PC, P-selectin expression in response to CRP-XL; and FC, fibrinogen binding in response to CRP-XL.

Data from Jones et al2  with permission.

*

SNP exhibiting lowest P value.

Variation in GP6, as anticipated, accounts for up to 40% of the variation in GPVI signaling.8  We have determined that the common minor allele of GPVI is associated with decreased Fyn/Lyn binding and enhanced CaM binding to the cytoplasmic tail.24 

For the remaining genes, 18 independent associations in 15 loci were identified. These included the cell surface receptor genes CD36, GP6, ITGA2, PEAR1, and P2RY12, adapter proteins FCERG1 and GNAZ, kinases JAK2, MA2K2, MAP2K4, and MAPK14, and the intracellular signaling molecules ITPR1 and VAV3.

Of note, an association was not found between ITGB3 and any of the indices of platelet response. There is a great deal of conflicting evidence in the literature regarding ITGB3 rs5918, which defines the HPA-1 system.16-19,32,33  One possible explanation is that the MAF is low (0.15), and the majority of studies were severely underpowered by small cohort sizes. In addition, the diversity of functional assays and platelet agonists used in these studies certainly contributed to the discrepant findings.

Additional major findings resulted from this study. First, PEAR1 was identified as a regulator of both CRP-XL and ADP signaling, with 2 distinct associations observed. One SNP, rs41299597, was associated with CRP-XL induced pathways, whereas the more common SNP rs11264579 was associated with the ADP pathway. Second, rs170414 in ITPR1, which encodes the IP3 receptor, was associated exclusively with the ADP signaling pathway. Lastly, rs17229705 in VAV3 was associated with P-selectin expression only in response to ADP.

At this time, a genome-wide association study (GWAS) using a direct measure of platelet function has not been performed. However, certain quantitative properties of platelets have been reported to significantly influence outcomes in cardiovascular disease and stroke, and these can represent indirect measures of platelet function in normal persons. Examples are mean MPV and platelet count.

MPV is a highly heritable quantitative trait in humans,59  and the genetic component accounts for approximately 51% of the variation in MPV in a large primate study.60  MPV was shown to be an independent predictor of negative outcome in CAD and ischemic stroke in 4 of 6 studies,61-65  and a positive correlation was confirmed in a recent meta-analysis.66  A biologic explanation for this association with MPV rests on the hypothesis that platelets with increased MPV are more active than smaller platelets, with a greater prothrombotic potential because of higher levels of intracellular TXA2 and an increased procoagulant surface.67  However, this has been disputed recently in a large cohort of normal persons, where a negative correlation was observed between MPV and 2 markers of platelet activation, binding of fibrinogen and annexin V.68 

Nonetheless, a progression of 3 collaborative European-based GWAS on the effect of MPV has provided relevant information.68-70  These investigations incorporated subjects from the United Kingdom Blood Services Common Control (UKBS-CC1 and -CC2), the TwinsUK cohort, the Kooperative Gesundheitsforschung in der Region Augsburg (KORA) F3, F4, and S4 cohorts, the Study of Health in Pomerania cohort, and the Cambridge BioResource.

In one study, Soranzo et al68  analyzed Affymetrix 500K Gene Chip data in the UKBS-CC1 (discovery n = 1221) and replicated these findings in TwinsUK (n = 1050), KORA (n = 1601), UKBS-CC2 (n = 1304), and Cambridge BioResource (n = 3410). This study identified one locus on chromosome 7q22.3 as the first locus to be associated with MPV. The strongest association signal was located at rs342293 within an intergenic region (Table 3). The most proximal hypothetical gene FLJ36031 is located at a distance of approximately 71 kb, and 5 additional genes are located within a 1-MB interval encompassing rs342293. Intriguingly, all 6 genes are transcribed in megakaryocytes (MKs), PIK3CG is robustly transcribed in MKs but not in erythroblasts, and transcripts of 2 of the 6 genes, PIK3CG and PRKAR2B, are detectable at high levels in platelet RNA.

Table 3

Loci and genes associated with MPV and platelet count determined by GWAS

Trait/geneProteinSNPPositionCytoband
MPV     
    DNM3 Dynamin-3 rs10914144 170 216 373 1q24.3 
    TPM1 Tropomyosin α-1 chain rs11071720 61 129 049 15q22.1 
    BET1L Unknown function rs11602954 192 856 11p15.5 
    ARHGEF3 Rho guanine nucleotide exchange factor 3 rs12485738 56 840 816 3p21-p13 
    TMCC2 Unknown function rs1668873 203 502 613 1q32.1 
    TAOK1 Serine/threonine-protein kinase TAO1 rs2138852 24 727 475 17q11.2 
    JMJD1C JmjC domain-containing histone demethylation protein 2C rs2393967 64 803 162 10q21.2-q21.3 
    PIK3CG Phosphatidylinositol-4,5-biphosphate 3-kinase catalytic subunit γ-isoform rs342293 106 159 455 7q22.3 
    SIRPA CD172a, tyrosine-protein phosphatase nonreceptor type substrate 1 rs6136489 1 871 734 20p13 
    EHD3 EH-domain containing protein 3 rs647316 31 318 333 2p21 
    WDR66 WD repeat-containing protein 66 rs7961894 120 849 966 12q24.31 
    CD226 CD226 antigen rs893001 65 667 825 18q22.3 
PLT     
    ATXN2 Unknown function rs11065987 110 556 807 12q24 
    SH2B3 SH2B adapter protein 3 rs3184504  12q24 
    PTPN11 Tyrosine-protein phosphatase nonreceptor type 11 rs11066301 111 355 755 12q24 
    BAK1 Bcl-2 homologous antagonist/killer function rs210135 33 648 670 6p21.3 
    AK3 Adenylate kinase 3 rs385893 4 753 176 9p24.1-p24.3 
Trait/geneProteinSNPPositionCytoband
MPV     
    DNM3 Dynamin-3 rs10914144 170 216 373 1q24.3 
    TPM1 Tropomyosin α-1 chain rs11071720 61 129 049 15q22.1 
    BET1L Unknown function rs11602954 192 856 11p15.5 
    ARHGEF3 Rho guanine nucleotide exchange factor 3 rs12485738 56 840 816 3p21-p13 
    TMCC2 Unknown function rs1668873 203 502 613 1q32.1 
    TAOK1 Serine/threonine-protein kinase TAO1 rs2138852 24 727 475 17q11.2 
    JMJD1C JmjC domain-containing histone demethylation protein 2C rs2393967 64 803 162 10q21.2-q21.3 
    PIK3CG Phosphatidylinositol-4,5-biphosphate 3-kinase catalytic subunit γ-isoform rs342293 106 159 455 7q22.3 
    SIRPA CD172a, tyrosine-protein phosphatase nonreceptor type substrate 1 rs6136489 1 871 734 20p13 
    EHD3 EH-domain containing protein 3 rs647316 31 318 333 2p21 
    WDR66 WD repeat-containing protein 66 rs7961894 120 849 966 12q24.31 
    CD226 CD226 antigen rs893001 65 667 825 18q22.3 
PLT     
    ATXN2 Unknown function rs11065987 110 556 807 12q24 
    SH2B3 SH2B adapter protein 3 rs3184504  12q24 
    PTPN11 Tyrosine-protein phosphatase nonreceptor type 11 rs11066301 111 355 755 12q24 
    BAK1 Bcl-2 homologous antagonist/killer function rs210135 33 648 670 6p21.3 
    AK3 Adenylate kinase 3 rs385893 4 753 176 9p24.1-p24.3 

Data from Meisinger et al70  and Soranzo et al68  with permission.

The minor G allele at rs342293 (MAF = 0.45) is associated with increased MPV (P = 1.08 × 10−24) and with decreased platelet reactivity, based on the level of fibrinogen binding and the proportion of platelets binding annexin V, induced by CRP-XL and measured by flow cytometry. The negative correlation between the minor allele and the proportion of platelets binding annexin V appears to contradict the positive correlation between platelet volume and thrombogenicity, which had been proposed earlier.67  However, this single association between MPV and annexin V binding, with incorrect directionality, explains only 1.5% of the observed population variation in MPV and therefore does not represent disproof of the probable positive correlation between MPV and prothrombotic tendency.

Both PICK3CG and PRKAR2B are plausible candidates for genes that would affect MPV and platelet function. PICK3CG encodes the γ-chain of PI3/PI4-kinase, responsible for the synthesis of phosphatidylinositol-3,4,5-trisphosphate,71  which is a key component in the initiation of megakaryopoiesis and proplatelet formation and an essential intermediary in collagen-induced regulation of phospholipase C in megakaryocytes and platelets.72  PI3K-deficient mice display impaired ADP-induced platelet aggregation and are protected from ADP-induced platelet-dependent thromboembolic vascular occlusion.73 PRKAR2B encodes the β-chain of cAMP-dependent protein kinase, which attenuates the release of intracellular calcium by phosphorylation of ITPR3, the receptor for phosphatidylinositol-3,4,5-trisphosphate,74,75  and would thus have an antagonsitic effect on megakaryopoiesis and proplatelet formation.

In another GWAS, Meisinger et al70  evaluated Affymetrix 500K Gene Chip data in KORA F3 (discovery n = 1644) and replicated these findings in the UKBC-CC1 (n = 1203), KORA S4 (n = 4261), and Study of Health in Pomerania (n = 3300) cohorts. They found that MPV is strongly associated with 3 common SNPs in 3 other loci (Table 3): rs7961894 located within intron 3 of WDR66 on chromosome 12q24.31, rs12485738 upstream of the ARHGEF3 on chromosome 3p13-p21, and rs2138852 located upstream of TAOK1 on chromosome 17q11.2. These SNPs had P values of 7.24 × 10−48 for rs7961894, 3.81 × 10−27 for rs12485738, and 7.19 × 10−28 for rs2138852.

Each of these 3 genes is also a plausible candidate gene involved in regulation of MPV. In the case of WDR66, previous studies have shown that WD-repeat proteins are involved in the regulation of various cellular functions, cell-cycle regulation, and apoptosis.76  Moreover, in the study of Meisinger et al,70  expression analysis indicated a direct correlation of WDR66 transcripts and MPV. ARHGEF3 encodes the rho guanine-nucleotide exchange factor 3 (RhoGEF3), which activates RhoGTPases, which play an important role in the regulation of cell morphology, cell aggregation, cytoskeletal rearrangements, and transcriptional activation.77,78 TAOK1 encodes the TAO kinase 1 peptide (hTAOK1, also known as MARKK or PSK2), a microtubule affinity-regulating kinase important in regulation of mitotic progression, chromosome congression, and checkpoint-induced anaphase delay.79 

The 7q22.3 locus discovered by Soranzo et al68  explains 1.5% of the variance in MPV attributable to genetic factors, and the 3 QTL identified by together account for only 4% to 5% of the variance in MPV. Clearly, MPV is regulated by the cumulative effects of many more genes, so it was not unexpected that additional QTLs might be discovered that are associated with this quantitative trait. The discovery sample sizes in these 2 GWAS (n ∼ 1200 in Soranzo et al68  and ∼ 1600 in Meisinger et al70 ) are relatively modest. As noted by Weedon et al80  in a GWAS of adult height, the power for the discovery of QTLs is very limited in samples sizes less than 10 000.

The third study, a meta-analysis by Soranzo et al,69  analyzed larger numbers of persons, from essentially the same population cohorts. Eighty-eight SNPs that showed a nominal association with MPV in the discovery phase (n ∼ 4600) were taken forward in a replication study of additional healthy subjects (n ∼ 9300). This identified 12 QTLs for MPV, including the aforementioned 4 and 8 new loci (Table 3). In addition, they reported the first 3 loci associated with platelet count. Nine of the 12 MPV loci were also associated with platelet count, and in all cases, those alleles that were associated with an increase in MPV were associated with a decrease in platelet count. Overall, the fraction of genetic variance explained by each locus in regression models adjusted for sex and age was 8.6% for MPV traits and 0.5% for platelet count traits.

Two candidate genes within the newly identified MPV-associated loci have a known role in MK development and are expressed at higher levels in MK relative to erythroblasts: DNM3 on chromosome 1q24.3 and CD226 on 18q22.3.81,82  Four of the other loci map in or near 4 genes that are indirectly involved in hematopoiesis in humans: jumonji gene JMJDIC, which encodes a probable histone demethylase83 ; TPM1, which encodes tropomyosin I84 ; SIRPA, which encodes a protein that is involved in cell adhesion85 ; and EHD3, which encodes a mediator of protein transport and endocytosis.86 

Three other independent loci contain genes with a known influence on platelet count: BAK1 within the 6p21.3 locus encodes a proapoptotic protein that controls platelet life span87 ; the locus at 12q24.12 harbors PTPN11, SH2B3, and BRAP,88,89  which play a regulatory role in a wide array of cell-signaling events important in MK and other cells; and the locus at 9p24.1-p24.3 may regulate the transcription of the proximal gene JAK2, a key regulator of MK maturation, but this remains to be proved.90 

The possibility always exists that heterogeneity between population cohorts can obscure real association signals or even create false positives in GWAS that include moderate samples sizes (eg, n < 10 000). For population-based studies of quantitative traits, discovery sample sizes in excess of 20 000 are often needed. Nonetheless, the findings made in a statistically robust meta-analysis with adequate power should confirm the authenticity of true associations, and this was apparently the case in this series of studies culminating in the meta-analysis of Soranzo et al.69 

In conclusion, it must be acknowledged that none of the SNPs claimed to be associated with platelet function in normal subjects has been proven to be risk loci for CAD.91-94  However, of the 15 QTLs associated with MPV and platelet count, one located at 12q24 is also a risk locus for CAD.69 

In the case of MPV and platelet count, the majority of the GWAS signals are in noncoding regions, supporting the contention that the principle mechanisms by which sequence variation alters phenotype are at the level of transcription or splicing events. This has been a major outcome of all GWAS to date.95 

The effect sizes of individual QTL are small, and this might lead one to question the value in conducting such massive GWAS of platelet function. But the results of those studies already completed have led to new insight into regulators of hematopoiesis and megakaryopoiesis that would otherwise remained unexplored. It also represents a means to identify cis-acting regulatory motifs, of which the chr7 MPV-QTL is an excellent example. In addition, platelet function QTL studies have identified key regulators of signaling pathways. Further meta-analysis of MPV and platelet count in tens of thousands of subjects will obviously increase the number of MPV/platelet count QTLs, and the completion of the ongoing 1000 genomes initiative (tens of thousands by 2014) will serve to identify rare coding and noncoding variants. Some of these may exert a large effect on platelet phenotypes.

We caution the reader that knowledge of platelet QTLs for function, MPV, or platelet count should not be used in clinical practice to infer risk of thrombotic or bleeding events.

We thank our colleagues, Drs Santhi Ganesh and David Ginsburg (University of Michigan, Ann Arbor, MI), and Dr Nicholas Schork (Scripps Translational Science Institute, La Jolla, CA), for their insights.

This work was supported by the National Heart, Lung, and Blood Institute (grants HL075821 and HL086904; T.J.K.) and CHOC Foundation of the Children's Hospital of Orange County (D.J.N.).

National Institutes of Health

Contribution: T.J.K. wrote the paper; and D.J.N. contributed to the writing and editing of the paper.

Conflict-of-interest disclosure: The authors declare no competing financial interests.

Correspondence: Thomas J. Kunicki, The Scripps Research Institute, Department of Molecular and Experimental Medicine, 10550 North Torrey Pines Rd, Maildrop MEM-150, La Jolla, CA 92037; e-mail: tomk@scripps.edu.

1
O'Donnell
 
CJ
Larson
 
MG
Feng
 
D
, et al. 
Genetic and environmental contributions to platelet aggregation: the Framingham heart study.
Circulation
2001
, vol. 
103
 
25
(pg. 
3051
-
3056
)
2
Jones
 
CI
Bray
 
S
Garner
 
SF
, et al. 
A functional genomics approach reveals novel quantitative trait loci associated with platelet signaling pathways.
Blood
2009
, vol. 
114
 
7
(pg. 
1405
-
1416
)
3
Dupont
 
A
Fontana
 
P
Bachelot-Loza
 
C
, et al. 
An intronic polymorphism in the PAR-1 gene is associated with platelet receptor density and the response to SFLLRN.
Blood
2003
, vol. 
101
 
5
(pg. 
1833
-
1840
)
4
Fontana
 
P
Dupont
 
A
Gandrille
 
S
, et al. 
Adenosine diphosphate-induced platelet aggregation is associated with P2Y12 gene sequence variations in healthy subjects.
Circulation
2003
, vol. 
108
 
8
(pg. 
989
-
995
)
5
Hetherington
 
SL
Singh
 
RK
Lodwick
 
D
Thompson
 
JR
Goodall
 
AH
Samani
 
NJ
Dimorphism in the P2Y1 ADP receptor gene is associated with increased platelet activation response to ADP.
Arterioscler Thromb Vasc Biol
2005
, vol. 
25
 
1
(pg. 
252
-
257
)
6
Yee
 
DL
Bergeron
 
AL
Sun
 
CW
Dong
 
JF
Bray
 
PF
Platelet hyperreactivity generalizes to multiple forms of stimulation.
J Thromb Haemost
2006
, vol. 
4
 
9
(pg. 
2043
-
2050
)
7
Yee
 
DL
Sun
 
CW
Bergeron
 
AL
Dong
 
JF
Bray
 
PF
Aggregometry detects platelet hyperreactivity in healthy individuals.
Blood
2005
, vol. 
106
 
8
(pg. 
2723
-
2729
)
8
Jones
 
CI
Garner
 
SF
Angenent
 
W
, et al. 
Mapping the platelet profile for functional genomic studies and demonstration of the effect size of the GP6 locus.
J Thromb Haemost
2007
, vol. 
5
 
8
(pg. 
1756
-
1765
)
9
Fontana
 
P
Gandrille
 
S
Remones
 
V
, et al. 
Identification of functional polymorphisms of the thromboxane A2 receptor gene in healthy volunteers.
Thromb Haemost
2006
, vol. 
96
 
3
(pg. 
356
-
360
)
10
Bray
 
PF
Mathias
 
RA
Faraday
 
N
, et al. 
Heritability of platelet function in families with premature coronary artery disease.
J Thromb Haemost
2007
, vol. 
5
 
8
(pg. 
1617
-
1623
)
11
Panzer
 
S
Hocker
 
L
Koren
 
D
Agonists-induced platelet activation varies considerably in healthy male individuals: studies by flow cytometry.
Ann Hematol
2006
, vol. 
85
 
2
(pg. 
121
-
125
)
12
Gaxiola
 
B
Friedl
 
W
Propping
 
P
Epinephrine-induced platelet aggregation: a twin study.
Clin Genet
1984
, vol. 
26
 
6
(pg. 
543
-
548
)
13
Richardson
 
M
Robishaw
 
JD
The alpha2A-adrenergic receptor discriminates between Gi heterotrimers of different betagamma subunit composition in Sf9 insect cell membranes.
J Biol Chem
1999
, vol. 
274
 
19
(pg. 
13525
-
13533
)
14
Siffert
 
W
Rosskopf
 
D
Siffert
 
G
, et al. 
Association of a human G-protein beta3 subunit variant with hypertension.
Nat Genet
1998
, vol. 
18
 
1
(pg. 
45
-
48
)
15
Siffert
 
W
G protein polymorphisms in hypertension, atherosclerosis, and diabetes.
Annu Rev Med
2005
, vol. 
56
 (pg. 
17
-
28
)
16
Lasne
 
D
Krenn
 
M
Pingault
 
V
, et al. 
Interdonor variability of platelet response to thrombin receptor activation: influence of PIA2 polymorphism.
Br J Haematol
1997
, vol. 
99
 
4
(pg. 
801
-
807
)
17
Goodall
 
AH
Curzen
 
N
Panesar
 
M
, et al. 
Increased binding of fibrinogen to glycoprotein IIIa-proline33 (HPA-1b, PlA2, Zwb) positive platelets in patients with cardiovascular disease.
Eur Heart J
1999
, vol. 
20
 
10
(pg. 
742
-
747
)
18
Michelson
 
AD
Furman
 
MI
Goldschmidt-Clermont
 
P
, et al. 
Platelet GP IIIa PI(A) polymorphisms display different sensitivities to agonists.
Circulation
2000
, vol. 
101
 
9
(pg. 
1013
-
1018
)
19
Theodoropoulos
 
I
Christopoulos
 
C
Metcalfe
 
P
Dimitriadou
 
E
Economopoulos
 
P
Loucopoulos
 
D
The effect of human platelet alloantigen polymorphisms on the in vitro responsiveness to adrenaline and collagen.
Br J Haematol
2001
, vol. 
114
 
2
(pg. 
387
-
393
)
20
Vijayan
 
KV
Bray
 
PF
Molecular mechanisms of prothrombotic risk due to genetic variations in platelet genes: enhanced outside-in signaling through the Pro33 variant of integrin beta3.
Exp Biol Med (Maywood)
2006
, vol. 
231
 
5
(pg. 
505
-
513
)
21
Kunicki
 
TJ
Kritzik
 
M
Annis
 
DS
Nugent
 
DJ
Hereditary variation in platelet integrin alpha2·beta1 copy number is associated with two silent polymorphisms in the alpha2 gene coding sequence.
Blood
1997
, vol. 
89
 
6
(pg. 
1939
-
1943
)
22
Ajzenberg
 
N
Berroeta
 
C
Philip
 
I
, et al. 
Association of the -92C/G and 807C/T polymorphisms of the alpha2 subunit gene with human platelets alpha2beta1 receptor density.
Arterioscler Thromb Vasc Biol
2005
, vol. 
25
 
8
(pg. 
1756
-
1760
)
23
Joutsi-Korhonen
 
L
Smethurst
 
PA
Rankin
 
A
, et al. 
The low-frequency allele of the platelet collagen signaling receptor glycoprotein VI is associated with reduced functional responses and expression.
Blood
2003
, vol. 
101
 
11
(pg. 
4372
-
4379
)
24
Trifiro
 
E
Williams
 
SA
Cheli
 
Y
, et al. 
The low frequency isoform of platelet GPVI (GPVIb) attenuates ligand-mediated signal transduction but not receptor expression or ligand binding.
Blood
2009
, vol. 
114
 
9
(pg. 
1893
-
1899
)
25
Lopez
 
JA
Ludwig
 
EH
McCarthy
 
BJ
Polymorphism of human glycoprotein Ib α results from a variable number of tandem repeats of a 13-amino acid sequence in the mucin-like macroglycopeptide region.
J Biol Chem
1992
, vol. 
267
 
14
(pg. 
10055
-
10061
)
26
Ishida
 
F
Furihata
 
K
Ishida
 
K
, et al. 
The largest variant of platelet glycoprotein Ib αhas four tandem repeats of 13 amino acids in the macroglycopeptide region and a genetic linkage with methionine 145.
Blood
1995
, vol. 
86
 
4
(pg. 
1356
-
1360
)
27
Afshar-Kharghan
 
V
Li
 
CQ
Khoshnevis-Asl
 
M
Lopez
 
JA
Kozak sequence polymorphism of the glycoprotein (GP) Ibalpha gene is a major determinant of the plasma membrane levels of the platelet GP Ib-IX-V complex.
Blood
1999
, vol. 
94
 
1
(pg. 
186
-
191
)
28
Kuijpers
 
RWAM
Faber
 
NM
Cuypers
 
HTM
Ouwehand
 
WH
Von dem Borne
 
AEGK
NH 2-terminal globular domain of human platelet glycoprotein Ib α has a methionine 145/threonine 145 amino acid polymorphism, which is associated with the HPA-2 (Ko) alloantigens.
J Clin Invest
1992
, vol. 
89
 
2
(pg. 
381
-
384
)
29
Yabe
 
M
Matsubara
 
Y
Takahashi
 
S
, et al. 
Identification of ADRA2A polymorphisms related to shear-mediated platelet function.
Biochem Biophys Res Commun
2006
, vol. 
347
 
4
(pg. 
1001
-
1005
)
30
Watkins
 
NA
O'Connor
 
MN
Rankin
 
A
, et al. 
Definition of novel GP6 polymorphisms and major difference in haplotype frequencies between populations by a combination of in-depth exon resequencing and genotyping with tag single nucleotide polymorphisms.
J Thromb Haemost
2006
, vol. 
4
 
6
(pg. 
1197
-
1205
)
31
Kunicki
 
TJ
The influence of platelet collagen receptor polymorphisms in hemostasis and thrombotic disease.
Arterioscler Thromb Vasc Biol
2002
, vol. 
22
 
1
(pg. 
14
-
20
)
32
Bennett
 
JS
Catella-Lawson
 
F
Rut
 
AR
, et al. 
Effect of the Pl(A2) alloantigen on the function of beta(3)-integrins in platelets.
Blood
2001
, vol. 
97
 
10
(pg. 
3093
-
3099
)
33
Payne
 
KE
Bray
 
PF
Grant
 
PJ
Carter
 
AM
Beta3 integrin haplotype influences gene regulation and plasma von Willebrand factor activity.
Atherosclerosis
2008
, vol. 
198
 
2
(pg. 
280
-
286
)
34
van der Pol
 
W
van de Winkel
 
JG
IgG receptor polymorphisms: risk factors for disease.
Immunogenetics
1998
, vol. 
48
 
3
(pg. 
222
-
232
)
35
Kaski
 
S
Kekomaki
 
R
Partanen
 
J
Systematic screening for genetic polymorphism in human platelet glycoprotein Ib alpha.
Immunogenetics
1996
, vol. 
44
 
3
(pg. 
170
-
176
)
36
Kritzik
 
M
Savage
 
B
Nugent
 
DJ
Santoso
 
S
Ruggeri
 
ZM
Kunicki
 
TJ
Nucleotide polymorphisms in the alpha 2 gene define multiple alleles which are associated with differences in platelet alpha 2 beta 1.
Blood
1998
, vol. 
92
 
7
(pg. 
2382
-
2388
)
37
Di Paola
 
J
Jugessur
 
A
Goldman
 
T
, et al. 
Platelet glycoprotein Ibalpha and integrin alphabeta polymorphisms: gene frequencies and linkage disequilibrium in a population diversity panel.
J Thromb Haemost
2005
, vol. 
3
 
7
(pg. 
1511
-
1521
)
38
Jacquelin
 
B
Tarantino
 
MD
Kritzik
 
M
, et al. 
Allele-dependent transcriptional regulation of the human integrin alpha 2 gene.
Blood
2001
, vol. 
97
 
6
(pg. 
1721
-
1726
)
39
Kalb
 
R
Santoso
 
S
Unkelbach
 
K
Kiefel
 
V
Mueller-Eckhardt
 
C
Localization of the Br polymorphism on a 144 bp Exon of the GPIa gene and its application in platelet DNA typing.
Thromb Haemost
1994
, vol. 
71
 
5
(pg. 
651
-
654
)
40
Lyman
 
S
Aster
 
RH
Visentin
 
GP
Newman
 
PJ
Polymorphism of human platelet membrane glycoprotein IIb associated with the Bak a/Bak b alloantigen system.
Blood
1990
, vol. 
75
 
12
(pg. 
2343
-
2348
)
41
Newman
 
PJ
Derbes
 
RS
Aster
 
RH
The human platelet alloantigens, PLA1 and PLA2, are associated with a leucine33/proline33 amino acid polymorphism in membrane glycoprotein IIIa, and are distinguishable by DNA typing.
J Clin Invest
1989
, vol. 
83
 
5
(pg. 
1778
-
1781
)
42
Jefferson
 
BK
Foster
 
JH
McCarthy
 
JJ
, et al. 
Aspirin resistance and a single gene.
Am J Cardiol
2005
, vol. 
95
 
6
(pg. 
805
-
808
)
43
Maree
 
AO
Curtin
 
RJ
Chubb
 
A
, et al. 
Cyclooxygenase-1 haplotype modulates platelet response to aspirin.
J Thromb Haemost
2005
, vol. 
3
 
10
(pg. 
2340
-
2345
)
44
Harrison
 
P
Mackie
 
I
Mathur
 
A
, et al. 
Platelet hyper-function in acute coronary syndromes.
Blood Coagul Fibrinolysis
2005
, vol. 
16
 
8
(pg. 
557
-
562
)
45
Ye
 
Z
Liu
 
EH
Higgins
 
JP
, et al. 
Seven haemostatic gene polymorphisms in coronary disease: meta-analysis of 66,155 cases and 91,307 controls.
Lancet
2006
, vol. 
367
 
9511
(pg. 
651
-
658
)
46
Freedman
 
ML
Reich
 
D
Penney
 
KL
, et al. 
Assessing the impact of population stratification on genetic association studies.
Nat Genet
2004
, vol. 
36
 
4
(pg. 
388
-
393
)
47
Motulsky
 
HJ
Shattil
 
SJ
Insel
 
PA
Characterization of alpha 2-adrenergic receptors on human platelets using [3H]yohimbine.
Biochem Biophys Res Commun
1980
, vol. 
97
 
4
(pg. 
1562
-
1570
)
48
Hoffman
 
BB
Michel
 
T
Brenneman
 
TB
Lefkowitz
 
RJ
Interactions of agonists with platelet alpha 2-adrenergic receptors.
Endocrinology
1982
, vol. 
110
 
3
(pg. 
926
-
932
)
49
Freeman
 
K
Farrow
 
S
Schmaier
 
A
Freedman
 
R
Schork
 
T
Lockette
 
W
Genetic polymorphism of the alpha 2-adrenergic receptor is associated with increased platelet aggregation, baroreceptor sensitivity, and salt excretion in normotensive humans.
Am J Hypertens
1995
, vol. 
8
 
9
(pg. 
863
-
869
)
50
Kondkar
 
AA
Bray
 
MS
Leal
 
SM
, et al. 
VAMP8/endobrevin is overexpressed in hyperreactive human platelets: suggested role for platelet microRNA.
J Thromb Haemost
2010
, vol. 
8
 
2
(pg. 
369
-
378
)
51
Ren
 
Q
Ye
 
S
Whiteheart
 
SW
The platelet release reaction: just when you thought platelet secretion was simple.
Curr Opin Hematol
2008
, vol. 
15
 
5
(pg. 
537
-
541
)
52
Ren
 
Q
Barber
 
HK
Crawford
 
GL
, et al. 
Endobrevin/VAMP-8 is the primary v-SNARE for the platelet release reaction.
Mol Biol Cell
2007
, vol. 
18
 
1
(pg. 
24
-
33
)
53
Di Paola
 
J
Federici
 
AB
Sacchi
 
E
, et al. 
Low platelet alpha 2 beta 1 levels in type I von Willebrand disease correlate with impaired platelet function in a high shear stress system.
Blood
1999
, vol. 
93
 
11
(pg. 
3578
-
3582
)
54
Cheli
 
Y
Kunicki
 
TJ
hnRNP L regulates differences in expression of mouse integrin alpha2-beta1.
Blood
2006
, vol. 
107
 
11
(pg. 
4391
-
4398
)
55
Gum
 
PA
Kottke-Marchant
 
K
Poggio
 
ED
, et al. 
Profile and prevalence of aspirin resistance in patients with cardiovascular disease.
Am J Cardiol
2001
, vol. 
88
 
3
(pg. 
230
-
235
)
56
Homoncik
 
M
Jilma
 
B
Hergovich
 
N
Stohlawetz
 
P
Panzer
 
S
Speiser
 
W
Monitoring of aspirin (ASA) pharmacodynamics with the platelet function analyzer PFA-100.
Thromb Haemost
2000
, vol. 
83
 
2
(pg. 
316
-
321
)
57
Chakroun
 
T
Gerotziafas
 
G
Robert
 
F
, et al. 
In vitro aspirin resistance detected by PFA-100 closure time: pivotal role of plasma von Willebrand factor.
Br J Haematol
2004
, vol. 
124
 
1
(pg. 
80
-
85
)
58
Kunicki
 
TJ
Williams
 
SA
Salomon
 
DR
, et al. 
Genetics of platelet reactivity in normal, healthy individuals.
J Thromb Haemost
2009
, vol. 
7
 
12
(pg. 
2116
-
2122
)
59
Evans
 
DM
Frazer
 
IH
Martin
 
NG
Genetic and environmental causes of variation in basal levels of blood cells.
Twin Res
1999
, vol. 
2
 
4
(pg. 
250
-
257
)
60
Mahaney
 
MC
Brugnara
 
C
Lease
 
LR
Platt
 
OS
Genetic influences on peripheral blood cell counts: a study in baboons.
Blood
2005
, vol. 
106
 
4
(pg. 
1210
-
1214
)
61
Mayda-Domac
 
F
Misirli
 
H
Yilmaz
 
M
Prognostic role of mean platelet volume and platelet count in ischemic and hemorrhagic stroke.
J Stroke Cerebrovasc Dis
2010
, vol. 
19
 
1
(pg. 
66
-
72
)
62
Martin
 
JF
Bath
 
PM
Burr
 
ML
Mean platelet volume and myocardial infarction.
Lancet
1992
, vol. 
339
 
8799
(pg. 
1000
-
1001
)
63
Boos
 
CJ
Lip
 
GY
Assessment of mean platelet volume in coronary artery disease: what does it mean?
Thromb Res
2007
, vol. 
120
 
1
(pg. 
11
-
13
)
64
Bath
 
P
Algert
 
C
Chapman
 
N
Neal
 
B
Association of mean platelet volume with risk of stroke among 3134 individuals with history of cerebrovascular disease.
Stroke
2004
, vol. 
35
 
3
(pg. 
622
-
626
)
65
De Luca
 
G
Santagostino
 
M
Secco
 
GG
, et al. 
Mean platelet volume and the extent of coronary artery disease: results from a large prospective study.
Atherosclerosis
2009
, vol. 
206
 
1
(pg. 
292
-
297
)
66
Chu
 
SG
Becker
 
RC
Berger
 
PB
, et al. 
Mean platelet volume as a predictor of cardiovascular risk: a systematic review and meta-analysis.
J Thromb Haemost
2010
, vol. 
8
 
1
(pg. 
148
-
156
)
67
Kamath
 
S
Blann
 
AD
Lip
 
GY
Platelet activation: assessment and quantification.
Eur Heart J
2001
, vol. 
22
 
17
(pg. 
1561
-
1571
)
68
Soranzo
 
N
Rendon
 
A
Gieger
 
C
, et al. 
A novel variant on chromosome 7q22.3 associated with mean platelet volume, counts, and function.
Blood
2009
, vol. 
113
 
16
(pg. 
3831
-
3837
)
69
Soranzo
 
N
Spector
 
TD
Mangino
 
M
, et al. 
A genome-wide meta-analysis identifies 22 loci associated with eight hematological parameters in the HaemGen consortium.
Nat Genet
2009
, vol. 
41
 
11
(pg. 
1182
-
1190
)
70
Meisinger
 
C
Prokisch
 
H
Gieger
 
C
, et al. 
A genome-wide association study identifies three loci associated with mean platelet volume.
Am J Hum Genet
2009
, vol. 
84
 
1
(pg. 
66
-
71
)
71
Hawkins
 
PT
Stephens
 
LR
PI3Kgamma is a key regulator of inflammatory responses and cardiovascular homeostasis.
Science
2007
, vol. 
318
 
5847
(pg. 
64
-
66
)
72
Pasquet
 
JM
Gross
 
BS
Gratacap
 
MP
, et al. 
Thrombopoietin potentiates collagen receptor signaling in platelets through a phosphatidylinositol 3-kinase-dependent pathway.
Blood
2000
, vol. 
95
 
11
(pg. 
3429
-
3434
)
73
Hirsch
 
E
Bosco
 
O
Tropel
 
P
, et al. 
Resistance to thromboembolism in PI3Kgamma-deficient mice.
FASEB J
2001
, vol. 
15
 
11
(pg. 
2019
-
2021
)
74
Tertyshnikova
 
S
Fein
 
A
Inhibition of inositol 1,4,5-trisphosphate-induced Ca2+ release by cAMP-dependent protein kinase in a living cell.
Proc Natl Acad Sci U S A
1998
, vol. 
95
 
4
(pg. 
1613
-
1617
)
75
Wojcikiewicz
 
RJ
Luo
 
SG
Phosphorylation of inositol 1,4,5-trisphosphate receptors by cAMP-dependent protein kinase: type I, II, and III receptors are differentially susceptible to phosphorylation and are phosphorylated in intact cells.
J Biol Chem
1998
, vol. 
273
 
10
(pg. 
5670
-
5677
)
76
Neer
 
EJ
Schmidt
 
CJ
Nambudripad
 
R
Smith
 
TF
The ancient regulatory-protein family of WD-repeat proteins.
Nature
1994
, vol. 
371
 
6495
(pg. 
297
-
300
)
77
Arthur
 
WT
Ellerbroek
 
SM
Der
 
CJ
Burridge
 
K
Wennerberg
 
K
XPLN, a guanine nucleotide exchange factor for RhoA and RhoB, but not RhoC.
J Biol Chem
2002
, vol. 
277
 
45
(pg. 
42964
-
42972
)
78
Thiesen
 
S
Kubart
 
S
Ropers
 
HH
Nothwang
 
HG
Isolation of two novel human RhoGEFs, ARHGEF3 and ARHGEF4, in 3p13-21 and 2q22.
Biochem Biophys Res Commun
2000
, vol. 
273
 
1
(pg. 
364
-
369
)
79
Draviam
 
VM
Stegmeier
 
F
Nalepa
 
G
, et al. 
A functional genomic screen identifies a role for TAO1 kinase in spindle-checkpoint signalling.
Nat Cell Biol
2007
, vol. 
9
 
5
(pg. 
556
-
564
)
80
Weedon
 
MN
Lango
 
H
Lindgren
 
CM
, et al. 
Genome-wide association analysis identifies 20 loci that influence adult height.
Nat Genet
2008
, vol. 
40
 
5
(pg. 
575
-
583
)
81
Reems
 
JA
Wang
 
W
Tsubata
 
K
, et al. 
Dynamin 3 participates in the growth and development of megakaryocytes.
Exp Hematol
2008
, vol. 
36
 
12
(pg. 
1714
-
1727
)
82
Ma
 
D
Sun
 
Y
Lin
 
D
, et al. 
CD226 is expressed on the megakaryocytic lineage from hematopoietic stem cells/progenitor cells and involved in its polyploidization.
Eur J Haematol
2005
, vol. 
74
 
3
(pg. 
228
-
240
)
83
Kitajima
 
K
Kojima
 
M
Kondo
 
S
Takeuchi
 
T
A role of jumonji gene in proliferation but not differentiation of megakaryocyte lineage cells.
Exp Hematol
2001
, vol. 
29
 
4
(pg. 
507
-
514
)
84
Sun
 
L
Gorospe
 
JR
Hoffman
 
EP
Rao
 
AK
Decreased platelet expression of myosin regulatory light chain polypeptide (MYL9) and other genes with platelet dysfunction and CBFA2/RUNX1 mutation: insights from platelet expression profiling.
J Thromb Haemost
2007
, vol. 
5
 
1
(pg. 
146
-
154
)
85
Barclay
 
AN
Signal regulatory protein alpha (SIRPalpha)/CD47 interaction and function.
Curr Opin Immunol
2009
, vol. 
21
 
1
(pg. 
47
-
52
)
86
Naslavsky
 
N
McKenzie
 
J
Altan-Bonnet
 
N
Sheff
 
D
Caplan
 
S
EHD3 regulates early-endosome-to-Golgi transport and preserves Golgi morphology.
J Cell Sci
2009
, vol. 
122
 
3
(pg. 
389
-
400
)
87
Mason
 
KD
Carpinelli
 
MR
Fletcher
 
JI
, et al. 
Programmed anuclear cell death delimits platelet life span.
Cell
2007
, vol. 
128
 
6
(pg. 
1173
-
1186
)
88
Tartaglia
 
M
Mehler
 
EL
Goldberg
 
R
, et al. 
Mutations in PTPN11, encoding the protein tyrosine phosphatase SHP-2, cause Noonan syndrome.
Nat Genet
2001
, vol. 
29
 
4
(pg. 
465
-
468
)
89
Merched
 
AJ
Chan
 
L
Absence of p21Waf1/Cip1/Sdi1 modulates macrophage differentiation and inflammatory response and protects against atherosclerosis.
Circulation
2004
, vol. 
110
 
25
(pg. 
3830
-
3841
)
90
Wickrema
 
A
Crispino
 
JD
Erythroid and megakaryocytic transformation.
Oncogene
2007
, vol. 
26
 
47
(pg. 
6803
-
6815
)
91
Wellcome Trust Case Control Consortium
Genome-wide association study of 14,000 cases of seven common diseases and 3000 shared controls.
Nature
2007
, vol. 
447
 
7145
(pg. 
661
-
678
)
92
Samani
 
NJ
Erdmann
 
J
Hall
 
AS
, et al. 
Genomewide association analysis of coronary artery disease.
N Engl J Med
2007
, vol. 
357
 
5
(pg. 
443
-
453
)
93
Coronary Artery Disease Consortium
Samani
 
NJ
Deloukas
 
P
, et al. 
Large scale association analysis of novel genetic loci for coronary artery disease.
Arterioscler Thromb Vasc Biol
2009
, vol. 
29
 
5
(pg. 
774
-
780
)
94
Myocardial Infarction Genetics Consortium
Kathiresan
 
S
Voight
 
BF
, et al. 
Genome-wide association of early-onset myocardial infarction with single nucleotide polymorphisms and copy number variants.
Nat Genet
2009
, vol. 
41
 
3
(pg. 
334
-
341
)
95
Johnson
 
AD
O'Donnell
 
CJ
An open access database of genome-wide association results.
BMC Med Genet
2009
, vol. 
10
 pg. 
6
 
Sign in via your Institution