P2Y12, the Gi-coupled platelet receptor for adenosine diphosphate (ADP), plays a central role in platelet function. Patients with congenital P2Y12 defects display a mild to moderate bleeding diathesis, characterized by mucocutaneous bleedings and excessive post-surgical and post-traumatic blood loss. Defects of P2Y12 should be suspected when ADP, even at high concentrations (≥ 10μM), is unable to induce full, irreversible platelet aggregation. Tests that evaluate the degree of inhibition of adenylyl cyclase by ADP should be used to confirm the diagnosis. Drugs that inhibit P2Y12 are potent antithrombotic drugs, attesting the central role played by P2Y12 in platelet thrombus formation. Clopidogrel, the most widely used drug that inhibits P2Y12, is effective both in monotherapy and in combination with acetylsalicylic acid. The most important drawback of clopidogrel is its inability to inhibit adequately P2Y12-dependent platelet function in approximately one-third of patients who are therefore not protected from major cardiovascular events. New drugs, such as prasugrel and ticagrelor, which effectively inhibit P2Y12 in the majority of patients, proved to be more efficacious than clopdidogrel in preventing major cardiovascular events. Although they increase the incidence of major bleedings, the net clinical benefit is in favor of the new P2Y12 inhibitors.

Adenosine diphosphate (ADP) plays a key role in platelet function. When it is secreted from the platelet-dense granules where it is stored, it amplifies the platelet responses induced by other platelet agonists1,2  and stabilizes platelet aggregates.3-5  The ADP-induced signal is mediated by P2Y receptors, which are G-coupled 7-membrane-spanning proteins that are present in almost any kind of cell, whose ligands are purine and pyrimidine nucleotides. From a phylogenetic and structural point of view, 2 distinct P2Y receptor subgroups with a relatively high level of structural divergence have been identified: the first subgroup includes the Gq-coupled subtypes (P2Y1, P2Y2, P2Y4, P2Y6, and P2Y11), and the second subgroup includes the Gi-coupled subtypes (P2Y12, P2Y13, and P2Y14).6  Human platelets express 2 distinct receptors for ADP: P2Y1 and P2Y12. The Gq-coupled P2Y1 receptor mediates a transient rise in cytoplasmic Ca2+, platelet shape change, and rapidly reversible aggregation, whereas the Gi-coupled P2Y12 receptor mediates inhibition of adenylyl cyclase and amplifies the platelet aggregation response.2  Concomitant activation of both the Gq and Gi pathways by ADP is necessary to elicit normal aggregation.2,7 

P2Y12, which maps to chromosome 3q21-q25, is present in platelets, endothelial cells, glial cells, and smooth muscle cells.8  It contains 342 amino acid residues, including 4 extracellular Cys residues at positions 17, 97, 175, and 270: Cys 97 and Cys 175, which are linked by a disulphide bridge and are important for receptor expression9-11 ; 2 potential N-linked glycosylation sites at its extracellular amino-terminus may modulate its activity (Figure 1).12 

Figure 1

Predicted secondary structure of the human P2Y12 receptor. Black circles represent the sites of amino acid substitution in patients with dysfunctional P2Y12 (Table 1). TM indicates transmembrane region; EL, extracellular loop; and IL, intracellular loop.

Figure 1

Predicted secondary structure of the human P2Y12 receptor. Black circles represent the sites of amino acid substitution in patients with dysfunctional P2Y12 (Table 1). TM indicates transmembrane region; EL, extracellular loop; and IL, intracellular loop.

Close modal

P2Y12 receptors exist predominantly as homo-oligomers situated in lipid rafts. On treatment with the active metabolite of clopidogrel (which inhibits P2Y12 function), the homo-oligomers are disrupted into nonfunctional dimers and monomers that are sequestered outside the lipid rafts.9 

P2Y12 signaling

ADP and some of its analogs, such as 2-methylthio-ADP and (N)-methanocarba-2-methylthio-ADP, stimulate P2Y12, whereas adenosine triphosphate and its triphosphate analogs act as antagonists.13  The P2Y12 receptor is coupled to inhibition of adenylyl cyclase activity mostly through activation of Gαi2 and has a critical requirement for lipid rafts.14  It must be noted, however that, although inhibition of adenylyl cyclase via Gαi2 is a key feature of platelet activation by ADP, it bears no causal relationship to platelet aggregation.15,16  Therefore, other signaling events downstream of Gαi2 are required for activation of integrin αIIbβ3 and subsequent platelet aggregation. Several studies demonstrated a crucial role for different isoforms of phosphoinositide 3-kinase (PI3K) in ADP-dependent P2Y12 receptor-mediated amplification of platelet activation.5,17-21  In addition, studies of platelets from P2Y1 knockout mice and of normal platelets in the presence of specific P2Y1 antagonists showed that ADP, at higher concentrations than those commonly used to activate platelets, induces slow and sustained PI3K-dependent platelet aggregation, which is not preceded by platelet shape change.22,23 

Role of P2Y12 in platelet function

Although ADP by itself is unable to cause the secretion of platelet-dense granules, its interaction with P2Y12 greatly amplifies platelet secretion induced by agonists, such as thromboxane A224-26  and thrombin receptor-activating peptide (Figure 2).27  This effect, which is probably mediated by PI3K,18,28  was observed both at physiologic and micromolar concentrations of extracellular Ca2+, in the presence of acetylsalicylic acid (ASA), and independently of the formation of large platelet aggregates, demonstrating that it is a direct effect of P2Y12, rather than secondary to P2Y12-mediated amplification of aggregation.24-26 

Figure 2

Central role of P2Y12 in platelet function. Green arrow represents activation; truncated red line, inhibition; blue line ending with ⊕, amplification; and dotted black line, secretion.

Figure 2

Central role of P2Y12 in platelet function. Green arrow represents activation; truncated red line, inhibition; blue line ending with ⊕, amplification; and dotted black line, secretion.

Close modal

P2Y12 plays an essential role in the stabilization of platelet aggregates induced by thrombin3-5  or thromboxane A2 (Figure 2),29  which is mediated by PI3K.14 

Studies of human platelets congenitally lacking P2Y12 and of P2Y12 knockout mice demonstrated that platelet aggregation and secretion induced by a range of platelet agonists were impaired.26,30-34  P2Y12 receptor stimulation by released ADP contributes to inhibition of adenylyl cyclase, activation of serine-threonine kinase Akt in platelets, tyrosine phosphorylation, extracellular signal-regulated kinase 2 activation, Rap1B activation, Rac activation, and Ca2+ mobilization induced by other agonists.8 

Early studies demonstrated that P2Y12 is an important mediator of shear-induced platelet aggregation using platelets from persons treated with the antithrombotic drug ticlopidine35  or from a patient with congenital P2Y12 deficiency.36  This effect of P2Y12, which was later confirmed using specific, direct P2Y12 antagonists,8  is dependent on PI3K activation.37 

As already mentioned, inhibition of AC via Gαi2 by ADP bears no causal relationship to platelet activation. However, it may substantially contribute to platelet thrombus formation in vivo by counteracting the antiplatelet effect of prostacyclin (Figure 2) or other substances that stimulate adenylyl cyclase.38 

P2Y12 shares with P2Y1 the ability to contribute to collagen-induced platelet microparticle formation in whole blood, and to contribute to the formation of platelet-leukocyte aggregates mediated by platelet surface P-selectin exposure, which results in tissue factor exposure at the surface of leukocytes.8  However, only the P2Y12 receptor was found to be involved in the exposure of phosphatidylserine by thrombin or other platelet agonists and in tissue factor-induced thrombin formation in platelet-rich plasma.8 

Congenital deficiency of P2Y12

Congenital P2Y12 deficiency is an autosomal recessive disorder. The first patient with severe P2Y12 deficiency (patient 1) was described in 1992.30  He had a lifelong history of excessive bleeding, prolonged bleeding time (15-20 minutes), reversible aggregation in response to weak agonists, and impaired aggregation in response to low concentrations of collagen or thrombin. However, the most typical feature was that ADP, even at very high concentrations (> 10μM), did not induce full and irreversible platelet aggregation. Other abnormalities of platelet function were: (1) no inhibition by ADP of prostaglandin E1 (PGE1)-stimulated platelet adenylyl cyclase, but normal inhibition by epinephrine; (2) normal shape change and borderline-normal mobilization of cytoplasmic Ca2+ induced by ADP; and (3) the presence of approximately 30% of the normal number of binding sites for [33P]2MeSADP on fresh platelets39  or [3H]ADP on formalin-fixed platelets (which are associated with the ADP receptor P2Y1).30  Five additional patients with severe P2Y12 deficiency, belonging to 4 kindreds, were subsequently described: one French man (patient 2),34  2 Italian sisters (patients 3 and 4),26  a Japanese woman (patient 5),40  and a British woman of Asian descent (patient 6; Table 1).41 

Table 1

Characteristics of the described patients with congenital P2Y12 defects

Patient no.ReferencesP2Y12 mutationsBinding sites for 2MeS-ADPPlatelet aggregation induced by ADP ≥ 10μM
30,43  p.[Gln98fs]+[Gln98fs] Severely reduced Reduced and reversible 
34,45  p.[Phe240fs]+[?]* Severely reduced Reduced and reversible 
3, 4 26,44  p.[0]+p.Thr126fs Severely reduced Reduced and reversible 
40  p.[0]+[0] Not tested Reduced and reversible 
41  p.[Gly12fs]+[Gly12fs] Not tested Reduced and reversible 
34,45  p.[Phe240fs]+[=] Intermediate Full and irreversible 
26,44  p.[0]+[=] Intermediate Full and irreversible 
32  p.[Arg256Gln]+[Arg265Trp] Normal Reduced and reversible 
10, 11 32  p.[Arg265Trp]+[=] Normal Full and irreversible 
12 48  p.[Pro258Thr]+[=] Not tested Reduced and reversible 
13 49  p.[Lys174Glu]+[=] Intermediate Reduced and reversible 
Patient no.ReferencesP2Y12 mutationsBinding sites for 2MeS-ADPPlatelet aggregation induced by ADP ≥ 10μM
30,43  p.[Gln98fs]+[Gln98fs] Severely reduced Reduced and reversible 
34,45  p.[Phe240fs]+[?]* Severely reduced Reduced and reversible 
3, 4 26,44  p.[0]+p.Thr126fs Severely reduced Reduced and reversible 
40  p.[0]+[0] Not tested Reduced and reversible 
41  p.[Gly12fs]+[Gly12fs] Not tested Reduced and reversible 
34,45  p.[Phe240fs]+[=] Intermediate Full and irreversible 
26,44  p.[0]+[=] Intermediate Full and irreversible 
32  p.[Arg256Gln]+[Arg265Trp] Normal Reduced and reversible 
10, 11 32  p.[Arg265Trp]+[=] Normal Full and irreversible 
12 48  p.[Pro258Thr]+[=] Not tested Reduced and reversible 
13 49  p.[Lys174Glu]+[=] Intermediate Reduced and reversible 

Patient 7 is the daughter of patient 2; patient 8 is the son of patient 4; patient 10 is the son of patient 9, and patient 11 is the daughter of patient 9.

*

No mutations were found in one allele of patient 2; however, the findings that the patient's platelets contained P2Y12 transcripts derived from the mutant allele only and that his daughter (patient 7) inherited the mutant allele from her father and a normal allele from her mother, suggest that patient 2 has an additional, as yet unknown, mutation that silences his normal allele.

p[0] was associated with partial or complete P2Y12 gene deletion in patients 3, 4, and 8.

Failure of expression of the P2Y12 protein (p.[0]) in patient 5 was associated with homozygous single nucleotide substitution in the transduction initiation codon (ATG to AGG).

The study of the son of patient 4 allowed the characterization of heterozygous P2Y12 deficiency26 : ADP-induced platelet aggregation was reversible for ADP concentrations less than or equal to 10μM but was full and irreversible for concentrations of ADP more than or equal to 10μM; the inhibition of PGE1-induced increase in platelet cyclic AMP was impaired, albeit not completely absent; the number of platelet binding sites for [33P]2MeS-ADP was intermediate between his mother's and normal subjects'; finally, the platelet secretion was impaired. Because the secretion defect in this patient's platelets was not associated with impaired production of thromboxane A2 or low concentrations of platelet granule contents, it is very similar to that described in patients with an ill-defined and probably heterogeneous group of congenital defect of platelet secretion, sometimes referred to with the general term “primary secretion defect.”25,42 

Molecular defects

The P2Y12 gene of patients 1 and 6 displayed homozygous base pair deletions in the open reading frame, resulting in frameshifts and premature truncation of the protein.41,43  The P2Y12 gene of sisters (patients 3 and 4) displayed an identical single base pair deletion (378delC), resulting in a frameshift and premature truncation of the protein.43  As only alleles encoding the mutated DNA sequence were found by polymerase chain reaction analysis, the patients were considered homozygous for the 378delC mutation. However, a subsequent study revealed that they have the P2Y12 deficiency because of haploinsufficiency and to the 378delC mutation in their remaining allele.44  Patient 5 is homozygous for a single nucleotide substitution in the transduction initiation codon (ATG to AGG).40  The molecular defect that is responsible for P2Y12 deficiency in patient 234  is less well defined45 : one mutant allele contains a deletion of 2 base pairs, resulting in a frameshift and early truncation of the protein. Surprisingly, the other allele did not display any mutation: the findings that the patient's platelets contained P2Y12 transcripts derived from the mutant allele only and that his daughter, who had a heterozygous phenotype, inherited the mutant allele from her father and a normal allele from her mother, suggest that patient 2 has an additional, as yet unknown, mutation that silences his normal allele.

Congenital dysfunction of P2Y12

One patient (patient 9) with congenital bleeding disorder associated with abnormal P2Y12-mediated platelet responses to ADP, whose platelets display normal number of dysfunctional P2Y12 receptors has been described.32  Platelets from this patient displayed reduced and reversible aggregation in response to 4μM ADP, similar to normal platelets with a blocked P2Y12 receptor. However, the response to 20μM ADP, albeit still decreased and reversible, was more pronounced and was further inhibited by a P2Y12 antagonist, indicating residual receptor function. ADP failed to lower adenylyl cyclase activity stimulated by PGE1 in the patient's platelets. Analysis of the patient's P2Y12 gene revealed, in one allele, a G to A transition changing the codon for Arg256 in TM6 to Gln and, in the other, a C to T transition changing the codon for Arg265 in EL3 to Trp (Table 1). Neither mutation interfered with receptor surface expression, but both altered receptor function because ADP inhibited the forskolin-induced increase of cyclic AMP markedly less in cells transfected with either mutant P2Y12 than in wild-type cells. These observations, in accordance with previous studies of the P2Y1 receptor,46,47  helped to identify regions in TM6 and EL3, whose structural integrity is necessary for normal receptor function (Figure 1). A heterozygous point mutation in the same region of the molecule, which changed codon 258 coding for proline (CCT) to threonine (ACT; Pro258Thr), was described in a patient with mild bleeding disorder and severely impaired ADP-induced platelet aggregation.48  Because the proline-to-threonine substitution alters the protein hydrophobicity, size, and rotational mobility, it probably affects the function of P2Y12.

Finally, a heterozygous mutation, predicting a lysine to glutamate (Lys174Glu) substitution in P2Y12 was identified in one patient with mild type 1 von Willebrand disease.49  Platelets from this patient showed reduced and reversible aggregation in response to ADP, up to 10μM. The reduced response was associated with an approximate 50% reduction in binding of [3H]2MeS-ADP. Considering that Lys174 is situated in the second extracellular loop of P2Y12, adjacent to Cys175, which may be important for the expression of the ADP binding site receptor, and that a hemagglutinin-tagged Lys174Glu P2Y12 variant showed surface expression in Chinese hamster ovary cells, it is probable that the Lys174Glu mutation is responsible for disruption of the ADP binding site of the receptor.

It is interesting to note that, for reasons that are presently unclear, 2 patients with heterozygous dysfunctional P2Y12 (Pro258Thr and Lys174Glu) display a much more severe impairment of ADP-induced platelet aggregation compared with the 2 patients who are heterozygous for P2Y12 deficiency26,34  and to the 2 children of patient 9, who are heterozygous for the Arg265Gln mutation32  (Table 1).

Bleeding diathesis

Patients with defects of P2Y12 experience mucocutaneous bleedings and excessive postsurgical or posttraumatic blood loss. The severity of their bleeding diathesis is variable. The bleeding scores of patient 1 and of the 2 sisters (patients 3 and 4), which was calculated using a standardized questionnaire that was developed to investigate patients with type 1 von Willebrand disease,50  were 8, 7, and 13, respectively, (normal values ≤ 3; unpublished data). After extensive investigation of hemostasis parameters, which included measurement of the activity of clotting and fibrinolytic factors and the search for known polymorphisms of hemostasis proteins, we found no explanation for the discrepancy in the severity of bleeding manifestations in the 2 sisters (patients 3 and 4).

The bleeding score of patient 8, the son of patient 4, was normal; however, it must be noted that this young boy had not yet experienced situations that could challenge the hemostatic system at the time of our investigation. His bleeding time, despite the mild defect of P2Y12, was prolonged (13 minutes).

Diagnosis and treatment

The diagnosis of P2Y12 defects is rather simple: they should be suspected when ADP, even at relatively high concentrations (≥ 10μM), is unable to induce full, irreversible platelet aggregation while inducing normal shape change. Tests that evaluate the degree of inhibition of adenylyl cyclase by ADP, by measuring either the platelet levels of cyclic AMP or the phosphorylation of vasodilator-stimulated phosphoprotein51  after the exposure of platelets to PGE1, should be used to confirm the diagnosis (Table 2).

Table 2

Diagnosis of severe congenital P2Y12 defects

Clinical hallmarks 
    Lifelong history of mucocutaneous bleedings; excessive postsurgical or posttraumatic blood loss 
Laboratory hallmarks 
    Screening test: inability of high concentrations of ADP (≥ 10μM) to induce full and reversible platelet aggregation (light transmission aggregometry) 
    Confirmatory test: inability of ADP to inhibit PGE1-stimulated adenylyl cyclase (measurement of platelet cyclic AMP or phosphorylation of vasodilator-stimulated phosphoprotein) 
Clinical hallmarks 
    Lifelong history of mucocutaneous bleedings; excessive postsurgical or posttraumatic blood loss 
Laboratory hallmarks 
    Screening test: inability of high concentrations of ADP (≥ 10μM) to induce full and reversible platelet aggregation (light transmission aggregometry) 
    Confirmatory test: inability of ADP to inhibit PGE1-stimulated adenylyl cyclase (measurement of platelet cyclic AMP or phosphorylation of vasodilator-stimulated phosphoprotein) 

The intravenous infusion of the vasopressin analog desmopressin (0.3 μg/kg) shortened the prolonged bleeding time of patient 1 from 20 minutes to 8.5 minutes.36  After the infusion of desmopressin, which was repeated twice at 24-hour intervals, the patient underwent a surgical intervention for disc hernia repair, which was not complicated by excessive bleeding. Although the efficacy of desmopressin in reducing bleeding complications of patients with defects of primary hemostasis is anecdotal,52  its administration is generally without serious side effects.

Drugs that target P2Y12 reduce the incidence of arterial thrombosis, as documented by the results of several randomized clinical trials (Table 3). Because of space limitations, this review will not analyze the results of these randomized clinical trials in detail: the interested reader is referred to recently published reviews.53,54 

Table 3

Main results of major double-blind, randomized, controlled clinical trials with ticlopidine or clopidogrel

Clinical trialReferencePatientsTreatmentsCardiovascular endpointsFollow-upEfficacy*Bleeding events
CATS 55  Recent thromboembolic stroke 1. Ticlopidine Stroke, MI, or vascular death 24 mo  6.5% 
   2. Placebo   23.3% (1.0-40.5) 3.0% 9.0% 
TASS 56  Recent transient or mild persistent focal cerebral or retinal ischemia 1. Ticlopidine Nonfatal stroke or death from any cause 3 y 12% (−2%-26%) 10.0% 
   2. ASA     
ISAR 57  CAD patients: PCI and stent implantation 1. Ticlopidine + ASA Cardiac death or AMI, CABG, or repeated PCI 4 wk 0.25 0.00 
   2. Heparin/VKA + ASA   (0.06-0.77) 0.00-0.019 
SAR 58  CAD patients: PCI and stent implantation 1. Ticlopidine + ASA AMI, death, repeat PCI, stent thrombosis at angiography 30 d 1 vs 2: 0.20 (0.07-0.61) 1 vs 2: 0.88 (0.55-1.43) 
   2. Heparin/VKA + ASA   1 vs 3: 0.15 (0.05-0.43) 1 vs 3: 3.06 (1.57-5.97) 
   3. ASA     
CAPRIE 59  Atherosclerotic vascular disease 1. Clopidogrel Ischemic stroke, AMI, or vascular death 1.9 y  9.27% vs 9.28% 
   2. ASA   8.7% (0.3-16.5)  
CHARISMA 60  Clinically evident CV disease or multiple risk factors 1. Clopidogrel + ASA MI, stroke, or 28 mo 0.93 F: 1.53 (0.83-2.82) 
   2. Placebo + ASA CV death  (0.83-1.05) M: 1.25 (0.97-1.61) 
       Mod: 1.62 (1.27-2.08) 
MATCH 61  Recent ischemic stroke or TIA and ≥ 1 risk factors 1. Clopidogrel + ASA Ischemic stroke, MI, vascular death, or rehospitalization for acute ischemia 18 mo 6.4% (−4.6-16·3) LT: 1.26 (0.62-1.88) 
   2. Clopidogrel + placebo    M: 1.36 (0.86-1.86) 
CURE 62  Acute coronary syndromes 1. Clopidogrel + ASA CV death, 12 mo 0.80 (0.72-0.90) LT: 1.21 (0.95-1.56) 
   2. Placebo + ASA Nonfatal AMI, or stroke   M: 1.38 (1.13-1.67) 
       T: 1.69 (1.48-1.94) 
COMMIT 63  Suspected AMI 1. Clopidogrel + ASA (1) Death, reinfarction, or stroke Up to 28 d (1) 0.91 (0.86-0.97) F: −0.1 (SE: 0.5) 
   2. Placebo + ASA (2) Death from any cause  (2) 0.93 (0.87-0·99) M: 0.4 (0.7) 
       m: 4.7 (1.7) 
PCI CURE 64  NSTE ACS undergoing PCI in the CURE study 1. Clopidogrel + ASA CV death, AMI, or urgent target vessel revascularization 30 d 0.70 LT: 0.92 (0.38-2.26) 
   2. Placebo + ASA   (0.50-0·97) M: 1.13 (0.61-2.10) 
       m: 1.33 (0.59-3.03) 
Clinical trialReferencePatientsTreatmentsCardiovascular endpointsFollow-upEfficacy*Bleeding events
CATS 55  Recent thromboembolic stroke 1. Ticlopidine Stroke, MI, or vascular death 24 mo  6.5% 
   2. Placebo   23.3% (1.0-40.5) 3.0% 9.0% 
TASS 56  Recent transient or mild persistent focal cerebral or retinal ischemia 1. Ticlopidine Nonfatal stroke or death from any cause 3 y 12% (−2%-26%) 10.0% 
   2. ASA     
ISAR 57  CAD patients: PCI and stent implantation 1. Ticlopidine + ASA Cardiac death or AMI, CABG, or repeated PCI 4 wk 0.25 0.00 
   2. Heparin/VKA + ASA   (0.06-0.77) 0.00-0.019 
SAR 58  CAD patients: PCI and stent implantation 1. Ticlopidine + ASA AMI, death, repeat PCI, stent thrombosis at angiography 30 d 1 vs 2: 0.20 (0.07-0.61) 1 vs 2: 0.88 (0.55-1.43) 
   2. Heparin/VKA + ASA   1 vs 3: 0.15 (0.05-0.43) 1 vs 3: 3.06 (1.57-5.97) 
   3. ASA     
CAPRIE 59  Atherosclerotic vascular disease 1. Clopidogrel Ischemic stroke, AMI, or vascular death 1.9 y  9.27% vs 9.28% 
   2. ASA   8.7% (0.3-16.5)  
CHARISMA 60  Clinically evident CV disease or multiple risk factors 1. Clopidogrel + ASA MI, stroke, or 28 mo 0.93 F: 1.53 (0.83-2.82) 
   2. Placebo + ASA CV death  (0.83-1.05) M: 1.25 (0.97-1.61) 
       Mod: 1.62 (1.27-2.08) 
MATCH 61  Recent ischemic stroke or TIA and ≥ 1 risk factors 1. Clopidogrel + ASA Ischemic stroke, MI, vascular death, or rehospitalization for acute ischemia 18 mo 6.4% (−4.6-16·3) LT: 1.26 (0.62-1.88) 
   2. Clopidogrel + placebo    M: 1.36 (0.86-1.86) 
CURE 62  Acute coronary syndromes 1. Clopidogrel + ASA CV death, 12 mo 0.80 (0.72-0.90) LT: 1.21 (0.95-1.56) 
   2. Placebo + ASA Nonfatal AMI, or stroke   M: 1.38 (1.13-1.67) 
       T: 1.69 (1.48-1.94) 
COMMIT 63  Suspected AMI 1. Clopidogrel + ASA (1) Death, reinfarction, or stroke Up to 28 d (1) 0.91 (0.86-0.97) F: −0.1 (SE: 0.5) 
   2. Placebo + ASA (2) Death from any cause  (2) 0.93 (0.87-0·99) M: 0.4 (0.7) 
       m: 4.7 (1.7) 
PCI CURE 64  NSTE ACS undergoing PCI in the CURE study 1. Clopidogrel + ASA CV death, AMI, or urgent target vessel revascularization 30 d 0.70 LT: 0.92 (0.38-2.26) 
   2. Placebo + ASA   (0.50-0·97) M: 1.13 (0.61-2.10) 
       m: 1.33 (0.59-3.03) 

Doses of the antiplatelet agents are as follows: ticlopidine, 250 mg twice a day55-58 ; ASA, 650 mg twice a day55 ; 325 mg daily58,59 ; 100 mg twice a day57 ; 160 mg daily63 ; 75 to 325 mg daily62,64 ; 75 to 160 mg daily60 ; 75 mg daily61 ; clopidogrel, 75 mg daily in all randomized clinical trials, plus 300 mg loading dose in the CURE trial.62,64 

MI indicates myocardial infarction; CAD coronary artery disease; AMI, acute myocardial infarction; VKA, vitamin K antagonist; CV, cardiovascular; and TIA, transient ischemic attack.

*

Results are reported as relative risk reduction55,56,59,60  and relative risk57,58,60,62-64  (95% confidence interval).

When available, data on major (M), life-threatening (LT), fatal (F), moderate (Mod), minor (m), and total (T) bleeding events are reported as total incidence,55,56,59  relative risk,57,58,60-62,64  and excess per 1000 patients63  (95% confidence interval).

Thienopyridines

First- and second-generation thienopyridines: ticlopidine and clopidogrel.

Ticlopidine and clopidogrel (Figure 3) are prodrugs that need to be converted in vivo by the hepatic cytochrome P-450 (CYP) enzymatic pathway to active metabolites, which covalently bind to P2Y12 by forming a disulfide bond with cysteine residues, thereby irreversibly inhibiting the receptor.53 

Figure 3

Metabolic pathways for the transformation of thienopyridines to their active metabolites.

Figure 3

Metabolic pathways for the transformation of thienopyridines to their active metabolites.

Close modal

Several randomized clinical trials documented the clinical efficacy of these drugs in the prevention of major adverse cardiovascular events (MACEs;55-64 Table 3). Their efficacy was essentially similar to that of ASA in patients with stable disease of cerebral or coronary arteries. For patients with acute coronary syndromes, undergoing medical treatment only or in combination with percutaneous coronary intervention (PCI), the addition of thienopiridines to ASA proved highly efficacious. In contrast, the combination of clopidogrel and ASA was not more effective than monotherapy in low- to moderate-risk patients with stable disease but increased the incidence of bleeding.

Because of its toxicity,53  ticlopidine has been almost completely replaced by clopidogrel in clinical practice.

Despite its proven antithrombotic efficacy, clopidogrel has some important drawbacks54 : (1) its antiplatelet effects are delayed because of the need for metabolism of the prodrug; (2) there is substantial interindividual variability in platelet inhibition; (3) its ability to irreversibly inhibit P2Y12 may represent a problem for patients who need to undergo coronary bypass (CABG) surgery because the incidence of postoperative bleeding complications is higher than in patients not treated with clopidogrel. Although the onset of action of clopidogrel can be accelerated by giving patients a loading dose of 300 to 600 mg, the solution of the other 2 problems appears more difficult.54 

The high interindividual variability of the response to clopidogrel is a clinically relevant issue, as it has been demonstrated that poor responders are not adequately protected from MACE.65  Approximately one-third of treated patients do not display adequate inhibition of P2Y12-dependent platelet function; this condition is associated with loss-of-function mutations of CYP,66-68  with the homozygous 3435C → T mutation of ABCB1, a gene encoding for the efflux pump P-glycoprotein, a key protein involved in thienopyridine absorption,69,70  and may be exacerbated by negative interference with common adjunctive medications, such as proton pump inhibitors.68 

Tailored treatment of patients, based on the results of platelet function tests or of CYP genotyping, has been proposed to solve the problem of clopidogrel resistance.54  This approach cannot be recommended in daily clinical practice yet because the best laboratory method to monitor the effects of clopidogrel on platelet function still needs to be identified, standardized (for preanalytical and analytical variables), and validated in the clinical setting. Several recent studies demonstrated that the agreement among different laboratory tests to identify poor responders is rather low and that assessment of platelet response to clopidogrel is highly test-specific.71-76  Recent studies showed that the search for loss of function mutations of CYP is not very accurate in predicting the response to clopidogrel.77,78  In addition, preliminary experiments that evaluated the effects of increasing the dose of clopidogrel in resistant patients gave results that are incompletely satisfactory because many patients remained “resistant” to clopidogrel, even after repeated administrations of high doses of the drug.79,80 

Mostly based on the aforementioned consideration, a recent consensus paper concluded that, until the results of large-scale trials of personalized antiplatelet therapy are available, the routine use of platelet function measurements in the care of patients with cardiovascular disease cannot be recommended.81  Therefore, the use of new P2Y12 antagonists that are able to induce predictable and adequate inhibition of platelet function in all patients is desirable.

Prasugrel, a new thienopyridine.

Prasugrel is a new thienopyridine, with much more rapid and consistent inhibitory effects on platelet aggregation than clopidogrel. It has a distinct chemical structure, which permits conversion to its active metabolite with less dependence on CYP enzymes than clopidogrel (Figure 3).54  Consequences of the different metabolism of prasugrel, compared with that of clopidogrel, are54 : (1) faster appearance and higher concentration of its active metabolite in circulating blood; (2) faster and greater mean inhibition of P2Y12-dependent platelet function; (3) no influence of the CYP genotype on its pharmacokinetics, pharmacodynamics, and antithrombotic activity; and (4) much lower interindividual variability in inhibition of P2Y12-dependent platelet responses and very low prevalence of subjects who display “resistance” to the drug.

The aforementioned more favorable characteristics of prasugrel compared with clopidogrel result in greater clinical benefit, as shown by the results of TRITON TIMI-38, which evaluated 13 608 high-risk patients with acute coronary syndromes who required PCI.82  Patients were randomized to receive prasugrel 60-mg loading dose followed by 10 mg/day or clopidogrel 300 mg followed by 75 mg/day for 6 to 15 months. Prasugrel was associated with fewer ischemic events but more non–CABG-related and CABG-related major and fatal bleedings (Table 4).

Table 4

Main results of two double-blind, randomized clinical trials that compared prasugrel and ticagrelor to clopidogrel for the treatment of patients with ACS

Clinical trialReferencePatientsTreatmentsPrimary endpointsFollow-upEfficacy HR (95% CI)TIMI major bleedings* HR (95% CI)
TRITON TIMI-38 82  ACS with scheduled PCI 1. Prasugrel + ASA CV death, nonfatal stroke, nonfatal AMI 6-15 mo 0.81 1.32 
   2. Clopidogrel + ASA   (0.73-0.90) (1.03-1.68) 
PLATO 91  ACS with or without ST elevation 1. Ticagrelor + ASA CV death, AMI, stroke 12 mo 0.84 1.25 
   2. Clopidogrel + ASA   (0.77-0.92) (1.03-1.53) 
Clinical trialReferencePatientsTreatmentsPrimary endpointsFollow-upEfficacy HR (95% CI)TIMI major bleedings* HR (95% CI)
TRITON TIMI-38 82  ACS with scheduled PCI 1. Prasugrel + ASA CV death, nonfatal stroke, nonfatal AMI 6-15 mo 0.81 1.32 
   2. Clopidogrel + ASA   (0.73-0.90) (1.03-1.68) 
PLATO 91  ACS with or without ST elevation 1. Ticagrelor + ASA CV death, AMI, stroke 12 mo 0.84 1.25 
   2. Clopidogrel + ASA   (0.77-0.92) (1.03-1.53) 

Although the investigators of the PLATO trial elaborated original criteria to classify the severity of bleeding episodes, they also calculated the incidence of bleedings based on TIMI criteria, which are reported in this table for easier comparison with the results of the TRITON TIMI-38 trial. Doses of the antiplatelet agents are as follows: prasugrel, 60 mg loading dose plus 10 mg daily; ticagrelor, 180 mg loading dose plus 90 mg twice a day; clopidogrel, 300 mg (TRITON TIMI-38) and 300 to 600 mg (PLATO) loading dose plus 75 mg daily; aspirin, 100 mg daily (TRITON TIMI-38) and 75 to 325 mg daily (PLATO).

HR indicates hazard ratio; CI, confidence interval; ACS, acute coronary syndrome; CV, cardiovascular; and AMI, acute myocardial infarction.

*

The non–CABG-related bleedings only are reported.

The incidence of death from any cause was significantly decreased by ticagrelor compared with clopidogrel.

Based on the results of the TRITON TIMI-38 trial, prasugrel is generally considered a more potent antiplatelet agent than clopidogrel, to be used only in high-risk patients or for a short period, whereas treatment with clopidogrel should be preferred in the remaining situations. However, it is incorrect to say that prasugrel is more potent than clopidogrel, as both ex vivo and in vitro studies demonstrated that the active metabolites of the 2 compounds have the same potency.83,84  The different clinical efficacy and safety of prasugrel compared with clopidogrel are mostly explained by the fact that very few treated patients are “resistant” to prasugrel. Because protection from thrombotic events and exposure to bleeding risk are a function of the degree of inhibition of P2Y12-dependent platelet function, the higher efficacy and the lower safety of prasugrel compared with clopidogrel are simply explained by the fact that prasugrel protects from MACE and exposes to the risk of bleeding more patients than clopidogrel. Based on the results of published studies, it can be predicted that, if tailored treatment with clopidogrel were successful in all patients displaying hyporesponsiveness to the drug, incidences of MACE and bleedings in patients given tailored clopidogrel treatment would be similar to those observed in patients given prasugrel.54 

Therefore, prasugrel appears an attractive solution to some of the problems that are associated with the use of clopidogrel. However, prasugrel does not solve the problem associated with the slow offset of action because, like clopidogrel, it causes irreversible inhibition of P2Y12. Some direct P2Y12 antagonists with such characteristics are currently under development.

Direct P2Y12 inhibitors

Ticagrelor.

Ticagrelor belongs to the new chemical class cyclopentyl-triazolo-pyrimidines (Figure 4): it does not require conversion to an active metabolite and has a half-life of 7 to 8.5 hours.85,86  After oral administration, it rapidly and reversibly inhibits P2Y12 via a mechanism that is noncompetitive with ADP, suggesting the existence of an independent receptor binding site.87  In phase 2 trials, ticagrelor more rapidly and effectively inhibited platelet aggregation and with less variability than clopidogrel.88,89  A study that compared the onset and offset of action of clopidogrel and ticagrelor showed that, despite the greater mean antiplatelet effect of ticagrelor, inhibition of platelet aggregation at 24 hours after the last dose was equivalent in ticagrelor- and clopidogrel-treated patients, which is indicative of a faster offset of effect.90  Considering that, because of its short half life, ticagrelor needs to be administered every 12 hours, which might negatively affect the compliance of patients under chronic treatment, these data have relevant practical implications because they suggest that patients who miss one dose of ticagrelor will have a level of platelet inhibition at 24 hours after the last dose that is not inferior to that of patients undergoing chronic clopidogrel therapy. Dyspnea was reported in 10% to 20% of patients treated with ticagrelor in phase 2 trials, although none of the incidents was considered to be serious.89,90  The pathogenesis of dyspnea during ticagrelor treatment is unclear, although it has been hypothesized that it may be mediated by adenosine.54 

Figure 4

Chemical structures of the direct P2Y12 inhibitors ticagrelor and cangrelor.

Figure 4

Chemical structures of the direct P2Y12 inhibitors ticagrelor and cangrelor.

Close modal

The results of PLATO trial, in which ticagrelor (180 mg loading dose [LD], 90 mg twice a day maintenance dose [MD]) was compared with clopidogrel (300-600 mg LD, 75 mg daily MD) for prevention of MACE in patients with non-ST or ST elevation acute coronary syndromes (two-thirds of them underwent PCI) showed that ticagrelor decreases the incidence of MACE compared with clopidogrel (Table 4).91  Very importantly, ticagrelor also decreased the incidence of cardiovascular and total mortality. There was a higher incidence of TIMI major non–CABG-related bleedings in patients who received ticagrelor compared with those treated with clopidogrel. The incidence of major CABG-related bleedings was similar in the 2 groups. Therefore, similarly to the TRITON-TIMI 38 trial, the PLATO trial showed that a more consistent, adequate inhibition of P2Y12-dependent platelet function than that achieved with standard doses of clopidogrel is associated with greater antithrombotic efficacy and higher risk of non–CABG-related major bleedings. The higher incidence of clinically irrelevant dyspnea in ticagrelor-treated patients was confirmed in the PLATO study.91 

Cangrelor.

Cangrelor (Figure 4) belongs to a family of analogs of adenosine triphosphate that are relatively resistant to breakdown by ectonucleotidases and display high affinity for the P2Y12 receptor, which is reversibly inhibited by the drug.85  Cangrelor does not require conversion to an active metabolite and is immediately active after intravenous infusion, with a half-life of 3 to 6 minutes.

Two trials, which compared cangrelor with clopidogrel in patients requiring PCI, were prematurely terminated because of insufficient evidence of superiority of cangrelor.92,93  Cangrelor is still being studied as a bridge for patients who need to suspend thienopyridines before surgery.

As already mentioned, patients with congenital P2Y12 defects have a bleeding diathesis of variable severity: it was therefore not unexpected that drugs targeting P2Y12 increase the incidence of bleedings. Although the value of laboratory tests of hemostasis for the prediction of bleeding events in patients with acute coronary syndromes under antithrombtic treatment is extensively being evaluated, a recent study showed that a simple bleeding score, based on 6 readily available clinical and laboratory variables (female sex, advanced age, elevated serum creatinine and white blood cell count, anemia, and type of acute coronary syndrome), plus the anticoagulation regimen used, may provide a rapid tool to predict the rate of major bleeding in these patients.94 

Major bleedings and the risk of mortality

Severe bleeding complications during antithrombotic therapy have negative consequences, not only because they may be fatal, disabling, and expose the patients to the risks that are associated with blood transfusion: they are also associated with poor prognosis of the patients, whose risk of death is increased during a follow-up of up to 1 year.94-96  The nature of the relationship between major bleeding complications and long-term mortality is unclear. Although this association remained statistically significant after adjustment for confounders, it is still possible that bleeding may simply be a marker of an underlying severe disease state, which exposes the patient to increased risk of mortality. A direct causal effect of major bleeding on the long-term risk of death is doubtful, and it is ruled out by the observation that major bleeding after CABG surgery is not associated with increased mortality.94,96  Yet, the possibility that non–CABG-related major bleeding may be indirectly causally associated with increased mortality of patients under treatment with antithrombotic drugs is biologically plausible. Antithrombotic drugs are usually withheld in patients who experience major bleeding, and this exposes them to high risk of MACE. In keeping with this hypothesis is the observation that major bleeding was also associated with increased risk of ischemic events, such as myocardial infarction and stroke.97 

Old thienopyridines versus ASA

Randomized clinical trials that compared old thienopyridines (ticlopidine or clopidogrel) with ASA showed that the risk of bleeding was not different between the 2 treatment arms (Table 3). This observation is somewhat surprising, considering that in vitro and in vivo experiments demonstrated that ADP plays a more important role in platelet thrombus formation than thromboxane A2. Although there are many plausible explanations for these unexpected results, the high prevalence of nonresponders to old thienopyridines, who are not exposed to the risk of bleeding, is the most plausible.

Combined treatment with clopidogrel and ASA

Combined treatment with clopidogrel and ASA is associated with increased bleeding compared with monotherapy. If this is a fair price to pay when treating patients with ACS, in consideration of the net clinical benefit associated with combined therapy, it is unacceptable for secondary prophylaxis of patients with stable disease or for primary prophylaxis of patients at risk because the higher bleeding risk is not counterbalanced by antithrombotic efficacy in these settings (Table 3).

New P2Y12 inhibitors versus clopidogrel

The higher incidence of bleeding complications that was observed in patients treated with the new P2Y12 antagonists prasugrel and ticagrelor, compared with clopidogrel (Table 4), is mostly explained by the fact that the new drugs effectively inhibit P2Y12-dependent platelet function in the great majority of treated patients, who will not only be protected from MACE, but will also be exposed to higher risk of bleeding, compared with patients treated with clopidogel who do not respond adequately to the drug. This hypothesis is biologically plausible and is supported by the observation that patients who respond adequately to clopidogrel not only are better protected from MACE, but also experience an increased incidence of bleedings, compared with poor responders.

P2Y12 inhibitors and CABG-related bleeding

A variable percentage of patients with acute coronary syndromes (< 10%) need to undergo CABG surgery. It has been demonstrated that clopidogrel treatment within approximately 4 days of the procedure is associated with increased blood loss, reoperation for bleeding, increased transfusion requirements, and prolonged intensive care unit and hospital stays.54  For this reason, when the clinical conditions of the patients allow it, clopidgrel is usually withheld for 5 days before CABG, to restore the hemostatic competency of the patient. This procedure was followed for all patients undergoing CABG in randomized clinical trials that compared the new P2Y12 antagonists to clopidogrel. For this reason, the incidence of CABG-related bleeding complications should not be considered when evaluating the risk of bleeding associated with the new anti-P2Y12 drug, for the simple reason that patients were off-treatment when they underwent CABG. Differences among P2Y12 antagonists in this setting should be evaluated on the basis of the time needed to withhold treatment before surgery to restore hemostatic competency. Considering that withholding antiplatelet treatment exposes patients to high risk of MACE, it is obvious that drugs with reversible mechanism of action and short half-life, such as ticagrelor, may be preferable to drugs that irreversibly inhibit the receptor.

The author thanks Dr Lidia Rota and Elena M. Faioni for critically reading his manuscript.

Contribution: M.C. searched and read the literature and wrote the paper.

Conflict-of-interest disclosure: M.C. participated in advisory board meetings and received lecture honoraria by AstraZeneca, Eli Lilli-Daiichi Sankyo, and The Medicines Company.

Correspondence: Marco Cattaneo, Unità di Medicina 3, Ospedale San Paolo, Università degli Studi di Milano, Via di Rudinì 8, 20142 Milano, Italy; e-mail: marco.cattaneo@unimi.it.

1
Packham
 
MA
Mustard
 
JF
Platelet aggregation and adenosine diphosphate/adenosine triphosphate receptors: a historical perspective.
Semin Thromb Hemost
2005
, vol. 
31
 
2
(pg. 
129
-
138
)
2
Cattaneo
 
M
Gachet
 
C
ADP receptors and clinical bleeding disorders.
Arterioscler Thromb Vasc Biol
1999
, vol. 
19
 
10
(pg. 
2281
-
2285
)
3
Cattaneo
 
M
Canciani
 
MT
Lecchi
 
A
, et al. 
Released adenosine diphosphate stabilizes thrombin-induced human platelet aggregates.
Blood
1990
, vol. 
75
 
5
(pg. 
1081
-
1086
)
4
Cattaneo
 
M
Akkawat
 
B
Kinlough-Rathbone
 
RL
Packham
 
MA
Cimminiello
 
C
Mannucci
 
PM
Ticlopidine facilitates the deaggregation of human platelets aggregated by thrombin.
Thromb Haemost
1994
, vol. 
71
 
1
(pg. 
91
-
94
)
5
Trumel
 
C
Payrastre
 
B
Plantavid
 
M
, et al. 
A key role of adenosine diphosphate in the irreversible platelet aggregation induced by the PAR1-activating peptide through the late activation of phosphoinositide 3-kinase.
Blood
1999
, vol. 
94
 
12
(pg. 
4156
-
4165
)
6
Abbracchio
 
MP
Burnstock
 
G
Boeynaems
 
JM
, et al. 
The recently deorphanized GPR80 (GPR99) proposed to be the P2Y15 receptor is not a genuine P2Y receptor.
Trends Pharmacol Sci
2005
, vol. 
26
 
1
(pg. 
8
-
9
)
7
Jin
 
J
Kunapuli
 
SP
Coactivation of two different G protein-coupled receptors is essential for ADP-induced platelet aggregation.
Proc Natl Acad Sci U S A
1998
, vol. 
95
 
14
(pg. 
8070
-
8074
)
8
Cattaneo
 
M
Michelson
 
AD
The platelet P2 receptors.
Platelets
2006
San Diego, CA
Academic Press
(pg. 
201
-
220
)
9
Savi
 
P
Zachayus
 
JL
Delesque-Touchard
 
N
, et al. 
The active metabolite of Clopidogrel disrupts P2Y12 receptor oligomers and partitions them out of lipid rafts.
Proc Natl Acad Sci U S A
2006
, vol. 
103
 
29
(pg. 
11069
-
11074
)
10
Savi
 
P
Pereillo
 
JM
Uzabiaga
 
MF
, et al. 
Identification and biological activity of the active metabolite of clopidogrel.
Thromb Haemost
2000
, vol. 
84
 
5
(pg. 
891
-
896
)
11
Ding
 
Z
Kim
 
S
Dorsam
 
RT
Jin
 
J
Kunapuli
 
SP
Inactivation of the human P2Y12 receptor by thiol reagents requires interaction with both extracellular cysteine residues, Cys17 and Cys270.
Blood
2003
, vol. 
101
 
10
(pg. 
3908
-
3914
)
12
Zhong
 
X
Kriz
 
R
Seehra
 
J
Kumar
 
R
N-linked glycosylation of platelet P2Y12 ADP receptor is essential for signal transduction but not for ligand binding or cell surface expression.
FEBS Lett
2004
, vol. 
562
 
1
(pg. 
111
-
117
)
13
Kauffenstein
 
G
Hechler
 
B
Cazenave
 
JP
Gachet
 
C
Adenine triphosphate nucleotides are antagonists at the P2Y receptor.
J Thromb Haemost
2004
, vol. 
2
 
11
(pg. 
1980
-
1988
)
14
Quinton
 
TM
Kim
 
S
Jin
 
J
Kunapuli
 
SP
Lipid rafts are required in Galpha(i) signaling downstream of the P2Y12 receptor during ADP-mediated platelet activation.
J Thromb Haemost
2005
, vol. 
3
 
5
(pg. 
1036
-
1041
)
15
Haslam
 
RJ
Interactions of the pharmacological receptors of blood platelets with adenylate cyclase.
Ser Haematol
1973
, vol. 
6
 
3
(pg. 
333
-
350
)
16
Savi
 
P
Pflieger
 
AM
Herbert
 
JM
cAMP is not an important messenger for ADP-induced platelet aggregation.
Blood Coagul Fibrinolysis
1996
, vol. 
7
 
2
(pg. 
249
-
252
)
17
Gratacap
 
MP
Herault
 
JP
Viala
 
C
, et al. 
FcgammaRIIA requires a Gi-dependent pathway for an efficient stimulation of phosphoinositide 3-kinase, calcium mobilization, and platelet aggregation.
Blood
2000
, vol. 
96
 
10
(pg. 
3439
-
3446
)
18
Dangelmaier
 
C
Jin
 
J
Smith
 
JB
Kunapuli
 
SP
Potentiation of thromboxane A2-induced platelet secretion by Gi signaling through the phosphoinositide-3 kinase pathway.
Thromb Haemost
2000
, vol. 
85
 
2
(pg. 
341
-
348
)
19
Hirsch
 
E
Bosco
 
O
Tropel
 
P
, et al. 
Resistance to thromboembolism in PI3Kgamma-deficient mice.
FASEB J
2001
, vol. 
15
 
11
(pg. 
2019
-
2021
)
20
Jackson
 
SP
Yap
 
CL
Anderson
 
KE
Phosphoinositide 3-kinases and the regulation of platelet function.
Biochem Soc Trans
2004
, vol. 
32
 
2
(pg. 
387
-
392
)
21
Garcia
 
A
Kim
 
S
Bhavaraju
 
K
Schoenwaelder
 
SM
Kunapuli
 
SP
Role of phosphoinositide 3-kinase beta in platelet aggregation and thromboxane A2 generation mediated by Gi signalling pathways.
Biochem J
2010
, vol. 
429
 
2
(pg. 
369
-
377
)
22
Jarvis
 
GE
Humphries
 
RG
Robertson
 
MJ
Leff
 
P
ADP can induce aggregation of human platelets via both P2Y(1) and P(2T) receptors.
Br J Pharmacol
2000
, vol. 
129
 
2
(pg. 
275
-
282
)
23
Kauffenstein
 
G
Bergmeier
 
W
Eckly
 
A
, et al. 
The P2Y(12) receptor induces platelet aggregation through weak activation of the alpha(IIb)beta(3) integrin: a phosphoinositide 3-kinase-dependent mechanism.
FEBS Lett
2001
, vol. 
505
 
2
(pg. 
281
-
290
)
24
Cattaneo
 
M
The P2 receptors and congenital platelet function defects.
Semin Thromb Hemost
2005
, vol. 
31
 
2
(pg. 
168
-
173
)
25
Cattaneo
 
M
Lombardi
 
R
Zighetti
 
ML
, et al. 
Deficiency of [33P]2MeS-ADP binding sites on platelets with secretion defect, normal granule stores and normal thromboxane A2 production: evidence that ADP potentiates platelet secretion independently of the formation of large platelet aggregates and thromboxane A2 production.
Thromb Haemost
1997
, vol. 
77
 
5
(pg. 
986
-
990
)
26
Cattaneo
 
M
Lecchi
 
A
Lombardi
 
R
Gachet
 
C
Zighetti
 
ML
Platelets from a patient heterozygous for the defect of P2CYC receptors for ADP have a secretion defect despite normal thromboxane A2 production and normal granule stores: further evidence that some cases of platelet ‘primary secretion defect’ are heterozygous for a defect of P2CYC receptors.
Arterioscler Thromb Vasc Biol
2000
, vol. 
20
 
11
(pg. 
E101
-
E106
)
27
Lecchi
 
A
Cattaneo
 
M
Mannucci
 
PM
ADP potentiates platelet-dense granule secretion induced by U46619 or TRAP through its interaction with the P2cyc receptor [abstract].
Haematologica
2000
, vol. 
85
 
suppl
pg. 
83
 
28
Li
 
Z
Zhang
 
G
Le Breton
 
GC
Gao
 
X
Malik
 
AB
Du
 
X
Two waves of platelet secretion induced by thromboxane A2 receptor and a critical role for phosphoinositide 3-kinases.
J Biol Chem
2003
, vol. 
278
 
33
(pg. 
30725
-
30731
)
29
Eckly
 
A
Gendrault
 
JL
Hechler
 
B
Cazenave
 
JP
Gachet
 
C
Differential involvement of the P2Y1 and P2YT receptors in the morphological changes of platelet aggregation.
Thromb Haemost
2001
, vol. 
85
 
4
(pg. 
694
-
701
)
30
Cattaneo
 
M
Lecchi
 
A
Randi
 
AM
McGregor
 
JL
Mannucci
 
PM
Identification of a new congenital defect of platelet function characterized by severe impairment of platelet responses to adenosine diphosphate.
Blood
1992
, vol. 
80
 
11
(pg. 
2787
-
2796
)
31
Foster
 
CJ
Prosser
 
DM
Agans
 
JM
, et al. 
Molecular identification and characterization of the platelet ADP receptor targeted by thienopyridine antithrombotic drugs.
J Clin Invest
2001
, vol. 
107
 
12
(pg. 
1591
-
1598
)
32
Cattaneo
 
M
Zighetti
 
ML
Lombardi
 
R
, et al. 
Molecular bases of defective signal transduction in the platelet P2Y12 receptor of a patient with congenital bleeding.
Proc Natl Acad Sci U S A
2003
, vol. 
100
 
4
(pg. 
1978
-
1983
)
33
Andre
 
P
Delaney
 
SM
LaRocca
 
T
, et al. 
P2Y12 regulates platelet adhesion/activation, thrombus growth, and thrombus stability in injured arteries.
J Clin Invest
2003
, vol. 
112
 
3
(pg. 
398
-
406
)
34
Nurden
 
P
Savi
 
P
Heilmann
 
E
, et al. 
An inherited bleeding disorder linked to a defective interaction between ADP and its receptor on platelets: its influence on glycoprotein IIb-IIIa complex function.
J Clin Invest
1995
, vol. 
95
 
4
(pg. 
1612
-
1622
)
35
Cattaneo
 
M
Lombardi
 
R
Bettega
 
D
Lecchi
 
A
Mannucci
 
PM
Shear-induced platelet aggregation is potentiated by desmopressin (DDAVP) and inhibited by ticlopidine.
Arterioscler Thromb
1993
, vol. 
13
 
3
(pg. 
393
-
397
)
36
Cattaneo
 
M
Zighetti
 
ML
Lombardi
 
R
Mannucci
 
PM
Role of ADP in platelet aggregation at high shear: studies in a patient with congenital defect of platelet responses to ADP.
Br J Haematol
1994
, vol. 
88
 
4
(pg. 
826
-
829
)
37
Resendiz
 
JC
Feng
 
S
Ji
 
G
Francis
 
KA
Berndt
 
MC
Kroll
 
MH
Purinergic P2Y12 receptor blockade inhibits shear-induced platelet phosphatidylinositol 3-kinase activation.
Mol Pharmacol
2003
, vol. 
63
 
3
(pg. 
639
-
645
)
38
Cattaneo
 
M
Lecchi
 
A
Inhibition of the platelet P2Y12 receptor for adenosine diphosphate potentiates the antiplatelet effect of prostacyclin.
J Thromb Haemost
2007
, vol. 
5
 
3
(pg. 
577
-
582
)
39
Gachet
 
C
Cattaneo
 
M
Ohlmann
 
P
, et al. 
Purinoceptors on blood platelets: further pharmacological and clinical evidence to suggest the presence of two ADP receptors.
Br J Haematol
1995
, vol. 
91
 
2
(pg. 
434
-
444
)
40
Shiraga
 
M
Miyata
 
S
Kato
 
H
, et al. 
Impaired platelet function in a patients with P2Y12 deficiency caused by a mutation in the translation initiation codon.
J Thromb Haemost
2005
, vol. 
3
 
10
(pg. 
2315
-
2323
)
41
Dawood
 
BB
Daly
 
M
Makris
 
M
Mumford
 
A
Watson
 
S
Wilde
 
J
Identification of a novel homozygous P2Y12 mutation in a patient with a mild platelet-based bleeding disorder [abstract].
J Thromb Haemost
2009
, vol. 
7
 
suppl 2
 
Abstract PP-MO-076
42
Cattaneo
 
M
Inherited platelet-based bleeding disorders.
J Thromb Haemost
2003
, vol. 
1
 
7
(pg. 
1628
-
1636
)
43
Conley
 
PB
Jurek
 
MM
Vincent
 
D
Lecchi
 
A
Cattaneo
 
M
Unique mutations in the P2Y12 locus of patients with previously described defects in ADP-dependent aggregation [abstract].
Blood
2001
, vol. 
98
 pg. 
43
 
44
Fontana
 
G
Ware
 
J
Cattaneo
 
M
Haploinsufficiency of the platelet P2Y12 gene in a family with congenital bleeding diathesis.
Haematologica
2009
, vol. 
94
 
4
(pg. 
581
-
584
)
45
Hollopeter
 
G
Jantzen
 
HM
Vincent
 
D
, et al. 
Identification of the platelet ADP receptor targeted by antithrombotic drugs.
Nature
2001
, vol. 
409
 
6817
(pg. 
202
-
207
)
46
Moro
 
S
Hoffmann
 
C
Jacobson
 
KA
Role of the extracellular loops of G protein-coupled receptors in ligand recognition: a molecular modeling study of the human P2Y1 receptor.
Biochemistry
1999
, vol. 
38
 
12
(pg. 
3498
-
3507
)
47
Moro
 
S
Guo
 
D
Camaioni
 
E
Boyer
 
JL
Harden
 
TK
Jacobson
 
KA
Human P2Y1 receptor: molecular modeling and site-directed mutagenesis as tools to identify agonist and antagonist recognition sites.
J Med Chem
1998
, vol. 
41
 
9
(pg. 
1456
-
1466
)
48
Remijn
 
JA
IJsseldijk
 
MJ
Strunk
 
AL
, et al. 
Novel molecular defect in the platelet ADP receptor P2Y12 of a patient with haemorrhagic diathesis.
Clin Chem Lab Med
2007
, vol. 
45
 
2
(pg. 
187
-
189
)
49
Daly
 
ME
Dawood
 
BB
Lester
 
WA
, et al. 
Identification and characterization of a novel P2Y 12 variant in a patient diagnosed with type 1 von Willebrand disease in the European MCMDM-1VWD study.
Blood
2009
, vol. 
113
 
17
(pg. 
4110
-
4113
)
50
Tosetto
 
A
Rodeghiero
 
F
Castaman
 
G
, et al. 
A quantitative analysis of bleeding symptoms in type 1 von Willebrand disease: results from a multicenter European study (MCMDM-1 VWD).
J Thromb Haemost
2006
, vol. 
4
 
4
(pg. 
766
-
773
)
51
Zighetti
 
ML
Carpani
 
G
Sinigaglia
 
E
Cattaneo
 
M
Usefulness of a flow cytometric analysis of intraplatelet VASP phosphorylation for the detection of patients with genetic defects of the platelet P2Y12 receptor for ADP.
J Thromb Haemost
2010
, vol. 
8
 
10
(pg. 
2332
-
2334
)
52
Cattaneo
 
M
Mannucci
 
PM
Michelson
 
AD
Desmopressin.
Platelets
2002
San Diego, CA
Academic Press
(pg. 
855
-
866
)
53
Cattaneo
 
M
Michelson
 
AD
ADP receptors antagonists.
Platelets
2006
San Diego, CA
Academic Press
(pg. 
1127
-
1144
)
54
Cattaneo
 
M
New P2Y(12) inhibitors.
Circulation
2010
, vol. 
121
 
1
(pg. 
171
-
179
)
55
Gent
 
M
Blakely
 
JA
Easton
 
JD
, et al. 
The Canadian American Ticlopidine Study (CATS) in thromboembolic stroke.
Lancet
1989
, vol. 
1
 
8649
(pg. 
1215
-
1220
)
56
Hass
 
WK
Easton
 
JD
Adams
 
HP
, et al. 
A randomized trial comparing ticlopidine hydrochloride with aspirin for the prevention of stroke in high-risk patients: Ticlopidine Aspirin Stroke Study Group.
N Engl J Med
1989
, vol. 
321
 
8
(pg. 
501
-
507
)
57
Schömig
 
A
Neumann
 
FJ
Kastrati
 
A
, et al. 
A randomized comparison of antiplatelet and anticoagulant therapy after the placement of coronary-artery stents.
N Engl J Med
1996
, vol. 
334
 
17
(pg. 
1084
-
1089
)
58
Leon
 
MB
Baim
 
DS
Popma
 
JJ
, et al. 
A clinical trial comparing three antithrombotic-drug regimens after coronary-artery stenting.
N Engl J Med
1998
, vol. 
339
 
23
(pg. 
1665
-
1671
)
59
CAPRIE Steering Committee
A randomised, blinded, trial of clopidogrel versus aspirin in patients at risk of ischaemic events (CAPRIE).
Lancet
1996
, vol. 
348
 
9038
(pg. 
1329
-
1339
)
60
Bhatt
 
DL
Fox
 
KA
Hacke
 
W
, et al. 
Clopidogrel and aspirin versus aspirin alone for the prevention of atherothrombotic events.
N Engl J Med
2006
, vol. 
354
 
16
(pg. 
1706
-
1717
)
61
Diener
 
HC
Bogousslavsky
 
J
Brass
 
LM
, et al. 
Aspirin and clopidogrel compared with clopidogrel alone after recent ischaemic stroke or transient ischaemic attack in high-risk patients (MATCH): randomised, double-blind, placebo-controlled trial.
Lancet
2004
, vol. 
364
 
9431
(pg. 
331
-
337
)
62
Yusuf
 
S
Zhao
 
F
Mehta
 
SR
, et al. 
Effects of clopidogrel in addition to aspirin in patients with acute coronary syndromes without ST-segment elevation.
N Engl J Med
2001
, vol. 
345
 
7
(pg. 
494
-
502
)
63
Chen
 
ZM
Jiang
 
LX
Chen
 
YP
, et al. 
Addition of clopidogrel to aspirin in 45 852 patients with acute myocardial infarction: randomised placebo-controlled trial.
Lancet
2005
, vol. 
366
 
9497
(pg. 
1607
-
1621
)
64
Mehta
 
SR
Yusuf
 
S
Peters
 
RJ
, et al. 
Effects of pretreatment with clopidogrel and aspirin followed by long-term therapy in patients undergoing percutaneous coronary intervention: the PCI-CURE study.
Lancet
2001
, vol. 
358
 
9281
(pg. 
527
-
533
)
65
Snoep
 
JD
Hovens
 
MM
Eikenboom
 
JC
van der Bom
 
JG
Jukema
 
JW
Huisman
 
MV
Clopidogrel nonresponsiveness in patients undergoing percutaneous coronary intervention with stenting: a systematic review and meta-analysis.
Am Heart J
2007
, vol. 
154
 
2
(pg. 
221
-
231
)
66
Hulot
 
JS
Bura
 
A
Villard
 
E
, et al. 
Cytochrome P450 2C19 loss-of-function polymorphism is a major determinant of clopidogrel responsiveness in healthy subjects.
Blood
2006
, vol. 
108
 
7
(pg. 
2244
-
2247
)
67
Brandt
 
JT
Close
 
SL
Iturria
 
SJ
, et al. 
Common polymorphisms of CYP2C19 and CYP2C9 affect the pharmacokinetic and pharmacodynamic response to clopidogrel but not prasugrel.
J Thromb Haemost
2007
, vol. 
5
 
12
(pg. 
2429
-
2436
)
68
Hulot
 
JS
Collet
 
JP
Silvain
 
J
, et al. 
Cardiovascular risk in clopidogrel-treated patients according to cytochrome P450 2C19*2 loss-of-function allele or proton pump inhibitor coadministration: a systematic meta-analysis.
J Am Coll Cardiol
2010
, vol. 
56
 
2
(pg. 
134
-
143
)
69
Simon
 
T
Verstuyft
 
C
Mary-Krause
 
M
, et al. 
Genetic determinants of response to clopidogrel and cardiovascular events.
N Engl J Med
2009
, vol. 
360
 
4
(pg. 
363
-
375
)
70
Mega
 
JL
Close
 
SL
Wiviott
 
SD
, et al. 
Genetic variants in ABCB1 and CYP2C19 and cardiovascular outcomes after treatment with clopidogrel and prasugrel in the TRITON-TIMI 38 trial: a pharmacogenetic analysis.
Lancet
2010
, vol. 
376
 
9749
(pg. 
1312
-
1319
)
71
Breet
 
NJ
van Werkum
 
JW
Bouman
 
HJ
, et al. 
Comparison of platelet function tests in predicting clinical outcome in patients undergoing coronary stent implantation.
JAMA
2010
, vol. 
303
 
8
(pg. 
754
-
762
)
72
Cuisset
 
T
Frere
 
C
Poyet
 
R
, et al. 
Clopidogrel response: head-to-head comparison of different platelet assays to identify clopidogrel non responder patients after coronary stenting.
Arch Cardiovasc Dis
2010
, vol. 
103
 
1
(pg. 
39
-
45
)
73
Gremmel
 
T
Steiner
 
S
Seidinger
 
D
Koppensteiner
 
R
Panzer
 
S
Kopp
 
CW
Comparison of methods to evaluate clopidogrel-mediated platelet inhibition after percutaneous intervention with stent implantation.
Thromb Haemost
2009
, vol. 
101
 
2
(pg. 
333
-
339
)
74
Paniccia
 
R
Antonucci
 
E
Maggini
 
N
, et al. 
Comparison of methods for monitoring residual platelet reactivity after clopidogrel by point-of-care tests on whole blood in high-risk patients.
Thromb Haemost
2010
, vol. 
104
 
2
(pg. 
287
-
292
)
75
Kim
 
IS
Jeong
 
YH
Kang
 
MK
, et al. 
Correlation of high post-treatment platelet reactivity assessed by light transmittance aggregometry and the VerifyNow P2Y(12) assay.
J Thromb Thrombolysis
2010
, vol. 
30
 
4
(pg. 
486
-
495
)
76
Madsen
 
EH
Saw
 
J
Kristensen
 
SR
Schmidt
 
EB
Pittendreigh
 
C
Maurer-Spurej
 
E
Long-term aspirin and clopidogrel response evaluated by light transmission aggregometry, VerifyNow, and thrombelastography in patients undergoing percutaneous coronary intervention.
Clin Chem
2010
, vol. 
56
 
5
(pg. 
839
-
847
)
77
Aleil
 
B
Léon
 
C
Cazenave
 
JP
Gachet
 
C
CYP2C19*2 polymorphism is not the sole determinant of the response to clopidogrel: implications for its monitoring.
J Thromb Haemost
2009
, vol. 
7
 
10
(pg. 
1747
-
1749
)
78
Hochholzer
 
W
Trenk
 
D
Fromm
 
MF
, et al. 
Impact of cytochrome P450 2C19 loss-of-function polymorphism and of major demographic characteristics on residual platelet function after loading and maintenance treatment with clopidogrel in patients undergoing elective coronary stent placement.
J Am Coll Cardiol
2010
, vol. 
55
 
22
(pg. 
2427
-
2434
)
79
Bonello
 
L
Camoin-Jau
 
L
Armero
 
S
, et al. 
Tailored clopidogrel loading dose according to platelet reactivity monitoring to prevent acute and subacute stent thrombosis.
Am J Cardiol
2009
, vol. 
103
 
1
(pg. 
5
-
10
)
80
Pena
 
A
Collet
 
JP
Hulot
 
JS
, et al. 
Can we override clopidogrel resistance?
Circulation
2009
, vol. 
119
 
21
(pg. 
2854
-
2857
)
81
Bonello
 
L
Tantry
 
US
Marcucci
 
R
, et al. 
Consensus and future directions on the definition of high on-treatment platelet reactivity to adenosine diphosphate.
J Am Coll Cardiol
2010
, vol. 
56
 
12
(pg. 
919
-
933
)
82
Wiviott
 
SD
Braunwald
 
E
McCabe
 
CH
, et al. 
Prasugrel versus clopidogrel in patients with acute coronary syndromes.
N Engl J Med
2007
, vol. 
357
 
20
(pg. 
2001
-
2015
)
83
Wallentin
 
L
Varenhorst
 
C
James
 
S
, et al. 
Prasugrel achieves greater and faster P2Y12receptor-mediated platelet inhibition than clopidogrel due to more efficient generation of its active metabolite in aspirin-treated patients with coronary artery disease.
Eur Heart J
2008
, vol. 
29
 
1
(pg. 
21
-
30
)
84
Sugidachi
 
A
Ogawa
 
T
Kurihara
 
A
, et al. 
The greater in vivo antiplatelet effects of prasugrel as compared to clopidogrel reflect more efficient generation of its active metabolite with similar antiplatelet activity to that of clopidogrel's active metabolite.
J Thromb Haemost
2007
, vol. 
5
 
7
(pg. 
1545
-
1551
)
85
van Giezen
 
JJ
Humphries
 
RG
Preclinical and clinical studies with selective reversible direct P2Y12 antagonists.
Semin Thromb Hemost
2005
, vol. 
31
 
2
(pg. 
195
-
204
)
86
Teng
 
R
Butler
 
K
Pharmacokinetics, pharmacodynamics, tolerability and safety of single ascending doses of ticagrelor, a reversibly binding oral P2Y(12) receptor antagonist, in healthy subjects.
Eur J Clin Pharmacol
2010
, vol. 
66
 
5
(pg. 
487
-
496
)
87
van Giezen
 
JJ
Nilsson
 
L
Berntsson
 
P
, et al. 
Ticagrelor binds to the human P2Y receptor independently from ADP but antagonizes ADP-induced receptor signaling and platelet aggregation.
J Thromb Haemost
2009
, vol. 
7
 
9
(pg. 
1556
-
1565
)
88
Husted
 
S
Emanuelsson
 
H
Heptinstall
 
S
Sandset
 
PM
Wickens
 
M
Peters
 
G
Pharmacodynamics, pharmacokinetics, and safety of the oral reversible P2Y12 antagonist AZD6140 with aspirin in patients with atherosclerosis: a double-blind comparison to clopidogrel with aspirin.
Eur Heart J
2006
, vol. 
27
 
9
(pg. 
1038
-
1047
)
89
Cannon
 
CP
Husted
 
S
Harrington
 
RA
, et al. 
Safety, tolerability, and initial efficacy of AZD6140, the first reversible oral adenosine diphosphate receptor antagonist, compared with clopidogrel, in patients with non-ST-segment elevation acute coronary syndrome: primary results of the DISPERSE-2 trial.
J Am Coll Cardiol
2007
, vol. 
50
 
19
(pg. 
1844
-
1851
)
90
Gurbel
 
PA
Bliden
 
KP
Butler
 
K
, et al. 
Randomized double-blind assessment of the ONSET and OFFSET of the antiplatelet effects of ticagrelor versus clopidogrel in patients with stable coronary artery disease: the ONSET/OFFSET study.
Circulation
2009
, vol. 
120
 
25
(pg. 
2577
-
2585
)
91
Wallentin
 
L
Becker
 
RC
Budaj
 
A
, et al. 
Ticagrelor versus clopidogrel in patients with acute coronary syndromes.
N Engl J Med
2009
, vol. 
361
 
11
(pg. 
1045
-
1057
)
92
Bhatt
 
DL
Lincoff
 
AM
Gibson
 
CM
, et al. 
Intravenous platelet blockade with cangrelor during PCI.
N Engl J Med
2009
, vol. 
361
 
24
(pg. 
2330
-
2341
)
93
Harrington
 
RA
Stone
 
GW
McNulty
 
S
, et al. 
Platelet inhibition with cangrelor in patients undergoing PCI.
N Engl J Med
2009
, vol. 
361
 
24
(pg. 
2318
-
2329
)
94
Mehran
 
R
Pocock
 
SJ
Nikolsky
 
E
, et al. 
A risk score to predict bleeding in patients with acute coronary syndromes.
J Am Coll Cardiol
2010
, vol. 
55
 
23
(pg. 
2556
-
2566
)
95
Eikelboom
 
JW
Mehta
 
SR
Anand
 
SS
Xie
 
C
Fox
 
KA
Yusuf
 
S
Adverse impact of bleeding on prognosis in patients with acute coronary syndromes.
Circulation
2006
, vol. 
114
 
8
(pg. 
774
-
782
)
96
Pocock
 
SJ
Mehran
 
R
Clayton
 
TC
, et al. 
Prognostic modeling of individual patient risk and mortality impact of ischemic and hemorrhagic complications: assessment from the Acute Catheterization and Urgent Intervention Triage Strategy trial.
Circulation
2010
, vol. 
121
 
1
(pg. 
43
-
51
)
97
van Hattum
 
ES
Algra
 
A
Lawson
 
JA
Eikelboom
 
BC
Moll
 
FL
Tangelder
 
MJ
Bleeding increases the risk of ischemic events in patients with peripheral arterial disease.
Circulation
2009
, vol. 
120
 
16
(pg. 
1569
-
1576
)
Sign in via your Institution