eISSN: 1897-4295
ISSN: 1734-9338
Advances in Interventional Cardiology/Postępy w Kardiologii Interwencyjnej
Current issue Archive Manuscripts accepted About the journal Editorial board Abstracting and indexing Subscription Contact Instructions for authors Publication charge Ethical standards and procedures
Editorial System
Submit your Manuscript
SCImago Journal & Country Rank
2/2021
vol. 17
 
Share:
Share:
Original paper

The influence of acute coronary syndrome on the levels of clopidogrel active metabolite and platelet inhibition in patients with and without CYP2C19 and ABCB1 gene polymorphisms

Tomasz Wójcik
1
,
Bożena Karolko
1
,
Jerzy Wiśniewski
2
,
Andrzej Mysiak
1
,
Krzysztof  Ściborski
1
,
Grzegorz Onisk
1
,
Arleta Lebioda
3
,
Anna Jonkisz
3
,
Marcin Protasiewicz
1

1.
Department and Clinic of Cardiology, Wroclaw Medical University, Wroclaw, Poland
2.
Department of Biochemistry, Wroclaw Medical University, Wroclaw, Poland
3.
Department of Molecular Techniques, Wroclaw Medical University, Wroclaw, Poland
Adv Interv Cardiol 2021; 17, 2 (64): 179–186
Online publish date: 2021/07/09
Article file
Get citation
 
 

Summary

There is evidence to support an association between the clinical status of patients and the bioavailability of clopidogrel, which is still widely used in cases of acute coronary syndrome (ACS). Additionally, clopidogrel bioavailability may be affected by the presence of CYP2C19 or ABCB1 gene polymorphisms. Our study showed that patients with ACS have similar levels of clopidogrel active metabolite as those with stable coronary artery disease. This association was observed both in individuals with ST-segment and non-ST-segment elevation myocardial infarction and was independent of the presence of CYP2C19 or ABCB1 gene polymorphisms.

Introduction

Although ticagrelor and prasugrel have become the standard antiplatelet treatments in acute coronary syndromes (ACSs), a large group of patients still present with indications for the use of clopidogrel. In particular, this refers to patients with stable coronary artery disease who are treated invasively as well as patients with ACS who have contraindications to the use of novel antiplatelet drugs (e.g., due to a high risk of bleeding associated with antithrombotic treatment) [13].

Clopidogrel is a prodrug that is metabolized to its active form by cytochrome p450 via complex biochemical processes that occur in the liver. The absorption of clopidogrel is regulated by glycoprotein P [4]. Only the active metabolite of clopidogrel can block the P2Y12 receptor, thus leading to inhibition of platelet activation and aggregation. It is believed that the clinical status of patients may affect clopidogrel metabolism [5, 6], which, in turn, may influence the antiplatelet effect of its active metabolite.

Aim

In the current study, we aimed to assess the levels of clopidogrel active metabolite with liquid chromatography–tandem mass spectrometry (LC-MS/MS) and to evaluate the effect of the drug on platelet inhibition in patients with ST-segment elevation myocardial infarction (STEMI) and non-STEMI (NSTEMI). To exclude genetic variability that might have affected the drug concentration and activity, patients were also assessed for the presence of the most common genetic polymorphisms that reduce the absorption (ABCB1) and activation (CYP2C19*2 and CYP2C19*3) of clopidogrel.

Material and methods

Study population

This single-center, open-label, prospective study included 199 patients hospitalized due to NSTEMI (n = 91) or STEMI (n = 108), who underwent percutaneous coronary intervention (PCI) with drug-eluting stent implantation. Patients with cardiogenic shock and those with Killip class IV were excluded. All patients provided written informed consent to participate in the study, and the study was approved by the local ethics committee.

The control group included 22 patients with stable coronary artery disease treated with PCI. All participants of the control group were tested to exclude the presence of ABCB1 as well as CYP2C19*2 and CYP2C19*3 gene polymorphisms.

All patients provided written informed consent to participate in the study, and the study was approved by the Wroclaw Medical University Ethics Committee.

Clopidogrel administration, blood collection, and plasma preparation

All participants received a loading dose of clopidogrel containing clopidogrel bisulfate as the active substance (600 mg), followed by a maintenance dose of 75 mg for the next 3 days. On day 3, blood samples were obtained 1 h after drug administration to assess the concentrations of the active metabolite.

Sample preparation

The method used to assess the clopidogrel active metabolite was described in our previous study [7]. Briefly, to stabilize the active clopidogrel metabolite, immediately after blood collection 2-bromo-30-methoxyacetophenone (MPB) was added to the sample. The MPB-derivatized clopidogrel active metabolite hydrochloride and [13C6]-(±)-clopidogrel carboxylic acid (internal standard) were purchased from Alsachim (Illkirch-Graffenstaden, France). Clopidogrel and clopidogrel carboxylic acid were obtained from Pharmaceutical Research Institute (Warsaw, Poland). Liquid chromatography–mass spectrometry (LC-MS) grade water, methanol, and acetonitrile were obtained from J.T. Baker (Deventer, the Netherlands). The formic acid (purity ≥ 98%), trichloroacetic acid (purity ≥ 99.5%), and MPB were purchased from Sigma-Aldrich (Poznan, Poland), while leucine-enkephalin was purchased from Waters (Warsaw, Poland). The serum concentrations of clopidogrel active metabolite hydrochloride, clopidogrel, and clopidogrel carboxylic acid were quantified by stable-isotope dilution LC-MS, according to a modified method adapted from Karaźniewicz-Łady et al. After additional centrifugation, the supernatant was analyzed using LC-MS/MS.

Liquid chromatography–tandem mass spectrometry

The LC-MS/MS analysis was performed using the nanoACQUITY UPLC system combined with a Xevo G2 QTof mass spectrometer (Waters, Warsaw, Poland). The analyzed compounds were separated in the HSS C18 column with a membrane inline filter (Waters, Milford, Massachusetts, United States) at 45°C. Mobile phase A consisted of 0.1% formic acid in water, while mobile phase B consisted of 0.1% formic acid in acetonitrile with an increasing gradient. The total run time of the method was 4 min, with a flow rate of 45 μl/min. Mass spectra for the analyzed compounds were acquired in positive ion mode electrospray ionization. Data acquisition was performed by means of the MassLynx Software (Waters, Warsaw, Poland), using the characteristic precursor and product ions. A quantitative analysis was performed using the QuanLynx software (Waters, Warsaw, Poland).

Platelet aggregation

Platelet aggregation was assessed on day 3 of clopidogrel treatment. Blood samples were obtained from a venous cannula into 2 tubes containing 0.109-mol/l trisodium citrate. Then, the blood was centrifuged at room temperature (800 × g for 15 min) to collect platelet-rich plasma. The sample was recentrifuged at 2400 × g for 15 min, and platelet-poor plasma was collected. Platelet reactivity was assessed within 2 h after collection by light transmission aggregometry developed by Born, using a single-channel Chrono Log 560CA lumi aggregometer (Chrono-log, Haverton, Pennsylvania, United States). A platelet-poor plasma sample with 100% light transmission was used as a reference. Consecutive platelet-rich plasma samples (light transmission, 0%) were placed in cuvettes and stimulated with adenosine diphosphate (ADP; 5 μmol/l and 10 μmol/l). The results were expressed as percentage of the maximum platelet aggregation (MPA) within 6 min. Calculations and platelet aggregation curves were performed using the dedicated AgroLink software (Chrono-log, Haverton, Pennsylvania, United States). All reagents and laboratory equipment were purchased from Biogenet (Piaseczno, Poland). They were stored and used according to the manufacturer’s instructions. The cutoff MPA values of higher than 46% and 67% were used to identify the high on-treatment platelet reactivity (HPR) to stimulation with 5-μmol/l and 10-μmol/l ADP, respectively.

Genetic studies

To identify ABCB1(rs1045642, c.3435C> T), CYP2C19*2 (rs4244285, c.681 G>A), and CYP2C19*3 (rs4986893, c.636 G>A) polymorphism, genetic material was extracted from 200 μl of whole-blood samples of each participant, using the High Pure PCR Template Preparation Kit (Roche Diagnostics GmbH, Mannheim, Germany). The lysate was centrifuged in a mini-column containing the silica membrane, which was then rinsed twice with a washing buffer. Finally, a mini-column elution buffer was applied to the membrane to recover the purified DNA.

Next, polymerase chain reaction (PCR) was carried out with 3 pairs of specific primers for CYP2C19*3(636 G>A), ABCB1 (C3435C> T), and CYP2C19*2 (681 G>A) using a Multiplex PCR Kit (QIAGEN, Hilden, Germany), according to the manufacturer’s instructions. A single nucleotide polymorphism of CYP2C19*2, CYP2C19*3, and ABCB1 was genotyped using a minisequencing technique, which is a modification of PCR. The SNaPshot Multiplex Kit (Applied Biosystems, Foster City, California, United States) was used for the analysis, according to the manufacturer’s instructions. The minisequencing reaction was performed with specific forward and reverse primers designed to hybridize to the template, ending before the designated polymorphic site. Dideoxynucleotide triphosphates, or fluorescent-labeled terminators, were involved in the reaction. Product detection was performed with capillary electrophoresis, using a 3130 Genetic Analyzer (Applied Biosystems). The results were analyzed with the GeneMapper ID v. 3.2 software (Applied Biosystems) against the internal GeneScan LIZ 120 standard.

Patients with at least one allele that negatively affects clopidogrel metabolism (CYP2C19*2, *3), were assigned to the loss-of-function (LoF) group, as opposed to the normal-function group (NoF), which included patients with CYP2C19*1. The study population was also divided into “good absorbers” (the GA group homozygous for the wild-type allele at position 3435CC) and “poor absorbers” (the PA group with heterozygous (TC) and homozygous (TT) mutations of the ABCB1 gene).

Statistical analysis

The statistical analysis was performed using the Statistica 9.0 PL program. The type of variable distribution was determined by the Shapiro-Wilk test. For the comparison of quantitative variables between two groups, depending on the distribution of the variable, the Mann-Whitney U or Student t test was used. For more than two groups and variables with an abnormal distribution the ANOVA rank Kruskal-Wallis test was used. The χ2 test was used to compare the frequency of the occurrence of the qualitative variable between the groups.

Results

There were no significant differences in the demographic data between the study and control groups (Table I). The results of exact genotype distributions are briefly summarized in Table II.

Table I

Demographic data

ParameterControl (n = 22)Study population (n = 199)P-value
Age [years]63.0 ±8.265.5 ±9.90.64
Male, n (%)14 (63.6)121 (60.8)0.98
Hypertension, n (%)16 (72.7)155 (77.8)0.78
Body mass index [kg/m2]24.7 ±5.427.2 ±3.20.55
Smoking, n (%)6 (27.2)52 (26.1)0.89
Hyperlipidemia, n (%)15 (68.1)132 (65.8)0.95
Previous MI, n (%)5 (22.7)42 (21.1)0.91
Diabetes, n (%)8 (36.3)64 (32.1)0.87
eGFR [ml/min/m2]46.5 ±12.641.3 ±10.10.63

[i] eGFR – estimated glomerular filtration rate, MI – myocardial infarction.

Table II

Genotype distributions in the study population

Genotype frequency
NACYPC19*2, n (%)
199NoFLoF
GGGAAA
49 (24.6)147 (73.9)3 (1.5)
199CYPC19*3, n (%)
NoFLoF
GGGAAA
199 (100)0 (0)0 (0)
199ABCB1, n (%)
GAPA
CCTCTT
32 (16)115 (57.9)52 (26.1)

[i] GA – good absorbers, NA – number of individuals analyzed, NoF – normal function, LoF – loss of function, PA – poor absorbers.

There were no CYPC19*3 allele carriers in the entire population.

The median (IQR) level of active clopidogrel in 22 controls was 16.7 (9.0–21.1) ng/ml, as compared with 14.9 (6.9–21.9) ng/ml for the whole study group (p < 0.59). A subgroup analysis revealed no differences in the levels of active clopidogrel between patients with STEMI and those with NSTEMI (median (IQR): 16.1 (7.8–23.4) ng/ml and 13.6 (9.5–19.1) ng/ml; p < 0.32). No differences were also noted between patients with STEMI and controls or between those with NSTEMI and controls (p < 0.33 and p < 0.54, respectively). Moreover, the levels of active clopidogrel did not differ between the NoF and LoF groups (p < 0.47) or between the PA and GA groups (p < 0.71). The simultaneous presence of the ABCB1 and CYP2C19*2 polymorphisms (PA + LoF) was not associated with lower drug levels in comparison with the GA+NoF group (p = 0.28). There were no significant differences between the individual genetic subgroups in the whole study population (Figure 1 A) or separately in patients with STEMI and NSTEMI (Figures 1 B, C).

Figure 1

Levels of clopidogrel active metabolite in: A – the whole study group and subgroups; B – patients with ST-segment elevation myocardial infarction (STEMI) and STEMI subgroups; C – patients with non-ST-segment elevation myocardial infarction (NSTEMI) and NSTEMI subgroups

Control – control group, GA – good absorbers, PA – poor absorbers, LoF – loss of function alleles, NoF – normal function alleles, NSTEMI – non-ST-segment elevation myocardial infarction, STEMI – ST-segment elevation myocardial infarction, Total – whole population.

/f/fulltexts/PWKI/44380/PWKI-17-44380-g001_min.jpg

Platelet inhibition, after stimulation with either 5-μmol/l or 10-μmol/l ADP, was similar in all study subgroups as well as controls (Figures 2 AC, 3 AC).

Figure 2

Maximum platelet aggregation after stimulation with 5-μmol/l adenosine diphosphate in: A – the whole study group and subgroups, B – patients with ST-segment elevation myocardial infarction (STEMI) and STEMI subgroups, C – patients with non-ST-segment elevation myocardial infarction (NSTEMI) and NSTEMI subgroups

Control – control group, GA – good absorbers, PA – poor absorbers, LoF – loss of function alleles, NoF – normal function alleles, NSTEMI – non-ST-segment elevation myocardial infarction, STEMI – ST-segment elevation myocardial infarction, Total – whole population.

/f/fulltexts/PWKI/44380/PWKI-17-44380-g002_min.jpg
Figure 3

Maximum platelet aggregation after stimulation with 10-μmol/l adenosine diphosphate: A – the whole study group and subgroups, B – patients with ST-segment elevation myocardial infarction (STEMI) and STEMI subgroups, C – patients with non-ST-segment elevation myocardial infarction (NSTEMI) and NSTEMI subgroups

Control – control group, GA – good absorbers, PA – poor absorbers, LoF – loss of function alleles, NoF – normal function alleles, NSTEMI – non-ST-segment elevation myocardial infarction, STEMI – ST-segment elevation myocardial infarction, Total – whole population.

/f/fulltexts/PWKI/44380/PWKI-17-44380-g003_min.jpg

High on-treatment platelet reactivity was noted in 26.7% of controls, as compared with 28.1% of the whole study population (p = 0.89). Moreover, there were no differences in HPR between the STEMI group (29.6%) and controls (p = 0.76) or between the NSTEMI group (27.1%) and controls (p = 0.83).

In a linear regression model including clinical data and genetic analysis, no relationship was found between these variables and the concentration of clopidogrel active metabolite.

Discussion

A short half-life of active clopidogrel complicates the quantitative assessment of its blood levels. To date, only a few high-performance liquid chromatography assays have proved sensitive enough to measure active clopidogrel levels in human plasma [8]. In our study, we used modified methods that allowed us to stabilize the clopidogrel active metabolite in blood [7]. To assess its concentrations, we decided to use LC-MS/MS because it can determine the drug concentrations in blood with high sensitivity and selectivity [9, 10].

The cutoff values for defining clopidogrel nonresponsiveness using aggregometry are often arbitrary. In 2010, Bonello et al. [11] published a consensus statement in which they proposed a 5-μmol/l ADP-induced MPA of 46% as a cutoff value to identify HPR. In our study, we accepted this recommended threshold for MPA induced by 5-μmol/l ADP, while for 10-μmol/l ADP, the cutoff value of more than 67% was used to identify an inadequate response to clopidogrel, similarly to a study by Cuisset et al. [12].

Multiple data suggest that insufficient active metabolite generation is a significant reason for variability in clopidogrel response. High on-clopidogrel platelet reactivity has been reported to be associated with a significantly higher incidence of ischemic adverse effects in patients undergoing PCI. There is evidence suggesting that the severity of the patient’s clinical status significantly affects clopidogrel metabolism, which consequently influences the concentrations and antiplatelet effects of its active metabolite. Among patients admitted to intensive care units, the number of poorly responding individuals was reported to range from 65% to 80% [5]. Generation of the clopidogrel active metabolite was markedly lower than in healthy volunteers or other patient groups [1315]. In a study by Součková et al. [6], including patients undergoing successful cardiopulmonary resuscitation who received clopidogrel because of a PCI, clopidogrel bioavailability and platelet inhibition were lower than in patients in a stable clinical condition [6].

Our findings show that in patients with lower cardiovascular burden (i.e., patients with STEMI and NSTEMI but excluding those with cardiogenic shock and those with Killip class IV), ACS is not associated with changes in the blood levels of clopidogrel active metabolite or platelet aggregation in comparison with patients with stable coronary artery disease. No differences in the analyzed parameters were also found between patients classified according to the type of ACS (STEMI vs. NSTEMI).

Our data on the incidence of HPR are in line with previous studies, which reported the occurrence of HPR in about 30% of the population [16, 17].

Considering the available data on the possible effect of the CYP2C19 and ABCB1 polymorphisms on the bioavailability of clopidogrel and its antiplatelet effects, we decided to assess platelet aggregation and the levels of clopidogrel active metabolite in combination with genotyping. Precise data on ABCB1 and CYP2C19 genotype distribution observed in our population (summarized in Table II) were published previously [7]. The results were very similar to previous observations in a Caucasian population [18, 19] and confirm poor inspection of the CYP2C19 *3 polymorphism in this part of Europe.

Our study revealed no correlations between ABCB1 and CYP2C19*2 polymorphisms and the drug concentrations or platelet aggregation. Available data on the effect of ABCB1 and CYP2C19 polymorphisms on the concentrations of clopidogrel active metabolite and platelet aggregation are contradictory [20]. The ABCB1 C3435T polymorphism has been associated with changes in the intestinal efflux of drugs and thus their bioavailability [21]. A clinical study in patients undergoing PCI reported that peak plasma concentrations and area under the curve of clopidogrel active metabolite following administration of a loading dose of clopidogrel were significantly lower in 3435T/T homozygotes than those in 3435C/T heterozygotes or 3435C/C (wild-type) homozygotes [22]. However, these results were not confirmed by subsequent studies for the maintenance clopidogrel doses of 75 mg or 150 mg [23, 24]. Our findings are in line with the results of 2 meta-analyses showing that the ABCB1 C3435T polymorphism is unlikely to play a major role in between-subject variability in response to clopidogrel [25, 26].

The available data on the CYP2C19 polymorphism on the blood levels of clopidogrel active metabolite were mostly derived from studies on young healthy volunteers. A major study on the effect of genes on the hepatic metabolism of clopidogrel, conducted in a group of healthy volunteers, demonstrated that the carriers of a mutant allele of the CYP2C19 gene had 32.4% lower plasma levels of clopidogrel and 25% higher platelet aggregation as compared with noncarriers [18]. Similar findings were reported by Kelly et al. [27], who studied 90 healthy young Chinese subjects. We lack big data in an older population with high cardiovascular burden. Only Wang et al. reported that the genetic polymorphisms of CYP2C19*2 and ABCB1 affect the pharmacokinetic and pharmacodynamic responses to clopidogrel in patients with ACS [14]. More convincing data refer only to the effect of the CYP2C19 polymorphism on the clinical course of ACS in patients receiving clopidogrel. However, there is no information on how the polymorphism affects drug concentrations and platelet aggregation. In the TRITON-TIMI 38 trial (Trial to Assess Improvement in Therapeutic Outcomes by Optimizing Platelet Inhibition with Prasugrel–Thrombolysis in Myocardial Infarction 38) the carriers of the CYP2C19 reduced-function allele had an increased risk of ischemic adverse events [18]. Similarly, studies reported that CYP2C19 reduced-function alleles were associated with worse clinical outcomes, including cardiovascular death, myocardial infarction, and stroke [2831]. This observation has been confirmed by several meta-analyses [3234].

Importantly, also in our research healthy volunteers without CYP2C19*2 and ABCB1 polymorphisms had higher levels of clopidogrel active metabolite (p < 0.001) and reduced platelet aggregation (p < 0.05), as compared with ACS patients (in press). This may suggest that the effect of genetic polymorphism on drug concentrations and platelet aggregation is much more notable in younger and more homogeneous populations, while other factors, such as age, diabetes, renal insufficiency, and the use of multiple medications, including other cytochrome p450-metabolized drugs [18], have a significantly greater effect on reduced bioavailability and antiplatelet effects of clopidogrel in older and more heterogeneous groups, thus masking the possible influence of genetic variability. This may partially explain why prospective randomized trials have failed to demonstrate the efficacy of personalized antiplatelet treatment based on platelet function in reducing the frequency of ischemic events [3539].

Our study has several limitations. First, we included a relatively small cohort of participants from a limited geographic region. We have to admit that the study’s sample was not powered sufficiently to formulate final conclusions, and thus our findings need to be corroborated with more statistically powerful studies. Second, we did not assess other genetic polymorphisms or other factors (diabetes, smoking, drug use) that might have affected the drug concentrations. Moreover, considering that the we did not find any CYP2C19*3 allele carriers, and homozygous carriers of the CYP2C19*2 allele constituted only 1.5% of our study population [8], it was difficult to assess the effect of this genetic variant on the analyzed parameters. Therefore, the relationship between these genotypes and exposure to the active metabolite of clopidogrel and the corresponding pharmacodynamic effects should be confirmed in larger cohorts.

Conclusions

Our study showed that ACS is not associated with the levels of clopidogrel active metabolite or platelet inhibition in patients with Killip class I-III, and this lack of association was not affected by the presence of ABCB1 and CYP2C19*2 gene polymorphisms. This suggests that clopidogrel may be safely used in this population in the presence of contraindications to prasugrel or ticagrelor treatment.

Conflict of interest

The authors declare no conflict of interest.

References

1 

Gargiulo G, Goette A, Tijssen J, et al. , authors. Safety and efficacy outcomes of double vs. triple antithrombotic therapy in patients with atrial fibrillation following percutaneous coronary intervention: a systematic review and meta-analysis of non-vitamin K antagonist oral anticoagulant-based randomized clinical trials. Eur Heart J. 2019. 40:p. 3757–67

2 

Bellin A, Berto P, Themistoclakis S, et al. , authors. New oral anti-coagulants versus vitamin K antagonists in high thromboembolic risk patients. PLoS One. 2019. 14:p. e0222762

3 

Protasiewicz M, Szymkiewicz P, Kuliczkowski W, et al. , authors. Modern antiplatelet therapy – opportunities and risks. Adv Clin Exp Med. 2013. 22:p. 875–85

4 

Saiz-Rodríguez M, Belmonte C, Caniego JL, et al. , authors. Influence of CYP450 ezymes, CES1, PON1, ABCB1, and P2RY12 polymorphisms on clopidogrel response in patients subjected to a percutaneous neurointervention. Clin Ther. 2019. 41:p. 1199–212.e2

5 

Schoergenhofer C, Hobl EL, Schellongowski P, et al. , authors. Clopidogrel in critically ill patients. Clin Pharmacol Ther. 2018. 103:p. 217–23

6 

Součková L, Opatřilová R, Suk P, et al. , authors. Impaired bioavailability and antiplatelet effect of high-dose clopidogrel in patients after cardiopulmonary resuscitation (CPR). Eur J Clin Pharmacol. 2013. 69:p. 309–17

7 

Wójcik T, Szymkiewicz p, Wiśniewski J, et al. , authors. Distribution of polymorphisms in the CYP2C19 and ABCB1 genes among patients with acute coronary syndrome in Lower Silesian population. Adv Clin Exp Med. 2019. 28:p. 1621–6

8 

Takahashi M, Pang H, Kawabata K, et al. , authors. Quantitative determination of clopidogrel active metabolite in human plasma by LC-MS/MS. J Pharm Biomed Anal. 2008. 48:p. 1219–24

9 

Peer CJ, Spencer SD, VanDenBerg DA, et al. , authors. A sensitive and rapid ultra HPLC-MS/MS method for the simultaneous detection of clopidogrel and its derivatized active thiol metabolite in human plasma. J Chromatogr B Analyt Technol Biomed Life Sci. 2012. 880:p. 132–9

10 

Karaźniewicz-Łada M, Danielak D, Teżyk A, et al. , authors. HPLC-MS/MS method for the simultaneous determination of clopidogrel, its carboxylic acid metabolite and derivatized isomers of thiol metabolite in clinical samples. J Chromatogr B Analyt Technol Biomed Life Sci. 2012. 911:p. 105–12

11 

Bonello L, Tantry US, Marcucci R, et al. , authors. Consensus and future directions on the definition of high on-treatment platelet reactivity to adenosine diphosphate. J Am Coll Cardiol. 2010. 56:p. 919–33

12 

Cuisset T, Frere C, Quilici J, et al. , authors. Predictive values of post-treatment adenosine diphosphate-induced aggregation and vasodilator-stimulated phosphoprotein index for stent thrombosis after acute coronary syndrome in clopidogrel-treated patients. Am J Cardiol. 2009. 104:p. 1078–82

13 

Laine M, Paganelli F, Bonello L , authors. P2Y12-ADP receptor antagonists: days of future and past. World J Cardiol. 2016. 8:p. 327–32

14 

Wang XQ, Shen CL, Wang BN, et al. , authors. Genetic polymorphisms of CYP2C19*2 and ABCB1 C3435T affect the pharmacokinetic and pharmacodynamic responses to clopidogrel in 401 patients with acute coronary syndrome. Gene. 2015. 558:p. 200–7

15 

Horenstein RB, Madabushi R, Zineh I, et al. , authors. Effectiveness of clopidogrel dose escalation to normalize active metabolite exposure and antiplatelet effects in CYP2C19 poor metabolizers. J Clin Pharmacol. 2014. 54:p. 865–73

16 

Warlo E, Arnesen H, Seljeflot I , authors. A brief review on resistance to P2Y12 receptor antagonism in coronary artery disease. Thromb J. 2019. 17:p. 11

17 

Aradi D, Storey RF, Komocsi A, et al. , authors. Expert position paper on the role of platelet function testing in patients undergoing percutaneous coronary intervention. Eur Heart J. 2014. 35:p. 209–15

18 

Mega JL, Close SL, Wiviott SD, et al. , authors. Cytochrome p-450 polymorphisms and response to clopidogrel. N Engl J Med. 2009. 360:p. 354–62

19 

Desta Z, Zhao X, Shin JG, Flockhart DA , authors. Clinical significance of the cytochrome P450 2C19 genetic polymorphism. Clin Pharmacokinet. 2002. 41:p. 913–58

20 

Jiang XL, Samant S, Lesko LJ, Schmidt S , authors. Clinical pharmacokinetics and pharmacodynamics of clopidogrel. Clin Pharmacokinet. 2015. 54:p. 147–66

21 

Kimchi-Sarfaty C, Oh JM, Kim IW, et al. , authors. A “silent” polymorphism in the MDR1 gene changes substrate specificity. Science. 2007. 315:p. 525–8

22 

Taubert D, von Beckerath N, Grimberg G, et al. , authors. Impact of P-glycoprotein on clopidogrel absorption. Clin Pharmacol Ther. 2006. 80:p. 486–501

23 

Simon T, Verstuyft C, Mary-Krause M, et al. French Registry of Acute ST-Elevation and Non-ST-Elevation Myocardial Infarction (FAST-MI) Investigators , authors. Genetic determinants of response to clopidogrel and cardiovascular events. N Engl J Med. 2009. 360:p. 363–75

24 

Frelinger AL 3rd, Bhatt DL, Lee RD, et al. , authors. Clopidogrel pharmacokinetics and pharmacodynamics vary widely despite exclusion or control of polymorphisms (CYP2C19, ABCB1, PON1), noncompliance, diet, smoking, co-medications (including proton pump inhibitors), and pre-existent variability in platelet function. J Am Coll Cardiol. 2013. 61:p. 872–9

25 

Roe MT, Goodman SG, Ohman EM, et al. , authors. Elderly patients with acute coronary syndromes managed without revascularization insights into the safety of long-term dual antiplatelet therapy with reduced-dose prasugrel vs. standard-dose clopidogrel. Circulation. 2013. 128:p. 823–33

26 

Su J, Xu J, Li X, et al. , authors. ABCB1 C3435T polymorphism and response to clopidogrel treatment in coronary artery disease (CAD) patients: a meta-analysis. PLoS One. 2012. 7:p. e46366

27 

Kelly RP, Close SL, Farid NA, et al. , authors. Pharmacokinetics and pharmacodynamics following maintenance doses of prasugrel and clopidogrel in Chinese carriers of CYP2C19 variants. Br J Clin Pharmacol. 2012. 73:p. 93–105

28 

Delaney JT, Ramirez AH, Bowton E, et al. , authors. Predicting clopidogrel response using DNA samples linked to an electronic health record. Clin Pharmacol Ther. 2012. 91:p. 257–63

29 

Collet JP, Hulot JS, Pena A, et al. , authors. Cytochrome P450 2C19 polymorphism in young patients treated with clopidogrel after myocardial infarction: a cohort study. Lancet. 2009. 373:p. 309–17

30 

Giusti B, Gori AM, Marcucci R, et al. , authors. Cytochrome P450 2C19 loss-of-function polymorphism, but not CYP3A4 IVS10 + 12G/A and P2Y12 T744C polymorphisms, is associated with response variability to dual antiplatelet treatment in high-risk vascular patients. Pharmacogenet Genomics. 2007. 17:p. 1057–64

31 

Sibbing D, Koch W, Massberg S, et al. , authors. No association of paraoxonase-1 Q192R genotypes with platelet response to clopidogrel and risk of stent thrombosis after coronary stenting. Eur Heart J. 2011. 32:p. 1605–13

32 

Zabalza M, Subirana I, Sala J, et al. , authors. Meta-analyses of the association between cytochrome CYP2C19 loss- and gain-of-function polymorphisms and cardiovascular outcomes in patients with coronary artery disease treated with clopidogrel. Heart. 2012. 98:p. 100–8

33 

Sorich MJ, Polasek TM, Wiese MD , authors. Challenges and limitations in the interpretation of systematic reviews: making sense of clopidogrel and CYP2C19 pharmacogenetics. Clin Pharmacol Ther. 2013. 94:p. 376–82

34 

Mega JL, Simon T, Collet JP, et al. , authors. Reduced-function CYP2C19 genotype and risk of adverse clinical outcomes among patients treated with clopidogrel predominantly for PCI: a meta-analysis. JAMA. 2010. 304:p. 1821–30

35 

Price MJ, Berger PB, Teirstein PS, et al. GRAVITAS Investigators , authors. Standard- vs. high-dose clopidogrel based on platelet function testing after percutaneous coronary intervention: the GRAVITAS randomized trial. JAMA. 2011. 305:p. 1097–105

36 

Collet JP, Cuisset T, Rangé G, et al. ARCTIC Investigators , authors. Bedside monitoring to adjust antiplatelet therapy for coronary stenting. N Engl J Med. 2012. 367:p. 2100–9

37 

Trenk D, Stone GW, Gawaz M, et al. , authors. A randomized trial of prasugrel versus clopidogrel in patients with high platelet reactivity on clopidogrel after elective percutaneous coronary intervention with implantation of drug-eluting stents: results of the TRIGGER-PCI (Testing Platelet Reactivity In Patients Undergoing Elective Stent Placement on Clopidogrel to Guide Alternative Therapy With Prasugrel) study. J Am Coll Cardiol. 2012. 59:p. 2159–64

38 

Cayla G, Cuisset T, Silvain J, et al. ANTARCTIC investigators , authors. Platelet function monitoring to adjust antiplatelet therapy in elderly patients stented for an acute coronary syndrome (ANTARCTIC): an open-label, blinded-endpoint, randomised controlled superiority trial. Lancet. 2016. 388:p. 2015–22

39 

Sibbing D, Aradi D, Jacobshagen C, et al. TROPICAL-ACS Investigators , authors. Guided de-escalation of antiplatelet treatment in patients with acute coronary syndrome undergoing percutaneous coronary intervention (TROPICAL-ACS): a randomised, open-label, multicentre trial. Lancet. 2017. 390:p. 1747–57

Copyright: © 2021 Termedia Sp. z o. o. This is an Open Access article distributed under the terms of the Creative Commons Attribution-NonCommercial-ShareAlike 4.0 International (CC BY-NC-SA 4.0) License (http://creativecommons.org/licenses/by-nc-sa/4.0/), allowing third parties to copy and redistribute the material in any medium or format and to remix, transform, and build upon the material, provided the original work is properly cited and states its license.
 
Quick links
© 2024 Termedia Sp. z o.o.
Developed by Bentus.