Open access peer-reviewed chapter

Pharmacogenetics of Cardiovascular Disease: Genetic Variation and Statin Intolerance

Written By

Jana Petrkova, Milos Taborsky and Martin Petrek

Submitted: 05 December 2017 Reviewed: 14 June 2018 Published: 05 November 2018

DOI: 10.5772/intechopen.79518

From the Edited Volume

Genetic Diversity and Disease Susceptibility

Edited by Yamin Liu

Chapter metrics overview

1,518 Chapter Downloads

View Full Metrics

Abstract

Statins are very effective for lowering low-density lipoprotein cholesterol for primary and secondary cardiovascular disease prevention. While statins are usually well tolerated, individual response to statin therapy varies and intolerance, predominantly muscle symptoms, may appear in a significant proportion of patients. Besides clinical factors, variation in genes coding for proteins with drug transporting, immune or enzymatic function have been implicated in the pathogenesis of statin intolerance. In this review, we will characterise the candidate gene variants for development of statin intolerance, describe their population distribution and summarise current knowledge on their biological plausibility. Clinical relevance and current guidelines/recommendations will be also discussed.

Keywords

  • genetic variation
  • pharmacogenetics
  • SLCO1B1
  • statin
  • statin-induced myopathy

1. Introduction

Statins (3-hydroxy-3-methylglutaryl coenzyme A reductase) inhibitors are highly effective drugs lowering plasmatic concentration of LDL-C cholesterol by 30–50% [1]. Despite the fact that they are usually considered safe and very well tolerated, a significant proportion of the treated patients does not tolerate the drug: they suffer from side effects, which may result in non-compliance of patients, drug dose-lowering and even discontinuation of therapy [2, 3, 4]. Undesirable effects of statins restrict their administration or reaching LDL-C cholesterol target values and limits effective treatment of patients at risk. Non-adherence or discontinuation of therapy is associated with an increased risk of cardiovascular events [5, 6, 7].

By analogy to individual nature of patients´ response to treatment [8, 9, 10] there are also inter-individual differences in occurrence and extent of statin intolerance and its symptoms. The knowledge of the risk factors predisposing for intolerance development including characteristic genetic background is crucial for its understanding and prevention. In this chapter we will review the polymorphic gene variants implicated in development of statin intolerance, briefly describe their biological plausibility and characterise clinical relevance.

Advertisement

2. Statin intolerance

Statin intolerance is the inability to tolerate sufficient dose of statin needed to reduce cardiovascular risk due to side effects or intolerance to treatment [11]. The most frequent are muscle symptoms characterised bellow.

2.1. Statin-associated muscle symptoms

Statin-induced muscle symptoms range from myalgia to mild or severe myopathy and even to rare rhabdomyolysis [12]. The symptoms appear in about 75% in the first 10–12 weeks and in 90% of cases in the first 6 months after treatment initiation or dose up titration [13]. The true frequency muscle related side effects has been widely debated: while an observational study reported as much as about 20% of patients on statins [14], clinical trial data suggests frequencies to be equal or lower than 5% [15], however there was a study reporting that clinical trials did not use a standard definition for statin myalgia [16], which may result in underestimated occurrence of statin-induced muscle symptoms. In any case, given very high usage of statins (the third most frequently prescribed drug), even lower relative frequency numbers would mean substantial absolute number of symptomatic patients.

2.2. Clinical-related risk factors

The available data shows that the side effects of statin therapy are group-dependent, time-dependent and dose-dependent; their frequency is greater at a higher statin dose [17].

Endogenous factors known to increase occurrence of side effects are as follows: another lipid-lowering therapy, alcohol abuse, surgery, heavy exercise. Importantly, interactions with medication may be serious [18]; particularly drug interactions likely contribute the susceptibility to statin related muscle symptoms [19].

Further factors predisposing to statin intolerance are: advanced age (>70 year), female sex, race/ethnicity, family history of muscle disorders, vitamin D deficiency, history of creatine elevation, hepatic and renal impartment, hypothyroidism, low body mass index [20].

2.3. Genetic factors

Besides the above characterised factors, genetic “make-up” of a given patient is important component in susceptibility to statin intolerance. Indeed, genetic variation represents the major factor responsible for inter-individual differences in patient responsiveness and their inclination towards undesirable side effects of statins.

Advertisement

3. Genes responsible for statin intolerance

The following section characterises the gene variants that have been implicated in mechanisms of statin intolerance represented by statin-induced myopathy. These are listed in the Table 1.

GeneChromosomeAllelers numberCoding variationStudy
SLCO1B112p12.2*5 ≠rs4149056521 T > C[21]
CYP2D622q13.1*3rs357426862549delA[22]
*4rs3892097splicing defect, G > A
*5gene deletion
CYP3A47q21.1*1Brs2740574-392A > G transition[23]
GATM15q15.3rs9806699G > A, cis-e QTL[24]
LILRB519q13.4rs12975366T > C: Asp247Gly[25]
COQ24q21.22-q21.23rs6335454synonymous[26]
rs4693075non-coding

Table 1.

Gene and their variants implicated in development of statin intolerance presented as statin myopathy.

Note: rs, reference sequence; ≠ denotes haplotype (not allele) designation, see Section 4, second paragraph.

3.1. SLCO1B1 gene

SLCO1B1 gene encodes the OATP1B1 (organic anion transporting polypeptide), which has been reported to regulate the hepatic uptakes of statins [27, 28]. Strong support for its nomination as a risk factor for statin intolerance came from the GWAS study which investigated genetic variation in 85 subject with myopathy and 90 controls, all taking 80 mg of simvastatin [21]: strong association was identified between statin-induced myopathy and single nucleotide polymorphism (SNP) rs4363657 located within the SLCO1B1 gene. This noncoding SNP was in nearly complete linkage disequilibrium with the nonsynonymous rs4149056 SNP variant, which has been linked to statin metabolism: the odds ratio for myopathy was 4.5% per one copy of the C allele and 16.9% in CC homozygotes compared with homozygotes for standard allele (TT). More than 60% of observed myopathy cases could be attributed to this particular genetic variation, [21], which is also due to its relatively high population prevalence - rs4149056 C allele frequency is 15%.

3.2. LILRB5 gene

A potential role for immune system genetic variation in development of statin-induced myopathy has been recently reported for a variant in leukocyte immunoglobulin-like receptor subfamily-B, LILRB5 gene (rs12975366:T > C:Asp247Gly) [25]. The missense variant Asp247Gly has been associated with serum creatine kinase (CK) levels; the mean levels of this enzyme were elevated in Asp247 homozygotes (TT). The LILRB5 Asp247 homozygous genotype has, therefore, been associated with increased risk of statin intolerance [25]. No independent replication data on this plausible new variant has been available so far.

3.3. GATM gene

Glycine amidinotransferase, GATM gene encodes a mitochondrial enzyme, which is involved in creatine biosynthesis. SNP rs9806699 within the GATM gene has been associated with statin induced myopathy, specifically minor allele A conferring a protective effect and reduced risk of myopathy [24]. However, as this result was not replicated [29], further investigations are required before a possible role for this variation in statin tolerance is clarified.

3.4. Family of cytochrome P450 genes

The cytochrome P450 family is a group of izoenzymes important for catalysing oxidation of xenobiotics. There is a wide spectrum of polymorphic variants affecting various pharmacogenetics aspects. Regarding cardiovascular setting, CYP gene variation plays role in warfarin and clopidogrel metabolism with clear clinical relevance (e.g. [30]). In context of statin adverse drug reaction, Mulder et al. [22] reported higher incidence of statin intolerance in the group of patients who carried two of the less effective CYP2D6 *3,*4,*5 alleles. Regarding another gene within cytochrome P450 system, namely CYP3A5, an association was observed between nonfunctional CYP3A5*3 allele and the magnitude of CK elevation in case of patients experiencing myalgia during atorvastatin treatment [23]. Importantly, patients who develop myalgia while taking atorvastatin were more likely to experience a greater degree of muscle damage if they express two copies of CYP3A5*3.

3.5. Other plausible gene variants

COQ2 gene encodes Coenzyme Q2, involved in synthesis of ubiquinon (Coenzyme Q10, CoQ10), a redox carrier in the mitochondrial respiratory chain and a lipid-soluble antioxidant. Two variants within the COQ2 gene (Table 1) have been associated with increased odds of statin intolerance, defined primarily through muscle symptomatology [26]. This observation has been subsequently replicated [31].

From other molecules functioning as drug transporters, ABCB1 gene variation may also participate in development of statin muscle symptoms. This gene encodes the P-glycoprotein, an independent efflux pump. From its variants, the 1236 T, 2677non-G, and 3435 T alleles were less frequent in cases undergoing statin therapy than in the control group [32]. The authors also demonstrated a reduced T-non-G-T haplotype frequency (20.0%) in patients in whom myalgia developed during simvastatin treatment, as compared with the control, non-myalgia group (41.4%).

Most recently, a variant of a UGT1 gene coding for uridine diphosphate glucuronosyltransferase, specifically UGT1A1*28 variant allele (rs8175347), was reported to possess plausible protective effect in development of statin intolerance [33], however again this finding must be replicated.

In the following text we will concentrate on the SLCO1B1 gene variation and describe its population distribution and clinical relevance. The reason for our focus is that to date, the rs4149056 SLCO1B1 variant has been repeatedly evidenced to possess the strongest effect in response to statin therapy.

Advertisement

4. Genetic variability and population distribution of SLCO1B1

More than 45 nonsynonymous variants in SLCO1B1 gene have been identified [34]. Some of the variants have altered function [35]. Genotypic frequencies of SLCO1B1 variants depend on ethnicity, and genetic difference between populations correlated with the geographical distances [34, 36, 37]. In particular, single nucleotide polymorphism the 521 T > C (rs4149056) appeared more commonly in European-Americans while it was less frequent in African-Americans. In opposite, single nucleotide polymorphism the 388A > G (rs2306283) was detected predominantly in African-Americans. Pasanen et al. [38] investigated the frequencies of 12 SNPs in SLCO1B1 in 941 persons from 52 populations across Europe, Asia, Africa, Middle East, Oceania and the Americas (Amerindians).

SLCO1B1 single nucleotide polymorphisms 521 T > C and 388A > G form four haplotypes: *1A (388A/521 T), *1B (388G/521 T), *5 (388A/521 T) and *15 (388G/521C) [38, 39]. The low activity haplotypes–*5 (388A/521C) and *15 (388G/521C) occur with combined haplotype frequency of approximately 15–20% in Europeans, 10–15% in Asians, 2% in sub-Saharian Africans. The *1B (388G/521 T) haplotype occurs in approximately 26% Europeans, in 39–63% Asians and in 77% sub-Saharian Africans. The haplotypes *5 and *15 are associated with significant reductions of statin hepatic uptake [40], resulting in increase of systemic substrates exposure.

Advertisement

5. Clinical relevance of the variation in the SCLO1B1 gene

Clinical relevance of the SCLO1B variation is based on biological role of its gene products in hepatic transport of statins. Statins are mainly delivered within hepatocytes to their site of actions by uptake transporters and eliminated into the bile by eflux transporters [41]. Many statins are substrates of hepatic uptake transporters including OATP1B1, OATP2B1 and OATP1B3 [28] with OATB1B1 as the main one. The loss of function the SLCO1B1*5 (Val174Ala, 521 T > C, rs4149056), located in exon 5, downregulates OATB1A1 transporter cell membrane and protein expression [42] which leads to decreased hepatic uptake, greater systemic statin plasma concentrations, and therefore greater muscle statin exposure, all these resulting in adverse effects [20, 36, 43, 44, 45].

Importantly, the impact of the rs4149056 variant on statin metabolism appears to differ between distinct statins. The effect of rs4149056 genotypes was much greater for simvastatin, less for atorvastatin and rosuvastatin in healthy volunteers [46, 47]: For simvastatin the area under curve, AUC (0-infinity) was increased by 221% in genotype CC individuals in compared with wild-type TT individuals. For atorvastatin this parameter was increased by 145% and for rosuvastatin by 62%. Individuals carrying C allele also reached maximum concentration (Cmax) earlier, and its value was 200% higher compared with TT individuals of rs4149056 [47]. Further, the rs4149056 polymorphism was significantly associated with simvastatin treatment cases of severe statin induced myopathy, which did not occur after atorvastatin [45] or pravastatin [48] treatment. Similar conclusions regarding simvastatin were obtained in animal model, again for simvastatin however not for pravastatin or atorvastatin [49]. Table 2 lists main studies investigating effect of simvastatin on statin intolerance/myopathy in humans.

StudySimvastatin doseOutcomeReference
SEARCH80 mgOR 4.5 per C allele[21]
OR 16.9 for CC homozygotes
Heart protection study40 mgOR 2.6 per C allele[21]
Brunham10-80 mgOR 2.3 per C allele[45]

Table 2.

Basic studies that have reported a strong association between the rs4149056 single nucleotide polymorphism and simvastatin induced myopathy.

Note: OR, odds ratio.

It is crucial that in agreement with the rules for performing association studies [50] the data obtained in the SEARCH study has been independently replicated within the Heart Protection Study, in 10,000 patients who received 40 mg Simvastatin [21]. The meaningful data obtained in this way provided starting point for reflection of the observations from genetic and pharmacokinetic studies into clinical practice, including formulation of treatment recommendations which will be subject of the next section of our chapter.

Advertisement

6. Testing for statin intolerance in clinic–current status and treatment recommendations

The spectrum of evidence supporting the association between the lead SNP rs4149056 and statin, namely simvastatin-induced myopathy prompted application of the SLCO1B1 genotyping for clinical usage. This translation to diagnostics aims mainly at reducing risk of simvastatin induced muscle toxicity and at increased adherence to therapy [43, 51, 52]. Another possible outcome of genotyping is the option to use alternate agents of the statin class.

In clinical practice, the adverse effect of SLCO1B1 polymorphism depends on the genotype (being highest in homozygotes), statin dose and statin type. This has been reflected in the guidance for prescribers provided primarily by the Clinical Pharmacogenetics Implementation Consortium (CPIC); the working group produces guidelines for simvastatin use in individual carriers risk allele in SLCO1B1 gene [43]; the guidelines have been recently updated [53]. In patients with one or two copies of SLCO1B1 rs4149056 C allele, simvastatin should be avoided or reduced dosage should be considered, pravastatin or rosuvastatin are preferred alternatives according to the CPIC guidelines, however other clinical and patients specific factors should be taken into account [43, 53].

KEY MESSAGE
  • Statins are the basis of dyslipidemia treatment in patients with increased cardiovascular risk.

  • Genetic factors may predispose to statin intolerance.

  • SCLO1B1 gene variation has strong association with statin-induced myopathy.

  • Guidance for prescribers reflects individual risk for SLCO1B1 rs4149056 C allele carriers.

  • Investigations of clinical relevance of other plausible gene variants are ongoing.

Apart from this general, non-compulsory guidance, there have been more systematic efforts to apply SLCO1B1 genotyping into practice. This direction is represented e.g. by pre-emptive programs performed from the initiative by U.S. Pharmacogenetics Research Network at eight sites [54]. Similarly, aiming of introducing genotype-guided prescribing, pre-emptive genotyping has also been investigated in Europe by the EU Horizon 2020-funded Ubiquitous Pharmacogenomics (U-PGx) Consortium (http://upgx.eu) in seven European countries [55]. Last but not least, recommendations were formulated also on a national level - in France: the French National Network of Pharmacogenetics (RNPGx) [56] is in favour of rs4149056 testing before starting therapy or early after treatment onset in patients with one or more risk factors. If the genotype is not known early, the RNPGx considers that a polymorphism test is potentially useful also in the event of already occurring muscle toxicity in patients treated with statins, in order to rule out or confirm a genetic cause. From the above examples it is clear that pharmacogenetic genotyping for prediction/confirmation of statin intolerance undergoes ongoing development and progress; further updates of the recommendations are expected. It should be noted that there have been opinions as well that the current status of knowledge has not been yet sufficient to allow clinical application of genotyping for risk of statin intolerance [52]. There have been several arguments, however especially those economical (“the tests are too costly”) are not substantiated; some “con” opinions have been also “traditionalistic”, from conservative point of view on doubting any new test or medical management measure including pharmacogenetics. However, this reluctant or at least “sceptical” attitude about pharmacogenetic contribution to routine statin usage, well known also from other applications of pharmacogenetics, has been gradually changing–it only takes time, systematic information on the evidence and particularly education to overcome it [57].

Advertisement

7. Future perspectives including economic aspects of genetic test for statin intolerance

Implementation of a genetic test for statin intolerance into routine practice definitely requires analysing its benefits not only for patients but also for health care providers. In this context, pharmacoecomic data on genetic testing statin intolerance have been scarce. The existing literature on cost-effectiveness of pharmacogenetic testing has been either general [58] or described economic savings solely due to hypolipidemic effect of statins [59]. The first specific data for statin intolerance and its genetic testing appeared only very recently [60]–the authors estimated 356 Canadian dollars as the cost limit for economic feasibility and at the same time dominant health effect for cardiovascular prevention. In extension of this very first report [60], this topic should be, therefore, addressed more intensively and also from other angles in the future. This has been the case with other pharmacogenetic applications (e.g. [61]), it will be also innovative to use new approaches which utilise alternative parameters for assessing effectiveness (e.g. [62, 63, 64]).

Though important, inclusion of economic criteria is the only one part of the future priorities in the field of application of genetic variation for testing statin intolerance. Other avenues for future may address (1) further search for and verification of other genetic markers than SLCO1B1 including providing pharmacokinetic data [65, 66], (2) reflection of ethnic differences in distribution of genetic markers between populations [64, 67], (3) inclusion of the results of genetic test into electronic medical records [68], (4) performing meta-analyses of studies reported so far, and last but not least, (5) performance of well-designed clinical studies implementing also other non-genetic criteria in order to propose a risk-score or clinically applicable algorithm. The existing examples from other pharmacogenetic applications (e.g. [69]) and above described initiatives such as U-PGX [55], RNPGX [56], or the recent idea to provide patients with their DNA (pharmacogenetic) passport [70], allow us to expect further developments targeted at patient benefit and innovation of medical care.

Advertisement

Acknowledgments

This work has been supported from the project LO1304.

Advertisement

Conflict of interest

The authors do not report any conflict of interest.

Advertisement

Thanks

The authors would like to dedicate this chapter to late Dr. Adéla Bártová (formerly of Dept. of Internal Medicine and HLA laboratory, Olomouc University Hospital) thanking her for her continuous support of their professional activities.

References

  1. 1. Cholesterol Treatment Trialists’ (CTT) Collaboration, Baigent C, Blackwell L, Emberson J, Holland LE, Reith C, Bhala N, Peto R, Barnes EH, Keech A, Simes J, Collins R. Efficacy and safety of more intensive lowering of LDL cholesterol: A meta-analysis of data from 170 000 participants in 26 randomised trials. The Lancet. 2010;376(9753):1670-1681. DOI: 10.1016/S0140-6736(10)61350-5
  2. 2. Zhang H, Plutzky J, Shubina M, Turchin A. Continued statin prescriptions after adverse reactions and patient outcomes. Annals of Internal Medicine. 2017;167(4):221-227. DOI: 10.7326/M16-0838
  3. 3. Mancini GBJ, Baker S, Bergeron J, Fitchett D, Frohlich J, Genest J, Gupta M, Hegele RA, Ng D, Pearson GJ, Pope J, Tashakkor AY. Diagnosis, prevention, and management of statin adverse effects and intolerance: Canadian consensus working group update (2016). Canadian Journal of Cardiology. 2016;32(7):S35-S65. DOI: 10.1016/j.cjca.2016.01.003
  4. 4. Nissen SE. Statin intolerance: An elusive but morbid disorder. Journal of the American College of Cardiology. 2017;69(11):1396-1398. DOI: 10.1016/j.jacc.2017.01.019
  5. 5. Serban M-C, Colantonio LD, Manthripragada AD, Monda KL, Bittner VA, Banach M, Chen L, Huang L, Dent R, Kent ST, Muntner P, Rosenson RS. Statin intolerance and risk of coronary heart events and all-cause mortality following myocardial infarction. Journal of the American College of Cardiology. 2017;69(11):1386-1395. DOI: 10.1016/j.jacc.2016.12.036
  6. 6. Reiner Ž, De Backer G, Fras Z, Kotseva K, Tokgözoglu L, Wood D, De Bacquer D, EUROASPIRE Investigators. Lipid lowering drug therapy in patients with coronary heart disease from 24 European countries – Findings from the EUROASPIRE IV survey. Atherosclerosis. 2016;246:243-250. DOI: 10.1016/j.atherosclerosis.2016.01.018
  7. 7. Colantonio LD, Huang L, Monda KL, Bittner V, Serban M-C, Taylor B, Brown TM, Glasser SP, Muntner P, Rosenson RS. Adherence to high-intensity statins following a myocardial infarction hospitalization among Medicare beneficiaries. JAMA Cardiology. 2017;2(8):890-895. DOI: 10.1001/jamacardio.2017.0911
  8. 8. Yip VLM, Hawcutt DB, Pirmohamed M. Pharmacogenetic markers of drug efficacy and toxicity. Clinical Pharmacology & Therapeutics. 2015;98(1):61-70. DOI: 10.1002/cpt.135
  9. 9. Petrek M. Personalized medicine in sarcoidosis. Current Opinion in Pulmonary Medicine. 2015;21(5):532-537. DOI: 10.1097/MCP.0000000000000194
  10. 10. Kitzmiller JP, Groen DK, Phelps MA, Sadee W. Pharmacogenomic testing: Relevance in medical practice: Why drugs work in some patients but not in others. Cleveland Clinic Journal of Medicine. 2011;78(4):243-257. DOI: 10.3949/ccjm.78a.10145
  11. 11. Banach M, Rizzo M, Toth PP, Farnier M, Davidson MH, Al-Rasadi K, Aronow WS, Athyros V, Djuric DM, Ezhov MV, Greenfield RS, Hovingh GK, Kostner K, Serban C, Lighezan D, Fras Z, Moriarty PM, Muntner P, Goudev A, Ceska R, Nicholls SJ, Broncel M, Nikolic D, Pella D, Puri R, Rysz J, Wong ND, Bajnok L, Jones SR, Ray KK, Mikhailidis DP. Position paper statin intolerance – An attempt at a unified definition. Position paper from an international lipid expert panel. Archives of Medical Science. 2015;11(1):1-23. DOI: 10.5114/aoms.2015.49807
  12. 12. Antons KA, Williams CD, Baker SK, Phillips PS. Clinical perspectives of statin-induced Rhabdomyolysis. The American Journal of Medicine. 2006;119(5):400-409. DOI: 10.1016/j.amjmed.2006.02.007
  13. 13. Jacobson TA. Toward “pain-free” statin prescribing: Clinical algorithm for diagnosis and Management of Myalgia. Mayo Clinic Proceedings. 2008;83(6):687-700. DOI: 10.4065/83.6.687
  14. 14. Cohen JD, Brinton EA, Ito MK, Jacobson TA. Understanding statin use in America and gaps in patient education (USAGE): An internet-based survey of 10,138 current and former statin users. Journal of Clinical Lipidology. 2012;6(3):208-215. DOI: 10.1016/j.jacl.2012.03.003
  15. 15. Thompson PD, Clarkson P, Karas RH. Statin-associated myopathy. Journal of the American Medical Association. 2003;289(13):1681-1690. DOI: 10.1001/jama.289.13.1681
  16. 16. Ganga HV, Slim HB, Thompson PD. A systematic review of statin-induced muscle problems in clinical trials. American Heart Journal. 2014;168(1):6-15. DOI: 10.1016/j.ahj.2014.03.019
  17. 17. Pedan A, Varasteh LT, Schneeweiss S. Analysis of factors associated with statin adherence in a hierarchical model considering physician, pharmacy, patient, and prescription characteristics. Journal of Managed Care Pharmacy. 2007;13(6):487-496. DOI: 10.18553/jmcp.2007.13.6.487
  18. 18. Mancini GBJ, Tashakkor AY, Baker S, Bergeron J, Fitchett D, Frohlich J, Genest J, Gupta M, Hegele RA, Ng DS, Pearson GJ, Pope J. Diagnosis, prevention, and management of statin adverse effects and intolerance: Canadian working group consensus update. Canadian Journal of Cardiology. 2013;29(12):1553-1568. DOI: 10.1016/j.cjca.2013.09.023
  19. 19. Tomita Y, Maeda K, Sugiyama Y. Ethnic variability in the plasma exposures of OATP1B1 substrates such as HMG-CoA reductase inhibitors: A kinetic consideration of its mechanism. Clinical Pharmacology & Therapeutics. 2012;94(1):37-51. DOI: 10.1038/clpt.2012.221
  20. 20. Stroes ES, Thompson PD, Corsini A, Vladutiu GD, Raal FJ, Ray KK, Roden M, Stein E, Tokgözoğlu L, Nordestgaard BG, Bruckert E, De Backer G, Krauss RM, Laufs U, Santos RD, Hegele RA, Hovingh GK, Leiter LA, Mach F, März W, Newman CB, Wiklund O, Jacobson TA, Catapano AL, Chapman MJ, Ginsberg HN. European atherosclerosis society consensus panel. Statin-associated muscle symptoms: Impact on statin therapy—European atherosclerosis society consensus panel statement on assessment, Aetiology and management. European Heart Journal. 2015;36(17):1012-1022. DOI: 10.1093/eurheartj/ehv043
  21. 21. SEARCH Collaborative Group, Link E, Parish S, Armitage J, Bowman L, Heath S, Matsuda F, Gut I, Lathrop M, Collins R. SLCO1B1 variants and statin-induced myopathy — A genomewide study. New England Journal of Medicine. 2008;359(8):789-799. DOI: 10.1056/NEJMoa0801936
  22. 22. Mulder AB, van Lijf HJ, Bon MA, van den Bergh FA, Touw DJ, Neef C, Vermes I. Association of polymorphism in the cytochrome CYP2D6 and the efficacy and tolerability of simvastatin. Clinical Pharmacology & Therapeutics. 2001;70(6):546-551. DOI: 10.1067/mcp.2001.120251
  23. 23. Wilke RA, Moore JH, Burmester JK. Relative impact of CYP3A genotype and concomitant medication on the severity of atorvastatin-induced muscle damage. Pharmacogenetics and Genomics. 2005;15(6):415-421
  24. 24. Mangravite LM, Engelhardt BE, Medina MW, Smith JD, Brown CD, Chasman DI, Mecham BH, Howie B, Shim H, Naidoo D, Feng QP, Rieder MJ, Chen Y-DI, Rotter JI, Ridker PM, Hopewell JC, Parish S, Armitage J, Collins R, Wilke RA, Nickerson DA, Stephens M, Krauss RM. A statin-dependent QTL for GATM expression is associated with statin-induced myopathy. Nature. 2013;502(7471):377-380. DOI: 10.1038/nature12508
  25. 25. Siddiqui MK, Maroteau C, Veluchamy A, Tornio A, Tavendale R, Carr F, Abelega N-U, Carr D, Bloch K, Hallberg P, Yue Q-Y, Pearson ER, Colhoun HM, Morris AD, Dow E, George J, Pirmohamed M, Ridker PM, Doney ASF, Alfirevic A, Wadelius M, Maitland-van der Zee A-H, Chasman DI, Palmer CNA; PREDICTION-ADR Consortium. A common missense variant of LILRB5 is associated with statin intolerance and myalgia. European Heart Journal. 2017;38(48):3569-3575. DOI: 10.1093/eurheartj/ehx467
  26. 26. Oh J, Ban MR, Miskie BA, Pollex RL, Hegele RA. Genetic determinants of statin intolerance. Lipids in Health and Disease. 2007;6(1):7. DOI: 10.1186/1476-511X-6-7
  27. 27. Niemi M. Role of OATP transporters in the disposition of drugs. Pharmacogenomics. 2007;8(7):787-802. DOI: 10.2217/14622416.8.7.787
  28. 28. Shitara Y, Maeda K, Ikejiri K, Yoshida K, Horie T, Sugiyama Y. Clinical significance of organic anion transporting polypeptides (OATPs) in drug disposition: Their roles in hepatic clearance and intestinal absorption. Biopharmaceutics & Drug Disposition. 2013;34(1):45-78. DOI: 10.1002/bdd.1823
  29. 29. Luzum JA, Kitzmiller JP, Isackson PJ, Ma C, Medina MW, Dauki AM, Mikulik EB, Ochs-Balcom HM, Vladutiu GD. GATM polymorphism associated with the risk for statin-induced myopathy does not replicate in case-control analysis of 715 dyslipidemic individuals. Cell Metabolism. 2015;21(4):622-627. DOI: 10.1016/j.cmet.2015.03.003
  30. 30. Petrek M, Kocourkova L, Zizkova V, Nosek Z, Taborsky M, Petrkova J. Characterization of three CYP2C19 gene variants by MassARRAY and point of care techniques: Experience from a Czech Centre. Archivum Immunologiae et Therapiae Experimentalis. 2016;64(S1):99-107. DOI: 10.1007/s00005-016-0440-8
  31. 31. Ruaño G, Windemuth A, Wu AHB, Kane JP, Malloy MJ, Pullinger CR, Kocherla M, Bogaard K, Gordon BR, Holford TR, Gupta A, Seip RL, Thompson PD. Mechanisms of statin-induced myalgia assessed by physiogenomic associations. Atherosclerosis. 2011;218(2):451-456. DOI: 10.1016/j.atherosclerosis.2011.07.007
  32. 32. Fiegenbaum M, da Silveira FR, Van der Sand CR, Van der Sand LC, Ferreira ME, Pires RC, Hutz MH. The role of common variants of ABCB1, CYP3A4, and CYP3A5 genes in lipid-lowering efficacy and safety of simvastatin treatment. Clinical Pharmacology & Therapeutics. 2005;78(5):551-558. DOI: 10.1016/j.clpt.2005.08.003
  33. 33. Willrich MAV, Kaleta EJ, Bryant SC, Spears GM, Train LJ, Peterson SE, Lennon VA, Kopecky SL, Baudhuin LM. Genetic variation in statin intolerance and a possible protective role for UGT1A1. Pharmacogenomics. 2018;19(2):83-94. DOI: 10.2217/pgs-2017-0146
  34. 34. Niemi M, Pasanen MK, Neuvonen PJ. Organic anion transporting polypeptide 1B1: A genetically polymorphic transporter of major importance for hepatic drug uptake. Pharmacological Reviews. 2011;63(1):157-181. DOI: 10.1124/pr.110.002857
  35. 35. Maeda K, Sugiyama Y. Impact of genetic polymorphisms of transporters on the pharmacokinetic, pharmacodynamic and toxicological properties of anionic drugs. Drug Metabolism and Pharmacokinetics. 2008;23(4):223-235. DOI: 10.2133/dmpk.23.223
  36. 36. Lee HH, Ho RH. Interindividual and interethnic variability in drug disposition: Polymorphisms in organic anion transporting polypeptide 1B1 (OATP1B1; SLCO1B1). British Journal of Clinical Pharmacology. 2017;83(6):1176-1184. DOI: 10.1111/bcp.13207
  37. 37. Alghalyini B, El Shamieh S, Salami A, Visvikis Siest S, Fakhoury HM, Fakhoury R. Effect of SLCO1B1 gene polymorphisms and vitamin D on statin-induced myopathy. Drug Metabolism and Personalized Therapy. 2018;33(1):41-47. DOI: 10.1515/dmpt-2017-0030
  38. 38. Pasanen MK, Neuvonen PJ, Niemi M. Global analysis of genetic variation in SLCO1B1. Pharmacogenomics. 2008;9(1):19-33. DOI: 10.2217/14622416.9.1.19
  39. 39. Pasanen MK, Backman JT, Neuvonen PJ, Niemi M. Frequencies of single nucleotide polymorphisms and haplotypes of organic anion transporting polypeptide 1B1 SLCO1B1 gene in a Finnish population. European Journal of Clinical Pharmacology. 2006;62(6):409-415. DOI: 10.1007/s00228-006-0123-1
  40. 40. Kameyama Y, Yamashita K, Kobayashi K, Hosokawa M, Chiba K. Functional characterization of SLCO1B1 (OATP-C) variants, SLCO1B1*5, SLCO1B1*15 and SLCO1B1*15+C1007G, by using transient expression systems of HeLa and HEK293 cells. Pharmacogenetics and Genomics. 2005;15(7):513-522. DOI: 10.1097/01.fpc.0000170913.73780.5f
  41. 41. Shitara Y, Sugiyama Y. Pharmacokinetic and pharmacodynamic alterations of 3-hydroxy-3-methylglutaryl coenzyme a (HMG-CoA) reductase inhibitors: Drug–drug interactions and interindividual differences in transporter and metabolic enzyme functions. Pharmacology & Therapeutics. 2006;112(1):71-105. DOI: 10.1016/j.pharmthera.2006.03.003
  42. 42. Tirona RG, Leake BF, Merino G, Kim RB. Polymorphisms in OATP-C: Identification of multiple allelic variants associated with altered transport activity among European- and African-Americans. Journal of Biological Chemistry. 2001;276(38):35669-35675. DOI: 10.1074/jbc.M103792200
  43. 43. Ramsey LB, Johnson SG, Caudle KE, Haidar CE, Voora D, Wilke RA, Maxwell WD, McLeod HL, Krauss RM, Roden DM, Feng Q, Cooper-DeHoff RM, Gong L, Klein TE, Wadelius M, Niemi M. The clinical pharmacogenetics implementation consortium guideline for SLCO1B1 and simvastatin-induced myopathy. Clinical Pharmacology & Therapeutics. 2014 Update, 2014;96(4):423-428. DOI: 10.1038/clpt.2014.125
  44. 44. Oshiro C, Mangravite L, Klein T, Altman R. PharmGKB very important pharmacogene: SLCO1B1. Pharmacogenetics and Genomics. 2010;20(3):211-216. DOI: 10.1097/FPC.0b013e328333b99c
  45. 45. Brunham LR, Lansberg PJ, Zhang L, Miao F, Carter C, Hovingh GK, Visscher H, Jukema JW, Stalenhoef AF, Ross CJD, Carleton BC, Kastelein JJP, Hayden MR. Differential effect of the rs4149056 variant in SLCO1B1 on myopathy associated with simvastatin and atorvastatin. The Pharmacogenomics Journal. 2012;12(3):233-237. DOI: 10.1038/tpj.2010.92
  46. 46. Pasanen MK, Fredrikson H, Neuvonen PJ, Niemi M. Different effects of SLCO1B1 polymorphism on the pharmacokinetics of atorvastatin and rosuvastatin. Clinical Pharmacology & Therapeutics. 2007;82(6):726-733. DOI: 10.1038/sj.clpt.6100220
  47. 47. Pasanen MK, Neuvonen M, Neuvonen PJ, Niemi M. SLCO1B1 polymorphism markedly affects the pharmacokinetics of simvastatin acid. Pharmacogenetics and Genomics. 2006;16(12):873-879. DOI: 10.1097/01.fpc.0000230416.82349.90
  48. 48. Voora D, Shah SH, Spasojevic I, Ali S, Reed CR, Salisbury BA, Ginsburg GS. The SLCO1B1*5Genetic variant is associated with statin-induced side effects. Journal of the American College of Cardiology. 2009;54(17):1609-1616. DOI: 10.1016/j.jacc.2009.04.053
  49. 49. Higgins JW, Bao JQ, Ke AB, Manro JR, Fallon JK, Smith PC, Zamek-Gliszczynski MJ. Utility of Oatp1a/1b-knockout and OATP1B1/3-humanized mice in the study of OATP-mediated pharmacokinetics and tissue distribution: Case studies with pravastatin, atorvastatin, simvastatin, and carboxydichlorofluorescein. Drug Metabolism and Disposition. 2013;42(1):182-192. DOI: 10.1124/dmd.113.054783
  50. 50. Little J, Higgins JPT, Ioannidis JPA, Moher D, Gagnon F, von Elm E, Khoury MJ, Cohen B, Davey-Smith G, Grimshaw J, Scheet P, Gwinn M, Williamson RE, Zou GY, Hutchings K, Johnson CY, Tait V, Wiens M, Golding J, van Duijn C, McLaughlin J, Paterson A, Wells G, Fortier I, Freedman M, Zecevic M, King R, Infante-Rivard C, Stewart A, Birkett N. STrengthening the REporting of genetic association studies (STREGA)–An extension of the STROBE statement. European Journal of Clinical Investigation. 2009;39(4):247-266. DOI: 10.1111/j.1365-2362.2009.02125.x
  51. 51. Wilke RA, Ramsey LB, Johnson SG, Maxwell WD, McLeod HL, Voora D, Krauss RM, Roden DM, Feng Q, Cooper-DeHoff RM, Gong L, Klein TE, Wadelius M, Niemi M. Clinical pharmacogenomics implementation consortium (CPIC). The clinical pharmacogenomics implementation consortium: CPIC guideline for SLCO1B1 and simvastatin-induced myopathy. Clinical Pharmacology & Therapeutics. 2012;92(1):112-117. DOI: 10.1038/clpt.2012.57
  52. 52. Wilke RA, Fanciullo J. Point-counterpoint: SLCO1B1 genotyping for statins. South Dakota Medicine: The Journal Of The South Dakota State Medical Association. 2017;70(3):102-104
  53. 53. Maxwell WD, Ramsey LB, Johnson SG, Moore KG, Shtutman M, Schoonover JH, Kawaguchi-Suzuki M. Impact of pharmacogenetics on efficacy and safety of statin therapy for dyslipidemia. Pharmacotherapy: The Journal of Human Pharmacology and Drug Therapy. 2017;37(9):1172-1190. DOI: 10.1002/phar.1981
  54. 54. Dunnenberger HM, Crews KR, Hoffman JM, Caudle KE, Broeckel U, Howard SC, Hunkler RJ, Klein TE, Evans WE, Relling MV. Preemptive clinical pharmacogenetics implementation: Current programs in five US medical centers. Annual Review of Pharmacology and Toxicology. 2015;55(1):89-106. DOI: 10.1146/annurev-pharmtox-010814-124835
  55. 55. Cecchin E, Roncato R, Guchelaar HJ, Toffoli G, Ubiquitous Pharmacogenomics Consortium. Ubiquitous pharmacogenomics (U-PGx): The time for implementation is now. An Horizon2020 program to drive pharmacogenomics into clinical practice. Current Pharmaceutical Biotechnology. 2017;18(3):204-209. DOI: 10.2174/1389201018666170103103619
  56. 56. Lamoureux F, Duflot T; French Network of Pharmacogenetics (RNPGX). Pharmacogenetics in cardiovascular diseases: State of the art and implementation-recommendations of the French National Network of Pharmacogenetics (RNPGx). Thérapie. 2017;72(2):257-267. DOI: 10.1016/j.therap.2016.09.017
  57. 57. Luzum JA, Luzum MJ. Physicians’ attitudes toward pharmacogenetic testing before and after pharmacogenetic education. Personalized Medicine. 2016;13(2):119-127. DOI: 10.2217/pme.15.57
  58. 58. Dervieux T, Bala MV. Overview of the pharmacoeconomics of pharmacogenetics. Pharmacogenomics. 2006;7(8):1175-1184. DOI: 10.2217/14622416.7.8.1175
  59. 59. Costa-Scharplatz M, Ramanathan K, Frial T, Beamer B, Gandhi S. Cost-effectiveness analysis of rosuvastatin versus atorvastatin, simvastatin, and pravastatin from a Canadian health system perspective. Clinical Therapeutics. 2008;30(7):1345-1357. DOI: 10.1016/S0149-2918(08)80061-6
  60. 60. Mitchell D, Guertin JR, Iliza AC, Fanton-Aita F, LeLorier J. Economic evaluation of a pharmacogenomics test for statin-induced myopathy in cardiovascular high-risk patients initiating a statin. Molecular Diagnosis & Therapy. 2017;21(1):95-105. DOI: 10.1007/s40291-016-0238-8
  61. 61. Plumpton CO, Roberts D, Pirmohamed M, Hughes DA. A systematic review of economic evaluations of pharmacogenetic testing for prevention of adverse drug reactions. PharmacoEconomics. 2016;34(8):771-793. DOI: 10.1007/s40273-016-0397-9
  62. 62. Spackman E, Hinde S, Bojke L, Payne K, Sculpher M. Using cost-effectiveness analysis to quantify the value of genomic-based diagnostic tests: Recommendations for practice and research. Genetic Testing and Molecular Biomarkers. 2017;21(12):705-716. DOI: 10.1089/gtmb.2017.0105
  63. 63. Payne K, Thompson AJ. Economics of pharmacogenomics: Rethinking beyond QALYs? Pharmacogenomics and Personalized Medicine. 2013;11(3):187-195
  64. 64. Mizzi C, Dalabira E, Kumuthini J, Dzimiri N, Balogh I, Başak N, Böhm R, Borg J, Borgiani P, Bozina N, Bruckmueller H, Burzynska B, Carracedo A, Cascorbi I, Deltas C, Dolzan V, Fenech A, Grech G, Kasiulevicius V, Kádaši Ľ, Kučinskas V, Khusnutdinova E, Loukas YL, Macek M, Makukh H, Mathijssen R, Mitropoulos K, Mitropoulou C, Novelli G, Papantoni I, Pavlovic S, Saglio G, Setric J, Stojiljkovic M, Stubbs AP, Squassina A, Torres M, Turnovec M, van Schaik RH, Voskarides K, Wakil SM, Werk A, del Zompo M, Zukic B, Katsila T, Lee MTM, Motsinger-Rief A, Mc Leod HL, van, der Spek PJ, Patrinos GP, Dubé M-PA. European spectrum of pharmacogenomic biomarkers: Implications for clinical pharmacogenomics. PLoS One. 2016;11(9):e0162866. DOI: 10.1371/journal.pone.0162866
  65. 65. Choi HY, Bae K-S, Cho S-H, Ghim J-L, Choe S, Jung JA, Jin S-J, Kim H-S, Lim H-S. Impact of CYP2D6, CYP3A5, CYP2C19, CYP2A6, SLCO1B1, ABCB1, and ABCG2 gene polymorphisms on the pharmacokinetics of simvastatin and simvastatin acid. Pharmacogenetics and Genomics. 2015;25(12):595-608. DOI: 10.1097/FPC.0000000000000176
  66. 66. Luzum JA, Theusch E, Taylor KD, Wang A, Sadee W, Binkley PF, Krauss RM, Medina MW, Kitzmiller JP. Individual and combined associations of genetic variants in CYP3A4, CYP3A5, and SLCO1B1 with simvastatin and simvastatin acid plasma concentrations. Journal of Cardiovascular Pharmacology. 2015;66(1):80-85. DOI: 10.1097/FJC.0000000000000246
  67. 67. Shah RR, Gaedigk A. Precision medicine: Does ethnicity information complement genotype-based prescribing decisions? Therapeutic Advances in Drug Safety. 2017;9(1):45-62. DOI: 10.1177/2042098617743393
  68. 68. Wei W-Q, Feng Q, Jiang L, Waitara MS, Iwuchukwu OF, Roden DM, Jiang M, Xu H, Krauss RM, Rotter JI, Nickerson DA, Davis RL, Berg RL, Peissig PL, McCarty CA, Wilke RA, Denny JC. Characterization of statin dose response in electronic medical records. Clinical Pharmacology & Therapeutics. 2014;95(3):331-338. DOI: 10.1038/clpt.2013.202
  69. 69. Lenzini P, Wadelius M, Kimmel S, Anderson JL, Jorgensen AL, Pirmohamed M, Caldwell MD, Limdi N, Burmester JK, Dowd MB, Angchaisuksiri P, Bass AR, Chen J, Eriksson N, Rane A, Lindh JD, Carlquist JF, Horne BD, Grice G, Milligan PE, Eby C, Shin J, Kim H, Kurnik D, Stein CM, McMillin G, Pendleton RC, Berg RL, Deloukas P, Gage BF. Integration of genetic, clinical, and INR data to refine warfarin dosing. Clinical Pharmacology & Therapeutics. 2010;87(5):572-578. DOI: 10.1038/clpt.2010.13
  70. 70. van Schaik R. Pharmacogenetics: Do you Have your DNA Passport? [Internet]. 2018. Available from: https://www.labqualitydays.fi/en/news/pharmacogenetics-do-you-have-your-dna-passport/ [Accessed: May 02, 2018]

Written By

Jana Petrkova, Milos Taborsky and Martin Petrek

Submitted: 05 December 2017 Reviewed: 14 June 2018 Published: 05 November 2018