A 30-years review on pharmacokinetics of antibiotics: is the right time for pharmacogenetics?

Lorena Baietto, Silvia Corcione, Giovanni Pacini, Giovanni Di Perri, Antonio D'Avolio, Francesco Giuseppe De Rosa, Lorena Baietto, Silvia Corcione, Giovanni Pacini, Giovanni Di Perri, Antonio D'Avolio, Francesco Giuseppe De Rosa

Abstract

Drug bioavailability may vary greatly amongst individuals, affecting both efficacy and toxicity: in humans, genetic variations account for a relevant proportion of such variability. In the last decade the use of pharmacogenetics in clinical practice, as a tool to individualize treatment, has shown a different degree of diffusion in various clinical fields. In the field of infectious diseases, several studies identified a great number of associations between host genetic polymorphisms and responses to antiretroviral therapy. For example, in patients treated with abacavir the screening for HLA-B*5701 before starting treatment is routine clinical practice and standard of care for all patients; efavirenz plasma levels are influenced by single nucleotide polymorphism (SNP) CYP2B6-516G>T (rs3745274). Regarding antibiotics, many studies investigated drug transporters involved in antibiotic bioavailability, especially for fluoroquinolones, cephalosporins, and antituberculars. To date, few data are available about pharmacogenetics of recently developed antibiotics such as tigecycline, daptomycin or linezolid. Considering the effect of SNPs in gene coding for proteins involved in antibiotics bioavailability, few data have been published. Increasing knowledge in the field of antibiotic pharmacogenetics could be useful to explain the high drug inter-patients variability and to individualize therapy. In this paper we reported an overview of pharmacokinetics, pharmacodynamics, and pharmacogenetics of antibiotics to underline the importance of an integrated approach in choosing the right dosage in clinical practice.

References

    1. Lesko L.J., Schmidt S. Individualization of drug therapy: history, present state, and opportunities for the future. Clin. Pharmacol. Ther. 2012;92(4):458–466.
    1. Cressey T.R., Lallemant M. Pharmacogenetics of antiretroviral drugs for the treatment of HIV-infected patients: an update. Infect. Genet. Evol. 2007;7(2):333–342. doi: 10.1016/j.meegid.2006.08.004.
    1. Wright J., Paauw D.S. Complications of antibiotic therapy. Med. Clin. North Am. 2013;97(4):667–679, xi. doi: 10.1016/j.mcna.2013.02.006. [xi.].
    1. Neuvonen P. J., Kivisto K. T., Lehto P. Interference of dairy products with the absorption of ciprofloxacin. Clin. Pharmacol. Ther. 1991.
    1. Meyer F.P., Specht H., Quednow B., Walther H. Influence of milk on the bioavailability of doxycycline--new aspects. Infection. 1989;17(4):245–246. doi: 10.1007/BF01639529.
    1. Minami R., Inotsume N., Nakano M., Sudo Y., Higashi A., Matsuda I. Effect of milk on absorption of norfloxacin in healthy volunteers. J. Clin. Pharmacol. 1993;33(12):1238–1240. doi: 10.1002/j.1552-4604.1993.tb03926.x.
    1. Gorski J.C., Jones D.R., Haehner-Daniels B.D., Hamman M.A., O’Mara E.M., Jr, Hall S.D. The contribution of intestinal and hepatic CYP3A to the interaction between midazolam and clarithromycin. Clin. Pharmacol. Ther. 1998;64(2):133–143. doi: 10.1016/S0009-9236(98)90146-1.
    1. Nicolau D.P. Pharmacokinetic and pharmacodynamic properties of meropenem. Clin. Infect. Dis. 2008;47(Suppl. 1):S32–S40. doi: 10.1086/590064.
    1. Kim K., Johnson J.A., Derendorf H. Differences in drug pharmacokinetics between East Asians and Caucasians and the role of genetic polymorphisms. J. Clin. Pharmacol. 2004;44(10):1083–1105. doi: 10.1177/0091270004268128.
    1. Matthews H.W. Racial, ethnic and gender differences in response to medicines. Drug Metabol. Drug Interact. 1995;12(2):77–91. doi: 10.1515/DMDI.1995.12.2.77.
    1. Mwinyi J., Johne A., Bauer S., Roots I., Gerloff T. Evidence for inverse effects of OATP-C (SLC21A6) 5 and 1b haplotypes on pravastatin kinetics. Clin. Pharmacol. Ther. 2004;75(5):415–421. doi: 10.1016/j.clpt.2003.12.016.
    1. Niemi M., Schaeffeler E., Lang T., Fromm M.F., Neuvonen M., Kyrklund C., Backman J.T., Kerb R., Schwab M., Neuvonen P.J., Eichelbaum M., Kivistö K.T. High plasma pravastatin concentrations are associated with single nucleotide polymorphisms and haplotypes of organic anion transporting polypeptide-C (OATP-C, SLCO1B1). Pharmacogenetics. 2004;14(7):429–440. doi: 10.1097/01.fpc.0000114750.08559.32.
    1. Hoffmeyer S., Burk O., von Richter O., Arnold H.P., Brockmöller J., Johne A., Cascorbi I., Gerloff T., Roots I., Eichelbaum M., Brinkmann U. Functional polymorphisms of the human multidrug-resistance gene: multiple sequence variations and correlation of one allele with P-glycoprotein expression and activity in vivo. Proc. Natl. Acad. Sci. USA. 2000;97(7):3473–3478. doi: 10.1073/pnas.97.7.3473.
    1. Maeda K., Sugiyama Y. Impact of genetic polymorphisms of transporters on the pharmacokinetic, pharmacodynamic and toxicological properties of anionic drugs. Drug Metab. Pharmacokinet. 2008;23(4):223–235. doi: 10.2133/dmpk.23.223.
    1. Zhang W., Yu B.N., He Y.J., Fan L., Li Q., Liu Z.Q., Wang A., Liu Y.L., Tan Z.R., Fen-Jiang, Huang Y.F., Zhou H.H. Role of BCRP 421C>A polymorphism on rosuvastatin pharmacokinetics in healthy Chinese males. Clin. Chim. Acta. 2006;373(1-2):99–103. doi: 10.1016/j.cca.2006.05.010.
    1. Sparreboom A., Gelderblom H., Marsh S., Ahluwalia R., Obach R., Principe P., Twelves C., Verweij J., McLeod H.L. Diflomotecan pharmacokinetics in relation to ABCG2 421C>A genotype. Clin. Pharmacol. Ther. 2004;76(1):38–44. doi: 10.1016/j.clpt.2004.03.003.
    1. Naesens M., Kuypers D.R., Verbeke K., Vanrenterghem Y. Multidrug resistance protein 2 genetic polymorphisms influence mycophenolic acid exposure in renal allograft recipients. Transplantation. 2006;82(8):1074–1084. doi: 10.1097/01.tp.0000235533.29300.e7.
    1. Rau T., Erney B., Göres R., Eschenhagen T., Beck J., Langer T. High-dose methotrexate in pediatric acute lymphoblastic leukemia: impact of ABCC2 polymorphisms on plasma concentrations. Clin. Pharmacol. Ther. 2006;80(5):468–476. doi: 10.1016/j.clpt.2006.08.012.
    1. Haas D.W., Ribaudo H.J., Kim R.B., Tierney C., Wilkinson G.R., Gulick R.M., Clifford D.B., Hulgan T., Marzolini C., Acosta E.P. Pharmacogenetics of efavirenz and central nervous system side effects: an Adult AIDS Clinical Trials Group study. AIDS. 2004;18(18):2391–2400.
    1. Rodriguez-Novoa S., Barreiro P., Rendón A., Jiménez-Nacher I., González-Lahoz J., Soriano V. Influence of 516G>T polymorphisms at the gene encoding the CYP450-2B6 isoenzyme on efavirenz plasma concentrations in HIV-infected subjects. Clin. Infect. Dis. 2005;40(9):1358–1361. doi: 10.1086/429327.
    1. Ikeda Y., Umemura K., Kondo K., Sekiguchi K., Miyoshi S., Nakashima M. Pharmacokinetics of voriconazole and cytochrome P450 2C19 genetic status. Clin. Pharmacol. Ther. 2004;75(6):587–588. doi: 10.1016/j.clpt.2004.02.002.
    1. Narita A., Muramatsu H., Sakaguchi H., Doisaki S., Tanaka M., Hama A., Shimada A., Takahashi Y., Yoshida N., Matsumoto K., Kato K., Kudo K., Furukawa-Hibi Y., Yamada K., Kojima S. Correlation of CYP2C19 phenotype with voriconazole plasma concentration in children. J. Pediatr. Hematol. Oncol. 2013;35(5):e219–e223. doi: 10.1097/MPH.0b013e3182880eaa.
    1. Becker B., Cooper M.A. Aminoglycoside antibiotics in the 21st century. ACS Chem. Biol. 2013;8(1):105–115. doi: 10.1021/cb3005116.
    1. Briskier A., Veyssier P. Aminocyclitol and aminoglycoside. In: Press A., editor. Antimicrobial agents; antibacterials and antifungals. 2005. pp. 453–457.
    1. Banerjee S.K., Jagannath C., Hunter R.L., Dasgupta A. Bioavailability of tobramycin after oral delivery in FVB mice using CRL-1605 copolymer, an inhibitor of P-glycoprotein. Life Sci. 2000;67(16):2011–2016. doi: 10.1016/S0024-3205(00)00786-4.
    1. Briskier A., Veyssier P. Tetracyclines. In: Press A., editor. Antimicrobial agents; antibacterials and antifungals. 2005. pp. 642–649.
    1. Agwuh K.N., MacGowan A. Pharmacokinetics and pharmacodynamics of the tetracyclines including glycylcyclines. J. Antimicrob. Chemother. 2006;58(2):256–265. doi: 10.1093/jac/dkl224.
    1. Amin A.R., Attur M.G., Thakker G.D., Patel P.D., Vyas P.R., Patel R.N., Patel I.R., Abramson S.B. A novel mechanism of action of tetracyclines: effects on nitric oxide synthases. Proc. Natl. Acad. Sci. USA. 1996;93(24):14014–14019. doi: 10.1073/pnas.93.24.14014.
    1. Gabler W.L., Creamer H.R. Suppression of human neutrophil functions by tetracyclines. J. Periodontal Res. 1991;26(1):52–58. doi: 10.1111/j.1600-0765.1991.tb01626.x.
    1. Whiteman M., Halliwell B. Prevention of peroxynitrite-dependent tyrosine nitration and inactivation of alpha1-antiproteinase by antibiotics. Free Radic. Res. 1997;26(1):49–56. doi: 10.3109/10715769709097783.
    1. Colovic M., Caccia S. Liquid chromatographic determination of minocycline in brain-to-plasma distribution studies in the rat. J. Chromatogr. B Analyt. Technol. Biomed. Life Sci. 2003;791(1-2):337–343. doi: 10.1016/S1570-0232(03)00247-2.
    1. Milane A., Fernandez C., Vautier S., Bensimon G., Meininger V., Farinotti R. Minocycline and riluzole brain disposition: interactions with p-glycoprotein at the blood-brain barrier. J. Neurochem. 2007;103(1):164–173.
    1. Sugie M., Asakura E., Zhao Y.L., Torita S., Nadai M., Baba K., Kitaichi K., Takagi K., Takagi K., Hasegawa T. Possible involvement of the drug transporters P glycoprotein and multidrug resistance-associated protein Mrp2 in disposition of azithromycin. Antimicrob. Agents Chemother. 2004;48(3):809–814. doi: 10.1128/AAC.48.3.809-814.2004.
    1. Garver E., Hugger E.D., Shearn S.P., Rao A., Dawson P.A., Davis C.B., Han C. Involvement of intestinal uptake transporters in the absorption of azithromycin and clarithromycin in the rat. Drug Metab. Dispos. 2008;36(12):2492–2498. doi: 10.1124/dmd.108.022285.
    1. Lemaire S., Van Bambeke F., Mingeot-Leclercq M.P., Tulkens P.M. Modulation of the cellular accumulation and intracellular activity of daptomycin towards phagocytized Staphylococcus aureus by the P-glycoprotein (MDR1) efflux transporter in human THP-1 macrophages and madin-darby canine kidney cells. Antimicrob. Agents Chemother. 2007;51(8):2748–2757. doi: 10.1128/AAC.00090-07.
    1. Schuetz E.G., Schinkel A.H., Relling M.V., Schuetz J.D. P-glycoprotein: a major determinant of rifampicin-inducible expression of cytochrome P4503A in mice and humans. Proc. Natl. Acad. Sci. USA. 1996;93(9):4001–4005. doi: 10.1073/pnas.93.9.4001.
    1. Giamarellou H., Poulakou G. Pharmacokinetic and pharmacodynamic evaluation of tigecycline. Expert Opin. Drug Metab. Toxicol. 2011;7(11):1459–1470. doi: 10.1517/17425255.2011.623126.
    1. Petersen P.J., Jacobus N.V., Weiss W.J., Sum P.E., Testa R.T. In vitro and in vivo antibacterial activities of a novel glycylcycline, the 9-t-butylglycylamido derivative of minocycline (GAR-936). Antimicrob. Agents Chemother. 1999;43(4):738–744.
    1. Meagher A.K., Ambrose P.G., Grasela T.H., Ellis-Grosse E.J. Pharmacokinetic/pharmacodynamic profile for tigecycline-a new glycylcycline antimicrobial agent. Diagn. Microbiol. Infect. Dis. 2005;52(3):165–171. doi: 10.1016/j.diagmicrobio.2005.05.006.
    1. Zhanel G.G., Karlowsky J.A., Rubinstein E., Hoban D.J. Tigecycline: a novel glycylcycline antibiotic. Expert Rev. Anti Infect. Ther. 2006;4(1):9–25. doi: 10.1586/14787210.4.1.9.
    1. Lemaire S., Van Bambeke F., Mingeot-Leclerq M.P., Tulkens P.M. Proceedings of the 18th ECCMID; April 19-22, 2008; Barcelona, Spain.
    1. Carbon C. Pharmacodynamics of macrolides, azalides, and streptogramins: effect on extracellular pathogens. Clin. Infect. Dis. 1998;27(1):28–32. doi: 10.1086/514619.
    1. Van Bambeke F., Tulkens P.M. Macrolides: pharmacokinetics and pharmacodynamics. Int. J. Antimicrob. Agents. 2001;18(Suppl.1):S17–S23. doi: 10.1016/S0924-8579(01)00406-X.
    1. Nightingale C.H., Murakawa T., Ambrose P.G. Macrolide, Azalide, and Ketolide Pharmacodynamics. In: Ag M.D., editor. Antimicrobial Pharmacodynamics in Theory and Clinical Practice. 2002. pp. 205–220.
    1. Lipinski C.A., Lombardo F., Dominy B.W., Feeney P.J. Experimental and computational approaches to estimate solubility and permeability in drug discovery and development settings. Adv. Drug Deliv. Rev. 2001;46(1-3):3–26. doi: 10.1016/S0169-409X(00)00129-0.
    1. He X.J., Zhao L.M., Qiu F., Sun Y.X., Li-Ling J. Influence of ABCB1 gene polymorphisms on the pharmacokinetics of azithromycin among healthy Chinese Han ethnic subjects. Pharmacol. Rep. 2009;61(5):843–850. doi: 10.1016/S1734-1140(09)70140-9.
    1. Baietto L., D’Avolio A., Ariaudo A., Corcione S., Simiele M., Cusato J., Urbino R., Di Perri G., Ranieri V.M., De Rosa F.G. Development and validation of a new UPLC-PDA method to quantify linezolid in plasma and in dried plasma spots. J. Chromatogr. B Analyt. Technol. Biomed. Life Sci. 2013;936:42–47. doi: 10.1016/j.jchromb.2013.08.003.
    1. Rivera A.M., Boucher H.W. Current concepts in antimicrobial therapy against select gram-positive organisms: methicillin-resistant Staphylococcus aureus, penicillin-resistant pneumococci, and vancomycin-resistant enterococci. Mayo Clin. Proc. 2011;86(12):1230–1243. doi: 10.4065/mcp.2011.0514.
    1. Pea F., Furlanut M., Cojutti P., Cristini F., Zamparini E., Franceschi L., Viale P. Therapeutic drug monitoring of linezolid: a retrospective monocentric analysis. Antimicrob. Agents Chemother. 2010;54(11):4605–4610. doi: 10.1128/AAC.00177-10.
    1. Rayner C.R., Forrest A., Meagher A.K., Birmingham M.C., Schentag J.J. Clinical pharmacodynamics of linezolid in seriously ill patients treated in a compassionate use programme. Clin. Pharmacokinet. 2003;42(15):1411–1423. doi: 10.2165/00003088-200342150-00007.
    1. Canut A., Isla A., Betriu C., Gascón A.R. Pharmacokinetic-pharmacodynamic evaluation of daptomycin, tigecycline, and linezolid versus vancomycin for the treatment of MRSA infections in four western European countries. Eur. J. Clin. Microbiol. Infect. Dis. 2012;31(9):2227–2235. doi: 10.1007/s10096-012-1560-7.
    1. Dryden M.S. Linezolid pharmacokinetics and pharmacodynamics in clinical treatment. J. Antimicrob. Chemother. 2011;66(Suppl. 4):iv7–iv15. doi: 10.1093/jac/dkr072.
    1. De Rosa F.G., Corcione S., Baietto L., Ariaudo A., Di Perri G., Ranieri V.M., D’Avolio A. Pharmacokinetics of linezolid during extracorporeal membrane oxygenation. Int. J. Antimicrob. Agents. 2013;41(6):590–591. doi: 10.1016/j.ijantimicag.2013.01.016.
    1. Gebhart B.C., Barker B.C., Markewitz B.A. Decreased serum linezolid levels in a critically ill patient receiving concomitant linezolid and rifampin. Pharmacotherapy. 2007;27(3):476–479. doi: 10.1592/phco.27.3.476.
    1. Andriole V.T. The quinolones: past, present, and future. Clin. Infect. Dis. 2005;41(Suppl. 2):S113–S119. doi: 10.1086/428051.
    1. Labreche M.J., Frei C.R. Declining susceptibilities of gram-negative bacteria to the fluoroquinolones: effects on pharmacokinetics, pharmacodynamics, and clinical outcomes. Am. J. Health Syst. Pharm. 2012;69(21):1863–1870. doi: 10.2146/ajhp110464.
    1. Drlica K., Malik M. Fluoroquinolones: action and resistance. Curr. Top. Med. Chem. 2003;3(3):249–282. doi: 10.2174/1568026033452537.
    1. Fish D.N. Levofloxacin: update and perspectives on one of the original ‘respiratory quinolones’. Expert Rev. Anti Infect. Ther. 2003;1(3):371–387. doi: 10.1586/14787210.1.3.371.
    1. Ambrose P.G., Bhavnani S.M., Owens R.C., Jr Clinical pharmacodynamics of quinolones. Infect. Dis. Clin. North Am. 2003;17(3):529–543. doi: 10.1016/S0891-5520(03)00061-8.
    1. Fish D.N., Chow A.T. The clinical pharmacokinetics of levofloxacin. Clin. Pharmacokinet. 1997;32(2):101–119. doi: 10.2165/00003088-199732020-00002.
    1. Wolfson J.S., Hooper D.C. Treatment of genitourinary tract infections with fluoroquinolones: activity in vitro, pharmacokinetics, and clinical efficacy in urinary tract infections and prostatitis. Antimicrob. Agents Chemother. 1989;33(10):1655–1661. doi: 10.1128/AAC.33.10.1655.
    1. Martinez M., McDermott P., Walker R. Pharmacology of the fluoroquinolones: a perspective for the use in domestic animals. Vet. J. 2006;172(1):10–28. doi: 10.1016/j.tvjl.2005.07.010.
    1. Nightingale C.H. Moxifloxacin, a new antibiotic designed to treat community-acquired respiratory tract infections: a review of microbiologic and pharmacokinetic-pharmacodynamic characteristics. Pharmacotherapy. 2000;20(3):245–256. doi: 10.1592/phco.20.4.245.34880.
    1. Weiner M., Burman W., Luo C.C., Peloquin C.A., Engle M., Goldberg S., Agarwal V., Vernon A. Effects of rifampin and multidrug resistance gene polymorphism on concentrations of moxifloxacin. Antimicrob. Agents Chemother. 2007;51(8):2861–2866. doi: 10.1128/AAC.01621-06.
    1. Yamaguchi H., Yano I., Hashimoto Y., Inui K.I. Secretory mechanisms of grepafloxacin and levofloxacin in the human intestinal cell line caco-2. J. Pharmacol. Exp. Ther. 2000;295(1):360–366.
    1. Ito T., Yano I., Tanaka K., Inui K.I. Transport of quinolone antibacterial drugs by human P-glycoprotein expressed in a kidney epithelial cell line, LLC-PK1. J. Pharmacol. Exp. Ther. 1997;282(2):955–960.
    1. Lowes S., Simmons N.L. Multiple pathways for fluoroquinolone secretion by human intestinal epithelial (Caco-2) cells. Br. J. Pharmacol. 2002;135(5):1263–1275. doi: 10.1038/sj.bjp.0704560.
    1. Maeda T., Takahashi K., Ohtsu N., Oguma T., Ohnishi T., Atsumi R., Tamai I. Identification of influx transporter for the quinolone antibacterial agent levofloxacin. Mol. Pharm. 2007;4(1):85–94. doi: 10.1021/mp060082j.
    1. Merino G., Alvarez A.I., Pulido M.M., Molina A.J., Schinkel A.H., Prieto J.G. Breast cancer resistance protein (BCRP/ABCG2) transports fluoroquinolone antibiotics and affects their oral availability, pharmacokinetics, and milk secretion. Drug Metab. Dispos. 2006;34(4):690–695. doi: 10.1124/dmd.105.008219.
    1. Ando T., Kusuhara H., Merino G., Alvarez A.I., Schinkel A.H., Sugiyama Y. Involvement of breast cancer resistance protein (ABCG2) in the biliary excretion mechanism of fluoroquinolones. Drug Metab. Dispos. 2007;35(10):1873–1879. doi: 10.1124/dmd.107.014969.
    1. Marquez B., Caceres N.E., Mingeot-Leclercq M.P., Tulkens P.M., Van Bambeke F. Identification of the efflux transporter of the fluoroquinolone antibiotic ciprofloxacin in murine macrophages: studies with ciprofloxacin-resistant cells. Antimicrob. Agents Chemother. 2009;53(6):2410–2416. doi: 10.1128/AAC.01428-08.
    1. Arakawa H., Shirasaka Y., Haga M., Nakanishi T., Tamai I. Active intestinal absorption of fluoroquinolone antibacterial agent ciprofloxacin by organic anion transporting polypeptide, Oatp1a5. Biopharm. Drug Dispos. 2012;33(6):332–341. doi: 10.1002/bdd.1809.
    1. Haslam I.S., Wright J.A., O’Reilly D.A., Sherlock D.J., Coleman T., Simmons N.L. Intestinal ciprofloxacin efflux: the role of breast cancer resistance protein (ABCG2). Drug Metab. Dispos. 2011;39(12):2321–2328. doi: 10.1124/dmd.111.038323.
    1. de Lange E.C., Marchand S., van den Berg D., van der Sandt I.C., de Boer A.G., Delon A., Bouquet S., Couet W. In vitro and in vivo investigations on fluoroquinolones; effects of the P-glycoprotein efflux transporter on brain distribution of sparfloxacin. Eur. J. Pharm. Sci. 2000;12(2):85–93. doi: 10.1016/S0928-0987(00)00149-4.
    1. Cormet-Boyaka E., Huneau J.F., Mordrelle A., Boyaka P.N., Carbon C., Rubinstein E., Tomé D. Secretion of sparfloxacin from the human intestinal Caco-2 cell line is altered by P-glycoprotein inhibitors. Antimicrob. Agents Chemother. 1998;42(10):2607–2611.
    1. Putnam W.S., Woo J.M., Huang Y., Benet L.Z. Effect of the MDR1 C3435T variant and P-glycoprotein induction on dicloxacillin pharmacokinetics. J. Clin. Pharmacol. 2005;45(4):411–421. doi: 10.1177/0091270004273492.
    1. Yin O.Q., Tomlinson B., Chow M.S. Effect of multidrug resistance gene-1 (ABCB1) polymorphisms on the single-dose pharmacokinetics of cloxacillin in healthy adult Chinese men. Clin. Ther. 2009;31(5):999–1006. doi: 10.1016/j.clinthera.2009.05.014.
    1. Baietto L., D’avolio A., De Rosa F.G., Cusato J., Pace S., Calcagno A., Pagani N., Montrucchio C., Simiele M., Di Perri G. Single Nucleotide Polymorphisms of ABCB1 Gene Influence Daptomycin Pharmacokinetics in Adult Patients, proceedings of the 52nd ICAAC, San Francisco, USA, September. 2012.
    1. Gill H.J., Tjia J.F., Kitteringham N.R., Pirmohamed M., Back D.J., Park B.K. The effect of genetic polymorphisms in CYP2C9 on sulphamethoxazole N-hydroxylation. Pharmacogenetics. 1999;9(1):43–53. doi: 10.1097/00008571-199902000-00007.
    1. Wang D., Curtis A., Papp A.C., Koletar S.L., Para M.F. Polymorphism in glutamate cysteine ligase catalytic subunit (GCLC) is associated with sulfamethoxazole-induced hypersensitivity in HIV/AIDS patients. BMC Med. Genomics. 2012;5:32. doi: 10.1186/1755-8794-5-32.
    1. Weiner M., Peloquin C., Burman W., Luo C.C., Engle M., Prihoda T.J., Mac Kenzie W.R., Bliven-Sizemore E., Johnson J.L., Vernon A. Effects of tuberculosis, race, and human gene SLCO1B1 polymorphisms on rifampin concentrations. Antimicrob. Agents Chemother. 2010;54(10):4192–4200. doi: 10.1128/AAC.00353-10.
    1. Kwara A., Cao L., Yang H., Poethke P., Kurpewski J., Tashima K.T., Mahjoub B.D., Court M.H., Peloquin C.A. Factors associated with variability in rifampin plasma pharmacokinetics and the relationship between rifampin concentrations and induction of efavirenz clearance. Pharmacotherapy. 2014;34(3):265–271. doi: 10.1002/phar.1388.
    1. Song S.H., Chang H.E., Jun S.H., Park K.U., Lee J.H., Lee E.M., Song Y.H., Song J. Relationship between CES2 genetic variations and rifampicin metabolism. J. Antimicrob. Chemother. 2013;68(6):1281–1284. doi: 10.1093/jac/dkt036.
    1. Chigutsa E., Visser M.E., Swart E.C., Denti P., Pushpakom S., Egan D., Holford N.H., Smith P.J., Maartens G., Owen A., McIlleron H. The SLCO1B1 rs4149032 polymorphism is highly prevalent in South Africans and is associated with reduced rifampin concentrations: dosing implications. Antimicrob. Agents Chemother. 2011;55(9):4122–4127. doi: 10.1128/AAC.01833-10.
    1. Parkin D.P., Vandenplas S., Botha F.J., Vandenplas M.L., Seifart H.I., van Helden P.D., van der Walt B.J., Donald P.R., van Jaarsveld P.P. Trimodality of isoniazid elimination: phenotype and genotype in patients with tuberculosis. Am. J. Respir. Crit. Care Med. 1997;155(5):1717–1722. doi: 10.1164/ajrccm.155.5.9154882.
    1. Kinzig-Schippers M., Tomalik-Scharte D., Jetter A., Scheidel B., Jakob V., Rodamer M., Cascorbi I., Doroshyenko O., Sörgel F., Fuhr U. Should we use N-acetyltransferase type 2 genotyping to personalize isoniazid doses? Antimicrob. Agents Chemother. 2005;49(5):1733–1738. doi: 10.1128/AAC.49.5.1733-1738.2005.
    1. Donald P.R., Sirgel F.A., Venter A., Parkin D.P., Seifart H.I., van de Wal B.W., Werely C., van Helden P.D., Maritz J.S. The influence of human N-acetyltransferase genotype on the early bactericidal activity of isoniazid. Clin. Infect. Dis. 2004;39(10):1425–1430. doi: 10.1086/424999.
    1. Craig W.A. Interrelationship between pharmacokinetics and pharmacodynamics in determining dosage regimens for broad-spectrum cephalosporins. Diagn. Microbiol. Infect. Dis. 1995;22(1-2):89–96. doi: 10.1016/0732-8893(95)00053-D.
    1. Joukhadar C., Frossard M., Mayer B.X., Brunner M., Klein N., Siostrzonek P., Eichler H.G., Müller M. Impaired target site penetration of beta-lactams may account for therapeutic failure in patients with septic shock. Crit. Care Med. 2001;29(2):385–391. doi: 10.1097/00003246-200102000-00030.
    1. Roberts J.A., Kirkpatrick C.M., Roberts M.S., Robertson T.A., Dalley A.J., Lipman J. Meropenem dosing in critically ill patients with sepsis and without renal dysfunction: intermittent bolus versus continuous administration? Monte Carlo dosing simulations and subcutaneous tissue distribution. J. Antimicrob. Chemother. 2009;64(1):142–150. doi: 10.1093/jac/dkp139.
    1. Roberts J.A., Roberts M.S., Robertson T.A., Dalley A.J., Lipman J. Piperacillin penetration into tissue of critically ill patients with sepsis--bolus versus continuous administration? Crit. Care Med. 2009;37(3):926–933. doi: 10.1097/CCM.0b013e3181968e44.
    1. Sime F.B., Roberts M.S., Peake S.L., Lipman J., Roberts J.A. Does Beta-lactam Pharmacokinetic Variability in Critically Ill Patients Justify Therapeutic Drug Monitoring? A Systematic Review. Ann. Intensive Care. 2012;2(1):35. doi: 10.1186/2110-5820-2-35.
    1. Paton D.M. Comparative bioavailability and half-lives of cloxacillin and flucloxacillin. Int. J. Clin. Pharmacol. Res. 1986;6(5):347–349.
    1. DRUGBANK, Dicloxacillin. DB00485 (accessed September 15, 2013).
    1. Campoli-Richards D.M., Brogden R.N. Sulbactam/ampicillin. A review of its antibacterial activity, pharmacokinetic properties, and therapeutic use. Drugs. 1987;33(6):577–609. doi: 10.2165/00003495-198733060-00003.
    1. Nahata M.C., Vashi V.I., Swanson R.N., Messig M.A., Chung M. Pharmacokinetics of ampicillin and sulbactam in pediatric patients. Antimicrob. Agents Chemother. 1999;43(5):1225–1229.
    1. Betrosian A.P., Douzinas E.E. Ampicillin-sulbactam: an update on the use of parenteral and oral forms in bacterial infections. Expert Opin. Drug Metab. Toxicol. 2009;5(9):1099–1112. doi: 10.1517/17425250903145251.
    1. DRUGBANK Amoxicillin. (accessed September 15, 2013).
    1. Sorgel F., Kinzig M. The chemistry, pharmacokinetics and tissue distribution of piperacillin/tazobactam. J. Antimicrob. Chemother., 1993, 31 Suppl A, 39-60. 1993.
    1. Kinzig M., Sörgel F., Brismar B., Nord C.E. Pharmacokinetics and tissue penetration of tazobactam and piperacillin in patients undergoing colorectal surgery. Antimicrob. Agents Chemother. 1992;36(9):1997–2004. doi: 10.1128/AAC.36.9.1997.
    1. Tjandramaga T.B., Mullie A., Verbesselt R., De Schepper P.J., Verbist L. Piperacillin: human pharmacokinetics after intravenous and intramuscular administration. Antimicrob. Agents Chemother. 1978;14(6):829–837. doi: 10.1128/AAC.14.6.829.
    1. Hayashi Y., Roberts J.A., Paterson D.L., Lipman J. Pharmacokinetic evaluation of piperacillin-tazobactam. Expert Opin. Drug Metab. Toxicol. 2010;6(8):1017–1031. doi: 10.1517/17425255.2010.506187.
    1. Ghibellini G., Bridges A.S., Generaux C.N., Brouwer K.L. In vitro and in vivo determination of piperacillin metabolism in humans. Drug Metab. Dispos. 2007;35(3):345–349. doi: 10.1124/dmd.106.012278.
    1. Singhvi S.M., Heald A.F., Schreiber E.C. Pharmacokinetics of cephalosporin antibiotics: protein-binding considerations. Chemotherapy. 1978;24(3):121–133. doi: 10.1159/000237771.
    1. Wright W.E., Line V.D. Biliary excretion of cephalosporins in rats: influence of molecular weight. Antimicrob. Agents Chemother. 1980;17(5):842–846. doi: 10.1128/AAC.17.5.842.
    1. Tsuji A., Yoshikawa T., Nishide K., Minami H., Kimura M., Nakashima E., Terasaki T., Miyamoto E., Nightingale C.H., Yamana T. Physiologically based pharmacokinetic model for beta-lactam antibiotics I: Tissue distribution and elimination in rats. J. Pharm. Sci. 1983;72(11):1239–1252. doi: 10.1002/jps.2600721103.
    1. Tsuji A. Impact of transporter-mediated drug absorption, distribution, elimination and drug interactions in antimicrobial chemotherapy. J. Infect. Chemother. 2006;12(5):241–250. doi: 10.1007/s10156-006-0478-3.
    1. Lister P.D. Carbapenems in the USA: focus on doripenem. Expert Rev. Anti Infect. Ther. 2007;5(5):793–809. doi: 10.1586/14787210.5.5.793.
    1. Majumdar A.K., Musson D.G., Birk K.L., Kitchen C.J., Holland S., McCrea J., Mistry G., Hesney M., Xi L., Li S.X., Haesen R., Blum R.A., Lins R.L., Greenberg H., Waldman S., Deutsch P., Rogers J.D. Pharmacokinetics of ertapenem in healthy young volunteers. Antimicrob. Agents Chemother. 2002;46(11):3506–3511. doi: 10.1128/AAC.46.11.3506-3511.2002.
    1. EUCAST. Imipenem: rationale for the EUCAST clinical breakpoints, version 1.3 2009. 2009.
    1. Moon Y.S., Chung K.C., Gill M.A. Pharmacokinetics of meropenem in animals, healthy volunteers, and patients. Clin. Infect. Dis. 1997;24(Suppl. 2):S249–S255. doi: 10.1093/clinids/24.Supplement_2.S249.
    1. Mouton J.W., Touzw D.J., Horrevorts A.M., Vinks A.A. Comparative pharmacokinetics of the carbapenems: clinical implications. Clin. Pharmacokinet. 2000;39(3):185–201. doi: 10.2165/00003088-200039030-00002.
    1. Breilh D., Texier-Maugein J., Allaouchiche B., Saux M.C., Boselli E. Carbapenems. J. Chemother. 2013;25(1):1–17. doi: 10.1179/1973947812Y.0000000032.
    1. Cirillo I., Mannens G., Janssen C., Vermeir M., Cuyckens F., Desai-Krieger D., Vaccaro N., Kao L.M., Devineni D., Redman R., Turner K. Disposition, metabolism, and excretion of [14C]doripenem after a single 500-milligram intravenous infusion in healthy men. Antimicrob. Agents Chemother. 2008;52(10):3478–3483. doi: 10.1128/AAC.00424-08.
    1. Nix D.E., Majumdar A.K., DiNubile M.J. Pharmacokinetics and pharmacodynamics of ertapenem: an overview for clinicians. J. Antimicrob. Chemother. 2004;53(2):ii23–ii28. doi: 10.1093/jac/dkh205.
    1. Matsumoto S., Saito H., Inui K. Transcellular transport of oral cephalosporins in human intestinal epithelial cells, Caco-2: interaction with dipeptide transport systems in apical and basolateral membranes. J. Pharmacol. Exp. Ther. 1994;270(2):498–504.
    1. Saito H., Okuda M., Terada T., Sasaki S., Inui K. Cloning and characterization of a rat H+/peptide cotransporter mediating absorption of beta-lactam antibiotics in the intestine and kidney. J. Pharmacol. Exp. Ther. 1995;275(3):1631–1637.
    1. Terada T., Saito H., Mukai M., Inui K. Characterization of stably transfected kidney epithelial cell line expressing rat H+/peptide cotransporter PEPT1: localization of PEPT1 and transport of beta-lactam antibiotics. J. Pharmacol. Exp. Ther. 1997;281(3):1415–1421.
    1. Ueo H., Motohashi H., Katsura T., Inui K. Human organic anion transporter hOAT3 is a potent transporter of cephalosporin antibiotics, in comparison with hOAT1. Biochem. Pharmacol. 2005;70(7):1104–1113. doi: 10.1016/j.bcp.2005.06.024.
    1. Jariyawat S., Sekine T., Takeda M., Apiwattanakul N., Kanai Y., Sophasan S., Endou H. The interaction and transport of beta-lactam antibiotics with the cloned rat renal organic anion transporter 1. J. Pharmacol. Exp. Ther. 1999;290(2):672–677.
    1. Uwai Y., Saito H., Inui K. Rat renal organic anion transporter rOAT1 mediates transport of urinary-excreted cephalosporins, but not of biliary-excreted cefoperazone. Drug Metab. Pharmacokinet. 2002;17(2):125–129. doi: 10.2133/dmpk.17.125.
    1. Kato S., Ito K., Kato Y., Wakayama T., Kubo Y., Iseki S., Tsuji A. Involvement of multidrug resistance-associated protein 1 in intestinal toxicity of methotrexate. Pharm. Res. 2009;26(6):1467–1476. doi: 10.1007/s11095-009-9858-6.
    1. Susanto M., Benet L.Z. Can the enhanced renal clearance of antibiotics in cystic fibrosis patients be explained by P-glycoprotein transport? Pharm. Res. 2002;19(4):457–462. doi: 10.1023/A:1015191511817.
    1. Luckner P., Brandsch M. Interaction of 31 beta-lactam antibiotics with the H+/peptide symporter PEPT2: analysis of affinity constants and comparison with PEPT1. Eur. J. Pharm. Biopharm. 2005;59(1):17–24. doi: 10.1016/j.ejpb.2004.07.008.
    1. EUROPA Cubicin powder for concentrate for solution for injection or infusion.
    1. Silverman J.A., Perlmutter N.G., Shapiro H.M. Correlation of daptomycin bactericidal activity and membrane depolarization in Staphylococcus aureus. Antimicrob. Agents Chemother. 2003;47(8):2538–2544. doi: 10.1128/AAC.47.8.2538-2544.2003.
    1. Pea F., Cojutti P., Sbrojavacca R., Cadeo B., Cristini F., Bulfoni A., Furlanut M. TDM-guided therapy with daptomycin and meropenem in a morbidly obese, critically ill patient. Ann. Pharmacother. 2011;45(7-8):e37. doi: 10.1345/aph.1P745.
    1. Bhavnani S.M., Rubino C.M., Ambrose P.G., Drusano G.L. Daptomycin exposure and the probability of elevations in the creatine phosphokinase level: data from a randomized trial of patients with bacteremia and endocarditis. Clin. Infect. Dis. 2010;50(12):1568–1574. doi: 10.1086/652767.
    1. Woodworth J.R., Nyhart E.H., Jr, Brier G.L., Wolny J.D., Black H.R. Single-dose pharmacokinetics and antibacterial activity of daptomycin, a new lipopeptide antibiotic, in healthy volunteers. Antimicrob. Agents Chemother. 1992;36(2):318–325. doi: 10.1128/AAC.36.2.318.
    1. Lee B.L., Sachdeva M., Chambers H.F. Effect of protein binding of daptomycin on MIC and antibacterial activity. Antimicrob. Agents Chemother. 1991;35(12):2505–2508. doi: 10.1128/AAC.35.12.2505.
    1. Sauermann R., Rothenburger M., Graninger W., Joukhadar C. Daptomycin: a review 4 years after first approval. Pharmacology. 2008;81(2):79–91. doi: 10.1159/000109868.
    1. Moise-Broder P.A., Forrest A., Birmingham M.C., Schentag J.J. Pharmacodynamics of vancomycin and other antimicrobials in patients with Staphylococcus aureus lower respiratory tract infections. Clin. Pharmacokinet. 2004;43(13):925–942. doi: 10.2165/00003088-200443130-00005.
    1. American Thoracic Society. Infectious Diseases Society of America Guidelines for the management of adults with hospital-acquired, ventilator-associated, and healthcare-associated pneumonia. Am. J. Respir. Crit. Care Med. 2005;171(4):388–416. doi: 10.1164/rccm.200405-644ST.
    1. Vandecasteele S.J., De Vriese A.S., Tacconelli E. The pharmacokinetics and pharmacodynamics of vancomycin in clinical practice: evidence and uncertainties. J. Antimicrob. Chemother. 2013;68(4):743–748. doi: 10.1093/jac/dks495.
    1. Pea F., Cojutti P., Petrosillo N., Furlanut M., Entenza J.M., Veloso T.R., Vouillamoz J., Giddey M., Moreillon P. Continuous infusion may improve the efficacy of vancomycin in treatment of experimental endocarditis due to heterogeneous vancomycin-intermediate Staphylococcus aureus. Antimicrob. Agents Chemother. 2011;55(9):4496. doi: 10.1128/AAC.00603-11.
    1. Reeves D. S. Therapeutic drug monitoring of aminoglycoside antibiotics. 1980.
    1. Ackerman B.H., Taylor E.H., Olsen K.M., Abdel-Malak W., Pappas A.A. Vancomycin serum protein binding determination by ultrafiltration. Drug Intell. Clin. Pharm. 1988;22(4):300–303.
    1. Albrecht L.M., Rybak M.J., Warbasse L.H., Edwards D.J. Vancomycin protein binding in patients with infections caused by Staphylococcus aureus. DICP. 1991;25(7-8):713–715.
    1. Brogden R.N., Peters D.H. Teicoplanin. A reappraisal of its antimicrobial activity, pharmacokinetic properties and therapeutic efficacy. Drugs. 1994;47(5):823–854. doi: 10.2165/00003495-199447050-00008.
    1. Fanos V., Cataldi L. Renal transport of antibiotics and nephrotoxicity: a review. J. Chemother. 2001;13(5):461–472. doi: 10.1179/joc.2001.13.5.461.
    1. del Moral R.G., Olmo A., Aguilar M., O’Valle F. P glycoprotein: a new mechanism to control drug-induced nephrotoxicity. Exp. Nephrol. 1998;6(2):89–97. doi: 10.1159/000020510.
    1. Koch-Weser J., Sidel V.W., Federman E.B., Kanarek P., Finer D.C., Eaton A.E. Adverse effects of sodium colistimethate. Manifestations and specific reaction rates during 317 courses of therapy. Ann. Intern. Med. 1970;72(6):857–868. doi: 10.7326/0003-4819-72-6-857.
    1. Beveridge E.G., Martin A.J. Sodium sulphomethyl derivatives of polymyxins. Br. Pharmacol. Chemother. 1967;29(2):125–135. doi: 10.1111/j.1476-5381.1967.tb01946.x.
    1. Biswas S., Brunel J.M., Dubus J.C., Reynaud-Gaubert M., Rolain J.M. Colistin: an update on the antibiotic of the 21st century. Expert Rev. Anti Infect. Ther. 2012;10(8):917–934. doi: 10.1586/eri.12.78.
    1. Bergen P.J., Li J., Rayner C.R., Nation R.L. Colistin methanesulfonate is an inactive prodrug of colistin against Pseudomonas aeruginosa. Antimicrob. Agents Chemother. 2006;50(6):1953–1958. doi: 10.1128/AAC.00035-06.
    1. Garonzik S.M., Li J., Thamlikitkul V., Paterson D.L., Shoham S., Jacob J., Silveira F.P., Forrest A., Nation R.L. Population pharmacokinetics of colistin methanesulfonate and formed colistin in critically ill patients from a multicenter study provide dosing suggestions for various categories of patients. Antimicrob. Agents Chemother. 2011;55(7):3284–3294. doi: 10.1128/AAC.01733-10.
    1. Mohamed A.F., Karaiskos I., Plachouras D., Karvanen M., Pontikis K., Jansson B., Papadomichelakis E., Antoniadou A., Giamarellou H., Armaganidis A., Cars O., Friberg L.E. Application of a loading dose of colistin methanesulfonate in critically ill patients: population pharmacokinetics, protein binding, and prediction of bacterial kill. Antimicrob. Agents Chemother. 2012;56(8):4241–4249. doi: 10.1128/AAC.06426-11.
    1. Plachouras D., Karvanen M., Friberg L.E., Papadomichelakis E., Antoniadou A., Tsangaris I., Karaiskos I., Poulakou G., Kontopidou F., Armaganidis A., Cars O., Giamarellou H. Population pharmacokinetic analysis of colistin methanesulfonate and colistin after intravenous administration in critically ill patients with infections caused by gram-negative bacteria. Antimicrob. Agents Chemother. 2009;53(8):3430–3436. doi: 10.1128/AAC.01361-08.
    1. Conly J., Johnston B. Colistin: the phoenix arises. Can. J. Infect. Dis. Med. Microbiol. 2006;17(5):267–269.
    1. Li J., Milne R.W., Nation R.L., Turnidge J.D., Smeaton T.C., Coulthard K. Use of high-performance liquid chromatography to study the pharmacokinetics of colistin sulfate in rats following intravenous administration. Antimicrob. Agents Chemother. 2003;47(5):1766–1770. doi: 10.1128/AAC.47.5.1766-1770.2003.
    1. Bergen P.J., Landersdorfer C.B., Zhang J., Zhao M., Lee H.J., Nation R.L., Li J. Pharmacokinetics and pharmacodynamics of ‘old’ polymyxins: what is new? Diagn. Microbiol. Infect. Dis. 2012;74(3):213–223. doi: 10.1016/j.diagmicrobio.2012.07.010.
    1. Li J., Milne R.W., Nation R.L., Turnidge J.D., Smeaton T.C., Coulthard K. Pharmacokinetics of colistin methanesulphonate and colistin in rats following an intravenous dose of colistin methanesulphonate. J. Antimicrob. Chemother. 2004;53(5):837–840. doi: 10.1093/jac/dkh167.
    1. Bosso J.A., Liptak C.A., Seilheimer D.K., Harrison G.M. Toxicity of colistin in cystic fibrosis patients. DICP. 1991;25(11):1168–1170.
    1. Li J., Nation R.L., Milne R.W., Turnidge J.D., Coulthard K. Evaluation of colistin as an agent against multi-resistant Gram-negative bacteria. Int. J. Antimicrob. Agents. 2005;25(1):11–25. doi: 10.1016/j.ijantimicag.2004.10.001.
    1. Jin L., Li J., Nation R.L., Nicolazzo J.A. Impact of p-glycoprotein inhibition and lipopolysaccharide administration on blood-brain barrier transport of colistin in mice. Antimicrob. Agents Chemother. 2011;55(2):502–507. doi: 10.1128/AAC.01273-10.
    1. Bryskier A. Dihydrofolate reductase inhibitors, nitroheterocycles (furans), and 8-Hydroxyquinolines. In: Press A., editor. Antimicrobial Agents: Antibacterials and Antifungals. 2005. pp. 941–945.
    1. Pirmohamed M., Alfirevic A., Vilar J., Stalford A., Wilkins E.G., Sim E., Park B.K. Association analysis of drug metabolizing enzyme gene polymorphisms in HIV-positive patients with co-trimoxazole hypersensitivity. Pharmacogenetics. 2000;10(8):705–713. doi: 10.1097/00008571-200011000-00005.
    1. Telenti A. Genetics and pulmonary medicine. 5. Genetics of drug resistant tuberculosis. Thorax. 1998;53(9):793–797. doi: 10.1136/thx.53.9.793.
    1. Zhang Y., Wade M.M., Scorpio A., Zhang H., Sun Z. Mode of action of pyrazinamide: disruption of Mycobacterium tuberculosis membrane transport and energetics by pyrazinoic acid. J. Antimicrob. Chemother. 2003;52(5):790–795. doi: 10.1093/jac/dkg446.
    1. Peloquin C.A. Therapeutic drug monitoring in the treatment of tuberculosis. Drugs. 2002;62(15):2169–2183. doi: 10.2165/00003495-200262150-00001.
    1. Lin M.Y., Lin S.J., Chan L.C., Lu Y.C. Impact of food and antacids on the pharmacokinetics of anti-tuberculosis drugs: systematic review and meta-analysis. Int. J. Tuberc. Lung Dis. 2010;14(7):806–818.
    1. Huang Y.S., Chern H.D., Su W.J., Wu J.C., Chang S.C., Chiang C.H., Chang F.Y., Lee S.D. Cytochrome P450 2E1 genotype and the susceptibility to antituberculosis drug-induced hepatitis. Hepatology. 2003;37(4):924–930. doi: 10.1053/jhep.2003.50144.
    1. Jamis-Dow C.A., Katki A.G., Collins J.M., Klecker R.W. Rifampin and rifabutin and their metabolism by human liver esterases. Xenobiotica. 1997;27(10):1015–1024. doi: 10.1080/004982597239994.
    1. Donald P.R., Maritz J.S., Diacon A.H. The pharmacokinetics and pharmacodynamics of rifampicin in adults and children in relation to the dosage recommended for children. Tuberculosis (Edinb.) 2011;91(3):196–207. doi: 10.1016/j.tube.2011.02.004.
    1. Ethambutol D.R. DRUGBANK Ethambutol. (accessed February 13,
    1. Ramachandran G., Swaminathan S. Role of pharmacogenomics in the treatment of tuberculosis: a review. Pharm. Genomics Pers. Med. 2012;5:89–98.
    1. Yee D., Valiquette C., Pelletier M., Parisien I., Rocher I., Menzies D. Incidence of serious side effects from first-line antituberculosis drugs among patients treated for active tuberculosis. Am. J. Respir. Crit. Care Med. 2003;167(11):1472–1477. doi: 10.1164/rccm.200206-626OC.
    1. Ben Mahmoud L., Ghozzi H., Kamoun A., Hakim A., Hachicha H., Hammami S., Sahnoun Z., Zalila N., Makni H., Zeghal K. Polymorphism of the N-acetyltransferase 2 gene as a susceptibility risk factor for antituberculosis drug-induced hepatotoxicity in Tunisian patients with tuberculosis. Pathol. Biol. (Paris) 2012;60(5):324–330. doi: 10.1016/j.patbio.2011.07.001.
    1. Cho H.J., Koh W.J., Ryu Y.J., Ki C.S., Nam M.H., Kim J.W., Lee S.Y. Genetic polymorphisms of NAT2 and CYP2E1 associated with antituberculosis drug-induced hepatotoxicity in Korean patients with pulmonary tuberculosis. Tuberculosis (Edinb.) 2007;87(6):551–556. doi: 10.1016/j.tube.2007.05.012.
    1. Hiratsuka M., Kishikawa Y., Takekuma Y., Matsuura M., Narahara K., Inoue T., Hamdy S.I., Endo N., Goto J., Mizugaki M. Genotyping of the N-acetyltransferase2 polymorphism in the prediction of adverse drug reactions to isoniazid in Japanese patients. Drug Metab. Pharmacokinet. 2002;17(4):357–362. doi: 10.2133/dmpk.17.357.
    1. Lee S.W., Chung L.S., Huang H.H., Chuang T.Y., Liou Y.H., Wu L.S. NAT2 and CYP2E1 polymorphisms and susceptibility to first-line anti-tuberculosis drug-induced hepatitis. Int. J. Tuberc. Lung Dis. 2010;14(5):622–626.
    1. Ohno M., Yamaguchi I., Yamamoto I., Fukuda T., Yokota S., Maekura R., Ito M., Yamamoto Y., Ogura T., Maeda K., Komuta K., Igarashi T., Azuma J. Slow N-acetyltransferase 2 genotype affects the incidence of isoniazid and rifampicin-induced hepatotoxicity. Int. J. Tuberc. Lung Dis. 2000;4(3):256–261.
    1. Wang P.Y., Xie S.Y., Hao Q., Zhang C., Jiang B.F. NAT2 polymorphisms and susceptibility to anti-tuberculosis drug-induced liver injury: a meta-analysis. Int. J. Tuberc. Lung Dis. 2012;16(5):589–595.
    1. Roy B., Ghosh S.K., Sutradhar D., Sikdar N., Mazumder S., Barman S. Predisposition of antituberculosis drug induced hepatotoxicity by cytochrome P450 2E1 genotype and haplotype in pediatric patients. J. Gastroenterol. Hepatol. 2006;21(4):784–786. doi: 10.1111/j.1440-1746.2006.04197.x.
    1. Gupta V.H., Singh M., Amarapurkar D.N., Sasi P., Joshi J.M., Baijal R., H R P.K., Amarapurkar A.D., Joshi K., Wangikar P.P. Association of GST null genotypes with anti-tuberculosis drug induced hepatotoxicity in Western Indian population. Ann. Hepatol. 2013;12(6):959–965.
    1. Huang Y.S., Su W.J., Huang Y.H., Chen C.Y., Chang F.Y., Lin H.C., Lee S.D. Genetic polymorphisms of manganese superoxide dismutase, NAD(P)H:quinone oxidoreductase, glutathione S-transferase M1 and T1, and the susceptibility to drug-induced liver injury. J. Hepatol. 2007;47(1):128–134. doi: 10.1016/j.jhep.2007.02.009.
    1. Kim S.H., Kim S.H., Yoon H.J., Shin D.H., Park S.S., Kim Y.S., Park J.S., Jee Y.K. TNF-α genetic polymorphism -308G/A and antituberculosis drug-induced hepatitis. Liver Int. 2012;32(5):809–814. doi: 10.1111/j.1478-3231.2011.02697.x.
    1. Kim S.H., Kim S.H., Lee J.H., Lee B.H., Kim Y.S., Park J.S., Jee Y.K. Polymorphisms in drug transporter genes (ABCB1, SLCO1B1 and ABCC2) and hepatitis induced by antituberculosis drugs. Tuberculosis (Edinb.) 2012;92(1):100–104. doi: 10.1016/j.tube.2011.09.007.
    1. Pachot J.I., Botham R.P., Haegele K.D., Hwang K. Experimental estimation of the role of P-Glycoprotein in the pharmacokinetic behaviour of telithromycin, a novel ketolide, in comparison with roxithromycin and other macrolides using the Caco-2 cell model. J. Pharm. Pharm. Sci. 2003;6(1):1–12.

Source: PubMed

3
구독하다