Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2012 Jul 1.
Published in final edited form as: Pharmacogenet Genomics. 2011 Jul;21(7):440–446. doi: 10.1097/FPC.0b013e32833ffb56

Doxorubicin pathways: pharmacodynamics and adverse effects

Caroline F Thorn a, Connie Oshiro a, Sharon Marsh e, Tina Hernandez-Boussard b, Howard McLeod d, Teri E Klein a, Russ B Altman a,c
PMCID: PMC3116111  NIHMSID: NIHMS286388  PMID: 21048526

Introduction

The goal of this study is to give a brief background on the literature supporting the PharmGKB pathway about doxorubicin action, and provides a summary of this active area of research. The reader is referred to recent in-depth reviews [14] for more detailed discussion of this important and complex pathway. Doxorubicin is an anthracyline drug first extracted from Streptomyces peucetius var. caesius in the 1970’s and routinely used in the treatment of several cancers including breast, lung, gastric, ovarian, thyroid, non-Hodgkin’s and Hodgkin’s lymphoma, multiple myeloma, sarcoma, and pediatric cancers [57]. A major limitation for the use of doxorubicin is cardiotoxicity, with the total cumulative dose being the only criteria currently used to predict the toxicity [4,8]. As there is evidence that the mechanisms of anticancer action and of cardiotoxicity occur through different pathways there is hope for the development of anthracycline drugs with equal efficacy but reduced toxicity [4]. Knowledge of the pharmacogenomics of these pathways may eventually allow for future selection of patients more likely to achieve efficacy at lower doses or able to withstand higher doses with lesser toxicity. We present here graphical representations of the candidate genes for the pharmacogenomics of doxorubicin action in a stylized cancer cell (Fig. 1) and toxicity in cardiomyocytes (Fig. 2), and a table describing the key variants examined so far.

Fig. 1.

Fig. 1

Graphical representation of the candidate genes involved in the pharmacodynamics of doxorubicin in a stylized cancer cell. A fully interactive version of this pathway is available online at PharmGKB at http://www.pharmgkb.org/do/serve?objId=PA165292163&objCls=Pathway. PK pathway, pharmacokinetics pathway; ROS, reactive oxygen species.

Fig. 2.

Fig. 2

Graphical representation of the candidate genes involved in the cardiotoxicity of doxorubicin. A fully interactive version of this pathway is available online at PharmGKB at http://www.pharmgkb.org/do/serve?objId=PA165292164&objCls=Pathway. NO, nitrogen oxide; RNS, reactive nitrogen species; ROS, reactive oxygen species.

Mechanisms of anticancer pharmacodynamics

There are two proposed mechanisms by which doxorubicin acts in the cancer cell (i) intercalation into DNA and disruption of topoisomerase-II-mediated DNA repair and (ii) generation of free radicals and their damage to cellular membranes, DNA and proteins (shown in Fig. 1) [9]. In brief, doxorubicin is oxidized to semiquinone, an unstable metabolite, which is converted back to doxorubicin in a process that releases reactive oxygen species. Reactive oxygen species can lead to lipid peroxidation and membrane damage, DNA damage, oxidative stress, and triggers apoptotic pathways of cell death [10]. Candidate genes that may modulate this pathway involve those capable of the oxidation reaction (NADH dehydrogenases, nitric oxide synthases, xanthine oxidase) [11,12] and those capable of deactivating the free radicals such as glutathione peroxidase, catalase, and superoxide dismutase. Alternatively, doxorubicin can enter the nucleus and poison topoisomerase-II, also resulting in DNA damage and cell death [13]. Candidate pharmacogenes for this part of the pathway include the enzymes involved in the DNA repair mechanisms and the cell cycle control (TOP2A, MLH1, MSH2, TP53, and ERCC2 genes). Although the evidence for some of these candidate genes is unrefutable (TOP2A) [13,14] others are included based on the data from model systems but the polymorphic nature may be worth exploring in PGx studies [15,16].

Mechanisms of cardiotoxicity

The exact mechanism of cardiotoxicity of doxorubicin is somewhat controversial. There are two main theories (i) iron-related free radicals and formation of doxorubicinol metabolite and (ii) mitochondrial disruption, which is somewhat intertwined (reviewed in [17] and [18] and shown in Fig. 2). One of the strongest pieces of supporting evidence for the iron hypothesis is that the iron chelator, dexrazoxane is protective against doxorubicin-induced toxicity in vivo [19]. The best evidence supporting the mitochondrial hypothesis is in the association of the genetic variants in several component genes of the mitochondrial NAD(P)H oxidase complex with doxorubicin cardiotoxicity in the pharmacogenetic studies (see pharmacogenetics section for details) [20,21].

In brief, doxorubicin can be reduced to doxorubicinol, a metabolite that interferes with iron (by ACO1) and calcium regulations (by the calcium pump of sarcoplasmic reticulum, ATP2A2 and the Na+/K+ pump of sarcolemma, RYR2) and the F0F1 proton pump of mitochondria (coded by the ATP5 gene family) [3,22,23]. Candidate genes for the formation of doxorubicinol are AKR1C3, AKR1A1, CBR1, and CBR3. (For more details of the doxorubicin metabolism see the doxorubicin pharmacokinetics (PK) pathway at PharmGKB). Candidate genes involved in the generation of reactive oxygen species or reactive nitrogen species from doxorubicin metabolism include nitric oxide synthases [24] and NAD(P)H oxidase complex genes NCF4, CYBA, and RAC2 [20]. Metabolism of doxorubicin within the mitochondria can disrupt respiration and leads to the release of cytochrome-C initiating apoptosis [25].

The mechanism of the action of dexrazoxane protection against cardiotoxicity may be by sequestration of iron preventing free radical formation. However, as other iron chelators, such as deferasirox, fail to exert the protective effects of dexrazoxane [26] an alternative mechanism is by the interaction with TOP2B that prevents doxorubicin from inducing DNA damage [3,27].

Finding a way to maintain the efficacy and reduce toxicity has been one of the major areas of focus of anthracycline research. Although dexrazoxane reduces toxicity when given with doxorubicin, efficacy is also reduced so this combination is used only after a cumulative dose threshold has been reached [28]. Other anthracyclines, daunorubicin, epirubicin, and idarubicin also result in cardiotoxicity to varying degrees [7]. Daunorubicin is also considered as cardiotoxic as doxorubicin [29]. Epirubicin was less toxic than doxorubicin in animal models [29] and some in-vivo data showed less cardiotoxicity for epirubicin [7]. Although a recent Cochrane review and metaanalysis concluded that there was no significant difference between occurrence of clinical heart failure between doxorubicin and epirubicin when looking at the randomized clinical trial data [28]. However, liposomal formulation of doxorubicin was shown to be less cardiotoxic than traditional doxorubicin without compromising efficacy in adults with solid tumors [28]. Idarubicin was also less cardiotoxic in animal models [30] but the Cochrane review did not find sufficient evidence from randomized studies to support any direct comparison in vivo [28].

As with most cancer treatments doxorubicin is rarely given in isolation. Most of the in-vivo studies involve cotreatment with other antineoplastic agents such as taxanes, platinum drugs, nitrogen mustard analogs, fluoropyrimidines, and vinca alkaloids, which can complicate the association of variants with a particular treatment. The reports of drug–drug interaction have been shown for doxorubicin with phenytoin [31,32] and cyclosporine [33], likely by ABCB1, and sorafenib by RALBP1 [34]. More clinically relevant are the drug–drug interactions resulting in the cardiotoxicity from cotreatment with doxorubicin and trastuzumab or taxanes such as paclitaxel and docetaxel [35]. Trastuzumab has Food and Drug Administration black box warning for cardiomyopathy that cautions against concurrent use of anthracyclines. Trastuzumab blocks the ERBB2 signaling of NRG1, a cardioprotective pathway that protects against stress. Without this endogenous cardioprotection, doxorubicin treatment can be more damaging [35]. The interaction with taxanes is through a different mechanism. In model systems, paclitaxel potentiates the cardiotoxicity of doxorubicin; this is less pronounced for docetaxel. Paclitaxel also potentiates the cardiotoxicity of epirubicin, with epirubicin plus docetaxel being the least toxic combination [29]. The mechanism for the taxane-induced increase in cardiotoxicity is by increased formation of doxorubicinol, by the modulation of the catalytic activity of aldehyde reductase [35].

Resistance

Although doxorubicin is a valuable clinical antineoplastic agent, in addition to problems with cardiotoxicity, resistance is also a problem limiting its use [7,36]. The mechanism of resistance involves ABCB1 (MDR1, Pgp) [37] and ABCC1 (MRP1) [38] and other transporters (ABCC2, ABCC3, ABCG2, and RALBP1) [1,39,40]. Another mechanism of doxorubicin resistance is the amplification of TOP2A [41], which has been shown to affect the treatment response [14,42]. The amplification of TOP2A has a complicated relationship to neighboring gene HER-2 (ERBB2), used as a marker for breast cancer treatments in particular the HER-2-targeted trastuzumab (reviewed in [14]). The amplification of ERBB2 gene also affects the doxorubicin response (see below) [14].

Pharmacogenomics

There are considerable interindividual variations in the pharmacokinetic parameters of doxorubicin and doxorubicinol [43]. However, the impact of genetic variants on doxorubicin response has only been studied recently. Cumulative anthracycline dose is the only confirmed significant risk factor for doxorubicin-induced cardiotoxicity [44]. So far, most studies have looked at the effects of variation on PK or resistance with only a few examining clinical outcomes such as cardiotoxicity or survival (summarized in Table 1). Below, we highlight variants associated with clinical phenotypes.

Table 1.

Pharmacogenomic studies of doxorubicin

Gene rs# Common names Variant Phenotype in vivo Phenotype in vitro PMIDs
ABCB1 rs1128503 1236C> T T Increased drug exposure, reduced clearancea 18377430 [45]
ABCB1 rs2032582 2677G>A/T A/T Increased drug exposure, reduced clearancea 18377430 [45]
ABCB1 rs1045642 3435C > T T Increased drug exposure, reduced clearancea 18377430 [45]
ABCB1 rs35810889 Met89Thr C Increased resistance 17352537b [46]
ABCB1 rs35657960 Leu662Arg G Increased resistance 17352537b [46]
ABCB1 rs35023033 Arg669Cys T Increased resistance 17352537b [46]
ABCB1 rs2229107 Ser1141Thr A Increased resistance 17352537b [46]
ABCB1 rs35730308 Trp1108Arg C Reduced resistance 17352537b [46]
ABCC1 rs45511401 Gly671Val T Cardiotoxicity No difference in transport 16330681 [20], 11721885 [47]
ABCC1 rs60782127 Arg433Ser T Increased resistance 12042670 [48]
ABCC1 rs4148356 Arg723Gln A Reduced resistance 19214144 [49]
ABCC2 rs17222723 Val1188Glu A Acute cardiotoxicitya 16330681 [20]
ABCC2 rs8187710 Cys1515Tyr A Acute cardiotoxicitya 16330681 [20]
ABCG2 rs2231142 421C > A C/A No difference in PK 18377430 [45]
CAT rs10836235 c.66 + 78C> T T Cardiotoxicity 19863340 [50]
CAT rs1001179 −262C > T T No association 19863340 [50]
CBR1 rs1143663 Val88Ile A Reduced activity 19204081 [51]
CBR1 rs41557318 Pro131Ser T Reduced activity 19204081 [51]
CBR1 rs20572 c.627C> T T Lower clearance, higher exposurea 19016765 [52]
CBR1 rs9024 1096G>A A Lower clearance, higher exposurea Reduced activity 19016765 [52], 19022938b [53]
CBR3 rs8133052 Cys4Tyr A Lower ratio of doxorubicinol plasma AUC/doxorubicin AUC; lower expression in breast tumor, no effect Higher Vmax lower Km 18551042 [54], 20007405 [55], 19016765 [52]
CBR3 rs1056892 Val244Met A Higher doxorubicinol AUC; higher expression in tumor tissue; risk of heart failure (trend); no effect Conflicting reports of reduced activity and higher activity 18551042 [54], 18457324 [56], 20007405 [55], 15537833 [57], 19016765 [52]
CBR3 rs2835285 Val93Ile T Reduced activity 20007405 [55]
CBR3 rs4987121 Met235Leu T Reduced activity 20007405 [55]
CYBA rs4673 His72Tyr T Acute cardiotoxicity, poorer event-free survival 16330681 [20], 19448608 [21]
GSTM1 0a Null No association 19863340 [50]
GSTT1 0a Null No association 19863340 [50]
ERBB2 HER-2 Amplification Multi Improved disease-free survival, improved overall survival 1 8258981 [42], 19758759 [14]
NCF4 rs1883112 −212A >G G Chronic cardiotoxicity, hematological toxicity 16330681 [20], 19448608 [21]
NOS3 rs2070744 −786T >C C Increased risk of cancer recurrence after chemotherapy that included doxorubicin 19671875 [58]
NOS3 rs1799983 894G> T Glu289Asp T Increased risk of cancer recurrence after chemotherapy that included doxorubicin 19671875 [58]
NQO1 rs1800566 NQO1*2 T No association Very low activity 18457324 [56], 15047100 [59]
NR1I2 n/a PXR*1B, 2654T >C C Lower doxorubicin clearance Reduced mRNA expression of PXR, CYP3A4, and ABCB1 18981011 [60]
RAC2 rs13058338 7508T > A A Acute cardiotoxicity 16330681 [20]
SLC22A16 rs714368 146A >G His49Arg G Increased drug exposure 17559346 [61]
SOD2 rs4880 Val16Ala C No association 19863340 [50]
TOP2A topo II alpha Amplification Multi Increased response 19758759 [14]

PMIDs, PubMed Identifiers.

a

As part of haplotype.

Variants in ABCC1 (rs45511401), ABCC2 (rs8187694, rs8187710), CAT (rs10836235), CBR3 (rs1056892), CYBA (rs4673), NCF4 (rs1883112), and RAC2 (rs13058338) are associated with cardiotoxicity in vivo [20,21,50,56]. Wojnowski et al. [20] in a study of single nucleotide polymorphisms of 82 genes from 1697 patients, 3.2% of whom developed either acute or chronic doxorubicin-induced cardiotoxicity, found five significant associations between cardiotoxicity and polymorphisms of the NAD(P)H oxidase complex (CYBA, NCF4, and RAC2), and doxorubicin transporters. Consistent with this, mice deficient in NAD(P)H oxidase activity, unlike wild-type mice, were resistant to the chronic doxorubicin treatment [20]. A recent study by Rossi et al. [21] in lymphoma patients treated with doxorubicin-containing chemotherapy also showed an association of CYBA (rs4673) and NCF4 (rs1883112) with toxicity.

In a small study, Blanco et al. [56] suggested the CBR3 Val244Met polymorphism (rs1056892) might have an impact on the risk of anthracycline-related congestive heart failure among childhood cancer survivors. This variant was also associated with higher doxorubicinol area under the curve and higher CBR3 expression in the tumor tissue from Asian breast cancer patients [54]. However, another study of Asian breast cancer patients showed no effect of this variant on PK [52]. The in-vitro studies of this variant are conflicting with some showing decreased activity of the variant protein with doxorubicin as a substrate [56,62] and others showing increased activity using menadione as a substrate [57].

A variant in PXR (NR1I3) affects doxorubicin clearance in vivo [60] and may be a candidate gene for the doxorubicin resistance because of its role in regulating the transporter expression [1]. Lal et al. [1] suggest additional transporters as candidate genes in their recent review, including ABCB5, ABCB8, and ABCC5, although none have yet been shown to have doxorubicin-related phenotypes.

In addition to studies of single nucleotide polymorphisms, there have also been PGx studies of gene copy number for ERBB2, TOP2A, and GSTs. Metaanalyses suggest that anthracycline-containing regimens provide more benefit than nonanthracycline-containing regimens in women whose tumors have multiple copies of ERBB2 gene [14,42] (approximately half of the studies examined had regimens that contained doxorubicin, the rest were epirubicin). Amplification of TOP2A gene is also proposed as a biomarker for doxorubicin response although some results have been contradictory [14,42].

Conclusion

There have been substantial studies on the doxorubicin mechanism of action and so we know many of the genes that modulate the doxorubicin response. However, PGx studies that implicate variants in these genes are still in their infancy. As with many antineoplastic drugs, the PGx can be complicated by combined treatments. However, there are clear benefits for identifying individuals at risk for toxicity and response. Assembling the PGx candidates should help us learn how to identify individuals at elevated risk for doxorubicin-related cardiotoxicity or treatment failure. There remains a need for large studies that can simultaneously examine expression and copy number biomarkers and single nucleotide polymorphisms in genes related to all aspects of doxorubicin PD and PK to clarify the complete picture of doxorubicin PGx.

Acknowledgments

This work is supported by the NIH/NIGMS (U01GM61374). The authors thank Fen Liu for assistance with the graphics.

References

  • 1.Lal S, Mahajan A, Chen WN, Chowbay B. Pharmacogenetics of target genes across doxorubicin disposition pathway: a review. Curr Drug Metab. 2010;11:115–128. doi: 10.2174/138920010791110890. [DOI] [PubMed] [Google Scholar]
  • 2.Simunek T, Sterba M, Popelova O, Adamcova M, Hrdina R, Gersl V. Anthracycline-induced cardiotoxicity: overview of studies examining the roles of oxidative stress and free cellular iron. Pharmacol Rep. 2009;61:154–171. doi: 10.1016/s1734-1140(09)70018-0. [DOI] [PubMed] [Google Scholar]
  • 3.Mordente A, Meucci E, Silvestrini A, Martorana GE, Giardina B. New developments in anthracycline-induced cardiotoxicity. Curr Med Chem. 2009;16:1656–1672. doi: 10.2174/092986709788186228. [DOI] [PubMed] [Google Scholar]
  • 4.Carvalho C, Santos RX, Cardoso S, Correia S, Oliveira PJ, Santos MS, et al. Doxorubicin: the good, the bad and the ugly effect. Curr Med Chem. 2009;16:3267–3285. doi: 10.2174/092986709788803312. [DOI] [PubMed] [Google Scholar]
  • 5.Arcamone F, Cassinelli G, Fantini G, Grein A, Orezzi P, Pol C, et al. Adriamycin, 14-hydroxydaunomycin: a new antitumor antibiotic from S. peucetius var. caesius. Biotechnol Bioeng. 1969;11:1101–1110. doi: 10.1002/bit.260110607. [DOI] [PubMed] [Google Scholar]
  • 6.Cortes-Funes H, Coronado C. Role of anthracyclines in the era of targeted therapy. Cardiovasc Toxicol. 2007;7:56–60. doi: 10.1007/s12012-007-0015-3. [DOI] [PubMed] [Google Scholar]
  • 7.Weiss RB. The anthracyclines: will we ever find a better doxorubicin? Semin Oncol. 1992;19:670–686. [PubMed] [Google Scholar]
  • 8.Swain SM, Whaley FS, Ewer MS. Congestive heart failure in patients treated with doxorubicin: a retrospective analysis of three trials. Cancer. 2003;97:2869–2879. doi: 10.1002/cncr.11407. [DOI] [PubMed] [Google Scholar]
  • 9.Gewirtz DA. A critical evaluation of the mechanisms of action proposed for the antitumor effects of the anthracycline antibiotics adriamycin and daunorubicin. Biochem Pharmacol. 1999;57:727–741. doi: 10.1016/s0006-2952(98)00307-4. [DOI] [PubMed] [Google Scholar]
  • 10.Doroshow JH. Role of hydrogen peroxide and hydroxyl radical formation in the killing of Ehrlich tumor cells by anticancer quinones. Proc Natl Acad Sci U S A. 1986;83:4514–4518. doi: 10.1073/pnas.83.12.4514. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Pawlowska J, Tarasiuk J, Wolf CR, Paine MJ, Borowski E. Differential ability of cytostatics from anthraquinone group to generate free radicals in three enzymatic systems: NADH dehydrogenase, NADPH cytochrome P450 reductase, and xanthine oxidase. Oncol Res. 2003;13:245–252. doi: 10.3727/096504003108748294. [DOI] [PubMed] [Google Scholar]
  • 12.Fogli S, Nieri P, Breschi MC. The role of nitric oxide in anthracycline toxicity and prospects for pharmacologic prevention of cardiac damage. FASEB J. 2004;18:664–675. doi: 10.1096/fj.03-0724rev. [DOI] [PubMed] [Google Scholar]
  • 13.Tewey KM, Rowe TC, Yang L, Halligan BD, Liu LF. Adriamycin-induced DNA damage mediated by mammalian DNA topoisomerase-II. Science. 1984;226:466–468. doi: 10.1126/science.6093249. [DOI] [PubMed] [Google Scholar]
  • 14.Oakman C, Moretti E, Galardi F, Santarpia L, Di Leo A. The role of topoisomerase-IIalpha and HER-2 in predicting sensitivity to anthracyclines in breast cancer patients. Cancer Treat Rev. 2009;35:662–667. doi: 10.1016/j.ctrv.2009.08.006. [DOI] [PubMed] [Google Scholar]
  • 15.Fedier A, Schwarz VA, Walt H, Carpini RD, Haller U, Fink D. Resistance to topoisomerase poisons due to loss of DNA mismatch repair. Int J Cancer. 2001;93:571–576. doi: 10.1002/ijc.1356. [DOI] [PubMed] [Google Scholar]
  • 16.Robles AI, Wang XW, Harris CC. Drug-induced apoptosis is delayed and reduced in XPD lymphoblastoid cell lines: possible role of TFIIH in p53-mediated apoptotic cell death. Oncogene. 1999;18:4681–4688. doi: 10.1038/sj.onc.1202862. [DOI] [PubMed] [Google Scholar]
  • 17.Minotti G, Recalcati S, Menna P, Salvatorelli E, Corna G, Cairo G. Doxorubicin cardiotoxicity and the control of iron metabolism: quinone-dependent and independent mechanisms. Methods Enzymol. 2004;378:340–361. doi: 10.1016/S0076-6879(04)78025-8. [DOI] [PubMed] [Google Scholar]
  • 18.Wallace KB. Adriamycin-induced interference with cardiac mitochondrial calcium homeostasis. Cardiovasc Toxicol. 2007;7:101–107. doi: 10.1007/s12012-007-0008-2. [DOI] [PubMed] [Google Scholar]
  • 19.Swain SM, Whaley FS, Gerber MC, Weisberg S, York M, Spicer D, et al. Cardioprotection with dexrazoxane for doxorubicin-containing therapy in advanced breast cancer. J Clin Oncol. 1997;15:1318–1332. doi: 10.1200/JCO.1997.15.4.1318. [DOI] [PubMed] [Google Scholar]
  • 20.Wojnowski L, Kulle B, Schirmer M, Schluter G, Schmidt A, Rosenberger A, et al. NAD(P)H oxidase and multidrug resistance protein genetic polymorphisms are associated with doxorubicin-induced cardiotoxicity. Circulation. 2005;112:3754–3762. doi: 10.1161/CIRCULATIONAHA.105.576850. [DOI] [PubMed] [Google Scholar]
  • 21.Rossi D, Rasi S, Franceschetti S, Capello D, Castelli A, De Paoli L, et al. Analysis of the host pharmacogenetic background for prediction of outcome and toxicity in diffuse large B-cell lymphoma treated with R-CHOP21. Leukemia. 2009;23:1118–1126. doi: 10.1038/leu.2008.398. [DOI] [PubMed] [Google Scholar]
  • 22.Minotti G, Recalcati S, Mordente A, Liberi G, Calafiore AM, Mancuso C, et al. The secondary alcohol metabolite of doxorubicin irreversibly inactivates aconitase/iron regulatory protein-1 in cytosolic fractions from human myocardium. FASEB J. 1998;12:541–552. doi: 10.1096/fasebj.12.7.541. [DOI] [PubMed] [Google Scholar]
  • 23.Olson RD, Mushlin PS, Brenner DE, Fleischer S, Cusack BJ, Chang BK, et al. Doxorubicin cardiotoxicity may be caused by its metabolite, doxorubicinol. Proc Natl Acad Sci U S A. 1988;85:3585–3589. doi: 10.1073/pnas.85.10.3585. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Weinstein DM, Mihm MJ, Bauer JA. Cardiac peroxynitrite formation and left ventricular dysfunction following doxorubicin treatment in mice. J Pharmacol Exp Ther. 2000;294:396–401. [PubMed] [Google Scholar]
  • 25.Clementi ME, Giardina B, Di Stasio E, Mordente A, Misiti F. Doxorubicin-derived metabolites induce release of cytochrome C and inhibition of respiration on cardiac isolated mitochondria. Anticancer Res. 2003;23:2445–2450. [PubMed] [Google Scholar]
  • 26.Hasinoff BB, Patel D, Wu X. The oral iron chelator ICL670A (deferasirox) does not protect myocytes against doxorubicin. Free Radic Biol Med. 2003;35:1469–1479. doi: 10.1016/j.freeradbiomed.2003.08.005. [DOI] [PubMed] [Google Scholar]
  • 27.Lyu YL, Kerrigan JE, Lin CP, Azarova AM, Tsai YC, Ban Y, et al. Topoisomerase-IIbeta mediated DNA double-strand breaks: implications in doxorubicin cardiotoxicity and prevention by dexrazoxane. Cancer Res. 2007;67:8839–8846. doi: 10.1158/0008-5472.CAN-07-1649. [DOI] [PubMed] [Google Scholar]
  • 28.Van Dalen EC, Michiels EM, Caron HN, Kremer LC. Different anthracycline derivates for reducing cardiotoxicity in cancer patients. Cochrane Database Syst Rev. 2010;5:CD005006. doi: 10.1002/14651858.CD005006.pub4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Robert J. Preclinical assessment of anthracycline cardiotoxicity in laboratory animals: predictiveness and pitfalls. Cell Biol Toxicol. 2007;23:27–37. doi: 10.1007/s10565-006-0142-9. [DOI] [PubMed] [Google Scholar]
  • 30.Platel D, Pouna P, Bonoron-Adele S, Robert J. Comparative cardiotoxicity of idarubicin and doxorubicin using the isolated perfused rat heart model. Anti-cancer Drugs. 1999;10:671–676. doi: 10.1097/00001813-199908000-00007. [DOI] [PubMed] [Google Scholar]
  • 31.Neef C, de Voogd-van der Straaten I. An interaction between cytostatic and anticonvulsant drugs. Clin Pharmacol Ther. 1988;43:372–375. doi: 10.1038/clpt.1988.45. [DOI] [PubMed] [Google Scholar]
  • 32.Cusack BJ, Tesnohlidek DA, Loseke VL, Vestal RE, Brenner DE, Olson RD. Effect of phenytoin on the pharmacokinetics of doxorubicin and doxorubicinol in the rabbit. Cancer Chemother Pharmacol. 1988;22:294–298. doi: 10.1007/BF00254234. [DOI] [PubMed] [Google Scholar]
  • 33.Colombo T, Zucchetti M, D’Incalci M. Cyclosporin A markedly changes the distribution of doxorubicin in mice and rats. J Pharmacol Exp Ther. 1994;269:22–27. [PubMed] [Google Scholar]
  • 34.Singhal SS, Sehrawat A, Sahu M, Singhal P, Vatsyayan R, Rao Lelsani PC, et al. Rlip76 transports sunitinib and sorafenib and mediates drug resistance in kidney cancer. Int J Cancer. 2010;126:1327–1338. doi: 10.1002/ijc.24767. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Gianni L, Salvatorelli E, Minotti G. Anthracycline cardiotoxicity in breast cancer patients: synergism with trastuzumab and taxanes. Cardiovasc Toxicol. 2007;7:67–71. doi: 10.1007/s12012-007-0013-5. [DOI] [PubMed] [Google Scholar]
  • 36.Kaye S, Merry S. Tumour cell resistance to anthracyclines: a review. Cancer Chemother Pharmacol. 1985;14:96–103. doi: 10.1007/BF00434344. [DOI] [PubMed] [Google Scholar]
  • 37.Germann UA. P-glycoprotein: a mediator of multidrug resistance in tumour cells. Eur J Cancer. 1996;32A:927–944. doi: 10.1016/0959-8049(96)00057-3. [DOI] [PubMed] [Google Scholar]
  • 38.Cole SP, Bhardwaj G, Gerlach JH, Mackie JE, Grant CE, Almquist KC, et al. Overexpression of a transporter gene in a multidrug-resistant human lung cancer cell line. Science. 1992;258:1650–1654. doi: 10.1126/science.1360704. [DOI] [PubMed] [Google Scholar]
  • 39.Young LC, Campling BG, Cole SP, Deeley RG, Gerlach JH. Multidrug resistance proteins MRP3, MRP1, and MRP2 in lung cancer: correlation of protein levels with drug response and messenger RNA levels. Clin Cancer Res. 2001;7:1798–1804. [PubMed] [Google Scholar]
  • 40.Singhal SS, Singhal J, Sharma R, Singh SV, Zimniak P, Awasthi YC, et al. Role of RLIP76 in lung cancer doxorubicin resistance: I. The ATPase activity of RLIP76 correlates with doxorubicin and 4-hydroxynonenal resistances in lung cancer cells. Int J Oncol. 2003;22:365–375. [PubMed] [Google Scholar]
  • 41.Burgess DJ, Doles J, Zender L, Xue W, Ma B, McCombie WR, et al. Topoisomerase levels determine chemotherapy response in vitro and in vivo. Proc Natl Acad Sci U S A. 2008;105:9053–9058. doi: 10.1073/pnas.0803513105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Pritchard KI, Messersmith H, Elavathil L, Trudeau M, O’Malley F, Dhesy-Thind B. HER-2 and topoisomerase II as predictors of response to chemotherapy. J Clin Oncol. 2008;26:736–744. doi: 10.1200/JCO.2007.15.4716. [DOI] [PubMed] [Google Scholar]
  • 43.Jacquet JM, Bressolle F, Galtier M, Bourrier M, Donadio D, Jourdan J, et al. Doxorubicin and doxorubicinol: intra- and inter-individual variations of pharmacokinetic parameters. Cancer Chemother Pharmacol. 1990;27:219–225. doi: 10.1007/BF00685716. [DOI] [PubMed] [Google Scholar]
  • 44.Torti FM, Bristow MR, Howes AE, Aston D, Stockdale FE, Carter SK, et al. Reduced cardiotoxicity of doxorubicin delivered on a weekly schedule: assessment by endomyocardial biopsy. Ann Intern Med. 1983;99:745–749. doi: 10.7326/0003-4819-99-6-745. [DOI] [PubMed] [Google Scholar]
  • 45.Lal S, Wong ZW, Sandanaraj E, Xiang X, Ang PC, Lee EJ, et al. Influence of ABCB1 and ABCG2 polymorphisms on doxorubicin disposition in Asian breast cancer patients. Cancer Sci. 2008;99:816–823. doi: 10.1111/j.1349-7006.2008.00744.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Jeong H, Herskowitz I, Kroetz DL, Rine J. Function-altering SNPs in the human multidrug transporter gene ABCB1 identified using a Saccharomyces-based assay. PLoS Genet. 2007;3:e39. doi: 10.1371/journal.pgen.0030039. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Conrad S, Kauffmann HM, Ito K, Deeley RG, Cole SP, Schrenk D. Identification of human multidrug resistance protein 1 (MRP1) mutations and characterization of a G671V substitution. J Hum Genet. 2001;46:656–663. doi: 10.1007/s100380170017. [DOI] [PubMed] [Google Scholar]
  • 48.Conrad S, Kauffmann HM, Ito K, Leslie EM, Deeley RG, Schrenk D, et al. A naturally occurring mutation in MRP1 results in a selective decrease in organic anion transport and in increased doxorubicin resistance. Pharmacogenetics. 2002;12:321–330. doi: 10.1097/00008571-200206000-00008. [DOI] [PubMed] [Google Scholar]
  • 49.Yin JY, Huang Q, Yang Y, Zhang JT, Zhong MZ, Zhou HH, et al. Characterization and analyses of multidrug resistance-associated protein 1 (MRP1/ABCC1) polymorphisms in Chinese population. Pharmacogenet Genomics. 2009;19:206–216. doi: 10.1097/FPC.0b013e328323f680. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Rajic V, Aplenc R, Debeljak M, Prestor VV, Karas-Kuzelicki N, Mlinaric-Rascan I, et al. Influence of the polymorphism in candidate genes on late cardiac damage in patients treated due to acute leukemia in childhood. Leuk Lymphoma. 2009;50:1693–1698. doi: 10.1080/10428190903177212. [DOI] [PubMed] [Google Scholar]
  • 51.Bains OS, Karkling MJ, Grigliatti TA, Reid RE, Riggs KW. Two nonsynonymous single nucleotide polymorphisms of human carbonyl reductase 1 demonstrate reduced in-vitro metabolism of daunorubicin and doxorubicin. Drug Metab Dispos. 2009;37:1107–1114. doi: 10.1124/dmd.108.024711. [DOI] [PubMed] [Google Scholar]
  • 52.Lal S, Sandanaraj E, Wong ZW, Ang PC, Wong NS, Lee EJ, et al. CBR1 and CBR3 pharmacogenetics and their influence on doxorubicin disposition in Asian breast cancer patients. Cancer Sci. 2008;99:2045–2054. doi: 10.1111/j.1349-7006.2008.00903.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Gonzalez-Covarrubias V, Zhang J, Kalabus JL, Relling MV, Blanco JG. Pharmacogenetics of human carbonyl reductase 1 (CBR1) in livers from black and white donors. Drug Metab Dispos. 2009;37:400–407. doi: 10.1124/dmd.108.024547. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Fan L, Goh BC, Wong CI, Sukri N, Lim SE, Tan SH, et al. Genotype of human carbonyl reductase CBR3 correlates with doxorubicin disposition and toxicity. Pharmacogenet Genomics. 2008;18:621–631. doi: 10.1097/FPC.0b013e328301a869. [DOI] [PubMed] [Google Scholar]
  • 55.Bains OS, Karkling MJ, Lubieniecka JM, Grigliatti TA, Reid RE, Riggs KW. Naturally occurring variants of human CBR3 alter anthracycline in-vitro metabolism. J Pharmacol Exp Ther. 2010;332:755–763. doi: 10.1124/jpet.109.160614. [DOI] [PubMed] [Google Scholar]
  • 56.Blanco JG, Leisenring WM, Gonzalez-Covarrubias VM, Kawashima TI, Davies SM, Relling MV, et al. Genetic polymorphisms in the carbonyl reductase 3 gene CBR3 and the NAD(P)H:quinone oxidoreductase 1 gene, NQO1 in patients who developed anthracycline-related congestive heart failure after childhood cancer. Cancer. 2008;112:2789–2795. doi: 10.1002/cncr.23534. [DOI] [PubMed] [Google Scholar]
  • 57.Lakhman SS, Ghosh D, Blanco JG. Functional significance of a natural allelic variant of human carbonyl reductase 3 (CBR3) Drug Metab Dispos. 2005;33:254–257. doi: 10.1124/dmd.104.002006. [DOI] [PubMed] [Google Scholar]
  • 58.Choi JY, Barlow WE, Albain KS, Hong CC, Blanco JG, Livingston RB, et al. Nitric oxide synthase variants and disease-free survival among treated and untreated breast cancer patients in a Southwest Oncology Group clinical trial. Clin Cancer Res. 2009;15:5258–5266. doi: 10.1158/1078-0432.CCR-09-0685. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Ross D, Siegel D. NAD(P)H:quinone oxidoreductase 1 (NQO1, DTdiaphorase), functions and pharmacogenetics. Methods Enzymol. 2004;382:115–144. doi: 10.1016/S0076-6879(04)82008-1. [DOI] [PubMed] [Google Scholar]
  • 60.Sandanaraj E, Lal S, Selvarajan V, Ooi LL, Wong ZW, Wong NS, et al. PXR pharmacogenetics: association of haplotypes with hepatic CYP3A4 and ABCB1 messenger RNA expression and doxorubicin clearance in Asian breast cancer patients. Clin Cancer Res. 2008;14:7116–7126. doi: 10.1158/1078-0432.CCR-08-0411. [DOI] [PubMed] [Google Scholar]
  • 61.Lal S, Wong ZW, Jada SR, Xiang X, Chen Shu X, Ang PC, et al. Novel SLC22A16 polymorphisms and influence on doxorubicin pharmacokinetics in Asian breast cancer patients. Pharmacogenomics. 2007;8:567–575. doi: 10.2217/14622416.8.6.567. [DOI] [PubMed] [Google Scholar]
  • 62.Bains OS, Karkling MJ, Lubieniecka JM, Grigliatti TA, Reid RE, Riggs KW. Naturally occurring variants of human CBR3 alter anthracycline in-vitro metabolism. J Pharmacol Exp Ther. 2010;332:755–763. doi: 10.1124/jpet.109.160614. [DOI] [PubMed] [Google Scholar]

RESOURCES