Topic Highlight Open Access
Copyright ©2014 Baishideng Publishing Group Inc. All rights reserved.
World J Gastroenterol. Aug 14, 2014; 20(30): 10316-10330
Published online Aug 14, 2014. doi: 10.3748/wjg.v20.i30.10316
FOLFOX/FOLFIRI pharmacogenetics: The call for a personalized approach in colorectal cancer therapy
Beatrice Mohelnikova-Duchonova, Pavel Soucek, Department of Toxicogenomics, National Institute of Public Health, 10042 Prague, Czech Republic
Beatrice Mohelnikova-Duchonova, Bohuslav Melichar, Department of Oncology, Palacky University Medical School and Teaching Hospital, 77525 Olomouc, Czech Republic
Author contributions: Mohelnikova-Duchonova B and Soucek P conceived the review; Mohelnikova-Duchonova B, Melichar B and Soucek P drafted the manuscript and performed its revisions.
Supported by Internal Grant Agency of the Czech Ministry of Health, No. NT14329-3; Research project Biomedreg, No. CZ.1.05/2.1.00/01.0030; Project from European Regional Development Fund, No. CZ.1.05/2.1.00/03.0076; and the project of the Palacky University, No. LF_2013_010
Correspondence to: Beatrice Mohelnikova-Duchonova, MD, PhD, Department of Oncology, Palacky University Medical School and Teaching Hospital, IP Pavlova 6, 77525 Olomouc, Czech Republic. d.beatrice@seznam.cz
Telephone: +420-775-270283
Received: October 29, 2013
Revised: March 5, 2014
Accepted: April 15, 2014
Published online: August 14, 2014

Abstract

While 5-fluorouracil used as single agent in patients with metastatic colorectal cancer has an objective response rate around 20%, the administration of combinations of irinotecan with 5-fluorouracil/folinic acid or oxaliplatin with 5-fluorouracil/folinic acid results in significantly increased response rates and improved survival. However, the side effects of systemic therapy such as myelotoxicity, neurotoxicity or gastrointestinal toxicity may lead to life-threatening complications and have a major impact on the quality of life of the patients. Therefore, biomarkers that would be instrumental in the choice of optimal type, combination and dose of drugs for an individual patient are urgently needed. The efficacy and toxicity of anticancer drugs in tumor cells is determined by the effective concentration in tumor cells, healthy tissues and by the presence and quantity of the drug targets. Enzymes active in drug metabolism and transport represent important determinants of the therapeutic outcome. The aim of this review was to summarize published data on associations of gene and protein expression, and genetic variability of putative biomarkers with response to therapy of colorectal cancer to 5-fluorouracil/leucovorin/oxaliplatin and 5-fluorouracil/leukovorin/irinotecan regimens. Gaps in the knowledge identified by this review may aid the design of future research and clinical trials.

Key Words: Colorectal cancer, Chemotherapy, 5-Fluorouracil, Oxaliplatin, Irinotecan

Core tip: 5-fluorouracil/leucovorin combined with oxaliplatin (FOLFOX) and irinotecan (FOLFIRI) represent the most effective chemotherapy regimens for colorectal carcinoma patients with distant metastases. Pharmacogenetics represents a promising strategy for the individualization of therapy, including identification of patients at increased risk of toxicity. This review summarizes contemporary knowledge about associations of gene and protein expression and genetic variability of putative biomarkers for the response of colorectal cancer to FOLFOX/FOLFIRI regimens. From the published data reviewed it is obvious that the problem is highly complex and the ultimate profile of the drug-sensitive or resistant patient will most probably be jointly defined by genetic, epigenetic, intracellular, extracellular, and extrinsic factors.



INTRODUCTION

Colorectal carcinoma is one of the most common causes of cancer mortality. While localized tumors are amenable to curative surgical resection, the curative potential of surgery in patients with metastatic disease is limited. Intuitively, the best approach to treat metastatic (systemic) disease is the systemic administration of therapy. Systemic therapy may also be used as adjuvant treatment after the resection of primary tumors in patients who have no evidence of metastatic disease, but are suspected to harbor microscopic metastases. Among the modalities of systemic treatment, chemotherapy has been the most widely used in patients with colorectal carcinoma, and in recent years several targeted agents have complemented the therapeutic armamentarium in patients with metastatic colorectal carcinoma.

Over the past half-century, fluoropyrimidines have constituted the backbone of chemotherapeutic regimens in colorectal carcinoma. Early randomized clinical trials demonstrated that the administration of fluoropyrimidine-based chemotherapy results in statistically significant prolongation of survival in patients with metastatic colorectal carcinoma[1,2]. Among fluoropyrimidines, 5-fluorouracil (5-FU) has been the most commonly used agent. Prospective studies have demonstrated the benefit of the administration of 5-FU in an infusional regimen and in combination with folinic acid[3]. In addition, in patients with isolated liver metastases the benefit of hepatic arterial infusion of fluoropyrimidines has also been demonstrated[4].

The next generation of regimens for the treatment of metastatic colorectal carcinoma has been introduced with the advent of two cytotoxic agents, irinotecan, a topoisomerase I inhibitor, and oxaliplatin, a platinum derivative, in the late 1990s. The activity of irinotecan and oxaliplatin was first demonstrated in patients that failed on fluoropyrimidines. While 5-FU alone has an objective response rate around 20%[3,5], the combinations of irinotecan with 5-FU/folinic acid and oxaliplatin with 5-FU/folinic acid result in significantly increased response rates and improved survival[6-8]. The activity of 5-FU/leucovorin combined with oxaliplatin (FOLFOX) or irinotecan (FOLFIRI) in the first-line treatment of metastatic colorectal carcinoma is comparable[9].

Another major step forward in the systemic management of colorectal carcinoma was the advent of monoclonal antibodies targeting vascular endothelial growth factor (VEGF) pathway or epidermal growth factor receptor (EGFR)[10-14]. Currently available targeted agents active in metastatic colorectal carcinoma include the anti-VEGF antibody bevacizumab, the anti-EGFR antibodies cetuximab and panitumumab, and the anti-VEGF agents aflibercept and regorafenib. These agents are active, mostly in combination with cytotoxic drugs, both in the first-line of therapy as well as in previously treated patients with metastatic colorectal cancer.

Besides causing morbidity and occasional mortality, the side effects of systemic therapy have a major impact on the quality of life of the patients. Common to fluoropyrimidines, irinotecan and oxaliplatin is myelotoxicity, the administration of fluoropyrimidines and irinotecan is frequently accompanied by gastrointestinal toxicity[15], while neurotoxicity regularly complicates the administration of oxaliplatin[16,17]. The administration of targeted agents is also not devoid of side effects that may be very annoying, e.g., skin toxicity associated with the administration of anti-EGFR agents[18] or hypertension after anti-VEGF therapy[11]. Liver toxicities that accompany the administration of agents used in colorectal carcinoma encompass non-alcoholic fatty liver disease after 5-FU, sinusoid obstruction syndrome after oxaliplatin or steatohepatitis after irinotecan[19]. These toxicities could result in postoperative complications in patients undergoing subsequent liver resection[19].

The role of pharmacogenetics in the management of patients with colorectal carcinoma has long been neglected. The significance of genetic polymorphisms of enzymes responsible for fluoropyrimidine degradation, e.g., dihydropyrimidine dehydrogenase, has been known for some time, but the assessment of these biomarkers has still not found routine use[20]. With the advent of targeted therapy, the presence or absence of RAS mutations has been identified as a predictive biomarker of efficacy of anti-EGFR antibodies[21].

PHARMACOGENETICS OF FOLFOX/FOLFIRI REGIMENS IN COLORECTAL CANCER

In general, the use of chemotherapy to treat cancers is limited by the inter-individual variability in drug response and by the development of resistance. Anticancer drugs are metabolized predominantly in liver, subsequently transported in conjugated or unconjugated form to the tumor microenvironment, where drug uptake/efflux transporters modulate intracellular levels of the drug or its active metabolites. The efficacy of anticancer drugs in tumor cells is dependent on the effective concentrations and on the presence and quantity of the drug targets. There are marked inter-individual differences in expression levels and activities of enzymes modifying efficacy, as well as toxicity, of anticancer drugs. From this point of view it seems obvious that biomarkers enabling prediction of optimal type, combination and dose of drugs for each patient may exist, and their use for individualization of therapy is envisaged. Such individualization would be highly cost-effective and socially desirable because of the prolonged survival and improved quality of life of large number of cancer patients.

However, this task is very complicated because of numerous factors determining the final functional phenotype of enzymes responsible for the drug metabolism, transport and targets. On the intracellular level, genotype vs phenotype relations must be considered along with epigenetic factors, such as methylation of regulatory DNA elements, histone acetylation, the presence of micro RNAs and other non-coding RNA species, protein-protein and DNA/RNA-protein interactions. Extracellular factors may include immune response or hormonal balance that could determine the expression pattern of intracellular enzymes, drug-drug and drug-environmental/alimentary interactions and other as yet unknown factors. Last, but not least, enzymes responding to drug administration may be induced to a high extent by repeated doses of the drug and the study of this phenomenon in vivo is very difficult. Investigation of such a complex system where the roles of a number of factors remain unknown is significantly limited by the currently available equipment and empirical approaches. Thus, from many published studies, very few of the most pertinent biomarkers emerged, which should be verified by upcoming controlled prospective clinical trials. This review summarizes the most promising predictive biomarkers for FOLFOX/FOLFIRI regimens in advanced colorectal cancers and highlights potential research trends.

5-FU

5-FU belongs to fluoropyrimidine drugs (Figure 1), which are widely used in the therapy of gastrointestinal cancers including colorectal cancer. Research on 5-FU pharmacogenetics (and pharmacogenomics) focused mainly on interindividual differences in 5-FU pharmacokinetics and genetic alterations in genes encoding transmembrane transporters and 5-FU-metabolizing enzymes, such as dihydropyrimidine dehydrogenase (DPD/DPYD, OMIM: 612779), thymidine phosphorylase (TYMP, OMIM: 131222), thymidine kinase 1 (TK1, OMIM: 188300), uridine monophosphate synthetase (UMPS/OPRT, OMIM: 613891)[22] (Table 1).

Table 1 Putative biomarkers of efficacy and/or toxicity of 5-fluorouracil in colorectal cancer.
GeneTranscriptProteinSNPEffectRef.
SLC29A1-High-Inferior clinical responsePhua et al[40] 2013
ABCC11Low--Inferior response and shorter DFSHlavata et al[28] 2012
TYMPHighHigh-Longer DFS and OSMeropol et al[48] 2006
High-Better clinical responseSadahiro et al[51] 2012
-High cytoplasmic-Longer OSMitselou et al[52] 2012
High--pCRChiorean et al[50] 2012
-No association-DFS and OSCiaparrone et al[42] 2006
No association--OSLassmann et al[43] 2006
-No association-OSSoong et al[44] 2008
No association--Clinical responseVallbohmer et al[45] 2007
No association--PFS and OSKoopman et al[49] 2009
UMPSNo association--Longer OSYanagisawa et al[54] 2007
-High-Longer OSTokunaga et al[55] 2007
-High (tumor cells)-Shorter OSKoopman et al[56] 2009
-High (stromal cells)-Longer OSKoopman et al[56] 2009
DPYDLow--Longer OSYanagisawa et al[54] 2007
Low--Longer OSVallböhmer et al[45] 2007
-Low-Longer OSTokunaga et al[55] 2007
-Low1-Longer OS1Koopman et al[56] 2009
-Low-Longer OSCiaparrone et al[42] 2007
-No association-DFSWestra et al[59] 2005
--Rs3918290Higher toxicityCaudle et al[27] 2013
Dose reduction recommendedSwen et al[62] 2011
--Rs55886062Higher toxicityCaudle et al[27] 2013
Dose reduction recommendedSwen et al[62] 2011
--Rs67376798Higher toxicityCaudle et al[27] 2013
Dose reduction recommendedSwen et al[62] 2011
TYMS-No association-Clinical responseJennings et al[57] 2012
--Rs45445694Lower protein expression, clinical benefit and higher toxicity improvedJennings et al[57] 2012
Figure 1
Figure 1 Chemical structures of components of 5-fluorouracil/leucovorin combined with oxaliplatin and irinotecan regimens.

A low accumulation of 5-FU in cancer cells may be caused by altered membrane transport, namely drug efflux, mediated mainly by ATP-binding cassette (ABC) transporters, or reduced drug uptake into the cancer cells, mediated mainly by solute carrier (SLC) transporters[23,24].

5-FU is phosphorylated to FdUMP by TYMP and TK1. FdUMP then inhibits an important enzyme for nucleotide synthesis - thymidylate synthase (TYMS, OMIM: 188350)[22]. 5-FU is also indirectly phosphorylated by UMPS via fluorouridine monophosphate (FUMP) to fluorouridine diphosphate (FUDP) and then converted by ribonucleotide reductase (RRM1 and 2, OMIM: 180410, OMIM: 180390, respectively) to FdUMP. The nucleotide diphosphate kinase (NME1/NM23, OMIM: 156490)-formed fluorodeoxyuridine triphosphate (FdUTP) incorporates into DNA and fluorouridine triphosphate (FUTP) incorporates into RNA and both cause chain termination[25]. It appears that the spectrum of 5-FU metabolites depends on the administration schedule, as bolus treatment favors RNA damage by FUTP and continuous regimen favors DNA damage by FdUTP[26]. DPYD catalyzes inactivation of 5-FU into inactive dihydrofluorouracil, mostly in the liver[27].

Biomarkers of 5-FU chemoresistance in colorectal cancer

Deregulation of ABC transporters in CRC tumors compared to non-malignant colon tissue has been reported recently[28]. Interestingly, the best-known ABC transporter, ABCB1 coding P-glycoprotein (OMIM: 171050), has not been shown to modify the sensibility of human-derived esophageal carcinoma cell lines to 5-FU[29]. Moreover, lack of relationship between the ABCB1 protein or transcript expression, genotype and long-term prognosis of patients treated by 5-FU was reported[28,30]. Transporters from the ABCC family can collectively confer resistance to anticancer drugs and their conjugated metabolites, platinum compounds, folate antimetabolites, nucleoside and nucleotide analogues in vitro[31]. In particular, the expression of ABCC2 (OMIM: 601107), ABCC3 (OMIM: 604323), ABCC4 (OMIM: 605250), ABCC5 (OMIM: 605251), ABCC6 (OMIM: 603234) and ABCC11 (OMIM: 607040) induced resistance to 5-FU in vitro[29,32-34]. Nevertheless, the results obtained using cell line models treated by the studied drug for a long time may not reflect the real situation in such a heterogeneous entity as a colorectal tumor. In breast cancer patients treated with neoadjuvant chemotherapy, ABCA1 (OMIM: 600046), ABCA12 (OMIM: 607800), ABCB6 (OMIM: 605452), ABCC5, ABCC11 and ABCC13 (OMIM: 608835) transcript levels were downregulated in patients with a complete pathological response compared with patients with residual disease[35]. Oguri et al[36] discovered that ABCC11 expression is induced by 5-FU and that ABCC11 is directly involved in resistance by the efflux transport of the active metabolite FdUMP in human small-cell lung cancer cell lines in vitro. High expression of ABCC11 has been associated with significantly better response and longer disease-free interval in colorectal cancer patients treated by first line 5-FU-based chemotherapy in either the palliative or adjuvant setting[28].

In humans, there are two major families of SLC transporters that transport nucleoside analogs including 5-FU: SLC28A (human concentrative transporters, namely SLC28A1, SLC28A2 and SLC28A3, OMIM: 606207, 606208, 608269, respectively) and SLC29A (human equilibrative transporters, namely SLC29A1, SLC29A2, SLC29A3 and SLC29A4, OMIM: 602193, 602110, 612373, 609149, respectively)[37,38]. A pilot study has indicated that colorectal tissue specimens from tumors that were resistant to 5-FU in an in vitro cell viability assay had higher expression of SLC29A1 mRNA[39]. These data were recently corroborated by another study showing correlation between high pre-treatment intratumoral SLC29A1 protein levels with worse clinical response to 5-FU[40]. The predictive significance of SLC29A1 has been studied extensively in pancreatic cancer. In contrast to colorectal cancer, the majority of studies on patients with pancreatic cancer have suggested that high SLC29A1 expression may be predictive of improved survival in patients treated with gemcitabine, but not for patients treated by 5-FU[41].

The issue of predictive value of phenotype and/or genotype of drug transporters remains open, mainly because of the complex nature of this phenomenon. No study investigated either the balance between uptake/efflux (status of both SLC and ABC transporters in the same cohort of patients) or the relationship between of drug transport and subsequent 5-FU metabolism and/or targets. The majority of available studies addressed single or isolated groups of biomarkers and contradictory results were often obtained.

Inside the tumor cells, 5-FU is converted to its monophosphate by TYMP, an angiogenic factor also known as platelet-derived endothelial cell growth factor (PD-ECGF). The question is whether TYMP acts as a predictor of poor prognosis because of its neoangiogenic activity in tumors or whether it may predict good prognosis because of activation of 5-FU. This dual role of TYMP may be one of the reasons for contradictory results of studies aiming to evaluate the role of TYMP as a prognostic biomarker. No association has been found in smaller studies with colorectal cancer patients treated by adjuvant 5-FU therapy[42-45]. TYMP protein expression also was not associated with disease-free survival and overall survival of the advanced colorectal cancer patients in other studies[46,47]. On the other hand, looking at the studies with capecitabine, high expression of TYMP mRNA and protein was associated with longer overall and disease-free survival in patients with advanced colorectal cancer treated by capecitabine plus irinotecan as in the first-line setting[48]. In 566 advanced colorectal cancer patients treated with capecitabine, irinotecan and oxaliplatin enrolled in the phase III CAIRO study, TYMP was neither a predictive nor prognostic factor[49]. In a phase II trial of neoadjuvant capecitabine plus irinotecan and radiation therapy for locally advanced rectal cancer (n = 22 patients), high tumor TYMP transcript expression was associated with complete pathological response[50]. Similarly, right-sided colon tumors with high TYMP transcript levels had higher response rate to neoadjuvant chemotherapy with oral fluoropyrimidine tegafur[51]. TYMP protein expression has been observed not only in tumor cells, but also in the stroma and in endothelium and tumor-associated macrophages[52]. Previously, higher TYMP expression was reported in stromal and tumor cells compared with normal tissue[53]. TYMP expression levels in the cytoplasm of tumor cells and stromal cells correlated with vascular endothelial growth factor (VEGF, OMIM: 192240) expression by tumor cells and vessels. However only high cytoplasmic expression of TYMP was associated with longer overall survival of these patients[52]. This study further supported the view concept of a differential role of TYMP in tumor cells compared with the stroma.

The reported data regarding UMPS and uridine phosphorylase (UPP1, OMIM: 191730) are again conflicting. UMPS transcript levels did not correlate with overall survival of colorectal cancer patients[54], but Tokunaga et al[55] reported that high expression of the UMPS protein is associated with longer overall survival of patients with advanced colorectal cancer. On the other hand, high expression of UMPS in tumor cells was an unfavorable prognostic parameter for overall survival in the CAIRO study; however, the opposite effect was observed in stromal cells, with high stromal cell UMPS expression was being associated with favorable prognosis[54]. Moreover, UPP1 bypasses UMPS and is suggested to be the key enzyme in conversion of 5-FU to the active metabolite FUMP[56].

It becomes obvious that further studies on the role of metabolic pathway of 5-FU in the treatment outcome of patients should also consider the localization of biomarker expression within the various cell types and even intracellular components.

A recent meta-analysis of 39 studies that investigated 2402 patients for the most commonly studied polymorphic biomarkers TYMS (rs45445694) and MTHFR (rs1801133) revealed that the TYMS polymorphism is significantly associated with protein expression, clinical benefit and adverse effects[57]. However, the authors concluded that the association between treatment effect and TYMS genotype and subsequently protein level is so small that it is of limited clinical relevance.

5-FU is deactivated mainly by DPYD, and this enzyme has also been shown to be an independent chemosensitivity predictive factor in vitro[58]. Low expression of both DPYD mRNA[54] and protein[42,47,55] was associated with longer overall survival or disease free survival in numerous studies on colorectal cancer patients treated by 5-FU in the adjuvant or advanced setting. Soong et al[44] observed only a trend of shorter overall survival in patients with low DPYD protein expression. On the other hand, no association of DPYD mRNA or protein levels with disease-free or overall survival was observed in other studies[45,59]. Retrospective analysis of DPYD protein expression in well-defined groups of patients in the phase III randomized CAIRO study did not confirm the predictive value of DPYD[49], except for the subgroup of patients treated with the combination of capecitabine plus irinotecan.

The B-CAST multicenter, prospective cohort study aims to analyze TYMP, DPYD and UMPS in a group of more than 2000 patients with stage III colon cancer treated with adjuvant 5-FU-based regimens. The results of this first prospective clinical trial studying predictive biomarkers of 5-FU will hopefully define the effect of these three enzymes on treatment outcome[60].

Biomarkers of 5-FU toxicity in colorectal cancer patients

The expression of genes involved in 5-FU metabolism mentioned above may affect not only drug efficacy and resistance, but also toxicity (Table 1). However, data about the majority of these genes and their association with 5-FU toxicity are inconsistent and cannot be applied in clinical practice. The only exception is DPYD, for which evidence for the dosing recommendations comes from two large prospective studies, small studies and case studies[27]. Patients with DPYD deficiency treated with fluoropyrimidines suffered from severe toxicity, including mucositis and diarrhea, myelosuppression, neurotoxicity and hand-food syndrome[61]. The United States Food and Drug Administration (FDA) has added statements to the drug labels for 5-FU that contraindicate its use in DPYD enzyme deficient individuals. The Dutch Pharmacogenetics Working group and Clinical Pharmacogenetics Implementation Consortium (CPIC) recommends the use of alternative drugs for heterozygous carriers of a decreased-activity allele[27,62] such as DPYD*2A (rs3918290), DPYD*13 (rs55886062) or rs67376798. For heterozygous carriers, it is recommended to start with at least a 50% reduction of initial dose and re-adjust dosing according to patient’s tolerability or pharmacokinetic tests[27]. Perhaps because of a number of other so far uncharacterized SNPs in DPYD, the presence of these variants does not always result in toxicity and the results of available studies are not consistent and have not been replicated.

In summary, the positive predictive value of DPYD*2A and DPYD*13 variants to predict a severe toxicity (grade III and IV) is 62% and negative predictive value is 95%[63,64]. Thus, the absence of these DPYD variants does not eliminate the risk of high-grade toxicity caused by additional rare variants in DPYD or from other factors including genetic, epigenetic, environmental or alimentary factors. Moreover, the combination of 5-FU with other anticancer drugs or bolus administration of 5-FU may further increase the risk of severe toxicity in heterozygous carriers[27,64].

The individual risk/benefit ratio should be estimated for all patients because for some patients, the potential benefit may still exceed clinically tolerable toxicity.

OXALIPLATIN

Oxaliplatin (trans-1-diaminocyclohexane oxalateplatinum, Figure 1) is a third-generation platinum derivative that is widely used in the therapy of colorectal cancer. Compared with cisplatin, oxaliplatin has enhanced water solubility[65]. Research into the pharmacogenetics of oxaliplatin focused mainly on interindividual differences in oxaliplatin pharmacokinetics and genetic alterations in genes coding ABC/SLC transporters, DNA damage repair machinery, such as excision cross-complementing genes (ERCC1, ERCC2, OMIM: 126380, OMIM: 126340, respectively) and X-ray repair cross-complementing protein 1 (XRCC1, OMIM: 194360), and conjugating enzymes glutathione S-transferases (GSTM1, GSTP1, GSTT1, OMIM: 138350, OMIM: 134660, OMIM: 600436, respectively)[66] (Table 2).

Table 2 Putative biomarkers of efficacy and/or toxicity of oxaliplatin in colorectal cancer.
GeneTranscriptProteinSNPEffectRef.
ATP7BLowLow-Longer PFSMartinez-Balibrea et al[78] 2009
ABCB1No association--DFS and OSHlavata et al[28] 2012
--AG genotype in rs1045642Longer DFSWu et al[79] 2013
--AA genotype in rs1128503Longer OSWu et al[79] 2013
--AA-GG-TT haplotype in rs1045642-rs1128503-rs2032582Shorter PFS and OSWu et al[79] 2013
ERCC1High--Shorter OSShirota et al[80] 2001
High--Shorter OSGrimminger et al[81] 2011
-High-Shorter DFS and OSHuang et al[82] 2013
-No association-PFS and OSKoopman et al[49] 2009
--GG genotype in rs11615Longer OSHuang et al[83] 2011
--Rs11615No association with OSMartinez-Balibrea et al[84] 2011
ERCC2-No associationDFS and OSHuang et al[110] 2013
--Variant allele in rs1052559Shorter OSLe Morvan et al[86] 2007
XRCC1-No associationDFS and OSHuang et al[82] 2013
--CC genotype in rs25487Longer OSHuang et al[83] 2011
GSTM1 deletion--One copyLonger OSFunke et al[87] 2010
--Null genotypeHigher neutropeniaMcLeod et al[90] 2010
GSTP1--AA genotype in rs1695Inferior responseKumamoto et al[88] 2013
--Rs1695No associationFunke et al[87] 2010
--Rs1695No associationHuang et al[85] 2008
--Rs1695No associationLe Morvan et al[86] 2011
--AA genotype in rs1695Higher neurotoxicityMcLeod et al[90] 2010
--Rs1695No association with toxicityPeng et al[91] 2013
GSTT1--DeletionNo associationFunke et al[87] 2010
--DeletionNo associationHuang et al[85] 2008
--DeletionNo associationLe Morvan et al[86] 2011

The principal mechanism of action of oxaliplatin is inhibition of DNA synthesis in cancer cells by the formation of crosslinks in DNA. There is no evidence of cytochrome P450-mediated metabolism in vitro. Inactivation of reactive oxaliplatin species is mediated by conjugation to glutathione, which is mediated by GSTs. DNA damage by oxaliplatin is repaired mainly by nucleotide excision repair (NER), base excision repair (BER) and by replicative bypass[67]. Interestingly, the mismatch repair complex important for resistance to other platinum drugs has not been shown as crucial for oxaliplatin resistance[68].

Oxaliplatin is transported mainly by the uptake transporters SLC22A1/OCT1 (OMIM: 602607) and SLC22A2/OCT2 (OMIM: 602608), which play a critical role not only in cellular uptake but also in the associated cytotoxicity[69]. Other important transporters are human copper transporters (SLC31A1 and 2 OMIM: 603085 and 603088, respectively) that mediate cellular uptake of oxaliplatin and P-type ATPases ATP7A and 7B (OMIM: 300011 and 606882, respectively) that promote efflux of oxaliplatin and its sequestration into subcellular compartments[70,71].

ABCC2, ABCC4 and ABCC5 transporters are potentially involved in disposition of platinum compounds[72,73]. However, the data on clinical implications of the genotype and/or phenotype of these transporters with regard to platinum efficacy or toxicity are scarce and inconclusive[71].

GSTP1 is involved in the detoxification of cisplatin by the formation of cisplatin-glutathione adducts[74] and, consequently, the putative role of GSTs in resistance to platinum compounds is generally accepted. The suggested link between GSTs and the MAP kinase pathway may also contribute to the common mechanisms of resistance towards anticancer drugs[75].

Biomarkers of chemoresistance to oxaliplatin in colorectal cancer

Enhanced expression of members of ABCC family of efflux transporters can lead to a decreased cellular glutathione level and thus indirectly cause decrease of oxaliplatin inactivation[76]. Overexpression of ABCC2 and ABCG2 (OMIM: 603756) resulted in increased activity of oxaliplatin in vitro[77]. Oxaliplatin is administered in combination with 5-FU, and 5-FU significantly suppressed ATP7B and SLC22A2 and simultaneously increased ABCC2 mRNA expression[77]. The synergistic action of these two drugs on transport direction has been demonstrated in vitro. Low ATP7B mRNA and protein expression is associated with longer time-to-progression of colorectal cancer patients receiving oxaliplatin-based chemotherapy[78]. The role of ABCB1/P-glycoprotein in the oxaliplatin pathway has not yet been proven. Lack of association of ABCB1 transcript level in tumors with therapy outcome of colorectal cancer patients treated by FOLFOX has been reported recently[28]. However, ABCB1 polymorphisms were shown to be significantly associated with survival of colorectal cancer patients treated by oxaliplatin[79]. Carriers of the AG genotype in rs1045642 had significantly longer time to recurrence than AA homozygotes, and carriers of AA genotype in rs1128503 had better overall survival compared to GG homozygotes. Moreover, carriers of the AA-GG-TT haplotype constructed from rs1045642-rs1128503-rs2032582 polymorphisms had an inferior progression-free and overall survival compared to carriers of other haplotypes[79].

Excision nucleases such as ERCC1 and ERCC2 play a major role in the repair of DNA adducts in tumor cells after chemotherapy. Thus, in theory, low ERCC1 gene expression leading to a decreased DNA repair should be a positive predictive factor of therapeutic effect of oxaliplatin. High ERCC1 mRNA expression was associated with shorter overall survival in patients treated with oxaliplatin-based chemotherapy[80,81]. Overexpression of ERCC1 protein has been shown to represent an independent predictor of early failure of adjuvant therapy by FOLFOX and short disease-free and overall survival[80]. In contrast to the above-mentioned results, the prognostic importance of ERCC1 protein expression has not been observed in the phase III CAIRO trial (n = 506)[49]. In another study, neither tumor ERCC2 nor XRCC1 protein expression was associated with overall survival of patients treated by adjuvant FOLFOX therapy[82].

Carriers of the GG genotype in rs11615 of ERCC1 were reported to have better progression-free and overall survival than carriers of the A allele[83] after FOLFOX therapy, but this was not been confirmed in another study[84]. The latter study also found no association of rs13181 in another DNA repair gene ERCC2 with the survival of colorectal patients treated by irinotecan. In an earlier study, Huang et al[85] observed a significant association of ERCC2 polymorphism rs13181 with increased risk of early relapse in patients with metastatic colorectal cancer treated by irinotecan-based chemotherapy. Carriers of the variant allele in ERCC2 polymorphism rs1052559 had shorter disease-free and overall survival after treatment by oxaliplatin than wild type carriers[86].

XRCC1 is involved in repair of DNA single-strand breaks formed by exposure to ionizing radiation and alkylating agents. Carriers of CC genotype in rs25487 of XRCC1 had better progression-free and overall survival after treatment by FOLFOX4 than T allele carriers[83].

GSTM1 DNA copy number was inversely associated with survival in colorectal cancer patients treated with chemotherapy[87]. Mortality was significantly reduced in patients with one GSTM1 copy (HR = 0.45, 95%CI: 0.23-0.90, P = 0.02) and non-significantly reduced in those with the null genotype (HR = 0.67, 95%CI: 0.35-1.27, P = 0.22) compared with carriers of two copies[87]. Neither functional GSTP1 polymorphism rs1695 nor GSTT1 deletion (null) were associated with survival of colorectal cancer patients in earlier studies[83,84,86,87]. However, a recent study reported that FOLFOX6-treated metastatic colorectal cancer patients (n = 63) with the GSTP1-rs1695 AA genotype had inferior responses to the treatment compared with G allele carriers[88].

Biomarkers of toxicity of oxaliplatin in colorectal cancer patients

A number of studies, including prospective clinical trials, provided conflicting results regarding the impact of polymorphisms in genes related to oxaliplatin mechanism of action and toxicity. No clear recommendations are currently ready for clinical practice. ERCC1 rs11615 and GSTP1 rs1695 are the most studied polymorphisms in relation to neutropenia. Carriers of the CC genotype in XRCC1 rs25487 had decreased the risk of neuropathy in colorectal cancer patients treated by adjuvant FOLFOX[89]. Carriage of the AA genotype in GSTP1 rs1695 predisposed patients with the TT genotype treated by FOFOX in a large N9741 clinical study to increased neurotoxicity, but the meta-analysis provided by McLeod et al[90] and Peng et al[91] (2013) showed no association between GSTP1 polymorphism rs1695 and the development of neurotoxicity (Table 2).

IRINOTECAN

Irinotecan is a camptothecin analog (Figure 1). Irinotecan is converted to an active metabolite, SN-38, by carboxylesterases CES1 and CES2 (OMIM: 114835 and 605278, respectively)[92]. SLCO1B1 (OMIM: 604843) mediates the uptake of irinotecan in hepatocytes[93]. SN-38 is transported by ABC transporters, namely by ABCC1, ABCC2, ABCB1, and ABCG2[94-97]. Inside a cell, the active metabolite SN-38 inhibits topoisomerase I and subsequently stalls both DNA replication and transcription[98].

SN-38 is further metabolized by glucuronidation by uridine diphosphate glycosyltransferase 1A1 (UGT1A1, OMIM: 191740) to its inactive glucuronide conjugate, SN-38G[99]. An alternative pathway of irinotecan inactivation is oxidation, mediated by members of cytochrome P450 3A subfamily[100]. Research on the pharmacogenetics of irinotecan focused mainly on interindividual differences in genetic alterations of genes coding transmembrane transporters and irinotecan-metabolizing enzymes, such as UGT1A1[98] (Table 3).

Table 3 Putative biomarkers of efficacy and/or toxicity of irinotecan in colorectal cancer.
GeneTranscriptProteinSNPEffectRef.
ABCB1--rs11288503Decreased clearance of irinotecanMathijssen et al[103] 2003
ABCB1--rs112503T-rs2032582T-rs1045642T haplotypeHigher levels of SN-38Sai et al[104] 2010
--rs112503T-rs2032582T-rs1045642T haplotypeShorter OSGlimelius et al[106] 2010
ABCG2--rs7699188Associate with RRDe Mattia et al[105] 2013
--GG genotype in rs425215Higher GI toxicityDi Martino et al[117] 2011
ABCC2--CC genotype in rs717620Longer PFSAkiyama et al[109] 2012
--CC genotype in rs562Higher GI toxicityDi Martino et al[117] 2011
SLCO1B1--GA/AA genotype in rs2306283Higher RRHuang et al[110] 2013
--GA genotype in rs 2306283Lower GI toxicityDi Martino et al[117] 2011
SLC19A1--GG genotype in rs1051266Higher RRHuang et al[82] 2013
UGT1A1--28 allele in rs81753471No association with RRDias et al[115] 2012
--28 allele in rs81753471Lower RRMarcuello et al[114] 2011
--28 allele in rs81753471Reduced activity of UGT1A1Swen et al[62] 2011
--28 allele in rs81753471Increased toxicity, dose reduction recommendedSwen et al[62] 2011
Biomarkers of irinotecan chemoresistance in colorectal cancer

Deficiency of the Abcc4 protein in Abcb1a/b; Abcg2 (-/-) mice in vivo model significantly increased the brain concentration of all camptothecin analogs, suggesting a possible role of ABCC4 in irinotecan efflux[101]. Interestingly, 5-FU significantly decreased the expression of ABCC2, ABCB1, and ABCG2 in the small intestine and increased the concentration of SN-38 in the blood of rats[102]. Thus, a synergistic action on efflux transporters of these two drugs used in FOLFIRI regimen has been described in the rat in vivo model. Colorectal tumors had significantly downregulated ABCB1, ABCC4 and ABCG2 transcripts compared with non-malignant tissues from the same patients before any chemotherapy[28]. Therefore, it appears that colorectal tumors have favorable expression profiles of ABC transporters active in irinotecan efflux.

ABCB1 polymorphism rs1128503 has been associated with a decreased clearance of irinotecan in cancer patients (mostly with gastrointestinal malignancies)[103]. Higher SN-38 levels in carriers of the *2 haplotype, which harbors 1236C>T, 2677G>T and 3435C>T, in ABCB1 (rs1128503T-rs2032582T-rs1045642T) among cancer patients receiving irinotecan have been observed[104]. On the other hand, ABCB1 polymorphism rs2032582 was not associated with pharmacokinetic parameters of metastatic colorectal cancer patients receiving first-line FOLFIRI treatment[105]. The association of carrying the most common ABCB1 haplotype (rs1128503T-rs2032582T-rs1045642T haplotype) with shorter overall survival was also reported[106]. The ABCG2 rs2231142 T-allele reduced expression in comparison with the GG genotype and led to resistance towards irinotecan in cancer cell lines[107]. 15622C>T and rs7699188 variants of ABCG2 were associated with the response rate of colorectal cancer patients treated with FOLFIRI[105]. The haplotype ABCC2*2 (rs717620C-rs2273697A-rs3740066C) was associated with lower irinotecan clearance in Caucasian cancer patients[108]. Akiyama et al[109] analyzed ABCC2 polymorphisms in a well-defined cohort of Japanese colorectal cancer patients who harbored UGTA1*1/*1, *1/*6, or *1/*28 genotypes, which are associated with similar irinotecan pharmacokinetics and responses to FOLFIRI. The ABCC2 rs717620 CC genotype has been associated with a higher response rate and longer progression-free survival, followed by CT and TT genotypes[109].

Genotype GA/AA of SNP rs2306283 of the gene SLCO1B1 and genotype GG of SNP rs1051266 of the gene SLC19A1 were associated with a higher response rate in colorectal cancer patients treated with irinotecan[110].

The expression of CES1 and CES2 is organ-specific. CES1 is expressed mainly in the liver, whereas CES2 is predominantly expressed in the small intestine. CES1 is 100-fold less effective than CES2 at drug activation of irinotecan, but plays a significant role, mainly in the liver[111]. Until recently, CES was believed to play only a minor role in irinotecan metabolism and no functional polymorphisms have been identified[112]. Sai et al[104] reported a gene-dosage effect of functional CES1A on SN-38 formation for the first time and suggested its potential role in irinotecan toxicity.

The reduced mRNA expression of UGT1A1, 1A3, 1A4, lA6, 1A9 and overexpression of UGT1A5, UGTlA8 and UGTlAl0 in colorectal tumors compared with non-malignant tissues from the same patient has recently been reported[113]. Interestingly, despite the fact that UGT1A1 is the most studied gene with regard to irinotecan toxicity, little is known about the association of its expression with the prognosis of colorectal cancer patients. Tailored dosing based on UGT1A1 genotype was tested in a prospective trial[114]. The dose of irinotecan in the biweekly schedule (standard dose 180 mg/m2) could be escalated to 450 mg/m2 in patients with the UGT1A1*1/*1 genotype and to 390 mg/m2 in patients with the UGT1A1*1/*28 genotype, but only to 150 mg/m2 in patients with the UGT1A1*28/*28 genotype. The maximum tolerated dose was 390 mg/m2, 340 mg/m2 and 130 mg/m2 for patients with the UGT1A1*1/*1, UGT1A1*1/*28 and UGT1A1*28/*28 genotypes, respectively. Significantly higher objective response rates was observed in patients with the UGT1A1*1/*1 (60%) and UGT1A1*1/*28 (39%) genotypes compared with patients with the UGT1A1*28/*28 genotype (13%). A marked difference in response rate was observed between patients treated with an irinotecan dose of 260 mg/m2 and higher (67%) vs patients treated by the lower dose (24%). In the multivariable logistic regression analysis irinotecan 260 mg/m2 or higher was the only predictor of objective response. The time-to-disease progression was also significantly increased in patients treated with higher irinotecan dose[114].

According to the meta-analysis published by Dias et al[115], the individual response to irinotecan is unlikely to be affected by the presence of the UGT1A1*28 variant because of the lack of statistically significant differences in objective response rates between irinotecan-administered cancer patients divided by UGT1A1 polymorphism rs8175347.

Biomarkers of toxicity in colorectal cancer patients treated with irinotecan

Studies on polymorphisms in ABC transporters have shown some promising results; several potential biomarkers have been describe such as ABCB1 (rs1045642), ABCC2 (rs3740066), ABCC5 (rs562), ABCG1 (rs425215) and ABCG2 (rs3832043) (Table 3)[106,112,116,117]. However, in terms of clinical practice, The Royal Dutch Pharmacists Association (RDPA) posted the only recommendation on genetic testing of colorectal cancer patients before treatment by irinotecan in FOLFIRI. The Pharmacogenetics Working Group of RDPA has assessed therapeutic dose recommendations for irinotecan pursuant to the UGT1A1 genotype. Dose reduction for UGT1A1*28 homozygous patients receiving more than 250 mg/m2 has been recommended[62]. This recommendation is not applicable for FOLFIRI regimen where a dose of 180 mg/m2 is prescribed. Homozygous carriers for the UGT1A1*28 allele have reduced UGT1A1 enzyme activity leading to an increased risk for neutropenia[62].

RESULTS OF CLINICAL TRIALS ON BIOMARKERS FOR FOLFOX/FOLFIRI REGIMENS

Studies on predictive biomarkers useful for deciding whether patients should be treated by FOLFOX or FOLFIRI are currently lacking. Only a few studies on colorectal carcinoma patients treated by regimens involving FOLFOX may be found in the recent literature. Watanabe et al[118] analyzed microarray expression profiles of 46 patients with metastatic or recurrent colorectal carcinoma who received modified FOLFOX6. As a result, a 27-gene FOLFOX response predictor with an overall accuracy of 92.5% was constructed for selection of patients who may benefit from therapy with this regimen in the adjuvant or advanced disease setting[118]. Another Japanese study suggested that the absence of ERCC1 or GSTP1, but not TYMS, as assessed by immunohistochemistry, predicts a favorable response of colorectal cancer patients to FOLFOX[119]. Interestingly, concurrent methylation of transcription factor neurogenin 1 (NEUROG1, OMIM: 601726) and tumor suppressor p16 (CDKN2A, OMIM: 600160) was recently associated with shorter DFS following adjuvant FOLFOX in Stages II/III colorectal cancer[120].

Carriers of the T allele in ERCC1 polymorphism rs11615 among colorectal cancer patients (n = 168) had significantly lower response to FOLFOX and shorter PFS than CC wild-type genotype carriers[121]. Similarly, FOLFOX4-treated colorectal cancer patients with Gln allele in ERCC2 rs13161 had a significantly shorter PFS and OS than Lys/Lys wild-type carriers[122]. The same authors also reported that Val allele (rs1695 in GSTP1) carriers had longer PFS and OS, but also a significantly higher incidence of grade 3/4 cumulative neuropathy after FOLFOX4 therapy than wild-type Ile/Ile carriers[123]. Metastatic colorectal cancer patients treated with FOLFOX4 carrying either the CC genotype in ERCC1 rs11615 or the GG genotype in XRCC1 polymorphism rs25487 had significantly longer PFS and OS than the respective variant allele carriers[83]. The combination of these two polymorphisms indicated an even higher survival benefit than carriage of a single predictor genotype[83].

There are currently no studies focusing on the evaluation of predictive biomarkers of FOLFIRI in colorectal cancer patients. Therefore, a retrospective study based on putative predictive biomarkers for 5-FU and irinotecan so far identified (Tables 1 and 3) are now highly desirable in a cohort of patients treated with this regimen.

Future directions

Despite remarkable progress during the past two decades that resulted from the introduction of new drugs, including targeted agents, the potential of systemic pharmacologic therapy has still not been fully translated into effective treatments in daily practice. Optimized algorithms for use of the available active agents are urgently needed. The pharmacogenomic approach has been promising in not only identifying the patients at increased risk of toxicity, but also in allowing a more tailored approach in pharmacotherapy of colorectal carcinoma.

From the reviewed information it is obvious that there are many important factors in resistance of colorectal tumor cells to FOLFOX/FOLFIRI regimens. The broad individual variability and cell- and tissue-specific pattern of expression, including complicated regulation, cause difficulties in drawing conclusive remarks on clinical application of currently identified putative biomarkers. The lack of functional studies on relations between genotype and phenotype, including activity and pharmacokinetics, of the majority of biomarkers also greatly complicates translation of results into biologically-relevant mechanisms that are necessary for recommendations on the further use of biomarkers.

The majority of studies reviewed provided conflicting results, which can be attributed to the different methodologies used, such as antibodies for immunohistochemistry, lack of statistical power, possible selection bias associated with the heterogeneity of the cohorts of patients in terms of ethnicity or clinical characteristics, including treatment, drug-drug interactions as well as drug-alimentary interactions.

The unprecedented increase in the knowledge concerning the pathogenesis of malignant tumors, together with the advent of new agents, has created a basis for clinical cancer pharmacogenetics. Enormous progress in the area of analysis of both genotype and phenotype that has been made possible by the advent of next generation sequencing, the application of strict methodical guidelines such as MIQE or REMARK, and the establishment of public databases and pharmacogenetics consortia for rigorous evaluation of existing studies may be instrumental for further advancement of our knowledge in this field.

The pursuit of research in the following directions may further accelerate the evolution of the field of pharmacogenetics: (1) studies on the functionality of haplotypes and the associations with cellular and clinical phenotypes should shed more light on the utility of the biomarkers identified so far. Such exploration is now made much easier with the advent of next generation sequencing; (2) very little is known about the regulation of expression and posttranslational processing of putative biomarkers. Also the role of epigenetic factors, such as methylation of promoter regions or noncoding RNAs will require more investigation. Interaction of drug transporters or metabolizing enzymes with oncogenic or tumor suppressor pathways is also a highly attractive topic because there are already several examples of such interactions, including p53, N-myc, or ERK/MAPK, but their clinical relevance remains unknown; and (3) the time for prospective clinical trials aimed at personalized chemotherapy defined by biomarkers has already come. Genotyping is not routinely used for prediction of therapeutic effect or toxicity of anticancer drugs in clinical practice. The lack of alternative treatments for patients carrying the risk genotype represents the main obstacle. It is also essential to clearly determine the sensitivity and specificity of genotyping tests and to define positive and negative predictive values before the introduction of these methods into routine practice.

Taken together, there are many opportunities to increase our insight into the importance of the above reviewed putative biomarkers of drug resistance and to look for others. The final profile of the drug-sensitive or resistant patient will most probably be a mixture of genetic, epigenetic, intracellular and extracellular factors, whose timing and interaction with extrinsic factors should also be considered.

Footnotes

P- Reviewer: Kauppila JH, Smaglo BG S- Editor: Qi Y L- Editor: Stewart GJ E- Editor: Ma S

References
1.  Scheithauer W, Rosen H, Kornek GV, Sebesta C, Depisch D. Randomised comparison of combination chemotherapy plus supportive care with supportive care alone in patients with metastatic colorectal cancer. BMJ. 1993;306:752-755.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 467]  [Cited by in F6Publishing: 453]  [Article Influence: 14.6]  [Reference Citation Analysis (1)]
2.  Nordic Gastrointestinal Tumor Adjuvant Therapy Group. Expectancy or primary chemotherapy in patients with advanced asymptomatic colorectal cancer: a randomized trial. J Clin Oncol. 1992;10:904-911.  [PubMed]  [DOI]  [Cited in This Article: ]
3.  de Gramont A, Bosset JF, Milan C, Rougier P, Bouché O, Etienne PL, Morvan F, Louvet C, Guillot T, François E. Randomized trial comparing monthly low-dose leucovorin and fluorouracil bolus with bimonthly high-dose leucovorin and fluorouracil bolus plus continuous infusion for advanced colorectal cancer: a French intergroup study. J Clin Oncol. 1997;15:808-815.  [PubMed]  [DOI]  [Cited in This Article: ]
4.  Buchler T, Pavlik T, Bortlicek Z, Poprach A, Vyzula R, Abrahamova J, Melichar B. Objective response and time to progression on sequential treatment with sunitinib and sorafenib in metastatic renal cell carcinoma. Med Oncol. 2012;29:3321-3324.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 7]  [Cited by in F6Publishing: 7]  [Article Influence: 0.6]  [Reference Citation Analysis (0)]
5.  Rougier P, Bugat R, Douillard JY, Culine S, Suc E, Brunet P, Becouarn Y, Ychou M, Marty M, Extra JM. Phase II study of irinotecan in the treatment of advanced colorectal cancer in chemotherapy-naive patients and patients pretreated with fluorouracil-based chemotherapy. J Clin Oncol. 1997;15:251-260.  [PubMed]  [DOI]  [Cited in This Article: ]
6.  Douillard JY, Cunningham D, Roth AD, Navarro M, James RD, Karasek P, Jandik P, Iveson T, Carmichael J, Alakl M. Irinotecan combined with fluorouracil compared with fluorouracil alone as first-line treatment for metastatic colorectal cancer: a multicentre randomised trial. Lancet. 2000;355:1041-1047.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2407]  [Cited by in F6Publishing: 2336]  [Article Influence: 97.3]  [Reference Citation Analysis (1)]
7.  Saltz LB, Cox JV, Blanke C, Rosen LS, Fehrenbacher L, Moore MJ, Maroun JA, Ackland SP, Locker PK, Pirotta N. Irinotecan plus fluorouracil and leucovorin for metastatic colorectal cancer. Irinotecan Study Group. N Engl J Med. 2000;343:905-914.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2273]  [Cited by in F6Publishing: 2190]  [Article Influence: 91.3]  [Reference Citation Analysis (0)]
8.  de Gramont A, Figer A, Seymour M, Homerin M, Hmissi A, Cassidy J, Boni C, Cortes-Funes H, Cervantes A, Freyer G. Leucovorin and fluorouracil with or without oxaliplatin as first-line treatment in advanced colorectal cancer. J Clin Oncol. 2000;18:2938-2947.  [PubMed]  [DOI]  [Cited in This Article: ]
9.  Tournigand C, André T, Achille E, Lledo G, Flesh M, Mery-Mignard D, Quinaux E, Couteau C, Buyse M, Ganem G. FOLFIRI followed by FOLFOX6 or the reverse sequence in advanced colorectal cancer: a randomized GERCOR study. J Clin Oncol. 2004;22:229-237.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2181]  [Cited by in F6Publishing: 2147]  [Article Influence: 102.2]  [Reference Citation Analysis (1)]
10.  Cunningham D, Humblet Y, Siena S, Khayat D, Bleiberg H, Santoro A, Bets D, Mueser M, Harstrick A, Verslype C. Cetuximab monotherapy and cetuximab plus irinotecan in irinotecan-refractory metastatic colorectal cancer. N Engl J Med. 2004;351:337-345.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3767]  [Cited by in F6Publishing: 3625]  [Article Influence: 181.3]  [Reference Citation Analysis (0)]
11.  Hurwitz H, Fehrenbacher L, Novotny W, Cartwright T, Hainsworth J, Heim W, Berlin J, Baron A, Griffing S, Holmgren E. Bevacizumab plus irinotecan, fluorouracil, and leucovorin for metastatic colorectal cancer. N Engl J Med. 2004;350:2335-2342.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 7832]  [Cited by in F6Publishing: 7523]  [Article Influence: 376.2]  [Reference Citation Analysis (1)]
12.  Van Cutsem E, Peeters M, Siena S, Humblet Y, Hendlisz A, Neyns B, Canon JL, Van Laethem JL, Maurel J, Richardson G. Open-label phase III trial of panitumumab plus best supportive care compared with best supportive care alone in patients with chemotherapy-refractory metastatic colorectal cancer. J Clin Oncol. 2007;25:1658-1664.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1444]  [Cited by in F6Publishing: 1424]  [Article Influence: 83.8]  [Reference Citation Analysis (0)]
13.  Van Cutsem E, Köhne CH, Hitre E, Zaluski J, Chang Chien CR, Makhson A, D’Haens G, Pintér T, Lim R, Bodoky G. Cetuximab and chemotherapy as initial treatment for metastatic colorectal cancer. N Engl J Med. 2009;360:1408-1417.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2901]  [Cited by in F6Publishing: 3021]  [Article Influence: 201.4]  [Reference Citation Analysis (1)]
14.  Giantonio BJ, Catalano PJ, Meropol NJ, O’Dwyer PJ, Mitchell EP, Alberts SR, Schwartz MA, Benson AB. Bevacizumab in combination with oxaliplatin, fluorouracil, and leucovorin (FOLFOX4) for previously treated metastatic colorectal cancer: results from the Eastern Cooperative Oncology Group Study E3200. J Clin Oncol. 2007;25:1539-1544.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1709]  [Cited by in F6Publishing: 1687]  [Article Influence: 99.2]  [Reference Citation Analysis (1)]
15.  Melichar B, Dvorák J, Hyspler R, Zadák Z. Intestinal permeability in the assessment of intestinal toxicity of cytotoxic agents. Chemotherapy. 2005;51:336-338.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 39]  [Cited by in F6Publishing: 42]  [Article Influence: 2.2]  [Reference Citation Analysis (0)]
16.  Maindrault-Goebel F, de Gramont A, Louvet C, André T, Carola E, Mabro M, Artru P, Gilles V, Lotz JP, Izrael V. High-dose intensity oxaliplatin added to the simplified bimonthly leucovorin and 5-fluorouracil regimen as second-line therapy for metastatic colorectal cancer (FOLFOX 7). Eur J Cancer. 2001;37:1000-1005.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 105]  [Cited by in F6Publishing: 110]  [Article Influence: 4.8]  [Reference Citation Analysis (0)]
17.  Taïeb J, Artru P, Paye F, Louvet C, Perez N, André T, Gayet B, Hebbar M, Goebel FM, Tournigand C. Intensive systemic chemotherapy combined with surgery for metastatic colorectal cancer: results of a phase II study. J Clin Oncol. 2005;23:502-509.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 62]  [Cited by in F6Publishing: 61]  [Article Influence: 3.2]  [Reference Citation Analysis (0)]
18.  Melichar B, Nemcová I. Eye complications of cetuximab therapy. Eur J Cancer Care (Engl). 2007;16:439-443.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 51]  [Cited by in F6Publishing: 52]  [Article Influence: 3.3]  [Reference Citation Analysis (0)]
19.  Zorzi D, Laurent A, Pawlik TM, Lauwers GY, Vauthey JN, Abdalla EK. Chemotherapy-associated hepatotoxicity and surgery for colorectal liver metastases. Br J Surg. 2007;94:274-286.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 369]  [Cited by in F6Publishing: 353]  [Article Influence: 20.8]  [Reference Citation Analysis (0)]
20.  Ciccolini J, Gross E, Dahan L, Lacarelle B, Mercier C. Routine dihydropyrimidine dehydrogenase testing for anticipating 5-fluorouracil-related severe toxicities: hype or hope? Clin Colorectal Cancer. 2010;9:224-228.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 66]  [Cited by in F6Publishing: 69]  [Article Influence: 5.3]  [Reference Citation Analysis (0)]
21.  Amado RG, Wolf M, Peeters M, Van Cutsem E, Siena S, Freeman DJ, Juan T, Sikorski R, Suggs S, Radinsky R. Wild-type KRAS is required for panitumumab efficacy in patients with metastatic colorectal cancer. J Clin Oncol. 2008;26:1626-1634.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2390]  [Cited by in F6Publishing: 2348]  [Article Influence: 146.8]  [Reference Citation Analysis (0)]
22.  Thorn CF, Marsh S, Carrillo MW, McLeod HL, Klein TE, Altman RB. PharmGKB summary: fluoropyrimidine pathways. Pharmacogenet Genomics. 2011;21:237-242.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 56]  [Cited by in F6Publishing: 79]  [Article Influence: 6.1]  [Reference Citation Analysis (0)]
23.  Mohelnikova-Duchonova B, Brynychova V, Oliverius M, Honsova E, Kala Z, Muckova K, Soucek P. Differences in transcript levels of ABC transporters between pancreatic adenocarcinoma and nonneoplastic tissues. Pancreas. 2013;42:707-716.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 71]  [Cited by in F6Publishing: 74]  [Article Influence: 6.7]  [Reference Citation Analysis (0)]
24.  Mohelnikova-Duchonova B, Brynychova V, Hlavac V, Kocik M, Oliverius M, Hlavsa J, Honsova E, Mazanec J, Kala Z, Melichar B. The association between the expression of solute carrier transporters and the prognosis of pancreatic cancer. Cancer Chemother Pharmacol. 2013;72:669-682.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 43]  [Cited by in F6Publishing: 55]  [Article Influence: 5.0]  [Reference Citation Analysis (0)]
25.  Longley DB, Harkin DP, Johnston PG. 5-fluorouracil: mechanisms of action and clinical strategies. Nat Rev Cancer. 2003;3:330-338.  [PubMed]  [DOI]  [Cited in This Article: ]
26.  Humeniuk R, Menon LG, Mishra PJ, Gorlick R, Sowers R, Rode W, Pizzorno G, Cheng YC, Kemeny N, Bertino JR. Decreased levels of UMP kinase as a mechanism of fluoropyrimidine resistance. Mol Cancer Ther. 2009;8:1037-1044.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 33]  [Cited by in F6Publishing: 36]  [Article Influence: 2.4]  [Reference Citation Analysis (0)]
27.  Caudle KE, Thorn CF, Klein TE, Swen JJ, McLeod HL, Diasio RB, Schwab M. Clinical Pharmacogenetics Implementation Consortium guidelines for dihydropyrimidine dehydrogenase genotype and fluoropyrimidine dosing. Clin Pharmacol Ther. 2013;94:640-645.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 252]  [Cited by in F6Publishing: 245]  [Article Influence: 22.3]  [Reference Citation Analysis (0)]
28.  Hlavata I, Mohelnikova-Duchonova B, Vaclavikova R, Liska V, Pitule P, Novak P, Bruha J, Vycital O, Holubec L, Treska V. The role of ABC transporters in progression and clinical outcome of colorectal cancer. Mutagenesis. 2012;27:187-196.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 146]  [Cited by in F6Publishing: 161]  [Article Influence: 13.4]  [Reference Citation Analysis (0)]
29.  Minegaki T, Takara K, Hamaguchi R, Tsujimoto M, Nishiguchi K. Factors affecting the sensitivity of human-derived esophageal carcinoma cell lines to 5-fluorouracil and cisplatin. Oncol Lett. 2013;5:427-434.  [PubMed]  [DOI]  [Cited in This Article: ]
30.  De Iudicibus S, De Pellegrin A, Stocco G, Bartoli F, Bussani R, Decorti G. ABCB1 gene polymorphisms and expression of P-glycoprotein and long-term prognosis in colorectal cancer. Anticancer Res. 2008;28:3921-3928.  [PubMed]  [DOI]  [Cited in This Article: ]
31.  Zhou SF, Wang LL, Di YM, Xue CC, Duan W, Li CG, Li Y. Substrates and inhibitors of human multidrug resistance associated proteins and the implications in drug development. Curr Med Chem. 2008;15:1981-2039.  [PubMed]  [DOI]  [Cited in This Article: ]
32.  Schmidt WM, Kalipciyan M, Dornstauder E, Rizovski B, Steger GG, Sedivy R, Mueller MW, Mader RM. Dissecting progressive stages of 5-fluorouracil resistance in vitro using RNA expression profiling. Int J Cancer. 2004;112:200-212.  [PubMed]  [DOI]  [Cited in This Article: ]
33.  Hagmann W, Jesnowski R, Faissner R, Guo C, Löhr JM. ATP-binding cassette C transporters in human pancreatic carcinoma cell lines. Upregulation in 5-fluorouracil-resistant cells. Pancreatology. 2009;9:136-144.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 65]  [Cited by in F6Publishing: 71]  [Article Influence: 4.4]  [Reference Citation Analysis (0)]
34.  Guo Y, Kotova E, Chen ZS, Lee K, Hopper-Borge E, Belinsky MG, Kruh GD. MRP8, ATP-binding cassette C11 (ABCC11), is a cyclic nucleotide efflux pump and a resistance factor for fluoropyrimidines 2’,3’-dideoxycytidine and 9’-(2’-phosphonylmethoxyethyl)adenine. J Biol Chem. 2003;278:29509-29514.  [PubMed]  [DOI]  [Cited in This Article: ]
35.  Park S, Shimizu C, Shimoyama T, Takeda M, Ando M, Kohno T, Katsumata N, Kang YK, Nishio K, Fujiwara Y. Gene expression profiling of ATP-binding cassette (ABC) transporters as a predictor of the pathologic response to neoadjuvant chemotherapy in breast cancer patients. Breast Cancer Res Treat. 2006;99:9-17.  [PubMed]  [DOI]  [Cited in This Article: ]
36.  Oguri T, Bessho Y, Achiwa H, Ozasa H, Maeno K, Maeda H, Sato S, Ueda R. MRP8/ABCC11 directly confers resistance to 5-fluorouracil. Mol Cancer Ther. 2007;6:122-127.  [PubMed]  [DOI]  [Cited in This Article: ]
37.  Baldwin SA, Beal PR, Yao SY, King AE, Cass CE, Young JD. The equilibrative nucleoside transporter family, SLC29. Pflugers Arch. 2004;447:735-743.  [PubMed]  [DOI]  [Cited in This Article: ]
38.  Gray JH, Owen RP, Giacomini KM. The concentrative nucleoside transporter family, SLC28. Pflugers Arch. 2004;447:728-734.  [PubMed]  [DOI]  [Cited in This Article: ]
39.  Yoshinare K, Kubota T, Watanabe M, Wada N, Nishibori H, Hasegawa H, Kitajima M, Takechi T, Fukushima M. Gene expression in colorectal cancer and in vitro chemosensitivity to 5-fluorouracil: a study of 88 surgical specimens. Cancer Sci. 2003;94:633-638.  [PubMed]  [DOI]  [Cited in This Article: ]
40.  Phua LC, Mal M, Koh PK, Cheah PY, Chan EC, Ho HK. Investigating the role of nucleoside transporters in the resistance of colorectal cancer to 5-fluorouracil therapy. Cancer Chemother Pharmacol. 2013;71:817-823.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 23]  [Cited by in F6Publishing: 19]  [Article Influence: 1.6]  [Reference Citation Analysis (0)]
41.  Mohelnikova-Duchonova B, Melichar B. Human equilibrative nucleoside transporter 1 (hENT1): do we really have a new predictive biomarker of chemotherapy outcome in pancreatic cancer patients? Pancreatology. 2013;13:558-563.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 9]  [Cited by in F6Publishing: 10]  [Article Influence: 0.9]  [Reference Citation Analysis (0)]
42.  Ciaparrone M, Quirino M, Schinzari G, Zannoni G, Corsi DC, Vecchio FM, Cassano A, La Torre G, Barone C. Predictive role of thymidylate synthase, dihydropyrimidine dehydrogenase and thymidine phosphorylase expression in colorectal cancer patients receiving adjuvant 5-fluorouracil. Oncology. 2006;70:366-377.  [PubMed]  [DOI]  [Cited in This Article: ]
43.  Lassmann S, Hennig M, Rosenberg R, Nährig J, Schreglmann J, Krause F, Poignee-Heger M, Nekarda H, Höfler H, Werner M. Thymidine phosphorylase, dihydropyrimidine dehydrogenase and thymidylate synthase mRNA expression in primary colorectal tumors-correlation to tumor histopathology and clinical follow-up. Int J Colorectal Dis. 2006;21:238-247.  [PubMed]  [DOI]  [Cited in This Article: ]
44.  Soong R, Shah N, Salto-Tellez M, Tai BC, Soo RA, Han HC, Ng SS, Tan WL, Zeps N, Joseph D. Prognostic significance of thymidylate synthase, dihydropyrimidine dehydrogenase and thymidine phosphorylase protein expression in colorectal cancer patients treated with or without 5-fluorouracil-based chemotherapy. Ann Oncol. 2008;19:915-919.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 102]  [Cited by in F6Publishing: 102]  [Article Influence: 6.4]  [Reference Citation Analysis (0)]
45.  Vallböhmer D, Yang DY, Kuramochi H, Shimizu D, Danenberg KD, Lindebjerg J, Nielsen JN, Jakobsen A, Danenberg PV. DPD is a molecular determinant of capecitabine efficacy in colorectal cancer. Int J Oncol. 2007;31:413-418.  [PubMed]  [DOI]  [Cited in This Article: ]
46.  van Triest B, Pinedo HM, Blaauwgeers JL, van Diest PJ, Schoenmakers PS, Voorn DA, Smid K, Hoekman K, Hoitsma HF, Peters GJ. Prognostic role of thymidylate synthase, thymidine phosphorylase/platelet-derived endothelial cell growth factor, and proliferation markers in colorectal cancer. Clin Cancer Res. 2000;6:1063-1072.  [PubMed]  [DOI]  [Cited in This Article: ]
47.  Tokunaga Y, Takahashi K, Saito T. Clinical role of thymidine phosphorylase and dihydropyrimidine dehydrogenase in colorectal cancer treated with postoperative fluoropyrimidine. Hepatogastroenterology. 2005;52:1715-1721.  [PubMed]  [DOI]  [Cited in This Article: ]
48.  Meropol NJ, Gold PJ, Diasio RB, Andria M, Dhami M, Godfrey T, Kovatich AJ, Lund KA, Mitchell E, Schwarting R. Thymidine phosphorylase expression is associated with response to capecitabine plus irinotecan in patients with metastatic colorectal cancer. J Clin Oncol. 2006;24:4069-4077.  [PubMed]  [DOI]  [Cited in This Article: ]
49.  Koopman M, Venderbosch S, van Tinteren H, Ligtenberg MJ, Nagtegaal I, Van Krieken JH, Punt CJ. Predictive and prognostic markers for the outcome of chemotherapy in advanced colorectal cancer, a retrospective analysis of the phase III randomised CAIRO study. Eur J Cancer. 2009;45:1999-2006.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 69]  [Cited by in F6Publishing: 76]  [Article Influence: 5.1]  [Reference Citation Analysis (0)]
50.  Chiorean EG, Sanghani S, Schiel MA, Yu M, Burns M, Tong Y, Hinkle DT, Coleman N, Robb B, LeBlanc J. Phase II and gene expression analysis trial of neoadjuvant capecitabine plus irinotecan followed by capecitabine-based chemoradiotherapy for locally advanced rectal cancer: Hoosier Oncology Group GI03-53. Cancer Chemother Pharmacol. 2012;70:25-32.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 12]  [Cited by in F6Publishing: 15]  [Article Influence: 1.3]  [Reference Citation Analysis (0)]
51.  Sadahiro S, Suzuki T, Tanaka A, Okada K, Nagase H, Uchida J. Association of right-sided tumors with high thymidine phosphorylase gene expression levels and the response to oral uracil and tegafur/leucovorin chemotherapy among patients with colorectal cancer. Cancer Chemother Pharmacol. 2012;70:285-291.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 10]  [Cited by in F6Publishing: 14]  [Article Influence: 1.2]  [Reference Citation Analysis (0)]
52.  Mitselou A, Ioachim E, Skoufi U, Tsironis C, Tsimogiannis KE, Skoufi C, Vougiouklakis T, Briasoulis E. Predictive role of thymidine phosphorylase expression in patients with colorectal cancer and its association with angiogenesis-related proteins and extracellular matrix components. In Vivo. 2012;26:1057-1067.  [PubMed]  [DOI]  [Cited in This Article: ]
53.  Walko CM, Lindley C. Capecitabine: a review. Clin Ther. 2005;27:23-44.  [PubMed]  [DOI]  [Cited in This Article: ]
54.  Yanagisawa Y, Maruta F, Iinuma N, Ishizone S, Koide N, Nakayama J, Miyagawa S. Modified Irinotecan/5FU/Leucovorin therapy in advanced colorectal cancer and predicting therapeutic efficacy by expression of tumor-related enzymes. Scand J Gastroenterol. 2007;42:477-484.  [PubMed]  [DOI]  [Cited in This Article: ]
55.  Tokunaga Y, Sasaki H, Saito T. Clinical role of orotate phosphoribosyl transferase and dihydropyrimidine dehydrogenase in colorectal cancer treated with postoperative fluoropyrimidine. Surgery. 2007;141:346-353.  [PubMed]  [DOI]  [Cited in This Article: ]
56.  Koopman M, Venderbosch S, Nagtegaal ID, van Krieken JH, Punt CJ. A review on the use of molecular markers of cytotoxic therapy for colorectal cancer, what have we learned? Eur J Cancer. 2009;45:1935-1949.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 68]  [Cited by in F6Publishing: 65]  [Article Influence: 4.3]  [Reference Citation Analysis (0)]
57.  Jennings BA, Kwok CS, Willis G, Matthews V, Wawruch P, Loke YK. Functional polymorphisms of folate metabolism and response to chemotherapy for colorectal cancer, a systematic review and meta-analysis. Pharmacogenet Genomics. 2012;22:290-304.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 39]  [Cited by in F6Publishing: 42]  [Article Influence: 3.5]  [Reference Citation Analysis (0)]
58.  Beck A, Etienne MC, Chéradame S, Fischel JL, Formento P, Renée N, Milano G. A role for dihydropyrimidine dehydrogenase and thymidylate synthase in tumour sensitivity to fluorouracil. Eur J Cancer. 1994;30A:1517-1522.  [PubMed]  [DOI]  [Cited in This Article: ]
59.  Westra JL, Hollema H, Schaapveld M, Platteel I, Oien KA, Keith WN, Mauritz R, Peters GJ, Buys CH, Hofstra RM. Predictive value of thymidylate synthase and dihydropyrimidine dehydrogenase protein expression on survival in adjuvantly treated stage III colon cancer patients. Ann Oncol. 2005;16:1646-1653.  [PubMed]  [DOI]  [Cited in This Article: ]
60.  Ishiguro M, Kotake K, Nishimura G, Tomita N, Ichikawa W, Takahashi K, Watanabe T, Furuhata T, Kondo K, Mori M. Study protocol of the B-CAST study: a multicenter, prospective cohort study investigating the tumor biomarkers in adjuvant chemotherapy for stage III colon cancer. BMC Cancer. 2013;13:149.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3]  [Cited by in F6Publishing: 4]  [Article Influence: 0.4]  [Reference Citation Analysis (0)]
61.  Fleming RA, Milano G, Thyss A, Etienne MC, Renée N, Schneider M, Demard F. Correlation between dihydropyrimidine dehydrogenase activity in peripheral mononuclear cells and systemic clearance of fluorouracil in cancer patients. Cancer Res. 1992;52:2899-2902.  [PubMed]  [DOI]  [Cited in This Article: ]
62.  Swen JJ, Nijenhuis M, de Boer A, Grandia L, Maitland-van der Zee AH, Mulder H, Rongen GA, van Schaik RH, Schalekamp T, Touw DJ. Pharmacogenetics: from bench to byte--an update of guidelines. Clin Pharmacol Ther. 2011;89:662-673.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 674]  [Cited by in F6Publishing: 703]  [Article Influence: 54.1]  [Reference Citation Analysis (0)]
63.  Morel A, Boisdron-Celle M, Fey L, Soulie P, Craipeau MC, Traore S, Gamelin E. Clinical relevance of different dihydropyrimidine dehydrogenase gene single nucleotide polymorphisms on 5-fluorouracil tolerance. Mol Cancer Ther. 2006;5:2895-2904.  [PubMed]  [DOI]  [Cited in This Article: ]
64.  Schwab M, Zanger UM, Marx C, Schaeffeler E, Klein K, Dippon J, Kerb R, Blievernicht J, Fischer J, Hofmann U. Role of genetic and nongenetic factors for fluorouracil treatment-related severe toxicity: a prospective clinical trial by the German 5-FU Toxicity Study Group. J Clin Oncol. 2008;26:2131-2138.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 297]  [Cited by in F6Publishing: 275]  [Article Influence: 17.2]  [Reference Citation Analysis (0)]
65.  Graham J, Mushin M, Kirkpatrick P. Oxaliplatin. Nat Rev Drug Discov. 2004;3:11-12.  [PubMed]  [DOI]  [Cited in This Article: ]
66.  Marsh S, McLeod H, Dolan E, Shukla SJ, Rabik CA, Gong L, Hernandez-Boussard T, Lou XJ, Klein TE, Altman RB. Platinum pathway. Pharmacogenet Genomics. 2009;19:563-564.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 32]  [Cited by in F6Publishing: 35]  [Article Influence: 2.3]  [Reference Citation Analysis (0)]
67.  Mamenta EL, Poma EE, Kaufmann WK, Delmastro DA, Grady HL, Chaney SG. Enhanced replicative bypass of platinum-DNA adducts in cisplatin-resistant human ovarian carcinoma cell lines. Cancer Res. 1994;54:3500-3505.  [PubMed]  [DOI]  [Cited in This Article: ]
68.  Vaisman A, Varchenko M, Umar A, Kunkel TA, Risinger JI, Barrett JC, Hamilton TC, Chaney SG. The role of hMLH1, hMSH3, and hMSH6 defects in cisplatin and oxaliplatin resistance: correlation with replicative bypass of platinum-DNA adducts. Cancer Res. 1998;58:3579-3585.  [PubMed]  [DOI]  [Cited in This Article: ]
69.  Zhang S, Lovejoy KS, Shima JE, Lagpacan LL, Shu Y, Lapuk A, Chen Y, Komori T, Gray JW, Chen X. Organic cation transporters are determinants of oxaliplatin cytotoxicity. Cancer Res. 2006;66:8847-8857.  [PubMed]  [DOI]  [Cited in This Article: ]
70.  Song IS, Savaraj N, Siddik ZH, Liu P, Wei Y, Wu CJ, Kuo MT. Role of human copper transporter Ctr1 in the transport of platinum-based antitumor agents in cisplatin-sensitive and cisplatin-resistant cells. Mol Cancer Ther. 2004;3:1543-1549.  [PubMed]  [DOI]  [Cited in This Article: ]
71.  Sprowl JA, Ciarimboli G, Lancaster CS, Giovinazzo H, Gibson AA, Du G, Janke LJ, Cavaletti G, Shields AF, Sparreboom A. Oxaliplatin-induced neurotoxicity is dependent on the organic cation transporter OCT2. Proc Natl Acad Sci USA. 2013;110:11199-11204.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 127]  [Cited by in F6Publishing: 130]  [Article Influence: 11.8]  [Reference Citation Analysis (0)]
72.  Cui Y, König J, Buchholz JK, Spring H, Leier I, Keppler D. Drug resistance and ATP-dependent conjugate transport mediated by the apical multidrug resistance protein, MRP2, permanently expressed in human and canine cells. Mol Pharmacol. 1999;55:929-937.  [PubMed]  [DOI]  [Cited in This Article: ]
73.  Wakamatsu T, Nakahashi Y, Hachimine D, Seki T, Okazaki K. The combination of glycyrrhizin and lamivudine can reverse the cisplatin resistance in hepatocellular carcinoma cells through inhibition of multidrug resistance-associated proteins. Int J Oncol. 2007;31:1465-1472.  [PubMed]  [DOI]  [Cited in This Article: ]
74.  Goto S, Iida T, Cho S, Oka M, Kohno S, Kondo T. Overexpression of glutathione S-transferase pi enhances the adduct formation of cisplatin with glutathione in human cancer cells. Free Radic Res. 1999;31:549-558.  [PubMed]  [DOI]  [Cited in This Article: ]
75.  Townsend DM, Tew KD. The role of glutathione-S-transferase in anti-cancer drug resistance. Oncogene. 2003;22:7369-7375.  [PubMed]  [DOI]  [Cited in This Article: ]
76.  Ballatori N, Hammond CL, Cunningham JB, Krance SM, Marchan R. Molecular mechanisms of reduced glutathione transport: role of the MRP/CFTR/ABCC and OATP/SLC21A families of membrane proteins. Toxicol Appl Pharmacol. 2005;204:238-255.  [PubMed]  [DOI]  [Cited in This Article: ]
77.  Theile D, Grebhardt S, Haefeli WE, Weiss J. Involvement of drug transporters in the synergistic action of FOLFOX combination chemotherapy. Biochem Pharmacol. 2009;78:1366-1373.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 47]  [Cited by in F6Publishing: 44]  [Article Influence: 2.9]  [Reference Citation Analysis (0)]
78.  Martinez-Balibrea E, Martínez-Cardús A, Musulén E, Ginés A, Manzano JL, Aranda E, Plasencia C, Neamati N, Abad A. Increased levels of copper efflux transporter ATP7B are associated with poor outcome in colorectal cancer patients receiving oxaliplatin-based chemotherapy. Int J Cancer. 2009;124:2905-2910.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 41]  [Cited by in F6Publishing: 47]  [Article Influence: 3.1]  [Reference Citation Analysis (0)]
79.  Wu H, Kang H, Liu Y, Xiao Q, Zhang Y, Sun M, Liu D, Wang Z, Zhao H, Yao W. Association of ABCB1 genetic polymorphisms with susceptibility to colorectal cancer and therapeutic prognosis. Pharmacogenomics. 2013;14:897-911.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 20]  [Cited by in F6Publishing: 26]  [Article Influence: 2.6]  [Reference Citation Analysis (0)]
80.  Shirota Y, Stoehlmacher J, Brabender J, Xiong YP, Uetake H, Danenberg KD, Groshen S, Tsao-Wei DD, Danenberg PV, Lenz HJ. ERCC1 and thymidylate synthase mRNA levels predict survival for colorectal cancer patients receiving combination oxaliplatin and fluorouracil chemotherapy. J Clin Oncol. 2001;19:4298-4304.  [PubMed]  [DOI]  [Cited in This Article: ]
81.  Grimminger PP, Shi M, Barrett C, Lebwohl D, Danenberg KD, Brabender J, Vigen CL, Danenberg PV, Winder T, Lenz HJ. TS and ERCC-1 mRNA expressions and clinical outcome in patients with metastatic colon cancer in CONFIRM-1 and -2 clinical trials. Pharmacogenomics J. 2012;12:404-411.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 29]  [Cited by in F6Publishing: 34]  [Article Influence: 2.6]  [Reference Citation Analysis (0)]
82.  Huang MY, Tsai HL, Lin CH, Huang CW, Ma CJ, Huang CM, Chai CY, Wang JY. Predictive value of ERCC1, ERCC2, and XRCC1 overexpression for stage III colorectal cancer patients receiving FOLFOX-4 adjuvant chemotherapy. J Surg Oncol. 2013;108:457-464.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 28]  [Cited by in F6Publishing: 33]  [Article Influence: 3.0]  [Reference Citation Analysis (0)]
83.  Huang MY, Huang ML, Chen MJ, Lu CY, Chen CF, Tsai PC, Chuang SC, Hou MF, Lin SR, Wang JY. Multiple genetic polymorphisms in the prediction of clinical outcome of metastatic colorectal cancer patients treated with first-line FOLFOX-4 chemotherapy. Pharmacogenet Genomics. 2011;21:18-25.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 43]  [Cited by in F6Publishing: 47]  [Article Influence: 3.6]  [Reference Citation Analysis (0)]
84.  Fariña Sarasqueta A, van Lijnschoten G, Lemmens VE, Rutten HJ, van den Brule AJ. Pharmacogenetics of oxaliplatin as adjuvant treatment in colon carcinoma: are single nucleotide polymorphisms in GSTP1, ERCC1, and ERCC2 good predictive markers? Mol Diagn Ther. 2011;15:277-283.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in F6Publishing: 1]  [Reference Citation Analysis (0)]
85.  Huang MY, Fang WY, Lee SC, Cheng TL, Wang JY, Lin SR. ERCC2 2251A& gt; C genetic polymorphism was highly correlated with early relapse in high-risk stage II and stage III colorectal cancer patients: a preliminary study. BMC Cancer. 2008;8:50.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 35]  [Cited by in F6Publishing: 39]  [Article Influence: 2.4]  [Reference Citation Analysis (0)]
86.  Le Morvan V, Smith D, Laurand A, Brouste V, Bellott R, Soubeyran I, Mathoulin-Pelissier S, Robert J. Determination of ERCC2 Lys751Gln and GSTP1 Ile105Val gene polymorphisms in colorectal cancer patients: relationships with treatment outcome. Pharmacogenomics. 2007;8:1693-1703.  [PubMed]  [DOI]  [Cited in This Article: ]
87.  Funke S, Timofeeva M, Risch A, Hoffmeister M, Stegmaier C, Seiler CM, Brenner H, Chang-Claude J. Genetic polymorphisms in GST genes and survival of colorectal cancer patients treated with chemotherapy. Pharmacogenomics. 2010;11:33-41.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 29]  [Cited by in F6Publishing: 31]  [Article Influence: 2.2]  [Reference Citation Analysis (0)]
88.  Kumamoto K, Ishibashi K, Okada N, Tajima Y, Kuwabara K, Kumagai Y, Baba H, Haga N, Ishida H. Polymorphisms of GSTP1, ERCC2 and TS-3’UTR are associated with the clinical outcome of mFOLFOX6 in colorectal cancer patients. Oncol Lett. 2013;6:648-654.  [PubMed]  [DOI]  [Cited in This Article: ]
89.  Lee KH, Chang HJ, Han SW, Oh DY, Im SA, Bang YJ, Kim SY, Lee KW, Kim JH, Hong YS. Pharmacogenetic analysis of adjuvant FOLFOX for Korean patients with colon cancer. Cancer Chemother Pharmacol. 2013;71:843-851.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 32]  [Cited by in F6Publishing: 34]  [Article Influence: 3.1]  [Reference Citation Analysis (0)]
90.  McLeod HL, Sargent DJ, Marsh S, Green EM, King CR, Fuchs CS, Ramanathan RK, Williamson SK, Findlay BP, Thibodeau SN. Pharmacogenetic predictors of adverse events and response to chemotherapy in metastatic colorectal cancer: results from North American Gastrointestinal Intergroup Trial N9741. J Clin Oncol. 2010;28:3227-3233.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 166]  [Cited by in F6Publishing: 180]  [Article Influence: 12.9]  [Reference Citation Analysis (0)]
91.  Peng Z, Wang Q, Gao J, Ji Z, Yuan J, Tian Y, Shen L. Association between GSTP1 Ile105Val polymorphism and oxaliplatin-induced neuropathy: a systematic review and meta-analysis. Cancer Chemother Pharmacol. 2013;72:305-314.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 28]  [Cited by in F6Publishing: 29]  [Article Influence: 2.6]  [Reference Citation Analysis (0)]
92.  Bencharit S, Morton CL, Howard-Williams EL, Danks MK, Potter PM, Redinbo MR. Structural insights into CPT-11 activation by mammalian carboxylesterases. Nat Struct Biol. 2002;9:337-342.  [PubMed]  [DOI]  [Cited in This Article: ]
93.  Nozawa T, Minami H, Sugiura S, Tsuji A, Tamai I. Role of organic anion transporter OATP1B1 (OATP-C) in hepatic uptake of irinotecan and its active metabolite, 7-ethyl-10-hydroxycamptothecin: in vitro evidence and effect of single nucleotide polymorphisms. Drug Metab Dispos. 2005;33:434-439.  [PubMed]  [DOI]  [Cited in This Article: ]
94.  Iyer L, Ramírez J, Shepard DR, Bingham CM, Hossfeld DK, Ratain MJ, Mayer U. Biliary transport of irinotecan and metabolites in normal and P-glycoprotein-deficient mice. Cancer Chemother Pharmacol. 2002;49:336-341.  [PubMed]  [DOI]  [Cited in This Article: ]
95.  Chen ZS, Furukawa T, Sumizawa T, Ono K, Ueda K, Seto K, Akiyama SI. ATP-Dependent efflux of CPT-11 and SN-38 by the multidrug resistance protein (MRP) and its inhibition by PAK-104P. Mol Pharmacol. 1999;55:921-928.  [PubMed]  [DOI]  [Cited in This Article: ]
96.  Chu XY, Kato Y, Niinuma K, Sudo KI, Hakusui H, Sugiyama Y. Multispecific organic anion transporter is responsible for the biliary excretion of the camptothecin derivative irinotecan and its metabolites in rats. J Pharmacol Exp Ther. 1997;281:304-314.  [PubMed]  [DOI]  [Cited in This Article: ]
97.  Nakatomi K, Yoshikawa M, Oka M, Ikegami Y, Hayasaka S, Sano K, Shiozawa K, Kawabata S, Soda H, Ishikawa T. Transport of 7-ethyl-10-hydroxycamptothecin (SN-38) by breast cancer resistance protein ABCG2 in human lung cancer cells. Biochem Biophys Res Commun. 2001;288:827-832.  [PubMed]  [DOI]  [Cited in This Article: ]
98.  Whirl-Carrillo M, McDonagh EM, Hebert JM, Gong L, Sangkuhl K, Thorn CF, Altman RB, Klein TE. Pharmacogenomics knowledge for personalized medicine. Clin Pharmacol Ther. 2012;92:414-417.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1216]  [Cited by in F6Publishing: 1224]  [Article Influence: 102.0]  [Reference Citation Analysis (0)]
99.  Hanioka N, Ozawa S, Jinno H, Ando M, Saito Y, Sawada J. Human liver UDP-glucuronosyltransferase isoforms involved in the glucuronidation of 7-ethyl-10-hydroxycamptothecin. Xenobiotica. 2001;31:687-699.  [PubMed]  [DOI]  [Cited in This Article: ]
100.  Haaz MC, Rivory L, Riché C, Vernillet L, Robert J. Metabolism of irinotecan (CPT-11) by human hepatic microsomes: participation of cytochrome P-450 3A and drug interactions. Cancer Res. 1998;58:468-472.  [PubMed]  [DOI]  [Cited in This Article: ]
101.  Lin F, Marchetti S, Pluim D, Iusuf D, Mazzanti R, Schellens JH, Beijnen JH, van Tellingen O. Abcc4 together with abcb1 and abcg2 form a robust cooperative drug efflux system that restricts the brain entry of camptothecin analogues. Clin Cancer Res. 2013;19:2084-2095.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 43]  [Cited by in F6Publishing: 43]  [Article Influence: 3.9]  [Reference Citation Analysis (0)]
102.  Shibayama Y, Iwashita Y, Yoshikawa Y, Kondo T, Ikeda R, Takeda Y, Osada T, Sugawara M, Yamada K, Iseki K. Effect of 5-fluorouracil treatment on SN-38 absorption from intestine in rats. Biol Pharm Bull. 2011;34:1418-1425.  [PubMed]  [DOI]  [Cited in This Article: ]
103.  Mathijssen RH, Marsh S, Karlsson MO, Xie R, Baker SD, Verweij J, Sparreboom A, McLeod HL. Irinotecan pathway genotype analysis to predict pharmacokinetics. Clin Cancer Res. 2003;9:3246-3253.  [PubMed]  [DOI]  [Cited in This Article: ]
104.  Sai K, Saito Y, Maekawa K, Kim SR, Kaniwa N, Nishimaki-Mogami T, Sawada J, Shirao K, Hamaguchi T, Yamamoto N. Additive effects of drug transporter genetic polymorphisms on irinotecan pharmacokinetics/pharmacodynamics in Japanese cancer patients. Cancer Chemother Pharmacol. 2010;66:95-105.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 51]  [Cited by in F6Publishing: 52]  [Article Influence: 3.5]  [Reference Citation Analysis (0)]
105.  De Mattia E, Toffoli G, Polesel J, D’Andrea M, Corona G, Zagonel V, Buonadonna A, Dreussi E, Cecchin E. Pharmacogenetics of ABC and SLC transporters in metastatic colorectal cancer patients receiving first-line FOLFIRI treatment. Pharmacogenet Genomics. 2013;23:549-557.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 37]  [Cited by in F6Publishing: 41]  [Article Influence: 4.1]  [Reference Citation Analysis (0)]
106.  Glimelius B, Garmo H, Berglund A, Fredriksson LA, Berglund M, Kohnke H, Byström P, Sørbye H, Wadelius M. Prediction of irinotecan and 5-fluorouracil toxicity and response in patients with advanced colorectal cancer. Pharmacogenomics J. 2011;11:61-71.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 85]  [Cited by in F6Publishing: 96]  [Article Influence: 6.9]  [Reference Citation Analysis (0)]
107.  Imai Y, Nakane M, Kage K, Tsukahara S, Ishikawa E, Tsuruo T, Miki Y, Sugimoto Y. C421A polymorphism in the human breast cancer resistance protein gene is associated with low expression of Q141K protein and low-level drug resistance. Mol Cancer Ther. 2002;1:611-616.  [PubMed]  [DOI]  [Cited in This Article: ]
108.  de Jong FA, Scott-Horton TJ, Kroetz DL, McLeod HL, Friberg LE, Mathijssen RH, Verweij J, Marsh S, Sparreboom A. Irinotecan-induced diarrhea: functional significance of the polymorphic ABCC2 transporter protein. Clin Pharmacol Ther. 2007;81:42-49.  [PubMed]  [DOI]  [Cited in This Article: ]
109.  Akiyama Y, Fujita K, Ishida H, Sunakawa Y, Yamashita K, Kawara K, Miwa K, Saji S, Sasaki Y. Association of ABCC2 genotype with efficacy of first-line FOLFIRI in Japanese patients with advanced colorectal cancer. Drug Metab Pharmacokinet. 2012;27:325-335.  [PubMed]  [DOI]  [Cited in This Article: ]
110.  Huang L, Zhang T, Xie C, Liao X, Yu Q, Feng J, Ma H, Dai J, Li M, Chen J. SLCO1B1 and SLC19A1 gene variants and irinotecan-induced rapid response and survival: a prospective multicenter pharmacogenetics study of metastatic colorectal cancer. PLoS One. 2013;8:e77223.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 19]  [Cited by in F6Publishing: 23]  [Article Influence: 2.1]  [Reference Citation Analysis (0)]
111.  Hatfield MJ, Tsurkan L, Garrett M, Shaver TM, Hyatt JL, Edwards CC, Hicks LD, Potter PM. Organ-specific carboxylesterase profiling identifies the small intestine and kidney as major contributors of activation of the anticancer prodrug CPT-11. Biochem Pharmacol. 2011;81:24-31.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 66]  [Cited by in F6Publishing: 68]  [Article Influence: 4.9]  [Reference Citation Analysis (0)]
112.  Marsh S, Hoskins JM. Irinotecan pharmacogenomics. Pharmacogenomics. 2010;11:1003-1010.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 80]  [Cited by in F6Publishing: 74]  [Article Influence: 5.7]  [Reference Citation Analysis (0)]
113.  Wang M, Sun DF, Wang S, Qing Y, Chen S, Wu D, Lin YM, Luo JZ, Li YQ. Polymorphic expression of UDP-glucuronosyltransferase UGTlA gene in human colorectal cancer. PLoS One. 2013;8:e57045.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 13]  [Cited by in F6Publishing: 14]  [Article Influence: 1.3]  [Reference Citation Analysis (0)]
114.  Marcuello E, Páez D, Paré L, Salazar J, Sebio A, del Rio E, Baiget M. A genotype-directed phase I-IV dose-finding study of irinotecan in combination with fluorouracil/leucovorin as first-line treatment in advanced colorectal cancer. Br J Cancer. 2011;105:53-57.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 85]  [Cited by in F6Publishing: 87]  [Article Influence: 6.7]  [Reference Citation Analysis (0)]
115.  Dias MM, McKinnon RA, Sorich MJ. Impact of the UGT1A1*28 allele on response to irinotecan: a systematic review and meta-analysis. Pharmacogenomics. 2012;13:889-899.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 43]  [Cited by in F6Publishing: 46]  [Article Influence: 3.8]  [Reference Citation Analysis (0)]
116.  Han JY, Lim HS, Park YH, Lee SY, Lee JS. Integrated pharmacogenetic prediction of irinotecan pharmacokinetics and toxicity in patients with advanced non-small cell lung cancer. Lung Cancer. 2009;63:115-120.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 75]  [Cited by in F6Publishing: 80]  [Article Influence: 5.0]  [Reference Citation Analysis (0)]
117.  Di Martino MT, Arbitrio M, Leone E, Guzzi PH, Rotundo MS, Ciliberto D, Tomaino V, Fabiani F, Talarico D, Sperlongano P. Single nucleotide polymorphisms of ABCC5 and ABCG1 transporter genes correlate to irinotecan-associated gastrointestinal toxicity in colorectal cancer patients: a DMET microarray profiling study. Cancer Biol Ther. 2011;12:780-787.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 70]  [Cited by in F6Publishing: 60]  [Article Influence: 4.6]  [Reference Citation Analysis (0)]
118.  Watanabe T, Kobunai T, Yamamoto Y, Matsuda K, Ishihara S, Nozawa K, Iinuma H, Ikeuchi H. Gene expression of vascular endothelial growth factor A, thymidylate synthase, and tissue inhibitor of metalloproteinase 3 in prediction of response to bevacizumab treatment in colorectal cancer patients. Dis Colon Rectum. 2011;54:1026-1035.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 17]  [Cited by in F6Publishing: 17]  [Article Influence: 1.3]  [Reference Citation Analysis (0)]
119.  Noda E, Maeda K, Inoue T, Fukunaga S, Nagahara H, Shibutani M, Amano R, Nakata B, Tanaka H, Muguruma K. Predictive value of expression of ERCC 1 and GST-p for 5-fluorouracil/oxaliplatin chemotherapy in advanced colorectal cancer. Hepatogastroenterology. 2012;59:130-133.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 13]  [Cited by in F6Publishing: 16]  [Article Influence: 1.3]  [Reference Citation Analysis (0)]
120.  Han SW, Lee HJ, Bae JM, Cho NY, Lee KH, Kim TY, Oh DY, Im SA, Bang YJ, Jeong SY. Methylation and microsatellite status and recurrence following adjuvant FOLFOX in colorectal cancer. Int J Cancer. 2013;132:2209-2216.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 42]  [Cited by in F6Publishing: 45]  [Article Influence: 3.8]  [Reference Citation Analysis (0)]
121.  Chang PM, Tzeng CH, Chen PM, Lin JK, Lin TC, Chen WS, Jiang JK, Wang HS, Wang WS. ERCC1 codon 118 C→T polymorphism associated with ERCC1 expression and outcome of FOLFOX-4 treatment in Asian patients with metastatic colorectal carcinoma. Cancer Sci. 2009;100:278-283.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 47]  [Cited by in F6Publishing: 54]  [Article Influence: 4.2]  [Reference Citation Analysis (0)]
122.  Lai JI, Tzeng CH, Chen PM, Lin JK, Lin TC, Chen WS, Jiang JK, Wang HS, Wang WS. Very low prevalence of XPD K751Q polymorphism and its association with XPD expression and outcomes of FOLFOX-4 treatment in Asian patients with colorectal carcinoma. Cancer Sci. 2009;100:1261-1266.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 30]  [Cited by in F6Publishing: 31]  [Article Influence: 2.1]  [Reference Citation Analysis (0)]
123.  Chen YC, Tzeng CH, Chen PM, Lin JK, Lin TC, Chen WS, Jiang JK, Wang HS, Wang WS. Influence of GSTP1 I105V polymorphism on cumulative neuropathy and outcome of FOLFOX-4 treatment in Asian patients with colorectal carcinoma. Cancer Sci. 2010;101:530-535.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 61]  [Cited by in F6Publishing: 67]  [Article Influence: 4.5]  [Reference Citation Analysis (0)]