Skip to main content
  • Research article
  • Open access
  • Published:

Genetic variability in drug transport, metabolism or DNA repair affecting toxicity of chemotherapy in ovarian cancer

Abstract

Background

This study aimed to determine whether single nucleotide polymorphisms (SNPs) in genes involved in DNA repair or metabolism of taxanes or platinum could predict toxicity or response to first-line chemotherapy in ovarian cancer.

Methods

Twenty-six selected SNPs in 18 genes were genotyped in 322 patients treated with first-line paclitaxel-carboplatin or carboplatin mono-therapy. Genotypes were correlated with toxicity events (anemia, neutropenia, thrombocytopenia, febrile neutropenia, neurotoxicity), use of growth factors and survival.

Results

The risk of anemia was increased for variant alleles of rs1128503 (ABCB1, C > T; p = 0.023, OR = 1.71, 95% CI = 1.07-2.71), rs363717 (ABCA1, A > G; p = 0.002, OR = 2.08, 95% CI = 1.32-3.27) and rs11615 (ERCC1, T > C; p = 0.031, OR = 1.61, 95% CI = 1.04-2.50), while it was decreased for variant alleles of rs12762549 (ABCC2, C > G; p = 0.004, OR = 0.51, 95% CI = 0.33-0.81). Likewise, increased risk of thrombocytopenia was associated with rs4986910 (CYP3A4, T > C; p = 0.025, OR = 4.99, 95% CI = 1.22-20.31). No significant correlations were found for neurotoxicity. Variant alleles of rs2073337 (ABCC2, A > G; p = 0.039, OR = 0.60, 95% CI = 0.37-0.98), rs1695 (ABCC1, A > G; p = 0.017, OR = 0.55, 95% CI 0.33-0.90) and rs1799793 (ERCC2, G > A; p = 0.042, OR = 0.63, 95% CI 0.41-0.98) associated with the use of colony stimulating factors (CSF), while rs2074087 (ABCC1, G > C; p = 0.011, OR = 2.09, 95% CI 1.18-3.68) correlated with use of erythropoiesis stimulating agents (ESAs). Homozygous carriers of the rs1799793 (ERCC2, G > A) G-allele had a prolonged platinum-free interval (p = 0.016).

Conclusions

Our data reveal significant correlations between genetic variants of transport, hepatic metabolism, platinum related detoxification or DNA damage repair and toxicity or outcome in ovarian cancer.

Peer Review reports

Background

Ovarian cancer is the fifth most common cause of cancer death in women and the leading cause of gynaecological cancer-related death in the developed world [1]. Despite optimization of debulking surgery and chemotherapy regimens, the overall 5-year survival in advanced stage disease is only 29% [2]. The current standard first-line chemotherapy is a combination of paclitaxel and carboplatin. This treatment is associated with serious hematologic toxicities including grade 3–4 anemia (incidence 4.3-6.6%), grade 3–4 thrombocytopenia (4.7-12.9%), grade 3–4 neutropenia (37-89%), febrile neutropenia (2.3-8%) [3-6] and grade 2–4 peripheral neuropathies (32-36%), resulting in dose reductions, treatment delays and representing an important physical, psychological and financial burden for the patient and society. Inter-individual differences in both toxicity and outcome related to treatment with paclitaxel-carboplatin are reported. A few patient-related risk factors for toxicity have been identified, such as elderly age (≥65 years), poor performance status and poor nutritional status [7]. Furthermore, tumor-related factors including advanced stage at diagnosis, high-grade serous disease and residual tumor after debulking surgery are associated with poor survival. Genetic variability represents another potential factor explaining this inter-individual variability.

Genes related to drug transport, metabolism, detoxification and DNA repair could influence the cytotoxic effects associated with chemotherapy, including those involved in the transport (e.g., ABCB1, ABCC1, ABCC2, ABCG2, and SLCO1B3) [8-24], hepatic metabolism (CYP3A4, CYP3A5, CYP2C8, CYP1B1) [8-12,14,18,25-27] and pharmacodynamics (e.g., MAPT, TUBB, TP53) [28,29] of paclitaxel. Likewise, genes involved in detoxification (e.g., GSTP1, GSTT1, GSTM1) [30-33] and base-excision DNA repair (e.g., ERCC1, ERCC2, XRCC1) [34,35] have previously been linked with cytotoxicity of platinum agents. In particular, genetic variants in these genes, which generally are supposed to reduce the function of the affected gene, have been proposed to underlie the inter-individual differences in chemotherapy related hematologic and neurotoxicity. Likewise, variants in other genes, including SLC12A6, SERPINB2, PPARD and ICAM have been proposed to contribute to chemotherapy-induced peripheral neurotoxicity [36]. Most studies identifying these candidate genes, however, have been performed in small study populations and were limited to testing only a few variants. Consequently, most of the reported associations have failed to be replicated in subsequent large-scale validation studies. Furthermore, most studies did not correlate genotypes with detailed clinical toxicity data.

In the current study, we therefore aimed to assess prior associations for 26 selected genetic variants in 18 genes, in a large cohort of 322 ovarian cancer patients treated with paclitaxel-carboplatin combination therapy or carboplatin mono-therapy of whom detailed clinical toxicity data were available.

Methods

Study population

All ovarian cancer patients presenting in participating hospitals of the Belgian and Luxembourg Gynaecological Oncology Group (BGOG) were recruited for this study. Collection of germ-line DNA and baseline patient characteristics were collected for each patient. Disease characteristics were recorded after histologic examination with registration of tumor stage according to the International Federation of Gynecology and Obstetrics (FIGO) classification, residual disease after debulking surgery, measurement of tumor size on computed tomography (CT) scans and determination of cancer antigen 125 (CA125) before, during and after chemotherapy. Response to treatment and disease progression were evaluated based on radiologic examination according to the Response Evaluation Criteria in Solid Tumors Group (RECIST) criteria [37]. Paclitaxel was administered at a starting dose of 175 mg/m2 and carboplatin at a starting area under the plasma concentration-versus time curve (AUC) of 5–7 mg/ml/min, with possible dose reductions after the occurrence of severe toxicity. During treatment, the use of erythropoiesis stimulating agents (ESAs) and colony stimulating factors (CSFs) was conform to uniform institutional standards; ESAs are given during treatment with chemotherapy in symptomatic patients with a hemoglobin level below 11 g/dl while CSFs are given if (1) neutropenia grade 4 (ANC <500/mm3) together with fever > 38°C or (2) neutropenia grade 4 (ANC < 500/mm3) during minimum 5 consecutive days. Toxicity during chemotherapy was systematically and routinely scored according to the Common Terminology for Adverse Events (CTCAE) version 4.0. Hematological toxicity was scored based on routinely performed weekly complete blood counts during treatment and before each cycle to determine the nadir of anemia, neutropenia and thrombocytopenia of each administered cycle, neurotoxicity was scored at each clinical-physical examination before each cycle. The scored toxicities for each patient together with all events of neutropenic fever and use of growth factors were systematically recorded in medical electronic records and for the purpose of the present study retrospectively collected by two independent investigators. The highest grade of toxicity over all courses within a patient was reported, if weekly performed blood counts were not available for each administered cycle or if neurotoxicity was not scored for every cycle, the patient was excluded from the analysis. The primary objective of this study was the correlation of genetic variation with the occurrence of hematologic toxicity or neurotoxicity in patients treated with first-line carboplatin with or without paclitaxel. Secondary objectives included the relation between genetic variation and the need for growth factors during treatment with chemotherapy, platinum-free interval (PFI) defined as the time between the last first-line platinum dose and progression, and overall survival (OS). Analyses for PFI and OS were performed in the population receiving a combination of carboplatin and paclitaxel (n = 266) with exclusion of the more favorable prognostic population receiving carboplatin alone, based on clinical prognostic parameters such as FIGO stage, tumor grade and histological subtype. All included patients provided written informed consent before enrollment. The Medical Ethics Committee of the Leuven University Hospitals approved the study (ML6541), serving as central site with the authority to approve the study for all participating sites.

Genotyping

We performed an extensive literature search before start of enrolment to identify genes associated with treatment outcome or toxicity after platinum and/or paclitaxel administration [8-34,36]. We then selected common missense or synonymous mutations in these genes, as well as a number of SNPs that were located in the promoter region of these genes, but have previously been correlated with toxicity after platinum. In addition, we selected 5 additional SNPs previously associated with thalidomide-related neuropathy to investigate their role in repair mechanisms and inflammation in the peripheral nervous system leading to altered neurotoxicity, rather than having a thalidomide-specific contribution to correlated neurotoxicity [36]. Genomic DNA was extracted from the leucocyte fraction of whole blood samples (Qiagen DNeasy blood and tissue kit). All selected SNPs were genotyped using Sequenom MassARRAY technology (Sequenom Inc., CA, USA), as reported previously [38]. Overall 26 SNPs in 18 genes (Table 1) were genotyped with an individual call rate >95% and an overall success rate >98.5%. We genotyped 15 duplicate samples revealing a genotype accuracy exceeding 99%.

Table 1 Overview of the 26 genotyped single nucleotide polymorphisms (SNPs)

Statistical analysis

We calculated median values and inter-quartile ranges for all continuous variables, while frequencies and percentages were calculated for categorical variables. Genotype frequencies were tested for Hardy-Weinberg equilibrium using a 1°-of-freedom χ2-test and considered significant at P < 0.05. Each of the variants were correlated with toxicity events (i.e., the primary objective) using binary logistic regression, while assuming an additive genotypic model. Per-allele odds ratios (OR) and their respective 95% confidence intervals (CI) are reported. Regression analyses were performed without correction for covariates and after correction for relevant covariates, including age and BMI at the time of treatment, dose of carboplatin per cycle (AUC), number of administered cycles and treatment regimen (paclitaxel/carboplatin versus carboplatin alone). For anemia, an additional covariate was included, i.e., use of ESAs, whereas for neutropenia and febrile neutropenia, use of CSFs was included as an additional covariate. Secondary objectives, PFI and OS, were analyzed for 26 variants using Cox-regression analysis, adjusted for age at diagnosis only or fully adjusted for age at diagnosis, FIGO stage, tumor grade, tumor histology and residual disease after debulking surgery and PFS and OS estimates were calculated using Kaplan-Meier method. Additionally, we investigated which of the variants could predict the need for ESAs or CSFs during treatment. All tests were two-sided and statistical significance was set at p = 0.05. The Bonferroni p-value threshold correcting for the multiple testing of 26 SNPs was p < 0.0019. Statistical analyses were performed using SPSS version 19 (SPSS for Windows, Rel. 19.0.0. 2010. Chicago, Illinois, USA: SPSS Inc.)

Results

Study population

Between January 2009 and December 2011 (pre-specified period of 2 years), we recruited 322 ovarian cancer patients treated with 3–6 cycles paclitaxel-carboplatin combination therapy (n = 266) or 3–6 cycles carboplatin mono-therapy (n = 56) (Additional file 1: Figure S1). Of all recruited patients, 99% was Caucasian (Table 2). Hematological toxicity was analyzed in 290 patients, after exclusion of patients for which weekly blood examinations were not available (n = 32). For neurotoxicity, 56 patients treated with carboplatin monotherapy were excluded since the incidence of sensory neuropathy was significantly lower in this population (p < 0.001). One patient with pre-existing sensory neuropathy before start of chemotherapy was additionally excluded, bringing the total number of patients eligible up to 265. For the secondary objectives, PFI and OS, all patients treated with paclitaxel-carboplatin (n = 266) were included. Patient, disease and toxicity characteristics are summarized in Table 2. Briefly, grade 3–4 anemia was present in 57 patients (19.7%), grade 3–4 thrombocytopenia in 57 patients (19.7%), grade 4 neutropenia in 202 patients (69.7%), whereas only 23 patients (7.9%) presented with grade 3–4 febrile neutropenia. In the group of patients selected for neurotoxicity analysis, 48 patients (18.1%) developed grade 2–3 sensory and none motor neuropathy following combination treatment with paclitaxel-carboplatin. Minor allele frequencies (MAF) were similar to those reported previously in Caucasians and adhered to Hardy-Weinberg equilibrium. Allele frequencies of all genotyped SNPs are shown in Additional file 2: Table S1.

Table 2 Patient and disease characteristics, hematologic and neuro-toxicity characteristics

Association with anemia, thrombocytopenia, neutropenia and sensory neuropathy

Among the 290 patients eligible for the hematological toxicity analysis, we observed significant associations for 5 variants (Table 3). In particular, rs1128503 (ABCB1, C > T), rs12762549 (ABCC2, C > G), rs363717 (ABCA1, A > G) and rs11615 (ERCC1, T > C) were significantly associated with grade 3–4 anemia (p = 0.035, OR 1.58; p = 0.005, OR 0.55; p = 0.001, OR 1.31 and p = 0.024, OR = 1.58). After correction for relevant covariates (as explained in the statistical methods), these variants were still significantly associated with toxicity (p = 0.023, OR 1.71; p = 0.004, OR 0.51; p = 0.002, OR 2.08; and p = 0.031, OR 1.61 respectively). Another variant rs4986910 (CYP3A4, T > C) correlated with thrombocytopenia grade 3–4, before and after correction for relevant covariates (p = 0.012, OR 5.61 and p = 0.025, OR 4.99 respectively; Table 3). When correlating each of the variants with grade 4 neutropenia and febrile neutropenia, we did not observe a significant association. Finally, we also correlated each of the variants to sensory neuropathy in the population that was eligible for neurotoxicity analyses, but failed to identify significant associations. None of the observed associations with hematologic toxicity survived correction for multiple testing.

Table 3 Association between genetic variants and hematologic toxicity

Association between genetic variants and use of growth factors

The use of ESAs or CSFs was also correlated with each of the 26 variants to examine whether they could predict the need for an ESA or CSF during treatment with chemotherapy in ovarian cancer. After correction for relevant covariates, a significant correlation for rs2074087 (ABCC1, G > C) and the use of ESA was noticed (p = 0.011, OR 2.09, Table 4). After correction for covariates, CSF use was significantly correlated with rs2073337 (ABCC2, A > G), rs1695 (GSTP1, A > G) and rs1799793 (ERCC2, G > A) (p = 0.039, OR 0.60; p = 0.017, OR 0.55; and p = 0.042, OR 0.63 respectively). None of the observed associations with use of growth factors survived correction for multiple testing.

Table 4 Association between genetic variants and erythropoiesis stimulating agents (ESA) or colony stimulating factor (CSF) use

Effects of variants on PFI and OS

The median follow-up of all patients participating to the study was 2.5 years (95% CI = 2.2-2.8 years) with 157 events for progression (59%) and 84 events for OS (31.6%). Uncorrected P-values were calculated using Cox regression analysis either adjusted for age at diagnosis only or fully adjusted for age at diagnosis, FIGO stage, tumor grade, tumor histology and residual disease after debulking surgery. Only one variant, rs1799793 (ERCC2 G > A), was significantly correlated with PFI in both cases (p = 0.003, HR = 0.71, 95% CI = 0.57-0.89, p = 0.016, HR = 0.75, 95% CI = 0.60-0.95). In particular, Kaplan-Meier survival analysis revealed a significant advantage in PFI for GG carriers of rs1799793 compared to AA or GA carriers (p = 0.016; Figure 1). Variants rs12762549 (ABCC2 A > G) and rs6104 (SERPINB2 C > G) were significantly associated with PFI in the fully-adjusted model (p = 0.037 and p = 0.040, respectively), but these associations were not statistically significant in model adjusted for age only (p = 0.402 and p = 0.219, respectively). No significant correlations were found for OS.

Figure 1
figure 1

Kaplan-Meier curve for platinum-free interval correlated with polymorphisms of rs1799793 in ERCC2. Kaplan-Meier survival analysis reveales a significant advantage in PFI for GG cariers of rs1799793 compared to AA or GA carriers (p = 0.016).

None of the observed associations with platinum-free interval survived correction for multiple testing.

Discussion

We correlated paclitaxel- and carboplatin-induced toxicity with genetic variation in genes involved in pharmacokinetics of these chemotherapeutics or DNA repair, and observed various correlations supporting a role for these genes in mediating toxicity and therapy outcome. We observed that several variants correlated either with grade ≥3 anemia or thrombocytopenia, use of CSFs or ESAs, as well as the platinum-free interval (see Table 5 for an overview of all significant associations). These variants are located in genes that play an important role in transport (ABCB1, ABCC2, ABCC1, ABCA1) and hepatic metabolism (CYP3A4) of paclitaxel and base-excision repair of platinum-induced DNA damage (ERCC1, ERCC2), thus confirming a role for these genes in mediating side-effects and efficacy of paclitaxel and carboplatin. Indeed, ATP-binding cassette transporters, which are expressed on the cell-membrane surface, play an important role in the transport of taxanes [39], whereas cytochrome P450 proteins, CYP2C8, CYP3A4 and CYP3A5, catalyze the oxidative metabolism of taxanes. Furthermore, ERCC1 and ERCC2 are subunits of the endonuclease complex that plays an essential role in DNA repair by removing platinum-induced intra-strand cross-links.

Table 5 Overview of all significant correlations per genetic variant

Most other studies assessing similar correlations have been performed in smaller populations, and typically evaluated only few variants. Our study evaluates a more systematically-selected panel of 26 variants in a large population of ovarian cancer patients, of which 290 were evaluable for hematologic toxicity and 265 for neurotoxicity. Another strength of our study is the availability of more detailed clinical toxicity data compared to previous pharmacogenetic association studies in ovarian cancer, allowing us to correlate specific entities of the hematologic toxicity spectrum, whereas other studies mostly grouped all hematologic > grade 3 toxicities into a single group [8,12,14,15] or focused on the occurrence of neutropenia alone [11,18].

Several of the previously published studies investigating the role of these variants with respect to toxicity and chemotherapy outcome confirmed the observations made in the present study. With respect to the ABC transporters, the rs1128503 (1236C > T) synonymous variant in ABCB1 has been associated with multidrug resistance in multiple studies, and with decreased docetaxel clearance in particular, for homozygous carriers of the variant T-allele in 92 patients [9]. However, its association with severe anemia observed in our study has not been reported before. The rs1045642 (3435C > T) synonymous variant in ABCB1 increased 3′p-hydroxy-paclitaxel metabolites in 23 ovarian cancer patients carrying the T-allele [8]. Vice versa, in a study of 26 patients, a significant greater percent decrease in absolute neutrophil count at nadir was reported for patients homozygous for the T-allele [11], Bergmann also reported a more pronounced neutrophil decrease in patients carrying the T-allele in 92 ovarian cancer patients carrying the T-allele. [18]. In our study, we failed to observe an association with grade ≥3 neutropenia or febrile neutropenia. Possibly, this is due to the fact that we analyzed grade ≥3 neutropenia whilst previous studies used absolute neutrophil decrease. A Japanese study demonstrated that carriers of the variant allele for rs12762549 (ABCC2,101620771 C > G) had an increased risk to develop docetaxel-induced leukopenia/neutropenia in 84 patients [19]. In the current study, no such association was found, but we did find a significant association with anemia and PFI, thereby confirming the potential importance of this variant in mediating taxane transport. Notably, another variant in this gene, rs2073337 (1668 + 148A > G), was significantly correlated with CSF use. For rs363717 (1896 A > G) in ABCA1, which was selected based on its association with thalidomide-related peripheral neuropathy [36], we did not observe a significant association with sensory neurotoxicity. We observed, however, a significant association of this variant with severe anemia (p = 0.001), suggesting that ABCA1 is involved in the transport and metabolism of platinum or carboplatin, similar to its role in the transport of cholesterol [40].

With respect to the CYP genes, low CYP3A4 enzyme activity increased the conversion of paclitaxel towards its metabolite, while heterozygous patients for CYP2C8*3 had a lower clearance of paclitaxel, suggesting the role of those genes in paclitaxel pharmacokinetics in a study of 38 patients [14]. In 93 patients with ovarian cancer, Bergmann et al. observed an 11% reduction in paclitaxel clearance in carriers of the rs10509681 (1196A > G) variant G-allele in CYP2C8 [16], whereas Leskelä et al. observed a correlation between neurotoxicity and these CYP2C8 and CYP3A5 variants in a study consisting of 118 patients [25]. In the present study, we failed however, to observe such associations. Another large study in ovarian cancer also failed to observe correlations with neurotoxicity for these variants in docetaxel or paclitaxel-treated patients [12]. We did observe, however, a significant association between rs4986910 (1331 T > C) in CYP3A4 and thrombocytopenia. Homozygous carriers of the rs1695 (313A > G) variant G-allele in GSTP1 have been associated with neuropathy in 90 patients receiving oxaliplatin-based chemotherapy [30]. This association was not confirmed in our study, although a correlation with febrile neutropenia and CSF use was observed.

Finally, with respect to the excision repair genes, previous studies reported a correlation between severe neutropenia and the rs1799793 (934G > A) variant A-allele in ERCC2 in 104 ovarian cancer patients receiving a cisplatin-cyclophosphamide regimen [32]. In the current study, no correlation with severe neutropenia was described, although the rs1799793 (934G > A) variant A-allele did correlate significantly with CSF use during treatment. Additionally, we observed an improved PFI for rs1799793 (934G > A) GG-carriers. The largest study to date exploring the association between 27 selected variants and ovarian cancer survival, which was performed by the ovarian cancer association consortium (OCAC) in >10,000 cases [41], rs1799793 (934G > A) was not tested. Nevertheless, in the high-grade serous sub-population of this large study, a significant correlation was found with another variant in ERCC2 (rs50872 A > G) and outcome, confirming the potential importance of ERCC2 in mediating chemotherapy outcome. Unfortunately, rs50872 was not linked with rs1799793 (934G > A) (r2 = 0.06), indicating that these variants represent different association signals with ERCC2.

In summary, we observed the strongest associations between variants in ABC-transporters and anemia. The mechanism explaining why altered transport of cytotoxic chemotherapy affects erythropoiesis rather than granulopoiesis or thrombopoiesis is not yet understood. Possibly, these variants alter intracellular concentrations of the transported cytotoxic drug in a cell type-specific manner. Another possibility is that some cell types might be more sensitive to specific changes in the concentration of certain metabolites. On the other hand, it is also possible that the effect on erythropoiesis is caused by a specific role of the affected gene during erythropoiesis. For example, a prominent role for ABCB6 during erythropoiesis as a mitochondrial porphyrin transporter essential for heme biosynthesis, has been established [42].

It is a limitation of the current study that the group for hematologic toxicity analysis included both single agent carboplatin as paclitaxel/carboplatin combination therapy although it is known that both regimens have a slightly different hematologic toxicity profile with more thrombocytopenia in carboplatin monotherapy compared to combination regimens, in our cohort the rate of grade 3–4 thrombocytopenia was 28% for carboplatin versus 18% for combination therapy. Apart from the fact that both regimens are included for hematologic analysis, this cohort is relatively homogenous including only data on first-line treatment in ovarian cancer patients with a uniform ethnicity (99% Caucasians), high number of optimal debulked patients and relatively uniform number of administered cycles of chemotherapy. To further reduce the problem of heterogeneity, pharmacogenetic research on prospective clinical trials including large populations of uniformly treated patients is warranted.

It should be noted that the candidate-gene approach employed so far selecting drug-related genes derived from platinum/taxane pharmacology only allows the analysis with candidate genes known to be involved in chemotherapy metabolism, transport or DNA repair. To discover novel genetic markers, other approaches such as whole-genome association studies or targeted re-sequencing of strong candidate genes to identify rare genetic variants, could be applied. In addition, other drug- or toxicity -related candidate genes relevant for paclitaxel-carboplatin treatment in ovarian cancer (such as GSTA-1 [32], MAD1L1 [43], OPRM1 [44], TRPV1 [44], …) could be selected based on pharmacogenetic knowledge bases such as pharmGKB (www.pharmgkb.org).

Conclusions

The current study revealed a correlation between SNPs in genes involved in DNA repair or metabolism or transport of taxanes or platinum and toxicity or response to first-line chemotherapy in ovarian cancer, using a candidate-gene approach. Variants reported in this study may serve as biomarkers and contribute to the clinical decision-making of chemotherapy dose reductions, feasibility of chemotherapy in patients at-risk based on age and/or performance status, and use of supportive medication such as ESA or CSF. However, as none of the identified associations survived correction for multiple testing, our data are only hypothesis-generating and still need independent validation. We plan to perform such a validation by performing genome-wide screens or targeted re-sequencing of candidate genes in a large multi-centered clinical trial.

Abbreviations

SNP:

Single nucleotide polymorphisms

DNA:

Deoxyribonucleic acid

OR:

Odds ratio

CI:

Confidence interval

HR:

Hazard Ratio

ATP:

Adenosine 5′-triphosphate

ABCA1:

ATP-binding cassette sub-family A, member 1

ABCB1:

ATP-binding cassette, sub-family B, member 1

ABCC1:

ATP-binding cassette, sub-family C, member 1

ABCC2:

ATP-binding cassette sub-family C, member 2

ABCG2:

ATP-binding cassette sub-familiy G, member 2

GSTP1:

Glutathione S-transferase, pi 1

GSTT1:

Glutathione S-transferase theta-1

GSTM1:

Glutathione S-transferase mu-1

CYP3A4:

Cytochrome P450, 3A4

CYP3A5:

Cytochrome P450, 3A5

CYP2C8:

Cytochrome P450, 2C8

CYP1B1:

Cytochrome P450, 1B1

ERCC1:

Excision repair cross-complementation group 1

ERCC2:

Excision repair cross-complementation group 2

XRCC1:

X-ray repair cross-complementing protein 1

SLCO1B3:

Solute carrier organic anion transporter family member 1B3

SLC12A6:

Solute carrier family 12 member 6

MAPT:

Microtubule-associated protein tau

TUBB:

Tubulin, beta class I

TP53:

Tumor protein p53

SERPINB2:

Serpin Peptidase Inhibitor, Clade B (Ovalbumin), Member 2

PPARD:

Peroxisome Proliferator-Activated Receptor Delta

ICAM 1:

Intercellular Adhesion Molecule 1

CSF:

Colony stimulating factor

ESA:

Erythropoiesis stimulating agent

BGOG:

Belgian and Luxembourg Gynaecologic Oncology Group (BGOG)

FIGO:

International Federation of Gynecology and Obstetrics

CT:

Computed tomography

CA 125:

Cancer antigen 125

CTCAE:

Common Terminology for Adverse Events

PFI:

Platinum-free interval

OS:

Overall survival

AUC:

Area under the curve

OC:

Ovarian Cancer

NSCLC:

Non-small cell lung carcinoma

FDR:

False discovery rate

References

  1. Ferlay J, Parkin DM, Steliarova-Foucher E. Estimates of cancer incidence and mortality in Europe in 2008. Eur J Cancer. 2010;46:765–81.

    CAS  PubMed  Google Scholar 

  2. Cannistra SA. Cancer of the ovary. N Engl J Med. 2004;351:2519–29.

    CAS  PubMed  Google Scholar 

  3. du Bois A, Luck HJ, Meier W, Adams HP, Mobus V, Costa S, et al. A randomized clinical trial of cisplatin/paclitaxel versus carboplatin/paclitaxel as first-line treatment of ovarian cancer. J Natl Cancer Inst. 2003;95:1320–9.

    CAS  PubMed  Google Scholar 

  4. du Bois A, Herrstedt J, Hardy-Bessard AC, Muller HH, Harter P, Kristensen G, et al. Phase III trial of carboplatin plus paclitaxel with or without gemcitabine in first-line treatment of epithelial ovarian cancer. J Clin Oncol. 2010;28:4162–9.

    CAS  PubMed  Google Scholar 

  5. Pfisterer J, Weber B, Reuss A, Kimmig R, du Bois A, Wagner U, et al. Randomized phase III trial of topotecan following carboplatin and paclitaxel in first-line treatment of advanced ovarian cancer: a gynecologic cancer intergroup trial of the AGO-OVAR and GINECO. J Natl Cancer Inst. 2006;98:1036–45.

    CAS  PubMed  Google Scholar 

  6. Ozols RF, Bundy BN, Greer BE, Fowler JM, Clarke-Pearson D, Burger RA, et al. Phase III trial of carboplatin and paclitaxel compared with cisplatin and paclitaxel in patients with optimally resected stage III ovarian cancer: a Gynecologic Oncology Group study. J Clin Oncol. 2003;21:3194–200.

    CAS  PubMed  Google Scholar 

  7. Aapro MS, Cameron DA, Pettengell R, Bohlius J, Crawford J, Ellis M, et al. EORTC guidelines for the use of granulocyte-colony stimulating factor to reduce the incidence of chemotherapy-induced febrile neutropenia in adult patients with lymphomas and solid tumours. Eur J Cancer. 2006;42:2433–53.

    CAS  PubMed  Google Scholar 

  8. Nakajima M, Fujiki Y, Kyo S, Kanaya T, Nakamura M, Maida Y, et al. Pharmacokinetics of paclitaxel in ovarian cancer patients and genetic polymorphisms of CYP2C8, CYP3A4, and MDR1. J Clin Pharmacol. 2005;45:674–82.

    CAS  PubMed  Google Scholar 

  9. Bosch TM, Huitema AD, Doodeman VD, Jansen R, Witteveen E, Smit WM, et al. Pharmacogenetic screening of CYP3A and ABCB1 in relation to population pharmacokinetics of docetaxel. Clin Cancer Res. 2006;12:5786–93.

    CAS  PubMed  Google Scholar 

  10. Henningsson A, Marsh S, Loos WJ, Karlsson MO, Garsa A, Mross K, et al. Association of CYP2C8, CYP3A4, CYP3A5, and ABCB1 polymorphisms with the pharmacokinetics of paclitaxel. Clin Cancer Res. 2005;11:8097–104.

    CAS  PubMed  Google Scholar 

  11. Sissung TM, Mross K, Steinberg SM, Behringer D, Figg WD, Sparreboom A, et al. Association of ABCB1 genotypes with paclitaxel-mediated peripheral neuropathy and neutropenia. Eur J Cancer. 2006;42:2893–6.

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Marsh S, Paul J, King CR, Gifford G, McLeod HL, Brown R. Pharmacogenetic assessment of toxicity and outcome after platinum plus taxane chemotherapy in ovarian cancer: the Scottish Randomised Trial in Ovarian Cancer. J Clin Oncol. 2007;25:4528–35.

    CAS  PubMed  Google Scholar 

  13. Johnatty SE, Beesley J, Paul J, Fereday S, Spurdle AB, Webb PM, et al. ABCB1 (MDR 1) polymorphisms and progression-free survival among women with ovarian cancer following paclitaxel/carboplatin chemotherapy. Clin Cancer Res. 2008;14:5594–601.

    CAS  PubMed  Google Scholar 

  14. Green H, Soderkvist P, Rosenberg P, Mirghani RA, Rymark P, Lundqvist EA, et al. Pharmacogenetic studies of Paclitaxel in the treatment of ovarian cancer. Basic Clin Pharmacol Toxicol. 2009;104:130–7.

    CAS  PubMed  Google Scholar 

  15. Kim HS, Kim MK, Chung HH, Kim JW, Park NH, Song YS, et al. Genetic polymorphisms affecting clinical outcomes in epithelial ovarian cancer patients treated with taxanes and platinum compounds: a Korean population-based study. Gynecol Oncol. 2009;113:264–9.

    CAS  PubMed  Google Scholar 

  16. Bergmann TK, Brasch-Andersen C, Green H, Mirza M, Pedersen RS, Nielsen F, et al. Impact of CYP2C8*3 on paclitaxel clearance: a population pharmacokinetic and pharmacogenomic study in 93 patients with ovarian cancer. Pharmacogenomics J. 2011;11:113–20.

    CAS  PubMed  Google Scholar 

  17. Tian C, Ambrosone CB, Darcy KM, Krivak TC, Armstrong DK, Bookman MA, et al. Common variants in ABCB1, ABCC2 and ABCG2 genes and clinical outcomes among women with advanced stage ovarian cancer treated with platinum and taxane-based chemotherapy: a Gynecologic Oncology Group study. Gynecol Oncol. 2012;124:575–81.

    CAS  PubMed  Google Scholar 

  18. Bergmann TK, Brasch-Andersen C, Green H, Mirza MR, Skougaard K, Wihl J, et al. Impact of ABCB1 variants on neutrophil depression: a pharmacogenomic study of paclitaxel in 92 women with ovarian cancer. Basic Clin Pharmacol Toxico. 2012;110(2):199–204.

    CAS  Google Scholar 

  19. Kiyotani K, Mushiroda T, Kubo M, Zembutsu H, Sugiyama Y, Nakamura Y. Association of genetic polymorphisms in SLCO1B3 and ABCC2 with docetaxel-induced leukopenia. Cancer Sci. 2008;99:967–72.

    CAS  PubMed  Google Scholar 

  20. Smith NF, Acharya MR, Desai N, Figg WD, Sparreboom A. Identification of OATP1B3 as a high-affinity hepatocellular transporter of paclitaxel. Cancer Biol Ther. 2005;4:815–8.

    CAS  PubMed  Google Scholar 

  21. Chew SC, Sandanaraj E, Singh O, Chen X, Tan EH, Lim WT, et al. Influence of SLCO1B3 haplotype-tag SNPs on docetaxel disposition in Chinese nasopharyngeal cancer patients. Br J Clin Pharmacol. 2012;73:606–18.

    CAS  PubMed  Google Scholar 

  22. Crouthamel MH, Wu D, Yang Z, Ho RJ. A novel MDR1 G1199T variant alters drug resistance and efflux transport activity of P-glycoprotein in recombinant Hek cells. J Pharm Sci. 2006;95:2767–77.

    CAS  PubMed  Google Scholar 

  23. Vanhoefer U, Cao S, Minderman H, Toth K, Scheper RJ, Slovak ML, et al. PAK-104P, a pyridine analogue, reverses paclitaxel and doxorubicin resistance in cell lines and nude mice bearing xenografts that overexpress the multidrug resistance protein. Clin Cancer Res. 1996;2:369–77.

    CAS  PubMed  Google Scholar 

  24. Huisman MT, Chhatta AA, van Tellingen O, Beijnen JH, Schinkel AH. MRP2 (ABCC2) transports taxanes and confers paclitaxel resistance and both processes are stimulated by probenecid. Int J Cancer. 2005;116:824–9.

    CAS  PubMed  Google Scholar 

  25. Leskela S, Jara C, Leandro-Garcia LJ, Martinez A, Garcia-Donas J, Hernando S, et al. Polymorphisms in cytochromes P450 2C8 and 3A5 are associated with paclitaxel neurotoxicity. Pharmacogenomics J. 2011;11:121–9.

    CAS  PubMed  Google Scholar 

  26. Baker SD, Verweij J, Cusatis GA, van Schaik RH, Marsh S, Orwick SJ, et al. Pharmacogenetic pathway analysis of docetaxel elimination. Clin Pharmacol Ther. 2009;85:155–63.

    CAS  PubMed  Google Scholar 

  27. Pearce CL, Near AM, Van Den Berg DJ, Ramus SJ, Gentry-Maharaj A, Menon U, et al. Validating genetic risk associations for ovarian cancer through the international Ovarian Cancer Association Consortium. Br J Cancer. 2009;100:412–20.

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Gadducci A, Di Cristofano C, Zavaglia M, Giusti L, Menicagli M, Cosio S, et al. P53 gene status in patients with advanced serous epithelial ovarian cancer in relation to response to paclitaxel- plus platinum-based chemotherapy and long-term clinical outcome. Anticancer Res. 2006;26:687–93.

    CAS  PubMed  Google Scholar 

  29. Schildkraut JM, Goode EL, Clyde MA, Iversen ES, Moorman PG, Berchuck A, et al. Single nucleotide polymorphisms in the TP53 region and susceptibility to invasive epithelial ovarian cancer. Cancer Res. 2009;69:2349–57.

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Lecomte T, Landi B, Beaune P, Laurent-Puig P, Loriot MA. Glutathione S-transferase P1 polymorphism (Ile105Val) predicts cumulative neuropathy in patients receiving oxaliplatin-based chemotherapy. Clin Cancer Res. 2006;12:3050–6.

    CAS  PubMed  Google Scholar 

  31. Mir O, Alexandre J, Tran A, Durand JP, Pons G, Treluyer JM, et al. Relationship between GSTP1 Ile(105)Val polymorphism and docetaxel-induced peripheral neuropathy: clinical evidence of a role of oxidative stress in taxane toxicity. Ann Oncol. 2009;20:736–40.

    CAS  PubMed  Google Scholar 

  32. Khrunin AV, Moisseev A, Gorbunova V, Limborska S. Genetic polymorphisms and the efficacy and toxicity of cisplatin-based chemotherapy in ovarian cancer patients. Pharmacogenomics J. 2010;10:54–61.

    CAS  PubMed  Google Scholar 

  33. Khrunin A, Ivanova F, Moisseev A, Khokhrin D, Sleptsova Y, Gorbunova V, et al. Pharmacogenomics of cisplatin-based chemotherapy in ovarian cancer patients of different ethnic origins. Pharmacogenomics. 2012;13:171–8.

    CAS  PubMed  Google Scholar 

  34. Kang S, Ju W, Kim JW, Park NH, Song YS, Kim SC, et al. Association between excision repair cross-complementation group 1 polymorphism and clinical outcome of platinum-based chemotherapy in patients with epithelial ovarian cancer. Exp Mol Med. 2006;38:320–4.

    CAS  PubMed  Google Scholar 

  35. Krivak TC, Darcy KM, Tian C, Bookman M, Gallion H, Ambrosone CB, et al. Single nucleotide polypmorphisms in ERCC1 are associated with disease progression, and survival in patients with advanced stage ovarian and primary peritoneal carcinoma; a Gynecologic Oncology Group study. Gynecol Oncol. 2011;122:121–6.

    CAS  PubMed  Google Scholar 

  36. Johnson DC, Corthals SL, Walker BA, Ross FM, Gregory WM, Dickens NJ, et al. Genetic factors underlying the risk of thalidomide-related neuropathy in patients with multiple myeloma. J Clin Oncol. 2011;29:797–804.

    CAS  PubMed  Google Scholar 

  37. Eisenhauer EA, Therasse P, Bogaerts J, Schwartz LH, Sargent D, Ford R, et al. New response evaluation criteria in solid tumours: revised RECIST guideline (version 1.1). Eur J Cancer. 2009;45:228–47.

    CAS  PubMed  Google Scholar 

  38. Reumers J, De Rijk P, Zhao H, Liekens A, Smeets D, Cleary J, et al. Optimized filtering reduces the error rate in detecting genomic variants by short-read sequencing. Nat Biotechnol. 2012;30:61–8.

    CAS  Google Scholar 

  39. Marsh S. Pharmacogenomics of taxane/platinum therapy in ovarian cancer. Int J Gynecol Cancer. 2009;19 Suppl 2:S30–4.

    PubMed  Google Scholar 

  40. Brooks-Wilson A, Marcil M, Clee SM, Zhang LH, Roomp K, van Dam M, et al. Mutations in ABC1 in Tangier disease and familial high-density lipoprotein deficiency. Nat Genet. 1999;22:336–45.

    CAS  PubMed  Google Scholar 

  41. White KL, Vierkant RA, Fogarty ZC, Charbonneau B, Block MS, Pharoah PD, et al. Analysis of Over 10,000 Cases Finds No Association between Previously Reported Candidate Polymorphisms and Ovarian Cancer Outcome. Cancer Epidemiol Biomarkers Prev. 2013;22:987–92.

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Krishnamurthy PC, Du G, Fukuda Y, Sun D, Sampath J, Mercer KE, et al. Identification of a mammalian mitochondrial porphyrin transporter. Nature. 2006;443:586–9.

    CAS  PubMed  Google Scholar 

  43. Santibanez M, Gallardo D, Morales F, Lopez A, Prada D, Mendoza J, et al. The MAD1 1673 G. A polymorphism alters the function of the mitotic spindle assembly checkpoint and is associated with a worse response to induction chemotherapy and sensitivity to treatment in patients with advanced epithelial ovarian cancer. Pharmacogenet Genomics. 2013;23:190–9.

    CAS  PubMed  Google Scholar 

  44. McWhinney-Glass S, Winham SJ, Hertz DL, Yen Revollo J, Paul J, He Y, et al. Cumulative genetic risk predicts platinum/taxane-induced neurotoxicity. Clin Cancer Res. 2013;19:5769–76.

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

This research project was financially supported by the Federal Public service of health, food chain safety and environment of Belgium, through the initiative ‘Cancer Plan – action 29’ Project KPC_29_054 (KankerPlanCancer@health.fgov.be).

Author information

Authors and Affiliations

Authors

Consortia

Corresponding author

Correspondence to Sandrina Lambrechts.

Additional information

Competing interests

The authors declare that they have no competing interests.

Authors’ contributions

SL: wrote manuscript, performed research, analyzed data; DL: designed research, performed research, analyzed data; ED: contributed new patients and data; EVN: contributed new patients and data; DS: performed research; PRD: contributed new patients and data; VR: contributed new patients and data; PV: contributed new patients and data; DL: contributed new patients and data; FA: contributed new patients and data; PN: contributed new patients and data; KL: contributed new patients and data; IV: designed research, contributed new patients and data. All authors read and approved the final manuscript.

Sandrina Lambrechts and Diether Lambrechts contributed equally to this work.

Additional files

Additional file 1: Figure S1.

Study design for pharmacogenetic analyses.

Additional file 2: Table S1.

Minor allele frequencies of the significant SNPs, calculated for all included patients (n = 322).

Rights and permissions

This article is published under license to BioMed Central Ltd. This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly credited. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lambrechts, S., Lambrechts, D., Despierre, E. et al. Genetic variability in drug transport, metabolism or DNA repair affecting toxicity of chemotherapy in ovarian cancer. BMC Pharmacol Toxicol 16, 2 (2015). https://doi.org/10.1186/s40360-015-0001-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40360-015-0001-5

Keywords