Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2010 Jun 1.
Published in final edited form as: Hum Genet. 2009 Mar 1;125(5-6):527–539. doi: 10.1007/s00439-009-0645-6

Bladder cancer SNP panel predicts susceptibility and survival

Angeline S Andrew 1, Jiang Gui 1, Arthur C Sanderson 2, Rebecca A Mason 3, Elaine V Morlock 3, Alan R Schned 4, Karl T Kelsey 5, Carmen J Marsit 5, Jason H Moore 6, Margaret R Karagas 6
PMCID: PMC2763504  NIHMSID: NIHMS123213  PMID: 19252927

Abstract

Bladder cancer is the fourth most common malignancy in men and the eighth most common in women in western countries. Single nucleotide polymorphisms (SNPs) in genes that regulate telomere maintenance, mitosis, inflammation, and apoptosis have not been assessed extensively for this disease. Using a population-based study with 832 bladder cancer cases and 1,191 controls, we assessed genetic variation in relation to cancer susceptibility or survival. Findings included an increased risk associated with variants in the methyl-metabolism gene, MTHFD2 (OR 1.7 95% CI 1.3–2.3), the telomerase TEP1 (OR 1.8 95% CI 1.2–2.6) and decreased risk associated with the inflammatory response gene variant IL8RB (OR 0.6 95% CI 0.5–0.9) compared to wild-type. Shorter survival was associated with apoptotic gene variants, including CASP9 (HR 1.8 95% CI 1.1–3.0). Variants in the detoxification gene EPHX1 experienced longer survival (HR 0.4 (95% CI 0.2–0.8). These genes can now be assessed in multiple study populations to identify and validate SNPs appropriate for clinical use.

Introduction

In western countries, bladder cancer is the fourth most common malignancy in men and the eighth most common in women (Kirkali et al. 2005). The majority of this disease is attributed to cigarette smoking; bladder cancer risk is up to fourfold higher among cigarette smokers compared with non-smokers (Kirkali et al. 2005). A hereditary component is likely since a family history of bladder cancer and variations in genes that detoxify aromatic amines are associated with increased risk (Garcia-Closas et al. 2005; Kantor et al. 1985; Murta-Nascimento et al. 2007b). We and others have also identified associated variations in DNA repair pathway genes, yet single nucleotide polymorphisms (SNPs) in important cancer regulatory pathways such as cell growth and apoptosis have not been comprehensively assessed for this disease (Andrew et al. 2006; Wu et al. 2006).

Bladder cancer generally carries a favorable prognosis, nevertheless, approximately 13,750 deaths occurred in the USA from bladder cancer in 2007 (Jemal et al. 2007). According to the US Statistics Epidemiology and End Results (SEER) program (1996–2002), the 5-year survival rate for localized disease is 94%, whereas for regional staged cancers the survival rate is only 47%. Those with distant metastasis at the time of diagnosis carry the lowest survival rate of 6% (Jemal et al. 2007). Established prognostic factors include tumor characteristics such as degree of invasion or carcinoma in situ (CIS) (Carroll et al. 2001). Higher grade of a tumor is also associated with a worse prognosis (Schned et al. 2007). Although these histopathologic factors can be used to predict prognosis, the behavior of each group is quite heterogeneous [reviewed in (Murta-Nascimento et al. 2007a)]. This inter-individual variation in survival highlights the potential for new genetic markers that can assist physicians in selecting appropriately aggressive treatment regimens.

We have assessed the role of genetic variation in bladder cancer risk and case survival using panel of SNPs in hypothesized cancer regulatory pathways, including cell cycle, cell growth, detoxification, telomerase, and apoptosis. Risk- or survival-associated single gene and multi-gene-combinations were ranked and tested using traditional multivariate statistical methods in a large, population-based case–control study in New Hampshire, USA.

Materials and methods

Study group

We identified all cases of bladder cancer among New Hampshire residents, ages 25–74 years from the State Cancer Registry. The study was conducted in two sequential phases based on cancer diagnosis date (Phase I 1 July 1994–30 June 1998, Phase II 1 July 1998–30 June 2001). Detailed methods have been described previously (Karagas et al. 1998). Briefly, we interviewed a total of 832 bladder cancer cases, which was 85% of the cases confirmed to be eligible for the study. Controls less than 65 years of age were selected using population lists obtained from the New Hampshire Department of Transportation. Controls 65 years of age and older were chosen from data files provided by the Centers for Medicare & Medicaid Services (CMS) of New Hampshire. For efficiency, we shared a control group with a study of non-melanoma skin cancer in New Hampshire covering an overlapping diagnostic period of 1 July 1993–30 June 1995 (Karagas et al. 1998). We selected additional controls for bladder cancer cases diagnosed from 1 July 1995 to 30 June 1998 frequency matched to these cases on age (25–34, 35–44, 45–54, 55–64, 65–69, 70–74 years) and gender. Most (>95%) of the subjects in this study are of Caucasian origin; and thus our analyses were not appreciably altered by restricting to Caucasians. We interviewed a total 1,191 controls (the total shared control group and additional controls), which was 70% of the controls confirmed to be eligible for the study.

Personal interview

Informed consent was obtained from each participant and all procedures and study materials were approved by the Committee for the Protection of Human Subjects at Dartmouth College. Consenting participants underwent a detailed in-person interview, usually at their home. Questions covered sociodemographic information (including level of education), lifestyle factors such as use of tobacco (including frequency, duration and intensity of smoking), family history of cancer, and medical history prior to the diagnosis date of the bladder cancer (cases) or reference date assigned to controls. Recruitment procedures for both the shared controls from the non-melanoma skin cancer study and additional controls were identical and ongoing concomitantly with the case interviews. Case–control status and the main objectives of the study were not disclosed to the interviewers. To ensure consistent quality of the study interviewer, interviews were tape recorded with the consent of the participants and routinely monitored by the interviewer supervisor. To assess comparability of cases and controls, we asked subjects if they currently held a driver's license or a Medicare enrollment card. Subjects were asked to provide a blood sample (buccal sample was requested in the case of a refusal). Samples were maintained at 4°C and sent via courier to the study laboratory at Dartmouth within 24 h for processing and analysis.

Genotyping

DNA was isolated using Qiagen genomic DNA extraction kits (QIAGEN Inc, Valencia, CA) from peripheral circulating blood lymphocyte specimens harvested at the time of interview. Genotyping was performed on all DNA samples of sufficient concentration (893 controls, 617 cases), using the Cancer Panel on the GoldenGate Assay system by Illumina's Custom Genetic Analysis service (Illumina Inc., San Diego, CA). The Cancer Panel contains 1,421 SNPs in approximately 400 hypothesized cancer-related genes from the SNP500 database (http://www.illumina.com). Genotype data were available on 75% of interviewed subjects and the genotyped population had similar characteristics as the overall study. Samples repeated on multiple GoldenGate plates yielded the same call for 99.9% of SNPs and 99.5% of samples submitted were successfully genotyped. Genotype calls were 99% concordant between genotyping platforms.

Statistical analysis

The main goals of the statistical analysis were to identify SNPs and combinations of SNPs that relate to bladder cancer susceptibility or survival. Selection of SNPs with dose–responsive relationships to bladder cancer risk was performed by ranking SNPs based on p-for-trend values associated with a dose response of variant alleles with FDR value ≤0.56.

Kullback–Leibler distance measures were used to compare discrete distributions of feature values for cases and controls (Duda et al. 2001). The Kullback–Leibler divergence provides an information-based measure of the diVerence between two probability distributions. The Kullback–Leibler divergence may be expressed as the number of bits of information required to encode samples based on the two distributions. The Kullback–Leibler divergence differs from a conventional distance metric on the space of probability distributions in that it is not symmetric and does not satisfy the triangle inequality. As an information measure, the Kullback–Leibler divergence may be thought of as a non-linear measure that often accentuates distribution difference properties that are not detected by conventional distance metrics (Kullback 1959, 1987). While the divergence cannot be easily generalized in its performance, it is often a valuable complement to conventional distance metrics in applications where well-defined models are not easily available. Kullback–Leibler divergence has recently been to be used in studies of genetic association [e.g. (Cussenot et al. 2008)]. The SNPs were ranked and the top 20 SNPs with the largest scores were selected.

Two, three, and four-way combinations of p-for-trend and Kullback–Liebler selected potentially interacting SNPs were identified using Multifactor Dimensionality Reduction (MDR) analysis (Ritchie et al. 2001). To allow for epistatic interactions in the absence of main effects, we used a more liberal p-for-trend cutoff for entry into the MDR model (FDR value 0.83). MDR is a data reduction (i.e. constructive induction) approach that seeks to identify combinations of multilocus genotypes and discrete environmental factors that are associated with either high risk or low risk of disease. Thus, MDR defines a single variable that incorporates information from several loci and/or environmental factors that can be divided into high risk and low risk combinations. This new variable can be evaluated for its ability to classify and predict outcome risk status using cross-validation and permutation testing (Ritchie et al. 2001, 2003).

We also used the tuned Relief F as a pre-filtering step to help identify epistatic or non-additive, interacting SNPs in MDR. ReliefF considers SNPs jointly rather than individually. In our experience, this method is better for detecting interactions than chi-square test. ReliefF estimates weights for each SNP “estimated based on whether the nearest neighbors (nearest hit, H) of a randomly selected instance from the same class and the nearest neighbors from the other class (nearest miss, M) have the same or different values” (Moore and White 2007). Tuned ReliefF uses vectors of SNPs, removes those of low quality, re-estimates, and then ranks them based on their ability to discriminate case–control status.

To assess the magnitude of the independent effects of the SNPs that were identified by the filters described above, we conducted logistic regression analyses for individuals with one or two variant alleles in comparison to those homozygous wild type for each individual SNP. These analyses were adjusted for age (less than or greater than median age 64), gender, and smoking status (never, former, current). We evaluated the nature of interactions between bladder cancer risk factors (gender, smoking, age) and/or the SNP allele combinations predicted by MDR, by including interaction terms in a logistic regression model. Statistical significances of the interactions were assessed using likelihood ratio tests comparing the models with and without the interaction terms (e.g. logit(p) = constant + SNP1 + SNP2 vs. logit(p) = constant + SNP1 + SNP2 + interaction). MDR identifies combinations of predictive factors, but does not distinguish between multiplicative and additive effects. Thus, we used the likelihood ratio test to assess the multiplicative nature of the interactions. p values represent two-sided statistical tests with statistical significance at P < 0.05. Our analytic strategy embraces both the algorithmic and traditional statistical approaches [discussed in (Breiman 2001)], allowing us to identify epistatic SNP interactions.

Selection of SNPs related to bladder cancer survival was performed using an FDR threshold <0.5, choosing the four individual SNPs with the lowest log-rank p values (≤0.1). Then we used Gene Set Assessment (GSA) to investigate groups of SNPs that are members of a common gene. The GSA method calculates an enrichment score based on whether the gene-based SNP sets contain a number of individual SNPs that are highly associated with cancer (over-represented), with adjustment for multiple testing (Holden et al. 2008). GSA was used to select the top 10 ranking genes (each containing multiple SNPs) by log-rank p values. Survival plots for bladder cancer cases were generated using the Kaplan-Meier method and differences between genotypes were assessed using the log-rank test. To adjust for additional factors related to patient survival, Cox-proportional hazards regression analysis was performed with age, gender, smoking (never, former, current), as well as AJCC tumor stage (0, 0is, 1–4), grade (1–4), size (<, ≥30 mm), treatment (immunotherapy with surgery; chemotherapy, radiotherapy, or immunotherapy only; chemotherapy or radiotherapy with cystectomy; transurethral resection only; cystectomy only) in the model. Results were not sensitive to removal of subjects with metastatic tumors from the analysis (data not shown). p values represent two-sided statistical tests with statistical significance at P < 0.05.

Results

The characteristics of the subjects who provided a DNA sample were similar to those of the overall study population. The majority of the population was Caucasian. The case group contained a higher percentage of current smokers than the controls (Table 1). A large proportion of the cases were non-invasive bladder cancer, reflecting the population-based nature of the study. Testing showed deviation from Hardy–Weinberg equilibrium for ADH1C_14, and MSH6_01.

Table 1.

Genotyped population characteristics

Controls n (%) Cases n (%) p value
Gender
     Female 338 (38) 150 (24) Ref.
     Male 555 (62) 467 (76) <0.0001
Reference age
     <40 43 (5) 17 (3) Ref.
     40–55 152 (17) 102 (17) 0.4
     55–70 480 (54) 337 (54) 0.3
     >70 218 (24) 161 (26) 0.2
Race
     White 889 (99) 607 (98) Ref.
     Non-white 4 (1) 10 (2) 0.06
Smoking status
     Never 305 (34) 106 (17) Ref.
     Former 447 (50) 305 (50) <0.0001
     Current 141 (16) 205 (33) <0.0001
AJCC Stage
     0 337 (56)
     0is 24 (4)
     1 170 (28)
     2 43 (7)
     3, 4 33 (5)

Data were missing for smoking on 1 case, stage on 10 cases

Risk models

The p-for-trend method that hypothesizes a dose–response effect by the number of variant alleles was used to select the individual SNPs shown in Table 2. Individuals homozygous for the variant forms of the alcohol dehydrogenase SNP ADH1C_14 had an increased risk of bladder cancer compared with wild type [OR 1.7 95% CI (1.3–2.2)]. Increased risk was also observed among individuals with the homozygous variant form of SNPs in MTHFD2, TEP1, AURKA, FZD7, IGF1, MSH6. Gene variants associated with reduced risk included IL8RB, MET, SLC4A2, MBL2 (Table 2). Of the SNPs selected in the p-for-trend analysis, only the risk associated with ADH1C_14 varied by smoking status. Never smokers had modestly higher bladder cancer risk OR 1.4 (95% CI 0.9–2.3) than current smokers OR 0.9 (95% CI 0.6–1.4) (gene—current smoker interaction p = 0.05). We also analyzed these SNPs separated by study phase (I, II). Results within each phase were consistent with our analysis of the overall population.

Table 2.

Individual SNPs selected for association with bladder cancer risk by p-for-trend <0.004

SNP (function) Genotype Controls
Cases
Bladder cancer
OR (95% CI)
N % N %
ADH1C_14 (Alcohol dehydrogenase) wt 440 51 256 43 1.0 Ref.
rs2241894 het 264 31 181 31 1.2 (0.9–1.5)
g.Ex5+106A>G var 156 18 152 26 1.7 (1.3–2.2)
p.T151T any var 420 49 333 57 1.4 (1.1–1.7)
MTHFD2_01 (Methyl-metabolism) wt 290 34 146 25 1.0 Ref.
rs1667627 het 389 45 292 49 1.5 (1.2–1.9)
g.IVS1+3323T>C var 183 21 152 26 1.7 (1.3–2.3)
any var 572 66 444 75 1.6 (1.2–2.0)
TEP1_08 (Telomerase) wt 452 51 261 42 1.0 Ref.
rs1760897 het 366 41 280 45 1.3 (1.0–1.6)
g.Ex1−222T>C var 75 8 75 12 1.8 (1.2–2.6)
p.S116P any var 441 49 355 58 1.4 (1.1–1.7)
IL8RB_01 (IL-8 Receptor, Angiogeneisis, CXCR2) wt 299 35 250 42 1.0 Ref.
rs1126579 het 414 48 255 43 0.7 (0.6–0.9)
Ex3+1235T>C var 150 17 84 14 0.6 (0.5–0.9)
any var 564 65 339 58 0.7 (0.6–0.9)
AURKA_15/STK6_15b (Cell cycle kinase, spindle poles in mitosis) wt 395 46 224 39 1.0 Ref.
rs8173 het 356 42 252 44 1.2 (1.0–1.6)
g.Ex11−347G>C var 99 12 92 16 1.6 (1.2–2.3)
any var 455 54 344 61 1.3 (1.1–1.6)
FZD7_10 (Wnt receptor, decreases function of antigen presenting cells) wt 453 51 274 44 1.0 Ref.
rs13034206 het 363 41 272 44 1.3 (1.0–1.6)
g.Ex1−1926C>T var 73 8 70 11 1.7 (1.1–2.4)
any var 436 49 342 55 1.3 (1.1–1.7)
MET_01 (Oncogene, hepatocyte growth factor receptor) wt 240 28 198 34 1.0 Ref.
rs41736 het 433 50 291 49 0.8 (0.6–1.0)
g.Ex20+60C>T var 192 22 102 17 0.6 (0.5–0.9)
p.D1304D any var 625 78 393 66 0.7 (0.6–0.9)
IGF1_11 (Somatomedin, growth factor) wt 492 57 294 50 1.0 Ref.
rs5742629 het 325 38 250 42 1.3 (1.0–1.6)
g.IVS2+12158A>G var 48 6 47 8 1.7 (1.1–2.6)
any var 373 44 297 50 1.4 (1.1–1.7)
FZD7_16 (Wnt receptor, decreases function of antigen presenting cells) wt 451 52 269 46 1.0 Ref.
rs12474408 het 352 41 261 44 1.3 (1.0–1.6)
g.*2710A>G var 62 7 60 10 1.7 (1.1–2.5)
any var 414 48 321 54 1.3 (1.1–1.6)
MSH6_01 (DNA mismatch recognition) wt 319 37 192 33 1.0 Ref.
rs3136228 het 376 44 248 42 1.1 (0.9–1.4)
g.−556G>T var 169 20 150 25 1.6 (1.2–2.1)
any var 545 64 398 67 1.3 (1.0–1.6)
GSK3B_45 (Kinase energy metabolism) wt 539 62 319 54 1.0 Ref.
rs3755557 het 266 31 222 38 1.5 (1.2–1.9)
−2675T>A var 58 7 49 8 1.3 (0.8–2.0)
any var 324 38 271 46 1.5 (1.2–1.8)
SLC4A2_04 (Solute exchange) wt 514 60 396 67 1.0 Ref.
rs13240966 het 312 36 177 30 0.8 (0.6–1.0)
g.IVS1−194C>G var 35 4 18 3 0.6 (0.3–1.1)
any var 347 40 195 33 0.7 (0.6–0.9)
MBL2_30 (Innate immunity) wt 457 54 346 59 1.0 Ref.
rs2099902 het 329 38 221 37 0.9 (0.7–1.1)
g.Ex4−710G>A var 66 8 23 4 0.5 (0.3–0.8)
any var 395 46 244 41 0.8 (0.6–1.0)
TERT_15 (Telomerase reverse transcriptase) wt 659 77 422 71 1.0 Ref.
rs13167280 het 182 21 152 26 1.4 (1.1–1.8)
g.IVS3−24T>C var 16 2 17 3 1.7 (0.8–3.4)
any var 198 23 169 29 1.4 (1.1–1.8)

Analyses were adjusted for gender, age, smoking. Genotype data were missing for ADH1C_14 on 5 controls, 28 cases; MTHFD2_01 3 controls, 27cases; TEP1_08 0 controls, 1 case; IL8RB_01 30 controls, 28 cases; AURKA_15 43 controls, 49 cases; FZD7_10 4 controls, 1 case; MET_01 28 controls, 26 cases; IGF1_11 28 controls, 26 cases; FZD7_16 28 controls, 27 cases; MSH6_01 29 controls; 27 cases; GSK3B_45 30 controls, 27 cases; SLC4A2_04 32 controls, 26 cases; MBL2_30 41 controls, 27 cases; TERT_15 36 controls, 26 cases

wt wildtype, het heterozygous, var variant, any variant heterozygous and variant, g. gene, Ex exon, p. protein

To identify SNPs that do not follow the p-for-trend risk model, we used the Kullback–Leibler distance algorithm to rank SNPs for an association with bladder cancer risk (heterozygous (het.) or variant (var.) vs. wildtype). Of the top 20 SNPs selected by Kullback–Leibler distance, five showed individual gene relationships with bladder cancer incidence when assessed by logistic regression analysis after adjustment for age, gender, and smoking status, GSTZ1_02, AKR1C3_35, TYR_02, SCARB1_03, and SLC23A1_05 (Table 3). The other SNPs that had top ranking Kullback–Leibler distances did not show a significant association with bladder cancer by logistic regression analysis with adjustment for potential confounders, including SCARB1_03, PLA2G6_08, PTGS1_02, EPHX1_17, SLC23A2_31, SLC6A3_14, TGM1_01, AKR1C3_11, KRAS_22, CTNNB1_02, IL10_06, HMGCR_01, BARD1_02, AXIN2_09, PTGS1_02, MSH2_06 (data not shown).

Table 3.

Individual SNPs selected for association with bladder cancer risk by Kullback–Leibler distance

Function SNP Genotype Controls
Cases
Bladder cancer
OR (95% CI)
N % N %
Detoxification GSTZ1_02 wt 636 86 367 80 1.0 Ref.
g.Ex5−12G>A het 99 14 88 19 1.5 (1.1–2.1)
p.G42R var 5 1 4 1 1.4 (0.4–5.3)
any var 104 15 92 20 1.5 (1.1–2.0)
Aldo/keto reductase AKR1C3_35 wt 509 59 312 53 1.0 Ref.
PAH metabolism g.*12259G>A het 315 37 238 41 1.3 (1.0–1.6)
var 40 4 38 6 1.5 (0.9–2.4)
any var 355 41 276 47 1.3 (1.0–1.6)
Pigmentation TYR_02 wt 463 53 299 51 1.0 Ref.
Melanin production g.IVS3−6895A>G het 322 37 255 43 1.2 (1.0–1.5)
var 80 9 36 6 0.7 (0.5–1.2)
any var 402 46 291 49 1.1 (0.9–1.4)
Lipid metabolism SCARB1_03 wt 383 44 235 40 1.0 Ref.
g.IVS1−18462G>A het 381 44 290 49 1.2 (0.9–1.5)
var 101 12 66 11 1.0 (0.7–1.4)
any var 482 56 356 60 1.1 (0.9–1.4)
Metabolism SLC23A1_05 wt 813 94 545 92 1.0 Ref.
g.Ex8+22G>A het 50 6 46 8 1.5 (1.0–2.3)
p.V264M var 2 0
any var 52 6 46 1.5 (0.9–2.2)

Analyses were adjusted for gender, age, smoking. Genotype data were missing for GSTZ1_02 on 153 controls, 158 cases; AKR1C3_35 on 29 controls, 29 cases; TYR_02 on 28 controls, 27 cases; SCARB1_03 on 28 controls, 26 cases; SLC23A1_05 on 28 controls, 26 cases

wt wildtype, het heterozygous, var variant, any variant heterozygous and variant, g. gene, Ex exon, p. protein

We then identified potential interactions between these selected SNPs using MDR, followed by quantification of the odds ratios associated with the high versus low risk combinations of alleles/characteristics using logistic regression analysis with adjustment for age, gender and smoking status. Of the p-for-trend selected genes using an FDR value cutoff of 0.8, MDR indicated that a combination of the high-risk genotypes for PIN1_21 rs889162 g.IVS3+2592T>C, MET_01 rs41736 g.Ex20+60C>T p.D1304D, and AMACR_17 rs10941112 g.Ex3-29A>G p.G175D was significantly associated with bladder cancer (MDR accuracy 0.56, cross-validation consistency 7/10). The logistic regression adjusted OR was 2.3 [95% CI 1.8–2.9)] for the high risk alleles, versus the low risk genotype combination. Of these PIN1_21 and AMACR_17 indicated the strongest evidence of synergy (likelihood ratio test gene–gene interaction p value = 0.02).

We also entered the top ranking 20 Kullback–Leibler distance selected genes, along with age, gender and smoking status, into MDR. The combination of high-risk genotypes for SLC23A2_31 rs12479919 g.IVS1+1312G>A, AKR1C3_11 rs2275928 g.IVS8+40A>G, SCARB1_03 rs4765621 g.IVS1-18462G>A, and PLA2G6_08 rs2016755 g.IVS3-309T>C was associated with elevated bladder cancer risk (logistic regression adjusted OR 2.6 (95% CI 2.1–3.2), MDR accuracy 0.53, cross-validation consistency 6/10). Within this MDR-selected group, we tested possible combinations of gene-gene interactions and found a significant relationship between SCARB1_03 and AKR1C3_11 (interaction p value = 0.04). This MDR analysis also picked the strongest two-factor interaction from the Kullback–Leibler selected genes, between AKR1C3_11 and IL10_06 (MDR accuracy 0.56, cross-validation consistency 6/10), adjusted OR 1.6 (95% CI 1.3–2.0), gene–gene interaction p value = 0.0004. Using the tuned ReliefF filter in MDR, we also identified CTNNB1_02 rs11564452 g.IVS7-562A>T, SCARB1_03 rs4765621 g.IVS1-18462G>A, SLC23A2_31 rs12479919 g.IVS1+1312G>A, and BARD1_02 rs1129804 g.Ex1+44C>G as a high risk combination (MDR accuracy 0.56, cross-validation consistency 10/10). The logistic regression adjusted OR was 2.3 (95% CI 1.7–2.9) for the high versus low risk allele combination. Of these, SCARB1_03 and SLC23A2_31 showed the strongest evidence of a gene–gene effect (interaction P = 0.02) on bladder cancer risk.

Additionally, we analyzed the data using a χ2 filter on the entire dataset. Entering the χ2 top-ranked 40 genes into MDR revealed an interaction between SLC19A1_01 rs1051266 g.Ex4-114T>C p.H27R, and IGF2AS_04 rs3741212 g.Ex1+112A>G, MDR accuracy 0.60, cross-validation consistency 10/10. The combined any variant adjusted odds ratio for this gene–gene combination from logistic regression was 1.4 (95% CI 1.1–1.7) (interaction P < 0.0001).

Survival models

We used the log-rank test to identify the top four individual SNPs likely to be related to bladder cancer case survival (Table 4). Cox-regression analysis (adjusted for age, gender, smoking status, stage/grade, and treatment) suggested that lower survival was associated with having SNPs in the T lymphocyte regulator CD80_04, the apoptosis regulator BCL2L1_03, and the ERCC4_01 nucleotide excision repair gene. The polycyclic aromatic hydrocarbon (PAH) metabolism SNP, EPHX1_15 was associated with better prognosis (Fig. 1). The effects did not vary significantly by smoking status (e.g. ERCC4 among never smokers HR 10.6 (95% CI 2.8–38), ever smokers HR 1.9 (95% CI 1.2–3.1), interaction P = 0.2).

Table 4.

Individual SNPs selected for relationship with bladder cancer survival

Function SNP Genotype Cases
Deaths
Bladder cancer
HR (95% CI)
Log rank test
p value
N % N %
Surface antigen CD80_04 wt 321 77 116 68 1.0 (ref) 0.008
rs9282638 het 93 22 54 31 1.9 (1.4–2.8)
g.IVS2−56G>A var 5 1 1 1 1.5 (0.2–10.9)
any var 98 23 55 32 1.9 (1.4–2.7)
Apoptosis BCL2L1_03 0.01
rs1994251 wt 265 63 85 50 1.0 (ref)
g.IVS2+22130A>C het 133 32 76 45 1.5 (1.1–2.1)
var 21 5 8 5 1.5 (0.7–3.3)
any var 154 37 84 50 1.5 (1.1–2.1)
Detoxify/activate polycyclic aromatic hydrocarbons EPHX1_15
rs2854461

wt

164

14

85

50

1.0 (ref)
0.02
g.−4786A>C het 198 47 74 43 0.8 (0.5–1.1)
var 57 14 13 7 0.5 (0.2–0.9)
any var 255 61 87 50 0.7 (0.5–1.0)
DNA repair ERCC4_01 0.1
rs1800067 wt 371 89 143 83 1.0 (ref)
g.Ex8+31G>A het 45 11 27 16 1.7 (1.1–2.8)
p.R415Q var 3 1 2 1 5.6 (1.4–23.4)
any var 48 12 29 17 1.8 (1.2–2.9)

Analyses were adjusted for gender, age, smoking, treatment, tumor stage, grade, tumor size. Genotype data were missing for CD80_04 27 cases; BCL2L1_0329 cases; EPHX1_15 26 cases; ERCC4_01 26 cases

wt wildtype, het heterozygous, var variant, any variant heterozygous and variant, g. gene, Ex exon, p. protein

Fig. 1.

Fig. 1

Bladder cancer survival in relation to SNPs. Kaplan–Meier plots show survival (y-axis) versus years from diagnosis (x-axis) by a EPHX1_15 genotype, b IL8RB_01 genotype, c PGR_05 genotype, and d CASP9_01 genotype. Black lines are homozygous wildtype, red are homozygous variant, green are heterozygous

Gene Set Assessment allowed us to use all the SNP data from each gene as a group to identify the genes that most strongly predicted survival by log-rank test (Table 5). We then evaluated the individual SNPs within the top 10 ranked genes using adjusted Cox-regression models. The variant form of the transcriptional regulator GATA3_29 was associated with poor survival HR 2.8 (95% CI 1.6–5.0). The same apoptotic gene BCL2L1 selected by this GSA analysis was also picked in the individual SNP log-rank test (BCL2L1_03). An additional SNP in this gene, BCL2L1_01, was also related to poor bladder cancer survival [any variant HR 1.6 (1.1–2.2)]. This GSA analysis also predicted poor survival for individuals with SNPs in the angiogenic gene IL8RB_01 [any variant HR 1.9 (1.3–2.7)] (Fig. 1), the mitotic genes PLK1_15 [any variant HR 1.3 (0.9–1.8)] and STK6_06 [any variant HR 1.4 (0.9–1.9)], the metabolism gene UGT1A1_24 [any variant HR 1.4 (1.0–1.9)], the apoptotic gene CASP9_01 [any variant HR 1.7 (1.2–2.4)] (Fig. 1) and better survival with the steroid receptor PGR_05 [any variant HR 0.7 (0.5–1.0)] (Fig. 1), and the telomerase TERT_02 [any variant HR 0.8 (0.6–1.2)]. The selected SNPs did not diVer significantly by smoking status (data not shown).

Table 5.

SNPs in genes selected by gene set analysis for relationship with bladder cancer survival

Function SNP Genotype Cases
Deaths
Bladder cancer
HR (95% CI)
Log rank test
p value
N % N %
Transription factor GATA3_29 <0.001
Growth control rs528778 wt 261 62 104 61 1.0 (ref)
Epithelial cell differentiation g.IVS4+582C>T het 150 34 51 30 0.9 (0.67–1.3)
var 8 2 16 9 2.8 (1.6–5.0)
any var 158 36 67 39 1.1 (0.8–1.5)
Interleukin 8 receptor IL8RB_01 0.003
rs1126579 wt 194 47 56 33 1.0 (ref)
g.Ex3+1235T>C het 173 42 82 48 1.7 (1.2–2.5)
var 50 12 34 20 1.7 (1.1–2.8)
any var 233 34 116 68 1.7 (1.2–2.5)
Mitotic kinase PLK1_15 0.1
Spindle formation rs40076 wt 260 62 91 53 1.0 (ref)
g.IVS3+26A>G het 138 33 70 41 1.2 (0.8–1.7)
var 21 5 10 6 1.0 (0.5–2.1)
any var 159 38 80 47 1.2 (0.8–1.6)
Glucuronidation UGT1A1_24 0.1
Lipophilic molecules rs1042640 wt 270 65 93 55 1.0 (ref)
g.Ex5−402G>C het 123 30 67 40 1.4 (1.0–2.0)
var 22 5 8 5 1.4 (0.6–3.1)
any var 155 35 75 45 1.4 (1.0–2.0)
Progesterone receptor PGR_05 0.005
rs613120 wt 120 29 34 20 1.0 (ref)
g.IVS2−1671T>C het 207 50 81 47 1.2 (0.8–1.9)
var 91 22 57 33 1.4 (0.9–2.3)
any var 298 77 138 80 1.3 (0.8–2.0)
Mitotic kinase AURKA/STK6_06 0.1
Chromosome distribution rs6024840 wt 176 56 68 47 1.0 (ref)
g.IVS7−80T>C het 123 39 70 48 1.4 (1.0–2.1)
var 15 5 8 5 1.1 (0.5–2.5)
any var 138 44 78 53 1.4 (1.0–2.0)
Telomerase TERT_02 0.1
rs6024840 wt 158 39 82 48 1.0 (ref)
g.IVS7−80T>C het 197 48 75 44 0.8 (0.6–1.1)
var 55 13 15 8 0.5 (0.3–1.0)
any var 252 61 90 52 0.8 (0.5–1.0)
Apoptosis CASP9_01 0.003
rs1052576 wt 203 50 55 34 1.0 (ref)
g.Ex5+32A>G het 148 36 77 46 1.3 (0.9–2.0)
p.Q221R var 59 14 35 21 1.4 (0.8–2.4)
any var 207 50 112 67 1.3 (0.9–2.0)

Analyses were adjusted for gender, age, smoking, treatment, tumor stage, grade, tumor size. Genotype data were missing for GATA3_29 27 cases; IL8RB_01 28 cases; PLK1_15 27 cases; UGT1A1_24 34 cases; PGR_05 27 cases; AURKA 157 cases; TERT_02 35 cases; CASP9_01 40 cases

wt wildtype, het heterozygous, var variant, any variant heterozygous and variant, g. gene, Ex exon, p. protein

Discussion

A genetic component for bladder cancer susceptibility is likely since individuals with a family history have a higher incidence. Furthermore, although cigarette smoking is the strongest risk factor, it is not clear why some smokers get cancer and others do not (Ahsan and Thomas 2004). The magnitude of the association between the SNPs that are well accepted to be predictive (such as the N-acetyltransferases and DNA repair genes) and bladder cancer risk is relatively small. Thus, there are likely to be other causal genetic factors that remain to be discovered. Bladder cancer survival also varies widely between individual patients, even within the same stage and grade of the tumor. We hypothesize that this survival is in part genetically controlled, yet reliable predictors have not been identified or validated.

The occurrence and prognosis for bladder cancer and other relatively common diseases is most likely controlled by a combination of multiple genetic factors and exposures, rather than a single polymorphism. Our strategy was to predict the single SNPs that have the strongest relationships with bladder cancer incidence or survival. The magnitude of these relationships was then evaluated using traditional logistic regression analysis procedures. We then identified the combinations of the top ranked SNPs and genes using multi-factor selection algorithms for the most predictive high risk sets of alleles. These models were then assessed statistically by traditional methods. Using this approach, we identified a number of SNPs and SNP combinations that were associated with bladder cancer incidence and survival. Modifying SNPs occurred in biologically plausible genes that regulate metabolism, DNA repair, telomere maintenance, mitosis, inflammation, and apoptosis.

Alcohol dehydrogenase 1C (ADH1C) metabolizes multiple substrates, most notably ethanol. Consistent with our data, a previous study in the Netherlands suggested an increased risk of bladder cancer among variants, but no interaction with alcohol consumption (van Dijk et al. 2001). UGT1A1 is a phase 2 enzyme which metabolizes lipophilic molecules (e.g. steroids, bilirubin, hormones, drugs, and aromatic amines such as benzidine) via glucuronidation for excretion (Zenser et al. 2002) and was related to worse survival. Polymorphisms in the one-carbon methyl-metabolism gene methylenetetrahydrofolate reductase (MTHFR) have been well studied in relation to bladder cancer risk (Karagas et al. 2005; Moore et al. 2004; Ouerhani et al. 2007; Vineis et al. 2007). The methyl-group metabolism pathway regulates methylation of DNA CpG islands, thus controlling expression and silencing of tumor suppressor and oncogenes (Paz et al. 2002). Our study found increased risk for variants in a related methyl-metabolism dehydrogenase enzyme that provides the 5,10 methylenetetrahydrofo-late pool (MTHFD2). This result is consistent with studies showing that MTHFD2 SNPs modify gastric and lung cancer risk (Liu et al. 2008; Wang et al. 2007).

Telomerase activity is critical for chromosomal stability and has been associated with bladder cancer risk (Alvarez and Lokeshwar 2007; Xing et al. 2007). TERT1 protein levels are a rate limiting step in telomerase activity and were associated with better survival in a previous study of 132 bladder cancer cases (P = 0.007) (Ito et al. 1998; Mavrommatis et al. 2005). Our study indicated better survival for TERT variants. Variants of telomerase-associated protein 1 (TEP1), a telomerase that adds new telomeres to the ends of chromosomes, protecting them from degradation, were associated with increased bladder cancer risk in our study, and in other cancers in a prior study (Savage et al. 2007).

Improper distribution of chromosomes during mitosis can result from modified expression of mitotic kinases, including Aurora kinase A (AURKA), and leads to chromosomal instability and aneuploidy in bladder cells (Fraizer et al. 2004). Subjects with the AURKA non-synonymous polymorphism in exon 11 had significantly greater susceptibility to bladder cancer in our study. Allelic variation in AURKA has been associated with AURKA expression levels and increased risk of multiple cancers in a meta-analysis of 9,549 cases (Ewart-Toland et al. 2005; Matarasso et al. 2007). AURKA amplification is very prevalent in bladder tumors (64%) and tumor AURKA gene and protein expression levels are associated with tumor recurrence and shorter survival time (Comperat et al. 2007; Denzinger et al. 2007; Schultz et al. 2007). The previously examined AURKA polymorphism T91A was not associated with bladder cancer prognosis in a study of 135 patients (Schultz et al. 2007). Our results suggest a lower survival rate for variants at another locus in AURKA intron 7 and a possible relationship with polo-like kinase 1 (PLK1). Aurora kinase A activates PLK1, which regulates spindle formation as DNA damage arrested cells move into mitosis (Macurek et al. 2008). High PLK1 expression levels are associated with bladder tumor chromosomal instability and progression (Yamamoto et al. 2006).

Insulin-like growth factor 1 (IGF1) variants had increased risk of bladder cancer in our study. SNPs in the promoter and intronic regions of IGF1 predict circulating IGF1 levels and increase colorectal cancer risk (Palles et al. 2008; Wong et al. 2008). Insulin-like growth factor 2 antisense (IGF2AS/PEG1) is aberrantly methylated in colorectal cancers and is overexpressed in Wilms' tumors (Nishihara et al. 2000; Okutsu et al. 2000). IGF2AS variants interacted with the folate transporter, SLC19A1, to increase bladder cancer risk in our study. Similarly, SLC19A1 was related to increased bladder cancer risk in the Spanish population (Moore et al. 2007).

Bladder cancer survival was negatively associated with variation in CD80, one of the surface antigens that increase in expression following treatment with the immunotherapy drug bacillus Calmette-Guerin (BCG) that is commonly used to treat bladder cancer patients (Ikeda et al. 2002). Interleukin-8 receptor (IL8RB/CXCR2) variants were at decreased risk of developing bladder cancer, but also experienced poorer survival than wildtype individuals in our study. IL8RB is associated with inflammatory response mediating angiogenesis, and neutrophil migration (Matheson et al. 2006). IL8RB polymorphisms have previously been associated with breast cancer progression (Kamali-Sarvestani et al. 2007).

Our data also suggest an increased risk of bladder cancer among variants for the mutS homolog 6 (MSH6), which recognizes mismatched nucleotides and induces apoptosis if repair fails. Reduced expression, mutation of MSH6 and concomitant loss of mismatch repair activity have been observed in bladder tumors (Mongiat-Artus et al. 2006; Thykjaer et al. 2001). As in our study, MSH6 variants were at higher risk for malignant bladder, an observation made for other cancers as well (Campbell et al. 2008; Sanyal et al. 2007). We observed shorter survival among ERCC4 variants. Although genetic variation in the DNA repair gene ERCC4 was not independently associated with bladder cancer susceptibility, few have investigated its role in survival (Garcia-Closas et al. 2006; Matullo et al. 2005).

Variants in apoptosis genes also modulated bladder cancer survival. The BCL-2 family member, BCL2L1/BCLX regulates the mitochondrial membrane potential controlling release of apoptotic inducers and expression levels have been related to cancer progression (Gazzaniga et al. 1998; Kirsh et al. 1998). Variants of the early apoptosis cascade member cysteine peptidase, Caspase 9 (CASP9) experienced shorter survival in our study. Both BCL2L1 and caspase 9 may be involved in response to some types of chemotherapy treatment in the bladder (Yuan et al. 2002).

In addition to main-gene eVects, we also detected a possible interaction between peptidylprolyl cis/trans isomerase (PIN1) and Alpha-methylacyl-CoA racemase (AMACR) variants. PIN1 is a conditional tumor suppressor, which transduces phosphorylation of P53 and promotes cell death by mediating dissociation from the apoptosis inhibitor iASPP (Mantovani et al. 2007). PIN1 plays a critical role in the DNA damage checkpoint, in part by protecting p53 from the inhibitor MDM2. Thus, Pin1−/− cells do not undergo mitosis in response to DNA damage (Takahashi et al. 2008). AMACR is involved in cellular energy metabolism by the oxidation of branched-chain fatty acids and in the metabolism of lipophilic drugs (Lloyd et al. 2008). AMACR expression is associated with higher bladder tumor histopathologic grade, implying that endogenous fatty acids may be used for energy as tumors grow (Gunia et al. 2008). AMACR variations have previously been associated with prostate cancer risk (Levin et al. 2007). The combination of inhibition of the DNA damage checkpoint by PIN1 with AMACR modulated energy availability for tumor growth and could result in the synergistic effect on tumor risk that we observed.

Another interaction involved scavenger receptor class B (SCARB1), which regulates cellular cholesterol uptake and sequesters chloesteryl esters from high-density lipoproteins (Miquel et al. 2003) with the aldo-keto reductase, AKR1C3, which is strongly expressed in the bladder. AKR1C3 regulates lipophilic ligand access to hormone receptors and activates PAHs, which are known bladder carcinogens (Azzarello et al. 2008). Our finding of modified bladder cancer risk in New Hampshire corroborates the recently reported association between AKR1C3 polymorphisms and bladder cancer risk in Spain (Figueroa et al. 2008).

The applicability of these results to other populations of bladder cancer patients needs to be explored before these markers can be used to guide clinical practice. These data will be used to inspire investigation of the mechanistic impact of these SNPs on tumor biology. Intronic SNPs may be acting through alternative splice sites, miRNAs, conformational changes or linkage disequilibrium with another functional SNP. Based on our results, a thorough investigation into SNPs in the metabolism, telomere maintenance, mitosis, inflammation, and apoptosis pathways is warranted. Nevertheless, these findings highlight a number of genetic variations in known carcinogenic pathway genes that modify bladder cancer risk and prognosis. These results may provide new insight into methods of predicting bladder cancer susceptibility using SNP panels, identify targets for bladder cancer prevention, and guide individualized choice of treatment methodologies based on molecular pathway alterations.

Acknowledgments

This publication was funded in part by grant numbers CA102327, CA121382, CA099500, CA82354, CA57494, CA078609, ES00002, P42 ES05947, RR018787, LM009012 and ES007373 from the National Cancer Institute, NIH and from the National Institute of Environmental Health Sciences, NIH.

References

  1. Ahsan H, Thomas DC. Lung cancer etiology: independent and joint eVects of genetics, tobacco, and arsenic. JAMA. 2004;292:3026–3029. doi: 10.1001/jama.292.24.3026. [DOI] [PubMed] [Google Scholar]
  2. Alvarez A, Lokeshwar VB. Bladder cancer biomarkers: current developments and future implementation. Curr Opin Urol. 2007;17:341–346. doi: 10.1097/MOU.0b013e3282c8c72b. [DOI] [PubMed] [Google Scholar]
  3. Andrew AS, Nelson HH, Kelsey KT, et al. Concordance of multiple analytical approaches demonstrates a complex relationship between DNA repair gene SNPs, smoking and bladder cancer susceptibility. Carcinogenesis. 2006;27:1030–1037. doi: 10.1093/carcin/bgi284. [DOI] [PubMed] [Google Scholar]
  4. Azzarello J, Fung KM, Lin HK. Tissue distribution of human AKR1C3 and rat homolog in adult genitourinary system. J Histochem Cytochem. 2008;56(9):853–861. doi: 10.1369/jhc.2008.951384. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Breiman L. Statistical modeling the two cultures. Stat Sci. 2001;16:199–231. [Google Scholar]
  6. Campbell PT, Curtin K, Ulrich C, et al. Mismatch repair polymorphisms and risk of colon cancer, tumor microsatellite instability, and interactions with lifestyle factors. Gut. 2008 doi: 10.1136/gut.2007.144220. doi:10.1136/gut.2007.144220. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Carroll P, Raghavan D, Stein J, et al. H0005 PM-the treatment of bladder cancer. Stage by Stage. AUA Annual Meeting 2001. 2001 [Google Scholar]
  8. Comperat E, Camparo P, Haus R, et al. Aurora-A/STK-15 is a predictive factor for recurrent behaviour in non-invasive bladder carcinoma: a study of 128 cases of non-invasive neoplasms. Virchows Arch. 2007;450:419–424. doi: 10.1007/s00428-007-0383-x. [DOI] [PubMed] [Google Scholar]
  9. Cussenot O, Azzouzi AR, Bantsimba-Malanda G, et al. EVect of genetic variability within 8q24 on aggressiveness patterns at diagnosis and familial status of prostate cancer. Clin Cancer Res. 2008;14:5635–5639. doi: 10.1158/1078-0432.CCR-07-4999. [DOI] [PubMed] [Google Scholar]
  10. Denzinger S, Stoehr R, Schwarz S, et al. Low level STK15 amplification in histologically benign urothelium of patients with bladder cancer adversely predicts patient outcome following cystectomy. Int J Oncol. 2007;31:793–802. [PubMed] [Google Scholar]
  11. Duda RO, Hart PE, Stork DG. Pattern classification. 2nd edn Wiley; New York: 2001. p. 632. [Google Scholar]
  12. Ewart-Toland A, Dai Q, Gao YT, et al. Aurora-A/STK15 T+91A is a general low penetrance cancer susceptibility gene: a meta-analysis of multiple cancer types. Carcinogenesis. 2005;26:1368–1373. doi: 10.1093/carcin/bgi085. [DOI] [PubMed] [Google Scholar]
  13. Figueroa JD, Malats N, Garcia-Closas M, et al. Bladder cancer risk and genetic variation in AKR1C3 and other metabolizing genes. Carcinogenesis. 2008;29:1955–1962. doi: 10.1093/carcin/bgn163. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Fraizer GC, Diaz MF, Lee IL, et al. Aurora-A/STK15/BTAK enhances chromosomal instability in bladder cancer cells. Int J Oncol. 2004;25:1631–1639. [PubMed] [Google Scholar]
  15. Garcia-Closas M, Malats N, Silverman D, et al. NAT2 slow acetylation, GSTM1 null genotype, and risk of bladder cancer: results from the Spanish Bladder Cancer Study and meta-analyses. Lancet. 2005;366:649–659. doi: 10.1016/S0140-6736(05)67137-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Garcia-Closas M, Malats N, Real FX, et al. Genetic variation in the nucleotide excision repair pathway and bladder cancer risk. Cancer Epidemiol Biomarkers Prev. 2006;15:536–542. doi: 10.1158/1055-9965.EPI-05-0749. [DOI] [PubMed] [Google Scholar]
  17. Gazzaniga P, Gradilone A, Silvestri I, et al. Variable levels of bcl-2, bcl-x and bax mRNA in bladder cancer progression. Oncol Rep. 1998;5:901–904. doi: 10.3892/or.5.4.901. [DOI] [PubMed] [Google Scholar]
  18. Gunia S, May M, Scholmann K, et al. Expression of alpha-methylacyl-CoA racemase correlates with histopathologic grading in noninvasive bladder cancer. Virchows Arch. 2008;453:165–170. doi: 10.1007/s00428-008-0638-1. [DOI] [PubMed] [Google Scholar]
  19. Holden M, Deng S, Wojnowski L, et al. GSEA-SNP: applying gene set enrichment analysis to SNP data from genome-wide association studies. Bioinformatics. 2008;24:2784–2785. doi: 10.1093/bioinformatics/btn516. [DOI] [PubMed] [Google Scholar]
  20. Ikeda N, Toida I, Iwasaki A, et al. Surface antigen expression on bladder tumor cells induced by bacillus Calmette-Guerin (BCG): a role of BCG internalization into tumor cells. Int J Urol. 2002;9:29–35. doi: 10.1046/j.1442-2042.2002.00415.x. [DOI] [PubMed] [Google Scholar]
  21. Ito H, Kyo S, Kanaya T, et al. Expression of human telomerase subunits and correlation with telomerase activity in urothelial cancer. Clin Cancer Res. 1998;4:1603–1608. [PubMed] [Google Scholar]
  22. Jemal A, Siegel R, Ward E, et al. Cancer statistics. CA Cancer J Clin. 2007;57:43–66. doi: 10.3322/canjclin.57.1.43. [DOI] [PubMed] [Google Scholar]
  23. Kamali-Sarvestani E, Aliparasti MR, Atefi S. Association of interleukin-8 (IL-8 or CXCL8) -251T/A and CXCR2 +1208C/T gene polymorphisms with breast cancer. Neoplasma. 2007;54:484–489. [PubMed] [Google Scholar]
  24. Kantor AF, Hartge P, Hoover RN, et al. Familial and environmental interactions in bladder cancer risk. Int J Cancer. 1985;35:703–706. doi: 10.1002/ijc.2910350602. [DOI] [PubMed] [Google Scholar]
  25. Karagas MR, Tosteson TD, Blum J, et al. Design of an epidemiologic study of drinking water arsenic exposure and skin and bladder cancer risk in a U.S. population. Environ Health Perspect. 1998;106(Suppl 4):1047–1050. doi: 10.1289/ehp.98106s41047. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Karagas MR, Park S, Nelson HH, et al. Methylenetetrahydrofolate reductase (MTHFR) variants and bladder cancer: a population-based case-control study. Int J Hyg Environ Health. 2005;208:321–327. doi: 10.1016/j.ijheh.2005.04.005. [DOI] [PubMed] [Google Scholar]
  27. Kirkali Z, Chan T, Manoharan M, et al. Bladder cancer: epidemiology, staging and grading, and diagnosis. Urology. 2005;66:4–34. doi: 10.1016/j.urology.2005.07.062. [DOI] [PubMed] [Google Scholar]
  28. Kirsh EJ, Baunoch DA, Stadler WM. Expression of bcl-2 and bcl-X in bladder cancer. J Urol. 1998;159:1348–1353. [PubMed] [Google Scholar]
  29. Kullback S. Information theory and statistics. Wiley; New York: 1959. [Google Scholar]
  30. Kullback S. The Kullback-Leibler distance. Am Stat. 1987;41:340–341. [Google Scholar]
  31. Levin AM, Zuhlke KA, Ray AM, et al. Sequence variation in alpha-methylacyl-CoA racemase and risk of early-onset and familial prostate cancer. Prostate. 2007;67:1507–1513. doi: 10.1002/pros.20642. [DOI] [PubMed] [Google Scholar]
  32. Liu H, Jin G, Wang H, et al. Association of polymorphisms in one-carbon metabolizing genes and lung cancer risk: a case-control study in Chinese population. Lung Cancer. 2008;61:21–29. doi: 10.1016/j.lungcan.2007.12.001. [DOI] [PubMed] [Google Scholar]
  33. Lloyd MD, Darley DJ, Wierzbicki AS, et al. Alpha-methylacyl-CoA racemase—an ‘obscure’ metabolic enzyme takes centre stage. FEBS J. 2008;275:1089–1102. doi: 10.1111/j.1742-4658.2008.06290.x. [DOI] [PubMed] [Google Scholar]
  34. Macurek L, Lindqvist A, Lim D, et al. Polo-like kinase-1 is activated by aurora A to promote checkpoint recovery. Nature. 2008;455:119–123. doi: 10.1038/nature07185. [DOI] [PubMed] [Google Scholar]
  35. Mantovani F, Tocco F, Girardini J, et al. The prolyl isomerase Pin1 orchestrates p53 acetylation and dissociation from the apoptosis inhibitor iASPP. Nat Struct Mol Biol. 2007;14:912–920. doi: 10.1038/nsmb1306. [DOI] [PubMed] [Google Scholar]
  36. Matarasso N, Bar-Shira A, Rozovski U, et al. Functional analysis of the Aurora Kinase A Ile31 allelic variant in human prostate. Neoplasia. 2007;9:707–715. doi: 10.1593/neo.07322. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Matheson MC, Ellis JA, Raven J, et al. Association of IL8, CXCR2 and TNF-alpha polymorphisms and airway disease. J Hum Genet. 2006;51:196–203. doi: 10.1007/s10038-005-0344-7. [DOI] [PubMed] [Google Scholar]
  38. Matullo G, Guarrera S, Sacerdote C, et al. Polymorphisms/Haplotypes in DNA repair genes and smoking: a bladder cancer case-control study. Cancer Epidemiol Biomarkers Prev. 2005;14:2569–2578. doi: 10.1158/1055-9965.EPI-05-0189. [DOI] [PubMed] [Google Scholar]
  39. Mavrommatis J, Mylona E, Gakiopoulou H, et al. Nuclear hTERT immunohistochemical expression is associated with survival of patients with urothelial bladder cancer. Anticancer Res. 2005;25:3109–3116. [PubMed] [Google Scholar]
  40. Miquel JF, Moreno M, Amigo L, et al. Expression and regulation of scavenger receptor class B type I (SR-BI) in gall bladder epithelium. Gut. 2003;52:1017–1024. doi: 10.1136/gut.52.7.1017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Mongiat-Artus P, Miquel C, Van der Aa M, et al. Microsatellite instability and mutation analysis of candidate genes in urothelial cell carcinomas of upper urinary tract. Oncogene. 2006;25:2113–2118. doi: 10.1038/sj.onc.1209229. [DOI] [PubMed] [Google Scholar]
  42. Moore JH, White BC. Tuning ReliefF for Genome-Wide Genetic Analysis. Lect Notes Comput Sci. 2007;4447:166–175. [Google Scholar]
  43. Moore LE, Wiencke JK, Bates MN, et al. Investigation of genetic polymorphisms and smoking in a bladder cancer case-control study in Argentina. Cancer Lett. 2004;211:199–207. doi: 10.1016/j.canlet.2004.04.011. [DOI] [PubMed] [Google Scholar]
  44. Moore LE, Malats N, Rothman N, et al. Polymorphisms in one-carbon metabolism and trans-sulfuration pathway genes and susceptibility to bladder cancer. Int J Cancer. 2007;120:2452–2458. doi: 10.1002/ijc.22565. [DOI] [PubMed] [Google Scholar]
  45. Murta-Nascimento C, Schmitz-Drager BJ, Zeegers MP, et al. Epidemiology of urinary bladder cancer: from tumor development to patient's death. World J Urol. 2007a;25:285–295. doi: 10.1007/s00345-007-0168-5. [DOI] [PubMed] [Google Scholar]
  46. Murta-Nascimento C, Silverman DT, Kogevinas M, et al. Risk of bladder cancer associated with family history of cancer: do low-penetrance polymorphisms account for the increase in risk? Cancer Epidemiol Biomarkers Prev. 2007b;16:1595–1600. doi: 10.1158/1055-9965.EPI-06-0743. [DOI] [PubMed] [Google Scholar]
  47. Nishihara S, Hayashida T, Mitsuya K, et al. Multipoint imprinting analysis in sporadic colorectal cancers with and without microsatellite instability. Int J Oncol. 2000;17:317–322. doi: 10.3892/ijo.17.2.317. [DOI] [PubMed] [Google Scholar]
  48. Okutsu T, Kuroiwa Y, Kagitani F, et al. Expression and imprinting status of human PEG8/IGF2AS, a paternally expressed antisense transcript from the IGF2 locus, in Wilms' tumors. J Biochem. 2000;127:475–483. doi: 10.1093/oxfordjournals.jbchem.a022630. [DOI] [PubMed] [Google Scholar]
  49. Ouerhani S, Oliveira E, Marrakchi R, et al. Methylenetetrahydrofolate reductase and methionine synthase polymorphisms and risk of bladder cancer in a Tunisian population. Cancer Genet Cytogenet. 2007;176:48–53. doi: 10.1016/j.cancergencyto.2007.03.007. [DOI] [PubMed] [Google Scholar]
  50. Palles C, Johnson N, Coupland B, et al. Identification of genetic variants that influence circulating IGF1 levels: a targeted search strategy. Hum Mol Genet. 2008;17:1457–1464. doi: 10.1093/hmg/ddn034. [DOI] [PubMed] [Google Scholar]
  51. Paz MF, Avila S, Fraga MF, et al. Germ-line variants in methyl-group metabolism genes and susceptibility to DNA methylation in normal tissues and human primary tumors. Cancer Res. 2002;62:4519–4524. [PubMed] [Google Scholar]
  52. Ritchie MD, Hahn LW, Roodi N, et al. Multifactor-dimensionality reduction reveals high-order interactions among estrogen-metabolism genes in sporadic breast cancer. Am J Hum Genet. 2001;69:138–147. doi: 10.1086/321276. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Ritchie MD, Hahn LW, Moore JH. Power of multifactor dimensionality reduction for detecting gene-gene interactions in the presence of genotyping error, missing data, phenocopy, and genetic heterogeneity. Genet Epidemiol. 2003;24:150–157. doi: 10.1002/gepi.10218. [DOI] [PubMed] [Google Scholar]
  54. Sanyal S, De Verdier PJ, Steineck G, et al. Polymorphisms in XPD, XPC and the risk of death in patients with urinary bladder neoplasms. Acta Oncol. 2007;46:31–41. doi: 10.1080/02841860600812693. [DOI] [PubMed] [Google Scholar]
  55. Savage SA, Chanock SJ, Lissowska J, et al. Genetic variation in five genes important in telomere biology and risk for breast cancer. Br J Cancer. 2007;97:832–836. doi: 10.1038/sj.bjc.6603934. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Schned AR, Andrew AS, Marsit CJ, et al. Survival following the diagnosis of noninvasive bladder cancer: WHO/International Society of Urological Pathology Versus WHO Classification Systems. J Urol. 2007;178:1196–1200. doi: 10.1016/j.juro.2007.05.126. [DOI] [PubMed] [Google Scholar]
  57. Schultz IJ, Kiemeney LA, Roelofs R, et al. The prognostic role of the STK15 T91A polymorphism and of STK15 mRNA expression in patients with urothelial cell carcinoma. Anticancer Res. 2007;27:1025–1030. [PubMed] [Google Scholar]
  58. Takahashi K, Uchida C, Shin RW, et al. Prolyl isomerase, Pin1: new findings of post-translational modifications and physiological substrates in cancer, asthma and Alzheimer's disease. Cell Mol Life Sci. 2008;65:359–375. doi: 10.1007/s00018-007-7270-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Thykjaer T, Christensen M, Clark AB, et al. Functional analysis of the mismatch repair system in bladder cancer. Br J Cancer. 2001;85:568–575. doi: 10.1054/bjoc.2001.1949. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. van Dijk B, van Houwelingen KP, Witjes JA, et al. Alcohol dehydrogenase type 3 (ADH3) and the risk of bladder cancer. Eur Urol. 2001;40:509–514. doi: 10.1159/000049827. [DOI] [PubMed] [Google Scholar]
  61. Vineis P, Veglia F, Garte S, et al. Genetic susceptibility according to three metabolic pathways in cancers of the lung and bladder and in myeloid leukemias in nonsmokers. Ann Oncol. 2007;18:1230–1242. doi: 10.1093/annonc/mdm109. [DOI] [PubMed] [Google Scholar]
  62. Wang L, Ke Q, Chen W, et al. Polymorphisms of MTHFD, plasma homocysteine levels, and risk of gastric cancer in a high-risk Chinese population. Clin Cancer Res. 2007;13:2526–2532. doi: 10.1158/1078-0432.CCR-06-2293. [DOI] [PubMed] [Google Scholar]
  63. Wong HL, Koh WP, Probst-Hensch NM, et al. Insulin-like growth factor-1 promoter polymorphisms and colorectal cancer: a functional genomics approach. Gut. 2008;57:1090–1096. doi: 10.1136/gut.2007.140855. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Wu X, Gu J, Grossman HB, et al. Bladder cancer predisposition: a multigenic approach to DNA-repair and cell-cycle-control genes. Am J Hum Genet. 2006;78:464–479. doi: 10.1086/500848. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Xing J, Zhu Y, Zhao H, et al. DiVerential induction in telomerase activity among bladder cancer patients and controls on gamma-radiation. Cancer Epidemiol Biomarkers Prev. 2007;16:606–609. doi: 10.1158/1055-9965.EPI-06-0615. [DOI] [PubMed] [Google Scholar]
  66. Yamamoto Y, Matsuyama H, Kawauchi S, et al. Overexpression of Polo-Like Kinase 1 (PLK1) and chromosomal instability in bladder cancer. Oncology. 2006;70:231–237. doi: 10.1159/000094416. [DOI] [PubMed] [Google Scholar]
  67. Yuan SY, Hsu SL, Tsai KJ, et al. Involvement of mitochondrial pathway in Taxol-induced apoptosis of human T24 bladder cancer cells. Urol Res. 2002;30:282–288. doi: 10.1007/s00240-002-0263-4. [DOI] [PubMed] [Google Scholar]
  68. Zenser TV, Lakshmi VM, Hsu FF, et al. Metabolism of N-acetylbenzidine and initiation of bladder cancer. Mutat Res. 2002;506–507:29–40. doi: 10.1016/s0027-5107(02)00149-5. [DOI] [PubMed] [Google Scholar]

RESOURCES