Genetic polymorphisms and breast cancer risk: evidence from meta-analyses, pooled analyses, and...

16
REVIEW Genetic polymorphisms and breast cancer risk: evidence from meta-analyses, pooled analyses, and genome-wide association studies Sihua Peng Bingjian Lu ¨ Wenjing Ruan Yimin Zhu Hongqiang Sheng Maode Lai Received: 14 February 2011 / Accepted: 15 March 2011 / Published online: 29 March 2011 Ó Springer Science+Business Media, LLC. 2011 Abstract To address the association between variants and breast cancer, an increasing number of articles on genetic association studies, genome-wide association studies (GWASs), and related meta- and pooled analyses have been published. Such studies have prompted an updated assess- ment of the associations between gene variants and breast cancer risk. We searched PubMed, Medline, and Web of Science and retrieved a total of 87 meta- and pooled anal- yses, which addressed the associations between 145 gene variants and breast cancer. Analyses met the following criteria: (1) breast cancer was the outcome, (2) the articles were all published in English, and (3) in the recent pub- lished meta- and pooled analyses, the analyses with more subjects were selected. Among the 145 variants, 46 were significantly associated with breast cancer and the other 99 (in 62 genes) were not significantly associated with breast cancer. The summary ORs for the 46 significant associa- tions (P \ 0.05) were further assessed by the method of false-positive report probability (FPRP). Our results demonstrated that 10 associations were noteworthy: CASP8 (D302H), CHEK2 (*1100delC), CTLA4 (?49G [ A), FGFR2 (rs2981582, rs1219648, and rs2420946), HRAS (rare alleles), IL1B (rs1143627), LSP1 (rs3817198), and MAP3K1 (rs889312). In addition, eight GWASs were identified, in which 25 loci were obtained (14 in nine genes, six near a gene or genes, and five intergenic loci). Of the 25 SNPs, 20 were noteworthy: C6orf97 (rs2046210 and rs3757318), FGFR2 (rs2981579, rs1219648, and rs2981582), LSP1 (rs909116), RNF146 (rs2180341), SLC4A7 (rs4973768), MRPS30 (rs7716600), TOX3 (rs3803662 and rs4784227), ZNF365 (rs10995190), rs889312, rs614367, rs13281615, rs13387042, rs11249433, rs1011970, rs614367, and rs1562430. In summary, in this review of genetic association studies, 31.7% of the gene- variant breast cancer associations were significant, and 21.7% of these significant associations were noteworthy. However, in GWASs, 80% of the significant associations were noteworthy. Keywords Breast cancer Polymorphism Meta-analysis Genome-wide association study Pooled analysis Introduction Breast cancer remains a major health problem around the world, and its incidence continues to increase [1]. It accounts for 26% of all new cancer cases among women in the United States [2]. During the last few decades, Electronic supplementary material The online version of this article (doi:10.1007/s10549-011-1459-5) contains supplementary material, which is available to authorized users. S. Peng H. Sheng M. Lai (&) Department of Pathology, Zhejiang University School of Medicine, 388 Yuhangtang Road, Hangzhou 310058, Zhejiang, People’s Republic of China e-mail: [email protected] B. Lu ¨ Department of Surgical Pathology, Affiliated Women’s Hospital, Zhejiang University School of Medicine, Hangzhou 310003, People’s Republic of China W. Ruan Department of Respiratory Diseases, Affiliated Sir Run Run Shaw Hospital, Zhejiang University School of Medicine, Hangzhou 310016, People’s Republic of China Y. Zhu Department of Public Health, School of Medicine, Zhejiang University, Hangzhou 310058, People’s Republic of China 123 Breast Cancer Res Treat (2011) 127:309–324 DOI 10.1007/s10549-011-1459-5

Transcript of Genetic polymorphisms and breast cancer risk: evidence from meta-analyses, pooled analyses, and...

REVIEW

Genetic polymorphisms and breast cancer risk: evidencefrom meta-analyses, pooled analyses, and genome-wideassociation studies

Sihua Peng • Bingjian Lu • Wenjing Ruan •

Yimin Zhu • Hongqiang Sheng • Maode Lai

Received: 14 February 2011 / Accepted: 15 March 2011 / Published online: 29 March 2011

� Springer Science+Business Media, LLC. 2011

Abstract To address the association between variants and

breast cancer, an increasing number of articles on genetic

association studies, genome-wide association studies

(GWASs), and related meta- and pooled analyses have been

published. Such studies have prompted an updated assess-

ment of the associations between gene variants and breast

cancer risk. We searched PubMed, Medline, and Web of

Science and retrieved a total of 87 meta- and pooled anal-

yses, which addressed the associations between 145 gene

variants and breast cancer. Analyses met the following

criteria: (1) breast cancer was the outcome, (2) the articles

were all published in English, and (3) in the recent pub-

lished meta- and pooled analyses, the analyses with more

subjects were selected. Among the 145 variants, 46 were

significantly associated with breast cancer and the other 99

(in 62 genes) were not significantly associated with breast

cancer. The summary ORs for the 46 significant associa-

tions (P \ 0.05) were further assessed by the method of

false-positive report probability (FPRP). Our results

demonstrated that 10 associations were noteworthy: CASP8

(D302H), CHEK2 (*1100delC), CTLA4 (?49G[A),

FGFR2 (rs2981582, rs1219648, and rs2420946), HRAS

(rare alleles), IL1B (rs1143627), LSP1 (rs3817198), and

MAP3K1 (rs889312). In addition, eight GWASs were

identified, in which 25 loci were obtained (14 in nine genes,

six near a gene or genes, and five intergenic loci). Of

the 25 SNPs, 20 were noteworthy: C6orf97 (rs2046210

and rs3757318), FGFR2 (rs2981579, rs1219648, and

rs2981582), LSP1 (rs909116), RNF146 (rs2180341),

SLC4A7 (rs4973768), MRPS30 (rs7716600), TOX3

(rs3803662 and rs4784227), ZNF365 (rs10995190),

rs889312, rs614367, rs13281615, rs13387042, rs11249433,

rs1011970, rs614367, and rs1562430. In summary, in this

review of genetic association studies, 31.7% of the gene-

variant breast cancer associations were significant, and

21.7% of these significant associations were noteworthy.

However, in GWASs, 80% of the significant associations

were noteworthy.

Keywords Breast cancer � Polymorphism �Meta-analysis � Genome-wide association study �Pooled analysis

Introduction

Breast cancer remains a major health problem around the

world, and its incidence continues to increase [1]. It

accounts for 26% of all new cancer cases among women in

the United States [2]. During the last few decades,

Electronic supplementary material The online version of thisarticle (doi:10.1007/s10549-011-1459-5) contains supplementarymaterial, which is available to authorized users.

S. Peng � H. Sheng � M. Lai (&)

Department of Pathology, Zhejiang University School

of Medicine, 388 Yuhangtang Road, Hangzhou 310058,

Zhejiang, People’s Republic of China

e-mail: [email protected]

B. Lu

Department of Surgical Pathology, Affiliated Women’s Hospital,

Zhejiang University School of Medicine, Hangzhou 310003,

People’s Republic of China

W. Ruan

Department of Respiratory Diseases, Affiliated Sir Run Run

Shaw Hospital, Zhejiang University School of Medicine,

Hangzhou 310016, People’s Republic of China

Y. Zhu

Department of Public Health, School of Medicine, Zhejiang

University, Hangzhou 310058, People’s Republic of China

123

Breast Cancer Res Treat (2011) 127:309–324

DOI 10.1007/s10549-011-1459-5

extensive efforts have identified sources of genetic sus-

ceptibility to breast cancer. With the help of continuing

technological advances, many observational studies eval-

uating the association between variants in different genes

and breast cancer risk have been published [3]. These

studies prompted us to estimate their general contribution

to our current understanding of the genetic susceptibility to

breast cancer [4, 5].

One of the drawbacks of genetic epidemiology is that the

results are not easily replicated. Many studies attempting to

explore previously published, statistically significant find-

ings for the variants have failed to repeat those findings,

suggesting that false-positive results may exist in certain

studies [6]. Recently, Burton et al. [7] emphasized that if

sample sizes are large enough, it is possible to identify and

replicate genetic associations with common complex dis-

eases. Therefore, the size issue is an important methodo-

logical concern in genetic association studies [8].

Meta-analysis was defined by Glass [9] as ‘‘the statis-

tical analysis of a large collection of analysis results from

individual studies for the purpose of integrating the find-

ings.’’ Blettner [10] gave a definition of pooled analysis:

‘‘prospectively planned, pooled analysis of several studies,

where pooling is already a part of the protocol. Data col-

lection procedures, definition of variables, questions and

hypotheses are standardized for the individual studies.’’

Clearly, meta-analyses and pooled analyses are very good

tools for combining results from different studies. To this

end, we collected evidence from articles published on

meta-analyses and pooled analyses to assess the associa-

tions between breast cancer risk and various polymor-

phisms [11].

As early as 1999, Dunning et al.[12] published a sys-

tematic review of the associations of genetic variants and

breast cancer risk, in which 46 case–control studies

involving 18 genes were identified. They reported 12

variants to be significantly associated with breast cancer

risk. Since then, more articles on the associations between

variants and breast cancer risk have been published.

In 2008, Dong et al. [13] published another insightful

systematic review on the associations between genetic

variants and cancer risk, in which 20 variants were reported

to be significantly related to breast cancer risk. However,

only 3 years later, more than 100 new meta-analyses and

pooled analyses have been published, including more than

90 meta-analyses published in 2010 and 2011. Thus, the

results have been thoroughly updated. To this end, an

updated and systematic review on this issue is urgently

needed to evaluate the associations between genetic vari-

ants and breast cancer risk.

In contrast, a genome-wide association study (GWAS)

examines all or most of the genes in the genome of dif-

ferent individuals of a particular species to identify the

extent to which the genes vary from individual to indi-

vidual [14]. In humans, GWASs can identify the associa-

tions of particular genes with diseases, such as type 2

diabetes and breast cancer [15, 16]. In a GWAS, at least

100,000 single nucleotide polymorphisms (SNPs) of hun-

dreds or thousands of individuals are tested. To date, more

than 812 human GWASs have examined approximately

150 diseases and traits and have found more than 3,977

SNP associations [17, 18]. Therefore, GWASs evaluating

the associations between the various SNPs and breast

cancer risk were also reviewed in this article.

Methods

We searched PubMed, Medline, and Web of Science up to

Jan. 2011 for meta-analyses and pooled analyses involving

the associations between various polymorphisms and breast

cancer risk. We performed a literature search of the above

databases using three search themes, which were combined

using the Boolean operator ‘‘and.’’ The first theme was

(‘‘breast neoplasms’’ OR ‘‘breast cancer’’); the second was

(‘‘gene’’ OR ‘‘polymorphism’’ OR ‘‘single nucleotide

polymorphisms’’ OR ‘‘variant’’); and the third was (‘‘meta-

analysis’’ OR ‘‘pooled analysis’’).

Using this search strategy, 282, 302, and 214 articles (a

total of 798) were retrieved from PubMed, Medline, and

the Web of Science, respectively. Finally, after reading the

abstracts and/or full text, we selected 87 articles to include

in this review (Fig. 1). To be included, studies had to meet

the following criteria: (1) the outcome investigated in the

meta-analyses and pooled analyses was breast cancer; (2)

the articles were published in English; and (3) to avoid

overlap of results from more than one meta-analysis or

pooled analysis addressing the same variant, the analyses

with more subjects were selected. Data extracted from each

meta-analysis or pooled analysis included gene name,

genetic variant, OR (odds ratio) and 95% CI (confidence

interval), the number of studies, the number of subjects

(cases/controls), the test for between-study heterogeneity

(e.g., Q test [19]), and the test for publication bias (Egger’s

test [20]).

In this review, associations were considered statistically

significant if the reported P value was\0.05 or if the 95%

CI exceeded 1.0. For each statistically significant associa-

tion reported, we estimated the false-positive report prob-

ability (FPRP) using the method described by Wacholder

et al. [21] The magnitude of the FPRP is determined by

three parameters: (1) prior probability p of a true associa-

tion of the tested genetic variant with a disease, (2) a level

or observed P value, and (3) statistical power to detect the

OR of the alternative hypothesis at the given a level or

P value. They suggested estimating statistical power based

310 Breast Cancer Res Treat (2011) 127:309–324

123

on the ability to detect an OR of 1.5, with an a level equal

to the observed P value. Because much smaller ORs were

reported recently, we think that this estimate is too con-

servative. Therefore, ORs of both 1.5 and 1.2 were chosen

to present results. To assess whether the associations were

noteworthy, an FPRP cutoff value of 0.2 was used [21].

Thereby FPRP values less than 0.2 indicate an association

and are referred to as noteworthy in this review. Statistical

power and FPRP values were calculated by the Excel

spreadsheet which can be downloaded from the web site of

Wacholder et al. [21].

GWAS articles were obtained by searching PubMed and

checking the web site at www.genome.gov/gwastudies

[18]. The inclusion/exclusion criteria were as follows: (1)

only significant associations between various SNPs and

breast cancer risk were included, with a cutoff value of

P \ 1910-5 and (2) if the OR and CI were not reported,

this GWAS was excluded. FPRP methodology was also

used to evaluate the significant associations.

Results

We identified 87 published articles with meta-analyses and

pooled analyses, encompassing 80 different genes. These 87

meta-analyses and pooled analyses addressed associations

between 145 gene variants and breast cancer, including 40

meta-analyses and pooled analyses that reported significant

associations between 46 variants (in 35 genes) and breast

cancer (Table 1) and 61 meta-analyses and pooled analyses

that reported non-significant associations between 99 vari-

ants (in 62 genes) and breast cancer (Tables S1 and S2). Of

these 87 articles, 76 (87.4%) were published after 2009, 67

(77.0%) after 2010, and 6 (6.9%) in 2011.

Significant gene-variant breast cancer associations

Among the 145 gene-variant breast cancer associations that

were evaluated, the summary ORs for 46 (31.7%) associ-

ations were statistically significant (Fig. 2; Table 1).

Inverse associations for the variants were found in 8

of these 46 associations: CASP8 (D302H) [22], ESR1

(rs2234693 and rs1801132) [23], ESR2 (rs4986938) [24],

NBS1 (8360G[C) [25], NOS3 (-786T[C) [26], TP53

(rs1042522) [27], and XRCC3 (rs1799796) [28], with a

mean OR of 0.86 (median, 0.89; range, 0.66-0.95). The

other 38 analyses reported ORs higher than 1.0, with a mean

OR of 1.39 (median, 1.15; range, 1.05–3.13) (Table 1).

To evaluate the robustness of these findings, we calcu-

lated FPRP values. Among the 46 significant associations,

36 gene-variant breast cancer associations had FPRP values

[0.2 at the pre-specified prior probability of 0.001; these

associations were not considered noteworthy. For example,

although the OR from the pooled analysis for CCND1

(G870A) indicated a significant association with breast

cancer (OR, 1.09; 95% CI, 1.01–1.18), FPRP values were

higher than 0.2; therefore, this result was not considered

noteworthy. At a prior probability level of 0.001 and sta-

tistical power to detect an OR of 1.5, ten associations were

noteworthy (FPRP \ 0.2): CASP8 (D302H), CHEK2

Records excluded (n=123)

Full-text articles excluded, with reasons: Not meta-analysis/pooled-analysis,

genetic polymorphism, or breast cancer (n1=42).

Not latest meta-analysis (n2=19).

Total n=61

Records screened by title (n=393)

Records screened by abstract (n=270)

Records excluded (n=122)

Additional records identified through other sources

(n=0)

Records identified through database searching (PubMed 282, Medline 302,

and Web of Science 214) (n=798)

Meta-analyses and pooled analyses (n=87)

Records excluded (n=405)

Records after overlaps removed (n=798)

Full-text articles assessed for eligibility (n=148)

Fig. 1 Selection of studies

Breast Cancer Res Treat (2011) 127:309–324 311

123

Table 1 Statistically significant gene-variant breast cancer overall associations and false-positive report probabilities (FPRP)

Gene (variant) Year Comparison OR (95% CI) Publication

bias/

heterogeneity

No. of

studies

Cases/control Power

OR,

1.5

Power

OR,

1.2

FPRP values

at prior

probability

of 0.001

Ref.

OR

1.5

OR

1.2

CASP8 (D302H) 2010 Dominant

model

0.87 (0.83–0.92) No/0.66 4 39,109 (18,791/20,318) 1.0 0.935 0.001 0.001 [22]

CCND1 (G870A) 2010 AA vs. GG 1.09 (1.01–1.18) 0.054/0.122 13 21,082 (9,911/11,171) 1.0 0.991 0.971 0.971 [145]

CHEK2

(*1100delC)

2008 Heterozygotes

vs.

noncarrier

2.40 (1.80–3.20) 0.08/NA 12 36,909 (18,329/18,580) 0.001 \0.000 0.004 0.678 [95]

CTLA4 (?49G[A) 2010 Dominant

model

1.24 (1.18–1.32) 0.998/0.11 32 24,452 (11,273/13,179) 1.0 0.152 0.000 0.000 [146]

CYP19 ((TTTA)10) 2010 (TTTA)10 1.52 (1.12–2.06) 0.58/0.06 14 16,104 (7,743/8,361) 0.466 0.064 0.937 0.991 [42]

ERCC2

(Lys751Gln)

2010 Dominant

model

1.13 (1.02–1.24) 0.036/0.0001 32 29,897 (14,545/15,352) 1.0 0.898 0.908 0.917 [47]

ESR1 (rs2234693) 2010 Dominant

model

0.95 (0.89–1.00) 0.97/0.39 11 26,920 (10,300/16,620) 1.0 1.0 0.980 0.980 [23]

(rs1801132) Dominant

model

0.92 (0.85–0.99) 0.86/0.66 7 12,505 (5,649/6,856) 1.0 0.996 0.963 0.963 [23]

ESR2 (rs4986938) 2010 Dominant

model

0.94 (0.90–0.99) 0.748/0.379 9 26,858 (10,837/16,021) 1.0 1.0 0.951 0.951 [24]

FGFR2 (rs2981582) 2010 GG vs. AA 1.48 (1.35–1.61) [0.05/

\ 0.000

11 91,890 (40,292/51,598) 0.0 0.0 0.000 0.000 [147]

(rs1219648) GG vs. AA 1.50 (1.36–1.65) [0.05/0.09 9 58,903 (13,930/44,973) 0.548 0.0 0.000 0.000 [147]

(rs2420946) CC vs. TT 1.52 (1.37–1.68) [0.05/0.20 6 15,329 (6,781/8,548) 0.413 0.0 0.000 0.000 [147]

GSTM1 (null/

present)

2010 Null vs.

present

1.10 (1.04–1.16) 0.445/0.003 59 46,281 (20,993/25,288) 1.0 0.999 0.303 0.303 [148]

GSTT1 (null/

present)

2010 Null vs.

present

1.11 (1.04–1.20) 0.048/0.059 41 36,584 (16,589/19,995) 1.0 0.975 0.897 0.899 [53]

For Chinese 2010 Null vs.

present

1.06 (0.87–1.29) 0.742/0.017 8 5,980 (2,450/3,530) – – – – [53]

HER2 (Ile655Val) 2010 Dominant

model

1.10 (1.01–1.20) 0.03/0.01 27 24,042 (11,504/12,538) 1.0 0.975 0.969 0.970 [149]

HRAS (rare alleles) 1997 Per-allele

comparison

2.7 (2.10–3.40) NA/NA 8 1,631 (694/937) 0.0 0.0 0.000 0.000 [150]

IGFBP3 (A-202C) 2010 CC vs. AA 1.06 (1.02–1.11) 0.63/0.60 27 78,811 (33,557/45,254) 1.0 1.0 0.93 0.93 [96]

IL1B (rs1143627) 2010 Recessive

model

1.40 (1.17–1.67) No/0.49 4 2,708 (1,543/1,165) 0.778 0.043 0.191 0.809 [58]

LSP1 (rs3817198) 2010 Allele contrast 1.06 (1.04–1.08) No/0.166 7 69,591 (33,920/35,671) 1.0 1.0 0.000 0.000 [151]

MAP3K1(rs889312) 2010 Dominant

model

1.10 (1.06–1.13) 0.696/0.43 7 59,977 (26,015/33,962) 1.0 1.0 0.000 0.000 [152]

MDM2(SNP309) 2010 GT vsTT 1.06 (1.00–1.12) No/0.77 16 25,979 (12,986/12,993) 1.0 1.0 0.974 0.974 [153]

For non-Chinese 2010 G vs. T 1.02 (0.97–1.08) No/No 13 23,652 (12,094/11,558) – – – – [153]

MMP-2 (-1306

C/T)

2010 CC vs.

CT ? TT

1.27 (1.10–1.47) No/\0.001 4 3,243 (1,590/1,653) 0.987 0.224 0.579 0.859 [61]

MTHFR (C667T) 2010 TT vs.

CC ? CT

1.12 (1.01–1.24) No/0.001 41 38,868 (16,480/22,388) 1.0 0.908 0.967 0.970 [154]

NBS1 (657del5) 2006 Carrier vs.

noncarrier

3.13 (1.40–7.00) No/No 2 2,406 (1,620/786) 0.037 0.010 0.993 0.993 [97]

(8360G[C) 2010 CC vs. GG 0.75 (0.74–0.98) 0.20/0.13 10 10,117 (4,452/5,665) 0.806 0.220 0.977 0.994 [25]

NOS3 (eNOS,894

G[T)

2010 TT vs. GG 1.29 (1.06–1.56) 0.110/0.223 11 9,507 (4,665/4,842) 0.94 0.228 0.902 0.974 [155]

(-786T[C) 2010 Dominant

model

0.66 (0.47–0.94) No/0.741 3 3,326 (1,856/1,470) 0.478 0.098 0.978 0.995 [26]

POR (Gly5Gly) 2007 GG vs. AA 1.58 (1.04–2.41) NA/0.85 4 1,915 (1,038/877) 0.405 0.101 0.988 0.997 [98]

312 Breast Cancer Res Treat (2011) 127:309–324

123

(*1100delC), CTLA4 (?49G[A), FGFR2 (rs2981582,

rs1219648, and rs2420946), HRAS (rare alleles), IL1B

(rs1143627), LSP1 (rs3817198), and MAP3K1 (rs889312).

This number was further reduced to eight when we cal-

culated the statistical power based on a lower OR of 1.2:

CASP8 (D302H), CTLA4 (?49G[A), FGFR2 (rs2981582,

rs1219648, and rs2420946), HRAS (rare alleles), LSP1

(rs3817198), and MAP3K1 (rs889312).

Non-significant gene-variant breast cancer associations

Besides the statistically significant associations, statisti-

cally non-significant associations were reported for 99

variants (with P values [ 0.05; Table S1) [22–24, 29–86].

Of these 99 associations, 49 (49.5%) were inverse associ-

ations for the variants, with a mean OR of 0.91 (median,

0.96; range, 0.39–0.99).

Table 1 continued

Gene (variant) Year Comparison OR (95% CI) Publication

bias/

heterogeneity

No. of

studies

Cases/control Power

OR,

1.5

Power

OR,

1.2

FPRP values

at prior

probability

of 0.001

Ref.

OR

1.5

OR

1.2

RAD51 (135G[C) 2010 Recessive

model

1.35 (1.05–1.74) 0.16/0.06 9 26,444 (13,241/13,203) 0.722 0.182 0.963 0.991 [156]

SULT1A1 (R213H) 2011 Dominant

model

1.13 (1.01–1.26) No/0.001 16 23,445 (9,881/13,564) 1.0 0.86 0.965 0.970 [157]

For Asians 2011 Dominant

model

1.11 (0.94–1.32) No/0.004 7 23,445 (4,966/7,770) – – – – [157]

TGFB1 (rs1982073) 2010 Dominant

model

1.05 (1.01–1.10) 0.675/0.195 30 47,817 (20,401/27,416) 1.0 1.0 0.975 0.975 [99]

TNF (rs1800629) 2009 Recessive

model

1.10 (1.04–1.17) \0.05/0.45 11 23,095 (10,184/12,911) 1.0 0.997 0.996 0.996 [158]

TOX3 (rs3803662) 2010 Dominant

model

1.15 (1.00–1.32) 0.249/0.000 8 29,505 (25,828/36,177) 1.0 0.727 0.979 0.985 [67]

TP53 (rs1042522) 2010 Dominant

model

0.90 (0.82–0.99) 0.015/0.001 39 55,720 (26,041/29,679) 1.0 0.943 0.968 0.970 [27]

(rs17878362) 2010 Ins/Ins vs. Del/

Del

1.81 (1.30–2.52) No/0.30 9 5,310 (2,715/2,595) 0.133 0.007 0.768 0.983 [159]

TYMS (TSER) 2010 2R/2R vs. non-

2R/

non-2R

1.30 (1.10–1.53) NA/0.74 6 6,141 (2,718/3,423) 0.957 0.168 0.625 0.905 [69]

VDR (FokI) 2009 ff vs. FF 1.14 (1.03–1.27) No/0.006 13 29,880 (12,464/17,416) 1.0 0.824 0.946 0.955 [71]

WDR79

(rs2287499)

2007 GG vs. CC 1.60 (1.04–2.47) NA/0.44 2 6,059 (2,692/3,367) 0.385 0.097 0.989 0.997 [31]

(rs2287498) 2007 CT vs. CC 1.15 (1.00–1.32) NA/0.19 2 6,009 (2,655/3,354) 1.0 0.727 0.979 0.985 [31]

XRCC1

(Arg399Gln)

2009 Gln/Gln vs.

Arg/Arg

1.14 (1.01–1.29) 0.006/0.002 37 46,386 (22,481/23, 905) 1.0 0.792 0.974 0.979 [100]

For Asians

(Arg399Gln)

2009 Recessive

model

1.26 (0.96–1.64) 0.006/0.571 4 3,275 (1,573/1,702) – – – – [100]

(-77T[C) 2011 CC vs. TT 2.55 (1.11–5.86) No/No 7 6,065 (2,888/3,177) 0.106 0.038 0.996 0.999 [160]

XRCC3

(rs1799794)

2010 Dominant

model

1.09 (1.01–1.17) No/0.15 4 12,866 (6,303/6,563) 1.0 0.996 0.945 0.945 [28]

(rs1799796) 2010 GG vs. AA 0.86 (0.77–0.96) No/0.21 4 12,952 (6,270/6,682) 1.0 0.713 0.878 0.910 [28]

(rs861539) 2010 TT vs. CC 1.07 (1.01–1.14) 0.116/0.200 23 43,028 (20,791/22,237) 1.0 1.0 0.979 0.979 [161]

For Chinese

(rs861539)

2010 TT vs. CC 0.57 (0.34–0.98) NA/0.819 3 2,328 (1,216/1,112) 0.285 0.085 0.993 0.998 [161]

ZNF350 (D35D) 2006 TC vs. TT 1.10 (1.01–1.20) NA/0.09 2 2,950 (1,554/1,396) 1.0 0.975 0.969 0.970 [86]

(S472P) 2006 Ser/Pro vs.

Ser/Ser

1.24 (1.05–1.48) NA/0.01 2 5,101 (2,719/2,382) 0.983 0.358 0.946 0.980 [86]

NA not available, No significant publication bias/heterogeneity was not found

Breast Cancer Res Treat (2011) 127:309–324 313

123

The other 50 analyses reported ORs higher than 1.0,

with a mean value of 1.08 (median, 1.04; range, 1.00–1.64)

(Table S1).

Significant gene-variant breast cancer associations

in genome-wide association studies (GWASs)

By searching PubMed and checking the web site (A Cat-

alog of Genome-Wide Association Studies), eight GWASs

were identified [11, 87–93], which included 25 SNPs, with

a threshold of P values \1 9 10-5 (Table 2; Fig. 3). Of

these 25 SNPs, 14 were in 9 genes, six were near a gene or

genes, and five were intergenic. These 25 associations

reported ORs [ 1.0, with a mean value of 1.19 (median,

1.17; range, 1.04–1.43). At a prior probability level of

0.001 and statistical power to detect an OR of 1.5/1.2, 20

SNPs maintained noteworthy associations with breast

cancer (FPRP \ 0.2) (Table 2).

Discussion

We found that 31.7% of all the gene-variant breast cancer

associations from published meta-analyses and pooled

analyses were statistically significant. With continuous

progress in genotyping technologies, a large number of

genetic variants can be tested. Therefore, many false-

positive results are likely to be published due to the widely

used significance threshold of P \ 0.05. Therefore, we also

used FPRP methodology, as did Dong et al. in their sys-

tematic review. FPRP is based not only on the observed

P value but also on both the power and the prior probability

of the hypothesis. By using this method, we can integrate

the prior knowledge of the tested variants [13]. Thomas and

Clayton [94] suggested that the prior probability will usu-

ally exceed 0.001. Thus, at a prior probability of 0.001,

only 10 of the associations were noteworthy, and these 10

may be true associations.

CASP8CCND1CHEK2CTLA4CYP19ERCC2ESR1ESR1ESR2FGFR2FGFR2FGFR2GSTM1GSTT1HER2HRASIGFBP3IL1BLSP1MAP3K1MDM2MMP-2MTHFRNBS1NBS1NOS3NOS3PORRAD51SULT1A1TGFB1TNFTNRC9TP53TP53TYMSVDRWDR79WDR79XRCC1XRCC1XRCC3XRCC3XRCC3ZNF350ZNF350

Gene

D302HG870A*1100delC+49G > A(TTTA)10Lys751Glnrs2234693rs1801132rs4986938rs2981582rs1219648rs2420946null/presentnull/presentIle655Valrare allelesA-202Crs1143627rs3817198rs889312SNP309-1306 C/TC667T657del58360G>C894 G>T-786T>CGly5Gly135G>CR213Hrs1982073rs1800629rs3803662rs1042522rs17878362TSERFokIrs2287499rs2287498Arg399Gln-77T>Crs1799794rs1799796rs861539D35DS472P

Polymorphism

2010201020082010201020102010201020102010201020102010201020101997201020102010201020102010201020062010201020102007201020102010200920102010201020102009200720072009201020102010201020062006

Year

39,10921,08236,90924,45216,10429,89726,92012,50526,85891,89058,90315,32946,28136,58424,0421,63178,8112,70833,92059,97725,9793,24338,8682,40610,1179,5073,3261,91526,44423,44547,81723,09529,50555,7205,3106,14129,8806,0596,00946,3861,99912,86612,95243,0282,9505,101

Subjects

0.87 (0.82, 0.92)1.09 (1.01, 1.18)2.40 (1.80, 3.20)1.24 (1.16, 1.32)1.52 (1.12, 2.06)1.13 (1.03, 1.24)0.95 (0.90, 1.00)0.92 (0.85, 0.99)0.94 (0.89, 0.99)1.48 (1.36, 1.61)1.50 (1.36, 1.65)1.52 (1.38, 1.68)1.10 (1.04, 1.16)1.11 (1.03, 1.20)1.10 (1.01, 1.20)2.70 (2.14, 3.40)1.06 (1.01, 1.11)1.40 (1.17, 1.67)1.06 (1.04, 1.08)1.10 (1.07, 1.13)1.06 (1.00, 1.12)1.27 (1.10, 1.47)1.12 (1.01, 1.24)3.13 (1.40, 7.00)0.75 (0.57, 0.98)1.29 (1.07, 1.56)0.66 (0.46, 0.94)1.58 (1.04, 2.41)1.35 (1.05, 1.74)1.13 (1.01, 1.26)1.05 (1.00, 1.10)1.10 (1.03, 1.17)1.15 (1.00, 1.32)0.90 (0.82, 0.99)1.81 (1.30, 2.52)1.30 (1.10, 1.53)1.14 (1.02, 1.27)1.60 (1.04, 2.47)1.15 (1.00, 1.32)1.14 (1.01, 1.29)2.55 (1.11, 5.86)1.09 (1.02, 1.17)0.86 (0.77, 0.96)1.07 (1.00, 1.14)1.10 (1.01, 1.20)1.24 (1.04, 1.48)

ratio (95% CI)Odds

1.2 1 1.5

Fig. 2 Risk of breast cancer

and genetic variants, as

evaluated in meta-analyses and

pooled analyses with significant

summary risk estimates

314 Breast Cancer Res Treat (2011) 127:309–324

123

Table 2 Genetic variants of breast cancer risk obtained from GWASs

Gene (variant) Year Variants OR (95% CI) MAFb P valuec Cases/control Power

OR,

1.5

Power

OR,

1.2

FPRP values

at prior

probability of

0.001

Ref.

OR,

1.5

OR,

1.2

C6orf97 2009 rs2046210 1.29

(1.21–1.37)

0.37 2 9 10-15 3,027 (1,505/

1,522)

1.0 0.018 0.000 0.000 [88]

2010 rs3757318 1.3

(1.17–1.46)

0.07 3 9 10-6 8,556 (3,659/

4,897)

0.992 0.088 0.009 0.096 [87]

FGFR2 2010 rs2981579 1.43

(1.35–1.53)

0.42 4 9 10-31 8,556 (3,659/

4,897)

0.866 0.000 0.000 0.000 [87]

2007 rs2981582 1.26

(1.23–1.30)

0.38 2 9 10-76 754 (390/364) 1.0 0.06 0.000 0.000 [11]

2010 rs1219648 1.32

(1.22–1.42)

0.42 2 9 10-13 8,428 (2,702/

5,726)

1.0 0.005 0.000 0.000 [92]

LSP1 2007 rs3817198 1.07

(1.04–1.11)

0.3 3 9 10-9 754 (390/364) 1.0 1.0 0.232 0.232 [11]

2010 rs909116 1.17

(1.10–1.24)

0.53 7 9 10-7 8,556 (3,659/

4,897)

1.0 0.803 0.000 0.000 [87]

SLC4A7 2010 rs4973768 1.16

(1.10–1.24)

0.47 6 9 10-7 8,556 (3,659/

4,897)

1.0 0.840 0.013 0.015 [87]

MRPS30 2010 rs7716600 1.24

(1.14–1.34)

0.23 7 9 10-7 8,428 (2,702/

5,726)

1.0 0.204 0.000 0.000 [92]

RNF146 2008 rs2180341 1.41

(1.25–1.59)

0.21 3 9 10-8 548 (249/299) 0.844 0.004 0.000 0.005 [89]

TOX3 2007 rs3803662 1.28

(1.21–1.35)

0.27 6 9 10-19 13,145 (1,599/

11,546)

1.0 0.017 0.000 0.000 [90]

2010 rs4784227 1.24

(1.20–1.29)

0.24 1 9 10-28 4,157 (2,073/

2,084)

1.0 0.11 0.000 0.000 [93]

ZNF365 2010 rs10995190 1.16

(1.10–1.22)

0.85 5 9 10-15 8,556 (3,659/

4,897)

1.0 0.906 0.000 0.000 [87]

RAD51L1 2009 rs999737 1.06

(1.01–1.14)

0.76 2 9 10-7 2,287 (1,145/

1,142)

1.0 1.0 0.991 0.991 [91]

(CDKN2A, CDKN2B) a 2010 rs1011970 1.09

(1.04–1.14)

0.17 3 9 10-8 8,556 (3,659/

4,897)

1.0 1.0 0.142 0.142 [87]

(MAP3K1)a 2010 rs889312 1.22

(1.14–1.30)

0.28 5 9 10-9 8,556 (3,659/

4,897)

1.0 0.305 0.000 0.000 [87]

2009 rs16886165 1.23

(1.12–1.35)

0.15 5 9 10-7 2,287 (1,145/

1,142)

1.0 0.302 0.013 0.042 [91]

(MYEOV, CCND1,

ORAOV1,

FGF19, FGF4,

FGF3)a

2010 rs614367 1.15

(1.10–1.20)

0.15 3 9 10-15 8,556 (3,659/

4,897)

1.0 0.975 0.000 0.000 [87]

(ZMIZ1)a 2010 rs704010 1.07

(1.03–1.11)

0.39 4 9 10-9 8,556 (3,659/

4,897)

1.0 1.0 0.232 0.232 [87]

(ANKRD16, FBXO18)a 2010 rs2380205 1.06

(1.02–1.10)

0.57 5 9 10-7 8,556 (3,659/

4,897)

1.0 1.0 0.672 0.672 [87]

Intergenic 2007 rs13281615 1.08

(1.05–1.11)

0.4 5 9 10-12 754 (390/364) 1.0 1.0 0.000 0.000 [11]

Intergenic 2009 rs11249433 1.16

(1.09–1.24)

0.39 7 9 10-10 2,287 (1,145/

1,142)

1.0 0.84 0.013 0.015 [91]

Breast Cancer Res Treat (2011) 127:309–324 315

123

In the review by Dong et al. [13], only 20 variants (in 20

genes) were reported to have significant associations with

breast cancer risk, and only three of the 20, were consid-

ered noteworthy. However, we found 46 significant asso-

ciations between the variants and breast cancer risk. Of

these 46 associations, 10 were noteworthy at a prior

probability of 0.001 and statistical power to detect an OR

of 1.5. Interestingly, 10 variants, which were reported to be

significantly associated with breast cancer risk in the

review by Dong et al., were not significantly associated

with breast cancer risk in the more recent meta-analyses (or

pooled analyses); these were ATP1B2 (-8852T?C),

COMT (Met108/158Val), CYP17 (rs4919687), CYP17

(rs4919682), CYP1A1 (A2455G), CYP1B1 (Leu432Val),

GATA3 (rs570613), PGR (PR, PROGINS), PTGS2

(Ex10?837), and TGFBR1 (*6A). Only 10 variants were

still reported to be significantly associated with breast cancer

risk based on the updated data (from novel meta-analyses or

pooled analyses); these were CASP8 (D302H) [22], CHEK2

(*1100delC) [95], CYP19 ((TTTA)10) [42], IGFBP3 (A-

202C) [96], NBS1 (657del5) [97], POR (Gly5Gly) [98],

TGFB1 (Leu10Pro) [99], WDR79 (Arg68Gly and Phe150-

Phe) [31], and XRCC1 (Arg399Gln) [100]. In addition,

another 36 polymorphisms (in 28 genes) that were signifi-

cantly associated with breast cancer risk in recent meta-

analyses or pooled analyses were added (Table 1). The

reasons for the discrepancies between results from recent

meta- and pooled analyses and previous published results

can be described as follows: (1) as technology advances,

fewer and fewer genotype errors in the subsequent individual

studies were produced; (2) with the lapse of time, the sample

sizes in the individual studies are increasing, with results of

higher power than that of previous studies; and (3) the sample

sizes in recent meta- and pooled analyses are bigger than

those of previous meta- and pooled analyses. Therefore, the

recent results are more plausible statistically than previous

results.

Recently, results from GWASs on breast cancer have

become available. In those studies, 25 SNPs (Table 2) were

significantly associated with breast cancer in eight

GWASs. Interestingly, in the eight GWASs, only three

SNPs overlapped with previous association studies, two in

FGFR2 (rs1219648 and rs2981582) and one in LSP1

(rs3817198).

Different genetic associations between polymorphisms

and breast cancer risk in various ethnic groups have been

found for many genetic variants. For example, statistically

significant associations were found between GSTT1 (null/

present), SULT1A1 (R213H), and XRCC1 (Arg399Gln)

and breast cancer risk in the overall population, but in

Asians, only non-significant associations were found

(Table 1). In particular, XRCC3 (rs861539) was signifi-

cantly associated with breast cancer in the overall popu-

lation (OR: 1.07, CI: 1.01–1.14), but an inverse (protective)

association was found in the Chinese population (OR: 0.57,

CI: 0.34–0.98) (Table 1). AURKA (T91a), ASP8 (-652

6N del), ERCC2 (Asp312Asn), HSD17B1 (Ser312Gly),

and MTR (A2756G) were not statistically associated with

breast cancer risk in overall population effects. However,

significant protective associations were found in Asians

(AURKA and ERCC2), Chinese (ASP8), Caucasians

(HSD17B1), and Europeans (MTR) (Table S1). In contrast,

CYP1A1 (A2455G), GSTP1 (Ile105Val), and MDM2

(SNP309) were not found to be significantly associated

with breast cancer risk, but significant associations were

found between breast cancer and CYP1A1 (A2455G),

GSTP1 (Ile105Val), and MTR (A2756G) in Caucasians,

Chinese, and European populations (with ORs and CI: 2.08

(1.19–3.61), 1.27 (1.01–1.61), and 0.90 (0.82–0.97)),

respectively.

Table 2 continued

Gene (variant) Year Variants OR (95% CI) MAFb P valuec Cases/control Power

OR,

1.5

Power

OR,

1.2

FPRP values

at prior

probability of

0.001

Ref.

OR,

1.5

OR,

1.2

Intergenic 2007 rs13387042 1.2

(1.14–1.26)

0.5 1 9 10-13 13,145 (1,599/

11,546)

1.0 0.50 0.000 0.000 [90]

Intergenic 2007 rs981782 1.04

(1.01–1.08)

0.53 9 9 10-6 754 (390/364) 1.0 1.0 0.977 0.977 [11]

Intergenic 2010 rs1562430 1.17

(1.10–1.25)

0.58 6 9 10-7 8,556 (3,659/

4,897)

1.0 0.773 0.003 0.004 [87]

a The SNP is near the gene/genes in bracketsb Minor Allele Frequency (MAF) in Controlsc The P values are all less than 1 9 10-5

316 Breast Cancer Res Treat (2011) 127:309–324

123

With regard to the 46 variants with statistically signifi-

cant associations, inconsistent results were reported in

many studies of the contribution of genetic polymorphisms

to breast cancer risk, e.g., of the 29 studies of TP53 (codon

72) concerning European populations, 13 ORs were greater

than 1.0 and 16 ORs were less than 1.0 [27]. In this case,

further studies are needed to clarify the contribution of

TP53 (codon 72) to breast cancer risk. From this perspec-

tive, further studies are also needed to clarify the associa-

tions between some other variants and breast cancer risk

due to many inconsistent results reported, including GSTT1

(null/present), HER2 (Ile655Val), MDM2 (SNP309),

MTHFR (C667T), RAD51 (135G[C), SULT1A1 (R213H),

TNF (rs1800629), TP53 (rs1042522 and rs17878362), VDR

(FokI), XRCC1 (Arg399Gln), and XRCC3 (rs861539).

CASP8, which was identified as noteworthy, belongs to

many key pathways, including p53 signaling, apoptosis,

and cancer [101]. CASP8 plays a important role in apop-

tosis [102]. The decreased risk for breast cancer with

CASP8 Asp302His was revealed in the pooled analysis and

has been replicated in a recent association study [103]. This

finding was also addressed in a recent meta-analysis [22].

CHEK2, located on chr22 and identified in our review as

being noteworthy at an OR of 1.5, is associated with DNA

repair, the cell cycle pathway and the p53 signaling pathway

[101]. It was reported that the protein encoded by CHEK2 is

a cell-cycle checkpoint regulator. And this protein is a

putative tumor suppressor containing a forkhead-associated

protein interaction domain, which is essential for activation

in response to DNA damage [104, 105]. When activated, the

CHEK2 protein inhibits CDC25C phosphatase, preventing

entry into mitosis, and also stabilizes the tumor suppressor

protein p53. Furthermore, the protein encoded by CHEK2

interacts with and phosphorylates BRCA1 [106].

CTLA4 is a member of the immunoglobulin super-

family, and this gene encodes a protein transmitting an

inhibitory signal to T cells. CTLA4 in humans maps to

chromosome 2q33 [107]. Schneider et al. concluded that

CTLA4 increases T-cell motility and overrides the T-cell

receptor-induced stop signal required for stable conjugate

formation between T cells and antigen-presenting cells

[108]. A series of SNPs were discovered by Ueda et al.

[109]. The polymorphisms in CTLA4 were tested by

Zhernakova et al. for association with type I diabetes [110].

The FGFR2 (fibroblast growth factor receptor 2) gene is

located on chromosome 10q26 and contains at least 22

exons [111]. FGFR2 is involved in the MAPK signaling

pathway, the endocytosis pathway, regulation of the actin

cytoskeleton pathway, pathways in cancer, and the prostate

cancer pathway [101]. The protein encoded by FGFR2 is a

member of the fibroblast growth factor receptor family,

which contributes to cell growth, invasiveness, motility,

and angiogenesis [112]. FGFR2 is over-expressed in both

breast cancer cell lines [113] and breast tumor tissues

[114]. Koziczak et al. identified the mechanism by which

FGFR promotes the proliferation of breast cancer cells

[115]. Moffa and Ethier concluded that aberrant expression

of FGFR2 in breast cancer cells results in sustained

C6orf97C6orf97FGFR2FGFR2FGFR2LSP1LSP1SLC4A7MRPS30RNF146TOX3TOX3ZNF365RAD51L1(CDKN2A,CDKN2B)(MAP3K1)(MAP3K1)(MYEOV,CCND1...)(ZMIZ1)(ANKRD16,FBXO18)IntergenicIntergenicIntergenicIntergenicIntergenic

Gene

rs2046210rs3757318rs2981579rs2981582rs1219648rs3817198rs909116rs4973768rs7716600rs2180341rs3803662rs4784227rs10995190rs999737rs1011970rs889312rs16886165rs614367rs704010rs2380205rs13281615rs11249433rs13387042rs981782rs1562430

Polymorphism

2009201020102007201020072010201020102008200720102010200920102010200920102010201020072009200720072010

Year

3,0278,5568,5567548,4287548,5568,5568,42854813,1454,1578,5562,2878,5568,5562,2878,5568,5568,5567542,28713,1457548,556

Subjects

1.29 (1.21, 1.37)1.30 (1.16, 1.46)1.43 (1.34, 1.53)1.26 (1.22, 1.30)1.32 (1.23, 1.42)1.07 (1.03, 1.11)1.17 (1.10, 1.24)1.16 (1.09, 1.24)1.24 (1.15, 1.34)1.41 (1.25, 1.59)1.28 (1.21, 1.35)1.24 (1.19, 1.29)1.16 (1.10, 1.22)1.06 (0.99, 1.14)1.09 (1.04, 1.14)1.22 (1.14, 1.30)1.23 (1.12, 1.35)1.15 (1.10, 1.20)1.07 (1.03, 1.11)1.06 (1.02, 1.10)1.08 (1.05, 1.11)1.16 (1.09, 1.24)1.20 (1.14, 1.26)1.04 (1.00, 1.08)1.17 (1.10, 1.25)

ratio (95% CI)Odds

1.2 1 1.5

Fig. 3 Risk of breast cancer

and genetic variants, as

evaluated in GWAS studies

Breast Cancer Res Treat (2011) 127:309–324 317

123

activation of signal transduction leading to transformation

[116].

HRAS was an additional variant identified in our review

as being noteworthy. HRAS is located on chromosome

11p15.5 and belongs to the Ras oncogene family. The

proteins encoded by these genes were reported to function

in signal transduction pathways. Defects in this gene are

implicated in many cancers, such as bladder cancer, fol-

licular thyroid cancer, and oral squamous cell carcinoma

[117, 118].

The IL1B (IL-1beta) protein, located at 5q11.2, is a

member of the interleukin 1 cytokine family [119]. This

cytokine is a pivotal mediator of the inflammatory

response, and it is involved in cell proliferation, differen-

tiation, and apoptosis. Microenvironmental IL1B and

IL1-alpha are required for the in vivo angiogenesis and

invasiveness of different tumor cells [120]. Furthermore,

antiangiogenic effects of IL1RN were reported, suggesting

a possible therapeutic role in cancer. Ben-Sasson et al.

[121] reported that IL1B signaling in T cells markedly

induces robust and durable primary and secondary CD4

responses.

LSP1 was mapped to 11p15.5 by May et al. [122] They

reported that LSP1 is an intracellular Ca2?- and F-actin

binding protein. Liu et al. [123] reported that LSP1

expressed in endothelium regulates neutrophil transendo-

thelial migration. Interestingly, the association between

rs3817198 (in LSP1) and breast cancer risk was noteworthy

in the genetic association studies but not noteworthy in the

GWAS. This finding may be due to the fact that to obtain

higher statistical power, the smaller the effect size is, the

bigger the sample size we required. With a small sample

size (390 cases and 364 controls) and a small OR: 1.07

(1.04–1.11) in the GWAS, the statistical power may be

small. Whereas with a bigger sample size (33,920 cases

and 35,671 controls) and a small OR: 1.06 (1.04–1.08) in

the genetic association studies, clearly, the statistical power

may be bigger than that in the GWAS. Thus, we think that

the noteworthy association between rs3817198 and breast

cancer is plausible.

MAP3K1, located at 5q11.2, is a serine/threonine

kinase. This kinase plays a key role in a network of

phosphorylating enzymes integrating cellular responses to

a number of mitogenic and metabolic stimuli [124]. Lu

et al. reported that MAP3K1 is an upstream activator of

ERK and JNK, and MAP3K1 is also an E3 ligase through

its PHD domain [125]. Roy et al. [126] demonstrated that

IFNG induces MEKK1 to up-regulate cellular responses.

C6orf97 is located at 6q25.1. The function of this gene

and its encoded protein is not known. Several GWASs have

suggested that the region around this gene is involved in

breast cancer [87, 88] and bone mineral density [127], but

no link to this specific gene has been found.

SLC4A7 is located on 3p22 [128]. Reiners et al. [129]

demonstrated that SLC4A7 and the USH2 (Usher syn-

drome 2A) proteins are partners in the supramolecular USH

protein network in the retina and inner ear. The visual and

auditory systems require H? disposal, which is mediated

by the sodium bicarbonate cotransporter NBC3 that is

encoded by the SLC4A7 gene [130].

MRPS30, located on 5p12-q11, is one of more than 70

protein components of mitochondrial ribosomes that are

encoded by the nuclear genome [131]. MRPS30 orthologs

were identified by Koc et al. in mouse, Drosophila, and

C. elegans but not in yeast or E. coli [132].

RNF146 is located on 6q22.1-q22.33 and contains 5

exons. It is up-regulated in the inferior temporal lobes of

Alzheimer disease patients compared with healthy controls

[133].

TOX3, located on 16q12.1 [90], belongs to the large and

diverse family of HMG-box proteins [134]. TOX3 is dif-

ferentially expressed in patients who experienced breast

cancer relapse to bone versus those patients who experi-

enced relapse elsewhere in the body [135]. Antoniou et al.

found that rs3803662 in TOX3 has multiplicative effects

on breast cancer risk in BRCA1 or BRCA2 mutation car-

riers [136].

The ZNF365 gene is located on 10q21.2 and contains 15

exons spanning approximately 300 kb. This gene has a

complex pattern of alternative splicing and transcriptional

start sites [137]. Wang et al. [138] found that ZNF365A

localizes to the centrosome throughout the cell cycle in

several human cell lines. A mutation analysis demonstrated

that the centrosomal localization of ZNF365A requires 2

coiled-coil subdomains but not the C-terminal zinc finger

structure. Over-expression of ZNF365A leads to abnormal

mitosis. A mutant form of ZNF365A lacking the C-termi-

nal region disruptes the localization of gamma-tubulin to

the centrosome.

As for the 99 variants with non-significant association

results, further studies are also needed to clarify the asso-

ciations with breast cancer risk, especially for those variants

with inconsistent results across different populations,

including AURKA (T91a), CYP1A1 (A2455G), ERCC2

(Asp312Asn), GSTP1 (Ile105Val), HSD17B1 (Ser312Gly),

MDM2 (SNP309), MTR (A2756G), and PGR (rs10895068).

We found 46 statistically significant associations

between gene variants and breast cancer, of which 10

(in eight genes) were considered noteworthy at a prior

probability of 0.001, and 99 statistically non-significant

gene-variant breast cancer associations. In addition, 25

statistically significant gene-variant breast cancer associa-

tions were reported in the GWASs, but only three SNPs

(rs2981582 and rs1219648 in FGFR2 and rs3817198 in

LSP1) were significantly associated with breast cancer risk

in both genetic association studies and GWASs, suggesting

318 Breast Cancer Res Treat (2011) 127:309–324

123

that the capacity of GWASs to detect disease susceptibility

genes is limited. To clarify the contribution of genetic

variants to breast cancer risk in different populations, fur-

ther studies are needed with large sample sizes and better

study design.

There are some limitations in this review, such as (1) the

associations were confined to those summarized in meta-

analyses, pooled analyses, and GWASs. Some individual

studies with larger sample sizes may have been missed, and

may possess more power to find significant associations

than some meta-analyses, pooled analyses, and GWASs;

(2) in the individual meta-analysis, the publication bias was

always addressed by the funnel plot and Egger’s tests. But

it is argued that the funnel plot can be misleading and

Egger’s test does not really test publication bias [139].

Actually, some studies with results of non-significant

associations tend to not be published. Furthermore, to

address the publication bias issue effectively in this review

was not realistic. Therefore, to some extent, the file drawer

problem may inevitably arise. In this case, even a small

number of studies lost ‘‘in the file drawer’’ can result in a

significant bias.

Undoubtedly, genetic association studies have been

helpful in the discovery of genetic susceptibility genes. In

particular, with the help of GWASs, many susceptibility

genes of several diseases have been confirmed. However,

several deficiencies of genetic association studies (or

GWASs) were observed, including their high cost and a

poor capacity to detect genetic susceptibility genes. Per-

haps the GWAS era will gradually wane, and with the

advance in technologies and price reductions, next-gener-

ation sequencing can open a new era in the discovery of

genetic susceptibility disease genes [140–144].

In summary, in this review of genetic association stud-

ies, 31.7% of the gene-variant breast cancer associations

were significant, and 21.7% of these significant associa-

tions were noteworthy. However, in GWASs, 80% of the

significant associations were noteworthy.

Acknowledgments This study was supported by The ‘‘Eleventh

Five-Year’’ Science and Technology Support Plan of the Ministry of

Science and Technology of China (MSTC, 2009BA180B00) and a

grant from the Natural Science Foundation of Zhejiang Province

(NSFZJ, Y2090081). We thank Dr. Iain Bruce for valuable comments

and English editing.

Conflict of interest None.

References

1. Parkin DM, Bray F, Ferlay J, Pisani P (2005) Global cancer

statistics, 2002. Cancer J Clin 55:74–108

2. Jemal A, Siegel R, Ward E, Hao YP, Xu JQ et al (2008) Cancer

statistics, 2008. Cancer J Clin 58:71–96. doi:10.3322/ca.2007.

0010

3. Pharoah PDP, Dunning AM, Ponder BAJ, Easton DF (2004)

Association studies for finding cancer-susceptibility genetic

variants. Nat Rev Cancer 4:850–860. doi:10.1038/nrc1476

4. Thompson D, Easton D (2004) The genetic epidemiology of

breast cancer genes. J Mammary Gland Biol Neoplasia 9:

221–236

5. Oldenburg RA, Meijers-Heijboer H, Cornelisse CJ, Devilee P

(2007) Genetic susceptibility for breast cancer: how many more

genes to be found? Crit Rev Oncol Hematol 63:125–149. doi:

10.1016/j.critrevonc.2006.12.004

6. Ioannidis JPA, Ntzani EE, Trikalinos TA, Contopoulos-Ioanni-

dis DG (2001) Replication validity of genetic association stud-

ies. Nat Genet 29:306–309

7. Burton PR, Hansell AL, Fortier I, Manolio TA, Khoury MJ et al

(2009) Size matters: just how big is BIG? Quantifying realistic

sample size requirements for human genome epidemiology. Int J

Epidemiol 38:263–273. doi:10.1093/Ije/Dyn147

8. Gordon D, Finch SJ (2005) Factors affecting statistical power in

the detection of genetic association. J Clin Invest 115:1408–1418

9. Glass GV (1976) Primary, secondary and meta-analysis of

research. Educ Res 5:3–8. doi:10.3102/0013189X005010003

10. Blettner M, Sauerbrei W, Schlehofer B, Scheuchenpflug T,

Friedenreich C (1999) Traditional reviews, meta-analyses and

pooled analyses in epidemiology. Int J Epidemiol 28:1–9

11. Easton DF, Pooley KA, Dunning AM, Pharoah PD, Thompson

D et al (2007) Genome-wide association study identifies novel

breast cancer susceptibility loci. Nature 447:1087–1093. doi:

10.1038/nature05887

12. Dunning AM, Healey CS, Pharoah PDP, Teare MD, Ponder BAJ

et al (1999) A systematic review of genetic polymorphisms

and breast cancer risk. Cancer Epidemiol Biomarkers Prev

8:843–854

13. Dong LM, Potter JD, White E, Ulrich CM, Cardon LR et al

(2008) Genetic susceptibility to cancer—the role of polymor-

phisms in candidate genes. JAMA 299:2423–2436

14. Wang WYS, Barratt BJ, Clayton DG, Todd JA (2005) Genome-

wide association studies: theoretical and practical concerns. Nat

Rev Genet 6:109–118. doi:10.1038/nrg1522

15. Hirschhorn JN, Daly MJ (2005) Genome-wide association

studies for common diseases and complex traits. Nat Rev Genet

6:95–108. doi:10.1038/nrg1521

16. McCarthy MI, Abecasis GR, Cardon LR, Goldstein DB, Little J

et al (2008) Genome-wide association studies for complex traits:

consensus, uncertainty and challenges. Nat Rev Genet 9:

356–369. doi:10.1038/nrg2344

17. Hindorff LA, Sethupathy P, Junkins HA, Ramos EM, Mehta JP

et al (2009) Potential etiologic and functional implications of

genome-wide association loci for human diseases and traits.

Proc Natl Acad Sci USA 106:9362–9367. doi:10.1073/pnas.

0903103106

18. Hindorff LA, Junkins HA, Hall PN, Mehta JP, Manolio TA. A

catalog of published genome-wide association studies. Available

at: www.genome.gov/gwastudies

19. Cochran WG (1954) The combination of estimates from dif-

ferent experiments. Biometrics 10:101–129

20. Egger M, Smith GD, Schneider M, Minder C (1997) Bias in

meta-analysis detected by a simple, graphical test. Br Med J

315:629–634

21. Wacholder S, Chanock S, Garcia-Closas M, El ghormli L,

Rothman N (2004) Assessing the probability that a positive

report is false: an approach for molecular epidemiology studies.

J Natl Cancer Inst 96:434–442. doi:10.1093/jnci/djh075

22. Sergentanis TN, Economopoulos KP (2010) Association of two

CASP8 polymorphisms with breast cancer risk: a meta-analysis.

Breast Cancer Res Treat 120:229–234. doi:10.1007/s10549-

009-0471-5

Breast Cancer Res Treat (2011) 127:309–324 319

123

23. Li N, Dong J, Hu Z, Shen H, Dai M (2010) Potentially func-

tional polymorphisms in ESR1 and breast cancer risk: a meta-

analysis. Breast Cancer Res Treat 121:177–184. doi:10.1007/

s10549-009-0532-9

24. Yu KD, Rao NY, Chen AX, Fan L, Yang C et al (2011) A

systematic review of the relationship between polymorphic sites

in the estrogen receptor-beta (ESR2) gene and breast cancer risk.

Breast Cancer Res Treat 126:37–45. doi:10.1007/s10549-010-

0891-2

25. Wang Z, Cui D, Lu W (2010) NBS1 8360G[C polymorphism is

associated with breast cancer risk: a meta-analysis. Breast

Cancer Res Treat 123:557–561. doi:10.1007/s10549-010-0772-8

26. Yao L, Fang F, Zhong Y, Yu L (2010) The association between

two polymorphisms of eNOS and breast cancer risk: a meta-

analysis. Breast Cancer Res Treat 124:223–227. doi:10.1007/

s10549-010-0800-8

27. Zhang Z, Wang M, Wu D, Tong N, Tian Y (2010) P53 codon 72

polymorphism contributes to breast cancer risk: a meta-analysis

based on 39 case-control studies. Breast Cancer Res Treat

120:509–517. doi:10.1007/s10549-009-0480-4

28. Qiu LX, Mao C, Yao L, Yu KD, Zhan P et al (2010) XRCC3 50-UTR and IVS5-14 polymorphisms and breast cancer suscepti-

bility: a meta-analysis. Breast Cancer Res Treat 122:489–493.

doi:10.1007/s10549-009-0726-1

29. Lu PH, Wei MX, Yang J, Liu X, Tao GQ et al (2011) Associ-

ation between two polymorphisms of ABCB1 and breast cancer

risk in the current studies: a meta-analysis. Breast Cancer Res

Treat 125:537–543. doi:10.1007/s10549-010-1033-6

30. The Breast Cancer Association Consortium (2006) Commonly

studied single-nucleotide polymorphisms and breast cancer:

results from the Breast Cancer Association Consortium. J Natl

Cancer Inst 98:1382–1396. doi:10.1093/jnci/djj374

31. Garcia-Closas M, Kristensen V, Langerod A, Qi Y, Yeager M

et al (2007) Common genetic variation in TP53 and its flanking

genes, WDR79 and ATP1B2, and susceptibility to breast cancer.

Int J Cancer 121:2532–2538. doi:10.1002/ijc.22985

32. Hao YJ, Montiel R, Li BH, Huang EY, Zeng LW et al (2010)

Association between androgen receptor gene CAG repeat poly-

morphism and breast cancer risk: a meta-analysis. Breast Cancer

Res Treat 124:815–820. doi:10.1007/s10549-010-0907-y

33. Ding H, Mao C, Li SM, Liu Q, Lin L et al (2011) Lack of

association between ATM C.1066-6T[G mutation and breast

cancer risk: a meta-analysis of 8,831 cases and 4,957 controls.

Breast Cancer Res Treat 125:473–477. doi:10.1007/s10549-010-

0977-x

34. Gao LB, Pan XM, Sun H, Wang X, Rao L et al (2010) The

association between ATM D1853N polymorphism and breast

cancer susceptibility: a meta-analysis. J Exp Clin Cancer Res 29.

doi:10.1186/1756-9966-29-117

35. Lu PH, Wei MX, Si SP, Liu X, Shen W et al (2010) Association

between polymorphisms of the ataxia telangiectasia mutated

gene and breast cancer risk: evidence from the current studies.

Breast Cancer Res Treat. doi:10.1007/s10549-010-1081-y

36. Sun HM, Bai J, Chen F, Jin Y, Yu Y et al (2011) Lack of an

association between AURKA T91A polymorphisms and breast

cancer: a meta-analysis involving 32,141 subjects. Breast Can-

cer Res Treat 125:175–179. doi:10.1007/s10549-010-0936-6

37. Gaudet MM, Milne RL, Cox A, Camp NJ, Goode EL et al

(2009) Five polymorphisms and breast cancer risk: results from

the Breast Cancer Association Consortium. Cancer Epidemiol

Biomarkers Prev 18:1610–1616. doi:10.1158/1055-9965.EPI-

08-0745

38. Qiu LX, Yao L, Xue K, Zhang JA, Mao C et al (2010) BRCA2

N372H polymorphism and breast cancer susceptibility: a meta-

analysis involving 44,903 subjects. Breast Cancer Res Treat

123:487–490. doi:10.1007/s10549-010-0767-5

39. Wang GY, Lu CQ, Zhang RM, Hu XH, Luo ZW (2008) The

E-cadherin gene polymorphism 160C?A and cancer risk: a

HuGE review and meta-analysis of 26 case-control studies. Am

J Epidemiol 167:7–14. doi:10.1093/aje/kwm264

40. Mao C, Wang XW, Qiu LX, Liao RY, Ding H et al (2010) Lack

of association between catechol-O-methyltransferase Val108/

158Met polymorphism and breast cancer risk: a meta-analysis of

25,627 cases and 34,222 controls. Breast Cancer Res Treat

121:719–725. doi:10.1007/s10549-009-0650-4

41. Yao L, Fang F, Wu Q, Yang Z, Zhong Y et al (2010) No

association between CYP17 T-34C polymorphism and breast

cancer risk: a meta-analysis involving 58,814 subjects. Breast

Cancer Res Treat 122:221–227. doi:10.1007/s10549-009-0679-4

42. Ma X, Qi X, Chen C, Lin H, Xiong H et al (2010) Association

between CYP19 polymorphisms and breast cancer risk: results

from 10,592 cases and 11,720 controls. Breast Cancer Res Treat

122:495–501. doi:10.1007/s10549-009-0693-6

43. Sergentanis TN, Economopoulos KP (2010) Four polymorphisms

in cytochrome P450 1A1 (CYP1A1) gene and breast cancer risk: a

meta-analysis. Breast Cancer Res Treat 122:459–469. doi:

10.1007/s10549-009-0694-5

44. Qiu LX, Yao L, Mao C, Yu KD, Zhan P et al (2010) Lack of

association of CYP1A2-164 A/C polymorphism with breast

cancer susceptibility: a meta-analysis involving 17,600 subjects.

Breast Cancer Res Treat 122:521–525. doi:10.1007/s10549-

009-0731-4

45. Yao L, Fang F, Wu Q, Zhong Y, Yu L (2010) No association

between CYP1B1 Val432Leu polymorphism and breast cancer

risk: a meta-analysis involving 40,303 subjects. Breast Cancer

Res Treat 122:237–242. doi:10.1007/s10549-009-0689-2

46. Economopoulos KP, Sergentanis TN (2010) Three polymor-

phisms in cytochrome P450 1B1 (CYP1B1) gene and breast

cancer risk: a meta-analysis. Breast Cancer Res Treat

122:545–551. doi:10.1007/s10549-009-0728-z

47. Pabalan N, Francisco-Pabalan O, Sung L, Jarjanazi H, Ozcelik H

(2010) Meta-analysis of two ERCC2 (XPD) polymorphisms,

Asp312Asn and Lys751Gln, in breast cancer. Breast Cancer Res

Treat 124:531–541. doi:10.1007/s10549-010-0863-6

48. Zhang Z, Xue H, Gong W, Wang M, Yuan L et al (2009) FAS

promoter polymorphisms and cancer risk: a meta-analysis based

on 34 case-control studies. Carcinogenesis 30:487–493. doi:

10.1093/carcin/bgp016

49. Liu Y, Wen QJ, Yin Y, Lu XT, Pu SH et al (2009) FASLG

polymorphism is associated with cancer risk. Eur J Cancer

45:2574–2578. doi:10.1016/j.ejca.2009.04.001

50. Garcia-Closas M, Troester MA, Qi Y, Langerod A, Yeager M

et al (2007) Common genetic variation in GATA-binding pro-

tein 3 and differential susceptibility to breast cancer by estrogen

receptor alpha tumor status. Cancer Epidemiol Biomarkers Prev

16:2269–2275. doi:10.1158/1055-9965.EPI-07-0449

51. Johnatty SE, Couch FJ, Fredericksen Z, Tarrell R, Spurdle AB

et al (2009) No evidence that GATA3 rs570613 SNP modifies

breast cancer risk. Breast Cancer Res Treat 117:371–379. doi:

10.1007/s10549-008-0257-1

52. Hu J, Zhou GW, Wang N, Wang YJ (2010) GPX1 Pro198Leu

polymorphism and breast cancer risk: a meta-analysis. Breast

Cancer Res Treat 124:425–431. doi:10.1007/s10549-010-0841-z

53. Sergentanis TN, Economopoulos KP (2010) GSTT1 and GSTP1

polymorphisms and breast cancer risk: a meta-analysis. Breast

Cancer Res Treat 121:195–202. doi:10.1007/s10549-009-0520-0

54. Zhao T, Lv J, Zhao J, Nzekebaloudou M (2009) Hypoxia-

inducible factor-1alpha gene polymorphisms and cancer risk: a

meta-analysis. J Exp Clin Cancer Res 28:159. doi:

10.1186/1756-9966-28-159

55. Yao L, Cao LH, Qiu LX, Yu L (2010) The association between

HSD17B1 Ser312Gly polymorphism and breast cancer risk: a

320 Breast Cancer Res Treat (2011) 127:309–324

123

meta-analysis including 31,053 subjects. Breast Cancer Res

Treat 123:577–580. doi:10.1007/s10549-010-0784-4

56. Chen X, Guan J, Song Y, Chen P, Zheng H et al (2008) IGF-I

(CA) repeat polymorphisms and risk of cancer: a meta-analysis.

J Hum Genet 53:227–238. doi:10.1007/s10038-007-0241-3

57. Li L, Huang X, Huo K (2010) IGFBP3 polymorphisms and risk

of cancer: a meta-analysis. Mol Biol Rep 37:127–140. doi:

10.1007/s11033-009-9552-0

58. Liu X, Wang Z, Yu J, Lei G, Wang S (2010) Three polymor-

phisms in interleukin-1beta gene and risk for breast cancer: a

meta-analysis. Breast Cancer Res Treat 124:821–825. doi:

10.1007/s10549-010-0910-3

59. Yu KD, Di GH, Fan L, Chen AX, Yang C et al (2010) Lack of

an association between a functional polymorphism in the

interleukin-6 gene promoter and breast cancer risk: a meta-

analysis involving 25,703 subjects. Breast Cancer Res Treat

122:483–488. doi:10.1007/s10549-009-0706-5

60. Yu KD, Chen AX, Yang C, Fan L, Huang AJ et al (2010) The

associations between two polymorphisms in the interleukin-10

gene promoter and breast cancer risk. Breast Cancer Res Treat.

doi:10.1007/s10549-010-1133-3

61. Zhou P, Du L, Lv G, Yu X, Gu Y et al (2010) Current evidence on

the relationship between four polymorphisms in the matrix

metalloproteinases (MMP) gene and breast cancer risk: a meta-

analysis. Breast Cancer Res Treat. doi:10.1007/s10549-010-1294-0

62. Qiu LX, Yao L, Mao C, Chen B, Zhan P et al (2010) Lack of

association between MnSOD Val16Ala polymorphism and

breast cancer risk: a meta-analysis involving 58,448 subjects.

Breast Cancer Res Treat 123:543–547. doi:10.1007/s10549-

010-0777-3

63. Chen Y, Pei J (2011) Possible risk modifications in the associ-

ation between MnSOD Ala-9Val polymorphism and breast

cancer risk: subgroup analysis and evidence-based sample size

calculation for a future trial. Breast Cancer Res Treat

125:495–504. doi:10.1007/s10549-010-0978-9

64. Qi XW, Zhang F, Yang XH, Fan LJ, Zhang Y et al (2010)

Transforming growth factor-beta 1 polymorphisms and breast

cancer risk: a meta-analysis based on 27 case-control studies.

Breast Cancer Res Treat 122:273–279. doi:10.1007/s10549-

010-0847-6

65. Colleran G, McInerney N, Rowan A, Barclay E, Jones AM et al

(2010) The TGFBR1*6A/9A polymorphism is not associated

with differential risk of breast cancer. Breast Cancer Res Treat

119:437–442. doi:10.1007/s10549-009-0395-0

66. Shen C, Sun H, Sun D, Xu L, Zhang X et al (2010) Polymor-

phisms of tumor necrosis factor-alpha and breast cancer risk: a

meta-analysis. Breast Cancer Res Treat. doi:10.1007/s10549-

010-1184-5

67. Chen MB, Wu XY, Shen W, Wei MX, Li C et al (2010)

Association between polymorphisms of trinucleotide repeat

containing 9 gene and breast cancer risk: evidence from 62,005

subjects. Breast Cancer Res Treat. doi:10.1007/s10549-010-

1114-6

68. Hu Z, Li X, Yuan R, Ring BZ, Su L (2010) Three common TP53

polymorphisms in susceptibility to breast cancer, evidence from

meta-analysis. Breast Cancer Res Treat 120:705–714. doi:

10.1007/s10549-009-0488-9

69. Wang J, Wang B, Bi J, Di J (2010) The association between two

polymorphisms in the TYMS gene and breast cancer risk: a

meta-analysis. Breast Cancer Res Treat. doi:10.1007/s10549-

010-1314-0

70. Yao L, Qiu LX, Yu L, Yang Z, Yu XJ et al (2010) The asso-

ciation between TA-repeat polymorphism in the promoter region

of UGT1A1 and breast cancer risk: a meta-analysis. Breast

Cancer Res Treat 122:879–882. doi:10.1007/s10549-010-0742-1

71. Raimondi S, Johansson H, Maisonneuve P, Gandini S (2009)

Review and meta-analysis on vitamin D receptor polymor-

phisms and cancer risk. Carcinogenesis 30:1170–1180. doi:

10.1093/carcin/bgp103

72. Qiu LX, Zhang J, Zhu XD, Zheng CL, Sun S et al (2010) The

p21 Ser31Arg polymorphism and breast cancer risk: a meta-

analysis involving 51,236 subjects. Breast Cancer Res Treat

124:475–479. doi:10.1007/s10549-010-0858-3

73. Yang DS, Sung HJ, Woo OH, Park KH, Woo SU et al (2010)

Association of a progesterone receptor gene ?331G/A poly-

morphism with breast cancer risk: a meta-analysis. Cancer

Genet Cytogenet 196:194–197. doi:10.1016/j.cancergencyto.

2009.10.005

74. Yu KD, Chen AX, Yang C, Qiu LX, Fan L et al (2010) Current

evidence on the relationship between polymorphisms in the

COX-2 gene and breast cancer risk: a meta-analysis. Breast

Cancer Res Treat 122:251–257. doi:10.1007/s10549-009-

0688-3

75. Qiu LX, Zhang J, Li WH, Zhang QL, Yu H et al (2010) Lack of

association between methylenetetrahydrofolate reductase gene

A1298C polymorphism and breast cancer susceptibility. Mol

Biol Rep. doi:10.1007/s11033-010-0361-2

76. Lu MP, Wang F, Qiu JR (2010) Methionine synthase A2756G

polymorphism and breast cancer risk: a meta-analysis involving

18,953 subjects. Breast Cancer Res Treat 123:213–217. doi:

10.1007/s10549-010-0755-9

77. Hu J, Zhou GW, Wang N, Wang YJ (2010) MTRR A66G

polymorphism and breast cancer risk: a meta-analysis. Breast

Cancer Res Treat 124:779–784. doi:10.1007/s10549-010-0892-1

78. Zhang J, Qiu LX, Wang ZH, Wang JL, He SS et al (2010) NAT2

polymorphisms combining with smoking associated with breast

cancer susceptibility: a meta-analysis. Breast Cancer Res Treat

123:877–883. doi:10.1007/s10549-010-0807-1

79. Yuan W, Xu L, Chen W, Wang L, Fu Z et al (2011) Evidence on

the association between NQO1 Pro187Ser polymorphism and

breast cancer risk in the current studies. Breast Cancer Res Treat

125:467–472

80. Gu D, Wang M, Zhang Z, Chen J (2010) Lack of association

between the hOGG1 Ser326Cys polymorphism and breast can-

cer risk: evidence from 11 case-control studies. Breast Cancer

Res Treat 122:527–531. doi:10.1007/s10549-009-0723-4

81. Tang C, Chen N, Wu M, Yuan H, Du Y (2009) Fok1 poly-

morphism of vitamin D receptor gene contributes to breast

cancer susceptibility: a meta-analysis. Breast Cancer Res Treat

117:391–399. doi:10.1007/s10549-008-0262-4

82. Qiu LX, Wang K, Yang S, Mao C, Zhao L et al (2010) Current

evidences on vascular endothelial growth factor polymorphisms

and breast cancer susceptibility. Mol Biol Rep. doi:10.1007/

s11033-010-0579-z

83. Francisco G, Menezes PR, Eluf-Neto J, Chammas R (2008) XPC

polymorphisms play a role in tissue-specific carcinogenesis: a

meta-analysis. Eur J Hum Genet 16:724–734. doi:10.1038/

ejhg.2008.6

84. Li H, Ha TC, Tai BC (2009) XRCC1 gene polymorphisms and

breast cancer risk in different populations: a meta-analysis.

Breast 18:183–191. doi:10.1016/j.breast.2009.03.008

85. Yu KD, Chen AX, Qiu LX, Fan L, Yang C et al (2010) XRCC2

Arg188His polymorphism is not directly associated with breast

cancer risk: evidence from 37,369 subjects. Breast Cancer Res

Treat 123:219–225. doi:10.1007/s10549-010-0753-y

86. Garcia-Closas M, Egan KM, Newcomb PA, Brinton LA, Titus-

Ernstoff L et al (2006) Polymorphisms in DNA double-strand

break repair genes and risk of breast cancer: two population-

based studies in USA and Poland, and meta-analyses. Hum

Genet 119:376–388. doi:10.1007/s00439-006-0135-z

Breast Cancer Res Treat (2011) 127:309–324 321

123

87. Turnbull C, Ahmed S, Morrison J, Pernet D, Renwick A et al

(2010) Genome-wide association study identifies five new breast

cancer susceptibility loci. Nat Genet 42:504–507. doi:10.1038/

ng.586

88. Zheng W, Long J, Gao YT, Li C, Zheng Y et al (2009) Genome-

wide association study identifies a new breast cancer susceptibility

locus at 6q25.1. Nat Genet 41:324–328. doi:10.1038/ng.318

89. Gold B, Kirchhoff T, Stefanov S, Lautenberger J, Viale A et al

(2008) Genome-wide association study provides evidence for a

breast cancer risk locus at 6q22 33. Proc Natl Acad Sci USA

105:4340–4345. doi:10.1073/pnas.0800441105

90. Stacey SN, Manolescu A, Sulem P, Rafnar T, Gudmundsson J

et al (2007) Common variants on chromosomes 2q35 and 16q12

confer susceptibility to estrogen receptor-positive breast cancer.

Nat Genet 39:865–869. doi:10.1038/ng2064

91. Thomas G, Jacobs KB, Kraft P, Yeager M, Wacholder S et al

(2009) A multistage genome-wide association study in breast

cancer identifies two new risk alleles at 1p11.2 and 14q24.1

(RAD51L1). Nat Genet 41:579–584. doi:10.1038/ng.353

92. Li J, Humphreys K, Heikkinen T, Aittomaki K, Blomqvist C

et al (2010) A combined analysis of genome-wide association

studies in breast cancer. Breast Cancer Res Treat. doi:10.1007/

s10549-010-1172-9

93. Long J, Cai Q, Shu XO, Qu S, Li C et al (2010) Identification of

a functional genetic variant at 16q12 1 for breast cancer risk:

results from the Asia Breast Cancer Consortium. PLoS Genet

6:e1001002. doi:10.1371/journal.pgen.1001002

94. Thomas DC, Clayton DG (2004) Betting odds and genetic

associations. J Natl Cancer Inst 96:421–423. doi:10.1093/jnci/

djh094

95. Weischer M, Bojesen SE, Ellervik C, Tybjaerg-Hansen A,

Nordestgaard BG (2008) CHEK2*1100delC genotyping for

clinical assessment of breast cancer risk: meta-analyses of

26,000 patient cases and 27,000 controls. J Clin Oncol

26:542–548. doi:10.1200/JCO.2007.12.5922

96. Qiu LX, Yao L, Yuan H, Mao C, Chen B et al (2010) IGFBP3

A-202C polymorphism and breast cancer susceptibility: a meta-

analysis involving 33,557 cases and 45,254 controls. Breast

Cancer Res Treat 122:867–871. doi:10.1007/s10549-010-0739-9

97. Steffen J, Nowakowska D, Niwinska A, Czapczak D, Kluska A

et al (2006) Germline mutations 657del5 of the NBS1 gene

contribute significantly to the incidence of breast cancer in Cen-

tral Poland. Int J Cancer 119:472–475. doi:10.1002/Ijc.21853

98. Haiman CA, Setiawan VW, Xia LY, Le Marchand L, Ingles SA

et al (2007) A variant in the cytochrome p450 oxidoreductase gene

is associated with breast cancer risk in African Americans. Cancer

Res 67:3565–3568. doi:10.1158/0008-5472.CAN-06-4801

99. Qiu LX, Yao L, Mao C, Chen B, Zhan P et al (2010) TGFB1 L10P

polymorphism is associated with breast cancer susceptibility:

evidence from a meta-analysis involving 47,817 subjects. Breast

Cancer Res Treat 123:563–567. doi:10.1007/s10549-010-0781-7

100. Huang Y, Li L, Yu L (2009) XRCC1 Arg399Gln, Arg194Trp

and Arg280His polymorphisms in breast cancer risk: a meta-

analysis. Mutagenesis 24:331–339. doi:10.1093/mutage/gep013

101. KEGG: http://www.genome.jp/kegg/pathway.html

102. Grenet J, Teitz T, Wei T, Valentine V, Kidd VJ (1999) Structure

and chromosome localization of the human CASP8 gene. Gene

226:225–232

103. Sigurdson AJ, Bhatti P, Doody MM, Hauptmann M, Bowen L et al

(2007) Polymorphisms in apoptosis- and proliferation-related

genes, ionizing radiation exposure, and risk of breast cancer among

US radiologic technologists. Cancer Epidemiol Biomarkers Prev

16:2000–2007. doi:10.1158/1055-9965.Epi-07-0282

104. Peng CY, Graves PR, Thoma RS, Wu ZQ, Shaw AS et al (1997)

Mitotic and G(2) checkpoint control: regulation of 14-3-3

protein binding by phosphorylation of Cdc25C on serine-216.

Science 277:1501–1505

105. Matsuoka S, Huang MX, Elledge SJ (1998) Linkage of ATM to

cell cycle regulation by the Chk2 protein kinase. Science

282:1893–1897

106. Entrez Gene: http://www.ncbi.nlm.nih.gov/gene/11200

107. Dariavach P, Mattei MG, Golstein P, Lefranc MP (1988) Human

Ig superfamily Ctla-4 gene—chromosomal localization and

identity of protein-sequence between murine and human Ctla-4

cytoplasmic domains. Eur J Immunol 18:1901–1905

108. Schneider H, Downey J, Smith A, Zinselmeyer BH, Rush C et al

(2006) Reversal of the TCR stop signal by CTLA-4. Science

313:1972–1975. doi:10.1126/science.1131078

109. Ueda H, Howson JMM, Esposito L, Heward J, Snook H et al

(2003) Association of the T-cell regulatory gene CTLA4 with

susceptibility to autoimmune disease. Nature 423:506–511. doi:

10.1038/Nature01621

110. Zhernakova A, Eerligh P, Barrera P, Weseloy JZ, Huizinga TWJ

et al (2005) CTLA4 is differentially associated with autoim-

mune diseases in the Dutch population. Hum Genet 118:58–66.

doi:10.1007/s00439-005-0006-z

111. Ingersoll RG, Paznekas WA, Tran AK, Scott AF, Jiang G et al

(2001) Fibroblast growth factor receptor 2 (FGFR2): genomic

sequence and variations. Cytogenet Cell Genet 94:121–126

112. Ricol D, Cappellen D, El Marjou A, Gil-Diez-de-Medina S,

Girault JM et al (1999) Tumour suppressive properties of

fibroblast growth factor receptor 2-IIIb in human bladder cancer.

Oncogene 18:7234–7243

113. Tannheimer SL, Rehemtulla A, Ethier SP (2000) Characteriza-

tion of fibroblast growth factor receptor 2 overexpression in the

human breast cancer cell line SUM-52PE. Breast Cancer Res

2:311–320

114. Adnane J, Gaudray P, Dionne CA, Crumley G, Jaye M et al

(1991) BEK and FLG, two receptors to members of the FGF

family, are amplified in subsets of human breast cancers.

Oncogene 6:659–663

115. Koziczak M, Holbro T, Hynes NE (2004) Blocking of FGFR

signaling inhibits breast cancer cell proliferation through

downregulation of D-type cyclins. Oncogene 23:3501–3508.

doi:10.1038/sj.onc.1207331

116. Moffa AB, Ethier SP (2007) Differential signal transduction of

alternatively spliced FGFR2 variants expressed epithelial in

human mammary cells. J Cell Physiol 210:720–731. doi:

10.1002/Jcp.20880

117. Tamir A, Fawzi AB, Northup JK (1990) Unique guanine-

nucleotide binding-properties of the human placental Gtp-

binding protein Gp. Biochemistry 29:6947–6954

118. Sakai E, Rikimaru K, Ueda M, Matsumoto Y, Ishii N et al (1992)

The P53 tumor-suppressor gene and Ras oncogene mutations in

oral squamous-cell carcinoma. Int J Cancer 52:867–872

119. March CJ, Mosley B, Larsen A, Cerretti DP, Braedt G et al

(1985) Cloning, sequence and expression of two distinct human

interleukin-1 complementary DNAs. Nature 315:641–647

120. Voronov E, Shouval DS, Krelin Y, Cagnano E, Benharroch D

et al (2003) IL-1 is required for tumor invasiveness and angio-

genesis. Proc Natl Acad Sci USA 100:2645–2650. doi:

10.1073/pnas.0437939100

121. Ben-Sasson SZ, Hu-Li J, Quiel J, Cauchetaux S, Ratner M et al

(2009) IL-1 acts directly on CD4 T cells to enhance their anti-

gen-driven expansion and differentiation. Proc Natl Acad Sci

USA 106:7119–7124. doi:10.1073/pnas.0902745106

122. May W, Korenberg JR, Chen XN, Lunsford L, Wood WJ

et al (1993) Human lymphocyte-specific Pp52 gene is a

member of a highly conserved dispersed family. Genomics

15:515–520

322 Breast Cancer Res Treat (2011) 127:309–324

123

123. Liu LX, Cara DC, Kaur J, Raharjo E, Mullaly SC et al (2005)

LSP1 is an endothelial gatekeeper of leukocyte transendothelial

migration. J Exp Med 201:409–418. doi:10.1084/Jem.20040830

124. Vinik BS, Kay ES, Fiedorek FT (1995) Mapping of the Mek

kinase gene (Mekk) to mouse chromosome-13 and human-

chromosome-5. Mamm Genome 6:782–783

125. Lu ZM, Xu SC, Joazeiro C, Cobb MH, Hunter T (2002) The

PHD domain of MEKK1 acts as an E3 ubiquitin ligase and

mediates ubiquitination and degradation of ERK1/2. Mol Cell

9:945–956

126. Roy SK, Hu JB, Meng QJ, Xia Y, Shapiro PS et al (2002)

MEKK1 plays a critical role in activating the transcription factor

gene expression in C/EBP-beta-dependent response to IFN-

gamma. Proc Natl Acad Sci USA 99:7945–7950

127. Styrkarsdottir U, Halldorsson BV, Gretarsdottir S, Gudbjartsson

DF, Walters GB et al (2008) Multiple genetic loci for bone

mineral density and fractures. N Engl J Med 358:2355–2365

128. Pushkin A, Abuladze N, Lee I, Newman D, Hwang J et al (1999)

Cloning, tissue distribution, genomic organization, and func-

tional characterization of NBCS, a new member of the sodium

bicarbonate cotransporter family. J Biol Chem 274:16569–

16575

129. Reiners J, van Wijk E, Marker T, Zimmermann U, Jurgens K

et al (2005) Scaffold protein harmonin (USH1C) provides

molecular links between Usher syndrome type 1 and type 2.

Hum Mol Genet 14:3933–3943. doi:10.1093/Hmg/Ddi417

130. Bok D, Galbraith G, Lopez I, Woodruff M, Nusinowitz S et al

(2003) Blindness and auditory impairment caused by loss of the

sodium bicarbonate cotransporter NBC3. Nat Genet 34:313–319.

doi:10.1038/Ng1176

131. Kenmochi N, Suzuki T, Uechi T, Magoori M, Kuniba M et al

(2001) The human mitochondrial ribosomal protein genes: map-

ping of 54 genes to the chromosomes and implications for human

disorders. Genomics 77:65–70. doi:10.1006/geno.2001.6622

132. Koc EC, Burkhart W, Blackburn K, Moseley A, Spremulli LL

(2001) The small subunit of the mammalian mitochondrial

ribosome—identification of the full complement of ribosomal

proteins present. J Biol Chem 276:19363–19374

133. von Rotz RC, Kins S, Hipfel R, von der Kammer H, Nitsch RM

(2005) The novel cytosolic RING finger protein dactylidin is up-

regulated in brains of patients with Alzheimer’s disease. Eur J

Neurosci 21:1289–1298. doi:10.1111/j.1460-9568.2005.03977.x

134. O’Flaherty E, Kaye J (2003) TOX defines a conserved subfamily

of HMG-box proteins. BMC Genomics 4:13

135. Smid M, Wang YX, Klijn JGM, Sieuwerts AM, Zhang Y et al

(2006) Genes associated with breast cancer metastatic to bone.

J Clin Oncol 24:2261–2267. doi:10.1200/Jco.2005.03.8802

136. Antoniou AC, Spurdle AB, Sinilnikova OM, Healey S, Pooley

KA et al (2008) Common breast cancer-predisposition alleles

are associated with breast cancer risk in BRCA1 and BRCA2

mutation carriers. Am J Hum Genet 82:937–948. doi:10.1016/

j.ajhg.2008.02.008

137. Gianfrancesco F, Esposito T, Ombra MN, Forabosco P, Mani-

nchedda G et al (2003) Identification of a novel gene and a

common variant associated with uric acid nephrolithiasis in a

Sardinian genetic isolate. Am J Hum Genet 72:1479–1491

138. Wang Q, Du XL, Meinkoth J, Hirohashi Y, Zhang HT et al

(2006) Characterization of Su48, a centrosome protein essential

for cell division. Proc Natl Acad Sci USA 103:6512–6517. doi:

10.1073/pnas.0601682103

139. Lau J, Ioannidis JPA, Terrin N, Schmid CH, Olkin I (2006)

Evidence based medicine—the case of the misleading funnel

plot. Br Med J 333:597–600

140. Shendure J, Ji HL (2008) Next-generation DNA sequencing. Nat

Biotechnol 26:1135–1145. doi:10.1038/nbt1486

141. Pfeifer GP, Hainaut P (2011) Next-generation sequencing:

emerging lessons on the origins of human cancer. Curr Opin

Oncol 23:62–68. doi:10.1097/CCO.0b013e3283414d00

142. Metzker ML (2010) Sequencing technologies—the next gener-

ation. Nat Rev Genet 11:31–46. doi:10.1038/nrg2626

143. Mardis ER (2008) The impact of next-generation sequencing

technology on genetics. Trends Genet 24:133–141

144. Meyerson M, Gabriel S, Getz G (2010) Advances in under-

standing cancer genomes through second-generation sequenc-

ing. Nat Rev Genet 11:685–696. doi:10.1038/nrg2841

145. Sergentanis TN, Economopoulos KP (2010) Cyclin D1 G870A

polymorphism and breast cancer risk: a meta-analysis com-

prising 9,911 cases and 11,171 controls. Mol Biol Rep. doi:

10.1007/s11033-010-0639-4

146. Zheng J, Yu X, Jiang L, Xiao M, Bai B et al (2010) Association

between the cytotoxic T-lymphocyte antigen 4 ?49G[A poly-

morphism and cancer risk: a meta-analysis. BMC Cancer

10:522. doi:10.1186/1471-2407-10-522

147. Jia C, Cai Y, Ma Y, Fu D (2010) Quantitative assessment of the effect

of FGFR2 gene polymorphism on the risk of breast cancer. Breast

Cancer Res Treat 124:521–528. doi:10.1007/s10549-010-0872-5

148. Qiu LX, Yuan H, Yu KD, Mao C, Chen B et al (2010) Gluta-

thione S-transferase M1 polymorphism and breast cancer sus-

ceptibility: a meta-analysis involving 46,281 subjects. Breast

Cancer Res Treat 121:703–708. doi:10.1007/s10549-009-0636-2

149. Lu S, Wang Z, Liu H, Hao X (2010) HER2 Ile655Val poly-

morphism contributes to breast cancer risk: evidence from 27

case–control studies. Breast Cancer Res Treat 124:771–778. doi:

10.1007/s10549-010-0886-z

150. Weston A, Godbold JH (1997) Polymorphisms of H-ras-1 and

p53 in breast cancer and lung cancer: a meta-analysis. Environ

Health Perspect 105(Suppl 4):919–926

151. Chen MB, Li C, Shen WX, Guo YJ, Shen W et al (2010)

Association of a LSP1 gene rs3817198T[C polymorphism with

breast cancer risk: evidence from 33,920 cases and 35,671

controls. Mol Biol Rep. doi:10.1007/s11033-010-0603-3

152. Lu PH, Yang J, Li C, Wei MX, Shen W et al (2010) Association

between mitogen-activated protein kinase kinase kinase 1

rs889312 polymorphism and breast cancer risk: evidence from

59,977 subjects. Breast Cancer Res Treat. doi:10.1007/s10549-

010-1151-1

153. Economopoulos KP, Sergentanis TN (2010) Differential effects

of MDM2 SNP309 polymorphism on breast cancer risk along

with race: a meta-analysis. Breast Cancer Res Treat

120:211–216. doi:10.1007/s10549-009-0467-1

154. Qi XW, Ma XY, Yang XH, Fan LJ, Zhang Y et al (2010)

Methylenetetrahydrofolate reductase polymorphisms and breast

cancer risk: a meta-analysis from 41 studies with 16,480 cases

and 22,388 controls. Breast Cancer Res Treat 123:499–506. doi:

10.1007/s10549-010-0773-7

155. Hao Y, Montiel R, Huang Y (2010) Endothelial nitric oxide

synthase (eNOS) 894 G[T polymorphism is associated with

breast cancer risk: a meta-analysis. Breast Cancer Res Treat

124:809–813. doi:10.1007/s10549-010-0833-z

156. Wang Z, Dong H, Fu Y, Ding H (2010) RAD51 135G[C

polymorphism contributes to breast cancer susceptibility: a

meta-analysis involving 26,444 subjects. Breast Cancer Res

Treat 124:765–769. doi:10.1007/s10549-010-0885-0

157. Sun YY, Zang ZH, Xu XH, Zhang ZL, Zhong L et al (2011) The

association of SULT1A1 codon 213 polymorphism and breast

cancer susceptibility: meta-analysis from 16 studies involving

23,445 subjects. Breast Cancer Res Treat 125:215–219. doi:

10.1007/s10549-010-0953-5

158. Fang F, Yao L, Yu XJ, Yu L, Wu Q (2010) TNF alpha-308 G/A

polymorphism is associated with breast cancer risk: a meta-

Breast Cancer Res Treat (2011) 127:309–324 323

123

analysis involving 10,184 cases and 12,911 controls. Breast

Cancer Res Treat 122:267–271. doi:10.1007/s10549-009-0698-1

159. Hu Z, Li X, Qu X, He Y, Ring BZ et al (2010) Intron 3 16 bp

duplication polymorphism of TP53 contributes to cancer sus-

ceptibility: a meta-analysis. Carcinogenesis 31:643–647. doi:

10.1093/carcin/bgq018

160. Liu L, Yuan P, Wu C, Zhang X, Guo H et al (2011) A functional

-77T[C polymorphism in XRCC1 is associated with risk of

breast cancer. Breast Cancer Res Treat 125:479–487. doi:

10.1007/s10549-010-0959-z

161. Economopoulos KP, Sergentanis TN (2010) XRCC3 Thr241Met

polymorphism and breast cancer risk: a meta-analysis. Breast

Cancer Res Treat 121:439–443. doi:10.1007/s10549-009-0562-3

324 Breast Cancer Res Treat (2011) 127:309–324

123